You are on page 1of 16

Molten Salt Power Towers Operating at 600°C to 650°C:

Salt Selection and Cost Benefits

Craig S. Turchi,1 Judith Vidal,2 and Matthew Bauer3


1
Senior Engineer, Thermal Systems R&D Group, National Renewable Energy Laboratory, Golden, Colorado, USA, 303-384-7565,
Corresponding author: craig.turchi@nrel.gov
2
Senior Scientist, Thermal Systems R&D Group, National Renewable Energy Laboratory, Golden, Colorado, USA
3
Technology Manager, ManTech International, on contract to U.S. Department of Energy, Solar Energy Technology Office,
Washington, DC, USA

Abstract
This analysis examines the potential benefit of adopting the supercritical carbon dioxide (sCO2)
Brayton cycle at 600°C to 650°C compared to the current state-of-the-art power tower operating a
steam-Rankine cycle with solar salt at approximately 574°C. The analysis compares a molten-salt
power tower configuration using direct storage of solar salt (60:40 weight% sodium nitrate :
potassium nitrate) or single-component nitrate salts at 600°C or alternative carbonate- or chloride-
based salts at 650°C.
The increase in power cycle efficiency offered by the sCO2 Brayton cycle is expected to reduce the
size and cost of the solar field required for a given thermal energy input. Power cycle capital cost is
expected to decrease compared to the superheated steam-Rankine cycle, based on projections from
sCO2 cycle developers. Maximizing the ∆T of the storage system is required for viable deployment
of sensible-salt TES. In this regard, the partial-cooling sCO2 cycle is noted as a better option than
the recompression sCO2 cycle. In the current analysis it is assumed that a ∆T = 180K can be
achieved with the partial-cooling cycle. Even with ∆T = 180K, the potential benefits of the sCO2
Brayton cycle are partially or completely offset by increased thermal storage cost, albeit for
reasons that differ for the different salts. An approximate 5% reduction in levelized cost of energy
(LCOE) is achieved with either solar salt at 600°C or ternary magnesium chloride salt at 650°C.
The potential of using pure sodium nitrate or potassium nitrate is considered because the cold tank
temperature for the sCO2 power cycle is estimated at 420°C, which would allow use of a salt with a
higher melting point than solar salt. Sodium nitrate is the most cost effective, resulting in an overall
LCOE reduction of 8.5%; however, sodium nitrate is known to have lower thermal stability than
potassium nitrate.
The strong influence of salt cost and hot-tank cost on overall economics led to the analysis of
single-tank thermocline options. The thermocline design significantly reduces salt inventory (by
50% or more) and in many cases also reduces the tank size versus the hot salt tank of the 2-tank
system. It is speculated that integration of encapsulated phase-change material (PCM) in the
thermocline could further increase the thermal-storage energy density and reduce storage tank
volume. The thermocline cases led to three scenarios with relative LCOE reductions of
approximately 10%; however, this must be tempered by possible operational inefficiencies of the
thermocline temperature profile.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.
Keywords
Molten salt; Concentrating Solar Power; Supercritical CO2 Brayton cycle

Highlights
• sCO2 Brayton cycle can decrease the cost of concentrating solar power
• Receiver efficiency and thermal storage cost may limit the advantages of the sCO2 cycle
• Pure NaNO3 salt offers cost advantages if its thermal stability can be enhanced
• Ternary chloride is the best of the alternative salts examined for T ≥ 650°C
• Best cases had ~10% reduction in levelized cost of energy vs. current state of the art

Nomenclature
Arec receiver area
Cp salt heat capacity
F view factor of receiver surface to ambient
Tmp salt melting or liquidus point
Tmax salt maximum bulk operating temperature based on thermal stability
Trec receiver surface temperature
𝑄𝑄̇𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖 power from the solar field intercepted by the receiver
𝑄𝑄̇𝑟𝑟𝑟𝑟𝑟𝑟 thermal power delivered by the receiver to the heat transfer fluid
α receiver surface absorptance
𝜖𝜖 receiver surface emittance
ηrec receiver overall efficiency based on intercepted power from the solar field and power
delivered to the heat transfer fluid
ηrec,opt receiver optical efficiency
ηrec,th receiver thermal efficiency
𝜂𝜂̅𝑟𝑟𝑟𝑟𝑟𝑟 receiver annual average efficiency
ρ salt density

CAPEX Capital equipment expense


CSP Concentrating solar power
DOE United States Department of Energy
HTF Heat transfer fluid
LCOE Levelized cost of energy
LPPA Levelized power purchase agreement
MSPT Molten salt power tower
NREL National Renewable Energy Laboratory
PCM Phase change material

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.
SAM System Advisor Model
sCO2 Supercritical carbon dioxide
TES Thermal energy storage
USD U.S. dollars

1. Introduction
The U.S. Department of Energy launched the SunShot Initiative in 2011 with the goal of making
solar electricity cost-competitive with conventionally generated electricity by 2020. The stated
metric of this initiative is a levelized cost of energy (LCOE) for utility-scale solar power of 0.06
USD/kWh (see, for example, Mehos et al., 2016).
The state-of-the-art concentrating solar power (CSP) system is assumed to be a molten-salt power
tower employing a 60:40 weight percent blend of sodium and potassium nitrate commonly known
as “solar salt” at a hot-salt temperature of about 570°C. The SunShot goal requires an additional
cost reduction of at least 50% from the current cost of this technology in the U.S. market (Mehos et
al., 2016, IRENA, 2016). In addition to solar-field cost reductions, analysis suggests that the
SunShot goal requires development of new heat-transfer fluids (HTFs) and power systems
operating at a temperature where the net power system thermal-to-electric conversion efficiency
will reach about 50%, for example, near 700°C. A molten-salt power tower is not the only possible
path for next-generation CSP; however, the operating flexibility, energy-storage efficiency, and
industry familiarity with this design makes it a leading contender. However, evolving from 570°C
to 700°C will necessitate a new HTF to be developed, owing to solar salt’s decomposition around
600°C. Furthermore, an advanced power cycle more amenable to CSP requirements than steam-
based turbines must be employed to achieve the LCOE objective. Each technology shift will have
several consequences on the CSP system.
Deploying a new CSP technology operating at approximately 700°C entails a level of risk that
makes financing such a technology difficult. Developing the necessary technologies in a step-wise
approach—first demonstrating system concepts and power technologies at 600°C to 650°C and
later evolving to higher-efficiency systems at 700°C—offers a lower-risk path. Furthermore,
financing high-risk technologies that can approach one billion dollars is highly challenging, and
progressing toward SunShot in steps that CSP industry members can support and implement is
essential for the health of the industry and the commercial viability of newly developed
technologies.

2. Approach
This study examines the benefits of operating a molten-salt power tower with an advanced power
cycle at 600°C to 650°C—temperatures that are low enough to use the same or similar alloys to
that in current CSP plants while allowing for increases in power-cycle efficiency. The proposed
power cycle is the supercritical carbon dioxide (sCO2) recompression Brayton cycle that is the
subject of international development activity.
NREL’s System Advisor Model (SAM) is a simulation tool with technology models for various
solar and other renewable energy systems. In this analysis, SAM version 2017.01.17 was used to
model a molten-salt power tower (MSPT). SAM’s default MSPT model was modified to simulate
higher salt temperatures and power-cycle efficiencies. Physical property data for the different salts
were added to the SAM model via the user-defined HTF feature.

2.1 Salt Selection and Properties


Current parabolic trough and power tower systems use solar salt to provide thermal energy storage.
Physical properties of this salt are well documented (SQM, 2016). One limitation of solar salt is a
thermal decomposition temperature in the range of 600°C, which limits the upper temperature of

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.
power tower systems employing solar salt as the HTF and thermal storage media. A number of
alternative salts have been proposed and explored; for this analysis we focus on the salts listed in
Table 1 and Table 2.
Table 1. Properties of solar salt and alternative salts. Tmp represents the melting point or approximate
liquidus point for non-eutectic salts. Physical properties shown at approximately 600ºC unless noted.

Tmp Tmax Heat Cap. Density Relative Visc.


Salt (°C) (°C) Cp (kJ/kg-K) ρ (kg/L) ρCp (cP) Refs.
Solar Salt
238 585 1.55 1.71 1.00 1.03 SQM, 2016
(baseline)
Bauer et al.,
NaNO3 306 520 1.62a 1.82a 1.11 -
2013a
Cordaro et
KNO3 334 600 1.40a 1.78a 0.94 -
al., 2011
Mohan et al.
2017,
KCl/MgCl2 426 >800 1.03 1.94 0.75 1.88
Williams,
2006
MgCl2/NaCl/ Mohan et al.
385 >800 1.14 1.93b 0.83 -
KCl 2016
ZnCl2/NaCl/
200 >800 0.92 2.08 0.72 4.5 Li et al., 2016
KCl
K2CO3/
Na2CO3/ 398 800 1.79 2.01 1.36 10.7 An, 2016
Li2CO3
a
Approximately 450ºC. b Value taken from binary NaCl/MgCl2 salt in reference (Mohan et al., 2017).

Table 2. Composition and estimated market price of salts. Estimated costs taken from (Mehos, et al.,
2017).

Estimated salt media


Composition by weight Estimated Cost cost with ∆T=200 K
Salt fraction ($/kg) ($/kWhth)
Solar Salt (NaNO3:KNO3) 0.60 : 0.40 800 10
NaNO3 1 680 8
KNO3 1 1000 13
KCl:MgCl2 0.625 : 0.375 350 6
MgCl2:NaCl:KCl 0.550 : 0.245 : 0.205 220 3
ZnCl2:KCl:NaCl 0.606 : 0.313 : 0.081 800 18
K2CO3:Na2CO3:Li2CO3 0.345 : 0.334 : 0.321 2500 28
In addition to showing specific heat capacity, density and viscosity, Table 1 highlights the
volumetric heat capacity, ρCp, relative to the value for solar salt. Volumetric heat capacity is an
important factor in determining the volume of the storage tanks, given that tank size is inversely
proportional to ρCp∆T. The only salts with a larger volumetric heat capacity than solar salt are
sodium nitrate and the ternary carbonate.
In addition to solar salt, this analysis considers pure sodium nitrate and potassium nitrate. Despite
its higher Tmp, pure sodium nitrate is considered a possible salt for use with the sCO2 Brayton cycle
because this power cycle optimizes to a higher cold-salt temperature (~400°C) than the steam-

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.
Rankine cycle (~300°C). While exhibiting cost and heat capacity benefits versus solar salt, sodium
nitrate is predicted to have lower thermal stability versus potassium nitrate (Bauer et al., 2013a ,
Bauer et al., 2013b). Thus, the study also considers pure potassium nitrate as an alternative with
similar physical properties but greater thermal stability than solar salt.
Beyond the nitrate salts we consider chloride and carbonate salts as options that exhibit greater
thermal stability and would allow for operation at temperatures in the range of 700ºC (Mehos, et
al., 2017). Three different chloride salt combinations are considered: a binary
potassium/magnesium chloride blend, a ternary magnesium/sodium/potassium blend, and a ternary
zinc/potassium/sodium blend. The binary MgCl2 salt has been studied fairly extensively for
application as a high-temperature HTF (Williams, 2006), although data with respect to its heat
capacity as a function of temperature are lacking. The ternary MgCl2 salt promises a lower melting
point, higher heat capacity, and lower cost that make it an attractive alternative to the binary blend
(Mohan et al., 2017, Williams, 2006, Mehos et al., 2017). Lastly, the listed ZnCl2 blend is one of
several ZnCl2-based salts that have been examined for CSP application (Li et al., 2016, Jonemann,
2013). ZnCl2 salts emphasize low melting point over cost and heat capacity considerations
compared to the MgCl2-bearing salts.
The ternary carbonate is a well-known eutectic salt blend (An, 2016), as alkali carbonate salts are
used in molten carbonate fuel cells at temperatures around 650°C. In this application they are
exposed to oxygen, H2O, and CO2 as part of operation, and they are inherently less corrosive than
chloride salts under such conditions. Although corrosion is still an issue of concern, the greatest
issue with the eutectic carbonate salt blend listed in Table 2 is the cost of lithium carbonate. If
Li2CO3 remains at or near its current market price, it is unlikely that the ternary carbonate will be
an economic option despite its attractive ρCp value.

2.2 Receiver Efficiency


One of the key metrics for the analysis is the calculated or assumed annual average receiver
efficiency, 𝜂𝜂̅𝑟𝑟𝑟𝑟𝑟𝑟 . Thermal power delivered by the receiver to the HTF is given by the product of the
overall receiver efficiency and the incoming energy intercepted by the receiver. The overall
receiver efficiency is often broken into the product of its optical and thermal efficiencies (Hirsch,
2017), thus:
𝑄𝑄̇𝑟𝑟𝑟𝑟𝑟𝑟 = 𝜂𝜂𝑟𝑟𝑟𝑟𝑟𝑟 𝑄𝑄̇𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖
𝜂𝜂𝑟𝑟𝑟𝑟𝑟𝑟 = 𝜂𝜂𝑟𝑟𝑟𝑟𝑟𝑟,𝑜𝑜𝑜𝑜𝑜𝑜 𝜂𝜂𝑟𝑟𝑟𝑟𝑟𝑟,𝑡𝑡ℎ
Within SAM, 𝜂𝜂𝑟𝑟𝑟𝑟𝑟𝑟,𝑜𝑜𝑜𝑜𝑜𝑜 is given by the coating absorptance, α, on the Tower and Receiver inputs
page, which is assumed to be a constant value during the simulation. The receiver thermal
efficiency, 𝜂𝜂𝑟𝑟𝑟𝑟𝑟𝑟,𝑡𝑡ℎ , varies with each timestep in the SAM simulation depending on the receiver
conditions; however, the annual average receiver efficiency can be calculated as:
Σ[𝑅𝑅𝑅𝑅𝑅𝑅. 𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖 𝑡𝑡ℎ𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝] − Σ[𝑅𝑅𝑅𝑅𝑅𝑅. 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐. +𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙]
𝜂𝜂̅𝑟𝑟𝑟𝑟𝑟𝑟 = 𝛼𝛼 𝜂𝜂̅𝑟𝑟𝑟𝑟𝑟𝑟,𝑡𝑡ℎ = 𝛼𝛼
Σ[𝑅𝑅𝑅𝑅𝑅𝑅. 𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖 𝑡𝑡ℎ𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝]
Where the terms in brackets are SAM output variables, and the Σ indicates use of the annual sum
of this variable from SAM’s Statistics page. (Note: SAM neglects heat loss due to conduction
through the tower components.) In this analysis we report 𝜂𝜂̅𝑟𝑟𝑟𝑟𝑟𝑟 for the various cases.

2.3 Supercritical CO2 Power Cycle


The sCO2 Brayton cycle has been identified as a potential replacement to the venerable
superheated steam-Rankine cycle that is used in almost all CSP plants (Mehos et al., 2017). sCO2
Brayton cycles can have many variations, but attention has focused on the recompression cycle as
the best combination of thermal efficiency and simplicity, and this configuration has been selected
for large-scale demonstration under the U.S. DOE’s Supercritical Transformational Electric Power
(STEP) program. A slight variation referred to as the partial-cooling recompression cycle has been

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.
proposed as well-suited to CSP plants because of the ability to increase the temperature differential
across storage with relatively little impact to thermal efficiency (Neises and Turchi, 2014). sCO2
Brayton cycles have been projected to have higher efficiency and lower capital cost than
comparable superheated steam cycles at turbine inlet temperatures greater than about 550°C
(Padilla et al. 2015).
Thermal efficiency of the sCO2 recompression cycle is based on input from power system
developer Echogen Power Systems (Akron, OH). Echogen is developing sCO2 turbomachinery and
system designs for multiple applications. In this study Echogen provided simulation results for
configurations using a recompression cycle operating with 590°C or 640°C turbine inlet
temperature and design-point ambient temperature of 35°C (compressor inlet temperature of 40°C).
The corresponding gross efficiency values were 0.437 and 0.468 and the design-point cooling fan
parasitic load was estimated at 2.1% and 2.0% of gross power, respectively. Pressure losses across
each heat exchanger were estimated at approximately 1% per unit. The sCO2 power cycle was
simulated with off-design tables developed by NREL for the recompression cycle and imported
into SAM via the user-defined power cycle option. While not part of the 2017.01.17 release, this
code will be included in future releases of SAM.
The capital cost of the power cycle is based on estimates made by Echogen Power Systems and
Abengoa Solar as part of a prior study exploring the design of a 10-MWe sCO2 turbine (Turchi,
2014). Under that project the team developed cost and performance estimates for a nominal 100
MWe partial-cooling cycle operating at 600°C and 700°C with a hypothetical molten salt that had
properties comparable to solar salt. These estimates are in overall agreement with more recent
estimates made by Carlson et al. (2017), although the distribution of cost between components in
the power block differ.
In the present analysis thermal efficiency estimates are based on a recompression cycle simulation;
whereas, the turbine temperature differential (180 K) and power cycle capital cost are based on the
partial cooling cycle. Thus, we assume that the two cycle designs have similar gross efficiency,
which is supported by prior analysis (Neises and Turchi, 2014).

2.3 Relative LCOE Comparison


The primary metric of this analysis is a relative comparison of system LCOE versus that for the
baseline MSPT technology. The baseline is assumed to be represented by the default MSPT case in
SAM 2017.01.17. Input variables for that case can be found by downloading SAM and viewing the
default MSPT case. Several of the key design and performance parameters are shown in Table 3.
The financial assumptions that exist in the SAM default case are retained, with the exception that
the federal investment tax credit is set to zero for all cases. Relative LCOE is selected to normalize
the influence of financial assumptions on the analysis. The nomenclature within SAM uses
levelized power purchase agreement (LPPA) price as the metric that accounts for market factors
such as time of delivery pricing. However, in the current analysis a flat price structure is used, and
relative LCOE and relative LPPA provide the same results.
Table 3. Primary parameters in the default MSPT model in SAM 2017.01.10 that serves as the baseline
case for this study.

Turbine gross output (MWe) 115


Hot-tank/cold-tank temp. (°C) 574 / 290
Thermal storage (h) 10
Solar multiple 2.4
Power cycle gross efficiency 0.412
Cooling method Air cooled
Location (USA) Daggett, CA

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.
3. Results
The initial runs were performed starting with the default MSPT model in SAM 2017.01.17. This
created the baseline for all subsequent cases. Initially, four alternative cases were created by
importing the alternative salt physical properties, adjusting the hot and cold salt temperatures, and
applying the sCO2 Brayton cycle via a user-defined power cycle. The results of these initial cases
are summarized in Table 4.
Table 4. Evaluation of alternative molten-salt/power-cycle systems that offer the potential for higher
operating temperature.

Baseline MgCl2-sCO2-
SS-Steam-600 SS-sCO2-600 Carb-sCO2-650
MSPT 650
Na2CO3/K2CO3/ MgCl2/NaCl/
Salt Solar Salt Solar Salt Solar Salt
Li2CO3 KCl
Hot-tank temp. 574°C 600°C 600°C 650°C 650°C
Cold-tank temp. 290°C 316°C 420°C 470°C 470°C
Power cycle Steam Steam sCO2 Brayton sCO2 Brayton sCO2 Brayton
Power cycle gross
0.412 0.422 0.437 0.468 0.468
efficiency
Annual HTF
pumping parasitic 2.5% 2.4% 3.1% 2.2% 3.9%
load (% of gross)
Annual avg.
0.882 0.874 0.863 0.828 0.853
receiver efficiency
Annual net
553,000 552,000 530,000 498,000 501,000
generation (MWh)
Overnight installed
6,760 6,720 6,130 6,490 5,900
cost ($/kW)
Relative LCOE – -0.4% -4.2% +6.8% -2.1%

3.1 Influence of Receiver Efficiency


The data in Table 4 show that receiver efficiency decreases with receiver temperature as one would
expect due to the dependence of emissive and convective losses on receiver temperature. In terms
of factors that can be affected by the receiver design, optical and thermal losses from the receiver
are described by:
4
𝑓𝑓(𝜖𝜖, 𝐹𝐹, 𝑇𝑇𝑟𝑟𝑟𝑟𝑟𝑟 ) + 𝑓𝑓(𝐴𝐴𝑟𝑟𝑟𝑟𝑟𝑟 , 𝑇𝑇𝑟𝑟𝑟𝑟𝑟𝑟 )
𝜂𝜂𝑟𝑟𝑟𝑟𝑟𝑟 = 𝛼𝛼 −
𝑄𝑄̇𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖

This falloff in receiver efficiency with temperature offsets the benefits of power cycle efficiency
increase with temperature. The SunShot Initiative established a target for annual average receiver
efficiency of 90%, acknowledging that operation at higher temperature must include advances in
receiver design to mitigate the potential for greater optical and thermal losses. With this in mind,
the analysis was repeated assuming an annual average receiver efficiency of 90% is achieved for
molten-salt receivers operating at an HTF outlet temperature of 650ºC. This value is obtained
within SAM by setting the absorptance, α = 0.97, and the emittance, 𝜖𝜖 = 0.40. The absorptance
value was selected based on the highest values reported in some selective surface research (Kim et
al., 2016). The emittance value was selected to force the lowest efficiency receiver from the SAM-
default simulations (Carb-sCO2-650) equal to 0.90. While this emittance may not be realistic for
the current-design molten salt receivers, we assume that other design changes, such as higher flux

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.
limits, smaller receivers, and/or cavity-type designs are implemented to achieve this “effective
emittance” value. These emittance and absorptance values were then applied to all the SAM cases
in the “ 𝜂𝜂̅𝑟𝑟𝑟𝑟𝑟𝑟 = 0.90” scenarios. The functional dependence of thermal loss with temperature is
retained such that receivers running at lower overall temperature experience annual average
efficiencies greater than 0.90. The results of this second analysis are compared to the initial results
in Table 5.
Table 5. Evaluation of alternative MSPT systems that offer the potential for higher operating
temperature with a hypothetical coating that sets receiver efficiency to the SunShot target of 90% for
the Carb-sCO2-650 case.

Baseline SS-Steam SS-sCO2 Carb-sCO2 MgCl2- KNO3- NaNO3-


MSPT 600 600 650 sCO2 650 sCO2 600 sCO2 600
Pyromark coating (SAM default)
Rec. efficiency 0.882 0.874 0.863 0.828 0.853 - -
Relative LCOE
– -0.4% -4.2% +6.8% -2.1% - -
from Table 4
SunShot Coating, 𝜼𝜼
� 𝒓𝒓𝒓𝒓𝒓𝒓 = 90% at 650°C for carbonate salt HTF
Rec. efficiency 0.929 0.924 0.919 0.900 0.913 0.919 0.919
Relative LCOE* – -0.8% -5.1% +1.9% -4.5% -1.9% -8.5%

* assumes Baseline MSPT also benefits from the SunShot coating

3.2 Pure Nitrate Salt as HTF


The cold tank temperature in the sCO2 cycle cases is approximately 420°C, which implies that the
melting point temperatures of NaNO3 (308°C) or KNO3 (334°C) would not prohibit their use as the
HTF/TES salt. In terms of cost and physical properties, NaNO3 would be favored; however, KNO3
is known to have greater thermal stability (Bauer et al., 2013a, Bauer et al. 2013b, Cordaro et al.,
2011). Each nitrate salt was examined to determine the impact to relative LCOE. The performance
simulation uses values for KNO3 as the user-defined HTF for KNO3-sCO2-600 and for solar salt in
the NaNO3-sCO2-600 case, due to lack of available pure-salt data for all properties. Solar salt
properties should be a conservative surrogate for NaNO3. The primary difference in the case inputs
is TES cost, which ranges from 27 to 35 to 41 $/kWh for NaNO3, solar salt, and KNO3
respectively, due to the relative cost and Cp of the different salts.
The use of pure NaNO3 has a cost benefit, but it is unclear if this salt could be used at the proposed
temperature without significant degradation. While potentially more stable, the cost benefit of
using KNO3 is rather small at the assumed cost of 1000 USD/tonne. Further study of the ability to
stabilize NaNO3 (or solar salt) for operation at 600°C could enable this relatively low-risk
alternative.

3.3 Interpretation of Initial Results


Figure 1 highlights how assumptions within the alternative cases influence certain key metrics.
Solar Field CAPEX reflects the impact of power cycle efficiency, because a more efficient power
cycle will require a smaller solar field. Thermal storage CAPEX is influenced by specific salt cost,
salt volumetric heat capacity, hot-tank/cold-tank temperature differential (a function of power
cycle), and tank material-of-construction. The high cost of the carbonate salt is apparent in the
116% increase in thermal storage CAPEX for that case. On the other hand, the low volumetric heat
capacity of the chloride salt is responsible for its 43% increase in thermal storage cost due to the
increase in tank volume and cost. Power cycle CAPEX changes with cycle design and temperature.
Total Installed Cost and LCOE show the overall system impact. In these analyses, the thermal
storage CAPEX cost reduces or overwhelms the advantages brought by the sCO2 power cycle.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.
Figure 1. Relative change (versus baseline MSPT) of key metrics for the alternative cases.

The annual average receiver efficiency was raised for all cases, including the Baseline MSPT, thus
the LCOE benefit is a moving target because the Baseline MSPT also experiences a significant
boost in receiver efficiency due this assumed technology advance.
Despite operation at the same nominal temperature, the influence of receiver efficiency on the
Carb-sCO2-650 case is more pronounced than the MgCl2-sCO2-650 case. This results from the
difference in heat capacity of the two salts. The greater ρCp of the carbonate salt leads to a slower
modeled velocity within the receiver. The slower velocity reduces the heat transfer coefficient,
which leads to a higher external surface temperature on the receiver. Hence, although the nominal
HTF temperatures are equal, the model predicts that the carbonate receiver loses more thermal
energy due to its higher skin temperature. In real terms, salt properties will affect receiver
efficiency even at the same bulk salt temperature and it may be more difficult to realize the higher
receiver efficiency with the carbonate salt.
Clearly the benefit of raising the power cycle temperature is contingent on other changes within the
system. With respect to receiver efficiency, one may need to redesign the receiver to reduce
exposed surface area via use of cavity designs or attainment of higher flux limits. Improving the
selective surface coating is another path to greater receiver efficiency, albeit one that becomes
more challenging as temperature increases. The effect of internal heat transfer coefficient is also
noted as influencing the receiver external skin temperature for a given nominal HTF temperature.
HTFs with greater thermal conductivity and receiver designs with internal fins or other heat
transfer enhancements are beneficial in this regard.

3.4 Influence of Thermal Storage Cost


Figure 1 illustrates that an increasing thermal storage cost can offset much or all of the potential
benefit of moving to higher temperature. The capital costs of the different thermal storage systems
are calculated using the methodology outlined in (Mehos et al., 2017). Starting from a detailed cost
analysis of the baseline MSPT (Abengoa Solar, 2010), the amount of storage salt and the size of
the salt tanks for the alternative cases are scaled based on thermal storage ∆T and the specific and
volumetric heat capacities of the alternative salts. This scaling factor is multiplied by an alloy
multiplier that accounts for differences in the alloy cost and alloy strength as a function of
temperature. Alloy cost estimates were obtained from discussion with Haynes International, a
provider of stainless steel and high-nickel alloys. Inconel 617 or 625 is selected for temperatures
between 600°C and 650°C due to greater strength than stainless steel. The cost ratio for Inconel
625 (IN625) to 347 stainless steel (347SS) is taken as 3.25:1. The greater tensile strength of IN625
reduces the multiplier from this value to 2.44. The general results are summarized in Table 6 and
Figure 2.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.
Table 6. TES assumptions and estimated cost inputs for SAM. The alloy multipliers account for
differences in alloy cost and alloy mass (based on tensile strength) as material and temperature change
from the baseline conditions. Storage capacity is 2700 MWhth in all cases. CS=carbon steel,
SS=stainless steel, IN=inconel

Salt Tank Hot Tank Cold Tank TES System


Hot / Cold
Case inventory Volume Alloy Alloy Cost
Tank Alloy
(kg) Multiplier Multiplier Multiplier ($/kWhth)
Baseline 2.51 x 107 1 1 1 347SS / CS 24
SS-Steam-600 2.51 x 107
1 1.23 1 347SS / CS 25
SS-sCO2-600 3.86 x 107
1.539 1.23 1.11 347SS / CS 35
Carb-sCO2-650 3.30 x 107
1.144 2.44 2.30 IN625 / 347SS 59
MgCl2-sCO2-650 5.15 x 107
1.804 2.44 2.30 IN625 / 347SS 39
ZnCl2-sCO2-650 6.41 x 107
2.151 2.44 2.30 IN625 / 347SS 59
KNO3-sCO2-600 4.20 x 107
1.690 1.23 1.11 347SS / CS 41
NaNO3-sCO2-600 3.56 x 107
1.401 1.23 1.11 347SS / CS 27

Figure 2. Breakdown of TES system cost for the baseline and alternative cases. Other category
includes site work, electrical, and misc. structural steelwork.

The combination of lower storage ∆T and higher specific cost (for carbonates) or lower ρCp (for
chlorides) leads to increased TES cost versus the baseline MSPT and offsets the potential
advantages of switching to the sCO2 Brayton cycle. Given the effort that has been dedicated to salt
exploration, it is unlikely that new salt formulations will be forthcoming, so this remains an
impediment to adoption of the new power cycle or advancement to higher temperature. An
alternative path for integration of the sCO2 Brayton cycle would be use of phase-change material
(PCM) for thermal storage. Use of PCM storage relaxes the incentive for a large ∆T across storage,
which may allow for greater flexibility in power cycle optimization. Alternatives to conventional
2-tank TES are discussed next.

3.5 Thermoclines to Reduce Salt Inventory and Tank Costs


Single-tank thermocline systems have long been considered as a pathway to lower-cost thermal
storage for CSP plants, and the concept was tested at pilot-scale by Sandia National Laboratories in
the 1990s (Pacheco et al., 2002). Thermocline systems use a single tank to hold a thermally
stratified fluid, rather than having separate hot and cold tanks. The primary benefit comes from

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.
having only a single tank. Addition of a solid fill material such as loose rock or structured packing
brings the added ability to reduce the required inventory of liquid HTF. The ability to reduce salt
inventory could bring significant cost benefits to TES systems that rely on relatively expensive
salts. Examining Figure 2, we postulate that the systems that would benefit most from a
thermocline are the carbonate-salt, ZnCl2, and KNO3 cases. Elimination of the cold tank is a benefit
in all cases, although the cold tank cost is small fraction of the TES system cost. These benefits
must be weighed against operational issues associated with non-constant temperature in the exiting
salt when the thermocline zone is discharged.
In addition to replacing a large fraction of the liquid inventory, a rock fill will substitute the heat
capacity of the rock for that of the salt. In the case of nitrate salts, the rock/mineral heat capacity is
generally lower than the salt; however, for chloride-based salts, potential rock materials have
greater heat capacity than the salt, which results in the added benefit of a smaller tank, see Table 7.
Unlike liquid salts, the heat capacity of rocks and minerals generally increase with temperature
(Waples and Waples, 2004), which is an important consideration in these systems.
Table 7. Comparison of volumetric heat capacity for molten salts, three earth-abundant rocks (Waples
and Waples, 2004), and an encapsulated PCM.

Heat Capacity Density Volumetric Heat


Material
(kJ/kg-K)* (kg/m3)* Capacity (kJ/m3-K)
KNO3 (properties at 450°C) 1.40 1780 2492
ZnCl2/NaCl/KCl 0.92 2080 1907
K2CO3/Na2CO3/Li2CO3 (Carb) 1.79 2010 3597
Quartzite 1.15 2640 a
3040
Basalt 1.35 2870a 3870
Amphibolite 1.75 3010a 5250
Encapsulated salt (NaCl/KCl, Tmp=659°C) 3.39b
1590 5390
* Measured or estimated at 600°C unless noted. a 20°C; b effective Cp across 180 K

As an initial estimate we assume that a thermocline will have an overall salt volume fraction
(including solids-free space at the top and bottom of the tank) of 35%. This is consistent with a
packed-bed void fraction of 25% (Libby, 2010). The cold tank is eliminated, and the hot tank
volume is adjusted by the difference in volumetric heat capacity of the salt and the fill material, as
well as a utilization factor of 80% to account for unusable fraction of the thermocline zone. The
volumetric heat capacity of the rock will impact whether the tank size increases or decreases with
the switch to a thermocline design. Figure 3 shows how the thermocline tank volume compares to
the hot tank volume of a 2-tank system based on the ratio of volumetric heat capacities of the rock
and the salt. A tank volume ratio of less than 1.0 indicates the thermocline tank is smaller than the
corresponding 2-tank hot tank, which is a desired result. Most of the combinations show some
benefit, with the notable exception of the ternary carbonate salt (Carb-Quartzite), which has a large
volumetric heat capacity.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.
Figure 3. Relative change in the tank volume and salt mass for a thermocline and 2-tank-system hot
tank as a function of the rock and salt heat capacity ratio. SS = solar salt, others from Table 7.

EPRI’s 2010 thermocline study indicated a 20% increase in thermocline volume, which is
generated in Figure 3 (EPRI) by assuming the heat capacity of quartzite is equal to its value at room
temperature. The other data here uses rock heat capacity at 600°C as estimated by Waples and
Waples, 2004.
The overall impact to TES system cost of the switch to a single-tank thermocline is shown in
Figure 4. Each of the salt systems experiences a beneficial reduction in cost due to elimination of
the cold tank, reduced salt inventory and, in most cases, a reduction in tank volume. Rock fill cost
is taken from (Libby, 2010).

Figure 4. Breakdown of TES system cost for the baseline and alternative cases assuming the use of a
single-tank thermocline with quartzite rock fill. Compare to 2-tank designs in Figure 2.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.
Lastly, the reduced TES system costs were introduced in the SAM cases to determine the relative
LCOE, see Figure 5. Comparing Figure 5 with Figure 1 highlights the potential benefits of using a
thermocline for reducing the TES system cost. This analysis oversized the thermocline by 20% to
account for operational inefficiencies (Yang and Garimella, 2010). The greatest benefit versus the
baseline MSPT is about a 10% decrease in LCOE for each nitrate-salt case at 600°C and the
MgCl2-sCO2-650 case.

Figure 5. Relative change (versus baseline MSPT) of key metrics for the alternative cases using a new
90% efficiency receiver and single-tank thermocline with quartzite fill.

4.0 Conclusions
The current analysis examines the potential benefit of adopting the sCO2 Brayton at 600°C to
650°C as an interim step towards the U.S. DOE SunShot goals. The analysis assumes a molten-salt
power tower configuration using direct storage of solar salt or single-component nitrate salts at
600°C or alternative carbonate- or chloride-based salts at 650°C.
The increase in power cycle efficiency offered by the sCO2 Brayton cycle is expected to reduce the
size and cost of the solar field required for a given thermal energy input. Power cycle CAPEX is
expected to decrease compared to the superheated steam-Rankine cycle, based on projections from
sCO2 cycle developers. Maximizing the ∆T of the storage system is required for viable deployment
of sensible-salt TES. In this regard, the partial-cooling sCO2 cycle is noted as a better option than
the recompression Brayton cycle. In the current analysis it is assumed that a ∆T = 180K can be
achieved with the partial-cooling cycle with a negligible change in efficiency versus the
recompression cycle. Even with ∆T = 180K, the potential benefits of the sCO2 Brayton cycle are
partially or completely offset by increased thermal storage cost, albeit for reasons that differ for the
different salts. The specific cost of the carbonate salt, in particular, the lithium carbonate
component, makes this salt cost prohibitive. The ternary magnesium chloride salt has an attractive
specific cost, but its low volumetric heat capacity, necessitates a large and costly increase in tank
size.
Receiver efficiency is a key metric, and in the absence of design changes, the receiver efficiency
decreases from the default case value as temperature increases in the alternative cases.
Subsequently, the study assumes that receiver efficiency achieves the SunShot target of an overall
annual average efficiency of 0.90 for a 650°C receiver. Under this condition, an approximate 5%
reduction in LCOE is achieved with either solar salt at 600°C or MgCl2/NaCl/KCl at 650°C. This
relative comparison assumes the baseline receiver efficiency is also increased by the hypothetical
new coating or receiver design from its current SAM value of 0.882 to 0.929 (at 574°C). This study
does not stipulate how to increase receiver efficiency, but a combination of improved selective

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.
surface coating, higher flux limits, reduced receiver aperture area or reduced view factor to
ambient are suggested.
The potential of using pure sodium nitrate or potassium nitrate is considered because the cold tank
temperature for the sCO2 power cycle is estimated at 420°C, which would allow use of a salt with a
higher melting point. Sodium nitrate is the most cost effective, resulting in an overall LCOE
reduction of 8.5%; however, sodium nitrate is the less stable of the two nitrate salts. Estimates with
pure potassium nitrate result in an LCOE reduction of only 1.9% due to its higher cost and lower
heat capacity versus solar salt.
The strong influence of salt cost and hot tank cost on overall economics led to the analysis of
single-tank thermocline options. The thermocline design significantly reduces salt inventory (by
50% or more) and in many cases also reduces the tank size versus the hot salt tank of the 2-tank
system, because in most cases the volumetric heat capacity of the rock fill exceeds that of the salt,
leading to smaller tank volume. It is speculated that integration of encapsulated PCM in the
thermocline could further increase the TES energy density and reduce storage tank volume. The
thermocline cases led to three scenarios with relative LCOE reductions of approximately 10%;
however, this must be tempered by possible operational inefficiencies of the thermocline
temperature profile.

4.1 Recommendations
• Develop and test selective coatings and receiver designs that minimize thermal losses for
higher operating temperatures.
• Explore means to increase the thermal stability of solar salt or sodium nitrate to allow
operation at 600°C. The greatest cost benefit was found with use of pure sodium nitrate,
assuming thermal degradation can be controlled.
• Thermoclines remain a potential path to reduce TES costs while retaining the benefits of a
direct-storage system. In addition to reduced salt inventory, thermoclines with high
volumetric-heat-capacity fill media will decrease required tank size.
o Test rock fills with potentially higher volumetric heat capacity for compatibility
with the candidate salts. Quartzite has been shown to be compatible with nitrate
salts, but other rocks have higher volumetric heat capacity and data are needed
with chloride salts.
o Simulate system operations with an sCO2 power cycle and thermocline TES to
determine the annual performance impacts of this system design.
o Thermal ratcheting of thermocline tanks remains a concern for loose-fill designs.
This phenomenon would have to be explored before large-scale, loose-fill
thermocline tanks could be implemented.
o Model the integration of encapsulated PCMs into a thermocline design to provide
higher energy-storage density and more stable temperature outlet profiles.
Structured packing could be used to avoid thermal ratcheting while providing
PCM containment.

Acknowledgements
Work performed under U.S. Department of Energy Contract No. DE-AC36-08GO28308.

References
Mehos, M. Turchi, C., Jorgenson, J., Denholm, P., Ho, C., Armijo, K., 2016. On the Path to SunShot:
Advancing Concentrating Solar Power Technology, Performance, and Dispatchability, National
Renewable Energy Laboratory, NREL/TP-5500-65688.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.
IRENA (2016), The Power to Change: Solar and Wind Cost Reduction Potential to 2025, available at
www.irena.org/publications
Kim, T.K. et al. 2016., Copper-alloyed spinel black oxides and tandem-structured solar absorbing layers for
high-temperature concentrating solar power systems, Solar Energy 132, 257–266.
Liu M., Saman, W., Bruno. F. 2012. Review on storage materials and thermal performance enhancement
techniques for high temperature phase change thermal storage systems, Renew Sustain Energy Rev 16,
2118–32.
SQM 2016. SQM’s Thermo-Solar Salts (Salt Factsheet), Antwerpen, Belgium: SQM Europe, N.V.
(www.sqm.com).
Bauer, T., Pfleger, N., Laing, D., Steinmann, W-D., Eck, M., Kaesche, S. 2013a. High-Temperature Molten
Salts for Solar Power Application, in Molten Salts Chemistry, Elsevier.
Bauer, T., Pfleger, N., Breidenbach, N., Eck, M., Laing, D., Kaesche, S. 2013b. Material aspects of Solar
Salt for sensible heat storage, Applied Energy, 111, 1114-1119.
Cordaro, J.G., Kruizenga, A.M., Altmaier, R., Sampson, M., Nissen, A. 2011. Thermodynamic Properties of
Molten Nitrate Salts, SolarPACES 2011.
Mohan, G., Venkataraman, M., Gomez-Vidal, J., Coventry, J. 2017. “Assessment of a novel ternary eutectic
chloride salt for next generation high-temperature sensible heat storage,” in preparation.
Williams, 2006, Assessment of Candidate Molten Salt Coolants for the NGNP/NHI Heat-Transfer Loop,
ORNL/TM-2006/69, Oak Ridge, TN.
Li, P., Molina, E., Wang, K., Xu, X., Dehghani, G., Kohli, A., Hao, Q., Kassaee, M.H., Jeter, S.M., Teja,
A.S., 2016. Thermal and Transport Properties of NaCl–KCl–ZnCl2 Eutectic Salts for New Generation
High-Temperature Heat-Transfer Fluids," J. Solar Energy Engg, 138.
An, X. 2016. Determination and Evaluation of the Thermophysical Properties of an Alkali Carbonate
Eutectic Molten Salt, Faraday Discuss., 190, p. 327.
Mehos, M. Turchi, C., Vidal, J., Wagner, M., Ma, Z., Ho, C., Kolb, W., Andraka, C., Kruizenga, A., 2017.
Concentrating Solar Power Gen3 Demonstration Roadmap, National Renewable Energy Laboratory,
Golden, Colorado, USA, NREL/TP-5500-67464.
Abengoa Solar, Inc. 2010. Advanced Thermal Storage for Central Receivers with Supercritical Coolants,
DOE Contract Report under Grant DE-FG36-08GO18149.
Neises, T., Turchi, C. 2014. A comparison of supercritical carbon dioxide power cycle configurations with
an emphasis on CSP applications, Energy Procedia 49, 1187-1196.
Turchi, C., 2014. 10 MW Supercritical CO2 Turbine Test, final report under DE-EE0001589, National
Renewable Energy Laboratory, Golden CO, USA.
Pacheco, J., Showalter, S.K., Kolb, W.J. 2002. Development of a Molten-Salt Thermocline Thermal Storage
System for Parabolic Trough Plants, Journal of Solar Energy Engineering 124.
Waples, D.W., Waples, J.S. 2004. A Review and Evaluation of Specific Heat Capacities of Rocks, Minerals,
and Subsurface Fluids, Part 1: Minerals and Nonporous Rocks, Natural Resources Research, Vol. 13,
No. 2.
Libby, C. 2010. Solar Thermocline Storage Systems, Preliminary Design Study, Electric Power Research
Institute, Palo Alto, California, USA, Final Report 1019581.
Li, P., Molina, E., Wang, K., Xu, X., Dehghani, G., Kohli, A., Hao, Q., Kassaee, M.H., Jeter, S.M., Teja,
A.S., 2016. Thermal and Transport Properties of NaCl–KCl–ZnCl2 Eutectic Salts for New Generation
High-Temperature Heat-Transfer Fluids, J. Solar Energy Engineering, 138.
Hirsch, T. 2017. Editor, “SolarPACES Guideline for Bankable STE Yield Assessment,” SolarPACES.
Jonemann, M., 2013. Advanced Thermal Storage System with Novel Molten Salt, National Renewable
Energy Laboratory, Golden, CO, NREL/SR-5200-58595.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.
Yang, Z., Garimella, S.V., 2010. Thermal analysis of solar thermal energy storage in a molten-salt
thermocline, Solar Energy, 84.
Padilla, R.V., Too, Y.C.S., Benito, R., Stein, W. 2015. Exergetic analysis of supercritical CO2 Brayton cycles
integrated with solar central receivers,” Applied Energy, 148, 348-365.
Carlson, M.D., Middleton, B.M., Ho C.K., 2017. Techno-Economic Comparison of Solar-Driven SCO2
Brayton Cycles Using Component Cost Models Baselined with Vendor Data and Estimates, Paper
ES2017-3590, Proceedings of the ASME 2017 11th International Conference on Energy Sustainability,
ES2017, June 26-30, 2017, Charlotte, North Carolina, USA.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted
manuscript. The published version of the article is available from the relevant publisher.

You might also like