Next Article in Journal
A Review of Gravitational Water Vortex Hydro Turbine Systems for Hydropower Generation
Previous Article in Journal
Energy Consumption and Grid Interaction Analysis of Electric Vehicles Based on Particle Swarm Optimisation Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Techno-Economic Assessment of CPVT Spectral Splitting Technology: A Case Study on Saudi Arabia

CCRC, King Abdullah University of Science and Technology (KAUST), Thuwal 23955, Saudi Arabia
*
Author to whom correspondence should be addressed.
Energies 2023, 16(14), 5392; https://doi.org/10.3390/en16145392
Submission received: 28 May 2023 / Revised: 4 July 2023 / Accepted: 12 July 2023 / Published: 14 July 2023
(This article belongs to the Topic Advances in Solar Technologies)

Abstract

:
Concentrating PV thermal (CPVT) collector with spectral splitting technology is a promising solution for heat and electricity production. To extend the use of this technology, a novel and cost-effective CPVT collector for harsh environments, such as those in Saudi Arabia, is presented and evaluated using theoretical energy, economy, and environmental analysis. Two questions are answered in this study, namely: which is the best operation strategy, and which is the best energy storage technology for CPVT. The potential of using a CPVT under the climate conditions of six cities in Saudi Arabia is also evaluated. It is found that a heat/electricity production strategy and a thermal energy storage are the most suitable for the CPVT technology. The economic assessment shows a levelized cost of electricity (LCOE) of $0.0847/kWh and a levelized cost of heat (LCOH) of $0.0536/kWh when water is used as a spectral filter, and a LCOE of $0.0906/kWh and a LCOH of $0.0462/kWh when ZnO nanoparticles are added. The CO2-equivalent emissions in a 20 MW CPVT plant are cut from 5675 tonnes to 7822 tonnes per year for Saudi Arabian weather and present power generation conditions.

1. Introduction

Solar energy can play an important role in a sustainable worldwide energy supply to address carbon emission and climate change. The range and the applications of solar energy conversion devices have expanded dramatically in recent years, with the objective of reducing reliance on fossil fuels. Solar energy can be converted into useful energy using thermal and photovoltaic (PV) collectors.
Traditional PV collectors convert part of the solar spectrum into electricity (typical efficiency of traditional PV panels is about 20%). The rest of the energy received by the PV panel is converted into heat, decreasing its performance. This constraint has led research groups all around the world to seek ways to use the solar radiation that cannot be converted by the PV cells into electricity, in other words, to be able to exploit the entire solar spectrum while preventing photovoltaic cell from overheating.
In this regard, spectrum beam splitting (SBS) has been the technique that has undergone the greatest progress in recent years. It employs filters that split the incoming solar radiation into different wavelengths. The solar radiation within the spectral window, useful for the photovoltaic effect, is directed to PV panels, while the unutilized energy by the PV panels is directed and absorbed by a heat transfer fluid (HTF) to generate heat. The PV panel and the solar thermal collector is combined into a single unit, which is known as, concentrating solar photovoltaic thermal (CPVT) collector.
There are three main methods to split solar radiation into different ranges of wavelengths: interference filtering, use of semi-transparent PV panels, and selective absorption. The challenge of using interference filters is their complicated manufacturability and high cost [1,2]. Some limitations of the semi-transparent PV panels include development of semi-transparent electrodes [3], insulation issues [4] and that some materials are made semi-transparent by reducing the semiconductor’s layer thickness; however, doing so results in a reduction in performance [5]. Alternatively, selective absorbers employing HTFs could be a more affordable approach. An HTF that is transparent to the desired wavelengths for PV cells is located in front of them, letting those wavelengths be transmitted to the cells. The HTF is highly absorbing in the rest of the spectrum [2]. From the economic point of view, selective absorption is a cost-effective technique since the working fluid can be water [6].
Several researchers have been working to develop the CPVT technology. One of the first studies was performed by Soule [7], who proposed, in 1987, a CPVT using dome-shaped linear Fresnel lenses as the concentrator with a dielectric-Au-dielectric multilayer filter. The system produced electricity, low-temperature thermal energy (50–70 °C), and high-temperature thermal energy (150–250 °C). The corresponding efficiencies are 9.5%, 41.9%, and 17.8%, respectively. A CPVT with SBS and a parabolic trough collector (PTC) has been proposed by Zhang et al. [8]. The system achieved a maximum electrical efficiency of 22.64%. Some studies showed that a CPVT with PTC can reach an overall thermal efficiency of 70% and an overall electrical efficiency of 25%, while a system with a Linear Fresnel Collector (LFC) can achieve a thermal efficiency of more than 60% and an electrical efficiency of more than 20% [9].
Ling et al. [10] investigated a CPVT with LFC and a selective filter and found a levelized cost of electricity (LCOE) of $0.20/kWh. Recently, Liew et al. [11] proposed a photovoltaic/concentrated solar power hybrid plant to increase the performance of a concentrated solar power plant currently operating in California, USA. The proposed hybrid system performed 9% better than the actual one and was also 4% more efficient than the virtual photovoltaic-alone scenario.
Instead of using solid filters for SBS, liquid absorptive filters can be used and have several advantages [12]. The absorptive liquid is often inexpensive and can perform numerous functions: it absorbs the unused spectral solar irradiance by PV cells; thermal energy can be transported and stored by absorptive fluids; and it could be used as the coolant of PV modules to extract the dissipated heat from the solar cells. Sabry et al. [13] theoretically demonstrated that an ideal liquid filter, which matches the spectral response of silicon solar cells, significantly reduces the solar cells’ operating temperature and increases their efficiency by 30%. The performance of a combined liquid and solid absorptive filter on a compact CPVT receiver for an LFC was investigated by Manfred et al. [14]. They found that, for Seville (Spain), the receiver can achieve an electrical efficiency of up to 6.2% and a thermal efficiency of up to 61.2%.
Advances in nanotechnology have resulted in nanoparticles that can selectively filter solar radiation and can be added to a base fluid to modify its optical characteristics. Meraje et al. [15] designed and validated a CPVT based on LFC and a nanofluid spectrum splitting filter. They evaluated several volume concentrations of ZnO nanoparticles. The closest spectrum match with a silicon solar cell was determined to be 0.00089 vol%. Recently, Barthwal et al. [16] examined the utilization of deionized water and ZnO nanoparticles as optical filters in a compound parabolic-concentrate-based CPVT. They evaluated it for conditions in New Delhi (India) and concluded that the cell temperature was kept near the standard test. Wang et al. [17] studied a CPVT with compact LFC and Ag/CoSO4-PG nanofluids. The performance estimation showed that the PV module has a photoelectric efficiency of 30.2%, and the receiver has a thermal efficiency up to 49.3%.
In terms of the applications for CPVT, Su et al. [18] investigated the feasibility of applying CPVT to boost biomethane generation in anaerobic digestion via biogas upgrading. They also proposed the use of CPVT for trigeneration (heat, cooling, and electricity) [19]. At Tucson (United States), Fernandes et al. [20] carried out a simulation for a small-scale nanofluid spectral filtering CPVT for domestic applications. The possibility of using CPVT for water desalination has also been investigated by several authors as reviewed by Anand. et al. [21]. Another recent application of a CPVT was proposed by Youssef et al. [22].
While many of the previous studies have investigated different types of CPVT collectors and highlighted their thermal performance, very few publications have reported on the operation strategy, the optimum heat versus electricity storage, or evaluated the benefits under harsh weather conditions, such as extremely high ambient temperatures and high levels of aerosols prevalent in places like Saudi Arabia. The objective of this paper is to address these shortcomings using Saudi Arabia as a case study.
To do so, a detailed techno-economic theoretical assessment is carried out. A CPVT with a novel receiver design, suitable for the harsh conditions, is investigated under the climate of six cities in Saudi Arabia. To provide a comprehensive analysis, a mathematical model is developed to investigate the optical and thermal performance of the proposed CVPT. For each location studied, a year-round performance assessment considering the hourly variation of solar radiation, sun position, ambient temperature, and wind speed is conducted. A comparison is then made for all cities and under all operating and storage scenarios.

2. Materials and Methods

2.1. Description of the CPVT

A Linear Fresnel Collector (LFC) with a hybrid receiver fitted at its focal axis is proposed in this study. As illustrated in Figure 1, the proposed system consists of mirrors, a thermal receiver with cooling channel, a heat transfer fluid that also plays the role of a filter, and a silicon bifacial PV module with a 22% nominal efficiency at 25 °C. The mirrors focus direct normal irradiance on the receiver’s front surface. The fluid is used as spectral filter, absorbing low and high-energy photons and converting them into useful heat. As a result, a suitable solar radiation spectrum for silicon PV cells reaches the PV module, which is placed above the nanofluid. Due to the bifaciality factor of the solar cell, the side with the highest efficiency faces the concentrated solar radiation to maximize energy production. The cooling channel is used to reduce the PV module temperature. The design values of the proposed system are presented in Table 1.
A detailed design of the receiver is presented in Figure 2a. It consists of the main liquid channel and the cooling channel together with the PV module. As illustrated in Figure 2b, these two channels are linked by a U-shaped pipe to enhance thermal efficiency [22]. The liquid initially flows at room temperature through the cooling channel to cool down the PV panel. As a result, the panel’s temperature drops, its efficiency increases, and the HTF is preheated before entering the main receiver channel. Figure 2c highlights the main parts of the receiver.
As highlighted in Figure 2c, concentrated light passes through the highly transparent glass and across the working fluid. The working fluid acts as a spectral filter, absorbing solar radiation with wavelengths less than 700 nm or greater than 1100 nm. As a result, only solar radiation within the spectral window of between 700 nm and 1100 nm reaches the PV module. The receiver’s side walls are painted with selective, highly absorbent materials.
In this study, two different working fluids, namely water and a water-based ZnO nanofluid (0.01 wt%), were examined. The introduction of nanoparticles into the water resulted in alterations within the thermophysical and spectral characteristics of the fluid, as documented in Table 2 and Table 3, respectively. The evaluation of the thermophysical properties was carried out under atmospheric pressure and at an approximate average fluid temperature of 62.5 °C, representing an average working fluid temperature of our system.
The present investigation focuses on the photovoltaic active range of 700 nm to 1100 nm for silicon solar cells, in accordance with prior research [26]. Notably, the study does not encompass the photovoltaic active spectrum spanning 400–700 nm, where energy states surpass the bandgap energy of silicon, resulting in the thermal relaxation of excess photon energy. Nevertheless, the examined fluids exhibit a notable degree of radiation transmission within the 400–700 nm range, as demonstrated in Table 3, thus signifying their potential efficacy in capturing solar energy from this specific region.

2.2. Design of the CPVT

The CPVT collector is north–south orientated and rotates along the east–west horizontal axis to increase the overall optical performance and reduce variation in energy delivery during the day [27].
Three parameters are important in the design of the LFC (see Figure 3): location ( M n ), tilt angle ( δ n ), and distance of adjacent mirrors ( S n ). These may be obtained using elementary geometrical optics by using the following formulas [28]:
δ n = a t a n M n f c r 2
S n = W m i r r o r 2 × [ sin δ n + sin δ n 1 × tan 2 δ n + cos δ n + c o s ( δ n 1 ) ]
M n = M n 1 + S n
where f c r is the focal length of the receiver, W m i r r o r the width of the primary mirrors, and the subscript n is the number of the primary mirror.

2.3. Optical and Thermal Modelling

2.3.1. Optical Efficiency

The following expression is used to estimate the optical efficiency of the LFC [29]:
η o p t = η o p t , n o m K T ( θ T ) K L ( θ L )
where η o p t , n o m is the nominal optical efficiency measured at solar noon, K T ( θ T ) is the transversal incidence angle modifier, θ T is the transversal incidence angle in degree, K L ( θ L ) is the longitudinal incidence angle modifier, and θ L is the longitudinal incidence angle in degree.
For a collector aligned along the north–south axis, the transversal and longitudinal angles are calculated as follows [27]:
θ T = t a n 1 ( sin A z × tan Z )
θ L = t a n 1 ( cos A z × tan Z )
where A z and Z are the Azimuth and Zenith angles, respectively.
In addition, the transversal and the longitudinal incidence angle modifiers are calculated using the following expressions, respectively [27]:
K T θ T = cos θ T 2 W f i e l d 4 f c r + f c r 2 + ( W f i e l d 4 ) 2 × s i n ( θ T 2 )
K L ( θ L ) = c o s ( θ L ) f c r L r × 1 + W f i e l d 4 f c r 2 × s i n ( θ L )
where L r is the receiver length, and W f i e l d is the field width.

2.3.2. Heat Transfer Model

To examine the heat flow inside the receiver, a heat transfer model is developed. The flowchart outlining the model’s structure and methodology can be found in Appendix A. The model takes into account the following set of assumptions:
  • Steady state heat transfer model
  • Thin PV module
  • Side walls of the receiver are adiabatic
  • Uniform temperature distribution
  • The nanofluid flow is uniform
Furthermore, considering the phenomenon of self-absorption exhibited by the fluid and the similarity in emissivity between the fluid and the glass window, it is assumed that the heat radiation losses can be directly attributed to the glass window.

Heat Transfer in the Receiver

According to Newton’s law of cooling, the convection heat transfer from the absorber’s interior surface to the HTF is:
Q c o n v , r f l = h f l × A r , i n × ( T r , i n T r , f l , m e a n )
where A r , i n is the inside surface of the thermal receiver, T r , i n is the temperature of the inside surface of the thermal receiver, T r , f l , m e a n is the mean temperature of the fluid in the receiver, and h f l is the fluid heat transfer coefficient defined in the following way:
h f l = N u f l × k f l D h r
where N u f l is the fluid Nusselt number, k f l is the fluid thermal conductivity, and D h r is the hydraulic diameter of the receiver. For the case of laminar flow, the Nusselt number is considered constant:
N u f l _ l a m i n a r = 4.36
For the case of turbulent flow, the following Nusselt number correlation is used:
N u f l _ t u r b u l e n t = 0.023 × R e f l 3 / 4 × P r f l 0.3
where R e f l   is the Reynolds number and P r f l is the fluid Prandtl number.
Conduction through the front and rear glass of the receiver can be represented as follows:
Q c o n d , r = k g l a s s × A r , g l a s s × ( T r , i n T r , o u t ) t g l a s s
where k g l a s s is the glass thermal conductivity, A r , g l a s s is the area of the front and rear glass, T r , o u t is the temperature of the outside surface of the thermal receiver and t g l a s s the glass thickness.
The rear glass surface of the receiver is connected to the cooling channel, and the walls are insulated, so convective heat exchange with the ambient air is only considered on the front glass surface of the receiver. Consequently, following Newton’s law of cooling, the convection heat transfer from the receiver’s outside surface to the atmosphere is:
Q c o n v , r a m b = h a i r × A r , o u t , f r o n t × ( T r , o u t T a m b )
where A r , o u t , f r o n t is the front glass surface of the thermal receiver, T a m b is the ambient temperature during sun hours, and h a i r is the air heat transfer coefficient defined in the following way:
h a i r = N u a i r × k a i r D h r
where N u a i r is the air Nusselt number, and k a i r is the air thermal conductivity. For laminar flow over a flat plate, the Nusselt number is expressed as follows:
N u a i r _ l a m i n a r = 0.664 × R e a i r 0.5 × P r a i r 1 / 3
For turbulent flow over a flat plate, the Nusselt number is expressed as follows:
N u a i r _ t u r b u l e n t = 0.037 × R e a i r 0.8 × P r a i r 1 / 3
Because the receiver’s front glass surface is in contact with the ambient air and the sidewalls are insulated, convective heat exchange with the cooling channel is only evaluated on the receiver’s rear glass surface. As a result, according to Newton’s law of cooling, the convection heat transfer from the outer surface of the receiver to the cooling channel is:
Q c o n v , r c h = h f l × A r , o u t , r e a r × ( T r , o u t T c h , f l , m e a n )
where A r , i n is the inside surface of the thermal receiver, T r , i n is the temperature of the inside surface of the thermal receiver, T c h , f l , m e a n is the mean temperature of the fluid in the cooling channel and h f l the fluid heat transfer coefficient.
According to the Stefan–Boltzmann law of radiation, the radiation heat transfer from the external surface of the receiver to the atmosphere is:
Q r a d , r a t m = σ × ε g l a s s × A r , o u t , f r o n t × ( T r , o u t 4 T s k y 4 )
where σ is the Stefan-Boltzmann constant ( 5.67 × 10 8   W m 2 K 4 ) , ε g l a s s is the glass emissivity, and T s k y is the sky temperature estimated using the following expression [30]:
T s k y = 0.0522 × T a m b 1.5
Radiation heat exchange with the PV panel is only evaluated on the receiver’s rear glass surface. As a result, the expression that estimates the radiation heat transfer between two parallel plates is used:
Q r a d , r P V = σ × A r , o u t , r e a r × ( T r , o u t 4 T P V 4 ) 1 ε g l a s s + 1 ε P V 1
where T P V is the temperature of the PV panel, and ε P V is its emissivity.

Heat Transfer in the PV Panel

The solar radiation on the rear surface of PV cell follows the Stefan–Boltzmann law of radiation:
Q r a d , P V a t m = σ × ε P V × A P V , r e a r × ( T P V 4 T s k y 4 )
where A P V , r e a r is the area of the PV panel rear surface.
Newton’s law of cooling states that the convective heat transfer from the PV panel to the cooling channel is:
Q c o n v , P V c h = h f l × A P V , f r o n t × ( T P V T c h , f l , m e a n )
where A P V , f r o n t is the front surface of the PV panel, and h f l is the fluid heat transfer coefficient.
The convective heat transfer from the PV panel to the ambient air is:
Q c o n v , P V a m b = h a i r × A P V , r e a r × ( T P V T a m b )
where h a i r is the air heat transfer coefficient.

Power, Efficiency and Energy

The efficiency of bifacial crystalline silicon PV cells can be estimated using the following expression, which considers a temperature coefficient of −0.45%/°C:
η P V = η P V , n o m [ 1 ( 0.0045 × ( T P V T P V , r e f ) ) ]
where η P V , n o m is the nominal efficiency of the PV panel at the reference temperature T P V , r e f . The electric energy produced by the PV panel can be calculated using the following equation:
Q u , P V , e l = D N I × A a p × η o p t × ( 1 f o p t ) × t r f l , 700 1100 n m × η P V + G H I × A P V , r e a r × η P V
where D N I is the direct normal irradiance, f o p t is the fraction of optical loss in the receiver, t r f l , 700 1100 n m is the average spectral transmittance of the fluid filter between the 700–1100 nm spectral window, A a p is the aperture area of the primary mirrors, and GHI is the global horizontal irradiance.
The power absorbed by the receiver is calculated using the following equation:
Q a b s , r = D N I × A a p × η o p t × f r
where f r is the fraction of radiation absorbed by the receiver.
The useful thermal power absorbed by the fluid in the receiver is:
Q u , f l , t h , r = m ˙ f l × C p f l × ( T r , f l , o u t T c h , f l , i n )
where C p f l is the specific heat capacity of the fluid, m ˙ f l is the fluid mass flow rate, T r , f l , o u t is the temperature of the fluid in the outlet of the receiver, and T c h , f l , i n is the temperature of the fluid in the inlet of the cooling channel. As a result, the thermal efficiency of the receiver may be calculated as follows:
η r = Q u , f l , t h , r Q a b s , r
The organic Rankine cycle (ORC) has received a great deal of attention as a well-accepted technology because it can make effective use of low-grade thermal energy sources, such as solar thermal [31]. In the present study, one of the scenarios examined considers that the thermal energy stored in the fluid is converted to electrical energy through an ORC. Therefore, the overall electrical efficiency of the system is defined as follows:
η t o t a l , e l = η P V + [ η r × η h e a t C a r n o t × 1 T a m b T r , f l , o u t ]
where η h e a t C a r n o t is the thermodynamic efficiency of heat engine to Carnot efficiency [32].
Lastly, the net solar-to-electric efficiency of the system, which incorporates the total incident solar power as a common denominator, is presented as:
η N S E = Q u , P V , e l + Q u , f l , t h , r × η h e a t C a r n o t × 1 T a m b T r , f l , o u t D N I × A a p

3. Results and Discussion

3.1. Ray Tracing and Optimum Geometric Concentration Ratio of the CPVT

A ray-tracing simulation of the LFC has been carried out using Tonatiuh software to assess the design of the proposed CPVT (see Figure 4). Figure 5 illustrates the heat flux distribution on the front glass of the receiver. As can be noticed, the flux distribution corresponds to that of a typical LFC.
A parametric study is carried out to determine the optimum concentration ratio for the CPVT collector. Average weather data for Tabuk was employed for this optimization process. Figure 6 illustrates the variation of the overall electric efficiency and the temperature of the PV module as a function of the geometric concentration ratio (GCR) of the CPVT collector. As can be seen, the optimum GCR that maximize the overall electric efficiency of the CPVT collector is about 20. At this GCR, the temperature of the PV module is less than 85 degree C (the maximum operating temperature of crystallin PV cells). Therefore, this value is used in this study.

3.2. Advantages of the Proposed Receiver Design

To highlight the advantages of the proposed receiver design, an annual performance comparison between a receiver with cooling the PV module (denoted C in this paper) and a receiver without cooling the PV module (denoted NC in this paper) has been conducted. Six different locations and two different HTFs—water (denoted W in this paper) and water with ZnO nanoparticles at 0.01 wt% concentration (denoted W+ZnO in this paper)—are considered.
As can be noticed in Figure 7, the average temperature of the PV module is lower for the case with cooling than for the case without cooling (more than 10 °C difference). This results in higher efficiency of the PV cells. The addition of ZnO nanoparticles to water improves the heat transfer, which further reduces the temperature of the PV module; thus, high electric efficiency is achieved.
Overall, the performance of the CPVT collector at Tabuk is better than other locations because of the low ambient temperature and the high solar irradiance (see Table 4).

3.3. Thermal Performance of the CPVT

The yearly energy production of the CPVT at different locations is illustrated in Figure 8. The PV electrical energy output and thermal energy output are higher when the CPVT is installed in Tabuk. When ZnO nanoparticles are added to water, the thermal energy increases in all the considered locations, but the electrical energy provided by the PV panel slightly decreases. Although the drop in electrical energy is small compared to the gain in thermal energy, if all the thermal energy is converted to electricity, less energy is obtained compared with the case of using water.
The monthly energy production in Tabuk is shown in Figure 9. Summer months always have the highest energy output. The amount of energy produced varies dramatically throughout the year, with the summer period producing twice as much electrical energy and up to four times more thermal energy compared to winter months. This is because solar radiation is higher in the summer than in the winter, and the optical efficiency of the system is also higher.
Figure 9 also highlights that, when ZnO nanoparticles are added to water, the thermal energy output increases. This is because the working fluid absorbs 6.2% more solar radiation, as shown in Table 5, due to the variation in the spectral transmittance property, when ZnO is added to the water. On the other hand, when the working fluid contains nanoparticles, the electrical production is slightly lower. The scientific reason behind it is that the nanofluid absorbs more solar radiation at wavelengths between 700 nm and 1100 nm; these wavelengths are used to generate energy through the photovoltaic effect for silicon-based PV panels. In this spectral window, water alone has an average spectral transmittance of 88.1% [2], which drops to 79.8% [24] when ZnO nanoparticles are added.
A monthly analysis of the efficiency for the PV panel and the receiver in Tabuk is illustrated in Figure 10. It is notable that the variation in the efficiency for the PV panel is not significant during the year. The PV efficiency is slightly better in the winter period compared to the summer period due to lower ambient temperature. In contrast, the net solar-to-electric and thermal receiver efficiencies follow the same trend as thermal energy generation, being higher in summer than in winter.

3.4. Economic Analysis

3.4.1. CAPEX of the CPVT

The estimation of the cost of the CPVT is based on the Hyperlight Energy project. Table 6 highlights the specific costs as well as the total CAPEX of the CPVT collector considered in this study [34,35]. The cost of bifacial photovoltaic panels has been evaluated based on an average of projects completed in the last few years following the IRENA report [36]. Estimation showed that the CPVT collector with nano-particles costs $8550, while a CPVT collector with water as a HTF costs $200 less.

3.4.2. The LCOE and LCOH for a 20 MW CPVT Plant

The LCOE and LCOH represent the average of the net present cost of energy production for the plant over its lifetime. The IRENA methodology is used in this paper [36]:
L C O E   o r   L C O H = t = 1 n I t + O M t 1 + r t t = 1 n E t 1 + r t
where I t are the investment expenditures in the year t , O M t are the operations and maintenance expenditures in the year t , E t is the energy generation in the year t , r is the discount rate, and n is the lifetime of the system.
The financial parameters that were used by IRENA are adapted in this study. These include: a 10% discount rate, a lifetime of 25 years, and 3% of the CAPEX were considered for the maintenance and operation costs.
Two operation strategies are considered, namely electricity production strategy and heat/electricity production strategy. In the former, the heat absorbed by the HTF is converted into electricity using an ORC cycle. The electricity may either be fed directly into the power grid or used to power an industry or households in remote places. A 20 MWh thermal energy storage (TES) consisting of a water tank is considered to store thermal energy with 82% round-trip-efficiency. In a heat/electricity production strategy, it is assumed that there is an industry nearby requiring hot water at 95 °C. This is highly feasible given that the industrial sector with low-temperature heat processes accounts for 7.1% of world energy consumption [38].
To calculate the LCOH and the LCOE of a large-scale CVPT power plant, it is important to estimate the CAPEX. Considering that the LFC uses an average of 70% of the total land, the cost of the land required was estimated at $5/m2. The cost of a TES system with a capacity of 20,000 kWh has been evaluated taking data from the European Association of Storage of Energy [39], and adding an extra cost of 15$/kWh for each system owing to the cost of building work and additional materials, like pipes. For the situation when the thermal energy is converted into electricity, the cost of the plant required has been determined using a Pratt and Whitney ORC catalogue [40]. Project efforts have also been considered and are estimated at 22.5% of the total cost of the solar plant.
Finally, an additional 5% has been added to the overall expenditures to compensate for any unanticipated occurrences throughout the project’s execution phase. The sum of all the expenditures is the capital expenditure (CAPEX), which is given in Table 7 and Table 8 for both operation strategies.
Table 9 illustrates the values of the LCOE, LCOH, and CAPEX in different locations. Overall, the LCOE is lower for the proposed design (C) compared with the traditional design (NC). This proves the advantages of using the novel design proposed in this study. The addition of ZnO nanoparticles to the water increases the LCOE but decreases the LCOH.
An important finding of this study is that the heat/electricity production strategy is much better than the electricity production strategy. For instance, at Tabuk, the LCOE for our proposed design with water as HTF is 0.2232 USD/kWh when the electricity production strategy is selected. However, it is only 0.0847 USD/kWh when heat/electricity production is selected. Indeed, the LCOE when the heat/electricity production strategy is selected is lower than that of CSP (and the heat is produced as a by-product for free).
The most suitable location for installing CPVT technology in Saudi Arabia is at Tabuk. The analysis shows a LCOE of $0.0847/kWh and a LCOH of $0.0536/kWh when water is used as a spectral filter, and a LCOE of $0.0906/kWh and a LCOH of $0.0462/kWh when ZnO nanoparticles are added.
The LCOE of CPVT systems has been investigated in a limited number of scientific papers. In this study, we compare our results with some previously published studies, as shown in Table 10. Some of these studies have used PTC, which is more expensive than the technology of LFC used in our proposed design. Additionally, Fernandez et al. [41] have employed more expensive materials, such as ITO nanocrystals and Au nanoparticles, instead of the ZnO nanoparticles used in our system. Furthermore, Ling et al. [10] have used a solid oxide fuel cell (SOFC) instead of an ORC to transform electrical energy into thermal energy, which requires the purchase of methanol, implying additional expenses. Moreover, according to the NREL database [42], the DNI and GHI in Shiraz are 7% and 10% lower, respectively, compared to Tabuk. Taking into account these differences and the novelty of our design for harsh environments, the proposed CPVT system offers a lower LCOE, making it a more cost-effective solution for the given geographical area.

3.4.3. CO2 Emission Analysis

According to the Brown to Green 2019 report, the national emissions in Saudi Arabia associated with electricity generation in 2019 were 0.723 kgCO2-equivalent for each kWh produced [44]. Moreover, according to the Ministry of Spain, emissions from stationary combustion equipment powered by natural gas (such as boilers) are 0.209 kgCO2-equivalent per kWh generated [45]. These two variables are used as the electricity and thermal emissions factors, respectively, to compute, based on the energy production, the emission savings due to the use of CPVT technology. Figure 11 and Figure 12 show the results for one single CPVT collector and for a 20 MW CPVT plant, respectively.
The implementation of the proposed CPVT can cut off annual emissions by 11.2 tCO2eq (Tabuk) per system and from 5675 tCO2eq (Makkah) to 7822 tCO2eq (Tabuk) per 20 MW plant if the heat/electricity production strategy is selected. However, if the electricity production strategy is selected, this technology can save a total of 8.5 tCO2eq per collector and 5968 tCO2eq per 20 MW CPVT plant annually.

3.4.4. Battery vs. TES

A large increase in battery production is expected in the coming years. However, the materials needed for their production are limited, so it is critical to look for other ways of storing energy. For this reason, a thermal energy storage (TES) system has been considered in this study, and its comparison in economic terms is shown below.
The cost of lithium-ion battery packs has increased for the first time since 2010 because of rising inflation and prices of raw materials and battery components, reaching an average of 151$/kWh in 2022 [46]. Moreover, in the most optimistic scenario, lithium-ion batteries have a lifetime of 15 years, so they would have to be replaced at least once to match the lifespan of the solar power system [47,48]. On the other hand, according to the European Association for Energy Storage, the price for a hot water storage tank is 15$/kWh with an average 30-year working life.
For the battery scenario, the heat is first converted to electricity using an ORC with an efficiency on the order of 10% for the working temperatures considered in this study (see Equation (30)). Afterwards, electricity is stored into a 20 MWh lithium-ion battery which nowadays reach up DC round-trip efficiency values as high as 95% [49]. Alternatively, the MWh TES system with 82% round-trip efficiency previously described in Section 3.4.2 can be used, with no need to convert the heat into electricity for storage. Table 11 compares the two suggested storage systems installed in Tabuk based on the LCOE, LCOH, and CAPEX for a 20 MW large-scale plant, in which an extra cost of 15$/kWh has been estimated for each system owing to the cost of building work and additional materials like cables or pipes.
The use of batteries compared to a TES system based on a water tank represents an additional increase of 5.74 million dollars, considering that the batteries will need to be replaced once during the lifetime of the solar plant. This represents an even greater increase in the CAPEX, which is reflected in the LCOE and LCOH costs, which increase up to 56% and 84% respectively.

3.5. Future Work

In terms of future work, there are several areas of research that could be explored to further enhance the performance of concentrated photovoltaic-thermal (CPVT) systems. Building a prototype to experimentally validate the theoretical outcomes of this study would be a valuable next step. This would provide a more accurate representation of the real-world performance of the CPVT system.
Another possible avenue of research would be to include the capability of joining multiple CPVT systems in series in the mathematical model presented. This would allow for a higher heat transfer fluid output temperature, expanding the range of potential applications beyond just low-temperature heating.
Furthermore, it may be useful to explore the recommendations of An et al. [50] and investigate the effectiveness of using two reflectors at the sides of the solar receiver to minimize the effect of imprecise sun tracking and receiver installation. This could potentially reduce the current optical losses of the proposed solar receiver (12.2%) and improve the overall performance of the CPVT system.
Overall, this study contributes to the body of knowledge on CPVT systems and provides valuable insights for future research in this area. There is still much to be explored in terms of optimizing the performance and applicability of CPVT systems, and the proposed future research directions could help to advance this field.

4. Conclusions

The design and performance evaluation of a novel CPVT with spectral beam splitting technology, a cooling channel, and nanofluid is presented in this paper. A raytracing simulation tool is used to assess optical performance of the proposed CPVT collector, while an optical-thermal model is used to estimate the performance of the system.
The investigation revealed that using fluids as a filter in the CPVT collector has numerous benefits, including a low operating temperature for the PV cells and a high energy output. By adding the cooling channel and ZnO nanoparticles, it is found that a significant decrease in the average and maximum temperature of the PV panel is achieved, where they are lowered by 16.6 °C and 43.4 °C, respectively. This allows conventional silicon photovoltaic panels which have a maximum operating temperature of 85 °C [51] to be used. Therefore, without these design improvements, we would have to resort to special high-temperature PV panels, which are in very limited supply from manufacturers and present lower efficiencies due to the increased temperature.
The calculated yearly average values of the efficiencies, with the addition of the cooling channel and nanofluid, are, for Tabuk, 19.74% for the photovoltaic panel, 35.65% for the thermal collector, and 22.65% for the total conversion to electricity.
The economic assessment showed that the CPVT system has great possibilities to lead the Saudi renewable energy production in the coming years. Under a heat/electricity production strategy, a LCOE of $0.0847/kWh and a LCOH of $0.0536/kWh when only water is used a HTF are obtained. At the same time, a LCOE of $0.0906/kWh and a LCOH of $0.0462/kWh are obtained when ZnO particles are added. The analysis showed that, due to the low performance and high costs of converting thermal energy into electricity, the CPVT technology is less competitive when the electricity production strategy is selected. The results showed an LCOE of $0.2232/kWh with water only and $0.2442/kWh with the addition of ZnO nanoparticles.
Furthermore, after comparing battery energy storage against a TES system, a large increase in the CAPEX was observed if batteries are used, reflected in the LCOE and LCOH costs, which increase up to 56% and 84% (compared with the case of TES), respectively. Thus, a CPVT plant with TES operating under a heat/electricity production strategy is better than a CPVT plant with battery operating under an electricity production strategy.
The study showed that a 20 MW CPVT plant cuts CO2-equivalent emissions up to 7822 tonnes every year under Saudi Arabian conditions. Another benefit in terms of sustainability is the ease of recycling the proposed CPVT technology, taking up less space, and requiring less photovoltaic material to capture the same sunlight as non-concentrating PV modules. Thus, the process is less dependent on the silicon supply chain.
Regarding the practicality of the technology presented, it has been demonstrated that the system is technically feasible through a series of rigorous computations and simulations. Specifically, the results indicate that the proposed design offers significant advantages when operating in harsh environments when compared to traditional designs. Additionally, an economic study was conducted which revealed that the system can be constructed at a relatively low cost in comparison to previous publications, resulting in an improved levelized cost of energy (LCOE) for this technology. Overall, these findings support the practicability of the technology, and suggest that it has the potential to be a viable cost-effective solution for a range of real-world energy applications.

Author Contributions

Conceptualization, C.L., O.B. and B.D.; Methodology, C.L., O.B. and B.D.; Software, C.L. and O.B.; Formal analysis, C.L.; Investigation, C.L., O.B. and B.D.; Resources, B.D.; Writing—original draft, C.L.; Writing—review & editing, O.B. and B.D.; Visualization, C.L.; Supervision, O.B. and B.D.; Project administration, B.D.; Funding acquisition, B.D. All authors have read and agreed to the published version of the manuscript.

Funding

Project is supported by Prof Bassam Dally KAUST baseline fund.

Data Availability Statement

Not applicable.

Acknowledgments

The authors gratefully acknowledge the funding of the King Abdullah University of Science and Technology.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

Abbreviations:
PVPhotovoltaic
CPVTConcentrating solar photovoltaic thermal
LCOELevelized cost of electricity
LCOHLevelized cost of heat
SBSSpectrum beam splitting
HTFHeat transfer fluid
PTCParabolic trough collector
LFCLinear Fresnel collector
TEThermoelectric generator
NRELNational Renewable Energy Laboratory
DNIDirect normal irradiance
GHIGlobal horizontal irradiance
ORCOrganic Rankine cycle
GCRGeometric concentration ratio
GMTGreenwich mean time
CSPConcentrated solar power
TESThermal energy storage
CCPVT with cooling channel
NCCPVT without cooling channel
CAPEXCapital expenditure
DRDiffuse irradiance
LSTLocal solar time
LTLocal time
EoTEquation of time
TCNet time correction factor
LSTMLocal standard time meridian
HRAHour angle
Nomenclature:
W Width m
H Height m
M Location of the mirrors m
S Distance of adjacent mirrors m
f c Focal length m
K Incidence angle modifier
A z Azimuth angle °
Z Zenith angle °
L Length m
A Area m 2
Q Heat flux W
T Temperature K
h Heat transfer coefficient W m 2 K
N u Nusselt number
k Thermal conductivity W m K
D h Hydraulic diameter m
R e Reynolds number
P r Prandtl number
t Thickness m
t r Average spectral transmittance
f Fraction of radiation absorbed
m ˙ Fluid mass flow rate k g s
C p Specific heat capacity J k g K
w t Mass fraction
d Day number of the year, ranging from 1 to 365
VVelocity m s
Greek letters:
δ Tilt angle °
η Efficiency
θ Incidence angle °
σStefan–Boltzmann constant W m 2 K 4
ε Emissivity
α Elevation angle °
φ Local latitude °
β Declination of the sun °
Subscripts:
nNumber of primary mirror
chChannel
rReceiver
optOptical
TTransversal
LLongitudinal
flFluid
ambAmbient
convConvection
condConduction
radRadiation
refReference
apAperture
absAbsorbed
atmAtmosphere
thThermal
elElectrical
NSENet solar-to-electric

Appendix A

A mathematical code is developed in Matlab to simulate the performance of the CPVT. The set of equations presented in the previous sections are solved using an iteration process. The flowchart of the model is illustrated in Figure A1. The output HTF temperature, location information, geometrical parameters, and fluid characteristics are used as input. The model uses the direct normal irradiance (DNI), global horizontal irradiance (GHI), wind speed, and ambient temperature from the NREL database [42]. The optical efficiency of the system is calculated using the location, solar angles, and geometrical data. Next, the variables to be determined are set up, and an iterative procedure based on energy balance is used. This process ends when all the energy and mass balance equations are satisfied. Lastly, the power and energy performance are determined.
Figure A1. Flowchart of calculation model.
Figure A1. Flowchart of calculation model.
Energies 16 05392 g0a1

References

  1. Han, X.; Tu, L.; Sun, Y. A spectrally splitting concentrating PV/T system using combined absorption optical filter and linear Fresnel reflector concentrator. Sol. Energy 2021, 223, 168–181. [Google Scholar] [CrossRef]
  2. Mojiri, A.; Stanley, C.; Rosengarten, G. Spectrally Splitting Hybrid Photovoltaic/thermal Receiver Design for a Linear Concentrator. Energy Procedia 2014, 48, 618–627. [Google Scholar] [CrossRef] [Green Version]
  3. Ramírez Quiroz, C.O.; Levchuk, I.; Bronnbauer, C.; Salvador, M.; Forberich, K.; Heumüller, T.; Hou, Y.; Schweizer, P.; Spiecker, E.; Brabec, C.J. Pushing efficiency limits for semitransparent perovskite solar cells. J. Mater. Chem. A 2015, 3, 24071–24081. [Google Scholar] [CrossRef] [Green Version]
  4. Jarimi, H.; Lv, Q.; Ramadan, O.; Zhang, S.; Riffat, S. Design, mathematical modelling and experimental investigation of vacuum insulated semi-transparent thin-film photovoltaic (PV) glazing. J. Build. Eng. 2020, 31, 101430. [Google Scholar] [CrossRef]
  5. Mujahid, M.; Chen, C.; Zhang, J.; Li, C.; Duan, Y. Recent advances in semitransparent perovskite solar cells. InfoMat 2021, 3, 101–124. [Google Scholar] [CrossRef]
  6. Joshi, S.S.; Dhoble, A.S.; Jiwanapurkar, P.R. Investigations of Different Liquid Based Spectrum Beam Splitters for Combined Solar Photovoltaic Thermal Systems. J. Sol. Energy Eng. 2016, 138, 021003. [Google Scholar] [CrossRef]
  7. Soule, D.E. Hybrid Solar Energy Generating System. 1987. Available online: https://www.osti.gov/biblio/5984591 (accessed on 28 May 2022).
  8. Zhang, J.J.; Qu, Z.G.; Zhang, J.F. MCRT-FDTD investigation of the solar-plasmonic-electrical conversion for uniform irradiation in a spectral splitting CPVT system. Appl. Energy 2022, 315, 119054. [Google Scholar] [CrossRef]
  9. Daneshazarian, R.; Cuce, E.; Cuce, P.M.; Sher, F. Concentrating photovoltaic thermal (CPVT) collectors and systems: Theory, performance assessment and applications. Renew. Sustain. Energy Rev. 2018, 81, 473–492. [Google Scholar] [CrossRef]
  10. Ling, Y.; Li, W.; Jin, J.; Yu, Y.; Hao, Y.; Jin, H. A spectral-splitting photovoltaic-thermochemical system for energy storage and solar power generation. Appl. Energy 2020, 260, 113631. [Google Scholar] [CrossRef]
  11. Liew, N.J.Y.; Yu, Z.; Holman, Z.; Lee, H.-J. Application of spectral beam splitting using Wavelength-Selective filters for Photovoltaic/Concentrated solar power hybrid plants. Appl. Therm. Eng. 2022, 201, 117823. [Google Scholar] [CrossRef]
  12. Ju, X.; Xu, C.; Han, X.; Du, X.; Wei, G.; Yang, Y. A review of the concentrated photovoltaic/thermal (CPVT) hybrid solar systems based on the spectral beam splitting technology. Appl. Energy 2017, 187, 534–563. [Google Scholar] [CrossRef]
  13. Sabry, M.; Gottschalg, R.; Betts, T.R.; Shaltout, M.A.M.; Hassan, A.F.; El-Nicklawy, M.M.; Infield, D.G. Optical filtering of solar radiation to increase performance of concentrator systems. In Proceedings of the Conference Record of the IEEE Photovoltaic Specialists Conference, New Orleans, LA, USA, 19–24 May 2002; pp. 1588–1591. Available online: https://www.scopus.com/inward/record.uri?eid=2-s2.0-0036957820&partnerID=40&md5=9aae59a3fe7add1b1f8cc02576b1774e (accessed on 29 May 2022).
  14. Hangweirer, M.; Höller, R.; Schneider, H. Design and analysis of a novel concentrated photovoltaic–thermal receiver concept. Jpn. J. Appl. Phys. 2015, 54, 08KE01. [Google Scholar] [CrossRef]
  15. Meraje, W.C.; Huang, C.-C.; Barman, J.; Huang, C.-Y.; Kuo, C.-F.J. Design and experimental study of a Fresnel lens-based concentrated photovoltaic thermal system integrated with nanofluid spectral splitter. Energy Convers. Manag. 2022, 258, 115455. [Google Scholar] [CrossRef]
  16. Barthwal, M.; Rakshit, D. Holistic opto-thermo-electrical analysis of a novel spectral beam splitting-based concentrating photovoltaic thermal system. J. Clean. Prod. 2022, 379, 134545. [Google Scholar] [CrossRef]
  17. Wang, G.; Ge, Z.; Lin, J. Design and performance analysis of a novel solar photovoltaic/thermal system using compact linear Fresnel reflector and nanofluids beam splitting device. Case Stud. Therm. Eng. 2022, 35, 102167. [Google Scholar] [CrossRef]
  18. Su, B.; Wang, H.; Zhang, X.; He, H.; Zheng, J. Using photovoltaic thermal technology to enhance biomethane generation via biogas upgrading in anaerobic digestion. Energy Convers. Manag. 2021, 235, 113965. [Google Scholar] [CrossRef]
  19. Su, B.; Han, W.; Qu, W.; Liu, C.; Jin, H. A new hybrid photovoltaic/thermal and liquid desiccant system for trigeneration application. Appl. Energy 2018, 226, 808–818. [Google Scholar] [CrossRef]
  20. Rodrigues Fernandes, M.; Schaefer, L.A. Long-term environmental impacts of a small-scale spectral filtering concentrated photovoltaic-thermal system. Energy Convers. Manag. 2019, 184, 350–361. [Google Scholar] [CrossRef]
  21. Anand, B.; Shankar, R.; Murugavelh, S.; Rivera, W.; Midhun Prasad, K.; Nagarajan, R. A review on solar photovoltaic thermal integrated desalination technologies. Renew. Sustain. Energy Rev. 2021, 141, 110787. [Google Scholar] [CrossRef]
  22. George, M.; Pandey, A.K.; Abd Rahim, N.; Tyagi, V.V.; Shahabuddin, S.; Saidur, R. Concentrated photovoltaic thermal systems: A component-by-component view on the developments in the design, heat transfer medium and applications. Energy Convers. Manag. 2019, 186, 15–41. [Google Scholar] [CrossRef] [Green Version]
  23. National Institute of Standards and Technology. Thermophysical Properties of Fluid Systems. Available online: https://webbook.nist.gov/chemistry/fluid/ (accessed on 14 May 2022).
  24. Han, X.; Lu, L.; Yan, S.; Yang, X.; Tian, R.; Zhao, X. Stability, Thermal Conductivity and Photothermal Conversion Performance of Water-Based ZnO Nanofluids. J. Therm. Sci. 2021, 30, 1581–1595. [Google Scholar] [CrossRef]
  25. Farzanehnia, A.; Sardarabadi, M. Exergy in Photovoltaic/Thermal Nanofluid-Based Collector Systems. Exergy and Its Application: Toward Green Energy Production and Sustainable Environment; IntechOpen: Rijeka, Croatia, 2019. Available online: https://www.intechopen.com/chapters/66854 (accessed on 3 June 2022).
  26. Hassani, S.; Taylor, R.A.; Mekhilef, S.; Saidur, R. A cascade nanofluid-based PV/T system with optimized optical and thermal properties. Energy 2016, 112, 963–975. [Google Scholar] [CrossRef]
  27. Rungasamy, A.E.; Craig, K.J.; Meyer, J.P. A review of linear Fresnel primary optical design methodologies. Sol. Energy 2021, 224, 833–854. [Google Scholar] [CrossRef]
  28. Zhu, Y.; Shi, J.; Li, Y.; Wang, L.; Huang, Q.; Xu, G. Design and thermal performances of a scalable linear Fresnel reflector solar system. Energy Convers. Manag. 2017, 146, 174–181. [Google Scholar] [CrossRef]
  29. Said, Z.; Ghodbane, M.; Hachicha, A.A.; Boumeddane, B. Optical performance assessment of a small experimental prototype of linear Fresnel reflector. Case Stud. Therm. Eng. 2019, 16, 100541. [Google Scholar] [CrossRef]
  30. Radwan, A.; Emam, M.; Ahmed, M. Chapter 2.15—Comparative Study of Active and Passive Cooling Techniques for Concentrated Photovoltaic Systems. In Exergetic, Energetic and Environmental Dimensions; Dincer, I., Colpan, C.O., Kizilkan, O., Eds.; Academic Press: Cambridge, MA, USA, 2018; pp. 475–505. [Google Scholar]
  31. Wang, H.; Li, H.; Wang, L.; Bu, X. Thermodynamic Analysis of Organic Rankine Cycle with Hydrofluoroethers as Working Fluids. Energy Procedia 2017, 105, 1889–1894. [Google Scholar] [CrossRef]
  32. An, W.; Li, J.; Ni, J.; Taylor, R.A.; Zhu, T. Analysis of a temperature dependent optical window for nanofluid-based spectral splitting in PV/T power generation applications. Energy Convers. Manag. 2017, 151, 23–31. [Google Scholar] [CrossRef]
  33. Huaxu, L.; Fuqiang, W.; Dong, L.; Jie, Z.; Jianyu, T. Optical properties and transmittances of ZnO-containing nanofluids in spectral splitting photovoltaic/thermal systems. Int. J. Heat Mass Transf. 2019, 128, 668–678. [Google Scholar] [CrossRef]
  34. Kincaid, N.; Mungas, G.; Kramer, N.; Wagner, M.; Zhu, G. An optical performance comparison of three concentrating solar power collector designs in linear Fresnel, parabolic trough, and central receiver. Appl. Energy 2018, 231, 1109–1121. [Google Scholar] [CrossRef]
  35. U.S. Department of Energy. PROJECT PROFILE: Hyperlight Energy (CSP: COLLECTS). Available online: https://www.energy.gov/eere/solar/project-profile-hyperlight-energy-csp-collects (accessed on 30 May 2022).
  36. IRENA. Renewable Power Generation Costs in 2020; International Renewable Energy Agency: Abu Dhabi, UAE, 2021; Available online: https://www.irena.org/publications/2021/Jun/Renewable-Power-Costs-in-2020 (accessed on 2 May 2022).
  37. Liang, H.; Wang, F.; Yang, L.; Cheng, Z.; Shuai, Y.; Tan, H. Progress in full spectrum solar energy utilization by spectral beam splitting hybrid PV/T system. Renew. Sustain. Energy Rev. 2021, 141, 110785. [Google Scholar] [CrossRef]
  38. Abarca, V. Energía Solar Térmica en la Industria: Potencial y Aplicaciones. Fundación de la Energía de la Comunidad de Madrid. Available online: https://www.fenercom.com/wp-content/uploads/2020/06/1_Energia_solar_termica_en_la_industria_potencial_y_aplicaciones_ASIT_fenercom-2020.pdf (accessed on 3 December 2022).
  39. Thermal Hot Water Storage. Available online: https://ease-storage.eu/wp-content/uploads/2016/03/EASE_TD_HotWater.pdf (accessed on 5 December 2022).
  40. Pratt & Whitney. Organic Rankine Cycle Power Generation. Available online: https://aeenewengland.org/images/downloads/AEE_ASHRAE_JointMeeting_March2013/orc_power_generation_pratt_and_whitney.pdf (accessed on 30 May 2022).
  41. Rodrigues Fernandes, M.; Schaefer, L.A. Levelized cost of energy of hybrid concentrating photovoltaic-thermal systems based on nanofluid spectral filtering. Sol. Energy 2021, 227, 126–136. [Google Scholar] [CrossRef]
  42. National Renewable Energy Laboratory (NREL). NSRDB: National Solar Radiation Database. Available online: https://nsrdb.nrel.gov/data-viewer (accessed on 27 November 2022).
  43. Abedanzadeh, A.; Borgheipour, H.; Fakouriyan, S.; Torabi, F. Economic Evaluation of Supplying Commercial Thermal Load by a New CPVT System: A Case Study in Iran. J. Appl. Comput. Mech. 2021. [CrossRef]
  44. Saudi Arabia, Country Profile 2019. Climate Transparency. Available online: https://www.climate-transparency.org/media/saudi-arabia-country-profile-2019#:~:text=This%20country%20profile%20is%20part,several%20studies%20by%20renowned%20institutions (accessed on 20 November 2022).
  45. Ministerio de España. Factores de Emisión. Available online: https://www.miteco.gob.es/ca/cambio-climatico/temas/mitigacion-politicas-y-medidas/factores_emision_tcm34-446710.pdf (accessed on 30 May 2022).
  46. Bloomberg. Lithium-Ion Battery Pack Prices Rise for First Time to an Average of $151/kWh. Available online: https://about.bnef.com/blog/lithium-ion-battery-pack-prices-rise-for-first-time-to-an-average-of-151-kwh (accessed on 10 December 2022).
  47. National Renewable Energy Laboratory. Battery Lifespan. Available online: https://www.nrel.gov/transportation/battery-lifespan.html (accessed on 11 December 2022).
  48. Sunrun. What Is the Life Expectancy of a Solar Battery? Available online: https://www.sunrun.com/go-solar-center/solar-articles/what-is-the-life-expectancy-of-a-solar-battery (accessed on 11 December 2022).
  49. Six Years and 26 Batteries Later. Available online: https://itpau.com.au/knowledge/itps-battery-test-centre/#:~:text=A%20typical%20lithium%20ion%20battery,losses%20for%20lead%2Dacid%20systems (accessed on 11 December 2022).
  50. An, W.; Zhang, J.; Zhu, T.; Gao, N. Investigation on a spectral splitting photovoltaic/thermal hybrid system based on polypyrrole nanofluid: Preliminary test. Renew. Energy 2016, 86, 633–642. [Google Scholar] [CrossRef]
  51. How Hot Do Solar Panels Get and How Does It Affect My System? Solar Reviews. Available online: https://www.solarreviews.com/blog/how-hot-can-solar-panels-get (accessed on 13 December 2022).
Figure 1. Basic design of the CPVT.
Figure 1. Basic design of the CPVT.
Energies 16 05392 g001
Figure 2. (a) CAD design of the receiver (cross-section). (b) U-shape pipe linking the cooling channel to the main channel. (c) Simplified design of the receiver.
Figure 2. (a) CAD design of the receiver (cross-section). (b) U-shape pipe linking the cooling channel to the main channel. (c) Simplified design of the receiver.
Energies 16 05392 g002aEnergies 16 05392 g002b
Figure 3. Schematic of the CPVT collector.
Figure 3. Schematic of the CPVT collector.
Energies 16 05392 g003
Figure 4. Ray tracing simulation with 250 rays using Tonatiuh software.
Figure 4. Ray tracing simulation with 250 rays using Tonatiuh software.
Energies 16 05392 g004
Figure 5. Front glass of the receiver flux distribution, simulation with 1x107 rays using Tonatiuh software.
Figure 5. Front glass of the receiver flux distribution, simulation with 1x107 rays using Tonatiuh software.
Energies 16 05392 g005
Figure 6. Optimum geometric concentration of the CPVT collector.
Figure 6. Optimum geometric concentration of the CPVT collector.
Energies 16 05392 g006
Figure 7. Annual average values of PV module efficiency ( η P V ), total efficiency of CPVT collector ( η t o t a l , e l ), temperature of PV module ( T P V ), and maximum temperature of PV module.
Figure 7. Annual average values of PV module efficiency ( η P V ), total efficiency of CPVT collector ( η t o t a l , e l ), temperature of PV module ( T P V ), and maximum temperature of PV module.
Energies 16 05392 g007
Figure 8. Annual energy production of the CPVT system.
Figure 8. Annual energy production of the CPVT system.
Energies 16 05392 g008
Figure 9. Energy production per month in Tabuk.
Figure 9. Energy production per month in Tabuk.
Energies 16 05392 g009
Figure 10. Energy efficiency per month in Tabuk.
Figure 10. Energy efficiency per month in Tabuk.
Energies 16 05392 g010
Figure 11. kgCO2-equivalent emissions saved in KSA with one CPVT system.
Figure 11. kgCO2-equivalent emissions saved in KSA with one CPVT system.
Energies 16 05392 g011
Figure 12. tCO2-equivalent emissions saved in KSA with a 20 MW plant.
Figure 12. tCO2-equivalent emissions saved in KSA with a 20 MW plant.
Energies 16 05392 g012
Table 1. Design data of the CPVT.
Table 1. Design data of the CPVT.
Component P a r a m e t e r V a l u e U n i t s
Linear Fresnel collectorLength10m
Receiver focal length1.5m
ReceiverHeight of the receiver0.08m
Wide of the mirrors0.1m
Wide of the receiver front surface0.2m
Wide of the receiver back surface0.33m
Table 2. Thermo-physical properties of water and ZnO water-based nanofluid.
Table 2. Thermo-physical properties of water and ZnO water-based nanofluid.
SymbolFluid PropertiesWaterZnO (0.01 wt%)
μ Dynamic viscosity, m P a s 0.47 [23]0.47 *
k Thermal conductivity, W m 1 K 1 0.65 [23]0.86 [24]
C p Specific heat capacity, J k g 1 K 1 4185 [23]4148 [25]
ρ Density, k g m 3 983.7 [23]976.9 [25]
* Due to a lack of data, it is presumed that the dynamic viscosity does not change due to the low concentration of ZnO particles.
Table 3. Average spectral transmittance of water and ZnO water-based nanofluid for specific spectral windows.
Table 3. Average spectral transmittance of water and ZnO water-based nanofluid for specific spectral windows.
Spectral Window200–700 nm700–1100 nm1100–2400 nm
Water [2]97.188.111.7
ZnO (0.01 wt%) [24]64.379.85.1
Table 4. Annual average values of direct normal irradiance (DNI), global horizontal irradiance (GHI), ambient temperature, wind speed, and optical efficiency of CPVT collector ( η o p t ).
Table 4. Annual average values of direct normal irradiance (DNI), global horizontal irradiance (GHI), ambient temperature, wind speed, and optical efficiency of CPVT collector ( η o p t ).
Location D N I   ( W   m 2 ) G H I   ( W   m 2 ) T a m b   ( ° C ) V w i n d   ( m s 1 ) η o p t   ( % )
Tabuk59952427.23.554.2
Riyadh45250130.63.755.3
Dammam44149431.13.655.3
Makkah42748734.64.455.0
Jeddah42649332.84.454.9
Medina51651232.54.154.8
Table 5. Percentage of light power absorbed by component using different fluid filters.
Table 5. Percentage of light power absorbed by component using different fluid filters.
Fluid FilterPV Module (%)Fluid (%)Receiver Walls (%)Thermal Unit (%)Optical Loss (%)
Water [2]31.523.133.256.312.2
Water-ZnO25.329.333.262.512.2
Values for water-ZnO nanofluid calculated using the spectral transmittance presented by Huaxu et al. [33] at 0.01 wt% concentration.
Table 6. CAPEX of a CPVT collector.
Table 6. CAPEX of a CPVT collector.
ComponentValueUnitCost ($)
Site improvement5 [34] $ / m 2 m i r r o r 200
Primary mirrors110 [34] $ / m 2 m i r r o r 4400
Thermal receiver (HTF, piping, etc.)60 [34] $ / m 2 m i r r o r 2400
Bifacial crystalline silicon cells (Total ins.)1.5 [36] $ / W p 1350
ZnO 0.01 wt% (preparation, product)5 [37] $ / m 2 m i r r o r 200
TOTAL CPVT COST 8550
Table 7. CAPEX for electricity production strategy, 20 MW CPVT plant with water + ZnO installed in Tabuk.
Table 7. CAPEX for electricity production strategy, 20 MW CPVT plant with water + ZnO installed in Tabuk.
ComponentValueUnitCost ($)
Design Type NCC
CPVT system8550$/system5,985,000
Land costs ( 60,060   m 2 )5$/m2300,300
Water storage system ( 20,000   k W h )30 [39]$/kWh600,000
Power plant unit (all included)2400 [40]$/kW4,451,0893,966,506
Project efforts (22.5% of solar plant costs)22.5%N/A2,550,6872,441,656
Uncertainties (5% of total costs)5%N/A694,353.8664,673
CAPEX 14,581,43113,958,136
Table 8. CAPEX for electricity + heat production strategy, 20 MW CPVT plant with water + ZnO installed in Tabuk.
Table 8. CAPEX for electricity + heat production strategy, 20 MW CPVT plant with water + ZnO installed in Tabuk.
ComponentValueUnitCost ($)
Design Type NCC
CPVT system8550$/system5,985,000
Land costs ( 60,060   m 2 )5$/m2300,300
Water storage system ( 20,000   k W h )30 [39]$/kWh600,000
Project efforts (22.5% of solar plant costs)22.5%N/A1,549,192
Uncertainties (5% of total costs)5%N/A421,724
CAPEX 8,856,217
Table 9. LCOE, LCOH, and CAPEX of the CPVT power plants for different scenarios. Note, system with cooling is denoted as C and without cooling as NC.
Table 9. LCOE, LCOH, and CAPEX of the CPVT power plants for different scenarios. Note, system with cooling is denoted as C and without cooling as NC.
LocationScenarioFluid Filter L C O E   ( $ / k W h ) L C O H   ( $ / k W h ) C A P E X   ( $ )
Design TypeNCCNCCNCC
TabukAll electricityWater0.24510.2232 13,533,99112,985,307
Water+ZnO0.26700.2442 14,581,43113,958,136
Electricity + heatWater0.09160.08470.04790.05368,676,1428,676,142
Water+ZnO0.09770.09060.04160.04628,856,2178,856,217
RiyadhAll electricityWater0.26970.2495 11,993,16411,555,130
Water+ZnO0.29420.2727 12,803,02712,298,728
Electricity + heatWater0.11090.10420.06690.07628,676,1428,676,142
Water+ZnO0.11930.11200.05740.06518,856,2178,856,217
DammamAll electricityWater0.27420.2539 11,878,37711,448,747
Water+ZnO0.29880.2771 12,663,56412,168,329
Electricity + heatWater0.11380.10680.06870.07848,676,1428,676,142
Water+ZnO0.12230.11480.05900.06708,856,2178,856,217
MakkahAll electricityWater0.27190.2552 11,540,49211,275,797
Water+ZnO0.29760.2794 12,301,94811,985,618
Electricity + heatWater0.11470.10850.07360.08088,676,1428,676,142
Water+ZnO0.12380.11700.06250.06858,856,2178,856,217
JeddahAll electricityWater0.26660.2500 11,647,52711,338,953
Water+ZnO0.29270.2744 12,439,09612,074,460
Electricity + heatWater0.11160.10570.07410.08248,676,1428,676,142
Water+ZnO0.12070.11420.06270.06968,856,2178,856,217
MedinaAll electricityWater0.25420.2359 12,401,73712,058,328
Water+ZnO0.27790.2585 13,295,35612,895,731
Electricity + heatWater0.10140.09510.05830.06388,676,1428,676,142
Water+ZnO0.10890.10210.05000.05468,856,2178,856,217
Table 10. Comparison of the LCOE.
Table 10. Comparison of the LCOE.
ReferencesLocationTechnology L C O E   ( $ / k W h )
Ling et al. [10]Not availableLFC with solid filter0.2000
Rodrigeus et al. [41]Blythe, CaliforniaPTC with nanofluid filter0.1783
Abedanzadeh et al. [43]Shiraz, IranPTC with pieces of mirrors0.1293
Present studyTabuk, Saudi ArabiaLFC with fluid filter0.0847
Present studyTabuk, Saudi ArabiaLFC with nanofluid filter0.0906
Table 11. LCOE, LCOH, and CAPEX of the CPVT with cooling for different energy storage types in Tabuk.
Table 11. LCOE, LCOH, and CAPEX of the CPVT with cooling for different energy storage types in Tabuk.
LocationScenarioFluid Filter L C O E   ( $ / k W h ) L C O H   ( $ / k W h ) C A P E X   ( $ )
Storage TypeTESBatteryTESBatteryTESBattery
TabukAll electricityWater0.22320.3502 12,985,30720,368,382
Water+ZnO0.24420.3734 13,958,13621,341,211
Electricity + heatWater0.08470.15680.05360.09918,676,14216,059,217
Water+ZnO0.09060.16610.04620.08488,856,21716,239,292
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

Lucio, C.; Behar, O.; Dally, B. Techno-Economic Assessment of CPVT Spectral Splitting Technology: A Case Study on Saudi Arabia. Energies 2023, 16, 5392. https://doi.org/10.3390/en16145392

AMA Style

Lucio C, Behar O, Dally B. Techno-Economic Assessment of CPVT Spectral Splitting Technology: A Case Study on Saudi Arabia. Energies. 2023; 16(14):5392. https://doi.org/10.3390/en16145392

Chicago/Turabian Style

Lucio, Cesar, Omar Behar, and Bassam Dally. 2023. "Techno-Economic Assessment of CPVT Spectral Splitting Technology: A Case Study on Saudi Arabia" Energies 16, no. 14: 5392. https://doi.org/10.3390/en16145392

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