Next Article in Journal
Advances in Triboelectric Nanogenerators for Sustainable and Renewable Energy: Working Mechanism, Tribo-Surface Structure, Energy Storage-Collection System, and Applications
Previous Article in Journal
A Full-State Reliability Analysis Method for Remanufactured Machine Tools Based on Meta Action and a Markov Chain Using an Exercise Machine (EM) as an Example
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Performance Improvement Overview of the Supercritical Carbon Dioxide Brayton Cycle

1
School of Energy and Building Environment Engineering, Henan University of Urban Construction, Pingdingshan 467036, China
2
School of Energy and Power Engineering, Shandong University, Jinan 250061, China
3
School of Energy and Environment, Zhongyuan University of Technology, Zhengzhou 451191, China
*
Author to whom correspondence should be addressed.
Processes 2023, 11(9), 2795; https://doi.org/10.3390/pr11092795
Submission received: 6 August 2023 / Revised: 14 September 2023 / Accepted: 15 September 2023 / Published: 20 September 2023
(This article belongs to the Section Energy Systems)

Abstract

:
Efficiency and compactness are core strengths of the supercritical carbon dioxide (sCO 2 ) Brayton cycle, which is considered an alternative to the steam Rankine cycle for moderate-temperature heat sources (350–800 °C). Numerical investigations on system design and analysis have received considerable attention, with the aim of improving the sCO 2 cycle from the viewpoint of thermodynamics. This paper reviews and compares previous studies in the literature to survey different cycle layouts, operating parameters, and working fluids of the sCO 2 cycle. Performance enhancement approaches are categorized into three classes according to characteristics: conventional methods, CO 2 mixtures, and combined cycles. The strengths, weaknesses, and limitations of each categorized method are discussed. This research is expected to provide a roadmap for performance improvement that meets the interests of researchers.

1. Introduction

The supercritical carbon dioxide (sCO 2 ) Brayton cycle employs supercritical CO 2 working fluid for electricity generation and is characterized by high cycle efficiency, small turbomachinery, high power density, and a simple cycle layout. This emerging technology is adaptable to the majority of heat sources, such as nuclear, solar, fossil fuels, and waste heat, and has attracted significant attention in the last couple of decades [1]. Figure 1 presents publication numbers on a yearly basis with respect to this hot topic since 2002. A body of raw data was collected by Wang [2]. The number of annual publications has risen exponentially. The number of articles published by Chinese researchers has grown since 2014. In 2019, China published more than 80 studies, which accounted for 36.6% of the total documented in an international journal and conference database. Research on nuclear power plants accounts for 42.9% of all sCO 2 cycle research reported in the literature, followed by research on concentrating solar power (CSP) (30.2%) and waste heat recovery (WHR) from gas turbines (9.6%) and fossil power plants (8.2%). Currently, research on the sCO 2 cycle is in the experimental and demonstration stages. The net power tests a 50 MWth plant to demonstrate the zero-emission Allam cycle, which is a variant of the sCO 2 cycle [3]. The Supercritical Transformational Electric Power (STEP) project is being developed by the Southwest Research Institute (SwRI), Gas Technology Institute (GTI), GE Global Research (GE), and the U.S. Department of Energy (DOE) to operate a 10 MWe indirect-fired sCO 2 cycle [4]. In Europe, components have been designed and tested for the sCO2-HeRo (supercritical CO 2 heat removal system), which is considered a backup cooling system for the reactor core [5]. In China, the Institute of Engineering Thermophysics (IET) built an MW-scale sCO 2 compressor test platform to demonstrate CO 2 compression under subcritical and supercritical conditions [6]. In addition, a printed circuit heat exchanger facility is under development by the IET, with CO 2 operating at a temperature and pressure of 550 °C and 32 MPa, respectively [7].
Compression near the critical point of CO 2 benefits the thermal efficiency of the sCO 2 cycle, but also constrains the application. Power plants operating in hot and/or arid environments are not uncommon, especially for CSP in deserts, taking advantage of the excellent year-round solar radiation. High ambient air temperatures in the range of 25–37 °C [8] and dry air-cooling technologies make it difficult and costly to achieve the desired condition at the inlet of the compressor. High compressor inlet temperatures (CITs) considerably reduce the thermal efficiency of the sCO 2 cycle [9]. In addition, daily and seasonal variations of heat sources and cooling conditions or the use of inefficient cooling systems can hinder the performance of the sCO 2 cycle. A cycle design with low CITs leads to long periods of part-load operation, whereas a high-temperature design penalizes the cycle efficiency [10,11].
Increasing the thermal efficiency of the sCO 2 cycle and the design of new cycles are among the interests of researchers in the field of thermodynamics. Moisseytsev and Sienicki [12] reported several improved methods for the sCO 2 cycle heated by a sodium-cooled fast reactor (SFR). These methods include broad operating ranges (temperature and pressure) and modifications of cycle configuration (reheating, intercooling, recompression, and condensation). Ahn et al. [13] reported on development progress with respect to the sCO 2 cycle and compared single-flow and split-flow cycle layouts. Crespi et al. [14] reviewed and categorized sCO 2 cycles according to their configurations, covering 42 varieties of stand-alone cycles and 38 combined layouts. However, they failed to present details of performance enhancement or include studies on the CO 2 mixture Brayton cycles. Li et al. [15] summarized the development of the sCO 2 cycle in the nuclear and solar industries and categorized designs into single-flow and split-flow layouts. They covered improvement in sCO 2 cycles (like using a bottoming cycle) but focused on the research status rather than the comparison of performance improvement. Wang et al. [16] reviewed and compared six sCO 2 cycles for solar tower CSP. Xu et al. [17] focused on barriers to current research on the sCO 2 cycle.
The research gap lies in a lack of a comprehensive review and interpretation of performance improvement options for the sCO 2 cycle, as well as a comparison of enhanced thermal efficiencies and analysis of the limitations of each method. CO 2 mixtures were previously employed to adjust the critical point of pure CO 2 for power cycles adapting to higher or lower heat sinks [18]. The features and influence of this approach may be aggressive or moderate with respect to other methods. It is not possible to draw a conclusion as to which method is preferable without comparison. In the present paper, we review previously published papers published on methods that could enhance the sCO 2 cycle. Applications, cycle layouts, methods, claimed thermal efficiencies, and operating conditions are presented. Each approach is interpreted and categorized according to its type and characteristics. the claimed performance changes of the sCO 2 cycle are compared from the viewpoints of thermodynamics and economics. In addition, we discuss the strengths, weaknesses, and limitations of each categorized method. This research is expected to provide an intuitive overview of the performance enhancement of sCO 2 cycles.

2. Conventional Method

A conventional method is referred to as a classical approach to improve a power cycle, such as with intercooling, reheating, recuperation, and high operating parameters, which have been applied to the Rankine cycle and Brayton cycle. In the early stages, these options were considered to be handy and extensive. Figure 2 presents four popular schematic diagrams of the sCO 2 cycle layout that have been improved by using the conventional method. Angelino [19] studied the various layouts of the sCO 2 cycle. However, he mainly focused on the transcritical CO 2 (tCO 2 ) cycle, i.e., the condensation cycle. Later, intercooling, reheating, recompression, partial cooling, and precompression were investigated [9,20,21,22]. Table 1 summarizes the methods, boundary conditions, and thermal efficiencies declared in the literature. It should be noted that these studies are theoretical and are based on specific component efficiency assumptions, forming many varieties of the sCO 2 cycle, as reviewed by Crespi et al. [14]. Since a heat exchanger contributes 54% to the capital cost of the recompression of the sCO 2 cycle [23], a reduction in pressure loss and the enhancement of heat transfer play important roles in enhancing the performance of the sCO 2 cycle from the point of view of exergy and energy [24,25,26]. This is also true for an increase in turbomachinery efficiency. The reason why Table 1 does not list the improved methods of the components is that these methods are less important than the cycle layout modifications and operating parameters.
Figure 3 shows the thermal efficiency of enhanced sCO 2 cycles as a function of turbine inlet temperature (TIT). Note that the efficiencies of various configurations were collected from published papers on the basis of very different boundary conditions. In order to establish a convincing comparison, the commonly used efficiencies of the sCO 2 cycle were plotted as the reference [9]. It is noted that the reference data were calculated without consideration of cycle losses. It can be seen that the enhanced recuperated sCO 2 cycle has an efficiency in the order of about 3% lower than the reference efficiency at the same TIT. The efficiencies of the improved recompression and partial cooling cycles are higher than the reference efficiencies, except for several points. This is mainly due to high CITs for these cycles, such as 50 °C [28], 51 °C [37], and 55.5 °C [32,33]. Figure 3 also presents the recompression cycle with thermal efficiencies of up to 50% at a TIT of 500 °C. This was achieved not only by reheating but also by the combination of intercooling, high-efficiency turbomachinery, and large recuperator conductance [36,38].
Figure 4 shows the cycle efficiency improvement at various TITs for each specific conventional method. Note that the improvements in efficiency achieved by the reheating or intercooling is obtained with respect to the sCO 2 cycle without modification. As for the methods for the CO 2 condensation cycle, CO 2 liquid-phase pumping, precompression, recompression, and partial cooling, the improvement in efficiency is the efficiency difference for the recuperated sCO 2 cycle. Efficiency gains from increasing TITs are calculated for every 100 °C rise in the TIT. Moreover, the turbine inlet pressure (TIP) is also a crucial factor in changing the efficiency.
For the sCO 2 cycle, the typical conditions at the turbine inlet are 550 °C and 20 MPa [9,39]. It is known that a higher temperature or pressure implies higher cycle efficiency. Figure 4 shows that an order of about 4% additional efficiency is obtained for each 100 °C rise in the TIT. At higher TITs, the efficiency improvement diminishes to 3% [9]. When increasing the TIT from 550 °C to 650 °C, efficiency was improved by 2.8% and 4% for the recuperated (with intercooling) and recompression sCO 2 cycles, respectively [27]. For a precompression layout with higher turbomachinery performance, increasing the temperature to 650 °C improved the efficiency by 5.3% to a value of 48.8% [29]. As for TIPs beyond 20 MPa, the gain in cycle efficiency is very modest. About a 0.3% increase in cycle efficiency was achieved when the TIP was raised from 20 to 22 MPa (at a TIT of 480 °C) [12]. Dostal [9] showed that increasing the pressure from 20 to 25 MPa yielded 1.4% (at a TIT of 650 °C) and 2.0% (at a TIT of 750 °C) efficiency improvements. In contrast, the increase in efficiency was less than 1% when the TIP rose from 25 MPa to 30 MPa. It should be noted that higher temperature and/or pressure conditions represent challenges in terms of material selection. Raising the pressure means increasing the thickness of the pipes, pressure-bearing casings, and heat exchangers, requiring additional capital costs. Moreover, material corrosion should be considered in high turbine inlet conditions.
The sCO 2 cycle takes advantage of the non-ideal properties of CO 2 near the critical point to reduce the compression work. It was shown that higher cycle efficiency could be obtained if the CO 2 is cooled below the pseudocritical temperature at a given supercritical pressure. When the temperature declines across the critical point, the cycle is called the “CO 2 liquid cycle” or “CO 2 condensation cycle”, depending on the compressor inlet pressure crossing the critical point (the latter) or not (the former). The liquid-phase cycle operating at supercritical pressures increased the thermal efficiency by 1.15% at pump inlet conditions of 30 °C and 7.4 MPa [12]. The CO 2 condensation cycle (under pump inlet conditions of 20 °C and 5.75 MPa) achieved an efficiency of 43.1% for a 483 °C SFR, having an efficiency improvement of 3.85% [12]. These two cycles should be carefully treated due to the requirement of a year-round supply of the cold-cooling medium. Moreover, in order to cool CO 2 below the critical temperature, the CO 2 has to pass through the peak specific heat, resulting in a significant increase in the heat transfer area of the cooler.
Reheating improves thermal efficiency by increasing the equivalent Carnot temperature, i.e., the average temperature at which the heat is added to the power cycle. As shown in Figure 4, reheating adds 1–2% to the thermal efficiency. The gains provided by a second reheating configuration, when compared to single reheating, stay below 0.5% and lower at high TITs [9,30,34]. Figure 4 also shows that single-stage reheating could obtain a 2.5% and 3.0% increase in efficiency at TITs of 500 °C and 550 °C, respectively, resulting from a high TIP of 25 MPa and a better component performance [31]. Several factors should be paid attention to regarding the reheating method: (1) The effect of reheating strongly depends on the pressure drop through the reheater. With an increase in pressure drop, the benefit from the reheat decreases until it is negative [9,33]. (2) In contrast to an ideal gas cycle, the pressure ratio split between sCO 2 turbines should be optimized to give equivalent temperatures of heat addition. It was observed that the optimum pressure–ratio split was slightly lower than an equal split [9]. (3) It was shown that using more than one stage of reheating is economically unattractive [9].
Intercooling reduces the average temperature of heat rejection from the sCO 2 cycle. However, the CO 2 compression work is already low, such that the benefit from intercooling is expected to be limited. As shown in Figure 4, intercooling offers a very modest efficiency improvement, which was 0.8% at a TIT of 550 °C (recuperated sCO 2 layout) [9]. With a significant increase in the TIT, the efficiency is slightly increased. It can be seen that efficiency gains can be up to 3% at TITs of higher than 800 °C [20,35]. This is due to the existence of large compression work in the original sCO 2 cycle. It should be noted that the properties of CO 2 are considerably affected by the critical point, and thus, the pressure–ratio split is not equal in order to achieve the same TIP. It was shown that the second compressor provided a 1.5 to 1.9 times higher pressure ratio than the equally split pressure ratio, giving the best cycle performance [9,21]. An equal pressure ratio for the two compressors results in a slightly low thermal efficiency [12]. The disadvantage of intercooling is the additional cost.
The cause of the irreversibility of the recuperator is the pinch-point problem, which largely reduces the performance of the sCO 2 cycle. In order to overcome this problem, compound cycles were introduced [19]. The precompression cycle is another way to increase regeneration [9,40]. As reported in [29], this cycle layout achieved an efficiency improvement of 4% over the recuperated sCO 2 cycle. The partial cooling cycle operates at pressures of about 12 MPa and temperatures of around 700–800 °C. Its efficiency is improved by reducing the average temperature of heat rejection. As can be seen in Figure 4, the efficiency improvement from the partial cooling cycle is larger than that from the precompression cycle. The recompression cycle improves efficiency by reducing the heat rejection from the cycle using an additional compressor before the cooler. The efficiencies of complex cycles (recompression and partial cooling) are consistently about 6% higher than those of the recuperated cycle [30]. At high CITs, the efficiency increase is reduced to a grade of around 4% [20,33].

3. Combined Cycle

Although the sCO 2 cycle has promising thermal efficiency, about 50% of heat addition is still rejected to the heat sink at temperatures of 100 °C to 200 °C. For the purpose of minimizing thermodynamic irreversibility as a consequence of the second law, WHR technologies were employed in the sCO 2 cycle either for power generation or cold supply [2,14]. The WHR technology, or a bottoming cycle, is introduced between the low-temperature recuperator (LTR) and the cooler with an additional heat exchanger, as shown in Figure 5. This approach is recognized as the combined cycle method, by which the recuperator in the topping cycle does not need to have high effectiveness. Table 2 presents the different configurations of the sCO 2 combined cycle in the literature. The bottoming cycles considered are the organic Rankine cycle (ORC), tCO 2 cycle, and the Kalina cycle. Statistics show that about 56% of combined sCO 2 cycles use the ORC bottoming cycle, and 38% employ the tCO 2 cycle. Only two papers are devoted to the complicated Kalina cycle. It should be noted that the combined cycle method requires a year-round supply of very cold cooling water. The WHR technologies operate at narrow temperature ranges between the heat source and heat sink. The working fluid needs a condensation process to maintain higher performance.
Figure 6 plots the declared thermal efficiencies of the combined sCO 2 system in the literature. Wide dispersion is observed in the chart, showing global efficiencies in the range of 35% to 60%. This is due to the layout and boundary conditions of each combined cycle, as shown in Table 2. The efficiencies, as well as the reference efficiencies, of the stand-alone sCO 2 cycle [9] are further plotted against TIT in Figure 6, showing the clear influence of this parameter on the performance of the combined cycle. The thermal efficiencies of the combined recuperated cycles are nearly 5% lower than reference efficiencies. Figure 7 shows the efficiency improvement after adding a bottoming cycle to the sCO 2 cycle at various TITs. In order to show more details, three common layouts of the topping sCO 2 cycle are distinguished using symbols and colors. The efficiency improvement is defined as the thermal efficiency of the combined sCO 2 cycle over the thermal efficiency of the initial sCO 2 cycle.
η i n i t i a l = W n e t , s C O 2 Q i n
η c o m b i n e d = W n e t , t o p p i n g + W n e t , b o t t o m Q i n
η i m p r o v e m e n t = η c o m b i n e d η i n i t i a l
ORC employs a refrigeration fluid as the working fluid to convert low-temperature heat into electricity, and this is applicable to the heat sources of geothermal energy [65], solar energy [66], and industrial waste heat [67,68,69]. There is an established industry with over 2000 MW installed. ORC technology was first added to a recuperated sCO 2 cycle (operating at the turbine inlet conditions of 727 °C and 30 MPa) in 2008, showing that the efficiency was improved by 6.1% when using cyclohexane as the working fluid [70]. Some interesting information from the literature review can be found as follows:
  • ORC is mainly combined with the recuperated sCO 2 cycle, followed by the recompression and partial cooling cycles. The recuperated sCO 2 -ORC combined cycle is slightly less complex than the other two combined cycles. Moreover, the recuperated sCO 2 cycle has a relatively larger amount of waste heat at high temperatures, which is the ideal source for WHR technologies.
  • The performance of the recuperated sCO 2 cycle is significantly improved by the additional ORC cycle. As shown in Figure 7, the gain in efficiency ranges from 6 to 13%, which is achieved by the combined cycle compared to the stand-alone recuperated sCO 2 cycle [42]. The CIT and organic fluid are the main contributors to the diversity of performance enhancement [44]. For the recompression sCO 2 cycle, the efficiency was improved by 4–4.6% at a TIT of 550 °C [48] and 4.2–5% at a TIT of higher than 700 °C [44]. Zhang et al. [47] studied a recompression sCO 2 -ORC combined cycle, having a thermal efficiency of 52.12% when using a liquefied natural gas (LNG) as the heat sink. About 6.8% of additional efficiency was achieved with respect to the stand-alone cycle due to the small power scale, i.e., 297 kWe, of which the bottoming cycle output accounted for 13%. As for the partial cooling sCO 2 cycle, the gain in efficiency ranged from 2–3% [44] depending on organic fluids.
  • The recompression sCO 2 -ORC cycle achieves the highest overall efficiency. It can be seen from Figure 6 that most combined recompression cycles have efficiencies that are higher than the reference data at the same TITs, although the cycle boundary conditions are not identical.
  • The distinct differences in the organic fluids show limited effects on the performance of the combined cycle. This is true, especially for large-scale sCO 2 -ORC-combined cycles [50]. In the literature, both pure substances and zeotropic mixtures [43,59,60] were investigated. The nature of organic fluids decides the efficiency improvement that rises with TIT, like R245fa [42] and Isopentane [44], or falls, like the isobutane and the n-butane/isopentane mixture [43], as shown in Figure 7.
One drawback of the ORC cycle is the pinch problem or constant temperature evaporation in the evaporator. Alternatively, the tCO 2 cycle and Kalina cycle are potentially useful technologies when applied to low-temperature heat sources. CO 2 is an ideal choice as the working fluid for transcritical operations due to its high performance, compactness, and non-flammability. Additionally, it is cost-effective, exhibits low toxicity, and has a negligible environmental impact. The transcritical process makes the tCO 2 cycle absorb heat at the gliding temperature instead of at the constant temperature evaporation in subcritical conditions. There are two objectives for the tCO 2 cycle, which are additional to the sCO 2 cycle.
  • For power generation, Yari and Sirousazar [41] first added a tCO 2 cycle to the recompression sCO 2 cycle, showing an efficiency improvement of around 5.1%. They also showed that the performance of the combined cycle was independent of TIT but was strongly affected by the ambient temperature. When increasing the ambient temperature from 15 °C to 25 °C, the gain in efficiency went down from 5.1% to 2.7%. Alsagri and Chiasson [58] reported that the recompression cycle using a split-flow tCO 2 bottoming cycle obtained 2% more efficiency than that with a non-recuperated tCO 2 cycle. As can be seen in Figure 7, the recuperated sCO 2 cycle, for which the heat source is a molten carbonate fuel cell, was improved by 4.8–6.5% in terms of efficiency, depending on the recuperator effectiveness at a CIT of 55 °C [45,46]. When the CIT was lowered to 32 °C, the gain in efficiency for the tCO 2 cycle was up to 7.2% [2].
  • For cold production, Akbari and Mahmoudi [54] proposed a recompression sCO 2 cycle combining a tCO 2 refrigeration cycle for both power and refrigeration production. In the combined cycle, a fraction of CO 2 leaving the cooler expends in the refrigeration cycle; after being heated and compressed, it then goes back to the cooler. They showed that the combined cycle produced 240 MW of power and 60 MW of cold simultaneously, or 562 MW of cold only. Manjunath et al. [61] studied a similar combined cycle utilizing the shipboard gas turbine exhaust. The proposed cycle generated a net power output of 3.0 MW (about 14.5% engine power) and a cooling output of 3.1 MW (about 15% engine power).
Although interesting studies have been published on the exergoeconomic performance of the combined sCO 2 cycle, as shown in Table 2, the techno-economic evaluation is still uncertain due to the lack of experimental facilities and standard equipment. Want et al. [23] conducted the thermoeconomic analysis of a combined recompression sCO 2 -tCO 2 cycle, showing that the capital cost per net power output was about 6% more expensive than that of the stand-alone sCO 2 cycle. The heat exchanger accounted for 53% of the total cost. Their cost estimation was based on the equipment used in the chemical process. Later, Wang [2] showed that the recompression sCO 2 cycle with a tCO 2 cycle was 5.3% more expensive than the stand-alone cycle using modified cost functions.
The Kalina cycle uses an ammonia/water solution as the working fluid for power generation. The boiling point of the ammonia/water mixture can be adjusted to suit the heat input temperature by the appropriate choice of the mass ratio. As shown in Figure 7, the Kalina cycle improves the recompression sCO 2 cycle by 3.2–4% in terms of efficiency when compared to the sCO 2 cycle in isolation, according to the ammonia concentration in the solution [51,52]. Note that the Kalina cycle itself is much more complicated and hard to operate, which alone and with any additional cost may offset the benefit.

4. CO 2 Mixture

The modification of the thermal–physical properties of the working fluid by adding a small amount of other gases is not uncommon to improve a power cycle. As for the helium Brayton cycle, the use of a He/Xe or He/N 2 binary mixture was successful in reducing the stages of the turbomachinery due to the increment in the molecular weight of the gas mixture [71,72]. For the ORC, zeotropic mixtures have a non-isothermal phase change during evaporation and condensation, which reduces temperature mismatching and exergy destruction [73,74,75,76,77]. In previous studies, CO 2 mixtures were considerably investigated in refrigeration and heat pump systems [78,79,80], as was the transcritical Rankine cycle (TRC) for low-to-medium temperature applications [81,82,83,84].
As for the sCO 2 cycle, the critical point of CO 2 plays a crucial role in the lowest operating boundary and reduction of compression work. The adjustment of the critical point changes the cycle temperature and pressure ranges, providing the possibility of using high-temperature heat sinks. The direction and range of the critical point variation of CO 2 depend on the mixed component and its amount. In this regard, research on CO 2 based mixtures has been performed for various purposes. In 2011, the Sandia National Laboratories (SNL) [85] investigated the effect of secondary gas additions, i.e., SF 6 , neon, n-butane, and methane, on the performance of a 50 kW full-scale CO 2 compressor. The Korea Advanced Institute of Science and Technology (KAIST) [18,86,87,88] studied the SFR using the sCO 2 cycle and first proposed the modification of the CO 2 critical point by mixing gases. Several potential gases were selected to consider their thermal stability, property database, and chemical stability in the temperature range of interest. Their goal was to shift the critical point of the pure CO 2 and, thus, expand the operating range of the cycle. The Czech Technical University (CTU) in Prague focused on the effect of gaseous admixtures on the pinch point [89] and cycle performance [90,91]. Both binary and multicomponent mixtures were considered.
Table 3 surveys gas additives in terms of the sCO 2 cycles, boundary conditions, applications, and thermal efficiencies declared in the original papers. For the Brayton cycles, including the basic, recuperated, recompression, precompression, and split expansion layouts, the CIT is always fixed at 1 °C above the critical temperature, and the TIP is fixed at a certain value. For TRC, the maximum power output is the focus of the purpose of WHR. As can be seen, although most articles considered the recompression cycle with a TIT of 550 °C, large discrepancies in thermal efficiencies were obtained from one paper to another.
Figure 8 plots the declared thermal efficiencies of the supercritical CO 2 mixture for both Brayton and Rankinecycles, as listed in Table 3. There are over 250 data, showing a wide range of cycle efficiencies. As can be seen, most Brayton cycles exhibit efficiencies in the range of 40% to 50%, and for the Rankine cycle, the efficiencies are from 6% to 10%. With low turbomachinery efficiency, the compound cycles have efficiencies of 25–32.5% [89,90]. The Brayton cycles without the recuperator have efficiencies well below 15% [88]. The wide dispersion of efficiencies is due to the cycle layout, as well as boundary conditions, such as TIT, turbomachinery efficiency, the effectiveness of heat exchangers, cooling conditions, and gas additives. Figure 8 also presents the efficiencies against TIT for all the cycles considered. In order to establish a convincing comparison, the commonly used efficiencies of the sCO 2 cycle were plotted as the reference [9]. In general, the recompression layout of the supercritical CO 2 mixture cycles outperforms other layouts and achieves efficiencies close to or even higher than the reference ones.
Figure 9 plots the influence of additives on cycle thermal efficiency. The efficiency change is defined as the thermal efficiency difference between the CO 2 mixture cycle and the pure CO 2 cycle. As can be seen, some additives do not improve cycle efficiency as might be anticipated, which are Ar, N 2 , O 2 [86,87], CO, CH 4 , and H 2 [89,91]. These additives alone with He, Xe [86,87], and Kr [18,91] decrease the critical temperature of CO 2 . However, the Xe and Kr additives improve the Brayton cycle by up to 2% in terms of efficiency, depending on the mass fraction. It was found that simultaneously lowering the critical temperature and critical pressure of the CO 2 mixture has a positive effect on the total cycle efficiency. The decrease in critical pressure leads to an increase in the cycle operating pressure ratio. It is interesting that the addition of small amounts of Kr leads to an increase and then a decrease in the critical pressure. The reported difference in cycle efficiency was down to −0.7% for the CO 2 /Kr binary mixture consisting of 1% mol impurities when compared to pure CO 2 [91]. For the CO 2 /He mixture, Jeong et al. [18] and Hu et al. [95] showed positive effects, whereas Vesely et al. [89] showed detrimental effects. They all used the REFPROP program, although Jeong et al. [18] pointed out that the properties of the CO 2 /He mixture were opposite to the experiment data.
The substances H 2 S, cyclohexane, n-butane, toluene, SO 2 , R123, R134a, R22, R32, and TiCl 4 have higher critical temperatures than pure CO 2 . For these gas additives, the benefit from the compression near the critical point is maintained at high cooling conditions. As can be seen in Figure 9, adding the substances increases the cycle efficiency, ranging from 0 to 2.0%. The CO 2 /TiCl 4 mixture makes the Brayton cycle efficiency increase in the range of 1.5% to 5%. The findings in the literature show that adding the above additives to CO 2 cannot stop cycle efficiency from decreasing in a warm environment, although the compression work is still reduced. Adding a gas to CO 2 just mitigates the extent of the efficiency reduction. At a CIT of 44 °C for a dry cooling system, the efficiency of a recompression sCO 2 cycle decreased from 45.4% to 41.4%, whereas adding H 2 S and cyclohexane to the working fluid reduced the efficiency to 43.2% and 42.6%, respectively [18]. In addition, the findings show that the higher the ambient temperature, the more obvious the advantages of the method of the CO 2 mixture. A higher ambient temperature requires a higher CIT; thus, a larger amount of gas should be added. Adding TiCl 4 may be a better choice in a high ambient temperature situation, showing an increase in efficiency of 5.5% higher than the pure CO 2 Brayton cycle. In addition, the CO 2 /TiCl 4 working fluid reduced the power cycle capital cost by 14.5–17.5% [98]. It is noted that several CO 2 based mixtures were tested using the sCO 2 equipment. The CO 2 /n-butane, CO 2 /neon, and CO 2 /SF 6 mixtures were tested in a full-scale compressor at SNL [85]. The performance of the CO 2 /R32 mixture with mass fractions of 0.88:0.12 was tested in the compressor, showing that a slightly higher pressure ratio was achieved [88].
The investigated multicomponent mixtures are Air, M-I, M-II, M-H, Air-H, and M-IH [90]. The basic multicomponent mixture is from the technology of carbon capture and storage. The rest of the mixtures are combinations of pure additives and basic multicomponents. As can be seen in Figure 9, the multicomponent mixtures have a negative effect on the sCO 2 cycle. The effect was small if the purity of CO 2 was over 99%. If the purity of CO 2 drops under 99%, the effect of the mixture deteriorates. The cycle output could be significantly decreased by the M-II mixtures, as shown in Figure 10, which shows the thermal efficiency and power output changes of the sCO 2 cycles when the CO 2 based mixture was applied as the working fluid with respect to the original cycles.
Mixing SF 6 with CO 2 increases the critical temperature, as calculated by the REFPROP program. However, it was found that the CO 2 /SF 6 mixture for the Brayton cycle had less efficiency than the pure sCO 2 cycle at the same CIT [88]. As shown in Figure 9, the SF 6 additive reduced the cycle efficiency by 2.1% at most (with a mass fraction of 0.4). This may indicate that the critical point predicted by REFPROP is not close to the experimental value. Earlier, Lewis et al. [85] showed that the addition of SF 6 to a CO 2 -dominated mixture reduced the critical temperature until SF 6 became the dominant component. A good prediction of CO 2 /SF 6 mixture behavior should be further investigated.
For the Rankine cycle, the substances R1270, R161, R152a, R1234ze, R1234yf, R290, R32, and R134a were blended with CO 2 , showing an efficiency gain in the range of −1.5% to 2.5%. The CO 2 /R161 mixture exhibited the best cycle efficiency [81] and the most effective economic performance [82]. Note that R161 is highly flammable until the CO 2 fraction is higher than 70% [103]. The CO 2 /R290 mixture is commonly used in refrigeration and heat pump systems with good performance, whereas it had decreased efficiency when used in the WHR system, as shown in Figure 9. Note that for the WHR system, the selection of the amount of additive is used to convert as much heat to power as possible; hence, the heat source temperature is not always very high.
The selection of a CO 2 mixture should simultaneously take into account thermal properties, cycle efficiency, heat transfer performance, safety, and environmental issues. To date, the choice of CO 2 mixtures has mainly considered the effect on Brayton cycle efficiency by using the basic law of thermodynamics. The thermal stability of the mixture has been less of a concern. The supercritical CO 2 mixture Brayton cycles operate at a temperature of 350 °C to 700 °C, raising the possibility of bond-breaking reactions and the further decomposition of the additive. Beyond the critical value of thermal stability, organic fluid pyrolysis yields gas, liquid, or solid products. The gaseous products (like non-condensing gases) circulate with the working fluid, whereas the solid products stick to the surface of the heat exchanger, both of which will worsen the heat transfer. In addition, the decomposition products are mixed with the working medium, which changes the thermal properties of the working medium and makes the system deviate from the design condition, resulting in a reduction in system output power [104,105]. More seriously, molecular polymerization may occur, resulting in the blockage of a pipeline and causing a hidden danger to the system safety [106].
Table 4 collects the CO 2 mixtures that have positive effects on the power cycle mentioned above. The physical properties, safety, and environmental data are also presented. H 2 S is a highly flammable and explosive gas. It was shown that its initial decomposition to hydrogen and sulfur occurred at a temperature below 444.6 °C [107,108]. The equilibrium concentration of hydrogen was less than 1% until the temperature rose up to 550 °C and about 6% at 800 °C [109]. SO 2 is chemically stable and was used as an early refrigerant in home refrigerators. Iron, steel, nickel, copper-nickel alloys, and Inconel nickel-chromium-iron are satisfactory for dry or hot SO 2 but are readily corroded by wet SO 2 gas [110]. The study of TiCl 4 as a heat pipe fluid in the intermediate temperature range of 400–700 K was proposed by Davarakonda and Olminsky [111], and the thermophysical properties were evaluated by NASA [112]. The experimental analysis showed that TiCl 4 was remarkably stable at temperatures up to 500 °C [113]. These considerations and studies support the application of TiCl 4 in high-temperature ORC [114]. N 2 O 4 is one of the chemically reactive gases that undergo dissociation reactions at high temperatures and recombination at low temperatures. At temperatures of 30–170 °C, N 2 O 4 is unstable and decomposes into NO 2 , which undergoes subsequent thermal decomposition into O 2 and NO, depending on the pressure and temperature [115].
Organic compounds are not thermally stable at high temperatures. Between 100 °C and 500 °C, many organic molecules break down, breaking the chemical bonds in their molecules. As can be seen in Table 4, propane, cyclohexane, benzene, R161, and R32 have relatively high decomposition temperatures. Solovyev [117] experimentally showed that propane began thermal decomposition at about 525 °C. The main products were hydrogen, methane, ethane, and ethylene. Tsang [119] proposed that the main initial processes of cyclohexane pyrolysis involved the isomerization of cyclohexane to 1-hexene, followed by the decomposition of 1-hexene. At around 800 °C, the extent of the reaction was extremely small, which is indicative of the great stability of cyclohexane. Zanetti and Egloff [120] showed that the decomposition of benzene began at temperatures as low as 500 °C. In cases below 600 °C, the amounts decomposed were small and chiefly turned into hydrogen and diphenyl at temperatures below 750 °C. As for flammability, benzene was suppressed by a volume fraction of 32% of CO 2 [128].
Several factors should also be considered when modifying the critical point of CO 2 .
  • The available mixture database is required. The NIST REFPROP is currently used to calculate the thermophysical properties of a CO 2 mixture. Experimental data are quite rare; only the properties of CO 2 /xenon can be used so far [129]. However, the REFPROP program suggests the equation of state (EOS) temperature and pressure limits as follows: 161–750 K and 700 MPa for CO 2 /Xe, 116–750 K and 200 MPa for CO 2 /Kr, 188–760 K and 170 MPa for CO 2 /H 2 S, and 279–700 K and 80 MPa for CO 2 /cyclohexane. In order to estimate the properties at higher temperatures, the temperature range of the REFPROP program has to be extended.
  • The critical phenomena of CO 2 mixtures should be considered carefully. The critical line has a continuous or discontinuous form among each critical point of the pure components. As for the discontinuous critical lines, the existence of phase separation could cause instability in the CO 2 mixtures. This could hardly further obtain the optimum design of the compressor.
  • The thermal stability of potential additives should be considered as a key selection criterion besides their thermodynamic performance. Otherwise, decomposition products, like non-condensable gases and deposits, may reduce the heat transfer rates, damage the components, and compromise cycle safety. Additional experiments are needed to identify the real phenomena of CO 2 mixtures at high temperatures.
  • The chemical effect on cycle components should be considered in the future.

5. Strengths and Weaknesses Analysis

In order to compare the three categorized methods discussed above, the strengths and weaknesses have been identified and are listed in Table 5.
The conventional method provides a moderate enhancement to the sCO 2 cycle. A promising power cycle always comes at the cost of high cycle complexity. More components and complex control strategies would make off-design operations less flexible than the recuperated layout of the sCO 2 cycle. It was found that the recompression sCO 2 cycle is able to yield a high efficiency while still retaining simplicity.
The combined cycle could largely improve the sCO 2 cycle by recovering waste heat from the cooler. At least four more components are required, which increases the capital cost by about 5% [2,23]. The challenging task of this approach is the condensation of the working fluid in the bottoming cycle. Wet cooling or a low-temperature heat sink is mandatory, especially for the tCO 2 bottoming cycle. If the terrestrial ambient temperature is too high or if dry cooling is employed, it is difficult or even impossible to cool the working fluid sufficiently.
The CO 2 mixture method increases the thermal efficiency of the sCO 2 cycle in an economical way. The adjustment of the critical temperature and pressure of the working fluid broadens the applicability of the sCO 2 cycle for a variety of possible heat sinks. The selection of additives plays a crucial role in cycle performance, heat transfer, thermal stability, control, and chemical stability.

6. Conclusions

This paper surveys the available methods of enhanced sCO 2 cycles and uses thermodynamics to gain insight into the effects of cycle modification. These approaches were classified into three categories by type and characteristic: the conventional method, CO 2 mixture, and combined cycle. The comparison of each method was based on the values declared in the original papers, which were investigated under very different operating conditions. An intuitive overview of the performance enhancement is provided for the sCO 2 cycle. In addition, the strengths and limitations of each method are discussed.
In general, the conventional method shows moderate efficiency improvements. Adding a bottoming cycle is a promising option that can significantly improve efficiency. The drawbacks are the requirement of a year-round supply of cold cooling water and about a 5% increase in the capital cost. The modification of the thermal–physical properties of CO 2 is an economical way to change cycle performance. The higher the ambient temperature, the more obvious the advantages of the method of critical point shifting. Thermal stability and the property database are crucial factors for gas mixture selections. Combining the potential methods is a better choice. Further steps toward a comparison of the three methods under the same operating conditions should be conducted.

Author Contributions

Conceptualization, X.W. and X.F.; methodology, J.W. and M.H.; resources, L.Z. and Z.Z.; writing—original draft preparation, L.Z. and Z.Z.; writing—review and editing, X.W. and X.F.; project administration, X.W.; funding acquisition, X.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (grant number 52106006), and was supported by the key scientific research projects of Henan Province Colleges and Universities (grant number 22B470001), the science and technology project of Henan Province (grant numbers 222102320254 and 212102310580), and the Chongqing Research Program of Basic Research and Frontier Technology (no. cstc2020jcyj-msxmX0840).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
sCO 2 supercritical carbon dioxide
tCO 2 transcritical carbon dioxide
TITturbine inlet temperature
TIPturbine inlet pressure
CITcompressor inlet temperature
CSPconcentrating solar power
WHRwaste heat recovery
ORCorganic Rankine cycle
TRCtranscritical Rankine cycle
EOSequation of state
LNGliquefied natural gas

References

  1. Brun, K.; Friedman, P.; Dennis, R. Fundamentals and Applications of Supercritical Carbon Dioxide (sCO2) Based Power Cycles; Woodhead Publishing: Sawston, UK, 2017. [Google Scholar]
  2. Wang, X. A Supercritical CO2 Brayton Cycle Based Cogeneration System. Ph.D. Thesis, Xi’an Jiaotong University, Xi’an, China, 2018. (In Chinese). [Google Scholar]
  3. Tollefson, J. Innovative zero-emissions power plant begins battery of tests. Nature 2018, 557, 622–623. [Google Scholar] [CrossRef] [PubMed]
  4. Southwest Research Institute. SwRI, GTI and GE Break Ground on $119 Million Supercritical CO2 Pilot Power Plant. Available online: https://www.swri.org/press-release/swri-gti-ge-supercritical-CO2-pilot-power-plant (accessed on 18 September 2023).
  5. Benra, F.; Brillert, D.; Frybort, O.; Hajek, P.; Rohde, M.; Schuster, S.; Seewald, M.; Starflinger, J. A supercritical CO2 low temperature Brayton cycle for residual heat removal. In Proceedings of the 5th International Symposium—Supercritical CO2 Power Cycles, San Antonio, TX, USA, 28–31 March 2016. [Google Scholar]
  6. Institute of Engineering Thermophysics. China’s First Large-Scale Supercritical Carbon Dioxide Compressor Experimental Platform Completed and Put Into Operation. Available online: http://www.bjb.cas.cn/kjdt2016/201809/t201809295110326.html (accessed on 18 September 2023). (In Chinese).
  7. Institute of Engineering Thermophysics. Construction of a Comprehensive Test Platform for Full-Temperature Full-Pressure Supercritical Carbon Dioxide Heat Exchanger. Available online: http://www.dnl.ac.cn/info/1025/1167.htm (accessed on 18 September 2023). (In Chinese).
  8. Singh, R. Dynamics and Control of a Closed Carbon-Dioxide Brayton Cycle. Ph.D. Thesis, The University of Queensland, Brisbane, Australia, 2013. [Google Scholar]
  9. Dostal, V. A Supercritical Carbon Dioxide Cycle for Next Generation Nuclear Reactors. Ph.D. Thesis, Massachusetts Institute of Technology, Cambridge, MA, USA, 2004. [Google Scholar]
  10. Dyreby, J.J. Modeling the Supercritical Carbon Dioxide Brayton Cycle with Recompression. Ph.D. Thesis, University of Wisconsin-Madison, Madison, WI, USA, 2014. [Google Scholar]
  11. de la Calle, A.; Bayon, A.; Too, Y.C.S. Impact of ambient temperature on supercritical CO2 recompression Brayton cycle in arid locations: Finding the optimal design conditions. Energy 2018, 153, 1016–1027. [Google Scholar] [CrossRef]
  12. Moisseytsev, A.; Sienicki, J. Performance Improvement Options for the Supercritical Carbon Dioxide Brayton Cycle; Technical Report; ANL-GenIV-103; Argonne National Laboratory: Argonne, IL, USA, 2007. [Google Scholar]
  13. Ahn, Y.; Bae, S.J.; Kim, M.; Cho, S.K.; Baik, S.; Lee, J.I.; Cha, J.E. Review of supercritical CO2 power cycle technology and current status of research and development. Nucl. Eng. Technol. 2015, 47, 647–661. [Google Scholar] [CrossRef]
  14. Crespi, F.; Gavagnin, G.; Sánchez, D.; Martínez, G.S. Supercritical Carbon Dioxide Cycles for Power Generation: A Review. Appl. Energy 2017, 195, 152–183. [Google Scholar] [CrossRef]
  15. Li, M.; Zhu, H.; Guo, J.; Wang, K.; Tao, W. The development technology and applications of supercritical CO2 power cycle in nuclear energy, solar energy and other energy industries. Appl. Therm. Eng. 2017, 126, 255–275. [Google Scholar] [CrossRef]
  16. Wang, K.; He, Y.L.; Zhu, H.H. Integration between supercritical CO2 Brayton cycles and molten salt solar power towers: A review and a comprehensive comparison of different cycle layouts. Appl. Energy 2017, 195, 819–836. [Google Scholar] [CrossRef]
  17. Xu, J.; Liu, C.; Sun, E.; Xie, J.; Li, M.; Yang, Y.; Liu, J. Perspective of S-CO2 power cycles. Energy 2019, 186, 115831. [Google Scholar] [CrossRef]
  18. Jeong, W.S.; Jeong, Y.H. Performance of Supercritical Brayton Cycle Using CO2-based Binary Mixture at Varying Critical Points for SFR Applications. Nucl. Eng. Des. 2013, 262, 12–20. [Google Scholar] [CrossRef]
  19. Angelino, G. Real Gas Effects in Carbon Dioxide Cycles. In Proceedings of the ASME 1969 Gas Turbine Conference and Products Show, Cleveland, OH, USA, 10–13 March 1969. [Google Scholar]
  20. Kato, Y.; Nitawaki, T.; Muto, Y. Medium Temperature Carbon Dioxide Gas Turbine Reactor. Nucl. Eng. Des. 2004, 230, 195–207. [Google Scholar] [CrossRef]
  21. Ma, Y.; Liu, M.; Yan, J.; Liu, J. Thermodynamic Study of Main Compression Intercooling Effects on Supercritical CO2 Recompression Brayton Cycle. Energy 2017, 140, 746–756. [Google Scholar] [CrossRef]
  22. Couso, G.B.; Vicencio, R.B.; Padilla, R.V.; Too, Y.C.S.; Pye, J. Dynamic Model of Supercritical CO2 Brayton Cycles Driven by Concentrated Solar Power. In Proceedings of the ASME 2017 11th International Conference on Energy Sustainability, American Society of Mechanical Engineers, Charlotte, NC, USA, 26–30 June 2017; p. V001T05A008. [Google Scholar]
  23. Wang, X.; Wu, Y.; Wang, J.; Dai, Y.; Xie, D. Thermo-economic Analysis of A Recompression Supercritical CO2 Cycle Combined With A Transcritical CO2 Cycle. In Proceedings of the ASME Turbo Expo 2015: Turbine Technical Conference and Exposition, Montréal, QC, Canada, 15–19 June 2015. [Google Scholar]
  24. Stamatellos, G.; Stamatelos, T. Effect of actual recuperators effectiveness on the attainable efficiency of supercritical CO2 Brayton cycles for solar thermal power plants. Energies 2022, 15, 7773. [Google Scholar] [CrossRef]
  25. Ren, Z.; Zhao, C.R.; Jiang, P.X.; Bo, H.L. Investigation on local convection heat transfer of supercritical CO2 during cooling in horizontal semicircular channels of printed circuit heat exchanger. Appl. Therm. Eng. 2019, 157, 113697. [Google Scholar] [CrossRef]
  26. Stamatellos, G.; Stamatellou, A.; Kalfas, A.I. CFD—Aided design methodology for PCHE-type recuperators in supercritical CO2 recompression power cycles. In Proceedings of the ASME Turbo Expo 2020: Turbomachinery Technical Conference and Exposition, New York, NY, USA, 21–25 September 2020. [Google Scholar]
  27. Dostal, V.; Driscoll, M.J.; Hejzlar, P.; Todreas, N.E. A Supercritical CO2 Gas Turbine Power Cycle for Next-generation Nuclear Reactors. In Proceedings of the 10th International Conference on Nuclear Engineering, Arlington, VA, USA, 14–18 April 2002. [Google Scholar]
  28. Sarkar, J.; Bhattacharyya, S. Optimization of Recompression S-CO2 Power Cycle With Reheating. Energy Convers. Manag. 2009, 50, 1939–1945. [Google Scholar] [CrossRef]
  29. Kulhanek, M.; Dostal, V. Thermodynamic Analysis and Comparison of Supercritical Carbon Dioxide Cycles. In Proceedings of the Supercritical CO2 Power Cycle Symposium, Boulder, CO, USA, 24–25 May 2011. [Google Scholar]
  30. Turchi, C.S.; Ma, Z.; Neises, T.; Wagner, M. Thermodynamic Study of Advanced Supercritical Carbon Dioxide Power Cycles for High Performance Concentrating Solar Power Systems. In Proceedings of the ASME 2012 6th International Conference on Energy Sustainability, San Diego, CA, USA, 23–26 July 2012. [Google Scholar]
  31. Turchi, C.S.; Ma, Z.; Dyreby, J. Supercritical Carbon Dioxide Power Cycle Configuration for Use in Concentrating Solar Power Systems. In Proceedings of the ASME Turbo Expo 2012: Turbine Technical Conference and Exposition, American Society of Mechanical Engineers, Copenhagen, Denmark, 11–15 June 2012; pp. 967–973. [Google Scholar]
  32. Padilla, R.V.; Benito, R.G.; Stein, W. An Exergy Analysis of Recompression Supercritical CO2 Cycles With and Without Reheating. Energy Procedia 2015, 69, 1181–1191. [Google Scholar] [CrossRef]
  33. Padilla, R.V.; Too, Y.C.S.; Beath, A.; McNaughton, R.; Stein, W. Effect of Pressure Drop and Reheating on Thermal and Exergetic Performance of Supercritical Carbon Dioxide Brayton Cycles Integrated With a Solar Central Receiver. J. Sol. Energy Eng. 2015, 137, 051012. [Google Scholar] [CrossRef]
  34. Mounir Mecheri, Y.L.M. Supercritical CO2 Brayton Cycles for Coal-fired Power Plants. Energy 2016, 103, 758–771. [Google Scholar] [CrossRef]
  35. Purjam, M.; Goudarzi, K.; Keshtgar, M. A New Supercritical Carbon Dioxide Brayton Cycle with High Efficiency. Heat Transf.—Asian Res. 2017, 46, 465–482. [Google Scholar] [CrossRef]
  36. Coco-Enriquez, L.; Munoz-Anton, J.; Martinez-Val, J. Thermodynamic Optimization of Supercritical CO2 Brayton Power Cycles Coupled to Line-Focusing Solar Fields. J. Sol. Energy Eng. 2017, 139, 061005. [Google Scholar] [CrossRef]
  37. Binotti, M.; Astolfi, M.; Campanari, S.; Manzolini, G.; Silva, P. Preliminary Assessment of sCO2 Cycles for Power Generation in CSP Solar Tower Plants. Appl. Energy 2017, 204, 1007–1017. [Google Scholar] [CrossRef]
  38. Coco-Enriquez, L.; Munoz-Anton, J.; Martinez-Val, J. Dual Loop Line-Focusing Solar Power Plants With Supercritical Brayton Power Cycles. Int. J. Hydrogen Energy 2017, 42, 17664–17680. [Google Scholar] [CrossRef]
  39. Moisseytsev, A. Passive Load Follow Analysis of the STAR-LM and STAR-H2 Systems. Ph.D. Thesis, Texas A&M University, College Station, TX, USA, 2003. [Google Scholar]
  40. Angelino, G. Carbon Dioxide Condensation Cycles for Power Production. J. Eng. Power Trans. ASME 1968, 90, 287–295. [Google Scholar] [CrossRef]
  41. Yari, M.; Sirousazar, M. A Novel Recompression S-CO2 Brayton Cycle With Pre-cooler Exergy Utilization. Proc. Inst. Mech. Eng. Part A J. Power Energy 2010, 224, 931–946. [Google Scholar] [CrossRef]
  42. Chacartegui, R.; de Escalona, J.M.; Sánchez, D.; Monje, B.; Sánchez, T. Alternative Cycles Based on Carbon Dioxide for Central Receiver Solar Power Plants. Appl. Therm. Eng. 2011, 31, 872–879. [Google Scholar] [CrossRef]
  43. Sánchez, D.; Brenes, B.M.; de Escalona, J.M.M.; Chacartegui, R. Non-conventional Combined Cycle for Intermediate Temperature Systems. Int. J. Energy Res. 2013, 37, 403–411. [Google Scholar] [CrossRef]
  44. Besarati, S.M.; Goswami, D.Y. Analysis of Advanced Supercritical Carbon Dioxide Power Cycles With a Bottoming Cycle for Concentrating Solar Power Applications. J. Sol. Energy Eng. 2014, 136, 010904. [Google Scholar] [CrossRef]
  45. Bae, S.J.; Ahn, Y.; Lee, J.; Lee, J.I. Hybrid System of Supercritical Carbon Dioxide Brayton Cycle and Carbon Dioxide Rankine Cycle Combined Fuel Cell. In Proceedings of the ASME Turbo Expo, Düsseldorf, Germany, 16–20 June 2014. [Google Scholar]
  46. Bae, S.J.; Ahn, Y.; Lee, J.; Lee, J.I. Various Supercritical Carbon Dioxide Cycle Layouts Study for Molten Carbonate Fuel Cell Application. J. Power Sources 2014, 270, 608–618. [Google Scholar] [CrossRef]
  47. Zhang, H.; Shao, S.; Zhao, H.; Feng, Z. Thermodynamic Analysis of a SCO2 Part-Flow Cycle Combined With an Organic Rankine Cycle With Liquefied Natural Gas as Heat Sink. In Proceedings of the ASME Turbo Expo 2014: Turbine Technical Conference and Exposition, American Society of Mechanical Engineers, Düsseldorf, Germany, 16–20 June 2014; p. V03BT36A012. [Google Scholar]
  48. Akbari, A.D.; Mahmoudi, S.M. Thermoeconomic Analysis & Optimization of the Combined Supercritical CO2 (Carbon Dioxide) Recompression Brayton/Organic Rankine Cycle. Energy 2014, 78, 501–512. [Google Scholar]
  49. Wang, X.; Wang, J.; Zhao, P.; Dai, Y. Thermodynamic comparison and optimization of supercritical CO2 Brayton cycles with a bottoming transcritical CO2 cycle. J. Energy Eng. 2015, 142, 04015028. [Google Scholar] [CrossRef]
  50. Wang, X.; Dai, Y. Exergoeconomic Analysis of Utilizing the Transcritical CO2 Cycle and the ORC for A Recompression Supercritical CO2 Cycle Waste Heat Recovery: A Comparative Study. Appl. Energy 2016, 170, 193–207. [Google Scholar] [CrossRef]
  51. S Mahmoudi, S.M.; D Akbari, A.; Rosen, M.A. Thermoeconomic Analysis and Optimization of A New Combined Supercritical Carbon Dioxide Recompression Brayton/Kalina cycle. Sustainability 2016, 8, 1079. [Google Scholar] [CrossRef]
  52. Li, H.; Wang, M.; Wang, J.; Dai, Y. Exergoeconomic Analysis and Optimization of A Supercritical CO2 Cycle Coupled With A Kalina Cycle. J. Energy Eng. 2016, 143, 04016055. [Google Scholar] [CrossRef]
  53. Mahmoudi, S.; Ghavimi, A. Thermoeconomic analysis and multi objective optimization of a molten carbonate fuel cell—Supercritical carbon dioxide—Organic Rankin cycle integrated power system using liquefied natural gas as heat sink. Appl. Therm. Eng. 2016, 107, 1219–1232. [Google Scholar] [CrossRef]
  54. Akbari, A.; Mahmoudi, S. Thermoeconomic Performance and Optimization of A Novel Cogeneration System Using Carbon Dioxide as Working Fluid. Energy Convers. Manag. 2017, 145, 265–277. [Google Scholar] [CrossRef]
  55. Cao, Y.; Ren, J.; Sang, Y.; Dai, Y. Thermodynamic analysis and optimization of a gas turbine and cascade CO2 combined cycle. Energy Convers. Manag. 2017, 144, 193–204. [Google Scholar] [CrossRef]
  56. Hou, S.; Wu, Y.; Zhou, Y.; Yu, L. Performance analysis of the combined supercritical CO2 recompression and regenerative cycle used in waste heat recovery of marine gas turbine. Energy Convers. Manag. 2017, 151, 73–85. [Google Scholar] [CrossRef]
  57. Wang, X.; Yang, Y.; Zheng, Y.; Dai, Y. Exergy and exergoeconomic analyses of a supercritical CO2 cycle for a cogeneration application. Energy 2017, 119, 971–982. [Google Scholar] [CrossRef]
  58. Alsagri, A.S.; Chiasson, A.D. Thermodynamic Analysis and Multi-Objective Optimizations of a Combined Recompression SCO2 Brayton Cycle-TCO2 Rankine Cycles for Waste Heat Recovery. Int. J. Curr. Eng. Technol. 2018, 8, 541–548. [Google Scholar]
  59. Hou, S.; Cao, S.; Yu, L.; Zhou, Y.; Wu, Y.; Zhang, F. Performance Optimization of Combined Supercritical CO2 Recompression Cycle and Regenerative Organic Rankine Cycle Using Zeotropic Mixture Fluid. Energy Convers. Manag. 2018, 166, 187–200. [Google Scholar] [CrossRef]
  60. Hou, S.; Zhou, Y.; Yu, L.; Zhang, F.; Cao, S.; Wu, Y. Optimization of A Novel Cogeneration System Including A Gas Turbine, A Supercritical CO2 Recompression Cycle, A Steam Power Cycle and An Organic Rankine Cycle. Energy Convers. Manag. 2018, 172, 457–471. [Google Scholar] [CrossRef]
  61. Manjunath, K.; Sharma, O.; Tyagi, S.; Kaushik, S. Thermodynamic Analysis of A Supercritical/transcritical CO2 Based Waste Heat Recovery Cycle for Shipboard Power and Cooling Applications. Energy Convers. Manag. 2018, 155, 262–275. [Google Scholar] [CrossRef]
  62. Singh, H.; Mishra, R. Performance Analysis of Solar Parabolic Trough Collectors Driven Combined Supercritical CO2 and Organic Rankine Cycle. Eng. Sci. Technol. Int. J. 2018, 21, 451–464. [Google Scholar]
  63. Song, J.; song Li, X.; dong Ren, X.; wei Gu, C. Performance Analysis and Parametric Optimization of Supercritical Carbon Dioxide (S-CO2) Cycle With Bottoming Organic Rankine Cycle (ORC). Energy 2018, 143, 406–416. [Google Scholar] [CrossRef]
  64. Cao, Y.; Rattner, A.S.; Dai, Y. Thermoeconomic analysis of a gas turbine and cascaded CO2 combined cycle using thermal oil as an intermediate heat-transfer fluid. Energy 2018, 162, 1253–1268. [Google Scholar] [CrossRef]
  65. Dipippo, R. Geothermal Power Plants. Principles Applications, and Case Studies; Elsevier Science: Amsterdam, The Netherlands, 2005. [Google Scholar]
  66. Harinck, J.; Turunen-Saaresti, T.; Colonna, P.; Rebay, S.; van Buijtenen, J. Computational Study of A High Expansion Ratio Radial Organic Rankine Cycle Turbine Stator. J. Eng. Gas Turbines Power 1996, 118, 359–367. [Google Scholar] [CrossRef]
  67. Drescher, U.; Bruggemann, D. Fluid Selection for the Organic Rankine Cycle (ORC) in Biomass Power and Heat Plants. Appl. Therm. Eng. 2007, 27, 223–228. [Google Scholar] [CrossRef]
  68. Lai, N.A.; Wendland, M.; Fischer, J. Working Fluids for High-Temperature Organic Rankine Cycles. Energy 2011, 36, 199–211. [Google Scholar] [CrossRef]
  69. Quoilin, S.; Aumann, R.; Grill, A.; Schuster, A.; Lemort, V.; Spliethoff, H. Dynamic Modeling and Optimal Control Strategy of Waste Heat Recovery Organic Rankine Cycles. Appl. Energy 2011, 88, 2183–2190. [Google Scholar] [CrossRef]
  70. Chacartegui, R.; Sánchez, D.; Jiménez-Espadafor, F.; Muñoz, A.; Sánchez, T. Analysis of Intermediate Temperature Combined Cycles With A Carbon Dioxide Topping Cycle. In Proceedings of the ASME Turbo Expo 2008: Power for Land, Sea and Air, Berlin, Germany, 9–13 June 2008; pp. 673–680. [Google Scholar]
  71. El-Genk, M.; Tournier, J. Noble Gas Mixtures for Gas-cooled Reactor Power Plants. Nucl. Eng. Des. 2008, 238, 1353–1372. [Google Scholar] [CrossRef]
  72. El-Genk, M.; Tournier, J. Performance Analyses of VHTR Plants With Direct and Indirect Closed Brayton Cycles and Different Working Fluids. Prog. Nucl. Energy 2009, 51, 556–572. [Google Scholar] [CrossRef]
  73. Wang, X.; Zhao, L. Analysis of zeotropic mixtures used in low-temperature solar Rankine cycles for power generation. Sol. Energy 2009, 83, 605–613. [Google Scholar] [CrossRef]
  74. Bao, J.; Zhao, L. A review of working fluid and expander selections for organic Rankine cycle. Renew. Sustain. Energy Rev. 2013, 24, 325–342. [Google Scholar] [CrossRef]
  75. Feng, Y.; Hung, T.; Zhang, Y.; Li, B.; Yang, J.; Shi, Y. Performance comparison of low-grade ORCs (organic Rankine cycles) using R245fa, pentane and their mixtures based on the thermoeconomic multi-objective optimization and decision makings. Energy 2015, 93, 2018–2029. [Google Scholar] [CrossRef]
  76. Sadeghi, M.; Nemati, A.; ghavimi, A.; Yari, M. Thermodynamic analysis and multi-objective optimization of various ORC (organic Rankine cycle) configurations using zeotropic mixtures. Energy 2016, 109, 791–802. [Google Scholar] [CrossRef]
  77. Modi, A.; Haglind, F. A review of recent research on the use of zeotropic mixtures in power generation systems. Energy Convers. Manag. 2017, 138, 603–626. [Google Scholar] [CrossRef]
  78. Dai, B.; Dang, C.; Li, M.; Tian, H.; Ma, Y. Thermodynamic performance assessment of carbon dioxide blends with low-global warming potential (GWP) working fluids for a heat pump water heater. Int. J. Refrig. 2015, 56, 1192145. [Google Scholar] [CrossRef]
  79. Wang, D.; Lu, Y.; Tao, L. Thermodynamic analysis of CO2 blends with R41 as an azeotropy refrigerant applied in small refrigerated cabinet and heat pump water heater. Appl. Therm. Eng. 2017, 125, 1490–1500. [Google Scholar] [CrossRef]
  80. Hu, J.; Liu, C.; Li, Q.; Shi, X. Molecular simulation of thermal energy storage of mixed CO2/IRMOF-1 nanoparticle nanofluid. Int. J. Heat Mass Transf. 2018, 125, 1345–1348. [Google Scholar] [CrossRef]
  81. Dai, B.; Li, M.; Ma, Y. Thermodynamic analysis of carbon dioxide blends with low GWP (global warming potential) working fluids-based transcritical Rankine cycles for low-grade heat energy recovery. Energy 2014, 64, 942–952. [Google Scholar] [CrossRef]
  82. Yang, M.H. The performance analysis of the transcritical Rankine cycle using carbon dioxide mixtures as the working fluids for waste heat recovery. Energy Convers. Manag. 2017, 151, 86–97. [Google Scholar] [CrossRef]
  83. Shu, G.; Yu, Z.; Tian, H.; Liu, P.; Xu, Z. Potential of the transcritical Rankine cycle using CO2-based binary zeotropic mixtures for engine’s waste heat recovery. Energy Convers. Manag. 2018, 174, 668–685. [Google Scholar] [CrossRef]
  84. Sánchez, C.J.; da Silva, A.K. Technical and environmental analysis of transcritical Rankine cycles operating with numerous CO2 mixtures. Energy 2018, 142, 180–190. [Google Scholar] [CrossRef]
  85. Lewis, T.; Wright, S.; Conboy, T. Supercritical CO2 Mixture Behavior for Advanced Power Cycles and Applications. In Proceedings of the Supercritical CO2 Power Cycle Symposium, Boulder, CO, USA, 24–25 May 2011. [Google Scholar]
  86. Jeong, W.S.; Jeong, Y.H.; Lee, J.I. Performance of S-CO2 Brayton Cycle with Additive Gases for SFR Application. In Proceedings of the Supercritical CO2 Power Cycle Symposium, Boulder, CO, USA, 24–25 May 2011. [Google Scholar]
  87. Jeong, W.S.; Lee, J.I.; Jeong, Y.H. Potential Improvements of Supercritical Recompression CO2 Brayton Cycle by Mixing Other Gases for Power Conversion System of A SFR. Nucl. Eng. Des. 2011, 241, 2128–2137. [Google Scholar] [CrossRef]
  88. Baik, S.; Lee, J.I. Preliminary Study of Supercritical CO2 Mixed With Gases for Power Cycle in Warm Environments. In Proceedings of the ASME Turbo Expo 2018: Turbomachinery Technical Conference and Exposition, Oslo, Norway, 11–15 June 2018; p. V009T38A017. [Google Scholar]
  89. Vesely, L.; Dostal, V.; Stepanek, J. Effect of Gaseous Admixtures on Cycles with Supercritical Carbon Dioxide. In Proceedings of the ASME Turbo Expo 2016: Turbomachinery Technical Conference and Exposition, GT2016, Seoul, Republic of Korea, 13–17 June 2016. [Google Scholar]
  90. Vesely, L.; Dostal, V. Effect of Multicomponent Mixtures on Cycles with Supercritical Carbon Dioxide. In Proceedings of the ASME Turbo Expo 2017: Turbomachinery Technical Conference and Exposition, GT2017, Charlotte, NC, USA, 26–30 June 2017. [Google Scholar]
  91. Vesely, L.; Manikantachari, K.R.V.; Vasu, S.; Kapat, J.; Dostal, V.; Martin, S. Effect of Mixtures on Compressor and Cooler in Supercritical Carbon Dioxide Cycles. In Proceedings of the ASME Turbo Expo 2018: Turbomachinery Technical Conference and Exposition, GT2018, Oslo, Norway, 11–15 June 2018. [Google Scholar]
  92. Jeong, W.S.; Lee, J.I.; Jeong, Y.H.; NO, H.C. Potential Improvements of Supercritical Recompression CO2 Brayton Cycle Coupled with KALIMER-600 by Modifying Critical Point of CO2. In Proceedings of the Transactions of the Korean Nuclear Society Autumn Meeting, Jeju, Republic of Korea, 21–22 October 2010. [Google Scholar]
  93. Invernizzi, C.M.; van der Stelt, T. Supercritical and real gas Brayton cycles operating with mixtures of carbon dioxide and hydrocarbons. Proc. Inst. Mech. Eng. Part A J. Power Energy 2012, 226, 682–693. [Google Scholar] [CrossRef]
  94. Vesely, L.; Dostal, V. Research on the Effect of the Pinch Point Shift in Cycles with Supercritical Carbon Dioxide. In Proceedings of the 4th International Symposium—Supercritical CO2 Power Cycles, Pittsburgh, PA, USA, 9–10 September 2014. [Google Scholar]
  95. Hu, L.; Chen, D.; Huang, Y.; Li, L.; Cao, Y.; Yuan, D.; Wang, J.; Pan, L. Investigation on the Performance of the Supercritical Brayton Cycle With CO2-based Binary Mixture as Working Fluid for An Energy Transportation System of A Nuclear Reactor. Energy 2015, 89, 874–886. [Google Scholar] [CrossRef]
  96. Bonalumi, D.; Lasala, S.; Macchi, E. CO2-TiCl4 working fluid for high-temperature heat source power cycles and solar application. Renew. Energy 2020, 147, 2842–2854. [Google Scholar] [CrossRef]
  97. Binotti, M.; Invernizzi, C.M.; Iora, P.; Manzolini, G. Dinitrogen tetroxide and carbon dioxide mixtures as working fluids in solar tower plants. Sol. Energy 2019, 181, 203–213. [Google Scholar] [CrossRef]
  98. Manzolini, G.; Binotti, M.; Bonalumi, D.; Invernizzi, C.; Iora, P. CO2 mixtures as innovative working fluid in power cycles applied to solar plants. Techno-economic assessment. Sol. Energy 2019, 181, 530–544. [Google Scholar] [CrossRef]
  99. Guo, J.Q.; Li, M.J.; Xu, J.L.; Yan, J.J.; Wang, K. Thermodynamic performance analysis of different supercritical Brayton cycles using CO2-based binary mixtures in the molten salt solar power tower systems. Energy 2019, 173, 785–798. [Google Scholar] [CrossRef]
  100. Yin, H.; Sabau, A.S.; Conklin, J.C.; McFarlane, J.; Qualls, A.L. Mixtures of SF6-CO2 as working fluids for geothermal power plants. Appl. Energy 2013, 106, 243–253. [Google Scholar] [CrossRef]
  101. Wu, C.; sen Wang, S.; Jiang, X.; Li, J. Thermodynamic analysis and performance optimization of transcritical power cycles using CO2-based binary zeotropic mixtures as working fluids for geothermal power plants. Appl. Therm. Eng. 2017, 115, 292–304. [Google Scholar] [CrossRef]
  102. Feng, L.; Zheng, D.; Chen, J.; Dai, X.; Shi, L. Exploration and Analysis of CO2 + Hydrocarbons Mixtures as Working Fluids for Trans-critical ORC. Energy Procedia 2017, 129, 145–151. [Google Scholar] [CrossRef]
  103. Zabetakis, M.G. Flammability Characteristics of Combustible Gases and Vapors; Technical Report; Bureau of Mines: Washington, DC, USA, 1965.
  104. Badr, O.; Probert, S.; O’Callaghan, P. Selecting a working fluid for a Rankine-cycle engine. Appl. Energy 1985, 21, 1–42. [Google Scholar] [CrossRef]
  105. Ginosar, D.M.; Petkovic, L.M.; Guillen, D.P. Thermal Stability of Cyclopentane as an Organic Rankine Cycle Working Fluid. Energy Fuels 2011, 25, 4138–4144. [Google Scholar] [CrossRef]
  106. Dai, X.; An, Q.; Shi, L. Experiment research for the thermal stability of isobutene and isopentane. J. Eng. Thermophys. 2013, 34, 1416–1419. (In Chinese) [Google Scholar]
  107. Barin, I.; Knacke, O. Thermochemical Properties of Inorganic Substances; Springer Berlin: Heidelberg, Germany, 1973. [Google Scholar]
  108. Fukuda, K.; Dokiya, M.; Kameyama, T.; Kotera, Y. Catalytic decomposition of hydrogen sulfide. Ind. Eng. Chem. Fundam. 1978, 17, 243–248. [Google Scholar] [CrossRef]
  109. Chivers, T.; Hyne, J.B.; Lau, C. The thermal decomposition of hydrogen sulfide over transition metal sulfides. Int. J. Hydrogen Energy 1980, 5, 499–506. [Google Scholar] [CrossRef]
  110. Weil, E.D.; Sandler, S.R.; Gernon, M. Kirk-Othmer Encyclopedia of Chemical Technology; John Wiley & Sons, Inc.: New York, NY, USA, 2006; Chapter Sulfur Compounds. [Google Scholar]
  111. Devarakonda, A.; Olminsky, J. An evaluation of halides and other substances as potential heat pipe fluids. In Proceedings of the 2nd International Energy Conversion Engineering Conference, Providence, RI, USA, 16–19 August 2004; p. 5575. [Google Scholar]
  112. Devarakonda, A.; Anderson, W.G. Thermo-Physical Properties of Intermediate Temperature Heat Pipe Fluids. AIP Conf. Proc. 2005, 746, 179–186. [Google Scholar]
  113. Invernizzi, C.; Iora, P.; Bonalumi, D.; Macchi, E.; Roberto, R.; Caldera, M. Titanium tetrachloride as novel working fluid for high temperature Rankine Cycles: Thermodynamic analysis and experimental assessment of the thermal stability. Appl. Therm. Eng. 2016, 107, 21–27. [Google Scholar] [CrossRef]
  114. Bombarda, P.; Invernizzi, C. Binary liquid metal-organic Rankine cycle for small power distributed high efficiency systems. Proc. Inst. Mech. Eng. Part A J. Power Energy 2015, 229, 192–209. [Google Scholar] [CrossRef]
  115. Krasin, A.; Nesterenko, V. Dissociating Gases: A New Class of Coolants and Working Substances for Large Power Plants. At. Energy Rev. 1971, 9, 177–194. [Google Scholar]
  116. Lemmon, E.W.; Bell, I.; Huber, M.L.; McLinden, M.O. NIST Standard Reference Database 23: Reference Fluid Thermodynamic and Transport Properties-REFPROP; Version 10.0; National Institute of Standards and Technology: Gaithersburg, MD, USA, 2018.
  117. Solovyev, E.; Kuvshinov, D.; Ermakov, D.; Kuvshinov, G. Production of hydrogen and nanofibrous carbon by selective catalytic decomposition of propane. Int. J. Hydrogen Energy 2009, 34, 1310–1323. [Google Scholar] [CrossRef]
  118. Dai, X.; Shi, L.; An, Q.; Qian, W. Screening of hydrocarbons as supercritical ORCs working fluids by thermal stability. Energy Convers. Manag. 2016, 126, 632–637. [Google Scholar] [CrossRef]
  119. Tsang, W. Thermal stability of cyclohexane and 1-hexene. Int. J. Chem. Kinet. 1978, 10, 1119–1138. [Google Scholar] [CrossRef]
  120. Zanetti, J.; Egloff, G. The Thermal Decomposition of Benzene. Ind. Eng. Chem. 1917, 9, 350–356. [Google Scholar] [CrossRef]
  121. Invernizzi, C.M.; Iora, P.; Manzolini, G.; Lasala, S. Thermal stability of n-pentane, cyclo-pentane and toluene as working fluids in organic Rankine engines. Appl. Therm. Eng. 2017, 121, 172–179. [Google Scholar] [CrossRef]
  122. Invernizzi, C.; Bonalumi, D. Thermal stability of organic fluids for Organic Rankine Cycle systems. In Organic Rankine Cycle (ORC) Power Systems; Macchi, E., Astolfi, M., Eds.; Woodhead Publishing: Sawston, UK, 2017; pp. 121–151. [Google Scholar]
  123. Dai, X.; Shi, L.; An, Q.; Qian, W. Thermal stability of some hydrofluorocarbons as supercritical ORCs working fluids. Appl. Therm. Eng. 2018, 128, 1095–1101. [Google Scholar] [CrossRef]
  124. Calderazzi, L.; di Paliano, P.C. Thermal stability of R-134a, R-141b, R-13I1, R-7146, R-125 associated with stainless steel as a containing material. Int. J. Refrig. 1997, 20, 381–389. [Google Scholar] [CrossRef]
  125. Angelino, G.; Invernizzi, C. Experimental investigation on the thermal stability of some new zero ODP refrigerants. Int. J. Refrig. 2003, 26, 51–58. [Google Scholar] [CrossRef]
  126. Okada, K.; Tschuikow-Roux, E.; Evans, P. Single-pulse shock-tube study of the thermal decomposition of ethyl fluoride and propyl chloride. J. Phys. Chem. 1980, 84, 467–471. [Google Scholar] [CrossRef]
  127. Ito, M.; Dang, C.; Hihara, E. Thermal decomposition of lower-GWP refrigerants. In Proceedings of the 15th International Refrigeration and Air Conditioning Conference at Purdue, Purdue University, West Lafayette, IN, USA, 14–17 July 2014. [Google Scholar]
  128. Lees, F.P. Loss Prevention in the Process Industries: Hazard Identification, Assessment, and Control; Butterworth-Heinemann: Oxford, UK, 1996; Volume 1. [Google Scholar]
  129. Ribeiro, N.; Casimiro, T.; Duarte, C.; Nunes da Ponte, M.; Aguiar-Ricardo, A.; Poliakoff, M. Vapor- Liquid Equilibrium and Critical Line of the CO2+ Xe System. Critical Behavior of CO2+ Xe versus CO2+ n-Alkanes. J. Phys. Chem. B 2000, 104, 791–795. [Google Scholar] [CrossRef]
Figure 1. Research on the sCO 2 cycle from 2002 to 2019. (a) Annual publications. (b) Application share.
Figure 1. Research on the sCO 2 cycle from 2002 to 2019. (a) Annual publications. (b) Application share.
Processes 11 02795 g001
Figure 2. The alternative sCO 2 cycle layouts. (a) Recuperated cycle. (b) Recompression cycle. (c) Precompression cycle. (d) Partial cooling cycle.
Figure 2. The alternative sCO 2 cycle layouts. (a) Recuperated cycle. (b) Recompression cycle. (c) Precompression cycle. (d) Partial cooling cycle.
Processes 11 02795 g002
Figure 3. The thermal efficiency of modified sCO 2 cycles listed in Table 1. (a) Summary of global thermal efficiencies. (b) Improved thermal efficiency as a function of TIT [9].
Figure 3. The thermal efficiency of modified sCO 2 cycles listed in Table 1. (a) Summary of global thermal efficiencies. (b) Improved thermal efficiency as a function of TIT [9].
Processes 11 02795 g003
Figure 4. Efficiency improvement of the sCO 2 cycle enhanced by the conventional method.
Figure 4. Efficiency improvement of the sCO 2 cycle enhanced by the conventional method.
Processes 11 02795 g004
Figure 5. The combined sCO 2 cycle layouts. (a) Combined recuperated cycle. (b) Combined recompression cycle. (c) Combined partial cooling cycle.
Figure 5. The combined sCO 2 cycle layouts. (a) Combined recuperated cycle. (b) Combined recompression cycle. (c) Combined partial cooling cycle.
Processes 11 02795 g005
Figure 6. Thermal efficiency of combined sCO 2 power cycles listed in Table 2. (a) Improved thermal efficiencies. (b) Global thermal efficiencies as functions of TIT [9].
Figure 6. Thermal efficiency of combined sCO 2 power cycles listed in Table 2. (a) Improved thermal efficiencies. (b) Global thermal efficiencies as functions of TIT [9].
Processes 11 02795 g006
Figure 7. The sCO 2 cycle thermal efficiency improvement with the combined cycle method. (a) +ORC. (b) +tCO 2 cycle. (c) +Kalina cycle.
Figure 7. The sCO 2 cycle thermal efficiency improvement with the combined cycle method. (a) +ORC. (b) +tCO 2 cycle. (c) +Kalina cycle.
Processes 11 02795 g007
Figure 8. Thermal efficiencies of the supercritical CO 2 mixture for both Brayton and Rankinecycles considered in Table 3. (a) Improved thermal efficiencies. (b) Efficiencies as a function of TIT [9].
Figure 8. Thermal efficiencies of the supercritical CO 2 mixture for both Brayton and Rankinecycles considered in Table 3. (a) Improved thermal efficiencies. (b) Efficiencies as a function of TIT [9].
Processes 11 02795 g008
Figure 9. Changes in the thermal efficiency of the CO 2 mixture’s Brayton and Rankine cycles with respect to the corresponding pure CO 2 cycles (magenta: Brayton cycle; blue: Rankine cycle).
Figure 9. Changes in the thermal efficiency of the CO 2 mixture’s Brayton and Rankine cycles with respect to the corresponding pure CO 2 cycles (magenta: Brayton cycle; blue: Rankine cycle).
Processes 11 02795 g009
Figure 10. The efficiency and power output changes for the sCO 2 cycles with additive gases.
Figure 10. The efficiency and power output changes for the sCO 2 cycles with additive gases.
Processes 11 02795 g010
Table 1. Numerical study on varied sCO 2 cycles enhanced by the conventional method.
Table 1. Numerical study on varied sCO 2 cycles enhanced by the conventional method.
Cycle LayoutInitial BehaviorImproved BehaviorApplicationRef.
CIT/°CTIP/MPaTIT/°C η /% Method η /%
Intercooled Recuperated312255039Recompression, Increase TIT41.8–47300 MWe Nuclear power2002 [27]
Recompression312055043.1, 45.8Increase TIT47, 49.9600 MWth Nuclear power2002 [27]
Recuperated322055040Reheating, Intercooling40.8–41.5600 MWth Nuclear power2004 [9]
Recompression3220550–75045.56–52Reheating, Increase TIT, Raise TIP46.775–57600 MWth Nuclear power2004 [9]
Recuperated357, 1280045.5, 45.8Intercooling, Partial cooling47.4–49.7Nuclear power2004 [20]
Partial cooling357, 1280049.2, 49.7Intercooling51.4, 51.9Nuclear power2004 [20]
Recompression31.252047239.1Condensation, Liquid phase pumping, Raise TIP39.5–43.1250 MWth Nuclear power2007 [12]
Recompression32, 5020, 30550, 75036.71–49.83Reheating37.65–50.78600 MWth Nuclear power2009 [28]
Recuperated3225550–75040.44–48.2Precompression, Recompression, Partial cooling, Increase TIT43.49–54.13600 MWth Nuclear power2011 [29]
Precompression322555043.49Increase TIT48.8, 52.43600 MWth Nuclear power2011 [29]
Recompression322555046.48Increase TIT49.9, 52.73600 MWth Nuclear power2011 [29]
Partial cooling322555046.12Increase TIT50.5, 54.13600 MWth Nuclear power2011 [29]
Recuperated3225550–70040.7–45.5Reheating, Increase TIT, Recompression, Partial cooling42.5–52.8CSP2012 [30]
Recompression3225550–70047.7–52.8Reheating, Increase TIT48.9–53.9CSP2012 [30]
Partial cooling3225550–70046.1–52Reheating, Increase TIT48.1–53.8CSP2012 [30]
Recompression3225500–60044.5–46Reheating, Increase TIT47–4910 MWe CSP2012 [31]
Recompression55.525500–85040–52Reheating, Increase TIT43.14–52.8CSP2015 [32]
Recuperated55.52560038.5Reheating, Recompression39.5, 42.7CSP2015 [33]
Recompression55.52560042.7Reheating, Intercooling43.2–45.3CSP2015 [33]
Recompression3225600–80045.1–52.4Reheating, Increase TIT46.5–53.42113 MWth Fossil energy2016 [34]
Recompression35, 5020500–80039.59–54.58Intercooling40.82–55.68Nuclear power2017 [21]
Recuperated with reheating272582746.3Intercooling48.7–49.4160 MWe Not specified2017 [35]
Recuperated-2570036.2Recompression, Precompression43.3104 MWe CSP2017 [22]
Recompression with reheating322550040.98–49.21Intercooling41.44–49.450 MWe CSP2017 [36]
Recompression512575047Intercooling48.625 MWe CSP2017 [37]
Table 2. Numerical study on sCO 2 combined cycles.
Table 2. Numerical study on sCO 2 combined cycles.
Cycle LayoutInitial BehaviorImproved BehaviorApplicationRef.
CIT/°CTIP/MPaTIT/°C η /% Cost Bottoming Cycle η /% Cost
Recompression3222.3550–75040.23–47.82-tCO 2 cycle44.25–52.98-90–165 kWe nuclear power2010 [41]
Recuperated3522.5527–72728.5–33.8-ORC (R245fa, isobutane, isopentane, cyclohexane)36.4–43.96-88–115 kWe CSP2011 [42]
Recuperated30-527–82728.1–38.7-ORC (isopentane, isobutane, n-Butane)35.3–47.5-500 kWe CSP2013 [43]
Recuperated5525700–85041.9–46.5-ORC (isopentane, n-butene, cis-butene)48–53-CSP2014 [44]
Recompression5525700-85046.3–50.6-ORC (R236ea, R245fa, isobutane)50.5–55.75-CSP2014 [44]
Partial cooling5525700–85046.2–51-ORC (R124, R245fa, isobutane)48.7–53.95-CSP2014 [44]
Recuperated3522.565033.8, 36.5-tCO 2 cycle39.5, 41.3-326 kWth Fuel cell2014 [45]
Recuperated3522.565039.6-tCO 2 cycle46.1-326 kWth Fuel cell2014 [46]
Recompression352052745.32-ORC (R123)52.12-570 kWth Nuclear power2014 [47]
Recompression3522.255039.59, 38.311.2 $/GJ, 10.87 $/GJORC (Isopentane, n-butane, isobutane, n-pentane, R123, R114, R141b, RC318)43.68–44.2210.73 $/GJ, 10.25 $/GJ600 MWth Nuclear power2014 [48]
Recompression322055044.296.2 $/kWtCO 2 cycle45.926.6 $/kW84 kWe Nuclear power2015 [23]
Recuperated322055037.46-tCO 2 cycle44.7-96 kWe Nuclear power2015 [49]
Recompression322055044.22-tCO 2 cycle47.69-81 kWe Nuclear power2015 [49]
Recompression3226.7, 21550--tCO 2 cycle44.99, 43.899.98 $/GJ, 9.75 $/GJ600 MWth Nuclear power2016 [50]
Recompression3228.12–29.53, 20.87–21.16550--ORC (R123, R245fa, toluene, isobutane, isopentane, cyclohexane)44.77–45.23, 43.78–44.089.88–9.96 $/GJ, 9.60–9.62 $/GJ600 MWth nuclear power2016 [50]
Recompression3521.555039.5711.2 $/GJ, 10.87 $/GJKalina cycle43.5510.73 $/GJ, 10.34 $/GJ600 MWth Nuclear power2016 [51]
Recompression352255039.6213.73 $/GJ, 13.64 $/GJKalina cycle42.813.04 $/GJ, 12.88 $/GJ600 MWth Nuclear power2016 [52]
Recuperated37-655--ORC (R245fa)-0.039 cents/kWh, 0.045 cents/kWhFuel cell2016 [53]
Recompression33-550--tCO 2 refrigeration49.9–50.5210.68–11.05 $/GJ600 MWth Nuclear power2017 [54]
Basic32.52038551.41-tCO 2 cycle52.53-9.25 MWe Exhaust gas2017 [55]
Recompression3121.2501.837.68-tCO 2 cycle50.060.046 $/kWh41.45 MWe Exhaust gas2017 [56]
Recompression322055041.47-tCO 2 cycle43.7511,243.15 $/h600 MWth Nuclear power2017 [57]
Recompression32-600–85044.5–51.9-tCO 2 cycle47.6–57.1-2.7–4.2 MWe Not specified2018 [58]
Recompression32-750--ORC (10 zeotropic mixtures)-10.93 $/GJ600 MWth nuclear power2018 [59]
Recuperated3131.06450.6--ORC (cyclopentane/R365mfc)-3.95–4.34 cents/kWhExhaust gas2018 [60]
Recuperated4020550--tCO 2 refrigeration42.3–44.5-2.4–3.0 MWe Exhaust gas2018 [61]
Recuperated-25380--ORC (R407c, R134a, R245fa, R1234yf, R1234ze)40.5–41.92-3.4–3.7 MWe CSP2018 [62]
Recuperated, recompression3616.4300--ORC (R123, R245fa, R600)17.7–19.1-200 kWe Not specified2018 [63]
Basic52–5520385--tCO 2 cycle49.39–51.440.042–0.0486 $/kWh8.886–9.255 MW Exhaust gas2018 [64]
Table 3. Numerical study on the supercritical power cycles based on CO 2 mixture.
Table 3. Numerical study on the supercritical power cycles based on CO 2 mixture.
Cycle LayoutInitial BehaviorImproved BehaviorApplicationRef.
CIT/°CTIP/MPaTIT/°C η /% Additives η /%
Recompression322050843.27He, Ar, N 2 , O 2 41–45.21529 MWth Nuclear power2011 [92]
Full-scale compressor test27–50---SF 6 , n-Butane, Neon-50 kW Compressor2011 [85]
Recompression322055045.37He, Ar, N 2 , O 2 44.02–47.1600 MWth Nuclear power2011 [87]
Recompression322055045.37Xe, Ar, N 2 , O 2 44.02–46.65600 MWth nuclear power2011 [86]
Recuperated403040021.5Benzene23.5–26.1Not specified2012 [93]
Recuperated322055039.77Xe, Kr, Ar, N 2 , O 2 39.75–41.46600 MWth Nuclear power2013 [18]
Recompression322055045.37Xe, Kr, Ar, N 2 , O 2 44–46.69600 MWth nuclear power2013 [18]
Recuperated442055037.97H 2 S, Cyclohexane37.88, 38.01600 MWth nuclear power2013 [18]
Recompression442055041.36H 2 S, Cyclohexane42.61, 43.23600 MWth nuclear power2013 [18]
Recuperated322060037.9N 2 , O 2 , Ar, Air, He37.3–37.6Not specified2014 [94]
Recompression322060042.6N 2 , O 2 , Ar, Air, He42.4–42.5Not specified2014 [94]
Recompression322065047.82O 2 , He, Ar, Kr45.9–50.5600 MWth nuclear power2015 [95]
Recompression522065041.1n-Butane, Cyclohexane42.55–43.4600 MWth nuclear power2015 [95]
Recompression342455032.45He, CO, O 2 , N 2 , Ar, H 2 , CH 4 , H 2 S31.39–32.4912.5 MWe Not specified2016 [89]
Precompression342455029.45He, CO, O 2 , N 2 , Ar, H 2 , CH 4 , H 2 S28.12–29.4816.6 MWe Not specified2016 [89]
Split expansion342455029.03He, CO, O 2 , N 2 , Ar, H 2 , CH 4 , H 2 S28.1–29.0611.2 MWe Not specified2016 [89]
Recompression342455032.45Air, M-I, M-II, M-H, Air-H, M-IH29.74–32.3112.5 MWe Not specified2017 [90]
Precompression342455029.45Air, M-I, M-II, M-H, Air-H, M-IH25.83–29.2916.6 MWe Not specified2017 [90]
Split expansion342455029.03Air, M-I, M-II, M-H, Air-H, M-IH26.62–28.9111.2 MWe Not specified2017 [90]
Recompression342755033.44He, H 2 , CO, O 2 , Ar, N 2 , CH 4 , H 2 S, Xe, Kr, SO 2 32.54–33.638 MWth not specified2018 [91]
Precompression342755044.44He, H 2 , CO, O 2 , Ar, N 2 , CH 4 , H 2 S, Xe, Kr, SO 2 43.14–44.6938 MWth not specified2018 [91]
Split expansion342755029.83He, H 2 , CO, O 2 , Ar, N 2 , CH 4 , H 2 S, Xe, Kr, SO 2 28.83–30.0338 MWth not specified2018 [91]
Basic402030013.15SF 6 , R32, R22, R123, R134a, Toluene13.86–14.491 MWth not specified2018 [88]
Recuperated402555131.3TiCl 4 36CSP2018 [96]
Recuperated402580040.7TiCl 4 44.7CSP2018 [96]
Recompression402555138.4TiCl 4 40.1CSP2018 [96]
Recompression402580047TiCl 4 49.9CSP2018 [96]
Recuperated5125550, 70040.4, 47.4N 2 O 4 40.8, 46.7100 MWth CSP2019 [97]
Recuperated5125550, 70040.44, 47.4N 2 O 4 , TiCl 4 40.63–49.1133–38 MWe CSP2019 [98]
Recuperated, recompression, partial cooling362555027.5–30.6Xe, n-Butane27.7–30.91 MWe CSP2019 [99]
TRC1510, 12175, 155-SF 6 13, 15Geothermal water2013 [100]
TRC2510120–2406.8–15.6R32, R1270, R161, R1234yf, R134a, R152a, R1234ze7.5–16.8520–45 kWe WHR2014 [81]
TRC2311–171708.97R32, R161, R290, R1234yf, R1234ze7.54–9.54112 kWe WHR2017 [82]
TRC20-1506.6R152a, R161, R290, R1270, R1234yf, R1234ze6.3–9223 kWe geothermal water2017 [101]
TRC20-1805.79Propane, n-Butane, isobutane, n-pentane, iso-pentane, neo-pentane7.25–7.6120 kWe Not specified2017 [102]
TRC2692776.55R290, R152a, R41, R32, R134a, R161, R1234yf, R1234ze8.7–11.513 kWe WHR2018 [83]
TRC30-200-R134a, R32, R152a, R41, R161, R1270, R1234ze(E), R1234yf-20–60 kWe WHR2018 [84]
Table 4. Physical, safety, and environmental data for promising additives.
Table 4. Physical, safety, and environmental data for promising additives.
TypeSubstancePhysical DataSafetyEnvironmental Data [116]
M /g·mol 1 P c /MPa T c /°C T D /°C ODP GWP
InorganicsH 2 S34.08999.95<444.6 [107,108]Highly Flammable0-
SO 2 64.067.884157.492000 [110]B10-
TiCl 4 189.74.661364.85>500 [113]Non Flammable0-
N 2 O 4 92.019.976158.230 [115]Non Flammable0-
HCPropane44.104.25196.74525 [117]Highly flammable03.3
n-Butane58.123.80152300–320 [118]A30~20
Pentane72.153.37196.55280–300 [118]A30~20
Cyclohexane84.164.08280.45700–800 [119]Flammable020
Benzene78.114.907288.87<500 [120]Highly flammable020
Toluene92.144.126318.6350–400 [121,122]Flammable03.35
HCFCR123152.9313.662183.68200–220 [123]B10.0177
R125120.023.617766.023396 [124]A103420
R127042.084.6692.4-A30~20
R134a102.0324.06101.1350–370 [122]A101430
R143a84.043.76172.71>350 [125]A204400
HFCR152a66.0514.52113.3160–180 [123]A20133
R16148.065.01102.1427 [126]A3012
R2286.4684.9996.145-A10.0551700
R227ea170.032.925101.75>425 [125]A103300
R236fa152.043.2124.92380–400 [123]A109820
R3252.0245.78278.1570–590 [127]A20675
R404a97.603.73572.12--03700
R407c86.204.63286.2-A101530
R4134.035.89744.13--0107
R410a72.584.90171.34-A101730
Table 5. Strengths and weaknesses of sCO 2 cycle improvement methods.
Table 5. Strengths and weaknesses of sCO 2 cycle improvement methods.
MethodStrengthsWeaknesses
Conventional method
  • Moderate efficiency rise
  • Technology proven (on steam Rankine cycles)
  • Adding component(s)
  • Complex layout
  • High TIT
  • Large pressure drop
Combined cycle
Good waste heat recovery
Large efficiency rise
Combined cold and power cogeneration
Additional components
Capital cost increase
Complex layout
Wet cooling needed
CO 2 mixture
  • Low efficiency rise
  • Cost-effective
  • Applicable to hot-arid environments
  • Thermal stability of additives
  • Control of additives
  • Chemical reaction with materials
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

Wang, X.; Zhang, L.; Zhu, Z.; Hu, M.; Wang, J.; Fan, X. Performance Improvement Overview of the Supercritical Carbon Dioxide Brayton Cycle. Processes 2023, 11, 2795. https://doi.org/10.3390/pr11092795

AMA Style

Wang X, Zhang L, Zhu Z, Hu M, Wang J, Fan X. Performance Improvement Overview of the Supercritical Carbon Dioxide Brayton Cycle. Processes. 2023; 11(9):2795. https://doi.org/10.3390/pr11092795

Chicago/Turabian Style

Wang, Xurong, Longwei Zhang, Zhenhua Zhu, Mingjiang Hu, Jing Wang, and Xiaowei Fan. 2023. "Performance Improvement Overview of the Supercritical Carbon Dioxide Brayton Cycle" Processes 11, no. 9: 2795. https://doi.org/10.3390/pr11092795

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