Next Article in Journal
An Electromagnetic Design of a Fully Superconducting Generator for Wind Application
Next Article in Special Issue
Experimental Validation and Numerical Simulation of a Hybrid Sensible-Latent Thermal Energy Storage for Hot Water Provision on Ships
Previous Article in Journal
Machine Learning Techniques in the Energy Consumption of Buildings: A Systematic Literature Review Using Text Mining and Bibliometric Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Mesoporous γ-Alumina Support for Water Composite Sorbents for Low Temperature Sorption Heat Storage

by
Manca Ocvirk
1,2,
Alenka Ristić
1,* and
Nataša Zabukovec Logar
1,3
1
National Institute of Chemistry, Hajdrihova 19, 1000 Ljubljana, Slovenia
2
Jožef Stefan International Postgraduate School, Jamova Cesta 39, 1000 Ljubljana, Slovenia
3
School of Science, University of Nova Gorica, Vipavska Cesta 13, 5000 Nova Gorica, Slovenia
*
Author to whom correspondence should be addressed.
Energies 2021, 14(22), 7809; https://doi.org/10.3390/en14227809
Submission received: 15 October 2021 / Revised: 15 November 2021 / Accepted: 19 November 2021 / Published: 22 November 2021
(This article belongs to the Special Issue New Trends in Thermal Energy Storage: Materials and Technologies)

Abstract

:
The efficiency of thermochemical heat storage is crucially determined by the performance of the sorbent used, which includes a high sorption capacity and a low regeneration temperature. The thermochemical salt hydrate– γ-alumina composite sorbents are promising materials for this application but lack systematic study of the influence of γ-alumina structural properties on the final storage performance. In this study, mesoporous γ-Al2O3 supports were prepared by solvothermal and hydrothermal synthesis containing a block copolymer (F-127) surfactant to design thermochemical CaCl2 and LiCl composite water sorbents. Altering the solvent in the synthesis has a significant effect on the structural properties of the γ-Al2O3 mesostructure, which was monitored by powder XRD, nitrogen physisorption, and SEM. Solvothermal synthesis led to a formation of mesoporous γ-Al2O3 with higher specific surface area (213 m2/g) and pore volume (0.542 g/cm3) than hydrothermal synthesis (147 m2/g; 0.414 g/cm3). The highest maximal water sorption capacity (2.87 g/g) and heat storage density (5.17 GJ/m3) was determined for W-46-LiCl containing 15 wt% LiCl for space heating, while the best storage performance in the sense of fast kinetics of sorption, without sorption hysteresis, low desorption temperature, very good cycling stability, and energy storage density of 1.26 GJ/m3 was achieved by W-46-CaCl2.

1. Introduction

Thermochemical energy storage (TCES), as one of the three technologies of thermal energy storage (TES), can reduce fossil fuel consumption, which still dominates building space heating, by shifting some of the solar thermal energy collected in summer to winter [1]. Thermochemical heat storage uses the reversible chemical reaction (reaction of water and CaCl2) and/or sorption processes of gases in solids or liquids (water sorption on porous solids). Water sorption heat storage consists of two phases. In the first phase, which is also called charging or desorption, gas or water vapour is desorbed from the material under solar radiation or waste heat. In the second phase or discharging (adsorption), the water vapour is adsorbed on the material with heat release, which can be used for space heating. The efficiency of the sorption technology, based on the alternating (ad)sorption (exothermic phenomenon) and desorption (endothermic phenomenon) of the working fluid on the sorption materials, is determined by the performance of the used sorbent, which should have a high sorption capacity, and consequently a high energy storage density, a low regeneration temperature of 80 to 120 °C, stability in humid conditions and also at temperatures up to 120 °C, and no hysteresis (of the sorption and desorption curves) during the sorption cycles. These requirements can be accomplished by the thermochemical (TCM) composite sorbents [2], formed of various porous supports and hygroscopic salt hydrates (halides, nitrates, sulphates) and are currently the most studied sorbents, especially when water is used as the working fluid. The process of water sorption in these composites can consist of three mechanisms: adsorption of water on the active support, chemical reaction between water and entrapped salt in the pores of the support, and absorption of water in salt solution in the pores [3]. The main advantage of these sorbents is tailoring their sorption properties by changing the structural properties and chemical composition of the support, and by entrapping of the particular hygroscopic salt hydrate nanoparticles in the support.
A brief review of the literature on the composites for thermochemical energy storage has shown that the increase in sorption capacity has mainly occurred by changing the salt hydrate and increasing its amount [4]. Fewer TCES studies have been carried out on the development of the active porous support, which can be involved in the adsorption process to further improve the storage performance in the context of increased sorption capacity by adjusting the chemical composition of the support and creating additional adsorption sites by using different synthesis routes [5,6,7,8]. Porous supports for TCM composites found in the literature are mainly commercial materials such as Zeolites [9], silica gels [10], activated alumina [11], and vermiculite [12], while recently, novel supports such as MOFs [13], polymeric foam [14], alginate [15], and porous carbon structures [16,17,18] have been used.
An important task in the preparation of salt hydrate TCM composites is the selection of the support for the salt inclusion. The support should meet certain requirements such as the ability to provide a stable nano-environment for salt confinement, hydrothermal stability, stability of the pore structure under working conditions, inertness to salt hydrates, and good thermal conductivity, to name a few, to enhance the performance of the composite. In addition, a simple and environmentally friendly synthesis that leads to the preparation of the composite and does not show hysteresis during sorption is of great importance [19].
To date, TCM composite sorbents based on mesoporous alumina (Al2O3) [20,21] or activated alumina [11,22,23,24] have been prepared from commercially available aluminas and combined with the following salt hydrates: MgSO4, LiCl, and CaCl2. These alumina composites have been investigated for thermochemical heat storage at low temperature (<100 °C). Recently, anodic alumina was prepared by anodizing of aluminium sheet and used for the salt composites [25,26]. In all these studies, only the physicochemical properties of the alumina support, such as specific surface area or pore volume and pore size, were shown, and it was not evident whether crystalline or amorphous alumina was used. Indeed, aluminas exist as amorphous or crystalline phases of various oxides and oxyhydroxides. The crystalline structures, which are mainly composed of γ, δ, ρ and α phases, exhibit different properties. All these structures can be prepared by irreversible thermal transformation of the crystalline aluminium hydroxides and salts or by colloidal gel precipitation. α-Al2O3 is the high-temperature end product of the thermal treatment [27]. For example, the γ-Al2O3 support is often described as robust, thermally stable, and easy to shape. In addition, it offers acid-base properties with polarity [28], favourable surface properties, such as high specific surface area (100–400 m2/g) and a tendency to disperse the active phase. Together with its low price, alumina is at the top of the list of the used supports in catalysis and adsorption [29]. The thermal and hydrothermal stability strongly depends on the synthetic methods and conditions.
Mesoporous alumina is most commonly obtained by sol-gel synthesis using soft or hard templates and by evaporation-induced self-assembly (EISA). Sol-gel synthesis in the presence of soft templates is one of the most common methods for preparing mesoporous alumina. The soft templates can be anionic (stearic acid, etc.), cationic (CTAB, etc.), nonionic (P123, etc.) surfactants, as well as non-surfactants (ionic liquids) [29]. The alumina precursor, usually aluminium isopropoxide, and the surfactant were dissolved in alcoholic solvent with the minimum amount of water added to promote hydrolysis of the isopropoxide and start the condensation reaction. Further evaporation of the solvent led to the formation of the ordered or disordered mesoporous inorganic substance [27,30,31]. Its final structural properties also depend significantly on the thermal treatment (e.g., calcination) during the removal of the template. Calcination is an important parameter in creating crystalline alumina with high surface area and has a great impact on the thermal, hydrothermal, and mechanical stability of alumina. The use of higher temperatures improves the crystallinity of porous materials but has a drawback, which can be seen in reduction in their mesostructural order as well as in the reduction in their pore volume and specific surface area [32].
The aim of this study is to design γ-alumina supports with mesoporous character, to determine the influence of the synthesis route using different solvents (water/ethanol/ethanolic solution) on the textural properties of the mesoporous γ-alumina support, to combine mesoporous γ-alumina supports with LiCl and CaCl2 salt hydrates, and to determine water sorption properties of these composites. In addition, the stability for the composites of the water-based γ-alumina (W-46) with CaCl2 and LiCl under humid conditions up to 100 °C has been evaluated. Crystalline γ-alumina has been selected as the support for the composite due to its thermal conductivity of 30 W/mK, which can further promote heat transfer in the system compared to silicate supports with a much lower thermal conductivity of 0.2 W/mK.
To the best of the author’s knowledge, this is the first report on the synthesis of crystalline mesoporous γ-alumina as a support for the TCM composite water sorbents.

2. Materials and Methods

2.1. Materials Synthesis

Pluronic F-127 (EO106PO70EO2106) and aluminium isopropoxide (98%) were purchased from Aldrich. Nitric acid (67%) and absolute anhydrous ethanol (99.5%) were purchased from Sigma-Aldrich (Darmstadt, Germany) and CARLO ERBA Reagents (Milan, Italy), respectively. Calcium chloride hexahydrate (98%) and lithium chloride hydrate (99%) were purchased from Sigma Aldrich (Darmstadt, Germany).

2.1.1. The γ-Alumina Support Synthesis

Mesoporous γ-alumina samples were synthesised in a two-step procedure. First, the evaporation-induced self-assembly (EISA) method was used. The following procedure is an adaptation of the synthesis described in [31]. The molar ratio of the reactants was: 1 AIP: 0.006 F-127: 2.3 HNO3: 68 ethanol or water and 1 AIP: 0.006 F-127: 2.3 HNO3: 34 ethanol: 34 water.
In all syntheses, two solutions were prepared: solution A containing solvent (an ethanol or a water or an ethanolic solution) and aluminium isopropoxide (AIP), and solution B containing a triblock copolymer non-ionic surfactant F-127, the same solvent as in solution A and nitric acid to aid solubility and hydrolysis.
Both types of solutions (A and B) were stirred at room temperature until completely dissolved and then mixed together and stirred at room temperature for 5 h. After solvent evaporation at 60 °C for 48 h in an oven, a semi-transparent solid was obtained. In the second step, the obtained solids were further solvothermally (ethanol) or hydrothermally (water, ethanolic solution) treated in Teflon-lined stainless steel autoclaves at 160 °C for 12 h.
The as-prepared samples were calcined in the furnace at 800 °C for 3 h under air flow with a temperature ramp of 1.6 °C/min to remove the surfactant and achieve the transformation of the precursor phase to crystalline γ-alumina.
The prepared support samples were named after the solvent used: water (W-46), ethanol (E-37), and ethanolic solution (ES-47).

2.1.2. Composite Preparation

The mesoporous γ-Al2O3 prepared with water and ethanolic solution was preheated in the oven at 120 °C for 2 h. A predetermined quantity of CaCl2·2H2O and LiCl·H2O was dissolved in 1 mL of deionized water in order to reach the required salt amount. The prepared salt solutions were then spread to the dehydrated mesoporous γ-Al2O3 using the incipient wetness impregnation method. The composites were dried over night at room temperature.

2.2. Methods

XRPD patterns were collected on a PANalytical X’Pert PRO high-resolution diffractometer (Almelo, The Netherlands) from 10 to 70° 2θ, and a fully opened X’Celerator detector.
Nitrogen physisorption was performed at 77 K on a Tristar3020 volumetric analyzer (Micromeritics, Norcross, GA, USA). The nonporous salt does not contribute to nitrogen adsorption to a large extent, so the amount of the salt was considered, and nitrogen isotherms, specific surface area, and pore volume values were corrected. The BET specific surface area was calculated from adsorption data ranging from 0.05 to 0.2 of relative pressure [33]. The total pore volume was evaluated at a relative pressure of 0.98. The KJS method [34] was used to determine the pore size distribution (PSD), and the maximum of the PSD was used for the pore diameter determination [35].
SEM images were recorded on Verios 4G HP field-emission gun microscope (Thermo Fisher Scientific, Waltham, MA, USA).
Energy dispersive X-ray analysis (EDAX) using Zeiss SupraTM 3VP microscope (Jena, Germany) and ICP-OES using spectrometer ES 715 (Varian, Palo Alto, Santa Clara, CA, USA) were performed for elemental analysis.
The desorption temperatures of W-46 and ES-47 composites were determined using a Q5000IR thermogravimetric analyser (the accuracy ±0.01 µg; ±0.1 °C) (TA Instruments, Inc., New Castle, DE, USA) where the maximum (Tpeak) of the DTG curve was considered the desorption temperature [36]. The samples were heated in air at a heating rate of 10 °C/min. Before thermogravimetric measurements, the samples were kept over a saturated potassium sulphate solution for 5 days.
The surface charge of the W-46 and ES-47 supports was measured using Zetasizer nano ZS (Malvern, UK) instrument.
Water sorption isotherms were measured using the IGA-100 gravimetric analyser (the accuracy ±0.1 µg) (Hiden Isochema Ltd., Warrington, UK) at 25 and 35 °C and used to the calculate the value of the differential enthalpy of sorption according to the well-known Clausius–Clapeyron equation. The equal pressure intervals of 1.6 mbar between vacuum and 40 mbar (saturation vapor pressure of 56.3 mbar) were set with an equilibrium time of 80 s. The samples were degassed at <10−5 mbar at 25 °C (the accuracy ±0.1 °C) for 4 h to a constant mass. The definition of the thermodynamic heat cycle is shown in [37], and the calculation of the amount of heat involved can be found in [38]. The integral enthalpy of sorption Qsor can be considered as the achievable heat storage density at the material level and was calculated by using the following equation: Qsor = ΔHsor (wsor − wdes) [kJ/kgsor]. ΔHsor [kJ/kgwater] is the differential enthalpy of sorption according to the sorbed water amount, and wsor and wdes [kgwater/kgsor] are the maximum and minimum sorbed water amount over the sorbent at the given boundary conditions. The specific heat of the γ-alumina is 94.6 kJ/mol K at 25 °C [39]. The integral heat of sorption was calculated for the application of heat storage.
The sorption/desorption cycling performance [38] of the W-46-CaCl2 and W-46-LiCl composites were carried out on IGAsorp-XT automated water sorption gravimetrical analyser (the accuracy ±0.1 µg) (Hiden Isochema Ltd., Warrington, UK.) with the sequential procedure of water uptake measurements at 35 °C (the accuracy ±0.1 °C) under moist nitrogen gas flow (flow rate of 250 mL/min and 80% relative humidity) for 3 h, followed by drying at 100 °C with dry nitrogen flow for 2 h. The sequence was repeated 25 times according to the literature [40] at a desorption temperature of 100 °C, which can be attained by solar thermal collectors. The sorption temperature was fixed to 35 °C, which is sufficient for space heating applications. The water vapor pressure during desorption and sorption of the samples was set to 12.3 mbar (a dew point temperature of 10 °C). The difference in the amount of sorbed water at 35 °C and 100 °C at 12.5 mbar is the cycle (water) loading lift of the composite.

3. Results and Discussion

3.1. Structural Properties of All γ-Alumina Supports and Composites

All mesoporous γ-alumina supports were prepared by the same synthesis procedure, which differs only in the use of different solvents (i.e., ethanol, water, and ethanolic solution), while the other conditions remained the same. The calcined samples exhibited mesostructures of all γ-alumina with low crystallinity, which can be due to the high calcination temperature (800 °C) [29] after the second step of hydrothermal or solvothermal treatment. Grant et al. [31] obtained cage-like ordered mesoporous γ-alumina by one-step synthesis, i.e., evaporation-induced self-assembly synthesis and calcination at 400 °C. Powder XRD was used to identify the crystal structure of the products after thermal treatment at 800 °C. The calcined powders in Figure 1a exhibited diffraction peaks corresponding to the γ-Al2O3 phase, which is consistent with JCPDS 00-050-0741. It can be observed that the intensity of the characteristic peak at 2θ 66.5° decreases in the sample prepared by hydrothermal synthesis compared to the sample prepared by solvothermal synthesis, indicating higher crystallinity of the solvothermal product. All the diffraction peaks are quite broad, which can be due to nanosized particles of the γ-Al2O3 products. The diffraction peak at 2θ 44.2° in the W-46 sample belongs to the XRD sample holder.
γ-Al2O3 supports prepared from water (W-46) and ethanolic solution (ES-47) were used to prepare the composites, as their crystal structure was preserved after treatment at 100 °C for 5 h in vacuum. Introduction of CaCl2 and LiCl salts into mesoporous γ-Al2O3 prepared in water and ethanolic solution revealed that the supports were inert to the salts, meaning that the crystal structure of the supports did not change or react with these salts (Figure 1b). Additional peaks due to the presence of the salts were not observed, indicating that the highly dispersed salts were successfully confined of in the mesopores of these supports and/or that the salts were dispersed in particles less than 5 nm in size, which are not detected by XRD.
The morphology of the hydrothermally prepared γ-Al2O3 supports (E-37, W-46, and ES-47) was investigated using SEM (Figure 2). The results showed that the particular solvent used in the synthesis had a significant effect on the morphology of the samples. In water, larger 3 µm γ-Al2O3 agglomerates were formed, while 100 nm and 25–50 nm round nanoparticles were obtained from the ethanol and ethanolic solution-containing, respectively.
The amount of impregnated CaCl2 after the incipient wetness impregnation procedure on W-46 and ES-47 γ-Al2O3 supports was lower than expected, considering the amounts of inorganic salt used in the composites with silica [8]. This is reported to be influenced by the high polarity of the alumina [28] and pH. The dependence of surface charge of γ-alumina (W-46 and ES-47) on pH is illustrated in Figure 3. It can be seen that the solvents affect the surface charge of the prepared supports. At the pH value of CaCl2 and the LiCl solutions (pH = 5.5), the W-46 γ-alumina support exhibits a positively charged surface (29 mV), as does the ES-47 γ-alumina support (28 mV).
Thus, the introduction of calcium chloride or lithium chloride solution into the mesopores of the W-46 or ES-47 support during incipient wetness impregantion was hindered by similarly charged species, compared to silica mesostructures with negatively charged surface (−20 mV) [41]. The negatively charged surface (−10 mV) of the γ-alumina support W-46 can be obtained at a lower pH than ES-47. In general, the surface charge can be changed by modifying the synthesis procedure.
EDX analysis determined 10 wt% CaCl2 in both supports, while determination of LiCl was performed using ICP-OES. In the W-46 and ES-47 supports, 15 wt% and 10 wt% LiCl were determined, respectively.
Nitrogen physisorption measurements were used to evaluate the textural properties of γ-Al2O3 supports prepared from different solvents. The nitrogen adsorption isotherms of the γ-Al2O3 supports and their pore size distributions (PSDs) are shown in Figure 4a, the composites W-46 in Figure 4b, and the composites of ES-47 in Figure 4c. The textural parameters determined based on these isotherms are listed in Table 1 and Table 2. All isotherms are of Type IV, which is typical of mesoporous materials [42]. The isotherms exhibit a capillary condensation step, indicating the presence of mesoporosity. The sample E-37 shows a steeper capillary condensation, which is typical of a narrow pore size distribution. It can be clearly seen that the hysteresis loops change from the H1 loop for samples E-37 and ES-47 to the H3 loop for W-46, indicating different γ-Al2O3 support mesostructures due to the solvent used. The H3 loop is typical of materials with slit-like mesopores [42]. Macroporosity is found in the W-46 sample (Figure 4a) but not in the ES-47 and E-37 samples. The capillary condensation step for the ES-47 sample is shifted towards higher relative pressures, indicating an increase in the size of the mesopores due to solvothermal synthesis [43].
The E-37 sample has the highest specific surface area and pore volume (Table 1), while samples ES-47 and W-46 have lower specific surface areas and pore volumes. The largest size of mesopores (11.3 nm) was determined from the maximum of pore size distribution (Figure 4a), for sample, ES-47 prepared from ethanolic solution. On the other hand, the smallest size of mesopores (9.1 nm) was determined for the W-46 sample prepared hydrothermally.
The inclusion of salts in both supports resulted in less intense hysteresis loops (Figure 4b,c) and pore size distributions (Figure 4b,c), accompanied by a decrease in pore sizes, pore volumes, and specific surface areas (Table 2). It can be concluded that the salt nanoparticles been dispersed inside of mesopores of the supports. For example, after the introduction of LiCl into the W-46 support, the specific surface area (68 m2/g) and pore volume (0.196 cm3/g) of the composite decreased significantly (Table 2). The pore diameter (8.7 nm) of this composite was larger than the pore diameter (8.3 nm) of the composite with CaCl2.
Desorption temperatures for the composites were determined from the DTG curves for composites W-46 and ES-47, which are shown in Figure 5. The desorption temperature of water was completed for the W-46 samples in the following order: the W-46-CaCl2 sample (98 °C), the W-46-LiCl (88 °C), and the W-46 support (29 °C) (Figure 5a). Additionally, 88.0% of the water was desorbed from the composite containing LiCl, while a smaller amount of water (74.9%) was removed from the composite containing CaCl2. The DTG curves show one water loss for all samples prepared in water and ethanolic solution. Water can be removed from W-46 composites at 120 °C. For ES-47 composites, water desorption can be completed even at a lower temperature (100 °C). Desorption temperature for the ES-47-LiCl composite was 68 °C, while a slightly lower desorption temperature (62 °C) was observed for ES-47-CaCl2 (Figure 5b). Additionally, 75.0% and 60.8% of water was removed from the ES-47-LiCl and ES-47-CaCl2 composites, respectively.

3.2. Water Sorption Properties

The mesoporous γ-Al2O3 and composites exhibit Type V water sorption isotherms (Figure 6a), measured at 35 °C. The γ-Al2O3 supports showed weak hydrophilic properties, which are also characteristic of mesoporous silica materials [8]. It was observed that the water uptake increased slowly up to 0.08 g/g at a relative pressure of 0.7. Thereafter, the water sorption capacity increased suddenly from 0.08 g/g to 0.31 g/g, which is due to the mechanism of capillary condensation and shows the active role of the supports. Jabbari-Hichri [21] reported that mesoporous alumina impregnated with 14 wt% CaCl2 showed a water sorption capacity of 0.17 g/g. The composite W-46-LiCl (2.89 g/g) exhibited the highest maximal water sorption capacity of all composites, as determined from water sorption isotherms measured at 35 °C. The ES-47-LiCl sample showed the maximal water uptake of 1.00 g/g. Composites containing CaCl2 showed lower maximal water sorption capacities: 0.76 g/g for W-46-CaCl2 and 0.46 g/g for ES-47-CaCl2. At a relative pressure of 0.4, the W-46 composite with 10 wt% CaCl2, showed 4.5 times greater water uptake (4% versus 18%) than the support W-46 (not shown), while the composite with 15 wt% LiCl revealed 18 times greater water uptake than its support (4% versus 71%). It can be concluded that the presence of the salts in the γ-Al2O3 support increases the water sorption capacity of the composites, and the salt content affects the sorption performance of these composites. It should be emphasized that, among all the composites, only the composite W-46-CaCl2 exhibits a water sorption isotherm without hysteresis loop (Figure 6a) up a relative pressure of 0.4, which may further improve the heat storage performance. Small hysteresis is seen in W-46-LiCl, while both ES-47 composites show larger hysteresis of the water sorption isotherm (Figure 6a inset).
The comparison of the water sorption isotherms of W-46-CaCl2 and ES-47-CaCl2 revealed differences in the range of 0 < p/p0 < 0.4, with the W-46-CaCl2 sample showing slightly higher uptake, which is most likely due to the preparation procedure of the support in water. The water uptake at 0.4 relative pressure of the composite ES-47-CaCl2, which has uniform mesopores with an average pore size of 10.2 nm and a total pore volume (0.207 cm3/g), was 1.7% lower than the water uptake of the composite W-46-CaCl2 with a smaller pore diameter (8.3 nm) and a larger total pore volume (0.252 cm3/g). The water uptake curve of the composite with W-46-γ-Al2O3 and 15 wt% LiCl showed a plateau at 0.038 p/po due to the formation of lithium chloride hydrate [44], while this plateau was not observed for the composites with lower salt content.
The kinetic curves (Figure 6b) show that the maximal water uptake on the composite ES-47-LiCl was reached in the shortest time. The composite ES-47-CaCl2 showed a slow sorption rate in the first two hours, increasing with a slow rate after 2 h. The water sorption on the composites W-46-LiCl and ES-47-LiCl was also slow in the first two hours and then gradually increased at a higher rate than that of the composites containing CaCl2. Water uptake occurred in the following order: ES-47-LiCl, W-46-CaCl2, ES-47-CaCl2, W-46-LiCl.
The integral heat of sorption is one of the most relevant parameters for evaluating the effectiveness of sorbent at the material level for TES. For sorption heat storage the adsorbents with high water loading lift between adsorption and desorption can reach high storage densities. The integral heat of sorption was calculated for the given boundary conditions for space heating [8]: sorption/condensation temperature of 30 °C, desorption temperature of 100 °C, and evaporation temperature of 10 °C. The integral heat of sorption Qsor of all composites is listed in Table 3. Composites containing similar amounts of LiCl and CaCl2 showed similar theoretical energy storage density values at material level ranging from 1.26 to 1.37 GJ/m3, while lower values were obtained with composites containing mesoporous silica and the same amount of CaCl2 [8].
The increased amount of the salt (15 wt% LiCl) in the composite increased the calculated water loading lifts, and consequently, theoretical energy storage density up to 5.17 GJ/m3 at material level. The physical and structural properties of the mesoporous γ-Al2O3 support, such as high density and the suitable mesostructure with the pore size of up to 10 nm, are advantageous for the sorption heat storage performance, which can be further enhanced by the high thermal conductivity of the γ-alumina support. A direct comparison of the energy storage density of these composites with other alumina-containing composites was difficult, because it strongly depends on the boundary conditions. The value of energy storage density of the composite with LiCl (14.7 wt.%) and activated alumina prepared by wet impregnation equalled 1.14 GJ/m3, considering a desorption temperature of 120 °C, sorption temperature of 20 °C, and 80% relative humidity [11]. This composite was also tested in the open sorption TES system for space heating at a desorption temperature of 110 °C and reached 0.69 GJ/m3 [22]. The influence of different boundary conditions on the integral heat of sorption was evaluated (Figure 7) for the conditions of the space heating storage cycle defined in [1] for the Central European climate: sorption temperature of 35 °C, evaporation temperature of 5 °C, desorption temperature of 90 °C, and condensation temperature of 30 °C. Frazzica [1] used the condensation temperature of 30 °C as the average ambient temperature during the day in summer in Central Europe. The evaporation temperature was set at 5 °C, which was explained by the situation of some rigid climatic conditions where geothermal boreholes were used to provide ambient heat for the evaporator and satisfactory heat transfer inside the component.
It can be seen that lower desorption temperature decreases the energy storage density of the composites and still promotes their use for low-temperature thermal energy storage.
The cycling stability of the W-46-CaCl2 (without hysteresis during the sorption cycle up to a relative pressure of 0.4) and W-46-LiCl (small hysteresis during the sorption cycle up to relative pressure of 0.4) composites was tested during 25 sorption and desorption cycles between 35 and 100 °C at 12 mbar, showing a small reduction in water uptake of 2.0% for W-46-CaCl2 and 1.5% for W-46-LiCl after the last cycle (Figure 8). After the cycle tests, no salt leakage was observed for these samples. This confirms that the mesoporous γ-Al2O3 support can be used as a stable nano-environment for the entrapment of calcium and lithium chloride. These composites are promising candidates for low-temperature thermal energy storage.

4. Conclusions

Mesoporous γ-alumina supports were synthesized from gels containing aluminium isopropoxide, the surfactant Pluronic F-127, and various solvents: ethanol, water, or ethanolic solution. This was undertaken by the two-step synthesis procedure, which included the evaporation-induced self-assembly method followed by solvothermal or hydrothermal treatment at 160 °C for 12 h and calcination at 800 °C to obtain the crystalline phase of γ-Al2O3. The solvents affected the uniformity of the mesopores, the specific surface area, and the morphology of the γ-Al2O3 supports. The highest degree of uniformity of mesopores and surface area was found in the sample E-37 prepared in etanol. The uniformity of the samples ranged from the E-37 (ethanol) and ES-47 (etanolic solution) to W-46 (water). The solvent also influenced the surface charge of the γ-Al2O3 support and, consequently, the amount of the salt inclusion in the support performed by incipient wetness impregnation. The desorption temperatures of all composites varied from 62 to 98 °C, indicating low temperature heat storage utilization.
Among all composites, the highest heat storage density (5.17 GJ/m3) was achieved for W-46-LiCl at boundary conditions for space heating at the sorption and condensation temperature of 30 °C, the desorption temperature of 100 °C, and the evaporation temperature of 10 °C. The composite W-46-CaCl2 showed the best heat storage performance in terms of fast kinetics of water sorption, no sorption–desorption hysteresis, low desorption temperature, and very good cycling stability and energy storage density of 1.26 GJ/m3. The fastest rate of water sorption was found for ES-47-LiCl composite, where the sorption of water reached its maximum after 9 h. Moreover, after 25 cycles between 35 °C and 100 °C at 12.5 mbar, small decreases in water uptake of 1.5 and 2.0 % were observed for the W-46-CaCl2 and W-46-LiCl composites, respectively, indicating the very good hydrothermal stability of these composites.
The properties of mesoporous γ-Al2O3 supports such as high density, particular mesostructure, high thermal conductivity, and structure stability are advantageous characteristics for the preparation and use of these TCM composites as low-temperature sorption heat storage materials. This study opens up several possibilities to tailor the porosity of mesostructured oxide supports by proper selection of solvents, reagents, and synthesis routes. Moreover, it represents an important step towards the development of mesostructured metal oxides with designed properties for TCM materials, which are of great importance for thermochemical energy storage.

Author Contributions

Conceptualization, A.R.; formal analysis, M.O., A.R.; investigation, M.O.; writing—original draft preparation, A.R.; writing—review and editing, N.Z.L.; visualization A.R.; supervision, A.R.; project administration, A.R.; funding acquisition, N.Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

The Slovenian Research Agency is thanked for the support of the research program P1-0021 (Nanoporous materials).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Suzana Mal for water sorption and nitrogen physisorption measurements, Andraž Krajnc for cycling stability test and Matjaž Mazaj for SEM pictures.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Frazzica, A.; Brancato, V.; Caprì, A.; Cannilla, C.; Gordeeva, L.G.; Aristov, Y.I. Development of “Salt in Porous Matrix” Composites Based on LiCl for Sorption Thermal Energy Storage. Energy 2020, 208, 118338. [Google Scholar] [CrossRef]
  2. Gunasekara, S.N.; Barreneche, C.; Inés, A.; Díaz, A.C.; Ristić, A.; Weinberger, P.; Paksoy, H.Ö.; Koçak, B.; Rathgeber, C.; Chiu, N.; et al. Thermal Energy Storage Materials ( TESMs)—What Does It Take to Make Them Fly ? Crystals 2021, 11, 1276. [Google Scholar] [CrossRef]
  3. Gordeeva, L.G.; Aristov, Y.I. Composites “salt inside Porous Matrix” for Adsorption Heat Transformation: A Current State-of-the-Art and New Trends. Int. J. Low-Carbon Technol. 2012, 7, 288–302. [Google Scholar] [CrossRef] [Green Version]
  4. Zbair, M.; Bennici, S. Survey Summary on Salts Hydrates and Composites Used in Thermochemical Sorption Heat Storage: A Review. Energies 2021, 14, 3105. [Google Scholar] [CrossRef]
  5. Tokarev, M.; Gordeeva, L.; Romannikov, V.; Glaznev, I.; Aristov, Y. New Composite Sorbent CaCl2 in Mesopores for Sorption Cooling/Heating. Int. J. Therm. Sci. 2002, 41, 470–474. [Google Scholar] [CrossRef]
  6. Ristić, A.; Maučec, D.; Henninger, S.K.; Kaučič, V. New Two-Component Water Sorbent CaCl2-FeKIL2 for Solar Thermal Energy Storage. Microporous Mesoporous Mater. 2012, 164, 266–272. [Google Scholar] [CrossRef]
  7. Silvester, L.; Touloumet, Q.; Kamaruddin, A.; Chassagneux, F.; Postole, G.; Auroux, A.; Bois, L. Influence of Silica Functionalization on Water Sorption and Thermochemical Heat Storage of Mesoporous SBA-15/CaCl2 Composites. ACS Appl. Energy Mater. 2021, 4, 5944–5956. [Google Scholar] [CrossRef]
  8. Ristić, A.; Zabukovec Logar, N. New Composite Water Sorbents CaCl2-PHTS for Low-Temperature Sorption Heat Storage: Determination of Structural Properties. Nanomaterials 2018, 9, 27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Nonnen, T.; Beckert, S.; Gleichmann, K.; Brandt, A.; Unger, B.; Kerskes, H.; Mette, B.; Bonk, S.; Badenhop, T.; Salg, F.; et al. A Thermochemical Long-Term Heat Storage System Based on a Salt/Zeolite Composite. Chem. Eng. Technol. 2016, 39, 2427–2434. [Google Scholar] [CrossRef]
  10. D’Ans, P.; Skrylnyk, O.; Hohenauer, W.; Courbon, E.; Malet, L.; Degrez, M.; Descy, G.; Frère, M. Humidity Dependence of Transport Properties of Composite Materials Used for Thermochemical Heat Storage and Thermal Transformer Appliances. J. Energy Storage 2018, 18, 160–170. [Google Scholar] [CrossRef]
  11. Zhang, Y.N.; Wang, R.Z.; Li, T.X. Thermochemical Characterizations of High-Stable Activated Alumina/LiCl Composites with Multistage Sorption Process for Thermal Storage. Energy 2018, 156, 240–249. [Google Scholar] [CrossRef]
  12. Brancato, V.; Gordeeva, L.G.; Caprì, A.; Grekova, A.D.; Frazzica, A. Experimental Comparison of Innovative Composite Sorbents for Space Heating and Domestic Hot Water Storage. Crystals 2021, 11, 476. [Google Scholar] [CrossRef]
  13. Permyakova, A.; Wang, S.; Courbon, E.; Nouar, F.; Heymans, N.; D’Ans, P.; Barrier, N.; Billemont, P.; De Weireld, G.; Steunou, N.; et al. Design of Salt-Metal Organic Framework Composites for Seasonal Heat Storage Applications. J. Mater. Chem. A 2017, 5, 12889–12898. [Google Scholar] [CrossRef]
  14. Calabrese, L.; Bonaccorsi, L.; Bruzzaniti, P.; Gullì, G.; Freni, A.; Proverbio, E. Zeolite Filled Siloxane Composite Foams: Compression Property. J. Appl. Polym. Sci. 2018, 135, 46145. [Google Scholar] [CrossRef]
  15. Kallenberger, P.A.; Posern, K.; Linnow, K.; Brieler, F.J.; Steiger, M.; Fröba, M. Alginate-Derived Salt/Polymer Composites for Thermochemical Heat Storage. Adv. Sustain. Syst. 2018, 2, 1–8. [Google Scholar] [CrossRef]
  16. Frazzica, A.; Brancato, V.; Dawoud, B. Unified Methodology to Identify the Potential Application of Seasonal Sorption Storage Technology. Energies 2020, 13, 1037. [Google Scholar] [CrossRef] [Green Version]
  17. Wang, L.W.; Metcalf, S.J.; Critoph, R.E.; Tamainot-Telto, Z.; Thorpe, R. Two Types of Natural Graphite Host Matrix for Composite Activated Carbon Adsorbents. Appl. Therm. Eng. 2013, 50, 1652–1657. [Google Scholar]
  18. Yan, T.; Li, T.; Xu, J.; Chao, J.; Wang, R.; Aristov, Y.I.; Gordeeva, L.G.; Dutta, P.; Murthy, S.S. Ultrahigh-Energy-Density Sorption Thermal Battery Enabled by Graphene Aerogel-Based Composite Sorbents for Thermal Energy Harvesting from Air. ACS Energy Lett. 2021, 6, 1795–1802. [Google Scholar] [CrossRef]
  19. An, G.; Zhang, Y.; Wang, L.; Zhang, B. An Advanced Composite Sorbent with High Thermal Stability and Superior Sorption Capacity without Hysteresis for a Better Thermal Battery. J. Mater. Chem. A 2020, 8, 11849–11858. [Google Scholar] [CrossRef]
  20. Dawoud, B.; Aristov, Y. Experimental Study on the Kinetics of Water Vapor Sorption on Selective Water Sorbents, Silica Gel and Alumina under Typical Operating Conditions of Sorption Heat Pumps. Int. J. Heat Mass Transf. 2003, 46, 273–281. [Google Scholar] [CrossRef]
  21. Jabbari-Hichri, A.; Bennici, S.; Auroux, A. CaCl2-Containing Composites as Thermochemical Heat Storage Materials. Sol. Energy Mater. Sol. Cells 2017, 172, 177–185. [Google Scholar] [CrossRef]
  22. Zhang, Y.N.; Wang, R.Z.; Li, T.X. Experimental Investigation on an Open Sorption Thermal Storage System for Space Heating. Energy 2017, 141, 2421–2433. [Google Scholar] [CrossRef]
  23. Xu, S.Z.; Wang, R.Z.; Wang, L.W.; Zhu, J. Performance Characterizations and Thermodynamic Analysis of Magnesium Sulfate-Impregnated Zeolite 13X and Activated Alumina Composite Sorbents for Thermal Energy Storage. Energy 2019, 167, 889–901. [Google Scholar] [CrossRef]
  24. Wu, H.; Salles, F.; Zajac, J. A Critical Review of Solid Materials for Low-Temperature Thermochemical Storage of Solar Energy Based on Solid-Vapour Adsorption in View of Space Heating Uses. Molecules 2019, 24, 945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Grekova, A.D.; Girnik, I.S.; Nikulin, V.V.; Tokarev, M.M.; Gordeeva, L.G.; Aristov, Y.I. New Composite Sorbents of Water and Methanol “Salt in Anodic Alumina”: Evaluation for Adsorption Heat Transformation. Energy 2016, 106, 231–239. [Google Scholar] [CrossRef]
  26. Henninger, S.K.; Ernst, S.J.; Gordeeva, L.; Bendix, P.; Fröhlich, D.; Grekova, A.D.; Bonaccorsi, L.; Aristov, Y.; Jaenchen, J. New Materials for Adsorption Heat Transformation and Storage. Renew. Energy 2017, 110, 59–68. [Google Scholar] [CrossRef]
  27. Afshar Taromi, A.; Kaliaguine, S. Synthesis of Ordered Mesoporous Γ-Alumina—Effects of Calcination Conditions and Polymeric Template Concentration. Microporous Mesoporous Mater. 2017, 248, 179–191. [Google Scholar] [CrossRef]
  28. Trueba, M.; Trasatti, S.P. γ-Alumina as a Support for Catalysts: A Review of Fundamental Aspects. Eur. J. Inorg. Chem. 2005, 2005, 3393–3403. [Google Scholar] [CrossRef]
  29. Xu, X.; Megarajan, S.K.; Zhang, Y.; Jiang, H. Ordered Mesoporous Alumina and Their Composites Based on Evaporation Induced Self-Assembly for Adsorption and Catalysis. Chem. Mater. 2020, 32, 3–26. [Google Scholar] [CrossRef]
  30. Žilková, N.; Zukal, A.; Čejka, J. Synthesis of Organized Mesoporous Alumina Templated with Ionic Liquids. Microporous Mesoporous Mater. 2006, 95, 176–179. [Google Scholar] [CrossRef]
  31. Grant, S.M.; Vinu, A.; Pikus, S.; Jaroniec, M. Adsorption and Structural Properties of Ordered Mesoporous Alumina Synthesized in the Presence of F127 Block Copolymer. Colloids Surf. A: Physicochem. Eng. Asp. 2011, 385, 121–125. [Google Scholar] [CrossRef]
  32. Gonçalves, A.A.S.; Costa, M.J.F.; Zhang, L.; Ciesielczyk, F.; Jaroniec, M. One-Pot Synthesis of MeAl2O4 (Me = Ni, Co, or Cu) Supported on γ-Al2O3 with Ultralarge Mesopores: Enhancing Interfacial Defects in γ-Al2O3 to Facilitate the Formation of Spinel Structures at Lower Temperatures. Chem. Mater. 2018, 30, 436–446. [Google Scholar] [CrossRef] [Green Version]
  33. Brunauer, S.; Emmett, P.H.; Teller, E. Adsorption of Gases in Multimolecular Layers. J. Am. Chem. Soc. 1938, 60, 309–319. [Google Scholar] [CrossRef]
  34. Kruk, M.; Jaroniec, M.; Sayari, A. Application of Large Pore MCM-41 Molecular Sieves to Improve Pore Size Analysis Using Nitrogen Adsorption Measurements. Langmuir 1997, 13, 6267–6273. [Google Scholar] [CrossRef]
  35. Jaroniec, M.; Solovyov, L.A. Improvement of the Kruk-Jaroniec-Sayari Method for Pore Size Analysis of Ordered Silicas with Cylindrical Mesopores. Langmuir 2006, 22, 6757–6760. [Google Scholar] [CrossRef]
  36. Ristić, A.; Fischer, F.; Hauer, A.; Zabukovec Logar, N. Improved Performance of Binder-Free Zeolite y for Low-Temperature Sorption Heat Storage. J. Mater. Chem. A 2018, 6, 11521–11530. [Google Scholar] [CrossRef] [Green Version]
  37. Nuhnen, A.; Janiak, C. A Practical Guide to Calculate the Isosteric Heat/Enthalpy of Adsorption: Via Adsorption Isotherms in Metal-Organic Frameworks, MOFs. Dalton Trans. 2020, 49, 10295–10307. [Google Scholar] [CrossRef] [PubMed]
  38. Krajnc, A.; Varlec, J.; Mazaj, M.; Ristić, A.; Logar, N.Z.; Mali, G. Superior Performance of Microporous Aluminophosphate with LTA Topology in Solar-Energy Storage and Heat Reallocation. Adv. Energy Mater. 2017, 7, 1601815. [Google Scholar] [CrossRef]
  39. Calvin, J.J.; Asplund, M.; Zhang, Y.; Huang, B.; Woodfield, B.F. Heat Capacity and Thermodynamic Functions of Γ-Al2O3. J. Chem. Thermodyn. 2017, 112, 77–85. [Google Scholar] [CrossRef] [Green Version]
  40. Frazzica, A.; Brancato, V. Verification of Hydrothermal Stability of Adsorbent Materials for Thermal Energy Storage. Int. J. Energy Res. 2019, 43, 6161–6170. [Google Scholar] [CrossRef]
  41. Šuligoj, A.; Ristić, A.; Dražić, G.; Pintar, A.; Logar, N.Z.; Tušar, N.N. Bimetal Cu-Mn Porous Silica-Supported Catalyst for Fenton-like Degradation of Organic Dyes in Wastewater at Neutral PH. Catal. Today 2020, 358, 270–277. [Google Scholar] [CrossRef]
  42. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of Gases, with Special Reference to the Evaluation of Surface Area and Pore Size Distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef] [Green Version]
  43. Tušar, N.N.; Ristić, A.; Mali, G.; Mazaj, M.; Arčon, I.; Arčon, D.; Kaučič, V.; Logar, N.Z. MnO x Nanoparticles Supported on a New Mesostructured Silicate with Textural Porosity. Chem. A Eur. J. 2010, 16, 5783–5793. [Google Scholar] [CrossRef] [PubMed]
  44. Sun, Y.; Spieß, A.; Jansen, C.; Nuhnen, A.; Gökpinar, S.; Wiedey, R.; Ernst, S.J.; Janiak, C. Tunable LiCl@UiO-66 Composites for Water Sorption-Based Heat Transformation Applications. J. Mater. Chem. A 2020, 8, 13364–13375. [Google Scholar] [CrossRef]
Figure 1. Powder XRD patterns of (a) γ-Al2O3 products obtained after thermal treatment at 800 °C and (b) CaCl2 and LiCl composites with W-46 and ES-47 γ-Al2O3 supports.
Figure 1. Powder XRD patterns of (a) γ-Al2O3 products obtained after thermal treatment at 800 °C and (b) CaCl2 and LiCl composites with W-46 and ES-47 γ-Al2O3 supports.
Energies 14 07809 g001
Figure 2. SEM pictures of hydrothermally prepared mesoporous γ-Al2O3 supports: (a) E-37, (b) W-46, and (c) ES-47.
Figure 2. SEM pictures of hydrothermally prepared mesoporous γ-Al2O3 supports: (a) E-37, (b) W-46, and (c) ES-47.
Energies 14 07809 g002
Figure 3. Surface charge of γ-alumina supports W-46 and ES-47.
Figure 3. Surface charge of γ-alumina supports W-46 and ES-47.
Energies 14 07809 g003
Figure 4. Nitrogen sorption isotherms of (a) γ-alumina supports with corresponding PSDs, (b) W-46 and its composites with corresponding PSDs, (c) ES-47 and its composites with corresponding PSDs.
Figure 4. Nitrogen sorption isotherms of (a) γ-alumina supports with corresponding PSDs, (b) W-46 and its composites with corresponding PSDs, (c) ES-47 and its composites with corresponding PSDs.
Energies 14 07809 g004
Figure 5. TG/DTG curves of the hydrothermally prepared support (a) W-46 and its composites and (b) ES-47 and its composites.
Figure 5. TG/DTG curves of the hydrothermally prepared support (a) W-46 and its composites and (b) ES-47 and its composites.
Energies 14 07809 g005
Figure 6. (a) Water uptake curves and (b) water sorption kinetics curves of the composites with W-46 and ES-47 supports at 35 °C.
Figure 6. (a) Water uptake curves and (b) water sorption kinetics curves of the composites with W-46 and ES-47 supports at 35 °C.
Energies 14 07809 g006
Figure 7. The integral heat of sorption of the composites for space heating storage cycle: sorption temperature of 35 °C, evaporation temperature of 5 °C, desorption temperature of 90 °C, and condensation temperature of 30 °C.
Figure 7. The integral heat of sorption of the composites for space heating storage cycle: sorption temperature of 35 °C, evaporation temperature of 5 °C, desorption temperature of 90 °C, and condensation temperature of 30 °C.
Energies 14 07809 g007
Figure 8. Cycling stability of 25 cycles of sorption and desorption for W-46-CaCl2 and W-46-LiCl between 35 °C and 100 °C at 12.5 mbar.
Figure 8. Cycling stability of 25 cycles of sorption and desorption for W-46-CaCl2 and W-46-LiCl between 35 °C and 100 °C at 12.5 mbar.
Energies 14 07809 g008
Table 1. Textural properties of the mesoporous γ-Al2O3 supports.
Table 1. Textural properties of the mesoporous γ-Al2O3 supports.
Sample
Support
SolventSBET
(m2/g)
Vp
(cm3/g)
Average Pore Size
(nm)
E-37-γ-Al2O3Ethanol213 ± 10.54210.7
W-46-γ-Al2O3Water147 ± 10.4149.1
ES-47-γ-Al2O3Ethanolic solution148 ± 10.36711.3
Table 2. Textural properties of the of the W-46 and ES-47 composites.
Table 2. Textural properties of the of the W-46 and ES-47 composites.
Sample
Support
SBET
(m2/g)
Vp
(cm3/g)
Average Pore Size
(nm)
W-46-LiCl68 ± 10.1968.7
W-46-CaCl281 ± 10.2528.3
ES-47-LiCl96 ± 10.2049.3
ES-47-CaCl283 ± 10.20710.2
Table 3. Water loading lift and the integral heat of adsorption for the composites.
Table 3. Water loading lift and the integral heat of adsorption for the composites.
Sample∆w
(kg/kg)
Qsor
(kWh/m3)
Qsor
(GJ/m3)
W-46-CaCl20.1273501.26
W-46-LiCl0.51414375.17
ES-47-LiCl0.1373801.37
ES-47-CaCl20.1353771.36
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ocvirk, M.; Ristić, A.; Zabukovec Logar, N. Synthesis of Mesoporous γ-Alumina Support for Water Composite Sorbents for Low Temperature Sorption Heat Storage. Energies 2021, 14, 7809. https://doi.org/10.3390/en14227809

AMA Style

Ocvirk M, Ristić A, Zabukovec Logar N. Synthesis of Mesoporous γ-Alumina Support for Water Composite Sorbents for Low Temperature Sorption Heat Storage. Energies. 2021; 14(22):7809. https://doi.org/10.3390/en14227809

Chicago/Turabian Style

Ocvirk, Manca, Alenka Ristić, and Nataša Zabukovec Logar. 2021. "Synthesis of Mesoporous γ-Alumina Support for Water Composite Sorbents for Low Temperature Sorption Heat Storage" Energies 14, no. 22: 7809. https://doi.org/10.3390/en14227809

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