Next Article in Journal
Realization of Bio-Coal Injection into the Blast Furnace
Previous Article in Journal
Influence of Processing Routes to Enhance the Mechanical Properties of Mg–6Zn–1Y–3.5CeMM (wt.%) Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental and Mathematical Investigation of Hydrogen Absorption in LaNi5 and La0.7Ce0.1Ga0.3Ni5 Compounds

by
Sihem Belkhiria
1,*,
Abdulrahman Alsawi
2,*,
Ibtissem Hraiech
1,
Mohamed Houcine Dhaou
1,2 and
Abdelmajid Jemni
1
1
Laboratory of Thermal and Energetic Systems Studies, LR99ES31, National Engineering School, University of Monastir, Monastir 5019, Tunisia
2
Department of Physics, College of Science, Qassim University, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Metals 2024, 14(9), 967; https://doi.org/10.3390/met14090967
Submission received: 20 June 2024 / Revised: 24 July 2024 / Accepted: 23 August 2024 / Published: 27 August 2024

Abstract

:
In the present study, the hydrogen-absorption properties of the LaNi5 and the La0.7Ce0.1Ga0.3Ni5 compounds were determined and compared. This work is therefore divided into two parts: an experimental part that presents and discusses the kinetics and isotherms of hydrogen absorption in the two compounds at two different temperatures (298 K and 318 K). In addition, the temperature variations inside the hydride bed were determined. In the second section, the experimental isotherms were compared to a numerical model processed using statistical physics. Following that, thanks to the perfect agreement between the experimental data and the proposed model, the stereographic and energetic parameters associated with the hydrogen absorption reaction, such as the number of hydrogen atoms per receptor site (n1, n2), the densities of the sites (Nm1, Nm2), the half-saturation pressures (P1, P2) and the absorption energies (ΔE1, ΔE2) for each receptor site, were calculated. All of these parameters are acquired by making numerical adjustments to the experimental data. Thermodynamic functions, such as internal energy and Gibbs energy, which regulate the absorption process, were then identified using these parameters. For both compounds, all of the aforementioned were compared and discussed in relation to initial temperature and pressure. The results demonstrated that the hydrogen-storage properties in LaNi5 are enhanced by more than 30% of stored mass and kinetics when Ce and Ga are substituted at the La sites.

1. Introduction

The world is going through an unprecedentedly deep and complicated global energy crisis [1]. As a result, decarbonization, electrification, and renewable energy sources are major themes in the World Energy Outlook for 2024, signaling a clear shift in the energy industry [2]. Hydrogen stands out among the many potential remedies as a novel, renewable and extremely potent energy vector that makes it possible to produce clean and silent electricity when it is needed [3]. However, hydrogen has a very low volume density despite having a very-high-mass energy density (1 kg of hydrogen has the same energy as around 3 kg of oil). Therefore, before being used, it must be stored in a usable volume under secure conditions [4]. By the way, a key market segment for hydrogen energy is storage [5]. Although various methods have been explored for hydrogen storage, including gas and liquid forms [5,6], solid storage in the form of metal hydrides appears to be a very promising solution, which is particularly interesting for transport and stationary applications [7,8]. Moreover, among the numerous families of metal hydrides [9,10], the AB5 hydride possesses intriguing and alluring features for reversibly storing hydrogen, including simple activation [11], high mass storage capacity [12], good recyclability [13] and storage under standard pressure and temperature conditions [14]. LaNi5 is specifically recognized as a hydrogen prototype for AB5 intermetallic compounds. Furthermore, several of its derivatives were among the first to be used electrochemically in nickel–metal hydride (Ni-MH) batteries [15]. They have been used as the negative electrode in Ni-MH batteries [16]. However, there are two main technical challenges related to solid-state hydrogen storage in LaNi5 [17]: The first concern is the thermal management during absorption and desorption processes since M-H has limited thermal conductivity; as a result, bulk MH-bed systems frequently have a Heat Transfer System (HTS) built into them. Another concern is efficient and effective scaling up. Actually, the poor hydrogen storage density of LaNi5-based compounds explains why Li-ion batteries are more common in portable electronics and rechargeable batteries for electric and hybrid cars [16].
Thus, one of the key techniques used to improve the hydrogen-storage characteristics inside the LaNi5 compound is the partial substitution of La and/or Ni sites [18,19]. Thereby, recent and innovative research has shown that partial or total substitution can significantly improve the cycle stability and high-rate capacity of the electrodes [20]. For example, we have taken a peek at an experimental work conducted by Zhida et al. [21], whereby the kinetic performances and hydrogen storage abilities of LaNi5−xCox alloys (x = 0, 0.25, 0.50, and 0.75) were examined through experimentation. The findings demonstrated that when the Co content increases, the cell volume increases as well, lowering the equilibrium plateau pressure and promoting a more stable hydride phase. The addition of Co is also found to increase mass and reaction kinetics. In another research work envisaged by Jingjing L et al. [22], hydrogen storage in the LaNi5−xAlx alloy (x = 0, 0.25, 0.5) is studied experimentally. The research results demonstrated that Al substitution increases the stability of the alloy’s crystal lattice by limiting atomic migration during cycling and reducing the rate of lattice volume expansion during hydrogenation. This improves the cycling stability of Al-doped alloys by around 10%.
Changpeng W. et al. [23] examined a combination substitution at the La and Ni sites in more recent research. By adjusting the Al concentration, this study attempted to enhance the (La0.33Y0.67)5Ni18.1−xMn0.9Alx alloys’ overall electrochemical hydrogen-storage capabilities. The findings demonstrated that Al doping leads to a 69.80% improvement in desorption capacity and a decrease in equilibrium plateau pressure. Moreover, this study demonstrates the alloy’s potential (La0.33Y0.67)5Ni17.5Mn0.9Al0.6) as a high-performing Ni-MH battery anode (La0.33Y0.67)5Ni17.5Mn0.9Al0.6. Similarly, the (LaCeCa)1(NiMnAl)5 alloy series was examined in a study planned by Xuqi L et al. [13] under the framework of a combined multi-substitution. Experiments have revealed that the maximum hydrogen-storage capacity of the alloy series (LaCeCa)1(NiMnAl)5 is approximately 0.3 weight percent higher than that of LaNi5, despite the fact that structural analysis by X-ray revealed that the substituted intermetallic crystallizes in a structure similar to that of the LaNi5 compound (CaCu5 structure).
Thus, building on earlier research, we propose investigating the hydrogen storage properties of the compound La0.7Ce0.1Ga0.3Ni5 both experimentally and numerically in this paper. Generally, elemental substitution at La sites aims to improve the overall qualities of hydrogen storage, such as the mass, kinetics, and stability of the hydride produced, while also reducing the cost of alloy production. Consequently, selecting doping materials requires significant thought for practical use in hydrogen storage systems. Studies have indicated that partial substitution of Ce for La in LaNi5 results in simpler activation, an increase in equilibrium plateau pressure, and an increase in reaction kinetics [24,25,26]. Furthermore, metallic Ce is less expensive than La. Thus, substitution can lower costs and encourage widespread commercial adoption [27,28]. Likewise, the Ga element has been used in several studies as a dopant in hydrogen storage materials [29,30]. These materials showed optimal hydrogen storage capacities and significant stability of the formed hydride. Thus, in this study, a combined substitution on the sites of La using the elements Ce and Ga is proposed. This study involves comparing experimentally and numerically the hydrogen-storage properties of LaNi5 and La0.7Ce0.1Ga0.3Ni5. For this reason, two parts will be presented in this paper: an experimental part, for which the absorption kinetics and isotherms of the two compounds will be traced. Then, the experimental isotherms will be compared to a numerical model using the statistical physics approach [31]. This model must be used to access the many macroscopic and microscopic phases of the system by providing insight into the molecular processes involved in the absorption reaction [32]. The suggested model’s analytical expressions will be utilized to adjust the experimental isotherms pertaining to LaNi5 and La0.7Ce0.1Ga0.3Ni5 compounds at varying temperatures (298 K and 318 K). Consequently, for a receptor site i, two categories of parameters are inferred from the adjustment: (i) stereographic parameters, which include the number of hydrogen atoms per receptor site (ni) and the densities of each receptor site (Nmi). (ii) Energy parameters, which include the half-saturation pressures for each receptor site (Pi). Ultimately, these parameters will be used to compute the absorption energies ∆Ei, the concentrations of hydrogen per unit of metal [H/M]i, and the thermodynamic functions governing the absorption.
Despite the various studies that have been carried out on these alloys, no numerical study using statistical physics has been considered to better describe and interpret the effect of substitution with the elements Ce and Ga on the Ni sites. Therefore, this research aims to provide new molecular investigations into the doping effect through theoretical work. The selected model was used to calculate the thermodynamic functions that control the hydrogen-storage process as well as the stereographic features that interfere with it.

2. Experimental Setup

The experimental device used is a Sievert apparatus (Figure 1a). It consists of a cylindrical metal hydride reactor M-H-R (Figure 1b) equipped with internal radial fins and connected to a gaseous hydrogen line via connection valves. A cold water bath is used to feed the hydride bed with cooling water during the absorption reaction through a helical heat exchanger. The reactor was filled with 30 g of purchased metal. A pressure sensor is installed between the reactor and the hydrogen tank, controlling the variation in hydrogen pressure during absorption processes. As well, a thermocouple is installed inside the reactor, controlling the internal temperature during absorption. These sensors are installed on an acquisition card and a microcomputer [33].

3. Results

3.1. Experimental Results

  • Absorption kinetics
Hydrogen absorption reaction kinetics refers to the rate at which hydrogen is absorbed by a material. Therefore, the task at hand is to depict how the mass of hydrogen absorbed changes over time. Figure 2 presents the variation of the mass of hydrogen absorbed as a function of time at T = 298 K and at T = 318 K for the LaNi5 and the La0.7Ce0.1Ga0.3Ni5 compounds.
Firstly, we observe that the mass of hydrogen steadily rises with time at a specific temperature until it reaches equilibrium. It is shown that, at a given temperature, the doped compound reaches equilibrium faster than LaNi5 with more stored hydrogen. This means that doping with Ce and Ga on the La sites improves the absorption kinetics and mass. We also point out that raising the initial temperature harms the absorption response by decreasing the mass of hydrogen absorbed and the absorption kinetics. For example, at T = 298 K, the mass of hydrogen absorbed at equilibrium for the LaNi5 is 7 g of hydrogen atoms, while at T = 318 K, it is 5 g of hydrogen atoms. For the La0.7Ce0.1Ga0.3Ni5 at T = 298 K, the mass of hydrogen absorbed at equilibrium for the LaNi5 is 8 g of hydrogen atoms while at T = 318 K it is 6.5 g of hydrogen atoms. This is explained by the exothermic nature of the hydrogen-absorption reaction.
  • Internal temperature
Temperature significantly affects absorption kinetics. It is important to regulate the internal temperature of an exothermic absorption reaction for multiple reasons. The process of exothermic absorption releases heat, which may raise the system’s temperature. The integrity of the absorbent material, the safety of the system, and the response kinetics may all be adversely impacted by improper management of this heat. A thermocouple was inserted on the acquisition card and inside the hydride bed to regulate the interior temperature.
Figure 3 depicts the variation of the temperature inside the hydride bed during absorption as a function of time at Ti = 298 K (a) and Ti = 318 K (b).
It is noted that in the initial stages of the absorption response, the internal temperature rises, which confirms the exothermic nature of the reaction. Then, the temperature decreases following the external cooling process by the cold water coming from the thermostatic bath, and it stabilizes at a temperature close to that set by the thermostatic bath (Ti). We see that La0.7Ce0.1Ga0.3Ni5 exhibits greater exothermic hydrogen absorption than does LaNi5. Since more heat is created during the doped compound’s absorption reaction, more hydrogen atoms are put into the La0.7Ce0.1Ga0.3Ni5’s interstitial sites under the same starting conditions.
  • Absorption isotherms
Pressure-composition isotherms (P-C-T) are one of the most significant considerations to assess thermodynamic properties for an alloy that absorbs hydrogen. It is feasible to track the development of the reversible storage reaction (absorption/desorption) thanks to these isotherms. The hydrogen absorption isotherms for the compounds LaNi5 and La0.7Ce0.1Ga0.3Ni5 at T = 298 K and T = 318 K are displayed in Figure 4.
It is shown, in Figure 4, that at a given temperature, an absorption isotherm is made up of three parts [34]:
α phase: This phase is characterized by a significant increase in pressure vs. 0 absorbed mass.
Hysteresis measurement. This indicates that the hydrogen molecules have not yet dissociated into atoms and are still present on the surface of the absorbent material.
α+β phase: This phase is associated with the plateau of equilibrium. It is distinguished by a pressure stabilization against an increasing rise in the hydrogen absorbed mass. This indicates that hydrogen can exist in both gaseous and solid forms.
β phase: This is related to the solid phase, in which the absorbent substance completely absorbs the hydrogen atoms.
First, it is demonstrated that raising the temperature causes the equilibrium plateau for both compounds to climb. This is explained by the exothermic nature of the hydrogen absorption reaction, which suggests that more pressure is required to store hydrogen at high temperatures. Secondly, through comparison, it is shown that, at a given temperature, the equilibrium plateau for the compound La0.7Ce0.1Ga0.3Ni5 is situated above that for LaNi5, and that the compound La0.7Ce0. 1Ga0.3Ni5 has a greater total mass absorbed than the compound LaNi5 by 0.2 wt%. In summary, doping increases the bulk capacity of stored hydrogen, even if it can destroy the stability of the hydride by slightly raising the plateau pressure. Consequently, doping with Ce and Ga elements on the La sites results in 1.6 weight percent of the storage capacity. Thus, this value represents, for example, theoretical storage capacity of the TiFe hydride, with better safety indicated by the normal working conditions of temperature and pressure [7].
  • Hysteresis measurement
The term “hysteresis” describes how the current state of a system is influenced by its past. When the hydrogen absorption and desorption curves on an isotherm do not overlap, a hysteresis phenomenon occurs. Understanding hysteresis is essential in developing materials with the best possible hydrogen-storage features since it shows structural changes and other phenomena within the material throughout the process of reversible hydrogen storage. The hysteresis cycles for the compounds LaNi5 and La0.7Ce0.1Ga0.3Ni5 are shown in Figure 5.
Consequently, we can use the following equation to estimate a hysteresis rate for each compound based on the experimental values of the equilibrium pressures (Peq) [35]:
H y s = ln ( P e q a b s P e q d e s )
Table 1 presents the hysteresis rate values for each compound versus temperatures.
First of all, the hysteresis rate calculation reveals that hysteresis falls with temperature. When considering the La0.7Ce0.1Ga0.3Ni5 compound, this reduction is much more impressive. In fact, the hysteresis is reduced by almost 50% when the temperature goes from 298 K to 318 K.
This can be explained by the prospect that, as the temperature rises, the desorption reaction will take place at temperatures that are closer to those of absorption, which will reduce the hysteresis. Then, it is shown that at a given temperature, the La0.7Ce0.1Ga0.3Ni5 compound has a lower hysteresis rate than the LaNi5 compound. Given that hysteresis and cyclability are interrelated aspects of material performance, low hysteresis improves efficiency, leading to high cyclability, which ensures long-term use. Thus, low hysteresis indicates that the material is working more efficiently, with less energy lost during absorption and desorption cycles. This increased efficiency can improve cyclability by reducing energy losses and preventing heat buildup or other negative effects that could degrade the material. Therefore, it is observed that doping the compound with Ce and Ga on the La sites increases its cyclability as well as its ability to regularly absorb and desorb hydrogen without degrading it, thus extending its useful life.
In the subsequent section, we will use the statistical physics formalism to physically examine the aforementioned results, starting at the microscopic level and working our way up to the macroscopic level. Given that the absorption process entails the penetration of atoms into the absorbent material’s volume, as opposed to the adsorption process, which involves the attachment of adsorbed molecules to the material’s surface, studies have demonstrated the energy similarity between these two processes because they arise from the same phenomenon of diffusional equilibrium [36]. Therefore, theoretical research by Wjihi et al. [32] suggests that adsorption and absorption phenomena can be modeled using the same statistical physics model. Thus, it is assumed in this study to solve the problem of hydrogen absorption in LaNi5 and La0.7Ce0.1Ga0.3Ni5 in the same way adopted for the adsorption process.

3.2. Modeling

  • Description of the model
The application of the statistical physics model to study the phenomenon of hydrogen storage in intermetallics and their derivatives is a fascinating subject because this model contains important physical parameters in its expressions that are different from those of empirical models like Langmuir and Freundlich, the parameters of which generally have no direct physical meaning [37,38]. Thereby, to use a statistical physics approach to examine the absorption process, three assumptions need to be stated [39]. First, the statistical grand canonical ensemble must be used because it calls for an interchange of particles from the free level to the absorbed level. Second, hydrogen must be regarded as an ideal gas. Third, all internal degrees of freedom of hydrogen—aside from translation and rotation—will be disregarded.
On the one hand, the general equation that describes the reaction of intermetallics’ reversible hydrogen storage is
M + n 2 H 2 M H n
where M is the absorbent intermetallic, n is the stoichiometric coefficient, Hn is defined as an aggregate of n hydrogen atoms in a receptor site, and the generated hydride is described by MHn.
On the other hand, the grand canonical partition function Zgc for one interstitial site is formulated in the following general formula [40]:
z g c = N j e β ( ε j μ ) N j
where −(−εj) represents the absorption energy of one receptor site, Nj is the occupation state of one receptor site and is either 0 or 1, µ denotes the chemical potential of one absorbed site, and β is expressed as follows:
β = 1 K B T
where T is the absolute temperature, and KB is the Boltzmann constant.
Assuming that the hydrogen absorption reaction occurs in this case on Nm independent and identical receptor sites, the grand canonical partition function can then be written as follows:
Z g c = ( Z g c ) N m
As mentioned by Belkhiria et al. [40], the expression of the average site occupancy number Nm is determined as
N 0 = k B T ln ( Z g c ) μ = N m k B T ln ( z g c ) μ
Therefore, considering that a variable quantity of Na hydrogen atoms is absorbed in the Nmi receptor sites distributed throughout the mass unit of the intermetallic, the above formulations make it possible to consider a certain number of modeling methods based on the formalism of statistical physics, such as the single-layer or double-layer model with identical sites, two types of sites, three sites and four sites. Thus, the choice of the appropriate model is based on the model, which gives a better fit of the experimental isotherms and better-adjusted coefficient R2. Applying the mono-layer model with two different types of receptor sites yields the optimal adjustment in the present case. Furthermore, the researchers have selected this model to represent the absorption process in intermetallics [41].
We suppose that the two energy levels can be allocated to phases 1 and 2 in the single-layer model with two distinct types of receptor sites. The values n1 and n2, respectively, represent the number of H atoms per site for phases 1 and 1. The two locations, 1 and 2, in Equation (2) are separated as follows:
n 1 2 H 2 + M M H n 1
n 2 2 H 2 + M M H n 2
First, we keep in consideration that the absorption is assumed to involve two distinct types of sites, n1 and n2. Second, we take into account that N1m provides the density of sites n1, and N2m provides the density of sites n2. Third, we employ ∆ε1 and ∆ε2 to determine the absorption energy at sites n1 and n2, respectively. Therefore, Zgc is represented as follows:
Z g c = ( z 1 g c ) N 1 m × ( z 1 g c ) N 2 m
where the partition functions of the first and second types of sites are denoted by Z1gc and Z2gc, respectively:
z 1 g c = N j = 0 , 1 e ( ε 1 μ ) N j k B T = 1 + e β ( ε 1 + μ )
z 2 g c = N j = 0 , 1 e ( ε 2 μ ) N j k B T = 1 + e β ( ε 2 + μ )
Therefore, using Equations (6) and (7), N0 can be expressed as follows:
N 0 = N 1 m k B T ln ( z 1 g c ) μ + N 2 m k B T ln ( z 2 g c ) μ = N 1 m 1 + e β ε 1 + μ + N 2 m 1 + e β ε 2 + μ
Considering in the one hand, that hydrogen is an ideal gas, the chemical potential of a free hydrogen molecule state is given as follows:
μ m = k B T ln N z g
Thus, at thermodynamic equilibrium, we have:
μ m = μ n
μ presents the chemical potential of the receptor site, and n presents the number of atoms per one site.
In the other hand, as previously noted, just the degrees of freedom of translation and rotation are considered; therefore Zgc is expressed as follows [40,41]:
z g = z g t r × z g r o t = V 2 π m k B T h 2 3 / 2 + T 2 θ r o t
where m is the mass of a hydrogen atom, V is the system’s volume, h is the Planck’s constant (h = 6.6260 × 10−34 m2 kg s−1), and θrot is the dihydrogen molecule’s typical rotational temperature.
The relationship between µ and µm is governed by Equation (3). Thus, in the thermodynamic equilibrium state (equilibrium of pressure and temperature), the law of mass action providing the equilibrium constant K is written as follows:
K T = M H n M × H 2 n 2 M H n = K × M × H 2 n 2
where [MHn], [M], and [H2] present the hydride, metal, and hydrogen concentrations, respectively. Thus, Equation (10) imposes the following:
μ M = n 2 μ m H 2
And
μ m H 2 = μ M n 2 = 2 n μ M
Consequently, the dihydrogen molecule’s absorption energy, εm, can also be represented as follows:
ε m H 2 = ε M n 2 = 2 n ε M
Assume that a fully absorbed molecule from the first and second site defined by n1/2 and n2/2 has energies of ε1m and ε2m, respectively. In this case, ε1m and ε2m are expressed by the following expressions:
ε 1 m = ε 1 n 1 2   and   ε 2 m = ε 2 n 2 2
As a result, based on Equation (9), the average number of the occupied sites N0 can be expressed by
N O = N 1 m 1 + e β ( ε 1 m + μ g ) n 1 2 + N 2 m 1 + e β ( ε 2 m + μ g ) n 2 2 = N 1 m 1 + e β ε 1 m n 1 2 z g β P 1 n 1 2 + N 2 m 1 + 1 + e β ε 2 m n 2 2 z g β P 2 n 2 2
where P1 and P2 present the pressures at halves saturation defining, respectively, the first and the second type of sites. They are expressed as follows:
P 1 = K B T Z g exp ( β ε 1 m )   and   P 2 = K B T Z g exp ( β ε 2 m )
As a result, N0 is involved as
N O = N 01 + N 02 = N 1 m 1 + P 1 P n 1 + N 2 m 1 + P 2 P n 2
Considering that the average number of hydrogen atoms absorbed is as follows:
H M O = n 1 N 01 + n 2 N 02
Thus, this gives
H M = H M 1 + H M 2 = n 1 N 1 m N 1 m + N 2 m 1 + P 1 P n 1 2 + n 2 N 2 m N 1 m + N 2 m 1 + P 2 P n 2 2 = H M 1 s a t 1 + P 1 P n 1 2 + H M 2 s a t 1 + P 2 P n 2 2
Lastly, the following formula provides the total amount of hydrogen absorbed as a function of pressure:
H M P = 1 N 1 m + N 2 m × n 1 N 1 m 1 + P 1 P n 1 2 + n 2 N 2 m 1 + P 2 P n 2 2

4. Results and Discussions

4.1. Adjustment of Experimental Isotherms

The two-site monolayer model previously developed was used to adapt the experimental isotherms for the LaNi5 and La0.7Ce0.1Ga0.3Ni5 compounds (Figure 6).
Figure 5 illustrates the excellent agreement between the suggested numerical model and the experimental data. Hence, the stereographic and energetic parameters involved in the hydrogen absorption reaction by the LaNi5 and the La0.7Ce0.1Ga0.3Ni5 are deduced using adjustment (Table 2).
The subsequent discussion will initially focus on these parameters in relation to temperature, providing an understanding of the microscopic level of the hydrogen absorption reaction via both the doped and parent compounds. The macroscopic level will then be accessible as they will be employed to determine the thermodynamic functions governing the process of hydrogen absorption.
Occupancy status: number of atoms per site.
A fundamental physicochemical parameter for understanding the absorption process is the number of atoms per site. Thus, n (= n 1 + n 2 2 ) denotes the absorption reaction’s stoichiometric coefficient. It represents the stoichiometric coefficient presented in Equation (3) [35]. Additionally, the degree of atoms’ aggregation is estimated by this coefficient.
Figure 7 depicts the number of hydrogen molecules absorbed on each type of site n1 and n2 as well as the average value of the molecules in the two types of sites n.
First, it is demonstrated that for both the parent (LaNi5) and doped (La0.7Ce0.1Ga0.3Ni5) compounds, n1 and n2 drop as temperature rises. As a result, n (= n 1 + n 2 2 ) decreases. This is explained by the exothermic nature of the hydrogen absorption process. Thus, the interstitial sites are blocked by the rise in temperature. Subsequently, the results indicated that, in the case of LaNi5, at a given temperature, n1 > n2. This indicates that more hydrogen atoms are being taken up by n1-type sites in the case of LaNi5. Moreover, n2 takes values close to 1. However, n1 takes values much greater than 1. This means that the phenomenon of the agglomeration of hydrogen atoms is more important at the n1 sites than n2. On the other hand, in the case of the La0.7Ce0.1Ga0.3Ni5 compound, the results indicated that n2 > n1. This indicates that n2-type sites are activated, and n1-type sites are blocked by Ce and Ga doping on the La sites. Furthermore, the fact that n1 takes values close to 1 while n2 takes values much higher than 1 indicates that the agglomeration phenomenon originates at n2 sites in the case of the doped compound. In addition, we notice that, at a given temperature, n takes on larger values in the case of the La0.7Ce0.1Ga0.3Ni5 compound than the LaNi5 compound. This means that the average number of hydrogen atoms absorbed at the La0.7Ce0.1Ga0.3Ni5 sites is greater than that absorbed at the LaNi5 sites. This gives a good explanation of the experimental findings. As previously noted, Figure 4 illustrates that the mass percentage of hydrogen stored is higher in the case of the La0.7Ce0.1Ga0.3Ni5 compound than the percentage of hydrogen stored in the case of the LaNi5 compound.
Interstitial sites refer to empty locations in a crystal structure where additional atoms or molecules can fit. While n in the LaNi5 and its derivatives is theoretically equal to 6, some statistical physics investigations have determined that n is much greater than 6 [35,42,43,44]. This is can indeed be attributed to the phenomena of confined agglomeration and decrepitating, which lead to the formation of additional sites for hydrogen absorption. These phenomena increase the material’s hydrogen-absorption capacity beyond the standard crystallographic interstitial sites.
The absorption energy associated with the bonds formed in these interstitial sites depends on several factors, such as the size of the interstitial sites, the nature of the atoms or molecules inserted, and the intermolecular or inter atomic interactions involved. In the section that follows, we propose determining the absorption energy of the hydrogen atoms in the n sites of LaNi5 as well as La0.7Ce0.1Ga0.3Ni5.

4.2. Energy Analysis: Absorption Energies

The study of absorption energies is crucial to better understand the phenomenon of hydrogen absorption in metal hydrides since it reveals the type of bond between the hydrogen atom and metal. In fact, hydrogen atoms are absorbed with energy ∆E1 on n1 sites and ∆E2 on n2 sites, respectively. ∆E1 and ∆E2 are calculated using the energetic parameters P1, P2, and the saturated vapor pressure of gaseous hydrogen Pvs. The relations provide them in the following manner [45]:
Δ E 1 = R T ln P v s P 1
Δ E 2 = R T ln P v s P 2
where P1 and P2 are determined from the model by fitting the hydrogen absorption isotherms.
The saturated vapor pressure is expressed as follows:
P v s = exp 12.69 94.896 T + 1.1125 ln T + 3.2915 × 10 4 T 2
Thus, absorption energies over n1 and n2 sites are presented in Figure 8.
The absorption energy values show that the bond between the hydrogen atoms and the intermetallics is of a chemical type. This implies that the bond formed between the hydrogen atoms and the interstitial sites (n1 and n2) of the intermetallics considered is relatively strong. It is observed that if the temperature increases, the absorption energy also increases. This suggests that higher temperatures require more energy due to the exothermic nature of the absorption reaction. Additionally, it is noted that the doped compound’s absorption energies are somewhat higher than those of LaNi5. This means that the amount of hydrogen stored in the doped compound is slightly greater than that stored in LaNi5, leading to a slightly higher energy expenditure. In addition, it is demonstrated that for the LaNi5 compound, ∆E1 > ∆E2, while for the La0.7Ce0.1Ga0.3Ni5 molecule, ∆E2 > ∆E1. This is explained by the fact that, as shown in Figure 8, the n1 sites are more occupied by hydrogen atoms in the case of LaNi5, whereas the n2 sites are more occupied in the case of La0.7Ce0.1Ga0.3Ni5.
It should be noted that the enthalpy change indicates the thermodynamic favorability of the absorption process, which is controlled by the system’s relaxation and stabilization, whereas the absorption energy shows the overall energy input required to absorb the hydrogen. The difference between the initial, transition, and final states of the system, as well as the additional energy needed to overcome activation barriers and create lattice distortions, can all be taken into account when explaining the discrepancy between absorption energies and absolute enthalpy change during hydrogen absorption.
  • Thermodynamic properties: Internal and Gibbs energies
Internal energy and Gibbs energy are two thermodynamic concepts crucial to understanding physical and chemical systems.
  • Internal energy
The internal energy (U) presents the sum of all microscopic energies. Thus, it is the kinetic and potential energies of the molecules or atoms that make up the system. It also includes the energy of chemical and physical interactions between molecules. A system’s pressure and temperature have a big impact on its internal energy. The internal energy of hydrogen absorption is expressed as follows [42]:
U int = ln ( z g c ) β + μ β ( ln ( z g c ) μ )
where
μ = K B T ln β P z g
Thus,
U int = k B ln ( β P z g ) N 1 m ( P P 1 ) n 1 1 + ( P P 1 ) n 1 + N 2 m ( P P 2 ) n 2 1 + ( P P 2 ) n 2 k B T N 1 m ( P P 1 ) n 1 1 + ( P P 1 ) n 1 + N 2 m ( P P 2 ) n 2 1 + ( P P 2 ) n 2
Figure 9 depicts the variation of the internal energy as a function of pressure when T = 298 K and T = 318 K.
It is shown that the internal energies for both compounds are negative. This indicates that energy is released throughout the process of hydrogen absorption. Therefore, the exothermic nature of the hydrogen-absorption process explains this. We additionally observe that when the starting temperature rises, the internal energy’s absolute value rises as well. This demonstrates how an increase in starting temperature causes an increase in energy dissipation. Furthermore, a comparison of the results at a given temperature revealed that the LaNi5 compound has a slightly lower absolute value of internal energy than that of La0.7 Ce0.1Ga0.3Ni5. This can be explained by the fact that the hydrogen atoms are more active in the interstitial sites of the compound La0.7Ce0.1Ga0.3Ni5 (as shown by experimental results), which boosts the system’s heat release.
  • Gibbs Energy
The Gibbs energy combines internal energy with pressure–volume effects (via enthalpy) and entropy, making it a powerful tool for predicting the spontaneity of processes at constant pressure and temperature. It is expressed as follows [42]:
G a = μ × n × N 0 = μ H M 0
Therefore, Ga can be expressed as a function of the fitted parameters as follows:
G a = K B ln β P z g n 1 N 1 m 1 + P 1 P n 1 2 + n 2 N 2 m 1 + P 2 P n 2 2
Figure 10 depicts the variation of the Gibbs energy as a function of pressure when T = 298 K and T = 318 K for both compounds.
First, we observe that, for a given temperature and pressure, the Gibbs energy takes negative values. This is a blatant sign that the intermetallics’ spontaneous and thermodynamically advantageous process of absorbing hydrogen is occurring. This means that the system will naturally evolve toward this state, often by releasing energy. Next, it is demonstrated that, up to a critical pressure, Pc, the absolute value of the Gibbs energy declines logarithmically as a function of pressure before rising and stabilizing. Pc corresponds to the establishment of thermodynamic equilibrium. This means that beyond Pc, the difference in chemical potential between the gaseous hydrogen and the absorbed hydrogen reaches equilibrium, where the absorption rate becomes zero. Then, the absorption of hydrogen stops, regardless of the increase in pressure. This pressure typically indicates that the material is getting close to its maximum absorption capacity and is discovered when the PCT curve starts to increase quickly after the plateau (Peq). According to the experimental results (Figure 5), we see that this pressure varies between 6 bar and 9 bar, which confirms the results of our model. At a given temperature, it is observed that Pc is the same for both compounds. This means that they start to absorb at the same pressure. This is because the energy required for hydrogen to insert into the interstitial sites is similar for both compounds, resulting in identical absorption pressures.
We observe that if the starting temperature rises, this pressure will increase even more. In fact, Pc = 6 bar at T = 298 K and Pc = 9 bar at T = 318 K. This indicates that in order to achieve thermodynamic equilibrium, a greater pressure is needed as the temperature rises.
We also see that the parent and doped compounds require the same pressure (Pc) to reach equilibrium at a given temperature.
This means that the doping did not disturb the thermodynamic equilibrium of the absorption reaction. Nonetheless, we note that the doped compound has a higher absolute value of Gibbs energy than the parent compound. This indicates that in the case of the compound La0.7Ce0.1Ga0.3Ni5, hydrogen absorption occurs more spontaneously. As a result, the La0.7Ce0.1Ga0.3Ni5 compound’s metallic matrix can hold more hydrogen atoms. This explains why doping with Ce and Ga on La sites enhances hydrogen storage efficiency.

5. Conclusions

In this work, an experimental and numerical study of the substitution effect on the sites of La in the compound LaNi5 by Ce and Ga was carried out. Therefore, the hydrogen-storage properties of the compounds LaNi5 and La0.7Ce0.1Ga0.3Ni5 were compared. The experimental results showed that doping with Ce and Ga elements on La sites improves the kinetics and the mass of stored hydrogen by more than 30%. Additionally, doping improves its cyclability and capacity to routinely absorb and desorb hydrogen without causing degradation and extends its usable life, resulting in a reduction in hysteresis of over 28%.
The simulated results demonstrated that for the two compounds, absorption occurs when hydrogen atoms are inserted into two different types of interstitial sites (n1, n2). In fact, these interstitial sites are more occupied by the hydrogen atoms in the doped compound. The numerical calculation of the Gibbs energy and internal energy further demonstrates that the hydrogen atoms are introduced into the interstitial sites by means of an exothermic and spontaneous reaction.

Author Contributions

Conceptualization, S.B. and A.A.; methodology, S.B. and A.J.; software, S.B. and A.A.; validation, S.B. and M.H.D.; formal analysis, S.B., I.H., and A.J.; investigation, S.B., A.J., and I.H.; resources, S.B.; data curation, S.B., A.J., and M.H.D.; writing—original draft preparation, S.B., I.H., and A.J.; writing—review and editing, S.B., A.A., and M.H.D.; visualization, S.B.; supervision, S.B; project administration, A.J.; funding acquisition, S.B. and A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Acknowledgments

The Researchers would like to thank the Deanship of Graduate Studies and Scientific Research at Qassim University for financial support (QU-APC-2024-9/1).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shabani, B.; Andrews, J. Standalone solar-hydrogen systems powering fire contingency networks. Int. J. Hydrogen Energy 2015, 27, 5509–5517. [Google Scholar] [CrossRef]
  2. Barahona, I.; Almulhim, T. Renewable energies and circular economies: A systematic literature review before the ChatGPT boom. Energy Rep. 2024, 11, 2656–2669. [Google Scholar] [CrossRef]
  3. Assaf, J.; Shabani, B. Economic analysis and assessment of a standalone solar-hydrogen combined heat and power system integrated with solar-thermal collectors. Int. J. Hydrogen Energy 2016, 41, 18389–18404. [Google Scholar] [CrossRef]
  4. Talukdar, M.; Blum, P.; Heinemann, N.; Miocic, J. Techno-economic analysis of underground hydrogen storage in Europe. iScience 2024, 27, 108771. [Google Scholar] [CrossRef]
  5. Yin, L.; Yang, H.; Ju, Y. Review on the key technologies and future development of insulation structure for liquid hydrogen storage tanks. Int. J. Hydrogen Energy 2024, 57, 1302–1315. [Google Scholar] [CrossRef]
  6. Morales-Ospino, R.; Celzard, A.; Fierro, V. Strategies to recover and minimize boil-off losses during liquid hydrogen storage. Renew. Sustain. Energy Rev. 2023, 182, 113360. [Google Scholar] [CrossRef]
  7. Drawer, C.; Lange, J.; Kaltschmitt, M. Metal hydrides for hydrogen storage—Identification and evaluation of stationary and transportation applications. J. Energy Storage 2024, 77, 109988. [Google Scholar] [CrossRef]
  8. Malleswararao, K.; Kumar, P.; Dutta, P.; Murthy, S.S. Experimental investigations on a coupled metal hydride based thermal energy storage system. Int. J. Hydrogen Energy 2024, 56, 1371–1383. [Google Scholar] [CrossRef]
  9. Zhu, Y.; Yang, X.; Xu, Z.; Tsui, G.; Zhou, Q.; Tang, R.; Xiao, F.; Chan, K. Development of AB2-type TiZrCrMnFeCoV intermetallic high-entropy alloy for reversible room-temperature hydrogen storage. J. Energy Storage 2024, 57, 109553. [Google Scholar] [CrossRef]
  10. Yartys, V.A.; Lototskyy, M.V. Laves type intermetallic compounds as hydrogen storage materials: A review. J. Alloys Compd. 2022, 916, 165219. [Google Scholar] [CrossRef]
  11. Ugaddan, E.; Violi, D.; Fiume, V.; Barale, J.; Luetto, C.; Rizzi, P.; Baricco, M. AB5-based metal hydride embedded in polyethylene and polymethylmethacrylate for hydrogen storage. Int. J. Hydrogen Energy 2024, 78, 952–961. [Google Scholar] [CrossRef]
  12. Lys, A.; Fadonougbo, J.O.; Faisal, M.; Suh, J.-Y.; Lee, Y.-S.; Shim, J.-H.; Park, J.; Cho, Y.W. Enhancing the Hydrogen Storage Properties of AxBy Intermetallic Compounds by Partial Substitution: A Short Review. Hydrogen 2020, 1, 38–63. [Google Scholar] [CrossRef]
  13. Li, X.; Li, B.; Zhao, Y.; Zhang, X.; Ma, H.; He, X.; Zhou, S.; Wang, L.; Yan, H. Composition design of (LaCeCa)1(NiMnAl)5 alloys by uniform design method and their hydrogen storage performance. J. Alloy. Metall. Syst. 2023, 2, 100006. [Google Scholar] [CrossRef]
  14. Chen, Z.; Ma, Z.; Zheng, J.; Li, X.; Akiba, E.; Li, H.-W. Perspectives and challenges of hydrogen storage in solid-state hydrides. Chin. J. Chem. Eng. 2021, 29, 1–12. [Google Scholar] [CrossRef]
  15. Joubert, J.-M.; Paul-Boncour, V.; Cuevas, F.; Zhang, J.; Latroche, M. LaNi5 related AB5 compounds: Structure, properties and applications. J. Alloys Compd. 2021, 862, 158163. [Google Scholar] [CrossRef]
  16. Spodaryk, M.; Gasilova, N.; Züttel, A. Hydrogen storage and electrochemical properties of LaNi5-xCux hydride-forming alloys. J. Alloys Compd. 2019, 775, 175–180. [Google Scholar] [CrossRef]
  17. Mansouri, M.; Shtender, V.; Tunsu, C.; Yilmaz, D.; Messaoudi, O.; Ebin, B.; Sahlberg, M.; Petranikova, M. Production of AB5 materials from spent Ni-MH batteries with further tests of hydrogen storage suitability. J. Power Sources 2022, 539, 231459. [Google Scholar] [CrossRef]
  18. Liu, Y.; Chabane, D.; Elkedim, O. Optimization of LaNi5 hydrogen storage properties by the combination of mechanical alloying and element substitution. Int. J. Hydrogen Energy 2024, 53, 394–402. [Google Scholar] [CrossRef]
  19. Kumar, A.; Muthukumar, P.; Sharma, P.; Kumar, E.A. Absorption based solid state hydrogen storage system: A review. Sustain. Energy Technol. Assess. 2022, 52, 102204. [Google Scholar] [CrossRef]
  20. Muhammed, N.S.; Gbadamosi, A.O.; Epelle, E.I.; Abdulrasheed, A.A.; Haq, B.; Patil, S.; Al-Shehri, D.; Kamal, M.S. Hydrogen production, transportation, utilization, and storage: Recent advances towards sustainable energy. J. Energy Storage 2023, 73, 109207. [Google Scholar] [CrossRef]
  21. Zhu, Z.; Zhu, S.; Lu, H.; Wu, J.; Yan, K.; Cheng, H.; Liu, J. Stability of LaNi5-xCox alloys cycled in hydrogen—Part 1 evolution in gaseous hydrogen storage performance. Int. J. Hydrogen Energy 2019, 44, 15159–15172. [Google Scholar] [CrossRef]
  22. Liu, J.; Li, K.; Cheng, H.; Yan, K.; Wang, Y.; Liu, Y.; Jin, H.; Zheng, Z. New insights into the hydrogen storage performance degradation and Al functioning mechanism of LaNi5-xAlx alloys. Int. J. Hydrogen Energy 2017, 42, 24904–24914. [Google Scholar] [CrossRef]
  23. Wan, C.; Zhao, S.; Wang, H. Tuning phase structure and electrochemical hydrogen storage properties of A5B19-type La–Y–Ni–Mn-based superlattice alloys by partial Al substitution. Int. J. Hydrogen Energy 2024, 49, 51–58. [Google Scholar] [CrossRef]
  24. Lv, W.; Yuan, J.; Zhang, B.; Wu, Y. Influence of the substitution Ce for La on structural and electrochemical characteristics of La0.75-xCexMg0.25Ni3Co0.5 (x=0, 0.05, 0.1, 0.15, 0.2 at. %) hydrogen storage alloys. J. Alloys Compd. 2018, 730, 360–368. [Google Scholar] [CrossRef]
  25. Tarasov, B.; Bocharnikov, M.; Yanenko, Y.; Fursikov, P.; Lototskyy, M. Cycling stability of RNi5 (R = La, La+Ce) hydrides during the operation of metal hydride hydrogen compressor. Int. J. Hydrogen Energy 2018, 43, 4415–4427. [Google Scholar] [CrossRef]
  26. Wen, X.; Wang, B.; Li, C.; Liu, T. Effect of La doping on the structural stability and hydrogen adsorption behavior of Ce-La alloys. J. Alloys Compd. 2024, 982, 173553. [Google Scholar] [CrossRef]
  27. Silvestri, L.; Forcina, A.; Silvestri, C.; Traverso, M. Circularity potential of rare earths for sustainable mobility: Recent developments, challenges and future prospects. J. Clean. Prod. 2021, 292, 26089. [Google Scholar] [CrossRef]
  28. Cha, S.; Kim, J.; Yun, J.; Rhyee, J. Magnetocaloric and hydrogen storage multi-functional properties of Eu4Ga8Ge16 compounds. J. Alloys Compd. 2023, 968, 17217. [Google Scholar] [CrossRef]
  29. Qin, C.; Wang, H.; Jiang, W.; Liu, J.; Ouyang, L.; Zhu, M. Comparative study of Ga and Al alloying with ZrFe2 for high-pressure hydrogen storage. Int. J. Hydrogen Energy 2022, 47, 13409–13417. [Google Scholar] [CrossRef]
  30. Janjua, M.R.S.A. Hydrogen as an energy currency: Encapsulation of inorganic Ga12N12 with alkali metals for efficient H2 adsorption as hydrogen storage materials. J. Phys. Chem. Solids 2022, 160, 110352. [Google Scholar] [CrossRef]
  31. Oueslati, K.; Sakly, A.; Lima, E.C.; Ayachi, F.; Ben Lamine, A. Statistical physics modeling of the removal of Resorcinol from aqueous effluents by activated carbon from avocado seeds. J. Mol. Liq. 2022, 63, 119386. [Google Scholar] [CrossRef]
  32. Wjihi, S.; Sellaoui, L.; Bouzid, M.; Dhaou, H.; Knani, S.; Jemni, A.; Ben Lamine, A. Theoretical study of hydrogen sorption on LaNi5 using statistical physics treatment: Microscopic and macroscopic investigation. Int. J. Hydrogen Energy 2017, 42, 2699–2712. [Google Scholar] [CrossRef]
  33. Prasad, J.S.; Muthukumar, P. Performance and energy efficiency of a solid-state hydrogen storage system: An experimental study on La0.7Ce0.1Ca0.3Ni5. Appl. Therm. Eng. 2022, 216, 119030. [Google Scholar] [CrossRef]
  34. Ge, Y.T. Characterisation of pressure-concentration-temperature profiles for metal hydride hydrogen storage alloys with model development. Energy Storage 2024, 6, 504. [Google Scholar] [CrossRef]
  35. Bouaziz, N.; Briki, C.; Dhahri, E.; Jemni, A.; Ben Lamine, A. Experimental and theoretical investigation of absorption and desorption of hydrogen in the LaNi4Co0.5Mn0.5 alloy. Chem. Eng. Sci. 2022, 251, 117453. [Google Scholar] [CrossRef]
  36. Schwarzer, M.; Schwarzer, M.; Hertl, N.; Hertl, N.; Nitz, F.; Nitz, F.; Borodin, D.; Borodin, D.; Fingerhut, J.; Fingerhut, J.; et al. Adsorption and Absorption Energies of Hydrogen with Palladium. J. Physic Chem. 2022, 126, 14500–14508. [Google Scholar] [CrossRef] [PubMed]
  37. Ziemiański, P.P.; Derkowski, A. Structural and textural control of high-pressure hydrogen adsorption on expandable and non-expandable clay minerals in geologic conditions. Int. J. Hydrogen Energy 2022, 47, 28794–28805. [Google Scholar] [CrossRef]
  38. Lam, S.T.; Ballinger, R.; Forsberg, C. Modeling and predicting total hydrogen adsorption in nanoporous carbon materials for advanced nuclear systems. J. Nucl. Mater. 2018, 511, 328–340. [Google Scholar] [CrossRef]
  39. Knani, S.; Mathlouthi, M.; Ben Lamine, A. Modeling of the psychophysical responsecurves using the grand canonical ensemble in statistical physics. Food Biophys. 2007, 2, 183–192. [Google Scholar] [CrossRef]
  40. Belkhiria, S.; Alsawi, A.; Briki, C.; Altarifi, S.M.; Dhaou, M.H.; Jemni, A. Physical Analysis and Mathematical Modeling of the Hydrogen Storage Process in the MmNi4.2Mn0.8 Compound. Materials 2024, 17, 2237. [Google Scholar] [CrossRef]
  41. Mechi, N.; Khemis, I.B.; Dhaou, H.; Zghal, S.; Lamine, A.B. Morphologic, Structural, Steric, Energetic and Thermodynamic Studies of the Mechanical Alloy Mg50Ni45Ti5 for Hydrogen Storage. J. Phys. Chem. J. Biophys. 2017, 7, 2–11. [Google Scholar] [CrossRef]
  42. Bouaziz, N.; Ben Manaa, M.; Ben Lamine, A. Theoretical study of hydrogen absorption-desorption on LaNi3.8Al1.2-xMnx using statistical physics treatment. Phys. B 2017, 525, 46–59. [Google Scholar] [CrossRef]
  43. Bouaziz, N.; Ben Torkia, Y.; Aouaini, F.; Nakbi, A.; Dhaou, H.; Ben Lamine, A. Statistical physics modeling of hydrogen absorption onto LaNi4.6Al0.4: Stereographic and energetic interpretations. Sep. Sci. Technol. 2019, 54, 1520–5754. [Google Scholar] [CrossRef]
  44. Wjihi, S.; Dhaou, H.; Ben Yahia, M.; Knani, S.; Jemni, A.; Ben Lamine, A. Statistical physics modeling of hydrogen desorption from LaNi4.75Fe0.25: Stereographic and energetic interpretations. Phys. B 2015, 497, 112–120. [Google Scholar] [CrossRef]
  45. Belkhiria, S.; Briki, C.; Dhaou, M.H.; Jemni, A. Experimental and numerical study of hydrogen adsorption by the Ni0.6Mg0.4Fe2O4 compound. Int. J. Hydrogen Energy 2024, 64, 990–1000. [Google Scholar] [CrossRef]
Figure 1. Schematic presentation of the experimental setup (a) and sectional depiction of the used M-HR in two dimensions (b).
Figure 1. Schematic presentation of the experimental setup (a) and sectional depiction of the used M-HR in two dimensions (b).
Metals 14 00967 g001
Figure 2. Hydrogen absorption kinetics at T = 298 K (a) and T = 318 K (b).
Figure 2. Hydrogen absorption kinetics at T = 298 K (a) and T = 318 K (b).
Metals 14 00967 g002
Figure 3. Evolution of the internal temperature for Ti = 298 K (a) and Ti = 318 K (b).
Figure 3. Evolution of the internal temperature for Ti = 298 K (a) and Ti = 318 K (b).
Metals 14 00967 g003
Figure 4. Absorption isotherms for the LaNi5 and La0.7Ce0.1Ga0.3Ni5 compounds at T = 298 K and T = 318 K.
Figure 4. Absorption isotherms for the LaNi5 and La0.7Ce0.1Ga0.3Ni5 compounds at T = 298 K and T = 318 K.
Metals 14 00967 g004
Figure 5. Hysteresis comparison between LaNi5 and La0.7Ce0.1Ga0.3Ni5 compounds at T = 298 K (a) and at T = 318 K (b).
Figure 5. Hysteresis comparison between LaNi5 and La0.7Ce0.1Ga0.3Ni5 compounds at T = 298 K (a) and at T = 318 K (b).
Metals 14 00967 g005
Figure 6. Adjustment of experimental isotherms for the LaNi5 and La0.7Ce0.1Ga0.3Ni5 compounds at T = 298 K and T =318 K.
Figure 6. Adjustment of experimental isotherms for the LaNi5 and La0.7Ce0.1Ga0.3Ni5 compounds at T = 298 K and T =318 K.
Metals 14 00967 g006
Figure 7. Number of atoms per site as a function of temperature for the LaNi5 (a) and La0.7Ce0.1Ga0.3Ni5 (b) compounds.
Figure 7. Number of atoms per site as a function of temperature for the LaNi5 (a) and La0.7Ce0.1Ga0.3Ni5 (b) compounds.
Metals 14 00967 g007
Figure 8. Energies of absorption for the LaNi5 (a) and La0.7Ce0.1Ga0.3Ni5 (b) compounds.
Figure 8. Energies of absorption for the LaNi5 (a) and La0.7Ce0.1Ga0.3Ni5 (b) compounds.
Metals 14 00967 g008
Figure 9. Internal energy versus pressure when T = 298 K and T = 318 K for the LaNi5 (a) and La0.7 Ce0.1Ga0.3Ni5 (b) compounds.
Figure 9. Internal energy versus pressure when T = 298 K and T = 318 K for the LaNi5 (a) and La0.7 Ce0.1Ga0.3Ni5 (b) compounds.
Metals 14 00967 g009
Figure 10. Gibbs energy versus pressure when T = 298 K and T = 318 K for the LaNi5 (a) and La0.7 Ce0.1Ga0.3Ni5 (b) compounds.
Figure 10. Gibbs energy versus pressure when T = 298 K and T = 318 K for the LaNi5 (a) and La0.7 Ce0.1Ga0.3Ni5 (b) compounds.
Metals 14 00967 g010
Table 1. Hysteresis rate calculation.
Table 1. Hysteresis rate calculation.
CompoundTemperature (K)Peq (Absorption) (bar)Peq (Desorption) (bar)Hys
LaNi52984.504.000.11
3185.605.200.07
La0.7Ce0.1Ga0.3Ni52984.884.450.09
3186.055.750.05
Table 2. Fitting parameters of the LaNi5- and La0.7Ce0.1Ga0.3Ni5-doped compounds.
Table 2. Fitting parameters of the LaNi5- and La0.7Ce0.1Ga0.3Ni5-doped compounds.
LaNi5 Compound
T (K)n1n2nNm1Nm2P1P2R-Square
29831.813.8817.840.040.0224.521.800.98
31822.290.1711.230.060.225.620.260.98
La0.7Ce0.1Ga0.3Ni5 Compound
2981.5044.6523.080.090.032.714.920.98
3180.5427.6214.080.120.050.876.070.99
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

Belkhiria, S.; Alsawi, A.; Hraiech, I.; Dhaou, M.H.; Jemni, A. Experimental and Mathematical Investigation of Hydrogen Absorption in LaNi5 and La0.7Ce0.1Ga0.3Ni5 Compounds. Metals 2024, 14, 967. https://doi.org/10.3390/met14090967

AMA Style

Belkhiria S, Alsawi A, Hraiech I, Dhaou MH, Jemni A. Experimental and Mathematical Investigation of Hydrogen Absorption in LaNi5 and La0.7Ce0.1Ga0.3Ni5 Compounds. Metals. 2024; 14(9):967. https://doi.org/10.3390/met14090967

Chicago/Turabian Style

Belkhiria, Sihem, Abdulrahman Alsawi, Ibtissem Hraiech, Mohamed Houcine Dhaou, and Abdelmajid Jemni. 2024. "Experimental and Mathematical Investigation of Hydrogen Absorption in LaNi5 and La0.7Ce0.1Ga0.3Ni5 Compounds" Metals 14, no. 9: 967. https://doi.org/10.3390/met14090967

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