Next Article in Journal
Data Embedding in SHVC Video Using Threshold-Controlled Block Splitting
Next Article in Special Issue
Thermal and Mechanical Behavior of Elastomers Incorporated with Thermoregulating Microcapsules
Previous Article in Journal
Mechanical Properties of Green Synthesized Graphene Nano-Composite Samples
Previous Article in Special Issue
Exploitation of Enzymes for the Production of Biofuels: Electrochemical Determination of Kinetic Parameters of LPMOs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermal Energy Storage Performance of Tetrabutylammonium Acrylate Hydrate as Phase Change Materials

1
Department of Mechanical Engineering, Keio University, 3-14-1 Hiyoshi Kohoku-ku, Yokohama 223-8522, Japan
2
Cyber Kobo LLC., 2-201, Kitabukuro-cho, Omiya-ku, Saitama-shi 330-0835, Japan
3
Department of System Design Engineering, Keio University, 3-14-1 Hiyoshi Kohoku-ku, Yokohama 223-8522, Japan
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(11), 4848; https://doi.org/10.3390/app11114848
Submission received: 17 April 2021 / Revised: 24 May 2021 / Accepted: 24 May 2021 / Published: 25 May 2021
(This article belongs to the Special Issue Phase Change Materials: Design and Applications)

Abstract

:
Kinetic characteristics of thermal energy storage (TES) using tetrabutylammonium acrylate (TBAAc) hydrate were experimentally evaluated for practical use as PCMs. Mechanical agitation or ultrasonic vibration was added to detach the hydrate adhesion on the heat exchanger, which could be a thermal resistance. The effect of the external forces also was evaluated by changing their rotation rate and frequency. When the agitation rate was 600 rpm, the system achieved TES density of 140 MJ/m3 in 2.9 h. This value is comparable to the ideal performance of ice TES when its solid phase fraction is 45%. UA/V (U: thermal transfer coefficient, A: surface area of the heat exchange coil, V: volume of the TES medium) is known as an index of the ease of heat transfer in a heat exchanger. UA/V obtained in this study was comparable to that of other common heat exchangers, which means the equivalent performance would be available by setting the similar UA/V. In this study, we succeeded in obtaining practical data for heat storage by TBAAc hydrate. The data obtained in this study will be a great help for the practical application of hydrate heat storage in the future.

1. Introduction

The use of sustainable, renewable energy sources is becoming increasingly important as global attention on environmental issues increases. However, the amount of electricity generated by renewable energy sources such as wind and solar power fluctuates greatly depending on the natural environment. In recent years, energy consumption has been increasing. It is essential to have a technology to fill the gap between the amount of electricity generated and the demand. These technologies would contribute to the load leveling [1].
In particular, the recent development in information and communication technology (ICT) (e.g., Internet of Things, 5G communications, cloud computing, big data) has produced a great demand for data centers (DCs). In 2020, the rapid spread of remote working due to the pandemic also boosted demand for DCs. The industrial use of these technologies is also driving this trend [2,3]. Masanet et al. [4] estimated that the electricity use in DCs accounted for 1% of worldwide electricity use in 2020. This trend will make DCs even more energy-intensive and will increase the cost of cooling the heat generated and the amount of electricity used [5]. Considering the generalization of ICT and the widespread use of the fifth-generation mobile communication system, the amount of communication will further increase worldwide. The importance of data centers will increase to support the exchange of huge amounts of data. A suitable cooling system will be able to keep the appropriate temperature. In addition, such systems are significant to realize the zero-downtime of a DC operation with low cost.
DC cooling systems are classified into two types. One is an air-based cooling system [6,7]. The other one is a liquid cooling system [8]. As for air-based system, the heat emitted from the computer is removed by air flow. The main drawback of this system is a low thermal conductivity, less than 0.03 W m−1 K−1 [9]. On the other hand, in a liquid cooling system, the heat is removed by the thermal conduction and convection into the liquid, mostly water. A liquid cooling system is inclined to be bigger and heavier because that system utilizes only sensible heat, which has low energy density. Thermal energy storage technology is used to store the heat energy with the heat capacity of substance. When the demand of the electricity is low, the surplus electricity is stored as cold energy. Then, the stored energy is changed to electricity when the demand increases, for example, daytime peak in power usage [10,11].
Phase Change Materials (PCMs) can be a competitive system for thermal energy storage because they can also utilize latent heat, which has a higher energy density [12,13,14,15,16]. Organic compounds and water are representative examples of PCMs for DC cooling. As for organic PCMs, they often have problems in terms of safety. For example, paraffin is known as an organic PCM, but it is not safe because of its flammability [17]. Water is not suitable because its phase change point is outside the range of operation temperature in DCs, 288 K to 305 K [18].
Utilizing clathrate hydrate for a thermal energy storage medium has been proposed as a better idea than water or an ice-based medium [19,20,21,22,23]. Clathrate hydrates, ice-like compounds, are formed when some gases have contact with water or ice under high pressure and/or low temperature [24]. The gases encapsulated in clathrate hydrates are called guest compounds. Clathrate hydrates have enormous potential for industrial usage thanks to their thermodynamic properties [19,22,25,26,27,28]. Considering the application for thermal energy storage medium, equilibrium temperature should be suitable for the DC operation temperature. Favorable equilibrium temperature can be obtained appropriately by selecting the guest compounds [29,30,31].
Until now, only tetrabutylammonium bromide (TBAB) hydrate has been commercially applied as a hydrate PCM for air conditioning. The thermodynamic properties of TBAB hydrate have been reported in the previous studies. TBAB hydrate has an equilibrium temperature at an atmospheric pressure of 285.9 K [32], a dissociation heat of 193 kJ/kg [33], and the thermal conductivity of 0.35 W m−1 K−1 [34]. As reported by ASHRAE Technical Committee [18], the appropriate temperature range of DC cooling is 288 K to 305 K. This is why TBAB hydrate is not favorable for DC cooling.
Sakamoto et al. [35] studied the thermodynamic properties of tetrabutylammonium acrylate (TBAAc) hydrate. According to the study, TBAAc hydrate had the highest equilibrium temperature of 291.5 K at w TBAAc = 0.36, where w TBAAc is the mass fraction of TBAAc. The greatest dissociation heat, a latent heat, of 195 kJ/kg was obtained at w TBAAc = 0.33 under ambient pressure. As for the thermal conductivity of ionic semiclathrate hydrates, Fujiura et al. [34] studied the value of tetrabutylammonium bromide (TBAB) hydrate and tetrabutylammonium chloride (TBAC) hydrate, which have similar characteristics with TBAAc hydrate. They reported that TBAB hydrate and TBAC hydrate have thermal conductivity of approximately 0.40 W m−1 K−1. According to the report by ASHRAE Technical Committee [18], the cooling system for DCs is required to be set from 288 K to 305 K. From the aspect of operation temperature in DCs, it can be said that TBAAc hydrates are suitable for utilizing in DC cooling as PCMs. Also, the dissociation heat of the TBAAc is larger than that of TBAB hydrate, which has been commercialized in the industrial area [33,35]. Given these advantages, TBAAc hydrates could be better PCMs for DC cooling. This is why it has significant meaning to practically evaluate the performance of TBAAc hydrates as PCMs. As a step for industrial use of TBAAc hydrate, we performed kinetic thermal energy storage experiments with TBAAc hydrate.
Based on these results, we obtained the data of energy storage density and energy storage rate during the experiments. During the experiments, we also found the adhesion of the hydrate to the surface of the heat exchanger. As the adhered hydrate is the thermal resistance to the further hydrate formation, we used an agitator or an ultrasonic transducer to break the adhesion. By comparing the data with and without external forces, such as the agitation and ultrasonic waves, we obtained the practical data for a hydrate-based thermal energy storage system in DC cooling.

2. Materials and Methods

2.1. Materials

The details on reagents used in this study were summarized in Table 1. Tetrabutylammonium Acrylate (TBAAc) aqueous solution was obtained by neutralizing Tetrabuthylammonium hydroxide (TBAOH) aqueous solution, 1.52 kg, (0.40 mass fraction, Sigma Aldrich Xo. LLC, Saint Louis, MO, USA) and Acrylic acid, 0.17 kg, (0.99 mass fraction in liquid reagent, Sigma Aldrich Xo. LLC, Saint Louis, State of MO, USA). Also, we added laboratory-made H2O, 0.35 kg. The mass fraction of TBAAc ( w T B A A c )) was 0.36. The masses of all reagents were measured by an electronic balance (GF-600, A&D Co. Ltd., Tokyo, Japan) with an expanded uncertainty of ± 0.004 g (coverage factor, k = 2).

2.2. Apparatus

The schematic diagram of the experimental apparatus is illustrated in Figure 1. The insulating container’s width, length and height are all 150 mm. Cooling water was circulated inside a copper coil tube (inner diameter 6.35 mm, outer diameter 4.35 mm) for heat exchange. This coil’s diameter, pitch and number of turns are 100 mm, 10 mm, and 7 times, respectively. The temperature and flow rate in the coil was controlled by a chiller (CTP-3000 EYELA, Tokyo Rikakikai Co. Ltd., Tokyo, Japan) and flowmeter with a precision needle valve (RK 1250, KOFLOC Kyoto, Kyoto, Japan). As shown in Figure 1, the temperature of cooling water and TBAAc were measured by three platinum resistance temperature detectors (222-055, Electronic Temperature Instrument Ltd., West Sussex, UK) with an uncertainty of ± 0.1 K (k = 2). The surface of the coil would be covered with hydrate during the experiment. This hydrate prevents the heat exchange between the heat exchanger and thermal energy storage medium. We used an ultrasonic transducer (HEC-45282, HONDA Electronics Co. Ltd., Tokyo, Japan) or an agitator (SM-103D, Kenis Ltd., Osaka, Japan) to prevent the hydrate deposition.

2.3. Procedures

We made 2.0 kg of TBAAc aqueous solution by mixing the above reagents and put them in the container. The entire system was confirmed to be stationary at 298 K. Tin (outlet of the chiller) and flow rate, V w ˙ , of the cooling water were controlled at 291 K and 5.0 × 10−4 m3/min. Then, we started the thermal energy storage with TBAAc hydrate. We set t = 0 when we started the circulation of the cooling water. The time evolution of Tin, Tout (temperature at inlet of the chiller) and TPCM were measured every 10 s to evaluate the kinetic aspect of TBAAc hydrate as a thermal energy storage medium.
We performed such experiments in three systems. First, we did not add any external forces to the apparatus. Second, we used a rotating-impeller agitator to break adhesions between the hydrate and the coil. Third, we used the ultrasonic wave instead of mechanical agitation. In the system where the agitator was used, we also investigated the effect of rotation rate of the agitator by changing the rate to 100 rpm, 300 rpm and 600 rpm. Similarly, we investigated the effect of the frequency of ultrasonic waves on the system by changing the frequency. The frequency was set at either 28 kHz or 56 kHz.

3. Results and Discussion

3.1. The System without Any External Forces (Static System)

Initially, we reported the thermographs of Tetrabutylammonium Acrylate (TBAAc) hydrate obtained by a differential scanning calorimeter (DSC). Figure 2 shows the thermographs, which were measured at w T B A A c = 0.33 and 0.36. A single peak was observed, and the existence of ice was not identified at w T B A A c = 0.33.
As for the system without any external forces (static system), we performed thermal energy storage experiments. We measured Tin, Tout, and TPCM. Figure 3 shows the time evolution of each temperature. Time zero indicates the time at which the circulating water is started to flow inside the coil. As shown in Figure 3, TPCM decreased until t = 1.5 h and increased from t = 1.5 to t = 4 h. When TPCM is increasing, TBAAc hydrate formation would occur because the TPCM was 10 K lower than its equilibrium temperature, 291.5 K [35]. After the hydrate formation, the heat generation from the hydrate formation occurred and exceeded the endothermic of the heat exchanger. This was why TPCM rose from t = 1.5 to t = 4 h. TPCM moderately decreased again at t = 4 h when TPCM was 288 K. At this moment, the formed TBAAc hydrate gradually grew around the heat exchanger as shown in Figure 4d. This adhesion would be the thermal resistance between the heat exchanger and thermal energy storage medium much lower than before. Then, the hydrate formation rate decreased due to the increased thermal resistance, and TPCM decreased because the endothermic of the heat exchanger exceeds the heat generation from hydrate formation.
Figure 4a–d illustrates the optical observation of the time evolution of the hydrate formation. The hydrate was formed on the surface of the heat exchanger. Heat exchange occurs on the heat exchanger surface, which leads to hydrate formation occurring there. From t = 1.5 h to t = 2 h, the hydrate gradually grew and started covering the coil surface. The surface of the coil was fully covered by the hydrate around t = 4 h. After t = 7 h, optical changes could not be observed. In addition, we found that the TBAAc hydrate kept growing from the surface of the coil to the bulk of the solution. There was a 3 cm gap between the surface of the coil and the wall of the container. This discovery of the hydrate growth length can be a tip to design a static thermal energy storage system. Thermal energy storage rate, q ˙ , and density, q ¯ , were calculated by the formulas (1) and (2). What every character stands for is written in Table 2. On the right-hand side of (2), the thermal energy storage density due to sensible heat is subtracted from the thermal energy storage density, which is obtained by integrating the thermal energy storage rate.
q ˙ = ρ w V ˙ w c w T o u t T i n / V P C M
q ¯ = t = 0 t q ˙ d t ρ P C M c P C M T P C M ,   0 T P C M
The obtained thermal energy storage density represents the heat storage density due to heat of hydrate formation and decomposition. The density of the PCM is assumed to be 1.0 kg m−3. The value of each parameter used in the calculation is shown in Table 2. The physical property of water was referred to REFPROP ver. 9.1 [36].
Figure 5a illustrates the time evolution of the thermal energy storage rate in the static system. Since q ˙ is proportional to the gap between Tin and Tout, q ˙ varied the same way with Tout. Immediately after the start of the experiment, q ˙ decreased as Tout decreased, and as Tout approached Tin, q ˙ approached 0. After that, q ˙ increased at t = 1.5 h and reached around 9 kW/m3. This is due to the increase in Tout caused by the heat generation from hydrate formation and growth, as described above. After t = 3 h, as Tout descended to be tangent to Tin, q ˙ descended to 0 with some fluctuations. Figure 5b illustrates the time evolution of thermal energy storage density. Since q ¯ indicates the thermal energy density by hydrate growth and decomposition, there is no change right after the experiments start. As the hydrate grew, thermal energy storage density increased. The rate of increase gradually became slower. The hydrate formation rate got slower due to the decreased heat conductivity around the heat exchanger coil. This is why the increasing rate of thermal energy storage density decreased. Thermal energy density reached 96 MJ/m3 in 10 h.
It is known that the solid phase fraction of ice thermal storage increases significantly with time up to around 30%, but the increasing rate decreases after about 40–50%. Therefore, it is favorable to design a thermal energy storage system that has a solid phase fraction of 40–50% [37]. For example, when the solid phase fraction is 45% in an ice thermal energy storage system, thermal energy storage density is approximately 140 MJ/m3. Oyama et al. [33] reported that TBAB hydrate, which is the only commercially available hydrate PCM, has the heat for formation and decomposition of approximately 200 MJ/m3. Assuming the TBAAc hydrate has 200 MJ/m3 of heat in hydrate formation and decomposition, the thermal energy storage density corresponds to a solid phase fraction of 48%.

3.2. In the System with Mechanical Agitation

As written in 3.1., the surface of the heat exchanger coil was covered by hydrate. This hydrate adhesion increased thermal resistance. To solve this problem, mechanical agitation was used. In addition, to investigate the effect of the rotation rate, we changed the rate to 100 rpm, 300 rpm and 600 rpm. Figure 6a–c shows the time evolution of Tin, Tout and TPCM in each rotation rate. Time zero was defined as the time at which the circulating water is started to be flown. In every rotation rate, TPCM decreased twice after the experiment started and temporarily increased. In the agitated system, TPCM increased earlier than the static system. As the rotation rate became quicker, TPCM increased earlier. When TPCM increased, the temperature was around 282 K, which is the same as the static system. This result indicates that the forced convection generated by the agitation increased the convective heat transfer coefficient, which resulted in faster cooling of the heat storage medium and thus faster nucleation. TPCM increased rapidly more than the static system, which was almost vertically to around 288 K. This would be due to the increase in heat and mass transfer rates caused by the forced convection generated by the agitation, which led to the increase in the hydrate growth rate.
TPCM changed from a rise to a fall around 288 K and then descended slowly. As the hydrate grew, the viscosity of the heat storage medium increased and the forced convection was weakened, which lowered the hydrate formation rate. TPCM decreased because the hydrate growth rate became smaller, and the heat absorption by the heat exchanger exceeded the heat generated from hydrate formation.
Figure 7a–c shows the pictures of crystal growth in the system when the mechanical agitation rate was 300 rpm. At t = 0.6 h, numerous small crystals began to float in the aqueous solution. After t = 0.7 h, hydrate was formed in the entire aqueous solution. The hydrate nucleation occurred in the entire aqueous solution. This was because of the forced convection generated by the agitation, which stripped the hydrate from the heat exchange coil and suspended it in the aqueous solution. This convection increased the hydrate growth rate as well. Within a few minutes, the heat exchange coil was only faintly visible by the hydrate, as shown in Figure 7b. The hydrate formation has progressed and the viscosity increased as shown in Figure 7c, after which little change was observed.
We visually observed that TBAAc hydrate formed and grew to the bulk of the solution. The gap between the heat transfer surface and the corner of the container was about 5 cm. In other words, when the cooling temperature was 10 K lower than the equilibrium temperature, TBAAc hydrate formed and grew at a distance of more than 5 cm from the heat transfer surface in the agitated system. The same tendency was observed at the rotation rate of 100 rpm and 600 rpm.
Figure 8a shows the time evolution of thermal energy storage density. To make a comparison with the static system, the time evolution of the static system is also shown in Figure 8a. Regarding the thermal energy storage rate, q ˙ is proportional to the difference between the inflow temperature Tin and the outflow temperature Tout. This was why q ˙ had almost the same trend as the Tout change in the static system. As Tout increased rapidly more than the static system, q ˙ increased rapidly more than the static system as well, rising almost vertically. The maximum thermal energy storage rate at that time was about 6 to 8 times faster than that of the static system. The larger the rotation rate of the agitation increased the value of q ˙ . After a rapid increase, the change in q ˙ suddenly turned into a downward movement and gradually descended to 0. This was because the hydrate had grown, and the viscosity had increased, which resulted in less convection.
As shown in Figure 8b, q ¯ in the agitated system increased earlier than that in the static system, and the higher rotation rate of the agitation increased q ¯ earlier. This was because the temperature of the thermal energy storage medium was lowered quickly by forced convection, and thus hydrate nucleation also occurred quickly. The thermal energy storage density q ¯ was larger than that of the static system and increased by a higher rotation rate. This was because the forced convection caused by the agitation allowed hydrate to be formed in the entire bulk of the solution. In the agitated system with the rotation rates of 100 rpm, 300 rpm and 600 rpm, thermal energy density reached 121 MJ/m3, 159 MJ/m3 and 183 MJ/m3 in 10 h, respectively. Similarly, assuming the TBAAc hydrate has 200 MJ/m3 of heat in hydrate formation and decomposition, the thermal energy storage density corresponds to a solid phase fraction of 61%, 80% and 92%, respectively.

3.3. In the System Using an Ultrasonic Transducer

In this section, we discuss the results in the system with ultrasonic vibration. The inflow temperature Tin, outflow temperature Tout and thermal energy storage medium temperature TPCM were measured. Figure 9a,b shows the time evolutions of Tin, Tout and TPCM in the ultrasonic vibration system (28 kHz and 56 kHz). Time zero was defined as the time at which the circulating water was started to flow.
The slope of the TPCM first decrease was similar to that of the static system. However, when the slope changed from negative to positive, the TPCM was about 284 K, which was higher than that of the static and agitated system. This is because the hydrate nucleation occurred at a higher temperature than in the static system. In an ultrasonic vibration system, hydrate nucleation would occur at a higher temperature due to the effects of cavitation or acceleration of ultrasonic vibration.
The slope of the TPCM increase is not much different from that of the static system at 28 kHz and appears to be more gradual at 56 kHz. This was because of relatively high TPCM, which results in a relatively slow hydrate growth rate and low heat of formation during hydrate nucleation. The increase in TPCM changed from a rise to a fall at around 288 K, and then TPCM slowly decreased. This was because the slope of the TPCM change turned from positive to negative because hydrate was generated and covered the heat exchange coil surface as the static system.
Figure 10a–c shows the pictures of the crystal growth during this system. As shown in Figure 10a–c, hydrate was formed on the heat transfer surface near the bottom of the container. As time went by, the formed hydrate piled up on the bottom of the container. Figure 11a shows time evolutions of thermal energy storage rate in the system using the ultrasonic vibration, which depends on the frequency of 28 kHz and 56 kHz. The thermal energy storage rate q ˙ in the static system is also shown for comparison. Time zero was defined as the time at which the circulating water is started to flow. Here, q ˙ was proportional to the difference between Tin and Tout and thus shows almost the same trend as Tout change in the static system. In addition, the slope of change in the thermal energy storage rate was similar to that of the static system. However, as for the 28 kHz ultrasonic vibration, the maximum value of q ˙ after the increase was about 1.5 times larger than that of the static system. This was because the hydrate was detached from the heat exchange coil by cavitation, which increased the growth rate of the hydrate. On the other hand, the maximum value of q ˙ at 56 kHz was almost the same as that of the stationary system. Considering the size of hydrate adhesion, the cavitation effect is stronger at lower frequencies. This was why the maximum value of q ˙ at 56 kHz was larger than that of q ˙ at 28 kHz. After the peak of q ˙ , the ultrasonic vibration system maintained a larger q ˙ for a relatively longer time than the static system. This was because the hydrate growth was continued for a relatively long time after the hydrate covered the heat exchange coil due to the effects of cavitation. Figure 11b shows the time evolution of the thermal energy storage density. The result of the static system is also shown in Figure 11b. Time zero was defined as the time at which the circulating water is started to flow. As shown in Figure 11b, q ¯ in the ultrasonic vibrated system started increasing a bit earlier than the static system. This was because the cavitation enabled the hydrate to be formed at a higher temperature. The slope of q ¯ change in the ultrasonic vibration system is almost the same as with the static system. However, the bigger slope could be kept longer than the static system. This was also because the hydrate growth could be kept by the effect of the cavitation. In the ultrasonic vibration system with a frequency of 28 kHz and 56 kHz, thermal energy storage density reached 126 MJ/m3, 111 MJ/m3 in 10 h, respectively. Similarly, assuming the TBAAc hydrate has 200 MJ/m3 of heat in hydrate formation and decomposition, the thermal energy storage density corresponds to a solid phase fraction of 63% and 55%, respectively.

3.4. Comparison in All Systems

Figure 12a shows a comparison in the time evolution of the thermal energy storage rate in all systems. In the agitated system, the thermal energy storage rate, q ˙ , increased earlier than the other systems; q ˙ increased at 1 h, 0.6 and 0.4 h for a rotation rate of 100 rpm, 300 rpm and 600 rpm, respectively. The forced convection generated by the agitation increased the convective heat transfer coefficient. This was why the temperature of the thermal energy storage medium dropped relatively earlier and hydrate formation occurred earlier, resulting in the early increase of q ˙ . In addition, this forced convection caused rapid hydrate formation in the entire container, which released huge heat in a short period. This was why q ˙ increased vertically. As the rotation rate increased, q ˙ increased more. The maximum values of q ˙ after the vertical increase were 47 kW m−3, 54 kW m−3 and 69 kW m−3 for a rotation rate of 100 rpm, 300 rpm and 600 rpm, respectively. These values are 5 to 8 times larger than that of a static system.
In the ultrasonic vibration system, the maximum of q ˙ was 14 kW m−3, which is about 1.5 times larger than that of the static system, when the frequency was 28 kHz. This was because cavitation detached the adhesion between hydrate and heat exchanger coil, which could be the thermal resistance. As for the system with 56 kHz frequency, the maximum values of q ˙ were almost the same as in the static system. This was because the effect of the cavitation was not as big as 28 kHz [38]. In summary, the agitation increased the rate of heat and mass transfer by creating forced convection, thereby shortening the hydrate formation time and increasing the hydrate growth rate. On the other hand, ultrasonic vibration kept the hydrate growth longer by the effect of cavitation.
Figure 12b shows a comparison in the time evolution of thermal energy storage density, q ¯ , in all systems. In the agitated system, q ¯ increased earlier and quicker than the other systems. This was because TPCM decreased quicker by the forced convection, which caused the quicker hydrate formation and growth for the entire container. In the ultrasonic vibration system, the q ¯ increase appeared to be relatively longer than the other systems. This was because the cavitation enabled hydrate to grow after hydrate covered the surface of the heat exchanger coil. The static system, the agitated system (rotation rate of 100 rpm, 300 rpm and 600 rpm), and the ultrasonic vibrated system (frequency of 28 kHz and 56 kHz) had the thermal energy storage density of 96 MJ/m3, 121 MJ/m3, 159 MJ/m3, 183 MJ/m3, 126 MJ/m3, and 111 MJ/m3, respectively. Assuming TBAAc hydrate has the heat for formation and decomposition of approximately 200 MJ/m3, each solid phase fraction is 48%, 61%, 80%, 92%, 63%, and 55%, respectively.

3.5. For Industrial Utilizing of TBAAc Hydrate as PCM

As written in Section 3.1., a range of the practical solid-phase fraction is from 40 to 50% [37]. The thermal energy storage density of ice is 140 MJ m−3 when the solid phase fraction is 45%. As for TBAB hydrate, it has a thermal energy storage density of 100 MJ/m3 when the solid phase fraction is 50% [33]. We regarded these values as a target thermal energy storage density. Given the thermal energy storage during nighttime of 10 h, the required time for thermal energy storage should be shortened. This was why the time to reach the target values was compared to each other system. In addition, the amount of the used electric power was estimated for more practical evaluation.
Table 3 shows the time that was needed to achieve the target thermal energy storage density. As for the target of 140 MJ/m3, it was achieved only when the rotation rates of agitation were 300 rpm and 600 rpm. When the rotation rates were 300 rpm or 600 rpm, it was 5.9 h or 2.9 h to achieve the target, respectively. Concerning the consumed electric power to achieve the target, it was 58 MJ/m3 and 29 MJ/m3, respectively. As for the target of 100 MJ/m3, it was achieved in all systems where an external force was applied. When the rotation rates were 300 rpm and 600 rpm, it took about a third of the time of the ultrasonic vibration system. However, the consumed electric power by the external force to achieve the target was 45 MJ/m3, 24 MJ/m3, 15 MJ/m3 for the agitation system (100 rpm, 300 rpm, 600 rpm), and 5 MJ/m3 for the ultrasonic vibration system at both frequencies (28 kHz, 56 kHz), respectively. Therefore, when the target of thermal energy storage density is 100 MJ/m3, the agitation system (300 rpm, 600 rpm) is more suitable for storing thermal energy in a short time, and the ultrasonic vibration system is more suitable as an external force for reducing the amount of electricity used.
In this study, the thermal energy storage density in the agitated system was greatly improved by adding agitation. On the other hand, the amount of electricity used for agitation became larger than other systems. The agitation intensity was almost the same as that in a general low-viscosity agitation. However, the agitation Reynolds number was 25,000 at 600 rpm. This was due to the fact that the impeller power number of the three propeller blades was 1, which was smaller than that of other agitation blades [39]. Therefore, even if the agitation intensity was a general value, the agitation Reynolds number became larger, and energy consumption also became larger. It will be possible to optimize the rotations and electric power consumption by selecting a more suitable agitator at the similar experimental apparatus. The thermal energy storage density in the ultrasonic vibration system was not as large as that in the agitation system. The reason for this result is that it was difficult to resonate the ultrasonic transducer. To resonate the ultrasonic transducer, it is necessary to provide a natural frequency, but this frequency varies slightly depending on the environment in which the ultrasonic transducer is placed, for example, the materials in contact. Also, the natural frequency is also expected to change during the transformation of the aqueous solution into hydrate. Therefore, it is possible that the ultrasonic transducer was not able to resonate properly in this experimental system. A possible solution to solve this problem is to use a device that automatically follows the resonant frequency and keeps giving the resonant frequency to the transducer. Since the power consumption for ultrasonic vibration is about one-tenth of the power used for agitation, it is thought that the ultrasonic vibration system could be the most favorable way by solving the above problems.

3.6. The Evaluation of the Heat Exchanger

The thermal energy storage rate by the heat exchanger can be calculated as in Equation (3), where U is the heat transfer coefficient, A is the surface area of the heat exchange coil, and VPCM is the volume of the thermal energy storage medium. The logarithmic mean temperature difference, ∆Tlm, is calculated by Equation (4). As can be seen from Equation (4), the logarithmic mean temperature difference, ∆Tlm, is the average of the temperature difference between the thermal energy storage medium and the cooling water using the logarithm. This logarithmic mean temperature difference was substituted into Equation (3) to calculate the heat transfer coefficient for the experiments under each condition. The heat transfer coefficient was calculated to be about 104–105 W m−2 K. The A/VPCM in this experiment was 15 m2/m3, and the UA/VPCM value was calculated to be around 103–104 W m−3 K. For instance, a plate heat exchanger has a UA/VPCM that is similar to this value [40]. As the UA/V increases, the amount of heat exchange will also increase. It realizes a higher thermal energy storage density. In this study, the UA/V of the heat exchanger was comparable to that of other common heat exchangers [40]. This result demonstrates that the comparable thermal energy storage density in this study can be obtained using other common heat exchangers.
For the practical use of thermal energy storage using TBAAc hydrate on a larger scale, it is also expected that the comparable thermal energy storage density in this study would be available by using a heat exchanger which has the same level UA/V.
q ˙ = U A T lm / V P C M
T lm T o u t T i n l n T P C M T i n T P C M T o u t

4. Conclusions

Kinetic characteristics of thermal energy storage using tetrabutylammonium acrylate (TBAAc) hydrate were practically and experimentally evaluated for practical use as Phase Change Materials (PCMs). During the experiments, the adhesion of hydrate on the surface of the heat exchanger coil was found, which increased the thermal resistance between the heat exchanger and thermal energy storage medium. This was why, as the external forces, we added a mechanical agitation (rotation rate of 100 rpm, 300 rpm, and 600 rpm) and ultrasonic vibration (frequency of 28 kHz and 56 kHz) on each system. It was revealed that the external forces improved the thermal energy storage kinetic characteristics. When the agitation rate was 600 rpm, the system achieved thermal energy storage of 140 MJ/m3 in 2.9 h. This value is comparable to the ideal performance of ice thermal energy storage when its solid phase fraction is 45%. The energy consumption for agitation was 28 MJ/m3 to achieve this value. The thermal energy storage of 100 MJ/m3, which is a TBAAc solid phase fraction of 50%, was achieved in 1.5 h and 5.8 h with mechanical agitation (600 rpm) and ultrasonic vibration (28 kHz), respectively. The energy consumptions to achieve this target value were 15 MJ/m3 and 5.2 MJ/m3, respectively. In summary, the agitation increased the rate of heat and mass transfer by creating forced convection, thereby shortening the hydrate formation time and increasing the growth rate of the hydrate. In addition, ultrasonic vibration could keep the hydrate growth time longer by the effect of cavitation.
In this system, the UA/V (U: thermal transfer coefficient, A: surface area of the heat exchange coil, V: volume of the thermal energy storage medium) was approximately 1.0 × 10−4–10−3 W m−3 K. This UA/V of the heat exchanger was comparable to that of other common heat exchangers. Even if the system would be scaled up and different from this study, the same level of thermal energy storage as this study could be obtained by using a heat exchanger of which UA/V is comparable to this study. As for the ultrasonic vibration system, thermal energy storage density was one-tenth of the agitated system. A device that enables an oscillator to resonate and break the hydrate adhesion is necessary to increase the thermal energy storage density. Then, that system would be most favorable in terms of the storing performance and energy power consumption.

Author Contributions

Conceptualization, H.K., H.T. and R.O.; methodology, H.T., S.I., H.N. and R.O.; formal analysis, H.T., S.I., H.N. and R.O.; investigation, H.K., H.T., S.I., H.N. and R.O.; resources, H.T. and R.O.; data curation, H.K., H.T., S.I., H.N. and R.O.; writing—original draft preparation, H.K. and R.O.; writing—review and editing, H.K. and R.O.; visualization, H.T. and R.O.; supervision, R.O.; project administration, R.O.; funding acquisition, R.O. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been supported by a Keirin-racing-based research promotion fund from the JKA Foundation (Grant Number: 2020M-195) and a part of Low Carbon Technology Research and Development Program for “Practical Study on Energy Management to Reduce CO2 emissions from University Campuses” from the Ministry of the Environment, Japan.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank Haruki Sato, professor emeritus and Masao Takeuchi, Keio University, for the encouragement on this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. IEA. World Energy Outlook; International Energy Agency (IEA): Paris, France, 2019. [Google Scholar]
  2. Fei, F.; Qi, Q.; Liu, A.; Kusiak, A.A. Data-driven smart manufacturing. J. Manuf. Syst. 2018, 48, 157–169. [Google Scholar]
  3. Król, A. The Application of the Artificial Intelligence Methods for Planning of the Development of the Transportation Network. Transp. Res. Procedia 2016, 14, 4532–4541. [Google Scholar] [CrossRef] [Green Version]
  4. Masanet, E.; Shehabi, A.; Lei, N.; Smith, S.; Koomey, J. Recalibrating global data center energy-use estimates. Science 2020, 367, 984–986. [Google Scholar] [CrossRef] [PubMed]
  5. Zimmermann, S.; Meijer, I.; Tiwari, M.K.; Paredes, S.; Michel, B.; Poulikakos, D. Aquasar: A hot water cooled data center with direct energy reuse. Energy 2012, 43, 237–245. [Google Scholar] [CrossRef]
  6. Beghi, A.; Cecchinato, L.; Mana, G.D.; Lionello, M.; Rampazzo, M.; Sisti, E. Modelling and control of a free cooling system for Data Centers. Energy Procedia 2017, 140, 447–457. [Google Scholar] [CrossRef]
  7. Ko, J.-S.; Huh, J.-H.; Kim, J.-C. Improvement of Energy Efficiency and Control Performance of Cooling System Fan Applied to Industry 4.0 Data Center. Electronics 2019, 8, 582. [Google Scholar] [CrossRef] [Green Version]
  8. Khalaj, A.H.; Halgamuge, S.K. A Review on efficient thermal management of air- and liquid-cooled data centers: From chip to the cooling system. Appl. Energy 2017, 205, 1165–1188. [Google Scholar] [CrossRef]
  9. Stephan, K.; Laesecke, A. The thermal conductivity of fluid air. J. Phys. Chem. Ref. Data 1985, 14, 227–234. [Google Scholar] [CrossRef] [Green Version]
  10. Pielichowska, K.; Pielichowski, K. Phase change materials for thermal energy storage. Prog. Mater. Sci. 2014, 65, 67–123. [Google Scholar] [CrossRef]
  11. Veerakumar, C.; Sreekumar, A. Phase change material based cold thermal energy storage: Materials, techniques and applications—A review. Int. J. Refrig. 2015, 67, 271–289. [Google Scholar] [CrossRef]
  12. Xie, N.; Huang, Z.; Luo, Z.; Gao, X.; Fang, Y.; Zhang, Z. Inorganic Salt Hydrate for Thermal Energy Storage. Appl. Sci. 2017, 7, 1317. [Google Scholar] [CrossRef] [Green Version]
  13. Porteiro, J.; Míguez, J.L.; Crespo, B.; Lara, J.; Pousada, J.M. On the Behavior of Different PCMs in a Hot Water Storage Tank against Thermal Demands. Materials 2016, 9, 213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Ndukwu, M.C.; Bennamoun, L.; Simo-Tagne, M. Reviewing the Exergy Analysis of Solar Thermal Systems Integrated with Phase Change Materials. Energies 2021, 14, 724. [Google Scholar] [CrossRef]
  15. Wong-Pinto, L.-S.; Milan, Y.; Ushak, S. Progress on use of nanoparticles in salt hydrates as phase change materials. Renew. Sustain. Energy Rev. 2020, 122, 109727. [Google Scholar] [CrossRef]
  16. Mitran, R.-A.; Ionita, S.; Lincu, D.; Berger, D.; Matei, C. A Review of Composite Phase Change Materials Based on Porous Silica Nanomaterials for Latent Heat Storage Applications. Molecules 2021, 26, 241. [Google Scholar] [CrossRef]
  17. Zalba, B.; Marín, J.; Cabeza, L.; Mehling, H. Review on thermal energy storage with phase change: Materials, heat transfer analysis and applications. Appl. Therm. Eng. 2003, 23, 251–283. [Google Scholar] [CrossRef]
  18. ASHRAE Technical Committee. ASHRAE TC9.9 Data Center Power Equipment Thermal Guidelines and Best Practices; ASHRAE: Peachtree Corners, GA, USA, 2016. [Google Scholar]
  19. Arai, Y.; Yamauchi, Y.; Tokutomi, H.; Endo, F.; Hotta, A.; Alavi, S.; Ohmura, R. Thermophysical property measurements of tetrabutylphosphonium acetate (TBPAce) ionic semiclathrate hydrate as thermal energy storage medium for general air conditioning systems. Int. J. Refrig. 2018, 88, 102–107. [Google Scholar] [CrossRef]
  20. Koyama, R.; Hotta, A.; Ohmura, R. Equilibrium temperature and dissociation heat of tetrabutylphosphonium acrylate (TBPAc) ionic semi-clathrate hydrate as a medium for the hydrate-based thermal energy storage system. J. Chem. Thermodyn. 2020, 144, 106088. [Google Scholar] [CrossRef]
  21. Miyamoto, T.; Koyama, R.; Kurokawa, N.; Hotta, A.; Alavi, S.; Ohmura, R. Thermophysical property measurements of tetrabutylphosphonium oxalate (TBPOx) ionic semiclathrate hydrates as a media for the thermal energy storage system. Front. Chem. 2020, 8, 547. [Google Scholar] [CrossRef]
  22. Koyama, R.; Arai, Y.; Yamauchi, Y.; Takeya, S.; Endo, F.; Hotta, A.; Ohmura, R. Thermophysical Properties of Trimethylolethane (TME) Hydrate as Phase Change Material for Cooling Lithium-ion Battery in Electric Vehicle. J. Power Sources 2019, 427, 70–76. [Google Scholar] [CrossRef]
  23. Nakane, R.; Shimosato, Y.; Gima, E.; Ohmura, R.; Senaha, I.; Yasuda, K. Phase equilibrium condition measurements inn carbo dioxide hydrate forming system coexisting with seawater. J. Chem. Thermodyn. 2021, 152, 106276. [Google Scholar] [CrossRef]
  24. Alavi, S.; Ohmura, R. Understanding decomposition and encapsulation energies of structure I and II clathrate hydrates. J. Chem. Phys. 2016, 145, 154708. [Google Scholar] [CrossRef] [PubMed]
  25. Horii, S.; Ohmura, R. Continuous separation pf CO2 from a H2 + CO2 gas mixture using clathrate hydrate. Appl. Energy 2018, 225, 78–84. [Google Scholar] [CrossRef]
  26. Kiyokawa, H.; Horii, S.; Alavi, S.; Ohmura, R. Improvement of continuous hydrate-based CO2 separation by forming structure II hydrate in the system of H2 + CO2 + H2O + Tetrahydropyran (THP). Fuel 2020, 278, 118330. [Google Scholar] [CrossRef]
  27. Hatsugai, T.; Nakayama, R.; Tomura, S.; Akiyoshi, R.; Nishitsuka, S.; Nakamura, R.; Takeya, S.; Ohmura, R. Development and Continuous Operation of a Bench-Scale System for the Production of O3 + O2 + CO2 Hydrates. Chem. Eng. Technol. 2020, 43, 2307–2314. [Google Scholar] [CrossRef]
  28. Nagashima, H.D.; Alavi, S.; Ohmura, R. Preservation of carbon dioxide clathrate hydrate in the presence of fructose or glucose and absence of sugars under freezer conditions. J. Ind. Eng. Chem. 2017, 54, 332–340. [Google Scholar] [CrossRef]
  29. Kondo, Y.; Alavi, S.; Takeya, S.; Ohmura, R. Characterization of the Clathrate Hydrate Formed with Fluoromethane and Pinacolone: The Thermodynamic Stability and Volumetric Behavior of the Structure H Binary Hydrate. J. Phys. Chem. B 2021, 125, 328–337. [Google Scholar] [CrossRef]
  30. Yamauchi, Y.; Yamasaki, T.; Endo, F.; Hotta, A.; Ohmura, R. Thermodynamic Properties of Ionic Semiclathrate Hydrate Formed with Tetrabutylammonium Propionate. Chem. Eng. Technol. 2017, 40, 1810–1816. [Google Scholar] [CrossRef]
  31. Yamauchi, Y.; Arai, Y.; Yamasaki, T.; Endo, F.; Hotta, A.; Ohmura, R. Phase equilibrium temperature and dissociation heat of ionic semiclathrate hydrate formed with tetrabutylammonium butyrate. Fluid Phase Equilib. 2017, 441, 54–58. [Google Scholar] [CrossRef]
  32. Sato, K.; Tokutomi, H.; Ohmura, R. Phase equilibrium of ionic semi-clathrate hydrates formed with tetrabutylammonium bromide and tetrabutylammonium chloride. Fluid Phase Equilib. 2013, 337, 115–118. [Google Scholar] [CrossRef]
  33. Oyama, H.; Shimada, W.; Ebinuma, T.; Kamata, Y.; Takeya, S.; Uchida, T.; Nagao, J.; Narita, H. Phase diagram, latent heat, and specific heat of TBAB semi-clathrate hydrate crystals. Fluid Phase Equilib. 2005, 234, 131–135. [Google Scholar] [CrossRef]
  34. Fujiura, K.; Nakamoto, Y.; Taguchi, Y.; Ohmura, R.; Nagasaka, Y. Thermal conductivity measurements of semi-clathrate hydrates and aqueous solutions of tetrabutylammonium bromide (TBAB) and tetrabutylammonium chloride (TBAC) by the transient hot-wire using parylene-coated probe. Fluid Phase Equilib. 2016, 413, 129–136. [Google Scholar] [CrossRef] [Green Version]
  35. Sakamoto, H.; Sato, K.; Shiraiwa, K.; Takeya, S.; Nakajima, M.; Ohmura, R. Synthesis, Characterization and thermal-property measurements of ionic semi-clathrate hydrates formed with tetrabutylphosphonium chloride and tetrabutylammonium acrylate. RSC Adv. 2011, 1, 315–322. [Google Scholar] [CrossRef]
  36. Lemmon, E.W.; McLinden, M.O. Huber M.L. NIST Reference Fluid Thermodynamic and Transport Properties (REFPROP). J. Res. Natl. Inst. Stand. Technol. 2013. [Google Scholar]
  37. Nakajima, M.; Hirata, A. High-Efficient Thermal Energy Storage Technique with a Clathrate Hydrate; IHI Technical Report; IHI: Tokyo, Japan, 2009; Volume 49, pp. 210–218. [Google Scholar]
  38. Iida, Y. Sono Process no Hanashi—Chemical Engineering Applications of Ultrasonic (SCIENCE AND TECHNOLOGY). Nikkan Kogyo Shimbun 2006, 7, 7–17. [Google Scholar]
  39. Hashimoto, K. Industrial Reaction Equipment: Selection, Design, and Examples; Baihukan: Tokyo, Japan, 1984; ISBN 9784563041618. [Google Scholar]
  40. The Japan Society of Mechanical Engineers. JSME Data Book: Heat Transfer Materials; The Japan Society of Mechanical Engineers: Tokyo, Japan, 1986; ISBN 9784888981842. [Google Scholar]
Figure 1. Schematic diagram of the apparatus.
Figure 1. Schematic diagram of the apparatus.
Applsci 11 04848 g001
Figure 2. The DSC heating curves measured at w T B A A c = 0.33 and 0.36. : w T B A A c = 0.36, : w T B A A c = 0.33.
Figure 2. The DSC heating curves measured at w T B A A c = 0.33 and 0.36. : w T B A A c = 0.36, : w T B A A c = 0.33.
Applsci 11 04848 g002
Figure 3. Time evolutions of the temperature at inlet of the chiller (Tin), outlet (Tout) of the chiller and thermal energy storage medium (TPCM) in the static system.
Figure 3. Time evolutions of the temperature at inlet of the chiller (Tin), outlet (Tout) of the chiller and thermal energy storage medium (TPCM) in the static system.
Applsci 11 04848 g003
Figure 4. Crystal growth of TBAAc hydrate on the surface of the heat exchanger coil. (ad) illustrate the optical measurement of the hydrate formation at t = 2, t = 2.5, t = 4 and t = 7, respectively.
Figure 4. Crystal growth of TBAAc hydrate on the surface of the heat exchanger coil. (ad) illustrate the optical measurement of the hydrate formation at t = 2, t = 2.5, t = 4 and t = 7, respectively.
Applsci 11 04848 g004
Figure 5. Time evolution of thermal energy storage rate and density in static system. (a,b) depict the thermal energy storage rate and density, respectively.
Figure 5. Time evolution of thermal energy storage rate and density in static system. (a,b) depict the thermal energy storage rate and density, respectively.
Applsci 11 04848 g005
Figure 6. Time evolutions of the temperature at inlet of the chiller (Tin), outlet (Tout) of the chiller and thermal energy storage medium (TPCM) in the system with the mechanical agitation. (ac) depict each curve at the agitation rate of 100 rpm, 300 rpm and 600 rpm, respectively.
Figure 6. Time evolutions of the temperature at inlet of the chiller (Tin), outlet (Tout) of the chiller and thermal energy storage medium (TPCM) in the system with the mechanical agitation. (ac) depict each curve at the agitation rate of 100 rpm, 300 rpm and 600 rpm, respectively.
Applsci 11 04848 g006
Figure 7. Optical observation changing in the container with mechanical agitation of 300 rpm. (ac) are the pictures at 0.6 h, 0.7 h and 1 h after cooling water started to be circulated, re-spectively.
Figure 7. Optical observation changing in the container with mechanical agitation of 300 rpm. (ac) are the pictures at 0.6 h, 0.7 h and 1 h after cooling water started to be circulated, re-spectively.
Applsci 11 04848 g007
Figure 8. Time evolution of thermal energy storage rate and density in the agitated system (a,b) depict the thermal energy storage rate and density, respectively.
Figure 8. Time evolution of thermal energy storage rate and density in the agitated system (a,b) depict the thermal energy storage rate and density, respectively.
Applsci 11 04848 g008
Figure 9. Time evolutions of the temperature at inlet of the chiller (Tin), outlet (Tout) of the chiller and thermal energy storage medium (TPCM) in the system with ultrasonic vibration. (a,b) depict each curve at the frequency of 28 kHz and 56 kHz, respectively.
Figure 9. Time evolutions of the temperature at inlet of the chiller (Tin), outlet (Tout) of the chiller and thermal energy storage medium (TPCM) in the system with ultrasonic vibration. (a,b) depict each curve at the frequency of 28 kHz and 56 kHz, respectively.
Applsci 11 04848 g009
Figure 10. Optical observation changing in the container with ultrasonic vibration of 28 kHz. (ad) are the pictures at 2 h, 2.5 h, 3 h and 5 h after cooling water started to be circulated, respectively.
Figure 10. Optical observation changing in the container with ultrasonic vibration of 28 kHz. (ad) are the pictures at 2 h, 2.5 h, 3 h and 5 h after cooling water started to be circulated, respectively.
Applsci 11 04848 g010
Figure 11. Time evolutions of the thermal energy storage rate and density in the system using an ultrasonic vibration. (a,b) depict the thermal energy storage rate and density, respectively.
Figure 11. Time evolutions of the thermal energy storage rate and density in the system using an ultrasonic vibration. (a,b) depict the thermal energy storage rate and density, respectively.
Applsci 11 04848 g011
Figure 12. Time evolutions of thermal energy storage rate and density in all systems. (a,b) depict the thermal energy storage rate and density, respectively.
Figure 12. Time evolutions of thermal energy storage rate and density in all systems. (a,b) depict the thermal energy storage rate and density, respectively.
Applsci 11 04848 g012
Table 1. Specification of the reagents used in this study.
Table 1. Specification of the reagents used in this study.
NameChemical FormulaSupplierPurity
Tetrabuthylammonium hydroxide(CH3CH2CH2CH2)4NOHSigma-Aldrich Co. LLC0.40 mass fraction in aqueous solution
Acrylic acidCH2=CHCOOHSigma-Aldrich Co. LLC0.99 mass fraction in liquid reagent
Tetrabuthylammonium Acrylate(CH3CH2CH2CH2)4NOOCHC=CH2Laboratory made from above solution0.36 mass fraction in aqueous solution after the neutralization process
The standard uncertainty of mass fraction was ± 1.0×10−4
WaterH2OLaboratory madeElectrical conductivity was less than 0.1 μS/cm
Table 2. The value of each parameter in the calculation.
Table 2. The value of each parameter in the calculation.
Parameter
Density, ρ w (Water)999.8 kg/m3
Volumetric flow rate, V w ˙ (Water)8.3 × 10−5 m3/s
Specific heat at constant pressure, c w (Water)4.20 kJ/(kg K)
Density, ρ P C M (PCM)1.0 kg/m3
Specific heat at constant pressure, c P C M (PCM)4.20 kJ/(kg K)
Table 3. The time to achieve the target thermal energy storage density.
Table 3. The time to achieve the target thermal energy storage density.
TargetStatic System Agitated System Ultrasonic Vibrated System
100 rpm300 rpm600 rpm28 kHz56 kHz
100 MJ/m3
Achieved time/h
-5.12.41.55.85.8
140 MJ/m3
Achieved time/h
--5.92.9--
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kiyokawa, H.; Tokutomi, H.; Ishida, S.; Nishi, H.; Ohmura, R. Thermal Energy Storage Performance of Tetrabutylammonium Acrylate Hydrate as Phase Change Materials. Appl. Sci. 2021, 11, 4848. https://doi.org/10.3390/app11114848

AMA Style

Kiyokawa H, Tokutomi H, Ishida S, Nishi H, Ohmura R. Thermal Energy Storage Performance of Tetrabutylammonium Acrylate Hydrate as Phase Change Materials. Applied Sciences. 2021; 11(11):4848. https://doi.org/10.3390/app11114848

Chicago/Turabian Style

Kiyokawa, Hitoshi, Hiroki Tokutomi, Shinichi Ishida, Hiroaki Nishi, and Ryo Ohmura. 2021. "Thermal Energy Storage Performance of Tetrabutylammonium Acrylate Hydrate as Phase Change Materials" Applied Sciences 11, no. 11: 4848. https://doi.org/10.3390/app11114848

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