1. Introduction
The shortage and intermittency of renewable energy sources and the pollution generated by the use of fossil fuels drives the search for energy process optimization methods. In this sense, the storage and release of thermal energy make a relevant contribution to the continuity and efficiency of thermodynamic processes. Within this scope, phase change materials (PCMs) have the characteristic of absorbing or releasing large amounts of energy in certain periods of time and under controlled operating conditions. These materials store this amount of thermal energy as latent heat during phase change processes. PCMs can store 5 to 14 times more energy per unit volume than other materials that store thermal energy by sensible heat, and they have a small ranges of phase change temperatures during the transformation of matter, allowing the storage process to be controlled within the loading and unloading cycles of these materials [
1]. However, the low thermal conductivity of PCMs limits their applicability, requiring the development of thermal flux enhancement systems in which PCMs interact with materials of a high thermal conductivity to intensify the heat flux [
2]. In this sense, several research studies have been carried out in order to develop methods to increase the heat flux to or from PCMs during thermal cycles [
3,
4,
5,
6,
7,
8].
Therefore, extended surfaces are applied to increase the interaction area between the PCM and the associated charge or discharge source. Beginning from the PCM container walls, thin structures are extended to increase the heat exchange area, intensifying the storage system’s heat transfer. Several research studies have been carried out to determine the influence of the fin geometry and type of material used to intensify the PCM thermal cycles [
9,
10,
11,
12].
Triplex tube heat exchangers (TTHXs) have been frequently applied in industry processes [
13] and are built with three concentric tubes (internal, intermediate and external). The volumes formed between these tubes are used to store the PCM and to drain the thermal exchange fluids that absorb or release energy into the material, causing it to melt or solidify during the charging and discharging processes [
10,
14,
15]. Fins made of materials with a high thermal conductivity are applied on the surfaces of these tubes so that the thermal interaction with the confined PCM intensifies the heat exchange process. Al-Abidi et al. [
15,
16] applied a TTHX in an energy storage system to thermally charge a PCM (paraffin), using a hot water flow from a storage tank connected to a solar collector.
To physically describe the thermal energy storage process via latent heat, it is necessary to develop a mathematical formulation that characterizes the transient phenomenon of phase change in which a phase transition zone is governed by a partial differential equation that can be solved analytically or numerically. The analytical solution of the governing equations for this process is problematic because of the nonlinear interfaces at the phase change boundary, which present a complex geometry and boundary condition; the few analytical studies available in this area are about one-dimensional cases with regular geometries and well-known boundary conditions [
9].
The conservation of energy, which governs the heat transfer phenomena in PCMs during the solid–liquid transition, can be solved based on temperature or enthalpy. In the first case, the temperature is treated as the main variable, the conservation equations are written separately for the solid and liquid phases and the interfaces between these phases can be traced to obtain an accurate solution. In the second case, the positions of the solid–liquid interfaces are not tracked, and the governing equations are treated as single-phase equations, with no explicit conditions at the solid–liquid interface; this enthalpy-based formulation involves the solution inside a transition mushy zone (MZ) in which both phases are present [
17].
In this sense, several research studies in the area of CFD have made efforts to characterize latent heat thermal energy storage systems (LHTESs), define optimal operating conditions and direct the applications of PCMs in the most varied applications [
9,
10,
14,
18,
19,
20,
21].
Mastani Joybari et al. [
14] realized a numerical analysis in order to evaluate the performance of a TTHX when it was subjected to simultaneous charge and discharge. The authors analyzed the continuous operating condition of the LHTES in which the PCM (Rubitherm
® RT31) was subjected to a melting process in part of its volume and a solidification process in elsewhere. This type of LHTES is a denominated simultaneous charging and discharging (SCD) system. The authors used Ansys FLUENT
® software, version 16.2, to numerically solve the governing equations, and the numerical results obtained were in accordance with the experimental results that consider the charging and discharging separately. The authors concluded that the initial condition of the PCM (completely in a liquid state or completely in a solid state) has a great impact on the final solid–liquid interface. They also observed that the only diffusive model presented low levels of error for the initial condition when considering a PCM in a completely liquid state, but neglecting natural convection in the case of a completely solidified initial condition produced relevant errors in the process description.
Youssef et al. [
21], who carried out an extensive numerical analysis of a PCM heat exchanger (HX), developed another example that is presented in the literature. This PCM HX was built with eight tubes and presents the external surfaces linked to spiral copper wires. Inside a metal container, these wires interact with an organic paraffin, PlusICE
® A16 (melts at 16 °C). The eight connected tubes form a serpentine structure which is used by the heat exchange fluid (air) to heat and cool the PCM. The conservation equations of the 3D model proposed to evaluate the performance of the PCM heat exchanger were solved using Ansys FLUENT
® software, and the numerical results obtained were in accordance with the experimental results. Moreover, the authors observed another important result: the PCM discharge time presented higher values than the PCM charge time. This heat transfer phenomenon is explained by the effect of natural convection, which is more effective during the melting process. The authors also found that the material charging and discharging times were inversely proportional to the cooling heat transfer fluid (CHTF) inlet velocity and the temperature difference between the fluid and the PCM.
Alkaabi et al. [
22] presented a numerical study to obtain the detailed heat transfer rate and pressure drop magnitude variations associated with changes in the demand for electrical energy that is produced through the operation of a nuclear reactor. The author proposed the optimization of the thermal energy generation system through coupling with an LHTES that releases or receives heat from the heat transfer fluid circulating in the nuclear reactor in accordance with the electric energy demand variation. This PCM-based LHTES works as a simple heat exchanger, receiving energy during periods of constant reactor operation, and operates in a secondary process as a supplier of thermal energy to the reactor heat transfer fluid to overcome the electrical power demand variations.
Mohamed et al. [
23] analyzed the coupled system of nuclear thermal energy generation with an LHTES in the Rankine and Brayton cycles of supercritical carbon dioxide (SCO
2). The authors presented results that indicated approximately 50% of the exergy was lost during the operation of the nuclear power plant without coupling, while less than 10% of the energy was lost when the system was coupled with the LHTES. They also noted that the advanced Brayton cycle (SCO
2) is more efficient than the Rankine cycle by up to 50%, depending on the effectiveness of the cycle components.
Mohamed et al. [
24] carried out a numerical investigation to evaluate the integration of a triplex-tube-type LHTES into a nuclear power plant system. In this research, the TTHX containers were vertically oriented in order to using the buoyancy forces developed during the PCM phase change to enhance the material melting/solidification cycles. The authors analyzed separate and simultaneous charging and discharging modes, evaluating the capacity of the LHTES to follow the nuclear power generation plant’s energic demand variations and avoid significant impacts on the operation of the reactor. For this purpose, two different heat transfer fluids were used for circulation in the TTHX: one for extreme charge/discharge conditions and another for the normal operating conditions of the nuclear power plant. The results showed that the LHTES with adaptive charging and discharging modes allowed the PCM to change from the solid to the liquid phase and vice versa in accordance with the plant’s energy demand variations.
Mohamed et al. [
25] numerically analyzed an LHTES based on vertically positioned axisymmetric triplex tube heat exchangers in which a PCM was stored that was simultaneously charged and discharged (SCD) through a thermal exchange with a hot heat transfer fluid (HHTF) and a CHTF that circulated simultaneously within the LHTES. This charging came from the interaction of the LHTES with a thermal nuclear power generation system that provided a constant load. With this load, the authors presented results in which an equilibrium between the melting and solidification (charge and discharge) of the PCM was established in a liquid fraction of 20%. Under conditions of variable loads that occurred as a function of the energy demand, the LHTES system was able to follow the variation through the variation in the PCM’s liquid fraction, filling the gap between the energy supplied by the reactor and the power demand and allowing for the constant operation of the reactor at its full rated capacity.
Therefore, the CFD technique is crucial for evaluating and improving thermal energy storage systems via latent heat. The number of works that analyze the LHTES in 3D geometries is small, given the complexity of the model and the generation of meshes for the heat exchangers. The works have shown that the models which consider the natural convection in the PCM along the phase change process present more accurate results than those that consider the process to be purely diffusive. As the numerical data are validated, they can be used in the development and improvement of phase change heat exchangers.
In this sense, unlike the numerical works that analyze the PCM solidification process with simplified, two-dimensional modeling, the present work presents a 3D numerical analysis that considers the entire LHTES, using a triplex tube heat exchanger with fins on the tubular surfaces during its discharging. The model considers the flows of heat transfer fluids, which exchange heat by forced convection through the walls of the pipes, and the conduction heat transfer through the finned pipes and the phase change process, considering the natural convection, buoyancy effects and thermal conduction. Thus, the three-dimensional analysis allows for a numerical analysis of all energetic and fluid dynamic phenomena occurring during the operation of the LHTES in all physical problem geometry.
2. Materials and Methods
The thermal energy release system analyzed in this work consisted of a triplex tube heat exchanger (TTHX) constructed with copper tubes and longitudinal fins installed on the walls in contact with the PCM, Rubitherm
® RT-82, which was completely in a liquid state after having stored thermal energy during periods of high availability (
Figure 1). The cooling heat transfer fluid (CHTF-Water) flowed in the pipes with a temperature lower than the PCM so that the energy contained in the material was released to the CHTF through the material’s solidification process. To physically describe this system, the TTXH was subdivided into five physical domains: the fluid volume (water), the solid material of the copper tubes (internal, intermediate and external) and the PCM, located between the intermediate tube and the inner tube.
During the energy release process, the water, which has a prescribed temperature (T
i) and mass flow rate (
), enters the exchanger through a copper tube with a diameter of 50.8 mm and a thickness of 1.2 mm. This tube is branched by another tube with a diameter of 32 mm and a thickness of 1.2 mm. In this way, the thermal energy is extracted from the PCM by the finned pipes which, through a process of thermal conduction, transfer this energy to the outer surfaces of the intermediate tube and the inner surface of the internal tube. These are in contact with the CHTF, which receives the energy through the thermal convection process. As the CHTF receives energy from the system, its outlet temperature (T
s) is increased, and the PCM solidifies over time. Furthermore, the external surfaces of the TTXH are considered isolated, causing the heat exchange with the external environment to be negligible. In
Figure 2,
Figure 3 and
Figure 4, the fluid domains, the inner and intermediate tubes and the PCM are illustrated with their respective geometric measurements.
Due to the small thickness of the pipe and the isolation considered on the external surfaces of the TTHX, the solid material referring to the external pipe was neglected in the model for simplification purposes. Thus, as illustrated in
Figure 5, only the CHTF, PCM, internal piping and intermediate piping were considered. Furthermore, the variation of 16 mm between the pipe length and the fins was also neglected, leaving only 480 mm of the TTHX length, which corresponded to the fins, considered in the simulation. These simplifications promoted the production of mesh of a reasonable quality.
Thus, the thermal exchange area (the surfaces of the finned pipes in contact with the PCM) was 308,832 mm
2, and the PCM volume was 0.00725707 m
3. Converting these values to mass as a function of the PCM density, there was 6.89 kg of phase change material stored in the TTHX.
Figure 5 also indicates the five planes (z = 0, 100, 240, 380 and 480 mm) in the z direction which were used to analyze the results obtained.
2.1. Mathematical Modeling
The CHTF flow in TTHX pipes occurs in turbulent, laminar and transition flow regimes, according to the considered fluid domain. As the PCM’s temperature changes, it presents density variations which first occur in the regions closest to the pipe walls. These variations, in the midst of the gravitational field, are expressed as buoyancy forces, which promote low velocity variations in the PCM particles. Therefore, the k-ω SST (shear stress transport) turbulence model, with corrections for low Reynolds numbers (Re), was used to analyze the flow in all fluid domains.
Therefore, the following equations were used:
- (b)
Linear momentum conservation
In these equations, ρ is the density of the fluid, μ is the dynamic viscosity, t is the time, P is the pressure, X is the position vector, u is the velocity vector, the sub-indices i and j represent the components (x, y and z) of the coordinate axes and the term represents the Reynolds stresses, derived from turbulent flow.
- (c)
Turbulence model: shear stress transport, k-ω SST
The stress transport model (shear stress transport, k-ω SST) applied in the Ansys FLUENT
® software was developed by Menter [
26]. The objective of the model is to unite the robustness and accuracy of the standard k-ω model for the results close to the wall [
27] with the accuracy and simplicity of the k-ε model [
28] in regions far from the wall. For this purpose, coupling functions are used so that the equations of the k-ω and k-ε models are activated in the cells near and far from the walls, respectively, in the computational domain. In short, the variable k represents the turbulent kinetic energy of a flow, and the variable ω represents the dissipation rate of this energy. Equations 3 and 4 describe the transport of these variables:
In these equations, k is the term referring to the turbulent kinetic energy, and ω represents the dissipation of this energy; Gk and Gω represent the generation of k and ω, and Γk and Γω represent the effective diffusivity; Yk and Yω represent the dissipation of k and ω; the term Dω represents cross-diffusion; and Wk and Wω represent the source terms of the referred equations.
- (d)
Energy conservation
To model the heat transfer between the thermal exchange fluid and the walls of the tubes and between the PCM in the liquid state and the walls of the tubes, the energy model used by Ansys FLUENT
® was applied, which deals with the conservation of energy as follows.
- (e)
Phase change model
To solve the transient problem of PCM solidification, the enthalpy–porosity model developed by Voller and Prakash [
29] was used. In this model, the calculated liquid fraction is what indicates the interface position along the phase transition. PCM heat transfer occurs by thermal diffusion by natural convection, which is a function of density variations, temperature and the PCM liquid fraction.
where:
- (f)
Fluid Properties: Density
In these equations, T represents the temperature, Wh represents the energy source term, E represents the total energy and γeff represents the effective thermal conductivity. The term ψ represents the liquid fraction; H represents the total enthalpy, the sum of the sensible enthalpy h and the latent enthalpy variation ∆H; L represents the latent heat of fusion; TS represents the solidification temperature; Tl represents the melting temperature; ρS and ρl are the material densities in the solid and liquid state, respectively, and η is the thermal expansion coefficient.
This mathematical modeling is another innovation of this research and was crucial to carrying out the LHTES simulations. For T < T
S, the density was considered constant (ρ = ρ
S); for T
l ≥ T ≥ T
S, the density was modeled as a mixing model that is a function of the liquid fraction calculated in the enthalpy model; for T > T
l, the Boussinesq natural convection model [
30], which is specific to buoyant flows in which the density variation is driven only by small temperature variations, was applied. Finally, the values obtained with Equation (10) and applied in the numerical simulations were ρ
l = 778.47 kg/m
3 and ρ
S = 950 kg/m
3.
2.2. Boundary Conditions
The mass flow rate at the CHTF inlet was defined as an inlet condition. In addition, the absolute reference system, the flow direction normal to the inlet surface, the turbulence intensity, I = 5%, and the viscosity ratio turbulent, Rμ = 10, were established. The output boundary condition defined for the CHTF was the outflow, which corresponded to a neglected diffusive flow at the output. Therefore, the outflow boundary conditions were extrapolated from the internal domain, and these conditions had no influence on the upstream flow. Furthermore, this boundary condition imposed a general mass balance correction for all flow variables. Wall boundary conditions (non-slip and negligible roughness) were used for the fluid–solid interfaces and the insolated external surfaces.
For the heat transfer, (a) a wall with non-heat flux at the CHTF domain external surfaces and (b) a wall coupled condition for the solid–liquid were considered. Thus, the heat flux was calculated at the interfaces as a function of the neighboring cell temperatures.
The initial moment of the solidification process occurs with the final condition of the melting process, which was presented by Porto et al. [
31]. In this situation, the PCM was brought to complete melting by being heated by water at a temperature of 90 °C. Thus, the solidification process initially occurred with the CHTF at 68 °C and 8.3 L/min, expelling the hot fluid found in the TTHX at the end of the melting process. In all simulations, a time step of 0.5 s was used.
The boundary conditions used in the simulations are described in
Table 1.
2.3. Numerical Mesh
The finite volume discretization method was applied. Hybrid meshes were developed using the Ansys Meshing
® software. This mesh allowed for the association of the molding capacity of the complex geometries of the tetrahedral elements, allocated in the CHTF domain, with the good quality of the results related to the hexahedral elements at the wall regions. The finned structures and the PCM domain were constructed only with hexahedral elements, maintaining the mesh quality for these geometries.
Figure 6 illustrates the isometric and frontal views of the mesh used for the heat exchanger, with all domains mounted.
Figure 7 illustrates the mesh for the PCM, internal and intermediate finned pipes: structured hexahedral elements in the pipes and unstructured hexahedral elements in the PCM volume.
The results were compared for two meshes, one with 1,155,528 elements and the other with 2,902,682 elements, and did not present significant variations. Therefore, the mesh with 1,155,528 elements was chosen. It presented a minimum and average orthogonality of 0.179 and 0.891, respectively, and a maximum and average deformation of 0.849 and 0.220, respectively. The influence of the time step was also verified by comparing the results obtained with ∆t = 0.1, 0.5 and 1.0. No significant variations between the results obtained for the three time steps were verified. Therefore, the 0.5 s time step was chosen. The results of numerical mesh and time step analyses were presented by Porto et al. [
31].
2.4. Physical Properties
The thermo-physical properties used in this research were experimentally obtained by Al-Abidi et al. [
32]. In this experimental work, the PCM density in the liquid state was obtained at 93 °C, and the PCM density in the solid state was obtained at 27 °C. In addition, the PCM latent heat of fusion (L = 201,643.8 J/kg) was also experimentally obtained for the phase change temperature range in which the solid temperature (TS) was 343.2775 K and liquid temperature (Tl) was 355.3263 K. These data are shown in
Table 2.
4. Conclusions
The present work proposed a numerical analysis able to predict the behavior of a PCM inserted into a triplex tube heat exchanger with finned walls, working as a latent heat thermal energy storage system, during the discharge process (solidification). The novelty of this research consists of the three-dimensional analysis of the solidification process, which proved to be efficient for describing the behavior of the entire LHTES and presented variations in the results along the volume that more simplified models, which consider constant boundary conditions along the material length, cannot determine. The numerical results were validated through a comparison with the experimental data of the mean temperature in the section at z = 100 mm from the entrance, which showed a maximum error of 5.02 °C (7.17%) and a mean error of 3.09 °C (4.2%).
When the solidification process begins, the TTHX is completely filled by the cold heat exchange fluid (CHTF), thus completely replacing, from its volume, the hot heat exchange fluid (HHTF) from the melting process of the material. This process causes the solidification of the PCM to occur in the first 512 s through an interaction with a mixture of the CHTF and HHTF. From that moment, the cooling starts to occur only through the interaction with the CHTF.
During the operation of the energy release system, it was verified that the CHTF heats up along the length of the heat exchanger and that the discharge potentials are also reduced in this direction, causing the sections most downstream from z = 0 mm to be the last to solidify. It was also verified that the cooling increased the PCM fraction densities and as the regions closest to the annular surface were cooled, the material fractions with a higher density were deposited in the lower parts of the lower and upper halves of the annular. However, the geometry of the fins retained some of the heavier fractions of PCM in the upper part of the annular throughout the solidification process. Finally, close to the state of complete solidification, the velocities developed with the buoyancy forces were reduced to negligible values. This occurred because all the fractions along the volume of the material presented very similar densities.
Quantitatively, it was possible to describe the heat exchange process along the PCM length and its corresponding influence on the heat exchange fluid coming out of the TTXH. While the PCM releases heat to the thermal heat exchange fluid and develops the solidification process, the CHTF begins to leave the heat exchanger more heated, reaching a temperature of 90 °C at the beginning of the process due to the presence of the mixture with hot fluid in the exchanger at the end of the melting process. It then drops to 69 °C within the first 10 min of unloading. The material is then slowly discharged until it reaches thermal equilibrium with the CHTF.
Regarding the heat fluxes, it was found that the temperature difference between the PCM (87 °C) and the CHTF (68 °C) promoted heat fluxes from the PCM to the CHTF of −6423 W/m
2 towards the inner surface of −742 W/m
2 for the external surface of the PCM annular. This deference shows that the fluid dynamic conditions developed on the LHTES promote different heat removal process for the annular surfaces, indicating that the TTHX modeling must be carried out using three-dimensional mesh to capture the boundary conditions that are modified as a function of the CHTF flow regime. Comparing the solidification heat flux with the melting heat flux presented by Porto et al. [
32] during the melting process, it was verified that the solidification heat fluxes were five times smaller due to the small difference between the CHTF temperature (68 °C) and the PCM solid temperature (70 °C), while the HHTF temperature was 90 °C and the PCM liquid temperature was 82 °C during the melting process.
The latent energy accumulated in the PCM is completely released to the CHTF during the solidification process, making the energy accumulated in the phase change material after complete solidification approximately 82.52 kJ/kg relative to the sensible heat of the PCM after complete solidification.
In this way, depending on the TTHX geometry, the three-dimensional numerical model must be used to describe the boundary conditions that are modified as a function of the CHTF flow regime. Symmetry simplifications using 2D numerical models and fixed boundary conditions for the external and internal surfaces of the PCM annular do not describe the fluid dynamic variations that promote the different boundary conditions during the solidification process.
The model used in this research did not consider some effects, such as the volumetric contraction due to temperature variation and the creation of air spaces at the exchange surfaces, which create resistant cavities at the heat exchange. These effects occur during the solidification and are verified in experimental works. Numerical modeling aims to describe, as reliably as possible, the real physical phenomenon for validation with experimental data. In this sense, in future research, these phenomena can be evaluated in order to continuously improve numerical models and reduce errors in relation to experimental data.