Next Article in Journal
Towards Achieving Zero Carbon Targets in Building Retrofits: A Multi-Parameter Building Information Modeling (BIM) Approach Applied to a Case Study of a Thermal Bath
Previous Article in Journal
Exergy Analysis and Off-Design Modeling of a Solar-Driven Supercritical CO2 Recompression Brayton Cycle
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

PV Temperature Prediction Incorporating the Effect of Humidity and Cooling Due to Seawater Flow and Evaporation on Modules Simulating Floating PV Conditions

1
Laboratory of Soft Energy Applications and Environmental Protection, University of West Attica, 12201 Athens, Greece
2
School of Engineering, Faculty of Science, University of East Anglia, Norwich Research Park, Norwich NR4 7TJ, UK
*
Author to whom correspondence should be addressed.
Energies 2023, 16(12), 4756; https://doi.org/10.3390/en16124756
Submission received: 15 May 2023 / Revised: 4 June 2023 / Accepted: 8 June 2023 / Published: 16 June 2023
(This article belongs to the Section A2: Solar Energy and Photovoltaic Systems)

Abstract

:
The temperature prediction for floating PV (FPV) must account for the effect of humidity. In this work, PV temperature prediction for steady-state Tpv and transient conditions Tpv(t) incorporates the effect of humidity and cooling due to seawater (s.w.) splashing and evaporation on PV modules. The proposed formulas take as main inputs the in-plane solar irradiance, wind speed, ambient temperature, relative humidity (RH), and s.w. temperature. The transient effects of s.w. splashing and the evaporation of the s.w. layer from the module are theoretically described considering the layer’s thickness using Navier–Stokes equations. Tpv and Tpv(t) measurements were taken before and after s.w. splashing on c-Si modules at the seashore and inland. PV temperature predictions compared to measured values showed very good agreement. The 55% RH at the seashore versus 45% inland caused the Tpv to decrease by 18%. The Tpv(t) at the end of the s.w. flow of 50–75 mL/s/m on the module at the seashore was 35–51% lower than the Tpv inland. This Tpv(t) profile depends on the s.w. splashing, lasts for about 1 min, and is attributed to higher convection, water cooling, and evaporation on the modules. The PV efficiency at FPV conditions was estimated to be 4–11.5% higher compared to inland.

1. Introduction

It is well established that when solar irradiance, IT, impinges on a PV module, a small part is reflected, while the main part is transmitted and absorbed. The latter is converted into power with efficiency ηpv and the rest is degraded into heat, which increases the PV cell temperature, Tpv, above the ambient temperature, Ta. The steady-state PV temperature, as well as the Tpv(t) profile at transient conditions for modules operating on land, has been extensively studied through simulation models taking into consideration wind speed and the effect of PV temperature through implicit functions [1,2,3], the varying environmental conditions [4,5,6,7], and also the PV module inclination and orientation with reference to the wind direction [8,9]. These models were primarily developed for land-based PV (LBPV), resulting in a generally good agreement between predicted and measured values. The effect of the environmental conditions on Tpv, and hence on Pm in c-Si modules, was further investigated theoretically and experimentally to cover a wide range of mounting configurations, PV technologies, and years of operation, and a model for Tpv prediction was developed in [10], achieving a very high accuracy for any PV installations that were free-standing, building-integrated PV (BIPV), or building-adapted PV (BAPV), and also incorporated the effect of aging. Other empirical and semi-empirical formulas that predict Tpv and Tpv(t) have been proposed and validated in a large diversity of PV configurations [11,12,13,14], while formulas based on artificial neural networks (ANNs) [15,16,17] predicted Tpv for FPV installations studied for each case. Tpv is a crucial factor that affects efficiency, ηpv, and evidently the PV power output, Pm, [18], while high Tpv and humidity accelerate the PV cell’s aging [19,20,21,22]. In the last decade, the study of the PV performance in lakes and reservoirs, nearshore and offshore, has been given special attention because of interest in floating PV (FPV) [23,24,25,26], with a focus on the effect of the environmental conditions, such as IT, wind speed, vw, ambient temperature, Ta, seawater (s.w.) temperature, Ts.w., humidity, hu, and salinity on Tpv, accounting for the different values those parameters take inland and on the seashore/offshore [27,28]. The induced PV cooling to partially recover Pm losses from the effect of Tpv has been studied using various techniques. Heat extraction by means of water flowing on the front PV side [29,30,31,32] was reported to cause a Tpv decrease of 20%, which increased with the flow rate. The water jet impingement or spray on the front and/or back PV sides [33,34] resulted in 14% recuperation of the power losses. PV cooling using either a soaked sponge or water spraying on the PV backside was reported to lead to the recuperation of Pm losses greater than 28% [35,36,37].
FPV installations in near-shore environments exposed to high hu and salinity showed a boost in performance vs. LBPV [23,24,26]. Recent works focus on the effect of the IT, vw, Ta, Ts.w., hu, salinity, and FPV installation geometries on Tpv. The prospects and challenges of various types of modules in FPV installations were discussed in [25,38,39,40], while the impact of climatic conditions, with a special emphasis on FPV, was estimated in [41,42] and in the tropics in [43,44,45] using simulation models. The estimation of Tpv in FPV plants was further studied in [46,47,48,49] with an emphasis on the impact of hu [50,51,52] and the effect of the solar spectral irradiance on the PV yield in [53].
Generally, the Tpv prediction formulae incorporated the heat loss coefficient, Upv, for FPV systems that were either air-cooled or air/water-cooled [45,46,47], while the hu effect was not considered explicitly. In [47], two FPV-mounting technologies were compared, and Tpv was found to be lower in air-cooled FPV than in the FPV in direct contact with water. The larger drop in the Tpv for the air-cooled FPV was attributed to the higher vw. In [41], Tpv was determined to be 9–14 °C lower in the sea environment (s.e.) compared to inland in two climatic conditions, the Netherlands (NL) and Singapore (SG). Tpv was found to be 13 °C lower in large- and medium-footprint FPV compared to rooftop PV in SG and by 11 °C lower in offshore FPV compared to LBPV in the NL due to higher vw and hu in the s.e. Other effects on Tpv due to solar irradiance spectrum [27], FPV mounting geometry [41], air/water environmental quantities [28] using appropriate databases, and the evaporation from the sea surface [54] and the PV module must be considered. It is important to note the lower albedo at s.e., about 8% compared to 15% on the ground [38,44,45]. However, due to the low tilt of the modules in the s.e., the effect of albedo is not significant.
The above literature review addresses the following main issues:
  • The Effect of RH on Tpv
Tpv decreases as RH increases [43,49], while [48] concludes that RH does not play an important role in the performance of the Tpv and PV. On the contrary, in [50,52], it is claimed that Tpv increases with RH and Pm decreases. However, in low RH, the PV performance may also increase with RH based on the combined effect of the environmental quantities given in [49,50,51]. The aforementioned findings are opposite, justifying the need for a further investigation of the effect of RH on Tpv.
2.
The Combined effect of wind and water on Tpv
Τpv in FPV was found to be 5–10 °C lower than in LBPV due to the cooling effect of the water evaporation and the higher vw [44,46,47,48]. Comparing FPV and LBPV in the NL and SG in [41], it was found that Tpv was 3.2 °C lower in FPV at s.e. vs. LBPV in the NL and 14.5 °C lower at s.e. vs. rooftop PV in SG. Additionally, in the aforementioned study, it was found that RH was 4.7% higher in s.e. than in rooftop PV in nearby sites, where Ta was 1–1.2 °C lower in s.e. and vw was 1 m/s higher due to lower air friction.
3.
Spectral effects
Tpv is affected by the spectrum shift towards the red due to solar light scattering on aerosol molecules. As the humid environment causes a spectrum redshift, which corresponds to the infrared part of the spectrum, the performance of c-Si FPV might be lower than the LBPV considering that c-Si PV have a spectral response in the 300–1100 nm range. Thin film technologies such as a-Si or CdTe are less affected by the redshift as their spectral response is mainly in the visible part of the spectrum [18,51].
4.
Atmospheric effects
The FPV performance was found to be 12.96% higher in the s.e. than the LBPV [49]. The global horizontal irradiation was about 8.54% higher in the s.e. because the sky above the sea is in general clearer than over the green land. The transparency of the atmosphere depends on the combined effect of Ta, Ts.w., vw, and air and water vapor pressure. Aerosols and air moisture shift the light spectrum towards the red and may attenuate IT in a rich s.w. evaporation area with low vw or when wind transfers water droplets in the air. This has a negative effect, decreasing Isc and Pm [44,46,52,53]. On the contrary, the convection coefficient, hc, increases with hu, causing a Tpv decrease (positive effect) [15,16,49]. These opposite effects may be the reason the Tpv results in the FPV installations do not always agree, as mentioned in point 1 above.
5.
Tpv modeling in FPV
A Tpv prediction model for FPV is outlined in [48], where the effect of Ts.w. is indirectly accounted for in the Upv estimation, whereas the effect of vw and hu is not incorporated. For FPV of various water footprints, Upv was estimated as 27 W/m2K inland, 37 W/m2K near-shore, and 57 W/m2K offshore [46,47]. In [16], it is stated that the existing models overestimate Tpv in s.e. because the cooling effect on PV through Ts.w. and hu is not accounted for. The correlation between Tpv and RH is considered through either regression analysis [43] or ANN where, depending on the PV cell type and the site conditions, RH has either a positive or negative effect on Tpv [15].

FPV Research Gaps and Objectives

Natural PV cooling due to the increased hu on the seashore/near-shore/offshore, along with the different values in the environmental parameters between inland and seashore/offshore, as well as the seawater splashing on the modules and the evaporation on their surface have not been rigorously studied, especially as it concerns their effect on Tpv and ηpv. This paper fills in gaps in the literature as it concerns the natural and induced PV cooling of modules operating in the s.e. A robust theoretical analysis on the effect of RH, s.w. temperature, s.w. splashing on the modules, and the s.w. layer evaporation on the front side of the modules on the transient Tpv(t) profile is outlined. A complete set of formulas is elaborated to predict Tpv, expanding the previous work of the authors [10], taking into account the effect of hu and indirectly of Ts.w. or freshwater temperature Tw. For the validation of the model, the predicted Tpv values are compared to the measured ones inland and on the seashore and compared further to Tpv values produced using other formulas proposed in [14,15,16,43,48].

2. Experimental Procedure for the Measurement of the Tpv Profiles on the Seashore and Inland

The steady-state temperature Tpv and the temperature profile Tpv(t) on the front side, also denoted as Tf(t), of mono c-Si M55 modules operating for 24 years with ηpv = 0.095 due to ageing, and mono c-Si SW80 modules with ηpv = 0.146, were measured at two sites. The modules were placed facing South with inclination β = 35° on the seashore at latitude φ = 38.311° N and L = 21.78° E and on the terrace of a building inland, φ = 38.22° N and L = 21.75° E. Tpv was measured in March under a bright environment with high air visibility and no water drops or water layer deposited on the PV modules, at RH equal to 55% on the seashore and 45% inland. The experiments were carried out at around solar noon under IT = 800 W/m2. The steady-state front side PV temperature, Tf = Tpv, was measured around the middle on one of the cells.
The study of the PV cooling included s.w. flow on the inclined module, simulating s.w. splashing on the module. Ts.w. was 14–15 °C during the experiments. The steady-state Tpv was measured prior to the s.w. flow on the module. The s.w. volume flow rate, Q, per unit width b, (Q/b), was 40, 50, and 75 mL/s/m and was sustained for a period of 2 to 6 s depending on the experiment. The measurement of the transient profile Tf(t) = Tpv(t) of the s.w. layer on the module began when the s.w. flow on the module stopped and continued until the layer was totally evaporated. The research focused on the effect of hu on Tpv under a clear sky and the effect of the s.w. flow on the module and the subsequent s.w. layer evaporation from its front side.
The environmental conditions at the time of the experiments were measured with the following sensors. A Kipp & Zonen CMP11 pyranometer (ISO 9060 & IEC 61724 Class A) coupled to a METEON data logger with an accuracy of 0.1% was used for the intensity, IT, on the PV plane. Ta and RH were measured using a thermohygrometer with an accuracy of ±0.5 °C for the Ta and ±2% for the RH. The PV temperature Tf(t) = Tpv(t) was measured using a TROTEC TP10 IR thermometer with an accuracy of ±1 °C. The wind speed vw was measured using a Sefram 9862 hot wire anemometer with an accuracy of ±3%. During the measurements, vw varied by 0.5 m/s which caused variation in the Tpv of ±1 °C. This band is the same as the accuracy of the temperature sensor TROTEC TP10 used in the measurements. An error analysis of Tpv prediction, taking into account the error in the measurement of the input parameters, is given in Section 4.3.

3. Theoretical Analysis of Tpv Profiles on the Seashore vs. Inland

The effect of hu on Tpv and also the Tpv(t) profiles due to s.w. splashing on a module and the s.w. evaporation from the module front side were studied theoretically and applied in the measurement scenarios for modules operating on the seashore and inland.
The steady-state PV temperature was predicted by incorporating a humidity correction factor improving the Tpv prediction formula previously proposed for free-standing, BIPV/BAPV [10], and adapting it to the s.e., freshwater environment, as well as humid inland environments.
The Tpv(t) = Tf(t) profile was studied considering conduction, convection, and s.w. layer evaporation on the module, while the s.w. layer’s thickness was estimated by solving the Navier–Stokes equation. These heat and mass flow processes were considered in the study of the Tpv(t) profiles as they contribute towards the heat extraction from the modules.

3.1. Steady-State Tpv Prediction Model

For modules operating inland or in the s.e. under the conditions of IT, Ta, vw, Ts.w., RH, and inclination β, Tpv may be predicted through the following set of formulas. Generally,
Tpv = Ta + f·IT
A generalized form of the coefficient f, known in its simplest form as the Ross coefficient in Equation (1), is predicted for any PV mounting configuration by Equations (2a,b) and (3), which hold for an average RH = 45%, based on the previous work of the authors [10]:
f = f v w 1 ϑ η p v ϑ T p v δ Τ p v + η p v I T δ I T 1 η p v , S O C 1 U f T p v δ Τ p v + U b T p v δ T p v + U f β δ β + U b β δ β U p v , S O C
f = f v w 1 ϑ η p v ϑ T p v δ T p v + η p v I T δ Ι Τ 1 η p v , S O C
Equation (2a) holds for natural flow or vw < 1.5 m/s, and Equation (2b) holds for air-forced flow or vw > 1.5 m/s.
f v w expresses the impact of wind speed on the f coefficient and is given by:
f v w = a + b v w 1 + c v w + d v w 2
where a = 0.0375, b = 0.0081, c = 0.2653, and d = 0.0492 [10].
Equation (2a,b) is expressed through the multiplication of f v w with correction factors due to the deviation of the efficiency ηpv from a reference value at standard operating conditions (SOCs) and the effect of the PV temperature and module inclination on the heat losses coefficients in the front and back side of the modules, Uf and Ub, respectively. The factors ηpv,soc, Upv,soc, dηpv/dT, dηpv/dIT, dUf/dT, dUb/dT, dUf/dβ, and dUb/dβ are defined and determined in [10].
In the present work, the applicability of the above expression is extended to the s.e., freshwater environment, and humid inland environment, with the introduction of a new correction factor in the model for f, expressing the effect of RH and Ts.w. or Tw on Tpv.
The humidity affects the convection coefficient hc due to the higher dynamic viscosity of H2O. The hc,a with dry air as a coolant vs. hc,s.w. with s.w. as a coolant may be expressed through the ratio of their corresponding Nu number [55]. The ratio hc,a/hc,s.w. is practically estimated by Equations (4)–(6) to account for the effect of hu on hc. The deviation of RH from its reference value 45% causes a change in δhc or δUpv to be introduced in the prediction of f and Tpv as outlined below. Equations (4) and (5) stand for turbulent and laminar forced flow over a flat surface, respectively, and Equation (6) holds for natural heat transfer.
hc,a/hc,s.w. = (ka/ks.w.)·(Pra/Prs.w.)1/3·(νas.w.)0.8 forced flow, turbulent, Re > 5 × 105
hc,a/hc,s.w. = (ka/ks.w.)·(Pra/Prs.w.)1/3·(νas.w.)1/2 laminar forced flow, Re < 5 × 105
hc,a/hc,s.w. = (ka/ks.w.)·(Pra/Prs.w.)1/3·(νas.w.)1/3 for natural heat flow
where ka and ks.w. are the thermal conductivity of air and s.w., respectively, in (W/mK); νa and νs.w. are the kinematic viscosity of air and s.w., respectively, in (m2/s); and Pra and Prs.w. are the Prandtl numbers of air and s.w., respectively.
The difference in hc is estimated by Equation (7a,b) for the following cases:
  • RH1 (phu,1% moles of dry air and qhu,1% moles of H2O), with phu,1 + qhu,1 = 1.
  • RH2 (phu,2% moles of dry air and qhu,2% moles of H2O), with phu,2 + qhu,2 = 1.
hc,2 − hc,1 = (phu,2 − phu,1)·hc,a + (qhu,2 − qhu,1)·hc,H2O
δhc,hu = δphu·hc,a + δqhu·hc,H2O
δphu and δqhu denote the difference in phu and qhu at two different conditions where RH differs. These are determined from the Mollier diagrams.
The new correction factor for coefficient f is presented in Equation (8). It accounts for the effect of the difference between RHs.e. and RHinl and gives the fs.e. for the s.e. in relation to finl for inland, where RH = 45%. The correction term δhc,hu/Upv is estimated by Equation (7b), while Upv = hc,f + hc,b + hr,f + hr,b is estimated as outlined in [8,10].
fs.e. = finl·(1 − δhc,hu/Upv)
The above correction factor is introduced through the last term (1 − δhc,hu/Upv) in Equation (9), improving the previous holistic model for f [10].
f = f v w 1 δ η p v 1 η p v , S O C 1 δ U p v U p v , S O C 1 δ η a g 1 η p v , S O C 1 δ η t e c 1 η p v , S O C 1 δ h c , h u U p v , S O C
Note that the second correction term 1 δ U p v U p v , S O C applies only for natural flow (or vw < 1.5 m/s) and is omitted for turbulent flow (or vw > 1.5 m/s). The third and fourth correction terms describe the impact of ageing and PV technology/efficiency, respectively, as outlined in [10].
The introduction of the last term 1 δ h c , h u U p v , S O C accounts for the effect of hu and Ts.w. or Tw as per case in the prediction of Tpv and applies to FPV both at s.w. and freshwater environments as well as LBPV at RH conditions different than 45%.

3.2. Transient Effects in Tpv Due to Water Splashing on the PV Module

3.2.1. Seawater Layer Thickness

Let Δxo be the thickness of the s.w. layer which flows steadily on an inclined module. Due to gravitational forces, when the flow stops, the thickness of the s.w. layer becomes thinner with time, Δx(t). Assuming fully developed laminar flow, both Δxo and the s.w. layer velocity profile, u(y), are determined by the Navier–Stokes equations for a Newtonian non-compressible fluid at steady-state flow on the module; see Equations (10)–(14) [55,56]. y is the distance measured from the module plane along a y-axis normal to the PV plane, where 0 < y < Δx.
0 = νd2u(y)/dy2 + ρ g sin(β) = 0
For boundary conditions u(y = 0) = 0 and du/dy = 0 at y = Δx:
u(y) = (gsin(β)/ν) (yΔx − y2/2)
The volume flow rate Q per unit width b, that is the width of the string of PV cells in a module over which s.w. flows, is given by Equation (12), while Δxo is given by Equation (13).
Q / b = 0 Δ x o u y d y
Δxo = [3ν·Q/(b·g·sin(β))]1/3
where ν is the s.w. kinematic viscosity obtained from the literature. Substituting ν, Q/b, and β into Equation (13) gives Δxo equal to 0.32 mm and 0.37 mm corresponding to Q/b = 50 and 75 mL/s/m, respectively, for s.w. at 14 °C, while Δxo = 0.28 mm and 0.32 mm, respectively, for the s.w. layer of Ts.w. = 30 °C. Δx(t) thins exponentially when the s.w. stops flowing according to Equation (14) [57], and so its contribution to heat capacity becomes negligible soon.
d(Δx)/Δx = −uav t/l
Δx(t + δt) = Δx(t)exp(−uav(t)·δt/l)
where l is the length along the s.w. flow with l = 0.1 m where Tf(t) was measured. uav is the average s.w. layer velocity at l when s.w. flows down the module and is a function of t. Equations (11)–(13) provide uav(t), where:
uav(t) = gsinβ(Δx(t))2/3νf
The subscript f denotes that the value of ν corresponds to Tf. Δx in Equation (15) holds when the viscous forces and surface tension, σ, are low compared to gravitational forces. This holds for Δx bigger than a critical value, hcr [57]:
hcr = 2dσ/dz(1 − cosθ)/ρg
where θ = 107° is the contact angle of the water–glass interface. dσ/dz is the derivative of the surface tension taken equal to σ/Lpv, with Lpv = 1.2 m corresponding to the length of the module, and z is an axis along the inclined module. Therefore, dσ/dz = 73.477 × 10−3 N/m/m at Ts.w. = 30 °C and salinity 35 g/kg. Substituting these values in Equation (17) gives hcr = 0.019 mm.
For t = 0 when Δx = Δxo, uav is estimated equal to 0.157 m/s at Ts.w. = 14 °C, and 0.175 m/s at Ts.w. = 30 °C. uav is negligible at thickness hcr. Δx(t) and uav(t) are determined at l = 0.1 m in steps of δt = 0.1 s following iteration between Equations (15) and (16). The results are shown in Figure 1. Δx reaches the hcr value at around 60 s, where the water film shows no motion because viscous forces and surface tension between water and glass prevail.

3.2.2. Tpv(t) Profile Taking into Consideration the s.w. Layer on the PV

The transient Tpv(t) profile due to s.w. splashing is expected to experience an initial steep drop. The time constant τs.w. of the heat convection due to the s.w. layer on the module is determined by Equation (18). Introducing into Equation (18), the s.w. density ρ, its heat capacity cp, the total heat transfer coefficient Upv, and Δxo, τs.w. is estimated to be 57 s at the beginning of the phenomenon.
τs.w. = ρ Δx cp/Upv
The heat conduction acts directly in full strength in such thin layers. The initial Tpv drop may be estimated applying the continuity in the heat conduction flow along a y-axis normal to the boundary between the PV glass and the s.w. layer, Equation (19), [56].
kgl dTf/dy = ks.w. dTs.w./dy
where kgl = 1 W/mK and ks.w. = 0.6 W/mK stand for the thermal conductivity of glass and s.w., respectively.
After the initial drop, the s.w. layer temperature Tf(t) is expected to increase according to (1 − exp(−t/τs.w.)) as the heat exchange due to convection on the module prevails after the splash, whereas the s.w. layer thins exponentially, Equation (15), and τs.w. consequently decreases fast.
Tpv(t), denoted here as Tf(t), is determined by Equations (20)–(26), taking into consideration convection and heat radiation. Here, the time constant τ refers both to the module and the s.w. layer as an ensemble. Based on the analysis in [4], Tf(t) is given by Equation (20):
[F1 − F2(U)(Tf(t + δt) − Ta)]/[F1 − F2(U)(Tf(t) − Ta)] = exp(−δt/τ)
Tf(t + δt) is the PV front side temperature at t + δt and Tf(t) at time t. To predict Tpv(t + δt) = Tf(t + δt), the following factors F1, F2(U) are required.
F1 = Ac((τα) − ηpv)IT/(mc)ef
F2(U) = Ac[Uf−a + Ub−a[(1 + (Uf−a/Uc−s.w.))/(1 + (Ub−a/Uc−b))]]/(mc)ef
(mc)ef = [(mc)f + (mc)EVA] + (mc)c[(Uc−f + Uf−a)/Uc−f] + [(mc)EVA + (mc)b][(Uc−b/(Uc−b + Ub−a)] [(Uc−f + Uf−a )/Uc−f]
(mc)ef is the effective heat capacity of the module determined from the heat capacities of the cell components; typical values are provided in [4]. Equations (22) and (23) must be corrected to take into account the heat capacity of the s.w. layer on the module by substituting (mc)f with ((mc)f + (mc)s.w.), Uc−f−1 with (Uc−f−1 + (Δx/ks.w.)), and Uf−a with Us.w.−a. This correction is important for about 10 s after the end of the s.w. splash on the module. Equation (22) may be simplified to:
F2(U) = Ac(Us.w.−a + Ub−a)/(mc)ef
The time constant of the Tpv(t) profile is given by:
τ = 1/F2(U)
According to Equations (24) and (25), τ increases with (mc)ef, i.e., with the glass and the s.w. layer thickness, and decreases with vw. Uf−a and Ub−a must take into account the increase in hc,f and hc,b due to RH, see Section 3.1. τ lies between 1.5 min in windy cases to 3.5 min in calm conditions. On the other hand, the s.w. layer with an initial thickness of 0.32–0.37 mm, as aforementioned, and cp = 4.0 J/gK exhibits initially a time constant τs.w. = 57 s which decreases fast as the s.w. layer thins. Tf undergoes a sudden drop (phase 1) determined by Equation (19) and then increases by exponential increments, Equation (20), mainly through heat convection, (phase 2). However, as Tf(t) increases, the s.w. layer evaporation starts prevailing, (phase 3), as analyzed in Section 3.3 and presented in the results in Section 4.

3.3. Evaporation Rate of Seawater Layer from the Module and the PV Cooling Effect

Evaporation of the s.w. layer from the module is a main cooling process when Ts.w.(t) is practically equal to Tf(t) = Tpv(t) and occurs at t > 10 s. The mass rate of the water evaporation from the module mev (g/s) may be given by converting the partial water vapor pressure to humidity ratio, hu, based on [58]:
mev = Uev Apv (hus − hu)/3.6
where Uev is an empirical evaporation coefficient (kg/m2h) given by Uev = 25 + 19vw, Apv is the module surface (m2), hus is the maximum humidity ratio of saturated air (kg H2O/kg dry air) at Ta = Ts.w., and hu is the humidity ratio (kg H2O/kg dry air) at Ta.
The heat rate q(W) required for the evaporation of mev is calculated from Equation (27).
q = hev mev
where hev is the evaporation heat 2370 J/g and 2345 J/g for s.w. with salinity 35 g/kg at Ts.w. = 20 °C and 30 °C, respectively. It is underlined that q(W) causes a Tpv(t) decrease with rate δTpv(t)/δt according to:
q = (mc)δTpv(t)/δt
The efficiency ηpv of the PV module is given by [59]:
ηpv = ηref(1 − γ(Tpv − Tref) + δ·log(IT/1000))
γ is the temperature coefficient 0.4–0.5%/°C, and δ for c-Si is 0.12. ηref is the efficiency of the module at standard test conditions, STC, (IT = 1000 W/m2, Tref = 25 °C, air mass AM = 1.5) given by the manufacturer.
Equation (30) gives the ηpv recuperation, δηpv, in relation to the decrease in Tpv as a result of the evaporation cooling and the increase in hc due to the higher hu in the s.e.
δηpv = −ηref·γ·δTpv

4. Results and Analysis

The steady-state Tpv and the transient Tpv(t) profiles for the two c-Si modules operating inland and on the seashore are measured and compared to the predicted profiles following the theoretical analysis presented in Section 3. Additionally, the steady-state Tpv is compared to the predicted values determined using seven other formulas from the literature as provided in Section 4.4.

4.1. Experimental Tpv(t) Profiles on the Seashore vs. Inland and Interpretation of the Seawater Splashing Effect

Measured Tpv(t) profiles at the seashore and inland sites are shown in Figure 2 and Figure 3, respectively. Both present the recorded Tpv at steady-state conditions and the transient Tpv(t) profile measured from t = 0, just when the water splashing on the module ends.
Figure 2 shows the Tpv(t) profile of the SW80 module operating on the seashore around solar noon under IT = 800 W/m2, vw = 0.5–1.0 m/s, Ta = 20 °C, Ts.w. = 15 °C, and RH = 55%, with module inclination β = 35°. The steady-state Tpv was measured 37 °C before s.w. splashing, as shown in Figure 2 for t < 0. Figure 3 shows the measured Tpv(t) profiles as well as the measured Tpv for the SW80 and M55 modules (45 °C and 47 °C, respectively) on a terrace inland under the same IT and vw as above, Ta = 21 °C and RH = 45%. The Tpv of the SW80 was 8 °C (45–37 °C), i.e., 18%, lower on the seashore vs. inland (see Figure 2 and Figure 3). This is in agreement with the results reported in [41,45]. The 8 °C difference is attributed mainly to the effect of hu on the seashore and is confirmed in Section 4.3.
Equations (2a,b) and (3) may predict Tpv for inland, and since the measured RH was 45%, there is no need for the hu correction. For the above experimental conditions, the predicted Tpv = Ta + f·IT = 21 + 0.031·800 = 45.8 °C, which is in good agreement with the measured Tpv = 45 °C for the SW80 module (Tpv,inl−2 curve in Figure 3).
In Figure 3, the steady-state Tpv of the two modules shows 2 °C difference which is due to the difference in their ηpv. This can be derived by theoretically combining Equation (1) and the simplified form f = (1 − ηpv)/Upv,inl which gives Equation (31).
δTpv = δf·IT = −(δηpv/Upv,inl)·IT
Introducing Equation (31), Upv,inl = 23 W/m2K [48], IT = 800 W/m2, and the values of ηpv (9.5% and 14.6% for the M55 and SW80, respectively) give δTpv = 1.8 °C ≈ 2 °C, which is the Tpv difference between the two curves in Figure 3.
For the s.w. volume flow rate per unit width Q/b = 50 mL/s/m, the Δxo = 0.32 mm, as estimated in Section 3.2.1. When s.w. stopped flowing, Δx(t) decreased (Figure 1). The Tpv(t) profiles in Figure 2 and Figure 3 at t = 0 show a sudden drop due to s.w. splashing on the module, with Ts.w. = 15 °C. For 0 < t < 1 s, phase 1 (see Figure 2), the phenomenon may be approximated by Equation (19). From the experimental data in the seashore environment, the steady-state Tpv = Tf = 37 °C and Ts.w.(t = 0) = 15 °C. Using kgl = 1 W/mK and ks.w. = 0.6 W/mK, Equation (19) gives Tf = 28.75 °C on the seashore which is in very good agreement with the measured minimum Tf = 29 °C (see Figure 2). The temperature drop is 8 °C or 22%. Similarly, Tf drops to 35 °C and 37 °C for the SW80 and M55, respectively (Figure 3). In this case, Equation (19) predicts Tf = 33.75 °C and 35 °C, respectively, which are in good agreement with the measured data.
During t = 1–10 s, phase 2 (see Figure 2), Δx thins fast according to Equation (15). At t = 2 s, Δx = 0.116 mm, and at t = 10 s, Δx = 0.056 mm, whereas τs.w. < 10 s. Tpv(t) increases fast following 29 °C (1−exp(−t/τs.w.)) as the heat exchange due to convection starts prevailing. The increasing Tpv(t) profile (see Section 3.3) may not reach the steady-state Tpv because as Tpv(t) increases, the s.w. evaporation on the module acts as an additional cooling mode. During phase 2, Tpv(t) is 3–5 °C lower than Tpv for about 15–20 s, as shown in Figure 2 and Figure 3.
Tpv(t) in Figure 4 was obtained under the same conditions as the profile in Figure 2. In this case, a lower flow rate Q/b = 40 mL/s/m results in a thinner s.w. layer, as described by Equation (13). In Figure 4, Tpv(t) shows a steep initial drop to 29 °C and then increases fast due to smaller τs.w. In phase 2, the Tpv(t) profile is higher than the profile in Figure 2 but shorter due to the lower Q/b and to the s.w. evaporation on the module which has an earlier onset (see Figure 2 and Figure 4). Then, Tpv(t) decreases, during phase 3, following the same profile as the curve in Figure 2.
A higher flow rate, Q/b = 75 mL/s/m split in 3 s.w. shots of 25 mL/s/m every 2 s, led to a Tpv drop of 19.2 °C, as shown in Figure 5. Applying Equation (19) three times, starting with Tpv = 37 °C and Ts.w. = 15 °C, allows the s.w. layer’s temperature to be theoretically determined at the end of each shot. In the third iteration, Tpv(t = 0) is determined equal to 20.4 °C, which is close to the experimental value of 17.8 °C (see Figure 5). The latter has a 52% decrease from the steady-state value of 37 °C. Then, Tpv(t) increases as outlined above during phase 2, and when it reaches 23 °C, it starts decreasing during phase 3 because the evaporation of the s.w. layer on the module prevails. In Figure 5, the rate of s.w. evaporation on the module, mev, is smaller because Tpv(t) is lower than in Figure 2 and Figure 4. The three cases presented in Figure 2, Figure 4, and Figure 5 are shown together for comparison in Figure 6.
In Figure 2, Figure 3, Figure 4 and Figure 5 and for around 10–15 s (phase 2), Tpv(t) increases but stays lower than the steady-state value due to s.w. layer evaporation. Thereafter, a significant Tpv(t) decrease appears because the s.w. layer evaporation prevails in phase 3. Specifically, for the case of Q/b = 75 mL/s/m, the Tpv(t) during phase 3 sustains an average value of 22 °C for more than 60 s. This is a 15 °C drop from the steady-state value measured on the seashore and corresponds to a 40% decrease (Figure 5). Compared to the steady-state Tpv measured inland (45 °C), this corresponds to a 51% decrease. On the other hand, for Q/b = 40–50 mL/s/m (Figure 2 and Figure 4), the average decrease in Tpv(t) from its steady-state Tpv value on the seashore, during phase 3, was about 20% and was sustained for 60 s (Figure 6) until the s.w. film evaporated. Compared to the steady-state Tpv measured inland (45 °C), this decrease corresponds to about 35%. Then, Tpv starts increasing during phase 4, following the function (1 − exp(−t/τ)) with a higher time constant, τ = 2 min, according to Equation (18).

4.2. Seawater Evaporation from the PV Module and Its Effect on Tpv

Figure 7 shows the Tpv(t) of the M55 at the inland site after s.w. splashing on the module. The steady-state Tpv = 42 °C underwent a steep drop by 12 °C after s.w. splashing with Ts.w. = 14 °C and under Ta = 17 °C, RH = 45% and vw = 2 m/s. Equation (19) predicts that Tpv drops initially by 10.6 °C, which is close to the experimental drop (Figure 7). Tpv(t) increases fast during phase 2 and then decreases to around 30 °C during phase 3. This is a 29% decrease from its Tpv value and remains at this lower temperature for 100 s until the water layer entirely evaporates, as theoretically confirmed below. Finally, Tpv increases towards the steady-state Tpv in 5τ, which is in total t = 5 × 2 min = 600 s.
The Mollier diagrams for the above conditions and for Ts.w. = 32 °C during phase 3 give hus = 0.032 g s.w./g dry air and hu = 0.0059 g s.w./g dry air. Then, Equation (26) for vw = 2 m/s gives mev/Apv = 0.45675 g/m2s and Equation (27) gives q = 1071.1 W/m2, while Equation (28) for (mc) = 3000 J/m2K for the PV module gives δTpv/δt = 0.357 °C/s. In Figure 7, Tpv(t = 10) = 35 °C and Tpv(t = 30) = 30 °C. Hence, the time for the temperature to decrease from 35 °C to 30 °C is estimated 5 °C/0.357 °C/s = 14 s. Therefore, Tpv(t) reaches its lower value during phase 3 in 10 + 14 = 24 s compared to the experimental 30 s in Figure 7.
The time period for the s.w. layer Δx(t = 10) to evaporate equals δtev = Δx(t = 10)/(mev/Apv). Section 3.2 gives Δx(t = 10) = 0.043 mm and, hence, δtev = 94 s. Therefore, the time needed for the s.w. layer to evaporate is equal to 10 s + 94 s = 104 s which is confirmed in Figure 7.
The Tpv,inl-2(t) in Figure 3 shows that in phase 3, the average Tpv(t) = 42 °C while Ta = 21 °C. The Mollier diagrams give hus = 0.051 and hu = 0.0059 g s.w./g dry air. Equation (26) gives mev/Apv = 0.56 g/m2s and δtev = 57 s. Considering that the start of phase 3 is at 10 s, where Tpv = 42 °C, then the end of phase 3 is estimated at 10 s + 57 s = 67 s, compared to around 60 s in the experimentally identified phase 3 in Figure 3.

4.3. Steady-State Tpv Prediction by the Proposed Model Taking into Account RH, Ta, Ts.w.

Equations (1)–(9) predict Tpv provided the difference in the heat convection coefficient for RH = 55% with reference to RH = 45% being estimated (Equation (7b)). The air flow in the experimental conditions was laminar forced flow. Substituting into Equation (5), the να, νs.w., Pra, Prs.w., ka, and ks.w. values corresponding to Ta and Ts.w. give hc,s.w. = 195hc,a. The Mollier diagrams for the SW80 on the seashore (RH = 55% and Ta = 20 °C) and inland site (RH = 45% and Ta = 21 °C) give the % concentration in g-mol of both H2O and dry air. Specifically,
  • The relative concentration is 0.98565% g-mol dry air and 0.014345% g-mol H2O in the humid air at RHs.e. = 55%.
  • The relative concentration is, correspondingly, 0.9885% g-mol dry air and 0.01147% g-mol H2O in the humid air at RHinl = 45%.
At saturated air, Ta = Ts.w. and hus = 0.014659 g H2O/g dry air or 17.3 g H2O/m3 air.
The heat convection coefficient hc,a is estimated to be 20 W/m2K, and Equation (7a) for laminar forced flow gives hc,55% = 75.6 W/m2K and hc,45% = 64.6 W/m2K.
Therefore, δhc,hu = 11.0 W/m2K and Upv,s.e. = Upv,inl + δhc,hu = 34 W/m2K. The hu correction term in Equation (9) (1−δhc,hu/Upv,s.e.) = 0.676. Equation (8), for finl = 0.031, gives fs.e.= 0.031·0.676 = 0.021 and Tpv = Ta + fs.e.·IT = 20 °C + 0.021 m2K/W·800 W/m2 = 36.8 °C, which is 0.5% lower than the measured value of 37 °C, in Figure 2.
Considering Equation (1), the error in the estimation of Tpv is the error in the measurement of Ta = ±0.5 °C (see Section 2), plus the error in the measurement in IT which is negligible (see Section 2) times f, plus IT times the error in the estimation of f. The latter was estimated in the third decimal digit. Hence, the error in the estimation of f·IT is ±(0.001 m2K/W)·800 W/m2 = ±0.8 °C. Therefore, the total error in the prediction of Tpv is ±1.3 °C while the accuracy in the Tpv measurement was ±1 °C. Similarly, in the prediction of Tpv using Equations (1)–(9), the analysis gives an additional contribution to the error due to the estimation of δhc,hu/Upv, which is equal to 3%, and that corresponds to an additional error in Tpv of ±0.7 °C. Therefore, the overall error in the estimation in Tpv accounting for the hu sums up to ±2 °C, which is an acceptable range in the estimation of Tpv.

4.4. Comparison with Other Tpv Prediction Models

Considering the experimental conditions on the seashore site, Ta = 20 °C, IT = 800 W/m2, vw = 1 m/s, and RH = 55%, the measured Tpv =37 °C at the steady-state is compared to the predicted Tpv by the model proposed in this study, Equations (1)–(9), and to other existing models for FPV systems, Equations (32)–(38), in Table 1.
In Equation (37), (τα) = 0.9 is the effective transmission–absorption coefficient of the PV module with ηref = 0.146 at STC. The values used for the heat losses coefficients Uf and Ub are 18.355 and 10.209 W/m2K, respectively, as proposed in [48].
This comparative analysis for the Tpv prediction by the proposed model and the seven other formulas in Table 1 confirms that the approach outlined in this paper, taking into account the RH and the Ts.w., gives results closer to those measured in the s.e. Apart from Equations (35) and (36), [16], which gave very good prediction results, the other Tpv prediction formulas overestimated Tpv.

5. Discussion

The recuperation in ηpv due to enhanced natural cooling is estimated at 4% by introducing in Equation (30) two steady-state Tpv values: 37 °C (Figure 5) on the seashore and 45 °C (Tpv,inl-2, Figure 3) inland. This recuperation in ηpv is attributed to the higher humidity in the s.e. with the module operating at SOC. Comparing the average Tpv(t) = 22 °C during phase 3 on the seashore with the steady-state Tpv = 45 °C inland leads to a 11.5% recuperation due to the combined effect of the humidity as well as the seawater cooling and evaporation on the modules. Similar efficiency gains are anticipated for FPV operation in freshwater environments due to the higher humidity present in the vicinity of lakes and reservoirs and the water evaporation on the modules.
While low wind speed conditions were present during the experiments on the seashore and inland, it is expected that higher wind speeds, which usually prevail in the s.e., would lead to a much higher efficiency recuperation for FPV. A higher vw = 4 m/s in Equation (3) leads to an overall fs.e. = 0.0136 m2K/W accounting for the effect of the higher wind speed and humidity and, therefore, Tpv= Ta + fs.e.·IT = 20 °C + 0.0136 m2K/W·800 W/m2 = 30.9 °C. This additional decrease in Tpv by 6.1 °C, accounting for the effect of wind, with vw = 4 m/s, would lead to a total recuperation of 7% at steady-state and 14.5% when s.w. splashing and evaporation on the modules are also considered.
While the aforementioned efficiency gains are significant, some losses may be encountered due to a thin layer of salt that remains after the s.w. layer evaporation on the PV module, which may lead to a reduction in the solar irradiance reaching the PV cells [60]. Long-term exposure to humid and saline environments can lead to corrosion, potential-induced degradation (PID), and other PV degradation effects reducing the module and system lifetime [19,21,60,61]. Biofouling effects, including algae growth on PV modules, may pose additional challenges [62,63]. The application of nanocoatings with self-cleaning and anti-fouling properties on PV glass [64], the use of anticorrosive PV material [63], and improved multilayer backsheets [61], as well as frequent cleaning, may mitigate some of these risks. Considering energy yield gains between 5 and 10% in FPV, a 3–9% higher LCOE is estimated for FPV systems in freshwater environments compared to LBPV in [65]. While FPV platforms at the s.e. are at the early stages of development, it is anticipated that the estimated efficiency recuperation due to higher humidity and s.w. cooling and evaporation on the PV modules may partially counterbalance the higher CAPEX costs. A full investigation into all the aforementioned effects would be needed for long-term performance predictions and estimation of the levelized cost of electricity for FPV operating at the s.e., which is outside the scope of the present article.

6. Conclusions

The PV cooling of c-Si modules operating on the seashore and inland sites was theoretically and experimentally studied through steady-state and transient temperature profiles. The research covered seawater splashing on the modules and the subsequent seawater layer evaporation as well as the effect of humidity. An improved model for the prediction of steady-state Tpv incorporating the effect of humidity is proposed, and complete theoretical analysis of the PV cooling phenomena after seawater splashing and the subsequent evaporation of the seawater layer from the module is provided. The Tpv prediction results were compared to measured values both inland and on the seashore, simulating FPV conditions, and were shown to be in very good agreement.
The research analysis showed that the Tpv profiles of the front side depend on the pattern of the seawater flow on the PV and the environmental conditions, including the humidity. Specifically:
  • Tpv depends on the humidity and decreases as hu increases from low to medium values in a clear sky. For relative humidity 55% on the seashore compared to 45% inland, the steady-state Tpv was both predicted and measured about 18% lower on the seashore. This corresponds to a 4% higher efficiency on the seashore compared to inland, which is mainly attributed to the difference in humidity as vw, IT, and Ta were almost the same on the seashore and inland sites.
  • The transient Tpv(t) profile depends on the pattern of seawater splashing on the module. After seawater splashing, a steep temperature drop of 22% lasting for 2 s was measured and theoretically confirmed. The drop depends on the seawater temperature and the mode it splashes on the modules. This reached 52% when the pattern of the s.w. flow on the module was three shots of 25 mL/s per unit width of the module.
  • Tpv is affected by the subsequent seawater layer evaporation on the module which caused an overall decrease between 20 and 40% (depending on the flow pattern) compared to the steady-state value on the seashore before the seawater splash. This decrease lasted for 60–100 s, depending on the seawater flow rate and mode of splashing, which was theoretically predicted and experimentally confirmed.
  • The Tpv profiles on the seashore with seawater splashing on the modules were 35–51% lower compared to the steady-state inland values.
  • Taking into consideration the effect of humidity as well as the seawater cooling and evaporation on the modules, it was estimated that the PV efficiency on the seashore was 11.5% higher than inland.
This research disclosed the importance of the effect of humidity on PV temperature as well as the effect of the seawater splashing and its evaporation on the modules, whose combined effect leads to significant recuperation in the operating efficiency on the seashore compared to inland. While low wind speed conditions were present during the experiments, it is expected that higher wind speeds, which are usual in the sea environment, would lead to the recuperation of much higher efficiency for FPV.

Author Contributions

Conceptualization, S.K. and E.K.; methodology, S.K.; software, E.K.; validation, S.K. and E.K.; formal analysis, S.K., E.K. and J.K.K.; investigation, S.K. and J.K.K.; resources, S.K.; data curation, S.K.; writing—original draft preparation, S.K., E.K. and J.K.K.; writing—review and editing, S.K. and E.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

ANNArtificial neural network
FPVFloating PV
ITGlobal solar radiation intensity on the PV plane (W/m2)
IT,SOCGlobal solar radiation intensity at SOC conditions, 800 W/m2
IT,refReference solar irradiance equal to 103 W/m2
LLength of the PV module in the direction of the seawater flow on the front side (m)
NuNusselt number of the air flow either in the front or back side of the PV module
PmPeak power of a PV module (W)
Pra, Prs.wPrandtl number of air and water (dimensionless)
QFlow rate (mL/s)
RHRelative humidity
ReReynolds number
SOCStandard operating conditions (IT = 800 W/m2, Ta = 20 °C, vw = 1 m/s)
STCStandard test conditions (IT = 1000 W/m2, Tpv = 25 °C, air mass AM1.5)
Tpv, Tf,Steady-state PV module temperature and PV front side temperature, respectively, considered equal in this paper
TaAmbient temperature (°C or K as specified)
Tpv(t)PV module temperature at transient conditions at time t
Ts.w. Seawater temperature (°C)
TwFreshwater temperature (°C)
Ub−a, Uc−s.w.Heat losses coefficients due to convection and IR radiation at the back side of the PV module (W/m2K), equal to hc,b + hr,b
Uev An empirical evaporation coefficient (kg/m2h)
Uf−a, Us.w.−aHeat losses coefficients due to convection and IR radiation at the front side of the PV module (W/m2K), equal to hc,f + hr,f
UpvThe overall heat losses coefficient in a PV (W/m2K), equal to Uf + Ub
bThe width of the string of PV cells in a module on which the water flows (m)
hc,aHeat convection coefficient with dry air as coolant (W/m2K)
hc,bHeat convection coefficient from PV back surface to air (W/m2K)
hc,fHeat convection coefficient from PV glass to air (W/m2K)
hc,s.w.Heat convection coefficient with s.w. as coolant (W/m2K)
hcrThe critical thickness of the water layer on the module (m)
hevEvaporation heat (J/g)
hr,bRadiative heat coefficient from the PV back side to environment (W/m2K)
hr,fRadiative heat coefficient from the front PV side (W/m2K)
huHumidity (kg H2O/kg dry air)
husHumidity ratio at saturation
kiThermal conductivity of material i (W/mK)
mevRate of mass evaporation (g/s)
(mc)ef, (mc)i Effective heat capacity of the PV cell or module and the heat capacity of a material i
phu,
qhu
Moles of dry air in the environment (%)
Moles of H2O in the environment (%)
qThe heat rate required for the evaporation (W)
s.w.Seawater
u(y), uavSeawater layer velocity at distance y off the module in an axis normal to its surface and the average speed, respectively
vwWind velocity (m/s)
ΔxSeawater layer thickness on a PV module (m)
βPV module inclination angle with reference to horizontal
δtevThe time the seawater layer evaporates
ηpvPV module efficiency
νfKinematic viscosity of the fluid (air, water) at temperature Tf (m2/s)
σSurface tension (N/m)
τ, τs.w., τgTemperature profile time constants. For the module, the seawater layer and the glass cover

References

  1. Vats, K.; Tiwari, G.N. Performance evaluation of a building integrated semitransparent photovoltaic thermal system for roof and façade. Energy Build. 2012, 45, 211–218. [Google Scholar] [CrossRef]
  2. Ghosh, A.; Sarmah, N.; Sundaram, S.; Mallick, T.K. Numerical studies of thermal comfort for semi-transparent building integrated photovoltaic (BIPV)-vacuum glazing system. Sol. Energy 2019, 190, 608–616. [Google Scholar] [CrossRef]
  3. Assoa, Y.B.; Gaillard, L.; Menezo, C.; Negri, N.; Sauzedde, F. Dynamic prediction of a building integrated photovoltaic system thermal behaviour. Appl. Energy 2018, 214, 73–82. [Google Scholar] [CrossRef]
  4. Kaplanis, S.; Kaplani, E. A New Dynamic Model to Predict Transient and Steady State PV Temperatures Taking into Account the Environmental Conditions. Energies 2019, 12, 2. [Google Scholar] [CrossRef] [Green Version]
  5. Kaldellis, J.; Kapsali, M.; Kavadias, K. Temperature and wind speed impact on the efficiency of PV installations. Experience obtained from outdoor measurements in Greece. Renew. Energy 2014, 66, 612–624. [Google Scholar] [CrossRef]
  6. Lobera, D.T.; Valkealahti, S. Dynamic thermal model of solar PV systems under varying climatic conditions. Sol. Energy 2013, 93, 183–194. [Google Scholar] [CrossRef]
  7. Luketa-Hanlin, A.; Stein, J.S. Improvement and Validation of a Transient Model to Predict Photovoltaic Module Temperature; Sandia National Laboratories: Albuquerque, NM, USA, 2012.
  8. Kaplani, E.; Kaplanis, S. Thermal modelling and experimental assessment of the dependence of PV module temperature on wind velocity and direction, module orientation and inclination. Sol. Energy 2014, 107, 443–460. [Google Scholar] [CrossRef] [Green Version]
  9. Kaplani, E.; Kaplanis, S. Dynamic Electro-Thermal PV Temperature and Power Output Prediction Model for Any PV Geometries in Free-Standing and BIPV Systems Operating under Any Environmental Conditions. Energies 2020, 13, 4743. [Google Scholar] [CrossRef]
  10. Kaplanis, S.; Kaplani, E.; Kaldellis, J.K. PV temperature and performance prediction in free-standing, BIPV and BAPV incorporating the effect of temperature and inclination on the heat transfer coefficients and the impact of wind, efficiency and ageing. Renew. Energy 2022, 181, 235–249. [Google Scholar] [CrossRef]
  11. Faiman, D. Assessing the outdoor operating temperature of photovoltaic modules. Prog. Photovolt. Res. Appl. 2008, 16, 307–315. [Google Scholar] [CrossRef]
  12. Ciulla, G.; Lo Brano, V.; Moreci, E. Forecasting the cell temperature of PV modules with an adaptive system. Int. J. Photoenergy 2013, 2013, 192854. [Google Scholar] [CrossRef]
  13. Schwingshackl, C.; Petitta, M.; Wagner, J.E.; Belluardo, G. Wind effect on PV module temperature: Analysis of different techniques for an accurate estimation. Energy Procedia 2013, 40, 77–86. [Google Scholar] [CrossRef] [Green Version]
  14. Skoplaki, E.; Boudouvis, A.G.; Palyvos, J.A. A simple correlation for the operating temperature of photovoltaic modules of arbitrary mounting. Sol. Energy Mater. Sol. Cells 2008, 92, 1393–1402. [Google Scholar] [CrossRef]
  15. TamizhMani, G.; Ji, L.; Tang, Y.; Petacci, L.; Osterwald, C. Photovoltaic module thermal/wind performance: Long term monitoring and model development for energy rating. In Proceedings of the NCPV and Solar Program Review Meeting Proceedings, Denver, CO, USA, 24–26 March 2003; pp. 936–939. [Google Scholar]
  16. Kamuyu, W.C.L.; Lim, J.R.; Won, C.S. Prediction model of Photovoltaic Module Temperature for Power Performance of Floating PVs. Energies 2018, 11, 447. [Google Scholar] [CrossRef] [Green Version]
  17. Graditi, G.; Ferlito, S.; Adinolfi, G.; Tina, G.M.; Ventura, C. Energy yield estimation of thin-film photovoltaic plants by using physical approach and artificial neural networks. Solar Energy 2016, 130, 232–243. [Google Scholar] [CrossRef]
  18. Dubey, S.; Sarvaiya, J.N.; Seshadri, B. Temperature dependent photovoltaic (PV) efficiency and its effect on PV production in the world—A review. Energy Procedia 2013, 33, 311–321. [Google Scholar] [CrossRef] [Green Version]
  19. Hacke, P.; Spataru, S.; Terwilliger, K.; Perrin, G.; Glick, S.; Kurtz, S.; Wohlgemuth, J. Accelerated testing and modeling of potential-induced degradation as a function of temperature and relative humidity. IEEE J. Photovolt. 2015, 5, 1549–1553. [Google Scholar] [CrossRef]
  20. Kempe, M.D.; Wohlgemuth, J.H. Evaluation of temperature and humidity on PV module component degradation. In Proceedings of the IEEE 39th Photovoltaic Specialists Conference (PVSC), Tampa, FL, USA, 16–21 June 2013; pp. 0120–0125. [Google Scholar] [CrossRef]
  21. Park, N.C.; Oh, W.W.; Kim, D.H. Effect of Temperature and Humidity on the Degradation Rate of Multicrystalline Silicon Photovoltaic Module. Int. J. Photoenergy 2013, 2013, 925280. [Google Scholar] [CrossRef] [Green Version]
  22. Kaplanis, S.; Kaplani, E.; Borza, P.N. PV defects identification through a synergistic set of non-destructive testing (NDT) techniques. Sensors 2023, 23, 3016. [Google Scholar] [CrossRef]
  23. Cazzaniga, R.; Rosa-Clot, M. The booming of floating PV. Sol. Energy 2021, 219, 3–10. [Google Scholar] [CrossRef]
  24. Rosa-Clot, M.; Tina, G.M.; Nizetic, S. Floating photovoltaic plants and wastewater basins: An Australian project. Energy Procedia 2017, 134, 664–674. [Google Scholar] [CrossRef]
  25. Tina, G.M.; Scavo, F.B.; Merlo, L.; Bizzarri, F. Comparative analysis of monofacial and bifacial photovoltaic modules for floating power plants. Appl. Energy 2021, 281, 116084. [Google Scholar] [CrossRef]
  26. Sahu, A.; Yadav, N.; Sudhakar, K. Floating photovoltaic power plant: A review. Renew. Sustain. Energy Rev. 2016, 66, 815–824. [Google Scholar] [CrossRef]
  27. Wild, M.; Folini, D.; Hakuba, M.Z.; Schär, C.; Seneviratne, S.I.; Kato, S.; Rutan, D.; Ammann, C.; Wood, E.F.; König-Langlo, G. The energy balance over land and oceans: An assessment based on direct observations and CMIP5 climate models. Clim. Dyn. 2015, 44, 3393–3429. [Google Scholar] [CrossRef] [Green Version]
  28. Lindfors, A.V.; Hertsberg, A.; Riihela, A.; Carlund, T.; Trentmann, J.; Mueller, R. On the Land-Sea Contrast in the Surface Solar Radiation (SSR) in the Baltic Region. Remote Sens. 2020, 12, 3509. [Google Scholar] [CrossRef]
  29. Govardhanan, M.S.; Kumaraguruparan, G.; Kameswari, M.; Saravanan, R.; Vivar, M.; Srithar, K. Photovoltaic Module with Uniform Water Flow on Top Surface. Int. J. Photoenergy 2020, 2020, 8473253. [Google Scholar] [CrossRef]
  30. Kerzika, A.H.; Barimah, B.; Aurelien, K.K.C. Photovoltaic Solar Panel Cooled with Runoff Water. Int. J. Energy Eng. 2020, 10, 41–45. [Google Scholar]
  31. Sukarno, K.; Hamid, A.S.A.; Razali, H.; Dayou, J. Evaluation on cooling effect on solar PV power output using Laminar H2O surface method. Int. J. Renew. Energy Res. 2017, 7, 1213–1217. [Google Scholar]
  32. Iqbal, S.; Afzal, S.; Mazhar, A.U.; Anjum, H.; Diyyan, A. Effect of Water Cooling on the Energy Conversion Efficiency of PV Cell. Am. Sci. Res. J. Eng. Technol. Sci. 2016, 20, 122–128. [Google Scholar]
  33. Nižetić, S.; Čoko, D.; Yadav, A.; Grubišić-Čabo, F. Water spray cooling technique applied on a photovoltaic panel: The performance response. Energy Convers. Manag. 2016, 108, 287–296. [Google Scholar] [CrossRef]
  34. Bahaidarah, H.M.S. Experimental performance evaluation and modelling of jet impingement cooling for thermal management of photovoltaics. Sol. Energy 2016, 135, 605–617. [Google Scholar] [CrossRef]
  35. 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]
  36. Chandrasekhar, M.; Suresh, S.; Senthilkumar, T.; Karthikeyan, M.G. Passive cooling of standalone flat PV module with cotton wick structures. Energy Convers. Manag. 2013, 71, 43–50. [Google Scholar] [CrossRef]
  37. Alami, A.H. Effects of evaporative cooling on efficiency of photovoltaic modules. Energy Convers. Manag. 2014, 77, 668–679. [Google Scholar] [CrossRef]
  38. Kumar, M.; Niyaz, H.M.; Gupta, R. Challenges and opportunities towards the development of floating photovoltaic systems. Sol. Energy Mater. Sol. Cells 2021, 233, 111408. [Google Scholar] [CrossRef]
  39. Trapani, K.; Millar, D.L. The thin film flexible floating PV (T3F-PV) array: The concept and development of the prototype. Renew. Energy 2014, 71, 43–50. [Google Scholar] [CrossRef]
  40. Liu, L.; Wang, Q.; Lin, H.; Li, H.; Sun, Q.; Wennersten, R. Power Generation Efficiency and Prospects of Floating Photovoltaic Systems. Energy Procedia 2017, 105, 1136–1142. [Google Scholar] [CrossRef]
  41. Dorenkamper, M.; Wahed, A.; Kumar, A.; de Jong, M.; Kroon, J.; Reindl, T. The cooling effect of floating PV in two different climate zones: A comparison of field test data from the Netherlands and Singapore. Sol. Energy 2021, 219, 15–23. [Google Scholar] [CrossRef]
  42. Bonkaney, A.L.; Madougou, S.; Adamou, R. Impact of Climatic Parameters on the Performance of Solar Photovoltaic (PV) Module in Niamey. Smart Grid Renew. Energy 2017, 8, 379–393. [Google Scholar] [CrossRef] [Green Version]
  43. Almaktar, M.; Rahman, H.A.; Hassan, M.Y.; Rahman, S. Climate Based Empirical Model for PV Module Temperature Estimation in Tropical Environment. Appl. Sol. Energy 2013, 49, 192–201. [Google Scholar] [CrossRef]
  44. Liu, H.; Krishna, V.; Leung, J.L.; Reindl, T.; Zhao, L. Field experience and performance analysis of floating PV technologies in the tropics. Progr. Photovolt. Res. Appl. 2018, 26, 957–967. [Google Scholar] [CrossRef]
  45. Luo, W.; Isukapalli, S.N.; Vinayagam, L.; Ting, S.A.; Pravettoni, M.; Reindl, T.; Kumar, A. Performance loss rates of floating photovoltaic installations in the tropics. Sol. Energy 2021, 219, 58–64. [Google Scholar] [CrossRef]
  46. Kjeldstad, T.; Lindholm, D.; Marstein, E.; Selj, J. Cooling of floating photovoltaics and the importance of water temperature. Sol. Energy 2021, 218, 544–551. [Google Scholar] [CrossRef]
  47. Lindholm, D.; Kjeldstad, T.; Selj, J.; Marstein, E.S.; Fjær, H.G. Heat loss coefficients computed for floating PV modules. Prog. Photovolt. Res. Appl. 2021, 29, 1262–1273. [Google Scholar] [CrossRef]
  48. Niyaz, H.M.; Kumar, M.; Gupta, R. Estimation of module temperature for water-base photovoltaic systems. J. Renew. Sustain. Energy 2021, 13, 053705. [Google Scholar] [CrossRef]
  49. Golroodbari, S.Z.; van Sark, W. Simulation of performance differences between offshore and land-based photovoltaic systems. Prog. Photovolt. Res. Appl. 2020, 28, 873–886. [Google Scholar] [CrossRef]
  50. Kazem, H.A.; Chaichan, M.T. Effect of humidity on Photovoltaic Performance based on experimental study. Int. J. Appl. Eng. Res. 2015, 10, 43572–43577. [Google Scholar]
  51. Chaicham, M.T.; Kazem, H.A. Experimental analysis of solar intensity on photovoltaic in hot and humid weather conditions. Int. J. Sci. Eng. Res. 2016, 7, 91–96. [Google Scholar]
  52. Choi, J.H.; Hyun, J.H.; Lee, W.; Bhang, B.-G.; Min, Y.K.; Ahn, H.-K. Power performance of high density photovoltaic module using energy balance model under high humidity environment. Sol. Energy 2021, 219, 50–57. [Google Scholar] [CrossRef]
  53. Dirnberger, D.; Blackburn, G.; Müller, B.; Reise, C. On the impact of solar spectral irradiance on the yield of different PV technologies. Sol. Energy Mater. Sol. Cells 2015, 132, 431–442. [Google Scholar] [CrossRef]
  54. Tashtoush, B.; Al-Oqool, A. Factorial analysis and experimental study of water-based cooling system effect on the performance of photovoltaic module. Int. J. Environ. Sci. Technol. 2019, 16, 3645–3656. [Google Scholar] [CrossRef]
  55. Lienhard, J.H., IV; Lienhard, J.H., V. A Heat Transfer Textbook, 3rd ed.; Phlogiston Press: Cambridge MA, USA, 2003. [Google Scholar]
  56. White, F.M. Heat and Mass Transfer; Addison-Wesley Publishing Co.: Boston, MA, USA, 1988. [Google Scholar]
  57. Ilha, A.; Doria, M.M.; Aibe, V.Y. Treatment of the Time Dependent Residual Layer and its Effects on the Calibration Procedures of Liquids and Gases Inside a Volume Prover. In Proceedings of the 15th Flow Measurement Conference (FLOMEKO), Taipei, Taiwan, 13–15 October 2010. [Google Scholar]
  58. Shah, M.M. Methods for calculation of evaporation from swimming pools and other water surfaces. ASHRAE Trans. 2014, 120 Pt 2, 3–17. [Google Scholar]
  59. Mattei, M.; Notton, G.; Cristofari, C.; Muselli, M.; Poggi, P. Calculation of the polycrystalline PV module temperature using a simple method of energy balance. Renew. Energy 2006, 31, 553–567. [Google Scholar] [CrossRef]
  60. Mannino, G.; Tina, G.M.; Cacciato, M.; Merlo, L.; Cucuzza, A.V.; Bizzarri, F.; Canino, A. Photovoltaic Module Degradation Forecast Models for Onshore and Offshore Floating Systems. Energies 2023, 16, 2117. [Google Scholar] [CrossRef]
  61. Zhang, J.-W.; Cao, D.-K.; Diaham, S.; Zhang, X.; Yin, X.-Q.; Wang, Q. Research on potential induced degradation (PID) of polymeric backsheet in PV modules after salt-mist exposure. Sol. Energy 2019, 188, 475–482. [Google Scholar] [CrossRef]
  62. Almeida, R.M.; Schmitt, R.; Grodsky, S.M.; Flecker, A.S.; Gomes, C.P.; Zhao, L.; Liu, H.; Barros, N.; Kelman, R.; McIntyre, P.B. Floating solar power could help fight climate change—Let’s get it right. Nature 2022, 606, 246–249. [Google Scholar] [CrossRef]
  63. Wang, J.; Lund, P.D. Review of recent offshore photovoltaics development. Energies 2022, 15, 7462. [Google Scholar] [CrossRef]
  64. Arabatzis, I.; Todorova, N.; Fasaki, I.; Tsesmeli, C.; Peppas, A.; Li, W.X.; Zhao, Z. Photocatalytic, self-cleaning, antireflective coating for photovoltaic panels: Characterization and monitoring in real conditions. Sol. Energy 2018, 159, 251–259. [Google Scholar] [CrossRef]
  65. World Bank Group; Energy Sector Management Assistance Program; The Solar Energy Research Institute of Singapore. Where Sun Meets Water: Floating Solar Handbook for Practitioners; World Bank: Washington, DC, USA, 2019. [Google Scholar]
Figure 1. Seawater layer thickness Δx vs. t when the flow stopped and the layer started thinning. The flow rate per unit width was 50 mL/s/m on a module inclined at 35° and Ts.w. = 30 °C.
Figure 1. Seawater layer thickness Δx vs. t when the flow stopped and the layer started thinning. The flow rate per unit width was 50 mL/s/m on a module inclined at 35° and Ts.w. = 30 °C.
Energies 16 04756 g001
Figure 2. Tpv(t) for the SW80 module on the seashore site. The steady-state Tpv values are shown at t < 0. The Tpv(t) profile begins at t = 0 at the end of the s.w. flow on the module with Q/b = 50 mL/s/m. The regions defined by the red dashed lines represent phase 1, phase 2, and phase 3.
Figure 2. Tpv(t) for the SW80 module on the seashore site. The steady-state Tpv values are shown at t < 0. The Tpv(t) profile begins at t = 0 at the end of the s.w. flow on the module with Q/b = 50 mL/s/m. The regions defined by the red dashed lines represent phase 1, phase 2, and phase 3.
Energies 16 04756 g002
Figure 3. Tpv(t) profile for the M55 (Tpv,inl-1) and the SW80 (Tpv,inl-2) modules measured on the site inland. The steady-state Tpv values are shown at t < 0. The Tpv(t) profile begins at t = 0 at the end of the s.w. flow on the module with Q/b = 50 mL/s/m.
Figure 3. Tpv(t) profile for the M55 (Tpv,inl-1) and the SW80 (Tpv,inl-2) modules measured on the site inland. The steady-state Tpv values are shown at t < 0. The Tpv(t) profile begins at t = 0 at the end of the s.w. flow on the module with Q/b = 50 mL/s/m.
Energies 16 04756 g003
Figure 4. Tpv(t) for the SW80 module on the seashore site. The steady-state Tpv values are shown at t < 0. The Tpv(t) profile begins at t = 0 at the end of the s.w. flow on the module with Q/b = 40 mL/s/m.
Figure 4. Tpv(t) for the SW80 module on the seashore site. The steady-state Tpv values are shown at t < 0. The Tpv(t) profile begins at t = 0 at the end of the s.w. flow on the module with Q/b = 40 mL/s/m.
Energies 16 04756 g004
Figure 5. Tpv(t) for the SW80 module on the seashore site. The steady-state Tpv values are shown at t < 0. The Tpv(t) profile begins at t = 0 at the end of the s.w. flow on the module with Q/b = 75 mL/s/m split in 3 shots of 25 mL/s/m each.
Figure 5. Tpv(t) for the SW80 module on the seashore site. The steady-state Tpv values are shown at t < 0. The Tpv(t) profile begins at t = 0 at the end of the s.w. flow on the module with Q/b = 75 mL/s/m split in 3 shots of 25 mL/s/m each.
Energies 16 04756 g005
Figure 6. Comparative presentation of Tpv(t) profiles as in Figure 2, Figure 4, and Figure 5 corresponding to Tpv,s-1, Tpv,s-2, and Tpv,s-3, respectively.
Figure 6. Comparative presentation of Tpv(t) profiles as in Figure 2, Figure 4, and Figure 5 corresponding to Tpv,s-1, Tpv,s-2, and Tpv,s-3, respectively.
Energies 16 04756 g006
Figure 7. Tpv(t) transient profile of the M55 module operating inland after s.w. splashing of Ts.w. = 14 °C.
Figure 7. Tpv(t) transient profile of the M55 module operating inland after s.w. splashing of Ts.w. = 14 °C.
Energies 16 04756 g007
Table 1. Predicted Tpv using the proposed model compared with other existing models for the s.e.
Table 1. Predicted Tpv using the proposed model compared with other existing models for the s.e.
Ref.ModelEquationTpv Predicted (°C)
Proposed model: Equations (1)–(9) 36.8
[43]Tpv = 26.97 + 0.77Ta + 0.023IT − 0.206RH −0.137vw(32)49.3
[15]Tpv = 0.961Ta + 0.029IT − 1.457vw + 0.000(°C/degree direction) + 0.109RH + 1.57 °C(33)48.5
[15]Tpv = 0.942Ta + 0.028IT − 1.509vw + 3.9 °C(34)43.6
[16]Tpv = 0.9458Ta + 0.0215IT − 1.2376vw + 2.0458 (35)36.9
[16]Tpv = 0.9282Ta + 0.021IT − 1.221vw + 0.0246Tw + 1.8081(36)36.3
[48]Tpv = [TaUf + TwUb + ((τα) −ηref − γηrefTref)IT]/(Uf + Ub − γηrefIT) (37)39.1
[14]Tpv = Ta + 0.32 IT/(8.91 + 2.0vw)(38)43.5
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kaplanis, S.; Kaplani, E.; Kaldellis, J.K. PV Temperature Prediction Incorporating the Effect of Humidity and Cooling Due to Seawater Flow and Evaporation on Modules Simulating Floating PV Conditions. Energies 2023, 16, 4756. https://doi.org/10.3390/en16124756

AMA Style

Kaplanis S, Kaplani E, Kaldellis JK. PV Temperature Prediction Incorporating the Effect of Humidity and Cooling Due to Seawater Flow and Evaporation on Modules Simulating Floating PV Conditions. Energies. 2023; 16(12):4756. https://doi.org/10.3390/en16124756

Chicago/Turabian Style

Kaplanis, Socrates, Eleni Kaplani, and John K. Kaldellis. 2023. "PV Temperature Prediction Incorporating the Effect of Humidity and Cooling Due to Seawater Flow and Evaporation on Modules Simulating Floating PV Conditions" Energies 16, no. 12: 4756. https://doi.org/10.3390/en16124756

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