Next Article in Journal
Comparative Study on Passive Film Formation Mechanism of Cast and PBF-LB/M-TC4 in Simulated Physiological Solution
Next Article in Special Issue
Additively-Manufactured Broadband Metamaterial-Based Luneburg Lens for Flexible Beam Scanning
Previous Article in Journal
An Investigation into PVA Fiber Modified with SiO2 for Improving Mechanical Properties of Oil-Well Cements
Previous Article in Special Issue
Effect of the Axial Profile of a Ceramic Grinding Wheel on Selected Roughness Parameters of Shaped Surfaces Obtained in the Grinding Process with a Dual-Tool Grinding Head
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Vapor Pressure and Enthalpy of Vaporization of Guanidinium Methanesulfonate as a Phase Change Material for Thermal Energy Storage

Laboratory of Electrolytes and Phase Change Materials, College of Materials Science and Engineering, Sichuan University, Chengdu 610065, China
*
Author to whom correspondence should be addressed.
Materials 2024, 17(11), 2582; https://doi.org/10.3390/ma17112582
Submission received: 9 May 2024 / Revised: 24 May 2024 / Accepted: 24 May 2024 / Published: 27 May 2024
(This article belongs to the Special Issue Obtaining and Characterization of New Materials (5th Edition))

Abstract

:
This paper reports the vapor pressure and enthalpy of vaporization for a promising phase change material (PCM) guanidinium methanesulfonate ([Gdm][OMs]), which is a typical guanidinium organomonosulfonate that displays a lamellar crystalline architecture. [Gdm][OMs] was purified by recrystallization. The elemental analysis and infrared spectrum of [Gdm][OMs] confirmed the purity and composition. Differential scanning calorimetry (DSC) also confirmed its high purity and showed a sharp and symmetrical endothermic melting peak with a melting point (Tm) of 207.6 °C and a specific latent heat of fusion of 183.0 J g−1. Thermogravimetric analysis (TGA) reveals its thermal stability over a wide temperature range, and yet three thermal events at higher temperatures of 351 °C, 447 °C, and 649 °C were associated with vaporization or decomposition. The vapor pressure was measured using the isothermogravimetric method from 220 °C to 300 °C. The Antoine equation was used to describe the temperature dependence of its vapor pressure, and the substance-dependent Antoine constants were obtained by non-linear regression. The enthalpy of vaporization (ΔvapH) was derived from the linear regression of the slopes associated with the linear temperature dependence of the rate of weight loss per unit area of vaporization. Hence, the temperature dependence of vapor pressures ln Pvap (Pa) = 10.99 − 344.58/(T (K) − 493.64) over the temperature range from 493.15 K to 573.15 K and the enthalpy of vaporization ΔvapH = 157.10 ± 20.10 kJ mol−1 at the arithmetic mean temperature of 240 °C were obtained from isothermogravimetric measurements using the Antoine equation and the Clausius–Clapeyron equation, respectively. The flammability test indicates that [Gdm][OMs] is non-flammable. Hence, [Gdm][OMs] enjoys very low volatility, high enthalpy of vaporization, and non-flammability in addition to its known advantages. This work thus offers data support, methodologies, and insights for the application of [Gdm][OMs] and other organic salts as PCMs in thermal energy storage and beyond.

1. Introduction

Thermal energy storage technologies employing phase change materials (PCMs) offer a promising solution for the intermittency of solar energy and industrial waste heat recovery and utilization [1,2,3,4,5]. PCMs can be divided simply into two categories: inorganic PCMs and organic PCMs [5]. Table 1 lists some PCMs with melting points in the intermediate temperature range between 100 °C and 230 °C. Among them, inorganic salt hydrates (e.g., MgCl2·6H2O) have poor cycling stability due to water separation and high supercooling degree [6]. Metals such as indium and selenium are unsuitable for use as large-scale PCMs due to their low latent heat of fusion (ΔfusH) [7]. Organic PCMs such as erythritol and D-dulcitol have high values of ΔfusH, but their significant supercooling is not conducive to heat release [8,9,10,11]. As a new kind of PCM, ionic liquids (or organic salts) have unique physicochemical properties, including non-flammability, low volatility, excellent thermal stability, tunable melting points, and high heat storage density, offering potential applications in thermal energy storage [12,13]. Currently, protic ionic liquids derived from pyridines [14], guanidines [15,16], pyrazoles [13], and imidazoles [12,17] have been explored as potential PCMs. Among them, guanidinium methanesulfonate ([Gdm][OMs], i.e., guanidinium mesylate, Figure 1) exhibits the highest ΔfusH to date (Tm = 208 °C, ΔfusH = 190 J g−1, Table 1) with total volumetric energy storage measured as 622 MJ m−3 (173 kWh m−3) and an excellent cyclic stability after 420 cycles between 150 °C and 215 °C, making it the state-of-the-art PCM based on protic organic salts [15,16]. Overall, [Gdm][OMs] has been identified as a very promising PCM for inexpensive renewable energy storage applications at intermediate temperatures [7,15,16,18]. Moreover, [Gdm][OMs] is a representative guanidinium organomonosulfonate (GMS), which displays a series of lamellar crystalline architectures and a two-dimensional hydrogen-bonding network of complementary guanidinium ions (G) and sulfonate moieties (S), the so-called GS sheet [19]. Previous studies on [Gdm][OMs] as a PCM in terms of thermal energy storage include melting temperature, latent heat of fusion, heat capacity, thermal conductivity, volume change, advanced thermal stability, long-term cycling, and economic analysis [15,16]. However, so far, there has been no report on the vapor pressure and vaporization enthalpy of [Gdm][OMs] to the best of our knowledge.
The determination of the temperature-dependent vapor pressure of ionic liquids (ILs) is useful for their industrial application. A higher vapor pressure of the storage medium can lead to thicker walls and the potential loss of PCMs, increasing cost [5], and safety concerns for thermal energy storage containers. Hence, vapor pressure (Pvap) and enthalpy of vaporization (ΔvapH) are not only important basic thermal properties but also key parameters for designing thermal energy storage systems, which are critical for the long-term safety of PCMs and thermal energy storage systems. In addition, ΔvapH reflects molecular interactions in the liquid state and can serve as a foundation for calibrating and validating force fields in molecular dynamics simulations and as an anchoring parameter in PVT equations for neat ILs [20,21]. For example, Liu et al. [21] calculated the values of ΔvapH of tetrabutylammonium bis(trifluoromethanesulfonyl)imide ([N4444][NTf2]) and investigated its cohesive energy, enthalpy, and entropy. The Gibbs free energy of the transition from condensed state to gas was successfully obtained using Born–Fajans–Haber cycles.
In this work, the protic organic salt [Gdm][OMs] was prepared and purified via recrystallization (Figure 1). Its composition and purity were verified by elemental analysis and Fourier transform infrared spectroscopy (FT-IR). The melting temperature (Tm), latent heat of fusion (ΔfusH), and thermal decomposition temperature (Td) were studied by differential scanning calorimetry (DSC) and thermogravimetric analysis (TGA). Pvap and ΔvapH of [Gdm][OMs] were determined using the isothermogravimetric method. The functional relationship between vapor pressure and temperature was fitted using the Antoine equation. [Gdm][OMs] exhibits very low vapor pressure, high enthalpy of vaporization, and non-flammability in addition to its reported merits like high thermal and cyclic stability, offering data support for its application as a PCM in thermal energy storage.

2. Experimental

2.1. Materials

The materials used in this study were as follows: guanidinium carbonate ((CH5N3)2·H2CO3, 99%, Acros Organics, Geel, Belgium), methanesulfonic acid (CH3SO3H, 99%, Sigma-Aldrich, Steinheim, Germany), ethanol (200 proof, anhydrous, ≥99.5%, Sigma-Aldrich, St. Louis, MO, USA), and glycerol (HOCH2CHOHCH2OH, ≥98%, Beijing Solarbio Science & Technology Co., Ltd., Beijing, China). They were all utilized without further purification.

2.2. Synthesis of the Organic Salt

The synthesis of guanidinium methanesulfonate ([Gdm][OMs], i.e., guanidinium mesylate) involves the slow addition of 22.61 mL of methanesulfonic acid to 15.69 g of guanidinium carbonate in a round bottom flask at room temperature. It should be noted that the adopted ratio of these two raw materials (CH3SO3H:(CH5N3)2·H2CO3) slightly exceeds the stoichiometric ratio of 2:1, ensuring excess methanesulfonic acid and thus the complete reaction of the guanidinium carbonate. Then, the resulting wet slurry was heated to 210 °C in an oil bath to form a homogenous melt, followed by stirring at 210 °C via a magnetic stirrer for 5 h to promote the complete formation of the expected salt. The reaction was accompanied by the production of large quantities of gas and water. Most of the water was evaporated off during the reaction while the excess methanesulfonic acid was removed under vacuum at 250 °C.

2.3. Recrystallization

The [Gdm][OMs] sample was purified by recrystallization before further use as follows: Dissolve the as-prepared [Gdm][OMs] in anhydrous ethanol in a sealed glass vial, and sonicate the solution in a water bath at 50 °C for 30 min until it is completely clear. The mass ratio of [Gdm][OMs] vs. anhydrous ethanol is about 1:15.0~15.4. Then, cool the solution and maintain it at 0 °C for 12 h. Afterward, the obtained crystals were filtered off and dried at 80 °C in an oven for 1 h.

2.4. Elemental Analysis

The composition of the [Gdm][OMs] sample was characterized using an elemental analyzer (UNICUBE, Elementar, Langenselbold, Germany) with a detection limit of less than 50 ppm using a thermally conductive detector.

2.5. Fourier Transform Infrared (FT-IR) Analysis

The FT-IR spectrum of the [Gdm][OMs] powders was recorded at room temperature in the range of 4000–400 cm−1 on a spectrometer (INVENIO-R, Bruker, Ettlingen, Germany) with a universal ATR accessory and was accumulated for 16 scans at a resolution of 4 cm−1.

2.6. Differential Scanning Calorimetry (DSC)

DSC measurements were conducted between 25 °C and 240 °C on a thermal analysis system (STARe DSC 3, Mettler-Toledo, Greifensee, Switzerland) at a heating and cooling rate of 5 °C min−1 to identify the phase transition temperatures and latent heat of fusion (ΔfusH). The values are reported using the second run of the DSC tests to eliminate possible thermal history. The melting point (Tm) was fixed as the peak temperature and ΔfusH as the integrated area of the melting peak. Measurements were carried out using sealed aluminum pans (40 μL) under a nitrogen atmosphere (60 mL min−1) with a sample weight of 8.34 mg.

2.7. Thermogravimetric Analysis (TGA)

All thermogravimetric tests were carried out on the same instrument (TGA 2, Mettler Toledo, Columbus, OH, USA), including variable temperature thermogravimetric analysis and isothermogravimetric analysis (IGA). Concerning the ramped TGA measurement, the sample was tested from 30 °C to 850 °C under N2 (50 mL min−1) and at a heating rate of 5 °C min−1 with a sample weight of 11.07 mg using a covered Al2O3 pan. The decomposition temperature (Td) is determined by the onset temperature.
The IGA experiments of [Gdm][OMs] and glycerol were carried out on the same instrument using the same batch of uncovered Al2O3 crucible (cross-sectional area (a): 0.204 cm2) under the same nitrogen atmosphere (50 mL min−1). For [Gdm][OMs], it was conducted at 220 °C, 240 °C, 260 °C, 280 °C, and 300 °C successively. It was held for 30 min at each measurement temperature. The heating rate for each measurement temperature was 20 °C min−1. Similarly, the IGA experiments for glycerol were conducted at 100 °C, 120 °C, 140 °C, and 160 °C successively, and at each temperature it was held for 30 min. The masses of [Gdm][OMs] and glycerol were 10.40 mg and 42.36 mg, respectively.

2.8. Flammability Test

The flammability of [Gdm][OMs] was tested by direct ignition in a battery case for 5 s.

3. Theoretical Basis

3.1. Vapor Pressure

The vapor pressures of [Gdm][OMs] can be derived based on the Langmuir equation using the thermogravimetric data [22,23]:
d m / d t / S = P vap α vap M w / 2 π R T 0.5
where (−dm/dt)/S is the rate of mass loss of the sample per unit area of vaporization (g min−1 cm−2), Pvap is the vapor pressure (Pa), αvap is the vaporization coefficient, Mw is the molecular weight of the effusing vapor (g mol−1), R is the gas constant (8.314 J mol−1 K−1), and T is the absolute temperature (K). αvap is considered to be dependent on the apparatus and experimental atmosphere rather than on the substance and temperature. Furthermore, it can be determined by using pure substances with reported precise vapor pressures [24,25]. Equation (1) is simplified as Equation (2):
P vap = k vap v
where kvap is a constant (Pa min cm2 g−0.5 mol−0.5 K−0.5), v is dependent on the material (g0.5 mol0.5 K0.5 min−1 cm−2), kvap = (2πR)0.5/αvap, and v = [(−dm/dt)/S](T/Mw)0.5.
Based on experimental data at several temperatures, vapor pressures in a given temperature range can be predicted by Antoine equation (Equation (3)) [26], which is an empirical equation that describes the PT relationship of pure liquid substances and offers various applications in engineering:
ln P vap = A B T + C
where A, B, and C are the substance-dependent Antoine constants (also known as Antoine coefficients) and Pvap is the vapor pressure (Pa) at temperature T (K).

3.2. Enthalpy of Vaporization

Enthalpy of vaporization of substances can be calculated by Equation (4), which is a regression equation derived from the Clausius–Clapeyron equation and was successfully used by Luo et al. [27] to calculate the enthalpy of vaporization of imidazolium-based ionic liquids:
ln ( d m / d t ) T 0.5 / S = A v a p H R T
where (−dm/dt)/S is the rate of mass loss of the sample per unit area of vaporization (g min−1 cm−2), T is the absolute temperature (K), R is the gas constant (8.314 J mol−1 K−1), ΔvapH is the molar enthalpy of vaporization at the arithmetic mean temperature of all tested temperature points, and A is an empirical parameter. This method has recently been widely used to calculate ΔvapH of various ionic liquids [28,29,30].

4. Results and Discussions

4.1. Basic Characterizations

The as-synthesized [Gdm][OMs] was examined by elemental analysis. The experimental values of the elemental contents of C, H, N, S, and O were as follows (the calculated values based on the molecular formula are given in parentheses): C-16.06% (15.48%), H-5.98% (5.81%), N-27.73% (27.10%), S-20.07% (20.65%), and O-30.16% (30.96%). It can be seen that the experimental values are in line with the calculated values for all the five elements. The chemical composition of the tested sample can be inferred to be C2H9N3O3S, indicating the high purity of the as-prepared sample.
The structure of the [Gdm][OMs] is also confirmed by FT-IR spectrum (Figure 2a). The appearance of new peaks that are different from the characteristic peaks of both guanidine carbonate and methanesulfonic acid, as well as the disappearance of some peaks, proved the formation of the proton transfer organic salt [31,32]. The as-prepared [Gdm][OMs] sample has a very low water content, indicated by the absence of the O−H stretching bands at 3400−3800 cm−1 and the H−O−H bending band at around 1650 cm−1 in the infrared spectrum [33,34,35]. Based on our previous work [32], the formation of CH3SO3 anions in protic organic salts is confirmed by the bands at 1416, 1337, 1185, 1153, 1044, 773, 739, and 723 cm−1, while the existence of guanidinium cations in [Gdm][OMs] is evidenced by the peaks at 3328, 3261, 3185, 1678, and 1582 cm−1 (Figure 2a) [31,36]. The sharp absorption bands at around 1185 and 1153 cm−1 are assigned to the SO3− asymmetric stretching vibration, while the peak at 1044 cm−1 relates to the SO3− symmetric stretching vibration [37]. This indicates a transformation of the −SO2OH group into the −SO3− anionic group, implying the occurrence of a proton transfer process from the acidic site to the basic site. This spectral feature is typical of CH3SO3 anions strongly involved in hydrogen bonding [38]. The peaks at 3328, 3261, and 3185 cm−1 correspond to the asymmetric and symmetric NH2 stretching vibrations, respectively [36,39]. In addition, the peaks at 1678 cm−1 and 1582 cm−1 are due to the asymmetric degenerated stretching of CN3 and the degenerated scissoring vibration of NH2, respectively [36,39]. Particularly, the peaks at 1351 cm−1 and 908 cm−1 associated with free CH3SO3H are not observed in [Gdm][OMs], implying no detectable excess acid in the product [40,41]. Moreover, the broadening of the vibrational bands between 3000 cm−1 and 3500 cm−1 implies a hydrogen bonding network [42,43]. The absorption bands at around 2825, 2755, and 2236 cm−1 may be attributed to the N−H or N+−H stretching vibration, thus also proving the salt formation [44].
The DSC thermogram of [Gdm][OMs] is shown in Figure 2b. The experimental conditions and results of DSC measurements in comparison with the literature [15,16] for [Gdm][OMs] are summarized in Table 2. [Gdm][OMs] exhibits distinct melting and solidification peaks during the heating and cooling cycles. It shows a very sharp and symmetrical endothermic melting peak with an accurate melting point (Tm) of 207.6 °C and a specific latent heat of fusion of 183.0 J g−1 (Figure 2b), indicating the high purity of as-prepared [Gdm][OMs]. The values of Tm and ΔfusH agree well with the values reported by MacFarlane et al. (Tm = 208 ± 1 °C; ΔfusH = 190 ± 9.5 J g−1) [15,16].

4.2. Vapor Pressure and Enthalpy of Vaporization

Glycerol is used as the calibrator for the isothermogravimetric method to determine the constant kvap (Equation (2)). Based on the IGA traces of glycerol at different temperatures (Figure 3a), the distinct mass loss rates of glycerol, the corresponding coefficient of determination R2 (all of them exceeds 0.98) for linear regression, and the values of v were obtained and are listed in Table 3. Using the reported vapor pressure data of glycerol [45,46], the curve of Pvap vs. v is plotted in Figure 3b. Obviously, there is a linear relationship between Pvap and v, and the constant kvap was evaluated from the slope as 5397 Pa min cm2 g−0.5 mol−0.5 K−0.5, with an excellent coefficient of determination (R2) of 0.999.
The TGA trace of [Gdm][OMs] at 5 K min−1 in nitrogen atmosphere (Figure 4a) indicates that the onset temperature of weight loss of [Gdm][OMs] is as high as 332 °C (605.15 K), revealing its thermal stability over a wide temperature range. It also confirms the ultra-low water content as no loss of water was observed. Furthermore, derivative thermogravimetry (DTG), in which the rate of weight changes of [Gdm][OMs] upon heating is plotted against temperature, demonstrates three thermal events associated with vaporization or decomposition at higher temperatures of 351 °C, 447 °C, and 649 °C successively. Figure 4b shows the IGA traces of [Gdm][OMs] at 220 °C, 240 °C, 260 °C, 280 °C, and 300 °C, respectively. Clearly, the weight loss rate of [Gdm][OMs] increases as the temperature increases, demonstrating that the higher the temperature is, the easier it is for [Gdm][OMs] to evaporate. Table 4 shows the linear fitting results of the IGA traces with R2 in the range of 0.939–0.999.
The measured vaporization rates and the calculated values of v (v = [(−dm/dt)/S](T/Mw)0.5) are listed in Table 4. The vapor pressures of [Gdm][Oms] at 220 °C, 240 °C, 260 °C, 280 °C, and 300 °C are then calculated via Equation (2) (Pvap = kvapv) thanks to the calibrated value of the constant kvap (5397 Pa min cm2 g−0.5 mol−0.5 K−0.5) using glycerol. Note that kvap is dependent on the apparatus and experimental atmosphere. The derived vapor pressure of [Gdm][OMs] at 533.15 K (260 °C) is 15.23 Pa (Table 4), which is much lower than that of 1-butyl-3-methylimidazolium bis(trifluoromethanesulfonyl)imide ([C4mim][NTf2]) (Pvap = 4660 Pa at 517.45 K), indicating that this protic organic salt has very low volatility even at high temperatures [47]. The ultra-low vapor pressure means less loss and better durability of [Gdm][OMs] as a PCM and also implies a simpler design, a lower cost, and higher safety for the containers in terms of pressure vessels.
The three Antoine constants are obtained by fitting Pvap vs. T using the Antoine equation (Figure 4d, R2 = 0.999), where A, B, and C are 10.99, 344.58, and −493.64, respectively. The temperature dependence of vapor pressures ln Pvap (Pa) = 10.99 − 344.58/(T (K) − 493.64) in the temperature range 493.15–573.15 K is thus acquired. The approximated semi-empirical equation indicates the exponential relationship between the vapor pressure and the temperature of [Gdm][OMs]. Moreover, the vapor pressures of [Gdm][OMs] between 220 °C and 300 °C can thus be interpolated.
Figure 5 shows a linear fit based on the linear weight loss rates at 220 °C, 240 °C, and 260 °C calculated from Figure 4c (also shown in Table 4) according to Equation (4). The calculated ΔvapH at the average temperature of 513.15 K (240 °C) for [Gdm][OMs] is 157.10 ± 20.10 kJ mol−1. The estimated value of ΔvapH is higher than those of [C4mim][NTf2] (ΔvapH = 118.5 ± 0.4 kJ mol−1 at 496 K) and guanidinium nonaflate ([Gdm][NfO], ΔvapH = 120.9 kJ mol−1 at 523 K), which again proves that [Gdm][OMs] has essentially low volatility [27,31].
Finally, the flammability test (Figure 6) indicates that [Gdm][OMs] cannot be ignited after 5 s of combustion in air and will not explode when exposed to open flames, either. This demonstrates that [Gdm][OMs] is non-flammable and has excellent thermal safety.

5. Conclusions

Pure guanidinium methanesulfonate ([Gdm][OMs]) was prepared by a slight deviation from the stoichiometric ratio of methanesulfonic acid to guanidinium carbonate (i.e., the acid is slightly excess), followed by heating, stirring, vacuum drying, and then recrystallization. The elemental analysis, infrared spectrum measurement, thermogravimetric analysis (TGA), and differential scanning calorimetry (DSC) all confirmed its high purity. TGA confirms its thermal stability over a wide temperature range while it also indicates three thermal events due to vaporization or decomposition between 350 °C and 650 °C. Using isothermogravimetric analysis, the Antoine equation was successfully employed to describe its vapor pressure as a function of temperature between 493.15 K and 573.15 K: ln Pvap (Pa) = 10.99 – 344.58/(T (K) – 493.64), indicating the vapor pressures ranging from 0.87 Pa at 220 °C to 899.98 Pa at 300 °C. Furthermore, its enthalpy of vaporization is derived as ΔvapH = 157.10 ± 20.10 kJ mol−1 at the average temperature of 240 °C via the Clausius–Clapeyron equation. In summary, [Gdm][OMs] exhibits a series of additional advantages, including essentially low volatility, high enthalpy of vaporization, and non-flammability. This work is expected to offer data support, methodologies, and insights for the application of organic salts as a new type of phase change materials in thermal energy storage and other industrial fields.

Author Contributions

J.L.: Conceptualization, supervision, resources, methodology, formal analysis, validation, writing—review & editing, funding acquisition, and project administration. W.B.: Investigation, formal analysis, visualization, data curation, and writing—original draft. S.L.: Formal analysis and writing—review & editing. X.R.: Formal analysis and writing—review & editing. G.M.: Formal analysis and writing—review & editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (project No.: 22378270 and 21776120), Sichuan Science and Technology Program (project No.: 2022ZYD0016 and 2023JDRC0013), Hohhot Science and Technology Program (project No.: 2023-JieBangGuaShuai-Gao-3), Natural Science Foundation of Fujian Province, China (project No.: 2023J01254), the starting grant (“One Hundred Talent Program”) from Sichuan University (project No.: YJ202089), the Fundamental Research Funds for Central Universities (project No.: 20826041G4185) and the Research Fund Program of Guangdong Provincial Key Laboratory of Fuel Cell Technology (project No.: FC202206 and FC202218).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tang, Z.; Gao, Y.; Cheng, P.; Jiang, Y.; Xu, J.; Chen, X.; Li, A.; Wang, G. Metal-organic framework derived magnetic phase change nanocage for fast-charging solar-thermal energy conversion. Nano Energy 2022, 99, 107383. [Google Scholar] [CrossRef]
  2. İnada, A.A.; Arman, S.; Safaei, B. A novel review on the efficiency of nanomaterials for solar energy storage systems. J. Energy Storage 2022, 55, 105661. [Google Scholar] [CrossRef]
  3. Zhang, Y.; Ren, J.; Pu, Y.; Wang, P. Solar energy potential assessment_ A framework to integrate geographic, technological, and economic indices for a potential analysis. Renew. Energy 2020, 149, 577–586. [Google Scholar] [CrossRef]
  4. Liu, D.; Lei, C.; Wu, K.; Fu, Q. A Multidirectionally Thermoconductive Phase Change Material Enables High and Durable Electricity via Real-Environment Solar–Thermal–Electric Conversion. ACS Nano 2020, 14, 15738–15747. [Google Scholar] [CrossRef] [PubMed]
  5. Alva, G.; Lin, Y.; Fang, G. An overview of thermal energy storage systems. Energy 2018, 144, 341–378. [Google Scholar] [CrossRef]
  6. Song, Z.; Deng, Y.; Li, J.; Nian, H. Expanded graphite for thermal conductivity and reliability enhancement and supercooling decrease of MgCl2⋅6H2O phase change material. Mater. Res. Bull. 2018, 102, 203–208. [Google Scholar] [CrossRef]
  7. Matuszek, K.; Kar, M.; Pringle, J.M.; MacFarlane, D.R. Phase change materials for renewable energy storage at intermediate temperatures. Chem. Rev. 2022, 123, 491–514. [Google Scholar] [CrossRef] [PubMed]
  8. Yuan, M.; Xu, C.; Wang, T.; Zhang, T.; Pan, X.; Ye, F. Supercooling suppression and crystallization behaviour of erythritol/expanded graphite as form-stable phase change material. Chem. Eng. J. 2021, 413, 127394. [Google Scholar] [CrossRef]
  9. del Barrio, E.P.; Godin, A.; Duquesne, M.; Daranlot, J.; Jolly, J.; Alshaer, W.; Kouadio, T.; Sommier, A. Characterization of different sugar alcohols as phase change materials for thermal energy storage applications. Sol. Energy Mater. Sol. Cells 2017, 159, 560–569. [Google Scholar] [CrossRef]
  10. Solé, A.; Neumann, H.; Niedermaier, S.; Martorell, I.; Schossig, P.; Cabeza, L.F. Stability of sugar alcohols as PCM for thermal energy storage. Sol. Energy Mater. Sol. Cells 2014, 126, 125–134. [Google Scholar] [CrossRef]
  11. Gasia, J.; Martin, M.; Solé, A.; Barreneche, C.; Cabeza, L. Phase Change Material Selection for Thermal Processes Working under Partial Load Operating Conditions in the Temperature Range between 120 and 200 °C. Appl. Sci. 2017, 7, 722. [Google Scholar] [CrossRef]
  12. Zhang, H.; Xu, W.; Liu, J.; Li, M.; Yang, B. Thermophysical properties of dicationic imidazolium-based ionic. J. Mol. Liq. 2019, 282, 474–483. [Google Scholar] [CrossRef]
  13. Matuszek, K.; Vijayaraghavan, R.; Forsyth, C.M.; Mahadevan, S.; Kar, M.; MacFarlane, D.R. Pyrazolium Phase-Change Materials for Solar-Thermal Energy Storage. Chemsuschem 2019, 13, 159–164. [Google Scholar] [CrossRef] [PubMed]
  14. Matuszek, K.; Hatton, C.; Kar, M.; Pringle, J.M.; MacFarlane, D.R. Molecular patterns in the thermophysical properties of pyridinium ionic liquids as phase change materials for energy storage in the intermediate temperature range. J. Non-Cryst. Solids X 2022, 15, 100108. [Google Scholar] [CrossRef]
  15. Matuszek, K.; Vijayaraghavan, R.; Kar, M.; Mahadevan, S.; MacFarlane, D.R. Guanidinium Organic Salts as Phase-Change Materials for Renewable Energy Storage. Chemsuschem 2021, 14, 2757–2762. [Google Scholar] [CrossRef] [PubMed]
  16. Vijayraghavan, R.; Rana, U.A.; Elliott, G.D.; MacFarlane, D.R. Protic Ionic Solids and Liquids Based on the Guanidinium Cation as Phase-Change Energy-Storage Materials. Energy Technol. 2013, 1, 609–612. [Google Scholar] [CrossRef]
  17. Zhang, H.; Li, M.; Yang, B. Design, synthesis, and analysis of thermophysical properties for imidazolium-based geminal dicationic ionic liquids. J. Phys. Chem. C 2018, 122, 2467–2474. [Google Scholar] [CrossRef]
  18. Piper, S.L.; Kar, M.; MacFarlane, D.R.; Matuszek, K.; Pringle, J.M. Ionic liquids for renewable thermal energy storage—A perspective. Green. Chem. 2022, 21, 102. [Google Scholar] [CrossRef]
  19. Horner, M.J.; Holman, K.T.; Ward, M.D. Architectural diversity and elastic networks in hydrogen-bonded host frameworks:  from molecular jaws to cylinders. J. Am. Chem. Soc. 2007, 129, 14640–14660. [Google Scholar] [CrossRef]
  20. Dunaev, A.M.; Motalov, V.B.; Kudin, L.S.; Butman, M.F. Thermodynamic properties of the ionic vapor species over EMImNTf 2 ionic liquid studied by Knudsen effusion mass spectrometry. J. Mol. Liq. 2016, 223, 407–411. [Google Scholar] [CrossRef]
  21. Liu, S.; Wei, R.; Ma, G.; Li, A.; Conrad, O.; Luo, J. The cohesive properties and pyrolysis mechanism of an aprotic ionic liquid tetrabutylammonium bis(trifluoromethanesulfonyl)imide. Soft Matter 2023, 19, 6458–6467. [Google Scholar] [CrossRef]
  22. Langmuir, I. The vapor pressure of metallic tungsten. Phys. Z. 1913, 11, 1273–1280. [Google Scholar] [CrossRef]
  23. Price, D.M.; Hawkins, M. Calorimetry of two disperse dyes using thermogravimetry. Thermochim. Acta 1998, 315, 19–24. [Google Scholar] [CrossRef]
  24. Shahbaz, K.; Mjalli, F.S.; Vakili-Nezhaad, G.; AlNashef, I.M.; Asadov, A.; Farid, M.M. Thermogravimetric measurement of deep eutectic solvents vapor pressure. J. Mol. Liq. 2016, 222, 61–66. [Google Scholar] [CrossRef]
  25. Verevkin, S.P.; Ralys, R.V.; Zaitsau, D.H.; Emel’yanenko, V.N.; Schick, C. Express thermo-gravimetric method for the vaporization enthalpies appraisal for very low volatile molecular and ionic compounds. Thermochim. Acta 2012, 538, 55–62. [Google Scholar] [CrossRef]
  26. Wisniak, J. Historical development of the vapor pressure equation from dalton to antoine. J. Phase Equilibria 2001, 22, 622–630. [Google Scholar] [CrossRef]
  27. Luo, H.; Baker, G.A.; Dai, S. Isothermogravimetric determination of the enthalpies of vaporization of 1-Alkyl-3-methylimidazolium ionic liquids. J. Phys. Chem. B 2008, 112, 10077–10081. [Google Scholar] [CrossRef] [PubMed]
  28. Liu, L.; Jing, L.-Q.; Liu, H.-C.; Fang, D.-W.; Tong, J. Measurement of vaporization enthalpy for amino acid ionic liquids [Cnmim][Thr](n = 3, 5) using the isothermogravimetrical analysis. J. Therm. Anal. Calorim. 2018, 134, 2247–2254. [Google Scholar] [CrossRef]
  29. Kelkar, M.S.; Maginn, E.J. Calculating the enthalpy of vaporization for ionic liquid clusters. J. Phys. Chem. B 2007, 111, 9424–9427. [Google Scholar] [CrossRef]
  30. Kabo, G.J.; Paulechka, Y.U.; Zaitsau, D.H.; Firaha, A.S. Prediction of the enthalpies of vaporization for room-temperature ionic liquids: Correlations and a substitution-based additive scheme. Thermochim. Acta 2015, 609, 7–19. [Google Scholar] [CrossRef]
  31. Chen, X.; Tang, H.; Putzeys, T.; Sniekers, J.; Wübbenhorst, M.; Binnemans, K.; Fransaer, J.; De Vos, D.E.; Li, Q.; Luo, J. Guanidinium nonaflate as a solid-state proton conductor. J. Mater. Chem. A 2016, 4, 12241–12252. [Google Scholar] [CrossRef]
  32. Luo, J.; Hu, J.; Saak, W.; Beckhaus, R.; Wittstock, G.; Vankelecom, I.F.J.; Agert, C.; Conrad, O. Protic ionic liquid and ionic melts prepared from methanesulfonic acid and 1H-1,2,4-triazole as high temperature PEMFC electrolytes. J. Mater. Chem. 2011, 21, 10426–10436. [Google Scholar] [CrossRef]
  33. Enev, V.; Sedláček, P.; Řihák, M.; Kalina, M.; Pekař, M. IR-supported thermogravimetric analysis of water in hydrogels. Front. Mater. 2022, 9, 931303. [Google Scholar] [CrossRef]
  34. Brubach, J.B.; Mermet, A.; Filabozzi, A.; Gerschel, A.; Roy, P. Signatures of the hydrogen bonding in the infrared bands of water. J. Chem. Phys. 2005, 122, 184509. [Google Scholar] [CrossRef] [PubMed]
  35. Luo, J.; Jensen, A.H.; Brooks, N.R.; Sniekers, J.; Knipper, M.; Aili, D.; Li, Q.; Vanroy, B.; Wübbenhorst, M.; Yan, F.; et al. 1,2,4-Triazolium perfluorobutanesulfonate as an archetypal pure protic organic ionic plastic crystal electrolyte for all-solid-state fuel cells. Energy Environ. Sci. 2015, 8, 1276–1291. [Google Scholar] [CrossRef]
  36. Roseanne, J.S.; Hudson, B.; Callis, P.R. Resonance Raman studies of guanidinium and substituted guanidinium ions. J. Phys. Chem. 1990, 94, 4015–4025. [Google Scholar]
  37. Ghazaryan, V.V.; Drebushchak, T.N.; Boldyreva, E.V.; Petrosyan, A.M. Bis-l-cysteinium sulfate and l-cysteinium methanesulfonate. Struct. Chem. 2020, 31, 1919–1925. [Google Scholar] [CrossRef]
  38. Langner, R.; Zundel, G. FTIR investigation of O···H···O hydrogen bonds with large proton polarizability in sulfonic acid–N-oxide systems in the middle and far-IR. J. Chem. Soc. Faraday Trans. 1998, 94, 1805–1811. [Google Scholar] [CrossRef]
  39. Drozd, M. Molecular structure and infrared spectra of guanidinium cation. Mater. Sci. Eng. B 2007, 136, 20–28. [Google Scholar] [CrossRef]
  40. Chackalackal, S.M.; Stafford, F.E. Infrared spectra of methane-, fluoro-, and vhlorosulfonic acids. J. Am. Chem. Soc. 1966, 88, 4815–4819. [Google Scholar] [CrossRef]
  41. Clarke, J.H.R.; Woodward, L.A. Raman spectrophotometric determination of the degrees of dissociation of methanesulphonic acid in aqueous solution at 25°C. Trans. Faraday Soc. 1966, 62, 2226–2233. [Google Scholar] [CrossRef]
  42. Luo, J.; Conrad, O.; Vankelecom, I.F.J. Physicochemical properties of phosphonium-based and ammonium-based. J. Mater. Chem. 2012, 22, 20574–20579. [Google Scholar] [CrossRef]
  43. Luo, J.; Tan, T.V.; Conrad, O.; Vankelecom, I.F.J. 1H-1,2,4-Triazole as solvent for imidazolium methanesulfonate. Phys. Chem. Chem. Phys. 2012, 14, 11441–11447. [Google Scholar] [CrossRef] [PubMed]
  44. Matsuoka, H.; Nakamoto, H.; Susan, M.A.B.H.; Watanabe, M. Brønsted acid–base and –polybase complexes as electrolytes for fuel cells under non-humidifying conditions. Electrochim. Acta 2005, 50, 4015–4021. [Google Scholar] [CrossRef]
  45. Takamura, K.; Fischer, H.; Morrow, N.R. Physical properties of aqueous glycerol solutions. J. Pet. Sci. Eng. 2012, 98–99, 50–60. [Google Scholar] [CrossRef]
  46. Ravula, S.; Larm, N.E.; Mottaleb, M.A.; Heitz, M.P.; Baker, G.A. Vapor pressure mapping of ionic liquids and low-volatility fluids using graded isothermal thermogravimetric analysis. ChemEngineering 2019, 3, 42. [Google Scholar] [CrossRef]
  47. Paulechka, Y.U.; Zaitsau, D.H.; Kabo, G.J.; Strechan, A.A. Vapor pressure and thermal stability of ionic liquid 1-butyl-3-methylimidazolium Bis(trifluoromethylsulfonyl)amide. Thermochim. Acta 2005, 439, 158–160. [Google Scholar] [CrossRef]
Figure 1. The chemical structure of [Gdm][OMs] (a) and the photograph of its crystals (b).
Figure 1. The chemical structure of [Gdm][OMs] (a) and the photograph of its crystals (b).
Materials 17 02582 g001
Figure 2. (a) FT-IR spectrum of [Gdm][OMs] in the region of 600–4000 cm−1. (b) DSC trace from the second heating/cooling cycle for [Gdm][OMs] (heating and cooling rate: 5 °C min−1).
Figure 2. (a) FT-IR spectrum of [Gdm][OMs] in the region of 600–4000 cm−1. (b) DSC trace from the second heating/cooling cycle for [Gdm][OMs] (heating and cooling rate: 5 °C min−1).
Materials 17 02582 g002
Figure 3. (a) IGA traces of glycerol at different temperatures; (b) the plot of Pvap vs. v for glycerol.
Figure 3. (a) IGA traces of glycerol at different temperatures; (b) the plot of Pvap vs. v for glycerol.
Materials 17 02582 g003
Figure 4. (a) TGA trace and DTG curve of [Gdm][OMs]; (b) IGA traces of [Gdm][OMs] at different temperatures; (c) Selected IGA traces and the linear fitting curves at 220 °C, 240 °C and 260 °C for [Gdm][OMs]; (d) The fitting curve using Antoine equation for [Gdm][OMs].
Figure 4. (a) TGA trace and DTG curve of [Gdm][OMs]; (b) IGA traces of [Gdm][OMs] at different temperatures; (c) Selected IGA traces and the linear fitting curves at 220 °C, 240 °C and 260 °C for [Gdm][OMs]; (d) The fitting curve using Antoine equation for [Gdm][OMs].
Materials 17 02582 g004
Figure 5. The linear fit based on the linear weight loss rates at 220 °C, 240 °C, and 260 °C, calculated from Figure 4c according to Equation (4). The fitting yields a slope of −ΔvapH/(1000R), where ΔvapH is the enthalpy of vaporization and R is the gas constant.
Figure 5. The linear fit based on the linear weight loss rates at 220 °C, 240 °C, and 260 °C, calculated from Figure 4c according to Equation (4). The fitting yields a slope of −ΔvapH/(1000R), where ΔvapH is the enthalpy of vaporization and R is the gas constant.
Materials 17 02582 g005
Figure 6. Photographs of the ignition tests of [Gdm][OMs].
Figure 6. Photographs of the ignition tests of [Gdm][OMs].
Materials 17 02582 g006
Table 1. Some PCMs with melting points in the intermediate temperature range (100–230 °C).
Table 1. Some PCMs with melting points in the intermediate temperature range (100–230 °C).
PCMsTm (°C)ΔfusH (J g−1)ΔT (°C) [b]
MgCl2·6H2O [6]12311741
Indium [7]15729/
Selenium [7]22086/
Erythritol [8,9]11834086
D-dulcitol [11]186334/
C3(mim)2(Br)2 [12]17311646
[Pzy][OMs] [13]168160/
[Pzy][OTf] [13]14717 ss, 24 ss, 27/
[Pzy][C6H5SO3] [13]137105/
[PyH][OMs] [14]18021 ss, 7716
[PyH][OTf] [14]22624 ss, 3841
[pyH][C6H5SO3] [14]13512380
[Gdm][OMs] [15,16]208190/
[Gdm][OTf] [15,16]1606 ss [a], 130/
[Gdm][C6H5SO3] [15,16]210138/
C2(mim)2(Br)2 [17]18811624
C2(mim)2(PF6)2 [17]191109/
[a] ss: solid–solid phase transition. [b] ΔT: degree of supercooling, i.e., difference between Tm and crystallization temperature.
Table 2. Experimental conditions and results of DSC measurements for [Gdm][OMs].
Table 2. Experimental conditions and results of DSC measurements for [Gdm][OMs].
Experimental Conditions and ResultsThis WorkLiterature [15,16]
Sample mass8.34 mg3~8 mg
Scanning rate5 °C min−110 °C min−1
Atmosphere and flow rateN2, 60 mL min−1N2 (flow rate not mentioned)
Measurement temperature range25~240 °C−40~220 °C
DSC instrumentSTARe DSC 3, Mettler-Toledo (with 120 thermocouples)DSC TA Q200 calorimeter (TA Instruments)
Sample conditionpowdersNOT MENTIONED
Crucible40 μL aluminum pan (sealed)NOT MENTIONED
Data selectionthe second run of the DSC cyclethe second run of the DSC cycle
Tm (peak maximum)207.6 °C208 ± 1 °C
ΔfusH183.0 J g−1190 ± 9.5 J g−1
Table 3. Vaporization rates, the values of v, and the reported vapor pressures of glycerol.
Table 3. Vaporization rates, the values of v, and the reported vapor pressures of glycerol.
T/°CT/K(−dm/dt)/S/g min−1 cm−2R2v/g0.5 mol0.5 K0.5 min−1 cm−2Pvap/Pa [45,46]
100373.153.32 × 10−30.9806.69 × 10−330.0
120393.157.42 × 10−30.9991.53 × 10−290.0
140413.152.58 × 10−20.9995.47 × 10−2320
160433.158.36 × 10−20.9991.81 × 10−1980
Table 4. Vaporization rates, the values of v, and derived vapor pressures of [Gdm][OMs].
Table 4. Vaporization rates, the values of v, and derived vapor pressures of [Gdm][OMs].
T/°CT/K(−dm/dt)/S/g min−1 cm−2R2v/
g0.5 mol0.5 K0.5 min−1 cm−2
Pvap/Pa
220493.159.01 × 10−50.9391.61 × 10−40.87
240513.155.39 × 10−40.9989.79 × 10−45.29
260533.151.52 × 10−30.9992.82 × 10−315.23
280553.151.10 × 10−30.9802.08 × 10−2112.26
300573.158.68 × 10−20.9961.67 × 10−1899.98
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

Bi, W.; Liu, S.; Rong, X.; Ma, G.; Luo, J. Vapor Pressure and Enthalpy of Vaporization of Guanidinium Methanesulfonate as a Phase Change Material for Thermal Energy Storage. Materials 2024, 17, 2582. https://doi.org/10.3390/ma17112582

AMA Style

Bi W, Liu S, Rong X, Ma G, Luo J. Vapor Pressure and Enthalpy of Vaporization of Guanidinium Methanesulfonate as a Phase Change Material for Thermal Energy Storage. Materials. 2024; 17(11):2582. https://doi.org/10.3390/ma17112582

Chicago/Turabian Style

Bi, Wenrong, Shijie Liu, Xing Rong, Guangjun Ma, and Jiangshui Luo. 2024. "Vapor Pressure and Enthalpy of Vaporization of Guanidinium Methanesulfonate as a Phase Change Material for Thermal Energy Storage" Materials 17, no. 11: 2582. https://doi.org/10.3390/ma17112582

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