Next Article in Journal
Pedestrian Crossings as a Means of Reducing Conflicts between Cyclists and Pedestrians in Shared Spaces
Next Article in Special Issue
Hydropower Planning in Combination with Batteries and Solar Energy
Previous Article in Journal
Remediation of Acid Mine Drainage in the Haizhou Open-Pit Mine through Coal-Gangue-Loaded SRB Experiments
Previous Article in Special Issue
Photovoltaic Systems Based on Average Current Mode Control: Dynamical Analysis and Chaos Suppression by Using a Non-Adaptive Feedback Outer Loop Controller
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exergo–Economic and Parametric Analysis of Waste Heat Recovery from Taji Gas Turbines Power Plant Using Rankine Cycle and Organic Rankine Cycle

1
Department of Mechanical Engineering, Faculty of Engineering, Karabuk University, Karabuk 78050, Turkey
2
Training and Workshop Center, University of Technology-Iraq, Ministry of Higher Education & Scientific Research, Baghdad 10066, Iraq
*
Authors to whom correspondence should be addressed.
Sustainability 2023, 15(12), 9376; https://doi.org/10.3390/su15129376
Submission received: 8 May 2023 / Revised: 4 June 2023 / Accepted: 5 June 2023 / Published: 9 June 2023
(This article belongs to the Special Issue Advanced Technologies Applied to Renewable Energy)

Abstract

:
This study focused on exergo–conomic and parametric analysis for Taji station in Baghdad. This station was chosen to reduce the emission of waste gases that pollute the environment, as it is located in a residential area, and to increase the production of electric power, since for a long time, Iraq has been a country that has suffered from a shortage of electricity. The main objective of this work is to integrate the Taji gas turbine’s power plant, which is in Baghdad, with the Rankine cycle and organic Rankine cycle to verify waste heat recovery to produce extra electricity and reduce emissions into the environment. Thermodynamic and exergoeconomic assessment of the combined Brayton cycle–Rankine cycle/Organic Rankin cycle (GSO CC) system, considering the three objective functions of the First- and Second-Law efficiencies and the total cost rates of the system, were applied. According to the findings, 258.2 MW of power is produced from the GSO CC system, whereas 167.3 MW of power is created for the Brayton cycle (BC) under the optimum operating conditions. It was demonstrated that the overall energy and exergy efficiencies, respectively, are 44.37% and 42.84% for the GSO CC system, while they are 28.74% and 27.75%, respectively, for the Brayton cycle. The findings indicate that the combustion chamber has the highest exergy degradation rate. The exergo–economic factor for the entire cycle is 37%, demonstrating that the cost of exergy destruction exceeds the cost of capital investment. Moreover, the cost of the energy produced by the GSO CC system is USD 9.03/MWh, whereas it is USD 8.24/MWh for BC. The results also indicate that the network of the GSO CC system decreases as the pressure ratio increases. Nonetheless, the GSO CC system’s efficiencies and costs increase with a rise in the pressure ratio until they reach a maximum and then decrease with further pressure ratio increases. The increase in the gas turbine inlet temperature and isentropic efficiency of the air compressor and gas turbine enhances the thermodynamic performance of the system; however, a further increase in these parameters increases the overall cost rates.

1. Introduction

Today, power is produced either through the combustion of fossil fuels in power cycles or via renewable energy sources, such as solar [1,2], wind, and water [3]. Increasing the efficiency of power generation systems is usually desirable for achieving energy sustainability and addressing the constraints of many conventional energy sources [4]. The use of fossil fuels to generate electricity was crucial to the success of the industrial revolution, and it has continued to play this role ever since, enabling important advances in society, science, and culture [5]. However, the widespread application of fossil fuels has contributed to the rising air pollution levels and CO2 emissions, leading to the greenhouse effect, and has exacerbated climate change generally [6]. Further, fossil fuel reserves are delimited and are eventually predicted to run out. As a result, it is crucial to lessen reliance on fossil fuels while boosting the effectiveness of energy infrastructure [7,8]. Combined cycles can assist in achieving such objectives by increasing net electrical power output, decreasing fuel consumption, and enhancing efficiency [9].
Exergy is a key term connected to the second law of thermodynamics and is defined as the greatest amount of useful work that can be extracted from a flow of matter or energy in a reference environment [10,11]. The engineering field known as exergoeconomics or thermo economics combines exergy analysis and economics [12]. This method assists the designer in discovering information that cannot be found with economic evaluation and regular energy analysis [13]. Modern Combined Cycle Gas Turbines (CCGTs) are a dependable method for producing energy with a thermal efficiency of at least 60% [14]. There have been several articles published over the years that provide energy, exergy, and economic analyses of the integration of ORC with the CCGT with the combined cycle power plant. According to Roy et al. [15], the integration of ORC with power plants resulted in lower emissions, improved thermal efficiency, and a lower power load. Parametric optimization of an SRC and ORC integrated with the GT for the variable turbine input pressure and temperature was carried out by Kose et al. [16]. According to the findings, R141b outperformed the other working fluids tested. Qui et al. [17] concluded that ORC improved energy efficiency, producing 14 kW of thermal output and 0.65 kW of electrical power output for combined heat and power application. Hemadri et al. [18] evaluated the influence of reheat on combined cycle performance in the context of repowering and found that integrating reheat in the ORC boosted specific thermal work output. Balanesscu et al. [19] conducted an analysis on the effect of ORC integration into the gas steam combined cycle and found that the thermal efficiency of the combined cycle improved by 1.1% due to decreased SFC (specific fuel consumption, kg/kWh). Exergo–environmental analysis of an integrated organic Rankine cycle for tri-generation purposes by Ahmadi et al. [20] showed that the input temperature, the compressor pressure ratio, and the isentropic efficiency of the gas turbine are the most influential factors in the system’s overall performance. Koç et al. [21] presented a GT-ORC hybrid power unit with and without HR for the Rankine portion. Several organic fluids were studied, and an exergy assessment was performed. The exergy efficiency of the organic Rankine cycle (ORC) with heat recovery is improved by approximately 13.5 percent, and the total power output is increased by almost 41.1 percent compared to the GT. Heat recovery for energy production was suggested by Salehi et al. [22] by combining the ORC with the Kalina cycle. Combined, the two organic fluids boost output from 250 kW (pure fluid) to 280 kW (mixed) under ideal circumstances. Cao et al. [23] performed a thermodynamic study of a gas turbine and ORC combined cycle with recuperators. They discovered that when the ORC turbine intake pressure increased, the power ratio of the GT-ORC combined cycle and thermal efficiency improved. Fernández-Guillamón et al. [24] describe the ORC under analysis considering a steady-state and two cycles, excluding and including irreversibility, and all of the data analyzed [25]. The summary in Table 1 illustrates the methods and relationships studied by many researchers.
Organic Rankine cycles (ORCs) were chosen because they are most suitable for converting low-quality thermal energy into electricity. In order to avoid the release of exhaust gases into the atmosphere, the main objective of this work is to integrate the Ghazi plant in Baghdad with the Rankine cycle and the organic Rankine cycle to increase the production of electricity quantities and reduce environmental emissions. In this novel power plant, the waste heat of the GT units is first utilized in the HRSG and later in the ORC as the heat source. The EES program is used to model the novel triple cycle. On the basis of the first and second laws, a comprehensive and exhaustive thermodynamic analysis and optimization of the GSO CC cycle was conducted with a focus on the irreversibility distribution within the plant. In addition, a parametric study was conducted to determine the effects of key parameters, such as compressor pressure ratio, compressor isentropic efficiency, turbine isentropic efficiency, gas turbine inlet temperature, boiler pressure, and condenser temperature, on cycle performance and cost.

2. Description of the GSOCC Cycle

A model is prepared and discussed for the steady-state analysis of the natural gas combined cycle, which is integrated with the organic Rankine cycle (GSO CC). It includes three main parts. The first part is the Brayton cycle, which consists of a gas turbine (GT), a combustion chamber (CC), and an air compressor (AC). The second part, which represents the Rankine cycle, consists of two pumps (P1 and P2), a deaerator, a condenser (CON1), a steam turbine (ST), and heat recovery steam generation (HRSG). The third and final part of the Organic Rankine cycle consists of a heat exchanger (HE), an organic Rankine turbine (ORT), a condenser (CON2), a pump (P3), and a heat recovery boiler (HRB) (see Figure 1). The principle of operation of the embedded system can be summarized as follows.
Air is compressed to operating pressure and heated as it enters an air compressor (AC). The air is then transported to the CC, reacting with the natural gas fuel to create high-pressure, high-temperature exhaust gases. Through the GT, the exhaust gases expand to produce mechanical power. The HRSG converts compressed water into steam at high temperatures using the temperature of the exhaust gases. To generate more mechanical power, the steam expands as it passes through the ST. The water is pressurized through the pump after entering the condenser, which turns all the vapor into saturated liquid.
The working principle of the organic Rankine cycle is similar to that of the Rankine cycle: the working fluid is pumped into a heat recovery boiler where it evaporates, passes through an expansion device (the turbine), then through a condenser heat exchanger where it is finally re-condensed.
The selection of the working fluid is essential. R123 was chosen as the working fluid in this current work because it is the most suitable fluid dynamically and economically. It is environmentally friendly with an acceptable value of ODP and low global warming value compared to other working fluids. Additionally, R123 falls into the category of dry working fluids with a higher critical temperature value (in relatively lower certainty ranges), which makes it a suitable working fluid in ORC applications. Moreover, R123 is the most suitable working fluid in ORCs for engine or gas turbine WHRT applications and is highly recommended by many researchers, given all of the above R123 environmental and technical advantages [26,27].

3. Mathematical Analysis

The EES program was used to solve the mass, energy, and exergy balance equations for all of the GSOCC system’s components. The basic input parameters of the GSOCC system are provided in Table 2, the parameters were adopted on the basis of the design values of the station, and the average weather rates in Baghdad were also adopted from the ambient temperature and relative humidity. These parameters selected depend on periodic statistics of the performance of the station’s units, where 4 units were stable in terms of their performance; therefore, they were chosen to conduct the analysis. Table 3 illustrates the thermodynamic modeling as well as the energetic and energetic relationships of the GSOCC system’s components. The general assumptions made for the simulation of the combined system are listed as follows:
  • All components of the combined system operate under steady-state conditions.
  • Compositions of air at the inlet of the AC are 79% N2 and 21% O2.
  • Natural gas is completely oxidized in the CC.
  • Ideal gas principles apply to the exhaust gases.
  • The CC is insulated completely.
  • The relations utilized to assess the thermodynamic performance of the triple combined cycle are shown in Table 4.
The cost balance and auxiliary equations for each part must be written as follows [28]:
e C ˙ e , k + C ˙ w , k = C ˙ q , k + i C ˙ i , k + Z ˙ k
C ˙ j = c j E ˙ j
where C is the cost rate (USD/h), and Z ˙ k represents the entire cost rate related to the capital investment and operation and maintenance costs component k. Table 5 presents the relevant parameters for the exergoeconomic assessment of the system, and Table 6 contains the cost balances and auxiliary equations for every system component. The total capital investment in the plant is given by [29,30]:
Z ˙ k = Z k × C R F × ϕ N × 3600
where PEC is the equipment purchase cost in US dollars, 𝜑 is the maintenance factor (1.06), and CRF is the Capital Recovery Factor, which can be calculated as follows:
C R F = i ( 1 + i ) n ( 1 + i ) n 1
The purchase equipment cost (PEC) for the GSO CC components is provided in Table 7.

4. Results and Discussion

This section describes the outcomes of thermodynamics, economic modeling, and the impact of different design factors on the performance of the GSO CC cycle. The suggested heat and power combined cycle system of this research consists of four 160 MW gas turbine cycles with exhaust gases directed into a single-pressure heat recovery steam generator (HRSG) to generate heat. The HRSG receives water, which exists as a superheated vapor. The superheated vapor is introduced into the ST to create additional electricity. Ultimately, a bottoming cycle of the ORC is added to boost system efficiency and maximize the benefit of heat losses [36,37], in which the gases ejected from the HRSG are sent to the ORC evaporator. The BC’s power production and thermal efficiency are verified after thermodynamic modeling of the GSO CC model, using the values presented in [21] as the standard [38]. Table 8 displays the results of the validation. Table 9 details the suggested model’s thermodynamic parameters, including each state’s mass, enthalpy, entropy, and exergy flow rates.
Table 10 presents the findings of an energy analysis performed on the components of the GSO CC model under the input parameters listed in Table 1. The Ẇnet of the BC model is 167.3 MW, with a First-Law efficiency ( η B C ) of 28.74% and a Second-Law efficiency ( Ψ B C ) of 27.74%. The RC/ORC model generates 258.2 MW of power by adding RC and ORC cycles. Therefore, the η o v e r a l l of the RC/ORC cycle increases to 44.37%, and Ψ o v e r a l l increases to 42.84%.
As shown in Table 11, the total exergy input and exergy destruction for the RC/ORC model are illustrated. The total exergy destruction for all components is approximately 315.3 MW, accounting for 52.31% of the total exergy input to the GT-RC/ORC. Therefore, the valuable work of the GSO CC is 258.2 MW, and its percentage is almost 42.84%. The remaining part of the exergy is released with the exhaust gases to the surrounding environment, and its percentage is nearly 4.85%.
Depending on the interaction analysis of the technical, economic, and social effects, new data can be collected. The thermodynamic quantities have been introduced in order to describe the socio-economic system as a biosystem, based on Grisolia et al. [39,40].
I = e x e r g y   l o s t q u a n t i t y   r e l a t e d   t o   r e q u i r e d   e f f e c t
when I is an active indicator.
The exergy analysis findings for the GSO CC system components under ideal conditions are shown in Table 12. This table demonstrates that the combustion chambers are where the greatest exergy is lost due to the highly irreversible nature of the combustion process (approx. 57.3%). The HRSG and condenser 1 have the second and third ranks, respectively, whereas pump 1 experiences the most negligible exergy loss (0.0004%). According to Table 12, the highest exergy efficiency is associated with the gas turbine (94.2%).
Exergoeconomic analysis is a valuable method for assessing the performance of a thermal system. The findings of the exergoeconomic study for the GSO CC system are shown in Table 13. The results indicate that the combustion chamber and steam generator had the highest Z ˙   K + C ˙   D values, correspondingly. It is clear that condensers have a more considerable relative cost difference than other components because they are less efficient. Evaluating the exergoeconomic factor demonstrates that 63% of this cost is attributable to the cost of exergy destruction, whereas only 37% is attributable to investment costs.
The effects of the changing pressure ratio (Pr), compressor isentropic efficiency ( η A C ), gas turbine isentropic efficiency ( η G T ), gas turbine inlet temperature (GTIT), boiler pressure ( P b o i l e r ), and condenser temperature ( T c o n d e n s e r ) on the performance and cost of the GSO CC system are analyzed here.
Figure 2a presents the impact of the pressure ratio (Pr) on the GSO CC system’s performance and cost. The findings demonstrate that ẆBC increases with an increase in Pr until it reaches a maximum point and then decreases with further increases in Pr. At high values of Pr, the power consumed by the compressors increases and affects ẆBC negatively. The maximum ẆBC was obtained at 10 bar (169.3 MW). The curves also illustrate that Ẇnet for the GSO CC system decreases with an increase in Pr. At the lower value of Pr, the exhaust temperature from the BC is very high and positively affects ẆRC and ẆORC. The findings reveal that when Pr increases from 4 to 18 bar, Ẇnet decreases from 288/6 MW to 229.3 MW for the GSO CC system.
Figure 2b reveals the effect of Pr on the efficiencies and cost of the GSO CC system. The GSO CC system’s efficiency and cost improve with a rise in Pr until it peaks and then decline with additional increases in Pr. At high values of Pr, Ẇnet decreases and negatively affects the efficiencies of the GSO CC system. The maximum η o v e r a l l and Ψ o v e r a l l were obtained at 13 bar (44.41% and 42.88%, respectively). The results also showed that the lower GSO CC system cost was obtained at lower Pr. For the GSO CC system, the lowest C ˙ electricity   is obtained at 7.5 bar (USD 8.67/MWh), and then C ˙ electricity   jumps to USD 10.24/MWh at 18 bar.
The influence of air compressor isentropic efficiency ( η A C ) on system performance and total cost rate is shown in Figure 3. The figures revealed that a rise in η A C results in an increase in both Ẇnet and the efficiencies of the GSO CC system. If the airflow rates remain constant, increasing the η A C will reduce the power consumption of the compressor, increasing the gas turbine’s power production. Figure 3a shows that a change in η A C from 70% to approximately 88% increases total Ẇnet from 215.6 MW to 275 MW. Figure 3b shows that when η A C was raised, the First- and Second-Law efficiencies of the cycle would improve. The results indicate that increasing η A C is needed in order to attain greater efficiency. However, this is not economical. Based on these findings, increasing η A C from 70% to approximately 84% results in a lower cycle’s overall cost. However, further raising η A C to beyond 84% increases the cycle’s overall cost [41,42].
Figure 4 shows how the isentropic efficiency of gas turbines ( η G T ) affects the performance of the GSO CC system and the rate of total costs. Ẇnet and the efficiencies of the suggested system improved by increasing η G T . By maximizing ηGT, the exergy destruction rate of the system was reduced and hence boosted the overall net and efficiencies of the proposed system. Figure 4a presents that when ηGT increases from 0.7 to 0.9, Ẇnet improves from 212.1 MW to 270 MW for the system. Additionally, the η o v e r a l l   of the GSO CC system increases from 36.45% to 46.37 % and Ψ o v e r a l l rises from 35.2% to 44.78%, as seen in Figure 4b. By increasing η G T , the overall cost rate of the system initially drops and subsequently rises. As the gas turbine’s isentropic efficiency rises, the investment cost rate of the ORC components increases by growing the mass flow rate of the working fluid. As a result of these variations, the overall cost rate will initially fall and then rise [43,44]. With an increase in η G T , C ˙ electricity   drops significantly from USD 12.4/MWh until it reaches a minimum value of USD 9.03/MWh at η G T of 86%, and then it increases to USD 9.24/MWh at η G T of 90%.
Figure 5 shows the variation in Ẇnet, η o v e r a l l , Ψ o v e r a l l , and C ˙ electricity with GTIT. These figures demonstrate that a change in GTIT has a substantial effect on the values of Ẇnet, η o v e r a l l , Ψ o v e r a l l , and C ˙ electricity . With a rise in GTIT, the temperature of the exhaust gases leaving the gas turbine rises, increasing the thermal energy transferred from the exhaust gases in the HRSG to the RC [38,45]. This produces an increase in the Ẇnet of the RC and ORC. Additionally, when the GTIT increases, the overall system becomes more efficient. Therefore, an increase in the GTIT from 1200 K to 1525 K dramatically increases Ẇnet from 133.2 MW to 282.6 MW, and η o v e r a l l improves substantially from 36.74% to 45.29%, while Ψ o v e r a l l increases from 35.74% to 43.73%. Figure 5b also illustrates how a change in the GTIT affects the entire cost of the cycle. C ˙ electricity   first declines drastically when the GTIT rises but subsequently considerably increases at higher GTIT levels [36,37,46]. C ˙ electricity   becomes minimal at a GTIT of 1475 K, leading to minimum C ˙ electricity   at 8.24 USD/MWh. With an increase in the GTIT, C ˙ electricity   reduces significantly from 14.86 USD/MWh until it falls to a minimum value of 8.24 USD/MWh at a GTIT of 1475 K and then increases to 8.86 USD/MWh at a GTIT of 1525 K.
Figure 6 shows how the boiler pressure ( P b o i l e r ) affects the performance of the GSO CC system and the rate of total costs. Since the network of the BC cycle remains constant regardless of the value of P b o i l e r , the network of the RC improves when P b o i l e r rises due to the higher enthalpy of the water leaving the boiler, as seen in Figure 6a. Consequently, when P b o i l e r climbs from 75 bar to 275 bar, Ẇnet for the GSO CC system increases from 244 MW to 262.3 MW. Likewise, the efficiencies of the GSO CC system also increase at high P b o i l e r . Figure 6b shows that as P b o i l e r increases, η o v e r a l l increases from 43.6% to 44.9%, Ψ o v e r a l l increases from 42.15% to 43.4%. The graph also shows that when Pboiler decreases from 275 bar to 75 bar, C ˙ electricity   rises from USD 8.87/MWh to USD 9.27/MWh. The decreased power output of the GSO CC system at a lower P b o i l e r is the fundamental cause of the rise in C ˙ electricity   .
The influence of Rankine cycle condenser temperature ( T c o n d e n s e r ) on system performance and the total cost rate is shown in Figure 7. The finding demonstrates that when T c o n d e n s e r rises, the RC’s power output decreases, limiting Ẇnet, η o v e r a l l , and Ψ o v e r a l l of the GSO CC system. Ẇnet drops from 2513 MW to 239.6 MW (approximately 11.7 MW) when T c o n d e n s e r increases from 303 K to 348 K, as seen in Figure 7a. Figure 7b also shows how η o v e r a l l falls from 44.77 to 42.7%, while Ψ o v e r a l l decreases from 43.23% to 41.23%. The results also show that the diminishment in the Ẇnet of the GSO CC leads to an increase in C ˙ electricity   at high T c o n d e n s e r . C ˙ electricity   increases from USD 8.83/MWh to USD 9.665/MWh with variations of T c o n d e n s e r between 303 K and 348 K.

5. Conclusions

In this paper, the Rankine and organic Rankine cycles are integrated with the Taji gas turbine’s power plant to produce extra electricity and reduce environmental emissions by using waste heat recovery from the existing power plant. The novel triple trigeneration cycle has been investigated from a thermodynamic and economic standpoint. The entire processes for the cycle were investigated parametrically and optimized using thermodynamic principles and EES software V10.561-3D. The power plant’s overall efficiency and cost as a function of the primary operational parameters are examined. The obtained results can be discussed in the following conclusions:
  • The combustion chamber is where the most exergy is lost (about 57.3%), and the pumps destroy the least amount of exergy compared to the other parts of the cycle.
  • The highest total cost rate parts are the combustion chamber, the gas turbine, the steam turbine, and the heat recovery steam generator. Therefore, from the point of view of energy–economic analysis, these are the most important parts.
  • Overall, the cost of exergy destruction for the cycle is 63%, indicating that 37% of the system’s total cost is attributable to the initial investment. This would mean that optimizing the system should focus on minimizing the cost of exergy destruction.
  • According to the findings of the parametric analysis of the base case, the thermodynamic performance and cost of the system are enhanced by increasing the gas turbine inlet temperature and the isentropic efficiencies of the gas turbine and the air compressor. However, as these parameters are increased, the total cost of the system increases. Additionally, the increase in the pressure ratio will benefit the system. Additionally, the increase in pressure ratio will be thermodynamically and exergo–economically beneficial to the system. The increase in boiler pressure increases the First- and Second-Law efficiencies and decreases the system’s total cost rate. The increase in condenser temperature decreases the First- and Second-Law efficiencies and raises the system’s total cost rate.
These results may be a suitable contribution to future investigations on the Taji gas turbine’s power plant development and enhancement.

Author Contributions

Conceptualization, A.F.K.; Methodology, A.F.K. and A.A. (Abdulrazzak Akroot); Software, A.F.K., A.A. (Abdulrazzak Akroot) and W.T.; Validation, A.F.K. and A.A. (Abdulrazzak Akroot); Formal analysis, H.A.A.W.; Resources, H.A.A.W.; Data curation, W.T.; Writing—original draft, A.F.K.; Writing—review & editing, A.A. (Abdulrazzak Akroot), R.M.G. and A.A. (Ali Alfaris). All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

SYMBOL
c F Average costs per exergy unit of fuel (USD/MJ)
c P Average costs per exergy unit of product (USD/MJ)
C ˙ D Cost rate of exergy destruction (USD/h)
E ˙ Exergy flows(MW)
E ˙ D Exergy destruction rate (MW)
E ˙ P Product exergy (MW)
E ˙ f Fuel exergy (MW)
fExergo–economic factor (%)
h specific enthalpy (kJ/kg)
iInterest rate (considered to be 10%)
LHVlower heating value (MJ/kg)
m ˙ i n Total mass flow entering per unit of time (kg/s)
m ˙ o u t Total mass flow exiting per unit time (kg/s)
nSystem lifetime (considered to be 20 years)
p pressure (bar)
𝑃𝐸𝐶Equipment purchase cost in US dollars
Q ˙ Heat transfer (kW)
s Specific entropy of mass (kJ/kg.K)
s 0 Specific entropy of mass entering the open system(kJ/kg.K)
T 0 Environment temperature (K)
W ˙ Work done (kW)
Z ˙ Capital cost rate (USD/h)
Greek symbols
ηEnergy efficiency
ε Exergy efficiency
𝜑Maintenance factor (1.06)
ψ Specific exergy
ABBREVITIONS
ACAir compressor
BCBrayton cycle
CCCombustion chamber
CONCondenser
CRFCapital Recovery Factor
GSO CCGas steam Organic combined cycle
GTGas turbine
GTITGas turbine inlet temperature
HEHeat exchanger
HRBHeat recovery boiler
HRSGHeat recovery steam generation
ORCOrganic Rankine cycle
ORTOrganic Rankine turbine
PPump
RCCycle
STSteam turbine

References

  1. Chennaif, M.; Zahboune, H.; Elhafyani, M.; Zouggar, S. Electric System Cascade Extended Analysis for Optimal Sizing of an Autonomous Hybrid CSP/PV/Wind System with Battery Energy Storage System and Thermal Energy Storage. Energy 2021, 227, 120444. [Google Scholar] [CrossRef]
  2. 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] [CrossRef]
  3. Abdelhafidi, N.; Bachari, N.E.I.; Abdelhafidi, Z.; Cheknane, A.; Mokhnache, A.; Castro, L. Modeling of Integrated Solar Combined Cycle Power Plant (ISCC) of Hassi R’mel, Algeria. Int. J. Energy Sect. Manag. 2020, 14, 505–526. [Google Scholar] [CrossRef]
  4. 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]
  5. Tutumlu, H.; Yumrutaş, R.; Yildirim, M. Investigating Thermal Performance of an Ice Rink Cooling System with an Underground Thermal Storage Tank. Energy Explor. Exploit. 2018, 36, 314–334. [Google Scholar] [CrossRef]
  6. Görgülü, A.; Yağlı, H.; Koç, Y.; Koç, A.; Öztürk, N.A.; Köse, Ö. Experimental Study of Butane Adsorption on Coconut Based Activated Carbon for Different Gas Concentrations, Temperatures and Relative Humidities. Environ. Technol. 2021, 42, 2122–2131. [Google Scholar] [CrossRef]
  7. Guvenc, M.A.; Kapusuz, H.; Mistikoglu, S. Experimental Study on Accelerometer-Based Ladle Slag Detection in Continuous Casting Process. Int. J. Adv. Manuf. Technol. 2020, 106, 2983–2993. [Google Scholar] [CrossRef]
  8. Kyritsis, D.C.; Rakopoulos, C.D.; Rakopoulos, D.C. Engine and Power Plant Combustion Technologies for Sustainability. J. Energy Eng. 2019, 145, 2019001. [Google Scholar] [CrossRef] [Green Version]
  9. Seyyedi, S.M.; Hashemi-Tilehnoee, M.; Rosen, M.A. Exergy and Exergoeconomic Analyses of a Novel Integration of a 1000 MW Pressurized Water Reactor Power Plant and a Gas Turbine Cycle through a Superheater. Ann. Nucl. Energy 2018, 115, 161–172. [Google Scholar] [CrossRef]
  10. Akroot, A.; Namli, L.; Ozcan, H. Compared Thermal Modeling of Anode- and Electrolyte-Supported SOFC-Gas Turbine Hybrid Systems. J. Electrochem. Energy Convers. Storage 2021, 18, 011001. [Google Scholar] [CrossRef]
  11. Cengel, Y.A.; Boles, M.A. Thermodynamics: An Engineering Approach, 8th ed.; McGraw-Hill Education: New York, NY, USA, 2015; ISBN 9788578110796. [Google Scholar]
  12. Liu, Z.; He, T. Exergoeconomic Analysis and Optimization of a Gas Turbine-Modular Helium Reactor with New Organic Rankine Cycle for Efficient Design and Operation. Energy Convers. Manag. 2020, 204, 112311. [Google Scholar] [CrossRef]
  13. Sahoo, P.K. Exergoeconomic Analysis and Optimization of a Cogeneration System Using Evolutionary Programming. Appl. Therm. Eng. 2008, 28, 1580–1588. [Google Scholar] [CrossRef]
  14. Woudstra, N.; Woudstra, T.; Pirone, A.; Stelt, T. van der Thermodynamic Evaluation of Combined Cycle Plants. Energy Convers. Manag. 2010, 51, 1099–1110. [Google Scholar] [CrossRef]
  15. Roy, J.P.; Mishra, M.K.; Misra, A. Parametric Optimization and Performance Analysis of a Waste Heat Recovery System Using Organic Rankine Cycle. Energy 2010, 35, 5049–5062. [Google Scholar] [CrossRef]
  16. Köse, Ö.; Koç, Y.; Yağlı, H. Performance Improvement of the Bottoming Steam Rankine Cycle (SRC) and Organic Rankine Cycle (ORC) Systems for a Triple Combined System Using Gas Turbine (GT) as Topping Cycle. Energy Convers. Manag. 2020, 211, 112745. [Google Scholar] [CrossRef]
  17. Qiu, K.; Entchev, E. Development of an Organic Rankine Cycle-Based Micro Combined Heat and Power System for Residential Applications. Appl. Energy 2020, 275, 115335. [Google Scholar] [CrossRef]
  18. Hemadri, V.B.; Subbarao, P.M.V. Thermal Integration of Reheated Organic Rankine Cycle (RH-ORC) with Gas Turbine Exhaust for Maximum Power Recovery. Therm. Sci. Eng. Prog. 2021, 23, 100876. [Google Scholar] [CrossRef]
  19. Bǎlǎnescu, D.T.; Homutescu, V.M. Performance Analysis of a Gas Turbine Combined Cycle Power Plant with Waste Heat Recovery in Organic Rankine Cycle. Procedia Manuf. 2019, 32, 520–528. [Google Scholar] [CrossRef]
  20. Ahmadi, P.; Dincer, I.; Rosen, M.A. Exergo-Environmental Analysis of an Integrated Organic Rankine Cycle for Trigeneration. Energy Convers. Manag. 2012, 64, 447–453. [Google Scholar] [CrossRef]
  21. Koç, Y.; Yağlı, H.; Kalay, I. Energy, Exergy, and Parametric Analysis of Simple and Recuperative Organic Rankine Cycles Using a Gas Turbine–Based Combined Cycle. J. Energy Eng. 2020, 146, 04020041. [Google Scholar] [CrossRef]
  22. Salehi, G.; Hasan, M.; Manesh, K.; Alahyari, A. Thermodynamic and Exergoeconomic Evaluation of Heat Recovery of Gas Refinery Steam Network Using Organic Rankine Cycle and Kalina Cycle with Different Fluids. J. Energy Eng. 2020, 146, 05020002. [Google Scholar] [CrossRef]
  23. Cao, Y.; Gao, Y.; Zheng, Y.; Dai, Y. Optimum Design and Thermodynamic Analysis of a Gas Turbine and ORC Combined Cycle with Recuperators. Energy Convers. Manag. 2016, 116, 32–41. [Google Scholar] [CrossRef]
  24. Fernández-Guillamón, A.; Molina-García, Á.; Vera-García, F.; Almendros-Ibáñez, J.A. Organic Rankine Cycle Optimization Performance Analysis Based on Super-Heater Pressure: Comparison of Working Fluids. Energies 2021, 14, 2548. [Google Scholar] [CrossRef]
  25. Macchi, E.; Astolfi, M. Organic Rankine Cycle (ORC) Power Systems: Technologies and Applications; Woodhead Publishing: Sawston, UK, 2016. [Google Scholar]
  26. Sadeghi, M.; Chitsaz, A.; Marivani, P.; Yari, M.; Mahmoudi, S.M.S. Effects of Thermophysical and Thermochemical Recuperation on the Performance of Combined Gas Turbine and Organic Rankine Cycle Power Generation System: Thermoeconomic Comparison and Multi- Objective Optimization. Energy 2020, 210, 118551. [Google Scholar] [CrossRef]
  27. Mago, P.J.; Chamra, L.M.; Srinivasan, K.; Somayaji, C. An Examination of Regenerative Organic Rankine Cycles Using Dry Fluids. Appl. Therm. Eng. 2008, 28, 998–1007. [Google Scholar] [CrossRef]
  28. Nourpour, M.; Khoshgoftar Manesh, M.H. Modeling and 6E Analysis of a Novel Quadruple Combined Cycle with Turbocompressor Gas Station. J. Therm. Anal. Calorim. 2022, 147, 5165–5197. [Google Scholar] [CrossRef]
  29. Li, X.; Gui, D.; Zhao, Z.; Li, X.; Wu, X.; Hua, Y.; Guo, P.; Zhong, H. Operation Optimization of Electrical-Heating Integrated Energy System Based on Concentrating Solar Power Plant Hybridized with Combined Heat and Power Plant. J. Clean. Prod. 2021, 289, 125712. [Google Scholar] [CrossRef]
  30. 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]
  31. Khaljani, M.; Khoshbakhti Saray, R.; Bahlouli, K. Comprehensive Analysis of Energy, Exergy and Exergo-Economic of Cogeneration of Heat and Power in a Combined Gas Turbine and Organic Rankine Cycle. Energy Convers. Manag. 2015, 97, 154–165. [Google Scholar] [CrossRef]
  32. Bejan, A.; Tsatsaronis, G.; Moran, M.J. Thermal Design and Optimization; John Wiley & Sons: Hoboken, NJ, USA, 1996; Volume 21, ISBN 0471584673. [Google Scholar]
  33. Wahhab, H.A.A.; Aziz, A.R.A.; Al-Kayiem, H.H.; Nasif, M.S.; Afolabi, L.O. Prediction of the phase distribution of diesel/CNG bubbly flow in a horizontal pipe under the influence of a magnetic field. J. Mech. Sci. Technol. 2017, 31, 5299–5309. [Google Scholar] [CrossRef]
  34. Pierobon, L.; Van Nguyen, T.; Larsen, U.; Haglind, F.; Elmegaard, B. Multi-Objective Optimization of Organic Rankine Cycles for Waste Heat Recovery: Application in an Offshore Platform. Energy 2013, 58, 538–549. [Google Scholar] [CrossRef] [Green Version]
  35. Cavalcanti, E.J.C. Exergoeconomic and Exergoenvironmental Analyses of an Integrated Solar Combined Cycle System. Renew. Sustain. Energy Rev. 2017, 67, 507–519. [Google Scholar] [CrossRef]
  36. Nami, H.; Mahmoudi, S.M.S.; Nemati, A. Exergy, Economic and Environmental Impact Assessment and Optimization of a Novel Cogeneration System Including a Gas Turbine, a Supercritical CO2 and an Organic Rankine Cycle (GT-HRSG/SCO2). Appl. Therm. Eng. 2017, 110, 1315–1330. [Google Scholar] [CrossRef]
  37. Ismaeel, A.A.; Wahhab, H.A.A.; Naji, Z.H. Performance Evaluation of Updraft Air Tower Power Plant Integrated with Double Skin Solar Air Heater. Evergreen 2021, 8, 296–303. [Google Scholar] [CrossRef]
  38. Abdul Wahhab, H.A.; Al-Kayiem, H.H. Environmental Risk Mitigation by Biodiesel Blending from Eichhornia crassipes: Performance and Emission Assessment. Sustainability 2021, 13, 8274. [Google Scholar] [CrossRef]
  39. Wahhab, H.A.A.; Aljabair, S.; Ayed, S.K. Assessment of Wind Fin Performance Run by Mixed Flows: Experimental and Numerical Investigation. Arab. J. Sci. Eng. 2021, 46, 12077–12088. [Google Scholar] [CrossRef]
  40. Grisolia, G.; Fino, D.; Lucia, U. Thermodynamic optimization of the biofuel production based on mutualism. Energy Rep. 2020, 6, 1561–1571. [Google Scholar] [CrossRef]
  41. Lucia, U.; Grisolia, G. Biofuels Analysis Based on the THDI Indicator of Sustainability. Front. Energy Res. 2021, 9, 794682. [Google Scholar] [CrossRef]
  42. Wahhab, H.A.A.; Aziz, A.R.A.; Al-Kayiem, H.H.; Nasif, M.S. Mathematical modeling of the flow of diesel-CNG fuel mixture in a pipe under the influence of a magnetic field. J. Appl. Fluid Mech. 2017, 10, 389–396. [Google Scholar]
  43. Abdul-Wahhab, H.A.; Aziz AR, A.; Al-Kayiem, H.H.; Nasif, M.S. Modeling of diesel/CNG mixing in a pre-injection chamber. IOP Conf. Ser. Mater. Sci. Eng. 2015, 100, 012044. [Google Scholar] [CrossRef] [Green Version]
  44. Al-Maamary, H.M.; Kazem, H.A.; Chaichan, M.T. Changing the energy profile of the GCC States: A review. Int. J. Appl. Eng. Res. (IJAER) 2016, 11, 1980–1988. [Google Scholar]
  45. Chaichan, M.T. Combustion and emission characteristics of E85 and diesel blend in conventional diesel engine operating in PPCI mode. Therm. Sci. Eng. Prog. 2018, 7, 45–53. [Google Scholar] [CrossRef]
  46. Khafaji, H.Q.A.; Wahhab, H.A.A.; Al-Maliki, W.A.K.; Alobaid, F.; Epple, B. Energy and Exergy Analysis for Single Slope Passive Solar Still with Different Water Depth Located in Baghdad Center. Appl. Sci. 2022, 12, 8561. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the GSOCC system.
Figure 1. Schematic diagram of the GSOCC system.
Sustainability 15 09376 g001
Figure 2. Variation of Ẇnet, η o v e r a l l , Ψ o v e r a l l , and C ˙ electricity with pressure ratio (Pr); (a) for performance and cost; (b) for efficiencies and cost.
Figure 2. Variation of Ẇnet, η o v e r a l l , Ψ o v e r a l l , and C ˙ electricity with pressure ratio (Pr); (a) for performance and cost; (b) for efficiencies and cost.
Sustainability 15 09376 g002
Figure 3. Variation of Ẇnet, η o v e r a l l , Ψ o v e r a l l , and C ˙ electricity with η A C . (a) for performance and cost; (b) for efficiencies and cost.
Figure 3. Variation of Ẇnet, η o v e r a l l , Ψ o v e r a l l , and C ˙ electricity with η A C . (a) for performance and cost; (b) for efficiencies and cost.
Sustainability 15 09376 g003
Figure 4. Variation of w ˙ net , η o v e r a l l , Ψ o v e r a l l , and C ˙ electricity with η G T . (a) for performance and cost; (b) for efficiencies and cost.
Figure 4. Variation of w ˙ net , η o v e r a l l , Ψ o v e r a l l , and C ˙ electricity with η G T . (a) for performance and cost; (b) for efficiencies and cost.
Sustainability 15 09376 g004
Figure 5. Variation of Ẇnet, η o v e r a l l , Ψ o v e r a l l , and C ˙ electricity with GTIT; (a) for performance and cost; (b) for efficiencies and cost.
Figure 5. Variation of Ẇnet, η o v e r a l l , Ψ o v e r a l l , and C ˙ electricity with GTIT; (a) for performance and cost; (b) for efficiencies and cost.
Sustainability 15 09376 g005
Figure 6. Variation of Ẇnet, η o v e r a l l , Ψ o v e r a l l , and C ˙ electricity with boiler pressure ( P b o i l e r ); (a) for performance and cost; (b) for efficiencies and cost.
Figure 6. Variation of Ẇnet, η o v e r a l l , Ψ o v e r a l l , and C ˙ electricity with boiler pressure ( P b o i l e r ); (a) for performance and cost; (b) for efficiencies and cost.
Sustainability 15 09376 g006
Figure 7. Variation in Ẇnet, η o v e r a l l , Ψ o v e r a l l  , and C ˙ electricity with condenser temperature ( T c o n d e n s e r ); (a) for performance and cost; (b) for efficiencies and cost.
Figure 7. Variation in Ẇnet, η o v e r a l l , Ψ o v e r a l l  , and C ˙ electricity with condenser temperature ( T c o n d e n s e r ); (a) for performance and cost; (b) for efficiencies and cost.
Sustainability 15 09376 g007
Table 1. Summary illustrates the methods and relationships studied by many researchers.
Table 1. Summary illustrates the methods and relationships studied by many researchers.
AuthorsMethod and Technique Remarked
Sahoo [13]Exergy analysis and economicsNew method for regular energy analysis
Woudstra et al. [14]Used modern Combined Cycle Gas Turbines (CCGTs)Thermal efficiency of at least 60%
Roy et al. [15]The integration of ORClower emissions, improved thermal efficiency, and a lower power load
Kose et al. [16] &
Qui et al. [17]
Used an SRC and ORC integrated with the GT for the variable turbine input pressure and temperatureImproved thermal efficiency
Hemadri et al. [18] &
Balanesscu et al. [19]
studied the influence of reheat on combined cycle performance and integrated with steamthermal efficiency of the combined cycle improved by 1.1%
Salehi et al. [22] &
Cao et al. [23]
Thermodynamic study of a gas turbine and ORCpressure increased, improved thermal efficiency
Table 2. Operation condition used for the GSOCC model.
Table 2. Operation condition used for the GSOCC model.
ParameterValue
GT cycleNumber of units4
Compression ratio12.1
Air mass flow rate (kg/s)142.6 × 4
GTIT (°C)1177
Ambient temperature (°C)27
LHV of fuel (Natural Gas (NG)) (kJ/kg)50,056
η A C (%)84
η G T (%)88
η C C (%)99.5
RC cycleST inlet pressure (bar)100
Condenser Temperature (°C)50
η S T (%)90
η P (%)80
Effectiveness of HRSG (%)70
ORC cycleORT inlet pressure (kPa)800
Condenser pressure (bar)1.2
Effectiveness of evaporator (%)70
Working fluidR123
Table 3. Energetic and exergetic relations for the subsystems of GSOCC [28].
Table 3. Energetic and exergetic relations for the subsystems of GSOCC [28].
SubsystemEnergy BalancesExergy Balances
Air compressor W ˙ A C = m ˙ air   ( h 2 h 1 ) E ˙ D , A C = ( E ˙ 1 E ˙ 2 ) + W ˙ A C
Combustion chamber m ˙ 2 h 2 + η C C m ˙ 3 LHV = m ˙ 4 h 4 E ˙ D ,   CC = E ˙ 2 + E ˙ 3 E ˙ 4
Gas turbine W ˙ G T = m ˙ g a s ( h 4 h 5 ) E ˙ D , G T = ( E ˙ 4 E ˙ 5 ) W ˙ G T
HRSG m ˙ 5 ( h 5 h 6 ) = m ˙ 8 ( h 9 h 8 ) E ˙ D ,   HRSG = E ˙ 5 E ˙ 6 + E ˙ 8 E ˙ 9
Steam Turbine W ˙ ST = m ˙ 9 ( h 9 h 13 ) + m ˙ 10 ( h 13 h 10 ) E ˙ D ,   ST = ( E ˙ 9 E ˙ 13 E ˙ 10 ) W ˙ ST
Condenser Q ˙ C O N D 1 = m ˙ 10 ( h 11 h 10 ) E ˙ D ,   C O N D 1 = ( E ˙ 10 E ˙ 11 ) + ( E ˙ 15 E ˙ 16 )
Pump1 W ˙ Pump 1   = m ˙ 11 ( h 12 h 11 ) E ˙ D ,   Pump 1   = W ˙ Pump 1   ( E ˙ 11 E ˙ 12 )
Pump2 W ˙ Pump 2   = m ˙ 8 ( h 14 h 8 ) E ˙ D ,   Pump 2   = W ˙ Pump 2   ( E ˙ 14 E ˙ 8 )
Deaerator m ˙ 13 h 13 + m ˙ 12 h 12 = m ˙ 14 h 14 E ˙ D ,   Deaerator   = E ˙ 13 + E ˙ 12 E ˙ 14
HRB Q ˙ H R B = m ˙ 18 ( h 19 h 18 ) E ˙ D , H R B = E ˙ 6 E ˙ 7 + E ˙ 18 E ˙ 19
ORT W ˙ ORT = m ˙ 19 ( h 20 h 19 ) E ˙ D ,   ORT = ( E ˙ 19 E ˙ 20 ) W ˙ ORT
ORC Condenser Q ˙ C O N D 2 = m ˙ 21 ( h 22 h 21 ) E ˙ D ,   C O N D 2 = ( E ˙ 21 E ˙ 22 ) + ( E ˙ 23 E ˙ 24 )
ORC Pump W ˙ Pump 3   = m ˙ 22 ( h 17 h 22 ) E ˙ D ,   Pump 3   = W ˙ Pump 3   ( E ˙ 22 E ˙ 17 )
Heat exchanger Q ˙ H E = m ˙ 17 ( h 18 h 17 ) E ˙ D , HE = E ˙ 17 E ˙ 18 + E ˙ 20 E ˙ 21
Table 4. Exergetic performance criteria of GSOCC system.
Table 4. Exergetic performance criteria of GSOCC system.
first-law efficiency η overall = ( W ˙ B C + W ˙ R C + W ˙ O R C ) / Q ˙ C C
Second-law efficiency Ψ o v e r a l l = ( W ˙ B C + W ˙ R C + W ˙ O R C ) / ( E ˙ 1 + W ˙ 3 )
Components exergy efficiency ε k = E ˙ P / E ˙ F
Exergy destructio ratio E d , k % = E ˙ d , k / E ˙ d ,   t o t a l
Table 5. Exergoeconomic evaluation parameters of GSO CC [31,32].
Table 5. Exergoeconomic evaluation parameters of GSO CC [31,32].
Average costs per exergy unit of fuel c F , k =   Ċ   F , k /   Ė   F , k
Average costs per exergy unit of product c P , k =   Ċ   P , k /   Ė   P , k
Cost rate of exergy destruction   Ċ   D , k = c F , k   Ė   D , k
Relative Cost Difference r k = ( c P , k c F , k ) / c F , k
Exergoeconomic factor f k =   Ż   k / (   Ż   k + c F , k   Ė   D , k +   Ė   L , k ) )
The overall cost C ˙ system = k = 1 N Z k + k = 1 N C ˙ D , k
Total electricity cost   Ċ   electricity ,   Total = C ˙ system / W ˙ net  
Table 6. Cost balance and auxiliary equations for the components of GSO CC system.
Table 6. Cost balance and auxiliary equations for the components of GSO CC system.
SubsystemExergy Cost Rate Balance EquationExergy Balances
Air compressor   Ċ   1 +   Ċ   AC +   Ż   AC =   Ċ   2   Ċ   1   W   C , LP =   Ċ   2   W   T          c 1   = 0
Combustion chamber   Ċ   2 +   Ċ   3 +   Ż   CC =   Ċ   4   Ċ   2 Ex 2 =   Ċ   4 Ex 4         c 2   =   c 4          c 3 =12
Gas turbine   Ċ   4 +   Ż   G T =   Ċ   5   +   Ċ   G T   Ċ   4 Ex 4 =   Ċ   5 Ex 5         c 4   =   c 5
HRSG   Ċ   5 +   Ċ   8 + Ż ˙ HRSG =   Ċ   6 +   Ċ   9   Ċ   5 Ex 5 =   Ċ   6 Ex 6           c 5 =   c 6
Steam Turbine   Ċ   9 +   Ż   ST =   Ċ   10 +   Ċ   13 +     Ċ   ST   Ċ   9 Ex 9 =   Ċ   10 Ex 10       c 9 =   c 10        c 10 =   c 13
Condenser   Ċ   10 +   Ċ   15 + Ż ˙ cond 1 =   Ċ   11 +   Ċ   16   Ċ   10 Ex 10 =   Ċ   11 Ex 11           c 10 =   c 11          c 15 = 0
Pump1   Ċ   11 +   Ċ   ST + Ż ˙ Pump 1 =   Ċ   12   Ċ   Pump 1   W   Pump 1 =   Ċ   ST   W   ST
Deaerator   Ċ   12 +   Ċ   13 + Ż ˙ Derea =   Ċ   14   Ċ   12 Ex 12 =   Ċ   14 Ex 14
Pump2   Ċ   14 +   Ċ   ST + Ż ˙ Pump 2 =   Ċ   8   Ċ   Pump 2   W   Pump 2 =   Ċ   ST   W   ST
HRB   Ċ   6 +   Ċ   18 + Ż ˙ HRB =   Ċ   7 +   Ċ   19   Ċ   6 Ex 6 =   Ċ   7 Ex 7            c 10 =   c 11
ORT   Ċ   19 + Ż ˙ ORT =   Ċ   20 +   Ċ   ORT   Ċ   19 Ex 19 =   Ċ   20 Ex 20            c 19 =   c 20
ORC Condenser   Ċ   21 +   Ċ   24 + Ż ˙ cond 2 =   Ċ   22 +   Ċ   23   Ċ   21 Ex 21 =   Ċ   22 Ex 22            c 21 =   c 22
ORP   Ċ   22 +   Ċ   ORP + Ż ˙ ORP =   Ċ   17   Ċ   ORP   W   ORP =   Ċ   ORT   W   ORT
Heat exchanger   Ċ   20 +   Ċ   17 + Ż ˙ HE =   Ċ   21 +   Ċ   18   Ċ   20 Ex 20 =   Ċ   21 Ex 21           c 20 =   c 21
Table 7. Purchased cost equations for the components of GSOCC system.
Table 7. Purchased cost equations for the components of GSOCC system.
SubsystemCost Equation
Air compressor ( 71.1 ×   m ˙ air / 0.9 η comp ) ( p 2 / p 1 ) Ln ( p 2 / p 1 ) [32]
Combustion chamber ( 46.08 ×   m ˙ air / 0.996 ( p 4 / p 2 ) ) ( 1 + exp ( 0.018 × T 4 26.4 ) ) [32]
Gas turbine ( 479.34 ×   m ˙ 4 / 0.92 η GT ) Ln ( p 4 / p 5 )   ( 1 + exp ( 0.036 × T 4 54.4 ) ) [32]
HRSG 6570   [   Q ˙ ec Δ T LMTD , ec 0.8 +   Q ˙ ev Δ T LMTD , ev 0.8 +   Q ˙ sup Δ T LMTD , sup 0.8 ] + 21276   m ˙ 8 + 1184.4   m ˙ 4 [32]
Steam Turbine 6000   W ˙ ST 0.7 [33]
Condenser 1773   m ˙ steam [34]
Pump1 2100   W ˙ pump 1 0.26 ( 1 η Pump / η Pump ) 0.5 [34]
Deaerator 52,000   m ˙ 9 [34]
Pump2 2100   W ˙ pump 2 0.26 ( 1 η Pump / η Pump ) 0.5 [34]
HRB 235   Q ˙ HRB [35]
ORT ( 479.3   m ˙ 19 / 0.92 η ORT ) Ln ( p 19 / p 20 )   ( 1 + exp ( 0.036 × T 4 54.4 ) ) [30]
ORC Condenser 1773   m ˙ organic [34]
ORP 2100   W ˙ ORP 0.26 ( 1 η ORP / η ORP ) 0.5 [34]
Heat exchanger 235   Q ˙ HE [35]
Table 8. Validation of the Brayton cycle model.
Table 8. Validation of the Brayton cycle model.
ParameterLiterature Model [21]Present Model
Pressure ratio 1515
Fuel type of gas turbine cycle Natural gasNatural gas
Air mass flow rate (kg/s) 21.421.4
Air mass flow rate (kg/s) 0.40.41
Exhaust Flow (kg/s) 21.821.9
Power output (kW) 62006203
Thermal efficiency 33.832.7
Table 9. Properties for each state for the GSO CC model at the optimum condition.
Table 9. Properties for each state for the GSO CC model at the optimum condition.
Statem
(kg/s)
P
(kPa)
T
(K)
h
(kJ/kg)
s
(KJ/kg K)
E
(MW)
c
($/GJ)
C ˙
($/h)
1498.4101.3300233.65.762000
2498.41226679.2627.85.886203.722.5316,522
311.2101.3288−467211.53518.21226,035
4509.611651450235.28.164508.818.942,578
5509.6107.7919.8−425.58.3200.7218.914,284
6509.6104.5500−9047.61741.5218.92873
7509.6101.3400−10127.38619.418.91364
867.5610,133445.7735.52.05511.7743.721931
967.569829869.836196.904140.525.4613,475
1054.0412.2632323397.29836.8925.463496
1154.0412.26323208.70.70192.88925.46273.3
1254.04100323208.80.7022.89325.48273.9
1313.51100372.826177.20313.3625.461269
1467.56100372.8690.42.0359.13646.831599
15637.110030020287.082000
16637.110031027819.5517.85499.853225
17177.2800306.42351.1190.112564.3426.12
18177.2800359.6292.91.2921.222140617.2
19177.28004805401.9211.8650.072141
20177.2121.6428.4499.31.9314.08850.07738.4
21177.2121.6306.1402.81.6690.816150.07147.4
22177.2121.6306.1234.41.1190.0284850.075.144
2339.6210030020287.082000
2439.6210030524048.3260.1238329.5147.2
Table 10. Performance of the GSO CC model.
Table 10. Performance of the GSO CC model.
BC ModelORC Model
Net output power167.3258.2
Overall exergy efficiency27.7542.84
Overall thermal efficiency28.7444.37
Table 11. Exergy input, output, losses of the model, and active indicator.
Table 11. Exergy input, output, losses of the model, and active indicator.
ExergyValues (MW)Percentage (%)StationI
Input602.7100%GT cycle7.2
Output (network)258.242.84%RC cycle9.6
Exergy destruction315.352.31%ORC cycle3.4
Exergy losses29.24.85%Combined 7.1
Total602.7100%
Table 12. Exergy analysis and the entropy change rate for each component of the (GSO CC) model.
Table 12. Exergy analysis and the entropy change rate for each component of the (GSO CC) model.
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
(%)
s ˙ g e n e r a t i o n
(Kw/K)
Δs
(KJ/Kg.K)
Exergy Effi.
(%)
AC224.8203.721.176.971.0411.65490.6
CC784.8608.7176.157.3590.99.25277.6
GT408384.323.737.779.30.13694.2
HRSG159.2128.830.399.88101.984.16680.9
ST90.382.697.62.52.57.59791.85
Condenser 1347.8526.168.587.784.127123.1
Deaerator16.259.147.122.3123.895.8756.22
Pump 10.0060.0050.0010.00040081.4
Pump 23.052.630.4110.1340086.5
ORB22.1410.6311.53.7438.590.0448
ORT7.777.210.550.181.8450.39792.86
HE3.271.112.160.77.250.08934
Condenser 20.790.1240.660.222.2151.79415.72
ORP0.1040.0840.020.00660080.6
Table 13. Exergoeconomic results of components of the GSO CC system.
Table 13. Exergoeconomic results of components of the GSO CC system.
Component c f
($/GJ)
c p
($/GJ)
C ˙   D
($/h)
Z ˙   K
($/h)
Z ˙   K + C ˙   D
($/h)
r
(%)
f
(%)
AC20.4122.27120190.4310.49.1261.33
CC14.6618.9735.820.1755.928.942.65
GT18.920.18124.2187.9312.16.7560.22
HRSG18.923.8172.9133305.925.9843.46
ST25.4628.6556.56268324.512.5582.75
Condenser 125.4699.85185.31.5186.8292.20.82
Deaerator25.4646.8351.9456.46108.483.9152.07
Pump 128.6535.990.00910.0140.023625.9561.18
Pump 228.6533.123.4390.0743.51315.582.1
ORB18.939.7360.5314.575.02110.219.31
ORT50.0755.287.72835.6543.3810.4282.19
HE50.07147.730.140.0830.221950.254
Condenser 250.07329.59.2514.914.1555834.62
ORP55.2869.180.3110.1810.49225.1336.82
Total system--1558.0912.42470.4-37
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

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. https://doi.org/10.3390/su15129376

AMA Style

Kareem AF, Akroot A, Abdul Wahhab HA, Talal W, Ghazal RM, 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(12):9376. https://doi.org/10.3390/su15129376

Chicago/Turabian Style

Kareem, Alaa Fadhil, Abdulrazzak Akroot, Hasanain A. Abdul Wahhab, Wadah Talal, Rabeea M. Ghazal, and Ali Alfaris. 2023. "Exergo–Economic and Parametric Analysis of Waste Heat Recovery from Taji Gas Turbines Power Plant Using Rankine Cycle and Organic Rankine Cycle" Sustainability 15, no. 12: 9376. https://doi.org/10.3390/su15129376

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