Next Article in Journal
An Efficient Computational Analysis and Modelling of Transferred Aerodynamic Loading on Direct-Drive System of 5 MW Wind Turbine and Results Driven Optimisation for a Sustainable Generator Structure
Previous Article in Journal
The Opening of High-Speed Rail and Economic Growth: The Moderating Role of Government Efficiency and Innovation Environment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Simulated Performance Analysis of a Hybrid Water-Cooled Photovoltaic/Parabolic Dish Concentrator Coupled with Conical Cavity Receiver

1
Mechanical and Energy Engineering Department, Imam Abdulrahman Bin Faisal University, Dammam 31441, Saudi Arabia
2
Energy and Thermal Systems Laboratory, National Engineering School of Monastir, University of Monastir, Ibn El Jazzar Street, Monastir 5019, Tunisia
3
Department of Mechanical Engineering, King Fahd University of Petroleum and Minerals, Dhahran 31261, Saudi Arabia
*
Author to whom correspondence should be addressed.
Sustainability 2024, 16(2), 544; https://doi.org/10.3390/su16020544
Submission received: 21 November 2023 / Revised: 23 December 2023 / Accepted: 2 January 2024 / Published: 8 January 2024

Abstract

:
The present research discloses a novel hybrid water-cooled Photovoltaic/Parabolic Dish Concentrator coupled with conical cavity receiver and spectral beam splitter (PV/PDC-CCR-BSF). In effect, a compact co-generating solar-concentrating PV system involving a subsequent optical interface has been fully developed and numerically tested. The optical performance of the proposed hybrid solar-concentrating system was modeled and assessed using the RT 3D-4R method while the thermal yield of the system was examined using the Finite Element Method. In addition to that, different configurations of serpentine-shape embedded water-cooling pipes (rectangle, semicircle, semi-ellipse and triangle) have been tested and optimized for maximum heat collection and minimum operating cell temperature. The performance of all the tested serpentine-shape embedded water-cooling pipes was evaluated with respect to conventional serpentine-shape water-cooling pipes. The outcomes indicated that the triangular cross-section outperforms other shapes in terms of heat dissipation capabilities, with about −446 W and maximum useful thermal power in the medium of the heat transfer fluid of 11.834 kW.

1. Introduction

Hybrid concentrating photovoltaic/thermal (CPV/T) solar systems are often used to further boost the useful work potential and the cost effectiveness of CPV technology [1,2,3]. The advantage of CPV/T’s usability is that it simultaneously produces heat and electrical output power from the incoming sun irradiation focused on the receivers at high power densities. The key component of this process is the Beam Splitter Filter (BSF) [4], which separates the reflected solar rays according to their wavelengths; solar rays with wavelengths in the bandpass zone are transmitted, otherwise they are reflected. To ensure the splitting of all reflected solar rays, the BSF’s position should be carefully optimized to ensure an acceptable optical efficiency. Several surveys on CPV/T systems coupled to a BSF have been examined in the literature [5,6,7,8,9,10,11,12]. Furthermore, solar hybrid systems using beam-splitter technology have attracted particular interest, as revealed through many investigations [4,5,6]. In addition, the design of cooling systems has been optimized with the aim of ensuring maximum thermos-electrical efficiencies and a longer lifespan for the entire system. In fact, micro-channel-based heat sinks (stepwise, manifold etc.) are the most investigated and employed cooling designs to ensure uniform temperature distribution across the solar cells. However, the excessive pressure drop should be considered and carefully optimized to maintain high exegetic efficiency. The micro-channel-based heat sink can be hybridized with impinging jets to further cool down the hot spots of the HCPV cells, but this solution is difficult to manufacture. The topic of advanced cooling technology has received significant attention from numerous researchers, such as Tyagi et al. [13], Sultan et al. [14], Lupu et al. [15] and Papis-Fraczek [16], who have extensively reviewed and analyzed it. Moreover, Akbarzadeh et al. [17] developed a theoretical study of a solar-concentrating parabolic trough system, and the results showed that the potential improvements that might be achieved by the suggested concentrating solar system were dependent on the optical properties of the reflector and the effectiveness of the cooling system. Mittelman et al. [18] analyzed solar cooling with CPV/T systems by combining heat and power to simultaneously cogenerate heat and electricity. The authors coupled the CPV/T with a single-effect absorption chiller. Moreover, Mittelman et al. [19] studied desalination units driven by CPV/T systems. According to the results, solar CPV/T combined with desalination systems has a promising potential that can compete with the traditional reverse osmosis (R.O.) desalination. Othman et al. [20] tested the proof of concept of a hybrid solar CPV/T consisting of a double-pass photovoltaic thermal (PV/T) solar air concentrator with a compound parabolic collector (CPC). The study aimed not only to increase the sunspot power density focused on the solar cells, but also to enhance the rate of heat transfer to the flowing air. Hedayatizadeh et al. [21] solved numerically the output power of a hybrid PV/T system based on CPC. The accuracy of the obtained outcomes was experimentally validated through data collection and analysis. He et al. [22] proposed an analytic model to study the CPV/T system’s characteristics and assess its overall performance. Meng et al. [23] suggested a novel design of CPV/T technology, aiming to achieve a high effective concentration ratio and a more uniform distribution of the solar cells across the receiver. Koronaki et al. [24] investigated a hybrid solar collector with a flat-plate receiver. The gold goal of the work was to estimate and enhance the overall performance of the proposed collector. The outcomes proved that the investigated solar collector could work efficiently throughout the year and record a seasonal capacity of generated power production of 2.2 kW in summer, 2.8 kW in spring and 2.6 kW in autumn. Wang et al. [25] studied the thermos-economic performance of a solar combined heat and power collector (SCHP) for the supply of electrical and heat loads. The outcomes showed that the selective spectrally coating filters employed with the hybrid PV/T collectors could significantly reduce the solar cells’ temperatures. Furthermore, they had the capability to yield thermal outputs at high temperatures for steam and hot water supply. Felsberger et al. [26] studied the pre-feasibility of an industrial thermal Parabolic Trough Collector (PTC). The findings revealed that the designed collector allowed a safe range of cell temperatures below 86 °C, which corresponds to an electrical cell efficiency of above 26.5% and Heat Transfer Fluid (HTF) temperatures of about 65 °C. Renno [27] investigated the performance of a line-focus CPV/T system applied to a residential user. In this study, the cooling fluid temperatures of the CPV/T system, the temperature stratification in the tank in winter and in summer, and the temperatures of the user loads were solved and calculated. The results demonstrated that the CPV/T system could supply a significant fraction of the annual heat loads. In addition, Renno et al. [28] developed a theoretical model of an Organic Rankine Cycle (ORC) system driven by a CPV/T collector. The results proved that the proposed CPV/T-ORC combined system could be an efficient alternative for power cogeneration. Nasseriyan et al. [29] conducted an experimental and numerical study to investigate the output power of the Solarus asymmetric CPC. The findings showed that maintaining a temperature difference with respect to the ambient temperature of around 23 °C can result in a rise in the electrical and thermal output power by 25% and 3%, respectively. Gakkhar et al. [30] reported and analyzed the cogenerated power of an experimental and numerical investigation of a hybrid CPV/T collector. The obtained results demonstrated that the mean overall efficiency of the system was 61.42%, 64.61% and 66.36% for a flow rate in the inner tube of 0.075 kg/s, 0.083 kg/s and 0.091 kg/s, respectively, for an annulus flow rate of 0.008 kg/s. Ustaoglu et al. [31] developed a new design for an enhanced CPV system, a compound hyperbolic concentrator-trumpet photovoltaic thermal system (CHCT-PV/T). The study reported that the non-imaging-light CHCT-PV/T can efficiently substitute CPV/T collectors. Lin et al. [32] explored the potential breakthroughs that can be achieved successfully by combining a CPV/T collector and a variable effect absorption chiller. Gomaa et al. [33] carried out a trade-off analysis to analyze the overall performance of a CPV/T collector to explore and define the operating boundaries within which the maximum output power can be produced. The outcomes revealed that the threshold electrical and thermal power of 170 W and 580 W, respectively, can be generated at a geometric concentration ratio of 3× and a water mass flow rate of 1 kg/min. Cabral et al. [34] analyzed an experimental study of a PV/T collector, and it was found that the proposed design was able to resolve the issue of the cells overheating when the sun was overhead, in otherwise normal incidence irradiation. Han et al. [35] presented a novel hybrid PV/T system based on a Silicon on Glass (SOG) Fresnel and beam splitting filter. Using a ray-tracing method, the optical efficiency of the proposed model was evaluated. The results depicted that beam filter coating characteristics have a consequential influence on the overall system performance and behavior. Gorouh et al. [36] developed a zero-dimensional thermal model to investigate a CPV/T collector at low optical concentration levels. In fact, the thermal and electrical peak efficiencies of the proposed collector were found to be 69.6% and 6.1%, respectively. According to the literature, many research studies have investigated the cooling mechanisms of the solar cells integrated into hybrid concentrated solar systems using numerical simulations tools or by conducting experimental studies, and then analyzing the collected data. The main purpose was to boost and optimize the thermos-optical efficiencies of such systems. However, there is a lack of sufficient research exploring through trade-off analysis the impact of the design of the active cooling-based embedded serpentine pipes on the cogenerated energies of such systems. In the present work, the effectiveness of a novel hybrid PV/PDC-CCR-BSF system is investigated. The examination revolves around understanding and contrasting the thermos-optical characteristics of the optimal cooling pipe designs with the conventional one using a numerical tool. The objective is first to model the opto-thermal behavior of the entire system, which has never been executed before. Then, the proposed hybrid system aims to identify the most effective and convenient alternative solution of conventional active cooling-based serpentine pipes. The performance of all the tested serpentine-shape embedded water-cooling pipes (rectangular, semicircular, semi-elliptical and triangular) was evaluated with respect to conventional serpentine-shape water-cooling pipes. The comparison of the performance of the tested serpentine-shape embedded water-cooling pipes indicated that the triangular cross-section outperformed other shapes in terms of heat dissipation capabilities and maximum useful thermal power in the medium of the heat transfer fluid.

2. Methodology

To design the entire system, the geometry modeling equations of the parabolic dish, conical cavity receiver and beam splitting filter were first formulated. The second step was dedicated to numerically investigating the spectral optical behavior of the hybrid system and then the heating load and the useful gain energy that would be removed within the domain of the working fluid for different cooling pipe cross-section shapes.

2.1. Geometry Modeling

The schematic of the investigated system is depicted in Figure 1 and Figure 2. Keeping the same focal point and the symmetric axis, the BSF is positioned above the parabolic dish collector (PDC) to concentrate almost all of the reflected solar rays. In the present study, the BSF is considered as the key element of the proposed hybrid solar system since it has a significant impact on the thermos-optical performance and responsiveness of the whole system. The principal characteristic of the BSF is the wide spectral selective coating which divides the concentrated solar rays according to their wavelengths. In fact, the spectral reflectivity (ρ) of the BSF selective coating is expressed as the following [5,7,8]:
ρ λ = 0         f o r   λ c / E g ρ λ = 1         f o r   λ < c / E g
where the solar cell’s bandgap is denoted by E g , the Planck’s constant is ℏ, and c denotes the speed of light in vacuum.
Equation (1) illustrates the splitting process of the reflected rays intercepted by the BSF surface. Solar rays having wavelengths below that of the semiconductor energy bandgap will be transmitted throughout the splitter filter placed at the focal plane of the concentrator. The other sun rays will be reflected back and projected on the plane of the solar cells, positioned at the reference zero-height of the considered vertical datum.
The parabolic dish concentrator is considered as a full tracker solar system to enable the exclusion of any further optical losses rather than that of the BSF.
Table 1 shows the main components and dimensions of the proposed hybrid system.

2.1.1. Parabolic Dish Collector

Several optical and geometric parameters can influence the processes for concentrating sunlight, such as the size of the reflectors and receivers, the spectral response of the parabolic collector and receiver materials, and the accuracy and resolution of the solar tracking system [37,38,39,40,41]. In fact, the parabolic collector equation is as follows [42]:
z = x 2 + y 2 4 f
where f is the focal length of the parabolic collector.
Thus, the reflector depth (dc) is deduced and can be written as follows [42]:
d c = w c 2 4 f
where wc is the reflector aperture radius.
The PDC system’s effectiveness is contingent upon the material and design selection of the receiver, which has an enormous influence on the yield of the whole system. The selection of the optimal design should consider the main purpose or application of the developed system [41]. The ratio of the intake aperture area to the exit receiver area is known as the geometric concentration ratio (GC) [42]:
G C = A i n l e t A o u t l e t  
where A i n l e t   and   A o u t l e t are the inlet aperture area and the outlet receiving area, respectively.
As presented by Kalogirou [43], for higher conversion efficiencies, in systems converting the solar radiation into heat by absorption, the GC should range between 600 and 2000×. Thus, the design of the receiver should take into account the geometrical concentration ratio and the optical properties of the involved materials [44].

2.1.2. Conical Cavity Receiver (CCR)

The performance of the parabolic solar concentrator system was studied and evaluated for different shapes of receivers [45,46,47,48,49,50,51,52,53]. According to the literature, these shapes can be categorized into two groups depending on their design: either a simple shape such as a cylindrical cavity [45], conical cavity [48,49,50], spherical cavity and flat disk [54], or a more intricate shape [51,52]. In fact, a comparison between these different shapes in terms of optical efficiency was conducted to determine the most efficient geometry [46,47]. This article focuses on the geometry of the conical receiver as the most efficient of all other simple receiver geometries [46,47].
The CCR equation In the Cartesian system coordinates is the following [54]:
z = z c r + [ r m a x x c r x 2 + y c r y 2 cot ( θ o p ) ]
where x c r , y c r   and   z c r are the center’s coordinates, rmax and rmin are the radius of the conical upper and lower bases, and θ o p is the CCR’s half opening angle, which is described as:
θ o p = tan 1 ( r m a x r m i n h r )  
where hr is the conical cavity height.
The formula of the geometric concentration ratio is then:
G C = w c 2 r m a x + r m i n * h r 2 + ( r m a x r m i n ) 2
The collector aperture area A c is determined from the relation:
A c = π w c 2
while the receiver surface area A r is expressed as follows:
A r = π r m a x + r m i n h r 2 + ( r m a x r m i n ) 2
The main objective of this article is to numerically model the opto-thermal behavior of the entire system (water-cooled PV/PDC-CCR-BSF) and to demonstrate the upper hand and superior performance of such novel technology to pave the avenues for future experimental prototyping and validation. A helical water tube is installed across the innermost portion of the conical cavity receiver. The conical height serves to carry out the circulation of cooling water from bottom to top.

2.1.3. PV Cells

A small-sized PV panel serves as a second receiver in the design of the proposed hybrid PV/PDC-CCR-BSF system. The objective is to generate electricity while simultaneously producing heat [55,56,57,58,59]. In fact, single-junction solar cells are placed at the vertical datum origin height, which is at the vertex of the parabolic dish. Since it is well known that solar cells are very sensitive to external influences (environmental and climatic conditions), the solar-cell panels are made up of various layers (Figure 1). These layers are as follows, from top to bottom: (1) the topmost layer of the solar panel is a glass layer, which is a protective layer made up of a highly transparent medium that is resistant to several conditions; (2) the solar cells are then covered with one ethylene vinyl acetate (EVA) layer (3) and then sandwiched between other two EVA layers, to protect the solar cells’ circuits from soft vibrations and shocks and to prevent the penetration of dirt and moisture; (4) the panel back sheet is a thin and opaque film that is composed of Thiophene-benzothiazole (TBT) to reduce the effects of the ultraviolet exposures from the bottom. For the PV/T panel, a cooling pipe (5) is embedded in the back sheet of the PV panel. In fact, once the reflected solar rays from the BSF reach the hybrid PV/T panel, a major fraction of the heat dissipated by the solar cells will be transferred to the water flow circulating inside the cooling pipe. The conventional cooling pipe PV system, carrying water, is a serpentine cylindrical shape [60].
Table 2 recapitulates the main optical and thermophysical properties of the different PV panel layers.
The aim of the serpentine-shaped water flow embedded in the back sheet of the PV panel is to ensure optimal operating conditions for the solar cells and to avoid any extreme heat loads and temperatures for optimum electrical efficiencies. While cooling pipes have long been used to remove heat from panels, research needs to be carried out on the ideal cross-section configuration of the tubing in relation to the rear sheet enabling efficient heat dispersal. In this regard, the computational fluid dynamics code ANSYS FLUENT was utilized in order to simulate and expose the serpentine-shaped water flow integrated into the heat sink, depicted in Figure 2, to perform evaluation with four distinct cross-sections.

2.1.4. Beam Splitter Filter

The paraboloidal equation is defined as follows [42]:
z = x 2 + y 2 4 f s + z f s
where z f s is the elevation of the filter over the dish reflector and f s is the focal length.
As mentioned before, the BSF splits the sunlight according to the wavelengths. Over the BSF area, a spectrally selective coating is disposed. The bandpass section of the spectral width is selected when the wavelength is above 1100 nm [8,63].

2.2. Simulation Methods

2.2.1. Optical RT 3D-4R Method

The investigation of the optical performance of the hybrid PV/PDC-CCR-BSF is carried out with numerical modeling based on the ray-tracing technique (RT 3D-4R) [54,64,65,66,67]. This method is developed to compute the optical path length and ray intensities for unpolarized electromagnetic waves. The solar disk theory is assumed, and the sun subtend angle σ s / 2 is set equal to 16 and 32 arcminutes [68]. The dependence of the refraction indexes on the temperatures is neglected.
Each integration node over the dish reflector area can be coordinated as the following:
x c i c , j c = r c i c   c o s ( φ c ( j c ) π 180 ) y c i c , j c = r c i c   s i n ( φ c ( j c ) π 180 ) z c i c , j c = x c i c , j c 2 + y c i c , j c 2 4   f
where r c i c = ( i c 1 ) d r c φ c j c = ( j c 1 ) d φ c , a n d   d r c , and d φ c are the two reflector spatial steps.
The conical cavity receiver coordinates x r ( i r , j r ) , y r ( i r , j r ) and   z r ( i r , j r ) could be expressed as follows:
x r i r , j r = x c r r r i r   c o s ( φ r ( j r ) π 180 ) y r i r , j r = y c r r r i r   s i n ( φ r ( j r ) π 180 ) z r i r , j r = z c r + ( r m a x r r ( i r ) ) cot ( α )
where r r i r = ( i r 1 ) d r r φ r j r = ( j r 1 ) d φ r , and d r r and d φ r are, respectively, the two reflector space increments.
The block diagram of the developed RT3D-4R algorithm is described in Figure 3.

2.2.2. Heat Transfer Modeling of the PV and Conical Cavity Receivers

The numerical model and simulation of the proposed solar system is based on Finite Element Method analysis. Figure 4 shows the generated mesh over both the conical cavity receiver and the PV solar cells. The physics-controlled mesh is adapted according to the physics settings for any boundary or domain condition and the constraints applied to the developed model geometry.

Heat Transfer on Sinusoidal Coiled Tube of the PV Cells Panels

The 4 different embedded serpentine-shaped water flow configurations were numerically modeled in the ANSYS FLUENT module. The front side of the PV panel was subjected to a constant heat flux of 6400   W / m 2 while the sides and the back of the panel were exposed to surrounding air at T s u r = 298   K   and   h = 5   W / m 2 · K , as shown in Figure 5. An emissivity of ε = 0.9 was considered for all surfaces exposed to surrounding air.
All cross-sectional areas, the pipe’s overall length, its shape (serpentine), its entry temperature, its velocity and the mass flow rate with which water passed through the pipe were maintained uniform for homogeneity across the board. The dimensional features of the tubing sectional areas are listed in Table 3, and Table 4 lists the boundary conditions common to all cases, as shown in Figure 2.
The photovoltaic module and tubing configuration computing domain was modeled in the Design Modeler module of ANSYS. The domain was then discretized using unstructured mesh made up of tetrahedral cells. To properly resolve the fluid boundary layer adjacent to the solid walls, the first height, the rate of increase and number of layers were set, to ensure that the wall y+ < 1. For the solution, the RANS equations representing the mass, momentum and energy conservation equations, and the steady pressure-based model employing the Spalart–Allmaras model were used.
The mass, momentum and energy conservation equations for a steady, three-dimensional flow are given by:
Mass :   · ρ V = 0
Momentum :   · ρ V V = p + · μ V + V T 2 3 · V I + ρ g
Energy :   · V ρ E + p = · k e f f T j h j J j + τ ̿ e f f · V
where transfers of energy via conduction, diffusion and dissipation through viscosity are represented by the initial three terms on the right-hand side of Equation (15).
A second-order upwind approach is employed to discretize Equations (13)–(15), which are subsequently resolved using a pressure–velocity coupling mechanism. The criteria for solution convergence are based on residual values below 1 × 10−5 and net heat flux and net mass flow rate values below 1 × 10−2.

3. Results of Simulation Methods

3.1. Validation of the RT3D-4R Technique

To validate the developed in-house optical model, the obtained results are compared against those obtained using other modeling and simulation tools such as the MATLAB program (R2020a) [53], COMPREC code [53] and experimental measurements by Jeter [69]. The same material and geometrical properties have been assumed to deliver accurate and fair validation:
  • Focal distance: f = 1 m;
  • Rim angle: 45° and 60°;
  • Sun subtend disk angle: θ s = 16′ and 32′;
  • Solar irradiance: I r = 1000   W / m 2 .
Figure 6 depicts the comparison between the experimental and numerical results in terms of LCR as a function of the normalized sunspot coordinates for two reference sun subtend angle values. The normalized sunspot coordinate represents the position of a point in the flat-disk receiver, normalized with respect to the focal distance of the parabolic dish collector. It is evident from Figure 6 that the optical approach currently being used yields a pattern that exhibits good agreement with both numerical and experimental results.

3.2. Hybrid Parabolic Dish Collector Using Spectral Beam Splitter Technology

3.2.1. Beam Splitter Filter and Concentrated Solar Rays

After the incident rays have been reflected by the paraboloidal mirrors of the PDC, they reach the boundaries of the BSF, and after that they are filtered to be either transmitted to the outer walls of the cavity receiver or reflected onto the solar cells array. The bending rate of the transmitted rays depends on the refractive index of the BSF and the angular displacement of the incident rays away from the normal to the element boundaries of the BSF. The reflected rays from the BSF are considered as a new set of incident rays and are projected onto the solar cells’ surface area with minimum interception angle. In fact, the BSF technology splits the impinged rays as function of their wavelengths. Almost 85% of the spectrum light density, which is reflected from the BSF, is concentrated on the PV/T panel. The rest of the spectrum light density is transmitted and reaches the CCR [8].

3.2.2. Concentration Ratio and Temperature Distribution on the CCR

In this section, the concentrated solar density (CSD) distribution on the conical cavity receiver is discussed. With the same case studies investigated in previous work [54], the simulation details are used with characteristics: wc = 2.5 m, f = 3 m, rmax = 0.1 m, rmin = 0.05 m, θS = 16′ and ES = 1 kW/m2. The simulation outcomes are illustrated in Figure 7. It is evident that the CSD distribution exhibits a non-uniformity aspect along the cavity’s height. Furthermore, the highest CSD value is around 40 kW/m2 situated at 0.1 m from the conical cavity bottom base. Compared with that found by Daabo et al. [46,47], this finding proves the reliability of the presented simulation method.
The optical efficiency is the ratio of the average local concentration ratio to the geometric concentration ratio [70]. The optimization of the hybrid PV/PDC-CCR-BSF system in terms of optical efficiency is presented in [54].
Figure 7b shows the temperature distribution on the CCR where an increase in the temperature with height is observed. The top part of the CCR is found to have the maximum surface temperature where maximum temperature values reach more than 170 °C. This temperature trend is influenced by the convective heat transfer rates of the water flow inside the helical tube, which are faster than the conduction heat transfer flow over the cavity inner walls exposed to the transmitted radiative sun power intensities. The cold-water inlet is fed in at the bottom base of the cavity, with the water temperature increasing from the inlet value of 20 °C to around 104 °C at the top of the CCR before leaving the helical tube.

3.2.3. Heat Load and Temperature Distribution of the Solar Cells with Conventional Coiled Pipe

Approximately 66% of the concentrated solar energy received by the BSF device is reflected toward the PV cells. As stated earlier, the BSF device is positioned in such a way that its focal point matches the primary parabolic reflector. Consequently, the solar rays bouncing off the BSF maintain a path parallel with the parabolic reflector’s symmetry axis and are then finished by being absorbed by the PV cell positioned on the parabolic reflector’s center. As a result, the CSD distribution over the PV cells area is spread out uniformly, reaching a maximum value of 8 kW/m2 (Figure 8). At the edges, the CSD has decreased since the sunspots are extending over a smaller area than that of the PV cells.
Figure 8b,c depict the temperature distribution of the cooled solar cells using conventional water-cooled pipe. For a full pipe flow, the maximum solar cell temperature has been decreased from 90.5 °C to 49.6 °C.

3.2.4. Heat Load and Temperature Distribution of the Solar Cells with Different Coiled Pipe Forms

This article aims to optimize the geometry of the serpentine-shaped water flow embedded in the heat sink placed underneath the solar cells of the PV panel. The conventional cooling pipe, characterized by a cylindrical shape, is in direct contact with the back sheet of the panel. Thus, the main idea is to examine the effect of different cooling pipe cross-sections with the same flow area but different contact areas that can lead to higher heat transfer to the cooling pipe. In this regard, four different cross-section area shapes are analyzed. The areas and pipe length are fixed at 0.0008 m2 and 12.637 m, respectively.
The distribution of heat flux on the Interface surface along the coolant conduit and panel rear sheet is depicted in Figure 9 for the four different cooling pipe cross-section cases. Since the contact sides of the cooling pipe channel have different widths, the heat transfer is directly proportional to the contact areas. However, a small difference is observed since the difference between the contact areas of the four different cross-sections is small and is listed in Table 4 as the width parameter for comparison.
Figure 10 depicts the temperature distribution on the serpentine shape of the different serpentine-shape embedded cooling pipes. The inlet water temperature is fixed at 298 °K for all cases. However, one can observe that the outlet temperature values differ for the four cases, see Table 5. In the case study of the conventional cooling pipe, the variation in temperature across the water at the point of entry and the outflow is about 19 °C, which is lower than that found for the other geometries. In fact, the temperature difference is about 28.6 °C for the rectangular cooling pipe cross-section, 28.2 °C for the semicircular cooling pipe case, 29.9 °C for the semi-elliptical cooling pipe case and 30.3 °C for the triangular cooling pipe case. Thus, the largest is found for the triangular cross-section cooling pipe by virtue of its larger width, see Table 3.
Table 5 summarizes the simulation results of the four different serpentine-shape embedded water flow cross-sections.
In Table 5, it is clear that the triangle cross-section appears to be the best choice in comparison to the others in terms of extracting heat from the rear of the sheet in that it results in the maximum thermal energy rate to the cooling water ( q ˙ o u t = 11.834 kW), which is the thermal power of the liquid out of the pipe, and the lowest back-sheet temperature (Tback = 333.1 K) and heat flux ( q ˙ b a c k = −446 W), meaning more cooling capability than the other cross-section cooling pipes.

4. Comparison of Previous Similar Studies

This article aims to optimize the geometry of the serpentine-shaped water flow embedded in the heat sink placed underneath the solar cells of the PV panel. The optical performance of the proposed hybrid solar-concentrating system was modeled and assessed using the RT 3D-4R method, while the thermal yield of the system was examined using the Finite Element Method. Different configurations of serpentine-shape embedded water-cooling pipes (rectangle, semicircle, semi-ellipse and triangle) have been tested and optimized for maximum heat collection and minimum operating cell temperature. The performance of all the tested serpentine-shape embedded water-cooling pipes (rectangle, semicircle, semi-ellipse and triangle) was evaluated with respect to conventional serpentine-shape water-cooling pipes. The outcomes indicated that the triangular cross-section outperforms other shapes in terms of heat dissipation capabilities and maximum useful thermal power in the medium of the heat transfer fluid. Table 6 provides a comparative overview in terms of improving both thermal and electrical efficiency. Table 6 provides various information like whether the analysis was conducted in various studies experimentally or numerically, the concentrator solar system used in these studies, the type of cooling system used, the optical concentration ratio, the heat transfer fluid, change in temperature, various efficiencies and the main outcomes of these studies.

5. Discussion

A crucial factor in ensuring optimal performance of the hybrid PV/PDC-CCR-BSF system is the integration of a sophisticated cooling pipe system to avoid cell overheating, which may result in an output voltage drop and a short lifespan of the solar cells. In this study, the design optimization, a serpentine-shape embedded water flow cooling pipe, was carried out by trade-off analyses using the optical ray-tracing technique and the Finite Element Methods, using ANSYS FLUENT software (version 2019R3) to simulate the relevance of the key designing parameters of the cooling system configurations.
To achieve this goal, four different serpentine-shape water-cooling pipes embedded in the plate of the underneath heat sink were investigated, while maintaining the same surface area; the four tested geometries were semicircular, semi-elliptical, rectangular and triangular pipe. Simulation was carried out to assess the thermo-optical efficiency of the selected serpentine-shape embedded cooling pipe configurations. A comparison between the proposed four different embedded water flow pipes’ cross-sections and the traditional circular cross-section cooling pipe was assessed. The heat exchange area between the panel and the pipe changed from one form to another.
The simulation’s findings indicated that the amount of heat dissipated is contingent upon the contact area utilized for transfer. Hence, the triangular cross-section shape showcases the utmost performance capability:
  • The triangular cross-section shape resulted in the largest increment in temperature, in other words the maximum useful gain energy, recording the highest increment in temperature of 30.3 °C. The triangular cross-section shape recorded the fastest rate at which heat was transferred to the water flow in the serpentine-shape embedded cooling pipes.
  • The same cooling pipe design, i.e., the triangular cross-section shape, recorded the lowest heat sink temperature resulting in the highest heat transfer rate to the cooling water flow of 11.834 kW.

6. Conclusions

This current work discloses a novel hybrid water-cooled Photovoltaic/Parabolic Dish Concentrator coupled with a conical cavity receiver and spectral beam splitter (PV/PDC-CCR-BSF). The optical modeling has been developed based on the ray-tracing method (RT3D-4R) and the thermal analysis has been elaborated using Finite Element Method software (COMSOL Multiphysics, version 5.2). In the current study, different configurations of serpentine-shape embedded water-cooling pipes (rectangle, semicircle, semi-ellipse and triangle) have been tested and optimized for maximum heat collection and minimum operating cell temperature. The performance of all the tested serpentine-shape embedded water-cooling pipes was evaluated with respect to the conventional serpentine-shape water-cooling pipes. The findings of this study lead to the following conclusions and suggestions:
  • First, the optical validity of the RT3D-4R method has been approved to simulate the performance of the parabolic dish concentrator for two case studies (flat receiver and conical receiver).
  • Second, the thermo-optical characteristics of the proposed PV/PDC-CCR-BSF system have been optimized and performed.
  • Third, the thermo-optical performance of the proposed system for four different cooling pipe cross-sections of serpentine-shape embedded water-cooling pipes have been numerically simulated and the results compared against those of the conventional circular cross-section cooling pipe placed underneath the PV panel rear sheet. The outcomes showed that:
    The heat sinks and the water-cooling pipe’s interface width determine the heat flux pattern and transfer rates in a comparative manner.
    The water temperature increment between the inlet and outlet ports of the embedded serpentine water-cooling pipe is equal to 28.6 °C for the rectangular cross-section cooling pipe, 28.2 °C for the semicircular cross-section cooling pipe, 29.9 °C for the semi-elliptical cross-section cooling pipe and 30.3 °C for the triangular cross-section cooling pipe.
    Amongst the four different cross-sections, the triangle cross-section is found to have a greater heat extraction capability than the other cross-sections. It gives the highest heat rate to the cooling water, the lowest back-sheet temperature and heat flux ( q ˙ b a c k = −446 W), yielding the highest amount of removed thermal power in the water flow medium.
    The outcomes demonstrated that the traditional cooling pipe, which is also placed underneath the solar cells, has poorer cooling performance in comparison with the four serpentine-shape embedded water-cooling configurations.
As an avenue of the present research, further analysis and enhancement of the serpentine-shape embedded water-cooling configurations is carried out in another concurrent study.

Author Contributions

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

Funding

This research received no external funding.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors would like to acknowledge the support provided by Imam Abdulrahman Bin Faisal University (IAU).

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

English alphabetsUnitsGreek symbolsUnits
DConcentrated solar density[kW/m2]αAbsorptivity-
dsSurface (differential) of the collector[m2]θAngle[rad]
dSSize (differential) of the sunspot[m2]ρReflectivity-
EIncident solar density[kWh/m2]σSolar angle[Arcminutes]
fFocal length[m]λSolar ray wavelength[m]
GGrid number-Subscripts
QSolar flux[kW/m2]cCollector
q ˙ Thermal power[W]crCenter receiver
TTemperature [°K]iIncidence
AbbreviationsmaxMaximum
DNIDirect Solar IrradianceminMinimum
CSDConcentrated Solar DensityopOpening
RT3D-4RRay-Tracing 3Dimensions-4RaysrReceiver
BSFBeam Splitter FilterrimRim
TBTThiophene-BenzothiazolesSolar
EVAEthylene Vinyl AcetatescSolar cell
PVPhotovoltaicsfSplitter filter
UVUltravioletgGap

References

  1. Ju, X.; Xu, C.; Liao, Z.; Du, X.; Wei, G.; Wang, Z.; Yang, Y. A review of concentrated photovoltaic-thermal (CPVT) hybrid solar systems with waste heat recovery (WHR). Sci. Bull. 2017, 62, 1388–1426. [Google Scholar]
  2. 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]
  3. Mathew, G.; Pandey, A.K.; Nasrudin, A.R.; Tyagi, V.V.; Syed, 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]
  4. Mojiri, A.; Taylor, R.; Thomsen, E.; Rosengarten, G. Spectral beam splitting for efficient conversion of solar energy—A review. Renew. Sustain. Energy Rev. 2013, 28, 654–663. [Google Scholar] [CrossRef]
  5. Imenes, A.G.; Mills, D.R. Spectral beam splitting technology for increased conversion efficiency in solar concentrating systems: A review. Sol. Energy Mater. Sol. Cells 2004, 84, 19–69. [Google Scholar] [CrossRef]
  6. Huang, G.; Curt, S.R.; Wang, K.; Markides, C.N. Challenges and opportunities for nanomaterials in spectral splitting for high-performance hybrid solar photovoltaic-thermal applications: A review. Nano Mater. Sci. 2020, 3, 183–203. [Google Scholar]
  7. Sahu, S.K.; Kopalakrishnaswami, A.S.; Natarajan, S.K. Historical overview of power generation in solar parabolic dish collector system. Environ. Sci. Pollut. Res. 2022, 29, 64404–64446. [Google Scholar]
  8. Jiang, S.; Chen, Z.; Peng, H. Three-dimension optical model of two-stage reflective spectral beam spliting concentrating PV system. Acta Energiae Solaris Sin. 2009, 30, 1188–1193. [Google Scholar]
  9. Jiang, S.; Hu, P.; Mo, S.; Chen, Z. Optical modeling for a two-stage parabolic trough concentrating photovoltaic/thermal system using spectral beam splitting technology. Sol. Energy Mater. Sol. Cells 2010, 94, 1686–1696. [Google Scholar] [CrossRef]
  10. Jiang, S.; Wang, G.; Hu, P.; Chen, Z.; Jia, L. The Design of Beam Splitter for Two-Stage Reflective Spectral Beam Splitting Concentrating PV/thermal System. In Proceedings of the Asia-Pacific Power and Energy Engineering Conference, IEEE 2011, Wuhan, China, 25–28 March 2011. [Google Scholar]
  11. Shou, C.-H.; Luo, Z.-Y.; Wang, T.; Shen, W.-D.; Gary, R.; Wang, C.; Ni, M.-J.; Cen, K.-F. A Dielectric Multilayer Filter for Combining Photovoltaics with a Stirling Engine for Improvement of the Efficiency of Solar Electricity Generation. Chin. Phys. Lett. 2011, 28, 128–402. [Google Scholar] [CrossRef]
  12. Wang, G.; Yao, Y.; Chen, Z.; Hu, P. Thermodynamic and optical analyses of a hybrid solar CPV/T system with high solar concentrating uniformity based on spectral beam splitting technology. Energy 2019, 166, 256–266. [Google Scholar] [CrossRef]
  13. Tyagi, V.V.; Kaushik, S.C.; Tyagi, S.K. Advancement in solar photovoltaic/thermal (PV/T) hybrid collector technology. Renew. Sustain. Energy Rev. 2012, 16, 1383–1398. [Google Scholar] [CrossRef]
  14. Sultan, S.M.; Efzan, M.N.E. Review on recent Photovoltaic/Thermal (PV/T) technology advances and applications. Sol. Energy 2018, 173, 939–954. [Google Scholar] [CrossRef]
  15. Lupu, A.G.; Homutescu, V.M.; Balanescu, D.T.; Popescu, A. A review of solar photovoltaic systems cooling technologies. In Proceedings of the 8th International Conference on Advanced Concepts in Mechanical Engineering, Iasi, Romania, 7–8 June 2018. [Google Scholar]
  16. Papis-Frączek, K.; Sornek, K. A Review on Heat Extraction Devices for CPVT Systems with Active Liquid Cooling. Energies 2022, 15, 6123. [Google Scholar] [CrossRef]
  17. Akbarzadeh, A.; Wadowski, T. Heat pipe-based cooling systems for photovoltaic cells under concentrated solar radiation. Appl. Therm. Eng. 1996, 16, 81–87. [Google Scholar] [CrossRef]
  18. Mittelman, G.; Kribus, A.; Dayan, A. Solar cooling with concentrating photovoltaic/thermal (CPVT) systems. Energy Convers. Manag. 2007, 48, 2481–2490. [Google Scholar] [CrossRef]
  19. Mittelman, G.; Kribus, A.; Mouchtar, O.; Dayan, A. Water desalination with concentrating photovoltaic/thermal (CPVT) systems. Sol. Energy 2009, 83, 1322–1334. [Google Scholar] [CrossRef]
  20. Othman, M.Y.; Yatim, B.; Sopian, K.; Bakar, M.N. Performance analysis of a double-pass photovoltaic/thermal (PV/T) solar collector with CPC and fins. Renew. Energy 2005, 30, 2005–2017. [Google Scholar] [CrossRef]
  21. Hedayatizadeh, M.; Ajabshirchi, Y.; Sarhaddi, F.; Safavinejad, A.; Farahat, S.; Chaji, H. Thermal and Electrical Assessment of an Integrated Solar Photovoltaic Thermal (PV/T) Water Collector Equipped with a Compound Parabolic Concentrator (CPC). Int. J. Green Energy 2013, 10, 494–522. [Google Scholar] [CrossRef]
  22. He, Y.; Xiao, L.; Li, L. Theoretical and experimental study on the application of diffuse-reflection concentrators in PV/T solar system. Int. J. Energy Res. 2016, 40, 963–970. [Google Scholar] [CrossRef]
  23. Meng, X.; Xia, X.; Sun, C.; Li, Y.; Li, X. A novel free-form Cassegrain concentrator for PV/T combining utilization. Sol. Energy 2016, 135, 864–873. [Google Scholar] [CrossRef]
  24. Koronaki, I.P.; Nitsas, M.T. Experimental and theoretical performance investigation of asymmetric photovoltaic/thermal hybrid solar collectors connected in series. Renew. Energy 2018, 118, 654–672. [Google Scholar] [CrossRef]
  25. Wang, K.; Pantaleo, A.M.; Herrando, M.; Pesmazoglou, I.; Franchetti, B.M.; Markides, C.N. Thermoeconomic assessment of a spectral splitting hybrid PV-T system in dairy farms for combined heat and power. In Proceedings of the 32nd International Conference on Efficiency, Cost, Optimization, Simulation and Environmental Impact of Energy Systems (ECOS), Wroclaw, Portland, 23–28 June 2019; pp. 23–28. [Google Scholar]
  26. Felsberger, R.; Buchroithner, A.; Gerl, B.; Wegleiter, H. Conversion and Testing of a Solar Thermal Parabolic Trough Collector for CPV-T Application. Energies 2020, 13, 6142. [Google Scholar] [CrossRef]
  27. Renno, C. Theoretical and Experimental Evaluation of the Working Fluid Temperature Levels in a CPV/T System. Energies 2020, 13, 3077. [Google Scholar] [CrossRef]
  28. Renno, C.; Petito, F.; D’Agostino, D.; Minichiello, F. Modeling of a CPV/T-ORC Combined System Adopted for an Industrial User. Energies 2020, 13, 3476. [Google Scholar] [CrossRef]
  29. Nasseriyan, P.; Afzali Gorouh, H.; Gomes, J.; Cabral, D.; Salmanzadeh, M.; Lehmann, T.; Hayati, A. Numerical and Experimental Study of an Asymmetric CPC-PVT Solar Collector. Energies 2020, 13, 1669. [Google Scholar] [CrossRef]
  30. Gakkhar, N.; Soni, M.S.; Jakhar, S. Experimental and theoretical analysis of hybrid concentrated photovoltaic/thermal system using parabolic trough collector. Appl. Therm. Eng. 2020, 171, 115069. [Google Scholar] [CrossRef]
  31. Ustaoglu, A.; Ozbey, U.; Torlaklı, H. Numerical investigation of concentrating photovoltaic/thermal (CPV/T) system using Compound Hyperbolic–Trumpet, V-trough and Compound Parabolic Concentrators. Renew. Energy 2020, 152, 1192–1208. [Google Scholar] [CrossRef]
  32. Lin, L.; Tian, Y.; Luo, Y.; Chen, C.; Jiang, L. A novel solar system integrating concentrating photovoltaic thermal collectors and variable effect absorption chiller for flexible co-generation of electricity and cooling. Energy Convers. Manag. 2020, 206, 112506. [Google Scholar] [CrossRef]
  33. Gomaa, M.R.; Al-Dhaifallah, M.; Alahmer, A.; Rezk, H. Design, Modeling, and Experimental Investigation of Active Water Cooling Concentrating Photovoltaic System. Sustainability 2020, 12, 5392. [Google Scholar] [CrossRef]
  34. Cabral, D.; Gomes, J.; Hayati, A.; Karlsson, B. Experimental investigation of a CPVT collector coupled with a wedge PVT receiver. Sol. Energy 2021, 215, 335–345. [Google Scholar] [CrossRef]
  35. 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]
  36. Gorouh, H.A.; Salmanzadeh, M.; Nasseriyan, P.; Hayati, A.; Cabral, D.; Gomes, J.; Karlsson, B. Thermal modelling and experimental evaluation of a novel concentrating photovoltaic thermal collector (CPVT) with parabolic concentrator. Renew. Energy 2022, 181, 535–553. [Google Scholar] [CrossRef]
  37. Howard, D.; Harley, R.G. Modeling of Dish-Stirling Solar Thermal Power Generation. In Proceedings of the IEEE PES General Meeting, Minneapolis, MN, USA, 25–29 July 2010; pp. 1–7. [Google Scholar]
  38. Fraser, P.R. Stirling Dish System Performance Prediction Model. Master’s Thesis, University of Wisconsin-Madison, Madison, WI, USA, 2008. [Google Scholar]
  39. Noor, N.; Muneer, S. Concentrating Solar Power (CSP) and Its Prospect in Bangladesh. In Proceedings of the 2009 1st International Conference on the Developements in Renewable Energy Technology (ICDRET), Dhaka, Bangladesh, 17–19 December 2009; 2009; pp. 1–5. [Google Scholar]
  40. Stine, W.B.; Diver, R.B. A Compendium of Solar Dish/Stirling Technology; Sandia National Laboratories, Albuquerque New Mexico 87185 and Livermore, California 94550, For the United States Department of Energy: Albuquerque, NM, USA, 1994. [Google Scholar]
  41. Geyer, M.A.; Stine, W.B. Power from the Sun. 2001. Available online: http://powerfromthesun.net/ (accessed on 24 December 2023).
  42. Gaith, F.; Taher, M.; Fahad, G. Performance Analysis of a Single Cell-Ultra-High Concentration Photovoltaic Thermal Module Based on Pin-Fins Cooling Microchannel. Int. J. Energy Res. 2021, 46, 2947–2969. [Google Scholar]
  43. Soteris, A.K. Solar thermal collectors and applications. Prog. Energy Combust. Sci. 2004, 30, 231–295. [Google Scholar]
  44. Li, H.; Huang, W.; Huang, F.; Hu, P.; Chen, Z. Optical analysis and optimization of parabolic dish solar concentrator with a cavity receiver. Sol. Energy 2013, 92, 288–297. [Google Scholar] [CrossRef]
  45. Wang, F.; Lin, R.; Liu, B.; Tan, H.; Shuai, Y. Optical efficiency analysis of cylindrical cavity receiver with bottom surface convex. Sol. Energy 2013, 90, 195–204. [Google Scholar] [CrossRef]
  46. Daabo, A.M.; Mahmoud, S.; Al-Dadah, R.K. The optical efficiency of three different geometries of a small-scale cavity receiver for concentrated solar applications. Appl. Energy 2016, 179, 1081–1096. [Google Scholar] [CrossRef]
  47. Daabo, A.M.; Mahmoud, S.; Al-Dadah, R.K. The effect of receiver geometry on the optical performance of a small-scale solar cavity receiver for parabolic dish applications. Energy 2016, 114, 513–525. [Google Scholar] [CrossRef]
  48. Shuai, Y.; Xia, X.-L.; Tan, H.-P. Radiation performance of dish solar concentrator/cavity receiver systems. Sol. Energy 2008, 82, 13–21. [Google Scholar] [CrossRef]
  49. Shuai, Y.; Xia, X.-L.; Tan, H.-P. Numerical Study of Radiation Characteristics in a Dish Solar Collector System. J. Sol. Energy Eng. 2008, 130, 021001. [Google Scholar] [CrossRef]
  50. Shuai, Y.; Xia, X.-L.; Tan, H.-P. Numerical simulation and experiment research of radiation performance in a dish solar collector system. Front. Energy Power Eng. China 2010, 4, 488–495. [Google Scholar] [CrossRef]
  51. Li, S.; Xu, G.; Luo, X.; Quan, Y.; Ge, Y. Optical performance of a solar dish concentrator/receiver system: Influence of geometrical and surface properties of cavity receiver. Energy 2016, 113, 95–107. [Google Scholar] [CrossRef]
  52. Reddy, K.S.; Natarajan, S.K.; Veershetty, G. Experimental performance investigation of modified cavity receiver with fuzzy focal solar dish concentrator. Renew. Energy 2015, 74, 148–157. [Google Scholar] [CrossRef]
  53. Johnston, G. Focal region measurements of the 20 m2 tiled dish at the Australian National University. Sol. Energy 1998, 63, 117–124. [Google Scholar] [CrossRef]
  54. Houcine, A.; Maatallah, T.; Saeed, F. Optothermal characterization of a hybrid Photovoltaic/Parabolic Dish Concentrator coupled with conical cavity receiver using beam splitter for steam and electrical power generation. Int. J. Energy Res. 2022, 46, 12341–12361. [Google Scholar] [CrossRef]
  55. Good, C. Environmental impact assessments of hybrid photovoltaic-thermal (PV/T) systems—A review. Renew. Sustain. Energy Rev. 2015, 55, 234–235. [Google Scholar] [CrossRef]
  56. Sreekumar, S.; Shah, N.; Mondol, J.; Hewitt, N.; Chakrabarti, S. Numerical Investigation and Feasibility Study on MXene/Water Nanofluid Based Photovoltaic/thermal System. Clean. Energy Syst. 2022, 103, 504–515. [Google Scholar] [CrossRef]
  57. Zenhäusern, D.; Bamberger, E.; Baggenstos, A.; Häberle, A. PVT Wrap-Up: Energy Systems with Photovoltaic-Thermal Solar Collectors. In Proceedings of the ISES Solar World Congress, Abu Dhabi, United Arab Emirates, 29 October–2 November 2017. [Google Scholar]
  58. Zeitouny, J.; Lalau, N.; Gordon, J.M.; Katz, E.A.; Flamant, G.; Dollet, A.; Vossier, A. Assessing high-temperature photovoltaic performance for solar hybrid power plants. Sol. Energy Mater. Sol. Cells 2018, 182, 61–67. [Google Scholar] [CrossRef]
  59. Ahliouati, M.; El Otmani, R.; Kandoussi, K.; Boutaous, M.H.; Louanate, A.; Louzazni, M.; Daya, A. Energetic and parametric studies of a basic hybrid collector (PV/T-Air) and a photovoltaic (PV) module for building applications: Performance analysis under El Jadida weather conditions. Mater. Sci. Energy Technol. 2023, 6, 267–281. [Google Scholar] [CrossRef]
  60. Alzaabi, A.A.; Badawiyeh, N.K.; Hantoush, H.O.; Hamid, A.K. Electrical/thermal performance of hybrid PV/T system in Sharjah, UAE. Int. J. Smart Grid Clean Energy 2014, 3, 385–389. [Google Scholar] [CrossRef]
  61. McIntosh, K.; Cotsell, J.; Cumpston, J.; Norris, A.; Powell, N.; Ketola, B. An optical comparison of silicone and EVA encapsulants for conventional silicon PV modules: A ray-tracing study. In Proceedings of the 34th IEEE Photovoltaic Specialists Conference, Philadelphia, PA, USA, 7–12 June 2009. [Google Scholar]
  62. DuPont Tedlar PVF General Properties. 2014. Available online: http://www.dupont.com/content/dam/dupont/products-and-services/membranes-and-films/pvffilms/documents/DEC_Tedlar_GeneralProperties.pdf (accessed on 1 September 2016).
  63. Ju, X.; Wang, Z.; Flamant, G.; Li, P.; Zhao, W. Numerical analysis and optimization of a spectrum splitting concentration photovoltaic-thermoelectric hybrid system. Sol. Energy 2012, 86, 1941–1954. [Google Scholar] [CrossRef]
  64. Houcine, A.; Maatallah, T.; El Alimi, S.; Nasrallah, S.B. The Performance Study of Parabolic Trough Concentrator Using a New RT3D-4R Method. Int. J. Control. Theory Appl. 2016, 9, 133–139. [Google Scholar]
  65. Maatallah, T.; Houcine, A.; El Alimi, S.; Nasrallah, S.B. Optical modeling and investigation of sun tracking parabolic trough solar collector basing on Ray Tracing 3Dimensions-4Rays. Sustain. Cities Soc. 2017, 35, 786–798. [Google Scholar]
  66. Taher, M.; Ahlem, H.; Souheil, E.A.; Sassi, B.N. A novel solar concentrating system based on a fixed cylindrical reflector and tracking receiver. Renew. Energy 2018, 117, 85–107. [Google Scholar]
  67. Maatallah, T.; Ammar, R. Design, modeling, and optimization of a dual reflector parabolic trough concentration system. Int. J. Energy Res. 2020, 44, 3711–3723. [Google Scholar] [CrossRef]
  68. Evans, D. On the Performance of Cylindrical Parabolic Solar Concentrators with Flat Absorbers. Sol. Energy 1977, 19, 379–385. [Google Scholar] [CrossRef]
  69. Jeter, S.M. The distribution of concentrated solar radiation in paraboloidal collectors. J. Sol. Energy Eng. 1986, 108, 108–219. [Google Scholar] [CrossRef]
  70. Soteris, A.K.; Constantinos, C.N.; Christos, N.S. A comparative study of methods for estimating intercept factor of parabolic trough collectors. In Proceedings of the Engineering Applications of Neural Networks (EANN’96) Conference, London, UK, 17–19 June 1996. [Google Scholar]
Figure 1. Schematics of the proposed system.
Figure 1. Schematics of the proposed system.
Sustainability 16 00544 g001
Figure 2. Different cooling pipe cross-sections considered in this study: (a) semicircle, (b) rectangle, (c) semi-ellipse, and (d) triangle.
Figure 2. Different cooling pipe cross-sections considered in this study: (a) semicircle, (b) rectangle, (c) semi-ellipse, and (d) triangle.
Sustainability 16 00544 g002
Figure 3. RT3D-4R flowchart diagram.
Figure 3. RT3D-4R flowchart diagram.
Sustainability 16 00544 g003
Figure 4. Grid mesh distribution generated over the conical cavity receiver and PV cells area.
Figure 4. Grid mesh distribution generated over the conical cavity receiver and PV cells area.
Sustainability 16 00544 g004aSustainability 16 00544 g004b
Figure 5. An illustration of the different temperature and heat transfer terms.
Figure 5. An illustration of the different temperature and heat transfer terms.
Sustainability 16 00544 g005
Figure 6. (a): Distribution (LCR) at 45° rim angle and 60° rim angle for a sun subtend angle of 16 arcminutes, and (b): Distribution (LCR) at 45° rim angle and 60° rim angle for a sun subtend angle of 32 arcminutes.
Figure 6. (a): Distribution (LCR) at 45° rim angle and 60° rim angle for a sun subtend angle of 16 arcminutes, and (b): Distribution (LCR) at 45° rim angle and 60° rim angle for a sun subtend angle of 32 arcminutes.
Sustainability 16 00544 g006
Figure 7. CSD distribution (a) and Temperature distribution on the CCR (b).
Figure 7. CSD distribution (a) and Temperature distribution on the CCR (b).
Sustainability 16 00544 g007
Figure 8. CSD distribution on the PV solar cell area (a) and temperature trending: (b) upper-side view and (c) rear-side view.
Figure 8. CSD distribution on the PV solar cell area (a) and temperature trending: (b) upper-side view and (c) rear-side view.
Sustainability 16 00544 g008aSustainability 16 00544 g008b
Figure 9. Heat flux distribution on the different serpentine-shape embedded cooling pipes.
Figure 9. Heat flux distribution on the different serpentine-shape embedded cooling pipes.
Sustainability 16 00544 g009aSustainability 16 00544 g009b
Figure 10. Temperature distribution on the different serpentine shapes of the cooled pipes.
Figure 10. Temperature distribution on the different serpentine shapes of the cooled pipes.
Sustainability 16 00544 g010
Table 1. Geometrical parameters.
Table 1. Geometrical parameters.
ItemSymbolValues [m]
Parabolic reflectorAperture radiuswc2.5
Focal lengthf3
Conical cavity receiverUpper radiusrmax0.1
Lower radiusrmin0.05
Beam splitterAperture radiuswsf0.1
Focal lengthfs0.05
PV cellsAperture widthwsc1.4
Table 2. Optical and thermophysical properties of the different PV panel layers.
Table 2. Optical and thermophysical properties of the different PV panel layers.
GlassEVACellBacksheet
MaterialSilica glassElvax 250 (28% VA, 25 MI)polycrystalline cellThiophene-Benzothiazole
Width (m)1.41.40.21.4
Length (m)1.41.40.21.4
Thickness (m)0.00320.00170.00020.02
Refractive index1.451.49 [61]4.0521.46 [62]
Thermal conductivity (W/m K)1.830.15600.15
Specific heat capacity (J/Kg K)73020903201250
Density (Kg/m3)220395053231170
Table 3. Channel/pipe cross-section geometry and dimensions for the same cross-sectional area.
Table 3. Channel/pipe cross-section geometry and dimensions for the same cross-sectional area.
NoGeometryDimensionsWidth
w (mm)
Perimeter *
p (mm)
Depth
d (mm)
Hydraulic Diameter
Dh (mm)
1.Sustainability 16 00544 i001Sustainability 16 00544 i00245.14141.822.5722.6
2.Sustainability 16 00544 i003Sustainability 16 00544 i00450.00132.016.0024.2
3.Sustainability 16 00544 i005Sustainability 16 00544 i00663.66158.316.0020.2
4.Sustainability 16 00544 i007Sustainability 16 00544 i00880.00169.420.0018.9
* Perimeter p is the sum of length of the sides of the channel minus the top side or width w.
Table 4. Boundary conditions common to all cases.
Table 4. Boundary conditions common to all cases.
ParameterValue
Pipe wall materialCopper
Pipe wall thickness (mm)0.5
Pipe inlet/outlet areas (m2)0.0008
Pipe length (m)12.637
Inlet temperature (K)298.0
Inlet flow velocity (m/s)0.125
Inlet mass flow rate (kg/s)0.10
Air temperature (K)298.0
Heat flux (W/m2)6400.0
Material surface emissivity0.9
Table 5. Temperature and thermal transfer rate description for the top, bottom, sides, and underneath of the sheet of the channel.
Table 5. Temperature and thermal transfer rate description for the top, bottom, sides, and underneath of the sheet of the channel.
Geometry T o u t
(K)
T
(°C)
T c h , t o p
(K)
T c h , s i d e s
(K)
T b a c k
(K)
q ˙ c h , t o p
(kW)
q ˙ c h , s i d e s
(W)
q ˙ b a c k
(W)
q ˙ o u t
(kW)
Sustainability 16 00544 i009326.228.2341.28311.93344.7611.690−136−85511.491
Sustainability 16 00544 i010326.628.6339.35311.72342.7011.765−154−77911.548
Sustainability 16 00544 i011327.929.9334.65312.10337.6011.941−149−60311.730
Sustainability 16 00544 i012328.330.3330.29313.82333.1012.098−195−44611.834
Table 6. A comparative overview of previous studies.
Table 6. A comparative overview of previous studies.
AuthorsNumerical/ExperimentalConcentrator Solar System DeviceType of Cooling SystemOptical Concentration RatioHeat Transfer FluidΔT (°C)Efficiencies/Electrical or Thermal PowerMain Outcomes/Milestones
Akbarzadeh et al. [17]Numerical & experimentalOne-axis tracked east-west parabolic trough concentratorThermosyphon external to the structure20 sunsWater & R-1121Electrical power = 20.6 WThe performance improvement of the proposed concentrating solar system is caused by enhancement of the primary optical stage properties by the effectiveness of the cooling mechanism.
Othman et al. [20]Numerical & experimentalParabolic concentratorFins peripheral to the structure1.95 sunsAir18
  • Thermal efficiency = 70%
  • Electrical efficiency = 8%
The surge in the cogeneration effectiveness of the system.
Hedayatiza-deh et al. [21]NumericalCompound parabolic concentratorDuct external to the system 2 sunsWater7.7
  • Thermal efficiency = 51.4584%
  • Electrical efficiency = 9.6113%
The investigation of the thermal and electrical performance of the hybrid PV/T-system-based compound parabolic concentrator.
He et al. [22]Experimental Non-imaging-based-diffuse-reflection PV/T concentrator solar systemChannel exterior to the arrangement≤2 sunsWater41Average output electrical power = 22.9 WThe proposed optical design boosted the productivity of the CPV/T system.
Koronaki et al. [24]Numerical & experimentalCompound parabolic concentratorDuct external to the system1.414 sunsWater7.8
  • Useful energy = 2278 W
  • Efficiency = 30%
  • Enhanced optical design of the hybrid solar collectors with flat-plate receiver.
  • Long-term collected data and performance analysis.
Wang et al. [25]ExperimentalParabolic trough concentratorChannel external to the system1.25 sunsWater0–30
  • Fraction of solar energy = 60% of annual hot water load
  • Net electrical output = 60% of the annual electric loads
Lower cell temperature by the implementation of a spectral beam splitter, higher steam production, and more significant hot water supply.
Felsberger et al. [26]Numerical & experimentalParabolic trough concentratorChannel900Water/glycol20.5
  • Electrical power = 3.87 W
The potential to combine the simultaneous generation of both heat and electricity by retrofitting traditional parabolic collector which is typically only employed in thermal systems with multijunction solar cells.
Renno et al. [27,28]Numerical & experimentalParabolic trough concentratorChannel100Water28–56
  • Thermal energy = 23 kWh–63 kWh
  • Electrical energy = 33 kWh–50 kWh
When considering energy options, an integrated CPV/T-ORC system proves to be the best choice for a commercial consumer.
Nasserian et al. [29]Numerical & experimentalCompound parabolic concentratorChannel1.52 Water35.1
  • Electrical power = 142.2 W
  • Thermal power = 511 W
The improvement of the thermal and electrical production of a Solarus Power Collector.
Gakkhar et al. [30]Numerical & experimentalParabolic trough concentratorPipe6Water28.7
  • Thermal efficiency = 49.48%
  • Electrical efficiency = 12%
The evaluation of the performance of hybrid system in terms of respective efficiencies.
Ustaoglu et al. [31]NumericalCompound hyperbolic trumpet concentratorPipe1.94Water1.2
  • Power output = 42.9%
Offer the CPV solution’s benefits in order to attain a more affordable design and superior performance.
Lin et al. [32]NumericalParabolic dish concentratorChannel10Water30
  • Exergy efficiency = 32–33%
The suggested system provides variable cogeneration of electricity and cooling.
Gomaa et al. [33] Numerical & experimentalFresnel flat mirror concentratorPipe3 suns Water8–13.5
  • Electric power = 170 W
  • Thermal power = 580 W
Both electrical and thermal efficiency was increased by raising the cooling water capacity in both setups. This maximized the heat recuperation from the photovoltaic module (PV).
Gabral et al. [34]ExperimentalParabolic trough concentratorChannel2Water25
  • Electrical efficiency = 8%
  • Thermal efficiency = 58.3%
Moving the highest power of energy at normal incidence angles to more extreme incidence angles.
Hen et al. [35]MathematicalFlat mirror concentratorChannel exterior to the system24 sunsLiquid 23.9
  • Optical efficiency = 84.82%
  • Overall efficiency = 46.77%
Enhancement of the optical behavior and overall system performance due to the optimized beam filter coating characteristics.
Gorouh et al. [36]Numerical & experimentalParabolic trough concentratorChannel 2.7Water40
  • Electrical efficiency = 6.1%
  • Thermal efficiency = 69.6%
Finding the essential CPVT collector design parameters for additional improvements.
Present workNumericalHybrid photovoltaic/parabolic dish concentratorVariously sized and positioned channels that are incorporated onto the PV rear sheet6.4 sunsWater29–33
  • Optical efficiency = 67%
  • Thermal power = 11.834 kW
PV module back sheet with 3D-printed and embedded channels of dissimilar cross-sections and positions with respect to the heat transfer source.
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

Maatallah, T.; Houcine, A.; Saeed, F.; Khan, S.; Ali, S. Simulated Performance Analysis of a Hybrid Water-Cooled Photovoltaic/Parabolic Dish Concentrator Coupled with Conical Cavity Receiver. Sustainability 2024, 16, 544. https://doi.org/10.3390/su16020544

AMA Style

Maatallah T, Houcine A, Saeed F, Khan S, Ali S. Simulated Performance Analysis of a Hybrid Water-Cooled Photovoltaic/Parabolic Dish Concentrator Coupled with Conical Cavity Receiver. Sustainability. 2024; 16(2):544. https://doi.org/10.3390/su16020544

Chicago/Turabian Style

Maatallah, Taher, Ahlem Houcine, Farooq Saeed, Sikandar Khan, and Sajid Ali. 2024. "Simulated Performance Analysis of a Hybrid Water-Cooled Photovoltaic/Parabolic Dish Concentrator Coupled with Conical Cavity Receiver" Sustainability 16, no. 2: 544. https://doi.org/10.3390/su16020544

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