Sei sulla pagina 1di 8

DEVELOPMENT OF MOLTEN SALT HEAT TRANSFER FLUID WITH LOW MELTING POINT AND HIGH THERMAL STABILITY

Justin W. Raade and David Padowitz


Halotechnics, Inc. 5980 Horton St. Suite 450, Emeryville, California 94608 USA. +1 (510) 547-2634
1

Corresponding author: jraade@halotechnics.com

Abstract This paper describes an advanced heat transfer fluid (HTF) consisting of a novel mixture of inorganic salts with a low melting point and high thermal stability. These properties produce a broad operating range molten salt and enable effective thermal storage for parabolic trough concentrating solar power plants. Previous commercially available molten salt heat transfer fluids have a high melting point, typically 140 C or higher, which limits their commercial use due to the risk of freezing. The advanced HTF exploits eutectic behavior with a novel composition of materials, resulting in a low melting point of 65 C and a thermal stability limit over 500 C. The advanced HTF described in this work was developed using advanced experiment design and data analysis methods combined with a powerful high throughput experimental workflow. Over 5000 unique mixtures of inorganic salt were tested during the development process. Additional work is ongoing to fully characterize the relevant thermophysical properties of the HTF and to assess its long term performance in realistic operating conditions for concentrating solar power applications or other high temperature processes. Keywords: molten salt, eutectic mixture, heat transfer fluid, thermal storage. 1. Introduction 1.1. Background This paper describes an advanced heat transfer fluid (HTF) consisting of a mixture of inorganic salts for use in concentrating solar power (CSP) applications or other high temperature processes. Previous commercially available molten salt heat transfer fluids have a high melting point, typically 140 C or higher. The advanced HTF described in this work exploits eutectic behavior with a novel composition of materials, resulting in a low melting point of 65 C and a thermal stability limit over 500 C. Concentrating solar power uses mirrors to focus solar energy to boil water and make high pressure steam. The steam subsequently drives a turbine and generator unit to generate electricity. There is a need to bring CSP electricity cost down to the point of being competitive with traditional fossil fuel-based electricity. An advanced low melting point HTF with a high thermal stability is a key technical advance necessary to reduce the cost of CSP electricity. This novel material would enable higher temperature operation and increased efficiency in converting solar energy to electricity. Increasing the maximum fluid output temperature of current CSP plants from 390 C to 500 C would increase the conversion efficiency of the Rankine power block, thereby reducing the levelized energy cost by 2 cents/kWh. Achieving 500 C operation would also double the effectiveness of sensible heat thermal storage systems, significantly reducing the capital cost of thermal storage [1]. The state of the art HTF used in CSP applications is an organic substance that is a eutectic mixture of biphenyl and diphenyl oxide, sold under the brand names VP-1 and Dowtherm A [2], [3]. It exhibits a low melting point of 12 C but is limited to an upper temperature of 390 C due to thermal degradation. This temperature ceiling limits the thermodynamic efficiency of the Rankine cycle steam turbines driven by the heat from the solar collector field. The organic HTF exhibits a high vapor pressure at elevated temperatures, approximately 10 bar at 390 C, which prevents its use as a thermal energy storage medium. The cost of commercially available organic HTF is quite high. Molten salts exhibit many desirable heat transfer qualities at high temperatures. They have high density, high

heat capacity, high thermal stability, and very low vapor pressure even at elevated temperatures. Their viscosity is low enough for sufficient pumpability at high temperatures, and many are compatible with common stainless steels. Salts of many varieties are currently available in large commercial quantities from several suppliers. There are several commercially available salt formulations, mixtures of nitrates or nitrites sold under the brand name Hitec [4], [5]. A drawback of these molten salts as heat transfer fluids however is their relatively high melting point, typically 140-240 C. The operational risk of a freeze-up if the process temperature drops unexpectedly adds cost to systems that use salt as a heat transfer fluid. Additional hardware must be installed, such as heat tracing, insulation, or emergency water-dilution systems [6]. This high melting point limits the practicality of molten salts as heat transfer fluids in CSP applications. The advanced HTF described in this paper exhibits both a low melting point and a high thermal stability, as shown in Figure 1. Table 1 shows the relevant properties of currently available high temperature heat transfer fluids relevant for CSP applications.

Advanced HTF traditional salt synthetic oil 0 100 200 300 400 500 C 600

Figure 1: Operating range of revelant CSP heat transfer fluids.

Name VP-1 or Dowtherm A

Manufacturer Solutia or Dow Chemical Coastal Chemical

Components biphenyl diphenyl oxide sodium nitrate potassium nitrate calcium nitrate sodium nitrate potassium nitrate sodium nitrite sodium nitrate potassium nitrate

Melting point (liquidus) 12 C

Maximum temperature 400 C

Hitec XL

140 C

500 C

Hitec

Coastal Chemical

142 C

538 C

Hitec Solar Salt

Coastal Chemical

240 C

593 C

Table 1: Currently available heat transfer fluids for CSP applications. 1.2. Eutectic mixtures A eutectic mixture exhibits the lowest melting point of any similar mixture with the same components. The change in Gibbs free energy G of a substance at the melting temperature T can be expressed in terms of the change in enthalpy H and the change in entropy S.

G = H T S
At equilibrium, G = 0 and the melting temperature can be expressed as

T=

H . S

Eutectic mixtures tend to disrupt intermolecular forces (reducing the change in enthalpy) or to increase the disorder generated upon melting (increasing the change in entropy). This leads to a reduction of the melting

temperature. Eutectic behavior is common with binary mixtures of salts, and can be even more dramatic with ternary mixtures. There has been significant work done both on modeling the phase behavior of binary and ternary mixtures of salts, as well as experimentally measuring their behavior [7], [8]. An example of a simple binary salt mixture is sodium nitrate and potassium nitrate (NaNO3 and KNO3). Sodium nitrate melts at 307 C and potassium nitrate melts at 337 C. This mixture has a eutectic point at 46 mol % NaNO3 and 54 mol % KNO3 which exhibits a drastically reduced melting point of 222 C. This represents a melting point suppression of 85 C from the lowest melting single component. An example of a ternary salt mixture is sodium nitrate, potassium nitrate, and lithium nitrate (LiNO3). Lithium nitrate melts at 253 C. The ternary mixture has a eutectic point at 18 mol % NaNO3, 44.5 mol % KNO3, and 37.5 mol % LiNO3 which exhibits a melting point of 120 C. The addition of lithium nitrate to the mixture achieves an additional melting point reduction of 102 C as compared to the binary mixture. Eutectic behavior and more drastic melting point reduction occurs with more complex salt mixtures, such as quaternary or higher order mixtures. There is limited experimental data of higher order mixtures; some recent work has been done on quaternary mixtures of nitrate salts (with Li, Na, K, and Ca cations) with melting points below 100 C [9]. It is difficult to accurately model phase behavior of higher order salt mixtures. Detailed material properties of each component must be known, some of which must be measured experimentally. Existing databases of thermodynamic salt properties are incomplete, with many salts of interest (such as nitrates and nitrites) missing. It is therefore typically more straightforward to rely on experimental methods and to directly measure the phase transitions of a system of salts. The large number of possible combinations with higher order mixtures makes experimental work burdensome, since the number of possible mixtures increases exponentially with the number of components. Eutectic behavior is quite sensitive to the weight percent of each component in the mixture; a deviation of only a few percentage points may have a significant effect on the resulting melting point. If one assumes that each component in a salt mixture can be controlled to the nearest percentage point, then with a two component system there are 101 possible combinations, with a three component system there are 5151 combinations, and with a four component system there are 176,851 combinations. The permutations increase correspondingly when one varies the components to explore different salt systems. 2. Experimental methods 2.1. Materials discovery workflow In order to overcome the challenge of the large number of possible combinations of salt mixtures, the advanced HTF was developing using high throughput experimentation techniques and apparatus. A materials discovery workflow capable of screening up to 100 mixtures per day was developed, shown in Figure 2. Salt mixtures were formulated using automated robotic systems for both powder dispense and liquid dispense. The powder dispense system was the MTM Powdernium from Symyx Technologies (Sunnyvale, California). This device measures each component as it is being dispensed and records the final weight with high accuracy. It can dispense many different components to many different mixtures. The liquid dispense system was the Synthesis Station Core Module from Symyx. The melting point of each mixture was measured with the Parallel Melting Point Workstation (PMP) from Symyx. The PMP allows the melting point for each mixture in the 96 well plate to be measured simultaneously. The PMP heats the plate at a controlled rate and uses an optical method to record the temperature at which each mixture transitions from opaque to clear. This transition corresponds to the liquidus temperature, which is defined as the temperature during heating at which the last remaining solid phase melts and becomes liquid. The liquidus temperature is also equivalent to the temperature during cooling at which a solid phase first appears in the melt (assuming no supercooling). However, supercooling is common with molten salts and therefore only data acquired during a heating mode was used to obtain the melting point. The melting point in this paper is defined as the liquidus temperature. Mixtures that exhibited a low melting point were subjected to further testing for thermal stability using a

thermogravimetric analysis (TGA) device, the Q500 TGA from TA Instruments (New Castle, Delaware). A TGA heats a sample in a controlled environment and continuously measures the sample weight, which typically decreases at higher temperatures as the sample decomposes into gaseous products. 2.2. Experimental procedure A standard procedure was developed to prepare and characterize salt mixtures. The first step was to prepare free flowing anhydrous salt components. Components were purchased in reagent grade purity, typically 99% pure, from Sigma Aldrich (St. Louis, Missouri). Each component that was available in anhydrous form was ground with a mortar and pestle and dehydrated in an oven at 115 C for at least 12 hours. Calcium nitrate was procured in a tetrahydrate form and was not dehydrated prior to dispensing. Components were typically dispensed as powder, but potassium nitrite was dispensed as an aqueous solution because it is very hygroscopic and tends to form clumps when in powder form. Calcium nitrite was only available from suppliers as an aqueous solution and was dispensed in this form. The mixtures were dispensed into a borosilicate glass plate containing 96 wells in an 8 by 12 array. Each mixture had a total mass of 250 mg. After dispensing, the plate was placed in a furnace purged with nitrogen gas and heated to 400 C for at least 8 hours in order to ensure complete melting and homogenization of each mixture. Mixtures that contained hydrated components or water were first held at 150 C for at least 4 hours to boil off the water without spitting. After melting the plate was allowed to cool and stored in a dessicator before subsequent testing. The plate was then inserted into the PMP and the temperature was set to 50 C and allowed to stabilize for 30 minutes. The temperature was then ramped to 200 C at 20 C/hour.

Salt blender

Mixing station

Melting point

Thermal stability

Advanced HTF
Figure 2: Materials discovery screening workflow.

Mixtures that exhibited a low melting point were subjected to further testing for thermal stability. Approximately 20 mg of each mixture was scraped from its well in the glass plate and loaded onto a platinum pan for TGA testing. The maximum temperature or thermal stability of a sample, termed T3, was defined for screening purposes as the temperature at which it has lost 3% of its anhydrous weight during a TGA test ramping at 10 C/min. The anhydrous weight of a salt sample was defined as the weight at 300 C during the TGA test. Initial weight loss below 300 C is due to absorbed water evaporating from the sample. Each mixture was tested in two atmospheres, one of air and one of nitrogen, in order to observe the effect of oxidation. The thermal stability using the T3 method typically produces similar results for each mixture in a given system; however significant differences are observed between systems. Therefore only a representative set of mixtures from each system were tested for thermal stability rather than every mixture in the system. Mixtures with only nitrate typically have similar T3 values for air and nitrogen atmospheres. Mixtures containing nitrate/nitrite typically have a higher T3 value when tested in air. This is likely due to oxidation of the nitrite ion to nitrate, producing an increase in weight and masking the effect of weight loss due to thermal decomposition. For this reason the T3 value in nitrogen is considered more useful for nitrate/nitrite mixtures. 2.3. Experiment design methodology A rigorous experiment design methodology for high dimensional salt mixtures was devised. Over 5000 unique mixtures of inorganic salts were tested during the screening process. The literature for available systems was first reviewed in order to provide a starting point for exploring uncharted phase space. Extensive data exists for binary and ternary phase diagrams of inorganic salts [10]. Data from ternary mixtures served as a tool to predict which salts would result in the greatest melting point reduction. However, there is minimal data available for quaternary (four component), quinary (five component), and higher order systems. Custom software tools were developed in order to rapidly design constrained sets of experiments with the desired salt components. These designs were fed into the high throughput materials discovery workflow. In this manner many feasible salt types (nitrates, nitrites, carbonates, sulfates, and others) were surveyed in

mixtures containing up to six components. 2.4. Plotting experimental results with high order phase diagrams The methodology and software tools for graphing six dimensional phase data were developed. The data can include five degrees of freedom in composition (typically six salts or seven ions) plus one for the melting point (depicted by color). The phase diagram is a graphical device that allows the composition and melting point of mixtures to be represented simultaneously (this type of phase diagram is called a polythermal projection). The typical phase diagram is triangular, which allows the plotting of a ternary system of three salts (typically four ions). Each corner of the triangle represents a pure salt and the interior area represents mixtures of varying proportions. The color represents the melting point. A quaternary system of four salts (typically five ions) may be plotted by a series of triangular phase diagrams. The location of each ternary diagram along a horizontal axis represents the proportion of the 5th ion. A quinary system of five salts (typically six ions) may be plotted by a two dimensional surface of ternary phase diagrams (Phase Diagrams for Ceramists, vol. 1 discusses this technique in the introductory material, for up to six ions). Each ternary phase diagram is located at the (x, y) coordinates corresponding to the level of the 5th and 6th ions (ion 5, ion 6). A system of six salts (typically seven ions) may be plotted by a series of two dimensional surfaces of ternary phase diagrams. Each surface represents a constant value of the 7th ion. Figure 3 shows examples of these high dimensional phase diagrams.

ion 6

ion 6

Ion 5 Ion 5 ion 7

ion 5

ion 5

Figure 3: Graphical representation of six and seven ion phase space. 3. Results and discussion 3.1. Melting point The advanced HTF developed as a result of the materials discovery screening process represents the lowest melting point known for a nitrate-only salt mixture that is stable to at least 500 C. Its properties are described below, including the composition, melting point, and thermal stability. Laboratory scale quantities of salt mixtures were tested during the course of this work, however a salt mixture of any desired size with the same properties (melting point and thermal stability) can be prepared by increasing the amount of each component but maintaining the relative proportions. The data shown here was collected from a ~250 mg salt mixture: 18.6 mg of LiNO3, 14.4 mg of NaNO3, 53.4 mg of KNO3, 102.7 mg of CsNO3 and 62.1 mg of Ca(NO3)2-4H2O. Table 2 summarizes this data. A U.S. provisional patent application has been filed (No. 61/325,725) that covers the composition of the advanced HTF.

Property Melting point Thermal stability limit in air Thermal stability limit in nitrogen

Value 65 C 561 C

Method Experimental (PMP) Experimental (TGA, 3% weight loss at 10 C/min) Experimental (TGA, 3% weight loss at 10 C/min)

Notes Temperature is liquidus To verify with field testing To verify with field testing

563 C

Table 2: Selected physical properties of the advanced HTF. Figure 4 shows the composition of the eutectic as well as a phase diagram of the mixtures in the immediate neighborhood of the eutectic. This data represents the best known eutectic (lowest melting point) mixture of the given salts, discovered after a thorough search of feasible salt mixtures. Figure 5 shows an expanded phase diagram of the area further from the eutectic composition. This diagram shows four dimensions of composition plus one more dimension for liquidus temperature.

Component

Weight percent 8% 6% 23% 44% 19%

Mole percent 15% 10% 30% 30% 15% eutectic

LiNO3 NaNO3 KNO3 CsNO3 Ca(NO3)2

Figure 4: Composition of the advanced HTF (patent pending) and phase diagram showing immediate neighborhood around eutectic with the liquidus temperature in degrees celsius.

eutectic

Figure 5: Phase diagram showing expanded region around the eutectic composition with the liquidus temperature in degrees celsius. 3.2. Thermal stability The thermal stability using the T3 method is shown in Figure 6 for a mixture similar in composition to that described in Figure 4. The large weight loss below 200 C is due to the evaporation of absorbed water. This data represents a laboratory screening measurement and may not provide a definitive prediction of the HTF long-term thermal stability in a commercial application. The T3 method ranks the salt mixtures in order of relative stability rather than acting as an absolute measurement of stability. It is primarily a laboratory-scale screening test that gives a comparative ranking of candidate salt mixtures. Field testing is planned with Sandia National Laboratories and industry partners to verify long-term performance.

Figure 6: Thermal stability behavior of Li-Na-K-Cs-Ca-NO3 mixture in air and nitrogen.

3.3. Cost of components Cost is important consideration for a commercially viable heat transfer fluid. The components of the advanced HTF described in this paper are expected to have the following relative cost (in order from least expensive to most expensive): Ca(NO3)2 NaNO3 KNO3 LiNO3 CsNO3 [9]. The per unit cost of a single component is generally reduced when produced in larger quantities and at lower purity levels. The cost of raw materials for this advanced HTF is likely to be considerably higher than the simple binary solar salt (sodium and potassium nitrate). The cost may be reduced by optimizing the mixture to limit the amount of cesium and lithium while still maintaining acceptable physical properties. Such properties in addition to the melting point are important for a commercially viable heat transfer fluid, including the viscosity (affects parasitic pumping losses), the long-term thermal cycling behavior (the composition may drift after many melt/freeze cycles), and chemical compatibility with common steels (corrosion tendency). The optimal combination of these properties at an acceptable cost may result in a composition differing from that given in Figure 4. This optimization is currently underway at Halotechnics and its partners. 5. Conclusion and future work This paper presents the melting behavior and high temperature stability of a novel heat transfer fluid with a broad operating range that may be suitable for applications in solar thermal power. Future work includes an economic optimization of the HTF composition as well as full thermophysical property characterization. Scale up testing will begin with component tests and pilot scale plants before full scale commercial deployment. The materials discovery methods described in this paper can be used to develop advanced thermal materials for other heat transfer and thermal storage applications. Acknowledgements This material is based upon work supported by the U.S. Department of Energy under Award Number DEFG36-08GO18144.

References [1] G. J. Kolb, Conceptual design of an advanced trough utilizing a molten salt working fluid, presented at SolarPACES Symposium, Las Vegas, Nevada, 2008. [2] Therminol VP-1 heat transfer fluid by Solutia, Technical Bulletin 7239115B, Solutia, Inc., 1999. [3] Dowtherm A heat transfer fluid, Form No. 176-1337-397 AMS, Dow Chemical Company, 1997. [4] Hitec heat transfer salt, technical bulletin, Coastal Chemical Co., LLC. [5] Hitec solar salt, technical bulletin, Coastal Chemical Co., LLC. [6] C. R. Barkhurst and W. P. Manning, Hitec heat transfer fluid, technical whitepaper, Coastal Chemical Co., LLC, 2004. [7] P.L. Lin, A. D. Pelton, and C. W. Bale, Computation of ternary molten salt phase diagrams, J. American Ceramic Soc., vol. 62, no. 7-8, pp. 414-422, 1979. [8] G. J. Janz et al., Physical properties data compilations relevant to energy storage, NSRDS-NBS 61, parts I, II, and IV, 1981. [9] R. W. Bradshaw and N. P. Siegel, Development of molten nitrate salt mixtures for concentrating solar power systems, paper no. 11538, proc. SolarPACES, Berlin, Germany, 2009. [10] Phase Diagrams for Ceramists, American Ceramic Society/NIST, vol. 1-4, 7, 1964-1989.

Potrebbero piacerti anche