Next Article in Journal
Model Predictive Control Strategies to Activate the Energy Flexibility for Zones with Hydronic Radiant Systems
Previous Article in Journal
An Innovative Metaheuristic Strategy for Solar Energy Management through a Neural Networks Framework
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molten Salts for Sensible Thermal Energy Storage: A Review and an Energy Performance Analysis

1
Department of Chemical and Environmental Engineering, ETSII, Universidad Politécnica de Madrid (UPM), 28006 Madrid, Spain
2
Department of Ceramics, Institute of Ceramics and Glass, Spanish National Research Council (CSIC), 28049 Madrid, Spain
*
Author to whom correspondence should be addressed.
Submission received: 3 February 2021 / Revised: 13 February 2021 / Accepted: 16 February 2021 / Published: 23 February 2021
(This article belongs to the Section D: Energy Storage and Application)

Abstract

:
A comprehensive review of different thermal energy storage materials for concentrated solar power has been conducted. Fifteen candidates were selected due to their nature, thermophysical properties, and economic impact. Three key energy performance indicators were defined in order to evaluate the performance of the different molten salts, using Solar Salt as a reference for low and high temperatures. The analysis provided evidence that nitrate-based materials are the best choice for the former and chloride-based materials are best for the latter instead of fluoride and carbonate-based candidates, mainly due to their low cost.

1. Introduction

Electricity generation is one of the main contributors to greenhouse gas (GHG) emissions due to CO2 being released from fossil fuels; additionally, electricity is also one of the energy vectors upon which many applications will be transformed in the near future [1,2]. As decarbonization is the objective of future energy systems, its stability and dispatchability at a reasonable cost must be ensured. Among all those available, solar energy is one of the most suitable alternatives: it is clean, abundant, and easily obtained anywhere on earth. Among the different alternatives, concentrated solar power (CSP) in combination with thermal energy storage (TES) allows for dispatched electricity to match peak demand and solves the supply−demand coupling problem, allowing energy release and its transformation to electricity whenever necessary and eluding the inherent instability of solar resource availability [3].
Although the International Energy Agency (IEA) estimates that CSP will provide 11% of all global electricity generated by 2050 [4], current plants in operation or in development mainly use sensible TES systems with nitrate-based materials. Other alternatives must be explored, since they have the potential to overcome several drawbacks of commercial TES materials in terms of cost reduction, thermal properties enhancement, and higher/broader operational. TES, together with CSP, still have a long road ahead of them to be considered consistent, robust, continuous, and competitive alternatives. Thus, the incorporation of both into future energy management and generation mix depends greatly upon the future developments of TES materials.
The authors of this work spot the need to present a comprehensive review of the most promising next generation TES materials in order to analyze their strengths and weaknesses, sum up the most relevant thermodynamical properties found in the narrative, and define and evaluate three different key performance indicators (KPIs) to help make the most suitable choice for a specific CSP application.

2. Thermal Energy Storage Materials

High-temperature TES is one of the cheapest forms of energy storage [5]. Although there are different alternatives, such as latent, thermochemical, or solid sensible heat storage [6,7,8], the most common TES materials are molten salts, which are classified as sensible heat storage [9]. Sensible storage implies that increasing the temperature of a substance increases its energy content; when the material is cooled, the stored energy is released, but without a phase change. The following characteristics are desirable in a TES material [10,11]:
  • The energy storage density must be high for a compact design.
  • For indirect TES systems, those where the energy storage medium is different from the heat transfer fluid (HTF) circulating through the solar field, a good heat transfer between HTF and TES material is needed to optimize the system’s efficiency.
  • The TES material should be thermally stable and possess a low vapor pressure in the operating temperature range to avoid undesirable side effects such as material ageing, performance decline of the system, or GHG emissions.
  • A high thermal, chemical, and cyclic stability for extended plant life should be expected.
  • A non-flammable and non-toxic nature are desirable.
  • Inexpensive and abundant.
The following thermal and transport properties are essential for a suitable selection of the proper material that complies with most of the previously mentioned characteristics:
  • Specific heat capacity: this property controls the capacity of the temperature rise that can be transferred or stored. It improves the TES system efficiency [12].
  • Melting temperature: this is directly related to operating and maintenance (O&M) costs since higher melting points need antifreeze protection.
  • Decomposition temperature: this is the theoretical maximum operating temperature that must not be exceeded in order to keep the TES material operative and in good condition; the higher the maximum operating temperature, the higher the TES system efficiency achieved.
  • Thermal conductivity: this property is related to the heat transfer behavior. Higher values are preferred in order to achieve higher heat exchange efficiency.
  • Viscosity: this is directly related to the cost of pumping energy through the system, and lower values are preferable. It is also linked to the previous property, and a compromise between them should be taken.
  • Density: this directly affects the specific energy that a selected TES material can store per unit of volume. The amount of heat carried by the HTF at working temperature relates to the density of a material and its specific heat capacity. Higher values are recommended.
Sensible TES has other relevant properties: the specific heat, the temperature variation of the material (see Equation (1)), as well as conductivity, diffusivity, thermal stability, vapor pressure, and the cost and compatibility of materials. Its main drawbacks are the temperature stability and its energy storage density, which is lower than latent and thermochemical TES [13].
Q = m c p T = ρ V c p T
For high temperature applications, such as CSP, molten salts are the most widely used material. This is due to their high volumetric heat capacity, a high boiling point, high temperature stability, and their vapor pressure being close to zero. Additionally, they are relatively cheap, readily available, neither toxic nor flammable, and can act as an HTF as well as a TES material. However, they have certain disadvantages: they have a relatively high melting point (generally > 200 °C), which results in them solidifying in pipes in the absence of a heat source and thus necessitates the installation of antifreeze systems; they also have high viscosity and low thermal conductivity compared to other fluids [9,13].
Focusing on reducing the levelized cost of electricity (LCOE) for CSP plants, which is currently slightly higher than other renewable alternatives [14], there is presently a strong interest in developing higher temperature power systems (around 700 °C) with better efficiency, such as supercritical CO2 Brayton cycles [15]. Current CSP plants are limited to relatively low temperatures (below 600 °C), which means they are enough for parabolic through collector (PTC) or linear Fresnel reflector (LFR) plants. However, they are inadequate for higher temperatures (up to 950 °C) that can be achieved in central tower (CT) plants [16]. Hence, there is substantial potential to increase the operating temperatures (increasing the Carnot efficiency) and the heat-to-electricity efficiency for the next CSP generation. Replacing commercial nitrate-based molten salts with other TES materials such as chloride, fluoride, or carbonate-based materials will achieve this [17,18]. Table 1 and Table 2 summarize the TES materials analyzed in the present work along with their thermal properties.

2.1. Nitrate-Based Materials

The most popular TES material used in CSP is Solar Salt: a non-eutectic mixture of 60% NaNO3–40% KNO3 (wt%), which has been investigated since the 1980s [46,47]. However, there are other candidates, mostly derived from Solar Salt. The main reasons for its popularity are its relatively low cost, its good chemical safety (it is neither toxic nor flammable), and its reasonable material compatibility which allows standard stainless steels to be used without incurring high corrosion rates [48,49]. However, its operating temperature range is restricted by a crystallization temperature of around 240 °C and a maximum operating temperature of around 565 °C, after which decomposition and salt degradation reactions begin [19,50]. In general, when considering an alkaline and alkaline earth nitrate generic mixture, the following main decomposition reactions can occur [21,51]:
MNO 3 MNO 2 + 1 2 O 2
5   MNO 2   3   MNO 3   +   M 2 O +   N 2
M ( NO 3 ) 2   M O   +   N 2 +   5 2 O 2
M ( NO 3 ) 2   M O   +   NO 2 +   NO + O 2
Being M alkaline and M′ alkaline earth metals. Oxides can be produced [20] and nitrates can in turn react with the atmospheric humidity and carbon dioxide:
M 2 O + H 2 O 2 MOH
M 2 O + CO 2 M 2 CO 3
M O + H 2 O M ( OH ) 2
M ( NO 3 ) 2 + CO 2   M CO 3   +   N 2 +   5 2 O 2
A realistic principle to determine the stability of nitrate mixtures is to measure the occurrence of reactions (3), (4) and (6)–(9) where oxides, hydroxides, and carbonates are produced [52]. In fact, the formation of nitrites from the equilibrium in reaction (2) does not prejudice neither the thermophysical properties of the mixtures nor the compatibility properties [51]. On the other hand, the solubility of oxides, hydroxides, and carbonates is limited so solids can be produced and, moreover, the corrosion phenomena from the thermal fluid are more acute [21]. From several investigations of different mixtures of alkaline and alkaline earth nitrates, the following considerations can be made:
  • If only sodium and potassium nitrates are present, the thermal stability limit is around 600 °C [51,53,54,55,56,57].
  • The presence of lithium nitrate leads to a thermal stability limit between 550 and 600 °C [51,58].
  • The addition of calcium nitrate produces a drastic decrease in long-term stability; an upper temperature value of around 450 °C has been proposed [21,22].
  • The presence of nitrites limits the thermal stability at about 450 °C under air atmosphere [59,60].
  • The decomposition kinetics depends on the concentration of chemical species; with respect to the nitrate mole fraction, a first order kinetics was found for sodium and potassium nitrates’ decomposition into nitrites [53,57].
  • The kinetics degradation is also related to the surrounding atmosphere, its humidity, and CO2 content [52].
Little effort has been devoted to increasing the stability limit of the upper temperature of nitrates, the chemical nature of the stability loss, nor the long periods of time in which the maximum temperature must be maintained. In addition, the results obtained using other cations different from alkali metals lead to a decrease in thermal stability [52,61,62]. To overcome this problem, some authors have proposed confinement of the molten salt in a controlled atmosphere that allows for an increase of the partial pressure of oxygen and for a shift to the nitrate side the reaction number 2 in order to decrease the nitrite concentrations at the same temperature [63]. Nonetheless, the current research into these materials is focused on decreasing the crystallization points of the mixtures and adding other alkali (i.e., lithium) or alkaline earth (i.e., calcium) cations, or other anions (i.e., nitrite) [51].
Regarding lithium nitrate, mixtures have been proposed in combination with sodium and potassium nitrates whose crystallization points can be found below 130 °C and report thermophysical properties comparable to those presented by Solar Salt [47,51,58,64,65,66,67]. Concerning calcium nitrate, several formulations related to its addition to sodium and potassium nitrates have been proposed [47,62,67,68]. The most researched is the one marketed as Hitec XL (15% NaNO3–43% KNO3–42% Ca(NO3)2 wt%), which closely corresponds to the ternary eutectic point of the published phase diagrams [68,69]. However, the presence of Ca(NO3)2 makes it difficult to accurately determine the crystallization point by calorimetric methods (around 130 °C) given the presence of slow transitions with low enthalpies [68], which explains the uncertainties and disparities presented in the literature. Similarly, the maximum operating temperature presents discrepancies in the published articles due to decomposition reactions [62,67,68]. On the other hand, Ca(NO3)2 increases the viscosity up to three times in comparison to the reference values of Solar Salt and this compromises the application of these mixtures at very low temperatures [62]. Both previously described nitrates have also been proposed to be used together, especially in ternary mixtures that include KNO3 [22,58,65,67].
The only anion proposed for these mixtures and applications with nitrate is nitrite, mainly sodium nitrite in a ternary mixture with sodium and potassium nitrates, and the composition most widely studied and used is the eutectic one (7% NaNO3–53% KNO3–40% NaNO2 wt%), also known as Hitec [20]. The use of nitrites leads to a decrease in the melting point of the liquid, but its presence limits the thermal stability at about 450 °C under air atmosphere due to decomposition and degradation reactions, limiting its application to mid-temperature applications [59,60].
Other high-order mixtures constituted by alkaline nitrates [20,26,47,58] or nitrate−nitrite mixtures [25,26,65] have been proposed, but they have received little attention, and as of yet there is no evidence of any commercial application.

2.2. Chloride-Based Materials

Molten mixtures of chlorides have been considered as coolants in nuclear reactor designs where fluids operate at about 525 °C and are stable at temperatures above 800 °C [45,70,71]. Due to their low cost, high decomposition temperatures, and natural abundance, molten chlorides are considered good TES material candidates even if their melting temperatures are generally higher than those of nitrate salts, but typically lower than fluoride and carbonate-based TES materials [72]. In recent years, research has focused on the development of multicomponent mixtures of lithium, sodium, potassium, magnesium, calcium, and zinc chlorides. Nonetheless, although lithium and zinc chlorides offer the mixtures the lowest melting point, their high cost implies a major disadvantage [15,72,73].
Regarding binary mixtures, the KCl–MgCl2 eutectic composition has the lowest melting temperature among all possible binaries considered above [27,74] and has been widely studied as a primary coolant in nuclear power [75]. Its vapor pressure is very low, even at high temperatures, resulting in good thermal stability [43]. In addition, if NaCl is added to the mixture, the melting temperature of the ternary eutectic is about 50 °C lower, which allows the thermal properties to remain almost unchanged [28].
The NaCl–MgCl2–CaCl2 mixture has been proposed due to its relatively low-temperature melting point (424 °C) and because it has a heat capacity higher than all the alkaline and alkaline earth chloride binary mixtures, excluding the ones with lithium chloride [29]. Moreover, the addition of potassium chloride to the latter mixture diminishes the melting point to 385 °C and increases the heat capacity by more than 25% [76]. By contrast, calcium chloride mixtures lack the same thermal stability as other chlorides, resulting in an operating temperature gap lower than other chloride mixtures [29].
Mixtures containing zinc chloride are very interesting due to the low melting temperature achieved. The NaCl–KCl–ZnCl2 mixture is one of the most promising candidates given its eutectic has a very low melting temperature of 204 °C [32]. The KCl–MgCl2–ZnCl2 mixture eutectic also has a reasonably low melting temperature of 356 °C [62] and high thermal stability of up to 730 °C [77]. However, as was mentioned earlier, the high cost of zinc chloride makes these mixtures less attractive.
Nevertheless, their main drawback is corrosion: the corrosion mechanisms are complex, with different interactions between atmosphere, molten salts, outer and inner corrosion layers, and the matrix of metallic alloys [78,79,80,81]. Chloride melts facilitate corrosion of steels in air, and this fact should be considered in the development of corrosion mitigation methods, such as inert atmospheres and gas purification of the melts prior to its operation [82]. Container materials should be nickel-based superalloys which have better corrosion resistance than stainless steels [83]. However, this leads to a more expensive solution for nitrate-based TES materials and should be taken into account for future applications.

2.3. Fluoride and Carbonate-Based Materials

Fluoride-based molten salts have been used as nuclear coolant fluids due to their relatively high specific heat capacity, thermal conductivity, and thermal stability compared to other molten salts, including experimental reactors on a megawatt scale and in test loops for hundreds of thousands of hours [33,71]. On the other hand, carbonate-based materials are widely used in fuel cells, operating between 600–700 °C [84], and for cleaning and processing uranium-containing alloys in the nuclear power industry [85].
In general, fluorides tend to be better HTF than chlorides and carbonates [45,70] and possess almost the same thermal stability over 700 °C [86]. It has been stated that at operating temperatures, the thermophysical properties of liquid fluoride-based salts are similar to those of room temperature water except for their very low vapor pressure [33]. However, carbonate-based TES materials are generally more stable; a limit decomposition temperature under air of 670 °C, 700 °C under argon, and no sign of decomposition under CO2 atmosphere, even at 1000 °C, has been reported [86].
The first fluoride material presented in this work is the ternary eutectic mixture of LiF–NaF–KF, which has good thermal properties, but a high melting temperature (see Table 2). This is also a problem for other fluoride-based mixtures, but fluoroborates mixtures can slightly decrease the melting temperature and maintain other thermal properties [34,35]. In addition, compositions based on zinc fluorides can reduce the temperature even further and present figures similar to those of chloride-based salts [34,35]. Other fluorides, such as beryllium fluoride, demonstrate good performance and different mixtures have been proposed, but their high cost makes them inappropriate for this application [87].
Single and binary mixtures of carbonate salts have high melting points above 500 °C [88]. However, the ternary eutectic mixture Li2CO3–Na2CO3–K2CO3 has been proposed as a TES material with a melting temperature of 397 °C [36]. In particular, the richer the potassium carbonate mixtures are, the higher the temperature value they will have (more than three times that of Solar Salt) [18]. Finally, some authors suggest mixtures of fluorides with carbonates or other fluorides, carbonates, and chlorides, but they have received little attention [89,90].
The main drawback for fluoride-based molten salt is corrosion, which is much lower in carbonate and chloride salt mixtures. They should be thermodynamically stable relative to the confinement materials because they tend to rapidly dissolve many of their oxide protective layers, which limits choices [33,91]. By contrast, it has been stated that carbonates are much less corrosive than fluorides and commercial stainless steels can be used in most practical situations [18,92]. Nevertheless, viscosity is their main problem, as it is higher in general than the other proposed salt mixtures in this work and therefore, its suitability for this application, due to pumping issues, must be revised [93].

3. Molten Salts Properties Enhancement

One of the main reasons to advocate for the use of the molten salts in CSP plants and other industrial facilities is their properties’ enhancement. This can be done by adding nanoparticles (NPs) to the fused system in order to create a nanofluid (NF) [94]. The basis for this solution is simple: the NPs can increase the specific heat capacity, precisely one of the main properties that is relatively low in all molten salts in comparison with other fluid-sensible materials like water or liquid metals, and therefore, it enhances the overall CSP efficiency [95,96,97]. However, the rheological behavior of NFs is a limiting factor because the viscosity of particulate suspensions increases as a function of particle volume fraction. Consequently, a maximum number of NPs can be added to the fluid without compromising the performance of the system exits [98,99].
Because of the high operating temperatures and aggressive nature of molten salts, the main requirements of NP are to be stable at high temperatures and to have a strong affinity to the adsorbate [100]. In addition, it is essential that they do not suffer deformations and/or chemical degradation in order to conserve the NF design features.
NFs’ complex nature implies that many factors affect their thermal behavior [101,102,103,104,105]:
  • Particle concentration, typically lower than 2%;
  • Particle size, from 2 to 90 nm;
  • Particle shape, mainly spherical or cylindrical;
  • Particle type, oxides being the most popular, followed by carbon-based materials such as graphite or multi-walled carbon nanotubes (MWCNTs) and solid metals;
  • Interaction between liquid and particles.
Among nitrate-based molten salts, Solar Salt is the most investigated base fluid. Different types and sizes of NPs like alumina, silica, iron, titanium, and copper or zinc oxides have been investigated [106,107,108,109], reporting a maximum specific heat capacity enhancement of 31.1% with 0.5%wt of silica NP [110]. Other nitrated-based mixtures mentioned in this work have also been studied, mainly silica NPs, and the authors reported similar enhancements compared to those obtained in the Solar Salt case [94,100,101,102,103,104,105,106,107,108,109,110,111,112].
Two different chloride-based NFs have been proposed. First, BaCl2–NaCl–CaCl2 mixtures with 1%wt of silica 20–30 nm NPs, which obtain a specific heat capacity enhancement of about 14.5% [94]. Second, KCl–CaCl2–LiCl mixtures have been proposed with 1%wt of silica 10 nm NPs that give a specific heat capacity enhancement of about 7.6% [113].
Regarding carbonate-based molten salts, the most investigated is the Li2CO3–K2CO3 eutectic mixture. It has been enhanced with silica, MWCNT, alumina, graphite, and TiO2 NPs [99,100,114,115,116,117,118,119,120] with different types and sizes. There is reportedly a very low specific heat capacity increase of about 1–2% in various composition ranges with silica, but up to 17% in mixtures containing ≥0.5% of MWCNT NPs [99,121]. Until now, no fluoride-based materials have been reported in the literature as NF candidates.

4. Energy and Cost Analysis of TES Materials

Many economic assessments have been made elsewhere both for CSP plants and TES systems [114,115,117,119,122,123,124], and it has been demonstrated that a CSP plant is more cost-effective if it has a TES system [125,126]. There are many key parameter indicators (KPIs) that can be used to carry out the techno-economic analysis for TES materials, and in this work, the following were used:
  • Specific mass energy density Em (in MJ/kg): quantifies the amount of energy stored by the TES material in its operating temperature range and is defined in Equation (9).
  • Specific volumetric energy density Ev (in MJ/m3): quantifies the same energy as the previous KPI, but in terms of volume (its definition is given in Equation (10)). This is helpful when estimating other KPIs in the TES system (i.e., confinement, land-area dimensions, and cost) and is useful in determining the flow rate and if it was going to be used as HTF.
  • Energy storage cost Ec (in $/MJ): represents the direct cost of the stored energy and is defined in Equation (11).
E m = T f u s T d e c c p d T
E v = T f u s T d e c ρ c p d T
E c = C E m
where Tfus and Tdec are the fusion and decomposition temperatures (in °C), ρ is the density (in kg/m3), cp is the specific heat (in J/kg °C), and C is the cost (in $/kg) of the molten salts. In this analysis, the temperature range considered will always be higher than real ones and narrower due to safety considerations such as avoiding solidification and decomposition. The values for the selected TES materials in this work are presented in Table 3. In this analysis, Solar Salt was used as a reference for comparison issues, and 700 °C was considered the maximum temperature allowed.
It is clear that for low temperature applications like PTC or LFR, nitrated-based TES materials are the only ones suitable, while the rest are reserved for high temperature CT plants operating up to 700 °C. Figure 1 and Figure 2 gathers KPIs comparison for selected candidates.
Regarding specific mass energy density, values for nitrate-based materials are the most attractive closely followed by carbonates, whereas chlorides and fluorides are in general relatively less interesting. It is worth to point out that Solar Salt has one of the highest values, only surpassed by nitrate mixtures containing lithium or lithium and calcium. In the case of other mixtures, LiNaKCO3, NaKZnCl, and NaBF have very close values to those reported for Solar Salt, giving high-temperature applications almost the same mass energy density. For specific volume energy density, the values are very similar to the previous ones; however, fluoride-based materials enhance their figures, almost to the same level as the nitrates, whereas LiNaKCO3 has an improved performance compared to Solar Salt. Finally, KPIs of previous chloride-based materials perform poorly in comparison to the other candidates, but its specific energy costs are more than half the Solar Salt ones, even achieving a reduction over 75%. This makes them one of the best choices for future high-temperature applications over fluorides and carbonate-based materials.

5. Conclusions

A comprehensive review of different thermal energy storage (TES) materials for concentrated solar power (CSP) has been completed: fifteen selected materials have been studied and compared and their nature, thermophysical properties, and economic impact have been considered. Three energy key performance indicators (KPIs) have been defined in order to evaluate the performance of the different molten salts by using Solar Salt as a reference.
The results of this analysis reveal that the most popular and sensible TES material, Solar Salt, scores reasonably well in comparison with other nitrate-based materials in terms of energy density and cost for its temperature interval of operation. However, other nitrate-based materials achieve almost the same figures, but their broader temperature ranges have the potential to be more interesting for very low-temperature applications.
In general, the studied chloride-based materials seem to be the favorite candidates for high-temperature applications compared to other fluoride and carbonate-based options. This is due to their very attractive low cost (excluding KMgZnCl), as even their energy KPIs give lower values than Solar Salt (except NaKZnCl).
For low-temperature applications, nitrate-based materials are more suitable due to their low melting temperature and good energy KPIs. On the other hand, for high-temperature applications, chloride-based materials are better choices over fluorides and carbonates, mainly due to their lower cost. However, the drawback that still remains is the corrosion of these materials, and this must be solved if they are to be the next generation of TES materials.

Author Contributions

Conceptualization, A.C., S.G.-C. and Á.C.; methodology, S.G.-C. and Á.C.; investigation, A.C., S.G.-C., Á.C. and S.S.; resources, A.C. and S.S.; writing—original draft preparation, A.C. and S.S.; writing—review and editing, A.C., S.G.-C., Á.C. and S.S.; supervision, S.G.-C. and Á.C.; project administration, S.G.-C. and Á.C.; funding acquisition, Á.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by CSIC under project 2017 60E028.

Acknowledgments

We acknowledge support of the publication fee by the CSIC Open Access Publication Support Initiative through its Unit of Information Resources for Research (URICI).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. de Sisternes, F.J.; Jenkins, J.D.; Botterud, A. The value of energy storage in decarbonizing the electricity sector. Appl. Energy 2016, 175, 368–379. [Google Scholar] [CrossRef] [Green Version]
  2. IPCC. Summary for Policymakers, Climate Change 2014: Mitigation of Climate Change; Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change; Intergovernmental Panel on Climate Change: Cambridge, UK, 2014. [Google Scholar]
  3. Herrmann, U.; Kearney, D.W. Survey of Thermal Energy Storage for Parabolic Trough Power Plants. J. Sol. Energy Eng. 2002, 124, 145–152. [Google Scholar] [CrossRef] [Green Version]
  4. IEA. Concentrating Solar Power, IEA Technology Roadmaps; OECD Publishing: Paris, France, 2010. [Google Scholar]
  5. Chen, H.; Cong, T.N.; Yang, W.; Tan, C.; Li, Y.; Ding, Y. Progress in electrical energy storage system: A critical review. Prog. Nat. Sci. 2009, 19, 291–312. [Google Scholar] [CrossRef]
  6. Xu, B.; Li, P.; Chan, C. Application of phase change materials for thermal energy storage in concentrated solar thermal power plants: A review to recent developments. Appl. Energy 2015, 160, 286–307. [Google Scholar] [CrossRef]
  7. Pardo, P.; Deydier, A.; Anxionnaz-Minvielle, Z.; Rougé, S.; Cabassud, M.; Cognet, P. A review on high temperature thermochemical heat energy storage. Renew. Sustain. Energy Rev. 2014, 32, 591–610. [Google Scholar] [CrossRef] [Green Version]
  8. Laing, D.; Steinmann, W.-D.; Tamme, R.; Richter, C. Solid media thermal storage for parabolic trough power plants. Sol. Energy 2006, 80, 1283–1289. [Google Scholar] [CrossRef]
  9. Gil, A.; Medrano, M.; Martorell, I.; Lázaro, A.; Dolado, P.; Zalba, B.; Cabeza, L.F. State of the art on high temperature thermal energy storage for power generation. Part 1—Concepts, materials and modellization. Renew. Sustain. Energy Rev. 2010, 14, 31–55. [Google Scholar] [CrossRef]
  10. Alva, G.; Liu, L.; Huang, X.; Fang, G. Thermal energy storage materials and systems for solar energy applications. Renew. Sustain. Energy Rev. 2017, 68, 693–706. [Google Scholar] [CrossRef]
  11. Alnaimat, F.; Rashid, Y. Thermal Energy Storage in Solar Power Plants: A Review of the Materials, Associated Limitations, and Proposed Solutions. Energies 2019, 12, 4164. [Google Scholar] [CrossRef] [Green Version]
  12. Fernández, A.G.; Gomez-Vidal, J.; Oró, E.; Kruizenga, A.; Solé, A.; Cabeza, L.F. Mainstreaming commercial CSP systems: A technology review. Renew. Energy 2019, 140, 152–176. [Google Scholar] [CrossRef]
  13. Alva, G.; Lin, Y.; Fang, G. An overview of thermal energy storage systems. Energy 2018, 144, 341–378. [Google Scholar] [CrossRef]
  14. IRENA. Renewable Power Generation Costs in 2017; International Renewable Energy Agency: Abu Dhabi, United Arab Emirates, 2018. [Google Scholar]
  15. Mohan, G.; Venkataraman, M.; Gomez-Vidal, J.; Coventry, J. Assessment of a novel ternary eutectic chloride salt for next generation high-temperature sensible heat storage. Energy Convers. Manag. 2018, 167, 156–164. [Google Scholar] [CrossRef]
  16. Hoffschmidt, B.; Tellez, F.M.; Valverde, A.; Fernandez, J.; Fernandez, V. Performance evaluation of the 200-kW(th) HiTRec-II open volumetric air receiver. J. Sol. Energy Eng. 2003, 125, 87–94. [Google Scholar] [CrossRef]
  17. Myers, P.D., Jr.; Goswami, D.Y. Thermal energy storage using chloride salts and their eutectics. Appl. Therm. Eng. 2016, 109, 889–900. [Google Scholar] [CrossRef] [Green Version]
  18. Wang, T.; Mantha, D.; Reddy, R.G. Novel high thermal stability LiF–Na2CO3–K2CO3 eutectic ternary system for thermal energy storage applications. Sol. Energy Mater. Sol. Cells 2015, 140, 366–375. [Google Scholar] [CrossRef] [Green Version]
  19. Bauer, T.; Pfleger, N.; Breidenbach, N.; Eck, M.; Laing, D.; Kaesche, S. Material aspects of Solar Salt for sensible heat storage. Appl. Energy 2013, 111, 1114–1119. [Google Scholar] [CrossRef]
  20. Raade, J.W.; Padowitz, D. Development of Molten Salt Heat Transfer Fluid with Low Melting Point and High Thermal Stability. J. Sol. Energy Eng. 2011, 133, 031013. [Google Scholar] [CrossRef]
  21. Bradshaw, R.; Meeker, D. High-temperature stability of ternary nitrate molten salts for solar thermal energy systems. Sol. Energy Mater. 1990, 21, 51–60. [Google Scholar] [CrossRef]
  22. Zhao, C.Y.; Wu, Z.G. Thermal property characterization of a low melting temperature ternary nitrate salt mixture for thermal energy storage systems. Sol. Energy Mater. Sol. Cells 2011, 95, 3341–3346. [Google Scholar] [CrossRef]
  23. Wang, T.; Mantha, D.; Reddy, R.G. Thermal stability of the eutectic composition in LiNO3–NaNO3–KNO3 ternary system used for thermal energy storage. Sol. Energy Mater. Sol. Cells 2012, 100, 162–168. [Google Scholar] [CrossRef]
  24. Coscia, K.; Nelle, S.; Elliott, T.; Mohapatra, S.; Oztekin, A.; Neti, S. Thermophysical properties of Li-NO3–NaNO3–KNO3 mixtures for use in concentrated solar power. J. Sol. Energy Eng. 2013, 135, 4024069. [Google Scholar] [CrossRef]
  25. Peng, Q.; Ding, J.; Wei, X.; Jiang, G. Thermodynamic Investigation of the Eutectic Mixture of the Li-NO3-NaNO3-KNO3-Ca(NO3)2 System. Int. J. Thermophys. 2017, 38, 142. [Google Scholar] [CrossRef]
  26. Wang, T.; Mantha, D.; Reddy, R.G. Novel low melting point quaternary eutectic system for solar thermal energy storage. Appl. Energy 2013, 102, 1422–1429. [Google Scholar] [CrossRef]
  27. Li, Y.; Xu, X.; Wang, X.; Li, P.; Hao, Q.; Xiao, B. Survey and evaluation of equations for thermophysical properties of binary/ternary eutectic salts from NaCl, KCl, MgCl2, CaCl2, ZnCl2 for heat transfer and thermal storage fluids in CSP. Sol. Energy 2017, 152, 57–79. [Google Scholar] [CrossRef]
  28. Vidal, J.C.; Klammer, N. Molten chloride technology pathway to meet the U.S. DOE sunshot initiative with Gen3 CSP. In AIP Conference Proceedings; AIP Publishing LLC: Melville, NY, USA, 2019; Volume 2126, p. 080006. [Google Scholar]
  29. Wei, X.; Song, M.; Wang, W.; Ding, J.; Yang, J. Design and thermal properties of a novel ternary chloride eutectics for high-temperature solar energy storage. Appl. Energy 2015, 156, 306–310. [Google Scholar] [CrossRef]
  30. Du, L.; Tian, H.; Wang, W.; Ding, J.; Wei, X.; Song, M. Thermal stability of the eutectic composition in NaCl-CaCl2-MgCl2 ternary system used for thermal energy storage applications. Energy Procedia 2017, 105, 4185–4191. [Google Scholar] [CrossRef]
  31. Du, L.; Ding, J.; Tian, H.; Wang, W.; Wei, X.; Song, M. Thermal properties and thermal stability of the ternary eutectic salt NaCl-CaCl2-MgCl2 used in high-temperature thermal energy storage process. Appl. Energy 2017, 204, 1225–1230. [Google Scholar] [CrossRef]
  32. Li, P.; Molina, E.; Wang, K.; Xu, X.; Dehghani, G.; Kohli, A.; Hao, Q.; Kassaee, M.H.; Jeter, S.M.; Teja, A.S. Thermal and Transport Properties of NaCl–KCl–ZnCl2 Eutectic Salts for New Generation High-Temperature Heat-Transfer Fluids. J. Sol. Energy Eng. 2016, 138, 054501. [Google Scholar] [CrossRef]
  33. 3Forsberg, C.W.; Peterson, P.F.; Zhao, H. High-Temperature Liquid-Fluoride-Salt Closed-Brayton-Cycle Solar Power Towers. J. Sol. Energy Eng. 2006, 129, 141–146. [Google Scholar] [CrossRef]
  34. Serrano-López, R.; Fradera, J.; Cuesta-López, S. Molten salts database for energy applications. Chem. Eng. Process. Process. Intensif. 2013, 73, 87–102. [Google Scholar] [CrossRef] [Green Version]
  35. Jerden, J. Molten Salt Thermophysical Properties Database Development: 2019 Update; 2019 Update (No. ANL/CFCT-19/6); Argonne National Laboratory (ANL): Argonne, IL, USA, 2019. [Google Scholar]
  36. An, X.-H.; Cheng, J.-H.; Su, T.; Zhang, P. Determination of thermal physical properties of alkali fluoride/carbonate eutectic molten salt. In AIP Conference Proceedings; AIP Publishing LLC: Melville, NY, USA, 2017; Volume 1850, p. 070001. [Google Scholar]
  37. Zavoico, A.B. Design Basis Document SAND2001–2100; Sandia National Laboratories: Albuquerque, NM, USA, 2001. [Google Scholar]
  38. Boerema, N.; Morrison, G.; Taylor, R.; Rosengarten, G. Liquid sodium versus Hitec as a heat transfer fluid in solar thermal central receiver systems. Sol. Energy 2012, 86, 2293–2305. [Google Scholar] [CrossRef]
  39. Siegel, N.P.; Bradshaw, R.W.; Cordaro, J.B.; Kruizenga, A.M. Thermophysical property measurement of nitrate salt heat transfer fluids. In Proceedings of the ASME 2011 5th International Conference on Energy Sustainability, Washington, DC, USA, 7–10 August 2011. [Google Scholar]
  40. Roget, F.; Favotto, C.; Rogez, J. Study of the KNO3–LiNO3 and KNO3–NaNO3–LiNO3 eutectics as phase change materials for thermal storage in a low-temperature solar power plant. Sol. Energy 2013, 95, 155–169. [Google Scholar] [CrossRef]
  41. Jiang, Z.; Leng, G.; Ye, F.; Ge, Z.; Liu, C.; Wang, L.; Ding, Y. Form-stable LiNO3–NaNO3–KNO3–Ca (NO3)2/calcium silicate composite phase change material (PCM) for mid-low temperature thermal energy storage. Energy Convers. Manag. 2015, 106, 165–172. [Google Scholar] [CrossRef]
  42. Cordaro, J.G.; Rubin, N.C.; Bradshaw, R.W. Multicomponent Molten Salt Mixtures Based on Nitrate/Nitrite Anions. J. Sol. Energy Eng. 2011, 133, 011014. [Google Scholar] [CrossRef]
  43. Xu, X.; Wang, X.; Li, P.; Li, Y.; Hao, Q.; Xiao, B.; Elsentriecy, H.; Gervasio, D. Experimental Test of Properties of KCl–MgCl2 Eutectic Molten Salt for Heat Transfer and Thermal Storage Fluid in Concentrated Solar Power Systems. J. Sol. Energy Eng. 2018, 140, 051011. [Google Scholar] [CrossRef]
  44. Xu, X.; Dehghani, G.; Ning, J.; Li, P. Basic properties of eutectic chloride salts NaCl-KCl-ZnCl2 and NaCl-KCl-MgCl2 as HTFs and thermal storage media measured using simultaneous DSC-TGA. Sol. Energy 2018, 162, 431–441. [Google Scholar] [CrossRef]
  45. Williams, D.F.; Toth, L.M.; Clarno, K.T. Assessment of Candidate Molten Salt Coolants for the Advanced High Temperature Reactor (AHTR); Oak Ridge National Laboratory: Oak Ridge, TN, USA, 2006; pp. 1–69. [Google Scholar]
  46. Bradshaw, R.W.; Siegel, N.P. Molten nitrate salt development for thermal energy storage in parabolic trough solar power systems. In Proceedings of the ASME 2008 2nd International Conference on Energy Sustainability, Jacksonville, FL, USA, 10–14 August 2008; ASME International: New York, NY, USA, 2008; pp. 631–637. [Google Scholar]
  47. Glatzmaier, G.N.; Siegel, N.P. Molten Salt Heat Transfer Fluids and Thermal Storage Technology Sandia Technical Report; SAND2010-3826C; Sandia National Laboratories: Albuquerque, NM, USA, 2010. [Google Scholar]
  48. Mehos, M.; Turchi, C.; Vidal, J.; Wagner, M.; Ma, Z.; Ho, C.; Kolb, W.; Andraka, C.; Kruizenga, A. Concentrating Solar Power Gen3 Demonstration Roadmap; (No. NREL/TP-5500-67464); National Renewable Energy Laboratory (NREL): Golden, CO, USA, 2017. [Google Scholar]
  49. Walczak, M.; Pineda, F.; Fernández-Ángel, G.; Mata-Torres, C.; Escobar, R.A. Materials corrosion for thermal energy storage systems in concentrated solar power plants. Renew. Sustain. Energy Rev. 2018, 86, 22–44. [Google Scholar] [CrossRef]
  50. Zhang, X.; Tian, J.; Xu, K.; Gao, Y. Thermodynamic evaluation of phase equilibria in NaNO3-KNO3 system. J. Phase Equilibria Diffus. 2003, 24, 441–446. [Google Scholar] [CrossRef]
  51. Sau, S.; Corsaro, N.; Crescenzi, T.; D’Ottavi, C.; Liberatore, R.; Licoccia, S.; Russo, V.; Tarquini, P.; Tizzoni, A. Techno-economic comparison between CSP plants presenting two different heat transfer fluids. Appl. Energy 2016, 168, 96–109. [Google Scholar] [CrossRef]
  52. Olivares, R.I. The thermal stability of molten nitrite/nitrates salt for solar thermal energy storage in different atmos-pheres. Sol. Energy 2012, 86, 2576–2583. [Google Scholar] [CrossRef]
  53. Freeman, E.S. The Kinetics of the Thermal Decomposition of Potassium Nitrate and of the Reaction between Potassium Nitrite and Oxygen1a. J. Am. Chem. Soc. 1957, 79, 838–842. [Google Scholar] [CrossRef]
  54. Kruizenga, A.M.; Cordaro, J.G. Preliminary Development of Thermal Stability Criterion for Alkali Nitrates; Sandia Technical Report SAND2011-5837C; Sandia National Laboratories: Albuquerque, NM, USA, 2011. [Google Scholar]
  55. Bauer, T.; Laing, D.; Tamme, R. Characterization of Sodium Nitrate as Phase Change Material. Int. J. Thermophys. 2012, 33, 91–104. [Google Scholar] [CrossRef]
  56. Vignarooban, K.; Xu, X.; Arvay, A.; Hsu, K.; Kannan, A. Heat transfer fluids for concentrating solar power systems—A review. Appl. Energy 2015, 146, 383–396. [Google Scholar] [CrossRef]
  57. Freeman, E.S. The Kinetics of the Thermal Decomposition of Sodium Nitrate and of the Reaction between Sodium Nitrite and Oxygen. J. Phys. Chem. 1956, 60, 1487–1493. [Google Scholar] [CrossRef]
  58. Fernández, A.; Ushak, S.; Galleguillos, H.R.; Pérez, F. Development of new molten salts with LiNO3 and Ca(NO3)2 for energy storage in CSP plants. Appl. Energy 2014, 119, 131–140. [Google Scholar] [CrossRef]
  59. Cordaro, J.G.; Rubin, N.C.; Sampson, M.D.; Bradshaw, R.W. Multi-component molten salt mixtures based on ni-trate/nitrite anions. In Proceedings of the 6th Electronic Proceedings SolarPaces, Perpignan, France, 21–24 September 2010. [Google Scholar]
  60. Villada, C.; Bonk, A.; Bauer, T.; Bolívar, F. High-temperature stability of nitrate/nitrite molten salt mixtures under different atmospheres. Appl. Energy 2018, 226, 107–115. [Google Scholar] [CrossRef]
  61. Wu, Y.-T.; Li, Y.; Ren, N.; Ma, C.-F. Improving the thermal properties of NaNO3-KNO3 for concentrating solar power by adding additives. Sol. Energy Mater. Sol. Cells 2017, 160, 263–268. [Google Scholar] [CrossRef]
  62. Bonk, A.; Sau, S.; Uranga, N.; Hernaiz, M.; Bauer, T. Advanced heat transfer fluids for direct molten salt line-focusing CSP plants. Prog. Energy Combust. Sci. 2018, 67, 69–87. [Google Scholar] [CrossRef]
  63. Bonk, A.; Braun, M.; Sötz, V.A.; Bauer, T. Solar Salt—Pushing an old material for energy storage to a new limit. Appl. Energy 2020, 262, 114535. [Google Scholar] [CrossRef]
  64. Fernández, A.; Carbajo, R.; Crutchik, M.; Galleguillos, H.; Fuentealba, E. Modular Solar Storage System for Medium Scale Energy Demand Mining in the North of Chile. Energy Procedia 2015, 69, 638–643. [Google Scholar] [CrossRef] [Green Version]
  65. Cáceres, G.; Montané, M.; Nasirov, S.; O’Ryan, R. Review of Thermal Materials for CSP Plants and LCOE Evaluation for Performance Improvement using Chilean Strategic Minerals: Lithium Salts and Copper Foams. Sustainability 2016, 8, 106. [Google Scholar] [CrossRef] [Green Version]
  66. Olivares, R.I.; Edwards, W. LiNO3–NaNO3–KNO3 salt for thermal energy storage: Thermal stability evaluation in different atmospheres. Thermochim. Acta 2013, 560, 34–42. [Google Scholar] [CrossRef]
  67. Zhang, P.; Uwaisg, J.; Jin, Y.; An, X. Evaluation of thermal physical properties of molten nitrate salts with low melting temperature. Sol. Energy Mater. Sol. Cells 2018, 176, 36–41. [Google Scholar] [CrossRef]
  68. Gomez, J.C.; Calvet, N.; Starace, A.K.; Glatzmaier, G.C. Ca(NO3)2—NaNO3—KNO3 Molten Salt Mixtures for Direct Thermal Energy Storage Systems in Parabolic Trough Plants. J. Sol. Energy Eng. 2013, 135, 021016. [Google Scholar] [CrossRef]
  69. Chen, M.; Shen, Y.; Zhu, S.; Li, P. Digital phase diagram and thermophysical properties of KNO3-NaNO3-Ca(NO3)2 ternary system for solar energy storage. Vacuum 2017, 145, 225–233. [Google Scholar] [CrossRef]
  70. Williams, D.F. Assessment of Candidate Molten Salt Coolants for the NGNP/NHI Heat-Transfer Loop; No. ORNL/TM-2006/69; Oak Ridge National Laboratory (ORNL): Oak Ridge, TN, USA, 2006. [Google Scholar]
  71. Le Brun, C. Molten salts and nuclear energy production. J. Nucl. Mater. 2007, 360, 1–5. [Google Scholar] [CrossRef] [Green Version]
  72. Mohan, G.; Venkataraman, M.B.; Coventry, J. Sensible energy storage options for concentrating solar power plants operating above 600 °C. Renew. Sustain. Energy Rev. 2019, 107, 319–337. [Google Scholar] [CrossRef]
  73. Badenhorst, H.; Böhmer, T. Enthalpy of fusion prediction for the economic optimisation of salt based latent heat thermal energy stores. J. Energy Storage 2018, 20, 459–472. [Google Scholar] [CrossRef]
  74. Li, C.J.; Li, P.; Wang, K.; Molina, E.E. Survey of properties of key single and mixture halide salts for potential ap-plication as high temperature heat transfer fluids for concentrated solar thermal power systems. AIMS Energy 2014, 2, 133–157. [Google Scholar] [CrossRef]
  75. Sohal, M.S.; Ebner, M.A.; Sabharwall, P.; Sharpe, P. Engineering Database of Liquid Salt Thermophysical and Ther-Mochemical Properties; No. INL/EXT-10-18297; Idaho National Laboratory: Idaho Falls, ID, USA, 2010. [Google Scholar]
  76. Wei, X.; Song, M.; Ding, J.; Yang, J. Quaternary chloride eutectic mixture for thermal energy storage at high tem-perature. Energy Procedia 2015, 75, 417–422. [Google Scholar] [CrossRef] [Green Version]
  77. Robelin, C.; Chartrand, P. Thermodynamic evaluation and optimization of the (NaCl+ KCl+ MgCl2+ CaCl2+ ZnCl2) system. J. Chem. Thermodyn. 2011, 43, 377–391. [Google Scholar] [CrossRef]
  78. Li, Y.S.; Spiegel, M. Models describing the degradation of FeAl and NiAl alloys induced by ZnCl2-KCl melt at 400–450 °C. Corros. Sci. 2004, 46, 2009–2023. [Google Scholar] [CrossRef]
  79. Liu, B.; Wei, X.; Wang, W.; Lu, J.; Ding, J. Corrosion behavior of Ni-based alloys in molten NaCl-CaCl2-MgCl2 eutectic salt for concentrating solar power. Sol. Energy Mater. Sol. Cells 2017, 170, 77–86. [Google Scholar] [CrossRef]
  80. Wang, J.W.; Zhang, C.Z.; Li, Z.H.; Zhou, H.X.; He, J.X.; Yu, J.C. Corrosion behavior of nickel-based superalloys in thermal storage medium of molten eutectic NaCl-MgCl2 in atmosphere. Sol. Energy Mater. Sol. Cells 2017, 164, 146–155. [Google Scholar] [CrossRef]
  81. Wang, J.W.; Zhou, H.X.; Zhang, C.Z.; Liu, W.N.; Zhao, B.Y. Influence of MgCl2 content on corrosion behavior of GH1140 in molten NaCl- MgCl2 as thermal storage medium. Sol. Energy Mater. Sol. Cells 2018, 179, 194–201. [Google Scholar] [CrossRef]
  82. Ding, W.; Bonk, A.; Bauer, T. Corrosion behavior of metallic alloys in molten chloride salts for thermal energy storage in concentrated solar power plants: A review. Front. Chem. Sci. Eng. 2018, 12, 564–576. [Google Scholar] [CrossRef] [Green Version]
  83. Dicks, A.L. Molten carbonate fuel cells. Curr. Opin. Solid State Mater. Sci. 2004, 8, 379–383. [Google Scholar] [CrossRef]
  84. Lageraaen, P.; Patel, B.; Kalb, P.; Adams, J. Treatability Studies for Polyethylene Encapsulation of INEL Low-Level Mixed Wastes. Treatability Studies for Polyethylene Encapsulation of INEL Low-Level Mixed Wastes; Final Report (No. BNL-62620); Brookhaven National Laboratory: Upton, NY, USA, 1995. [Google Scholar]
  85. Ambrosek, J.; Anderson, M.; Sridharan, K.; Allen, T. Current Status of Knowledge of the Fluoride Salt (FLiNaK) Heat Transfer. Nucl. Technol. 2009, 165, 166–173. [Google Scholar] [CrossRef]
  86. Olivares, R.I.; Chen, C.; Wright, S. The Thermal Stability of Molten Lithium–Sodium–Potassium Carbonate and the Influence of Additives on the Melting Point. J. Sol. Energy Eng. 2012, 134, 041002. [Google Scholar] [CrossRef]
  87. Chen, C.; Tran, T.; Olivares, R.; Wright, S.; Sun, S. Coupled Experimental Study and Thermodynamic Modeling of Melting Point and Thermal Stability of Li2CO3-Na2CO3-K2CO3 Based Salts. J. Sol. Energy Eng. 2014, 136, 031017. [Google Scholar] [CrossRef]
  88. Wu, Y.-T.; Ren, N.; Wang, T.; Ma, C.-F. Experimental study on optimized composition of mixed carbonate salt for sensible heat storage in solar thermal power plant. Sol. Energy 2011, 85, 1957–1966. [Google Scholar] [CrossRef]
  89. Li, X.; Wu, S.; Wang, Y.; Xie, L. Experimental investigation and thermodynamic modeling of an innovative molten salt for thermal energy storage (TES). Appl. Energy 2018, 212, 516–526. [Google Scholar] [CrossRef]
  90. Guo, S.; Zhang, J.; Wu, W.; Zhou, W. Corrosion in the molten fluoride and chloride salts and materials development for nuclear applications. Prog. Mater. Sci. 2018, 97, 448–487. [Google Scholar] [CrossRef]
  91. Frangini, S. Corrosion of metallic stack components in molten carbonates: Critical issues and recent findings. J. Power Sources 2008, 182, 462–468. [Google Scholar] [CrossRef]
  92. Kim, S.W.; Uematsu, K.; Toda, K.; Sato, M. Viscosity analysis of alkali metal carbonate molten salts at high tem-perature. J. Ceram. Soc. Jpn. 2015, 123, 355–358. [Google Scholar] [CrossRef] [Green Version]
  93. Keblinski, P.; Eastman, J.A.; Cahill, D.G. Nanofluids for thermal transport. Mater. Today 2005, 8, 36–44. [Google Scholar] [CrossRef]
  94. Shin, D.; Banerjee, D. Enhancement of specific heat capacity of high-temperature silica-nanofluids synthesized in alkali chloride salt eutectics for solar thermal-energy storage applications. Int. J. Heat Mass Transf. 2011, 54, 1064–1070. [Google Scholar] [CrossRef]
  95. Arthur, O.; Karim, M. An investigation into the thermophysical and rheological properties of nanofluids for solar thermal applications. Renew. Sustain. Energy Rev. 2016, 55, 739–755. [Google Scholar] [CrossRef] [Green Version]
  96. Awad, A.; Navarro, H.; Ding, Y.; Wen, D. Thermal-physical properties of nanoparticle-seeded nitrate molten salts. Renew. Energy 2018, 120, 275–288. [Google Scholar] [CrossRef] [Green Version]
  97. Wang, X.; Xu, X.; Choi, S.U.S. Thermal Conductivity of Nanoparticle—Fluid Mixture. J. Thermophys. Heat Transf. 1999, 13, 474–480. [Google Scholar] [CrossRef]
  98. Murshed, S.; Leong, K.; Yang, C. Investigations of thermal conductivity and viscosity of nanofluids. Int. J. Therm. Sci. 2008, 47, 560–568. [Google Scholar] [CrossRef]
  99. Thoms, M.W. Adsorption at the Nanoparticle Interface for Increased Thermal Capacity in Solar Thermal Systems. Ph.D. Thesis, Massachusetts Institute of Technology, Cambridge, MA, USA, 2012. [Google Scholar]
  100. Shin, D.; Banerjee, D. Enhanced Specific Heat of Silica Nanofluid. J. Heat Transf. 2010, 133, 024501. [Google Scholar] [CrossRef]
  101. Lu, M.-C.; Huang, C.-H. Specific heat capacity of molten salt-based alumina nanofluid. Nanoscale Res. Lett. 2013, 8, 292. [Google Scholar] [CrossRef] [Green Version]
  102. Das, S.K.; Putra, N.; Thiesen, P.; Roetzel, W. Temperature Dependence of Thermal Conductivity Enhancement for Nanofluids. J. Heat Transf. 2003, 125, 567–574. [Google Scholar] [CrossRef]
  103. Zhang, X.; Gu, H.; Fujii, M. Effective thermal conductivity and thermal diffusivity of nanofluids containing spherical and cylindrical nanoparticles. Exp. Therm. Fluid Sci. 2007, 31, 593–599. [Google Scholar] [CrossRef]
  104. Ding, Y.; Chen, H.; Wang, L.; Yang, C.-Y.; He, Y.; Yang, W.; Lee, W.P.; Zhang, L.; Huo, R. Heat Transfer Intensification Using Nanofluids. KONA Powder Part. J. 2007, 25, 23–38. [Google Scholar] [CrossRef] [Green Version]
  105. Muñoz-Sánchez, B.; Nieto-Maestre, J.; Iparraguirre-Torres, I.; García-Romero, A.; Sala-Lizarraga, J.M. Molten salt-based nanofluids as efficient heat transfer and storage materials at high temperatures. An overview of the literature. Renew. Sustain. Energy Rev. 2018, 82, 3924–3945. [Google Scholar] [CrossRef]
  106. Chieruzzi, M.; Cerritelli, G.F.; Miliozzi, A.; Kenny, J.M. Effect of nanoparticles on heat capacity of nanofluids based on molten salts as PCM for thermal energy storage. Nanoscale Res. Lett. 2013, 8, 448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Schuller, M.; Shao, Q.; Lalk, T. Experimental investigation of the specific heat of a nitrate–alumina nanofluid for solar thermal energy storage systems. Int. J. Therm. Sci. 2015, 91, 142–145. [Google Scholar] [CrossRef] [Green Version]
  108. Andreu-Cabedo, P.; Mondragon, R.; Hernandez, L.; Martinez-Cuenca, R.; Cabedo, L.; Julia, J.E. Increment of specific heat capacity of solar salt with SiO2 nanoparticles. Nanoscale Res. Lett. 2014, 9, 582. [Google Scholar] [CrossRef] [Green Version]
  109. Zhang, Y.; Li, J.; Gao, L.; Wang, M. Nitrate based nanocomposite thermal storage materials: Understanding the enhancement of thermophysical properties in thermal energy storage. Sol. Energy Mater. Sol. Cells 2020, 216, 110727. [Google Scholar] [CrossRef]
  110. Devaradjane, R.; Shin, D. Nanoparticle Dispersions on Ternary Nitrate Salts for Heat Transfer Fluid Applications in Solar Thermal Power. J. Heat Transf. 2016, 138, 051901. [Google Scholar] [CrossRef]
  111. Seo, J.; Shin, D. Enhancement of specific heat of ternary nitrate (LiNO3-NaNO3-KNO3) salt by doping with SiO2 nanoparticles for solar thermal energy storage. Micro Nano Lett. 2014, 9, 817–820. [Google Scholar] [CrossRef]
  112. Changla, S. Experimental Study of Quaternary Nitrate/Nitrite Molten Salt as Advanced Heat Transfer Fluid and Energy Storage Material in Concentrated Solar Power Plant. Ph.D. Thesis, The University of Texas, Arlington, TX, USA, 2015. [Google Scholar]
  113. Shin, D.; Banerjee, D. Enhanced Specific Heat Capacity of Molten Salt-Metal Oxide Nanofluid as Heat Transfer Fluid for Solar Thermal Applications; (No. 2010-01-1734); SAE Technical Paper; SAE: Warrendale, PA, USA, 2010. [Google Scholar]
  114. Dowling, A.W.; Zheng, T.; Zavala, V.M. Economic assessment of concentrated solar power technologies: A review. Renew. Sustain. Energy Rev. 2017, 72, 1019–1032. [Google Scholar] [CrossRef]
  115. Mohan, G.; Venkataraman, M.; Gomez-Vidal, J.; Coventry, J. Thermo-economic analysis of high-temperature sensible thermal storage with different ternary eutectic alkali and alkaline earth metal chlorides. Sol. Energy 2018, 176, 350–357. [Google Scholar] [CrossRef]
  116. Tiznobaik, H.; Shin, D. Enhanced specific heat capacity of high-temperature molten salt-based nanofluids. Int. J. Heat Mass Transf. 2013, 57, 542–548. [Google Scholar] [CrossRef]
  117. Parrado, C.; Marzo, A.; Fuentealba, E.; Fernández, A. 2050 LCOE improvement using new molten salts for thermal energy storage in CSP plants. Renew. Sustain. Energy Rev. 2016, 57, 505–514. [Google Scholar] [CrossRef]
  118. Shankar, S. Thermal Cycling Effect on the Nanoparticle Distribution and Specific Heat of a Carbonate Eutectic with Alumina Nanoparticles. Ph.D. Thesis, Texas A&M University, College Station, TX, USA, 2011. [Google Scholar]
  119. Romaní, J.; Gasia, J.; Solé, A.; Takasu, H.; Kato, Y.; Cabeza, L.F. Evaluation of energy density as performance indicator for thermal energy storage at material and system levels. Appl. Energy 2019, 235, 954–962. [Google Scholar] [CrossRef]
  120. Jo, B.; Banerjee, D. Viscosity measurements of multi-walled carbon nanotubes-based high temperature nanofluids. Mater. Lett. 2014, 122, 212–215. [Google Scholar] [CrossRef]
  121. Shin, D.; Jo, B.; Kwak, H.-E.; Banerjee, D. Investigation of High Temperature Nanofluids for Solar Thermal Power Conversion and Storage Applications. In Proceedings of the 2010 14th International Heat Transfer Conference, Washington, DC, USA, 8–13 August 2010; pp. 583–591. [Google Scholar]
  122. Turchi, C.S.; Vidal, J.; Bauer, M. Molten salt power towers operating at 600–650 °C: Salt selection and cost benefits. Sol. Energy 2018, 164, 38–46. [Google Scholar] [CrossRef]
  123. Yumashev, A.; Ślusarczyk, B.; Kondrashev, S.; Mikhaylov, A. Global Indicators of Sustainable Development: Evaluation of the Influence of the Human Development Index on Consumption and Quality of Energy. Energies 2020, 13, 2768. [Google Scholar] [CrossRef]
  124. Hrifech, S.; Agalit, H.; Bennouna, E.G.; Mimet, A. Selection methodology of potential sensible thermal energy storage materials for medium temperature applications. In MATEC Web of Conferences; EDP Sciences: Les Ulis, France, 2020; Volume 307, p. 01026. [Google Scholar]
  125. Madaeni, S.H.; Sioshansi, R.; Denholm, P. How thermal energy storage enhances the economic viability of concen-trating solar power. Proc. IEEE 2011, 100, 335–347. [Google Scholar] [CrossRef]
  126. Wagner, S.J.; Rubin, E.S. Economic implications of thermal energy storage for concentrated solar thermal power. Renew. Energy 2014, 61, 81–95. [Google Scholar] [CrossRef]
  127. González-Roubaud, E.; Pérez-Osorio, D.; Prieto, C. Review of commercial thermal energy storage in concentrated solar power plants: Steam vs. molten salts. Renew. Sustain. Energy Rev. 2017, 80, 133–148. [Google Scholar] [CrossRef]
  128. Ding, W.; Bonk, A.; Bauer, T. Molten chloride salts for next generation CSP plants: Selection of promising chloride salts & study on corrosion of alloys in molten chloride salts. In AIP Conference Proceedings; AIP Publishing LLC: Melville, NY, USA, 2019; Volume 2126, p. 200014. [Google Scholar]
  129. Williams, D.F.; Clarno, K.T. Evaluation of Salt Coolants for Reactor Applications. Nucl. Technol. 2008, 163, 330–343. [Google Scholar] [CrossRef]
Figure 1. Nitrate-based TES materials KPIs comparison (Solar Salt = 1).
Figure 1. Nitrate-based TES materials KPIs comparison (Solar Salt = 1).
Energies 14 01197 g001
Figure 2. Chloride-based TES materials KPIs comparison (Solar Salt = 1).
Figure 2. Chloride-based TES materials KPIs comparison (Solar Salt = 1).
Energies 14 01197 g002
Table 1. Composition, fusion, and decomposition temperatures for selected molten salt thermal energy storage (TES) materials.
Table 1. Composition, fusion, and decomposition temperatures for selected molten salt thermal energy storage (TES) materials.
%WeightFusion Temperature (°C)Decomposition Temperature (°C)References
Nitrate-based
Solar Salt60 NaNO3–40 KNO3240565[19]
Hitec7 NaNO3–53 KNO3–40 NaNO2142450[20]
Hitec XL15 NaNO3–43 KNO3–42 Ca(NO3)2130450[21,22]
LiNaKNO330 LiNO3–18 NaNO3–52 KNO3118550[23,24]
LiNaKCaNO315.5 LiNO3–8.2 NaNO3–54.3 KNO3–22 Ca(NO3)293450[25]
LiNaKNO3NO29 LiNO3–42.3 NaNO3–33.6 KNO3–15.1 KNO297450[26]
Chloride-based
KMgCl62.5 KCl–37.5 MgCl2430>700[27]
NaKMgCl20.5 NaCl–30.9 KCl–48.6 MgCl2383>700[27,28]
NaMgCaCl39.6 NaCl–39 MgCl2–21.4 CaCl2407650[29,30,31]
NaKZnCl7.5 NaCl–23.9 KCl–68.6 ZnCl2204>700[31,32]
KMgZnCl49.4 KCl–15.5 MgCl2–35.1 ZnCl2356>700[31,32]
Fluoride-based
LiNaKF29.2 LiF–11.7 NaF–59.1 KF454>700[33]
NaBF3 NaF–97 NaBF4385>700[34]
KBF13 KF–87 KBF4460>700[35]
KZrF32.5 KF–67.5 ZrF4420>700[34]
Carbonate-based
LiNaKCO332.1 Li2CO3–33.4 Na2CO3–34.5 K2CO3397670[36]
Table 2. Properties of selected molten salt TES materials.
Table 2. Properties of selected molten salt TES materials.
Density (kg/m3)Specific Heat Capacity (J/kg°C)References
Nitrate-based
Solar Salt2090 − 0.636T1443 + 0.172T[37]
Hitec1938 − 0.732T1560 − 0.001T[34,38]
Hitec XL2240 − 0.827T1542.3 − 0.322T[39]
LiNaKNO32088 − 0.612T1580[40]
LiNaKCaNO31993 − 0.700T1518[41,42]
LiNaKNO3NO22074 − 0.720T1135.3 + 0.071T[26]
Chloride-based
KMgCl2125.1 − 0.474T999[32,43]
NaKMgCl1899.2 − 0.4253T1023.8[27]
NaMgCaCl4020.57 − 2.7697T12,382.2 + 0.040568T^2–42.78T[29,30]
NaKZnCl2625.44 − 0.926T911.4 − 0.0227T[32,44]
KMgZnCl2169.6 − 0.5926T866.4[27]
Fluoride-based
LiNaKF2530 − 0.73T976.78 + 1.0634T[33,45]
NaBF2252.1 − 0.711T1506.0[34]
KBF2258 − 0.8026T1305.4[35]
KZrF3041.3 − 0.6453T1000[34]
Carbonate-based
LiNaKCO32270 − 0.434T1610[36]
Table 3. Specific cost and energy of selected molten salt TES materials.
Table 3. Specific cost and energy of selected molten salt TES materials.
Specific Cost
$/kg
Em
MJ/kg
Ev
MJ/m3
E
$/MJ
References
Nitrate-based
Solar Salt1.30.491901.12.65[127]
Hitec1.930.480826.94.02[127]
Hitec XL1.660.464928.13.58[127]
LiNaKNO31.10.6831285.71.61[56]
LiNaKCaNO30.70.542977.11.29[56]
LiNaKNO3NO2N/A0.408764.9N/A-
Chloride-based
KMgCl0.350.271431.31.29[48]
NaKMgCl0.220.325541.60.68[48]
NaMgCaCl0.170.289739.70.57[128]
NaKZnCl0.80.447986.61.79[48]
KMgZnCl10.298553.43.36[48]
Fluoride-based
LiNaKF20.391824.15.11[128]
NaBF4.880.474885.410.29[129]
KBF3.680.313833.311.75[129]
KZrF4.850.280750.317.32[129]
Carbonate-based
LiNaKCO32.020.44899124.15[18]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Caraballo, A.; Galán-Casado, S.; Caballero, Á.; Serena, S. Molten Salts for Sensible Thermal Energy Storage: A Review and an Energy Performance Analysis. Energies 2021, 14, 1197. https://0-doi-org.brum.beds.ac.uk/10.3390/en14041197

AMA Style

Caraballo A, Galán-Casado S, Caballero Á, Serena S. Molten Salts for Sensible Thermal Energy Storage: A Review and an Energy Performance Analysis. Energies. 2021; 14(4):1197. https://0-doi-org.brum.beds.ac.uk/10.3390/en14041197

Chicago/Turabian Style

Caraballo, Adrián, Santos Galán-Casado, Ángel Caballero, and Sara Serena. 2021. "Molten Salts for Sensible Thermal Energy Storage: A Review and an Energy Performance Analysis" Energies 14, no. 4: 1197. https://0-doi-org.brum.beds.ac.uk/10.3390/en14041197

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