Open Access
13 September 2024 Techno-economic analysis of a solar thermophotovoltaic system for a residential building
Manish Mosalpuri, Fatima Toor, Mark Mba-Wright
Author Affiliations +
Abstract

Thermophotovoltaics (TPV) is a technology that converts heat to electricity using a thermal emitter and a matched photovoltaic (PV) cell. TPV is becoming increasingly popular due to its advantages of silent power generation, higher power density (>2.5 W/cm2), reduced cost, no moving parts (thus, low maintenance costs), reaching full power in less time as compared to turbines, operating at high temperatures, and suitability for long-duration energy storage applications. This study conducts a techno-economic analysis (TEA) of a solar energy conversion (using TPV) and storage system (using phase-change materials). We optimize the levelized cost of consumed energy (LCOE) and electricity (LCOEel) using the Nelder-Mead algorithm for four scenarios (as identified in the reference study). These scenarios differ in nominal-weighted average cost of capital (WACCnom), fuel and electricity inflation rate, and capital cost factor (CAPEX) of high-temperature energy storage (HTES), power generation unit (PGU), and PV systems. We perform a sensitivity analysis that predicts a modest decrease in LCOE and LCOEel from the mean values of $0.038/kWh and $0.128/kWh, respectively. We perform a Monte Carlo uncertainty assessment and fit a probability distribution based on input variables’ historical data from the literature. The fitted probability distribution for outputs (mean, the standard deviation in brackets) is LCOE ($/kWh)—general extreme value (0.035, 0.009), and LCOEel ($/kWh)—t (0.132, 0.016). The reduced mean values for the optimized system indicate a massive potential for TPV to be economically feasible; however, the LCOEel is higher than the current average electricity price of $0.124/kWh. The box plot shows that lifetime, PV CAPEX, inflation rate, natural gas price, and WACCnom significantly impact LCOE, and future research focused on them would lead to a better adoption of TPV technology.

1.

Introduction

The adverse effects of climate change and global warming are showing up in numerous sectors, including agriculture, decrease in crop productivity;1 food security, increase in food prices,1 food inequalities, and risks and uncertainty in local and global food systems;2 biodiversity;35 water resources;6 and extreme weather events7 and are set to worsen in the coming decades. To tackle these pressing impacts of climate change, a near-total decarbonization (carbon dioxide, CO2 emission reduction by 50% to 90%) of the energy sector is recommended.8 One of the best and the most prominent strategies to decarbonize the energy grid and limit global warming to 1.5°C (and thus reduce its harmful effects) is increasing the share of renewable energy technologies such as hydropower, solar photovoltaic (PV), and wind.9,10 Some economists argue that supporting low-carbon technologies is the second-best strategy after carbon pricing.11

Thermophotovoltaics (TPV) is an emerging decarbonization and renewable energy technology that converts heat to electricity. Figure 1 shows the TPV process schematic. It requires a heat source that can be any of these four: solar, chemical, nuclear, or electrical.1315 The heat from these sources (in the range of 1300 to 2500 K)13,16 or waste heat sources17 are incident on the absorber/emitter, which converts it into radiation. The absorber/emitter acts as an infrared (IR) bandpass filter that absorbs IR photons from the heat source as a broadband blackbody spectrum and emits filtered IR photons matching the bandgap of the PV cell for efficient power conversion. An ideal emitter also serves as a heat barrier for the PV cell, blocking sub-bandgap photons emitted from the heat source. The overall TPV system efficiency (ηTPV) is defined in Eq. (1).12,18

Eq. (1)

ηTPV=PoutQh=(SE.IQE)(VF)(FF)(CE)=  (JSC*VgQabs)*(VOCVG)*[Jmpp*VmppJSC*VOC]*(QabsQh),
where Pout is the generated electrical power, Qh is the heat flow out of the emitter surface, SE is the spectral efficiency, IQE is the internal quantum efficiency (SE and IQE combined represent the amount of absorbed heat by the cell, Qabs converting to the product of short-circuit current, Jsc and cell bandgap voltage, Vg), VF is the voltage factor (a measure of bandgap utilization, the ratio of open-circuit voltage, VOC and bandgap voltage, Vg), FF is the fill factor (ratio of generated power, Pout, the product of maximum power point voltage, Vmpp and current, Jmpp and the product of Jsc and VOC), and CE is the cavity efficiency (represents the emitter-cell combination effectiveness as a ratio of Qabs and Qh).12

Fig. 1

TPV process schematic showing heat conversion to electrical power by the emitter and a PV cell (adapted from Burger et al.12).

JPE_14_4_042404_f001.png

1.1.

Spectral Control

Spectral control refers to the employed techniques that increase the spectral efficiency by increasing the in-band radiation/photons that reach the PV cell. Spectral control in the TPV system is achieved using an IR emitter. There are two types of emitters: broadband and selective narrower band emitters. A broadband emitter emits a broad spectrum of blackbody radiation, which includes the undesired sub-bandgap photons. The sub-bandgap photons are reflected back into the cell absorber using either a back-surface reflector (BSR) (silver (Ag), gold (Au), aluminum (Al),19 etc.) or a front-surface filter (FSF) to enhance photon recycling in the PV cell, a common strategy to increase power efficiency of PV cells.20 The sub-bandgap photons’ reflection transfers the sub-bandgap light energy and prevents the PV cell from heating, which increases spectral efficiency and thus ηTPV[Eq. (1)]. Instead of using a BSR or an FSF, the emitter itself can be modified (called a selective emitter) using rare earth metals/oxides21 to radiate only the valuable part of the spectrum. Selective emitter allows using a low-temperature emitter (400 to 1400 K), lower thermal stresses, a wide range of available materials, and improved material stability.22,23 Selective emitters help reduce thermalization in PV cells but reduce the in-band power density.24 The broadband emitters have higher in-band power density but reduce the overall efficiency; thus, both the broadband and selective emitters have tradeoffs, and a suitable emitter is chosen based on the application.

1.2.

TPV Advantages

TPV is increasingly becoming popular as an energy conversion pathway due to its numerous applications, some of which include energy storage, combined heat and power25 (CHP), waste energy recovery,17,26 and in-space programs.12,27 It can be used for CHP applications as it provides electricity and heat (it can also store heat due to its compatibility with storage applications). Due to its higher power density12 (>2.5  W/cm2) and reduced cost, there has been an increasing focus on energy storage using TPV, called thermal energy electrical storage (TEES) or thermal energy grid storage (TEGS) that is competitive with electrochemical batteries.12,28 It also provides modular and silent power generation,25 is compatible with a variety of fuels/heat sources (including waste heat), and has no moving parts (thus low maintenance costs). TPV is often compared with the most widespread method of power generation nowadays, power cycles using turbines. It reaches full power in less time as compared to turbines (starting time in order of seconds as compared with tens of minutes to an hour), operates at high temperatures29 (due to its contact-less nature),30 and is suitable for long-duration energy storage applications.12,25 Due to material and technological constraints, turbines have an upper limit on the operating temperature, which is 1500°C for the Brayton cycle and 700°C for the Rankine cycle. The efficiency upper limit is closely tied to this temperature limitation, around 38% for the Brayton cycle and 23% for the Rankine cycle.27 TPV can operate at much higher temperatures (around 1700°C) and has an efficiency approaching 40%.28 New TPV thermal energy storage technologies have the potential for commercialization and deployment at a large scale.25 At the same time, turbine-based heat engines are high-risk applications due to their low efficiencies and thus require a significant capital investment.29 The literature has several new studies related to the spectral control, prototyping, modeling, economic analysis, and energy storage applications of TPV systems.31,32

1.3.

TPV Prototype and Modeling

There are a few models in the literature to predict experimental results from TPV technology and help it scale up and build a practical design for a prototype. Woolf et al.33 modeled a TPV system consisting of a selective emitter, a dielectric filter, and a PV cell. They predicted accurate photovoltaic conversion efficiency with the model. Hassan et al.34 modeled the optical (emitter-side spectral and spatial subsystem) and photovoltaic subsystems for a TPV system and observed the effects of emitter features on the optimum bandgap and overall efficiency. A recent innovative TPV study used machine learning to automatically model and optimize multilayer thin film structures to design an ideal spectral response.35 Their deep learning-based model is efficient and accurate (to predict target spectra) and reduces resource consumption compared with complex numerical models.

1.4.

TPV Economic Analyses

Some researchers have worked on evaluating the economic feasibility of TPV systems; however, it is still in its developing stages. Bianchi et al. (2012)21,36 established an energy and economic evaluation framework for a TPV system for combined heat and power (CHP) application in a residential building. They calculate parameters such as primary and electrical energy consumption, primary energy savings, cost of electricity, and money savings. They conclude that the TPV system has immense potential as it reduces purchased grid electricity and increases monetary savings. Fraas et al.37 performed a preliminary techno-economic analysis (TEA) for a TPV furnace-generator system suitable for CHP applications. It consisted of a silicon carbide (SiC) radiant tube burner surrounded by a cylindrical silicon (Si)/gallium antimonide (GaSb) PV converter array, along with an antireflection-coated tungsten foil and an infrared filter over the cells to reduce out-of-band absorption. The system’s electrical efficiency with a Si array and a standard furnace (without any furnace modifications) was found as 1.5% (and a power density of 0.3  Wcm2), and for a system with the low-bandgap GaSab cells, it was found as 15% (higher due to better spectral control, power density = 1  Wcm2). Their proposed TPV CHP unit would be economically feasible for integration into homes when the GaSb cells and arrays reach the desired low production cost. Their fuel-powered TPV system is suitable for distributed CHP applications and can be installed in homes for Winter, and renewable energy sources can be utilized in summer. Durisch and Bitnar propose a novel design of a self-powered residential electric heating system based on TPV technology.38 Their system consists of a ytterbium oxide (Yb2O3) selective emitter and thin film solar cell materials (such as Si) or copper indium gallium selenide (Cu(In,Ga)Se2) (CIGS). Thin film cells are typically 100 to 1000 times thinner than the conventional crystalline Si solar cells, thus saving extensive material and fabrication costs. Yb2O3 produces a selective emission spectrum and has a suitable bandgap to be used with Si or CIGS photocells. Their novel TPV system requires a production cost reduction to be feasible. However, it has the potential to be a basis for next-generation TPV technology to power off-grid homes and large-scale industries. The following section focuses on previous efforts in the literature on integrating energy storage and TPV and its economic feasibility.

1.5.

Energy Storage Using TPV

Renewable energy sources such as solar PV and wind have the disadvantage of intermittency: they are only sometimes available throughout the day, which creates an imbalance between the load and energy generation (especially during peak hours). Energy storage technologies help tackle this problem by storing excess renewable energy that can be used when the energy required by the load is higher than the energy generated and thus would be an integral part of future grid networks to help increase power system flexibility and stability and ensure a smooth transition for decarbonization.39,40 Electrochemical batteries (such as Li-ion) offer a cheap solution for energy storage due to the abundance of Li and other metals like cobalt and nickel used in electric vehicles and electricity storage batteries.41 Low prices are due to excess supply and low demand, which did not meet the high demand expectations after the coronavirus disease (COVID-19) pandemic in 2020. Low demand might not encourage their mining in the future, but the demand is expected to go up; thus, there are high chances of high prices of these metals (and batteries). In this paper, we consider storing renewable and grid electricity as an emerging alternative to electrochemical batteries.

Storing electricity as heat, sometimes called electrothermal energy storage (ETES), is a possible solution to decarbonize industrial heating systems and can reduce up to 40% of 2022 global natural gas use and 14% of global energy greenhouse gas (GHG) emissions.42 Many commercially available technologies for storing energy as sensible heat for industrial processes require high temperatures. Rondo Energy has designed a way to store excess renewable energy as high-temperature heat in insulated material at more than 1200°C for industrial applications.43 Using renewable energy for industrial processes would help reduce its significant GHG emissions, especially from the two processes that release most emissions—the iron and steel industry44 (ISI) and cement manufacturing.45 The automotive and clean energy company Tesla mentions in its master plan for sustainable energy that thermal batteries (with an annual energy capacity of 41 TWh) can meet global electricity and transportation storage demand, along with renewables and other batteries.46 Octopus energy uses artificial intelligence, machine learning, and its software system, Kraken, to balance power demand and supply using renewable energy storage technologies.47 PV cell thermal energy storage technologies have the potential for commercialization and deployment at a large scale. In literature, there has been a focus on a few storage mediums, such as Si/Si alloys (phase-change materials, PCMs having high-energy densities >1  MWh/m3 and low cost <4/kWh=4.31  $/kWh,25,48 graphite,49 Fe-Si-B alloy,25,50 alumina fiber board.51 For long-duration energy storage systems (10 to 100 h), it is essential to reduce the cost per unit energy (CPE) more than the cost per unit power (CPP) and round trip efficiency (RTE).25 Datas et al. analyzed a storage-integrated solar thermophotovoltaic (SISTPV) system, combining the TPV and PCMs.52 They perform a steady-state analysis for a system using PCM to store incoming heat as latent heat, which is then used to produce electricity via the TPV system. Sensible heating systems need temperatures of more than 1500°C to operate, thus more challenging insulation requirements, but not latent heat units, as they operate at a constant discharge temperature. In a subsequent paper, they performed a transient analysis for the same SISTPV system.53 They obtained promising results but mentioned that future technological improvements must be made to the model to perform an in-depth analysis. Datas et al.30 designed a thermal energy storage system using phase change materials (PCMs) that convert electricity to latent heat (stored in PCM), which is then converted to electrical energy via TPV cells. Electricity is stored as heat in the PCM, which is inefficient, but it is economical to store heat that need not be converted to electricity due to heating loads.25,49 As opposed to sensible heating storage mediums, latent heat provides the benefit of almost constant discharge temperature, which ensures the decoupling of energy storage and power generation capacity and operation at lower temperatures. They considered ideal TPV cells with 100% internal luminescent efficiency ηint (the fraction of electron-hole pairs that recombine radiatively within the semiconductor), a bandgap of 0.5 eV and BSR reflectivity of 80% and 100% (which denote the practical and ideal values, respectively). They obtained a maximum total energy density (thermal + electrical) of 1 MWh/m3 and electrical energy densities of 200 to 600  kWh/m3, comparable to the state-of-the-art lithium-ion (Li-ion) batteries. Datas et al.25 estimate the levelized cost of electricity (LCOE), CPE, and CPP for a latent heat TPV (LHTPV) system. Their latent heat TPV (LHTPV) system stores the incoming power as heat with the help of a PCM (Si or FeSiB). First, the incoming power (from the grid or solar panels) is converted to heat via a heater (ohmic/induction/microwave/arc system) using an electric switch. The heat then melts the PCM (initially a solid), and the power is stored as latent heat in it. When the TPV cells are introduced into the system, the heat is transferred to the emitter of the TPV arrangement, and the PCM starts to solidify. For a detailed system embodiment, readers are referred to the original paper.25 The PCM then converts heat to electricity as and when required with the help of a TPV cell. Amy et al.29 presented a TEGS unit consisting of integrating energy storage with multi-junction PV (MPV) cells. In their proposed system, they first convert electricity to heat via joule heating of liquid Si. The heat is stored at 2400°C in a storage tank, then converted back to electricity on demand via MPV cells. They introduce a framework to evaluate economic feasibility using cost per unit energy (CPE), power (CPP), and the RTE.29 They conclude that their TEGS system using MPV cells has the potential to be economically feasible and can enable high renewable energy penetration in the grid. Kelsall et al. modified the TEGS system presented by Amy et al. to include a multi-junction TPV cell, a solid graphite storage tank, and tin (Sn) as a heat-transfer fluid instead of liquid Si tanks.19 This replacement by graphite and Sn is because of their advantages of low melting point and low solubility of Sn in graphite. Their TEA concludes that the CPE for their system can be below the target value of 20/kWhe54 (= $21.55/kWhe) after the production capacity reaches a 100 MW target. A bottom-up TEA for highly concentrated PV modules suggests that the present PV cell cost can be reduced from a present value of 2/cm2 to 0.2/cm2 by reusing the substrate and reducing metallization costs, among other improvements.55 Thus, there is an immense potential in energy storage using TPV, as seen from literature studies. Section 2 describes our methodology to obtain our economic analysis and uncertainty assessment results.

This study performs a techno-economic analysis and uncertainty assessment (using Monte Carlo simulations) and calculates the CO2eq. emissions of energy storage using TPV. We are not aware of any literature studies performing Monte Carlo simulations for the TEA and CO2eq. emissions of a TPV system. However, Monte Carlo simulations are essential for evaluating the uncertainty in input parameters, which is essential for evaluating the robustness of emerging TPV systems. The findings from this study could inform performance improvement targets and metrics for various grid resilience applications.

2.

Methodology

We calculate and optimize the levelized cost of consumed energy (LCOE) and electricity (LCOEel) for a TPV system for a typical residential building in Boone, Iowa, based on a model built by Datas et al.56 Their energy storage system is depicted in Fig. 2. It includes three energy sources—solar PV, electricity grid, and natural gas; a boiler (running on natural gas); two heat storages: low-temperature energy storage (LTES) and high-temperature energy storage (HTES); a power generation unit (PGU, here, it is a TPV unit); and a heat pump (either electric or thermal). Their energy management algorithm is shown in Fig. 10, along with our modifications highlighted in yellow. It computes the energy going into each storage based on the electric and heat demand and generation interplay. The algorithm details are provided in Appendix A.

Fig. 2

System model showing energy sources and energy transfers to satisfy electricity consumption and heating and cooling loads (adapted from Datas et al.56).

JPE_14_4_042404_f002.png

We model this algorithm in Python (code available on GitHub57 and explanation provided in Appendix E) and minimize LCOE and LCOEel, using the Nelder-Mead algorithm by varying the power capacity of the PV (Pnom), HTES (HTESmax), and PGU (PGUmax) units. Nelder-Mead is a classic, direct search (no gradient knowledge needed) function minimization method that uses a simplex (function evaluations at different test points) and finds the following test points (the next simplex) using operations of reflection, expansion, contraction, and shrinkage.58

The inputs to the model are mentioned in Tables 1 and 2. LCOE and LCOEel are defined in Eqs. (3) and (4). Minimizing LCOEel also minimizes Pgrid and increases the amount of electricity supplied from the HTES unit, as the LCOEel’s operating expenditure (OPEX) term increases with Pgrid, as observed in Eqs. (3) and (4).

Table 1

List of important input model parameters from the reference study56 and a report by Lawrence Berkeley National Laboratory and the National Renewable Energy Laboratory.59

Input parameterValueUnit
HTES maximum capacity (optimized)80kWhth
LTES maximum capacity (optimized)80kWhth
PGU efficiency25%
PGU maximum generation capacity1.5kWel
Building annual electricity consumption7.73MWh
Building annual heat consumption39.63MWh

Table 2

List of important input model economic parameters (capital cost of LTES unit, fixed cost of grid electricity, and fixed cost of fuel obtained from the reference study56 and the rest from Monte Carlo distributions—see Fig. 3).

Input parameterValueUnit
Capital cost of PV unit914.98$kWel
Capital cost of PGU unit1184.66$kWel
Capital cost of HTES unit118.4$kWhth
Capital cost of LTES unit32.32$kWhth
Variable cost of grid electricity0.124$kWel
Fixed cost of grid electricity53.87$kWel/year
Variable cost of fuel0.015$kWhth
Fixed cost of fuel64.65$kWhth/year
Lifetime24.34years

To consider parameter uncertainty, we perform a scenario analysis (as performed in the reference study56) that calculates LCOE and LCOEel based on current and future improvements and tradeoffs in parameters such as HTES, PGU, PV capital cost, fuel (Inflfuel), and electricity (Inflel) inflation rates, and the nominal weighted average cost of capital (WACCnom). It tells how the LCOE and LCOEel would change if the present parameters change in the future due to production scale-up and other changes. The scenarios are taken from the reference study and are mentioned in Table 3, where the original prices in € are converted to $.60 From scenarios 1 to 4, the WACCnom and capital cost factor (CAPEX) values increase, and inflation rates decrease. The brackets in the first column mention an easier way of scenario representation. For example, scenario 1 is named W2I4C32,323,970 as for this scenario, WACCnom=2%, Inflfuel=Inflel=4%, and CAPEX of HTES (CAPEXHTES), PGU (CAPEXPGU), and PV (CAPEXPV) = $32/kWhth, $323/kWel, and $970/kWel, respectively.

Table 3

Scenario analysis inputs (prices converted to $ from €).

ScenarioWACCnom(%)Inflel and Inflfuel (%)CAPEXHTES($/kWhth)CAPEXPGU($/kWel)CAPEXPV($/kWel)
1 (W2I4C32,323,970)2432.325323.25969.75
2 (W2I4C108,970,1077)24107.751077.51293
3 (W3I2C108,1077,1293)32107.751077.51293
4 (W3I2C215,2155,1293)32215.521551293

Sensitivity or scenario analysis does an excellent job of performing a preliminary uncertainty analysis by considering a range of values for the uncertain parameters. However, it does not adequately represent the uncertainty as they are still supplied as a fixed data point to the model at a time (there are still uncertainties within each parameter).61,62 We thus perform a Monte Carlo uncertainty assessment to better model the uncertainty in some input parameters. The Monte Carlo method is an analysis tool that uses random numbers (generated using a probability distribution for each parameter) to calculate the output for a modeled stochastic process.63

The Monte Carlo method is already being used for advanced energy management algorithms for household systems,64 renewable energy hubs,65 and hospitals,66 to name a few. It has also been used for TPV applications since 1997 when Gethers et al.67 used it to track millions of photons from birth to death (by absorption or escape) and simulated individual probabilistic events sequentially to evaluate the TPV array’s heat flux absorption and electrical performance. Since then, it has also been used to model TPV efficiency68 and radiative energy transfer,69 among other aspects.

The steps involved in the Monte Carlo method are listed in Appendix C.

Our modifications to the original LCOE and LCOEel minimization done in Datas et al.56 are listed below:

  • a. Replacing the PGU unit with a TPV unit and using its actual efficiency values

  • b. Modifications to the original energy management algorithm, as shown in Fig. 10

  • c. Updating input parameter values obtained from the literature and using actual building energy consumption data

  • d. Providing a better estimate of LCOE and LCOEel after considering the uncertainty in input parameters via a Monte Carlo uncertainty analysis

  • e. Performing a sensitivity analysis and quantifying the total CO2eq. emissions

3.

Results and Discussion

First, we discuss the results obtained from the TEA of the TPV system for the scenarios in the reference study (scenario analysis results). We then conduct a TEA from the updated scenario inputs from the literature. We calculate the base case and optimized values for various parameters. The base-case value is the one without LCOE/LCOEel optimization and thus has a higher cost than the optimized case where the system size is varied for cost minimization.

3.1.

Scenario Analysis Results

The results for scenario 1, W2I4C32,323,970, are displayed in Table 4 based on the inputs listed in Tables 1, 2, and 3. We obtain the minimum base-case LCOE and LCOEel as $0.046/kWh and $0.17/kWh, respectively.

Table 4

Base-case and optimized inputs and results for scenario 1 (W2I4C32,323,970).

ParameterBase-case valueOptimized valueUnit
Solar PV size9.55.0kW
HTES maximum capacity40.015.0kWhth
PGU maximum generation capacity1.00.504kW
Power obtained from the grid5.134.99MW
Power supplied from PGU/TPV by stored energy in HTES0.180.31MWh
Power supplied from PV3.933.93MW
Power supplied by the system9.24MW
Heat supplied from LTES0.380.57MWh
Heat supplied from the external boiler39.2539.06MWh
Heat supplied by the system39.63MWh
Power generated from PV7.73MW
Total system losses3.312.96MWh
Self-consumption ratio57.1861.68%
LCOE0.0460.041$/kWh

From Table 4, for the base-case scenario, the power obtained from the grid = 5.13 MW (= 55% of the total power demand), from solar PV = 3.93 MW (= 43% of the total power demand), and from the PGU (by stored energy in HTES) = 0.18 MWh (= 2% of the total power demand). These three sources satisfy entirely the power demand of 9.24 MW. HTES energy losses constitute 79% of the incoming solar PV energy into HTES (=3.59 MW), and the rest is mainly passed to the PGU (20%) or stored in HTES at the end of the year (less than 1%). These numbers change for the optimized case, represented in the fourth column of Table 4. HTES energy losses decrease to 66%, and the heat passed to the PGU increases to 33%. Less than 1% of the energy is stored in HTES at the end of the year, as most of it is passed to the PGU to satisfy the electrical load. LTES supplies only 0.96% of the heating demand due to the algorithm’s priority of HTES for energy storage over LTES (as defined in the reference study56). LTES stores energy only when it is not fully charged, and the HTES is either almost entirely (>99%) or fully charged. The external boiler supplies the rest (99%) of the heating demand. These numbers improve modestly after optimizing LCOE, and the LTES supplies 1.5% of heating demand, and the external boiler supplies less energy = 98.5%. Thus, optimizing the system or searching for the least LCOE increases the energy supply from both the energy storages. Using the energy management algorithm is beneficial as it reduces the LCOE and increases solar PV’s self-consumption ratio (SCR) from 57.2% to 61.7%. Less PV energy is lost, and the consumer utilizes more.

The SCR [defined in Eq. (2)] represents the amount of PV energy generated being utilized by the load either directly as electricity or indirectly as stored energy in HTES or LTES.

Eq. (2)

Selfconsumption ratio(SCR)=1Total system lossesTotal PV power generated,
where total system losses = sum of energy losses from HTES, LTES, PGU, and PV systems.

The optimal system configuration is nominal PV capacity (Pnom)=5  kWel, HTES maximum capacity (HTESmax)=15  kWhth, and PGU maximum capacity (PGUmax)=0.504  kWel. The base-case system configuration (Pnom=9.5  kWel, HTESmax=40  kWhth, and PGUmax=1.0  kW) converges to the minimum bounds for two optimized variables and is very close to the minimum for the third one, as seen in Table 5.

Table 5

Scenario analysis results showing optimized system size (nominal PV capacity, HTES max. Capacity, and PGU max. generation capacity) and optimum LCOE and LCOEel values.

ScenarioOptimized system size = [nominal PV capacity (kWel), HTES max. capacity (kWhth), PGU max. generation capacity (kWel)]Optimum values
LCOE minimizationLCOEel minimizationLCOELCOEel
1 (W2I4C32,323,970)[5, 15, 0.504][5, 15, 0.519]0.04030.144
2 (W2I4C108,970,1077)[5, 15,0.5][5, 15, 0.5]0.04200.152
3 (W3I2C108,1077,1293)[5, 15, 0.5][5, 15, 0.5]0.0360.139
4 (W3I2C215,2155,1293)[5, 15, 0.5][5, 15, 0.5]0.0380.149

We observe that the optimal solution for the economically favorable scenarios (having lower WACCnom and CAPEX values) converges to a somewhat larger PGU system, with its size being 0.504  kWel for scenario 1 (W2I4C32,323,970). It obtains a lower Pgrid(=4.99  MWh), whereas small systems (nominal PV capacity = 5  kWel, HTES max. capacity = 15  kWhth, and PGU max. generation capacity = 0.5  kWel) are obtained for unfavorable economic conditions, including high WACCnom and high capital cost of HTES, PGU, and PV systems (their values mentioned in Table 5, column 1). Table 5 presents the scenario analysis results—the optimum system size (HTES, PGU, and nominal PV capacity) and the optimum LCOE and LCOEel values. The base case and optimum LCOE and LCOEel values are also shown in Fig. 3.

Fig. 3

Base-case and optimum (a) LCOE and (b) LCOEel variation with scenarios.

JPE_14_4_042404_f003.png

Unfavorable economic conditions produce a high optimum Pgrid (=5 MWh for scenario 4:W3I2C215,2155,1293). As economic conditions become favorable (low WACCnom and low capital cost), we obtain a large PGU system, which becomes favorable to accommodate a large system (a large system is needed to store the produced PV energy). The same is true for LCOEel minimization—scenario 1: W2I4C32,323,970 (having the least WACCnom and CAPEX) produces a system with PGU size = 0.519  kWel. In comparison, all the rest scenarios converge on the lower bounds and give the result as nominal PV capacity = 5  kWel, HTES max. capacity = 15  kWhth, and PGU max. generation capacity = 0.5  kWel (a small system). This phenomenon is illustrated in Appendix D, and an explanation is provided in Table 6.

Table 6

Mean parameter values, as obtained from the Monte Carlo simulation fitted probability distributions.

ParameterMean value
Lifetime (years)24.34
WACCnom (%)2.17
PGU efficiency (%)27.45
Electricity price ($/kWh)0.124
CAPEXPGU ($/kW)1184.66
CAPEXPV ($/kW)914.98
CAPEXHTES ($/kWhth)118.4
Inflation rate (%)2.31
Natural gas price ($/kWh)0.015

The optimized LCOE and LCOEel values mentioned in Table 5 seem counterintuitive: their values for the favorable scenario are higher than some unfavorable ones. This is due to the higher value of Inflfuel and Inflel for favorable cases, as assumed in the reference study. The LCOE equation [Eq. (3)] has two terms involving inflation rates, and LCOEel [Eq. (4)] just has one. This is why the effect is less pronounced for the LCOEel optimization case and more for LCOE. Scenario 3 (W3I2C108,1077,1293) produces the least optimized LCOE and LCOEel, as it has the right tradeoff between WACCnom and inflation rates.

3.2.

Sensitivity Analysis and Monte Carlo Risk Assessment Results

Based on historical data from the literature, the fitted distribution for the non-triangular distributed parameters—natural gas price is a general extreme value, Inflel and Inflfuel is t; CAPEXPV70 is a general extreme value; and electricity price is Weibull distribution, as shown in Fig. 4, along with their 10 and 90% confidence interval limits.

Fig. 4

Fitted probability distributions and 90% confidence limits for Monte Carlo parameters. Their distributions, mean, and standard deviation are as follows: (a) Lifetime (triangular distribution: 24.34 years, 2.47 years), (b) WACCnom (triangular distribution: 2.17%, 0.31%), (c) PGU efficiency (triangular distribution: 27.45%, 5.08%), (d) electricity price (Weibull distribution: $0.124/kWh, $0.015/kWh), (e) CAPEXPGU (triangular distribution: $1184.66/kW, $373.18/kW), (f) CAPEXPV (general extreme value distribution: $914.98/kW, $145.41/kW), (g) CAPEXHTES (triangular distribution: $118.40/kWh, $37.60/kWh), (h) inflation rates—Inflfuel and Inflel (t distribution: 2.31%, 2.11%), and (i) natural gas price (general extreme value distribution: $0.015/kWh, $0.011/kWh).

JPE_14_4_042404_f004.png

Now, we mention the results from the scenario analysis we used for our study. First, the base-case and optimized LCOE and LCOEel for a mean scenario are illustrated in Fig. 5. For this scenario, we chose the mean parameter values (listed in Table 6 and Fig. 4) from the fitted Monte Carlo distributions. The values in Table 6 differ from the ones used in the four scenario analyses previously as those values were obtained from the reference study,56 and these updated values are obtained from the fitted probability distributions.

Fig. 5

Mean scenario (obtained from Monte Carlo results) analysis results for this study. The number on top of the bar denotes the total LCOEel/LCOE, and the one below denotes the capital cost. (a) Base-case LCOEel=$0.165/kWh. (b) Optimized LCOEel=$0.129/kWh. (c) Base-case LCOE = $0.045/kWh. (d) Optimized LCOE = $0.038/kWh.

JPE_14_4_042404_f005.png

LCOE and LCOEel are defined in Eqs. (3) and (4). LCOE represents the cost of supplying both heat and electricity; however, LCOEel only takes into account electricity costs. LCOEel was defined in the reference study as an alternative to LCOE to lower the payback period (LCOE minimization converges to a larger system, thus high investment and longer payback period) at the cost of obtaining a higher value for LCOEel. For our study, Fig. 5 shows the base-case LCOEel is $0.165/kWh, which is reduced by 22% to $0.129/kWh on optimization. The optimized value is very close to the average price of electricity obtained in Monte Carlo simulations (Fig. 4) = $0.124/kWh. Energy storage using TPV can thus be profitable in the future. To assess future scenarios, we performed a sensitivity analysis and Monte Carlo simulations on the optimized system, which reduced LCOE and LCOEel, and the results are mentioned in the next section.

LCOE has a much smaller value of $0.045/kWh and optimizes to $0.038/kWh. For both LCOE and LCOEel, the optimum system converges to the smallest system size: nominal PV capacity = 5 kWel, HTES max. capacity = 15  kWhth, and PGU max. generation capacity kWel=0.5  kW. Based on this optimized system, we perform a sensitivity analysis based on the same parameters we considered for the uncertainty assessment. The results are displayed in Fig. 6 for parameters with values ±25% from the mean values listed in Table 6. The parameters having the most significant impact on LCOE and LCOEel are Inflfuel and Inflel, electricity price, CAPEXPV, lifetime, and natural gas price (only for LCOE), with a potential cost reduction of around 10% by the most dominant parameter (natural gas price for LCOE and Inflfuel and Inflel for LCOEel). LCOEel does not change with natural gas prices because it only considers electricity, not heating requirements.

Fig. 6

Sensitivity analysis results for (a) LCOE and (b) LCOEel with parameter values ±25% from the mean values listed in Table 6.

JPE_14_4_042404_f006.png

Figure 7 shows the Monte Carlo uncertainty assessment results for LCOE and LCOEel (for an optimized system: nominal PV capacity = 5  kWel, HTES max. capacity = 15  kWhth, PGU max. generation capacity kWel=0.5  kW). We depict the confidence intervals for our large sample analysis and do not rely on the p-values as they offer little insight for large samples.71 The fitted distribution for LCOE is a general extreme value distribution, with a mean of $0.035/kWh and a standard deviation of $0.009/kWh, and that for LCOEel is a t distribution, with a mean of $0.132/kWh and a standard deviation of $0.016/kWh. The previously obtained optimized LCOE and LCOEel values listed in Table 5 (scenario analysis results) are within the 10% and 90% confidence limits. Thus, we obtained similar cost numbers to those in reference study56 even after updating the assumptions and input values. The optimized LCOE from Fig. 6 is $0.033/kWh, close to the mean ($0.035/kWh) and median ($0.039/kWh) values obtained here, and is much less than the electricity price mean ($0.124/kWh), but is higher than the natural gas price mean ($0.015/kWh). LCOE represents the cost of consumed energy (heat and electricity) and thus is compared with both mean values. Adding storage units and PV energy sources increases the price of the supplied energy, but it can be reduced, as shown in sensitivity analysis and uncertainty assessment results.

Fig. 7

Monte-Carlo risk uncertainty assessment results (probability density function, cumulative distribution function, 10% and 90% confidence limits) for (a) LCOE: mean = 0.035, median = 0.039, standard deviation = 0.009, 10% and 90% confidence limits = (0.028,0.064) (all in $/kWh) and (b) LCOEel: mean = 0.132, median = 0.132, standard deviation = 0.016, 10% and 90% confidence limits = (0.098, 0.165) (all in $/kWh).

JPE_14_4_042404_f007.png

The optimized LCOEel, as mentioned in Fig. 5, is $0.129/kWh. The mean value calculated from the uncertainty analysis is close, $0.132/kWh. Figure 7 depicts other statistical parameters such as standard deviation = $0.016/kWh, 10%, and 90% confidence limits = (0.098, 0.165) $/kWh. It also shows the probability density function and the cumulative distribution function for LCOE and LCOEel.

Box plots in Fig. 8 show the effects of each parameter’s uncertainty on LCOE and LCOEel. The boxes’ left and right edges denote the 0.25 and 0.75 quartiles of the distributions (the width of the box is termed inter-quartile region, IQR), the dotted line inside the box indicates the median, and the whiskers and caps at the ends indicate the region from the box to the outermost data point within 1.5 times the IQR. Figure 8 illustrates that lifetime, CAPEXPV, inflation rates, natural gas, and electricity prices have the most considerable impact on LCOE (the same parameters for sensitivity analysis). LCOEel has all the same dominant parameters (except natural gas price). A way to reduce LCOE and LCOEel would be to improve these parameters, as the boxes primarily lean towards the left (which implies lower cost). Thus, an improvement would lead to favorable economic results. The slightest uncertainty is observed in CAPEXPGU and PGU efficiency; in other words, their variation does not have much effect on LCOE and LCOEel.

Fig. 8

(a) LCOE and (b) LCOEel box plot representing parameter uncertainty.

JPE_14_4_042404_f008.png

3.3.

CO2eq. Emission Results

Based on the results in Fig. 5, we calculate the CO2eq. emissions from the energy system before and after optimization using emission factors listed in Table 7. Here, we use the emission factors for natural gas, grid electricity, and PV electricity and assume negligible emissions from the HTES, LTES, and PGU/TPV units due to a lack of literature data. The results are shown in Fig. 9. We observe that before using energy storage (using natural gas for heating and grid electricity), natural gas CO2eq. emissions are the highest (=7.166 tons/year) and reduce modestly for the base case and optimized cases (both use energy storage). The total CO2eq. emissions for the system with no energy storage are 9.98 tons/year. The total emissions reduce for the base case (b) = 8.824 tons/year due to lesser dependence on natural gas and grid electricity and more utilization of cleaner solar PV energy to supply heat and electricity. The emissions are the lowest for the optimized case (c) (=8.75 tons/year) due to an optimized system with less dependence on grid electricity and an external boiler. Thus, according to our results, utilizing the energy management algorithm for energy storage and optimizing the system reduces CO2eq. emissions by a modest amount (12%), primarily due to less dependence on grid electricity. Renewable energy integration in future grids would reduce their environmental impact; thus, TPV energy storage inclusion might not significantly reduce emissions. However, a detailed life cycle assessment must be performed to understand the process emissions better.

Table 7

Emission factors for natural gas, grid electricity, and PV electricity.

ParameterValue
Natural gas emission factor (gCO2eq/kWhth)180.7872
Grid electricity emission factor (gCO2eq/kWhel)304.1473
PV electricity emission factor (gCO2eq/kWhel)43.0074

Fig. 9

CO2eq. emissions in tons/year for (a) without energy storage (heat and electricity obtained from an external boiler and the grid, respectively), (b) base case, and (c) optimized case. For each column, the numbers represent the emissions from natural gas, grid electricity, and solar PV electricity (going from bottom to top).

JPE_14_4_042404_f009.png

4.

Conclusions and Future Work

Decarbonization using renewable energy is vital to achieving the environmental targets set by the IPCC to keep global warming below 1.5°C. This paper performs a techno-economic analysis and uncertainty assessment for a renewable energy technology, thermophotovoltaics (TPV). TPV research is still in the beginning stages, with no working large-scale prototypes. Our analysis shows that TPV still needs to be economically feasible based on previous scenarios from the literature and updated inputs from the results of Monte Carlo simulations, but it has immense potential. According to the authors’ knowledge, this is the first study that performs an uncertainty assessment of the TPV technology. The critical results, conclusions, and future work from the study are listed below:

  • 1. From the scenario analysis (based on reference study data), we obtain a minimum LCOE of $0.036/kWh and a minimum LCOEel of $0.139/kWh. These values are for the scenario W3I2C108,1077,1293, where WACCnom=3%, Inflfuel=Inflel=2%, CAPEXHTES=$108/kWhth, CAPEXPGU=$1077/kWel, and CAPEXPV=$1293/kWel.

  • 2. For the mean scenario (obtained from Monte Carlo results), base-case LCOEel=$0.165/kWh, optimized LCOEel=$0.129/kWh; base-case LCOE = $0.045/kWh, and optimized LCOE = $0.038/kWh. For both LCOE and LCOEel, the optimum system converges to the smallest system size: nominal PV capacity = 5 kWel, HTES max. capacity = 15 kWhth, PGU max. generation capacity kWel=0.5  kW.

  • 3. The sensitivity analysis results show that the parameters having the most significant impact on LCOE and LCOEel are inflation rates (Inflfuel and Inflel), electricity price, CAPEXPV, lifetime, and natural gas price (only for LCOE), with a potential cost reduction of around 10% by the most dominant parameter (natural gas price for LCOE and Inflfuel and Inflel for LCOEel).

  • 4. According to the Monte Carlo simulations, the fitted distribution for LCOE is a general extreme value distribution, with a mean of $0.035/kWh and a standard deviation of $0.009/kWh, and that for LCOEel is a t distribution, with a mean of $0.132/kWh and a standard deviation of $0.016/kWh.

  • 5. Box plot results show that the lifetime, CAPEXPV, inflation rates, natural gas price, and electricity price significantly impact LCOE (the same parameters for sensitivity analysis). LCOEel has all the same dominant parameters (except natural gas price). A way to reduce LCOE and LCOEel would be to improve these parameters; thus, there is an immense potential for economic feasibility for TPV in the future.

  • 6. According to the CO2eq. emissions results, the total CO2eq. emissions for the system with no energy storage are 9.98 tons/year, for the base case are 8.824 tons/year, and for the optimized case are 8.75 tons/year. Utilizing the energy management algorithm for energy storage and optimizing the system reduces CO2eq. emissions by a modest amount (12%) due to lesser dependence on natural gas for heating and grid electricity and higher utilization of renewable solar PV energy.

5.

Appendix A: Energy Management Algorithm

The energy management algorithm is shown in Fig. 10, along with our modifications highlighted in yellow. It computes the energy going into each storage based on the electric and heat demand and generation interplay.

Fig. 10

Energy management algorithm from the reference study,56 with our modifications highlighted in yellow.

JPE_14_4_042404_f010.png

The algorithm by Datas et al.56 follows this approach (which requires an input of power consumed by the load and that produced by the solar PV system per hour): first, it computes if there is excess electricity produced by the solar PV system [Pg(t)] compared with the electricity consumption [Pc(t)]. If yes, the excess energy is stored in HTES (which stores heat at high temperature to supply to the PGU unit) if it is not fully charged; if it is fully charged and there is a heating load, it is stored in LTES (which stores heat at low temperature and supplies it for space and domestic hot water heating); if even this is full, this energy is lost.

We modify their algorithm as the energy storage terms (EHTES and ELTES) for the (t+1)th interval turn negative when their respective loss terms (Qloss,HTES and Qloss,LTES) begin to dominate in the mass balance equation. In this modification, we make these assumptions:

  • 1. When the loss term starts to dominate, it is possible to cut off the respective energy storage unit. When Qloss,LTES dominates, we shut off the LTES unit for that iteration (all the inputs and outputs are turned down to zero), the stored energy, ELTES (t), is carried forward to the (t+1)th interval, and the heat loads are supplied energy from the external boiler. Similarly, for the HTES unit, when Qloss,HTES dominates, we shut it off for that iteration (all the inputs and outputs are turned down to zero). The stored energy, EHTES (t), is carried forward to the (t + 1)th interval, and the required electricity for that interval is supplied from the grid (= Pgrid). There were a few instances when Pgrid turned out negative because electricity generation was more significant than consumption. Solar PV supplies all the electricity in that case, and Pgrid is reduced to 0.

  • 2. When the heat output from LTES (Qout,LTES) is calculated as negative (due to the large value of Eloss,LTES), it is possible to obtain the required heat energy from the external boiler and not use LTES for that iteration.

The modifications we made are for the building energy values we considered, thus might not work very well for other buildings. We suggest that researchers first use the original code from the reference study and then decide whether our modifications are required for their building. A benefit of using the modifications is that we would not obtain negative values for stored energy if the loss terms dominate. However, a limitation is that cutting off the storage units or connecting them back to the system in real time might be challenging.

6.

Appendix B: LCOE, LCOEel, and Operating Expenditure (OPEX) Equations

LCOE and LCOEel are defined in Eqs. (3) and (4).

Eq. (3)

LCOE=CAPEX+t=1T[OPEX(t)(1+WACCnom)t]t=1T[t=18760(ce(t)+ch(t))Δt(1+WACCreal)t],
where
OPEX(t)=(1+Inflel)t{OPEXelec,fix*max[Pgrid(t)+OPEXelec,var*t=18760Pgrid(t)Δt}+(1+Inflfuel)t{OPEXfuel,fix*+OPEXfuel,var*t=18760Qext(t)Δt}

Eq. (4)

LCOEel=CAPEX+t=1T[OPEXel(t)(1+WACCnom)t]t=1T[t=18760ce(t)Δt(1+WACCreal)t],
where
OPEXel(t)=(1+Inflel)t{OPEXelec,fix*max[Pgrid(t)+OPEXelec,var*t=18760Pgrid(t)Δt}
where CAPEX is equal to the total capital expenditure of the system, Ce(t) is the electricity consumption, Ch(t) is heat consumption, WACCreal is the real weighted average cost of capital, OPEX*elec,fix, OPEX*elec,var, OPEX*fuel,fix, and OPEX*fuel,var represent the fixed and variable cost factors for fuel and electricity, and Qext(t) is the heat obtained from the external boiler.

7.

Appendix C: Monte Carlo Uncertainty Assessment Methodology

The steps involved in the Monte Carlo method are listed below:

  • 1. The first step is identifying the key parameters to participate in the method. We use the same parameters as we did for the scenario analysis. We also consider these additional parameters—natural gas price, electricity price, PGU/TPV efficiency, and project lifetime.

  • 2. The second step is identifying a best-fit probability distribution for each parameter. We fit a triangular distribution for WACCnom, CAPEXHTES, CAPEXPGU, PGU efficiency, and lifetime. We specify their lowest, most probable (mode), and highest values to generate a triangular distribution, which attempts to cover the range of values. For other parameters (natural gas price, Inflel, and Inflfuel, CAPEXPV, and electricity price), we obtain their historical data from the literature, and the distfit python package v1.4.0 identifies and fits their probability distributions (shown in Fig. 6 of the main text).75 Here, we assume the historical parameters’ variation would guide future variation in the form of probability distributions.

  • 3. The next step is to generate random numbers (5000 for our study) from the specified distributions for all the parameters and supply them, one at a time, to the Python algorithm model, which generates an empirical distribution for the output variables (LCOE and LCOEel). A data frame stores the LCOE and LCOEel values for each parameter input from the random data set. Here, we choose the number of samples as 5000 as the results (mean and standard deviation) converged for numbers greater than it (tolerance limit = 103). We ran each simulation ten times.

  • 4. The last step is to analyze the mean, standard deviation, significance, and uncertainty levels of LCOE and LCOEel.

8.

Appendix D: Scenario Analysis Results

Figure 11 shows that the smallest bounds have the lowest LCOE for scenario 4: W3I2C215,2155,1293, whereas for scenario 1, W2I4C32,323,970, the optimum is not always at the smallest system size. The PGU maximum capacity has an optimum LCOE at around 1 kWel—favorable economic conditions can accommodate a larger system with lower cost. This is the reason we obtain larger systems for favorable economic conditions. The variation of LCOE with PGU max. generation capacity for scenario 1 shows a discontinuous variation for the most part. It is due to the direct effect of PGU capacity on Pgrid (see the code), which affects the operating expenditure—an exponent function (summed over the entire lifetime of the project) to calculate LCOE (the non-linear variation of operating expenditure dominates the linear variation of capital expenditure). PGU capacity thus updates Pgrid and the power output from PGU at each time interval (and is responsible for LCOE fluctuations), which is not valid for the other two variables plotted in Fig. 5 (main text)—PV and HTES capacities are not directly involved in the calculation of the operating expenditure (an exponent function); they have a linear relationship with capital expenditure, and thus, they have a continuous variation. Scenario 1 variation of LCOE with PGU capacity is discontinuous. In contrast, scenario 4 is smooth and continuous as it has higher capital expenditure (and low-operating expenditure), which dominates LCOE and produces a linear variation.

Fig. 11

LCOE ($/kWh) variation with economic scenarios and system size.

JPE_14_4_042404_f011.png

9.

Appendix E: GitHub Code

Our code allows anyone with their own input parameters to calculate the optimized LCOE and LCOEel for their building’s energy storage system. Readers who wish to use our code as a template are encouraged to go through the comments in the code for a better understanding.

In the GitHub link,57 follow these instructions:

  • 1. The file “Optimization_and_CO2_emissions.py” contains the code to calculate and optimize the levelized cost of energy (LCOE) and electricity (LCOEel) by following the energy management algorithm from the original study (ref: https://doi.org/10.1016/j.apenergy.2019.113935) along with our modifications according to our needs. It imports the Excel file “TPV.xlsx” that contains the hourly electricity consumption and solar PV generation data for a residential building in Boone, Iowa. The Python file also plots the base case and optimized LCOE and LCOEel. It also calculates the CO2eq. emissions for a building with no energy storage system and base case and optimized energy storage systems with a TPV unit.

  • 2. The file “Monte_Carlo_uncertainty_analysis.py” performs the Monte Carlo uncertainty assessment on some input parameters by importing their historical data from the literature (from the Excel file “Monte_Carlo_parameters.xlsx”). It then fits a probability distribution to each parameter and supplies their values to the model to calculate LCOE. Then, it plots the probability distribution and cumulative distribution functions for LCOE (for an optimized system).

Disclosures

We have no relevant financial interests in the paper and no other potential conflicts of interest to disclose.

Code and Data Availability

The data presented in this article are publicly available in the ThermoPhotoVoltaics repository at https://zenodo.org/doi/10.5281/zenodo.10988244

Acknowledgments

The authors gratefully acknowledge the funding for the work by the Decarb 2040 Community Feasibility Grant Program.

References

1. 

A. Calzadilla et al., “Climate change impacts on global agriculture,” Clim. Change, 120 (1–2), 357 –374 https://doi.org/10.1007/s10584-013-0822-4 CLCHDX 0165-0009 (2013). Google Scholar

2. 

T. Wheeler, “Climate change impacts on global food security,” Science, 341 508 –513 https://doi.org/10.1126/science.1239402 SCIEAS 0036-8075 (2013). Google Scholar

3. 

S. C. Doney et al., “Climate change impacts on marine ecosystems,” Annu. Rev. Mar. Sci., 4 (1), 11 –37 https://doi.org/10.1146/annurev-marine-041911-111611 1941-1405 (2012). Google Scholar

4. 

C. Bellard et al., “Impacts of climate change on the future of biodiversity,” Ecol. Lett., 15 (4), 365 –377 https://doi.org/10.1111/j.1461-0248.2011.01736.x ECLEFU 1461-023X (2012). Google Scholar

5. 

J. Gross, “Coral reefs are in trouble. One lab in the desert is trying to help,” The New York Times, (19 December 2023). https://www.nytimes.com/2023/12/13/climate/emirates-coral-restoration.html?searchResultPosition=10 Google Scholar

6. 

B. Sivakumar, “Global climate change and its impacts on water resources planning and management: assessment and challenges,” Stoch. Environ. Res. Risk Assess., 25 (4), 583 –600 https://doi.org/10.1007/s00477-010-0423-y (2011). Google Scholar

7. 

D. Erdenesanaa, “A year of ‘unreal’ fire and warming in the arctic,” The New York Times, (19 December 2023). https://www.nytimes.com/2023/12/12/climate/arctic-report-card-climate-change.html?searchResultPosition=18 Google Scholar

8. 

P. J. Loftus et al., “A critical review of global decarbonization scenarios: what do they tell us about feasibility?,” WIREs Clim. Change, 6 (1), 93 –112 https://doi.org/10.1002/wcc.324 (2015). Google Scholar

9. 

Intergovernmental Panel on Climate Change, Global warming of 1.5°C. An IPCC special report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty, Cambridge University Press( (2022). https://www.ipcc.ch/sr15/ Google Scholar

10. 

E. Papadis and G. Tsatsaronis, “Challenges in the decarbonization of the energy sector,” Energy, 205 118025 https://doi.org/10.1016/j.energy.2020.118025 ENGYD4 0149-9386 (2020). Google Scholar

11. 

J. Meckling, T. Sterner and G. Wagner, “Policy sequencing toward decarbonization,” Nat. Energy, 2 (12), 918 –922 https://doi.org/10.1038/s41560-017-0025-8 (2017). Google Scholar

12. 

T. Burger et al., “Present efficiencies and future opportunities in thermophotovoltaics,” Joule, 4 (8), 1660 –1680 https://doi.org/10.1016/j.joule.2020.06.021 (2020). Google Scholar

13. 

Z. Zhou et al., “Solar thermophotovoltaics: reshaping the solar spectrum,” Nanophotonics, 5 (1), 1 –21 https://doi.org/10.1515/nanoph-2016-0011 (2016). Google Scholar

14. 

B. Wernsman, “Advanced thermophotovoltaic devices for space nuclear power systems,” AIP Conf. Proc., 746 1441 –1448 APCPCS 0094-243X (2005). Google Scholar

15. 

K. Qiu and A. C. S. Hayden, “Implementation of a TPV integrated boiler for micro-CHP in residential buildings,” Appl. Energy, 134 143 –149 https://doi.org/10.1016/j.apenergy.2014.08.016 (2014). Google Scholar

16. 

H. Daneshvar, R. Prinja and N. P. Kherani, “Thermophotovoltaics: fundamentals, challenges and prospects,” Appl. Energy, 159 560 –575 https://doi.org/10.1016/j.apenergy.2015.08.064 (2015). Google Scholar

17. 

A. Licht et al., “A review of advances in thermophotovoltaics for power generation and waste heat harvesting,” MRS Adv., 4 (41–42), 2271 –2282 https://doi.org/10.1557/adv.2019.342 (2019). Google Scholar

18. 

R. Sakakibara et al., ““Practical emitters for thermophotovoltaics: a review,” J. Photonics Energy, 9 (3), 032713 https://doi.org/10.1117/1.JPE.9.032713 ENGYD4 0149-9386 (2019). Google Scholar

19. 

J. Fernández et al., “Back-surface optimization of germanium TPV cells,” AIP Conf. Proc., 890 190 –197 APCPCS 0094-243X (2007). Google Scholar

20. 

T. Bauer, Thermophotovoltaics: Basic Principles and Critical Aspects of System Design, Springer Berlin Heidelberg, Berlin, Heidelberg (2011). Google Scholar

21. 

J. M. Baker and T. J. Griffis, “Feasibility of recycling excess agricultural nitrate with electrodialysis,” J. Environ. Qual., 46 (6), 1528 –1534 https://doi.org/10.2134/jeq2017.05.0215 JEVQAA 0047-2425 (2017). Google Scholar

22. 

W. R. Chan et al., “Toward high-energy-density, high-efficiency, and moderate-temperature chip-scale thermophotovoltaics,” Proc. Natl. Acad. Sci. U. S. A., 110 (14), 5309 –5314 https://doi.org/10.1073/pnas.1301004110 (2013). Google Scholar

23. 

D. Fan et al., “Near-perfect photon utilization in an air-bridge thermophotovoltaic cell,” Nature, 586 (7828), 237 –241 https://doi.org/10.1038/s41586-020-2717-7 (2020). Google Scholar

24. 

S. Molesky, C. J. Dewalt and Z. Jacob, “High temperature epsilon-near-zero and epsilon-near-pole metamaterial emitters for thermophotovoltaics,” Opt. Express, 21 (S1), A96 https://doi.org/10.1364/OE.21.000A96 OPEXFF 1094-4087 (2013). Google Scholar

25. 

A. Datas et al., “Latent heat thermophotovoltaic batteries,” Joule, 6 (2), 418 –443 https://doi.org/10.1016/j.joule.2022.01.010 (2022). Google Scholar

26. 

K. Park et al., “Performance analysis of near-field thermophotovoltaic devices considering absorption distribution,” J. Quantum Spectrosc. Radiat. Transf., 109 (2), 305 –316 https://doi.org/10.1016/j.jqsrt.2007.08.022 (2008). Google Scholar

27. 

A. Datas and A. Martí, “Thermophotovoltaic energy in space applications: review and future potential,” Sol. Energy Mater. Sol. Cells, 161 285 –296 https://doi.org/10.1016/j.solmat.2016.12.007 SEMCEQ 0927-0248 (2017). Google Scholar

28. 

A. LaPotin et al., “Thermophotovoltaic efficiency of 40%,” Nature, 604 (7905), 287 –291 https://doi.org/10.1038/s41586-022-04473-y (2022). Google Scholar

29. 

C. Amy et al., “Thermal energy grid storage using multi-junction photovoltaics,” Energy Environ. Sci., 12 (1), 334 –343 https://doi.org/10.1039/C8EE02341G EESNBY 1754-5692 (2019). Google Scholar

30. 

A. Datas et al., “Ultra high temperature latent heat energy storage and thermophotovoltaic energy conversion,” Energy, 107 542 –549 https://doi.org/10.1016/j.energy.2016.04.048 ENGYD4 0149-9386 (2016). Google Scholar

31. 

M. M. A. Gamel et al., “A review on thermophotovoltaic cell and its applications in energy conversion: issues and recommendations,” Materials, 14 (17), 4944 https://doi.org/10.3390/ma14174944 MATEG9 1996-1944 (2021). Google Scholar

32. 

K. F. Mustafa et al., “A review of combustion-driven thermoelectric (TE) and thermophotovoltaic (TPV) power systems,” Renew. Sustain. Energy Rev., 71 572 –584 https://doi.org/10.1016/j.rser.2016.12.085 (2017). Google Scholar

33. 

D. N. Woolf et al., “High-efficiency thermophotovoltaic energy conversion enabled by a metamaterial selective emitter,” Optica, 5 (2), 213 https://doi.org/10.1364/OPTICA.5.000213 (2018). Google Scholar

34. 

S. Hassan, C. F. Doiron and G. V. Naik, “Optimum selective emitters for efficient thermophotovoltaic conversion,” Appl. Phys. Lett., 116 (2), https://doi.org/10.1063/1.5131367 APPLAB 0003-6951 (2020). Google Scholar

35. 

Q. Pan et al., “Deep learning-based inverse design optimization of efficient multilayer thermal emitters in the near-infrared broad spectrum,” Opt. Express, 31 (15), 23944 https://doi.org/10.1364/OE.490228 OPEXFF 1094-4087 (2023). Google Scholar

36. 

M. Bianchi et al., “Feasibility study of a thermo-photo-voltaic system for CHP application in residential buildings,” Appl. Energy, 97 704 –713 https://doi.org/10.1016/j.apenergy.2012.01.049 (2012). Google Scholar

37. 

L. M. Fraas, J. E. Avery and H. X. Huang, “Thermophotovoltaic furnace-generator for the home using low bandgap GaSb cells,” Semicond. Sci. Technol., 18 (5), S247 https://doi.org/10.1088/0268-1242/18/5/316 SSTEET 0268-1242 (2003). Google Scholar

38. 

W. Durisch and B. Bitnar, “Novel thin film thermophotovoltaic system,” Sol. Energy Mater. Sol. Cells, 94 (6), 960 –965 https://doi.org/10.1016/j.solmat.2010.01.024 SEMCEQ 0927-0248 (2010). Google Scholar

39. 

A. Castillo and D. F. Gayme, “Grid-scale energy storage applications in renewable energy integration: a survey,” Energy Convers. Manag., 87 885 –894 https://doi.org/10.1016/j.enconman.2014.07.063 ECMADL 0196-8904 (2014). Google Scholar

40. 

M. Jafari, A. Botterud and A. Sakti, “Decarbonizing power systems: a critical review of the role of energy storage,” Renew. Sustain. Energy Rev., 158 112077 https://doi.org/10.1016/j.rser.2022.112077 (2022). Google Scholar

41. 

Y. Khan, “Low battery metal prices set to persist in 2024, adding friction to energy transition,” (20231 April 2024). https://www.wsj.com/articles/low-battery-metal-prices-set-to-persist-in-2024-adding-friction-to-energy-transition-3773ba00?&mod=djemSustainableBusinessPro Google Scholar

42. 

M. Guan et al., “Catalysing the global opportunity for electrothermal energy storage,” (202421 March 2024). https://www.systemiq.earth/wp-content/uploads/2024/03/Global-ETES-Opportunity-Report-240227.pdf Google Scholar

43. 

E. Ballard, “Startup wants to store spare renewable electricity to power heavy industry,” (202221 October 2023). https://www.wsj.com/articles/startup-wants-to-store-spare-renewable-electricity-to-power-heavy-industry-11644328800?mod=djemclimate Google Scholar

44. 

R. Wu, Z. Tan and B. Lin, “Does carbon emission trading scheme really improve the CO2 emission efficiency? Evidence from China’s Iron and Steel Industry,” Energy, 277 127743 https://doi.org/10.1016/j.energy.2023.127743 ENGYD4 0149-9386 (2023). Google Scholar

45. 

R. Kajaste and M. Hurme, “Cement industry greenhouse gas emissions–management options and abatement cost,” J. Clean. Prod., 112 4041 –4052 https://doi.org/10.1016/j.jclepro.2015.07.055 (2016). Google Scholar

46. 

“Master plan Part 3- sustainable energy for all of Earth,” (202321 December 2023). https://www.tesla.com/ns_videos/Tesla-Master-Plan-Part-3.pdf Google Scholar

47. 

R. Toplensky, “Octopus energy has Texas-size ambitions,” (20244 January 2024). https://www.wsj.com/articles/octopus-energy-has-texas-size-ambitions-6f11cfb8?mod=djemSustainableBusinessPro Google Scholar

48. 

C. Amy et al., “High-temperature pumping of silicon for thermal energy grid storage,” Energy, 233 121105 https://doi.org/10.1016/j.energy.2021.121105 ENGYD4 0149-9386 (2021). Google Scholar

49. 

C. C. Kelsall, K. Buznitsky and A. Henry, “Technoeconomic analysis of thermal energy grid storage using graphite and tin,” (202125 July 2023). http://arxiv.org/abs/2106.07624 Google Scholar

50. 

J. Jiao et al., “The use of eutectic Fe-Si-B alloy as a phase change material in thermal energy storage systems,” Materials, 12 (14), 2312 https://doi.org/10.3390/ma12142312 MATEG9 1996-1944 (2019). Google Scholar

51. 

S. Lang, H. Drück and D. Bestenlehner, “Ultrahigh temperature thermal insulation,” Ultra-high temperature thermal energy storage, transfer and conversion, 201 –219 Elsevier( (2021). Google Scholar

52. 

A. Datas, D. L. Chubb and A. Veeraragavan, “Steady state analysis of a storage integrated solar thermophotovoltaic (SISTPV) system,” Sol. Energy, 96 33 –45 https://doi.org/10.1016/j.solener.2013.07.002 (2013). Google Scholar

53. 

A. Veeraragavan, L. Montgomery and A. Datas, “Night time performance of a storage integrated solar thermophotovoltaic (SISTPV) system,” Sol. Energy, 108 377 –389 https://doi.org/10.1016/j.solener.2014.07.005 (2014). Google Scholar

54. 

P. Albertus, J. S. Manser and S. Litzelman, “Long-duration electricity storage applications, economics, and technologies,” Joule, 4 (1), 21 –32 https://doi.org/10.1016/j.joule.2019.11.009 (2020). Google Scholar

55. 

K. A. W. Horowitz et al., “A bottom-up cost analysis of a high concentration PV module,” AIP Conf. Proc., 1679 100001 APCPCS 0094-243X (1679). Google Scholar

56. 

A. Datas, A. Ramos and C. del Cañizo, “Techno-economic analysis of solar PV power-to-heat-to-power storage and trigeneration in the residential sector,” Appl. Energy, 256 113935 https://doi.org/10.1016/j.apenergy.2019.113935 (2019). Google Scholar

57. 

M. Mosalpuri, “Python code for techno-economic analysis and uncertainty assessment of thermophotovoltaics technology,” (16 April 2024). https://github.com/manishm211/ThermoPhotoVoltaics Google Scholar

58. 

J. A. Nelder and R. Mead, “A simplex method for function minimization,” Comput. J., 7 (4), 308 –313 https://doi.org/10.1093/comjnl/7.4.308 (1965). Google Scholar

59. 

N. M. Frick et al., “End-use load profiles for the U.S. building stock: market needs, use cases, and data gaps,” (201910 April 2024). https://www.nrel.gov/docs/fy20osti/75215.pdf Google Scholar

60. 

Google Finance, “Euro to United States dollar,” (8 February 2024). https://www.google.com/finance/quote/EUR-USD?sa=X&ved=2ahUKEwjr-5i7opyEAxXhj4kEHTtdCtQQmY0JegQIBxAv Google Scholar

61. 

S. Greenland, “Sensitivity analysis, Monte Carlo risk analysis, and Bayesian uncertainty assessment,” Risk Anal., 21 (4), 579 –584 https://doi.org/10.1111/0272-4332.214136 RIANDF 0272-4332 (2001). Google Scholar

62. 

B. Li et al., “Techno-economic and uncertainty analysis of in situ and ex situ fast pyrolysis for biofuel production,” Bioresour. Technol., 196 49 –56 https://doi.org/10.1016/j.biortech.2015.07.073 BIRTEB 0960-8524 (2015). Google Scholar

63. 

M. H. Kalos and P. A. Whitlock, “What is Monte Carlo?,” Monte Carlo Methods, 1 –5 Wiley-VCH Verlag GmbH & Co. KGaA( (2008). Google Scholar

64. 

S. Arens et al., “Monte-Carlo evaluation of residential energy system morphologies applying device agnostic energy management,” IEEE Access, 10 7460 –7475 https://doi.org/10.1109/ACCESS.2021.3138549 (2022). Google Scholar

65. 

A. Tavakoli and A. Karimi,, “Development of Monte-Carlo-based stochastic scenarios to improve uncertainty modelling for optimal energy management of a renewable energy hub,” IET Renew. Power Gener., 17 (5), 1139 –1164 https://doi.org/10.1049/rpg2.12671 (2023). Google Scholar

66. 

G. Mavrotas, K. Florios and D. Vlachou, “Energy planning of a hospital using mathematical programming and Monte Carlo simulation for dealing with uncertainty in the economic parameters,” Energy Convers. Manage., 51 (4), 722 –731 https://doi.org/10.1016/j.enconman.2009.10.029 ECMADL 0196-8904 (2010). Google Scholar

67. 

C. K. Gethers et al., “TPV efficiency predictions and measurements for a closed cavity geometry,” in Third NREL Conf. thermophotovolt. Gener. of Electr., 471 –486 (1997). Google Scholar

68. 

R. Liu et al., “Simulation of edge effects in thermophotovoltaic efficiency by Monte Carlo calculation,” AIP Adv., 13 (2), 025263 https://doi.org/10.1063/5.0132045 AAIDBI 2158-3226 (2023). Google Scholar

69. 

J. Aschaber, C. Hebling and J. Luther, “Realistic modelling of TPV systems,” Semicond. Sci. Technol., 18 (5), S158 –S164 https://doi.org/10.1088/0268-1242/18/5/304 SSTEET 0268-1242 (2003). Google Scholar

70. 

National Renewable Energy Laboratory (NREL), “Residential PV annual technology baseline,” (1 June 2024). https://atb.nrel.gov/electricity/2022/residential_pv Google Scholar

71. 

M. Lin, H. C. Lucas and G. Shmueli, “Too big to fail: large samples and the p-value problem,” Inf. Syst. Res., 24 (4), 906 –917 https://doi.org/10.1287/isre.2013.0480 ISYREH (2013). Google Scholar

72. 

United States Environmental Protection Agency (EPA), “Greenhouse gases equivalencies calculator–calculations and references,” (22 March 2024). https://www.epa.gov/energy/greenhouse-gases-equivalencies-calculator-calculations-and-references#:∼:text=The%20average%20carbon%20dioxide%20coefficient,cubic%20foot%20(EIA%202023 Google Scholar

73. 

United States Environmental Protection Agency (EPA), “eGRID CO2 total output emission rate (lb/MWh),” (22 March 2024). https://www.epa.gov/egrid/data-explorer Google Scholar

74. 

S. Nicholson and G. Heath, “Life cycle emissions factors for electricity generation technologies,” (2023). Google Scholar

75. 

E. Taskesen, “Distfit is a Python library for probability density fitting,” (202016 April 2024). https://erdogant.github.io/distfit Google Scholar

Biography

Manish Mosalpuri is a fourth-year PhD student in mechanical engineering at Iowa State University. His research focuses on techno-economic analysis (TEA) and life cycle assessment (LCA) of renewable energy technologies. He has worked on electrochemical conversion of waste nitrates to hydroxylamine, energy conversion and storage using thermophotovoltaics, and solar thermal methane pyrolysis. With experience in modeling and optimization during PhD and masters’ research, he aims to advance sustainable energy solutions.

Fatima Toor is an associate professor of electrical and computer engineering and the Lowell G. Battershell Chair in Laser Engineering at the University of Iowa. She also holds a secondary appointment in physics and astronomy. She specializes in semiconductor optoelectronics, with research areas including chemical sensing, biomedical imaging, and energy generation. She is a senior member of Optica and SPIE and has published numerous papers in her field.

Mark Mba-Wright is an associate professor in mechanical engineering at Iowa State University. He specializes in techno-economic and life-cycle analysis of energy systems, focusing on carbon-negative energy technologies. With over 60 publications, his research is sponsored by ARPA-e, NSF, and other federal and private agencies. He chairs the American Institute of Chemical Engineers Sustainable Engineering Forum.

© 2024 Society of Photo-Optical Instrumentation Engineers (SPIE)
Manish Mosalpuri, Fatima Toor, and Mark Mba-Wright "Techno-economic analysis of a solar thermophotovoltaic system for a residential building," Journal of Photonics for Energy 14(4), 042404 (13 September 2024). https://doi.org/10.1117/1.JPE.14.042404
Received: 26 April 2024; Accepted: 20 August 2024; Published: 13 September 2024
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Electroluminescence

Solar energy

Solar cells

Photovoltaics

Monte Carlo methods

Nanooptomechanical systems

Power consumption

Back to Top