Next Article in Journal
Public Sentiment toward Solar Energy—Opinion Mining of Twitter Using a Transformer-Based Language Model
Next Article in Special Issue
Optimal Allocation of a Hybrid Photovoltaic Biogas Energy System Using Multi-Objective Feasibility Enhanced Particle Swarm Algorithm
Previous Article in Journal
The Idea of Justice in Innovation: Applying Non-Ideal Political Theory to Address Questions of Sustainable Public Policy in Emerging Technologies
Previous Article in Special Issue
Latent Heat Phase Change Heat Transfer of a Nanoliquid with Nano–Encapsulated Phase Change Materials in a Wavy-Wall Enclosure with an Active Rotating Cylinder
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Evaluation of the Melting Performance in a Conical Latent Heat Thermal Unit Having Variable Length Fins

1
Metamaterials for Mechanical, Biomechanical and Multiphysical Applications Research Group, Ton Duc Thang University, Ho Chi Minh City 758307, Vietnam
2
Faculty of Applied Sciences, Ton Duc Thang University, Ho Chi Minh City 758307, Vietnam
3
Young Researchers and Elite Club, Yasooj Branch, Islamic Azad University, Yasooj 7591493686, Iran
4
Mechanical Engineering Department, Gannon University, 109 University Square, Erie, PA 16541, USA
5
Department of Mechanical Engineering, College of Engineering at Wadi Addwaser, Prince Sattam Bin Abdulaziz University, Wadi Addwaser 11991, Saudi Arabia
6
Department of Mechanical Engineering, Faculty of Engineering, University of Khartoum, Khartoum 11111, Sudan
7
School of Engineering, Design and Built Environment, Western Sydney University, Kingswood, NSW 2747, Australia
*
Author to whom correspondence should be addressed.
Sustainability 2021, 13(5), 2667; https://doi.org/10.3390/su13052667
Submission received: 29 January 2021 / Revised: 21 February 2021 / Accepted: 23 February 2021 / Published: 2 March 2021

Abstract

:
A conical shell-tube design with non-uniform fins was addressed for phase change latent heat thermal energy storage (LHTES). The shell was filled with nano-enhanced phase change material (NePCM). The cone aspect ratio of the shell and the fins aspect ratio were adopted as the geometrical design parameters. The type and volume fraction of the nanoparticles were other design parameters. The investigated nanoparticles were alumina, graphite oxide, silver, and copper. The finite element method was employed to solve the natural convection flow and phase change thermal energy equations in the LHTES unit. The Taguchi optimization method was utilized to maximize the melting rate in the unit. Two cases of ascending and descending conical shells were investigated. The outcomes showed that the shell-aspect ratio and fin aspect ratio were the most important design parameters, followed by the type and concentration of nanoparticles. Both ascending and descending designs could lead to the same melting rate at their optimum design. The optimum design of LHTES could improve the melting rate by up to 18.5%. The optimum design for ascending (descending) design was a plain tube (a cone aspect ratio of 1.17) filled by 4.5% alumina-Bio-PCM (1.5% copper-Bio-PCM).

1. Introduction

Offering a high energy density and nearly isothermal operation, a latent heat thermal energy storage (LHTES) system is often a preferable choice over other thermal energy storage systems. LHTES systems use a phase change material (PCM) as the storage medium. Thermal energy is stored when the solid PCM undergoes a phase change to liquid, and then when the phase change process is reversed, i.e., the molten PCM solidifies, the stored energy is released. The performance of an LHTES system might be degraded due to the low thermal conductivity of the PCMs; therefore, implementation of heat transfer enhancement techniques may become necessary. The heat transfer enhancement techniques can be divided into two categories: (1) improving the thermal physical properties of the PCM by dispersion of highly conductive micro- and nano-sized particles and (2) using highly conductive inserts such as metal fins, heat pipes, and porous structures [1,2,3].
In recent years, the dispersion of highly conductive nanoparticles into the PCM has been utilized to improve the unit performance during the charging and discharging processes [4]. Various nanoparticles such as pure metals [5,6], metal oxides [7,8], and carbon-based nanoparticles [9] have been considered. The results of these studies have denoted that the volume fraction of nanoparticles plays an important role in controlling the melting and solidification processes. The impact of graphene nanoparticles on the melting process of various PCMs in a square cavity was studied by Kant et al. [10]. It was found that using nanoparticles improves the conductive heat transfer while weakening the natural convection heat transfer. The overall results, however, indicated that there is a reduction in the melting time.
Lin and Al-Kayiem [11] have experimentally evaluated the effects of incorporating Cu nanoparticles on the thermal characteristics of a paraffin wax during the melting and solidification processes. They have reported that the dispersion of copper nanoparticles with a concentration of 2.0% by weight leads to a 46.3% increase in the thermal conductivity of the composite PCM.
A numerical study performed by Mahdi and Nsofor [12] indicated that the discharging time is reduced by 8–20% of alumina nanoparticles with volume concentrations of 8–13%, respectively, are used. The melting of paraffin in the presence of copper nanoparticles in an elliptical capsule was numerically studied by Boukani et al. [13]. Three volumetric nanoparticle concentrations of 0%, 1%, and 3% and three container aspect ratios of 2, 1, and 0.5 were employed. The results revealed that, at a fixed aspect ratio, the presence of nanoparticles improves the performance of the system.
As mentioned earlier, fins improve the distribution of heat through high thermal conductive extended surfaces, and thus, they reduce a thermal resistance in an LHTES. Gharebaghi and Sezai [14] studied the influence of aluminum fin arrays on the melting process of paraffin wax in a rectangular container. The results showed that the melting time and the stored energy were impacted by the spacing and the height of the fin arrays. The performance of a fin and tube LHTES system was experimentally investigated by Rahimi et al. [15]. The heat transfer fluid (HTF) flowed inside a spiral tube, and plate fins were used to increase the heat transfer rate. The results showed that the presence of fins notably reduces the melting time, and the closer the fins, the shorter the melting time. Yang et al. [16] simulated the phase change heat transfer in an annularly finned LHTES. The authors discussed the fin quantity, height, and thickness for optimal charging of a shell-tube LHTES unit. The authors considered the contribution of natural convection on the overall melting. It was found that the presence of fins diminishes the charging time by 65%.
In addition to plate and circular fins, the use of longitudinal fins has also been studied by researchers [17,18]. In a research study conducted by Kazemi et al. [19], the effect of the longitudinal fin angle was evaluated. The results showed that a lower melting time was obtained when the angle was increased from 60° to 120° for the case with three fins. However, when two fins were used, increasing the angle from 45° to 150° deteriorated the performance of the system. The effect of fin angle was also investigated by Deng et. al. [20]. In an experimental study carried out by Rathod and Banerjee [21], the augmentation in heat transfer was studied in a shell and tube heat exchanger. Three longitudinal fins were installed at the outer surface of the inner tube to increase the heat transfer surface. The results showed that the decreased percentage in melting time relies on the heat transfer fluid (HTF) inlet temperature.
The available studies in the literature have reported that combining heat transfer enhancement techniques can lead to a better heat transfer performance of the storage unit [22,23]. Studies can be found using a combination of fins and heat pipes [24], nanoparticles and fins [25,26], nanoparticles and metal foams [27], metal foams, and heat pipes [28]. Ren et al. [29] studied the application of triangular fins in a rectangular LHTES with and without the presence of nanoparticles in the PCM. It was reported that using triangular fins was more efficient than solely using nanoparticles. In addition, the arrangement with long and narrow fins that were appropriately spaced provided a higher energy storage efficiency. Li et al. [30] studied the melting of graphene nanoplatelets in 1-tetradecanol in a rectangular enclosure. They investigated the impact of various mass fraction of nanoparticles and enclosure aspect ratio on the melting behavior of nano-enhanced phase change material (NePCM). They found that a slender enclosure could be better for application of NePCMs.
Darzi et al. [31] studied the influence of copper nanoparticles in three horizontal double pipes LHTES systems. The three configurations included a circular shell with an elliptical inner pipe, a cylindrical inner pipe, and a finned cylindrical inner pipe. The effect of the number of fins, as well as the orientation of the ellipse, were evaluated for different concentrations of nanoparticles. Mahdi and Nsofor [22] studied the combined effects of nanoparticles and longitudinal fins in a shell and tube LHTES unit. Fins were added both on the outer surface of the inner pipe and the inner wall of the shell. It was found that better solidification time was achieved for the case with only fins when compared to the cases with only nanoparticles or with a combination of both.
The literature review of the available studies shows that natural convection heat transfer plays a role in the system’s performance, especially during the melting process. It should be noted that the significance of these effects is influenced by various parameters such as PCM type, container configuration, and container orientation. Shell and tube latent heat thermal energy storage systems are one of the common types of LHTES systems and are considered for the current study. The heat transfer fluid flows inside the inner tube, while the PCM is located in the shell. As the container is located vertically, the melting process will be governed by natural convection heat transfer. Therefore, due to the upward motion of the molten PCM, more melting will occur at the top of the container as compared to the lower section. To provide a more uniform and accelerated melting process, a conical shell is considered. To further accelerate the melting process, combinations of fins and dispersed nanoparticles are used as the heat transfer enhancement technique.
As mentioned earlier, fins contribute to the enhanced heat transfer process by providing more surface area. The level of improvement varies based on the design of the shell and the fins geometry. In some cases, the presence of the fins may dampen the natural convection flows, which may suppress some of the potential improvements. On the other hand, using dispersed nanoparticles increases the effective thermal conductivity of the NePCM, but it also increases the viscosity of NePCM, which consequently limits the liquid motion and natural convection. By considering these factors, finding an ideal configuration of an LHTES system and enhancement techniques is a crucial task. Thus, the present study aims to address the melting heat transfer in a conical shape LHTES and its optimum design for the shell’s aspect ratio, fins’ aspect ratio, type of nanoparticles, and volume fractions of nanoparticles for the first time.

2. Mathematical Model

A conical latent heat thermal energy storage (LHTES) system, filled with a nano-enhanced phase change material (NePCM), is illustrated in Figure 1. The general shape of the LHTES’ shell is a conical shape that could be a descending or ascending conical enclosure. The fins can also be uniform, ascending, or descending. Based on the shape of the LHTES unit shell here, two types of the general design were adopted. A view of the ascending conical and the descending conical LHTES units are depicted in Figure 1a,c, respectively. The cylindrical design is a special case of each of these two designs and is illustrated in Figure 1b. Assuming the LHTES volume is constant, the radius of the cross-section of the system along the z-direction can be descending, constant, or increasing. Furthermore, the height of the connected copper fins can be changed according to the shape of the system. The height of the LHTES is L = 200 mm, and the radius of the normal system with a uniform circular cross-section (depicted in Figure 1b) is Ro = 50 mm. In the conical LHSS, the lower radius of the cross-section, i.e., Rdc, is defined base on the aspect ratio, i.e., ARc = Rdc/ Ro. Also, the length of the lowest fin, i.e., wdf, is ARf × wuf in which wuf is the length of identical fins. It is worth noting that the volume of all fins is considered a constraint. Figure 2 shows the structural specification of the unit. Water, as the heat transfer fluid has a high temperature, enters the tube and the energy is transferred to the composite NePCM through the tube wall. The NePCM contains the highly thermal conductive nano-additives added to the coconut oil as the base PCM. The thermophysical properties of the base PCM, HTF, and the fins are listed in Table 1. As tabulated in Table 2, the nano additives dispersed in the PCM can separately be Al2O3, GO, Ag, and Cu. Furthermore, the axisymmetric structure of the system allows the problem to be solved in 2D space.
For the NePCM domain, the enthalpy-porosity technique is used to model the melting flow. The controlling equations, i.e., the equations describing mass, momentum, and energy conservations, are:
D ρ D t = 0 ;
ρ L N e P C M D V D t = p + η L N e P C M 2 V + A m 1 s T 2 10 3 + s T 3 V + g ρ L N e P C M β L N e P C M T T m e l t i n g ,
where the velocity vector ( V ) and the pressure field (p) are the dependent variables. The density (ρ), volume expansion coefficient (β), and dynamic viscosity (μ) are the thermophysical properties. The gravity direction ( g ) is toward the bottom, and its component in the horizontal direction is zero. The subscripts LNePCM shows the liquid NePCM and Am (O(106)) is a large number to procurance the value of the source term. The liquid fraction (s) is a temperature dependent-function:
s T = 0          T < T m e l t i n g 0.5 T w i n d o w T T m e l t i n g T w i n d o w + 0.5     T m e l t i n g 0.5 T w i n d o w < T < T m e l t i n g + 0.5 T w i n d o w 1          T > T m e l t i n g + 0.5 T w i n d o w ;
controls the velocity components in the momentum domain. The symbols T m e l t i n g and T w i n d o w are the melting and window temperatures.
[ s T ρ C p L N e P C M + 1 s T ρ C p S N e P C M ] D T D t =                     s T k L N e P C M + 1 s T k S N e P C M T + V F n a 1 ρ P P C M L P P C M s T t
where temperature (T) is the dependent variable. Subscripts SNePCM and PPCM refer to the properties of the solid NePCM, and the pure PCM (without nanoparticles). Thermal conductivity coefficient (k), latent heat of phase transition (L), and heat capacity (Cp) are the thermophysical properties. VFna represents the volume fraction of nanoparticles.
For the solid fin domain, the energy balance can be defined as the following:
ρ C p s T t = k s T ,
where the subscript s denotes the solid fin.
Assuming the fluid flow passing the tube is laminar and incompressible, the governing equations are:
D V D t = 0 ,
ρ H T F D V D t = p + η H T F 2 V ,
ρ C p H T F D T D t = k H T F T .
The subscript HTF points to heat transfer fluid (water) in the tube. The thermophysical characteristics of the nano-enhance phase change material are reached based on the relations listed below:
Density:
ρ N e P C M = 1 V F n a ρ P P C M + V F n a ρ n a ;
ρ P P C M = s T ρ L P P C M ρ S P P C M + ρ S P P C M .
PPCM denotes the properties of the pure phase change material.
Dynamic viscosity:
η L N e P C M = η L P P C M 1 V F n a 2.5 .
Coefficient of thermal expansion
β L N e P C M = ρ L P P C M β L P P C M V F n a ρ L P P C M β L P P C M ρ n a β n a ρ L N e P C M .
Thermal conductivity:
k L / S N e P C M k L / S P P C M = k n a + 2 k L / S P P C M 2 V F n a k L / S P P C M k n a k n a + 2 k L / S P P C M + V F n a k L / S P P C M k n a .
Effective heat capacity:
C p , L N e P C M = ρ P P C M C p , P P C M V F n a ρ P P C M C p , P P C M ρ n a C p , n a ρ L N e P C M .
C p , P P C M = s T ρ L P P C M C p , L P P C M ρ S P P C M C p , S P P C M + ρ S P P C M C p , S P P C M ρ P P C M .
The relevant initial and boundary conditions for the above equations are as follows.
At the interface of the tube and NePCM domain:
k T r N e P C M = k T r H T F ,   T H T F = T N e P C M .
At the entrance of the tube:
T   H T F = 313 K ,   v r   H T F = 0 ,   v z   H T F = 0.01   m / s .
At the outlet tube:
v r   H T F = 0 ,   T z H T F   = v z z H T F = 0 .
Also, the initial temperature of HTF, NePCM and fins is T i n   = 293   K where the subscripts r and z are the coordinates in vertical (axial) and horizontal (radial) directions, respectively. At the vertical outer wall of the NePCM domain:
v r   N e P C M = v z   N e P C M = 0 ,   T r N e P C M   = 0 .
At the horizontal walls of the NePCM domain:
v r   N e P C M = v z   N e P C M = 0 ,   T z N e P C M   = 0 .
The pertinent initial condition of the NePCM zone:
v r   N e P C M = v z   N e P C M = 0 ,   T i n , N e P C M   = 293   K .
In the LHTES unit, the total stored energy is defined as the following:
E S t = A ρ L N e P C M C p , L N e P C M T T i n + 1 V F n a ρ P P C M L P P C M   d A .
The liquid fraction of NePCM during melting flow is expressed as:
M V F = A L N e P C M A L N e P C M + A S N e P C M .
A L N e P C M is the area that NePCM is in liquid phase, and A S N e P C M is the non-melted NePCM area. To evaluate the uniformity of the temperature distribution, σ (t) can be defined as follows:
σ 2 t = A T T a v e r a g e 2 d A / A d A .
T a v e r a g e indicates the average temperature of the PCM domain.

3. Numerical Approach and Grid Dependency

3.1. Solution Method

To find the fields of MVF, temperature, and velocity, we employed the numerical analysis by user defined finite element codes based on the Petrov–Galerkin approach. The governing equations were transformed into a weak form, and linear shape function (for heat equation) and quadratic Lagrange (for momentum equation) were applied. Then, the equations were integrated over mesh elements to obtain the residual equations. The residual equations were solved in the fully coupled form using the PARallel DIrect SOlver (PARDISO) [37,38] with a relative tolerance of 1 × 10−4. A Newtonian damping factor of 0.9 was applied to enhance the convergence of the solver. The selection of the time-step is crucial in phase change problems with natural convection effects. Here, a backward differentiation formula (BDF) was used to control the computations’ accuracy and convergence automatically. The details of the utilized finite element method and solution method can be found in [39,40].
To reach the high accuracy results, the mesh density was optimized through sensitivity analysis of the numerical results to the element numbers. Here a size parameter, NE, was introduced to control the size of mesh at various regions. The NE parameter was changed in the range of 0.75–1.75. A larger NE shows more elements and thus a fanner mesh. The details of the mesh elements in each region have been reported in Table 3. The mesh type in all domains is quad, and the mesh in the fins is structured, leading to rectangular mesh in the solid-fins domain. The reset of domains was filled with un-structured quad elements. The results of optimization can be found in Figure 3. As seen, the mesh with NE = 1.25 can be selected as the optimized grid.

3.2. Validation of the Model with Literature Data

The implemented numerical model can be verified by employing a numerical analysis performed by Costa [41] and an experimental work conducted by Kumar et al. [42]. Costa [41] studied the influence of the thermal conductivity of the separators on the free convection in a partially divided cavity. Figure 4 shows the temperature distribution of the current model and that simulated by Costa [41]. The working fluid filling the cavity is air with a Prandtl number of 0.71. The left and right walls were subject to cold and hot isothermal temperatures. The domain temperatures were scaled for the temperature difference between the hot and cold walls. Thus, the isotherms are in the range of 0–1. The natural convection Ryleigh number was 106. The partition walls were ten times more thermally conductive than the fluid inside the enclosure (air), and the size of the partition was 0.6 size of the enclosure, and the distance between the partition and side wall was 0.3 enclosure size.
In another validation, the liquid fraction field observed by Kumar et al. [42] which was also simulated by the current model are compared in Figure 5. The authors investigated the melting of lead in an enclosure with a height of 50 mm and an aspect ratio of 0.75 (height/width). The enclosure walls were well insulated except the right wall, which was subject to a constant heat flux (16.3 kW/m2). The thermal neutron radiography method was used to map the melting interface. As seen, the melting starts from the right wall and advances toward the left. Moreover, due to the natural convection effects, the melting at the top is faster than at the bottom.
In this comparison, the right side-wall of the cavity occupied by lead was exposed to a constant heat flux. The comparative analyses presented in Figure 4 and Figure 5 confirm the reliability of the utilized code.

4. Results and Discussion

The present study aims to maximize the melting rate at seven hours of charging time. Here, the Taguchi optimization method is utilized to optimize four design parameters for the descending and ascending conical designs. The design parameters are the volume fraction of nanoparticles V F n a ( 0 V F n a 0.045 ), the ratio of the radius of the lower fin to the radius of the base fin ( 0.1 A R f 1.0 for ascending conical and 1.0 A R f 1.6 for descending conical design), the ratio of the lower radius of conic to the radius of the regular cylinder ( 0.4 A R c 1.0 , 1.0 A R c 1.5 ), and type of nanoparticles. The type of nanoparticles could be Al2O3, GO (graphene oxide), Ag, and Cu. The amount of molten PCM is selected as the target design. Thus, “the higher, the better” design strategy was set for the Taguchi method. Each of the design parameters is divided into four levels. The details of the design parameters and their levels are summarized in Table 4 and Table 5. These tables show the design for ascending and descending conical shells, respectively.
The combination of four design parameters and four levels leads to 44 possible design combinations for each shell geometry. Simulating all possible combinations requires a tremendous computational budget. Thus, Taguchi uses an orthogonal table to reduce the required simulations. Here, the standard orthogonal L16 table has been selected for each design geometry. Table 6 and Table 7 show the L16 design cases for each of the ascending and descending conical shells, respectively.
The melting phase change simulations were executed for each of the design cases of L16 Table 6 and Table 7. Then, the values of MVF and Total stored energy (ES) after seven-hours of charging (melting) were reported in the Tables. The Taguchi method was used to compute the values of the S/N ratio based on the values of MVF. Since the target is “the higher, the better,” a larger value of S/N ratio shows a better-designed case. The Taguchi method uses these S/N ratios to find the best level of each design parameter and reach a final optimum design.
Figure 6 and Figure 7 show the values of the S/N ratio for each level of each design parameter for two geometrical designs of ascending and descending conical shells. From Figure 6, it can be found that the maximum S/N values correspond to VFna = 0.045, ARf = 0.4, ARc = 1.0, and Al2O3 nanoparticles for ascending design. The optimum design levels for descending design could be found in Figure 7, as follows: VFna = 0.015, ARf = 1.4, ARc = 1.17, and Cu nanoparticles. The optimum design parameters and the estimated values of MVF by the Taguchi method are summarized in Table 8. Numerical simulations were also executed for these optimum cases, and the computed MVF is reported in Table 8. The simulated cases show an MVF = 0.92 after seven hours of thermal charging for both ascending and descending designs. These values are greater than equal to the simulated case of L16 tables. Thus, they have been selected as optimum designs.
It should be noted that the range of the variation of S/N in Figure 6 and Figure 7 shows the impact of a design parameter on the MVF. As seen, the most significant parameter is ARc for both cases. This parameter indicates the ratio of the lower radius of conic to the radius of the regular cylinder. The next important design parameter is ARf, which represents the ratio of the radius of the lower fin to the radius of the base fin. The other less important design variables are the type of nanoparticles and the volume fraction of nanoparticles. Thus, special attention should be made to selecting shell ratio (ARc) and fin ratio (ARf). Then, the LHTES unit’s charging rate can be fine-tuned by selecting the type and volume fraction of nanoparticles.
For ascending conical shell geometry, the minimum MVF corresponds to case 16 Table 6 with MVF = 0.75. Thus, the optimum case with MVF = 0.92 increases the melting rate by 18.5% compared to this case. The lowest MVF = 0.76 corresponds to case 13 of Table 7 for a descending design. Since the optimum case is MVF = 0.92, a 17.3% increase of MVF can be achieved by an optimum design.
Figure 8 shows the melting maps of NEPCM for the optimum case of ascending conical shell. The results are reported at seven different time steps during the melting process. Figure 9 depicts the isotherm maps for the same snapshots of Figure 6. As seen, the melting commences around the tube wall and fins. Then, it advances in the enclosure toward the adiabatic shell. At t = 3750 s, small circulation cells between fins can be seen in molten regions. The circulation extends to a large portion of the enclosure by the advancement of melting, and a single central circulation occurs at t = 15,000 s and beyond. The solid zones quickly disappear at the top of the enclosure due to the circulation flows and long fins at the top. The only un-melted region at 26,250 s is at the bottom of the enclosure. This region slowly melts down as time passes. Figure 9 shows that the HTF in the tube is mostly at a uniform hot temperature. Only at the initial charging stages, a smooth low-temperature boundary layer next to the tube wall can be seen. At 7500 s and larger, the heat transfer rate at the PCM sides drops notability, and the temperature of HTF in the tube is uniformly hot. It is interesting that the temperature of the molten PCM reaches the HTF temperature at the top regions of the enclosure, where there is a small amount of solid PCM at the bottom.
Figure 10 and Figure 11 show melting fraction maps and the temperature in the optimum design of an ascending conical shell. As seen in this design, the fins are long at the bottom and short at the top. The shell is also wide at the bottom and narrow at the top. In the descending case, the same as the ascending one, the melting commences from the tube wall and around the fins and advances into the enclosure toward the shell. However, in this design, there is a narrow space between the tube and con-shell at the top. This narrow space breaks the main circulation flow into two distinct circulation flows. The PCM melts uniformly next to the cone shell. Thus, a smaller piece of solid PCM can be seen at the bottom at t = 22,500 s compared to the ascending design. Moreover, the hot molten PCM is in the middle of the enclosure and advances toward the corners.
Figure 12 shows the temperature variation of eight monitor points in each of the ascending and descending designs. The monitor points are depicted in Table 9. The points start from the bottom and move upward. As seen, all of the temperature points are initially at a supper cold temperature of 293 K. Then, the temperature sharply increases to about 295 K. This sharp rise of temperature corresponds to the solid region’s conduction heat transfer. Then the phase change occurs, and the rise of the temperature limits to the phase change temperature bond. In this situation, the heat transfer does not increase the temperature notably, and the heat stores in the PCM in the form of latent heat. Once the PCM at the monitor point melts down, the temperature rises again. The rise of temperature at the end of the melting process is due to the increase of temperature in the molten PCM. The monitor points at the top of the enclosure tend to quickly reach about 313 K, which is the temperature of HTF. The monitor points show a short phase change history for the case of ascending design. The melting maps also showed that the bottom of the enclosure remains at solid state during most of the charging time. Figure 12b shows the monitor points have almost similar behavior since the melting front advances in the enclosure uniformly toward the enclosure shell.
Figure 13a–c show the melted portion (MVF), total stored energy (ES), and temperature non-uniformity (σ) during the melting process, respectively. The results are plotted for two ascending and descending designs. As seen, the descending designs lead to a higher MVF during the charging process.
Only at the final stages of melting, the ascending design is better than the descending design. Thus, for applications in which the full charging is not very important and about 85% of charging is adequate, then the descending design can be selected. However, when the full melting is the main target of design, the ascending design should be selected.
Figure 13b shows that the total stored energy for the case of ascending design is also higher around the final stage of charging (just before full charging). The temperature non-uniformity for the ascending design is higher than the case of the descending design at the end of the charging process. This is because the top of the ascending design at the end of the charging process is at a hot temperature and subject to natural convection while its bottom is not melted yet. Thus, the LHTES unit is subject to high-temperature differences.
The results of Figure 6 show that the presence of nanoparticles could induce a minor impact on MVF, and the increase of the concentration of nanoparticles does not improve MVF monastically. These findings are in agreement with the results of the experimental investigation of Li et al. [30]. The experimental results show that the presence of nanoparticles increases both thermal conductivity and dynamic viscosity. Thus, using NePCMs could benefit conduction-dominant regimes where the heat transfer is under the direct influence of thermal conductivity while the dynamic viscosity is not important. However, as the melting advances and natural convection flows occur the dynamic viscosity influences the intensity of liquid circulation. Thus, the growth of viscosity due to the presence of nanoparticles could slow down the natural convection effects. It should also be noted that the natural convection effects are under the influence of geometrical design (enclosure shape and fins), and therefore, the combination of these effects should be taken into account in the heat transfer design of LHTES systems.
Currently, various types of NePCMs are being synthesized by various researchers. Some of these works have been reviewed by Leong et al. [43] and Nižetić [44]. The analysis of synthesized NePCM samples shows that the presence of nanoparticles changes the host PCM’s thermophysical properties. It should be noted that the heat transfer performance of LHTES systems is a nonlinear function of the thermophysical properties and their geometrical design. Thus, numerical simulations/experimental studies are demanded to investigate the advantage/drawback of using NePCMs. The present study took into account the combination of geometrical factors and the presence of nanoparticles to analyze the heat transfer behavior of LHTES systems for various concentrations and types of nanoparticles.

5. Conclusions

The melting heat transfer and thermal energy storage in a conical-shell LHTES unit filled with nano-enhanced phase change materials were addressed. The melting heat transfer was enhanced by applying a non-uniform array of fins. Two possible shell-designs, ascending and descending designs, were investigated. Four types of nanoparticles were examined. For each design, the volume fraction and type of nanoparticles, the shell aspect ratio, and the fin aspect ratio were optimized using the Taguchi method for maximum melting rate. The major finding of the current research can be summarized as follows:
  • The most important parameter influencing the melting rate is the aspect ratio of the shell (ARc). The optimum ARc for the ascending design was 1.0 (a cylindrical shell), and for the descending design, it was 1.17.
  • The fins arrangement was the second most important design parameter, influencing the melting rate. The best fin aspect ratios were the cases with a smooth variation of fin sizes. Thus, the middle size values of ARf = 0.4 and 1.4 for ascending and descending designs were led to optimum melting rate, respectively. The melting maps showed that the fins significantly contribute to heat transfer at the initial stages of melting, where the heat transfer is conduction dominant. However, at the final stages of melting, the natural convection circulations play an important role in melting heat transfer. The fins’ arrangement is crucial since long fins could suppress the natural convection flows.
  • The type of nanoparticles was the third most important design parameter, influencing the melting rate. The optimum type of nanoparticles depends on the designed geometry. Alumina nanoparticles were the best for an ascending design, while copper nanoparticles were a better choice for a descending design. Indeed, the presence of nanoparticles influences the heat capacity and thermal conductivity of the NePCM. Thus, the type of these nano-additives should be selected carefully.
  • The presence of nanoparticles could improve the melting rate. However, the best volume fraction of nanoparticles for a descending design was only 1.5% of nanoparticles, while the maximum possible volume fraction could be 4.5%. This shows that the increase in the volume fraction of the nanoparticles is not always advantageous. The nanoparticles increase the dynamic viscosity of molten NePCM, and thus, they can weaken the natural convection flows.

Author Contributions

Conceptualization, M.G., S.A.M.M. and M.M.; methodology, M.G., S.A.M.M., M.M., O.Y. and M.A.A.; software, S.A.M.M.; validation, S.A.M.M.; formal analysis, M.G. and S.A.M.M.; and M.A.A.; investigation, M.G., S.A.M.M., M.M., O.Y. and M.A.A.; resources, M.G., O.Y. and M.A.A.; writing—original draft preparation, M.G., S.A.M.M., M.M., O.Y. and M.A.A.; writing—review and editing, M.G., S.A.M.M., M.M., O.Y. and M.A.A.; visualization, S.A.M.M. and M.G; supervision, M.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

Asurface area of domain, m2
aslope of the outer surface of the shell, (RucRdc)/L
A mmushy parameter, kg m−3 s−1
ARcratio of the lower radius of the conical shell to the radius of the uniform cylindrical shell, ARc = Rdc/Ro
ARfratio of the length of lowest fin to the length of identical fin, wdf/wuf
C pspecific heat capacity, J kg−1 K−1
ddistance between two adjacent fins, m
EStotal stored energy, kg m2 s−2
ggravitational acceleration, m s−2
hsfdistance between the lowest fin and the lower surface of the shell, mm
kthermal conductivity, kg m2 s−3 K−1
Lmelting latent, m2 s−2; heat storage height, m
MVFmelting volume fraction
Nanotype of nanoparticle
NEparameter for mesh size control
ppressure, kg m−1 s−2
Pimonitor points
r, zcoordinate system, m
Rdclower radius of the cylindrical shell, mm
Riheating tube radius, m
Roradius of the uniform cylindrical shell, m
Rucupper radius of the cylindrical shell, mm
slocal liquid fraction
S/Nsignal to noise ratio
ttime s; fin’s thickness, m
Ttemperature, K
Tininitial temperature, K
Tmeltingmelting temperature, K
V velocity vector, m
VFnaconcentration of nanoparticles
vr, vzr, z velocities directions, m s−1
wdflength of the lowest fin, m
wffin length, m
wuflength of the top fin, m
Greek
µdynamic viscosity, kg s−1 m−1
βthermal expansion coefficient, K−1
ρdensity, kg m−3
σparameter of the temperature distribution, K
Subscripts
averageaverage property
HTFheat transfer fluid
LNePCMnano-enhanced phase change material at liquid phase
NePCMnano-enhanced phase change material
PPCMpure phase change material
ssold fins
SNePCMnano-enhanced phase change material at solid phase

References

  1. Mahdi, J.M.; Lohrasbi, S.; Nsofor, E.C. Hybrid heat transfer enhancement for latent-heat thermal energy storage systems: A review. Int. J. Heat Mass Transf. 2019, 137, 630–649. [Google Scholar] [CrossRef]
  2. Maldonado, J.M.; De Gracia, A.; Cabeza, L.F. Systematic review on the use of heat pipes in latent heat thermal energy storage tanks. J. Energy Storage 2020, 32, 101733. [Google Scholar] [CrossRef]
  3. Shabgard, H.; Bergman, T.L.; Sharifi, N.; Faghri, A. High temperature latent heat thermal energy storage using heat pipes. Int. J. Heat Mass Transf. 2010, 53, 2979–2988. [Google Scholar] [CrossRef]
  4. Xiong, T.; Zheng, L.; Shah, K.W. Nano-enhanced phase change materials (NePCMs): A review of numerical simulations. Appl. Therm. Eng. 2020, 178, 115492. [Google Scholar] [CrossRef]
  5. Hosseinizadeh, S.F.; Darzi, A.A.R.; Tan, F.L. Numerical investigations of unconstrained melting of nano-enhanced phase change material (NEPCM) inside a spherical container. Int. J. Therm. Sci. 2012, 51, 77–83. [Google Scholar] [CrossRef]
  6. Elbahjaoui, R.; El Qarnia, H. Thermal analysis of nanoparticle-enhanced phase change material solidification in a rectangular latent heat storage unit including natural convection. Energy Build. 2017, 153, 1–17. [Google Scholar] [CrossRef]
  7. Pahamli, Y.; Hosseini, M.J.; Ranjbar, A.A.; Bahrampoury, R. Effect of nanoparticle dispersion and inclination angle on melting of PCM in a shell and tube heat exchanger. J. Taiwan Inst. Chem. Eng. 2017, 81, 316–334. [Google Scholar] [CrossRef]
  8. Parsazadeh, M.; Duan, X. Numerical study on the effects of fins and nanoparticles in a shell and tube phase change thermal energy storage unit. Appl. Energy 2018, 216, 142–156. [Google Scholar] [CrossRef]
  9. Iachachene, F.; Haddad, Z.; Oztop, H.F.; Abu-Nada, E. Melting of phase change materials in a trapezoidal cavity: Orientation and nanoparticles effects. J. Mol. Liq. 2019, 292, 110592. [Google Scholar] [CrossRef]
  10. Kant, K.; Shukla, A.; Sharma, A.; Henry Biwole, P. Heat transfer study of phase change materials with graphene nano particle for thermal energy storage. Solar Energy 2017, 146, 453–463. [Google Scholar] [CrossRef]
  11. Lin, S.C.; Al-Kayiem, H.H. Evaluation of copper nanoparticles—Paraffin wax compositions for solar thermal energy storage. Solar Energy 2016, 132, 267–278. [Google Scholar] [CrossRef]
  12. Mahdi, J.M.; Nsofor, E.C. Solidification of a PCM with nanoparticles in triplex-tube thermal energy storage system. Appl. Therm. Eng. 2016, 108, 596–604. [Google Scholar] [CrossRef]
  13. Hajighafoori Boukani, N.; Dadvand, A.; Chamkha, A.J. Melting of a Nano-enhanced Phase Change Material (NePCM) in partially-filled horizontal elliptical capsules with different aspect ratios. Int. J. Mech. Sci. 2018, 149, 164–177. [Google Scholar] [CrossRef]
  14. Gharebaghi, M.; Sezai, I. Enhancement of Heat Transfer in Latent Heat Storage Modules with Internal Fins. Numer. Heat Transf. Part A Appl. 2007, 53, 749–765. [Google Scholar] [CrossRef]
  15. Rahimi, M.; Ranjbar, A.A.; Ganji, D.D.; Sedighi, K.; Hosseini, M.J.; Bahrampoury, R. Analysis of geometrical and operational parameters of PCM in a fin and tube heat exchanger. Int. Commun. Heat Mass Transf. 2014, 53, 109–115. [Google Scholar] [CrossRef]
  16. Yang, X.; Lu, Z.; Bai, Q.; Zhang, Q.; Jin, L.; Yan, J. Thermal performance of a shell-and-tube latent heat thermal energy storage unit: Role of annular fins. Appl. Energy 2017, 202, 558–570. [Google Scholar] [CrossRef]
  17. Khan, Z.; Khan, Z.A. An experimental investigation of discharge/solidification cycle of paraffin in novel shell and tube with longitudinal fins based latent heat storage system. Energy Convers. Manag. 2017, 154, 157–167. [Google Scholar] [CrossRef] [Green Version]
  18. Mahdi, J.M.; Lohrasbi, S.; Ganji, D.D.; Nsofor, E.C. Accelerated melting of PCM in energy storage systems via novel configuration of fins in the triplex-tube heat exchanger. Int. J. Heat Mass Transf. 2018, 124, 663–676. [Google Scholar] [CrossRef]
  19. Kazemi, M.; Hosseini, M.; Ranjbar, A.; Bahrampoury, R. Improvement of longitudinal fins configuration in latent heat storage systems. Renew. Energy 2018, 116, 447–457. [Google Scholar] [CrossRef]
  20. Deng, S.; Nie, C.; Wei, G.; Ye, W.-B. Improving the melting performance of a horizontal shell-tube latent-heat thermal energy storage unit using local enhanced finned tube. Energy Build. 2019, 183, 161–173. [Google Scholar] [CrossRef]
  21. Rathod, M.K.; Banerjee, J. Thermal performance enhancement of shell and tube Latent Heat Storage Unit using longitudinal fins. Appl. Therm. Eng. 2015, 75, 1084–1092. [Google Scholar] [CrossRef]
  22. Mahdi, J.M.; Nsofor, E.C. Solidification enhancement of PCM in a triplex-tube thermal energy storage system with nanoparticles and fins. Appl. Energy 2018, 211, 975–986. [Google Scholar] [CrossRef]
  23. Mahdi, J.M.; Lohrasbi, S.; Ganji, D.D.; Nsofor, E.C. Simultaneous energy storage and recovery in the triplex-tube heat exchanger with PCM, copper fins and Al2O3 nanoparticles. Energy Convers. Manag. 2019, 180, 949–961. [Google Scholar] [CrossRef]
  24. Tiari, S.; Qiu, S.; Mahdavi, M. Numerical study of finned heat pipe-assisted thermal energy storage system with high temperature phase change material. Energy Convers. Manag. 2015, 89, 833–842. [Google Scholar] [CrossRef]
  25. Nakhchi, M.E.; Hatami, M.; Rahmati, M. A numerical study on the effects of nanoparticles and stair fins on performance improvement of phase change thermal energy storages. Energy 2021, 215, 119112. [Google Scholar] [CrossRef]
  26. u Mekrisuh, K.; Giri, S.; Udayraj; Singh, D.; Rakshit, D. Optimal design of the phase change material based thermal energy storage systems: Efficacy of fins and/or nanoparticles for performance enhancement. J. Energy Storage 2021, 33, 102126. [Google Scholar] [CrossRef]
  27. Mahdi, J.M.; Mohammed, H.I.; Hashim, E.T.; Talebizadehsardari, P.; Nsofor, E.C. Solidification enhancement with multiple PCMs, cascaded metal foam and nanoparticles in the shell-and-tube energy storage system. Appl. Energy 2020, 257, 113993. [Google Scholar] [CrossRef]
  28. Liang, L.; Diao, Y.H.; Zhao, Y.H.; Wang, Z.Y.; Bai, F.W. Numerical and experimental investigations of latent thermal energy storage device based on a flat micro-heat pipe array–metal foam composite structure. Renew. Energy 2020, 161, 1195–1208. [Google Scholar] [CrossRef]
  29. Ren, Q.; Xu, H.; Luo, Z. PCM charging process accelerated with combination of optimized triangle fins and nanoparticles. Int. J. Therm. Sci. 2019, 140, 466–479. [Google Scholar] [CrossRef]
  30. Li, Z.-R.; Hu, N.; Liu, J.; Zhang, R.-H.; Fan, L.-W. Revisiting melting heat transfer of nano-enhanced phase change materials (NePCM) in differentially-heated rectangular cavities using thermochromic liquid crystal (TLC) thermography. Int. J. Heat Mass Transf. 2020, 159, 120119. [Google Scholar] [CrossRef]
  31. Rabienataj Darzi, A.A.; Jourabian, M.; Farhadi, M. Melting and solidification of PCM enhanced by radial conductive fins and nanoparticles in cylindrical annulus. Energy Convers. Manag. 2016, 118, 253–263. [Google Scholar] [CrossRef]
  32. Al-Jethelah, M.; Ebadi, S.; Venkateshwar, K.; Tasnim, S.H.; Mahmud, S.; Dutta, A. Charging nanoparticle enhanced bio-based PCM in open cell metallic foams: An experimental investigation. Appl. Therm. Eng. 2019, 148, 1029–1042. [Google Scholar] [CrossRef]
  33. Choi, S.-K.; Kim, S.-O.; Lee, T.-H.; Dohee-Hahn. Computation of the natural convection of nanofluid in a square cavity with homogeneous and nonhomogeneous models. Numer. Heat Transf. Part A Appl. 2014, 65, 287–301. [Google Scholar] [CrossRef]
  34. Bondareva, N.S.; Buonomo, B.; Manca, O.; Sheremet, M.A. Heat transfer inside cooling system based on phase change material with alumina nanoparticles. Appl. Therm. Eng. 2018, 144, 972–981. [Google Scholar] [CrossRef]
  35. Turkyilmazoglu, M.; Pop, I. Heat and mass transfer of unsteady natural convection flow of some nanofluids past a vertical infinite flat plate with radiation effect. Int. J. Heat Mass Transf. 2013, 59, 167–171. [Google Scholar] [CrossRef]
  36. Öğüt, E.B. Natural convection of water-based nanofluids in an inclined enclosure with a heat source. Int. J. Therm. Sci. 2009, 48, 2063–2073. [Google Scholar] [CrossRef]
  37. Schenk, O.; Gärtner, K. Solving unsymmetric sparse systems of linear equations with PARDISO. Future Gener. Comput. Syst. 2004, 20, 475–487. [Google Scholar] [CrossRef]
  38. Bollhöfer, M.; Eftekhari, A.; Scheidegger, S.; Schenk, O. Large-scale sparse inverse covariance matrix estimation. SIAM J. Sci. Comput. 2019, 41, A380–A401. [Google Scholar] [CrossRef]
  39. Girault, V.; Raviart, P.-A. Finite Element Methods for Navier-Stokes Equations: Theory and Algorithms; Springer Science & Business Media: Berlin, Germany, 2012; Volume 5. [Google Scholar]
  40. Thomasset, F. Implementation of Finite Element Methods for Navier-Stokes Equations; Springer Science & Business Media: Berlin, Germany, 2012. [Google Scholar]
  41. Costa, V. Natural convection in partially divided square enclosures: Effects of thermal boundary conditions and thermal conductivity of the partitions. Int. J. Heat Mass Transf. 2012, 55, 7812–7822. [Google Scholar] [CrossRef]
  42. Kumar, L.; Manjunath, B.; Patel, R.; Markandeya, S.; Agrawal, R.; Agrawal, A.; Kashyap, Y.; Sarkar, P.; Sinha, A.; Iyer, K. Experimental investigations on melting of lead in a cuboid with constant heat flux boundary condition using thermal neutron radiography. Int. J. Therm. Sci. 2012, 61, 15–27. [Google Scholar] [CrossRef]
  43. Leong, K.Y.; Rahman, M.R.A.; Gurunathan, B.A. Nano-enhanced phase change materials: A review of thermo-physical properties, applications and challenges. J. Energy Storage 2019, 21, 18–31. [Google Scholar] [CrossRef]
  44. Nižetić, S.; Jurčević, M.; Arıcı, M.; Arasu, A.V.; Xie, G. Nano-enhanced phase change materials and fluids in energy applications: A review. Renew. Sustain. Energy Rev. 2020, 129, 109931. [Google Scholar] [CrossRef]
Figure 1. Latent heat thermal energy storage (LHTES) unit: (a) ascending conical; (b) cylinder; (c) descending conical.
Figure 1. Latent heat thermal energy storage (LHTES) unit: (a) ascending conical; (b) cylinder; (c) descending conical.
Sustainability 13 02667 g001
Figure 2. Structural specifications of the LHTES; coordinate of Pi is R d c + a i 2 / 3 h s f 2 R i , i 2 / 3 h s f .
Figure 2. Structural specifications of the LHTES; coordinate of Pi is R d c + a i 2 / 3 h s f 2 R i , i 2 / 3 h s f .
Sustainability 13 02667 g002
Figure 3. The effect of grid size for descending conical for a design with VFna = 0.015, ARf = 1.0, ARc = 0.8, and GO nanoparticle. (a) Total energy stored; (b) melt volume fraction (MVF); (c) temperature uniformity; (d) temperature at P4.
Figure 3. The effect of grid size for descending conical for a design with VFna = 0.015, ARf = 1.0, ARc = 0.8, and GO nanoparticle. (a) Total energy stored; (b) melt volume fraction (MVF); (c) temperature uniformity; (d) temperature at P4.
Sustainability 13 02667 g003
Figure 4. Temperature distributions of the work conducted by Costa [41] and the current isotherms.
Figure 4. Temperature distributions of the work conducted by Costa [41] and the current isotherms.
Sustainability 13 02667 g004
Figure 5. Melting fields of the present work and study conducted by [42].
Figure 5. Melting fields of the present work and study conducted by [42].
Sustainability 13 02667 g005
Figure 6. The S/N ratios for all the levels of the design parameters for ascending conical shell.
Figure 6. The S/N ratios for all the levels of the design parameters for ascending conical shell.
Sustainability 13 02667 g006
Figure 7. The S/N ratios for all the levels of the design parameters for descending conical shell.
Figure 7. The S/N ratios for all the levels of the design parameters for descending conical shell.
Sustainability 13 02667 g007
Figure 8. Melting maps and streamlines of ascending conical shell. The purple region is solid nano-enhanced phase change material (NePCM) and the blue region is molten NePCM at various t (s).
Figure 8. Melting maps and streamlines of ascending conical shell. The purple region is solid nano-enhanced phase change material (NePCM) and the blue region is molten NePCM at various t (s).
Sustainability 13 02667 g008
Figure 9. Isotherm maps of an ascending conical shell. From left to right hand at various t (s).
Figure 9. Isotherm maps of an ascending conical shell. From left to right hand at various t (s).
Sustainability 13 02667 g009
Figure 10. Melting maps and streamlines of descending conical shell. The purple region is solid NePCM and the blue region is molten NePCM. From left to right hand at various t (s).
Figure 10. Melting maps and streamlines of descending conical shell. The purple region is solid NePCM and the blue region is molten NePCM. From left to right hand at various t (s).
Sustainability 13 02667 g010
Figure 11. Isotherm maps of descending conical shell. From left to right hand at various t (s).
Figure 11. Isotherm maps of descending conical shell. From left to right hand at various t (s).
Sustainability 13 02667 g011
Figure 12. The temperature variation of the monitor points for two cases of: (a) optimum ascending design; (b) optimum descending design.
Figure 12. The temperature variation of the monitor points for two cases of: (a) optimum ascending design; (b) optimum descending design.
Sustainability 13 02667 g012
Figure 13. The variation of characteristic parameters for the ascending and descending designs. (a) MVF; (b) total stores energy; (c) temperature uniformity.
Figure 13. The variation of characteristic parameters for the ascending and descending designs. (a) MVF; (b) total stores energy; (c) temperature uniformity.
Sustainability 13 02667 g013
Table 1. Thermophysical properties of the coconut oil and the nano-additives [32,33].
Table 1. Thermophysical properties of the coconut oil and the nano-additives [32,33].
PropertiesBio-Based PCM Water as HTFSolid Fin
LiquidSolid
Cp (JkgK−1)201037504178386
η (Nsm−2)326 × 10−4Non-defined705 × 10−6Non-defined
ρ (kgm−1)914920993.738900
k (Wm−1K−1)166 × 10−3228 × 10−3623 × 10−3401
Tmelting (K) Non-defined297Non-definedNon-defined
LPPCM (Jkg−1)103e3Non-definedNon-definedNon-defined
Table 2. Nanoparticles dispersed in the phase change material [34,35,36].
Table 2. Nanoparticles dispersed in the phase change material [34,35,36].
PropertiesGOAl2O3CuAg
ρ (kgm−1)18003600896010,500
Cp (JkgK−1)717765385235
k (Wm−1K−1)500036400429
β (K−1)28.4 × 10−57.8 × 10−616.7 × 10−618.9 × 10−6
Table 3. Details of mesh study and element sizes.
Table 3. Details of mesh study and element sizes.
CaseNEElement Number in Different Domains
HTFFinsPCM
10.75183255512,485
21.003080126021,078
31.255432154027,730
41.507472221438,935
51.7510,222348051,090
Table 4. The range and levels of design parameters for ascending conical shell.
Table 4. The range and levels of design parameters for ascending conical shell.
FactorsDescriptionLevel 1Level 2Level 3Level 4
AVFNa (volume fraction)00.0150.030.045
BARf (ratio of the radius of lower fin to the radius of the base fin)0.10.40.71.0
CARc (the ratio of the lower radius of conic to the radius of the regular cylinder)0.40.60.81
DNano (type of nanoparticle)Al2O3GOAgCu
Table 5. The range and levels of design parameters for descending conical shell.
Table 5. The range and levels of design parameters for descending conical shell.
FactorsDescriptionLevel 1Level 2Level 3Level 4
AVFNa (volume fraction)00.0150.030.045
BARf (ratio of the radius of lower fin to the radius of the base fin)11.21.41.6
CARc (the ratio of the lower radius of conic to the radius of the regular cylinder)11.171.331.5
DNano (type of nanoparticle)Al2O3GOAgCu
Table 6. The L16 Taguchi table for ascending conical shell geometry.
Table 6. The L16 Taguchi table for ascending conical shell geometry.
CaseABCDValue at 7 h
VFNaARfARcNanoMVFES (kJ)S/N Ratio
10.0000.10.410.77149.4−2.27019
20.0000.40.620.83160.3−1.61844
30.0000.70.830.88170.0−1.11035
40.0001.01.040.90165.0−0.91515
50.0150.10.630.85161.6−1.41162
60.0150.40.440.77149.6−2.27019
70.0150.71.010.92167.9−0.72424
80.0151.00.820.87168.3−1.20961
90.0300.10.840.91161.9−0.81917
100.0300.41.030.91173.1−0.81917
110.0300.70.420.77146.2−2.27019
120.0301.00.610.80154.6−1.93820
130.0450.11.020.88166.3−1.11035
140.0450.40.810.95180.6−0.44553
150.0450.70.640.85162.8−1.41162
160.0451.00.430.75142.1−2.49877
Table 7. The L16 Taguchi table for descending conical shell geometry.
Table 7. The L16 Taguchi table for descending conical shell geometry.
CaseABCDValue at 7 h
VFNaARfARcNanoMVFES (kJ)S/N Ratio
10.0001.01.0010.90165.1−0.91515
20.0001.21.1720.90164.3−0.91515
30.0001.41.3330.86163.7−1.31003
40.0001.61.5040.79151.2−2.04746
50.0151.01.1730.90173.0−0.91515
60.0151.21.0040.92173.2−0.72424
70.0151.41.5010.78150.2−2.15811
80.0151.61.3320.88160.7−1.11035
90.0301.01.3340.84159.8−1.51441
100.0301.21.5030.78147.8−2.15811
110.0301.41.0020.90172.6−0.91515
120.0301.61.1710.91164.7−0.81917
130.0451.01.5020.76142.7−2.38373
140.0451.21.3310.88165.4−1.11035
150.0451.41.1740.96176.8−0.35458
160.0451.61.0030.87166.1−1.20961
Table 8. The optimum values of the design parameters and MVF.
Table 8. The optimum values of the design parameters and MVF.
TypeOptimum FactorsOptimum MVF at 7 h
VFnaARfARcNanoTaguchi PredictionNumerical Simulations
Ascending0.0450.41.0Al2O30.930.92
Descending0.0151.41.17Cu0.950.92
Table 9. The coordinate of temperature monitor points in optimum ascending and descending designs.
Table 9. The coordinate of temperature monitor points in optimum ascending and descending designs.
PointAscending DesignDescending Designs
r (mm)z (mm)r (mm)z (mm)
1408488
240334633
340584358
440834183
54010839108
64013337133
74015835158
84018332183
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ghalambaz, M.; Mehryan, S.A.M.; Mahdavi, M.; Younis, O.; Alim, M.A. Evaluation of the Melting Performance in a Conical Latent Heat Thermal Unit Having Variable Length Fins. Sustainability 2021, 13, 2667. https://doi.org/10.3390/su13052667

AMA Style

Ghalambaz M, Mehryan SAM, Mahdavi M, Younis O, Alim MA. Evaluation of the Melting Performance in a Conical Latent Heat Thermal Unit Having Variable Length Fins. Sustainability. 2021; 13(5):2667. https://doi.org/10.3390/su13052667

Chicago/Turabian Style

Ghalambaz, Mohammad, S.A.M. Mehryan, Mahboobeh Mahdavi, Obai Younis, and Mohammad A. Alim. 2021. "Evaluation of the Melting Performance in a Conical Latent Heat Thermal Unit Having Variable Length Fins" Sustainability 13, no. 5: 2667. https://doi.org/10.3390/su13052667

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