Next Article in Journal
Calculation of Processes in the Electric Drive of the Sucker-Rod Pumping Unit with Reactive Power Compensation
Next Article in Special Issue
The Role of Education and Science-Driven Tools in Scaling Up Photovoltaic Deployment
Previous Article in Journal
Identification of Key Factors for the Development of Agricultural Biogas Plants in Poland
Previous Article in Special Issue
Two-Dimensional Geothermal Model of the Peruvian Andes above the Nazca Ridge Subduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Power Generation with Renewable Energy and Advanced Supercritical CO2 Thermodynamic Power Cycles: A Review

by
Xinyu Zhang
and
Yunting Ge
*
School of the Built Environment and Architecture, London South Bank University, 103 Borough Road, London SE1 0AA, UK
*
Author to whom correspondence should be addressed.
Energies 2023, 16(23), 7781; https://doi.org/10.3390/en16237781
Submission received: 26 October 2023 / Revised: 20 November 2023 / Accepted: 24 November 2023 / Published: 26 November 2023
(This article belongs to the Collection Renewable Energy and Energy Storage Systems)

Abstract

:
Supercritical CO2 (S-CO2) thermodynamic power cycles have been considerably investigated in the applications of fossil fuel and nuclear power generation systems, considering their superior characteristics such as compactness, sustainability, cost-effectiveness, environmentally friendly working fluid and high thermal efficiency. They can be potentially integrated and applied with various renewable energy systems for low-carbon power generation, so extensive studies in these areas have also been conducted substantially. However, there is a shortage of reviews that specifically concentrate on the integrations of S-CO2 with renewable energy, encompassing biomass, solar, geothermal and waste heat. It is thus necessary to provide an update and overview of the development of S-CO2 renewable energy systems and identify technology and integration opportunities for different types of renewable resources. Correspondingly, this paper not only summarizes the advantages of CO2 working fluid, design layouts of S-CO2 cycles and classifications of renewable energies to be integrated but also reviews the recent research activities and studies carried out worldwide on advanced S-CO2 power cycles with renewable energy. Moreover, the performance and development of various systems are well grouped and discussed.

1. Introduction and Motivation

Global energy demand is on the rise in numerous countries as the population increases and the economy grows. One of the primary and representative facts of economic growth in all countries worldwide is the increased electricity demand and generation. From 2019 to 2022, total primary energy consumption experienced a 3% increase, while the proportion of fossil fuel consumption in relation to primary energy consumption remained high at approximately 82% [1]. Growing energy consumption makes it a significant challenge to transition our energy sources and supplies away from fossil fuels and move towards low-carbon ones. Extensive fossil fuel consumption has caused environmental issues such as global warming, ozone depletion and atmospheric pollution. Therefore, it is critically important to achieve energy conversion efficiency improvement and increase renewable energy utilization and thus diminish the reliance on fossil fuels.
A decentralized and localized power generation system is an effective option to reduce energy consumption since energy loss from long-distance transmission or transportation can be avoided. Steam Rankine and gas turbine cycles have been significantly involved in large-scale power plants for electricity generation. The steam Rankine cycle can, however, achieve relatively higher energy conversion efficiency at lower turbine inlet temperature, considering the fact that the working fluid is pumped at a liquid state before being heated by a steam boiler. This is significant because liquid water is essentially incompressible, which can reduce the work required for pumping compared to compressing a gas by a compressor. According to the theory of the Carnot cycle for a heat engine, the higher the temperature at which heat is supplied and the lower the temperature at which heat is rejected, the greater efficient the cycle will be. It reveals the importance of a power cycle operating at a higher heat supply temperature whenever possible to maximize its thermal efficiency. The gas turbine cycle, also known as the Brayton Cycle, utilizes air as a working fluid at a higher turbine inlet temperature. However, in the Brayton Cycle, the compression process consumes more work compared to the steam Rankine cycle since air is compressible. This leads to lower efficiency for the gas turbine cycle. On the other hand, over the last decades, as a mature energy conversion technology, the organic Rankine cycle (ORC) has been generally used in small-scale power generation. Industrial waste heat, biomass combustion, solar energy and geothermal heat can be utilized as heat sources for ORCs [2,3]. The ORC is superior to the traditional steam Rankine cycle in terms of performance when employing a low-temperature heat source. However, the selection of a working fluid for an ORC is a challenge due to the fact that some organic compounds have high global warming potentials (GWPs). Therefore, the investigation of alternative working fluids in applicable power generation cycles attracted more attention [4,5]. CO2 is a natural, non-toxic, non-flammable, abundant and zero ozone depletion potential (ODP) working fluid, making it a noteworthy competitive candidate to be utilized in a power generation system [6]. According to its low critical point, the CO2 power cycle can easily extend to the supercritical region and turn it into a transcritical or supercritical CO2 power cycle. The supercritical CO2 (S-CO2) power cycle was originally proposed as a format of partial condensation Brayton cycle by Sulzer [7] in 1940. A simple recuperated supercritical Brayton cycle was propositioned by Feher [8] in 1967, and thereafter, Angelino [9,10,11] conducted a thorough study of CO2 power cycles. The S-CO2 Brayton cycle integrates the strengths of the steam Rankine and the gas turbine cycles by compressing the working fluid in the incompressible region and achieving higher thermal efficiency at higher turbine inlet temperatures.
S-CO2 power cycles equipped with advantages of simplicity, compactness, sustainability and cost-effectiveness have therefore been significantly researched for various applications, including fossil fuel power plants, nuclear power plants, and integrations with renewable energies. The existing power cycles, such as Rankine and gas turbine cycles, face the challenges of working fluid selections and less cycle thermal efficiencies. Driven by the better thermal performance of energy conversion systems, CO2 has attracted more attention to be utilized as an alternative working fluid. Another reason is the lower critical temperature at 31 °C which enables CO2 working fluid and its associated system to be applicable in a variety range of heat sources. A power cycle with CO2 working fluid could be classified as either a direct or indirect Brayton cycle. The semi-closed direct-fire oxyfuel Brayton cycle was suitable for fossil fuel power generation systems, while the indirect Brayton cycle was adapted for applications involving nuclear and renewable energy sources [12]. There have been many studies investigating the applications of direct S-CO2 cycles. A design was conceptualized by Le Moullec [13] for a coal-fired power plant with 90% post-combustion CO2 capture. It showed that the cost of electricity and the cost of CO2 reduction could be reduced by 15% and 45%, respectively. Mecheri and Le Moullec [14] conducted additional simulations on S-CO2 coal-fired power plants, concluding that by combining a coal-fired power plant with a double reheated single recompression Brayton cycle, a net thermal efficiency of 47.8% could be attained. This efficiency was significantly higher than that of traditional coal-fired power plants. Xu et al. [15] reached a similar conclusion, stating that the net power generation efficiency of a triple-compression S-CO2 cycle could attain 49.01%, compared to 48.12% achieved by the water–steam Rankine cycle. In addition, although the fabrication cost could be increased, the levelized cost of electricity achieved a reduction of 1.32% compared to the water–steam Rankine cycle. Optimizing the recuperator in the system is crucial to reducing the total cost of the entire S-CO2 cycle. To further enhance the performance of the S-CO2 coal-fired power plant, employing a partial flow mode in the S-CO2 boiler was demonstrated to effectively decrease pressure drop, thereby increasing the system’s thermal efficiency [16]. As for the indirect Brayton cycle, significant attention was paid to the S-CO2 applications in nuclear reactors, as they could offer higher turbine inlet temperatures, potentially enhancing the system’s thermal efficiency. Various countries have undertaken efforts to develop Generation IV nuclear reactors, given that these reactors operate in the temperature range of 500 °C to 900 °C, compared to water-cooled reactors with operating temperatures at around 300 °C [17]. Dostal [18] conducted an in-depth analysis of the S-CO2 cycle for nuclear reactors. The study revealed that the S-CO2 cycle exhibited competitive system efficiency compared to the helium Brayton cycle at the same operating condition. Additionally, it was concluded that the S-CO2 power cycle was apt for all nuclear reactors with a core CO2 gas heater outlet temperature exceeding 500 °C. Moisseytsev and Sienicki [19] investigated several alternative cycle layouts for S-CO2 Brayton cycles coupled to the sodium-cooled fast reactor (SFR). It was found that no advantages could be gained from utilizing a double recompression cycle, incorporating intercooling between the main compressor stages or applying reheating between the high-pressure and low-pressure turbine cycles. However, optimizing the minimum cycle temperature down to 20 °C could lead to improvement in cycle efficiency. Wright et al. [20] demonstrated that integrating light water reactors (LWRs) with an S-CO2 cycle could potentially improve the LWR power cycle efficiency and save capital costs. In order to further evaluate the benefits of coupling an S-CO2 Brayton cycle with a small- and medium-sized water-cooled nuclear reactor (SWR), the effects of different operating conditions on system performance were studied by Yoon et al. [21]. The results revealed that even though the S-CO2 cycle was previously recognized for having superior efficiency in high operating temperature regions, the efficiency of S-CO2 with SMR cycle still demonstrated competitive efficiency to the existing steam-Rankine cycle with SMRs (around 30%) at the optimum pressure ratio. Additionally, a conclusion aligning with Wright [20] was reached that by combining the S-CO2 cycle with SWR, capital costs could also be reduced. In 2021, about 440 nuclear power reactors were in operation in 32 countries worldwide with a total generated power of 2653 TWh, which took about 10% of the global electricity supply [22].
In light of growing environmental concerns, the utilization of the S-CO2 power cycle in the renewable energy sector has recently emerged as an appealing option. While several researchers have provided extensive reviews of S-CO2 cycles, as outlined in Table 1, the majority have focused on the development status of S-CO2 cycles and the progress in optimizing components for applications in fossil fuel, nuclear, and solar power regions. Renewable energy resources are also known as alternative, sustainable or nonconventional energy supplies, including solar, geothermal, biomass and waste heat. However, reviews exclusively addressing S-CO2 cycles in the context of these renewable energies are limited. Therefore, the objective of this paper is to review the most recent development of supercritical CO2 cycles in power generation systems with renewable energy by offering a comprehensive view of the advantages of supercritical CO2 working fluid, the landscapes of renewable energy, the options of S-CO2 cycles, and application status of S-CO2 renewable energies. Below are the key points highlighted in each section:
(i)
In Section 2, the superior thermal–physical properties of CO2 are outlined, along with the benefits of incorporating CO2 in supercritical power generation cycles.
(ii)
Section 3 demonstrates the advantage characteristics and categorizations of renewable energy as a promising source of heat, encompassing biomass, solar, geothermal and waste heat.
(iii)
In Section 4, representative S-CO2 cycles are summarized with emphasis on features of each layout, T-S diagrams and thermodynamic equations.
(iv)
In Section 5, a review of recent applications of S-CO2 renewable power systems is presented, including S-CO2 for biomass power systems, S-CO2 cycle for concentrating solar power systems, S-CO2 cycle for geothermal power systems and S-CO2 cycle for waste heat recovery. This focuses on various technologies, operating conditions and efficiencies. In addition, the barriers to S-CO2 technology are also concluded.

2. Superior Thermal–Physical Properties of CO2

CO2 is a natural, non-toxic and non-flammable working fluid that possesses excellent thermophysical properties, including higher density, latent heat, specific heat, thermal conductivity and volumetric cooling capacity, along with lower viscosity [28]. These attributes make CO2 a significant player in various energy conversion systems. It has a low critical temperature of 31 °C but quite a high critical pressure of 7.4 MPa. As shown in Figure 1, near the pseudocritical region, the thermophysical properties of CO2 undergo rapid changes due to its high density and low compressibility factor close to its critical point. This leads to significant fluctuations in density and specific heat capacity with only slight variations in pressure or temperature. The compressibility factor is characterized as the ratio of the actual volume of a substance to its ideal volume. As observed in Figure 2, around the critical point, the compressibility factor of CO2 fluctuates between 0.2 and 0.5, which leads to a reduction in the power consumption of the compressor. Consequently, the S-CO2 cycle is distinguished by its high thermal efficiency, simple cycle configuration and compactness of system components. At the same power generation, the overall size of a steam Rankine cycle is estimated to be approximately four times larger than that of an S-CO2 Brayton cycle [17]. The benefits of utilizing CO2 working fluid in thermodynamic cycles include (1) environmentally friendly nature, with no ozone-depleting potential (ODP) and neglectable global warming potential (GWP = 1), (2) abundance, non-toxicity and non-combustibility, (3) non-reactivity with component materials, and (4) superb thermodynamic and transport properties. According to the low critical temperature of CO2 working fluid, its associated thermodynamic cycle can easily traverse both subcritical and supercritical regions with high-temperature heat sources, naming the cycles as either transcritical or supercritical cycles. Three options exist for utilizing CO2 in the Brayton cycle as shown in Figure 3: (1) the classic cycle (‘SN’), which operates entirely under critical pressure, known as the subcritical cycle; (2) the transcritical cycle (‘TN’), the highest pressure operates above critical pressure, allowing CO2 to pass through both subcritical and supercritical regions; and (3) the supercritical cycle (‘S’), which operates entirely above the critical pressure. However, the difference between the transcritical and supercritical Rankine cycles is not strict. The transcritical Rankine cycle, under some conditions, can also be called the supercritical Rankine cycle. The concept of the supercritical CO2 Rankine cycle pertains to the situation where heat addition happens at CO2 pressure above the critical point, while heat rejection occurs at CO2 pressure below the critical point, as illustrated in Figure 4. It has been demonstrated that the efficiency of the S-CO2 Brayton cycle surpasses that of the superheated steam Rankine cycle when the turbine inlet temperature exceeds 470 °C [29]. The potential for maximizing power output in an ORC is hindered by the evaporation process with constant temperature, making it a less favourable option for sensible heat sources. By avoiding the isothermal boiling process, CO2 in the transcritical/supercritical cycle achieves a more effective thermal match with the heat source, thereby attaining higher thermal efficiency compared to ORC, as depicted in Figure 5. The issue of pinching points between the heat source and working fluid temperature along the heat exchanger is effectively addressed by utilizing a transcritical or supercritical CO2 cycle.

3. Superior Characteristics of Renewable Energy

3.1. Biomass

Biomass, a type of non-fossilized and biodegradable organic material originating from plants, animals and microorganisms, has emerged as a global frontrunner for the development of low-carbon energy. It encompasses products, by-products, residues and waste from agriculture, industry and forestry. The energy stored in biomass is initially derived from the sun, with photosynthesis being the primary process through which plants convert the sun’s radiant energy into chemical energy, stored as glucose or sugar. Biomass is categorized into two main groups: virgin biomass, which includes forest biomass, energy crops and grasses; and waste biomass, which comprises municipal solid waste, agricultural crop residues and leaves [33]. Serving as a renewable and sustainable energy source, biomass can be utilized to generate electricity or other forms of energy. Unlike fossil fuels, the CO2 produced from the complete combustion of biomass is equivalent to the amount it absorbs from the atmosphere, resulting in no net contribution to atmospheric carbon dioxide levels and emitting low levels of SOx and NOx. There are two main thermochemical conversion routes to use biomass for supplying electricity and heating, i.e., gasification and combustion. Combustion is the most mature technology to convert biomass to useful electricity and heating. Using biomass as a heat source can not only effectively reduce biomass waste but also provide high temperatures definitely higher than 900 °C [34] during the combustion process. Figure 6 shows the biomass energy conversion processes and temperature ranges of the thermochemical route.
Biomass has already become a crucial element in the UK’s energy supply, contributing to 11% of the total electricity generated in 2022. According to the most recent energy statistics from the same year, it was estimated that bioenergy made up approximately 8.6% of the UK’s overall energy supply [35]. The key advantages and disadvantages of biomass applications are outlined in Table 2 below.

3.2. Solar Power

Solar energy is the most abundant and widely distributed form of renewable energy available for utilization. Out of the 1.75 × 105 terawatts (TW) representing the total energy from the sun reaching Earth’s atmosphere, approximately 1 × 105 TW consistently reaches the Earth’s surface [36]. In 2022, solar energy production reached a total of 1289.27 terawatt-hours (TWh), contributing to approximately 4.6% of the global electricity generation [37].
There are two approaches to converting solar energy for electricity generation: the photovoltaic (PV) cell system and the solar thermal system. Generally, multiple PV cells are linked in a series to capture sunlight and transform it into direct current (DC) electrical power. Nevertheless, PV systems are primarily employed on a smaller scale and can be affected by weather conditions. In contrast, solar thermal systems are suitable for larger-scale applications. Conventional concentrating solar power (CSP) is a well-known sunlight conversion technology. Typically, a CSP consists of a central receiver system, parabolic trough, dish Stirling unit and integrated gas cycles. The CSP plant generates electricity by using a linear or punctual collector to focus on radiation energy and convert it into high-temperature heat. This high-temperature heat is then transferred into a working fluid, and the absorbed heat is thus converted into electricity by a generator via a power cycle. The Rankine cycle is the most common technology for converting solar energy to electricity [38]. The main components of a solar ORC system are the solar collector and evaporator, energy storage, turbine, generator and condenser, as shown in Figure 7. However, as explained previously, the ORC is more appropriate for low-temperature heat sources [39]. In addition, oil, salt and steam are typical heat transfer fluids used in CSP systems for converting solar energy into electricity. Nevertheless, the properties of these heat transfer mediums can impose constraints on the performance of CSP plants. For instance, synthetic oil and salt can only withstand temperatures up to 400 °C and 560 °C, respectively, and direct steam generation has limitations in terms of storage capacity [40]. From the study of Dostal et al. [29], the S-CO2 cycle outperformed the steam Rankine cycle when the inlet temperature of the turbine exceeded 550 °C. This temperature range falls within the attainable parameters of solar power systems, with the hot-end temperatures having reached between 800 °C and 1000 °C [41] or even above 1000 °C [42]. The challenges of the development of solar thermal power systems are the lower efficiency and high capital cost. As an alternative promising technology, the S-CO2 cycle has become more favourable for converting solar power to electricity since it has higher thermal efficiency.

3.3. Geothermal Resource

Geothermal energy is globally available, constituting a clean, plentiful and sustainable energy source originating primarily from the natural decay of radioactive isotopes during the Earth’s formation, deeply embedded within its layers. There is approximately 43,000,000 EJ of geothermal energy stored at depths reaching as far as 3000 m below the surface [44].
Hydrothermal resources represent the dominant form of geothermal energy harnessed for large-scale electricity generation. In addition to hydrothermal sources, there are five other categories of geothermal energy: hot dry rock, geopressured, magma energy, deep hydrothermal and low-temperature systems [45]. The temperature range of the geothermal heat resources is between 50 °C and 350 °C [46]. Geothermal resources can also be classified into various temperature grades, which are high-temperature (>180 °C), intermediate temperature (100–180 °C) and low-temperature (<100 °C) [39]. In order to extract heat at a suitable temperature, it is typically necessary to drill holes into the ground, creating both production and injection wells. Table 3 shows the potential of different heat source temperature ranges of geothermal energy in Europe.

3.4. Waste Heat Resource

Low-temperature exhaust stream emission is still a notable issue. In many manufacturing industries, 20~50% of the energy consumed by manufacturing is lost as waste heat [48]. However, this heat cannot be recovered completely on-site and used for district heating. It is then discharged into the ambient, which has a significant negative impact on human health, biodiversity, and the environment. Generally, waste heat can be categorized as low-temperature (<230 °C), medium-temperature (230–650 °C) and high-temperature (>650 °C) [3]. Recovery of low-grade waste heat for electricity production is a promising technology to protect the environment and meet electricity demand, although it is a big challenge for power plants. There are several technologies for waste heat recovery based on the heat transfer between working fluid and waste heat. The organic Rankine cycle is a common method utilized for low-grade waste heat recovery, in which heat is immediately recovered by a heat transfer loop to evaporate the working fluid. Persichilli et al. [11,25] discovered that the transcritical CO2 power cycle could attain superior efficiency across a broad spectrum of heat source temperatures, ranging from 204 °C to 650 °C, and be more cost effective in comparison to ORC and steam Rankine cycles. Similarly, Chen et al. [32] conducted a comparative analysis between transcritical CO2 Rankine and R123 ORC cycles, utilizing a low-grade heat source with a temperature of 150 °C. Their investigation established that the transcritical CO2 cycle outperformed the subcritical R123 ORC one in terms of efficiency. However, for the low-grade heat source temperature of around 100 °C, the power output of the transcritical R125 cycle was approximately 14% higher than that of the transcritical CO2 cycle. Moreover, if the heat temperature is as low as approximately 112 °C, the transcritical CO2 Rankine cycle exhibits greater efficiency than that of the transcritical CO2 Brayton cycle due to reduced compression work while achieving the same temperature increase. The supercritical CO2 Brayton cycle is expected to replace the organic Rankine cycle to improve thermal efficiency for waste heat recovery. In addition, the S-CO2 cycle can recover waste heat from small turbines.

4. S-CO2 Layouts

The progressions of the five representative S-CO2 cycles are delineated in Figure 8, including the recuperation cycle [8], recompression cycle [18,29,49,50], pre-compression cycle [10,51], intercooling cycle [52] and reheating cycle [53]. Various configurations have been explored to enhance the performance of S-CO2 cycles, building upon the fundamental S-CO2 power system depicted in Figure 8a. Thermodynamic equations of each component and thermal efficiencies of different layouts are summarized in Table 4 and Table 5, respectively. The introduction of a recuperator allows for the recovery of more waste heat, resulting in what is termed the recuperation S-CO2 cycle, as illustrated in Figure 8b. However, the adoption of an internal heat exchanger or recuperator, while boosting electrical efficiency, has revealed a significant internal irreversibility issue within the recuperator. This is primarily due to the substantial difference in specific heat capacity between the cold and hot fluid sides of CO2, leading to a pinch-point problem [49]. The smaller the pinch-point, the better the heat transfer efficiency; however, a larger heat transfer area could be required. To address the pinch-point problem, a recompression S-CO2 cycle was developed, as shown in Figure 8c. In this configuration, an additional recuperator and compressor are introduced. A portion of CO2 is cooled by a gas cooler and then compressed to the highest pressure by the main compressor. The remaining amount of CO2 is compressed by the secondary compressor. These two streams are combined at the inlet of the high-temperature recuperator (HTR or Recuperator 1) on the high-pressure side. This minimizes the temperature difference between the hot and cold fluid sides of CO2, thereby improving heat transfer performance and resolving the pinch-point issue. Recompression S-CO2 is the most efficient cycle compared to internal cooling, reheating and pre-compression cycles [18]. The pre-compression configuration offers an alternative method to mitigate the impact of the pinch-point problem and enhance regeneration, as seen in Figure 8d. The pre-compression cycle was initially introduced by Angelino [10]. It involves the placement of a pre-compressor between the high-temperature recuperator (HTR) and the low-temperature recuperator (LTR or Recuperator 2). This pre-compression cycle works by narrowing the gap in specific heat capacity between the low-pressure and high-pressure streams via the elevation of pressure in the low-pressure stream. Intercooling is the traditional method to minimize the load of the compressor and increase the thermal efficiency of the S-CO2 cycle. Increasing the number of stages results in the compression process approaching near-isothermal conditions at the compressor’s inlet temperature [54]. However, intercooling (Figure 8e) does not hold much appeal in S-CO2 cycles due to the minimal efficiency enhancement it provides [18]. Reheating (Figure 8f) is the technology to improve the turbine work and thus enhance the thermal efficiency of the S-CO2 cycle since work output from the turbine could be increased without increasing the maximum temperature in the cycle and keeping the compressor work constant. Reheating holds substantial promise for the development of the S-CO2 cycle. Nonetheless, it is exclusively applicable to indirect cycles, and utilizing more than one reheat stage is not economically practicable.

5. Application Status of S-CO2 Renewable Power Systems

5.1. S-CO2 for Biomass Power Systems

Typical biomass power generation technologies have been listed in Table 6. Biomass can be combusted directly within waste-to-energy facilities for electricity production. Biomass co-firing is the substitution of a portion of the fuel with biomass within coal-fired thermal power plants. It is an economical approach to convert biomass effectively and environmentally into electricity and involves integrating biomass as a partial replacement for fuel in high-efficiency coal boilers [55]. Sweden has successfully operated the initial integrated gasification combined cycle (IGCC) plant that utilizes 100% biomass, specifically straw [56]. However, biomass poses a constraint on the adoption of large-scale steam cycles or IGCCs, which are designed to achieve higher efficiencies. The majority of biomass plants are typically small-scale and rely on internal combustion engines and ORCs. The electrical efficiency of a biomass-fired ORC system was between 7.5% and 13.5% [57]. Subsequently, in an experimental study conducted by Qiu et al. [58], it was observed that the electricity generation efficiency of this biomass-fired ORC system was 1.41%, primarily attributed to the lower efficiency of the expander and alternator during the experiments.
For the further development of biomass conversion technologies, a supercritical CO2 power system is a promising option for the efficient utilization of abundant renewable resources. Biomass has the potential to be used for power generation and bio-synthetic production with zero CO2 emission. Chitsaz et al. [59] proposed a novel tri-generation system of bio-synthetic nature gas, fresh water and power, as illustrated in Figure 9. The operational concept of this system involves introducing biomass into the fluidized bed gasifier, where it undergoes a reaction with steam to produce a mixture of highly combustible species such as H2, CO and CH4. Subsequently, this high-temperature hydrogen-rich syngas flows through a CO2 gas heater, elevating the temperature of CO2 to drive the S-CO2 Brayton cycle, and subsequently enters the methanation reactor. By utilizing heat from heat exchangers 3 and 4 within the humidification–dehumidification loop, seawater can be transformed into freshwater. By carrying out a comprehensive simulation of this system, it was found that the power production of 172.6 kW could be achieved under multi-objective optimum conditions. An alternative approach to utilize biomass gasification combined with the S-CO2 Brayton power cycle is the combustion of biofuel. Cao et al. [60] introduced a heat and power system that integrates biomass gasification with an advanced solid oxide fuel cell-CO2 supercritical Brayton cycle, as shown in Figure 10. After the biofuel (solid oxide fuel cell) is produced via the peach stone gasification process, it is burned in post-combustion to generate high-temperature gas up to 600 °C~700 °C. The results indicated that the optimal conditions for achieving maximum power (138 kW) and heat (195 kW) were as follows: an equivalence ratio of 4, a fuel cell temperature of 680 °C, a fuel utilization factor of 0.82 and a pressure ratio of 5.11. To enhance the electrical power output even further, biomass gasification with CHP plants can be equipped with multiple power cycles. Ji-chao and Sobhani [61] conducted a mathematical modelling study on an integrated power and heat system that merged biomass gasification with both the S-CO2 Brayton cycle and the Kalina cycle. They achieved a peak net power output of 7.375 MW when the pressure ratio was set at 3.5. Moradi et al. [62] conducted a comparison study between a gas turbine and an S-CO2 cycle, both of which were coupled with bottom ORCs and heated via biomass gasification. The results indicated that the average net electric power output of the entire integrated S-CO2 system was approximately 126 kW, which was about 25% higher than the power output of the gas turbine system.
For the combustion biomass conversion route, a small-scale power generation test system with biomass and CO2 transcritical Brayton cycles has been designed and constructed with purposely selected and manufactured system components [63,64,65], as depicted in Figure 11. Based on their modelling findings, it was discovered that there was an ideal pressure ratio that maximizes the thermal efficiency of the system. Furthermore, a higher temperature of the biomass flue gas was associated with higher thermal efficiency. As a single S-CO2 cycle may utilize high-temperature flue gas insufficiently, a concept of cascaded supercritical CO2 Brayton cycles was proposed to optimize the conversion efficiency of biomass energy into electricity, with a potential maximum efficiency of up to 36% [66]. Wang et al. [67] conducted a thermodynamic analysis of a biomass-solar combined with S-CO2 Brayton power generation system. The combined use of solar energy and biomass enabled the continuous operation of the system. This is because biomass steps in to supply heat, either partially or entirely, when solar irradiation falls short. This system consists of a solar island, a biomass burner, a recompression S-CO2 cycle and a simple recuperation S-CO2 cycle. Results showed that the solar-to-electric efficiency could achieve up to 27.85%. More studies of biomass energy conversion technologies based on S-CO2 cycles have been summarized in Table 7. Although higher efficiency in biomass conversion to electricity can be obtained by using the supercritical CO2 Brayton cycle, some barriers still exist for bioenergy development. The limited large-scale application of biomass can be attributed to several factors, including the need for logistics for feedstock collection and transportation, elevated feedstock costs compared to fossil fuels and greater upfront capital investment requirements.

5.2. S-CO2 Cycle for Concentrating Solar Power Systems

Globally, there are 144 concentrated solar power (CSP) projects running in 22 countries, where Spain, the United States and China are at the forefront in terms of construction and operation [73]. However, among those projects, steam Rankine cycles or organic Rankine cycles are still the primary technologies for converting solar energy to electricity. In comparison to photovoltaic (PV) panel technologies, CSP exhibits an inherent capacity to retain thermal energy for a short time, allowing for its subsequent conversion into electricity. CSP plants, when equipped with thermal storage capacity, have the capability to generate electrical power even in situations where sunlight is blocked by cloud cover or during post-sunset hours. CSP can be classified into three categories, including the first generation with a receiver temperature range of 250 °C~450 °C, the second generation with a receiver temperature range of 500 °C~720 °C, and the third generation with a receiver temperature above 700 °C. Moreover, based on the collector’s type, CSP technologies can be divided into four types, including linear Fresnel reflector (LFR), solar power tower (SPT), parabolic power dish collector (PDC) and parabolic trough collector (PTC) [74], as illustrated in Figure 12. Detailed information regarding the CSP category is listed in Table 8.
The levelized cost of electricity (LCOE) is widely recognized as a crucial metric for assessing different power generation systems, as it encompasses all relevant aspects, such as initial capital investments, installation expenses, continuous operational costs and maintenance expenditures across the full lifespan of a power station [77]. The development of S-CO2 technology in CSP applications is crucial for improving the efficiency of solar plants. Replacing the steam Rankine power block with an S-CO2 cycle can enhance the LCOE of conventional molten salt tower technology. Operating at salt temperatures close to 600 °C is projected to yield an 8% enhancement in LCOE, while further advancements in temperature and efficiency can be achieved by employing S-CO2 power cycles in CSP systems [78].
Comprehensive modelling of various SPT systems that are coupled with S-CO2 Brayton cycles was conducted by Wang et al. [79]. Results showed that the intercooling cycle delivered the highest level of efficiency, with the partial-cooling cycle ranking the next, followed by the recompression cycle. Similarly, Neises and Turchi [80] concluded that the implementation of a partial cooling cycle had the potential to generate a larger temperature difference across the primary heat exchanger, thus leading to cost savings on the heat exchanger and enhancing the efficiency of the CSP receiver. Zhu et al. [81] further investigated the effects of different turbine inlet temperatures on the thermal efficiency of this system which was the same as Wang et al.’s [79]. The findings indicated that the turbine inlet temperature had a parabolic impact on the overall efficiencies of each S-CO2 cycle. For further improving the efficiency of S-CO2 SCP, modified cycles have been investigated at different operating conditions. The thermal efficiency of the supercritical CO2 Brayton cycle consistently rises with the temperature of the cycle, as seen in Figure 13, by comparing with different S-CO2 cycles, including recuperation, recompression, partial cooling with recompression and recompression with main compression. The recompression cycle with main compression intercooling achieved the best thermal efficiency of 55.2% at 850 °C [82]. Binotti et al. [83] also arrived at similar findings, indicating further that the recompression with main compression intercooling S-CO2 cycle could attain a solar-to-electric efficiency of 24.5%. The studies mentioned previously relied on steady-state design conditions. However, other researchers have conducted dynamic analyses of S-CO2 CSP performance with varied solar irradiance levels during different seasons, including spring, summer, fall and winter [84,85]. More research conducted by researchers to assess the performance of S-CO2 CSP power systems has been summarized and presented in Table 9. Despite the extensive analysis conducted on the feasibility and efficacy of the S-CO2 CSP solar power system, its implementation on a commercial scale has not yet been achieved due to the challenges associated with operating under high-temperature and high-pressure conditions, which are necessary for optimal performance and analysis. Attaining and sustaining these conditions can pose difficulties to material selection and component production. In addition, the initial capital cost may exhibit a considerably higher magnitude in comparison to alternative renewable energy sources and technologies such as PV panels or conventional fossil fuel power plants.

5.3. S-CO2 Cycle for Geothermal Power Systems

Geothermal power generation is the utilization of underground thermal energy to generate electricity. In general, brine is the most common working fluid to extract heat from underground earth and convert it into electricity via thermodynamic power cycles. However, conventional geothermal energy conversion systems are constrained by size and location, while brines have the potential for scaling and erosion of injection systems and heat exchangers. Therefore, improved monitoring and system management are essential [91]. Furthermore, the majority of geothermal energy is trapped within rocks characterized by low fracture permeability and a lack of fluid circulation. Consequently, the development of new technologies is imperative.
To extract energy from hard dry rock (HDR) to generate electricity, enhanced geothermal systems (EGS) involve the extraction of thermal energy via the creation of artificial geothermal reservoirs [92]. Typically, an EGS requires hydrofracturing of rock with low natural permeability [93]. The performance of EGS based on working fluid of CO2 and water was investigated, and it was found that a CO2-EGS was more efficient than a water-EGS [94,95,96,97]. A CO2-Plume Geothermal (CPG) system was proposed by Randolph et al. [98], as shown in Figure 14, CO2 was sent to the subsurface to recover the energy, and then a small portion of it was piped back to the surface to undergo turbine, generator and heat exchangers for electricity generation, or it was used to provide heat for a power cycle. A CPG was applied as an alternative technology to EGS without hydrofracturing, and the amount of CO2 stored in a CO2-CPG system was significantly greater than what was typically seen in CO2-EGS [93]. Adams et al. [99] conducted a comprehensive comparison between CPG and brine geothermal systems under the conditions of different reservoirs. The results showed that in comparison to brine systems, CO2 direct systems generated a higher net power output when dealing with reservoir depths and permeabilities that fall within the low to moderate range. Wang [92] compared the performance of different S-CO2 cycles, including pre-compression, inter-cooling and reheating for utilizing geothermal energy via model simulations. Subsequently, the S-CO2 Cycle with reheating has the highest net power output of 6.9 MW. Similarly, with low-grade geothermal heat sources, Ruiz-Casanova et al. [100] numerically analyzed the performance of four S-CO2 Brayton cycles, including simple, recuperation, intercooling and intercooled recuperation. The results revealed that the intercooled recuperated Brayton cycle achieved the best performance with the highest power output and highest thermal efficiency. Furthermore, an enhanced natural gas recovery (EGR) reservoir has a larger size than an EGS reservoir, and EGR-CPG can be a promising technology to efficiently extract heat from geothermal. In order to enhance the total producible energy from the gas field while mitigating electricity costs, Ezekiel et al. [101] explored the potential of CPG and the enhanced natural gas recovery (EGR) in a high-temperature reservoir using S-CO2 thermodynamic cycles, as depicted in Figure 15. In this process, external CO2 was introduced into the deep natural gas reservoir. As it travels from the injection well to the production well, the CO2 fluid undergoes heating due to the presence of hot natural gas. The resulting mixture of high-temperature gases was then transported to the land surface, where they were separated for use in a combined system consisting of an S-CO2 cycle and an ORC. The simulation results demonstrated that under conditions of a low CO2 circulation rate in the CPG stage, a net electricity generation of 0.656 MWe could be sustained over 42 years. Conversely, when the CO2 circulation rate is high, the system could generate 1.187 MWe over a period of 32 years. The higher flow rate of CO2 contributed to higher power output with less timeframe and thus fewer capital costs could be achieved. In addition to the investigation of different S-CO2 configurations, researchers also carried out the comparison between pure CO2 and CO2 mixtures. Wright et al. [102] developed an S-CO2 cycle for low-temperature geothermal heat sources and conducted a system performance comparison when either CO2-mixture (CO2/butane) or pure CO2 as a working fluid. The findings indicated that, at a turbine inlet temperature of 160 °C and a dry heat rejection temperature of 46.7 °C, the efficiency of the CO2-mixture cycle was 18.1%, while the efficiency of the pure CO2 cycle was 15%. Another investigation on CO2 mixture containing SF6 was conducted by Yin et al. [103]. It was observed that CO2 concentrations of 15 mol% led to the highest Brayton cycle efficiency, while 20 mol% resulted in the highest Rankine cycle efficiency. Some recent S-CO2 geothermal systems studies are listed in Table 10.
While geothermal energy is characterized by its cleanliness, sustainability and low operating costs, its utilization is constrained by geographic factors. Only a few countries with ample geothermal resources can effectively harness this energy source. The construction of a geothermal power plant demands substantial initial investments since the drilling and exploration stages are the main project-related risks. Additionally, when assessing a technology’s potential, the technology readiness level (TRL) is often utilized [104]. Although S-CO2-CPG demonstrates superior thermal performance compared to conventional hydrothermal geothermal power systems, it is important to note that the TRL of CO2-CPG is relatively low and less mature in comparison to traditional approaches.
Figure 14. Simplified schematic of a CO2-plume geothermal (CPG) system [98].
Figure 14. Simplified schematic of a CO2-plume geothermal (CPG) system [98].
Energies 16 07781 g014
Figure 15. The integrated CO2 EGR-CPG system designed for electricity generation [101].
Figure 15. The integrated CO2 EGR-CPG system designed for electricity generation [101].
Energies 16 07781 g015
Table 10. S-CO2 geothermal systems studies in recent years.
Table 10. S-CO2 geothermal systems studies in recent years.
Ref.YearTechnologyTsource
(°C)
Pmax (MPa)Pmin
(Mpa)
Wnet (kW)th
Ezekiel et al. [101]2019CO2 EGR-CPG15030 1187
Ruiz-Casanova et al. [100]2020Simple Brayton cycle15023.9427.904725.3410.71%
Recuperated Brayton cycle17.9198748.9511.1%
Intercooled Brayton cycle24.747.939719.3110.62%
Intercooled recuperated Brayton cycle18.3328.13779.9911.51%
Wang [92]2018Simply sCO₂ cycle19522.57.8275813.92%
Recuperative sCO₂ cycle25848.92%
sCO2 cycle with pre-compression and inter-cooling319411.02%
sCO2 cycle with reheating597010.3%
sCO2 Cycle with pre-compression, inter-cooling and reheating690411.91%
Wright [102]2017sCO2 cycle with recompression, reheating and intercooling>16022.28.62236915%
sCO2/10% Butane cycle with recompression, reheating and intercooling267518%
Glos et al. [105]2019s-CO2 Rankine cycle>10224.5632405%
Levy et al. [106]2018Direct turbine expansion system22514.58.3430,000-
Tagliaferri et al. [107]2022Direct sCO2 cycleDistrict heating system located between turbine:
T hot water = 35 °C
T cold water = 60 °C
Recovery heat exchanger located after the production well:
T hot water = 50 °C
T cold water = 80 °C
35 °C ≤ Tinjection ≤ 55 °C
1630-
Indirect sCO2 cycle with ORC (binary cycle)2612-
Direct S-CO2 with cogenerationDistrict heating system located between turbine stages1556-
District heating system located after the production well1055-
Combined direct sCO2 with ORCRecovery heat exchanger located before the injection well2918-
Recovery heat exchanger located after the production well2663-
Sun et al. [108]2023T-CO2 Rankine Cycle + power and heat generation unit1966205.7367.542.5%

5.4. S-CO2 Cycle for Waste Heat Recovery

Plenty of waste heat exists in industries, which can be potentially recovered, as shown in Table 11. Waste heat recovery represents a promising approach for minimizing energy consumption and channelling additional energy towards end-use applications. Given the substantial abundance of waste heat in process industries, advanced thermodynamic cycles present greater potential to increase power generation and improve energy efficiency [109].
Waste heat to power (WHP) involves harnessing the heat that is typically wasted by an ongoing process and converting it into electrical energy. The primary conversion route is shown in Figure 16, with applicable technologies and thermodynamic cycles such as steam Rankine, ORC, Stirling, Kalina, and S-CO2 Brayton, as shown in Figure 17. As the S-CO2 Brayton cycle outperforms the existing Rankine cycles in terms of adapting for higher grade heat sources, higher efficiency, system compactness and environmentally friendly working fluid, it has been considered in numerous applications, including waste heat recovery. The initial commercial implementation of a supercritical CO2 power cycle emerged in Alberta. This system captured residual heat from a gas-fired turbine and subsequently transformed this heat into electricity utilizing an innovative S-CO2 power cycle based on the patented technology developed by Echogen and General Electric [111,112], as shown in Figure 18. The power generation system consists of two cycles: a gas turbine cycle and a steam Rankine cycle. Typically, the exhaust gas temperature emanating from a gas turbine exceeds 450 °C, and the conventional steam Rankine cycle makes use of this exhaust gas to enhance thermal efficiency. Ahnv et al. [113] conducted research on the application of a basic S-CO2 cycle for recovering waste heat from a gas turbine shipboard. Their study revealed that it was possible to recover 16.7% of the wasted energy. Zhang et al. [114] carried out a novel recompression S-CO2 system to recover the waste heat from an internal combustion engine in which one more heater and one more turbine were added in the cycle to continuously recover the waste heat. Results revealed that the recovery efficiency of this novel system increased by about 18%. Song et al. [115] pointed out that preheating the S-CO2 cycle can improve net power output in the waste heat recovery system. Wright et al. [116] compared the performance of four different layouts of S-CO2 for waste heat recovery from a gas turbine, including simple recuperation, cascaded, dual recuperated and recuperated with preheating Brayton cycles, as shown in Figure 19. The results pointed out that the simple recuperation cycle had the lowest waste recovery efficiency of 61.2%, while the cascaded cycle had the highest efficiency of 85.64%. Nonetheless, the cascaded cycle demonstrates the lowest thermal efficiency and net cycle efficiency, at 26.5% and 24.7%, respectively. In contrast, the simple recuperated cycle attains the highest values, with thermal efficiency and net cycle efficiency reaching 30.42% and 28.3%, respectively. Manente et al. [117] further investigated three other S-CO2 layouts for recovering waste heat of temperature at 600 °C, finding that the S-CO2 duo-expansion cycle had the highest energy conversion efficiency. More research regarding waste heat recovery technologies has been summarized in Table 12.
Although waste heat recovery systems are attracting more and more attention, there are limitations due to factors such as temperature constraints and the expenses associated with recovery equipment. The majority of waste heat recovery research focuses on medium temperatures ranging from 230 °C to 650 °C because low temperatures offer limited thermal energy, lower electricity generation and thus lower economic viability. However, recovering high temperatures of waste heat can lead to increased thermal stress on materials used in heat exchangers.

5.5. Barriers to Take Up of the S-CO2 Technology

The application of supercritical CO2 in both power cycles or heating and cooling systems [124,125,126,127] have potential advantages in efficiency, size and environmental impact. However, the barriers to taking up this technology can be concluded as follows:
  • Although a small power generation system can be constructed owing to the high density of CO2, it allows for more compact turbine, compressor and heat exchanger components. The design, production and selection of turbomachinery is still a challenge for operating CO2 in a wide range of temperatures and pressures.
  • Enhancing the efficiency of the system via the improvement in heat exchangers remains a compelling aspect for the successful operation of a S-CO2 Brayton cycle, given the presence of at least two heat exchangers in the basic cycle.
  • Insufficient practical experience and performance data from both experimental and commercial applications to provide solid support for applying this technology in the area of renewable energy.

6. Conclusions

The S-CO2 cycle development has attracted considerable attention for its applications in renewable energy sectors. This paper reviews advanced S-CO2 technologies and cycles integrated with renewables such as biomass, solar and geothermal, as well as waste heat for power generation. The S-CO2 system can achieve higher efficiency compared to other conventional power generation systems, as CO2 working fluid can achieve a more effective thermal match with the applicable heat source. Different renewable energies have been categorized and reviewed according to their characteristics, including temperature ranges, energy conversion routes, and suitable technologies.
Biomass has the capacity to reach temperatures as high as 1400 °C via combustion, while solar energy can achieve temperatures up to 1000 °C, allowing its associated power generation system to achieve higher net power output and enhanced energy conversion efficiency in the range of 0.1~68 MWe by adopting different S-CO2 layouts. By reviewing recent years’ studies, thermal efficiencies of S-CO2 biomass systems are in the range of 21~78%, and the S-CO2 solar system is in the range of 15.2~58%. Combined multiple cycles for biomass energy conversion can achieve higher power output. Partial cooling, intercooling and recompression cycles are considered the most favourable cycle configurations for concentrated solar power. The majority of waste heat recovery research focuses on medium temperatures ranging from 230 °C to 650 °C. From the reviewed latest publications, the thermal efficiency and energy conversion efficiency can be achieved in the range of 17.5~39.9% and 18~85%, respectively. The simple recuperation S-CO2 Brayton cycle for waste heat recovery is less attractive due to its lower net power output. In terms of a relatively lower temperature heat source geothermal, supercritical CO2-Plume Geothermal demonstrates superior thermal performance compared to conventional hydrothermal geothermal power systems. The thermal efficiency of different S-CO2 Brayton cycles for geothermal utilization ranges from 8.92% to 18%.
Advanced S-CO2 cycles for renewable energy are the promising approach to improve power output and increase thermal efficiency than the conventional Rankine cycle. There are still some barriers that need to be considered. The limited large-scale application of biomass can be attributed to the need for logistics for feedstock collection and transportation, elevated feedstock costs compared to fossil fuels and greater upfront capital investment requirements. The majority of current CSP solar thermal projects worldwide are primarily based on the steam Rankine cycle. This is due to the Rankine cycle being a mature technology with lower initial capital cost compared to the S-CO2 Brayton cycle. Geothermal power plants are more constrained by location, which must be situated in close proximity to or directly above geothermal resources. The effectiveness of waste heat recovery is heavily dependent on the temperature of the waste heat source. Waste heat recovery faces limitations due to factors such as temperature constraints and the expenses associated with recovery equipment. Moreover, there is a pronounced need for the advancement of turbomachinery and heat exchanger technology. Future research on S-CO2 renewable energy power generation systems is expected to focus on achieving both economic viability and high efficiency.

Author Contributions

Conceptualization, Y.G. and X.Z.; methodology, X.Z.; validation, X.Z.; formal analysis, X.Z.; investigation, X.Z.; resources, X.Z.; data curation, X.Z.; writing—original draft preparation, X.Z.; writing—review and editing, Y.G.; visualization, Y.G.; supervision, Y.G.; project administration, Y.G.; funding acquisition, Y.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Research Councils UK (RCUK), grant number EP/R000298/1.

Data Availability Statement

No new data were created.

Acknowledgments

The authors would like to acknowledge the support received from Ashwell Biomass Ltd. for their technical support.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

CpSpecific heat at constant pressure, J/(kg.K)
hEnthalpy, J/kg
m ˙ Mass flow rate, kg/s
PPressure, Pa
QHeat transfer, W
TTemperature, K
WPower, W
Greek letters
ηEfficiency
Effectiveness
Subscripts
CCooler
HHeater, High
LLow
PCompressor
RecRecuperator
sIsentropic
th Thermal
TTurbine

References

  1. Energy Institute. Statistical Review of World Energy; Energy Institute: London, UK, 2023. [Google Scholar]
  2. Schuster, A.; Karellas, S.; Kakaras, E.; Spliethoff, H. Energetic and Economic Investigation of Organic Rankine Cycle Applications. Appl. Therm. Eng. 2009, 29, 1809–1817. [Google Scholar] [CrossRef]
  3. Tchanche, B.F.; Lambrinos, G.; Frangoudakis, A.; Papadakis, G. Low-Grade Heat Conversion into Power Using Organic Rankine Cycles—A Review of Various Applications. Renew. Sustain. Energy Rev. 2011, 15, 3963–3979. [Google Scholar] [CrossRef]
  4. Coco-Enríquez, L.; Muñoz-Antón, J.; Martínez-Val, J.M. New Text Comparison between CO2 and Other Supercritical Working Fluids (Ethane, Xe, CH4 and N2) in Line- Focusing Solar Power Plants Coupled to Supercritical Brayton Power Cycles. Int. J. Hydrogen Energy 2017, 42, 17611–17631. [Google Scholar] [CrossRef]
  5. 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]
  6. Besarati, S.M.; Yogi Goswami, D. Analysis of Advanced Supercritical Carbon Dioxide Power Cycles with a Bottoming Cycle for Concentrating Solar Power Applications. J. Sol. Energy Eng. 2013, 136, 010904. [Google Scholar] [CrossRef]
  7. Sulzer, G. Verfahren zur erzeugung von arbeit aus warme. Swiss Pat. 1950, 269599, 15. [Google Scholar]
  8. Feher, E.G. The supercritical thermodynamic power cycle. Energy Convers. 1968, 8, 85–90. [Google Scholar] [CrossRef]
  9. Angelino, G. Perspectives for the Liquid Phase Compression Gas Turbine. J. Eng. Power 1967, 89, 229–236. [Google Scholar] [CrossRef]
  10. Angelino, G. Carbon Dioxide Condensation Cycles for Power Production. J. Eng. Power 1968, 90, 287–295. [Google Scholar] [CrossRef]
  11. Angelino, G. Real Gas Effects in Carbon Dioxide Cycles. In Proceedings of the ASME 1969 Gas Turbine Conference and Products Show, Cleveland, OH, USA, 9–13 March 1969. [Google Scholar] [CrossRef]
  12. Brun, K.; Friedman, P.; Dennis, R. Fundamentals and Applications of Supercritical Carbon Dioxide (SCO2) Based Power Cycles; Woodhead Publishing: Cambridge, UK, 2017. [Google Scholar]
  13. Le Moullec, Y. Conceptual Study of a High Efficiency Coal-Fired Power Plant with CO2 Capture Using a Supercritical CO2 Brayton Cycle. Energy 2013, 49, 32–46. [Google Scholar] [CrossRef]
  14. Mecheri, M.; Le Moullec, Y. Supercritical CO2 Brayton Cycles for Coal-Fired Power Plants. Energy 2016, 103, 758–771. [Google Scholar] [CrossRef]
  15. Xu, J.; Wang, X.; Sun, E.; Li, M. Economic Comparison between SCO2 Power Cycle and Water-Steam Rankine Cycle for Coal-Fired Power Generation System. Energy Convers. Manag. 2021, 238, 114150. [Google Scholar] [CrossRef]
  16. Wang, Z.; Xu, J.; Wang, T.; Miao, Z.; Wang, Q.; Liu, G. Performance of SCO2 Coal-Fired Power Plants at Various Power Capacities. J. Clean. Prod. 2023, 416, 137949. [Google Scholar] [CrossRef]
  17. 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]
  18. Dostal, V.; Driscoll, M.J.; Hejzlar, P. A Supercritical Carbon Dioxide Cycle for Next Generation Nuclear Reactors. Doctoral Dissertation, Department of Nuclear Engineering, Massachusetts Institute of Technology, Cambridge, MA, USA, 2004. [Google Scholar]
  19. Moisseytsev, A.; Sienicki, J.J. Investigation of Alternative Layouts for the Supercritical Carbon Dioxide Brayton Cycle for a Sodium-Cooled Fast Reactor. Nucl. Eng. Des. 2009, 239, 1362–1371. [Google Scholar] [CrossRef]
  20. Wright, S.A.; Conboy, T.M.; Radel, R.; Rochau, G.E. Modeling and Experimental Results for Condensing Supercritical CO2 Power Cycles; OSTI OAI (U.S. Department of Energy Office of Scientific and Technical Information): Oak Ridge, TN, USA, 2011. [CrossRef]
  21. Yoon, H.J.; Ahn, Y.; Lee, J.I.; Addad, Y. Potential Advantages of Coupling Supercritical CO2 Brayton Cycle to Water Cooled Small and Medium Size Reactor. Nucl. Eng. Des. 2012, 245, 223–232. [Google Scholar] [CrossRef]
  22. World Nuclear Performance Report 2023. Available online: https://www.world-nuclear.org/getmedia/0156a8d7-01c6-42d9-97be-3f04f34cb8fa/performance-report-2023-final.pdf.aspx (accessed on 1 September 2023).
  23. Kumar, P.; Srinivasan, K. Carbon Dioxide Based Power Generation in Renewable Energy Systems. Appl. Therm. Eng. 2016, 109, 831–840. [Google Scholar] [CrossRef]
  24. 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]
  25. Marchionni, M.; Bianchi, G.; Tassou, S.A. Review of Supercritical Carbon Dioxide (SCO2) Technologies for High-Grade Waste Heat to Power Conversion. SN Appl. Sci. 2020, 2, 611. [Google Scholar] [CrossRef]
  26. White, M.T.; Bianchi, G.; Chai, L.; Tassou, S.A.; Sayma, A.I. Review of Supercritical CO2 Technologies and Systems for Power Generation. Appl. Therm. Eng. 2021, 185, 116447. [Google Scholar] [CrossRef]
  27. Guo, J.Q.; Li, M.J.; He, Y.L.; Jiang, T.; Ma, T.; Xu, J.L.; Cao, F. A Systematic Review of Supercritical Carbon Dioxide(S-CO2) Power Cycle for Energy Industries: Technologies, Key Issues, and Potential Prospects. Energy Convers. Manag. 2022, 258, 115437. [Google Scholar] [CrossRef]
  28. Ge, Y.T.; Tassou, S.A.; Santosa, I.D.; Tsamos, K. Design Optimisation of CO2 Gas Cooler/Condenser in a Refrigeration System. Appl. Energy 2015, 160, 973–981. [Google Scholar] [CrossRef]
  29. Dostal, V.; Hejzlar, P.; Driscoll, M.J. The Supercritical Carbon Dioxide Power Cycle: Comparison to Other Advanced Power Cycles. Nucl. Technol. 2006, 154, 283–301. [Google Scholar] [CrossRef]
  30. Garg, P.; Kumar, P.; Srinivasan, K. Supercritical Carbon Dioxide Brayton Cycle for Concentrated Solar Power. J. Supercrit. Fluids 2013, 76, 54–60. [Google Scholar] [CrossRef]
  31. Chen, H.; Goswami, D.Y.; Stefanakos, E.K. A Review of Thermodynamic Cycles and Working Fluids for the Conversion of Low-Grade Heat. Renew. Sustain. Energy Rev. 2010, 14, 3059–3067. [Google Scholar] [CrossRef]
  32. Chen, Y.; Lundqvist, P.; Johansson, A.; Platell, P. A Comparative Study of the Carbon Dioxide Transcritical Power Cycle Compared with an Organic Rankine Cycle with R123 as Working Fluid in Waste Heat Recovery. Appl. Therm. Eng. 2006, 26, 2142–2147. [Google Scholar] [CrossRef]
  33. Basu, P. Biomass Gasification, Pyrolysis and Torrefaction: Practical Design and Theory; Elsevier Academic Press: London, UK, 2018. [Google Scholar]
  34. Pospíšil, J.; Lisý, M.; Špiláček, M. Optimalization of Afterburner Channel in Biomass Boiler Using CFD Analysis. Acta Polytech. 2016, 56, 379–387. [Google Scholar] [CrossRef]
  35. Biomass Strategy. 2023. Available online: https://assets.publishing.service.gov.uk/government/uploads/system/uploads/attachment_data/file/1178897/biomass-strategy-2023.pdf (accessed on 1 September 2023).
  36. Aboelwafa, O.; Fateen, S.E.K.; Soliman, A.; Ismail, I.M. A Review on Solar Rankine Cycles: Working Fluids, Applications, and Cycle Modifications. Renew. Sustain. Energy Rev. 2018, 82, 868–885. [Google Scholar] [CrossRef]
  37. Global Solar Energy Share in Electricity Mix. 2022. Statista. Available online: https://www.statista.com/statistics/1302055/global-solar-energy-share-electricity-mix/ (accessed on 1 September 2023).
  38. Tzivanidis, C.; Bellos, E.; Antonopoulos, K.A. Energetic and Financial Investigation of a Stand-Alone Solar-Thermal Organic Rankine Cycle Power Plant. Energy Convers. Manag. 2016, 126, 421–433. [Google Scholar] [CrossRef]
  39. Tchanche, B.F.; Pétrissans, M.; Papadakis, G. Heat Resources and Organic Rankine Cycle Machines. Renew. Sustain. Energy Rev. 2014, 39, 1185–1199. [Google Scholar] [CrossRef]
  40. Turchi, C.S.; Ma, Z.; Neises, T.W.; Wagner, M.J. Thermodynamic Study of Advanced Supercritical Carbon Dioxide Power Cycles for Concentrating Solar Power Systems. J. Sol. Energy Eng. 2013, 135, 041007. [Google Scholar] [CrossRef]
  41. Caccia, M.; Tabandeh-Khorshid, M.; Itskos, G.; Strayer, A.R.; Caldwell, A.S.; Pidaparti, S.; Singnisai, S.; Rohskopf, A.D.; Schroeder, A.M.; Jarrahbashi, D.; et al. Ceramic–Metal Composites for Heat Exchangers in Concentrated Solar Power Plants. Nature 2018, 562, 406–409. [Google Scholar] [CrossRef] [PubMed]
  42. Patil, V.R.; Kiener, F.; Grylka, A.; Steinfeld, A. Experimental Testing of a Solar Air Cavity-Receiver with Reticulated Porous Ceramic Absorbers for Thermal Processing at above 1000 °C. Sol. Energy 2021, 214, 72–85. [Google Scholar] [CrossRef]
  43. Mahlia, T.M.I.; Syaheed, H.; Abas, A.E.P.; Kusumo, F.; Shamsuddin, A.H.; Ong, H.C.; Bilad, M.R. Organic Rankine Cycle (ORC) System Applications for Solar Energy: Recent Technological Advances. Energies 2019, 12, 2930. [Google Scholar] [CrossRef]
  44. Loni, R.; Mahian, O.; Najafi, G.; Sahin, A.Z.; Rajaee, F.; Kasaeian, A.; Mehrpooya, M.; Bellos, E.; le Roux, W.G. A Critical Review of Power Generation Using Geothermal-Driven Organic Rankine Cycle. Therm. Sci. Eng. Prog. 2021, 25, 101028. [Google Scholar] [CrossRef]
  45. DiPippo, R. Geothermal Power Plants; Butterworth-Heinemann: Oxford, UK, 2012. [Google Scholar]
  46. Liu, Q.; Duan, Y.; Yang, Z. Performance Analyses of Geothermal Organic Rankine Cycles with Selected Hydrocarbon Working Fluids. Energy 2013, 63, 123–132. [Google Scholar] [CrossRef]
  47. Quoilin, S.; Broek, M.V.D.; Declaye, S.; Dewallef, P.; Lemort, V. Techno-Economic Survey of Organic Rankine Cycle (ORC) Systems. Renew. Sustain. Energy Rev. 2013, 22, 168–186. [Google Scholar] [CrossRef]
  48. Reddy, C.; Naidu, S.V.; Rangaiah, G.P. Waste heat recovery methods and technologies: There is significant potential for recovering some of the wasted heat in the CPI. Key requirements, benefits and drawbacks for numerous techniques are reviewed. Chem. Eng. 2013, 120, 28–39. [Google Scholar]
  49. Wu, P.; Ma, Y.; Gao, C.; Liu, W.; Shan, J.; Huang, Y.-P.; Wang, J.; Zhang, D.; Ran, X. A Review of Research and Development of Supercritical Carbon Dioxide Brayton Cycle Technology in Nuclear Engineering Applications. Nucl. Eng. Des. 2020, 368, 110767. [Google Scholar] [CrossRef]
  50. Yun, S.; Zhang, D.; Li, X.; Zhou, X.; Jiang, D.; Lv, X.; Wu, W.; Feng, Z.; Min, X.; Tian, W.; et al. Design, Optimization and Thermodynamic Analysis of SCO2 Brayton Cycle System for FHR. Prog. Nucl. Energy 2023, 157, 104593. [Google Scholar] [CrossRef]
  51. Khan, Y.; Mishra, R.S. Performance Evaluation of Solar Based Combined Pre-Compression Supercritical CO2 Cycle and Organic Rankine Cycle. Int. J. Green Energy 2020, 18, 172–186. [Google Scholar] [CrossRef]
  52. Li, M.J.; Zhu, H.H.; Guo, J.Q.; Wang, K.; Tao, W.Q. 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]
  53. Wang, K.; Li, M.J.; Guo, J.Q.; Li, P.; Liu, Z.B. A Systematic Comparison of Different S-CO2 Brayton Cycle Layouts Based on Multi-Objective Optimization for Applications in Solar Power Tower Plants. Appl. Energy 2018, 212, 109–121. [Google Scholar] [CrossRef]
  54. Saeed, M.; Kim, M.H. A Newly Proposed Supercritical Carbon Dioxide Brayton Cycle Configuration to Enhance Energy Sources Integration Capability. Energy 2022, 239, 121868. [Google Scholar] [CrossRef]
  55. Biomass Co-Firing: A Renewable Alternative for Utilities. 2000. Available online: https://www.nrel.gov/docs/fy00osti/28009.pdf (accessed on 1 September 2023).
  56. Biomass for Power Generation and CHP. 2007. Available online: https://iea.blob.core.windows.net/assets/1028bee0-2da1-4d68-8b0a-9e5e03e93690/essentials3.pdf (accessed on 1 September 2023).
  57. Liu, H.; Shao, Y.; Li, J. A Biomass-Fired Micro-Scale CHP System with Organic Rankine Cycle (ORC)—Thermodynamic Modelling Studies. Biomass Bioenergy 2011, 35, 3985–3994. [Google Scholar] [CrossRef]
  58. Qiu, G.; Shao, Y.; Li, J.; Liu, H.; Riffat, S.B. Experimental Investigation of a Biomass-Fired ORC-Based Micro-CHP for Domestic Applications. Fuel 2012, 96, 374–382. [Google Scholar] [CrossRef]
  59. Chitsaz, A.; Khalilarya, S.; Mojaver, P. Supercritical CO2 Utilization in a CO2 Zero Emission Novel System for Bio-Synthetic Natural Gas, Power and Freshwater Productions. J. CO2 Util. 2022, 59, 101947. [Google Scholar] [CrossRef]
  60. Cao, Y.; Dhahad, H.A.; Rajhi, A.A.; Alamri, S.; Anqi, A.E.; El-Shafay, A.S. Combined Heat and Power System Based on a Proton Conducting SOFC and a Supercritical CO2 Brayton Cycle Triggered by Biomass Gasification. Int. J. Hydrogen Energy 2022, 47, 5439–5452. [Google Scholar] [CrossRef]
  61. Ji-chao, Y.; Sobhani, B. Integration of Biomass Gasification with a Supercritical CO2 and Kalina Cycles in a Combined Heating and Power System: A Thermodynamic and Exergoeconomic Analysis. Energy 2021, 222, 119980. [Google Scholar] [CrossRef]
  62. Moradi, R.; Cioccolanti, L.; Zotto, L.D.; Renzi, M. Comparative Sensitivity Analysis of Micro-Scale Gas Turbine and Supercritical CO2 Systems with Bottoming Organic Rankine Cycles Fed by the Biomass Gasification for Decentralized Trigeneration. Energy 2023, 266, 126491. [Google Scholar] [CrossRef]
  63. Zhang, X.; Ge, Y.; Ling, C.; Lang, P. Power Generation and Heat Recovery from Biomass with Advanced CO2 Thermodynamic Power Cycles: Modelling Development and Simulation. In Proceedings of the IIR Rankine Conference, Glasgow, UK, 27–31 July 2020. [Google Scholar]
  64. Zhang, X. Experimental and Theoretical Investigation of Biomass-CO2 Transcritical Brayton Cycles and Heat Exchanger Optimizations. Doctoral Dissertation, London South Bank University, London, UK, 2022. [Google Scholar]
  65. Ge, Y.T.; Zhang, X. Performance Optimization of Supercritical CO2 Gas Heater in a Biomass-CO2 Power Generation System. J. Enhanc. Heat Transf. 2023, 30, 1–28. [Google Scholar] [CrossRef]
  66. Manente, G.; Lazzaretto, A. Innovative Biomass to Power Conversion Systems Based on Cascaded Supercritical CO2 Brayton Cycles. Biomass Bioenergy 2014, 69, 155–168. [Google Scholar] [CrossRef]
  67. Wang, X.; Liu, Q.; Bai, Z.; Lei, J.; Jin, H. Thermodynamic Investigations of the Supercritical CO2 System with Solar Energy and Biomass. Appl. Energy 2017, 227, 108–118. [Google Scholar] [CrossRef]
  68. Nkhonjera, L.; Ansari, S.H.; Liu, X. Development of Hybrid CSP Biomass Gasification Process with Supercritical Carbon Dioxide Cycle for Power Generation. In Proceedings of the SOLARPACES 2019: International Conference on Concentrating Solar Power and Chemical Energy Systems, Daegu, Republic of Korea, 1–4 October 2019. [Google Scholar] [CrossRef]
  69. Chein, R.Y.; Chen, W.H. Thermodynamic Analysis of Integrated Adiabatic Chemical Looping Combustion and Supercritical CO2 Cycle. Energy Convers. Manag. X 2021, 10, 100078. [Google Scholar] [CrossRef]
  70. Wang, Y.; Zhu, L.; He, Y.; Yu, J.; Zhang, C.; Jin, S. High-Efficient and Carbon-Free Biomass Power Generation Hybrid System Consisting of Chemical Looping Air Separation and Semi-Closed Supercritical CO2 Cycle with a Bottoming Organic Rankine Cycle. Fuel Process. Technol. 2022, 236, 107393. [Google Scholar] [CrossRef]
  71. Zhang, W.; Chen, F.; Shen, H.; Cai, J.; Yi, L.; Zhang, J.; Wang, X.; Heydarian, D. Design and Analysis of an Innovative Biomass-Powered Cogeneration System Based on Organic Flash and Supercritical Carbon Dioxide Cycles. Alex. Eng. J. 2023, 80, 623–647. [Google Scholar] [CrossRef]
  72. Zhang, Q.; Chen, H.; Li, B.; Pan, P.; Xu, G.; Zhao, Q.; Jiang, X. A Novel System Integrating Water Electrolysis and Supercritical CO2 Cycle for Biomass to Methanol. Appl. Therm. Eng. 2023, 225, 120234. [Google Scholar] [CrossRef]
  73. Concentrating Solar Power Projects. National Renewable Energy Laboratory 2023. Available online: https://solarpaces.nrel.gov/ (accessed on 1 September 2023).
  74. Bijarniya, J.P.; Sudhakar, K.; Baredar, P. Concentrated Solar Power Technology in India: A Review. Renew. Sustain. Energy Rev. 2016, 63, 593–603. [Google Scholar] [CrossRef]
  75. Răboacă, M.S.; Badea, G.; Enache, A.; Filote, C.; Răsoi, G.; Rata, M.; Lavric, A.; Felseghi, R.-A. Concentrating Solar Power Technologies. Energies 2019, 12, 1048. [Google Scholar] [CrossRef]
  76. He, Y.L.; Qiu, Y.; Wang, K.; Yuan, F.; Wang, W.Q.; Li, M.J.; Guo, J.Q. Perspective of Concentrating Solar Power. Energy 2020, 198, 117373. [Google Scholar] [CrossRef]
  77. Persichilli, M.; Kacludis, A.; Zdankiewicz, E.; Held, T. Supercritical CO2 power cycle developments and commercialization: Why sCO2 can displace steam. In Proceedings of the Power-Gen India & Central Asia, New Delhi, India, 19–21 April 2012. [Google Scholar]
  78. Turchi, C. 10 MW Supercritical CO2 Turbine Test; OSTI OAI (U.S. Department of Energy Office of Scientific and Technical Information): Oak Ridge, TN, USA, 2014. [CrossRef]
  79. 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]
  80. Neises, T.; Turchi, C. A Comparison of Supercritical Carbon Dioxide Power Cycle Configurations with an Emphasis on CSP Applications. Energy Procedia 2014, 49, 1187–1196. [Google Scholar] [CrossRef]
  81. Zhu, H.H.; Wang, K.; He, Y.L. Thermodynamic Analysis and Comparison for Different Direct-Heated Supercritical CO2 Brayton Cycles Integrated into a Solar Thermal Power Tower System. Energy 2017, 140, 144–157. [Google Scholar] [CrossRef]
  82. Padilla, R.V.; Too, Y.C.S.; Benito, R.; Stein, W. Exergetic Analysis of Supercritical CO2 Brayton Cycles Integrated with Solar Central Receivers. Appl. Energy 2015, 148, 348–365. [Google Scholar] [CrossRef]
  83. 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]
  84. Osorio, J.D.; Hovsapian, R.; Ordóñez, J.C. Dynamic Analysis of Concentrated Solar Supercritical CO2-Based Power Generation Closed-Loop Cycle. Appl. Therm. Eng. 2016, 93, 920–934. [Google Scholar] [CrossRef]
  85. Dyreby, J.J.; Klein, S.A.; Nellis, G.F.; Reindl, D.T. Modeling Off-Design and Part-Load Performance of Supercritical Carbon Dioxide Power Cycles. In Volume 8: Supercritical CO2 Power Cycles; Wind Energy; Honors and Awards, Proceedings of the ASME Turbo Expo 2013: Turbine Technical Conference and Exposition, San Antonio, TX, USA, 3–7 June 2013; ASME: New York, NY, USA, 2013. [Google Scholar] [CrossRef]
  86. Iverson, B.D.; Conboy, T.M.; Pasch, J.J.; Kruizenga, A.M. Supercritical CO2 Brayton Cycles for Solar-Thermal Energy. Appl. Energy 2013, 111, 957–970. [Google Scholar] [CrossRef]
  87. Khan, M.S.; Abid, M.; Ali, H.M.; Amber, K.P.; Bashir, M.A.; Javed, S. Comparative Performance Assessment of Solar Dish Assisted S-CO2 Brayton Cycle Using Nanofluids. Appl. Therm. Eng. 2019, 148, 295–306. [Google Scholar] [CrossRef]
  88. Sun, Y.; Duniam, S.; Guan, Z.; Gurgenci, H.; Dong, P.; Wang, J.; Hooman, K. Coupling Supercritical Carbon Dioxide Brayton Cycle with Spray-Assisted Dry Cooling Technology for Concentrated Solar Power. Appl. Energy 2019, 251, 113328. [Google Scholar] [CrossRef]
  89. Liu, Y.; Wang, Y.; Zhang, Y.; Hu, S. Design and Performance Analysis of Compressed CO2 Energy Storage of a Solar Power Tower Generation System Based on the S-CO2 Brayton Cycle. Energy Convers. Manag. 2021, 249, 114856. [Google Scholar] [CrossRef]
  90. Chen, J.; Cheng, K.; Li, X.; Huai, X.; Dong, H. Thermodynamic Evaluation and Optimization of Supercritical CO2 Brayton Cycle Considering Recuperator Types and Designs. J. Clean. Prod. 2023, 414, 137615. [Google Scholar] [CrossRef]
  91. Matek, B. The Manageable Risks of Conventional Hydrothermal Geothermal Power Systems; Geothermal Energy Association: Washington, DC, USA, 2014. [Google Scholar]
  92. Wang, X. Investigation of Geothermal Heat Extraction Using Supercritical Carbon Dioxide (sCO2) and Its Utilization in sCO2-based Power Cycles and Organic Rankine Cycles—A Thermodynamic & Economic Perspective. Doctoral Dissertation, Lehigh University, Bethlehem, PA, USA, 2018. [Google Scholar]
  93. Randolph, J.B.; Saar, M.O. Carbon-dioxide plume geothermal (CPG) systems, an alternative engineered geothermal system (EGS) that does not require hydrofracturing: Comparison with traditional EGS regarding geologic reservoir heat energy extraction and potential for inducing seismicity. In AGU Fall Meeting Abstracts; American Geophysical Union: Washington, DC, USA, 2010; p. H23I-06. [Google Scholar]
  94. Zhang, F.Z.; Jiang, P.X. Thermodynamic Analysis of a Binary Power Cycle for Different EGS Geofluid Temperatures. Appl. Therm. Eng. 2012, 48, 476–485. [Google Scholar] [CrossRef]
  95. Mohan, A.R.; Turaga, U.; Shembekar, V.; Elsworth, D.; Pisupati, S.V. Utilization of Carbon Dioxide from Coal-Based Power Plants as a Heat Transfer Fluid for Electricity Generation in Enhanced Geothermal Systems (EGS). Energy 2013, 57, 505–512. [Google Scholar] [CrossRef]
  96. Xu, T.; Feng, G.; Hou, Z.; Tian, H.; Shi, Y.; Lei, H. Wellbore–Reservoir Coupled Simulation to Study Thermal and Fluid Processes in a CO2-Based Geothermal System: Identifying Favorable and Unfavorable Conditions in Comparison with Water. Environ. Earth Sci. 2015, 73, 6797–6813. [Google Scholar] [CrossRef]
  97. Zhong, C.; Xu, T.; Gherardi, F.; Yuan, Y. Comparison of CO2 and Water as Working Fluids for an Enhanced Geothermal System in the Gonghe Basin, Northwest China. Gondwana Res. 2022, 122, 199–214. [Google Scholar] [CrossRef]
  98. Randolph, J.B.; Saar, M.O. Coupling Carbon Dioxide Sequestration with Geothermal Energy Capture in Naturally Permeable, Porous Geologic Formations: Implications for CO2 Sequestration. Energy Procedia 2011, 4, 2206–2213. [Google Scholar] [CrossRef]
  99. Adams, B.M.; Kuehn, T.H.; Bielicki, J.M.; Randolph, J.B.; Saar, M.O. A Comparison of Electric Power Output of CO2 Plume Geothermal (CPG) and Brine Geothermal Systems for Varying Reservoir Conditions. Appl. Energy 2015, 140, 365–377. [Google Scholar] [CrossRef]
  100. Ruiz-Casanova, E.; Rubio-Maya, C.; Pacheco-Ibarra, J.J.; Ambriz-Díaz, V.M.; Romero, C.E.; Wang, X. Thermodynamic Analysis and Optimization of Supercritical Carbon Dioxide Brayton Cycles for Use with Low-Grade Geothermal Heat Sources. Energy Convers. Manag. 2020, 216, 112978. [Google Scholar] [CrossRef]
  101. Ezekiel, J.; Ebigbo, A.; Adams, B.M.; Saar, M.O. On the use of supercritical carbon dioxide to exploit the geothermal potential of deep natural gas reservoirs for power generation. In Proceedings of the European Geothermal Congress, Hague, The Netherlands, 11–14 June 2019; p. 292. [Google Scholar]
  102. Wright, S.A.; Conboy, T.M.; Ames, D.E. CO2-Based Mixtures as Working Fluids for Geothermal Turbines; OSTI OAI (U.S. Department of Energy Office of Scientific and Technical Information): Oak Ridge, TN, USA, 2012. [CrossRef]
  103. Yin, H.; Sabau, A.S.; Conklin, J.; 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]
  104. De Rose, A.; Buna, M.; Strazza, C.; Olivieri, N.; Stevens, T.; Peeters, L.; Tawil-Jamault, D. Technology Readiness Level: Guidance Principles for Renewable Energy Technologies; European Commission: Ispra, Italy, 2017; pp. 17–27. [Google Scholar] [CrossRef]
  105. Glos, S.; Grotkamp, S.; Wechsung, M. Assessment of performance and costs of CO2 based next level geothermal power (NLGP) systems. In Proceedings of the 3rd European Supercritical CO2 Conference, Paris, France, 19–20 September 2019; pp. 19–20. [Google Scholar] [CrossRef]
  106. Levy, E.K.; Wang, X.; Pan, C.; Romero, C.E.; Maya, C.R. Use of Hot Supercritical CO2 Produced from a Geothermal Reservoir to Generate Electric Power in a Gas Turbine Power Generation System. J. CO2 Util. 2018, 23, 20–28. [Google Scholar] [CrossRef]
  107. Tagliaferri, M.; Gładysz, P.; Ungar, P.; Strojny, M.; Talluri, L.; Fiaschi, D.; Manfrida, G.; Andresen, T.; Sowiżdżał, A. Techno-Economic Assessment of the Supercritical Carbon Dioxide Enhanced Geothermal Systems. Sustainability 2022, 14, 16580. [Google Scholar] [CrossRef]
  108. Sun, K.; Zhang, W.; Liang, Y.F.; Habila, M.A.; Chen, X.; Zheng, J.; Xie, S.; Project, S. Multi-Variable Investigation of an Innovative Multigeneration Process Based on Geothermal Energy and Allam Power Unit for Yielding Electric Power, Cooling, Heating, and Liquid CO2 with Zero CO2 Footprint. Sep. Purif. Technol. 2023, 326, 124731. [Google Scholar] [CrossRef]
  109. 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]
  110. Johnson, I.; Choate, W.T.; Davidson, A. Waste Heat Recovery. Technology and Opportunities in U.S. Industry; United States Department of Energy: Washington, DC, USA, 2008. [CrossRef]
  111. Lehar, M.A.; Michelassi, V. General Electric Co. System and Method for Recovery of Waste Heat from Dual Heat Sources. U.S. Patent 9,038,391, 26 May 2015. [Google Scholar]
  112. Held, T.J.; Hostler, S.; Miller, J.D.; Vermeersch, M.; Xie, T. Echogen Power Systems LLC. Heat Engine and Heat to Electricity Systems and Methods with Working Fluid Mass Management Control. U.S. Patent 8,613,195, 24 December 2013. [Google Scholar]
  113. Ahnb, Y.; Seoa, H.; Chaa, J.E.; Chunga, H.J. Design of Supercritical CO2 Waste Heat Recovery System for Shipboard Applications. Available online: https://sco2symposium.com/papers2018/098_Paper.pdf (accessed on 1 September 2023).
  114. Zhang, R.; Su, W.; Lin, X.; Zhou, N.; Zhao, L. Thermodynamic Analysis and Parametric Optimization of a Novel S-CO2 Power Cycle for the Waste Heat Recovery of Internal Combustion Engines. Energy 2020, 209, 118484. [Google Scholar] [CrossRef]
  115. Song, J.; Li, X.; Ren, X.; Gu, C. Performance Improvement of a Preheating Supercritical CO2 (S-CO2) Cycle Based System for Engine Waste Heat Recovery. Energy Convers. Manag. 2018, 161, 225–233. [Google Scholar] [CrossRef]
  116. Wright, S.A.; Davidson, C.S.; Scammell, W.O. Thermo-economic analysis of four sCO2 waste heat recovery power systems. In Proceedings of the Fifth International SCO2 Symposium, San Antonio, TX, USA, 28–31 March 2016. [Google Scholar]
  117. Manente, G.; Costa, M. On the Conceptual Design of Novel Supercritical CO2 Power Cycles for Waste Heat Recovery. Energies 2020, 13, 370. [Google Scholar] [CrossRef]
  118. Liu, L.; Yang, Q.; Cui, G. Supercritical Carbon Dioxide(S-CO2) Power Cycle for Waste Heat Recovery: A Review from Thermodynamic Perspective. Processes 2020, 8, 1461. [Google Scholar] [CrossRef]
  119. Ahmadi, M.H.; Mohammadi, A.; Pourfayaz, F.; Mehrpooya, M.; Bidi, M.; Valero, A.; Uson, S. Thermodynamic Analysis and Optimization of a Waste Heat Recovery System for Proton Exchange Membrane Fuel Cell Using Transcritical Carbon Dioxide Cycle and Cold Energy of Liquefied Natural Gas. J. Nat. Gas Sci. Eng. 2016, 34, 428–438. [Google Scholar] [CrossRef]
  120. Manjunath, K.; Sharma, O.P.; Tyagi, S.K.; Kaushik, S.C. 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]
  121. Bonalumi, D.; Giuffrida, A.; Sicali, F. Thermo-Economic Analysis of a Supercritical CO2-Based Waste Heat Recovery System. In Proceedings of the E3S Web of Conferences, Strasbourg, France, 5–7 May 2021; Volume 312, p. 08022. [Google Scholar] [CrossRef]
  122. Sanchez, D.; De Escalona, J.M.; Chacartegui, R.; Munoz, A.; Sanchez, T. A Comparison between Molten Carbonate Fuel Cells Based Hybrid Systems Using Air and Supercritical Carbon Dioxide Brayton Cycles with State of the Art Technology. J. Power Sources 2011, 196, 4347–4354. [Google Scholar] [CrossRef]
  123. Marchionni, M. Supercritical Carbon Dioxide Power Cycles for Waste Heat Recovery Applications. Doctoral Dissertation, Brunel University London, London, UK, 2021. [Google Scholar]
  124. Wang, J.; Belusko, M.; Semsarilar, H.; Evans, M.; Liu, M.; Bruno, F. An Optimisation Study on a Real-World Transcritical CO2 Heat Pump System with a Flash Gas Bypass. Energy Convers. Manag. 2022, 251, 114995. [Google Scholar] [CrossRef]
  125. Ji, W.; Evans, M.D.; Belusko, M.; Zhao, C.; Liu, M.; Bruno, F. Subcooling Effect on the Optimal Performance for a Transcritical CO2 Heat Pump with Cold Thermal Energy Storage. Heat Mass Transf. 2022, 59, 1257–1275. [Google Scholar] [CrossRef]
  126. Wang, J.; Belusko, M.; Evans, M.; Liu, M.; Zhao, C.; Bruno, F. A Comprehensive Review and Analysis on CO2 Heat Pump Water Heaters. Energy Convers. Manag. X 2022, 15, 100277. [Google Scholar] [CrossRef]
  127. Wang, J.; Belusko, M.; Liu, M.; Semsarilar, H.; Liddle, R.; Alemu, A.; Evans, M.; Zhao, C.; Hudson, J.; Bruno, F. A Comprehensive Study on a Novel Transcritical CO2 Heat Pump for Simultaneous Space Heating and Cooling—Concepts and Initial Performance. Energy Convers. Manag. 2021, 243, 114397. [Google Scholar] [CrossRef]
Figure 1. CO2 thermal–physical properties vary with temperature at different pressures.
Figure 1. CO2 thermal–physical properties vary with temperature at different pressures.
Energies 16 07781 g001
Figure 2. CO2 compressibility factor [17].
Figure 2. CO2 compressibility factor [17].
Energies 16 07781 g002
Figure 3. P-H diagram and T-S diagram of subcritical, transcritical and supercritical CO2 Brayton cycles: SN–subcritical; TN–transcritical; S–supercritical [30].
Figure 3. P-H diagram and T-S diagram of subcritical, transcritical and supercritical CO2 Brayton cycles: SN–subcritical; TN–transcritical; S–supercritical [30].
Energies 16 07781 g003
Figure 4. Configuration and processes of a CO2 supercritical Rankine cycle. (a) The configuration. (b) The process in a T–S diagram [31].
Figure 4. Configuration and processes of a CO2 supercritical Rankine cycle. (a) The configuration. (b) The process in a T–S diagram [31].
Energies 16 07781 g004
Figure 5. Diagram illustrating the heat transfer between the heat source and working fluid in the high-temperature main heat exchanger: (a) ORC cycle. (b) CO2 transcritical/supercritical power cycle [32].
Figure 5. Diagram illustrating the heat transfer between the heat source and working fluid in the high-temperature main heat exchanger: (a) ORC cycle. (b) CO2 transcritical/supercritical power cycle [32].
Energies 16 07781 g005
Figure 6. Biomass energy conversion processes.
Figure 6. Biomass energy conversion processes.
Energies 16 07781 g006
Figure 7. Solar thermal power generation system [43].
Figure 7. Solar thermal power generation system [43].
Energies 16 07781 g007
Figure 8. Layouts of S-CO2 Brayton cycle and T-S diagrams.
Figure 8. Layouts of S-CO2 Brayton cycle and T-S diagrams.
Energies 16 07781 g008
Figure 9. A S-CO2 biomass system for bio-synthetic natural gas, power and freshwater productions [59].
Figure 9. A S-CO2 biomass system for bio-synthetic natural gas, power and freshwater productions [59].
Energies 16 07781 g009
Figure 10. The integration of proton-conducting solid oxide fuel cell and supercritical CO2 Brayton cycle [60].
Figure 10. The integration of proton-conducting solid oxide fuel cell and supercritical CO2 Brayton cycle [60].
Energies 16 07781 g010
Figure 11. A small-scale power generation and heat recovery test rig based on biomass with CO2 transcritical Brayton cycle [64].
Figure 11. A small-scale power generation and heat recovery test rig based on biomass with CO2 transcritical Brayton cycle [64].
Energies 16 07781 g011
Figure 12. Main CSP technologies [75].
Figure 12. Main CSP technologies [75].
Energies 16 07781 g012
Figure 13. Variations in thermal and exergy efficiency with different turbine inlet temperatures for five S-CO2 CSP cycles [81].
Figure 13. Variations in thermal and exergy efficiency with different turbine inlet temperatures for five S-CO2 CSP cycles [81].
Energies 16 07781 g013
Figure 16. Basic process of a WHP system.
Figure 16. Basic process of a WHP system.
Energies 16 07781 g016
Figure 17. Thermodynamic cycles for waste heat recovery at different temperatures and scales [118].
Figure 17. Thermodynamic cycles for waste heat recovery at different temperatures and scales [118].
Energies 16 07781 g017
Figure 18. A schematic diagram showing waste recovery cycle using supercritical carbon dioxide [111,112].
Figure 18. A schematic diagram showing waste recovery cycle using supercritical carbon dioxide [111,112].
Energies 16 07781 g018
Figure 19. Different layouts of S-CO2 waste heat recovery power systems for gas turbine waste heat recovery [116].
Figure 19. Different layouts of S-CO2 waste heat recovery power systems for gas turbine waste heat recovery [116].
Energies 16 07781 g019
Table 1. Recent reviews on S-CO2 power generation technologies.
Table 1. Recent reviews on S-CO2 power generation technologies.
Ref.YearMain EnergyThermodynamic EquationsSummary PointsLimitations
Ahn et al. [17]2015
  • Coal-fired power plant
  • Nuclear
  • Waste heat recovery
n/a
  • Performance of various layouts of S-CO2 Braton cycles were compared.
  • Progress in the development of S-CO2 was introduced.
Application reviews of various heat sources are not sufficient.
Kumar and Srinivasan [23]2016
  • Solar power
n/a
  • The potential and limitation of thermodynamics were reviewed when CO2 was used either alone or as a component in a mixture of the working fluid.
  • Heat transfer issue in recuperator was reviewed.
The study mainly focuses on the application of solar power generation without specifically considering other renewable energies.
Crespi et al. [24]2017
  • Nuclear
  • Solar power
  • Waste heat
n/a
  • 42 standalone layouts of the S-CO2 cycle and 38 combined cycle configurations were reviewed.
  • Operating conditions of different layouts were reviewed.
The limited thermodynamic equations or T-S diagrams aim to assist readers in comprehending distinctions among different S-CO2 configurations.
Marchionni et al. [25]2020
  • Waste heat
n/a
  • Technological challenges of high-grade waste heat recovery were reviewed.
  • Main components of S-CO2 cycle were reviewed.
The illustrations of various cycles are not sufficient.
White et al. [26]2021
  • Fossil fuel
  • Nuclear
  • Solar power
n/a
  • State-of-the-art S-CO2 cycles, along with their technical and operational issues, were reviewed.
  • Development status of turbomachinery, heat exchanger, material selection and control system designs were reviewed.
The advantages of integrating the S-CO2 cycle with renewable energies are not thoroughly outlined.
Guo et al. [27]2022
  • Coal-fired power plant
  • Nuclear
  • Solar power
n/a
  • The challenges of state-of-the-art S-CO2 technologies were reviewed.
  • Research progress of S-CO2 power cycles and components was reviewed.
  • The review explores the thermodynamic, economic, environmental and flexible feasibility of the technology.
The thermodynamic performance of system application needs to be further analyzed.
Table 2. Major advantages and disadvantages of biomass.
Table 2. Major advantages and disadvantages of biomass.
AdvantagesDisadvantages
Renewable and inexhaustible sourceLow energy density
Low content of ash, C, S, N and trace elementsPotential competition with food and feed production
During combustion, ash can capture some hazardous componentsGreat harvesting, collection, transportation and storage cost
Cheap resourceCould lead to global warming if burned directly
Table 3. Potential of different heat source temperature ranges of geothermal energy in terms of heating (MWth) and electricity (MWe) in Europe [47].
Table 3. Potential of different heat source temperature ranges of geothermal energy in terms of heating (MWth) and electricity (MWe) in Europe [47].
Temperature °CMWthMWe
65–90147,73610,462
90–12075,4217503
120–15022,8191268
150–22542,7034745
225–35066,89711,150
Table 4. Thermodynamic equations of each component.
Table 4. Thermodynamic equations of each component.
ComponentThermodynamic Equation
HeaterEnergies 16 07781 i001 Q H = m ( ˙ h 4 h 3 )
TurbineEnergies 16 07781 i002 W T = m ( ˙ h 1 h 2 ) ,
T = ( h 1 h 2 ) / ( h 1 h 2 , s )
RecuperatorEnergies 16 07781 i003 Q r e c , m a x = m i n m ˙ h 1 h 2 , a s s u m m i n g T 2 = T 3 m ˙ h 4 h 3 , a s s u m m i n g T 4 = T 1
r e c = h 1 h 2 Q r e c , m a x = h 4 h 3 Q r e c , m a x
Gas coolerEnergies 16 07781 i004 Q c = m ( ˙ h 1 h 2 )
CompressorEnergies 16 07781 i005 W P = m ( ˙ h 2 h 1 )
P = ( h 2 , s h 1 ) / ( h 2 h 1 )
Table 5. Thermal efficiencies of each S-CO2 layout.
Table 5. Thermal efficiencies of each S-CO2 layout.
LayoutFirst Law Efficiency Equation
Basic S-CO2 t h = W T W P Q H
Recuperation S-CO2 t h = W T W P Q H
Recompression S-CO2 t h = W T W P Q H = W T ( W c o m p r e s s o r + W r e c o m p r e s s o r ) Q H
Pre-compression S-CO2 t h = W T W P Q H = W T ( W p r e c o m p r e s s o r + W c o m p r e s s o r ) Q H
Intercooling S-CO2 t h = W T W P Q H = W T ( W P , 1 + W P , 2 ) Q H
Reheating S-CO2 t h = W T W P Q i n = W T , 1 + W T , 2 W P Q H , 1 + Q H , 2
Table 6. Typical data for power generation from biomass [56].
Table 6. Typical data for power generation from biomass [56].
TechnologiesEfficiency % (LHV)Typical Size (MWe)Typical Costs
Capital Costs ($/kW)Electricity ($/kWh)
Co-firing35–4010–501100–13000.05
Dedicated steam cycles30–355–253000–50000.11
IGCC30–4010–302500–55000.11–0.13
Gasification + engine CHP25–300.2–13000–40000.11
Stirling engine CHP11–20<0.15000–70000.13
Table 7. Reviews in biomass conversion technologies.
Table 7. Reviews in biomass conversion technologies.
RefsYearBiomass ConversionThermodynamic CycleOptimum Power Production (kWe)Energy Conversion Efficiency
Manente et al. [66]2014CombustionCascaded supercritical CO2 Brayton cycles535936%
Wang et al. [67]2018CombustionRecompression S-CO2 Brayton cycle combined with Recuperation S-CO2 Brayton cycle11,25021%
Ge et al. [63,64,65]2020CombustionRecuperation T-CO2 Brayton cycle11.922%
Nkhonjera et al. [68]2020GasificationRecuperation S-CO2 Brayton cycle combined with Steam Rankine cycle-60%
Ji-chao et al. [61]2021GasificationRecuperation S-CO2 Brayton cycle combined with Kalina cycles740078.15%
Chein et al. [69]2021GasificationRecuperation S-CO2 Brayton cycle-21%
Chitsaz et al. [59]2022GasificationRecuperation S-CO2 Brayton cycle172.6-
Cao et al. [60]2022GasificationRecuperation S-CO2 Brayton cycle138-
Wang et al. [70]2022Gasificationsemi-closed S-CO2 cycle with a bottom ORC70,21038.76%
Moradi et al. [62]2023GasificationGas turbine, S-CO2 Brayton cycle, ORC12648%
Zhang et al. [71]2023GasificationGas turbine cycle, S-CO2 cycle, Organic Flash Cycle integrated821075.80%
Zhang et al. [72]2023GasificationRecuperation S-CO2 Brayton cycle68,20067.98%
Table 8. Categories of CSP technologies [76].
Table 8. Categories of CSP technologies [76].
GenerationFirst GenerationSecond GenerationThird Generation
Receiver outlet temperature (°C)250–450500–720>700
Typical technologyParabolic trough collector
Solar power tower
Linear Fresnel reflector
Parabolic trough collector
Solar power tower
Linear Fresnel reflector
Power dish collector
Particles
Gas
Heat transfer mediumOil
Steam
Salt
Steam
Gas
Salt
Air
Helium
CO2
Thermodynamic cycleSteam Rankine cycleSteam Rankine cycle/StirlingBrayton cycle
Cycle efficiency (%)28–3838–44>50
Table 9. Summary of typical applications of S-CO2 in CSP and modelling/experimental analysis.
Table 9. Summary of typical applications of S-CO2 in CSP and modelling/experimental analysis.
Ref.YearApproachCSP typeThermodynamic CyclesPmax (MPa)Pmin (MPa)TH (°C)TL (°C)thWnet (MW)
Dyreby et al. [85]2013Modelling-Recuperation;
Recompression
258.14~9.177004547.6~49.4%10
Iverson et al. [86]2013ExperimentSix immersion heatersSplit-flow recompression14.0917.68853832.415.2%0.176
Neises and Turchi [80]2014ModellingSolar power towerRecuperation;
Recompression;
Partial Cooling
25 6505044.6~49.5%35
Padilla et al. [82]2015ModellingSolar power towerRecuperation;
Recompression;
Partial cooling with recompression; Recompression with main compression intercooling
256.25~16.1500~80055.535.1~55.2%-
Osorio et al. [84]2016ModellingSolar power towerRecompression with multi-stage expansion and intercooling208497.1~515.238.2~44.944.3~48.1%1.516~1.855
Binotti et al. [83]2017Modellingmolten salts solar tower plantsRecompression;
Partial-cooling;
Recompression with main compression intercooling
255.23~9.37740~78051-23.81~24.78
Wang et al. [53,79,81]2017ModellingMolten salt solar power towersRecompression;
Intercooling;
Partial-cooling;
Split-expansion
257.6450~8003538%~58%1
Khan et al. [87]2019ModellingParabolic
dish solar
Recompression with reheat207.6549.931.8533.7%-
Sun et al. [88]2019ExperimentHeaterRecuperation with spray-assisted dry cooling20861042~5739.4~40.9%0.79~0.9
Liu et al. [89]2021ModellingMolten salt solar power towerSplit-recompression with bottom Rankine Cycle31.818.14893.23544.5~49.5%50
Chen et al. [90]2023Modelling-Recompression257.615~7.646500~7003244.5~53.7%-
Table 11. Waste heat sources and end uses [110].
Table 11. Waste heat sources and end uses [110].
Waste Heat SourcesEnergy End Use
  • Combustion exhausts:
Glass melting furnace;
Cement kiln;
Fume incinerator;
Aluminum reverberatory furnace;
Boiler.
  • Process off-gas:
Steel electric arc furnace.
  • Cooling water from
Furnaces;
Air compressors;
Internal combustion engines.
  • Conductive, convective and radiative losses from equipment and heat products
  • Combustion air preheating;
  • Boiler feedwater preheating;
  • Load preheating;
  • Power generation;
  • Steam generation for use in
power generation;
mechanical power;
process steam;
  • Space heating;
  • Water preheating;
  • Transfer to liquid or gaseous process streams.
Table 12. Research on waste heat recovery in recent years.
Table 12. Research on waste heat recovery in recent years.
Ref.YearSourceCyclePower of Source (kW)Tsource (°C)Pmax (Mpa)Pmin
(Mpa)
Wnet
(kW)
threcovery
Ahnb et al. [113]-Gas turbineRecuperation S-CO225,000566--4175-16.70%
Ahmadi et al. [119]2016Proton exchange membrane fuel cellS-CO2 Rankine cycle combined with liquefied natural gas cycle->70100.61413-66.39%
Wright et al. [116]2016Gas turbineRecuperation S-CO240,731549247.7701730.42%61.20%
Cascaded S-CO2 cycle821426.50%85.64%
Dual recuperated sCO2832228.17%78.36%
Recuperated Brayton cycle with preheating860127.80%82.10%
Manjunath et al. [120]2018Gas turbineRecuperation S-CO2 with T-CO2 vapour compression cycle20,60057220>7.37313838.70%44.50%
Song et al. [115]2018Engine waste heatPreheating S-CO2996300157.864--
Preheating S-CO2 with regeneration68
Zhang et al. [114]2020Internal combustion enginesRecompression S-CO2235.851925833.0635.86%58.70%
Manente et al. [117]2020Steel industryS-CO2 dual expansion2010600207.63100026.62%22.30%
Gas turbineS-CO2 dual recuperation231228.40%19.39%
Fuel cellS-CO2 partial heating207325.82%21.63%
Bonalumi et al. [121]2021Gas turbinePartial heating S-CO24710511269.56155025%70%
Sanchez et al. [122]2011Molten carbonate fuel cellSimple recuperation S-CO2->65022.57.5583.639.90%59.40%
Marchionni et al. [123]2021Simulated waste heat source–Air heaterSimple recuperation S-CO2830650207.48423%-
Reheating S-CO28725%-
Recompression S-CO28524%-
Recompression reheating S-CO28827%-
Preheating S-CO215526%-
Preheating Split-Expansion S-CO214023%-
Split-heating Split-Expansion S-CO211017.50%-
Preheating pre-compression S-CO215025-
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

Zhang, X.; Ge, Y. Power Generation with Renewable Energy and Advanced Supercritical CO2 Thermodynamic Power Cycles: A Review. Energies 2023, 16, 7781. https://doi.org/10.3390/en16237781

AMA Style

Zhang X, Ge Y. Power Generation with Renewable Energy and Advanced Supercritical CO2 Thermodynamic Power Cycles: A Review. Energies. 2023; 16(23):7781. https://doi.org/10.3390/en16237781

Chicago/Turabian Style

Zhang, Xinyu, and Yunting Ge. 2023. "Power Generation with Renewable Energy and Advanced Supercritical CO2 Thermodynamic Power Cycles: A Review" Energies 16, no. 23: 7781. https://doi.org/10.3390/en16237781

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