Next Article in Journal
Numerical Study on Flow and Noise Characteristics of High-Temperature and High-Pressure Steam Ejector
Previous Article in Journal
Experimental Characterization of an Additively Manufactured Inconel 718 Heat Exchanger for High-Temperature Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Dual Receiver–Storage Design for Concentrating Solar Thermal Plants Using Beam-Down Optics

1
School of Photovoltaics and Renewable Energy Engineering, University of New South Wales, Sydney, NSW 2052, Australia
2
School of Mechanical and Manufacturing Engineering, University of New South Wales, Sydney, NSW 2052, Australia
*
Author to whom correspondence should be addressed.
Energies 2023, 16(10), 4157; https://doi.org/10.3390/en16104157
Submission received: 14 April 2023 / Revised: 5 May 2023 / Accepted: 9 May 2023 / Published: 17 May 2023
(This article belongs to the Section D: Energy Storage and Application)

Abstract

:
Advanced power cycles—such as the supercritical carbon dioxide (sCO2) cycle—have the potential to reduce the levelized cost of energy (LCOE) of concentrated solar thermal power (CST) plants by significantly boosting their overall solar-to-electric efficiency. To successfully integrate these cycles into CST plants, the industry may need to transition away from liquid working fluids (e.g., synthetic oils and molten salts) to solid and/or gaseous heat transfer media, which are more stable at high temperatures. To address this challenge, this study investigates a novel rotating receiver–storage unit that could enable high-temperature CST plants. A validated numerical model is presented for the charging and discharging processes of the proposed design. It was found that with cast steel as the storage medium in the proposed design, it is possible to achieve >70% receiver efficiency for operation temperatures of 850–1000 K. The overall plant model shows this design is best for relatively small CST systems as modularized units of 10 m diameter (reaching an energy density around 80 kWh/m3), which can be used to drive a 5 MWe sCO2 CST plant. These findings suggest that such a design would have up to 9 h of storage and could be effectively employed as an efficient peaking plant.

1. Introduction

Transitioning to an economy powered by renewables represents the key to mitigating the impacts of climate change and achieving the United Nations’ Sustainable Development Goals [1]. Concentrated solar thermal (CST) electricity generation plants are expected to play a vital role in the future sustainable energy portfolio due to their inherent capacity to incorporate low-cost thermal energy storage. In this low-carbon emissions future, energy storage is critical to cover the natural intermittencies of renewable resources and shift generation to peaking periods [2]. Nowadays, CST plants have a higher levelized cost of electricity (LCOE) than other renewables [3,4]. One way to reduce LCOE in CST plants is by transitioning towards higher temperatures and correspondingly higher thermal efficiency cycles, which would require replacing conventional CST heat transfer mediums (synthetic oils and molten salts), which are constrained to relatively low temperatures (i.e., below 600 °C) [5,6]. Numerous solid mediums are available without such temperature limitations, but the few proposed designs (e.g., volumetric and falling particle receivers) have had limited success in the market [7]. As an alternative—and potentially simpler and more efficient—option, this paper investigates an innovative modular concept of bringing concentrated light down onto a rotating solid medium. Although this concept introduces some engineering challenges, bringing the light directly onto the storage medium essentially bypasses the use of a liquid working fluid (i.e., no solar loop) and can be designed to allow the high-temperature gaseous working fluid for the power cycle (i.e., pressurized sCO2) to extract heat directly from the receiver–storage module to increase the overall efficiency of the CSP plant.
Beam-down receivers (BDRs) work by redirecting the radiation onto the ground, where the final receiver is located. This is accomplished by a secondary reflector that is located at the top of the tower, where the conventional receiver would have been originally located. Although this idea was proposed in the seventies and eighties [8], the seminal works in this technology were developed in the early 2000s, when the optics, possible applications, and proof-of-concept tests were carried out by Segal and Epstein [9,10,11]. They compared two secondary reflectors: an ellipsoid (concave) and a hyperboloid (convex), concluding that the hyperboloid provides the best performance. In theory, BDRs can achieve higher concentration ratios and correspondingly higher thermal efficiencies in the ground-mounted thermal receiver. The main benefits of BDRs are their simpler structure, lower parasitic consumption, and lower impact of wind-associated losses (advective and convective). However, their main drawback is their lower overall optical efficiency due to additional optical devices. Additionally, the same authors found that adding a tertiary (CPC) receiver can increase the overall optical efficiency; however, they also showed how a BDR geometry has an upper bound on the size of the plant since it limits the thermal power up to ~50 MWth [10].
Several experimental facilities have tested the beam-down receiver (BDR) concept. The first experimental facility built with this concept was developed by the Weizmann Institute of Science in Israel [11] in 1999. The plant has a nominal capacity of 0.7 MWth with a 75 m2 hyperboloid mirror and a 2.2 m-aperture diameter CPC. On other hand, the Masdar Institute built and tested a beam-down facility in the UAE since 2011 [12]. This unit, called CSPonD (concentrated solar power on demand demonstration), includes 280.7 m2 of heliostats and is designed for a nominal heat input of 100 kWth with a concentration ratio of up to 600 and an optical efficiency of up to 77% [13]. In an additional stage, a 600 kWh molten salt storage tank was incorporated [13,14], with an overall thermal performance in the range of 24–28% (using an idealized Carnot cycle) [15]. A scale-up analysis predicts a LCOE of ~0.1 USD/kWh for a plant of 50 MWe [16]. Finally, the University of Miyazaki (Japan), developed another BDR facility in 2012, which also has a 100 kWth nominal power output and a 176 m2 primary collection area. The main difference between this system and the previous ones is that it uses an ellipsoid mirror, and its purpose is thermochemical water splitting [17].
On the other hand, a couple of commercial plants have been installed. The Italian company Magaldi Industrie installed the first documented commercial plant using beam-down optics. In 2012, a 100 kWth experimental plant was built in Buccino (Italy) [18]. Then, in 2016, the company started operating a 2 MWth plant in S. Filippo Del Mela (Italy) designed to generate 20.5 tons/day of steam [19]. The system uses a four-facet hyperboloid or flat mirror as a secondary reflector, 786 heliostats of 7 m2 each, heating a 270-ton sand tank of 8.2 MWh of thermal capacity. The Yumen Xinneng, Xinchen, in, China, is the only commercial plant currently in operation. The 50 MWe plant is a multi-tower plant that opened its first unit in 2021 and consists of 5 towers that are 70 m each and heat molten salt [20]. The plant has nine hours of storage capacity to drive a conventional steam cycle with dry cooling. The solar field has 208240 m2 of mirror area installed on a site with a total area of 3.69 km2. Finally, a second beam-down CSP project is under construction in Akesai Jiuquan, China [21]. The project, called Huidong New Energy Akesai, is a hybrid 110 MW CSP + 640 MW PV and is part of the massive portfolio of hybrid projects currently under construction in China. It consists of a conventional steam cycle with 8 h of molten salt thermal storage. It is expected to be operational by the end of 2023 [21].
Despite this continued interest in the beam-down receiver, there has been little investigation into how a beam-down receiver could incorporate solid storage materials. In one example, a numerical simulation of a dual beam-down receiver–storage unit with conductive rods in a sand-basalt mixture has been shown to reach a utilisation ratio of up to 73% [22]. In addition, a gravity-fed receiver concept for sand particles was proposed by Iniesta et al. [23] and tested experimentally, which reached up to 900 °C for a 2 kW BDR. However, the agglomeration of the particles at operation temperatures above 600 °C created a reliability issue for this type of sand. Finally, a beam-down Fresnel-like linear receiver was proposed by Gómez-González et al. [24] in 2018. To date, although the BDR represents a straightforward design that is capable of reaching high concentration ratios, the literature on beam-down receivers has mostly been dedicated to either the optics or their use for conventional heat transfer mediums (such as molten salts). Therefore, the high temperature potential of BDRs (e.g., via solid materials) has not been explored in detail.
The proposed system considers a directly irradiated receiver–storage device that absorbs and stores solar energy simultaneously. Since the receiver is located on the ground and contains a lot of thermal mass, it requires beam-down optics to deliver concentrated radiation to the absorber surface. The thermal transport subsystem is simplified with this concept (relative to a volumetric air receiver or a falling particle receiver), which enables a reduction in parasitic pumping power consumption and in the exergetic losses associated with transferring heat between separate receiver and storage components. The proposed design uses a simple, solid cylindrical unit that is insulated on all sides except for a small window on the top that receives concentrated irradiance, as shown in Figure 1 (left). Candidate materials for the cylinder include concrete, cast steel, magnesia fire bricks, ceramics, etc.
The main difference with this design over a traditional high-temperature sensible thermal storage unit is that charging relies upon radiation (and subsequent conduction) rather than convective heat transfer with a circulating heat transfer fluid. Thus, even more so than a shell-tube sensible storage unit, this design depends on the thermal conductivity of the material. Low thermal conductivity would result in high stratification in the storage unit and, consequently, low storage effectiveness (e.g., only a small proportion of total storage capacity can be effectively used). In addition, low thermal conductivity would result in hotspots at the focal spot of the receiver, which would induce higher radiative losses.
To distribute the heat, we considered the use of conductive rods and a recirculating gas but settled on a novel alternative: rotating the spot relative to the storage tank. That is, as shown in Figure 1, if the solar focal spot is not concentrically located with the central axis of the cylinder, then the heat can be spread in the theta direction by slowly rotating the focal spot relative to the storage unit (or by manually rotating the cylinder relative to a fixed focal spot). This should reduce heat losses, increase efficiency, and promote a more homogeneous charging process.
Clearly, the geometry of this type of storage unit critically affects the effectiveness and temperature distribution during charging/discharging. To obtain a good balance between storage volume (total storage capacity) and charging effectiveness (usage of the stored energy), which are inversely related, multiple modular storage cylinders can be used to obtain the desired capacity. As shown in Figure 1 (right), once one unit is charged, the following unit can be rotated in as the active receiver. This modular concept also serves as a form of control to ensure optimal charging and avoid overheating zones. Moreover, the modular design allows heat extraction from the modules to occur in series and/or parallel with the charging process to better match the solar resource.
As was alluded to previously, it is envisioned that this storage system would drive a high-temperature supercritical CO2 Brayton cycle. Therefore, the heat extraction will require supercritical CO2 fluid to pass through a piping array inside the storage unit. One of the challenges for the discharging process of a solid storage medium is to ensure a stable fluid mass flow rate and temperature output for the cycle when the sensible storage unit’s temperature is decreasing over time. A stable temperature and mass flow rate output from the storage system are some of the biggest advantages of the conventional two-tank configuration. The proposed system will be compared with that conventional configuration. In order to reach that feature, several possible configurations are explored. The discharging process can also be done in series or in parallel to precisely control the inlet conditions in advanced power cycles, with no need for thermal stratification since each module would naturally be at different states of charge (e.g., average temperature). Two operation modes are proposed after testing different alternatives. A general setup for the different configurations between two tanks was developed and can be checked in the Appendix B. With the proposed modular design, two different operation modes, which are shown in Figure 2, are proposed to extract thermal energy:
(a)
Stand-alone operation of one cylinder. If one storage cylinder can provide the desired extracted energy, then it can operate independently. In this case, only a fraction of the total mass flow rate would circulate inside this cylinder, while the remaining fraction would be mixed with the hotter stream. Over time, the fraction increases until all the flow goes through this cylinder. At this point, it is necessary to change the operation mode because the cylinder is unable to provide the desired output.
(b)
Partial-series flow between low-temperature (LT) and high-temperature (HT) cylinders. In this strategy, the whole flow passes first through one unit, the low-temperature (LT) cylinder. As required, a fraction of this flow is diverted to subsequently circulate through a high-temperature (HT) cylinder to obtain a high enough temperature that when the two flows are remixed at the outlet manifold, they provide the required CO2 Brayton cycle inlet temperature (and flow rate). During this process, the temperature of both cylinders decreases, so a gradually higher flow fraction must be diverted into the HT cylinder. The process finishes when all the flow circulates in a series mode, which means no more energy can be extracted from this layout. After that, a new HT cylinder is required, and the previous HT cylinder can become the LT cylinder.
It should be noted that several additional strategies were considered for this design, but these were found to provide the best results and were the most easily controlled. For example, using parallel flow between two cylinders and adjusting the splitting fraction could also be used to ensure a constant output temperature and flow. However, the above mode (b) would extract the most possible energy from the LT cylinder because the total mass flow rate always circulates through this cylinder, giving it the most time/opportunity to discharge. This enables higher energy extraction, more operation time, and a higher stored energy density in each cylinder. Thus, the discharge process strategy in the present work will consist of operation mode (a) when possible, followed by operation mode (b).
It should be noted that several additional strategies were considered for this design, but these were found to provide the best results and were the most easily controlled. A detailed discussion of the general setup and the different possible configurations is given in the Appendix B. For example, using parallel flow between two cylinders and adjusting the splitting fraction could also be used to ensure a constant output temperature and flow. However, the above mode (b) would extract the most possible energy from the LT cylinder because the total mass flow rate always circulates through this cylinder, giving it the most time/opportunity to discharge. This enables higher energy extraction, more operation time, and a higher stored energy density in each cylinder. Thus, the discharge process strategy in the present work will consist of operation mode (a) when possible, followed by operation mode (b).
In addition, to increase the overall temperature difference, a third storage cylinder can be used as a preheater (PH). A preheater can operate in the same way as the low and high temperature units but requires a lower temperature level. The use of a PH unit produces a cascade effect, which allows a more efficient charging process because PH and LT units work at lower temperatures than HT. These operation modes are presented schematically in Figure 2. Consequently, mode (b) requires at least three storage units: a preheater (PH), a low-temperature (LT) cylinder, and a high-temperature (HT) cylinder. This operation mode will be analyzed in detail in the overall plant analysis section.

2. Materials and Methods

To test and analyze this receiver–storage concept, a numerical model was built. The numerical model was based on heat transfer equations for conduction, convection, and radiation. The storage system assumes the two different processes occur independently: (1) rotational charging from a concentrated solar flux and (2) discharging via conduction and convection to a flowing supercritical CO2 fluid.

2.1. Charging Process

For the charging process, the storage unit was assumed to be a solid cylinder. Therefore, inside the volume of the unit, conduction represents the only heat transfer mechanism. In order to ensure that this assumption is acceptable, a design constraint was introduced (see Equation (24)) to ensure the volume occupied by internal piping (needed for the discharging) was limited to less than 10% of the total volume. To simulate this, the energy conservation equation was used in its differential three-dimensional transient form, assuming constant thermophysical properties:
2 T s t x 2 + 2 T s t y 2 + 2 T s t z 2 = 1 α T s t t ,
where  T s t  is the storage material temperature,  α = λ / ρ c p  is the thermal diffusivity of the material,  x  and  y  are the horizontal coordinates,  z  is the vertical coordinate, and  t  represents the temporal variable. At first glance, a Cartesian coordinate system seems more complicated than cylindrical coordinates for a cylindrical block. However, Cartesian coordinates have the same element sizes throughout the volume and are useful for the implementation of the boundary conditions of this problem. Equation (1) can be numerically solved using a finite volume method for spatial discretisation and the fully implicit scheme for temporal discretisation. For each discrete (non-boundary) element, the governing numerical equation can be written as:
1 + 2 r x + 2 r y + 2 r z T i , j , k - r x T i - 1 , j , k + T i + 1 , j , k - r y T i , j - 1 , k + T i , j + 1 , k - r z T i , j , k - 1 + T i , j , k + 1 m + 1 = T i , j , k m ,
where  r n = α Δ t / Δ n , for  n = x , y , z , and  m  is the temporal step. This is valid for  i = 1 N x j = 1 N y , and  k = 1 N z .
The cylinder’s outer boundary was assumed to be well-insulated, so adiabatic boundaries were employed on the bottom and around the outside surface of the cylinder. For the top surface, an adiabatic boundary was also applied, with the exception of the receiver window area, which was matched to the focal spot diameter, which transmits the incident light.
A Cartesian coordinate system was used, so it was necessary to identify the elements of storage belonging to the curved surface of the cylinder. To handle this, a function ( f s t ) was defined to evaluate if the elements belong ( f s t = 1.0 ) or do not ( f s t = 0.0 ) to the storage unit. For most boundary elements,  f s t  takes a fractional value. A brief explanation of this function is presented in Appendix A. A similar function ( f r ) was used to define the elements under the receiver.
For the elements exposed to concentrated radiation, Equation (2) is not valid because the external energy input and heat losses must be considered. For these receiver elements ( k = 0 ; f r > 0.0 ), the energy equation becomes:
1 + 2 r x + 2 r y + r z T i , j , 0 - r x T i - 1 , j , 0 + T i + 1 , j , 0 - r y T i , j - 1,0 + T i , j + 1,0 - r z T i , j , 1 m + 1 = T i , j , 0 m + r z d z k Q i n , i j 0 - Q i j 0 - g ,
where  Q i n , i j 0  and  Q i j 0 - g , are the incident solar flux and the heat losses for element  i j 0 . For elements under the receiver, the inlet flux of energy can be calculated from Equation (4).
Q i n , i j 0 = q c s r τ g α s t A z , i j 0 ,
where  q c s r  is concentrated solar radiation in the beam-down receiver (in  W / m 2 ),  τ g  is the transmittance of the quartz receiver,  α s t  is the absorptance of the storage medium, and  A z , i j 0  is the surface area of the element. In order to ensure simplicity, the solar flux was considered to be uniformly distributed.
The proposed design includes a quartz cover to reduce heat losses, which was considered in the numerical model as only one element. The heat losses were calculated considering convection and radiation heat transfer between the surface elements and the quartz cover. The heat loss flux, a combined convection–radiation coefficient, and the linearized radiation coefficient are presented in Equations (5)–(7).
Q i j 0 - g = h r c , i j 0 - g A z , i j 0 T i j 0 - T g
h r c , i j 0 - g = h r , i j 0 - g + h c , i j 0 - g
h r , i j 0 - g = σ T g + T i j 0 T g 2 + T i j 0 2 1 ϵ s t + A z , i j 0 A g 1 ϵ g - 1
The convective heat transfer between the cylinder’s surface and the quartz was calculated by assuming an air gap at atmospheric pressure via correlations for natural convection between horizontal plates [25]:
N u i j k - g = 0.195 R a 1 / 4 i f 0.5 < P r < 2.0 10 4 < R a < 4 10 5 0.068 R a 1 / 3 i f 0.5 < P r < 2.0 4 10 5 R a < 10 7 0.069 R a 1 / 3 P r 0.074 o t h e r   c a s e
For the heat transfer between the quartz cover and the environment, both convective and radiative heat transfer must be considered. In this case, the radiative heat transfer is between the quartz cover and the sky, while the convective heat transfer is between the quartz cover and the surroundings (e.g., the ambient air temperature). In addition, the natural convection correlation for the horizontal flat plate was used to define hc,g-amb.
Q g - a m b = h r , g - s k y A g T g - T s k y + h c , g - a m b A g T g - T a m b
h r , g - s k y = ϵ g σ A g T g + T s k y T g 2 + T s k y 2
N u g - a m b = 0.54 R a 1 / 4 i f 10 4 < R a < 10 7 0.15 R a 1 / 3 i f 10 7 R a < 10 9 1.52 T s - T a m b 1 / 3 L / λ o t h e r   c a s e

2.2. Discharging Process

The discharging process considers the heat transfer between the supercritical fluid and the wall of the tube array, located inside the storage cylinder. The model assumes that each tube has the same mass flow rate, which extracts thermal energy from the surrounding volume of storage material. In addition, it was assumed that when the storage material reached a thermodynamic equilibrium, there would only be conduction in the vertical direction because the discharge process would extract energy first from the bottom part of the cylinder. In order to take this phenomenon into consideration, the cylinder was discretised vertically, and each element of the storage was assumed to be lumped system (e.g., each element has a uniform temperature).
The equations that represent this phenomenon are:
m ˙ f c p , f , k T f , o , k - T f , i , k = Q ˙ c , k
m s t , k c p , s t , k T s t , k t = - k s t A s t T s t , o , k z + k s t A s t T s t , i , k z - Q ˙ c , k
Q ˙ c , k = h c , k A c T s t , k - T f , k ,
where subscripts  f s t i o  represent flow, storage, inlet, and outlet, respectively.  A c = π D p Δ z  represents the convective area, with  D p  as pipe diameter and  Δ z  as the vertical discretisation. Subscript k in Equation (12) represents the element in the vertical discretisation in these equations as well.  Q ˙ c , k  is the convective heat transfer. Replacing  Q ˙ c , k  with the right-hand-side of Equation (14) in both Equations (12) and (13) and considering a linear temperature increase in the flow side  T f , k = T f , k , o + T f , k , i / 2  results in an implicit system valid for  k = 1 N z , which can be represented as follows:
T f , k m + 1 = 2 m ˙ f c p , f , k T f , k m + h c , k A c T s t , k m + 1 2 m ˙ f c p , f , k + h c , k A c
1 + 2 r 1 + r 2 T s t , k m + 1 - r 1 T s t , k + 1 m + 1 - r 1 T s t , k - 1 m + 1 - r 2 T f m , k m + 1 = T s t , k m ,
where  m  is the timestep,  r 1 = k Δ t / Δ z 2 ρ s t c p , s t , and  r 2 = h c π D p Δ t / A ρ c p s t . The convective heat transfer coefficient is calculated using the well-known Gnielinski correlation for transition and turbulent flow [26] and a developing flow correlation for laminar flow developed by Kakaç et al. [26]:
N u f - s t = 3.66 + 0.0018 x D p P e 1 / 3 0.04 + x D p P e 2 / 3 2 i f R e < 2300 f / 8 R e - 1000 P r 1 + 12.7 f / 8 0.5 P r 2 / 3 - 1 i f R e > 2300 ,
where  P e  and  R e  are the Peclet and Reynolds numbers for circular pipes, and  f  is the Darcy friction factor, calculated with the Haaland approximation of the Colebrook–White equation [26]. All the properties for sCO2 were calculated using the Cantera module for Python [27]. This model also uses the lumped approximation for each element, neglecting the conduction inside the storage section. This approximation is valid for  B i < 0.1 , which is not met by typical solid storage materials. To fix this issue, Xu et al. [28] proposed an extension of the lumped method that defined an equivalent heat transfer coefficient,  h e f f , that takes into account both convection and internal conduction. For storage material with fluid flowing inside a tube, the Xu et al. method for equivalent HTC is:
h e f f = 1 h c + 1 k r a 3 4 b 2 - a 2 + a b 4 4 ln b / a - 3 4 b 2 - a 2 2 - 1 ,
Here,  a = D p / 2  and  b = D s t / 4 N p  for the present system.  D p  and  N p  are the diameter and number of pipes inside the storage cylinder. This correction reduces the effective heat transfer coefficient, and it depends strongly on the storage material’s thermal conductivity. As an example, for  h c = 100   W m - 2 K - 1  and the geometry used in the discussion (Section 3.1), the correction ranges from a 3.2% reduction for cast steel (i.e.,  h e f f = 96.8   W m - 2 K - 1  at the highest thermal conductivity tested) to 48.2% (i.e.,  h e f f = 51.8   W m - 2 K - 1  at the lowest thermal conductivity tested).
During the discharge process, the two cylinders will operate in the “partial-series” operation mode (b) described above. Thus, the discharge output will be the combination of the flows from both the HT and the LT cylinders, which can be determined mathematically with:
m ˙ f c p , f T f , o = m ˙ L T c p , f T L T , o + m ˙ H T c p , f T H T , o
m ˙ f = m ˙ L T + m ˙ H T
Note that for each timestep, the proportion  m ˙ L T / m ˙ H T  will be changed to ensure a desired  T f , o .

2.3. Non-Dimensional and Performance Parameters

The model is designed to work with different storage unit volumes, solar field sizes, and power block capacities. In order to simulate this system over a broad range and achieve meaningful results, several non-dimensional parameters were defined via normalizing based on geometry. To define the receiver size for a given storage volume, a receiver-to-storage surface ratio ( R S S R ) was defined, which provides the ratio between the receiver area and the transversal storage area. In addition, the height-to-diameter ratio for the storage cylinder was used ( H D R ), along with the eccentricity receiver ratio ( E R ), a parameter that defines how far off centre the irradiance spot is located. Further, the ratio of unused storage,  R u n , was defined as the proportion of storage volume that is occupied by the pipe bundle. For the purpose of this work, this ratio was limited to  R u n 0.1 . So, for a given pipe diameter ( D p ), with  R u n  set, it is possible to define the number of pipes inside each unit.
R S S R = A r A s t = π r r 2 π D s t 2 / 4 = 2 r r D s t 2
H D R = H s t D s t
E R = d e c c r r
R u n = N p π 4 D p 2 π 4 D s t 2 = N p D p D s t 2
These parameters are expected to impose geometric limitations on the system’s performance. For example, the receiver radius is fixed for a given receiver capacity. Then a large storage diameter (i.e., a low  R S S R ) would mean a smaller fraction of storage volume is used. This means the energy density would decrease, and the charging time would increase. Following similar arguments,  H D R  and  R u n  would impact charging and discharging times, respectively, imposing additional constraints.
On the other hand, five performance parameters were used to assess the storage behaviour. The receiver–storage efficiency ( η r ) was calculated as the ratio between input energy and stored energy (Equation (25)). The stored energy density ( S E D ) was defined as volume-specific stored energy (Equation (26) below). The solar multiple ( S M ) was calculated as the ratio between solar field capacity and the power cycle’s nominal capacity (Equation (27) below), while the time ratio ( R t ) is a parameter that relates the charging and discharging times (see Equation (28) below). Lastly, the daily delivered energy ( E e l ) is defined as the energy that can be dispatched by the power plant considering both the time of discharge and the nominal power capacity of the plant (Equation (29)).
η r = Q s t Q i n n = 1 N t ρ c p V n T n - T i n i q s u n A r
S E D = Q s t V s t n = 1 N t ρ c p V n T n , s t - T i n i , s t n = 1 N t V n
S M = q c s r A r P e l = G b n C A r P t h η e l = G b n A s f m C O 2 c p , f T f , o u t - T f , i n η e l
R t = t c h t d s
E e l = P e l t d s
In these equations,  N t  is the total number of elements in spatial discretisation,  T i n i  is the initial storage temperature,  T s t -  is the average storage temperature,  G b n = 1000   W / m 2 ,   a n d   C = 1000 -  represents the direct normal beam irradiance and the concentration ratio, respectively.  A s f  is the solar field area, which was calculated for a given ( q c s r A r ).

2.4. Validation

The charging process model cannot be validated directly for high temperature ‘on-sun’ desired conditions since this is a new design. Thus, validation was done for a situation in which a similar phenomenon exists. In this case, the process of stir welding was used, which includes several common characteristics with the current process: a moving heat (circular spot) source over a horizontal surface, conductive heat transfer inside the medium, and natural convective and radiative losses at high temperatures in a horizontal face-up surface. Zhu and Chao [29] have modelled and experimentally tested a friction stir welding process using 304L steel. The steel plates have dimensions of  304.8 × 101.6 × 3.18   m m , the pin tool is  19.05   m m  of diameter, and it moves at  1.693   m m / s  with a thermal power input of  760   W . Three thermocouples were located in the middle of the length plate at different distances from the longitudinal axis (e.g., at 18 mm, 21 mm, and 26.5 mm). Figure 3 shows the temperature profiles over time between Zhou’s experimental testing and our simulation. The results are in good agreement in terms of both the heating rate (slope) and the maximum temperatures.

2.5. Discretisation Test

A discretisation test was conducted to determine the number of elements needed to attain model independence with respect to temporal and spatial discretisation. For this simulation, the number of elements varied from 10 elements per side ( 1 × 10 3  in total) to 200 per side ( 8 × 10 6  in total), while four temporal steps were considered: 360 s, 120 s, 60 s, and 30 s. Figure 4 shows efficiency as a function of spatial and temporal discretisation. For temporal discretisation lower than 60 s, there is almost no difference, whilst for spatial discretisation, the elements quickly converge above 50,000 elements.

3. Results

The results are presented separately for the charging and discharging processes. The charging process is also used to select the proper material from several candidates. Then, a coupled integrated analysis of the overall performance of the plant is presented in the Discussion section.

3.1. Charging Process

The charging process of one storage unit was tested using the conditions defined in Table 1. One unit with a storage diameter  D s t = 4   m  and an initial homogeneous temperature of  850   K  was heated under a receiver with homogeneous radiation of  q c s r = 1   M W m - 2  while it was rotating at  2   r p h  (i.e., two revolutions per hour). The final situation is presented in Figure 5 and Figure 6, which show how the heat is spread inside the storage unit. The final position of the receiver after half an hour of heating (one rotation) is shown in Figure 5 with a black circle. The  R S S R  and  E R  are selected to allow the receiver to be located inside the storage and to avoid overlap (e.g., double heating of areas). For this reason, the central axis and the mantle are the coldest regions. The centre is a useless region, as the rotating system’s structure would be located there, while the mantle region can work as a self-insulation area. In addition, the maximum local temperature is around  1600   K , which is lower than the melting point (i.e.,  1800   K  for cast steel). When this limit is reached, a new unit can be located under the receiver while this unit stabilizes its temperature through conductive heat transfer.
In this study, four materials were tested, and their efficiencies and stored energy densities were assessed. Cast steel, castable ceramics, solid NaCl, and magnesia fire bricks were chosen because they are frequently proposed as solid sensible storage mediums in the literature [30,31]. Their thermodynamic properties are listed in Table 2.
Figure 7 presents the results for these four materials under different rotational speeds, while all geometric parameters were kept constant. The material with the highest efficiency was cast steel, which can be explained by its higher thermal conductivity, which enables the surface under the receiver to stay cooler since the heat is spread relatively faster. It was found that while the rotational speed has a positive effect on the efficiency, above 2 rph; this effect provides diminishing returns. These results are promising compared with thermal efficiency for particle receivers, which operate over a similar temperature range in the literature but have efficiencies in the range of 50–70% for on-sun tests [7].
As cast steel is the only material to reach above 70% efficiency, this material was selected as the solid medium for use in the following analysis. Figure 5 and Figure 6 present the detailed results for cast steel at a speed of 2 rph. Whilst the efficiency and stored energy density are important, the storage unit must be able to deliver this energy towards the power cycle; therefore, it is also necessary to analyse the behaviour of the discharging process.

3.2. Discharging Process

Table 3 presents the initial, geometrical, and operational conditions for a storage unit during the discharging process, which must meet the constraints of the sCO2 cycle. The parameters for the temperature and flow constraints for a sCO2 Brayton cycle were obtained from Reyes-Belmonte et al. [33]. The sCO2 temperature difference is  Δ T = T f o - T f i = 950 - 775 = 175   K  and the thermal-to-power efficiency is set to  η e l = 0.50 - . To simplify the problem, a fixed mass flow rate of  10   k g   s - 1  was defined, and the storage size was set accordingly to the charging process defined in the preceding section; the discharge process considered the operation mode discussed in the proposed design description in the Introduction.
Figure 8a–d show results for the discharging process at various times (0.5, 1.0, 1.5, and 2.0 h). The temperatures of both the storage cylinder and the sCO2 flow were plotted as a function of storage height. At first, at 0.5 h, the LT cylinder discharged faster, which happens because most of the flow circulates inside this unit. It is important to note that in this configuration, the outlet temperature in each cylinder is the inlet temperature in the next one. During the process, the HT cylinder output temperature is higher than what is required ( T f o = 950   K ), but the combined output is always constant and defined by Equation (19). Figure 9 presents the same process over time. The average temperature (left) and mass flow rate (right) of each cylinder are also plotted in the figures. The minimum average temperature of any cylinder determines the initial temperature of the charging process and the stored energy density (through Equation (26)). In this example, the initial charging temperature is close to  820   K  for PH and around 890 K and 960 K for LT and HT cylinders, respectively. This means that the total  Δ T 180   K  for the storage medium during the whole process. Figure 9 (right) shows the evolution of the mass flow rate in the HT and LT cylinders. With this strategy, it is possible to provide a constant output—both in temperature and mass flow rate—for the power cycle over a long period of time, even with transient temperature conditions in the individual storage units. However, in order to find an optimal design, it is necessary to consider both the charging and discharging processes in an integrated model.

4. Discussion

In this section, a design process is discussed. The objective is to maximize both the daily average generated energy and the specific energy density. The relationship between the power block size and storage capacity is a key parameter in this regard. Clearly, an increase in storage size would result in an increase in the discharge time and/or the power block capacity. However, it is not possible to indefinitely increase the size of storage cylinders in this design, nor is it possible for a beam-down receiver to increase the size of a solar field beyond a certain area (e.g., the ~50 MWth limit for a BDR system defined by [10]). The capacity of the power cycle is defined by the mass flow rate (at constant temperature), while the size and number of storage units determine the storage capacity and required solar field size.
The storage unit was assessed in the operation of a typical CST plant using a sCO2 Brayton power plant with the conditions defined in the previous section. To analyse this in detail, multi-factorial simulations were run to obtain the performance parameters of the coupled system. The variables considered are sCO2 mass flow rates ( m f ) (from the power block side), storage unit diameter ( D s t ), height-to-diameter ratio ( H D R ), and maximum storage average temperature ( T a v m x ). The solar field size was determined by the storage diameter and the receiver-storage surface ratio ( R S S R ) via Equations (21) and (27). The number of storage units was set at three, working as indicated in Section 2.2. This configuration represents the minimal system, but the simulations would be similar to the situation in which additional storage units are incorporated until a nominal day (8 h) of charging is achieved. Additionally, the maximum temperature allowed at any point in the storage material is set to  T m a x = 1800   K  to stay below the melting point of cast steel. It was assumed that before starting the charging or discharging processes, the cylinders had been in a standby state long enough to reach thermal equilibrium. Table 4 shows the different parameter values considered in simulations in which a total of 150 conditions were simulated.
The mass flow rate was limited to  100   k g   s - 1  because, with a higher mass flow rate, the discharge time is considerably shorter than the charging time for any condition tested. In addition, the maximum average temperature was limited to  T a v m x = 1100   K  because it is not possible to charge the units to higher temperatures without surpassing the maximum temperature limit of the material (recall ~1800 K for cast steel). The two main parameters of interest are the energy released by the power block ( E e l , see Equation (29)) and the stored energy density ( S E D , see Equation (26)). The first one indicates the amount of energy the system can deliver (and therefore sell), while the last one indicates the amount of material required to store that energy (which is directly related to the capital cost of the storage system).
It should be noted that not all the simulations produce useful results. For all the results in which the discharge process did not occur (i.e., the storage unit was not able to deliver the required output) or if it was too fast (less than 1 h), those results were filtered out. The same was done when the charging process was longer than a nominal day (8 h) or the discharge process was longer than a whole day (24 h). In addition, it was expected that the discharging time would be longer or on the same order as the charging time, so only the results with  R t > 0.75  were considered. These constraints reduced the simulation set to 26 useful alternatives, presented in Figure 10, where each point represents a simulated condition. The different average maximum temperatures are represented by different symbols, while the ellipses indicate conditions with the same storage diameter. The other two parameters (mass flow rate and HDH) are indicated in the labels. Due to the fact that the tested conditions have differences in charging times, they are best compared in terms of the delivered energy over 8 h of available charging time.
Broadly speaking, a higher  T a v m x  allows the system to reach higher SED, but its influence on the total energy delivered is not clear. Nevertheless, for  T a v m x = 1100   K , there were no simulations with delivery energy above  17   M W h ; therefore, a maximum storage temperature of  1050   K  was allowed. The diameter has a clear and direct influence on the delivered energy. For a  4   m  diameter, the energy delivered was around  5   M W h ; for  6   m , it was between  14   a n d   16   M W h . For  8   m , the energy delivered ranged from  25   t o   30   M W h ; for  10   m , it varied between  40  and  46   M W h . No successful simulations were obtained for the  12   m  diameter units. This suggests that  10   m  units are the biggest size allowed. The influence of  H D H  is related to the diameter. For small storage units (4 m and 6 m), both short ( H D H = 0.25 ) and tall ( H D H = 0.50 ) units work; however, for larger units (above 6 m), only short units are useful ( H D H = 0.25 ), except in one case with low SED. The reason for that is, for larger units, the charging process cannot heat up the whole unit (i.e., the heat does not transfer via conduction to the bottom of the cylinder fast enough). Finally, the influence of the power block size (defined by mass flow rate) is not clear. A larger power block would deliver more energy but for a shorter period of time; therefore, these two phenomena compensate each other. A plausible value for stored energy density is  S E D > 70   k W h m - 3 . Therefore, Table 5 presents the detailed results for simulations with  S E D > 70   k W h m - 3  and  E e l > 20   M W h . These results suggest that the solar field size is limited by the storage unit size rather than beam-down receiver optics (~5 MWth here versus 50 MWth, as noted in previous works [10]).

5. Conclusions

A novel concept of a dual receiver–storage unit for concentrated solar thermal plants was presented. This idea uses the beam-down receiver to concentrate solar radiation onto the proposed unit, which would be located on the ground. This concept allows direct, the radiative heat of the solid storage medium, eliminating the need for a solar loop with a liquid working fluid. While this does add some optical complexity, it reduces the mechanical complexity of the tower while also reducing pumping parasitic losses and exergy losses due to fewer heat exchanges. In addition, this work presents a clever discharge strategy to ensure a constant input temperature and flow rate from this storage system for a sCO2 power cycle. To achieve this, at least three storage modules were required to work together in a ‘partial-series discharge’ mode.
A numerical model was built and validated that can simulate both charging and discharging processes under different conditions. The charging process analysis showed that high thermal conductivity is indispensable to spreading the concentrated radiation quickly inside each storage unit. Therefore, among the proposed materials, cast steel was selected since it achieved the highest charging efficiency (>70%).
The simulation of these storage units in a CSP plant indicated some constraints about the expected size this configuration can have. Several conditions were tested, exploring different storage sizes and power plant capacities. It was found that there was no viable solution for plants above  5   M W e m f = 50   k g / s  and storage units above  10   m  in diameter. However, the specific size would depend on the purpose of the plant since multiple smaller towers could be combined into a common power block (as is done in other commercial systems, such as the Vast Solar demonstration facility [34]). Thus, a single BDR receiver could be envisioned to deliver up to a nominal  45.9   M W h  of electricity (with up to 9 h of storage), using  588   m 3  of storage and  15.707   m 2  of solar field to drive a  5   M W e  power plant.
The size limitations found in this work show that BDR is suitable for modular power generation. Additional research should be conducted to study other possible applications where a modular design is also desired, such as industrial processes (SHIP), thermochemical reactions, and desalination. Although the results shown here can be extrapolated to other applications, a specific study for each case should be performed, specifically in the discharge process, using a different HTF. Finally, additional work should be conducted in experimental analysis to further the validation of the existing and techno-economic analyses to optimise the solar field, beam-down optics, and power block sizes.
These conclusions, together with the ability of a sCO2 turbine to ramp quickly [35], suggest that this kind of plant could work well as a small solar-peaking plant. The modular nature of this concept could reduce investment risks and would allow multiple levels of dispatch if modules were distributed to stabilise the grid or if they were co-located with large solar PV installations.

Author Contributions

Conceptualization, D.S. and R.A.T.; methodology, D.S. and R.A.T.; software, D.S.; validation, D.S.; formal analysis, D.S. and R.A.T.; data curation, D.S.; writing—original draft preparation, D.S.; writing—review and editing, D.S. and R.A.T.; visualization, D.S.; supervision, R.A.T.; project administration, D.S. and R.A.T.; funding acquisition, D.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by ANID Chile (Chilean Scientific Agency)’s program “PFCHA: Doctorado Becas Chile 2018” (No. 72190387).

Data Availability Statement

The coding used in this research is available in a Github repository [36]. The data used in the Figure can be reproduced from these scripts.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. Identification of Storage Elements

The charging process is modelled using cartesian coordinate system, while the tanks are cylindrical. Therefore, it is important to identify those elements that belong to the storage, those that are outside the storage domain, and those within the boundary.
The transverse area of the cylinder is a circle, and each element is defined by its central point. Therefore, it is possible to define if a volume element is completely inside the storage, completely out of storage, or if it is partially inside storage (i.e., a boundary element) by assessing the number of element corners ( N c ) inside the storage circumference. A function is built to define this for each element:
f s t = 0.0 i f   N c = 0   ( o u t   s t o r a g e ) 0.0 < f s t < 1.0 i f   1 N c 3   ( b o u n d a r y ) 1.0 i f   N c = 4   ( f u l l   s t o r a g e )
For elements in the boundary, the fraction of each boundary (north, south, east, and west) belonging to the storage unit is computed. These values are proportional to the equivalent area of conduction in each direction. The percentage of volume under storage is calculated approximating the circumference arc of the storage surface with trapezes or triangles, according to the number of nodes belonging to storage. A schematic of this process is shown in Figure A1.
Figure A1. Examples of elements and how function  f s  works. Element A is completely inside storage, and B and C have two and three corners inside storage, respectively. Element D has one corner inside storage. The corners inside storage are solid dots. The grey area with segmented lines indicates the approximation used to the define fraction of volume. The common border between elements is labelled in green.
Figure A1. Examples of elements and how function  f s  works. Element A is completely inside storage, and B and C have two and three corners inside storage, respectively. Element D has one corner inside storage. The corners inside storage are solid dots. The grey area with segmented lines indicates the approximation used to the define fraction of volume. The common border between elements is labelled in green.
Energies 16 04157 g0a1
In Figure A1, the element A has all its boundaries inside storage, then  f s t , A = 1 . For elements with one corner inside storage (Element D in Figure A1), the fraction of the volume of D inside the storage is approximated by the triangle defined by the boundaries with elements C and D ( b C D  and  b B D ), then for element D:
f s t , D = b C D b B D 2
For elements with two corners inside the boundary (element B), the storage fraction is proportional to the trapeze defined by the sides ( b A B , b B D ,  and the north boundary  b B N ).
f s t , B = b B D + b B N 2
Note that  b A B = 1 , so it is not part of the calculation. Finally, for elements with three corners, the fraction is calculated as the square minus the triangle that is not part of the storage. Using Element C as an example:
f s t , C = 1 - 1 - b C S 1 - b C D 2
The elements outside the storage are considered perfect thermal insulators ( λ = α = 0.0 ), which ensure adiabatic conditions and numerical consistency. A similar function ( f r ) is defined to determine which elements of the upper surface are under the receiver. The only difference is that this function must be evaluated in each timestep because the storage is rotating, and therefore, the elements under the receiver are changing.

Appendix B. General Description for Discharging Process Configurations

Two different discharge strategies were presented in Section 1: denominated stand-alone operation and partial-series operation. However, an array of different strategies could be proposed. Therefore, a general design was developed that allows testing for any possible configuration. Later, some of the possibilities are analysed, and the best options are selected. The two options selected are presented in the main text.
The general configuration between two tanks at different temperatures (low temperature and high temperature) is presented in Figure A2. The different states are indicated with Arabic numbers, while the nodes with mass flow rate mixes are indicated with Roman numbers. In this configuration, it is assumed that the initial fluid temperature  T f i n , the required outlet temperature ( T f o u t ), and the overall mass flow rate ( m f ) are known. There are 9 states with unknown mass flow rates and temperatures ( m i , T i ), and the flows on the tanks  m f L  and  m f H  are also unknown, comprising 20 unknown variables to be solved. The mass conservation and energy conservation on each node and the storage units are presented in Table A1.
Figure A2. General configuration for discharge process with two tanks. The different states are shown in red Arabic numbers, while the different nodes are shown in blue Roman numbers.
Figure A2. General configuration for discharge process with two tanks. The different states are shown in red Arabic numbers, while the different nodes are shown in blue Roman numbers.
Energies 16 04157 g0a2
Table A1. Mass conservation and Energy conservation equations for each node, LT tank and HT tank.
Table A1. Mass conservation and Energy conservation equations for each node, LT tank and HT tank.
NodeMass Conservation Energy Conservation
  ( i )   m 1 = m f L + m 7 (A5)   T 1 m 1 = m f L T f i n + m 7 T 7 (A6)
  ( i i )   m 3 = m f H + m 8 + m 9 (A7)   T 3 m 3 = m f H T f i n + m 8 T 8 + m 9 T 9 (A8)
  ( i i i )   m f = m 6 + m 4 - m 9 (A9)   T f o u t m f + m 9 = m 6 T 6 + m 4 T 4 (A10)
  ( i v )   m 2 = m 6 + m 5 (A11)   T 2 = T 5 (A12)   T 2 = T 6 (A13)
  ( v )   m 5 = m 7 + m 8 (A14)   T 5 = T 7 (A15)   T 5 = T 8 (A16)
  ( L T )   m 1 = m 2 (A17)   T 2 = f s g T s t L , m 1 , T 1 (A18)
  ( H T )   m 3 = m 4 (A19)   T 4 = f s g T s t H , m 3 , T 3 (A20)
Equations (A18) and (A20) are the energy balance in the storage units described in Section 2 and can be represented by a non-linear storage function,  f s g . There are four additional equations required, which come from four decisions needed to split the mass flow rates. These decisions are expressed as ratios:
(a)
Ratio among HT and LT mass flow rates:
R H - L = m f H m f
(b)
Ratio between parallel and series operation:
R p - s = m 6 m 2
(c)
Ratio of LT tank recirculation:
R r L = m 7 m 5
(d)
Ratio of HT tank recirculation:
R r H = m 9 m f = m 9 m 6 + m 4
The combinations of these four ratios allow to set the different operation modes and solve the system to ensure the operation goal, which is to obtain constant mass flow rate ( m f ) and outlet temperature ( T f o u t ) to the power system.
Six modes of operation were programmed and tested. Those more adequate to operate and extract the most heat possible from the storage units were selected and presented in the main text. The operation modes are depicted in Figure A3, followed by a brief explanation of them.
  • Operation Mode 0: One tank with recirculation. In this case, only one storage tank is working. The outlet temperature is higher than the required one, so it operates with a higher mass flow rate, and part of it is recirculated. While the tank is operating, its temperature decreases, so over time a lower mass flow rate is required until it reaches a point where the nominal sCO2 mass flow rate is reached, and no more heat can be extracted from the unit without losing stable output. In this configuration, three ratios are constant:  R H - L = 1.0 R r L = 0.0 R p - s = 0.0 ; the stable output is reached by modifying the ratio  R r H . The system stops when  R r H = 1.0 . In other words, when there is no more recirculation.
  • Operation Mode 1: Two tanks in parallel without recirculation. This is a common parallel configuration. It requires that both tanks are in different temperatures. The combination of both flows should give a stable output in mass and temperature. In this configuration,  R p - s = 1.0  and  R r H = 0.0 R r L  is not needed to be defined ( m 5 = m 7 = 0 ) . The stable output is reached with the ratio  R H - L . The process stops when  R H - L = 1.0 .
  • Operation Mode 2: Partial-series operation. In this case, all the flow goes through the LT tank, but only a fraction goes through HT. The fraction that goes through HT is calculated to ensure the required output. The fixed ratios are  R H - L = 0.0 R r L = 0.0 R r H = 0.0 , and  R p - s  and vary to ensure a stable output. When  R p - s = 0.0 , the system transposes to a full series mode and a new tank is required.
  • Operation Mode 3: Series with recirculation in LT. In this case, the LT and HT tanks are in a series, but recirculation is included in the LT tank to extract more energy from that tank. The fixed ratios are  R H - L = 0.0 R p - s = 0.0 R r H = 0.0 , and the stable outputs are reached by modifying  R r L . When  R r L = 0.0 , it transposes to a full series mode; the LT tank becomes the preheater, and a new tank is needed in order to keep the desired output.
  • Operation Mode 4: Series with recirculation in HT. This case is similar to the previous one, but the recirculation is imposed on the HT tank instead of the LT. The fixed ratios are  R H - L = 0.0 R p - s = 0.0 R r L = 0.0 , and the stable outputs are reached by modifying  R r H . Similar to the previous one, when  R r H = 0.0 , it transposes to a full series mode, and a new tank is needed. Although this mode is implemented, it is clear that this extracts more heat from the HT than the LT, which is not what is expected, as the idea is to use most of the heat from the LT tank.
  • Operation Mode 5: Stand-alone with bypass. This is an improvement to mode 0, with only one tank in operation. Instead of recirculating flow in the tank, part of the flow is bypassed and goes directly to the power cycle. The fraction that circulates in the tank reaches a higher temperature, and the combined temperature and mass flow rate at the end are stable. In this case, the ratios are similar to mode 0:  R p - s = 1.0 R r L = 0.0 R r H = 0.0 ,  and the stable output is reached by modifying the ratio  R H - L . The system stops when  R r H = 1.0 , when there is no more bypass flow.
Initially, when all the tanks are fully charged, only those modes with one tank in operation (modes 0 or 5) can operate. From those, mode 5 is more efficient and is used. Then, when two tanks are required, mode 2 showed better results, as it extracts most of the heat from LT, and it was selected. Once the first LT cannot be used anymore, it becomes a pre-heater tank (PH), the HT becomes the new LT, and a new HT tank is required. This detailed operation mode is explained in the main text.
Figure A3. Different operation modes tested. Each mode can be programmed by setting some of the flows and nodes in the general configuration.
Figure A3. Different operation modes tested. Each mode can be programmed by setting some of the flows and nodes in the general configuration.
Energies 16 04157 g0a3

References

  1. Guterres, A. The Sustainable Development Goals Report 2017; United Nations: San Francisco, CA, USA, 2017. [Google Scholar]
  2. Islam, M.T.; Huda, N.; Abdullah, A.B.; Saidur, R. A Comprehensive Review of State-of-the-Art Concentrating Solar Power (CSP) Technologies: Current Status and Research Trends. Renew. Sustain. Energy Rev. 2018, 91, 987–1018. [Google Scholar] [CrossRef]
  3. Zhuang, X.; Xu, X.; Liu, W.; Xu, W. LCOE Analysis of Tower Concentrating Solar Power Plants Using Different Molten-Salts for Thermal Energy Storage in China. Energies 2019, 12, 1394. [Google Scholar] [CrossRef]
  4. Renewable Power Generation Costs in 2021; IRENA: Abu Dhabi, United Arab Emirates, 2021; p. 204.
  5. Stein, W.H.; Buck, R. Advanced Power Cycles for Concentrated Solar Power. Sol. Energy 2017, 152, 91–105. [Google Scholar] [CrossRef]
  6. Giaconia, A.; Tizzoni, A.C.; Sau, S.; Corsaro, N.; Mansi, E.; Spadoni, A.; Delise, T. Assessment and Perspectives of Heat Transfer Fluids for CSP Applications. Energies 2021, 14, 7486. [Google Scholar] [CrossRef]
  7. Ho, C.K. Advances in Central Receivers for Concentrating Solar Applications. Sol. Energy 2017, 152, 38–56. [Google Scholar] [CrossRef]
  8. Stern, T.G.; Cornwall, M.; Kaincz, B.; Mildice, J.W. U.S. Patent No. 4,784,700, 15 November 1988.
  9. Segal, A.; Epstein, M. The Optics of the Solar Tower Reflector. Sol. Energy 2001, 69, 229–241. [Google Scholar] [CrossRef]
  10. Segal, A.; Epstein, M. Practical Considerations in Designing Large Scale “Beam Down” Optical Systems. J. Sol. Energy Eng. 2008, 130, 011009. [Google Scholar] [CrossRef]
  11. Segal, A. Solar Energy at High Temperatures; Researches at the Weizmann Institute of Science, Israel; 25 Years of Success. Renew. Energy Environ. Sustain. 2016, 1, 1. [Google Scholar] [CrossRef]
  12. Calvet, N.; Martins, M.; Grange, B.; Perez, V.G.; Belasri, D.; Ali, M.T.; Armstrong, P.R. The Masdar Institute Solar Platform: A New Research Facility in the UAE for Development of CSP Components and Thermal Energy Storage Systems. AIP Conf. Proc. 2016, 1734, 100003. [Google Scholar] [CrossRef]
  13. Grange, B.; Kumar, V.; Gil, A.; Armstrong, P.R.; Codd, D.S.; Slocum, A.; Calvet, N. Preliminary Optical, Thermal and Structural Design of a 100 KWth CSPonD Beam-down On-Sun Demonstration Plant. Energy Procedia 2015, 75, 2163–2168. [Google Scholar] [CrossRef]
  14. Lahlou, R.; Armstrong, P.; Grange, B.; Almheiri, S.; Calvet, N.; Slocum, A.; Shamim, T. Thermal Modeling of a Secondary Concentrator Integrated with an Open Direct-Absorption Molten-Salt Volumetric Receiver in a Beam-down Tower System. In Proceedings of the AIP Conference Proceedings, Cape Town, South Africa, 13–16 October 2016; p. 020012. [Google Scholar]
  15. Mokhtar, M.; Meyers, S.A.; Armstrong, P.R.; Chiesa, M. Performance of a 100 KWth Concentrated Solar Beam-Down Optical Experiment. J. Sol. Energy Eng. 2014, 136, 041007. [Google Scholar] [CrossRef]
  16. Musi, R.; Grange, B.; Diago, M.; Topel, M.; Armstrong, P.; Slocum, A.; Calvet, N. Techno-Economic Optimization of a Scaled-up Solar Concentrator Combined with CSPonD Thermal Energy Storage. AIP Conf. Proc. 2017, 1850, 110010. [Google Scholar] [CrossRef]
  17. Kodama, T.; Gokon, N.; Matsubara, K.; Yoshida, K.; Koikari, S.; Nagase, Y.; Nakamura, K. Flux Measurement of a New Beam-down Solar Concentrating System in Miyazaki for Demonstration of Thermochemical Water Splitting Reactors. Energy Procedia 2014, 49, 1990–1998. [Google Scholar] [CrossRef]
  18. Chirone, R.; Salatino, P.; Ammendola, P.; Solimene, R.; Magaldi, M.; Sorrenti, R.; Michele, G.D.; Donatini, F. Development of a Novel Concept of Solar Receiver/Thermal Energy Storage System Based on Compartmented Dense Gas Fluidized Beds. In Proceedings of the 14th International Conference on Fluidization—From Fundamentals to Products, Noordwijkerhout, The Netherlands, 26–31 May 2013; p. 10. [Google Scholar]
  19. Miller, S.A. To Beam or Not to Beam Down. In Proceedings of the ISES Solar World Congress 2017 and IEA SHC Solar Heating and Cooling Conference for Buildings and Industry 2017, Abu Dhabi, United Arab Emirates, 29 October–2 November 2017; pp. 146–157. [Google Scholar]
  20. Yumen Xinneng/Xinchen—50MW Beam-Down|Concentrating Solar Power Projects|NREL. Available online: https://solarpaces.nrel.gov/project/yumen-xinneng-xinchen-50mw-beam-down (accessed on 20 December 2022).
  21. SolarPACES Huidong New Energy Akesai 110MW Beam-Down Tower + 640MW PV|Concentrating Solar Power Projects|NREL. Available online: https://solarpaces.nrel.gov/project/huidong-new-energy-akesai-110mw-beam-down-tower-640mw-pv (accessed on 20 December 2022).
  22. Kiwan, S.; Soud, Q.R. Numerical Investigation of Sand-Basalt Heat Storage System for Beam-down Solar Concentrators. Case Stud. Therm. Eng. 2019, 13, 100372. [Google Scholar] [CrossRef]
  23. Iniesta, A.C.; Diago, M.; Delclos, T.; Falcoz, Q.; Shamim, T.; Calvet, N. Gravity-Fed Combined Solar Receiver/Storage System Using Sand Particles as Heat Collector, Heat Transfer and Thermal Energy Storage Media. Energy Procedia 2015, 69, 802–811. [Google Scholar] [CrossRef]
  24. Gómez-Hernández, J.; González-Gómez, P.Á.; Ni-Song, T.; Briongos, J.V.; Santana, D. Design of a Solar Linear Particle Receiver Placed at the Ground Level. In Proceedings of the AIP Conference Proceedings, Santiago, Chile, 26–29 September 2017; p. 170005. [Google Scholar]
  25. Cengel, Y. Heat and Mass Transfer—A Practical Approach, 3rd ed.; McGraw-Hill: New York, NY, USA, 2006. [Google Scholar]
  26. Kakaç, S.; Shah, R.K.; Aung, W. (Eds.) Handbook of Single-Phase Convective Heat Transfer; Wiley: New York, NY, USA, 1987; ISBN 978-0-471-81702-4. [Google Scholar]
  27. Goodwin, D.G.; Speth, R.L.; Moffat, H.K.; Weber, B.W. Cantera: An Object-Oriented Software Toolkit for Chemical Kinetics, Thermodynamics, and Transport Processes, Version 2.2.0. 2018. Available online: https://zenodo.org/record/1174508#.ZGRMiKVBxPZ (accessed on 14 April 2023).
  28. Xu, B.; Li, P.-W.; Chan, C.L. Extending the Validity of Lumped Capacitance Method for Large Biot Number in Thermal Storage Application. Sol. Energy 2012, 86, 1709–1724. [Google Scholar] [CrossRef]
  29. Zhu, X.K.; Chao, Y.J. Numerical Simulation of Transient Temperature and Residual Stresses in Friction Stir Welding of 304L Stainless Steel. J. Mater. Process. Technol. 2004, 146, 263–272. [Google Scholar] [CrossRef]
  30. Gil, A.; Medrano, M.; Martorell, I.; Lázaro, A.; Dolado, P.; Zalba, B.; Cabeza, L.F. State of the Art on High Temperature Thermal Energy Storage for Power Generation. Part 1—Concepts, Materials and Modellization. Renew. Sustain. Energy Rev. 2010, 14, 31–55. [Google Scholar] [CrossRef]
  31. Kuravi, S.; Trahan, J.; Goswami, D.Y.; Rahman, M.M.; Stefanakos, E.K. Thermal Energy Storage Technologies and Systems for Concentrating Solar Power Plants. Prog. Energy Combust. Sci. 2013, 39, 285–319. [Google Scholar] [CrossRef]
  32. Bansal, N.P.; Doremus, R.H. Handbook of Glass Properties; Elsevier: Amsterdam, The Netherlands, 1986. [Google Scholar]
  33. Reyes-Belmonte, M.A.; Díaz, E.; Romero, M.; González-Aguilar, J. Particles-Based Thermal Energy Storage Systems for Concentrated Solar Power. In Proceedings of the AIP Conference Proceedings, Santiago, Chile, 26–29 September 2017; p. 210013. [Google Scholar]
  34. Wood, C.; Drewes, K. Vast Solar: Improving Performance and Reducing Cost and Risk Using High Temperature Modular Arrays and Sodium Heat Transfer Fluid. In Proceedings of the SolarPaces Conference, Daegu, Republic of Korea, 1–4 October 2019; p. 8. [Google Scholar]
  35. Luu, M.T.; Milani, D.; McNaughton, R.; Abbas, A. Advanced Control Strategies for Dynamic Operation of a Solar-Assisted Recompression Supercritical CO2 Brayton Power Cycle. Appl. Therm. Eng. 2018, 136, 682–700. [Google Scholar] [CrossRef]
  36. Saldivia, D. MonteCarlo Ray Tracing Module for Beam Down Receivers. Available online: https://github.com/DavidSaldivia/BDR_MCRT (accessed on 14 April 2023).
Figure 1. (left) Scheme of proposed solution during charging stage of one storage unit. (right) Layout of proposed system with eight storage units.
Figure 1. (left) Scheme of proposed solution during charging stage of one storage unit. (right) Layout of proposed system with eight storage units.
Energies 16 04157 g001
Figure 2. Discharge control operation modes.
Figure 2. Discharge control operation modes.
Energies 16 04157 g002
Figure 3. Comparison between experimental results (extracted from Zhu and Chao [29]) and simulations for welding process. Simulation in lines, experimental data in dots.
Figure 3. Comparison between experimental results (extracted from Zhu and Chao [29]) and simulations for welding process. Simulation in lines, experimental data in dots.
Energies 16 04157 g003
Figure 4. Discretisation test for base case described in Section 3.
Figure 4. Discretisation test for base case described in Section 3.
Energies 16 04157 g004
Figure 5. Detailed result for an example case after half hour of simulation. Upper view for  z = H s t  (upper surface); a black circumference shows the final position of receiver.
Figure 5. Detailed result for an example case after half hour of simulation. Upper view for  z = H s t  (upper surface); a black circumference shows the final position of receiver.
Energies 16 04157 g005
Figure 6. Detailed result for an example case after half hour of simulation. Four different side views are shown. (a y = 0 , (b y = x , (c x = 0 , and (d y = - x . In each case, vertical axis is height (z) and horizontal axis is diameter.
Figure 6. Detailed result for an example case after half hour of simulation. Four different side views are shown. (a y = 0 , (b y = x , (c x = 0 , and (d y = - x . In each case, vertical axis is height (z) and horizontal axis is diameter.
Energies 16 04157 g006
Figure 7. Comparison between four different materials. The rotational speed (rph, revolutions per hour) is varied, efficiency is shown in solid lines, and stored energy density in dashed lines.
Figure 7. Comparison between four different materials. The rotational speed (rph, revolutions per hour) is varied, efficiency is shown in solid lines, and stored energy density in dashed lines.
Energies 16 04157 g007
Figure 8. Detailed results of the discharging process for two cylinders and a preheater using the mode “partial-series” strategy (discussed in Section 1) at different times. (a) 0.5 h, (b) 1.0 h, (c) 1.5 h, and (d) 2.0 h. For all figures, the modules are label as HT = red lines, LT = blue lines, and PH = green lines. The CO2 flow temperature is denoted with solid lines, and the average storage temperature is represented with dashed lines.
Figure 8. Detailed results of the discharging process for two cylinders and a preheater using the mode “partial-series” strategy (discussed in Section 1) at different times. (a) 0.5 h, (b) 1.0 h, (c) 1.5 h, and (d) 2.0 h. For all figures, the modules are label as HT = red lines, LT = blue lines, and PH = green lines. The CO2 flow temperature is denoted with solid lines, and the average storage temperature is represented with dashed lines.
Energies 16 04157 g008
Figure 9. Average temperature profile (left) and mass flow rate (right) for three-cylinder discharge process. HT (red solid line), LT (blue dashed dots), and PH cylinders (green dashed).
Figure 9. Average temperature profile (left) and mass flow rate (right) for three-cylinder discharge process. HT (red solid line), LT (blue dashed dots), and PH cylinders (green dashed).
Energies 16 04157 g009
Figure 10. Daily delivered energy and stored energy density for successful simulations. Average storage temperatures indicated with different symbols, ellipses indicate diameters, HDR indicated by colours (dark for 0.25 and light for 0.5), and mass flow rate indicated with labels.
Figure 10. Daily delivered energy and stored energy density for successful simulations. Average storage temperatures indicated with different symbols, ellipses indicate diameters, HDR indicated by colours (dark for 0.25 and light for 0.5), and mass flow rate indicated with labels.
Energies 16 04157 g010
Table 1. Initial, geometrical, and operational conditions of the storage unit during the charging process.
Table 1. Initial, geometrical, and operational conditions of the storage unit during the charging process.
GeometryValueReceiver OpticsValueInitial Conditions and DiscretizationValue
Storage Diameter,  D s t     4.00   m Quartz Transmittance
(Solar-weighted),  τ g  
  0.90 Initial Temp,  T i n i     850   K
Eccentricity Ratio,  E R     1.22 - Quartz Absorptance
(Solar-weighted),  α g  
  0.05 Ambient Temp,  T a m b     300   K
Receiver–Storage Surface Ratio,  R S R     0.20 - Quartz Absorptance
(Black-body-weighted),  α g  
  0.35 Solar Inlet Heat,  q c s r     1.0   MW m - 2
Height-to-Diameter Ratio,  H D R     0.50 - Quartz Reflectivity
(Solar-weighted),  ρ g  
  0.05 Charging time per storage unit,  t     1800   s
Rotational speed,  ω s t     2.0   r p h Storage and quartz emissivity,  ϵ     0.80 Temp Discretization,  Δ t     30   s
Table 2. Materials Properties at 300 °C. Sources indicated in each row.
Table 2. Materials Properties at 300 °C. Sources indicated in each row.
MaterialRef.Density
ρ k g m - 3
Conductivity
λ W m - 1 K - 1
Heat Capacity
c p J   k g - 1 K - 1
Thermal Diff.
α m 2   s - 1
Cast steel[31]780040.0600   8.5 × 10 - 6
Castable ceramic[31]35001.4866   4.5 × 10 - 7
Magnesia fire bricks[31]30005.01150   1.4 × 10 - 6
NaCl[31]21607.0850   3.8 × 10 - 6
Quartz (cover)[32]22001.71000   7.3 × 10 - 7
Table 3. Initial, geometrical, and operational conditions of the receiver–storage unit during the discharging process.
Table 3. Initial, geometrical, and operational conditions of the receiver–storage unit during the discharging process.
ConditionsSymbolValueConditionsSymbolValue
Pipes diameter   D p   0.05   m HT block Initial temperature   T i n i , H T   1000   K
Ratio non-use   R n u   0.10 - LT block Initial temperature   T i n i , L T   970   K
Fluid mass flow rate   m f   10   k g   s - 1 PH block Initial temperature   T i n i , P H   900   K
Inlet fluid temperature   T f i   775   K Outlet fluid temperature   T f o   950   K
Table 4. Parameter values used in simulations.
Table 4. Parameter values used in simulations.
ConditionsSymbolValueConditionsSymbolValue
Pipes diameter   D p   0.05   m HT block initial temperature   T i n i , H T   1000   K
Ratio non-use   R n u   0.10 - LT block initial temperature   T i n i , L T   970   K
Fluid mass flow rate   m f   10   k g   s - 1 PH block initial temperature   T i n i , P H   900   K
Inlet fluid Temperature   T f i   775   K Outlet fluid temperature   T f o   950   K
Table 5. Summary of performance results for selected simulations.
Table 5. Summary of performance results for selected simulations.
  D s t
m
  H D R
-
  T s t m x
K
  m C O 2
k g   s - 1
  E e l
M W h
  S E D
k W h m - 3
  V s t
m 3
  R t
-
  P e l
M W
  S M
-
  A S F
m 2
80.2510002527.976.83301.21.392.51.9310053.1
80.251050528.181.99301.27.025.09.6510053.1
100.2510005044.082.82588.31.100.51.5115707.9
100.2510501044.281.00588.35.521.07.5415707.9
100.2510005045.9113.65588.31.155.01.5115707.9
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

Saldivia, D.; Taylor, R.A. A Novel Dual Receiver–Storage Design for Concentrating Solar Thermal Plants Using Beam-Down Optics. Energies 2023, 16, 4157. https://doi.org/10.3390/en16104157

AMA Style

Saldivia D, Taylor RA. A Novel Dual Receiver–Storage Design for Concentrating Solar Thermal Plants Using Beam-Down Optics. Energies. 2023; 16(10):4157. https://doi.org/10.3390/en16104157

Chicago/Turabian Style

Saldivia, David, and Robert A. Taylor. 2023. "A Novel Dual Receiver–Storage Design for Concentrating Solar Thermal Plants Using Beam-Down Optics" Energies 16, no. 10: 4157. https://doi.org/10.3390/en16104157

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