1. Introduction
Energy storage is crucial for balancing energy demand and intermittent renewable generation. The increase in renewable energy deployment is driven by efforts to reduce greenhouse gas emissions and enhance energy security. According to the International Energy Agency (IEA), renewable sources generated 29% of global electricity in 2020 [
1,
2]. This percentage is expected to increase to 49% by 2030 [
3,
4]. To address the intermittency of renewable energy technologies, such as solar and wind power, Thermal Energy Storage systems (TESs) will be essential [
4]. TESs also have applications in waste heat recovery.
The storage of thermal energy can be classified into three distinct methods, each tailored for specific applications.
Sensible heat storage (SHS) is a widely used technique which stores and releases thermal energy by varying the temperature of a medium without changing its phase. In this method, temperature changes occur linearly with the amount of heat stored, depending on the specific heat capacity of the material. Traditional SHS systems commonly utilise water storage tanks. Latent heat storage (LHS) involves the energy absorbed or released during a phase transition. This method is generally applied in low-temperature settings, using phase-change materials selected based on the temperature requirements. Thermochemical energy storage employs reversible chemical reactions to facilitate the storage and release of thermal energy. This approach is characterised by its high energy density and versatility across various contexts [
2].
Packed beds have applications in a range of thermal systems, such as the storage of renewable solar energy, industrial waste heat recovery, and geothermal energy [
5] and can be more compact and cost-effective and compact compared to alternative storage systems. In addition, depending on the storage medium and heat transfer fluid (HTF) used, packed beds often utilise more abundant, locally sourced, environmentally friendly, and non-reactive materials than other storage technologies [
6].
As illustrated in
Figure 1, the PBSS system has three main components: a storage vessel (container), storage media (filler), and heat transfer fluid (HTF). The storage vessel is filled with solid packing material known as the “filler”. These solid particles receive energy through an HTF, which may also function as the working fluid within the system that incorporates the packed bed. Various packed-bed designs can be found in numerous publications [
7,
8,
9,
10], showcasing differences in shape, the materials used for storage, and the HTF employed.
Storage units are typically cylindrical, with the HTF flowing along the axis. These packed beds are classified as sensible heat storage (SHS) because energy is stored by changing the temperature of the filler material. An alternative approach could involve using latent heat storage systems with phase-change materials [
11], but this may lead to higher costs and increased system complexity [
12,
13]. HTFs can be gases, such as air, carbon dioxide, or argon [
14], or they can be liquids, like thermal oils [
15] or molten salts [
16].
Several applications of thermal energy storage have proposed the utilisation of packed-bed storage systems (PBSSs). These applications include Pumped Thermal Energy Storage (PTES) [
1,
2,
3,
4,
5], Advanced-Adiabatic Compressed-Air Energy Storage (AA-CAES) [
6,
7], and Liquid Air Energy Storage (LAES) [
4], specifically for large-scale systems.
Packed-bed thermal tanks are an essential component of the Pumped Thermal Energy Storage (PTES) system, which is an emerging technology for industrial applications [
17,
18,
19]. Since this technology is not yet commercially available and requires further research [
19,
20,
21], optimising the design of the packed bed—both hot and cold—is essential for enhancing the overall PTES system. This optimisation can also contribute valuably to the existing literature in this area.
Pumped Thermal Energy Storage (PTES) is a promising technology for large-scale energy storage. Compared to other thermal energy storage methods, PTES provides high round-trip efficiency (RTE), substantial capacity, a lifespan of up to 30 years, short response times [
2,
3,
4], and quick start-up times [
5,
6]. Additionally, PTESs are environmentally friendly and have a smaller carbon footprint compared to other large-scale energy storage technologies, such as Compressed Air Energy Storage (CAES) and Pumped Hydro Energy Storage Systems (PHES) [
6,
7]. Moreover, PTES can achieve reasonable round-trip efficiencies of up to 70% and competitive energy and power densities at affordable costs, as shown in
Table 1.
The PTES system consists of five main components: two storage tanks where the PTES stores the electrical energy as thermal energy, a heat pump, a heat engine, and a motor or generator [
2].
Figure 2 illustrates the traditional configuration of the PTES.
The charging cycle involves a heat pump that uses electrical energy to move heat from a low-temperature reservoir (cold tank) to a high-temperature reservoir (hot tank). Electricity is generated during discharge by a heat engine, using the thermal energy stored in the reservoirs. While leakage losses typically do not affect large systems, the storage tanks are designed with insulation to minimise heat loss during storage [
23].
It is important to note that sensible heat storage is the simplest and most cost-effective method for storing thermal energy. While there are some advantages to phase-change energy storage, due to these technological and economic factors, sensible heat storage tends to be more suitable and appealing for large-scale energy storage systems compared to phase-change energy storage [
22].
1.1. Literature Review and Research Gap
Desrues et al. [
24] investigated a system utilising the Brayton cycle with argon gas in HTF, aiming to develop PTESs for large-scale applications using MATLAB programming. An RTE of 66.7% was achieved; however, this required a maximum temperature of 1000 °C, which exceeds the current compressor’s operating temperature of approximately 900 K. Adding a second electrical heater can partially mitigate this limitation. Howes [
25] analysed the heat transfer mechanisms and losses associated with three PTES prototypes using mathematical modelling, where reciprocating devices replaced conventional compressor/turbine pairs. Furthermore, White et al. [
26] used the numerical analysis to explore how the geometry of storage tanks, operational modes, and temperature variations impact thermodynamic losses, noting that long-term storage results in considerable losses, while intermittent operation generates manageable losses.
Furthermore, numerous researchers have refined the system’s configuration through the thermodynamic numerical analyses of packed-bed effects, incorporating buffer vessels and characterising energy loss and generation mechanisms [
1,
3,
27,
28,
29]. Wang suggested a method for transient analysis, as outlined by Wang et al. [
30], who studied the influence of numerous factors and contrasted systems involving helium and argon. Helium demonstrated a higher RTE, as determined by the study, with energy storage performance increasing at isentropic conditions and higher pressure ratios. Charging was achieved using an electric heater in a plant developed by Benato et al. [
22,
31,
32]. Furthermore, a one-dimensional, two-phase packed-bed model was developed to evaluate the performance of nine heat storage materials with two heat transfer fluids.
Ge et al. [
17] utilised MATLAB to construct an analytical model of a 10.5 MW, 5 h packed-bed latent heat/cold storage system. The study conducted an energy and exergy analysis of various components. It also investigated whether packed-bed latent heat or cold storage systems could effectively replace packed-bed sensible heat or cold storage systems in pumped thermal electricity storage. The research systematically evaluated thermodynamic performance under optimal conditions by analysing the effects of porosity, compression ratios, inlet velocities, and isentropic efficiency. The findings revealed that replacing latent heat and cold storage with packed-bed sensible heat and cold storage could increase the energy storage density of the system from 232.5 kWh/m
3 to 245.4 kWh/m
3. Furthermore, this new system achieved a round-trip efficiency of 84.7% and a power density of 216.5 kW/m
3, demonstrating its competitiveness and efficiency compared to large-scale electrical energy storage systems.
A novel PTES system integrated with a CSP was presented and evaluated by Petrollese et al. [
19]. The system incorporated the same working fluid (argon) as the CSP, allowing for both the simultaneous and independent operation of its various components. The Thermal Energy Storage (TES) system is made up of three packed-bed tanks that utilise thermoclines. In this research, the functioning PTES-CSP plant that integrates TES tanks was simulated in MATLAB, employing particular mathematical models under standard conditions.
A research study was carried out to examine how the primary design parameters of TESs, such as operating temperatures and pressure ratios, influence the main performance metrics. It was determined that a pressure ratio of roughly 5:2 is ideal for maximising the exergetic round-trip efficiency of the integrated system. Consequently, an effective PTES-CSP system was established, achieving an exergetic round-trip efficiency of about 60% [
19].
Table 2 and
Table 3 show a summary of different packed-bed systems in the literature and in operations, respectively.
1.2. Research Novelty and Research Structure
In previous work, studies have used a single HTF and focused on other aspects of the design and operation of the packed-bed storage system.
Rabi et al. [
6] used air as the HTF and performed a study to determine optimal parameters for the storage system. This study suggests using Magnesia as the packed-bed particles with a diameter of 0.004 m and a porosity of 0.2 with a mass flow rate of 13.7 kg/m
3, along with an aspect ratio of 1. These values were used in the study presented here.
Alva et al. [
7] presented a review of storage systems for solar applications. Here, the focus, aligning with the majority of work in this area, focused on considering different storage materials for both SHS and LHS. It identified air as being a common HTF but did not consider comparisons between different fluids.
Meier et al. [
45] investigated rock bed storage systems using air as the HTF and focused on the properties and performance of the storage without considering alternative HTFs. Similarly, the work of White et al. [
1] focused on the storage system and only considered a single HTF: argon.
Other studies have also considered only a single HTF. For gas-based systems, this is typically air or argon, but there has been no consideration of the effect the HTF has on the operation and performance of the packed-bed storage system. Here, we present an analysis using six different HTFs, air, argon, carbon dioxide, helium, hydrogen, and nitrogen, to investigate and analyse how the different gases affect the performance and operation of the packed-bed system.
The methodology and details of the simulation model are set out in
Section 2. The results are presented and discussed in
Section 3. Conclusions drawn for this study and ideas for future work are then presented in
Section 4.
2. Model Set-up and Methodology
A cylindrical model with 2D axisymmetric has been developed for a fully insulated tank positioned with its axis vertical. This model uses a porous media formulation available in COMSOL 5.6.
COMSOL Multiphysics is a high-performance software package suitable for simulating a range of physical processes, including fluid flow through porous materials and heat transfer phenomena. Its strength lies in its ability to combine various physics models to simulate complex multiphysics systems. Considering the specific requirements and preferences of this simulation project, COMSOL has been chosen.
A predefined triangular mesh has been employed to discretise the geometry of the model. The nine available mesh options in COMSOL were evaluated [
6] for the physical model used in this study using air as the HTF. The temperature was recorded at three different heights within the packed bed at 0.5 m, 1.5 m, and 3.5 m following an 8 h charging cycle with a time step of 0.2 s. The results indicate that the three finest mesh options (referred to as Finer, Extra Fine and Extremely Fine) all produce comparable results with variation of no more than 3%. Here, we used the Extra Fine option.
The boundary conditions set included a no-slip wall where the velocity was zero at the solid surfaces (
. At the inlet, a condition of fully developed flow was established with a constant volume flow rate (
m
3/s), while a pressure condition was applied at the outlet to avoid backflow. Adiabatic boundary conditions were used at the outer boundary of the insulating material (
). The simulation examines incompressible fluid flow with operational pressures of 10.5 bar for the hot-storage reservoir and 1.05 bar for the cold-storage reservoir. The boundary conditions are shown in
Figure 3.
The classification of the flow regime in a packed bed is determined by the bed’s Reynolds number, which was computed using the size of the particles shown in Equation (1):
where Re denotes the Reynold number,
and
μ represents the density and viscosity and,
is the superficial velocity vector that the different regimes show in
Table 4.
In this study, Re = 352.4 and the flow is fully turbulent. This is simulated within COMSOL using the Brinkman equation with a turbulent viscosity based on the RANS k-ε (turbulent kinetic energy and dissipation) model [
6,
50].
This study employs a modified version of the one-dimensional Schumann equations [
51] to simulate packed beds. This approach is an extension of the Schumann model and relies on the following assumptions:
The interfacial thermal resistance at the particle surface limits heat transfer between the gas and the solid. The small particle size and the resultant low Biot numbers underlie this assumption.
The solid allows for convenient axial direction longitudinal conduction through the bed. Heat leakage to the surroundings also occurs from the solid.
The gas flow’s kinetic and potential energy terms are negligible.
A numerical solution to the heat transfer problem was obtained using COMSOL Multiphysics. The solution was derived using the local thermal non-equilibrium equation, Fourier’s law, and the non-Darcian flow model. Heat transfer in porous media is described using the local thermal non-equilibrium hypothesis.
The standard equations for momentum and mass conservation are presented in Equations (2) and (3) [
52] below:
CF is the parameter associated with Forchheimer. Moreover, the Schumann method independently addresses two coupled energy equations: one of the solids and the other of the heat transfer fluid (HTF). The principles of this theory can be explained using Equations (4) and (5) [
53].
where
represents the volume fraction of the solid; ε is the void fraction (material porosity);
indicates the density of the solid;
refers to the density of the fluid;
is the heat capacity of the solid at constant pressure;
is the heat capacity of the fluid at constant pressure;
qs denotes the conductive heat flux in the solid;
qf signifies the conductive heat flux in the fluid;
qsf represents the coefficient for interstitial convective heat transfer;
Qs stands for the heat source in the solid;
Qf is the heat source in the fluid;
refers to the contribution of thermoelastic damping that comes straight from the interfaces of the solid mechanics;
q is the conductive heat flux; and ∇ is the gradient operator.
COMSOL numerically addresses the heat transfer through Equations (6)–(8).
Following [
6], the heat transfer is included in Equations (9) and (11):
and Fourier conduction is simulated using Equations (12) and (13) [
6]:
where
is the porosity;
is the viscous dissipation in the fluid;
is the tensor that describes viscous stress;
S is the second Piola–Kirchhoff tensor; and
Ts and
Tf and
are the temperatures and thermal conductivities, respectively, for the of the solid and fluid, respectively. The pressure is uncoupled from the velocity, and the solution is that of the second order. The convergence criteria were set at 10
−4 for all variables.
This study analysed a hot and a cold cylindrical packed-bed storage reservoir. The numerical approach employed in this paper was previously utilised in reference [
6], where it was validated through the experimental research conducted by Meier et al. [
45]. The parameters used in [
6] are detailed in
Table 5, while the optimal parameters, which were determined, are shown in
Table 6. These were applied throughout this study, with Magnesia as the storage material.
Table 7 and
Table 8 [
33] show the thermodynamic properties of the considered gases at the average operational temperature for the hot and cold storage-packed beds, respectively.
The following sections provide a comprehensive overview of the considered optimisation criteria and comparative analyses of the selected heat transfer fluids (HTFs) during the charging and discharging cycles of the hot and cold storage packed bed.
Modelling Hot/Cold Thermal Energy Storage Simulations
The charging process for hot and cold storage involved the use of HTF based on the charging temperatures outlined in
Table 5; specifically, 476 °C for hot storage and −154 °C for cold storage. During this process, the HTF was circulated through the SHS tank. The HTF exchanged energy with Magnesia particles, which had a particle diameter of 4 mm and a void fraction of 0.2, while the tanks maintained an aspect ratio of 1.
Initially, the temperature of the Magnesia and HTF within the packed bed was set to 25 °C. The charging cycle was simulated until the majority of the Magnesia reached temperatures approaching the charging temperature of the HTF (476 °C and −154 °C for the hot and cold storage, respectively). This was then taken as the initial condition for the discharge process, where the flow direction of the HTF was reversed. Simulations were performed for each of the six selected HTFs to evaluate the performance of each gas during the charging and discharging processes, leading to a comparative study presented in the results and analysis section based on the established performance criteria.
The thermal performance of the packed bed during the charging and discharging processes was assessed using three criteria: the total energy stored
the capacity factor (CF), which represents the fraction of the total storage capacity that is filled [
6,
29],
and the thermal power [
6]
where the subscript
i corresponds to the initial temperature.
4. Conclusions and Future Work
Packed-bed systems have emerged as a prominent solution for large-scale thermal energy storage applications, owing to their impressive thermal efficiency and economic viability. The packed beds examined in this study employed optimised parameters identified in previous research using air as the HTF, involving a solid particle diameter of 0.004 m, a material porosity of 0.2, a mass flow rate of 13.7 kg/m3, and an aspect ratio of 1, using Magnesia as the storage medium.
This paper presents a comparative analysis to determine the influence of different heat transfer gases, namely air, Ar, CO2, He, H2, and N2, on the heat transfer efficiency of the studied hot/cold packed beds. The findings have been categorised into three groups: Argon/Helium, Nitrogen/Hydrogen/Air, and CO2. Comparative analysis of hot and cold storage systems indicated that CO2 emerged as the most efficient HTF among the studied gases, with the highest rate of charge and discharge. Analysis of the maximum operating range for both reservoirs, at 80% of their fully charged storage capacity, indicates that CO2 also provided the largest operating range compared to the others. The results showed that carbon dioxide (group 3) had an operating range of 72% for hot storage and 76% for cold storage. In comparison, group 2, which is hydrogen, nitrogen and air (group 2), had an operating range of 55% for hot storage and 75% for cold storage. Helium and argon (group 1) had an operating range of 50% for hot storage and 66% for cold storage. Additionally, the results indicate that carbon dioxide gas was the most promising heat transfer fluid for the packed beds studied.
The study also indicated that the maximum power level was inversely proportional to the ratio of specific heats, , and proportional to This suggests that when considering alternative HTFs, fluids with a low value and high may perform well.
The results also indicate that Ar was in the poorest performing group, suggesting that despite its existing use [
1,
22], it is not a good choice for HTF.
The following practical recommendations are made for the operation of backed bed storage tanks:
CO2 was identified as the preferred HTF for use in a system in terms of its performance.
Air performed less well but may be preferred on the basis of cost and availability.
When CO2 is used, the storage system should be used within the charging range of between 16% and 88% of full charge for the hot storage system and 11% and 87% for the cold tank. This ensures that the system is operating within 80% of its maximum power.
For alternative fluids, the corresponding ranges are given in
Table 9. For air, they are 24–79% and 13–87% for the hot and cold tanks, respectively.
A key limitation of this work is that the tank was assumed to be perfectly insulated, and heat losses to the environment were not considered. While this is appropriate when the storage is being constantly charged and discharged, it will not account for losses which occur, particularly when storage occurs over a longer time period. Future work should investigate the effect of these heat losses over both a short and a long storage period.