Applied Thermal Engineering 168 (2020) 114848
Contents lists available at ScienceDirect
Applied Thermal Engineering journal homepage: www.elsevier.com/locate/apthermeng
Performance enhancement of pyramid solar distiller using nanofluid integrated with v-corrugated absorber and wick: An experimental study
T
⁎
Swellam W. Sharshira,b, , M.R. Elkadeemc,d, An Menge a
State Key Laboratory of Coal Combustion, School of Energy and Power Engineering, Huazhong University of Science and Technology, Wuhan 430074, China Mechanical Engineering Department, Faculty of Engineering, Kafrelsheikh University, Kafrelsheikh, Egypt c School of Electrical and Electronic Engineering, Huazhong University of Science and Technology, Wuhan, China d Electrical Power and Machines Engineering Department, Faculty of Engineering, Tanta University, Tanta 31521, Egypt e College of Mechanical & Electrical Engineering, Shaanxi University of Science and Technology, Xian 710021, China b
H I GH L IG H T S
combination of v-corrugated basin with wick and CuO nanofluid (DPDW + CuO) was investigated. • AThenewthermal efficiency of the DPDW + CuO was 60.5% compared to only 34% for TPD. • The daily exergy of the DPDW + CuO was increased by 93% compared to the TPD. • The DPDW + CuOefficiency get the least price of one-liter of the freshwater (0.015$) with 28% less than the TPD. •
A R T I C LE I N FO
A B S T R A C T
Keywords: Solar desalination Pyramid distiller Wick materials v-corrugated Copper oxide nanofluid Cost analysis
This paper aims to improve the performance of the pyramid distiller (PD) through an experimental investigation of three proposed modifications applied to the traditional pyramid distiller (TPD). First modification was called developed pyramid distiller (DPD) using v-corrugated absorbers to increase the surface area of evaporation, while the second modification (named DPDW) was implemented by inclusion of wick materials with the vcorrugated absorbers to minimize the rate of feed water, thus increasing the productivity to an appropriate value. In the third modification, which was designated (DPDW + CuO), copper oxide nanofluid with wick and vcorrugated absorbers was added to increase the thermal conductivity and absorptivity as well as minimize the specific heat of base fluid. The performance of three adaptations was compared to TPD, to characterize the thermal performance enhancement. Moreover, a cost analysis study was conducted to evaluate the economic performance of the three proposed systems and TPD. The results show that the three modified systems have a good thermo-economic performance compared to TPD. The use of DPD, DPDW, and DPDW + CuO enhanced the total freshwater productivity by about 28.38%, 45%, and 72.95%, respectively, compared to the TPD. Further, the DPDW + CuO has the best improvement in terms of daily energy efficiency and exergy by 77.9% and 93%, respectively, compared to TPD. Finally, the cost analysis results reveal that the DPDW + CuO gives the lowest cost among all modifications and TPD.
1. Introduction Safe, freshwater is essential for urban expansion, human civilization, and industrial development [1]. Billions of people are suffering from freshwater scarcity; however, seawater covers approximately 75% of the area of our planet [2]. Given both of the environmental pollution and water scarcity, it is desirable to develop eco-friendly technologies for seawater desalination [1,3]. As a result of the exacerbation of the
water shortage crisis and its many causes over time, water drops will be very expensive in the future. So, there are many attempts to look for technological solutions to this problem [4]. One of these technologies is desalinated water using sunlight, thus it has the benefits of using the cheap solar energy and minimizing the environmental impacts [5]. On a small scale, this technologies have been heavily investigated, but large-scale deployment is limited by efficiency and cost [6]. Different types of solar desalination technologies such as
⁎
Corresponding author at: Mechanical Engineering Department, Faculty of Engineering, Kafrelsheikh University, Kafrelsheikh, Egypt. E-mail addresses:
[email protected],
[email protected] (S.W. Sharshir),
[email protected] (M.R. Elkadeem),
[email protected] (A. Meng). https://doi.org/10.1016/j.applthermaleng.2019.114848 Received 14 May 2019; Received in revised form 22 November 2019; Accepted 24 December 2019 Available online 28 December 2019 1359-4311/ © 2019 Published by Elsevier Ltd.
Applied Thermal Engineering 168 (2020) 114848
S.W. Sharshir, et al.
Nomenclatures Al2O3 CuO C T TiO2 hfg ρsw I (t ) mew
MPSS n OCP S SFF TPD TPPL wt
aluminum oxide nanoparticles copper oxide nanoparticles specific heat, J/kg k temperature, °C titanium nanoparticles evaporation phase change, kJ/kg density, kg/m3 insolation, W/m2 hourly freshwater, kg/m2 h
Greek letters k E x output E xsun ηd ρ ϕ ηEX i
Abbreviations PV AFP AMOP ASV ATP AVS CB CP CPSS CRF CSS DPD DPDW DPDW + FP
modified pyramid solar still system life overall constant salvage value sinking fund factor traditional pyramid distiller Total price per liter of freshwater weight concentration, %
pyramid distiller annual fixed price annual maintenance and operating price annual salvage value annual total price annual salvage value carbon black constant price conventional pyramid solar still capital recovery factor conventional solar still developed pyramid distiller developed pyramid distiller with wick CuO developed pyramid distiller with wick and copper oxide nanofluid fixed price
thermal conductivity, W/m K output exergy input exergy of daily efficiency density, kg/m3 particle weight fraction, % exergy efficiency interest per year
Subscripts bf g nf P s sw w
base fluid glass nanofluid solid nanoparticle sun saline water water
best methods that improve the freshwater productivity using solar energy [31], which is the upper glazier cover in the form of a pyramid. There are three types of hierarchy; the first one has a rectangular shape, and the second takes the square shape, while the third one takes the form of the triangle. The main advantages of the pyramid distiller over the conventional one as follows:
humidification–dehumidification [7,8], solar stills distillation [6,9], solar chimneys [10], membrane processes [11] and so on have been investigated. Among all the technologies applied for desalination, solar stills, which are usually small-scale solar desalination systems, provide a solution for the shortage of freshwater in remote and arid areas, due to low-cost, simple-to-operate, self-reliant water supply and eco-friendly systems [9,12]. Although the all merits of these systems, it suffers from low efficiency and freshwater productivity [12,13]. As a result, numerous works have been investigated in the literature to improve solar distiller performance. In this sense, to maintain extra energy during bad weather situations (i.e., cloudy days and nighttime periods, different types of wick and energy storing materials were used in the basin of distiller [14,15], corrugated and finned basin in the conventional distiller [16]. Also, to enhance the solar distiller performance, a preheating process of saline water using external collector [17], solar pond [18], and internal reflector with different inclination angles [19], gravel black rubber [20], horizontal tubular solar still [21–23] vertical tubular solar still [24], sponge with and without metallic wiry and black rocks [25] phase change materials [26], using stone particle with different size as a porous absorber [27], multi-effect solar still [28], single and double effect [29], solar still integrated with solar collector and phase change materials [30] and so on. Along with the same line, increasing the temperature difference between the glass and water in the basin can augment the thermal performance of solar distiller system by increasing the area subject to the solar radiation using different methods such as evacuated tubes [31], flat collector [32], cylindrical parabolic concentrator [33], solar dish concentrator [34], and hybrid PVT collectors [35]. Unfortunately, these methods can increase the productivity, but they need high costs and system efficiency is relatively low due to the large surface area exposed to solar radiation. Recently, the pyramid distiller has been recognized as one of the
• There is no need to specify the direction of solar radiation, where all • •
upper parts are exposed to solar radiation at anytime and anywhere. In contrast, the traditional type must be determining the trends of solar radiation to maximize the solar radiation, which will enter into the distiller [36]. When pyramid distiller is used, the shading results for walls are much lower than the traditional type [37]. Considering the same area of the basin, the vapor condensation in the pyramid distiller is larger than the traditional one due to the larger condensation area [38].
Kabeel et al. [39] examined the modifications in the pyramid distiller integrated with corrugated basin as well as combined with phase change materials. The results illustrated that the freshwater improved by about 87.4% compared to the traditional one. Recently, to enhance the solar energy conversion, nanofluids have been used [40]. In contrast to the bulk water, nanofluids have enormous thermal-physical properties, which include large absorptivity, low specific heat with large surface area [41], high thermal conductivity [42], and accordingly great heat mass transfer capabilities [43]. To increase the water evaporation and water production different types of nanomaterials were utilized, such as dispersing metallic nanoparticles [44], nanomicro particles [45,46], carbon nanomaterials [47], plasmonic nanofluid [48], thin film heat [13] heat localization [49], and plasmonic metals [50]. Also, Kabeel et al. [51] performed experiments with a new basin of the triangle pyramid distiller coated with TiO2 nanoparticles 2
Applied Thermal Engineering 168 (2020) 114848
S.W. Sharshir, et al.
maxed with black paint with different water depth the results illustrated that the freshwater improved by about 6% compared to the traditional one at the minimum water depth. Sharshir et al. [52] conducted a theoretical and experimental comparative study by using two types of micro-and nanoparticles graphite or copper oxide to improve the performance of solar distiller. Results illustrated that the still performance with micro-nano particles was much better than the traditional one. Furthermore, the use of graphite gives higher performance than CuO. Sahota and Tiwari [53] used three different nanofluids (CuO, TiO2, and Al2O3) with various concentrations to investigate the performance of passive DSSS and identify the best concentration of nanofluid. The results reveal that the nanofluids achieved the best energy and exergy efficiencies compared with the pure water. From the above literature review, there are different types of designs such as a conventional, pyramid, inclined wick, stepped, double slope and so on. All these systems have different productivities and thermal efficiencies. Furthermore, some designs need external parts such as reflectors, condensers, collectors and so on to improve the productivity and efficiency. But unfortunately, their investment cost will be high. Also, there are some systems that have high productivity, but the thermal efficiency is relatively low due to increasing in the area subjected to solar radiation and the overall system cost becomes more expensive. Also, there still lack of works that study the impact of using nanofluid combined wick and v- corrugated basin on enhancing the thermal performance of the pyramid distiller, and it will be the main concern of this paper. In this study, the performance of the traditional pyramid distiller was enhanced by developing three modifications on the system structure. Each modification was experimentally tested and investigated. The first modification was implemented by using of v- corrugated basin to increase the evaporation surface area subjected to solar radiation. The second modification used wick materials in integrated with vcorrugated basin to reduce the rate of feed water, hence increase the productivity and overcome the water depth problem. Lately, in the third developed modification, CuO nanofluid was added to improve the thermo-physical properties of the base fluid (i.e. water), which is expected to increase the heat transfer as well as improve the rate of evaporation which will enhance the freshwater production and the thermal performance. Besides evaluating the thermal performance of the all modified system, the price per liter of the freshwater was calculated for the different modifications and compared with the traditional pyramid distiller.
5:00 pm in thirty different days of August and September 2017, where the experimental investigation of each developed modification was carried out through ten days. Two similar pyramid solar distillers were used to compare the performance of desalination systems. A detailed drawing of solar desalination setup is shown in Fig. 1 and depicted in Fig. 2. The setup consists of traditional pyramid distiller (TPD), and developed pyramid distiller (DPD). At first, the TPD basin area was 0.56 m2 (0.75 m × 0.75 m), which is a square area. The pyramid distillers were made of 1.5 mm galvanized sheets. To increase the solar absorption of the PD and thereby improve the rate of evaporation as well as productivity, black paint was used to paint the inner sides of all walls and basin. Furthermore, to minimize loss of heat as much as potential, all outside walls and bottom of the basin were perfect isolated using fiberglass with a thickness of 0.05 m and 0.02 wood. Additionally, the trough was installed in the underside of the glazier cover inside the aquarium and used to collect the distillate water produced in the external calibration flask through a rubber hose integrated with the flow control valve. The saline brine was discharged outside the distiller passing through another outlet. The basin was covered by upper glazier cover in the form of a pyramid (four sides) with 3 mm thickness inclined at 30° on horizontal this was the best angle according to [54]. This tilt angle was selected to maximize the insolation received by the absorber and minimize reflection losses. Regarding the DPD, it has the same dimensions as the TPD. But also has an additional v-corrugated square basin 0.74 × 0.74 m with a height of 0.05 m. The square basin was manufactured in the form of vcorrugated basin made of steel. This square v-corrugated basin put in the large main square area of the pyramid distiller, and it represents the first modification in this work. In the second modification, the v-corrugated basin was covered by a double layer of black wick materials to increase the overall heat transfer coefficient. Finally, CuO nanofluid with 1% wt concentrations was added to enhance the absorbance of solar radiations and thermal conductivity as well as decreasing the specific heat and this is considered as the third modifications. A constant amount of water was maintained at 5.6 kg in the TPD and DPD to make a fair comparison at different modifications, which is equivalent to a height of about 1 cm. Moreover, to keep the same amount of the brackish water during the experimental operation, the hourly amount of the freshwater was fed into the distiller again.
2. Experimental setup and procedure
It’s crucial to record the conditions of the experiment for a better understanding of the results and to evaluate the enhancement of them. Different parameters were observed using multiple tools and devices that were used in the perpetration of the proposed model. First, the most important measurement parameter was the temperatures. Type K
2.1. Measurements devices and error analysis
The experimental setup was designed and manufactured in Burullus City, Kafrelsheikh government, Egypt (Latitude 31° E 56°N and longitude 30° E 93° N). The experiments were carried out from 8:00 am to
Fig. 1. Layout of the experimental setup of the traditional and developed PDs. 3
Applied Thermal Engineering 168 (2020) 114848
S.W. Sharshir, et al.
Fig. 2. Experimental setup.
saline water as a base fluid.
thermocouples with a range of 0–120 °C, accuracy of ± 0.1 °C and 1.5% error were used to measure temperature in multiple points in the system such as basin water temperature, vapor temperature, glass inner temperatures glass outer temperature, air temperature). Second, solar radiation parameter was measured in real time for keeping track of the solar radiation levels during the day head, where solar radiation meter (called, Solar meter) with a maximum range of 0–5000 W/m2, accuracy ± 10 W/m2 and error of 1.5% error was used for this purpose. Third, the wind speed was measured using an anemometer with an accuracy of ± 0.1 m/s, and an allowable value of error equals 2.7%, and scale ranged from 0.230 m/s to 30 m/s. The same device can also be used to show the level of relative humidity. Finally, the freshwater productivity was observed via freshwater flask (range: 0–1500 mL, accuracy: ± 5 mL and error: 1.2%). It should mention that the errors in the total freshwater distillates and daily energy efficiency were ± 0.48% and 0.76%, respectively. In the current experimental study, an uncertainty analysis was performed for all the experimental results, in which the uncertainty of the experimental measuring devices ( X function) such as thermocouples, solar meter, and measuring beaker are computed of according to the following relation: 2
wx =
⎟
⎜
⎟
The saline water is consisting of freshwater and sea salts. Most physical properties of saline water can be described by functions of salinity (which is defined as the ratio of dissolved salts mass to water mass), temperature and pressure. Naturally, the salinity of seawater differs from one place to another; it may reach about 50 g/kg in the Arabian Gulf, 70 g/kg in Australian Shark Bay and may reach saturation concentration in the Dead Sea [55,56]. Therefore, the salinity of the saline water should be considered during studying the performance of the solar distiller. The physical and thermal characteristics of saline water such as density, specific heat, and thermal conductivity, have been measured by many researchers. The measured data are available over a wide range of salinity and temperature. The correlations for these measurements are given as follows [57]:
• Saline water density (ρ
sw)
⎜
is given by [58]:
ρsw
2
2
⎛ ∂X ⎞ wx12 + ⎛ ∂X ⎞ wx 22 + ............+⎛ ∂X ⎞ wx n2 ⎝ ∂x1 ⎠ ⎝ ∂x2 ⎠ ⎝ ∂x n ⎠ ⎜
3.1. Thermophysical properties of saline water
⎟
= (a1 + a2 Tw + a3 Tw2 + a4 Tw3 + a5 Tw4 )(b1 S + b2 STw + b3 STw2 + b4 STw3
(1)
+ b5 S 2Tw2 )
where wx is the propagation of uncertainty for X value, w is the uncertainty of the measured parameter and x n is the parameter of interest.
(2) 3
where ρsw in kg/m ; 0≺Tw ≺180 °C ;
• Saline water specific heat (C
sw)
3. Thermophysical properties
is given by [59]:
Csw = (A + B (Tw + 273) + C (Tw + 273)2 + D (Tw + 273)3) × 103 The knowledge of the saline water and nanofluid properties is essential for studying the performance of the solar stills. The heat transfer properties of the fluid depend on its thermophysical properties, which enhanced significantly by adding nanoparticles [52]. In this study, the copper-oxide was used as a type of nanoparticles, which dispersed in
where Csw in J/kg.k; 0≺Tw ≺180
°C ;
0≺S≺180 g/kg
• Saline water thermal conductivity (K 4
sw)
is given by [60]:
(3)
Applied Thermal Engineering 168 (2020) 114848
S.W. Sharshir, et al.
scale. Finally, the lower absorption value (~10%) was occurred in the case of the bulk water, particularly at a low wavelength from 400 to 750 nm. It was noted that the bulk water absorptivity was increased reaching 44% at 975 nm and then decreased again to 24% at the end. Fig. 3-(c) show the thermal conductivity of the bulk water and 1% wt of CuO nanofluid based on the mathematical correlations we have mentioned above. Notably, the higher thermal conductivity was accounted for the nanofluid and also increase with the increase of temperature, and this completely in agreement with [63–65]. Also, Fig. 3-(d) illustrates the specific heat of the base fluid (water) and mixture of nanofluid (water + 1% wt of CuO). Fundamentally, based on Eq. (5) the density of nanofluid is the summation of water density (1000 kg m−3) and the density of CuO nanoparticles (6500 kg m−3) [62]. Consequently, the nanofluid density is greater than that of water. So, according to the physical principle of mixture rule given in Eq. (6), the specific heat of CuO nanoparticles measured at room temperature (0.531 kJ kg−1 K−1) is lesser than the specific heat of water (4.18 kJ kg−1 K−1). Moreover, the base fluid specific heat is marginally higher that of mixture (i.e., CuO nanofluid) at various temperatures because of the sizable heat capacity of base fluid compared to CuO nanoparticles. Also, it is clearly seen that the base fluid specific heat and mixture marginally rise with the temperature increase. So, the achieved results are totally in agreement with [41,66].
log10 (Ksw ) = 10−3 × log10 (240 + 0.0002S ) + 0.000434 ⎛2.3 − ⎝ 0.333 + T 27 w ⎞ × ⎛1 − 647 + 0.03S ⎠ ⎝ ⎜
343.5 + 0.037S ⎞ Tw + 273 ⎠ ⎟
(4)
where Ksw in W/m.°C; 0≺Tw ≺180 °C ; 0≺S≺160 g/kg The variables of the above equations are defined in the Appendix. 3.2. Nanofluid thermophysical properties Nanofluids have abundant specific features compared to the pure water, for example, large insolation absorptivity, large thermal conductivity, large surface area, and low specific heat as well, which improve the freshwater production. The nanofluids thermophysical properties can be defined as functions of the nanoparticles and water. The main thermophysical properties of nanofluids are given by the equations as follows:
• Density of nanofluids (ρ
nf)
is given by [61]
φ φ ⎞ ρp + ⎛1 − ⎞ρ ρnf = ⎛ 100 ⎠ sw ⎝ 100 ⎠ ⎝
• Specific heat of nanofluids (C Cnf =
(5) nf)
is given by [61] 4.2. Effect of using v-corrugated basin
(φ 100) ρp Cp + (1 − (φ 100)) ρsw Csw ρnf
• Thermal conductivity of nanofluids (K Knf = Ksw
The ambient temperature, the intensity of solar radiation, and the wind speed factors on August 13, 2017, for the DPD and TPD are illustrated in Fig. 4-(a). From the figure, we can observe that the temperature of ambient and the intensity of solar radiation reaches the largest value at noon (about 12:00 pm), then it reduced to the minimum value at the end of the day at 6:00 pm. However, the air speed has fluctuated between 0.5 and 3.6 m/s. Fig. 4-(b) show that the water basin temperature of the DPD was larger than that of TPD. Fig. 4-(b) show that the water basin temperature of the DPD was larger than that of TPD. This because of the DPD has a considerable amount of heat transfer obtained by the inclusion of v-corrugated shape in the basin, which will increase the effective absorption area. Consequently, this will increase the amount of the absorbed solar energy. Basically, this will improve the heat transfer coefficients (i.e., radiative, evaporative, and convective), which increase the rate of evaporation as well as the water production. The largest water basin temperatures of the DPD and TPD were 74 °C and 68.5 °C respectively at 1:00 pm as illustrated in Fig. 4-(b). Also, Fig. 4-(b) demonstrated the temperature of the south-direction of the glazier cover for the DPD and TPD. Obviously, the DPD has a large temperature of south glazier reached 49 °C, and for the traditional pyramid distiller, the substantial temperatures of glazier reached 47 °C for south direction. Furthermore, the temperature difference between the glazier for the DPD and traditional one varies in range between 0.5 and 3 °C, 0.5 to 3.5 °C for the south and east directions, respectively, while the temperature variation recorded 0.5–2 °C, and 0.5–2 °C, respectively for west and north, as shown in Fig. 4-(c). The significant difference in temperature between the brackish water and south glazier was about 25 and 21.5 for the DPD and TPD, respectively. Also, it is worth to mention that the difference in temperature between the
(6) nf)
is given by [62]
Kp + (n − 1) Ksw − (n − 1) φ (Ksw − Kp) Kp + (n − 1) Ksw + φ (Ksw − Kp)
(7)
where φ is the percentage weight concentration is expressed:
mp ⎞ φ = ⎜⎛ ⎟ × 100 m ⎝ p + msw ⎠
(8)
The base fluid properties (saline water) are calculated by the equations mentioned in the previous section, and properties of the nanoparticles are given in Table 1. 4. Experimental results 4.1. Characterization of copper oxide nanoparticles and nanofluid The nanofluid was prepared by adding CuO nanoparticles to the water and thoroughly mixed by hand before entering to the DPD over the wick layer on the v-corrugated absorbers surfaces at the beginning of the experimental operation. Then, the freshwater is collected out of the distillation process every hour. After that, a similar amount to the collected freshwater was fed into the distiller every hour to maintain a constant amount of water as well as a constant value of the nanomaterial’s concentration. The morphology characteristics of nanomaterials were quantified via transmission electron microscopy (TEM) and given in Fig. 3-(a). Obviously, the CuO nanoparticles were consistently dispersed in the water and have a spherical shape of average diameter equals 40 nm. The absorption spectra of CuO nanoparticles, water, and CuO nanofluid, which are given in see Fig. 3-(b), were calculated through UV–Vis at 400 to 1100 nm wavelength and showing the variation of the spectra with various wavelengths. With a small wavelength (400 to 550 nm), the maximum value was recorded at nearly 97%. Then, it rapidly declines at 950 nm to 72.5%, followed by a very slight variation at larger wavelength from 950 up to 1100 nm. Also, the CuO nanofluid has an excellent absorption features near to 99% at small wavelengths, then it declined to its minimum value of 95% at end of the
Table 1 Properties of copper oxide nanoparticles.
5
Item
Properties
Thermal Conductivity, W/m K Density, g/cm3 Particle size average, nm Particles shape
76 6.4 40 Spherical
Applied Thermal Engineering 168 (2020) 114848
S.W. Sharshir, et al.
Fig. 3. (a) TEM photo of the mixture nanofluids (water + CuO), (b) absorption of pure CuO nanoparticles, CuO nanofluid with 1% wt and pure water, (c) thermal conductivity and (d) specific heat.
the difference in water temperature between the DPDW and TPD increases in afternoon more than in the early hours during the morning, where the wick materials work as thermal storage materials with high heat transfer capability. Accordingly, this will increase the heat transfer coefficients, improve the rate of evaporation and increase the productivity as well as overcome the water depth problem, particularly during afternoon periods. The largest water basin temperatures of DPDW and TPD were 76 °C and 69 °C, respectively, at 1:00 pm as illustrated in Fig. 5-(b). Also, Fig. 5-(b) showed the temperature of the south-direction of the glazier for the DPDW and TPD. Notably, in the case of DPDW, the large temperature of south glazier was reached to 50 °C, while it reached 48 °C for the TPD. Additionally, the disparities in the temperature between the glazier for the DPDW and TPD change between 1 and 4 °C, 1 to 3 °C for the south and east, respectively. For the west and north directions, the temperature differences were the same and equaled 1–3 °C, as illustrated in Fig. 5-(c). As a result of increasing the basin temperature of DPDW than TPD caused by increasing the absorption area of the corrugated basin covered with wick materials, the large temperature differences between the water and south glazier measured for the DPDW and TPD were about 26 °C and 21 °C, respectively. Fig. 5-(d) illustrated the hourly and accumulated freshwater output from DPDW and TPD. It can observe that the large hourly freshwater value of 0.76 L m−2 h−1 obtained at 1:00 pm for the DPDW and 0.65 L m−2 h−1 for the TPD. Also, the large accumulated freshwater reached to 4.51 L m−2 h−1 for DPDW and 3.11 L m−2 h−1 in the case of TPSS. The results reveal that the hourly and accumulated freshwater for DPDW was higher than that of TPSS. The use of a v-corrugated basin enhanced the freshwater output of the DPDW by about 45% compared
brackish water and glazier for the DPD was higher than the case of TPD. This attributed to increasing the temperature of the brackish water of the DPD over the TPD resulted from increasing the absorption area using the v-corrugated basin. Fig. 4-(d) show the hourly and accumulated freshwater output from DPD and TPD. It is clearly observed that the large hourly freshwater value of 0.7 L m−2 h−1 obtained at 1:00 pm for the DPD, while the larger value was recorded 0.6 L m−2 h−1 for the TPD at the same time. Also, the large accumulated freshwater reached 3.89 L m−2 h−1 for DPD and 3.03 L m−2 h−1 for TPD. The results show that the hourly and accumulated freshwater for DPD was more significant than that of TPD. Accordingly, the uses of the v-corrugated basin as a first modification in the DPD enhanced the freshwater output by about 28.38% compared to the TPD due to the larger heat transfer from the v-corrugated basin to brackish water in the basin obtained through increasing the heat transfer area as mentioned above. 4.3. Effect of using v-corrugated basin combined with wick material The meteorological factors on August 26, 2017, for the DPDW materials (named DPDW) and TPD are illustrated in Fig. 5-(a). From the figure, it is observed that the temperature of ambient and the intensity of solar radiation reached the largest value at 12:00 pm, then reduced to the minimum values at the end of the day at 18:00. However, the airspeed has fluctuated between 0.4 and 2.1 m/s. Fig. 5-(b) illustrates that the water basin temperature of the DPDW was greater than that of TPD. Where, the wick materials added with v- corrugated basin were used to increase efficiency of basin water utilization, in which the capillary wick effect will help to overcome the dry spots problem. Besides, 6
Applied Thermal Engineering 168 (2020) 114848
S.W. Sharshir, et al.
Fig. 4. Thermal performance of the DPD and TPD (a) ambient temperature, the intensity of solar radiation, and the wind speed, (b) water and glass temperatures, (c) glass temperature difference between DPD and TPD for four sides and (d) hourly and accumulated freshwater.
storage materials, thus increase the evaporation rate and productivity from noon until sunset. The largest water basin temperature of DPDW + CuO and TPD were 79.5 °C and 68 °C respectively at 1:00 pm as presented in Fig. 6-(b). Also, the same figure give the temperature of the south-direction of the glazier for the DPDW + CuO and TPD. The figure proves that the DPDW + CuO has a large temperature of the south glazier, which was 51 °C, while the large temperature of glazier measured with TPD was 48 °C. Moreover, the variations in temperature between the glazier for the DPDW + CuO and TPD change between 1.5 and 6 °C and 1.5 to 5.5 °C for the south and east directions, respectively, as illustrated in Fig. 6(c). On the other side, the differences in temperature were 1–4.5 °C and 0.5–5.5 °C when measured for the west, and north directions, respectively. The maximum temperature difference between the water and south glazier was about 25 °C and 21.5 °C for the DPDW + CuO and TPD, respectively. The obtained results indicate that the temperature difference between the glazier and basin water for the DPDW + CuO was higher than that of TPD pyramid distiller. This mainly because the temperature of basin water for DPDW + CuO was higher than that of TPD due to the combination and inclusion of the wick materials and CuO nanoparticles in the v-corrugated basin. As expected from the high thermal conductivity and large absorptivity as well as the low specific heat of nanofluid as investigated and explained in Section 4.1, the nanofluid is of advantages to improve the overall performance of DPDW + CuO system than TPD. Fig. 6-(d) illustrates the hourly and accumulated freshwater output from DPDW + CuO and TPD. It can notice that the large hourly freshwater value was 0.89 L m−2 h−1 obtained at 1:00 pm for the DPDW + CuO, and 0.6 L m−2 h−1 for the TPD. Also, the large accumulated freshwater was reached to 5.5 L m−2 h−1 for DPDW + CuO, while accounted for 3.18 L m−2 h−1 TPD. The results show that the values of the hourly and accumulated freshwater for the DPDW + CuO
to the TPD. This attributed to increasing the area of the basin water for DPDW in comparison with TPD, which was obtained due to the high intensity of solar radiation absorbed by the v-corrugated basin and the large heat transfer from the v-corrugated basin with wick to brackish water in the basin. 4.4. Effect of using v-corrugated basin combined with wick and copper oxide nanofluid The ambient temperatures, intensity of solar radiation, and wind speed on September 15, 2017, for the DPD with wick materials and copper oxide nanoparticles (named DPDW + CuO) and TPD are illustrated in Fig. 6-(a). From the figure, it is clear that the temperature of ambient and the intensity of solar radiation reach the top values at noon, then declined to their minimum values at the end of the day at 18:00. Also, the airspeed was varied between 1.5 and 2.1 m/s. It can be concluded from Fig. 6-(b) that the water basin temperature of the DPDW + CuO is larger than that of TPD. This mainly attributed to the effect of substantial amount of heat transfer taken from v-corrugated to the water in the basin in the case of DPDW + CuO. Where the proposed modifications will lead to increasing the area subjected to water evaporation and enhancing efficiency of basin water utilization by means of the capillary action and increase the heat transfer in the air-water interface. Additionally, the adding of CuO nanofluid result in improving the thermo-physical properties of the base fluid (i.e. water), which increase the thermal conductivity and absorption as well as decrease the specific heat which enhance the radiative, evaporative, and convective heat transfer coefficients [52]. So, this will improve the rate of evaporation and enhance the freshwater production and the thermal performance. Furthermore, the difference in water temperature between the DPDW + CuO and TPD increases at afternoon more than the beginning of the day. In this case, the wick materials act as thermal 7
Applied Thermal Engineering 168 (2020) 114848
S.W. Sharshir, et al.
Fig. 5. Thermal performance of the DPDW and TPD (a) ambient temperature, the intensity of solar radiation, and the wind speed, (b) water and glass temperatures, (c) glass temperature difference between DPDW and TPD for four sides, and (d) hourly and accumulated freshwater.
about 34%. Furthermore, the daily energy efficiency for DPDW and DPDW + CuO was significantly improved compared to TPD by 46.5% and 77.9%. Therefore, nanoparticles are recommended to be integrated with pyramid distiller because their capabilities to enhance the freshwater and the thermal energy performance of the distiller system.
were more significant than that of TPD. The use of a v-corrugated basin enhanced the freshwater output of the DPDW + CuO by 72.95% when compared to the TPD. The final results concluded that the increased in thermal conductivity and absorptivity and reduction in the specific heat of the fluid as explained in Section 4.1, help the fluid to heat up rapidly, subsequently increasing the evaporation rate and the production of freshwater.
4.5.2. Exergy efficiency for solar still The exergy of the solar distiller system can be defined as the maximum work attained from the system to reaches the state of thermodynamic balance under certain climate conditions and it can be obtained through the second law of thermodynamics [67]. The exergy efficiency (ηEX ) of solar distiller as a function of the system output exergy (E x output ) and input exergy (E xsun ) is expressed as follows:
4.5. Thermal efficiency 4.5.1. Thermal energy efficiency Thermal energy efficiency is one of the main crucial criteria that can be utilized to estimate the performance of the pyramid distiller systems. The overall efficiency (η) is obtained by Eq. (9):
%η = 100 ×
ηEX (%) =
∑ mew × hfg ∑ I (t ) × As × 3600
E xinput
(9)
E x output = E x evap =
where mew is the hourly freshwater, hfg is the evaporation phase change, I (t ) is the daily insolation, and As is the whole area of the device. The latent heat evaporation hfg can be calculated at the average basin water temperature (Tw ) as mentioned by Kabeel and Abdelgaied [33]:
× 100 (11)
mew × hfg (3600s. h−1)
T + 273.15 ⎞ × ⎛1 − a Tw + 273.15 ⎠ ⎝ ⎜
⎟
(12) 4
4 ⎛ Ta + 273.15 ⎞ 1 T + 273.15 ⎞ ⎤ + ×⎛ a E xsun = As × I (t ) s ⎡ ⎢1 − 3 × ⎥ T 3 Ts s ⎠⎦ ⎠ ⎝ ⎝ ⎣ ⎜
⎟
⎜
⎟
(13)
hfg
where mew is distillate water per hour, hfg denotes the latent heat of phase change, Ta is the ambient temperature and finally Tw is the temperature of the water basin. Also, As is the overall area of the plate, I (t ) s is the solar radiation intensity, while Ts is the sun temperature in 6000 K. The calculations indicate that the daily thermal exergy efficiencies for DPD, DPDW, and DPDW + CuO were approximately 3.7%, 4.5%,
= 103 × [2501.9 − 2.40706 × Tw + 1.192217 × 10−3 × Tw 2 − 1.5863 × 10−5]
E x evap
(10)
From the calculations, the daily thermal energy efficiency for DPD, DPDW, and DPDW + CuO were about 43.2%, 49.8%, and 60.5% respectively. On the other hand, the thermal energy efficiency for TPD is 8
Applied Thermal Engineering 168 (2020) 114848
S.W. Sharshir, et al.
Fig. 6. Thermal performance of the DPDW + CuO and TPD (a) ambient temperature, intensity of solar radiation, and the wind speed, (b) water and glass temperatures, (c) glass temperature difference between DPDW + CuO and TPD for four sides, and (d) hourly and accumulated freshwater.
The annual fixed price ( AFP ) was estimated through the fixed price (FP ) and the capital recovery factor (CRF ) as deduced from Eq. (16) [39].
and 5.6%, respectively. In contrast, the thermal exergy efficiency for TPD is about 2.9%. Furthermore, the daily exergy efficiency for DPDW and DPDW + CuO was magnificently higher than that of the TPD by about 93%. This shows the positive impacts of the wick and nanoparticles on the exergy performance of the pyramid distiller.
AFP = FP × CRF The CRF is estimated through Eq. (17) [51].
5. Cost analysis
CRF =
In this section, a cost analysis study in terms of the total price per one liter of freshwater for the traditional system (i.e., TPD) and the three different modifications (i.e., DPD, DPDW, and DPDW + CuO) was conducted. Firstly, the average total freshwater distillate per square meter per day during the year was estimated to be 2.8, 3.5, 4, and 5 L/m2 for the TPD, DPD, DPDW, and DPDW + CuO, respectively. Based on the sunshine periods throughout the year in Egypt, we assumed that the operating days of distillers per year are 340 days. The detailed cost analysis calculations The total price per liter of freshwater (TPPL ) was calculated through Eq. (14).
TPPL =
Annual total price Annual total distillate water productivity
i × (i + 1)n (i + 1)n − 1
(17)
where i is the interest per year assumed as 12% in this work and n is the years of the system life assume 10 years according to [39]. The value of AMOP is calculated by Eq. (18)
AMOP = 20%AFP
(18)
Also, the annual salvage value (ASV ) was estimated by Eq. (19) as in [51].
ASV = S × SFF
(19)
where S is the salvage value (assumed 20% of the fixed cost in this study), while SFF represents the sinking fund factor and is expressed by Eq. (20) [39].
SFF =
(14)
i (i + 1)n − 1
(20)
Table 2 illustrates the comparative evaluations of cost analysis results of the different types of the distillers considered in this work. The obtained results demonstrate that the total price of one liter of freshwater distilled was about 0.021$, 0.018$, 0.017$, and 0.015$ for the traditional, DPD, DPDW, and DPDW + CuO respectively. This clearly indicates that the inclusion of wick materials plus the nanoparticles not
The annual total price (ATP ) can be calculated as from Eq. (15) [39].
ATP = AFP + AMOP − AVS
(16)
(15)
where AFP is the annual fixed price, AMOP is the annual maintenance and operating price, while AVS is annual salvage value. 9
Applied Thermal Engineering 168 (2020) 114848
S.W. Sharshir, et al.
proposed in order to improve the thermo-economic performance of the pyramid distiller system. The modifications have been applied through the use of v-corrugated absorbers, wick materials, and CuO nanofluid. The developed modifications were named DPD (where v-corrugated absorber has been used), DPDW (where wick material has been integrated with v-corrugated absorbers), and DPDW + CuO (where CuO nanofluid has been added to the DPDW system). Experiential investigation of the three modified systems was performed, and results were compared to assess the thermal characteristics of each modification in comparison with the TPD. Moreover, the economic performance of TPD, DPD, DPDW, and DPDW + CuO was examined through conducting cost analysis study. The experimental results displayed that the freshwater yield of DPD was enhanced by 28.38% compared to TPD. While the freshwater yield of a DPDW was improved by about 45% compared to TPD. Finally, the largest enhancement in freshwater yield was obtained when DPDW + CuO was used with 72.95% increase compared to TPD. Also, the thermal efficiency DPD, DPDW, and DPDW + CuO were 43.2%, 49.8%, and 60.5% respectively, while the TPD was only 34%. Further, the daily exergy efficiencies of the DPD, DPDW, and DPDW + CuO were increased by almost 27.6%, 55%, and 93%, respectively compared to TPD. The comparative cost analysis results for the different studied systems indicate the DPDW + CuO system successfully gets the least price of one-liter of the freshwater distilled (0.015 $) with 28% reduction compared to the TPD system. In our opinion, the inclusion of CuO nanofluid can significantly improve the thermo-economic performance of pyramid distiller system in terms of increasing the thermal conductivity and absorptivity and reduction of the specific heat, which will enhance the rate of evaporation and freshwater production, as well as minimizing the one-liter price of the freshwater production.
Table 2 Results of a comparative cost analysis for the different studied systems. Cost item ($)
TPD
DPD
DPDW
DPDW + CuO
Fixed price Annual fixed price Annual maintenance and operating price Annual salvage value Total annual price Total amount of freshwater during one year Total price per one liter of freshwater
95 16.81 3.36
105 18.58 3.71
110 19.46 3.89
125 22.12 4.42
1.08 20.16 952
1.19 22.28 1190
1.25 23.35 1360
1.424 26.53 1700
0.021
0.018
0.017
0.015
only improves the performance of the distiller system in terms of freshwater productivity and the thermal energy but also minimize the cost per one liter of freshwater. 6. Thermo-economic performance comparison and water quality Fig. 7 provides a comparison between different thermal-economic metrics of the three developed modifications performed on PD together with the reference case (i.e., TPD). The figure showed the enhancement of freshwater productivity, energy and exergy efficiencies, and system cost with all modifications proposed in this work. The use of DPD, DPDW, and DPDW + CuO enhanced the total freshwater productivity by about 28.38%, 45%, and 72.95%, respectively, compared to the TPD. Also, the DPDW + CuO has the best improvement in terms of daily energy efficiency and exergy by 77.9% and 93%, respectively, when compared to TPD. Further, cost saving of 14.3%, 19%, and 28.5% of the total price of one liter of freshwater was obtained by DPD, DPDW DPDW + CuO, respectively, in comparison to TDP. Finally, among the proposed modifications, the last case with DPDW + CuO was of optimal results to be the winning system. The system has more abundant energy and exergy efficiencies and large production, as well as the lower price of one-liter of freshwater. Besides, the total dissolved solids (TDS) in brackish water collected from Burullus Lake was decreased from 1350 to 95 mg/L. Thus, the value of water’s pH was declined from 8.6 to 7.1 which is satisfactory based on WHO [68].
CRediT authorship contribution statement Swellam W. Sharshir: Conceptualization, Methodology, Writing original draft, Writing - review & editing, Funding acquisition. M.R. Elkadeem: Methodology, Formal analysis, Writing - review & editing. An Meng: Visualization, Investigation. Declaration of Competing Interests
7. Conclusions The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
In summary, three modifications on the TPD system have been
Fig. 7. Thermo-economic performance comparison of TPD, DPD, DPDW, and DPDW + CuO. 10
Applied Thermal Engineering 168 (2020) 114848
S.W. Sharshir, et al.
Foundation of China (NSFC) (Grant No. 51950410592).
influence the work reported in this paper. Acknowledgement This work was supported by the National Natural Science Appendix A
a1 = 9.999 × 102, a2 = 2.034 × 10−2, a3 = −6.162 × 10−3, a4 = 2.261 × 10−5, a5 = −4.57 × 10−8 b1 = 8.020 × 102, b2 = −2.001, b3 = 1.677 × 10−2, b4 = −3.060 × 10−5, b3 = 1.613 × 10−5
A = 5.328 − 9.76 × 10−2S + 4.04 × 10−4S 2 B = −6.913 × 10−3 + 7.351 × 10−4S − 3.15 × 10−6S 2
C = 9.6 × 10−6 − 1.927 × 10−6S + 8.23 × 10−9S 2
D = 2.5 × 10−9 + 1.666 × 10−9S − 7.125 × 10−12S 2 A1 = 1.474 × 10−3 + 1.5 × 10−5 × Tw − 3.927 × 10−8 × Tw2
B1 = 1.474 × 10−3 + 1.5 × 10−5 × Tw − 3.927 × 10−8 × Tw2 Appendix B. Supplementary material Supplementary data to this article can be found online at https://doi.org/10.1016/j.applthermaleng.2019.114848.
ponds: a review, Sol. Energy 174 (2018) 189–206. [19] M.R. Karimi Estahbanati, A. Ahsan, M. Feilizadeh, K. Jafarpur, S.S. Ashrafmansouri, M. Feilizadeh, Theoretical and experimental investigation on internal reflectors in a single-slope solar still, Appl. Energy 165 (2016) 537–547. [20] A.S. Nafey, M. Abdelkader, A. Abdelmotalip, A.A. Mabrouk, Enhancement of solar still productivity using floating perforated black plate, Energy Convers. Manage. 43 (2002) 937–946. [21] S.W. Sharshir, Y.M. Ellakany, A.M. Algazzar, A.H. Elsheikh, M.R. Elkadeem, E.M.A. Edreis, A.S. Waly, R. Sathyamurthy, H. Panchal, M.S. Elashry, A mini review of techniques used to improve the tubular solar still performance for solar water desalination, Process Saf. Environ. Prot. 124 (2019) 204–212. [22] A.E. Kabeel, S.W. Sharshir, G.B. Abdelaziz, M.A. Halim, A. Swidan, Improving performance of tubular solar still by controlling the water depth and cover cooling, J. Cleaner Prod. 233 (2019) 848–856. [23] M. Elashmawy, Effect of surface cooling and tube thickness on the performance of a high temperature standalone tubular solar still, Appl. Therm. Eng. 156 (2019) 276–286. [24] J. Hou, J. Yang, Z. Chang, H. Zheng, Y. Su, The mass transfer coefficient assessment and productivity enhancement of a vertical tubular solar brackish water still, Appl. Therm. Eng. 128 (2018) 1446–1455. [25] S. Abdallah, M.M. Abu-Khader, O. Badran, Effect of various absorbing materials on the thermal performance of solar stills, Desalination 242 (2009) 128–137. [26] S.W. Sharshir, G. Peng, L. Wu, F.A. Essa, A.E. Kabeel, N. Yang, The effects of flake graphite nanoparticles, phase change material, and film cooling on the solar still performance, Appl. Energy 191 (2017) 358–366. [27] A.F. Mohamed, A.A. Hegazi, G.I. Sultan, E.M.S. El-Said, Augmented heat and mass transfer effect on performance of a solar still using porous absorber: Experimental investigation and exergetic analysis, Appl. Therm. Eng. 150 (2019) 1206–1215. [28] E. Shanazari, R. Kalbasi, Improving performance of an inverted absorber multieffect solar still by applying exergy analysis, Appl. Therm. Eng. 143 (2018) 1–10. [29] R. Kalbasi, A.A. Alemrajabi, M. Afrand, Thermal modeling and analysis of single and double effect solar stills: an experimental validation, Appl. Therm. Eng. 129 (2018) 1455–1465. [30] M. Al-harahsheh, M. Abu-Arabi, H. Mousa, Z. Alzghoul, Solar desalination using solar still enhanced by external solar collector and PCM, Appl. Therm. Eng. 128 (2018) 1030–1040. [31] S.W. Sharshir, A.W. Kandeal, M. Ismail, G.B. Abdelaziz, A.E. Kabeel, N. Yang, Augmentation of a pyramid solar still performance using evacuated tubes and nanofluid: Experimental approach, Appl. Therm. Eng. 160 (2019) 113997. [32] Y. Tong, H. Lee, W. Kang, H. Cho, Energy and exergy comparison of a flat-plate solar collector using water, Al2O3 nanofluid, and CuO nanofluid, Appl. Therm. Eng. 159 (2019) 113959. [33] A.E. Kabeel, M. Abdelgaied, Observational study of modified solar still coupled with oil serpentine loop from cylindrical parabolic concentrator and phase changing material under basin, Sol. Energy 144 (2017) 71–78. [34] M. Bahrami, V. Madadi Avargani, M. Bonyadi, Comprehensive experimental and theoretical study of a novel still coupled to a solar dish concentrator, Appl. Therm. Eng. 151 (2019) 77–89. [35] D.B. Singh, Improving the performance of single slope solar still by including N identical PVT collectors, Appl. Therm. Eng. 131 (2018) 167–179. [36] V.K. Dwivedi, G.N. Tiwari, Comparison of internal heat transfer coefficients in
References [1] M. Elimelech, W.A. Phillip, The future of seawater desalination: energy, Technol. Environ. 333 (2011) 712–717. [2] K.G. Zhou, K.S. Vasu, C.T. Cherian, M. Neek-Amal, J.C. Zhang, H. GhorbanfekrKalashami, K. Huang, O.P. Marshall, V.G. Kravets, J. Abraham, Y. Su, A.N. Grigorenko, A. Pratt, A.K. Geim, F.M. Peeters, K.S. Novoselov, R.R. Nair, Electrically controlled water permeation through graphene oxide membranes, Nature 559 (2018) 236–240. [3] V.G. Gude, N. Nirmalakhandan, S. Deng, Desalination using solar energy: Towards sustainability, Energy 36 (2011) 78–85. [4] M. Elimelech, W.A.J.S. Phillip, The future of seawater desalination: energy, Technol. Environ., 333 (2011) 712–717. [5] H. Sharon, K.S. Reddy, A review of solar energy driven desalination technologies, Renew. Sustain. Energy Rev. 41 (2015) 1080–1118. [6] G. Xiao, X. Wang, M. Ni, F. Wang, W. Zhu, Z. Luo, K. Cen, A review on solar stills for brine desalination, Appl. Energy 103 (2013) 642–652. [7] M.H. Hamed, A.E. Kabeel, Z.M. Omara, S.W. Sharshir, Mathematical and experimental investigation of a solar humidification–dehumidification desalination unit, Desalination 358 (2015) 9–17. [8] R. Santosh, T. Arunkumar, R. Velraj, G. Kumaresan, Technological advancements in solar energy driven humidification-dehumidification desalination systems – a review, J. Cleaner Prod. 207 (2019) 826–845. [9] S.W. Sharshir, N. Yang, G. Peng, A.E. Kabeel, Factors affecting solar stills productivity and improvement techniques: a detailed review, Appl. Therm. Eng. 100 (2016) 267–284. [10] L. Zuo, L. Ding, J. Chen, X. Zhou, B. Xu, Z. Liu, Comprehensive study of wind supercharged solar chimney power plant combined with seawater desalination, Sol. Energy 166 (2018) 59–70. [11] Y. Zhang, L. Liu, K. Li, D. Hou, J. Wang, Enhancement of energy utilization using nanofluid in solar powered membrane distillation, Chemosphere 212 (2018) 554–562. [12] T. Arunkumar, K. Raj, D. Dsilva Winfred Rufuss, D. Denkenberger, G. Tingting, L. Xuan, R. Velraj, A review of efficient high productivity solar stills, Renew. Sustain. Energy Rev., 101 (2019) 197–220. [13] A.H. Elsheikh, S.W. Sharshir, M.K. Ahmed Ali, J. Shaibo, E.M.A. Edreis, T. Abdelhamid, C. Du, Z. Haiou, Thin film technology for solar steam generation: a new dawn, Sol. Energy 177 (2019) 561–575. [14] K. Kalidasa Murugavel, S. Sivakumar, J. Riaz Ahamed, K.K.S.K. Chockalingam, K. Srithar, Single basin double slope solar still with minimum basin depth and energy storing materials, Appl. Energy 87 (2010) 514–523. [15] H. Al-Hinai, M.S. Al-Nassri, B.A. Jubran, Effect of climatic, design and operational parameters on the yield of a simple solar still, Energy Convers. Manage. 43 (2002) 1639–1650. [16] Z.M. Omara, M.H. Hamed, A.E. Kabeel, Performance of finned and corrugated absorbers solar stills under Egyptian conditions, Desalination 277 (2011) 281–287. [17] M. Feilizadeh, M.R. Karimi Estahbanati, K. Jafarpur, R. Roostaazad, M. Feilizadeh, H. Taghvaei, Year-round outdoor experiments on a multi-stage active solar still with different numbers of solar collectors, Appl. Energy 152 (2015) 39–46. [18] A. Kasaeian, S. Sharifi, W.-M. Yan, Novel achievements in the development of solar
11
Applied Thermal Engineering 168 (2020) 114848
S.W. Sharshir, et al.
[37] [38] [39]
[40] [41]
[42] [43]
[44] [45]
[46]
[47]
[48]
[49]
[50]
[51]
[52] S.W. Sharshir, G. Peng, A.H. Elsheikh, E.M.A. Edreis, M.A. Eltawil, T. Abdelhamid, A.E. Kabeel, J. Zang, N. Yang, Energy and exergy analysis of solar stills with micro/ nano particles: a comparative study, Energy Convers. Manage. 177 (2018) 363–375. [53] L. Sahota, G.N. Tiwari, Effect of nanofluids on the performance of passive double slope solar still: A comparative study using characteristic curve, Desalination 388 (2016) 9–21. [54] A.E. Kabeel, M. Abdelgaied, N. Almulla, Performances of pyramid-shaped solar still with different glass cover angles: Experimental study, in: 2016 7th International Renewable Energy Congress (IREC), 2016, pp. 1–6. [55] B.W. Logan, G.R. Davies, J.F. Read, D.E. Cebulski, Carbonate Sedimentation and Environments, Shark Bay, Western Australia, American Association of Petroleum Geologists, 1970. [56] A. Elhakeem, W. Elshorbagy, Evaluation of the long-term variability of seawater salinity and temperature in response to natural and anthropogenic stressors in the Arabian Gulf, Mar. Pollut. Bull. 76 (2013) 355–359. [57] M.H. Sharqawy, J.H. Lienhard, S.M. Zubair, Thermophysical properties of seawater: a review of existing correlations and data, Desalin. Water Treat. 16 (2010) 354–380. [58] J.D. Isdale, R. Morris, Physical properties of sea water solutions: density, Desalination 10 (1972) 329–339. [59] D.T. Jamieson, J.S. Tudhope, R. Morris, G. Cartwright, Physical properties of sea water solutions: heat capacity, Desalination 7 (1969) 23–30. [60] D.T. Jamieson, J.S. Tudhope, Physical properties of sea water solutions: thermal conductivity, Desalination 8 (1970) 393–401. [61] S.Z. Heris, S.G. Etemad, M.N. Esfahany, Convective heat transfer of a Cu/Water nanofluid flowing through a circular tube, Exp. Heat Transf. 22 (2009) 217–227. [62] R.S. Vajjha, D.K. Das, Experimental determination of thermal conductivity of three nanofluids and development of new correlations, Int. J. Heat Mass Transf. 52 (2009) 4675–4682. [63] S. Akilu, A.T. Baheta, A.A. Minea, K.V. Sharma, Rheology and thermal conductivity of non-porous silica (SiO2) in viscous glycerol and ethylene glycol based nanofluids, Int. Commun. Heat Mass Transf. 88 (2017) 245–253. [64] N.N. Esfahani, D. Toghraie, M. Afrand, A new correlation for predicting the thermal conductivity of ZnO–Ag (50%–50%)/water hybrid nanofluid: an experimental study, Powder Technol. 323 (2018) 367–373. [65] M.H. Ahmadi, A. Mirlohi, M. Alhuyi Nazari, R. Ghasempour, A review of thermal conductivity of various nanofluids, J. Mol. Liq. 265 (2018) 181–188. [66] M. Fakoor Pakdaman, M.A. Akhavan-Behabadi, P. Razi, An experimental investigation on thermo-physical properties and overall performance of MWCNT/heat transfer oil nanofluid flow inside vertical helically coiled tubes, Exp. Therm. Fluid Sci. 40 (2012) 103–111. [67] S.W. Sharshir, A.H. Elsheikh, G. Peng, N. Yang, M.O.A. El-Samadony, A.E. Kabeel, Thermal performance and exergy analysis of solar stills – a review, Renew. Sustain. Energy Rev. 73 (2017) 521–544. [68] G. World Health Organization %J World Health Organization Guidelines for drinking water quality. 3rd, (2004).
passive solar stills by different thermal models: an experimental validation, Desalination 246 (2009) 304–318. R. Sathyamurthy, H.J. Kennady, P.K. Nagarajan, A. Ahsan, Factors affecting the performance of triangular pyramid solar still, Desalination 344 (2014) 383–390. S. Yadav, K.J.R. Sudhakar, S.E. Reviews, Different domestic designs of solar stills: a review, 47 (2015) 718–731. A.E. Kabeel, M.A. Teamah, M. Abdelgaied, G.B. Abdel Aziz, Modified pyramid solar still with v-corrugated absorber plate and PCM as a thermal storage medium, J. Cleaner Prod. 161 (2017) 881–887. A. Zeiny, H. Jin, L. Bai, G. Lin, D. Wen, A comparative study of direct absorption nanofluids for solar thermal applications, Sol. Energy 161 (2018) 74–82. E. De Robertis, E.H.H. Cosme, R.S. Neves, A.Y. Kuznetsov, A.P.C. Campos, S.M. Landi, C.A. Achete, Application of the modulated temperature differential scanning calorimetry technique for the determination of the specific heat of copper nanofluids, Appl. Therm. Eng. 41 (2012) 10–17. G. Xu, J. Fu, B. Dong, Y. Quan, G. Song, A novel method to measure thermal conductivity of nanofluids, Int. J. Heat Mass Transf. 130 (2019) 978–988. D. Singh, E.V. Timofeeva, M.R. Moravek, S. Cingarapu, W. Yu, T. Fischer, S. Mathur, Use of metallic nanoparticles to improve the thermophysical properties of organic heat transfer fluids used in concentrated solar power, Sol. Energy 105 (2014) 468–478. N.J. Hogan, A.S. Urban, C. Ayala-Orozco, A. Pimpinelli, P. Nordlander, N.J. Halas, Nanoparticles heat through light localization, Nano Lett. 14 (2014) 4640–4645. S. Rashidi, S. Akar, M. Bovand, R. Ellahi, Volume of fluid model to simulate the nanofluid flow and entropy generation in a single slope solar still, Renewable Energy 115 (2018) 400–410. G. Peng, H. Ding, S.W. Sharshir, X. Li, H. Liu, D. Ma, L. Wu, J. Zang, H. Liu, W. Yu, H. Xie, N. Yang, Low-cost high-efficiency solar steam generator by combining thin film evaporation and heat localization: Both experimental and theoretical study, Appl. Therm. Eng. 143 (2018) 1079–1084. X. Wang, Y. He, G. Cheng, L. Shi, X. Liu, J. Zhu, Direct vapor generation through localized solar heating via carbon-nanotube nanofluid, Energy Convers. Manage. 130 (2016) 176–183. X. Wang, Y. He, X. Liu, L. Shi, J. Zhu, Investigation of photothermal heating enabled by plasmonic nanofluids for direct solar steam generation, Sol. Energy 157 (2017) 35–46. G. Peng, S. Deng, S.W. Sharshir, D. Ma, A.E. Kabeel, N. Yang, High efficient solar evaporation by airing multifunctional textile, Int. J. Heat Mass Transf. 147 (2020) 118866. . Gao, P.K.N. Connor, G.W.J.E. Ho, E. Science, Plasmonic photothermic directed broadband sunlight harnessing for seawater catalysis and desalination, 9 (2016) 3151–3160. A.E. Kabeel, R. Sathyamurthy, S.W. Sharshir, A. Muthumanokar, H. Panchal, N. Prakash, C. Prasad, S. Nandakumar, M.S. El Kady, Effect of water depth on a novel absorber plate of pyramid solar still coated with TiO2 nano black paint, J. Cleaner Prod. 213 (2019) 185–191.
12