Next Article in Journal
A Method for the Optimization of Daily Activity Chains Including Electric Vehicles
Next Article in Special Issue
Anti-Reflective Coating Materials: A Holistic Review from PV Perspective
Previous Article in Journal
A Wireless Power Transfer Based Implantable ECG Monitoring Device
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Investigations Conducted for the Characteristic Study of OM29 Phase Change Material and Its Incorporation in Photovoltaic Panel

by
Rajvikram Madurai Elavarasan
1,*,
Karthikeyan Velmurugan
2,3,*,
Umashankar Subramaniam
4,
A Rakesh Kumar
5 and
Dhafer Almakhles
4
1
Department of Electrical and Electronics Engineering, Sri Venkateswara College of Engineering, Chennai, 602117 Tamil Nadu, India
2
School of Renewable Energy and Smart Grid Technology, Naresuan University, 65000 Phitsanulok, Thailand
3
Center for Alternative Energy Research and Development, Khon Kaen University, 40002 Khon Kaen, Thailand
4
Renewable Energy Laboratory, Faculty of Engineering, Prince Sultan University, 11586 Riyadh, Saudi Arabia
5
Department of Electrical and Electronics Engineering, National Institute of Technology, 620015 Tiruchirappalli, India
*
Authors to whom correspondence should be addressed.
Submission received: 17 January 2020 / Revised: 8 February 2020 / Accepted: 12 February 2020 / Published: 18 February 2020
(This article belongs to the Special Issue Nano-Structured Solar Cells 2020-2022)

Abstract

:
The solar photovoltaic (PV) system is emerging energetically in meeting the present energy demands. A rise in PV module temperature reduces the electrical efficiency, which fails to meet the expected energy demand. The main objective of this research was to study the nature of OM29, which is an organic phase change material (PCM) used for PV module cooling during the summer season. A heat transfer network was developed to minimize the experimental difficulties and represent the working model as an electrical resistance circuit. Most existing PV module temperature (TPV) reduction technology fails to achieve the effective heat transfer from the PV module to PCM because there is an intermediate layer between the PV module and PCM. In this proposed method, liquid PCM is filled directly on the back surface of the PV module to overcome the conduction barrier and PCM attains the thermal energy directly from the PV module. Further, the rear side of the PCM is enclosed by tin combined with aluminium to avoid any leakages during phase change. Experimental results show that the PV module temperature decreased by a maximum of 1.2 °C using OM29 until 08:30. However, after 09:00, the OM29 PCM was unable to lower the TPV because OM29 is not capable of maintaining the latent heat property for a longer time and total amount of the PCM experimented in this study was not sufficient to store the PV module generated thermal energy for an entire day. The inability of the presented PCM to lower the temperature of the PV panel was attributed to the lower melting point of OM29. PCM back sheet was incapable of dissipating the stored PCM’s thermal energy to the ambient, and this makes the experimented PCM unsuitable for the selected location during summer.

1. Introduction

The global energy demand has increased by 2.3% compared to 2010, and this increase is mainly due to modernization and industrialization [1]. Air pollution has increased significantly because of the forest to agricultural land transformation, burning of fossil fuels, dumping waste into water, burning waste, and vehicle smoke emission. This increases by 1.7% the CO2 concentration every year, leading to a threatening global climate change [2,3,4]. To overcome this issue, renewable energy systems are widely engaged to meet energy demand [5]. Among other renewable energy sources, solar photovoltaic (PV) systems gain attention for their low maintenance and payback period. A great amount of research work is also carried out on PV materials to improve the efficiency of PV panel [6,7]. The PV system can be operated on a standalone basis or in grid connected mode [8,9,10,11,12]. Crystalline silicon solar cell converts 1/5th of the photon into electrical energy, while some is reflected, and this reflection can be controlled by coating a thin layer of Al2O3 and Ta2O5 [13], and the rest is almost converted into thermal energy. Excess thermal energy in the PV module results in the drop of efficiency by 0.3–0.4% for every 1 °C more than the nominal operating cell temperature [14]. Further reduction in TPV can increase the performance of the system.
Earlier PV module cooling was conducted using water or air as a cooling agent. Water has a high specific heat capacity (4.2 J/g·K), and it can absorb higher thermal energy from the PV module. Water-assisted PV modules are mostly performed as an active method by spraying water over the PV module [15] or using heat pipe behind the PV module for both the electrical and thermal applications [16,17]. Other than active method, thermosiphon concept has also been tested but the result lowers the TPV reduction [18]. Air-based PV module cooling is mostly examined for hybrid applications where hot air is used for thermal comfort and space heating [19,20]. In most cases, water and air-based TPV reductions are not convenient due to lack of resource.
In recent years, phase change material (PCM)-assisted PV module cooling has gained great attention in terms of its striking latent heat of fusion [21,22]. PCM can extract the generated thermal energy from the PV module, and it is stored in the form of specific heat capacity and latent heat of fusion; this is as expressed in Equation (1) [23]. PCM-aided PV modules are prominent and they depend on the application of cooling methods chosen, either active or passive.
Q = T i T m m C p , s d T + m L + T m T e m C p , l d T
where Ti is initial temperature, Tm is melting temperature, Te is ending temperature, m is total mass of PCM, Cp,s is solid specific heat capacity of PCM, Cp,l is liquid specific heat capacity of PCM, L is latent heat capacity of the PCM, and Q is the total amount of energy stored in the PCM.
Active cooling methods are implemented for the performance enhancement of electrical and thermal energy using working fluids [24,25]. This active method requires less PCM because it is assisted with working fluid and periodically PCM temperature is removed by working fluid, which enables the heat transfer between the PV module and PCM. Fayaz et al. performed COMSOL simulation for a photovoltaic–thermal (PVT) system with and without PCM using heat exchanger-assisted model and compared it with the PV module [26]. They found that numerical simulation results are accurate compared to the experimental result and the remarkable increase in flow rate leads to reduce the TPV maximum by 13.77 °C. Al-Waeli et al. studied a nano PCM and nanofluid-assisted PVT collector, a copper absorber tube entrenched with nano PCM, to extract the heat from PCM container by flowing nanofluid, which enhances the thermal energy transfer from PV to PCM container [27]. Silicon oil is placed between the PV module rear surfaces and the PCM container to make perfect contact. The result shows that nano PCM assisted TPV reduced from 68.46 °C to 36.04 °C with the help of nanofluid.
Modjinou et al. comparatively analyzed macro encapsulated PCM, micro-channel heat pipe (MCHP), and regular PVT with water [28]. They found that PVT-PCM and PVT-MCHP enhanced both electrical and thermal performance. Notably, PCM-based PV module maintains constant temperature whenever there is a sudden drop in ambient temperature. Acetone influenced MHCP gains higher performance than regular PVT setup since refrigerant fluid leads to enhance the heat transfer.
PCM-only (passive cooling) incorporated PV modules are widely performed to enhance electrical efficiency without using an external source, and this system is totally autonomous and maintenance-free. Sandro et al. examined RT 38 and pork fat as a PCM in Croatia. Both PCM performed well throughout the year and there was no noticeable difference in TPV reduction except in June when ambient temperature is higher [29]. Zhao et al. analyzed different melting temperatures of PCM (PCM15, PCM20, PCM25, and PCM30) using MATLAB to find the role of PCM melting temperature [30]. A higher melting range of PCM sustains a longer time in summer, and a lower melting range of PCM performs well in winter (PCM20). This study enhanced by 2.4% the annual electric energy production.
The orientation of the PV module along with PCM plays an energetic role in TPV reduction because, internally, PCM gains thermal energy by both the conduction and the convection [31]. When the PV-PCM tilt increases, PCM internal convection also increases and it helps to reduce the higher TPV. During PCM melting, thermal energy gains by convection and it activates latent heat property, and conduction occurs during PCM solid state. Thermal variation on the PV module was simulated using ANSYS fluent under different climatic conditions such as 1–5 m/s wind variation. It showed an increase in wind speed and wind azimuth angle lead to dissipate higher thermal energy from both the PV module and the PCM container, which encourages heat transfer between the PV module and the PCM. A 22.6 °C melting range of PCM extracts higher heat than a 30.6 °C melting range of PCM [32].
In the passive cooling system, PCM temperature has been dissipated naturally without using any working fluid. Mostly, PCM filled in a container that is made up of high thermal conducting material. Considering cost-effectiveness and high thermal conductivity, aluminium is more widely used as a PCM container than nickel, stainless steel, and iron. Thin aluminium is difficult to weld because of its melting temperature, which is lower than other metals. During sunny days, PCM changes its phase and, considering this phase change, PCM container is fabricated in the thickness of 4–8 mm metal [32,33,34,35,36]. In TPV reduction, low operating PCM temperature is used, which is sensitive in thermal conduction whenever the thickness of PCM container material increases. To facilitate thermal dissipation, the high conducting heatsink has been attached behind the PCM container.
Rajvikram et al. conducted a comparative analysis of PV with heatsink, PCM, and PCM–heatsink [37]. The results show that PV only and heatsink-assisted PV module maintain 1 °C reduction. PCM integrated PV module regulates better TPV than heatsink by almost 7.7 °C because PCM enables high heat transfer from PV module. Finally, PCM–heatsink controls better than only heatsink or PCM, because it dissipates the stored thermal energy to the ambient during sunshine hours, which leads to maintaining a lower PCM temperature than only PCM and this discharge helps to transfer the heat from PV module and PCM.
It was noticed that using heatsink for PCM-based passive cooling technique enhances TPV reduction [38]; however, the PV module backside (tedlar) is made up of non-metal, the whole PV module is fragile, and it is not advisable to keep the PCM container or PCM container with heat sink. To obtain perfect physical contact, PCM container is attached to the PV module back surface using high-temperature adhesive material or conducting oil, even though perfect physical contact is not justified. It was reported that total weight of the system could increase due to the thickness of PCM container material and fastening heatsink behind the PCM container. Attaching a PCM container or PCM container with heatsink reduces TPV, but it damages the PV module’s physical structure.
To prevent this contact loss and physical damage, PCM is filled on the back surface of the PV module directly [39]. Nada et al. used 10-cm PCM and its rear side was sealed using a 5-mm galvanized sheet to prevent leaks [40]. The comparative study of RT55 as pure PCM and Al2O3 nanoparticle composited PCM shows better TPV reduction and the result shows that pure PCM reduced a maximum of 8.1 °C and composite PCM reduced 10.6 °C. The reason behind the TPV reduction of composite PCM is its enhanced thermal conductivity. Even though PCM is an effective thermal energy storage material, it has a lack of thermal conductivity. This low thermal conductivity creates thermal conduction barrier within the PCM.
Stropnik et al. conducted TRNSYS simulation for Ljubljana location using 3.5-cm RT28HC and they validated it with experimental results [41]. Filling PCM directly behind the PV module enhances heat transfer because PV module back surface temperature is directly in contact with PCM and there is no intermediate layer. The rear side of the PCM is covered by 5-mm acrylic glass, and the direct enforcement of PCM leads to reduce the TPV by a maximum of 35.6 °C. The relative error of the simulation is less than 3.8% compared to the experimental results. Total energy production was enhanced from 242.36 to 260.17 kWh/year and, especially from March to September, better performance is achieved.
Hasan Mahamudul et al. conducted the performance analysis of RT 35 PCM under Malaysian climatic condition with a thickness of 2 cm, which is filled and fitted within the PV module frame height, and the rear side of PCM is closed by using fiber optic glass [42,43]. The results show that a maximum of 10–12 °C reduction is achieved, but this system yields TPV reduction for only 4 h because the amount of PCM used in this system is less. Some of the literature works are summarized in Table 1 for the understanding of the PCM behavior in PV module cooling under different climatic condition. Selecting appropriate melting range of PCM is one of the most crucial decisions in PV module cooling because thermal absorption of the PCM relies on its melting range and natural factors, mainly ambient temperature, humidity, and wind speed [44].
The existing literature makes clear that PCM is an effective thermal absorber and storage medium for TPV reduction. PCM only and composite PCM gains great attention because of its zero maintenance and free from external energy. Systematically, PCM stores thermal energy during sunshine hours and it discharges at night, which makes this system free from maintenance and external sources. The only drawback in this system is physical contact between the PV module and the PCM container because it could damage the physical structure of the PV module due to forging hard metal onto the soft material of the PV module. Considering contact loss and weight, filling liquid PCM behind the PV module is the most effective and safest method to reduce the TPV.
The main objective of this research was to analyze the effectiveness of the direct filling PCM on the back surface (tedlar) of the PV module for the summer season. OM29 is used as PCM for the location of Madurai, India (9.88° N, 78.08° E) to regulate the excess TPV. In this current work, thermal absorption and dissipation of the PCM are exclusively discussed in the following section.

2. Materials and Methods

2.1. Phase Change Material

OM29 is a commercial organic PCM purchased from Pluss [43], India, and it is a combination of several fatty acids. The physical appearance of OM29 is a white wax at 25 °C (solid-state). OM29 was tested by PLUSS T-History method, resulting in 29 °C onsets melting temperature with high latent heat of fusion, which makes this PCM able to store thermal energy effectively. In addition, OM29 is chemically and thermally stable for 2000 cycles, which means OM29 can perform effectively for 2000 days. It is considered that every single day TPV reduction is equivalent to a single thermal cycle. The experimental location is rich in strong sunshine of approximately 300 days per year, which means OM29 can perform effectively for 6.5 years without degrading its thermal property and performance. The thermophysical properties of OM29 are listed in Table 2.

2.2. Experimental Setup

Two identical 20 Wp polycrystalline PV modules (Loom solar with the dimension 450 mm × 350 mm × 22 mm) were used in this experiment, one for PV without PCM and another for PV with PCM, as shown in Figure 1. Without further processing, 2.8 kilograms of liquid OM29 PCM was filled on the backside of the PV module with 5% space left for volume change during phase transition or volume expansion. Over the PCM layer, the rear side of the PCM was enclosed by using a 0.5-mm tin combined with aluminium sheet and, thus, leakages were protected during phase change.
The experiment was conducted in summer (6 July 2019) at the location of 9.88 N, 78.08 E. To make an effective comparison of OM29, which was incorporated for the PV module cooling, the front and back surface of the PV module temperature was captured using FLIR thermal imaging camera every 30 min for PV without PCM. For the PV with PCM front surface was captured, but it was difficult to capture the back surface of the PV module because PV module back surface was not exposed to the camera as it was incorporated with the PCM. Thus, PCM back sheet temperature was captured to analyze the thermal dissipation of the PCM.

3. Thermal Heat Transfer Model

3.1. Heat Transfer Model for PV without PCM

Conventional PV module thermal characteristics are represented in the form of a heat transfer model, as shown in Figure 2. In the conventional system, the PV module generated thermal energies and they are mostly dissipated to the surrounding, depending on the ambient temperature (Tamb) and wind speed.
In hot region, less thermal dissipation occurs because of high Tamb and low wind speed, which increases the TPV because excess thermal energy has been dissipated by radiation (Rrad, PV→sky and Rrad, PV→surr) and natural convection (Rconv, PV→amb and Rconv, PV→surr) on both the top and the bottom surface of the PV module.
Thermal energy dissipations are exclusively represented in the form of an energy balance equation in the following subsection.

Energy Balance for PV without PCM

The solar PV module is a specific heat storage material, and it stores thermal energy during the photovoltaic effect. Equation (2) denotes the total amount of thermal energy stored in the PV module corresponding to the solar irradiance, which is absorbed and transmitted to the solar cell for the energy conversion process, and during that time the PV module generates thermal energy. That energy is transferred to the ambient and sky by the mode of convection and radiation over the flat surface of the PV module, respectively. The rear side of the PV module transfers thermal energy to the surrounding by natural convection and radiation. PV module electrical efficiency is directly correlated with the TPV.
m P V C P V d T P V d t = [ α g τ g I ( t ) h c o n v , g a m b ( T P V , g T a m b ) h r a d , g s k y ( T P V , g T s k y ) h c o n v , t s u r r ( T P V , t T s u r r ) h r a d , t s u r r ( T P V , t T s u r r ) η P V I ( t ) ] β A P V
where mPV is the weight of the PV module; CPV is the specific heat capacity of the PV module; α g is the absorbance of the PV module glass; τ g is the transmittance of the PV module; I is the solar irradiance in W/m2; t is the time in s; h c o n v , g a m b is the convection heat transfer between PV module glass and ambient (W/m2°C); h r a d , g s k y is the radiation heat transfer between PV module glass and sky (W/m2°C); h c o n v , t s u r r is the convection heat transfer between PV module tedlar and surrounding (W/m2°C); h r a d , t s u r r is the radiation heat transfer between PV module tedlar and surrounding (W/m2°C); TPV,g is the PV module glass temperature in °C; Tamb is the ambient temperature in °C; Tsky is the sky temperature in °C; TPV,t is the PV module tedlar temperature in °C, Tsurr is the surrounding temperature in °C, η P V is the efficiency of the PV module in %; β is the packing factor; and APV is the area of the PV module.
Convection over the PV module majorly occurs with the influence of wind (v) and an increase in wind speed over the surface of the PV module dissipates higher thermal energy to the ambient, as represented by Equation (3).
h c o n v , g a m b = 2.8 + 3 v , 0 < v < 7 m / s
PV module front surface radiates thermal energy to the sky and this radiation truly depends on the emissivity of the solar PV module glass [55]. It is represented by Equation (4).
h r a d , g s k y = ϵ g σ ( T P V , g + T s k y ) ( T P V , g + T s k y ) 2
where σ is the Stefan–Boltzmann constant and ϵ g is the emissivity of the PV module glass.
Long-range thermal radiation to the sky is calculated using Equation (5) [56],
T s k y = 0.0522 T a m b 0.5
The backside of the PV module dissipates thermal energy to the surroundings. Usually, this convection is poor due to the thermal resistance. PV module dissipated thermal energy stagnates in the area between the PV module backsurface and the ground. Especially the hot and low wind area suffers in thermal stagnation. This stagnation leads to reduce the thermal conductivity of air because convection depends on the surrounding air thermal conductivity, as expressed in Equation (6), and the increase in air temperature causes to reduce the thermal conductivity of the air.
h c o n v , t s u r r = K a i r L P V N u a i r
where K a i r is the thermal conductivity of air; LPV is the PV module length; and Nuair is Nusslets Number.
The rear side of the PV module Nusselt number is calculated using Equation (7) and radiation to the surrounding using Equation (8).
N u a i r = 0.825 + 0.387 R a 1 / 6 [ 1 + ( 0 . 492 P r ) 9 / 6 ] 8 / 27 2
where Ra is the Rayleigh number; Pr is the Prandtl number; and ϵ t is the Emissivity of the PV module tedlar.
h r a d , t s u r r = ϵ t σ ( T P V , t + T s u r r ) ( T P V , t + T s u r r ) 2
Electrical efficiency of the PV module is calculated using Equation (9) [57],
η P V = η S T C [ 1 β S T C ( T P V T S T C ) ]
where η P V is the Efficiency of the PV module in %; η S T C is the PV module efficiency at standard test condition in %; TPV is the PV module temperature in °C; TSTC is the Standard test condition temperature in °C; and β S T C is the Temperature Coefficient.

3.2. Heat Transfer Model for PV with PCM

Excess thermal energy from the PV module is absorbed by PCM, which is filled up on the back surface. The conventional PV module rear surface dissipates thermal energy to the surrounding with the help of temperature difference and wind. In this model, PCM plays a major role during energy conversion time by absorbing the excess thermal energy and, simultaneously, it is expected to discharge. The rear surface of the PV module thermal energy is transferred to the PCM by conduction (Rcond, PV→PCM) and it is stored in it. Stored PCM thermal energy is transferred to the PCM back-sheet by conduction (Rcond, PCM→Al). From PCM back sheet, thermal energy is dissipated to the surrounding by natural convection (Rconv, Al→surr) and radiation (Rrad, Al→surr). Default front side of the PV module dissipates thermal energy by convection (Rconv, PV→amb) and radiation (Rrad, PV→sky). The working principle of the PCM integrated PV module is represented in the form of a heat transfer model, as shown in Figure 3.

Energy Balance for PCM Integrated PV Module

Equation (10) denotes the amount of thermal energy stored in the PV module after PCM absorption. The front side of the PV module absorbs and transmits the solar irradiance to the PV module for energy conversion. During that time, the PV module generates thermal energy, and it is absorbed by the PCM, which is incorporated on the back surface by direct conduction. The rest is dissipated to the sky and the ambient the same as for PV without PCM. In this case, the electrical efficiency of the PV module is directly correlated with the amount of thermal energy absorbed by PCM.
m P V C P V d T P V d t = [ α g τ g I ( t ) h c o n v , g a m b ( T P V , g T a m b ) h r a d , g s k y ( T P V , g T s k y ) h c o n d , t P C M ( T P V , t T P C M ) η P V I ( t ) ] β A P V
where hcond,t→PCM is the conduction heat transfer between PV module tedlar and PCM (W/m2°C) and TPCM is the PCM temperature in °C.
Moreover, regulating TPV for longer and effectively depends on the ratio of PCM thermal conductivity and PCM thickness, as expressed in Equation (11).
h c o n d , t P C M = K P C M Δ X P C M
where KPCM is the PCM thermal conductivity and Δ XPCM is the PCM thickness.

3.3. Energy Balance for Conduction Sourced PCM

Equation (12) denotes the amount of thermal energy stored in the PCM during energy conversion. Excess thermal energy from the PV module is transferred to the PCM by conduction and this energy is stored in three different phases, as expressed in Equation (13). During charging and off sunshine hours, PCM thermal energy is transferred to the PCM back sheet by conduction. Thus, dissipating PCM thermal energy to the surrounding leads to enhance the heat transfer from the PV module to PCM. PCM back sheet transfers thermal energy to the surrounding by natural convection and radiation because the PCM backside is not insulated. PCM should discharge the stored thermal energy before the next sunshine (overnight), else the next day PCM would not be able to initiate the absorption effectively.
m P C M C P C M d T P C M d t = [ h c o n d , t P C M ( T P V , t T P C M ) h c o n d , P C M A l ( T P C M T A l ) h c o n v , A l s u r r ( T A l T s u r r ) h r a d , A l s u r r ( T A l T s u r r ) ] β A P C M
where mPCM is the mass of PCM; CPCM is the storage capacity of PCM; hcond,PCM→Al is the conduction heat transfer between PCM and PCM back sheet (W/m2°C); hconv,Al→surr is the convection heat transfer between PCM back sheet and surrounding (W/m2°C); hrad,Al→surr is the radiation heat transfer between PCM back sheet and surrounding (W/m2°C); TPCM is the PCM temperature (°C); TAl is the PCM back sheet temperature (°C); and Tsurr is the surrounding temperature (°C).
Equation (13) represents three different phases of PCM during thermal energy storage. When the PCM temperature is less than the melting point, thermal energy stores in the form of solid specific heat capacity. When it is equal to the melting point, PCM stores thermal energy in the form of latent heat of fusion. It is recommended to maintain the PCM to be in latent heat of fusion state because the liquid specific heat capacity of the PCM is not effective in storing thermal energy and thus thermal conductivity of the liquid PCM is low.
C P C M = s o l i d s p e c i f i c h e a t c a p a c i t y , T P C M < T m e l t l a t e n t h e a t o f f u s i o n , T P C M = T m e l t l i q u i d s p e c i f i c h e a t c a p a c i t y , T P C M > T m e l t
where Tmelt is the melting temperature.

4. Results and Discussions

4.1. Temperature Profile of PV Module

The experimental setup of the PV with and without PCM was developed and examined in the summer (6 July 2019) under the direct sunlight to observe the thermal behavior. During every 30 min interval, thermal images of the PV module and PCM back sheet were taken. Figure 4 and Figure 5 show the meteorological data of solar irradiance and ambient temperature on the corresponding experimental day.
The examined OM29 organic PCM enabled thermal energy transfer from PV module to PCM by conduction. Figure 6a shows a thermal image of the PV module at 8:00. There was a noticeable TPV reduction of 0.4 °C because, during that time, PCM had solid specific heat capacity as well semi mushy region.
PCM absorbed the TPV effectively when the PCM attained its melting point. The results show that PCM utilized its thermal absorbing and storing capability until 8:30 because TPV reduction achieved up to 8:30 maximum of 1.2 °C. From 9:00, TPV started to rise slightly less than the PV without PCM, as shown in Figure 6b,c. Even though the latent heat capacity of the PCM was high, TPV reduction sustained for a lower period due to its low melting point, which is inappropriate for the selected location for the summer season. Further following this, TPV was not achieved during the peak sunshine hours, as shown in Figure 7 and Figure 8. After 9:00, PCM was completely liquid completely, as listed in Table 3. Liquid specific heat capacity was not effective for storing the thermal energy such as latent heat of fusion (mushy state).
Secondly, in this experiment, PCM failed to discharge the stored thermal energy during sunshine hours, because thermal conductivity of the PCM was low, increasing the thermal conduction resistance within the liquid PCM, and the PCM back sheet (tin combined with aluminium) was not an effective thermal dissipater as compared to pure aluminium [39]. Increasing the thickness of PCM enhanced the heat transfer between the PV module and the PCM [58,59,60]. In this system, 2-cm PCM was used, which melted completely within 1 h from and, during the rest of the period, TPV increased to a maximum of 7 °C, which was more than the conventional PV module (PV without PCM), as shown in Figure 8 and Figure 9.
The reason behind this TPV rise is that the PCM integrated PV module back surface was not interacting with the ambient and wind. Since PCM was filled on the back surface of the PV module, the PCM integrated PV module thermal dissipation depended on the PCM and PCM container material. For PV without PCM, thermal energy could be easily dissipated to the surrounding because there was no storage medium between the PV module back surface and the surrounding. However, PCM integration reduced TPV effectively when the PCM sustained the mushy state for a longer time, but, once it became liquid, PCM acted as an insulating material compared to PV without PCM. Thus, PV without PCM had a lower TPV than PV with PCM. Since the back surface of the PV module had PCM, it had stored thermal energy in it. Note that the experimental day ambient temperature was higher than the PCM melting temperature.
The 2.2-cm PCM layer contained only 2.8 kilograms of PCM. It could not maintain the mushy state for a long time and had difficulty reducing the PCM temperature lower than the melting temperature using passive cooling because it depended on the natural factor. If the PCM back sheet also had a good thermal dissipating material, PCM would not reach its melting point or mushy state again during the experimental day, because ambient temperature was higher than the PCM melting point. Thus, the only possible way is active cooling method, which could lower the PCM temperature forcibly and help to maintain the mushy state for longer time. Many researchers addressed this ineffective TPV reduction by categorizing inappropriate PCM melting temperature selection [44], PCM thickness [61], and PCM thermal conduction barrier [35]. Existing experimental investigation states that PCM selection is the crucial term in the whole process of the PV module cooling; selecting a lower or higher melting range of PCM will make the TPV reduction ineffective because TPV reduction is only achieved when the PCM reaches its melting point. Lower melting range of PCM will utilize its latent heat property before the peak sunshine hours and higher melting range of PCM will not use its latent heat property during peak sunshine hours. The only solution for this issue is to select PCM with melting range 2–3 °C higher than the peak ambient temperature so that the PCM will not gain thermal energy from ambient during sunshine hours.
During the experimental day, the recorded high ambient temperature was 33.5 °C. If the selected PCM melting range were 35–36 °C, a better TPV reduction would have been achieved. A general rule of thumb indicates that an increase in the thickness of PCM leads to sustainability of the TPV reduction for longer time, as well as creates the thermal conduction barrier during thermal absorption. To overcome this issue, several researchers have incorporated thermal additives to enhance the thermal conductivity of the PCM by adding metal foam [62], metal scrap [63], copper powder [64], graphite [64], expandable graphite [65], and nanoparticles [48].

4.2. Temperature Corrected Electrical Efficiency

The electrical efficiency of the PV module was directly correlated with the TPV; the increase in TPV led to reduce electrical efficiency by −0.43%/°C. However, this experiment enhanced the electrical efficiency until 8:30; afterwards, PCM integrated PV module electrical efficiency was lower than PV without PCM, as shown in Figure 10.
The corresponding increase in TPV in the PCM integrated PV module reduced the electrical efficiency by a maximum of 0.38 %, which shows that PCM acted as a thermal source to the PV module when it turned completely liquid.

5. Conclusions

The heat transfer capability from the PV module to the PCM can be greatly enhanced by the direct filling of PCM behind the PV panel without the use of an intermediate layer. It is an efficient way of keeping the temperature of the PV panel within a reasonable limit by means of passive cooling methodology. However, f an OM29 PCM tends to behave differently in the summer season because of the atmospheric condition. On the day of experimentation, TPV dropped up to 8:30; after that, the OM29 was unable to absorb the PV module generated thermal energy due to the impact of a lower melting range of the PCM. The experimental results reveal that the OM29 PCM was not capable of PV module cooling, especially for a hot and humid region such as India. Many researchers addressed this issue because selecting PCM melting temperature is the most challenging task for the module cooling for an entire year. The selected OM29 PCM performed well in winter, while, in summer, it did not dissipate the stored thermal energy; in addition, OM29 lost its latent heat property within 8:30 and after that PCM started to act as an insulating material for the PV module. Hence, it was concluded that OM29 was not a suitable PCM for cooling the PV module in the summer season. Further, it is recommended to select an appropriate PCM for this desired location. In addition, it is necessary to increase the thickness of the PCM and thermal conductivity in order to attain better absorption and dissipation.

Author Contributions

R.M.E. and K.V. carried out the experimental research, analysis of results and writing the draft of the manuscript. A.R.K. carried out the proofreading and formatting of the research work and paper writing. U.S. and D.A. carried out the proof reading of the manuscript and funding the publication of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The APC was funded by the College of Engineering, Prince Sultan University, Riyadh-11586, Saudi Arabia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zedalis, R.J. Global Energy &amp; CO2 Status Report; Technical Report; Routledge: London, UK, 2017. [Google Scholar] [CrossRef]
  2. Qazi, A.; Hussain, F.; Rahim, N.A.; Hardaker, G.; Alghazzawi, D.; Shaban, K.; Haruna, K. Towards Sustainable Energy: A Systematic Review of Renewable Energy Sources, Technologies, and Public Opinions. IEEE Access 2019, 7, 63837–63851. [Google Scholar] [CrossRef]
  3. Elavarasan, R.M. Comprehensive Review on India’s Growth in Renewable Energy Technologies in Comparison With Other Prominent Renewable Energy Based Countries. J. Sol. Energy Eng. 2020, 142, 1–11. [Google Scholar] [CrossRef]
  4. Elavarasan, R.M.; Shafiullah, G.M.; Kumar, N.M.; Padmanaban, S. A State-of-the-Art Review on the Drive of Renewables in Gujarat, State of India: Present Situation, Barriers and Future Initiatives. Energies 2020, 13, 40. [Google Scholar] [CrossRef] [Green Version]
  5. Elavarasan, R.M. The Motivation for Renewable Energy and its Comparison with Other Energy Sources: A Review. Eur. J. Sustain. Dev. Res. 2019, 3, 1–19. [Google Scholar] [CrossRef]
  6. Das, N.K.; Islam, S.M. Conversion Efficiency Improvement in GaAs Solar Cells. In Large Scale Renewable Power Generation. Green Energy and Technology; Springer: Singapore, 2014; pp. 53–75. [Google Scholar] [CrossRef]
  7. Das, N.; Wongsodihardjo, H.; Islam, S. Modeling of multi-junction photovoltaic cell using MATLAB/Simulink to improve the conversion efficiency. Renew. Energy 2015, 74, 917–924. [Google Scholar] [CrossRef]
  8. Subramaniam, U.; Ganesan, S.; Bhaskar, M.S.; Padmanaban, S.; Blaabjerg, F.; Almakhles, D.J. Investigations of AC Microgrid Energy Management Systems Using Distributed Energy Resources and Plug-in Electric Vehicles. Energies 2019, 12, 2834. [Google Scholar] [CrossRef] [Green Version]
  9. Kumar, A.R.; Thangavelusamy, D.; Padmanaban, S.; Kothari, D.P. A Modified PWM Scheme to improve AC Power Quality for MLIs using PV Source. Int. J. Power Energy Syst. 2019, 39, 34–41. [Google Scholar] [CrossRef]
  10. Sridhar, V.; Umashankar, S.; Sanjeevikumar, P.; Ramachandaramurthy, V.K.; Mihet-Popa, L.; Fedák, V. Control Architecture for Cascaded H-Bridge Inverters in Large-Scale PV Systems. Energy Procedia 2018, 145, 549–557. [Google Scholar] [CrossRef]
  11. Elavarasan, M.; Mallick, G.; Saravanan, K. Investigations on Performance Enhancement Measures of the Bidirectional Converter in PV-Wind Interconnected Microgrid System. Energies 2019, 12, 2672. [Google Scholar] [CrossRef] [Green Version]
  12. Bhukya, M.N.; Kota, V.R.; Depuru, S.R. A Simple, Efficient, and Novel Standalone Photovoltaic Inverter Configuration With Reduced Harmonic Distortion. IEEE Access 2019, 7, 43831–43845. [Google Scholar] [CrossRef]
  13. Rajvikram, M.; Leoponraj, S. A method to attain power optimality and efficiency in solar panel. Beni-Suef Univ. J. Basic Appl. Sci. 2018, 7, 705–708. [Google Scholar] [CrossRef]
  14. Alami, A.H. Effects of evaporative cooling on efficiency of photovoltaic modules. Energy Convers. Manag. 2014, 77, 668–679. [Google Scholar] [CrossRef]
  15. Schiro, F.; Benato, A.; Stoppato, A.; Destro, N. Improving photovoltaics efficiency by water cooling: Modelling and experimental approach. Energy 2017, 137, 798–810. [Google Scholar] [CrossRef]
  16. Liang, R.; Wang, P.; Zhou, C.; Pan, Q.; Riaz, A.; Zhang, J. Thermal performance study of an active solar building façade with specific PV/T hybrid modules. Energy 2019, 116532. [Google Scholar] [CrossRef]
  17. Bahaidarah, H.; Subhan, A.; Gandhidasan, P.; Rehman, S. Performance evaluation of a PV (photovoltaic) module by back surface water cooling for hot climatic conditions. Energy 2013, 59, 445–453. [Google Scholar] [CrossRef]
  18. Sutanto, B.; Indartono, Y.S. Computational fluid dynamic (CFD) modelling of floating photovoltaic cooling system with loop thermosiphon. In Proceedings of the AIP Conference, Bali, Indonesia, 14–16 August 2019; Volume 2062, p. 020011(1-6). [Google Scholar] [CrossRef]
  19. Sellami, R.; Amirat, M.; Mahrane, A.; Slimani, M.E.A.; Arbane, A.; Chekrouni, R. Experimental and numerical study of a PV/Thermal collector equipped with a PV-assisted air circulation system: Configuration suitable for building integration. Energy Build. 2019, 190, 216–234. [Google Scholar] [CrossRef]
  20. Shahsavar, A.; Khanmohammadi, S.; Khaki, M.; Salmanzadeh, M. Performance assessment of an innovative exhaust air energy recovery system based on the PV/T-assisted thermal wheel. Energy 2018, 162, 682–696. [Google Scholar] [CrossRef]
  21. Krishnan, S.; Garimella, S.; Kang, S. A novel hybrid heat sink using phase change materials for transient thermal management of electronics. IEEE Trans. Components Packag. Technol. 2005, 28, 281–289. [Google Scholar] [CrossRef] [Green Version]
  22. Afshari-Bavil, M.; Dong, M.; Li, C.; Feng, S.; Zhu, L. Thermally Controllable High-Efficiency Unidirectional Coupling in a Double-Slit Structure Filled With Phase Change Material. IEEE Photonics J. 2019, 11, 1–8. [Google Scholar] [CrossRef]
  23. Karthikeyan, V.; Sirisamphanwong, C.; Sukchai, S. Thermal investigation of Paraffin Wax for Low-Temperature Application. J. Adv. Res. Dyn. Control. Syst. 2019, 11, 1437–1443. [Google Scholar]
  24. Stupar, A.; Drofenik, U.; Kolar, J.W. Optimization of Phase Change Material Heat Sinks for Low Duty Cycle High Peak Load Power Supplies. IEEE Trans. Components Packag. Manuf. Technol. 2012, 2, 102–115. [Google Scholar] [CrossRef]
  25. Alawadhi, E. Thermal Management of Blocks in a Channel Using Phase Change Material. IEEE Trans. Components Packag. Technol. 2009, 32, 89–99. [Google Scholar] [CrossRef]
  26. Fayaz, H.; Rahim, N.; Hasanuzzaman, M.; Nasrin, R.; Rivai, A. Numerical and experimental investigation of the effect of operating conditions on performance of PVT and PVT-PCM. Renew. Energy 2019, 143, 827–841. [Google Scholar] [CrossRef]
  27. Al-Waeli, A.H.; Chaichan, M.T.; Sopian, K.; Kazem, H.A.; Mahood, H.B.; Khadom, A.A. Modeling and experimental validation of a PVT system using nanofluid coolant and nano-PCM. Sol. Energy 2019, 177, 178–191. [Google Scholar] [CrossRef]
  28. Modjinou, M.; Ji, J.; Yuan, W.; Zhou, F.; Holliday, S.; Waqas, A.; Zhao, X. Performance comparison of encapsulated PCM PV/T, microchannel heat pipe PV/T and conventional PV/T systems. Energy 2019, 166, 1249–1266. [Google Scholar] [CrossRef]
  29. Nižetić, S.; Arıcı, M.; Bilgin, F.; Grubišić-Čabo, F. Investigation of pork fat as potential novel phase change material for passive cooling applications in photovoltaics. J. Clean. Prod. 2018, 170, 1006–1016. [Google Scholar] [CrossRef]
  30. Zhao, J.; Ma, T.; Li, Z.; Song, A. Year-round performance analysis of a photovoltaic panel coupled with phase change material. Appl. Energy 2019, 245, 51–64. [Google Scholar] [CrossRef]
  31. Al-Jethelah, M.S.M.; Al-Sammarraie, A.; Tasnim, S.H.; Mahmud, S.; Dutta, A. Effect of convection heat transfer on thermal energy storage unit. Open Phys. 2018, 16, 861–867. [Google Scholar] [CrossRef]
  32. Khanna, S.; Reddy, K.; Mallick, T.K. Performance analysis of tilted photovoltaic system integrated with phase change material under varying operating conditions. Energy 2017, 133, 887–899. [Google Scholar] [CrossRef]
  33. Huang, M.; Eames, P.; Norton, B. Thermal regulation of building-integrated photovoltaics using phase change materials. Int. J. Heat Mass Transf. 2004, 47, 2715–2733. [Google Scholar] [CrossRef]
  34. Nehari, T.; Benlakam, M.; Nehari, D. Effect of the Fins Length for the Passive Cooling of the Photovoltaic Panels. Period. Polytech. Mech. Eng. 2016, 60, 89–95. [Google Scholar] [CrossRef] [Green Version]
  35. Khanna, S.; Reddy, K.; Mallick, T.K. Optimization of finned solar photovoltaic phase change material (finned pv pcm) system. Int. J. Therm. Sci. 2018, 130, 313–322. [Google Scholar] [CrossRef]
  36. Biwole, P.; Eclache, P.; Kuznik, F. Improving the Performance of Solar Panels by the Use of Phase-Change Materials. In Proceedings of the World Renewable Energy Congress, Linköping, Sweden, 8–13 May 2011; Volume 57, pp. 2953–2960. [Google Scholar] [CrossRef] [Green Version]
  37. Rajvikram, M.; Sivasankar, G. Experimental study conducted for the identification of best heat absorption and dissipation methodology in solar photovoltaic panel. Sol. Energy 2019, 193, 283–292. [Google Scholar] [CrossRef]
  38. Wongwuttanasatian, T.; Sarikarin, T.; Suksri, A. Performance enhancement of a photovoltaic module by passive cooling using phase change material in a finned container heat sink. Sol. Energy 2020, 195, 47–53. [Google Scholar] [CrossRef]
  39. Rajvikram, M.; Leoponraj, S.; Ramkumar, S.; Akshaya, H.R.; Dheeraj, A. Experimental investigation on the abasement of operating temperature in solar photovoltaic panel using PCM and aluminium. Sol. Energy 2019, 188, 327–338. [Google Scholar] [CrossRef]
  40. Nada, S.; El-Nagar, D.; Hussein, H. Improving the thermal regulation and efficiency enhancement of PCM-Integrated PV modules using nano particles. Energy Convers. Manag. 2018, 166, 735–743. [Google Scholar] [CrossRef]
  41. Stropnik, R.; Stritih, U. Increasing the efficiency of PV panel with the use of PCM. Renew. Energy 2016, 97, 671–679. [Google Scholar] [CrossRef]
  42. Mahamudul, H.; Silakhori, M.; Henk Metselaar, I.; Ahmad, S.; Mekhilef, S. Development of a temperature regulated photovoltaic module using phase change material for Malaysian weather condition. Optoelectron. Adv. Mater. Rapid Commun. 2014, 8, 1243–1245. [Google Scholar]
  43. Polymers, P. Technical Data Sheet of savE® OM29; PLUSS: Haryana, India, 2017. [Google Scholar]
  44. Waqas, A.; Ji, J. Thermal management of conventional PV panel using PCM with movable shutters—A numerical study. Sol. Energy 2017, 158, 797–807. [Google Scholar] [CrossRef]
  45. Hasan, A.; Sarwar, J.; Alnoman, H.; Abdelbaqi, S. Yearly energy performance of a photovoltaic-phase change material (PV-PCM) system in hot climate. Sol. Energy 2017, 146, 417–429. [Google Scholar] [CrossRef]
  46. Li, Z.; Ma, T.; Zhao, J.; Song, A.; Cheng, Y. Experimental study and performance analysis on solar photovoltaic panel integrated with phase change material. Energy 2019, 178, 471–486. [Google Scholar] [CrossRef]
  47. Siahkamari, L.; Rahimi, M.; Azimi, N.; Banibayat, M. Experimental investigation on using a novel phase change material (PCM) in micro structure photovoltaic cooling system. Int. Commun. Heat Mass Transf. 2019, 100, 60–66. [Google Scholar] [CrossRef]
  48. Abdollahi, N.; Rahimi, M. Potential of water natural circulation coupled with nano-enhanced PCM for PV module cooling. Renew. Energy 2020, 147, 302–309. [Google Scholar] [CrossRef]
  49. Preet, S.; Bhushan, B.; Mahajan, T. Experimental investigation of water based photovoltaic/thermal (PV/T) system with and without phase change material (PCM). Sol. Energy 2017, 155, 1104–1120. [Google Scholar] [CrossRef]
  50. Li, D.; Xuan, Y.; Yin, E.; Li, Q. Conversion efficiency gain for concentrated triple-junction solar cell system through thermal management. Renew. Energy 2018, 126, 960–968. [Google Scholar] [CrossRef]
  51. Klemm, T.; Hassabou, A.; Abdallah, A.; Andersen, O. Thermal energy storage with phase change materials to increase the efficiency of solar photovoltaic modules. Energy Procedia 2017, 135, 193–202. [Google Scholar] [CrossRef]
  52. Gaur, A.; Ménézo, C.; Giroux-Julien, S. Numerical studies on thermal and electrical performance of a fully wetted absorber PVT collector with PCM as a storage medium. Renew. Energy 2017, 109, 168–187. [Google Scholar] [CrossRef]
  53. Sardarabadi, M.; Passandideh-Fard, M.; Maghrebi, M.J.; Ghazikhani, M. Experimental study of using both ZnO/water nanofluid and phase change material (PCM) in photovoltaic thermal systems. Sol. Energy Mater. Sol. Cells 2017, 161, 62–69. [Google Scholar] [CrossRef]
  54. Soares, N.; Costa, J.; Gaspar, A.; Matias, T.; Simões, P.; Durães, L. Can movable PCM-filled TES units be used to improve the performance of PV panels? Overview and experimental case-study. Energy Build. 2020, 210, 109743. [Google Scholar] [CrossRef]
  55. Reiter, C.N.; Trinkl, C.; Zörner, W.; Hanby, V.I. A Dynamic Multinode Model for Component-Oriented Thermal Analysis of Flat-Plate Solar Collectors. J. Sol. Energy 2015, 2015, 1–16. [Google Scholar] [CrossRef] [Green Version]
  56. Chaabane, M.; Mhiri, H.; Bournot, P. Thermal performance of an integrated collector storage solar water heater (ICSSWH) with phase change materials (PCM). Energy Convers. Manag. 2014, 78, 897–903. [Google Scholar] [CrossRef]
  57. Skoplaki, E.; Palyvos, J. On the temperature dependence of photovoltaic module electrical performance: A review of efficiency/power correlations. Sol. Energy 2009, 83, 614–624. [Google Scholar] [CrossRef]
  58. Hendricks, J.H.C.; Sark, W.G.J.H.M. Annual performance enhancement of building integrated photovoltaic modules by applying phase change materials. Prog. Photovoltaics Res. Appl. 2011, 20. [Google Scholar] [CrossRef] [Green Version]
  59. Klugmann-Radziemska, E.; Wcisło-Kucharek, P. Photovoltaic module temperature stabilization with the use of phase change materials. Sol. Energy 2017, 150, 538–545. [Google Scholar] [CrossRef]
  60. Waqas, A.; Jie, J. Effectiveness of Phase Change Material for Cooling of Photovoltaic Panel for Hot Climate. J. Sol. Energy Eng. 2018, 140, 1–19. [Google Scholar] [CrossRef]
  61. Khanna, S.; Reddy, K.; Mallick, T.K. Optimization of solar photovoltaic system integrated with phase change material. Sol. Energy 2018, 163, 591–599. [Google Scholar] [CrossRef]
  62. Abdulmunem, A.R. Passive Cooling By Utilizing the Combined PCM /Aluminum Foam Matrix To Improve Solar Panels Performance: Indoor Investigation. Iraqi J. Mech. Mater. Eng. 2017, 17, 712–723. [Google Scholar]
  63. Maiti, S.; Banerjee, S.; Vyas, K.; Patel, P.; Ghosh, P.K. Self regulation of photovoltaic module temperature in V-trough using a metal-wax composite phase change matrix. Sol. Energy 2011, 85, 1805–1816. [Google Scholar] [CrossRef]
  64. Hachem, F.; Abdulhay, B.; Ramadan, M.; El Hage, H.; El Rab, M.G.; Khaled, M. Improving the performance of photovoltaic cells using pure and combined phase change materials—Experiments and transient energy balance. Renew. Energy 2017, 107, 567–575. [Google Scholar] [CrossRef]
  65. Karthikeyan, V.; Prasannaa, P.; Sathishkumar, N.; Emsaeng, K.; Sukchai, S.; Sirisamphanwong, C. Selection and preparation of suitable composite phase change material for PV module cooling. Int. J. Emerg. Technol. 2019, 10, 385–394. [Google Scholar]
Figure 1. Experimental setup for PV with and without PCM using OM29.
Figure 1. Experimental setup for PV with and without PCM using OM29.
Energies 13 00897 g001
Figure 2. Heat transfer model for PV without PCM.
Figure 2. Heat transfer model for PV without PCM.
Energies 13 00897 g002
Figure 3. Heat transfer model for PV with PCM. Rconv,PV→amb is the convection resistance between the PV front side and ambient; Rrad,PV→sky is the radiation resistance between the PV front side PV and sky; Rconv,PV→surr is the convection resistance between the PV rear side and ambient; Rrad,PV→surr is the radiation resistance between the PV rear side and surrounding; Rcond,PV→PCM is the conduction resistance between the PV rear side and PCM; Rcond,PCM→Al is the conduction resistance between the PCM rear side and PCM back sheet; Rconv,Al→surr is the convection resistance between the rear side of the PCM back sheet and surrounding; and Rrad,Al→surr is the radiation resistance between the rear side of the PCM back sheet and surrounding.
Figure 3. Heat transfer model for PV with PCM. Rconv,PV→amb is the convection resistance between the PV front side and ambient; Rrad,PV→sky is the radiation resistance between the PV front side PV and sky; Rconv,PV→surr is the convection resistance between the PV rear side and ambient; Rrad,PV→surr is the radiation resistance between the PV rear side and surrounding; Rcond,PV→PCM is the conduction resistance between the PV rear side and PCM; Rcond,PCM→Al is the conduction resistance between the PCM rear side and PCM back sheet; Rconv,Al→surr is the convection resistance between the rear side of the PCM back sheet and surrounding; and Rrad,Al→surr is the radiation resistance between the rear side of the PCM back sheet and surrounding.
Energies 13 00897 g003
Figure 4. Experimental day solar irradiance.
Figure 4. Experimental day solar irradiance.
Energies 13 00897 g004
Figure 5. Experimental day ambient temperature.
Figure 5. Experimental day ambient temperature.
Energies 13 00897 g005
Figure 6. Thermal imaging of PV module spot 1 without PCM and spot 2 with PCM: (a) 8:00; (b) 8:30; (c) 9:00; (d) 9:30; (e) 10:00; and (f) 10:30.
Figure 6. Thermal imaging of PV module spot 1 without PCM and spot 2 with PCM: (a) 8:00; (b) 8:30; (c) 9:00; (d) 9:30; (e) 10:00; and (f) 10:30.
Energies 13 00897 g006
Figure 7. Thermal imaging of PV module spot 1 without PCM and spot 2 with PCM: (a) 11:00; (b) 11:30; (c) 12:00; (d) 12:30; (e) 13:00; and (f) 13:30.
Figure 7. Thermal imaging of PV module spot 1 without PCM and spot 2 with PCM: (a) 11:00; (b) 11:30; (c) 12:00; (d) 12:30; (e) 13:00; and (f) 13:30.
Energies 13 00897 g007
Figure 8. Thermal imaging of PV module spot 1 without PCM and spot 2 with PCM: (a) 14:00; (b) 14:30; (c) 15:00; (d) 15:30; (e) 16:00; and (f) 16:30.
Figure 8. Thermal imaging of PV module spot 1 without PCM and spot 2 with PCM: (a) 14:00; (b) 14:30; (c) 15:00; (d) 15:30; (e) 16:00; and (f) 16:30.
Energies 13 00897 g008
Figure 9. Overall PV module temperature with and without PCM.
Figure 9. Overall PV module temperature with and without PCM.
Energies 13 00897 g009
Figure 10. Temperature corrected electrical efficiency of the PV module with and without PCM.
Figure 10. Temperature corrected electrical efficiency of the PV module with and without PCM.
Energies 13 00897 g010
Table 1. Literature review of PCM integrated PV module.
Table 1. Literature review of PCM integrated PV module.
AuthorPCM/LocationSystem DescriptionTPV without
PCM (°C)
TPV with
PCM (°C)
Waqas et al., [44]RT44/PVPCMs are filled in a movable
container to detach from the PV
module back surface when it
turns to be liquid. Yearly
experimental results reveal that
in summer there
occurred a higher TPV reduction.
6442
Hasan et al., [45]RT42/PVYearly performance of RT42
integrated PV module enhances
the electrical efficiency by about
5.9% and higher TPV reduction
was achieved during April.
7161
Zhenpeng et al., [46]Paraffin 35/
China
One month experimental analysis
shows that, on 19 July,
the highest electrical energy conversion
noticed in the range of 499– 524 Wh.
68.753
Leila et al., [47]Sheep fat/
indoor
Sheep fat as PCM performed
better than paraffin wax and
cooled water flows in the channel
to enhance the heat transfer.
8761.5
Nasrin et al., [48]Composed oil/
indoor
Zigzag PCM container along
with the combination of composed
oil (coconut oil+sunflower oil)
with boehmite. The final result
aids in better TPV reduction for
composed oil with boehmite.
7242.5
Sajan et al., [49]RT30/
India
Hybrid application of water-based
PCM integration enhances the
efficiency more than water-based
cooling technique.
7857
Nada et al., [40]RT55/
Egypt
Al2O3 nanoparticle improves the
PCM thermal absorption capability
with an increase in thermal
conductivity.
7549.3
Dianhong et al., [50]Paraffin
50/China
The thermoelectric generator
attached heatsink incorporated behind
the PCM container to utilize the
stored thermal energy of PCM.
6551
Torsten et al., [51]RT54HC/
Qatar
Metal fiber porous foam is
impregnated with PCM to enhance the
higher heat transfer from the PV
module to PCM.
8060
Ankita et al., [52]OM39/
France
A fully wetted absorber channel
removes the thermal energy from the
PV module and it transfers to the
PCM container.
69.1753.86
Sardarabadu et al., [53]Paraffin
42/Iran
Deionized water and ZnO nanofluid
used as working fluid to carry out the
PCM temperature for PVT application.
6245
Modjinou et al., [28]PCM 45/
China
A microchannel heat pipe is used to
circulate the working fluid for both
PVT and PVT- PCM.
7560
Soares et al., [54]RT22/
Portugal
Five days of experimental results
reveal that movable thermal energy
storage system claims negative impact
on peak sunshine hours because
selected RT22 PCM is not capable
of this desired location.
5872
Table 2. Thermophysical properties of PCM [43].
Table 2. Thermophysical properties of PCM [43].
Sl.PropertyRange
1Melting temperature (°C)29
2Freezing temperature (°C)26
3Latent heat of fusion (kJ/kg)194
4Liquid density (Kg/m3)870
5Solid density (Kg/m3)976
6Liquid specific heat capacity (kJ/kgK)2.71
7Solid specific heat capacity (kJ/kgK)2.32
8Liquid thermal conductivity (W/mK)0.172
9Solid thermal conductivity (W/mK)0.293
10Congruent meltingYES
11Thermal stability (No.)∼2000
12Maximum operating temperature120
Table 3. Temperature profile of PV only and PV with PCM.
Table 3. Temperature profile of PV only and PV with PCM.
Time (h)PV without PCM (°C)PV with PCM (°C)
Front SurfaceBack SurfacePV Front SurfacePCM Backsheet
8:0042.244.341.827.3
8:3041.843.840.628.1
9:0042.344.642.928.9
9:3038.743.644.132.5
10:0039.544.144.833.8
10:3041.746.848.337.5
11:0040.746.147.138.6
11:3039.245.845.339.1
12:0036.442.842.638.4
12:3038.144.544.138.7
13:0038.545.344.739.1
13:3041.847.148.541.9
14:0041.647.948.541.6
14:304147.247.741.3
15:0041.246.947.940.9
15:3041.547.148.141.5
16:0040.646.847.241.1
16:3039.346.445.440.9

Share and Cite

MDPI and ACS Style

Elavarasan, R.M.; Velmurugan, K.; Subramaniam, U.; Kumar, A.R.; Almakhles, D. Experimental Investigations Conducted for the Characteristic Study of OM29 Phase Change Material and Its Incorporation in Photovoltaic Panel. Energies 2020, 13, 897. https://0-doi-org.brum.beds.ac.uk/10.3390/en13040897

AMA Style

Elavarasan RM, Velmurugan K, Subramaniam U, Kumar AR, Almakhles D. Experimental Investigations Conducted for the Characteristic Study of OM29 Phase Change Material and Its Incorporation in Photovoltaic Panel. Energies. 2020; 13(4):897. https://0-doi-org.brum.beds.ac.uk/10.3390/en13040897

Chicago/Turabian Style

Elavarasan, Rajvikram Madurai, Karthikeyan Velmurugan, Umashankar Subramaniam, A Rakesh Kumar, and Dhafer Almakhles. 2020. "Experimental Investigations Conducted for the Characteristic Study of OM29 Phase Change Material and Its Incorporation in Photovoltaic Panel" Energies 13, no. 4: 897. https://0-doi-org.brum.beds.ac.uk/10.3390/en13040897

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