Active Forced Convection Photovoltaic/Thermal Panel Efficiency Optimization Analysis by Bradley Fontenault An Engineering Project Submitted to the Graduate Faculty of Rensselaer Polytechnic Institute in Partial Fulfillment of the Requirements for the degree of Master of Engineering in Mechanical Engineering Approved: _________________________________________ Ernesto Gutierrez-Miravete, Engineering Project Advisor Rensselaer Polytechnic Institute Hartford, CT April, 2012 © Copyright 2012 by Bradley Fontenault All Rights Reserved ii CONTENTS Active Forced Convection Photovoltaic/Thermal Panel Efficiency Optimization Analysis……………………………………………………………………………….i LIST OF TABLES ............................................................................................................ iv GLOSSARY ...................................................................................................................... v LIST OF FIGURES .......................................................................................................... vi NOMENCLATURE ........................................................................................................ vii ACKNOWLEDGMENT .................................................................................................. ix ABSTRACT ...................................................................................................................... x 1. Introduction.................................................................................................................. 1 1.1 Background ......................................................................................................... 1 1.2 Project Scope ...................................................................................................... 6 2. Methodology/Approach ............................................................................................... 7 2.1 PV/T Test Model and Arrangement ................................................................... 7 2.2 PV/T Panel Thermal Model ................................................................................ 8 2.2.1 Assumptions ........................................................................................... 8 2.2.2 Theory and Governing Equations .......................................................... 9 2.3 Test Cases ......................................................................................................... 12 2.4 Meshing ............................................................................................................ 14 2.5 Material Properties............................................................................................ 15 2.6 Solving .............................................................................................................. 16 3. Results and Discussion .............................................................................................. 18 4. Conclusions................................................................................................................ 29 5. References.................................................................................................................. 31 Appendix A: Test Case Supplementary Data .................................................................... 1 iii LIST OF TABLES Table 1: Summary of test cases ....................................................................................... 13 Table 2: Input data for test cases ..................................................................................... 14 Table 3: PV/T panel materials and study-relevant properties ......................................... 15 iv GLOSSARY EVA Ethylene Vinyl-Acetate FEA Finite Element Analysis FGM Functionally Graded Material COP Coefficient of Performance CPV Concentrated Photovoltaic PV Photovoltaic PV/T Photovoltaic and Thermal SQ Shockley-Queisser v LIST OF FIGURES Figure 1: Solar irradiance per day given in kWh per square meter per day [5] ................ 3 Figure 2: PV/T system proposed by Yang et al. [4] .......................................................... 5 Figure 3: Conceptual PV/T design analyzed in this study ................................................. 5 Figure 4: PV/T solar panel simulation test set-up ............................................................. 8 Figure 5: Thermal model of a PV panel [12] ................................................................... 10 Figure 6: PV/T solar panel meshed in COMSOL using the physics controlled mesh sequence ................................................................................................................... 15 Figure 7: Laminar flow profile common to all test cases ................................................ 19 Figure 8: Test Case “1a” two-dimensional surface plot of temperature at steady state .. 20 Figure 9: Test Case “2a” two-dimensional surface plot of temperature at steady stat ... 20 Figure 10: Test Case “3a” two-dimensional surface plot of temperature at steady state 21 Figure 11: Test Case “1d” two-dimensional surface plot of temperature at steady state 21 Figure 12: Water Velocity Impact to Average PV/T Cell Surface Temperature – Inlet Water Temperature, Tin = 298.15 K ........................................................................ 23 Figure 13: PV/T Average Surface Temperature Impact to Average PV/T Output Efficiency – Inlet Water Temperature, Tin = 298.15 K ........................................... 24 Figure 14: Water Flow Velocity Impact to Average Water Outlet Temperature – Inlet Water Temperature, Tin = 298.15 K ........................................................................ 25 Figure 15: Water Flow Velocity Impact to Average PV/T Thermal Efficiency – Inlet Water Temperature Tin = 298.15 K ......................................................................... 26 Figure 16: PV/T Panel Average Total Efficiency vs. Flow Velocity of Coolant Water – Inlet Water Temperature, Tin = 298.15 K ................................................................ 27 vi NOMENCLATURE Symbol A A flowpath Cp C pwater Dh Ein E pv Ewater hc, forced I mp k mwater Pout Q q'' q''' qconv '' qheat qin qlw qout '' qsw qsw '' qrad Re T Tamb Tin Tout Tpv Tref t Uwater u Description Units 2 PV cell Area m Cross-sectional area of the reservoir (assuming unit m2 depth) Heat capacity at constant pressure J/(kg*K) Specific heat of water 4.18 J/(g*C) Hydraulic diameter Total solar energy into the cell Electrical energy m W W Thermal energy extracted by water Forced convection heat transfer coefficient W W/m2 Maximum power point current Thermal conductivity Mass of water Electrical power out of PV cell Flow rate of water Heat flux Heat source Convective heat loss Heat flux going into PV cell A W/(m*K) g W ml/s W/m2 W/m2 W W/m2 Heat or energy into PV cell Longwave energy Heat or energy out of PV cell Shortwave energy flux W W W W/m2 Shortwave energy Solar irradiance W W/m2 Reynolds Number Temperature Ambient temperature Water inlet temperature Water outlet temperature Temperature of PV cell [] K K K K K Reference temperature K Time Inlet water flow velocity Fluid velocity seconds m/s m/s vii Maximum power point voltage Temperature coefficient at reference temperature V 1/K e h pv Surface emissivity Average cell electrical efficiency [] [] h pv htot hth hTref m r PV cell electrical efficiency [] Total PV/T panel efficiency Thermal efficiency of the cell PV cell electrical efficiency at reference temperature [] [] [] Dynamic viscosity Density Stefan-Boltzmann constant Kinematic viscosity of water at 25 C Pa*s kg/m3 5.67e-8 W/(m2*K4) 9.137e-7 (m2/s) Vmp bref s u viii ACKNOWLEDGMENT I would like to thank Rensselaer Polytechnic Institute for the resources and professionals that have provided me with an invaluable education that I plan on using throughout the rest of my career as an engineer. I would like to especially thank Dr. Ernesto GutierrezMiravete for all his help and patience he provided throughout the completion of this project. Also, this couldn’t have been possible without the financial support of Electric Boat Corporation. Most importantly, I would like to give my upmost thanks to my family and to my wife, Katelyn, for providing endless and unconditional support throughout my education program at RPI. ix ABSTRACT The electrical output efficiency of a photovoltaic (PV) cell, similar to any other semiconductor, is negatively impacted as its temperature increases; however, PV cells must be arranged in direct sunlight to produce electricity. To optimize the capability of PV panels and decrease accelerated PV cell wear by controlling their operating temperatures measures must be taken. A conceptual design for a photovoltaic thermal (PV/T) solar panel has been developed and analyzed to control the inherent temperature increase of PV cells to increase electrical efficiency, while also carrying the absorbed heat away from the panels for a myriad of applications. The cooling system consists primarily of a thin rectangular aluminum reservoir that is mounted to the backside of PV/T panels through which water flows. Several water flow rates and reservoir thicknesses were analyzed to determine which combinations of these factors produced optimal PV/T panel cooling results, thus the greatest PV/T panel electrical efficiency. Thermal efficiency of the PV/T panel was also calculated and compared for the various combinations of flow velocity and reservoir thickness. The total panel efficiency is resultantly an additive efficiency of thermal and electrical efficiencies. It was found that higher total efficiencies were achieved in configurations utilizing the highest flow rates and largest reservoir thickness. However, high flow rates translated to minimal net temperature differences between the PV/T panel inlet and outlet, which is undesirable for practical use of the coolant water in secondary applications mentioned above. Heat transfer between the PV/T panels and the cooling reservoirs was modeled and simulated in COMSOL. x 1. Introduction 1.1 Background Global warming, rising fuel oil prices, and other environmental factors in the world today are influencing organizations to develop green energy technologies to be used by corporations and individual consumers. One such technology was invented in 1894 by Charles Fritts and is referred to as the photovoltaic (PV) cell [1]. PV cells convert sunlight into electrical energy, which can be harnessed and supplied to the electrical grid. The first PV cell created by Fritts was only 1% efficient; however, modern day single junction solar cells have efficiencies up to 23% [2], which is approaching the Shockley Queisser (SQ) Limit of 33% [3]. The SQ Limit was developed in 1961 by William Shockley and Hans Queisser and represents the maximum theoretical efficiency of a perfect solar cell using a p/n junction to extract electrical power at standard test conditions. New PV technologies have been shown to improve energy utilization efficiency, such as multi-junction cells, optical frequency shifting, and concentrated photovoltaic (CPV) systems, among others; however, these technologies are expensive both in acquisition and maintenance costs and resultantly have not been utilized on a wide scale at this time [4]. By visual inspection of the aforementioned SQ limit, most (67%) of solar energy is lost, or unable to be harnessed to produce usable electricity. Of this 67% of lost energy, approximately 47% can be attributed to solar energy being converted to heat, which directly correlates to an increase in PV cell operating temperature. Approximately 18% of photons pass through the PV cell, 2% of the solar energy is lost to other factors, and the remaining 33% of solar energy accounts for maximum theoretical energy that can be converted to electricity [3]. It should be noted that for the statements above to be valid, it is assumed that there is only one semiconductor material be used per cell; only one P/N junction per cell is employed; the sunlight is not concentrated or magnified; and all solar energy is converted to heat from photons greater than the band gap of the semiconductor 1 material be utilized. Multi-junction PV cells, CPV systems, and other advanced PV systems can exceed the SQ limit, thus it is not applicable to those types of systems. The term band gap refers to the energy difference between the top of valence (outer electron) band and the bottom of the conduction (free electron flow) band. In the valence band, electrons are tightly held in their orbits by the nuclear forces of a single atom. In the conduction band, electrons have enough energy to move freely and are not tied to any one atom. Materials with a small band gap which behave as insulators (large band gaps) at absolute zero but allow excitation of electrons into their conduction bands (at temperatures below their melting point) are called semiconductors. Electrical voltage is created when electrons flow from one band to the other (valence to conduction), which is brought upon by an external force: photons from sunlight with the energy required to cross the band gap [3]. Photons with energy less than the band gap will pass through the PV cell. Photons with energy greater than the band gap energy do excite electrons and make them flow to produce usable voltage, however, excess photon energy is dissipated to generate heat. PV cell semiconductor materials are chosen to be able to absorb most of the solar light spectrum as possible, in order to generate the most amount of electricity possible, which results in a low band gap. However, materials with larger band gaps produce more voltage, which consequently relates to a narrower spectrum of sunlight that can be used. Therefore, a balance between the amount of usable sunlight and voltage output must be found when choosing PV cell semiconductor materials. PV cell semiconductor materials with smaller band gaps may be more subject to higher operating temperatures, because the energy content of photons above the material band gap will be generated into heat as a byproduct [3]. Most solar cells today are made from silicon, the second most abundant element in the earth's crust, which happens to have one of the lowest band gaps among other materials used to make PV cells [2]. In commercial and residential applications, PV cells are assembled into modules, which are then assembled into panels. PV panels are then assembled to form arrays. The most 2 applicable regions to use PV panels are in environments with plentiful amounts of sun exposure, which most often are regions with warm climates (i.e. high ambient temperatures). This is especially true for the United States, which is shown by Figure 1 below. Figure 1 graphically shows the average daily solar radiation in Kilowatt-hours per square meter per day that falls on the United States. Figure 1: Solar irradiance per day given in kWh per square meter per day [5] PV panel temperature increases due to low solar energy-to-electricity efficiencies because not all energy absorbed by PV cells can be converted to electrical energy; therefore, to satisfy conservation of energy laws, the remaining energy must be converted to heat. Also, regions with warm climates (i.e. high ambient temperatures), where solar energy potential is at its greatest as shown above, reduce the efficiency of PV panels by limiting the heat dissipation from the panels. Therefore, it is relevant and essential to develop methods of cooling the PV cells to increase output efficiency. This is especially true for CPV systems, which essentially magnify the solar illumination being received by PV panels [6]. The temperature of adequately ventilated CPV systems 3 can reach upwards of 70 degrees Celsius, which can be harmful to the PV panels and reduce their operating lifetime. Oh et al. [7] has found that poorly ventilated PV panels in environments with high ambient temperatures can reach temperatures greater than 90 degrees Celsius. These conditions can be extremely dangerous to people who may be in the vicinity of the panels and can also decrease the service life of the panels. Both active and passive methods of cooling PV panels have been researched and analyzed to date. Conclusions of these studies have been made that controlling the temperature increase of PV panels results in significant gains in electrical output of the panels. In many of these investigations, the thermal energy extracted from the PV panels has been utilized for a variety of low temperature applications (i.e. residential water heating, radiant floor heating, swimming pool heating, etc.). These systems are referred to as hybrid photovoltaic and thermal (PV/T) systems. Teo et al. [8] investigated an active PV panel cooling system, in which the PV panels were cooled by forced convection, with air being the heat carrying fluid, and resulted in a 4-5% efficiency increase. Chen et al. [9] developed a hybrid PV/heat pump system using refrigerant fluid R134a as the heat carrying fluid. The coefficient of performance (COP) of the heat pump and electrical efficiency of the PV panel was measured at different condensing fluid flow rates and temperatures. Bakker et al. [10] analyzed a PV/T panel array where heat was extracted from the panels and stored underground in a heat exchanger. In winter, the heat could be extracted from the ground via a heat pump and used to heat potable water and support a floor heating system while increasing the electrical efficiency of the solar panel. The largest PV/T solar panel installation in the United States was brought online in February 2012 in Rhode Island at Brown University’s Katherine Moran Coleman Aquatics Center. In this application, 168-patented PV/T panels were installed on the center’s roof, which under full sun exposure can provide enough electricity to light the building and heat the one million gallon swimming pool inside of it [11]. Yang et al. [4] developed a functionally graded material (FGM) with copper water pipes cast into it that was bonded with thermal paste to the backside of a PV/T panel. Cooling 4 water was pumped at various flow rates through the cast copper pipes to decrease the PV/T panel temperature thus increasing PV/T panel electrical efficiency by up to 2%. This study also analyzed the thermal efficiency of the FGM/copper tube design, and reported that a combined thermal and electrical efficiency of 71% could be achieved, compared to 53-68% total efficiency of other PV/T concepts. The system design proposed by Yang et al. is shown below in Figure 2. Figure 2: PV/T system proposed by Yang et al. [4] This study leverages off work previously accomplished by Yang et al. [4], by utilizing the identical PV/T panel and thermal glue properties; however, a thin rectangular reservoir through which water will flow through and carry heat away from the panel will be developed and analyzed in lieu of the FGM with cast copper water coolant pipes explored by Yang as shown below in Figure 3. This is accomplished in COMSOL Multiphysics finite element software package. Increases in electrical efficiency of the panel at various operating temperatures will be calculated from the subject PV/T panel thermal coefficient, bref . Figure 3: Conceptual PV/T design analyzed in this study 5 1.2 Project Scope The objective of this study is to investigate a preliminary PV/T panel design using finite element analysis (FEA) to determine the anticipated gains in electrical efficiency that can be achieved, as well as to quantitatively determine the amount of thermal energy that can be captured. The subject design includes a reservoir that is mounted to the backside of a predefined PV/T panel through which water will flow and carry heat energy away from the panel. The electrical efficiency of the PV/T panel is expected to increase as the cell operating temperature decreases, while the thermal energy carried away from the PV/T panel be utilized for numerous applications (i.e. heating a swimming pool, heating potable water, use in radiant floor heating systems, etc.). The scope of this project will be limited to the analysis of the heat transfer between the PV panel and the reservoir bonded to the panel, the resulting increase in PV/T panel electrical efficiency, and the thermal energy extracted from the panel. The total PV/T panel efficiency will be calculated as a function of both the electrical and thermal efficiency of the panel. This study will not consider the entire recirculating water system that delivers water to the panel reservoir and carries the heated water away from the panel; this project rather, is focused primarily on the affect to electrical efficiency of the PV/T panel and the thermal energy that can be extracted from it. 6 2. Methodology/Approach In this project, a single PV/T panel will be evaluated using COMSOL Multiphysics FEA software, from which results could be extrapolated for an array of identical PV/T panels. An aluminum reservoir will be modeled in COMSOL for the subject PV/T panel, through which water at a predetermined inlet temperature will flow. Three different reservoir thicknesses will be analyzed to determine the impact they have on cooling the panel. As this study investigates forced convection using water as the coolant, or heat carrying fluid, various water flow rates will also be analyzed for impact to cooling performance with qualitative consideration given to the size of pump needed for the flow rates (greater size correlates to more electricity needed to operate pump). 2.1 PV/T Test Model and Arrangement As mentioned in Section 1.1, work being done in this study will utilize the PV materials defined in the work done by Yang et al. [4], and will use the uncooled PV/T panel data as a reference point since an experimental analysis of the subject PV/T design was not feasible for this project. The PV cells making up the PV panel assemble into an approximate length, width, and thickness of 30.5 cm X 30.5 cm X 0.27 mm, respectively. For simplicity, it is assumed that the whole panel is covered in PV cells, with no packing material (material used to fill in gaps between the cells on a panel). The PV cells are commercial grade monocrystalline silicon cells with electrical efficiency, hT , of 13% and have a thermal coefficient, bref , of 0.54% [1/K] [4]. The thermal ref coefficient represents the degradation of PV cell output per degree of temperature increase. The cooling reservoir is bond to the back of the panel by a thermal paste with an approximate uniform thickness of 0.3 mm over the whole surface area of the reservoir – PV panel interface. The reservoir walls are approximately 1 mm in thickness and are constructed from aluminum. Aluminum was chosen as the reservoir enclosure material, due to its high thermal conductivity, which promotes heat transfer across its boundary, its availability, and its relatively low cost in comparison to other conductive metals such 7 as copper. A cross section view of the assembly is shown below in Figure 4. Similar to Yang et al. [4], it is also assumed that the conceptual PV/T panel being considered in this project is not coated with a protective glass and/or a layer(s) of ethylene vinylacetate (EVA), both of which are typically used to protect PV cells in real world applications. However, while protecting the delicate silicon PV panels, these encapsulation materials hinder the performance of PV panels by affecting the panel’s absorptivity of solar irradiance. Teo et al. [8] found that the highest temperatures experienced in a PV panel are on the backside of the panel due to the high thermal conductivity of the silicon PV material; therefore, precedence exists for cooling the panel from the backside rather than using water to cool the panel on the topside. Figure 4: PV/T solar panel simulation test set-up 2.2 PV/T Panel Thermal Model 2.2.1 Assumptions Several assumptions must be made to perform this study regarding the conceptual PV/T panel construction, atmospheric conditions, water flow characteristics, and other factors, which impact this thermal analysis. 8 1. The solar irradiance imparted on the entire surface of the PV/T panel is 1000 W/m2. 2. All solar irradiance that is not used to produce electricity in the PV/T panel will be developed into heat. 3. The subject conceptual PV/T panel is not constructed with a glass and/or EVA encapsulating layer, which would result in decreased PV/T absorption of solar irradiance. 4. No dust or any other agent is deposited on the PV/T surface affecting the absorptivity of the PV/T panel. 5. The coolant water at the inlet of the conceptual PV/T reservoir will have uniform temperature. 6. The flow through the coolant reservoir is considered to be fully laminar and incompressible. 7. The ambient temperature surrounding the PV/T panel is 298.15 K. 8. An average wind speed of 1 m/s exists throughout the simulations. 2.2.2 Theory and Governing Equations All three modes of heat transfer are involved when considering a basic PV panel. Heat is transferred within the PV cell and its structure by conduction and heat is transferred to the PV/T panel surroundings by both free and forced convection. Heat is also removed from the panel in the form of long-wave radiation [12]. Heat transfer by conduction to the panel structural framework is often ignored due to the small area of contact points; however, it will be considered throughout the COMSOL simulations from the PV/T panel surface and through the reservoir casing. Steady state heat conduction through from the PV/T cell surface to the reservoir enclosure casing is given by Equation [1] below. Ñ× (kÑT ) = 0 [1] The PV panel shown in Figure 5 receives energy from the solar irradiance, converts some of it into electricity through the PV effect and the rest is transformed into heat. 9 The objective of attaching the thermal panel underneath the PV cell is to remove as much as this heat as possible in order to increase the efficiency. Figure 5: Thermal model of a PV panel [12] The heat loss due to forced convection on the top and bottom surfaces of a PV cell is given by Equation [2] below [12]. qconv = -hc, forced × A × (Tpv - Tamb ) [2] In this study, forced convection does not only occur on the top surface of the PV panel, but also through the reservoir mounted to the backside of the panel. Therefore, the total convective heat transfer is a combination of the heat transfer at the top and bottom surfaces of the PV/T panel and the heat transfer from the flowing water in the reservoir. The FEA software being used in this study, COMSOL, already contains a nonisothermal laminar flow and conjugate heat transfer physics module, which was used to model conduction heat transport within the cell as well as the convective heat transfer in the water reservoir on the backside of the PV panel. This package is appropriate for this study, because of the inhomogeneous temperature field that is created as water flows from the inlet to the outlet of the reservoir. COMSOL numerically solves the continuity and momentum equations, which are the governing equations for the fluid flow, and are shown below in Equations [3] and [4], respectively [15]. 10 Ñ × (ru) = 0 [3] (( ru × Ñu = -Ñp + Ñ × m Ñu + ( Ñu) T )) [4] The conduction-convection equation is also solved for the heat transfer in the flowing water, which is shown in Equation [5]. rCpu× ÑT = Ñ× ( kÑT ) [5] The longwave radiation heat loss can be calculated from Equation [6] below [12]. 4 qlw = e × s × (Tpv4 - Tamb ) [6] The thermal model analyzed in this study is similar to that modeled by Jones and Underwood [12], although the amount of energy applied to the PV cell that is converted to heat energy is calculated using the method of Kerzmann and Schaefer [13]. The amount of energy converting into electric power in the PV cell is a function of the PV cell efficiency, h pv , as shown below in Equation [7], which satisfies Assumption 2 above. '' '' qheat = qrad × (1- h pv ) [7] The PV cell electrical efficiency, h pv , is given by Equation [8] below as a function of its efficiency at reference temperature, the PV cell temperature, and the PV cell thermal coefficient [14]. h pv = hT éë1- bref (Tpv - Tref )ùû ref 11 [8] The PV cell electrical output efficiency can also be expressed as a function of PV cell power output, solar irradiance, and the PV cell surface area as shown in Equation [9] below [8]. h pv = Vmp I mp '' qrad A [9] In Equation [8], hTref is the PV cell efficiency at reference conditions (i.e. Tref = 25°C , '' qrad =1000 W ), and bref is the PV thermal coefficient. m2 The thermal model of the PV/T system being analyzed is complex, in which the PV cell temperature, PV cell electrical output efficiency, and the amount of energy being converted into electricity and into heat to increase the PV cell temperature are all interrelated. Similar studies such as the study being presented in this work have been performed by Kerzmann and Schaefer [13], who numerically solved a similar PV/T complex system, and Yang et al. [4], who analyzed another type of PV/T system using a different commercial FEA software package, ABAQUS. 2.3 Test Cases Numerous test cases will be simulated in COMSOL, in which water inlet velocity and reservoir thickness will be varied to determine some of the best designs for the PV/T system. A constant ambient temperature of 298.15 K will be used for all test cases to simulate laboratory test conditions and to concentrate the focus of performance impacts of the aforementioned variables. A summary of the test cases selected in this study is shown in Table 1 below. It is shown that for each test case, the flow was confirmed to be laminar to satisfy Assumption 6 by confirming that the dimensionless Reynolds number, Re, was less than 2300. The Reynolds number is given by Equation [10] below. 12 Table 1: Summary of test cases In Table 2 below, the input data used in all test cases are shown. Many of the values shown in the table have been discussed thus far. One however, that has not, is the heat transfer coefficient used for the heat convection occurring at top and bottom of the PV/T panel. Jones and Underwood [12] source values for the forced convection heat transfer coefficient at a nominal wind speed of 1 various sources, which range from 1.2 m on the top surface of a PV panel from s W W to 9.6 2 . Similar correlations were 2 m ×K m ×K made for heat transfers coefficients on the bottom side of a PV panel. An average of the subject values of approximately 6.5 W for forced convection will be used in this m2 × K study for convection occurring at the top and bottom surfaces of the PV/T panel, which satisfies Assumption 8 above. 13 Table 2: Input data for test cases 2.4 Meshing The PV/T solar panel was meshed in COMSOL using the built-in physics controlled mesh sequence setting. As shown in Figure 6 below, the number of mesh elements increase at each boundary so that the heat transfer and flow fields can be resolved accurately. A normal mesh setting was used, which resulted in 15,240 elements being developed in this model. A normal mesh setting was chosen for this study to decrease the physical time of running all of the required simulations in COMSOL, while ensuring that accurate results were obtained. Finer meshes were experimented with but were found to increase simulation time reasonably, while subtle changes were observed in the solutions for cell and cooling water temperatures for the same initial conditions. 14 Figure 6: PV/T solar panel meshed in COMSOL using the physics controlled mesh sequence 2.5 Material Properties The materials used to construct and analyze the conceptual PV/T system shown above in Figure 4, consist of silicon, silicone thermal paste, aluminum, and water for the PV/T cell, binding agent, reservoir enclosure, and coolant fluid, respectively. Below, in Table 3, the properties of these materials that are relevant to this study are given. The values shown for water that are indicated to be temperature dependent are actively provided by COMSOL at each time step while it is performing the computations. Table 3: PV/T panel materials and study-relevant properties 15 2.6 Solving As mentioned before, the FEA software package, COMSOL, was used to simulate and solve the flow and heat transfer model described thus far using various equations defined in Section 2.2 of this paper. All of the simulations run were steady state studies solved in two dimensions as shown by Figure 6, in which the conjugate heat transfer and laminar flow physics modules were utilized. It was verified that all the flow velocities used would produce laminar flows, rather than turbulent flows, by calculating the Reynolds number, Re, from Equation [10] shown below. For this study, the flow in the channel can be characterized by flow between parallel planes, in which the hydraulic diameter, Dh , becomes twice the plate spacing [16]. Re = Uwater Dh [10] u The software modeled the flow through the PV/T reservoir by solving the continuity, momentum, and energy equations. At each iteration in the simulations performed, the PV cell efficiency, h pv , is calculated from Equation [8] from the user input values for bref , hTref , Tref , and from the COMSOL solved value for the cell temperature, Tpv . The '' '' amount of solar irradiance, qrad , that transforms into heat, qheat , is then calculated from Equation [7]. Similarly, values for the thermal efficiency of the cell were calculated iteratively by COMSOL using Equations [11], [12], and [14] below. Convergence of the steady state solution was monitored throughout the simulation, which on average, converged in approximately 30-40 seconds (real time). Post processing of the data recorded in the simulations is required to calculate the thermal efficiency, hth , of the PV/T panel [4]. First, the total amount of energy (solar irradiance) into the cell must be calculated, which is given by Equation [11] below. '' Ein = qrad ×A [11] 16 Next, the thermal energy extracted by the water per second must be calculated from Equation [12]. · Ewater = mwater Cpwater (Tout - Tin ) [12] The mass flow rate of the water passing through the reservoir can be calculated from the density and flow rate of the water, assuming unit depth of the reservoir. For threedimensional analysis, the depth could be extrapolated to any desired depth. The mass flow rate is given by Equation [13] below. · m = rQ = rUwater A flowpath [13] The thermal efficiency is simply given by Equation [14]. hth = Ewater Ein [14] Similarly, the quantity of the total input energy converted to electrical energy can be approximated from the solution data by obtaining the average electrical efficiency of the PV/T panel, and multiplying it by the total energy into the panel. As mentioned above, the electrical efficiency was calculated in COMSOL at each iteration in the simulations by Equation [8] from which an average of the efficiency for the entire PV cell was obtained at the completion of the simulation. This is shown by Equation [15]. E pv = h pv × Ein [15] The total efficiency of the PV/T panel is then computed from Equation [16]. htot = (E water + E pv ) Ein [16] 17 3. Results and Discussion Using the initial values shown in Table 2, each of the test cases in Table 1 was simulated in COMSOL, which solved the governing equations for the PV/T system discussed in Section 2.2.2, above. Data extracted from COMSOL following each simulation for post processing included the average cell surface temperature, Tpv , average water outlet temperature, Tout , average PV/T panel thermal efficiency, hth , and the average PV/T panel electrical efficiency, h pv . The values for hth after each simulation were also manually calculated in Microsoft Excel using Equations [12] and [14] and the extracted values for Tout . Comparison of the manually calculated values for thermal efficiency and the COMSOL calculated values differed significantly. This is due to the fact that the average water temperature at the outlet was used to calculate the thermal efficiency; however, COMSOL iteratively solved for the thermal efficiency at each iteration in the simulation, which is a more accurate method. For the remainder of this paper, discussion of thermal efficiency values will correspond to values calculated by the software. Using the data extracted from the COMSOL solutions, the results in the following sections were compiled. The simulations performed assumed that the water inlet temperature was of uniform temperature equal to 298.15 K (25 °C), which is the same temperature specified for the ambient temperature. This water temperature was chosen to imitate a scenario in which the cooling water may reach ambient air temperature before entering the cooling reservoir to carry heat away from the PV/T panel. For the remainder of this section, flow reservoir thickness and flow channel thickness are to be considered interchangeable terms. In Figure 7, the velocity profile of the water in the flow channel is shown. A similar laminar flow profile was achieved in each of the test cases and is only shown once here for information. One can see the no-slip boundary condition invoked on the interior walls of the reservoir, and the parabolic flow profile that is created. As expected, the maximum flow velocity is at the center of the flow channel. 18 Figure 7: Laminar flow profile common to all test cases In Figure 8 through Figure 10 below, two-dimensional plots for the steady state solution of the temperature distribution for test cases 1a, 2a, and 3a are shown. As specified in Table 1, the only difference between these cases is the reservoir flow thickness, while the inlet flow velocity is invariable. As shown in the figures below, the reservoir flow thickness has a large impact on the overall temperature of all materials that the PV/T system is comprised of. From inspection of the figure legends, it can be deduced that the greater the flow channel thickness at low flow velocities, the cooler the system will remain. A compilation of temperature gradient plots and additional results data for all test cases run is displayed in Appendix A. As shown below, for the largest flow channel thickness of 0.015 m, the maximum temperature reached on the PV/T surface is approximately 319 K, while the maximum temperature reached for the thinnest flow channel is approximately 327 K. As anticipated, one can see that the water temperature is warmest towards the top of the 19 flow channel and almost linearly decreases as distance from the top of the channel increases. Figure 8: Test Case “1a” two-dimensional surface plot of temperature at steady state Figure 9: Test Case “2a” two-dimensional surface plot of temperature at steady stat 20 Figure 10: Test Case “3a” two-dimensional surface plot of temperature at steady state In Figure 11 below, the temperature gradient across the PV/T system is shown for a high flow velocity test case. In comparison to low velocity temperature gradient plots, one can see the effect velocity has on the temperature gradient of the cooling water. Figure 11: Test Case “1d” two-dimensional surface plot of temperature at steady state This phenomenon is shown further by visual inspection of the raw data extracted from the COMSOL solutions to develop Figure 12, which represents the results obtained from Test Cases “1a” through “3d” shown in Table 1. From Figure 12, it is also inferred that the higher the cooling water velocity, the lower the average PV/T surface temperature will be. This was also expected, since more water flows through the channel at higher 21 velocities. As the water temperature increases in the reservoir, the difference in temperature between the PV/T surface and the water decreases, resulting in decreased heat transfer across the reservoir-thermal paste interface. From the data shown in Figure 12, it is shown that the average PV/T surface temperature can approach temperatures of approximately 318 K, when it is equipped with the narrowest flow channel thickness of 0.005 m at the lowest flow velocity of 0.0002 m/s. However, the surface temperature can be expected to approach 312 K when the PV/T panel is equipped with the largest flow reservoir at the lowest flow velocity analyzed in this study. At the highest flow velocity examined, the PV/T surface temperature actually reaches its lowest value when the narrowest reservoir is utilized. It is also shown that the largest flow channels at high velocities result in higher average surface temperatures, whereas the highest surface temperatures at lower flow rates are obtained using the smallest flow channel. This is due to the physical mass of water in the channel and the amount of time it remains in the channel (i.e. mass flow rate) to absorb heat. Also, the warmer the temperature of the water near the bottom surface of the PV panel, the less of a cooling effect it will have on the panel. Heat cannot be dissipated throughout the smaller channel as well as in a larger flow channel, due to the smaller mass of the fluid in the narrowest flow channel at the slowest flow velocity. At the highest input flow velocity analyzed of 0.01 m/s, it is shown in Appendix A (Test Cases 1d and 3d) that the velocity, especially near the bottom surface of the PV cell layer, is greater in the small channel than in the larger channel. At higher flow velocities, heat from the PV cell can be carried away faster than at slower velocities. It is also shown in Appendix A (Test Cases 1d and 3d), that the amount of heat input into the PV cell for the smaller channel is less than for the larger channel, since the heat input into the cell is a function of PV cell electrical efficiency and PV cell temperature as defined in Section 2.2.2. Therefore, it can be deduced that slightly higher flow internal flow velocities result in higher heat transfer rates and lower surface PV cell temperatures, thus higher cell efficiency and less heat input into the system. Appendix A contains plots of the steady state surface temperature as a function of the length of the PV/T panel for each of the cases simulated. 22 Figure 12: Water Velocity Impact to Average PV/T Cell Surface Temperature – Inlet Water Temperature, Tin = 298.15 K From a cell electrical output efficiency standpoint, the cooler the PV/T surface is, the greater the efficiency will be, which is shown in Figure 13. Results show that cell output efficiency of approximately 11.2% can be expected from the PV/T system with the smallest reservoir thickness with water flowing at the lowest inlet velocity examined. For this case, the output efficiency of the PV/T panel is comparative to the output efficiency of an uncooled (i.e. negligible flow velocity) PV/T panel, which was shown to reach a saturated average temperature of approximately 328 K (55 °C), resulting in an output efficiency of 10.9%. It should be noted that the simulated uncooled average PV/T surface temperature of 328 K corresponds well to the uncooled PV/T surface temperature obtained by Yang et al. [4]. 23 Figure 13: PV/T Average Surface Temperature Impact to Average PV/T Output Efficiency – Inlet Water Temperature, Tin = 298.15 K As previously mentioned, the smallest reservoir thickness at the high inlet velocity results in the lowest average PV/T surface temperature, thus the highest PV/T output efficiency of approximately 12.9%, which approaches the NTOC PV output efficiency of 13%. It should be noted, however, that the difference in cell efficiency and surface temperature at the three higher flow velocities is minor, and a significant difference is only given at the lowest flow velocity of 0.0002 m/s. Individual plots of PV/T output efficiency for each of the cases simulated are shown in Appendix A. A similar trend for the PV/T surface temperature shown in Figure 12 is displayed in Figure 14 for the cooling water outlet temperature. For low water flow velocities and narrow reservoir thicknesses, the outlet temperature is much warmer than for larger flow channels. This is due to the larger volume of fluid in the thicker reservoirs, which naturally heats up slower than the smaller volume in the narrower reservoirs. However, at high velocities the outlet temperature for the narrower flow thickness is slightly cooler than the largest flow thicknesses, although the temperature of the fluid hardly increases from inlet temperature. It is shown that the maximum water outlet temperature obtained was approximately 326 K, which occurred during the simulation of the narrowest flow channel and the slowest flow velocity. In Appendix A, plots are 24 shown for the water outlet temperature as a function of the flow channel thickness for all cases run. Figure 14: Water Flow Velocity Impact to Average Water Outlet Temperature – Inlet Water Temperature, Tin = 298.15 K From Figure 15 it is shown that highest thermal efficiency for any PV/T test case is achieved by the largest reservoir thickness, at the fastest flow velocity. However, as shown above in Figure 14, the temperature increase from the inlet to the outlet of the PV/T panel is minimal. If the user of such a PV/T system prefers higher temperature water, a slower flow velocity would be desired so that a significant change in water temperature could be realized. Even with a slower water velocity, an increase in the PV/T average electrical efficiency is achieved, compared to a PV panel without a cooling feature as shown in Figure 13. A significant change in water temperature from the PV/T panel inlet to the outlet results in high thermal efficiencies for the configurations with flow channel thicknesses of 0.01 m and 0.005 m and at a flow velocity of 0.001 m/s. More heat can be carried away from the cell at the faster flow velocity of 0.001 m/s in comparison to the slowest flow velocity tested, even though a higher change in temperature was obtained. Figure 15 shows that a noticeable dip in the efficiency is obtained for flow channels of 0.01 m and 0.005 m at a flow velocity of 0.005 m/s, which is due to the slight change in temperature from the inlet to the outlet of the flow channel. At a flow velocity of 0.01 25 m/s, the temperature change in the water is still small, however, an increase in the velocity (i.e. mass flow rate) results in an increase in the thermal efficiency of the PV/T system. Figure 15: Water Flow Velocity Impact to Average PV/T Thermal Efficiency – Inlet Water Temperature Tin = 298.15 K In Figure 16 the combined efficiency of the PV/T panel is plotted with respect to the inlet flow velocity, which shows that the largest reservoir thickness, combined with the highest flow velocity is the most efficient option with a water inlet temperature equivalent to the ambient temperature of 298.15 K. As mentioned previously, however, this option does not result in an optimal water outlet temperature that would be useful for a secondary application, such as heating potable water, heating a swimming pool, or other functions. All options, except for the smallest flow channel at the lowest flow velocity, yield a measurable gain in PV/T electrical output efficiency. Also worth noting is the relatively high thermal efficiency of the 0.01 m and 0.015 m flow channel configurations at higher water flow velocities. This is mainly due to Assumptions 1 - 4 stated in Section 2.2.1 and the extreme sensitivity of thermal efficiency to temperature change at high flow velocities. It was conservative to assume that all solar 26 irradiance energy not converted to electrical energy would be developed into heat. It is reasonable to consider that a functional PV/T panel would not absorb a percentage of the solar irradiance imparted on the panel due to the fact some of the solar irradiance is of the incorrect wavelength for any given PV cell material. Also, after being in use for a period of time, the surface of the PV/T panel would likely get dirty or dusty, which would impact the absorptivity of the panel. With that mentioned, a proportional reduction in thermal efficiency could be expected with a decrease in the absorptivity of the panel. Although a different type of PV/T application was analyzed, Kerzmann and Schaefer [13] employed an absorptivity reduction constant of 15% in their work to account for unabsorbed solar irradiance, which would otherwise contribute heat to their application. Figure 16: PV/T Panel Average Total Efficiency vs. Flow Velocity of Coolant Water – Inlet Water Temperature, Tin = 298.15 K The thermal efficiency of the PV/T panel at high flow velocities (i.e. high mass flow rates) was found to be extremely sensitive with calculated values of the difference in water temperature at the inlet and outlet of the panel. As mentioned previously, the average temperature calculated in COMSOL was used to determine the thermal efficiency of the panel. Using Equation [14], it was found that at the highest flow 27 velocity evaluated (0.01 m/s) and largest flow channel thickness (0.015 m), variations in the outlet temperature, Tout , as small as 0.1 K resulted in a 20% difference in thermal efficiency. 28 4. Conclusions In this work, a conceptual photovoltaic thermal panel design was modeled and analyzed using a commercial finite element software package, COMSOL Multiphysics: Version 4.2a. The PV/T panel evaluated consisted of monocrystalline silicon PV cells that were bound with a silicone thermal paste to an aluminum reservoir through which coolant water flowed. Numerous simulations were completed to model the heat transfer across the PV/T panel and ultimately to determine the PV/T electrical output and thermal efficiencies of the panel. Water flow velocity and flow channel thickness were varied and analyzed to determine which combinations yielded not only the highest total PV/T efficiency, but also the most useful thermal and electrical output. It was found that the highest total PV/T panel efficiencies were achieved for test cases involving combinations of high flow velocity and large flow channel thicknesses. The highest total cell efficiency obtained of 95.7% was obtained from Test Case “1d”, in which the flow thickness was 0.015 m and the inlet flow velocity was 0.01 m/s. Test Case “3c” was found to be the least efficient configuration, which recorded a total efficiency of 27.5% and consisted of a flow channel thickness of 0.005 m and inlet flow velocity of 0.005 m/s. The highest efficiency obtained is unrealistic, which is due conservative assumptions and the extreme sensitivity of thermal efficiency to temperature change at high water flow velocities. It was deduced that at high flow velocities (i.e. high mass flow rates), slight changes in temperature result in drastic differences in thermal efficiency; therefore, precise temperature measurement is essential for accurate results. It was also concluded that the PV/T system with the highest efficiency is most likely not the most desirable configuration for practical use. Although high inlet velocities result in the lowest PV/T surface temperatures, thus the highest electrical efficiency, the coolant water exiting the panel experiences no significant temperature change. Therefore, it would not be entirely beneficial to utilize the water exiting the PV/T panel for any practical application, such as heating a swimming pool, for use in a radiant floor 29 heating system, etc. Also, high flow velocities would require larger pumps and more electrical power to operate them depending on the size of the PV/T panel array, which may reduce or negate the electrical gains experienced in such a system. A cost savings study would be required to determine the optimal balance of electrical efficiency and thermal efficiency; however, that is beyond the scope of this study. Future study of this PV/T system could include work evaluating the performance of this system in different climates, utilization of different coolant fluids, and evaluation of various inlet water temperatures. 30 5. References [1] Nelson, Jenny. The Physics of Solar Cells. London: Imperial College Press, 2003. [2] Turner, Wayne and Steve Doty. Energy Management Handbook: Sixth Edition. Lilburn: The Fairmont Press, 2007. [3] “Solar Cell Central: Four Peaks Technologies Inc., Scottsdale, AZ,” Design Copyright 2010, < http://solarcellcentral.com/index.html>. [4] D. J. Yang, Z. F. Yuan, P. H. Lee, and H. M. Yin, Simulation and experimental validation of heat transfer in a novel hybrid solar panel, International Journal of Heat and Mass Transfer 55 (2012) 1076-1082. [5] “Arizona Solar Center, Inc.,” Copyright 1999-2012, <http://www.azsolarcenter.org> [6] Li Zhu, Robert F. Boehm, Yiping Wang, Christopher Halford, and Yong Sun, Water immersion cooling of PV cells in a high concentration system, Solar Energy Materials and Solar Cells 95 (2011) 538-545. [7] Jaewon Oh, Govinda Samy, and Tamizh Mani, Temperature Testing and Analysis of PV Modules Per ANSI/UL 1703 and IEC 61730 Standards, Conference Record of the IEEE Photovoltaic Specialists Conference, p 984-988, 2010, Program - 35th IEEE Photovoltaic Specialists Conference, PVSC 2010. [8] H. G. Teo, P. S. Lee, and M. N. A. Hawlader, An active cooling system for photovoltaic modules, Applied Energy 90 (2012) 309-315. [9] H. Chen, Saffa B. Riffat, Yu Fu, Experimental study on a hybrid photovoltaic/heat pump system, Applied Thermal Engineering 31 (2011) 4132-4138. [10] M. Bakker, H. A. Zondag, M. J. Elswijk, K. J. Strootman, M. J. M. Jong, Performance and costs of a roof-sized PV/thermal array combined with a ground coupled heat pump, Solar Energy 78 (2005) 331-339. [11] Sharp, Rowan. “Brown’s Hybrid Solar Panels a First for RI.” Eco RI News: Rhode Island’s Environmental News Source 26 February 2012. Providence, RI. <www.ecoRI.org>. [12] A.D. Jones and C.P. Underwood, A Thermal Model For Photovoltaic Systems, Solar Energy 70 (2001) 349-359. [13] Tony Kerzmann and Laura Schaefer, System simulation of a linear concentrating photovoltaic system with an active cooling system, Renewable Energy 41 (2012) 254-261. [14] E. Skoplaki and J.A. Palyvos, On the temperature dependence of photovoltaic module electrical performance: A review of efficiency/power correlations, Solar Energy 83 (2009) 614-624. [15] R. Byron Bird, Warren E. Stewart, and Edwin N. Lightfoot, Transport Phenomena: Second Edition, New York: John Wiley and Sons, Inc., 2007. [16] W. M. Kays, M.E. Crawford, and Bernhard Weigand, Convective Heat and Mass Transfer: Fourth Edition, New York: McGraw-Hill International Edition 2005, page 106. 31 Appendix A: Test Case Supplementary Data 1. Test Case 1a Figure A - 1: Test Case 1a two-dimensional temperature plot Figure A - 2: Heat flux into PV/T panel vs. panel length 1 Figure A - 3: Cell electrical output efficiency vs. panel length 2. Test Case 1b Figure A - 4: Steady state solution two-dimensional temperature plot of PV/T panel 2 Figure A - 5: Heat flux into PV/T panel vs. panel length Figure A - 6: Cell electrical output efficiency vs. panel length 3 3. Test Case 1c Figure A - 7: Steady state solution two-dimensional temperature plot of PV/T panel Figure A - 8: Heat flux into PV/T panel vs. panel length 4 Figure A - 9: Cell electrical output efficiency vs. panel length 4. Test Case 1d Figure A - 10: Steady state solution two-dimensional temperature plot of PV/T panel 5 Figure A - 11: Heat flux into PV/T panel vs. panel length Figure A - 12: Cell electrical output efficiency vs. panel length 6 Figure A - 13: Cooling reservoir internal velocity profile 5. Test Case 2a Figure A - 14: Steady state solution two-dimensional temperature plot of PV/T panel 7 Figure A - 15: Heat flux into PV/T panel vs. panel length Figure A - 16: Cell electrical output efficiency vs. panel length 6. Test Case 2b 8 Figure A - 17: Steady state solution two-dimensional temperature plot of PV/T panel Figure A - 18: Heat flux into PV/T panel vs. panel length 9 Figure A - 19: Cell electrical output efficiency vs. panel length 7. Test Case 2c Figure A - 20: Steady state solution two-dimensional temperature plot of PV/T panel 10 Figure A - 21: Heat flux into PV/T panel vs. panel length Figure A - 22: Cell electrical output efficiency vs. panel length 11 8. Test Case 2d Figure A - 23: Steady state solution two-dimensional temperature plot of PV/T panel Figure A - 24: Heat flux into PV/T panel vs. panel length 12 Figure A - 25: Cell electrical output efficiency vs. panel length 9. Test Case 3a Figure A - 26: Steady state solution two-dimensional temperature plot of PV/T panel 13 Figure A - 27: Heat flux into PV/T panel vs. panel length Figure A - 28: Cell electrical output efficiency vs. panel length 14 10.Test Case 3b Figure A - 29: Steady state solution two-dimensional temperature plot of PV/T panel Figure A - 30: Heat flux into PV/T panel vs. panel length 15 Figure A - 31: Cell electrical output efficiency vs. panel length 11.Test Case 3c Figure A - 32: Steady state solution two-dimensional temperature plot of PV/T panel 16 Figure A - 33: Heat flux into PV/T panel vs. panel length Figure A - 34: Cell electrical output efficiency vs. panel length 17 12.Test Case 3d Figure A - 35: Steady state solution two-dimensional temperature plot of PV/T panel Figure A - 36: Heat flux into PV/T panel vs. panel length 18 Figure A - 37: Cell electrical output efficiency vs. panel length Figure A - 38: Cooling reservoir internal flow velocity profile 19