Next Article in Journal
Evaluation of Greenhouse Gas Emission Levels during the Combustion of Selected Types of Agricultural Biomass
Next Article in Special Issue
Preparation and Thermal Model of Tetradecane/Expanded Graphite and A Spiral Wavy Plate Cold Storage Tank
Previous Article in Journal
Reproducibility of Small-Format Laboratory Cells
Previous Article in Special Issue
Study on the Flow Boiling Heat Transfer Characteristics of the Liquid Film in a Rotating Pipe
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Solar Thermochemical CO2 Splitting Integrated with Supercritical CO2 Cycle for Efficient Fuel and Power Generation

1
School of Energy and Power Engineering, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, China
2
Integrated Energy Institute, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, China
*
Author to whom correspondence should be addressed.
Energies 2022, 15(19), 7334; https://doi.org/10.3390/en15197334
Submission received: 6 August 2022 / Revised: 23 September 2022 / Accepted: 30 September 2022 / Published: 6 October 2022
(This article belongs to the Special Issue Heat and Cold Storage for a Net-Zero Future)

Abstract

:
Converting CO2 into fuels via solar-driven thermochemical cycles of metal oxides is promising to address global climate change and energy crisis challenges simultaneously. However, it suffers from low energy conversion efficiency ( η en ) due to high sensible heat losses when swinging between reduction and oxidation cycles, and a single product of fuels can hardly meet multiple kinds of energy demands. Here, we propose an alternative way to upsurge energy conversion efficiency by integrating solar thermochemical CO2 splitting with a supercritical CO2 thermodynamic cycle. When gas phase heat recovery (εgg) is equal to 0.9, the highest energy conversion efficiency of 20.4% is obtained at the optimal cycle high pressure of 260 bar. In stark contrast, the highest energy conversion efficiency is only 9.8% for conventional solar thermochemical CO2 splitting without including a supercritical CO2 cycle. The superior performance is attributed to efficient harvesting of waste heat and synergy of CO2 splitting cycles with supercritical CO2 cycles. This work provides alternative routes for promoting the development and deployment of solar thermochemical CO2 splitting techniques.

1. Introduction

Concentrations of carbon dioxide in the atmosphere have been continuously increasing by 50% since the beginning of the industrial era and set a new record value of 421 ppm in 2022 [1]. Unprecedented CO2 has led to severe climate change and global warming problems, which threatens the sustainable development of human beings. The dominant contribution of excess CO2 emissions is related with massive utilization of fossil fuels. Employing sustainable energy to convert CO2, the major products of carbon-containing fuels combustion, into fuels such as carbon monoxide (CO) is one of the most effective routes to reduce CO2 concentrations [2]. On the other hand, CO is one of the main components of syngas, which can be further converted into liquid fuels through the Fischer-Tropsch process [3,4,5]. Therefore, converting CO2 into fuels is promising to tackle both climate change problems by reducing CO2 in the atmosphere and energy crisis challenges via providing sustainable carbon fuels.
Several methods of using solar energy to produce CO are on the list. Photocatalysis [5,6,7,8], electrolysis [9,10,11], and thermochemical CO2 splitting [12,13,14] have been investigated extensively. Recently, thermochemical CO2 splitting has attracted widespread attention due to its high theoretical energy conversion efficiency ( η en ) due to the capability of utilizing the entire solar spectrum [15,16,17]. For one-step CO2 decomposition, an extremely high temperature over 3000 K is required [18], and mixture gas (possible explosive) needs to be separated, which consumes extra energy. Two-step non-volatile metal oxide cycles, on the other hand, can reduce the required decomposition temperature down to 1773 K and produce O2 and CO in separate steps [19], thus avoiding energy-consuming gas separation issues. Meanwhile, metal oxide cycles enable rather simple design and operation and can achieve high solar-to-fuel efficiency theoretically [20,21,22]. The process of two-step metal oxide cycles is shown as Equations (1)–(3),
M x O y δ ox M x O y δ red + Δ δ 2 O 2
M x O y δ red + Δ δ CO 2 M x O y δ ox + Δ δ CO
Δ δ = δ red δ ox
where M x O y δ red and M x O y δ ox represent the reduced and oxidized metal oxides, respectively. Δ δ = δ red δ ox is the nonstoichiometric swing of the redox material between the reduced and oxidized states.
To achieve a high solar-to-fuel efficiency, developing new materials and optimizing cyclic systems have been extensively investigated in a parallel effort. Ceria has been investigated from a wide range of oxygen pressures and temperatures due to its rapid reaction kinetics at high temperatures [23]. On this basis, the transition metals such as Ca2+, Cr3+ and Zr4+ were doped into cerium dioxide to further increase the CO yield to 156, 251, and 315.4 μmol/g, respectively [24,25,26]. Recently, perovskite oxides have received much attention due to their lower reaction temperatures and high CO yield. Gao et al. reported a high CO yield of 595.6 μmol/g based on Sm0.6Ca0.4Mn0.8Al0.2O3 [27]. Different reactor designs are also investigated to achieve high efficiencies. For example, researchers proposed a rotary-type reactor by using the reactive ceramics of ceria and Ni0.5Mn0.5Fe2O4 to produce CO continuously, and the highest average efficiency of 0.66% was obtained [28,29]. Marxer et al. designed a high-temperature solar reactor containing a reticulated porous ceramic, which was made of pure ceria, and obtained a maximum solar-to-fuel efficiency of 1.73% [30]. Haeussler et al. designed and tested a monolithic solar reactor using cerium dioxide foam with a peak solar fuel efficiency higher than 8% [2]. However, energy conversion efficiency based on traditional metal oxide cycles is still limited due to high sensible heat losses when switching between reduction and oxidation processes. Such heat losses can be harvested via a high-temperature particle-particle heat exchanger but suffer from low efficiencies due to a poor heat transfer coefficent between particles. In addition, the single product of fuels can hardly meet multiple kinds of energy demands for practical applications.
In this paper, we propose to combine thermochemical CO2 splitting with a supercritical CO2 Bryton cycle and to use cerium dioxide as an intermediate medium. The heat dissipated by cerium dioxide from a high reduction temperature down to a low oxidation temperature is used to heat the supercritical CO2, which is then fed to the turbine to export power. Simultaneously, cerium dioxide is oxidized at low temperatures to produce CO, which can be used as solar fuel. This combined cycle enriches products of traditional thermochemical CO2 splitting cycles and makes it possible to reach a higher overall energy efficiency of 20.4% compared with 9.8% otherwise. More detailed thermodynamic analysis and economic analysis, including the energy distribution of sub-systems and effects of different operating parameters on the system efficiency, are discussed.

2. Thermodynamic Model

The system of combining a solar thermochemical CO2 splitting system [31,32] and a supercritical CO2 Brayton cycle [33,34] is shown in Figure 1, in which ceria is used as intermediate medium to transport energy. Sunlight (I = 1 kW/m−2) shines directly on the heliostat which concentrates solar energy into the solar receiver via reflection. Cerium dioxide particles are irradiated directly by sunshine inside the solar receiver and are rapidly heated to reduction temperature to release O2. The solar receiver is maintained at a low pressure using a vacuum pump to ensure that the cerium dioxide reduction reaction continues. O2 cleaved from the fluorite phase is removed by the vacuum pump. Meanwhile, CO2 enters the compressor at room temperature (T0) and is compressed to circulating high pressure (P7), passing through a heat exchanger at the turbine outlet to reach an intermediate temperature (T8), and then flows into the reaction chamber. In the reaction chamber, CO2 and ceria exchange heat and undergo an oxidation reaction, during which CO is produced. Therefore, the CO2 flowing out of the reaction chamber also contains CO, which will be split up by the separator. The pure CO2 leaves the separator and enters the turbine to expand and export power and then flows into the heat exchanger to recover the waste heat.
It is noted that the functions given require temperatures in Kelvin and pressures in bar for a uniform form of expression. We assume that the system operates in steady-state and thermodynamic equilibrium, meaning that all heat and mass fluxes are independent of time. This facilitates us to calculate the optimal energy conversion efficiency of the system, since the continuous production avoids the start-up and cooling process of the material, thus minimizing losses.

2.1. Thermodynamic Efficiency

We introduce thermodynamic efficiency ( η th ) to calculate energy uniformly, which depends on the energy obtained by solar receiver ( Q ˙ rec ), radiation loss ( Q ˙ rad ), convection loss ( Q ˙ conv ), and reflected energy ( Q ˙ ref ). The energy analysis of the fixed heliostat field and the solar receiver is shown below. The design parameters are shown in Table 1 as well.
Q ˙ rec = η h I C
where I is the direct normal solar irradiation intensity, C is the solar concentration ratio, and η h is the optical efficiency. It is noted that optical efficiency varies widely depending on specific designs, taking values from 60% [35] to 90% [36,37]. Since the decrease in optical efficiency suppresses the overall system efficiency, we choose the upper value of the interval to maximize the energy conversion efficiency.
The radiation and convection loss rates of the solar receiver are represented by Equations (5) and (6), respectively [38,39],
Q ˙ rad = A ape ε σ ( T surf 4 T 0 4 )
Q ˙ conv = h nc ( T surf T 0 ) A surf + h fc ( T surf T 0 ) A ape
where A ape is the receiver aperture size, A surf is the surface area of the receiver, ε is the average emissivity of the solar receiver (ε = 0.85), and σ is the Stefan–Boltzmann constant. The convection loss of the solar receiver is the sum of the forced convection loss and the natural convection loss. The forced convection loss comes from the flat plate and is related to the size of the aperture. The natural convection cavity is similar to the flat plate. Here, the wall temperature (Tsurf) of the receiver is 20 K higher than the reduction temperature (T1) [39,40]. We noted that there is a critical point for supercritical CO2 (7.38 Mpa, 304 K), so here, we set T0 to 305K [41].
Since the solar receiver reflectivity changes with surface temperature, the numerical simulation of reflection loss is difficult to implement. Therefore, we introduce the view factor with reflectance to simplify the calculation
Q ˙ ref = Q ˙ rec ρ F r
wherein ρ is the reflectance and Fr is the view factor. The view factor is defined as the ratio of the solar receiver aperture area to the solar receiver surface area [42].
According to the first law of thermodynamics, the remaining absorbed heat can be computed as:
Q ˙ abs = η h I C Q ˙ rad Q ˙ conv Q ˙ ref
The thermodynamic efficiency thus can be counted due to the analysis of heat loss from Equation (9).
η th = η h I C Q ˙ rad Q ˙ conv Q ˙ ref I C

2.2. Ceria’s Reduction and Heating

We set T1 as the reduction temperature (Tred) and T4 as the oxidation temperature (Tox) to describe the loop conveniently. Equations (1) and (2) describe the cyclic reaction of cerium dioxide. Ceria is reduced in a low oxygen atmosphere at a temperature of T1 and a partial pressure of oxygen, P red . δ red is the oxygen vacancy concentration after reduction reaction. Cerium dioxide is then oxidized in CO2 at a pressure of P ox , where the oxygen vacancy concentration is δ ox . The yield of each cycle is the difference in oxygen vacancy concentration Δ δ . For ease of understanding, δ is defined as a unitless measure:
δ = [ O ] [ Ce ]
where [O] is the concentration of oxygen vacancies and [Ce] is the concentration of cerium atoms.
Factors affecting the oxygen vacancy concentration of cerium dioxide have long been the subject of research [23,43]. After continuous exploration, it has been found that the oxygen vacancy concentration is only related to the temperature and the amount of oxygen pressure in the environment. The properties of cerium dioxide have also been investigated over a wide range of oxygen partial pressures (10−2 to 10−8 bar) and temperatures (1273–2173 K) [44]. Bulfin et al. fitted the curve based on the accumulated experimental data and obtained Equation (11) [44].
( δ 0.35 δ ) = 8700 P O 2 0.217 exp ( 195.6   kJ   mol 1 R T )
The units of temperature in Equation (11) are Kelvin and the units of partial pressure of oxygen are bar. We bring T1 into Equation (11) with the adjusted oxygen partial pressure P red in the solar receiver to calculate δ red . For calculating δ ox , the oxygen partial pressure in T4 and 200 bar CO2 is brought into Equation (11), where parameters used in the calculation of δ ox are computed via HSC.
Figure 2 shows the values of nonstoichiometric coefficient δ in a low oxygen partial pressure and a CO2 atmosphere. The δ red and δ ox we calculate are consistent with previous thermodynamic studies [45]. Meanwhile, cerium dioxide requires energy ( Δ H red ) to undergo reduction to remove the oxygen atoms from the lattice. The change in enthalpy has been found to only depend on the nonstoichiometric coefficient [23]. A polynomial curve of the enthalpy change was fitted.
Δ H = ( 478 - 1158 δ + 1790 δ 2 + 23368 δ 3 64929 δ 4 ) 10 3
The heat required to reduce 1 mol of cerium dioxide from δ red to δ ox can be obtained via variable integration using Equation (12). However, oxidizing cerium dioxide to δ red requires an excess of oxidant. Therefore, we introduce a stopping point for the oxidation reaction:
δ ox ( α ) = δ ox + ( 1 - α ) Δ δ ,   0 <   α < 1
where α is the fraction of the reaction completed and is a number between 0 and 1. Based on a previous calculation [46,47], we take the value of α to be 0.95. Combining Equations (12) and (13), the energy required for the reduction of cerium dioxide can be counted.
Q ˙ red = n ˙ CeO 2 δ ox ( α ) δ red Δ H   d δ
Since the re-oxidation process of cerium dioxide involves the decomposition of CO2 and the generation of CO, the energy ( Q ˙ HOX ) required for the oxidation of cerium dioxide can be obtained according to the empirical formula [22]
Q ˙ hox = n ˙ CeO 2 ( Δ H red Δ H CO 2 f + Δ H CO f )
where Δ H red represents the heat required for 1 mol of cerium dioxide to be reduced, and this can be calculated using Equations (12) and (13). The Δ H CO 2 f and Δ H CO f are the mole heat of the formation of CO2 and CO, respectively, which can be found from the NIST Chemistry Webbook.
Ceria will be heated from T4 to T1 after entering the solar receiver. The heat capacity for ceria is taken to be 80 J/mol K [48], because the specific capacity heat of ceria dioxide has little change in the range of 1400–2100 K.
Q ˙ CeO 2 = n ˙ CeO 2 C p CeO 2 ( T 4 T 1 )
The heat ( Q ˙ reco ) recovered from the reheater at the outlet of the turbine is used to heat more supercritical CO2. Although this heat is not counted as solar energy, it is also one of the energy sources worth investigating.
Q ˙ reco = n ˙ CO 2 , t , out ε gg ( h 11 h 12 ) = n ˙ CO 2 ( h 8 h 7 )
The local heat capacity of supercritical CO2 varies greatly, thus, the return heat exchanger is generally of segmented design [49], and discontinuous design allows for faster heat transfer [50]. For the convenience of calculation, we use ε gg to denote the heat transfer efficiency of the return heater.

2.3. Storage Tank and Reaction Chamber

Cerium dioxide pellets are transported to a hot tank for storage and subsequently enter the reaction chamber to exchange heat with CO2 and react. Both components have heat losses to the environment, so we discuss them together. The main sources of energy loss in a particle storage facility are convection from the surrounding environment and conduction from the bin walls [51], which can be expressed by Equation (18).
Q ˙ loss , tan k = Q ˙ foundation + 0 l p h ( T tan k T 0 )   d x + Q ˙ top
Q ˙ loss , tan k = n ˙ CeO 2 C p CeO 2 ( T 2 T 3 )
where Q ˙ foundation and Q ˙ top are the heat loss at the bottom and top of the heat storage tank, respectively. The perimeter p is for a round silo to 2 π A , and A is the silo cross sectional area. The heat loss of the silo is integrated along the silo height l. According to the second law of thermodynamics, the outlet temperature of the heat storage tank can be easily calculated using Equation (19).
In the reaction chamber, cerium dioxide reacts with CO2 in an oxidation process and heat exchange takes place. This can be achieved using either a fluidized bed or a cyclone heat exchanger [52,53]. In order to facilitate the calculations, the pressure drop in the reaction chamber is ignored and the forced heat balance is reached at the outlet.
Q ˙ loss , camber + n ˙ CeO 2 ( h 3 h 4 ) + Q ˙ hox = n ˙ CO 2 ( h 9 h 8 ) + n ˙ CO H H V
The energy that can be released by the fuel is expressed in terms of the high heat value (HHV), because CO is gaseous and there is no latent heat of vaporization. Thus, here, the HHV is equal to the value of the low heat value (i.e., heat of combustion).

2.4. Auxiliary Energy

In exception to chemical reactions that require heat to complete the reaction, the system requires additional energy to complete the cycle, including removal of oxygen from the solar receiver ( Q ˙ pump ), transport of cerium dioxide between cold/hot tanks ( Q ˙ mech ), compression of supercritical CO2 from low to high pressure in the cycle ( W ˙ c ), and separation of the mixture of CO2 and CO ( Q ˙ sep ).
Other methods such as inert gas sweeping and chemical removal can remove the generated oxygen [54]; here, we use a vacuum pump to remove the excess oxygen in the solar receiver for convenience of calculation and give the formula
Q ˙ pump = n ˙ O 2 R T 0 ln ( P 0 P red )   1 η pump
wherein the value of P red is equal to the partial pressure of oxygen at point 1 and the pumping efficiency ( η pump ) is as shown in Table 1.
The particles in the thermochemical CO2 splitting system need to be transported to the solar receiver and sent to the thermal storage tank for cycling. We assume that the transport height H for one cycle is 10 m. The mechanical work is obtained by dividing the gravitational potential energy by the mechanical efficiency ( η mech ). The energy share of Q ˙ mech in the system is very low (less than 1%) and almost negligible.
Q ˙ mech = n ˙ CeO 2 M × g × H η mech
In a supercritical CO2 Brayton cycle, the compressor works on the same principle as a turbine, in which CO2 is compressed or expanded, resulting in a change in enthalpy that can be calculated as power. The flow rate in the compressor and the turbine is not the same due to the fact that the CO2 flows through the separator and into the turbine.
W ˙ c = n ˙ CO 2 ( h 7 h 6 )
W ˙ t = n ˙ CO 2 , t , in ( h 10 h 11 )
The energy required for the separator can be determined using the second law of thermodynamics with the entropy of the separated gas divided by the separation efficiency ( η sep ) to derive Equation (25). Here, n ˙ mix and Δ S mix correspond to the flow rate and the entropy of the unseparated gas mixture at state 9, respectively. We assume that the value of T sep is equal to the oxidation temperature.
Q ˙ sep = Δ S unmix T sep η sep = T sep ( n ˙ CO Δ S 14 + n ˙ CO 2 , sep , out Δ S 10 n ˙ mix Δ S mix ) 1 η sep

2.5. System Efficiency

The total amount of energy essential for the operation of a ceria-based thermochemical CO2 splitting system integrating supercritical CO2 is computed via Equation (26).
Q ˙ tc = Q ˙ CeO 2 + Q ˙ red + Q ˙ pump + Q ˙ mech + Q ˙ sep + W ˙ c
Based on our calculated thermodynamic efficiency, we can work out the solar energy needed for the system.
Q ˙ solar = Q ˙ tc η th
The total energy losses of the system are also computed using Equation (28).
Q ˙ loss = Q ˙ solar ( 1 η th )
The system fuel efficiency ( η fuel ) of the process is defined as the ratio of the high heating value of the production fuel to the solar energy needed to drive the cycle. Analogously, the system’s turbine efficiency ( η W t ) can be defined as the ratio of power exported by the turbine to the total solar energy.
η fuel = n ˙ CO H H V Q ˙ solar
η W t = W ˙ t Q ˙ solar
To count the energy conversion of the system, we write Equation (31) by adding the values of η fuel and η W t .
η en = n ˙ CO H H V + W ˙ t Q ˙ solar

3. Results

3.1. Oxygen Partial Pressure during Reduction

In terms of the degree of δ red created in the ceria crystal structure, Figure 3a reports the amount of O2 released during reduction conducted at different Tred (from 1400–2100 K) and Pred (from 10−3–10−5 bar). The results presented show that, at all Tred, the δ red upsurged with the reduction in the Pred. For instance, the δ red was increased by 0.077, 0.037, 0.009, and 0.001 at 2100 K, 1900 K, 1700 K, and 1500 K, respectively, when the Pred was decreased from 10−3–10−5 bar. Simultaneously, the growth in the Tred was beneficial towards improving the capacity of O2 released during reduction. In terms of values, the δred was increased by 0.12, 0.16, and 0.21 when the Tred was enhanced from 1400–2100 K at stable Pred = 10−3 bar, Pred = 10−4 bar, and Pred = 10−5 bar, separately.
The effect of Tred on n ˙ CO 2 , n ˙ CO 2 , sep , out , and n ˙ CO , sep , out is shown in Figure 3b. The n ˙ CO 2 and the n ˙ CO 2 , sep , out were reduced due to the increment in the Q ˙ loss . For example, as the Tred was upsurged from 1400–2100 K, the n ˙ CO 2 and n ˙ CO 2 , sep , out decreased by 16.7 mol/s and 17.9 mol/s, respectively. On the contrary, the n ˙ CO , sep , out rose by 1.3 mol/s due to the escalation in the Tred from 1400–2100 K. The variations in the W ˙ t , n ˙ CO H H V , and Q ˙ solar due to the decrease of Pred are presented in Figure 3c. As the released O2 was increased due to the decreased Pred from 10−3 to 10−5 bar, a greater quantity of CO was produced during the oxidation reaction, i.e., the n ˙ CO H H V rose from 307.6–508.9 kW at 1900 K. In contrast, the W ˙ t decreased from 138.1–107.0 kW due to the upsurged CO when the Pred decreased from 10−3–10−5 bar. In terms of numbers, the value and growth of n ˙ CO H H V far exceeds that of W ˙ t . Simultaneously, the Q ˙ solar was increased from 3680.4–4048.6 kW due to the increment in the Pred from 10−3 to 10−5 bar.
The efficiency analysis was initiated by exploring the effect of oxygen partial pressure (Pred) on the various process parameters at constant Tox = 1300 K, Pc,in = 75 bar, and Pt,in = 200 bar. By using Equation (31), the η en was computed. As shown in Figure 3d, the numbers obtained indicated that each curve of η en had a peak at high temperatures at all Pred as the Tox, Pc,in, and Pt,in were kept constant. All these peak values were observed to be increased with the increased Pred. In terms of numerical values, the value of the peak escalated from 12.2–15.2%, and the peak temperatures went down from 1950–1850 K due to the decrease in the Pred from 10−3–10−5 bar.

3.2. Reduction Temperature

In this part, the effect of oscillation in the Tred from 1400–2100 K at constant Pred = 10−4 bar and Tox = 1300 K on the energy distribution of the cycle was examined. Firstly, the values associated with η th were computed using Equation (9). As the Tox was steady at 1300 K, the η th was reduced by 33% due to the rise in the Tred from 1400–2100 K.
Figure 4b shows the percentage of energy required for each part of the system. The heat losses due to radiation, convection, and reflection were estimated as Equations (5)–(7). As expected according to the principles of heat transfer, Q ˙ loss was observed to be increased with the rise in Tred. For instance, as the Tred rose from 1400–2100 K, the percentage of Q ˙ loss upsurged from 0.26–0.59. Meanwhile, with the increase of Tred, a higher quantity of O2 was released. Thus, as the δ red upsurged, the energy needed to drive the reduction reaction also increased considerably. In terms of ratios, when the Q ˙ rec was constant and the Tred was increased from 1400–2100 K, the proportion of Q ˙ CeO 2 decreased from 0.67–0.16 because of the increased Q ˙ loss and Q ˙ red .In contrast, the ratio of Q ˙ red increased from 0.03–0.18 when Tred upsurged from 1400–2000 K, due to the rise in the conversion rate, and decreased by 0.02 as Tred was 2100 K because of the increment of Q ˙ loss . The summary of Q ˙ mech , Q ˙ pump , Q ˙ sep , and W ˙ c accounted for 0.04 of Q ˙ solar when the Tred was 1400 K, and the proportion rose to 0.09 as the Tred reached 2100 K. The change in proportion in this process was tiny enough to be ignored.
The effect of ascension in the Tred on the η fuel , η W t , and η en of the system are presented in Figure 4a. Based on the substantial loss in the Q ˙ solar and the increment in the conversion rate, the trend of η fuel was similar to that of Q ˙ red . In terms of numbers, the highest η fuel = 11% in the cycle was obtained at Tred = 2000 K. Due to the ascension of the conversion rate, a higher quantity of CO was released during the oxidation reaction, reducing the amount of CO2 which was delivered into the turbine. The η W t , thus, dropped from 7.8% to 1.9% when Tred upsurged from 1400–2100 K. According to Equation (31), the η en rose to 13.8%, as Tred increased from 1400–1900 K and decreased by 1.4% when Tred rose to 2100 K.

3.3. Oxidation Temperature

After investigating the influence of Tred, this part further explores the consequence of a rise in Tox from 700–1500 K on energy distribution associated with the cycle. The energy trend of each part is reported in Figure 5b at constant Tred = 1900 K and εgg = 0.9 (at stable Pred = 10−4 bar). According to Equation (9), the η th was stable when Tred was a constant.
Namely, the percentage of Q ˙ loss was maintained at 0.46 due to the fixed thermodynamic efficiency. In the case of the solar receiver, the inlet and outlet of ceria was Tox and Tred, individually. Thus, while the Tox was increased, the temperature gap between the inlet and outlet ceria temperature was diminished. This decrease in the temperature gap resulted in an upturn in the Q ˙ red , as a higher quantity of cerium dioxide can be heated to the reduction temperature. Conversely, the ratio of the Q ˙ CeO 2 diminished from 1220–828.3 kW when the Tox increased from 700–1500 K. This consequence was caused by the upsurged Q ˙ pump and Q ˙ sep because more O2 and CO was released with the increase of the ceria. For instance, the proportion of Q ˙ pump and Q ˙ sep rose by 0.01 and 0.07, respectively, due to the upsurge in the Tox from 700 to 1500 K.
The η fuel , η W t , and η en were calculated as functions of Tox, meanwhile, the results are presented in Figure 5a. The finding shown in the figure indicates that the η fuel had a continuous growth due to the increased Q ˙ red . In terms of numbers, the η fuel rose from 6.9–11.6% when the Tox increased from 700–1500 K. The upturned Q ˙ red resulted in the reduction of Q ˙ CeO 2 , diminishing the quantity of CO2 imported into the turbine. The η W t , thus, reduced from 11.1–7.4%, as the Tox upsurged from 760–1500 K. It should be noted that there was an ascension of the η W t with the increment of Tox because the enthalpy of CO2 increases rapidly during this period. The trend of η en about Tred is similar to that about Tox. The highest η en = 19.05% in this case was obtained at Tox = 1300 K, and the η en upsurged from 18.95–18.87% with the increment of the Tox from 1460–1500 K due to the considerable rise in the η fuel .

3.4. Heat Recovery

At Tred = 1900 K and Tox = 1300 K, the effect of variation in the εgg from 0 to 1 on the η en was investigated. In case of the heat exchanger, as the εgg was increased, the Q ˙ reco and the quantity of CO2 upsurged due to the constant Q ˙ rec . Figure 6a indicates that the Q ˙ reco had an ascension from 0 to 2140 kW when the εgg increased from 0 to 1. The consequence of the increment in the εgg for the Q ˙ sep and Q ˙ solar was explored using Equations (25) and (27), respectively. With the increment in the εgg from 0 to 1, the Q ˙ sep was enhanced from 333.6–470.3 kW. This ascension in the Q ˙ sep reflected an increase in the Q ˙ solar from 3870–4210 kW.
Figure 6b represents the variation associated with the η fuel , η W t , and η en as a function of εgg. Because the Q ˙ rec was at a constant and the ceria can be converted completely, the value of n ˙ CO H H V was kept at 406.9 kW. As previously studied, the Q ˙ solar was escalated with the increment in the εgg. The η fuel , thus, diminished from 10.5–9.7% when the εgg increased from 0 to 1. In contrast, the η W t upsurged from 3.2–11.6%, as a larger amount of CO2 was fed into the turbine. Overall, the η en improved by 7.5%, as calculated via Equation (31).

3.5. Cycle High Pressure

The effect of upswing in the cycle high pressure (i.e., Pt,in) on the n ˙ CO 2 and W ˙ t of the cycle is presented in Figure 7b, when Tred = 1900 K, Tox = 1300 K, and εgg = 0.9.
In case of the compressor, the inlet and the outlet pressure of CO2 was 75 bar and Pt,in, respectively. Hence, as the Pt,in was increased, the pressure gap between the inlet and the outlet of the compressor upsurged. Thus, the outlet temperature of the compressor rose accordingly by about a dozen degrees, reducing the temperature increase required for CO2. However, Figure 8 indicates that the average heat capacity of CO2 was increased due to the increment in the Pt,in. Thus, the n ˙ CO 2 had a decrease with the upturned Pt,in. In terms of numbers, the n ˙ CO 2 diminished from 45.5–33.9 mol/s, as the Pt,in improved from 180–260 bar. In addition, the ht,in can be reduced due to the ascension in the Pt,in, making it possible for the turbine to export more power under a limited flow rate. Thence, the highest W ˙ t = 437.9 kJ/s in case of the cycle was obtained at Pt,in = 260 bar, reaching to the maximum value, which was 10.6%, and then diminished in the range of 260–300 bar. Opposite to this consequence, the η fuel remained at about 9.8% due to the stable CO generation and Q ˙ solar . Overall, it seems that the oscillation of the system high pressure mainly impacted the η en when Tred and Tox were constant, as the trend of η en was similar to that of η W t . In terms of values, the maximum η en = 20.4% was attained at Pt,in = 260 bar.

3.6. Cycle Low Pressure

Figure 9a presents the effect of cycle low pressure (i.e., Pc,in) on system performance. It can be seen that the increasing cycle low pressure was profitless in terms of both η W t and η en in the range of 75–87 bar, but the η fuel was maintained at 9.9%, as the increase in cycle low pressure caused the compressor outlet temperature to decrease. However, unlike the cycle high pressure, the increase in cycle low pressure from 72 bar to 90 bar will decrease the compressor outlet temperature by 50 degrees. Near the critical point of supercritical CO2, a temperature difference of a few tens of degrees will cause a wide range of fluctuations in the heat capacity of supercritical CO2. In addition, the average heat capacity of supercritical CO2 was upsurged with the increment in cycle low pressure. Therefore, as shown in Figure 9b, our calculated CO2 flow rate and W ˙ t are not regular with the change of cycle low pressure.

3.7. Economic Analysis

The economic evaluation can derive the production cost of solar thermal chemical CO production and electricity generation [55,56]. This cost takes into account the time value cost of money and calculates the total cost of the system over its life cycle [57,58]. Using the annual rate method to derive annualized values, the total life cycle cost (TLCC) can be used to derive the cost per kilogram of CO. TLCC is computed by Equation (32), wherein V denotes the investments, M the tax rate, PVDEP the present value of the series of occurring depreciation, and PVO&M the present value of operation and maintenance costs. And the present value of reducing recurring U over a time period is found with Equation (33). In case of constant annual payments, the equation can be simplified using the annuity factor N = i   ( 1 -   ( 1 + i ) b ) 1 , wherein i is the interest rate and b is the lifetime of the system.
T L C C = V -   ( M P V DEP ) + ( 1 M )   P V O & M 1 M
P V x = { DEP , O & M } = j = 1 U j U j 1 + i j = U j N
The production costs, or levelized costs per kilogram of CO (LCO), can then be determined by dividing the TLCC by the annual factor times the annual production rate Q (kg CO per year): L C O = T L C C Q N . Finally, Equation (32) can be simplified to T L C C = V + P V O & M , when the project is supported by the government and no taxes have to be paid.
Regardless of weather changes and the intensity of light, the system’s working time is set to 12 h a day, and a year is 365 days. At the highest energy conversion efficiency, the value of the annual output of CO is 634,819.68 kg, while the annual output of electricity is 76,734.97 kWh. The lifetime of the system is 25 years, the average interest rate is 6%, and other parameters are obtained from the previous literature [59,60]. According to Equations (32) and (33), the unit cost of solar fuel is $ 5.86/kg, which is lower than the average market price of CO. The unit cost further decreases to $ 5.62/kg when economic benefits of electricity generation are deducted. The electricity produced by the supercritical CO2 cycle reduces the cost of the conventional thermochemical cycle by 4%, improving the economics of the system.

4. Conclusions

In summary, we proposed an alternative way to upsurge energy conversion efficiency by integrating solar thermochemical CO2 splitting with a supercritical CO2 thermodynamic cycle. For a traditional thermochemical CO2 splitting cycle, the optimal reduction temperature is around 1900 K at a partial pressure of oxygen of 10−4 bar, where the highest energy conversion efficiency is 11%. For an integrated system, the optimal reduction and oxidation temperatures were found to be 1900 and 1300 K, respectively. Simultaneously, the fuel efficiency was constant at 9.8% and the energy efficiency was upturned to 20.4% under the cycle high pressure of 260 bar. The superior performance is attributed to efficient harvesting of waste heat and synergy of CO2 splitting cycles with supercritical CO2 cycles. This work provides alternative routes for improving low efficiency of traditional solar thermochemical CO2 splitting cycles while also enriching products beyond single fuels.

Author Contributions

Conceptualization, X.Y. and W.L.; methodology, X.Y. and K.G.; validation, X.Y., Z.J. and C.T.; formal analysis, X.Y. and N.S.; data curation, X.Y., X.W. and H.Z.; writing—original draft preparation, X.Y.; writing—review and editing, X.Y., W.L. and X.L.; funding acquisition, X.Y., W.L., C.S. and X.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (No. 52076106, 51888103). X.L. also wants to thank the support from Jiangsu Province (No. BE2022024, BK20220077 and BK20202008).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

Aarea (m−2)
Csolar concentration
Cpspecific heat capacity (J mol−1 K−1)
Frview factor
hspecific enthalpy (kJ/mol) or heat
transfer coefficient (W m−2 K−1)
Hheight (m)
HHVhigh heating value (kJ/mol)
Isolar radiation intensity (kW/m2)
n ˙ mole flow rate (mol/s)
Ppressure (bar)
Q ˙ heat rate (kW)
Runiversal gas constant
Ttemperature (K)
W ˙ work rate (kW)
Greek
αfraction completed for oxidation reaction
δnon-stoichiometric coefficient
Δδnon-stoichiometric coefficient difference
ΔHchange in enthalpy or
ΔSchange in entropy
εemissivity or heat recovery effectiveness
ηefficiency
ρreflectivity
Subscripts
0ambient
1,2, …state point
absabsorb
apeaperture
ccompressor
convconvection
enenergy
ggas
hoxheat in exothermic oxidation reaction
ininlet
mechmechanically moving objects
outoutlet
oxoxidation
pumpvacuum pump
radradiation
recreceiver
recorecovery
redreduction
refreflection
sepseparation
surfsurface
tturbine
tctotal
ththermodynamic

References

  1. NOAA Administration. Carbon Dioxide Now More Than 50% Higher Than Pre-Industrial Levels. Available online: www.noaa.gov (accessed on 11 July 2022).
  2. Haeussler, A.; Abanades, S.; Julbe, A.; Jouannaux, J.; Drobek, M.; Ayral, A.; Cartoixa, B. Remarkable performance of microstructured ceria foams for thermochemical splitting of H2O and CO2 in a novel high–temperature solar reactor. Chem. Eng. Res. Des. 2020, 156, 311–323. [Google Scholar] [CrossRef]
  3. Zang, G.; Sun, P.; Elgowainy, A.A.; Bafana, A.; Wang, M. Performance and cost analysis of liquid fuel production from H2 and CO2 based on the Fischer-Tropsch process. J. CO2 Util. 2021, 46, 101459. [Google Scholar] [CrossRef]
  4. Gong, W.; Ye, R.-P.; Ding, J.; Wang, T.; Shi, X.; Russell, C.K.; Tang, J.; Eddings, E.G.; Zhang, Y.; Fan, M. Effect of copper on highly effective Fe-Mn based catalysts during production of light olefins via Fischer-Tropsch process with low CO2 emission. Appl. Catal. B Environ. 2020, 278, 119302. [Google Scholar] [CrossRef]
  5. Sonal; Ahmad, E.; Upadhyayula, S.; Pant, K.K. Biomass-derived CO2 rich syngas conversion to higher hydrocarbon via Fischer-Tropsch process over Fe–Co bimetallic catalyst. Int. J. Hydrog. Energy 2019, 44, 27741–27748. [Google Scholar] [CrossRef]
  6. Li, D.; Kassymova, M.; Cai, X.; Zang, S.-Q.; Jiang, H.-L. Photocatalytic CO2 reduction over metal-organic framework-based materials. Coord. Chem. Rev. 2020, 412, 213262. [Google Scholar] [CrossRef]
  7. Xiong, X.; Mao, C.; Yang, Z.; Zhang, Q.; Waterhouse GI, N.; Gu, L.; Zhang, T. Photocatalytic CO2 Reduction to CO over Ni Single Atoms Supported on Defect-Rich Zirconia. Adv. Energy Mater. 2020, 10, 2002928. [Google Scholar] [CrossRef]
  8. Chen, Z.; Hu, Y.; Wang, J.; Shen, Q.; Zhang, Y.; Ding, C.; Bai, Y.; Jiang, G.; Li, Z.; Gaponik, N. Boosting Photocatalytic CO2 Reduction on CsPbBr3 Perovskite Nanocrystals by Immobilizing Metal Complexes. Chem. Mater. 2020, 32, 1517–1525. [Google Scholar] [CrossRef]
  9. Huan, T.N.; Corte DA, D.; Lamaison, S.; Karapinar, D.; Lutz, L.; Menguy, N.; Foldyna, M.; Turren-Cruz, S.-H.; Hagfeldt, A.; Bella, F.; et al. Low-cost high-efficiency system for solar-driven conversion of CO2 to hydrocarbons. Proc. Natl. Acad. Sci. USA 2019, 116, 9735–9740. [Google Scholar] [CrossRef] [Green Version]
  10. Sriramagiri, G.M.; Ahmed, N.; Luc, W.; Dobson, K.D.; Hegedus, S.S.; Jiao, F. Toward a Practical Solar-Driven CO2 Flow Cell Electrolyzer: Design and Optimization. ACS Sustain. Chem. Eng. 2017, 5, 10959–10966. [Google Scholar] [CrossRef]
  11. Küngas, R. Review—Electrochemical CO2 Reduction for CO Production: Comparison of Low- and High-Temperature Electrolysis Technologies. J. Electrochem. Soc. 2020, 167, 044508. [Google Scholar] [CrossRef]
  12. Chen, Z.; Jiang, Q.; An, H.; Zhang, J.; Hao, S.; Li, X.; Cai, L.; Yu, W.; You, K.; Zhu, X.; et al. Platinum Group Metal Catalyst (RuOx, PtOx, and IrOx)-Decorated Ceria-Zirconia Solid Solution as High Active Oxygen Carriers for Solar Thermochemical CO2 Splitting. ACS Catal. 2022, 12, 7719–7736. [Google Scholar] [CrossRef]
  13. Riaz, A.; Kreider, P.; Kremer, F.; Tabassum, H.; Yeoh, J.S.; Lipiński, W.; Lowe, A. Electrospun Manganese-Based Perovskites as Efficient Oxygen Exchange Redox Materials for Improved Solar Thermochemical CO2 Splitting. ACS Appl. Energy Mater. 2019, 2, 2494–2505. [Google Scholar] [CrossRef]
  14. Liu, X.; Wang, T.; Gao, K.; Meng, X.; Xu, Q.; Song, C.; Zhu, Z.; Zheng, H.; Hao, Y.; Xuan, Y. Ca- and Ga-Doped LaMnO3 for Solar Thermochemical CO2 Splitting with High Fuel Yield and Cycle Stability. ACS Appl. Energy Mater. 2021, 4, 9000–9012. [Google Scholar] [CrossRef]
  15. Lu, Y.; Zhu, L.; Agrafiotis, C.; Vieten, J.; Roeb, M.; Sattler, C. Solar fuels production: Two-step thermochemical cycles with cerium-based oxides. Prog. Energy Combust. Sci. 2019, 75, 100785. [Google Scholar] [CrossRef]
  16. Xiao, L.; Wu, S.-Y.; Li, Y.-R. Advances in solar hydrogen production via two-step water-splitting thermochemical cycles based on metal redox reactions. Renew. Energy 2012, 41, 1–12. [Google Scholar] [CrossRef]
  17. Kodama, T.; Gokon, N. Thermochemical Cycles for High-Temperature Solar Hydrogen Production. Chem. Rev. 2007, 107, 4048–4077. [Google Scholar] [CrossRef]
  18. Yutaka, N.; Bernard, C. Production of Carbon Monoxide by Direct Thermal Splitting of Carbon Dioxide at High Temperature. Bull. Chem. Soc. Jpn. 1986, 59, 1997–2002. [Google Scholar]
  19. Smestad, G.P.; Steinfeld, A. Review: Photochemical and Thermochemical Production of Solar Fuels from H2O and CO2 Using Metal Oxide Catalysts. Ind. Eng. Chem. Res. 2012, 51, 11828–11840. [Google Scholar] [CrossRef]
  20. Kang, Y.; Han, Y.; Wei, C.; Liu, K.; Tian, M.; Huang, C.; Wang, C.; Lin, J.; Hou, B.; Pan, X.; et al. A novel carbon cycle process assisted by Ni/La2O3 catalyst for enhanced thermochemical CO2 splitting. J. Energy Chem. 2021, 61, 297–303. [Google Scholar] [CrossRef]
  21. Kong, H.; Kong, X.; Wang, H.; Wang, J. A strategy for optimizing efficiencies of solar thermochemical fuel production based on nonstoichiometric oxides. Int. J. Hydrog. Energy 2019, 44, 19585–19594. [Google Scholar] [CrossRef]
  22. Muhich, C.L.; Blaser, S.; Hoes, M.C.; Steinfeld, A. Comparing the solar-to-fuel energy conversion efficiency of ceria and perovskite based thermochemical redox cycles for splitting H2O and CO2. Int. J. Hydrog. Energy 2018, 43, 18814–18831. [Google Scholar] [CrossRef]
  23. Panlener, R.J.; Blumenthal, R.N.; Garnier, J.E. A thermodynamic study of nonstoichiometric cerium dioxide. J. Phys. Chem. Solids 1975, 36, 1213–1222. [Google Scholar] [CrossRef]
  24. Le Gal, A.; Abanades, S.; Flamant, G. CO2 and H2O Splitting for Thermochemical Production of Solar Fuels Using Nonstoichiometric Ceria and Ceria/Zirconia Solid Solutions. Energy Fuels 2011, 25, 4836–4845. [Google Scholar] [CrossRef]
  25. Zhu, L.; Lu, Y.; Li, F. Reactivity of Ni, Cr and Zr doped ceria in CO2 splitting for CO production via two-step thermochemical cycle. Int. J. Hydrog. Energy 2018, 43, 13754–13763. [Google Scholar] [CrossRef]
  26. Zhu, L.; Lu, Y. Reactivity and Efficiency of Ceria-Based Oxides for Solar CO2 Splitting via Isothermal and Near-Isothermal Cycles. Energy Fuels 2018, 32, 736–746. [Google Scholar] [CrossRef]
  27. Gao, K.; Liu, X.; Jiang, Z.; Zheng, H.; Song, C.; Wang, X.; Tian, C.; Dang, C.; Sun, N.; Xuan, Y. Direct solar thermochemical CO2 splitting based on Ca- and Al- doped SmMnO3 perovskites: Ultrahigh CO yield within small temperature swing. Renew. Energy 2022, 194, 482–494. [Google Scholar] [CrossRef]
  28. Kaneko, H.; Miura, T.; Fuse, A.; Ishihara, H.; Taku, S.; Fukuzumi, H.; Naganuma, Y.; Tamaura, Y. Rotary-Type Solar Reactor for Solar Hydrogen Production with Two-step Water Splitting Process. Energy Fuels 2007, 21, 2287–2293. [Google Scholar] [CrossRef]
  29. Johnson, T.A.; Hogan, J.; Roy, E.; McDaniel, A.H.; Siegel, N.P.; Dedrick, D.E.; Stechel, E.B.; Diver, J.; Richard, B.; Miller, J.E.; et al. Reimagining Liquid Transportation Fuels: Sunshine to Petrol; Sandia National Laboratories (SNL): Albuquerque, NM, USA; Livermore, CA, USA, 2012; p. 33. [Google Scholar]
  30. Marxer, D.; Furler, P.; Scheffe, J.; Geerlings, H.; Falter, C.; Batteiger, V.; Sizmann, A.; Steinfeld, A. Demonstration of the Entire Production Chain to Renewable Kerosene via Solar Thermochemical Splitting of H2O and CO2. Energy Fuels 2015, 29, 3241–3250. [Google Scholar] [CrossRef]
  31. Farsi, A.; Dincer, I. Thermodynamic assessment of a hybrid particle-based concentrated solar power plant using fluidized bed heat exchanger. Sol. Energy 2019, 179, 236–248. [Google Scholar] [CrossRef]
  32. Alqahtani, B.J.; Patiño-Echeverri, D. Integrated Solar Combined Cycle Power Plants: Paving the way for thermal solar. Appl. Energy 2016, 169, 927–936. [Google Scholar] [CrossRef] [Green Version]
  33. Wang, K.; He, Y.-L. Thermodynamic analysis and optimization of a molten salt solar power tower integrated with a recompression supercritical CO2 Brayton cycle based on integrated modeling. Energy Convers. Manag. 2017, 135, 336–350. [Google Scholar] [CrossRef]
  34. Wang, X.; Liu, Q.; Lei, J.; Han, W.; Jin, H. Investigation of thermodynamic performances for two-stage recompression supercritical CO2 Brayton cycle with high temperature thermal energy storage system. Energy Convers. Manag. 2018, 165, 477–487. [Google Scholar] [CrossRef]
  35. Pitz-Paal, R.; Botero, N.B.; Steinfeld, A. Heliostat field layout optimization for high-temperature solar thermochemical processing. Sol. Energy 2011, 85, 334–343. [Google Scholar] [CrossRef]
  36. Ehrhart, B.; Gill, D. Evaluation of Annual Efficiencies of High Temperature Central Receiver Concentrated Solar Power Plants with Thermal Energy Storage. Energy Procedia 2014, 49, 752–761. [Google Scholar] [CrossRef] [Green Version]
  37. Dähler, F.; Wild, M.; Schäppi, R.; Haueter, P.; Cooper, T.; Good, P.; Larrea, C.; Schmitz, M.; Furler, P.; Steinfeld, A. Optical design and experimental characterization of a solar concentrating dish system for fuel production via thermochemical redox cycles. Sol. Energy 2018, 170, 568–575. [Google Scholar] [CrossRef]
  38. Duffie, J.A.; Beckman, W.A.; Worek, W.M. Solar Engineering of Thermal Processes. J. Sol. Energy Eng. 1994, 116, 67–68. [Google Scholar] [CrossRef]
  39. Li, X.; Kong, W.; Wang, Z.; Chang, C.; Bai, F. Thermal model and thermodynamic performance of molten salt cavity receiver. Renew. Energy 2010, 35, 981–988. [Google Scholar] [CrossRef]
  40. Ho, C.K.; Christian, J.M.; Yellowhair, J.; Armijo, K.; Kolb, W.J.; Jeter, S.; Golob, M.; Nguyen, C. Performance evaluation of a high-temperature falling particle receiver. In Energy Sustainability; ASME: New York, NY, USA, 2016. [Google Scholar]
  41. Liu, Y.; Wang, Y.; Huang, D. Supercritical CO2 Brayton cycle: A state-of-the-art review. Energy 2019, 189, 115900. [Google Scholar] [CrossRef]
  42. Xu, C.; Wang, Z.; Li, X.; Sun, F. Energy and exergy analysis of solar power tower plants. Appl. Therm. Eng. 2011, 31, 3904–3913. [Google Scholar] [CrossRef]
  43. Zinkevich, M.; Djurovic, D.; Aldinger, F. Thermodynamic modelling of the cerium–oxygen system. Solid State Ion. 2006, 177, 989–1001. [Google Scholar] [CrossRef]
  44. Bulfin, B.; Lowe, A.J.; Keogh, K.A.; Murphy, B.E.; Lübben, O.; Krasnikov, S.A.; Shvets, I.V. Analytical Model of CeO2 Oxidation and Reduction. J. Phys. Chem. C 2013, 117, 24129–24137. [Google Scholar] [CrossRef]
  45. Bader, R.; Venstrom, L.J.; Davidson, J.H.; Lipiński, W. Thermodynamic Analysis of Isothermal Redox Cycling of Ceria for Solar Fuel Production. Energy Fuels 2013, 27, 5533–5544. [Google Scholar] [CrossRef]
  46. Bulfin, B.; Call, F.; Lange, M.; Lübben, O.; Sattler, C.; Pitz-Paal, R.; Shvets, I.V. Thermodynamics of CeO2 Thermochemical Fuel Production. Energy Fuels 2015, 29, 1001–1009. [Google Scholar] [CrossRef]
  47. Bulfin, B.; Lange, M.; de Oliveira, L.; Roeb, M.; Sattler, C. Solar thermochemical hydrogen production using ceria zirconia solid solutions: Efficiency analysis. Int. J. Hydrog. Energy 2016, 41, 19320–19328. [Google Scholar] [CrossRef]
  48. Ricken, M.; Nölting, J.; Riess, I. Specific heat and phase diagram of nonstoichiometric ceria (CeO2−x). J. Solid State Chem. 1984, 54, 89–99. [Google Scholar] [CrossRef]
  49. Guo, J. Design analysis of supercritical carbon dioxide recuperator. Appl. Energy 2016, 164, 21–27. [Google Scholar] [CrossRef]
  50. PraveenKumar, S.; Agyekum, E.B.; Velkin, V.I.; Yaqoob, S.J.; Adebayo, T.S. Thermal management of solar photovoltaic module to enhance output performance: An experimental passive cooling approach using discontinuous aluminum heat sink. Int. J. Renew. Energy Res. 2021, 11, 1700–1712. [Google Scholar]
  51. Ma, Z.; Glatzmaier, G.C.; Mehos, M. Development of Solid Particle Thermal Energy Storage for Concentrating Solar Power Plants that Use Fluidized Bed Technology. Energy Procedia 2014, 49, 898–907. [Google Scholar] [CrossRef] [Green Version]
  52. Shimizu, A.; Yokomine, T.; Nagafuchi, T. Development of gas–solid direct contact heat exchanger by use of axial flow cyclone. Int. J. Heat Mass Transf. 2004, 47, 4601–4614. [Google Scholar] [CrossRef]
  53. Ma, Z.; Martinek, J. Analysis of a Fluidized-Bed Particle/Supercritical-CO2 Heat Exchanger in a Concentrating Solar Power System. J. Sol. Energy Eng. 2020, 143, 031010. [Google Scholar] [CrossRef]
  54. Lin, M.; Haussener, S. Solar fuel processing efficiency for ceria redox cycling using alternative oxygen partial pressure reduction methods. Energy 2015, 88, 667–679. [Google Scholar] [CrossRef] [Green Version]
  55. Baloch, A.A.B.; Bahaidarah, H.M.S.; Gandhidasan, P.; Al-Sulaiman, F.A. Experimental and numerical performance analysis of a converging channel heat exchanger for PV cooling. Energy Convers. Manag. 2015, 103, 14–27. [Google Scholar] [CrossRef]
  56. Praveenkumar, S.; Gulakhmadov, A.; Agyekum, E.B.; Alwan, N.T.; Velkin, V.I.; Sharipov, P.; Safaraliev, M.; Chen, X. Experimental Study on Performance Enhancement of a Photovoltaic Module Incorporated with CPU Heat Pipe—A 5E Analysis. Sensors 2022, 22, 6367. [Google Scholar] [CrossRef]
  57. Agyekum, E.B.; PraveenKumar, S.; Alwan, N.T.; Velkin, V.I.; Shcheklein, S.E.; Yaqoob, S.J. Experimental Investigation of the Effect of a Combination of Active and Passive Cooling Mechanism on the Thermal Characteristics and Efficiency of Solar PV Module. Inventions 2021, 6, 63. [Google Scholar] [CrossRef]
  58. Agyekum, E.B.; PraveenKumar, S.; Alwan, N.T.; Velkin, V.I.; Adebayo, T.S. Experimental Study on Performance Enhancement of a Photovoltaic Module Using a Combination of Phase Change Material and Aluminum Fins—Exergy, Energy and Economic (3E) Analysis. Inventions 2021, 6, 69. [Google Scholar] [CrossRef]
  59. Falter, C.; Sizmann, A. Solar Thermochemical Hydrogen Production in the USA. Sustainability 2021, 13, 7804. [Google Scholar] [CrossRef]
  60. Yang, J.; Yang, Z.; Duan, Y. Load matching and techno-economic analysis of CSP plant with S–CO2 Brayton cycle in CSP-PV-wind hybrid system. Energy 2021, 223, 120016. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of ceria-based solar thermochemical CO2 splitting system integrated with a supercritical CO2 Brayton cycle.
Figure 1. Schematic diagram of ceria-based solar thermochemical CO2 splitting system integrated with a supercritical CO2 Brayton cycle.
Energies 15 07334 g001
Figure 2. Nonstoichiometric coefficient, δ red , as a function of temperature, T, for a reduction with P O 2 = 10−4 bar. δ ox as a function of T, for an oxidation with P CO 2 = 200 bar.
Figure 2. Nonstoichiometric coefficient, δ red , as a function of temperature, T, for a reduction with P O 2 = 10−4 bar. δ ox as a function of T, for an oxidation with P CO 2 = 200 bar.
Energies 15 07334 g002
Figure 3. (a) Effect of Pred and Tred on δred of ceria, (b) influence of Tred on n ˙ CO 2 , n ˙ CO 2 , sep , out , and n ˙ CO , sep , out (Pred = 10−4 bar), (c) effect of Pred on W ˙ t , n ˙ CO H H V , and Q ˙ solar (Tred = 1900 K), (d) influence of Tred on η en (Tox = 1300 K, Pred = P, α = 0.95, εgg = 0, Pc,in = 75 bar, and Pt,in = 200 bar).
Figure 3. (a) Effect of Pred and Tred on δred of ceria, (b) influence of Tred on n ˙ CO 2 , n ˙ CO 2 , sep , out , and n ˙ CO , sep , out (Pred = 10−4 bar), (c) effect of Pred on W ˙ t , n ˙ CO H H V , and Q ˙ solar (Tred = 1900 K), (d) influence of Tred on η en (Tox = 1300 K, Pred = P, α = 0.95, εgg = 0, Pc,in = 75 bar, and Pt,in = 200 bar).
Energies 15 07334 g003
Figure 4. (a) Effect of Tred on η fuel , η W t , and η en ( Pred = 10−4 bar), (b) energy balance chart (Tox = 1300 K, Pred = 10−4 bar, α = 0.95, εgg = 0, Pc,in = 75 bar, and Pt,in = 200 bar).
Figure 4. (a) Effect of Tred on η fuel , η W t , and η en ( Pred = 10−4 bar), (b) energy balance chart (Tox = 1300 K, Pred = 10−4 bar, α = 0.95, εgg = 0, Pc,in = 75 bar, and Pt,in = 200 bar).
Energies 15 07334 g004
Figure 5. (a) Effect of Tox on η fuel , η W t , and η en ( Pred = 10−4 bar), (b) energy balance chart (Tred = 1900 K, Pred = 10−4 bar, α = 0.95, εgg = 0.9, Pc,in = 75 bar, and Pt,in = 200 bar).
Figure 5. (a) Effect of Tox on η fuel , η W t , and η en ( Pred = 10−4 bar), (b) energy balance chart (Tred = 1900 K, Pred = 10−4 bar, α = 0.95, εgg = 0.9, Pc,in = 75 bar, and Pt,in = 200 bar).
Energies 15 07334 g005
Figure 6. (a) Effect of εgg on Q ˙ solar , Q ˙ reco , and Q ˙ sep , (b) effect of εgg on η fuel , η W t , and η en (Tred = 1900 K, Tox = 1300 K, Pred = 10−4 bar, α = 0.95, Pc,in = 75 bar, and Pt,in = 200 bar).
Figure 6. (a) Effect of εgg on Q ˙ solar , Q ˙ reco , and Q ˙ sep , (b) effect of εgg on η fuel , η W t , and η en (Tred = 1900 K, Tox = 1300 K, Pred = 10−4 bar, α = 0.95, Pc,in = 75 bar, and Pt,in = 200 bar).
Energies 15 07334 g006
Figure 7. (a) Effect of cycle high pressure on η fuel , η W t , and η en ( Pred = 10−4 bar), (b) influence of cycle high pressure on n ˙ CO 2 and W ˙ t (Tred = 1900 K, Tox = 1300 K, Pred = 10−4 bar, α = 0.95, εgg = 0.9, Pc,in = 75 bar).
Figure 7. (a) Effect of cycle high pressure on η fuel , η W t , and η en ( Pred = 10−4 bar), (b) influence of cycle high pressure on n ˙ CO 2 and W ˙ t (Tred = 1900 K, Tox = 1300 K, Pred = 10−4 bar, α = 0.95, εgg = 0.9, Pc,in = 75 bar).
Energies 15 07334 g007
Figure 8. Effect of temperature on supercritical CO2 heat capacity.
Figure 8. Effect of temperature on supercritical CO2 heat capacity.
Energies 15 07334 g008
Figure 9. (a) Effect of cycle low pressure on η fuel , η W t , and η en ( Pred = 10−4 bar), (b) influence of cycle low pressure on n ˙ CO 2 and W ˙ t (Tred = 1900 K, Tox = 1300 K, Pred = 10−4 bar, α = 0.95, εgg = 0.9, Pt,in = 200 bar).
Figure 9. (a) Effect of cycle low pressure on η fuel , η W t , and η en ( Pred = 10−4 bar), (b) influence of cycle low pressure on n ˙ CO 2 and W ˙ t (Tred = 1900 K, Tox = 1300 K, Pred = 10−4 bar, α = 0.95, εgg = 0.9, Pt,in = 200 bar).
Energies 15 07334 g009
Table 1. Properties and values for the integrated system.
Table 1. Properties and values for the integrated system.
PropertyValue(s)
C3000
Fr0.0757
Tred1400–2100 K
Tox700–1500 K
P01 bar
Pc,in72–90 bar
Pt,in180–300 bar
ηmech0.1
ηO2-rem0.15
ηsep0.15
ρ0.05
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yu, X.; Lian, W.; Gao, K.; Jiang, Z.; Tian, C.; Sun, N.; Zheng, H.; Wang, X.; Song, C.; Liu, X. Solar Thermochemical CO2 Splitting Integrated with Supercritical CO2 Cycle for Efficient Fuel and Power Generation. Energies 2022, 15, 7334. https://doi.org/10.3390/en15197334

AMA Style

Yu X, Lian W, Gao K, Jiang Z, Tian C, Sun N, Zheng H, Wang X, Song C, Liu X. Solar Thermochemical CO2 Splitting Integrated with Supercritical CO2 Cycle for Efficient Fuel and Power Generation. Energies. 2022; 15(19):7334. https://doi.org/10.3390/en15197334

Chicago/Turabian Style

Yu, Xiangjun, Wenlei Lian, Ke Gao, Zhixing Jiang, Cheng Tian, Nan Sun, Hangbin Zheng, Xinrui Wang, Chao Song, and Xianglei Liu. 2022. "Solar Thermochemical CO2 Splitting Integrated with Supercritical CO2 Cycle for Efficient Fuel and Power Generation" Energies 15, no. 19: 7334. https://doi.org/10.3390/en15197334

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