Next Article in Journal
Recent Progress of Energy-Storage-Device-Integrated Sensing Systems
Next Article in Special Issue
Sensor to Electronics Applications of Graphene Oxide through AZO Grafting
Previous Article in Journal
Enhancing Anticorrosion Resistance of Aluminum Alloys Using Femtosecond Laser-Based Surface Structuring and Coating
Previous Article in Special Issue
Design and Implementation of Graphene-Based Tunable Microwave Filter for THz Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study on Thermal Effect of Aluminum-Air Battery

1
School of Materials Science and Engineering, Tianjin Key Laboratory of Composite and Functional Materials, Key Laboratory of Advanced Ceramics and Machining Technology (Ministry of Education), Tianjin University, Tianjin 300072, China
2
Huadian Water Technology Co., Ltd., Beijing 100160, China
3
China Huadian Co., Ltd., Beijing 100031, China
4
School of Materials Science and Engineering, Kunming University of Science and Technology, Kunming 650093, China
5
Joint School of National University of Singapore and Tianjin University, International Campus of Tianjin University, Fuzhou 350207, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(4), 646; https://doi.org/10.3390/nano13040646
Submission received: 30 December 2022 / Revised: 28 January 2023 / Accepted: 4 February 2023 / Published: 6 February 2023
(This article belongs to the Special Issue Current Advances in Nanoelectronics, Nanosensors and Devices)

Abstract

:
The heat released from an aluminum-air battery has a great effect on its performance and operating life during the discharge process. A theoretical model was proposed to evaluate the resulting thermal effect, and the generated heat was divided into the following sources: anodic aluminum oxidation reaction, cathodic oxygen reduction reaction, heat production against the battery internal resistance, and hydrogen-evolution reaction. Quantitative analysis was conducted on each part, showing that all heat production sources increased with discharge current density. It should be noted that the heat caused by hydrogen evolution accounted for the most, up to 90%. Furthermore, the regulation strategy for inhibiting hydrogen evolution was developed by addition of hybrid additives to the electrolyte, and the hydrogen-evolution rate was greatly reduced by more than 50% as was the generated heat. This research has important guidance for the thermal effect analysis of aluminum–air batteries, together with control of the thermal management process by inhibiting hydrogen evolution, thus promoting their practical application.

1. Introduction

With rapid construction of clean energy networks and development of new energy industries such as photovoltaic, wind power, and electrical vehicles, modern society has put forward higher requirements for energy density and security of battery energy systems [1,2]. Among the newly proposed electrochemical devices, metal-air batteries are a promising alternative to lithium-based batteries. The most important feature of a metal–air battery is a metal anode coupling characterized by high-energy density, together with an open-structure-catalyzed cathode that continuously draws oxygen from the atmosphere. This feature is responsible for the intrinsically high theoretical energy density generally associated with this type of device [3]. Aqueous aluminum-air batteries show great potential due to their high theoretical electrochemical capacity (2980 mAh·g−1 [4,5]) and specific energy (8100 Wh·kg−1 [6,7]), the richest Earth reserves with low cost (8.21 wt. %) [8], and intrinsic safety (noncombustible). Therefore, aluminum-air batteries have been recognized as an attractive candidate for next-generation batteries [5]. The choice of electrolytes used in aluminum-air batteries is fairly flexible. The electrolyte can be common salt (NaCl), sea water, or alkaline electrolytes such as sodium hydroxide (NaOH) [9,10]. Alkaline aqueous solution was usually applied as its electrolyte to achieve high conductivity and to eliminate surface passivation of the metal anode [11]. The corresponding discharge reactions consist of an aluminum oxidation reaction at the anode, an oxygen reduction reaction (ORR) at the cathode, and a side reaction of hydrogen evolution. The specific reaction equations (Equations (1)–(4)) follow:
Anode   reaction :   A l 3 e A l 3 +
Cathode   reaction :   O 2 + 2 H 2 O + 4 e 4 O H
Total   reaction : 4 A l + 3 O 2 + 6 H 2 O = 4 A l ( O H ) 3
Side   reaction : A l + 3 H 2 O + O H = 3 / 2 H 2 + A l ( O H ) 4
The thermal effect of batteries could significantly affect their performance and operating life, resulting in serious safety issues, such as electrolyte decomposition [12] and thermal runaway of the battery [13,14]. Therefore, it is necessary to evaluate the thermal characteristics, and a series of related research has been conducted on lithium batteries. Generally, the heat generated by a lithium-ion battery could be divided into four parts consisting of reaction heat induced by Li+ transfer (QR), joule heat against ohmic internal resistance (QJ), polarization heat from the electrode polarization effect (QP), and side-reaction heat caused by the self-discharge reaction of the battery (QS) [15,16,17]. Further, the proportion of the abovementioned parts was investigated at different discharge rates [18], which indicated that the heat production of the battery was mainly joule heat at high discharge rates, while reaction heat accounted for the majority at low rates. Furthermore, a heat generation model of lithium-ion batteries was established by coupling electrochemical principles to verify the accuracy of the model through the temperature rise under different charge and discharge rates [19]. The thermal behavior of bagged lithium-ion batteries at different charging and discharging rates (0.5-1-2 C) was studied at operating temperatures of 30, 40, and 54 °C [20]. It was shown that the rates of charge and discharge inversely affect battery performance. In addition, the heat generation equation of the battery system was also proposed according to the law of energy conservation, which has been used to predict battery discharge performance at different operating temperatures [21]. Various studies have been carried out to explore the heating mechanism of lithium-ion batteries and established the relationship between performance and operating temperature. Obviously, evaluating the thermal effect of aluminum-air batteries is crucial for their practical application. However, there has been no relevant study on aluminum-air batteries until now.
Owing to the similar discharge principle of a metal battery, an evaluation model for the thermal effect of an aluminum–air battery was proposed, referring to lithium-ion battery [15]. The corresponding schematic diagram is exhibited in Figure 1, including QR (referring to the discharge reaction consisting of aluminum oxidation at the anode, QAl [18], and the oxygen reduction reaction at the cathode, QORR [22]), QJ (joule heat against ohmic internal resistance of the aluminum-air battery) [18,22], QP (polarization heat from the electrode polarization effect) [23], and QH2 [22] (referring to the side reaction of hydrogen evolution). The specific values for each part and their changing trends with discharge rates were evaluated quantitatively in this work, conducive to establishing a certain data model and theoretical basis for the thermal effect of aluminum-air batteries.
Moreover, the strategy for solving thermal runaway of the battery was also provided to optimize the thermal management process and subsequently promote large-scale commercial application of aluminum-air batteries. Currently, two main strategies have been reported to reduce thermal runaway in batteries. The first strategy is aluminum anode alloying. Alloying aluminum with specific elements is one of the effective ways to reduce the oxidation overpotential of aluminum in an alkaline medium, inhibit the rate of parasitic corrosion, and reduce thermal runaway. For example, adding Sb [24], Mg [25], Mn [26], or Li [27] to pure aluminum alloy with high hydrogen overpotential can inhibit hydrogen evolution, thus reducing thermal runaway of the battery and improving the safety of the battery. However, most alloying elements occur as a solid solution in aluminum, which performs an activation or corrosion inhibition role. The electrochemical performance of the aluminum anode will be affected if the alloying elements precipitate and form a second phase at the grain boundary. The other strategy is to control the side reaction of the aluminum-air battery by electrolyte modification, so as to control the thermal runaway problem. For example, using polyvinyl alcohol (PVA) as the anode liquid and neutral salt xanthan gum as the cathode liquid, dual electrolyte (DE) aluminum-air batteries were prepared without a separator. In this system, the hydrogen-evolution reaction can be reduced by decreasing the water content of the anode, thus controlling the thermal runaway of the battery [28]. We modulated the thermal effect by adding additives directly to the aluminum-air battery electrolyte, which is a more common strategy for scaled applications.

2. Experimental

2.1. Construction of Thermal Effect Test System

A self-designed thermal effect test system was assembled as shown in Figure 2. The experimental environment temperature was controlled at 25 °C ± 1 °C. The battery was subjected to constant-current discharge tests using a Neware battery test system (CT-3008-5V3A-164, Neware Technology Limited, Shenzhen, China). The distance between the anode (pure aluminum, 99.9 wt. %) and air electrode consisting of commercial MnO2/C catalyst is 1 cm with 4 M NaOH solution as the electrolyte [28,29]. Prior to testing, the aluminum electrode, with an active area of 50 mm × 50 mm, was abraded with silicon carbide paper and then rinsed with deionized water.
The discharge current densities were 2.5, 5, 10, 15, and 20 mA cm−2, with discharge time of 1 h. The total heat released (Qtol) during the discharge process was assessed directly by a temperature measuring device (test accuracy of ±0.1°C), while QJ and QP were determined by an internal resistance test device (M2111, Smacq, Beijing, China) and QH2 by a hydrogen-evolution test device. In addition, the sample was weighed to obtain weight loss before and after discharge.

2.2. Hydrogen-Evolution Test

The hydrogen-evolution rate was calculated by collecting H2 gas volume at different discharge densities by the water drainage method. All the measurements were repeated at least three times in order to achieve convincing results. The equation used for calculating the hydrogen-evolution rate (Equation (5)) was obtained by referring to previous studies [30]:
R = V H 2 A × t
where R is the hydrogen-evolution rate (mL·cm−2·h−1), VH2 is the volume of hydrogen (mL), A is the working area of the aluminum anode (cm2); and t is discharge time (1 h).

2.3. Internal Resistance Test of Battery

The internal resistance of the battery (Rint) consisted of ohmic resistance (Rohm) and polarization resistance (Rp), based on previous research by Pastor-Fernandez et al. [30]. To evaluate the values of QJ and QP, the Rint of the aluminum-air battery in this work was measured by a battery internal resistance meter (RC3563, Hopetcsh, Nanjing, China) with a sampling accuracy of 0.01 Ω.

2.4. Specific Heat Capacity Test of the Electrolyte

Differential scanning calorimetry (Q2000, TA, New Castle, DE, USA) was used to determine the specific heat capacity of the electrolyte at different temperatures. The test was conducted within the temperature range 16−90 °C with a heating rate of 5 °C/min and a sampling accuracy of 0.1 °C.

2.5. Electrolyte Additives

Hybrid additives were introduced into the electrolyte to regulate the hydrogen-evolution reaction and the thermal effect of the battery. Dodecyl trimethyl ammonium bromide (DTAB, 0.16 M) and ZnO (0.2 M) were added to the basic electrolyte of 4 M NaOH.

3. Results and Discussion

3.1. Composition of Heat Release during the Discharge Process

3.1.1. Qtol

The total heat release (Qtol) of the aluminum-air battery during the discharge process could be directly measured by monitoring temperature change (∆T) during discharge, and then calculated by Equation (6):
Q t o l = c   m   Δ T
where c is the specific heat capacity of battery electrolyte (J/g·°C), m is the mass of electrolyte (g), and ∆T is the temperature difference of electrolyte before and after discharging for 1 h (°C). The formula shows that the total heat release depends on the specific heat capacity of the battery electrolyte and the temperature difference between the charge−discharge processes for the aluminum-air battery. It should be noted that the specific heat capacity of the electrolyte varied with temperature, and thus the integral form of Equation (6) could be expressed as Equation (7). T1 and T2 corresponded to the electrolyte temperature before and after discharge, respectively.
Q t o l = T 1 T 2 c   m   dT
According to the thermal effect theory of a lithium-ion battery [15,31], Qtol of the aluminum-air battery can also be expressed as Equation (8):
Q t o l = Q R + Q J + Q P + Q H 2

3.1.2. QR

QR refers to the heat released from the discharge reaction, which consists of heat generated by the aluminum oxidation reaction (QAl) and the oxygen reduction reaction (QORR). The calculation of QAl is mainly based on the enthalpy change in the aluminum electrode reaction, and the corresponding reaction formula and enthalpy change are expressed as Equations (9) and (10), respectively:
A l 3 e A l 3 +
Δ r H m θ ( 298   K , A l ) = Δ f H m θ [ A l 3 + ] Δ f H m θ [ A l ]
According to Faraday’s Law, the amount of electricity consumed can be used to represent the amount of reactant nr [32]:
n r = I t / F
where I is the current density of the battery discharge (A) and F is the Faraday constant (96,485 C·mol−1). However, the change in molar enthalpy is no longer a fixed value during reaction process. According to Kirchhoff’s formula [33], the change in enthalpy for the reaction at different temperatures can be obtained by using Equation (12):
Δ r H m θ ( T , A l ) = Δ r H m θ ( 298   K , A l ) + 298 T   V B Δ C p , m A l ( B ) d T
where VB is the stoichiometric number and Δ C p , m A l ( B ) is the specific heat capacity at constant pressure of the substance. The specific value can be found in Table 1. Equation (12) can then be written as follows:
Δ r H m θ ( T ) = Δ r H m θ ( 298   K , A l ) + Δ C p , m A l   ( T 298 )
where Δ C p , m A l ( B ) can be expressed as Equation (14):
Δ C p , m A l = C p , m [ A l 3 + ] C p , m [ A l ]
QAl can be expressed as Equation (15) [14]:
Q A l = n r   Δ r H m θ ( T )
Q A l = I t / F · { Δ f H m θ [ A l 3 + ] Δ f H m θ [ A l ] + ( C p , m [ A l 3 + ] C p , m [ A l ] )   ( T 298 ) }
Similarly, the QORR calculation is also based on the enthalpy change in ORR at the cathode, and the cathodic reaction and standard enthalpy change are shown as Equations (17) and (18), respectively:
O 2 + 2 H 2 O + 4 e 4 O H  
Δ r H m θ ( 298   K , O R R ) = 4   Δ f H m θ [ [ O H ] Δ f H m θ [ O 2 ] 2   Δ f H m θ [ H 2 O ] ]
The enthalpy change at different temperatures can be expressed as Equation (19):
Δ r H m θ ( T , O R R ) = Δ r H m θ ( 298   K , O R R ) + Δ C p , m O R R   ( T 298 )
where Δ C p , m O R R can be expressed as Equation (20):
Δ C p , m O R R = 4   Δ f H m θ [ [ O H ] Δ f H m θ [ O 2 ] 2   Δ f H m θ [ H 2 O ] ]
QORR can be expressed as:
Q O R R = n r   Δ r H m θ ( T )
Q O R R = I t / F · { 4   Δ f H m θ [ [ O H ] Δ f H m θ [ O 2 ] 2   Δ f H m θ [ H 2 O ] ] + ( 4   Δ f H m θ [ [ O H ] Δ f H m θ [ O 2 ] 2   Δ f H m θ [ H 2 O ] ] ) ( T 298 ) }

3.1.3. QJ and QP

QJ is the heat generated by ohm resistance when current flows through the aluminum-air battery, which is subject to Joule’s law and can be expressed as Equation (23):
Q J = I 2 R o h m
QP is the polarization heat induced by the electrode polarization effect, and the specific value can be achieved by Equation (24):
Q P = I 2 R P
According to Pastor-Fernandez et al. [30], it can be shown that the internal resistance of the battery (Rint) consists of ohmic resistance (Rohm) and polarization resistance (Rp). Additionally, Rp always changes dynamically and is difficult to monitor; thus, Rint is measured directly to investigate the sum of QJ and QP instead of quantitative analysis of them separately. The corresponding Qint is calculated by Equation (25) [34,35]:
Q i n t = I 2 R i n t

3.1.4. QH2

The side reaction of the aluminum–air battery is mainly a process of hydrogen evolution and QH2 is also calculated based on the enthalpy change in the hydrogen-evolution reaction. The detailed reaction formula and standard enthalpy change are expressed as Equations (26)–(28):
A l + 3 H 2 O + O H = 3 / 2 H 2 + A l ( O H ) 4
Δ r H m θ ( 298   K , H 2 ) = 3 2 Δ f H m θ [ H 2 ] + Δ f H m θ [ A l ( O H ) 4 ] Δ f H m θ [ A l ] 3   Δ f H m θ [ H 2 O ] Δ f H m θ [ O H ]
Δ r H m θ ( T , H 2 ) = Δ r H m θ ( 298   K ) + Δ C p , m H 2   ( T 298 )
where Δ C p , m H 2 can be expressed as Equation (29):
Δ C p , m H 2 = 3 / 2 C p , m [ H 2 ] + C p , m [ A l ( O H ) 4 ] 3 C p , m [ H 2 O ] C p , m [ A l ] C p , m [ O H ]
The amount of H2 (mol) produced in unit time t (h) is obtained by Equation (30):
n H 2 = V H 2 / 22.4
The heat generated by the hydrogen-evolution reaction, QH2, is expressed by Equation (31):
Q H 2 = Δ r H m θ ( T , H 2 ) · V H 2 / 22.4

3.1.5. QA

At the aluminum anode, the discharge reaction occurs simultaneously with the side reaction of hydrogen evolution, and they both consumed electrons released by the dissolution of aluminum electrode. The heat generated by the aluminum anode is introduced as QA, which obviously consists of QAl and QH2. It can be determined by the mass-loss method [35,36] and calculated by Equation (15) with nr referred with nA, based on Equation (32):
n A = m A l   / M A l
where mAl is recorded as the weight loss before and after the discharge reaction (g), and MAl is the molar mass of the aluminum electrode (g/mol). Therefore, the aluminum anode heat production QA can also be expressed as Equation (33):
Q A = Q A l + Q H 2

3.2. Quantitative Analysis for Heat Generated by Each Part

The discharge curves for the aluminum-air battery at different discharge current densities for 1 h are shown in Figure 3. It can be seen that the discharge voltage decreases with the increase in current density. The heat generation results for each part of the aluminum-air battery were obtained by discharging at different current densities.
QR reflects the heat generated by the discharge reaction of the aluminum-air battery, consisting of QAl and QORR; the specific values are listed in Table 2 and exhibited in Figure 4. It can be found that QAl, QORR, and QR all increased linearly with discharge current density. Obviously, fast electron transmission with a raised discharge rate was accompanied by more heat release.
Qint was introduced to directly characterize the sum of QJ and QP. The internal resistance (Rint) of the battery was measured at different discharge densities, as shown in Figure 5a, and the corresponding values of Qint are exhibited in Figure 5b. It can be observed that Rint increased slightly (from 0.35 and 0.4) with current density, exhibiting little change with discharge time. Therefore, the values of Qint calculated by Equation (25) are mainly determined by discharge current density.
QH2 is the heat released by the hydrogen-evolution reaction, and the corresponding evolution rate of hydrogen from the aluminum electrode was recorded during the discharge process (Figure 6a). The hydrogen-evolution rate maintained an upward tendency to 1400 mL/h with a current density of 20 mA/cm2. The increasing hydrogen-evolution rate slowed when the current density was greater than 10 mA/cm2. This observation is mainly attributed to the increase in hydrogen-evolution potential induced by aluminum electrode polarization and the resulting inhibition effect on the hydrogen-evolution reaction [25]. In accordance with Equations (28) and (31), the values for QH2 also exhibited a similar trend, as shown in Figure 6b.
The heat generated by the aluminum anode reaction is given as QA, which consists of QAl and QH2 and related to the heat from the discharge reaction of the aluminum electrode and the hydrogen-evolution reaction, respectively. By measuring the weight loss of the aluminum anode after discharge process (Table 3), QA can also be directly obtained based on Equations (15) and (32), indicating good consistency with that calculated by summing QAl and QH2, as shown in Figure 7a. Subsequently, the proportion of QAl and QH2 was further analyzed quantitatively (Figure 7b), showing that QH2 accounted for most of the heat, up to 95%. Additionally, the ratio (QH2/QA) decreased gradually with discharge current density, but remained greater than 80%.
According to Equation (8) and relevant analysis of QR, QJ, and QP, Qtol can be expressed by Equation (34):
Q t o l = Q A l + Q O R R + Q i n t + Q H 2
The value of Qtol can also be directly calculated based on Equation (7) by monitoring the temperature change (∆T listed in Table 3), which was continuously recorded during the discharge process (Figure 8a), and the specific heat capacity of electrolyte (Figure 8b). This experimental value is also in good agreement with the calculated value according to Equation (34), as illustrated in Figure 8c. It can be seen clearly that the temperature of the aluminum-air battery increased with discharge current density, as shown in Figure 7a. When the aluminum-air battery was discharged at 2.5 mA/cm2, the temperature increased by 28.6 °C. When the aluminum-air battery was discharged at 5.0 mA/cm2, the temperature increased by 36.3 °C. When the aluminum-air battery was discharged at 10.0 mA/cm2, the temperature increased by 44.3 °C. At 15.0 mA/cm2 discharge, the temperature increased by 47.8 °C. At 20.0 mA/cm2 discharge, the temperature increased by 56.5 °C. Thus, the larger the current density, the higher the temperature rise of the aluminum-air battery. It should be further noted that the temperature rising rate also accelerated gradually. The larger the discharge current density, the faster the temperature rose, and more easily the battery became unstable in a short period, resulting in a thermal runaway of the battery [33]. The specific values of QAl, Qint, QORR, QH2, and Qtol are listed in Table 2, and the relevant proportions of each part are shown in Figure 8. It can be shown that QH2 accounted for more than 60%, especially up to 90% at relatively lower discharge rate (2.5 mA/cm2), resulting in the enhanced heat release. In conclusion, QH2 played the dominant role in the thermal effect during the discharge process, and effective inhibition of the hydrogen-evolution reaction is essential to avoid the thermal runaway of the aluminum-air battery.

3.3. Regulating the Thermal Effect of the Aluminum–Air Battery

It can be seen from Figure 9a that hydrogen-evolution heat generation accounted for a very large amount of the overall heat generation. A regulation strategy for inhibiting the hydrogen-evolution reaction was proposed by means of hybrid additives (0.16 M DTAB and 0.2 M ZnO) to the electrolyte. The corresponding values of R’H2, Q’H2, and Q’tol are listed in Table 4 and then compared with that free from additives, as shown in Figure 10a,b. It can be seen clearly that the hydrogen-evolution rate (R’H2) of the aluminum–air battery was greatly reduced by more than 50% under different discharge current densities, and the reduction in the hydrogen-evolution rate increased with the increase in current density. When the discharge current density was 20 mA/cm−2, the hydrogen-evolution rate decreased by 58.4%. Hydrogen-evolution heat production was also greatly reduced by more than 50%; the maximum could be reduced by 53.8%, and the overall heat production was also reduced by more than 30%, the maximum could be reduced by 44.2%. The results showed that the hybrid addition of DTAB and ZnO could effectively reduce the heat production of the hydrogen-evolution reaction and the overall heat production, which could effectively improve the stability of the battery. In addition, the detailed proportional analyses of heat generation for each part are shown in Figure 10c. The heat-generation ratio of the hydrogen-evolution reaction decreased significantly, and the decrease in heat generation also increased with the increase in current density. This also provides an effective strategy to regulate the thermal effect and promotes the practical application of aluminum-air batteries.

4. Conclusions

(1)
The thermal effect of an aluminum-air battery was investigated to provide a theoretical model by combining experimental measurement and mathematical calculation. The generated heat during the discharge process (Qtol) was divided into the following parts: aluminum oxidation reaction heat at the anode (QAl), oxygen reduction reaction at the cathode (QORR), heat production against the battery’s internal resistance (Qint, sum of QJ and QP), and the hydrogen-evolution heat (QH2).
(2)
The heat of each part noted above was quantitatively analyzed with different discharge rates. It could be shown that all heat production increased with the increase in current density, accelerating the rise in temperature of the battery. Obviously, the battery became more unstable at high discharge rates, resulting in thermal runaway in a short period.
(3)
Hydrogen-evolution heat (QH2) accounted for most of the total heat, more than 60% and up to 90% at the low discharge rate of 2.5 mA/cm2. It can be concluded that QH2 plays the dominant role in thermal effect during the discharge process of the battery. The hydrogen-evolution reaction should be inhibited to avoid thermal runaway of the battery.
(4)
Hybrid additives of DTAB and ZnO were introduced into the electrolyte to inhibit the hydrogen-evolution reaction at the aluminum electrode. The hydrogen-evolution rate was greatly reduced, lowering its contributed heat by more than 50% and its portion in Qtol. This also provides an effective way to solve the crucial challenge of battery thermal runaway.

Author Contributions

Conceptualization, Z.W. and Y.C.; methodology, Y.T.; software, Y.L. (Yingjie Liu); validation, Z.W., Y.C., and X.L.; formal analysis, Y.T.; investigation, B.Z.; resources, Y.L. (Yichun Liu); data curation, Z.W. and B.C.; writing—original draft preparation, Y.C.; writing—review and editing, Z.W.; visualization, Z.Q.; project administration, F.L. and W.H.; funding acquisition, Z.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (No. 52171078, 51971155) and Natural Science Foundation of Tianjin (No. 21JCZDJC00440).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hopkins, B.J.; Shao-Horn, Y.; Hart, D.P. Suppressing corrosion in primary aluminum air batteries via oil displacement. Science 2018, 362, 658–661. [Google Scholar] [CrossRef]
  2. Luan, J.; Zhang, Q.; Yuan, H.; Sun, D.; Peng, Z.; Tang, Y.; Ji, X.; Wang, H. Plasmalogtrengthened lithiophilicity of copper oxide nanosheet-decorated Cu foil for stable lithium metal anode. Adv. Sci. 2019, 6, 1901433. [Google Scholar] [CrossRef]
  3. Wu, Z.; Zhang, H.; Zou, J.; Shen, X.; Qin, K.; Ban, C.; Cui, J.; Nagaumi, H. Enhancement of the discharge performance of Al-0.5Mg-0.1Sn-0.05Ga (wt.%) anode for Al-air battery by directional solidification technique and subsequent rolling process. J. Alloys Compd. 2020, 827, 154272. [Google Scholar] [CrossRef]
  4. Li, F.; Li, J.; Feng, Q.; Yan, J.; Tang, Y.; Wang, H. Significantly enhanced oxygen reduction activity of Cu/CuNxCy co-decorated ketjenblack catalyst for Al-air batteries. J. Energy Chem. 2018, 27, 419–425. [Google Scholar] [CrossRef]
  5. Wang, H.; Leung, D.; Leung, M. Energy analysis of hydrogen and electricity production from aluminum-based processes. Appl. Energy 2012, 90, 100–105. [Google Scholar] [CrossRef]
  6. Driscoll, C.T.; Schecher, W.D. The chemistry of aluminum in the environment. Environ. Geochem. Health 1990, 12, 28–49. [Google Scholar] [CrossRef] [PubMed]
  7. Rahman, M.A.; Wang, X.J.; Wen, C.E. High Energy Density Metal-Air Batteries: A Review. J. Electrochem. Soc. 2013, 160, A1759. [Google Scholar] [CrossRef]
  8. Buckingham, R.; Asset, T.; Atanassov, P. Aluminum-air batteries: A review of alloys, electrolytes and design. J. Power Sources 2021, 498, 229762. [Google Scholar] [CrossRef]
  9. Mauger, A.; Julien, C.M.; Paolella, A.; Armand, M.; Zaghib, K. A comprehensive review of lithium salts and beyond for rechargeable batteries: Progress and perspectives. Mater. Sci. Eng. R Rep. 2018, 134, 1–21. [Google Scholar] [CrossRef]
  10. Nestoridi, M.; Pletcher, D.; Wood, R.; Wang, S.; Jones, R.L.; Stokes, K.R. The study of aluminium anodes for high power density al/air batteries with brine electrolytes. J. Power Sources 2007, 178, 445–455. [Google Scholar] [CrossRef] [Green Version]
  11. Wang, Y.F.; Kwok, H.Y.H.; Pan, W.D.; Zhang, H.M.; Lu, X.; Leung, D.Y.C. Parametric study and optimization of a low-cost paper-based Al-air battery with corrosion inhibition ability. Appl. Energy 2019, 251, 113342. [Google Scholar] [CrossRef]
  12. Shim, J.; Kostecki, R.; Richardson, T.; Song, X.; Striebel, K. Electrochemical analysis for cycle performance and capacity fading of a lithium-ion battery cycled at elevated temperature. J. Power Sources 2002, 112, 222–230. [Google Scholar] [CrossRef]
  13. Coman, P.T.; Darcy, E.C.; Veje, C.T.; White, R.E. Numerical analysis of heat propagation in a battery pack using a novel technology for triggering thermal runaway. Appl. Energy 2017, 203, 189–200. [Google Scholar] [CrossRef]
  14. Kvasha, A.; Gutiérrez, C.; Osa, U.; Meatza, I.D.; Blazquez, J.A.; Macicior, H.; Urdampilleta, I. A comparative study of thermal runaway of commercial lithium-ion cells. Energy 2018, 159, 547–557. [Google Scholar] [CrossRef]
  15. Sato, N. Thermal behavior analysis of lithium-ion batteries for electric and hybrid vehicles. J. Power Sources 2001, 99, 70–77. [Google Scholar] [CrossRef]
  16. Wang, Q.; Jiang, B.; Li, B.; Yan, Y.Y. A critical review of thermal management models and solutions of lithium-ion batteries for the development of pure electric vehicles. Renew. Sustain. Energy Rev. 2016, 64, 106–128. [Google Scholar] [CrossRef]
  17. Liu, H.; Wei, Z.; He, W.; Zhao, J. Thermal issues about li-ion batteries and recent progress in battery thermal management systems: A review. Energy Convers. Manag. 2017, 150, 304–330. [Google Scholar] [CrossRef]
  18. Jiang, F.M.; Peng, P.; Sun, Y.Q. Thermal analyses of LiFePO4 /graphite battery discharge processes. J. Power Sources 2013, 243, 181–194. [Google Scholar] [CrossRef]
  19. Pan, D.; Guo, H.; Tang, S.; Li, X.; Wang, Z.; Peng, W.; Wang, J.; Yan, G. Evaluating the accuracy of electro-thermal coupling model in lithium-ion battery via altering internal resistance acquisition methods. J. Power Sources 2020, 463, 228174. [Google Scholar] [CrossRef]
  20. Chatterjee, K.; Majumdar, P.; Schroeder, D.; Kilaparti, S.R. Performance analysis of Li-ion battery under various thermal and load conditions. J. Electrochem. Energy 2019, 16, 021006. [Google Scholar] [CrossRef]
  21. Bernadi, D.; Pawlikowski, E.; Newman, J. A general energy balance for battery systems. J. Electrochem. Soc. 1985, 132, 5–12. [Google Scholar] [CrossRef]
  22. Chen, L.D.; Nørskov, J.K.; Luntz, A.C. Al-air batteries: Fundamental thermodynamic limitations from first-principles theory. J. Phys. Chem. Lett. 2014, 6, 175–179. [Google Scholar] [CrossRef]
  23. Nyman, A.; Zavalis, T.G.; Elger, R.; Behm, M.; Lindbergh, G. Analysis of the Polarization in a Li-Ion Battery Cell by Numerical Simulations. J. Electrochem. Soc. 2010, 157, A1236–A1246. [Google Scholar] [CrossRef]
  24. Zhang, P.; Liu, X.; Xue, J.; Wang, Z. Evaluating the discharge performance of heat-treated Al-Sb alloys for al-air batteries. J. Mater. Eng. Perform. 2019, 28, 5476–5484. [Google Scholar] [CrossRef]
  25. Pk, A.; Vm, B.; Cp, B.; Ga, A.; Pl, C. Study of some basic operation conditions of an al-air battery using technical grade commercial aluminum. J. Power Sources 2020, 450, 227624. [Google Scholar]
  26. Wang, D.; Li, H.; Liu, J.; Zhang, D.; Gao, L.; Tong, L. Evaluation of AA5052 alloy anode in alkaline electrolyte with organic rare-earth complex additives for aluminium-air batteries. J. Power Sources 2015, 293, 484–491. [Google Scholar] [CrossRef]
  27. Xiong, H.; Wang, Z.; Yu, H.; Chen, T.; Ye, L. Performances of al-xli alloy anodes for al-air batteries in alkaline electrolyte. J. Alloys Compd. 2021, 889, 161677. [Google Scholar] [CrossRef]
  28. Gaele, M.F.; Migliardini, F.; Palma, T.M.D. Dual solid electrolytes for aluminium-air batteries based on polyvinyl alcohol acidic membranes and neutral hydrogels. J. Solid State Electrochem. 2021, 25, 1207–1216. [Google Scholar] [CrossRef]
  29. Wu, Z.; Zhang, H.; Yang, D.; Zou, J.; Qin, K.; Ban, C.; Cui, J.; Nagaumi, H. Electrochemical behaviour and discharge characteristics of an Al-Zn-In-Sn anode for Al-air batteries in an alkaline electrolyte. J. Alloys Compd. 2020, 837, 155599. [Google Scholar] [CrossRef]
  30. Pastor-Fernandez, C.; Uddin, K.; Chouchelamane, G.H.; Widanage, W.D.; Marco, J. Dataset to support: “A comparison between electrochemical impedance spectroscopy and incremental capacity-differential voltage as li-ion diagnostic techniques to identify and quantify the effects of degradation modes within battery management systems”. J. Power Sources 2017, 360, 301–318. [Google Scholar] [CrossRef]
  31. Yang, X.G.; Zhang, G.; Wang, C.Y. Computational design and refinement of self-heating lithium-ion batteries. J. Power Sources 2016, 328, 203–211. [Google Scholar] [CrossRef]
  32. Smith, J.M.; Ness, V.; Abbott, M.M. Introduction to Chemical Engineering Thermodynamics; McGraw-Hill: New York, NY, USA, 2001; pp. 476–509. [Google Scholar]
  33. Fan, Y.X.; Yan, G.; Zou, H.Q.; Fu, J. Development of real-time simulation application software for four-quadrant converter system based on matlab. Int. J. Softw. Eng. Knowl. Eng. 2018, 28, 523–535. [Google Scholar] [CrossRef]
  34. Wang, D.P.; Zhang, D.Q.; Lee, K.Y.; Gao, L.X. Performance of AA5052 alloy anode in alkaline ethylene glycol electrolyte with dicarboxylic acids additives for aluminium air batteries. J. Power Sources 2015, 297, 464–471. [Google Scholar] [CrossRef]
  35. Jingling, M.; Jiuba, W.; Hongxi, Z.; Quanan, L. Electrochemical performances of Al–0.5Mg–0.1Sn–0.02In alloy in different solutions for Al–air battery. J. Power Sources 2015, 293, 592–598. [Google Scholar] [CrossRef]
  36. Ryu, J.; Park, M.; Cho, J. Advanced technologies for high-energy aluminum-air batteries. Adv. Mater. 2019, 31, 1804784. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram showing the evaluation model for the thermal effect of an aluminum-air battery.
Figure 1. Schematic diagram showing the evaluation model for the thermal effect of an aluminum-air battery.
Nanomaterials 13 00646 g001
Figure 2. Schematic diagram (a) and optical photo (b) of the aluminum-air battery structure and thermal effect test system.
Figure 2. Schematic diagram (a) and optical photo (b) of the aluminum-air battery structure and thermal effect test system.
Nanomaterials 13 00646 g002
Figure 3. The discharge curves for the aluminum-air battery at different discharge current densities in 4 M NaOH electrolyte for 1 h.
Figure 3. The discharge curves for the aluminum-air battery at different discharge current densities in 4 M NaOH electrolyte for 1 h.
Nanomaterials 13 00646 g003
Figure 4. Calculation results for QAl, QORR, and QR with different discharge current densities.
Figure 4. Calculation results for QAl, QORR, and QR with different discharge current densities.
Nanomaterials 13 00646 g004
Figure 5. (a) Rint for the aluminum–air battery at different current densities with discharge time and (b) Qint with different discharge current densities.
Figure 5. (a) Rint for the aluminum–air battery at different current densities with discharge time and (b) Qint with different discharge current densities.
Nanomaterials 13 00646 g005
Figure 6. (a) Hydrogen-evolution rate and (b) QH2 with different discharge current densities.
Figure 6. (a) Hydrogen-evolution rate and (b) QH2 with different discharge current densities.
Nanomaterials 13 00646 g006
Figure 7. (a) QA with different discharge current densities measured by direct and indirect methods and (b) the proportion of QAl and QH2 in QA of the aluminum–air battery.
Figure 7. (a) QA with different discharge current densities measured by direct and indirect methods and (b) the proportion of QAl and QH2 in QA of the aluminum–air battery.
Nanomaterials 13 00646 g007
Figure 8. (a) The change in temperature of the aluminum–air battery with discharge time, (b) the specific heat capacity of the electrolyte with temperature, and (c) Qtol with different discharge current densities obtained by direct and indirect methods.
Figure 8. (a) The change in temperature of the aluminum–air battery with discharge time, (b) the specific heat capacity of the electrolyte with temperature, and (c) Qtol with different discharge current densities obtained by direct and indirect methods.
Nanomaterials 13 00646 g008
Figure 9. The proportion of QAl, QORR, Qint, and QH2 in Qtol under different discharge current densities.
Figure 9. The proportion of QAl, QORR, Qint, and QH2 in Qtol under different discharge current densities.
Nanomaterials 13 00646 g009
Figure 10. Comparison of (a) heat generated by hydrogen-evolution reaction, (b) total heat, and (c) the proportion of each part with modification by electrolyte additives.
Figure 10. Comparison of (a) heat generated by hydrogen-evolution reaction, (b) total heat, and (c) the proportion of each part with modification by electrolyte additives.
Nanomaterials 13 00646 g010
Table 1. Thermodynamic data for ∆fHmθ (298 K) and ∆Cp,m (B) at 298.15 K.
Table 1. Thermodynamic data for ∆fHmθ (298 K) and ∆Cp,m (B) at 298.15 K.
SubstancefHmθ (298 K) (kJ·mol−1)Cp,m (J·deg−1·mol−1)
Al024.4
H2O−285.8337.11
OH−230.015−148.5
H2028.84
Al(OH)4−1502.55.12 *
Al3+−538.43.98 *
O200.92
* experimental data.
Table 2. Heat production for each part at different discharge current densities.
Table 2. Heat production for each part at different discharge current densities.
Current Density/mA·cm−1QR/kJQAl/kJQORR/kJQH2/kJQtol/kJQint/kJQA/kJQAl + QH2/kJQAl + QORR QH2 + Qint/kJ
2.51.230.420.8111.13 ± 0.8813.38 ± 0.090.00711.03 ± 0.2011.55 ± 0.8812.36 ± 0.88
52.460.841.6214.25 ± 0.6416.98 ± 0.120.02415.31 ± 0.1615.09 ± 0.6416.74 ± 0.64
104.931.683.2515.52 ± 1.0521.01 ± 0.060.1017.24 ± 0.1717.20 ± 1.0520.55 ± 1.05
157.382.524.8615.92 ± 0.8122.36 ± 0.180.2118.78 ± 0.2418.44 ± 0.8123.51 ± 0.81
209.863.366.5016.45 ± 0.9126.43 ± 0.090.3519.78 ± 0.1319.81 ± 0.9126.66 ± 0.91
Table 3. The hydrogen-evolution rate (RH2), internal resistance (Rint), temperature change (∆T), and weight loss (∆m) of aluminum anode at different discharge current densities.
Table 3. The hydrogen-evolution rate (RH2), internal resistance (Rint), temperature change (∆T), and weight loss (∆m) of aluminum anode at different discharge current densities.
Current Density/mA·cm−12.55101520
RH2/mL·cm−2·h−1947.90 ± 81.301168.72 ± 78.411257.48 ± 52.741289.40 ± 50.841332.20 ± 54.52
Rint0.3752 ± 0.01210.3982 ± 0.01240.4055 ± 0.00740.4138 ± 0.01340.5063 ± 0.0296
∆T/°C28.62 ± 2.1036.35 ± 1.3244.90 ± 1.8947.80 ± 2.3156.50 ± 1.54
∆mAl/g0.55 ± 0.020.77 ± 0.010.87 ± 0.010.94 ± 0.030.99 ± 0.01
Table 4. The hydrogen-evolution reaction, the heat generated by the hydrogen-evolution reaction and total heat after regulating the hydrogen-evolution reaction.
Table 4. The hydrogen-evolution reaction, the heat generated by the hydrogen-evolution reaction and total heat after regulating the hydrogen-evolution reaction.
Current Density/mA cm−12.55101520
R’H2/mL·h−1445.23 ± 39.30478.76 ± 45.63522.74 ± 34.34552.40 ± 52.21610.23 ± 34.52
Q’H2/kJ6.43 ± 0.596.79 ± 0.437.16 ± 0.397.41 ± 0.5117.93 ± 0.38
Q’tol/kJ7.68 ± 0.149.34 ± 0.1812.23 ± 0.1015.02 ± 0.1418.17 ± 0.09
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cai, Y.; Tong, Y.; Liu, Y.; Li, X.; Chen, B.; Liu, F.; Zhou, B.; Liu, Y.; Qin, Z.; Wu, Z.; et al. Study on Thermal Effect of Aluminum-Air Battery. Nanomaterials 2023, 13, 646. https://doi.org/10.3390/nano13040646

AMA Style

Cai Y, Tong Y, Liu Y, Li X, Chen B, Liu F, Zhou B, Liu Y, Qin Z, Wu Z, et al. Study on Thermal Effect of Aluminum-Air Battery. Nanomaterials. 2023; 13(4):646. https://doi.org/10.3390/nano13040646

Chicago/Turabian Style

Cai, Yajun, Yunwei Tong, Yingjie Liu, Xinyu Li, Beiyang Chen, Feng Liu, Baowei Zhou, Yichun Liu, Zhenbo Qin, Zhong Wu, and et al. 2023. "Study on Thermal Effect of Aluminum-Air Battery" Nanomaterials 13, no. 4: 646. https://doi.org/10.3390/nano13040646

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