Open Access
26 February 2019 Practical emitters for thermophotovoltaics: a review
Reyu Sakakibara, Veronika Stelmakh, Walker R. Chan, Michael Ghebrebrhan, John D. Joannopoulos, Marin Soljacic, Ivan Čelanović
Author Affiliations +
Abstract
Thermophotovoltaic (TPV) systems are promising for harnessing solar energy, waste heat, and heat from radioisotope decay or fuel combustion. TPV systems work by heating an emitter that emits light that is converted to electricity. One of the key challenges is designing an emitter that not only preferentially emits light in certain wavelength ranges but also simultaneously satisfies other engineering constraints. To elucidate these engineering constraints, we first provide an overview of the state of the art, by classifying emitters into three categories based on whether they have been used in prototype system demonstrations, fabricated and measured, or simulated. We then present a systematic approach for assessing emitters. This consists of five metrics: optical performance, ability to scale to large areas, stability at high temperatures, ability to integrate into the system, and cost. Using these metrics, we evaluate and discuss the reported results of emitters used in system demonstrations. Although there are many emitters with good optical performance, more studies on their practical attributes are required, especially for those that are not yet used in prototype systems. This framework can serve as a guide for the development of emitters for long-lasting, high-performance TPV systems.

1.

Brief Introduction to Thermophotovoltaics

A thermophotovoltaic (TPV) system converts heat to electricity using light as an intermediary and consists of (at least) three components: a heat source, an emitter, and a photovoltaic (PV) cell with a low bandgap. The heat source brings the emitter to high temperature (1000  K), causing the emitter to emit thermal radiation, which is absorbed and converted to electricity by the PV cell. Some advantages of this energy conversion scheme include the static and quiet conversion process, the physically separated paths of heat conduction and electricity generation, and the lack of fundamental temperature gradients across materials.1 In addition, several heat sources can be used, of which there are three major kinds: radioisotope decay, chemical fuel, and sunlight that is concentrated and absorbed (Fig. 1). The radiated power density from the TPV emitter is fundamentally limited only by Planck’s law for blackbody emission.2 However, high-performance TPV systems are particularly challenging to realize in part because of the need to coordinate multiple subsystems and the difficulties in designing a good emitter.

Fig. 1

The basic three components of a TPV system are a heat source, an emitter, and a PV cell (sometimes the PV cell is known as the TPV cell). The hot side is made up of a heat source in thermal contact with an emitter and converts heat to light. On the cold side, the PV cell converts the thermal radiation from the emitter into electricity. Sometimes, the cold side includes a front side filter or back surface reflector, to be explained later. A near-field TPV device has a subwavelength gap between the emitter and the PV cell, but we only focus on standard TPV systems.

JPE_9_3_032713_f001.png

The scope of our review is the TPV emitter, which we consider the critical component toward high system performance, and specifically on its practical implementation in TPV systems. The emitter is any material that is heated up to a high temperature. A particularly useful emitter is a selective emitter, which preferentially emits in a specific wavelength region. There is no one way to make an emitter; there are many types of emitters that can each involve a different geometric configuration and a separate set of materials, some of which we touch upon in Sec. 2.

In our review, we propose five metrics to evaluate the practicality of TPV emitters, and in particular, we examine emitters that have been used in system demonstrations of TPV prototypes. Most work on TPV emitters has focused on achieving good optical performance, but there has been little consideration of the challenges associated with implementing the emitter in and operating a TPV system.

Our review is organized as follows: in Sec. 2, we classify TPV emitters from the literature into three categories. In Sec. 3, we present and discuss the five metrics for practical TPV emitters as well as their submetrics. In Sec. 4, we provide an at-a-glance evaluation, based on the five metrics, of the five types of emitters used in prototype system demonstrations. The evaluations are summarized in two tables, with one more detailed table in the Appendix (Sec. 6).

2.

Classification of TPV Emitters

Before evaluating the practicality of TPV emitters, we first classify emitters in the literature into the following categories: (i) used in system demonstrations of TPV prototypes, (ii) fabricated and optical performance measured, and (iii) optical performance simulated. These categories are shown in Fig. 2. In this figure, we only include emitters with emission in the range of 1 to 3  μm, which corresponds approximately to the peak emission wavelengths of typical emitters at temperatures of 1000 to 1500 K. It is not within the scope of this review to discuss the mechanisms behind different emitters, nor have we provided a complete list of all emitters that have been proposed and fabricated. Detailed discussion of emitter types, mechanisms, and more examples of emitters can be found in reviews elsewhere, such as that by Pfiester and Vandervelde.3

Fig. 2

Three categories of TPV emitters include those that have been (i) used in published system demonstrations of TPV prototypes, (ii) fabricated and measured, and (iii) simulated. The emitters in this figure emit in 1 to 3  μm range. Abbreviations and some terminology: atomic symbols are used, ARC is an antireflection coating, a photonic crystal is a periodic structure,4 a metamaterial is a manmade material that has optical properties not usually found in nature, and metasurfaces are a class of metamaterials that consist of a 2-D array of metal features on a dielectric spacer on a metal substrate.590

JPE_9_3_032713_f002.png

Our focus for this review is on the emitters in the first category that have been used for system demonstrations: in Sec. 4, we evaluate the practical aspects of these emitters, following the discussion of our metrics.

3.

Metrics for Practical TPV Systems

Although the primary purpose to develop an emitter is for its optical performance, the emitter with the best optical performance is not necessarily the best emitter for practical implementation.

For this reason, we present five practical metrics, as shown in Fig. 3: (1) optical performance, (2) ability to scale to large areas, (3) long-term high-temperature stability, (4) ease of integration within the TPV system, and (5) cost.

Fig. 3

The proposed five metrics for practical TPV emitters include (1) optical performance, (2) ability to scale to large areas, (3) long-term high-temperature stability, (4) ease of integration within the TPV system, and (5) cost. Each has submetrics as shown.

JPE_9_3_032713_f003.png

In the following subsections, we discuss each metric, including some approaches that researchers have taken to address key challenges.

3.1.

Optical Performance

An emitter with good optical performance has, at all angles, preferential emission of in-band photons and suppression of out-of-band photons. Optical performance refers to the emission of photons as a function of both angle and photon energy, in particular, in two regimes for the latter, in-band photons that have energy higher than the PV cell bandgap and out-of-band photons that have energy lower. Spectral control refers to the methods that enable preferential in-band emission. Some designs of spectral control are designed for broadband emission while others for narrow-band emission [the ideal cases which are shown in Fig. 4(a)]. In the latter case, the emitted photons have energies slightly above the PV cell bandgap. Typically, broadband emitters yield higher output electrical power density while narrow-band emitters can increase the TPV conversion efficiency.91

Fig. 4

An emitter with good optical performance may have (a) either broadband emission, where any in-band photons (energy greater or wavelength shorter than the PV cell bandgap, where EPV and λPV are the bandgap energy and wavelength, respectively) are preferentially emitted, or narrow-band emission, where only photons with energy slightly above the bandgap are emitted. Note that we refer to photons or radiation both in terms of energy and wavelength. (b) The goal of angular control is to ensure good spectral control (preferential in-band emission) over a wide range of angles, as thermal radiation can be off-normal. (c) View factor loss, where photons are lost through the emitter-PV cell gap, is a significant source of loss.

JPE_9_3_032713_f004.png

The purpose of angular control, which is often an implicit aspect of spectral control, is to ensure spectral control over all angles (polar and azimuthal, θ and ϕ), because an emitter radiates photons over a wide range of angles [see Fig. 4(b)]. This is especially important because most thermal radiation is off-normal as according to Lambert’s law.

For TPV, the wavelength regions of interest are around 1 to 3  μm, approximately the regions of peak emission. For an emitter heated to realistic temperatures of 1000 to 1500 K, the peak emission wavelengths are 1.9 to 2.9  μm, as according to Wien’s displacement law. As such, one of the main requirements of TPV is to have low-bandgap PV cells, with typical bandgaps in the range of about 0.50 to 0.74 eV or 1.7 to 2.3  μm.

Although a PV cell can generate electricity only from in-band photons, a real emitter emits both in-band and out-of-band photons at any given angle. This leads to the following problems: (a) if out-of-band photons reach the PV cell, they overheat the PV cell and reduce the PV cell efficiency and (b) when out-of-band photons are emitted and not recovered, this leads to both reduced heat-to-radiation efficiency and emitter temperature.

While we have initially defined good optical performance as that achieved by designing selective emitters, there are actually two main approaches of spectral control. The first is to enhance in-band and suppress out-of-band emission via selective emitters. The second is to reflect out-of-band photons back to the emitter, or photon recycling, via cold side filters or reflectors (CSFR) in front of (front side filter) or behind the PV cell (back surface reflector). (These are shown in Fig. 1.) It is also possible to combine both approaches, for example, to have a selective emitter and a filter or reflector, or even all three in theory.

Although an emitter with a CSFR performs better than a blackbody or graybody (relatively higher temperature and mitigated PV cell efficiency reduction), it suffers from view factor and absorption losses. In view factor loss, which is inherent to systems with diffuse emitters, photons are lost in the finite gap between the emitter and the filter/reflector [Fig. 4(c)], and in absorption loss, photons are absorbed at any interface (at the filter, reflector, or PV cell). Although it is possible to reduce the view factor loss by reducing the emitter area relative to the PV cell area, keeping an emitter arbitrarily small decreases its absolute radiated power.

On the other hand, selective emitters suppress out-of-band emission relative to in-band emission. This reduces view factor and absorption losses for out-of-band photons, although both losses, especially view factor losses, remain significant for in-band photons.

For the approach of selective emitters, the objective is to find or engineer a TPV emitter that emits mostly in-band photons and little to no out-of-band photons. While this is not within the scope of this review, there are many design questions regarding specific emission characteristics:

  • Is it better to prioritize high in-band emission, even if the out-of-band emission is moderately high, or to prioritize low out-of-band emission, even if the in-band emission is only moderately high?

  • Is it better to prioritize broadband or narrow-band emission? Narrow-band emitters are intended to prevent thermalization82 in PV cells, in which the excess energy (difference between photon and bandgap energies) is absorbed and lost. However, this comes at the cost of a reduction in the radiated in-band power density. One proposed way to mitigate thermalization, then, is to use PV cells of multiple bandgaps.91

  • For engineered emitters, which is better: (a) select a material with naturally high emission, and suppress it in out-of-band wavelength regions or (b) select a material with naturally low emission, and enhance it in the in-band wavelength regions? Generally, suppression of naturally high emission works only for a limited wavelength range;27,92 ideally, the emission should be suppressed for wavelengths up to about 15  μm, which accounts for >96% of the energy emitted by a blackbody at 1000 K.

However, an emitter is not better than others simply because it has reached high temperatures, because temperatures beyond 1500 K are hard to achieve, the amount of input power required to heat an emitter may vary widely, and it is unclear if a given emitter can sustain high optical performance at high temperature for prolonged periods. The issue of stability at high temperature is discussed as another metric later on.

One submetric is the in-band radiated power density Mrad,in because the ultimate system goal is to have high-output electrical power, which is enabled by maximizing the in-band power density that is emitted and can be converted.

The radiated in-band power density, Mrad,in, can be calculated from the hemispherical emittance ε, the cutoff or bandgap wavelength λPV, and the blackbody spectrum eb(λ,T):93

Eq. (1)

Mrad,in=π0λPVε(λ,T)eb(λ,T)dλ.

The hemispherical emittance ε is the emittance across all angles, where the emittance is a measure of how close the emission is to that of a blackbody. Ideally, the in-band emittance εin is close to 1, while the out-of-band emittance εout is close to 0. In addition, the hemispherical emittance is temperature dependent, as the optical properties of a material change with temperature:

Eq. (2)

ε(λ,T)=1π02π0π/2ε(λ,T,θ,ϕ)cosθsinθdθdϕ.

However, many papers often report only the emittance at a single angle at room temperature, since it is difficult to measure the emittance across all angles as well as at high temperature.

The common metric spectral selectivity or spectral efficiency ηsp, which is the fraction of the radiated energy that is convertible by the PV cell, can be calculated using the hemispherical emittance:94,95

Eq. (3)

ηsp=0λPVε(λ,T)eb(λ,T)dλ0ε(λ,T)eb(λ,T)dλ.

It is important to point out that spectral selectivity describes in-band emission relative to the total or out-of-band emission and is distinct from the absolute values of in-band and out-of-band emittance. In other words, it is possible to have a highly selective emitter with low absolute in-band emittance or an emitter that has high in-band emittance but low selectivity (such as a graybody emitter).

Finally, we are also interested in the efficiency of the radiated in-band power to the input power, ηri-input, which is calculated from the radiated in-band power density Mrad,in, the area of the emitter Aemitter, and the input power Pinput:

Eq. (4)

Prad,in=Mrad,inAemitter,

Eq. (5)

ηriinput=Prad,in/Pinput.

One qualitative metric, which we do not use in our evaluation, is that the design of the emitter itself should be robust to fabrication imperfections across a large area, such as the lack of uniformity in critical feature dimensions. For example, a very thin film with 30-nm thickness is less robust to effects of surface roughness compared to a much thicker film.

3.2.

Scalability to Large Areas

Because the fundamental limit for emitters is on the power radiated per unit area, one way to increase the absolute radiated power is by increasing the emitter area (its macroscopic exterior dimensions).

In terms of practical implementation, it is important to consider the following: (1) the substrates must be available in large sizes and (2) the fabrication methods should accommodate large-area samples relatively easily. For example, for 1, the single-crystalline substrates of tungsten and tantalum are typically available in small diameters 1 to 1.5 cm (area 3 to 7  cm2),33,57,58 while naturally selective emitters made of rare earth metals can be on the order of tens of cm2.2325 An example for 2 is that electron beam lithography, which is typically used for features <500  nm, is both costly and time-consuming. The overall complexity of the fabrication process, including the number of steps and the complexity of each individual step, can impact the scalability as well as the cost. However, many of the fabrication techniques used in papers may be those best suited for proof-of-concept demonstrations, rather than mass production.

3.3.

Long-Term High-Temperature Stability

The TPV emitter must sustain its optical performance at high temperatures for extended periods of time, either continuously or over multiple thermal cycles.

However, at high temperatures, the kinetic energy of atoms increases and atoms diffuse more easily, leading to a number of potential thermodynamic effects:1

  • Sharp edges and features can become more rounded.51,67,68,96100

  • A phase change may occur (e.g., the emitter might melt),101 accompanied also by changes in morphology and optical properties.51 This can happen also for crystalline phases.45 However, it is important to keep in mind that the melting point of a material at nanometer scale is lower than for bulk.101

  • The sizes of grains can grow in polycrystalline materials.51,58,96100103 However, this can actually stabilize the material, so some substrates such as polycrystalline tantalum are annealed prior to use.58 It is also possible to use large-grain or single-crystal substrates.51,57,102,104,105

  • Chemical degradation may occur, such as the formation of tungsten oxides96,97,99,100 and tantalum carbide.102,105 This can necessitate that the emitter operate in inert atmosphere or vacuum,97,99,100 which requires special packaging and complicates the TPV system integration. Chemical degradation of 2-D and 3-D tungsten and 2-D tantalum photonic crystals can be mitigated by capping the surface with a 20 to 40 nm protective coating of hafnium dioxide (HfO2).1,67,68,102104 One comparison of HfO2 and Al2O3 in 3-D photonic crystals67 has found HfO2 to be more thermally robust than Al2O3, but Al2O3 is less expensive and has been used to protect a metasurface emitter.73

  • Thermal expansion could lead to the cracking of a material.67,68,95,103 Also, emitters with interfaces between different materials are at risk of delamination because different materials have different thermal expansion coefficients.

Some strategies for improving the high-temperature stability include selecting materials that are known to have good high temperature properties, alloying to promote a solute drag effect,97,99,100 and modifying the geometry of a structure to change diffusion rates.100,106

There do not appear to be any published long-term (>1000  h) studies; in some cases, it appears the emitter is only heated to measure its high-temperature optical properties. One long study is 168 h (7 days) at 1000°C (1273 K) for a 2-D structure made with tungsten and carbon nanotubes.61 We have included some studies of emitter stability at high temperatures in Table 1.

Table 1

A few (this list is not exhaustive) studies of high-temperature stability of TPV emitters.

EmitterUsed in system demo.?LengthTemperature (K)EnvironmentResult
Al2O3-coated W 3-D inverse opal photonic crystal6712 h1273Forming gas, 5% H2 in ArLittle to no degradation
Same as above12 h1673Forming gas, 5% H2 in ArCollapse of structure
HfO2-coated W 3-D inverse opal photonic crystal6712 h1273Forming gas, 5% H2 in ArLittle to no degradation
Same as above12 h1673Forming gas, 5% H2 in ArSome grain growth, increased surface roughness
W-coated 3-D inverse opal photonic crystal6812 h1273Little to no degradation
Same as above>1273Collapse of structure
W inverse 3-D inverse opal photonic crystal6812 h1273Little to no degradation
20  nm-HfO2-coated W inverse 3-D inverse opal photonic crystal681 h1673Little to no degradation
2-D structure made with W, carbon nanotubes61168 h, 7 days12731×103  Torr with He protectionLittle to no degradation
Single crystal Er doped yttrium aluminum garnet21Yes300 hIn sunIn sunEmitter cracked and darkened, but may be due to water leak
2-D photonic crystal made of Ta3%W alloy w/HfO2 coating1Yes31,32300 h1273Vacuum (5×106  Torr)Little to no degradation
Same as above104Yes31,321 h1473Vacuum (5×106  Torr)Little to no degradation
Yb2O3 foam ceramic25Yes200 cyclesLittle to no degradation
2-D photonic crystal made of polycrystalline Ta w/HfO2 coating102Yes29,30144 h, 6 days1173Ar, 100  mTorrLittle to no degradation
Same as aboveYes29,301 h1273Ar, 100  mTorrLittle to no degradation
A multilayer stack made of W and HfO2107Yes34>1  h1473Vacuum, <5×102  PaDegradation
Same as aboveYes34Two rapid thermal cycles1250 maxLittle to no degradation

The longest reported studies we know of are 300 h each for an erbium-doped yttrium aluminum garnet (Er-YAG) crystal used in an solar TPV system21 and a 2-D photonic crystal made of tantalum-tungsten alloy and capped with 20 to 40 nm HfO21 used in a radioisotope TPV prototype.31 The former cracked and darkened after 300 h in the sun, although the authors attribute it potentially to a water leak. The 2-D photonic crystal showed little to no degradation in optical performance after annealing for 300 h at 1000°C (1273 K)1 and also 1 h at 1200°C (1473 K).104

The only other emitters we know of that were used in both system demonstrations and some high-temperature stability experiments include a Yb2O3 foam ceramic,25 a 2-D photonic crystal made of polycrystalline tantalum and coated with 20 to 40 nm HfO2,102 and a multilayer stack made of tungsten and HfO2.34 The foam ceramic was robust under 200 thermal cycles, the 2-D tantalum photonic crystal showed no visible degradation after 144 h at 900°C (1173 K) and 1 h at 1000°C (1273 K),102 and the multilayer stack showed little to no degradation after at least 1 h at 1423 K in vacuum <5×102  Pa and two rapid thermal cycles up to 1250 K, but showed degradation after 1 h at 1473 K.

3.4.

Ease of Integration within the TPV System

The design choices for the emitter can present several challenges for its integration within the TPV system, in particular, when (1) putting the emitter and heat source physically together for thermal contact and (2) packaging the system for operation in vacuum or inert gas environment.

In some cases, the emitter and heat source are made out of the same material, such as silicon carbide12 or platinum,12 or the emitter is directly fabricated onto the heat source, for example, through the deposition of emitter materials, such as silicon and silicon dioxide6,27 or tantalum59 onto a microcombustor, or the fabrication of a combined absorber/emitter for solar TPV.16,33,34

In other cases, it may be required to cut the emitter to the correct size and to machine and weld it onto the heat source, such as a microcombustor. It is possible to use foil, sputtered coating, or a solid-state substrate. In the last case especially, the mechanical properties of the emitter material become significant. As an example, three different substrates have been explored in the development of 2-D photonic crystal emitters. These include single crystalline tungsten, polycrystalline tantalum, and tantalum-tungsten alloy.57,58,104 Tungsten is brittle and difficult to machine and weld,58 while polycrystalline tantalum is more compliant and easier to weld and machine but is soft so needs to be thicker than tungsten to achieve the same mechanical stability. Tantalum-tungsten alloy combines the better thermomechanical properties of tungsten with tantalum’s ability to be more easily machined and welded.1,104

In addition, high temperature stability concerns also apply: operating the heat source and emitter and high temperatures can lead to cracking and delamination of the emitter or heat source.

Also, the emitter often must be in vacuum or inert gas environment in order to prevent chemical degradation processes, such as oxidation and heat losses, due to convective heat transfer processes.

In our comparison of system demonstrations of prototype TPV, we have noted where the system was operated under vacuum or inert atmosphere, and whether the emitter was fabricated onto or together with the heat source.

3.5.

Cost

The overall cost of emitter production depends on the cost of the raw materials, fabrication, as well as system integration. In particular, many emitters make use of relatively scarce materials, such as hafnium and rare earth metals. The cost of fabrication may increase with increased number of processing steps or complexity of fabrication processes.

For our evaluation, because it is generally difficult to project the cost, considering the emitter area can be scaled and improvements in fabrication technology may reduce costs, we have not done any assessment.

4.

At-a-Glance Evaluation of Emitters Used in Prototype System Demonstrations

In Table 2, we summarize reported features of the emitters that have been implemented in prototype system demonstrations. In this table, we show the cutoff wavelength (which corresponds to the bandgap of the PV cell used in the system demonstration), the average in-band emittance εin, the average out-of-band emittance εout, the emittance measurement angle θ, the emitter temperature Temitter, whether we have found high temperature studies, whether the system is operated in vacuum or an inert gas environment, and whether the emitter is fabricated onto or with the heat source.

Table 2

At-a-glance summary of prototype TPV system demonstrations.

EmitterCutoff (μm)εinεoutθTemitter (K)Area (cm2)High temperature studies?Vacuum/inert?Emitter fab. w/heat src?
Si62.2710131YesYes
SixNy-coated Si51.720.70.7All1043Yes
SiC, with filter71.8101.4
SiC, with filter81.70.50.545 deg1053
Commercial SiC burner9,101.81323Yes
SiC, with filter112.07131229.16
Combined emitter-combustor: SiC121.81120Yes
ARC on W on SiC131.80.860.271548469
ARC-coated Pt foil141.81287>38Yes
HfO2-coated W foil151.881300–15002.83Yes
Absorber/emitter: Si3N4-coated, textured W161.80.6110.05214566.25YesYes
W-coated SiSiC171.81463118.8YesYes
Combined emitter-combustor: Pt121.80.0990.0441450Yes
Burner made of unspecified “super alloy”191.7314581282Yes
Absorber/Emitter: W (Ta?) foil181.820.690.4217201.67
Same as above, larger area181.820.740.5214005.09
Commercial gas mantle20,1081.11570–1970
Absorber/emitter: SiC plate, Emitter: Er-YAG211.1, 1.651133–1473Yes21
Al2O3/Er3Al5O12 composite, with pinhole221.80.410.2512541
Yb2O3 mantle, with filter231.1231.7
Yb2O3 fibrous mantle261.1
Yb2O3 mantle24,251.10.63173575
Yb2O3-coated Al2O3 or SiC foam ceramic25,1091.11735?Yes25
1-D PhC of Si, SiO2 on Si microreactor62.2710731YesYes
1-D PhC of Si, SiO2 on Si272.250.860.3212851YesYes
1-D PhC of Si, SiO2, on Si with filter282.254YesYes
HfO2-coated, 2-D PhC made in polycrystalline Ta292.30.520.29All1270Yes102YesYes
HfO2-coated, 2-D PhC made in polycrystalline Ta302.00.580.18All13274.41Yes30YesNo
HfO2-coated, 2-D PhC made in Ta3%W alloy31,322.250.750.2612331Yes1,104YesNo
Combined emitter-absorber: W, yttria-stabilized zirconia multilayer stack331.850.840.2116401.77YesYes
Combined emitter-absorber: Mo, HfO2 multilayer stack341.850.710.15All16400.636Yes107YesYes
Note: PhC, photonic crystal.

Section 6 contains tables with more details, including estimations of the in-band radiated power density (power per area) Mrad,in and the absolute in-band radiated power to input power efficiency, ηri-input. Because spectral selectivity ηsp is a commonly reported metric, and also in part because many papers do not report the long-wavelength emittance, we have not included this in our evaluation.

Broadly, there are five types of emitters that have been implemented in prototype system demonstrations (the emitters in Table 2 are organized by these types, in this order):

  • 1. Bulk emitters

    • a. Graybody emitters, such as silicon and silicon carbide,512 are typically inexpensive, easy to fabricate in large areas, and often can be fabricated onto or with the heat source. However, they have both high in-band emission and out-of-band emission, which is why they are often coupled with cold side filters.

    • b. Metals with1316 or without antireflection coating (ARC)12,1719 can be easy and inexpensive to fabricate in large areas. The emission depends on the metal optical properties; the role of the ARC layer is typically to enhance emittance in a narrow band around the bandgap.

  • 2. Naturally, selective emitters2026,108,109 have been made primarily from rare earth metals, especially erbium and ytterbium. They are easy to fabricate in large areas and with high-temperature stability, especially by doping high-temperature ceramics. However, the emission wavelength range of naturally selective emitters is not tunable and narrow-band, which can lead to low in-band emitted power density.

  • 3. 1-D photonic crystals,6,27,28 also known as dielectric mirrors, consist of alternating layers of materials with a high contrast in indices of refraction. Interference effects in this structure lead to a fairly broad reflection bandwidth, which can be used to suppress high natural emittance of a material for a wavelength region. They are easy and inexpensive to fabricate at large areas, but have multiple interfaces and are not typically made from high-temperature materials though can be directly fabricated onto a heat source. They may have high out-of-band emission outside of the region of suppression.

  • 4. 2-D photonic crystals2932 for TPV typically are a 2-D array of features on top of or in a substrate, such as cylindrical posts or air cavities, with feature sizes on the order of the wavelength of interest. For photonic crystals with air cavities, each individual cavity acts as a waveguide to enhance emission of wavelengths below a cutoff (half a wavelength corresponds roughly to the cavity diameter). Whether an emitter can be fabricated inexpensively with large area and can be integrated with the heat source depends largely on the substrate, for which a high-temperature material is often used.

  • 5. The multilayer stacks33,34 (which differ from 1-D photonic crystals in that there is no periodicity in the thicknesses of the layers) featured in this review are combined absorber/emitters for solar TPV that consist of alternating metal and dielectric layers of varying thicknesses. The layer thicknesses, which are sometimes subwavelength, are optimized to enable both broadband absorption and emission. Although the optical performance is good and the fabrication costs likely low, the ability to fabricate large areas depends on the available sizes of the metal. Although high temperature materials are used, there are many interfaces and the long-term high-temperature stability is unclear.

5.

Conclusion

In our review, we have presented five metrics for evaluating the practicality of TPV emitters and have used this approach to evaluate the five broad categories of emitters that have been used in TPV system prototypes. We have categorized emitters in the literature based on whether they have been used in prototype system demonstrations, fabricated, or simulated, and have evaluated the emitters in the first category according to our metrics. The emitters in this category include bulk emitters, naturally selective emitters, 1-D photonic crystals, 2-D photonic crystals, and multilayer stacks. The metrics for a practical TPV emitter include optical performance, scalability to large areas, long-term high temperature stability, ease of integration within a TPV system, and cost.

However, none of the emitters or types of emitters identified have yet satisfied all five criteria for practical TPV emitters. An emitter with good optical performance shows high, broadband, preferential in-band emittance. Some emitters with promising optical performance include the 1-D photonic crystal, 2-D photonic crystal, and multilayer stacks. For the latter two types, there have been some studies on the high-temperature stability and system integration. The multilayer stack is fabricated directly with the absorber (heat source), but its high-temperature stability has not been shown beyond 1 h. On the other hand, a 2-D photonic crystal made in refractory metals that can be fabricated on the order of cm2 has some slightly high out-of-band emittance but has been successfully integrated with microcombustor heat sources and has shown promising high-temperature stability of a few hundred hours at 900°C to 1000°C.

Moving forward, studies on the practical aspects of each TPV emitter, such as the five metrics we have presented in this review, are critical for the maturation of TPV technology. We have yet to find an emitter that satisfies the following performance metrics: spectral selectivity exceeding 90% to 95%, high in-band emittance of 0.9 to 0.95, and high-temperature stability on the order of thousands of hours and hundreds of cycles (a typical Li ion battery lasts 2 to 3 years and 500  cycles110). Both of these as well as large-area fabrication can be addressed independently of the overall TPV system setup. One promising class of emitter is metamaterials, which show high optical performance; however, studies of high-temperature stability are at the moment limited. The ability of TPV systems to be commercialized hinges on the practical attributes of selective emitters, for which our five-metric approach can serve as a framework.

6.

Appendix: Evaluation of Emitters Used in Published System Demonstrations of Prototype TPV

Because there is no consistency in the optical performance figures of merit reported by papers, we have used the following methods to find or calculate Mrad,in (or P) and ηri-input, as shown in Tables 3Table 45:

  • 1. Reported the numerical values for power provided in the paper:

    • a. Prad,in or Prad,out,

    • b. Ptotal,in-band.

  • 2. Calculation is based on radiation spectrum (power density as a function of photon energy):

    • a. Radiation spectrum at high temperature,

    • b. Radiation spectrum at unknown or room temperature.

  • 3. Calculation is based on the measured or simulated spectrum of emittance or absorptance, emitter temperature Temitter, and PV cell bandgap or cutoff:

    • a. Hemispherical emittance spectrum at TTemitter,

    • b. Hemispherical emittance spectrum at a temperature that is unknown or not close to Temitter,

    • c. Emittance at a single or unknown angle, at TTemitter,

    • d. Absorptance at a single or unknown angle, at a temperature that is not close to Temitter,

    • e. Emittance at a single or unknown angle, at a temperature that is unknown or not close to Temitter.

  • 4. Calculation is based on assumption of graybody emitter behavior with ε=0.96 unless otherwise stated in the paper, ignoring any effects of a cold side filter if included.

  • 5. No calculation:

    • a. Temitter or emitter temperature range is provided, but no emittance spectrum is provided,

    • b. Neither emittance nor Temitter are provided.

Table 3

Details on prototype TPV system demonstrations, part I. Bulk emitters: Si, SiC, metals with and without ARC.

EmitterCutoff μm/(eV)Temitter (K)Mrad,in (W cm−2)Prad,in (W)Pinput (W)ηri-inputMethodAreaHigh temp. studiesVacuum/inert?Emitter fab. w/heat src?
Si62.27 (0.547)10130.6880.68813.70.0502410  mm×10  mmYesYes
SixNy-coated Si51.7210430.1831.33bYes
SiC, with filter71.85bD 1.5 in., L 3.5 in.
Double-chambered SiC cylinder, with filter81.710531.62; 0.246 W transmit. through filter4970.03261aYes
Commercial SiC radiant burner9,101.813232.284Yes
SiC, with filter112.07 (0.6)13123.3798.345.4  cm×5.4  cm
Combined emitter-combustor: SiC121.811200.5934Yes
ARC on W on SiC131.815485.224391a469  cm2, cylindrical
ARC-coated Pt foil141.81287694970.09941a>38  cm2Designed for inert gas
Same as above, different test condition1.812871429520.1491a>38  cm2Designed for inert gas
HfO2-coated W foil151.881300–15005aD 12 mm, L 25 mmAr flow
Absorber/emitter: Si3N4-coated, laser-textured W161.8 (0.67)14562.7417.168.80.2493eAbsorber: 7.7  mm×7.7  mm, Emitter: 25  mm×25  mm<10  mTorr
W-coated SiSiC cylinder171.8146318005aD 55 mm; L>50  mmAr>20  mbarYes
Combined emitter-combustor: Pt121.814500.3623eYes
Burner made of unspecified “super alloy”191.73 (0.72)14585.8611.77 kW0.2501bD 8.0 cm, L 25.5 cm; effective emitter area: 502  cm2Yes
Absorber/Emitter: W (Ta?) foil cylinder181.8217208.900.6064.8; (180  Wm2)0.00933dConcentrator: 0.6  m×0.6  m, Emitter: D, L 15 mm
Same as above, but longer cylinder181.8214002.310.4764.8; (180  Wm2)0.00723dConcentrator: 0.6  m×0.6  m, Emitter: D 12 mm, L 45 mm
Note: D, diameter; L, length.

Table 4

Details on prototype TPV system demonstrations, part II. Naturally selective emitters based on rare earth materials, 1-D photonic crystals, 2-D photonic crystals.

EmitterCutoff μm/(eV)Temitter (K)Mrad,in (W cm−2)Prad,in (W)Pinput (W)ηri-inputMethodAreaHigh temperature studiesVacuum/inert?Emitter fab. w/heat src?
Gas mantle in a commercial campinglantern20,1081.1 (1.1)1570–19705a
Absorber/emitter: SiC plate, Emitter: single crystal Er doped yttrium aluminum garnet211.1,1.651133–14731006  kWm25aPost 300 h in sun, emitter cracked and darkened, but may be due to water leak21
Al2O3/Er3Al5O12 composite, w/polished Mo pinhole outsideemitter221.812540.6340.63100 to 4500.0014 to 0.00633e10  mm×10  mm
Yb2O3 mantle, w/filter231.111,7235bD 1.5 in., L 8 in.
Yb2O3 fibrous mantle261.15b
Yb2O3 mantle24,251.1 (1.1)17351.3271.120000.03562bActual 75  cm2, Effective 54  cm2
Yb2O3-coated Al2O3 or SiC foam ceramic25,1091.1 (1.1)17351.272b200 thermalcycles25
1-D photonic crystal (PhC), 5 alternating layers of Si, SiO2 directly deposited on Si microreactor62.27 (0.457)10730.5840.58413.70.04262a10  m×10  mYesYes
Absorber: vertically aligned multiwall carbon nanotubes, emitter: 1-D PhC of Si, SiO2 on Si272.25 (0.55)12853.443.4416(48  Wm2)0.2153cAbsorber 0.1  cm2, emitter: 1  cm2<0.5  PaYes
As above, absorber: carbon nanotubes, emitter: 1-D Si/SiO2 PhC; w/cold side rugate filter282.25 (0.55)5bEmitter: 4  cm2, 2 ratios Aemitter/Aabsorber: 12 and 7, filter: 4  cm2<0.5  PaYes
Absorber & emitter each: 2-D photonic crystal made of polycrystalline Taw/HfO2 coating292.3 (0.54)12701.37130  kWm2 (130 suns), if Aa=Ae0.1063a1 h at 1000°C, 144 h at 900°C102<0.3  pa
2-D photonic crystal made of polycrystalline Taw/HfO2 coating302.013274.08181000.181a and 3b21  mm×21  mm>100  h at 900°C, no degradation in structure or opticalperf.30YesNo
2-D photonic crystal made of Ta3%W alloy w/HfO2 coating31,322.25 (0.55)12332.012.01530.03791b and 3-d1  cm×1  cm300 h at 1000°C1, 1 h at 1200°C1041×105  TorrNo
Note: D, diameter; L, length.

Table 5

Details on prototype TPV system demonstrations, part III. Multilayer stacks. D=diameter, L=length.

EmitterCutoff μm/(eV)Temitter (K)Mrad,in (W cm−2)Prad,in (W)Pinput (W)ηri-inputMethodAreaHigh temperature studiesVacuum/inert?Emitter fab. w/heat src?
Combined emitter absorber: multilayer stack w/subwavelength layers of W, yttria-stabilized zirconia331.85 (0.67)16409.7917.3790.2193eAbsorber D: 15 mm, emitter D: 15 mm1.0×105  PaYes
Combined emitter absorber: multilayer stack w/subwavelength layers of Mo, HfO2341.85 (0.67)16406.624.210.26016.23bAbsorber D: 6 mm, Emitter D: 9 mmNo degradation post 1 h at 1423 K, <5×102  Pa, 2 rapid thermal cycles 500–1250 K, but degradation post 1 h at 1473 K1074.6×103  PaYes

Acknowledgments

This work was supported by the U.S. Army Research Office through the Institute for Soldier Nanotechnologies under Contract No. W911NF-13-D-0001, the Micro Autonomous Systems and Technology Collaborative Technology Alliance under Contract No. 892730, the Solid-State Solar-Thermal Energy Conversion Center (S3TEC), an Energy Frontier Research Center funded by the U.S. Department of Energy (DOE), Office of Science, Basic Energy Sciences (BES), under Award No. DE-SC0001299/DE-FG02-09ER46577. R.S. would like to acknowledge Samantha Dale Strasser of the MIT EECS Communication Lab for help with the manuscript.

References

1. 

V. Stelmakh, “A practical high temperature photonic crystal for high temperature thermophotovoltaics,” (2017). Google Scholar

2. 

V. Rinnerbauer et al., “Recent developments in high-temperature photonic crystals for energy conversion,” Energy Environ. Sci., 5 8815 –8823 (2012). https://doi.org/10.1039/c2ee22731b EESNBY 1754-5692 Google Scholar

3. 

N. Pfiester and T. Vandervelde, “Selective emitters for thermophotovoltaic applications,” Phys. Status. Solidi A, 214 (1), 1600410 (2017). https://doi.org/10.1002/pssa.201600410 Google Scholar

4. 

J. Joannopoulos et al., Photonic Crystals, 2nd ed.Princeton University Press, Princeton (2008). Google Scholar

5. 

O. Nielsen et al., “Thermophotovoltaic micro-generator for portable power applications,” in 12th Int. Conf. Transducers, Solid-State Sens., Actuators and Microsyst., 714 –717 (2003). https://doi.org/10.1109/SENSOR.2003.1215573 Google Scholar

6. 

W. 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 (2013). https://doi.org/10.1073/pnas.1301004110 PNASA6 0027-8424 Google Scholar

7. 

L. Fraas, “Small efficient thermophotovoltaic power supply using infrared-sensitive gallium antimonide cells,” (1996). Google Scholar

8. 

E. Horne, “Hybrid thermophotovoltaic power systems,” (2002). Google Scholar

9. 

C. Astle, G. Kovacik and T. Heidrick, “Design and preliminary testing of a prototype thermophotovoltaic system,” in Proc. IMECE, (2003). Google Scholar

10. 

C. Astle, G. Kovacik and T. Heidrick, “Design and performance of a prototype thermophotovoltaic system,” J. Sol. Energy Eng., 129 340 –342 (2007). https://doi.org/10.1115/1.2735351 Google Scholar

11. 

B. Wernsman et al., “Greater than 20% radiant heat conversion efficiency of a thermophotovoltaic radiator/module system using reflective spectral control,” IEEE Trans. Electron Devices, 51 (3), 512 –515 (2004). https://doi.org/10.1109/TED.2003.823247 IETDAI 0018-9383 Google Scholar

12. 

W. Yang et al., “Research on micro-thermophotovoltaic power generators with different emitting materials,” J. Micromech. Microeng., 15 S239 (2005). https://doi.org/10.1088/0960-1317/15/9/S11 JMMIEZ 0960-1317 Google Scholar

13. 

L. Fraas et al., “Antireflection coated refractory metal matched emitters for use with GaSb thermophotovoltaic generators,” in Conf. Rec. 28th IEEE Photovoltaic Spec. Conf., (2000). https://doi.org/10.1109/PVSC.2000.916059 Google Scholar

14. 

E. Doyle, K. Shukla and C. Metcalfe, “Development and demonstration of a 25 watt thermophotovoltaic power source for a hybrid power system,” (2001). Google Scholar

15. 

A. Datas and C. Algora, “Development and experimental evaluation of a complete solar thermophotovoltaic system,” Prog. Photovoltaics Res. Appl., 21 1025 –1039 (2013). https://doi.org/10.1002/pip.2201 Google Scholar

16. 

C. Ungaro, S. Gray and M. Gupta, “Solar thermophotovoltaic system using nanostructures,” Opt. Express, 23 (19), A1149 –A1156 (2015). https://doi.org/10.1364/OE.23.0A1149 OPEXFF 1094-4087 Google Scholar

17. 

T. Aicher et al., “Development of a novel TPV power generator,” AIP Conf. Proc., 738 71 –78 (2004). https://doi.org/10.1063/1.1841881 APCPCS 0094-243X Google Scholar

18. 

V. Andreev et al., “Solar thermophotovoltaic converters based on tungsten emitters,” J. Sol. Energy Eng., 129 298 –303 (2007). https://doi.org/10.1115/1.2734576 Google Scholar

19. 

K. Qiu, A. Hayden and E. Entchev, “TPV power generation system using a high temperature metal radiant burner,” AIP Conf. Proc., 890 27 –36 (2007). https://doi.org/10.1063/1.2711717 APCPCS 0094-243X Google Scholar

20. 

H. Kolm, “Solar-battery power source,” (1956). Google Scholar

21. 

K. Stone et al., “Testing and modeling of a solar thermophotovoltaic power system,” AIP Conf. Proc., 199 199 –209 (1996). https://doi.org/10.1063/1.49687 APCPCS 0094-243X Google Scholar

22. 

H. Yugami et al., “Solar thermophotovoltaic using Al2O3/Er3Al5Ol2 eutectic composite selective emitter,” in Conf. Rec. 28th IEEE Photovoltaic Specialists Conf., 1214 –1217 (2000). https://doi.org/10.1109/PVSC.2000.916107 Google Scholar

23. 

K. Chen et al., “Small, efficient thermophotovoltaic power supply,” (1999). Google Scholar

24. 

B. Bitnar et al., “Characterisation of rare earth selective emitters for thermophotovoltaic applications,” Sol. Energy Mater. Sol. Cells, 73 221 –234 (2002). https://doi.org/10.1016/S0927-0248(01)00127-1 SEMCEQ 0927-0248 Google Scholar

25. 

S. Bitnar et al., “Practical thermophotovoltaic generators,” Semiconductors, 38 941 –945 (2004). https://doi.org/10.1134/1.1787116 SMICES 1063-7826 Google Scholar

26. 

R. Nelson, “TPV systems and state-of-art development,” AIP Conf. Proc., 653 3 –17 (2003). https://doi.org/10.1063/1.1539359 APCPCS 0094-243X Google Scholar

27. 

A. Lenert et al., “A nanophotonic solar thermophotovoltaic device,” Nat. Nanotechnol., 9 126 –130 (2014). https://doi.org/10.1038/nnano.2013.286 NNAABX 1748-3387 Google Scholar

28. 

D. Bierman et al., “Enhanced photovoltaic energy conversion using thermally based spectral shaping,” Nat. Energy, 1 16068 (2016). https://doi.org/10.1038/nenergy.2016.68 Google Scholar

29. 

V. Rinnerbauer et al., “Metallic photonic crystal absorber-emitter for efficient spectral control in high-temperature solar thermophotovoltaics,” Adv. Energy. Mater., 4 1400334 (2014). https://doi.org/10.1002/aenm.201400334 ADEMBC 1614-6840 Google Scholar

30. 

W. Chan et al., “Enabling efficient heat-to-electricity generation at the mesoscale,” Energy Environ. Sci., 10 1367 –1371 (2017). https://doi.org/10.1039/C7EE00366H EESNBY 1754-5692 Google Scholar

31. 

X. Wang et al., “Prototype of radioisotope thermophotovoltaic system using photonic crystal spectral control,” J. Phys. Conf. Ser., 660 012034 (2015). https://doi.org/10.1088/1742-6596/660/1/012034 JPCSDZ 1742-6588 Google Scholar

32. 

X. Wang, “Toward high efficiency radioisotope thermophotovoltaic system by spectral control,” (2017). Google Scholar

33. 

M. Shimizu, A. Kohiyama and J. Yugami, “High-efficiency solar-thermophotovoltaic system equipped with a monolithic planar selective absorber/emitter,” J. Photonics Energy, 5 053099 (2015). https://doi.org/10.1117/1.JPE.5.053099 Google Scholar

34. 

A. Kohiyama, M. Shimizu and H. Yugami, “Unidirectional radiative heat transfer with a spectrally selective planar absorber/emitter for high-efficiency solar thermophotovoltaic systems,” Appl. Phys. Express, 9 112302 (2016). https://doi.org/10.7567/APEX.9.112302 APEPC4 1882-0778 Google Scholar

35. 

L. Ferguson and L. Fraas, “Matched infrared emitters for use with GaSb TPV cells,” AIP Conf. Proc., 401 169 –179 (1997). https://doi.org/10.1063/1.53279 APCPCS 0094-243X Google Scholar

36. 

L. Ferguson and F. Dogan, “A highly efficient NiO-doped MgO matched emitter for thermophotovoltaic energy conversion,” Mater. Sci. Eng. B, 83 35 –41 (2001). https://doi.org/10.1016/S0921-5107(00)00795-9 MSBTEK 0921-5107 Google Scholar

37. 

L. Ferguson and F. Dogan, “Spectral analysis of transition metal-doped MgO ‘matched emitters’ for thermophotovoltaic energy conversion,” J. Mater. Sci., 37 1301 –1308 (2002). https://doi.org/10.1023/A:1014599924372 JMTSAS 0022-2461 Google Scholar

38. 

G. Guazzoni, “High-temperature spectral emittance of oxides of erbium, samarium, neodymium and ytterbium,” Appl. Spectrosc., 26 (1), 60 –65 (1972). https://doi.org/10.1366/000370272774352597 APSPA4 0003-7028 Google Scholar

39. 

P. Adair and M. Rose, “Composite emitters for TPV systems,” AIP Conf. Proc., 321 245 –262 (1995). https://doi.org/10.1063/1.47028 APCPCS 0094-243X Google Scholar

40. 

A. Licciulli et al., “Sol-gel preparation of selective emitters for thermophotovoltaic conversion,” J. Sol-Gel Sci. Technol., 26 1119 –1123 (2003). https://doi.org/10.1023/A:1020719109969 JSGTEC 0928-0707 Google Scholar

41. 

D. Chubb et al., “Rare earth doped high temperature ceramic selective emitters,” J. Eur. Ceram. Soc., 19 2551 –2562 (1999). https://doi.org/10.1016/S0955-2219(99)00096-5 JECSER 0955-2219 Google Scholar

42. 

M. G. Krishna et al., “Spectral emissivity of ytterbium oxide-based materials for application as selective emitters in thermophotovoltaic devices,” Sol. Energy Mater. Sol. Cells, 59 337 –348 (1999). https://doi.org/10.1016/S0927-0248(99)00049-5 SEMCEQ 0927-0248 Google Scholar

43. 

A. Licciulli et al., “Porous garnet coatings tailoring the emissivity of thermostructural materials,” J. Sol-Gel Sci. Technol., 32 247 –251 (2004). https://doi.org/10.1007/s10971-004-5796-2 JSGTEC 0928-0707 Google Scholar

44. 

R. A. Lowe et al., “Rare-earth garnet selective emitter,” Appl. Phys. Lett., 64 3551 –3553 (1994). https://doi.org/10.1063/1.111221 APPLAB 0003-6951 Google Scholar

45. 

V. Tomer et al., “Selective emitters for thermophotovoltaics: erbia-modified electrosopun titania nanofibers,” Sol. Energy Mater. Sol. Cells, 85 477 –488 (2005). https://doi.org/10.1016/j.solmat.2004.04.019 SEMCEQ 0927-0248 Google Scholar

46. 

R. Teye-Mensah et al., “Erbia-modified electrospun titania nanofibres for selective infrared emitters,” J. Phys. Condens. Matter., 16 7557 –7564 (2004). https://doi.org/10.1088/0953-8984/16/41/032 Google Scholar

47. 

D. Diso et al., “Erbium containing ceramic emitters for thermophotovoltaic energy conversion,” Mater. Sci. Eng. B, B98 144 –149 (2003). https://doi.org/10.1016/S0921-5107(03)00036-9 MSBTEK 0921-5107 Google Scholar

48. 

H. Wang, H. Ye and Y. Zhang, “Preparation and performance evaluation of Er2O3 coating-type selective emitter,” Sci. China Technol. Sci., 57 332 –338 (2014). https://doi.org/10.1007/s11431-014-5456-x Google Scholar

49. 

W. Tobler and W. Durisch, “Plasma-spray coated rare-earth oxides on molybdenum disilicide–high temperature stable emitters for thermophotovoltaics,” Appl. Energy, 85 371 –383 (2008). https://doi.org/10.1016/j.apenergy.2007.07.015 Google Scholar

50. 

B. Lee and Z. Zhang, “Design and fabrication of planar multilayer structures with coherent thermal emission characteristics,” J. Appl. Phys., 100 063529 (2007). https://doi.org/10.1063/1.2349472 JAPIAU 0021-8979 Google Scholar

51. 

H. Sai, Y. Kanamori and H. Yugami, “High-temperature resistive surface grating for spectral control of thermal radiation,” Appl. Phys. Lett., 82 (11), 1685 –1687 (2003). https://doi.org/10.1063/1.1560867 APPLAB 0003-6951 Google Scholar

52. 

H. Sai and H. Yugami, “Thermophotovoltaic generation with selective radiators based on tungsten surface gratings,” Appl. Phys. Lett., 85 3399 –3401 (2004). https://doi.org/10.1063/1.1807031 APPLAB 0003-6951 Google Scholar

53. 

M. U. Pralle et al., “Photonic crystal enhanced narrow-band infrared emitters,” Appl. Phys. Lett., 81 4685 –4687 (2002). https://doi.org/10.1063/1.1526919 APPLAB 0003-6951 Google Scholar

54. 

N. Jovanić, I. Čelanović and J. Kassakian, “Two-dimensional tungsten photonic crystals as thermophotovoltaic selective emitters,” AIP Conf. Proc., 890 47 –55 (2008). https://doi.org/10.1063/1.2711719 APCPCS 0094-243X Google Scholar

55. 

M. Araghchini et al., “Fabrication of two-dimensional tungsten photonic crystals for high-temperature applications,” J. Vac. Sci. Technol. B, 29 061402 (2011). https://doi.org/10.1116/1.3646475 JVTBD9 1071-1023 Google Scholar

56. 

I. Čelanović, N. Jovanić and J. Kassakian, “Two-dimensional tungsten photonic crystals as selective thermal emitters,” Appl. Phys. Lett., 92 193101 (2008). https://doi.org/10.1063/1.2927484 APPLAB 0003-6951 Google Scholar

57. 

Y. Yeng et al., “Enabling high-temperature nanophotonics for energy applications,” Proc. Natl. Acad. Sci. U. S. A., 109 (7), 2280 –2285 (2012). https://doi.org/10.1073/pnas.1120149109 PNASA6 0027-8424 Google Scholar

58. 

V. Rinnerbauer et al., “Large-area fabrication of high aspect ratio tantalum photonic crystals for high-temperature selective emitters,” J. Vac. Sci. Technol. B, 31 (1), 011802 (2013). https://doi.org/10.1116/1.4771901 JVTBD9 1071-1023 Google Scholar

59. 

V. Stelmakh et al., “Sputtered tantalum photonic crystal coatings for high-temperature energy conversion applications,” IEEE Trans. Nanotechnol., 15 (2), 303 –309 (2016). https://doi.org/10.1109/TNANO.2016.2522423 ITNECU 1536-125X Google Scholar

60. 

V. Stelmakh et al., “Fabrication of an omnidirectional 2D photonic crystal emitter for thermophotovoltaics,” J. Phys. Conf. Series, 773 012037 (2016). https://doi.org/10.1088/1742-6596/773/1/012037 JPCSDZ 1742-6588 Google Scholar

61. 

K. Cui et al., “Tungsten-carbon nanotube composite photonic crystals as thermally stable spectral-selective absorbers and emitters for thermophotovoltaics,” Adv. Energy. Mater., 8 1801471 (2018). https://doi.org/10.1002/aenm.v8.27 ADEMBC 1614-6840 Google Scholar

62. 

A. Heinzel et al., “Radiation filters and emitters for the NIR based on periodically structured metal surfaces,” J. Mod. Opt., 47 (13), 2399 –2419 (2000). https://doi.org/10.1080/09500340008230522 JMOPEW 0950-0340 Google Scholar

63. 

H. Qiao et al., “Femtosecond laser direct writing of large-area two-dimensional metallic photonic crystal structures on tungsten surfaces,” Opt. Express, 23 (20), 26617 –26627 (2015). https://doi.org/10.1364/OE.23.026617 OPEXFF 1094-4087 Google Scholar

64. 

S. Lin, J. Moreno and J. Fleming, “Three-dimensional photonic-crystal emitter for thermal photovoltaic power generation,” Appl. Phys. Lett., 83 (2), 380 –382 (2003). https://doi.org/10.1063/1.1592614 APPLAB 0003-6951 Google Scholar

65. 

P. Nagpal et al., “Efficient low-temperature thermophotovoltaic emitters from metallic photonic crystals,” Nano Lett., 8 (10), 3238 –3243 (2008). https://doi.org/10.1021/nl801571z NALEFD 1530-6984 Google Scholar

66. 

M. Garin et al., “Three-dimensional metallo-dielectric selective thermal emitters with high-temperature stability for thermophotovoltaic applications,” Sol. Energy Mater. Sol. Cells, 134 22 –28 (2015). https://doi.org/10.1016/j.solmat.2014.11.017 SEMCEQ 0927-0248 Google Scholar

67. 

K. Arpin, M. Losego and P. Braun, “Electrodeposited 3D tungsten photonic crystals with enhanced thermal stability,” Chem. Mater., 23 4783 –4788 (2011). https://doi.org/10.1021/cm2019789 CMATEX 0897-4756 Google Scholar

68. 

K. Arpin et al., “Three-dimensional self-assembled photonic crystals with high temperature stability for thermal emission modification,” Nat. Commun., 4 2630 (2013). https://doi.org/10.1038/ncomms3630 NCAOBW 2041-1723 Google Scholar

69. 

J. Hao et al., “High performance optical absorber based on a plasmonic metamaterial,” Appl. Phys. Lett., 96 251104 (2010). https://doi.org/10.1063/1.3442904 APPLAB 0003-6951 Google Scholar

70. 

B. Zhang, J. Hendrickson and J. Guo, “Multispectral near-perfect metamaterial absorbers using spatially multiplexed plasmon resonance metal square structures,” J. Opt. Soc. Am. B, 30 656 (2013). https://doi.org/10.1364/JOSAB.30.000656 JOBPDE 0740-3224 Google Scholar

71. 

C. Shemelya et al., “Stable high temperature metamaterial emitters for thermophotovoltaic applications,” Appl. Phys. Lett., 104 201113 (2014). https://doi.org/10.1063/1.4878849 APPLAB 0003-6951 Google Scholar

72. 

W. Li et al., “Refractory plasmonics with titanium nitride: Broadband metamaterial absorber,” Adv. Mater., 26 7959 –7965 (2014). https://doi.org/10.1002/adma.v26.47 ADVMEW 0935-9648 Google Scholar

73. 

D. Woolf et al., “Heterogeneous metasurface for high temperature selective emission,” Appl. Phys. Lett., 105 081110 (2014). https://doi.org/10.1063/1.4893742 APPLAB 0003-6951 Google Scholar

74. 

D. Woolf et al., “High-efficiency thermophotovoltaic energy conversion enabled by a metamaterial selective emitter,” Optica, 5 (2), 213 –218 (2018). https://doi.org/10.1364/OPTICA.5.000213 Google Scholar

75. 

P. Dyachenko et al., “Controlling thermal emission with refractory epsilon-near-zero metamaterials via topological transitions,” Nat. Commun., 7 11809 (2016). https://doi.org/10.1038/ncomms11809 NCAOBW 2041-1723 Google Scholar

76. 

N. Nguyen-Huu, Y.-B. Chen and Y.-L. Lo, “Development of a polarization-insensitive thermophotovoltaic emitter with a binary grating,” Opt. Express, 20 (6), 5882 –5890 (2012). https://doi.org/10.1364/OE.20.005882 OPEXFF 1094-4087 Google Scholar

77. 

Y.-B. Chen and Z. Zhang, “Design of tungsten complex gratings for thermophotovoltaic radiators,” Opt. Commun., 269 411 –417 (2007). https://doi.org/10.1016/j.optcom.2006.08.040 OPCOB8 0030-4018 Google Scholar

78. 

J. Song et al., “1D trilayer films grating with W/SiO2/W structure as a wavelength-selective emitter for thermophotovoltaic applications,” J. Quantum Spectrosc. Radiat. Transf., 158 136 –144 (2015). https://doi.org/10.1016/j.jqsrt.2015.02.002 Google Scholar

79. 

H. Ye, H. Wang and Q. J. Cai, “Two-dimensional VO2 photonic crystal selective emitter,” J. Quantum Spectrosc. Radiat. Transf., 158 119 –126 (2015). https://doi.org/10.1016/j.jqsrt.2015.01.022 Google Scholar

80. 

G. B. Farfan et al., “High-efficiency photonic crystal narrowband thermal emitters,” Proc. SPIE, 7609 76090V (2010). https://doi.org/10.1117/12.849045 PSISDG 0277-786X Google Scholar

81. 

E. Sakr, Z. Zhou and P. Bermel, “High efficiency rare-earth emitter for thermophotovoltaic applications,” Appl. Phys. Lett., 105 111107 (2014). https://doi.org/10.1063/1.4895932 APPLAB 0003-6951 Google Scholar

82. 

S. Molesky, C. Dewalt and Z. Jacob, “High temperature epsilon-near-zero and epsilon-near-pole metamaterial emitters for thermophotovoltaics,” Opt. Express, 21 (S1), A96 –A110 (2013). https://doi.org/10.1364/OE.21.000A96 OPEXFF 1094-4087 Google Scholar

83. 

J.-Y. Chang, Y. Yang and L. Wang, “Tungsten nanowire based hyperbolic metamaterial emitters for near-field thermophotovoltaic applications,” Int. J. Heat Mass Transf., 87 237 –247 (2015). https://doi.org/10.1016/j.ijheatmasstransfer.2015.03.087 Google Scholar

84. 

H. Deng et al., “Metamaterial thermal emitters based on nanowire cavities for high-efficiency thermophotovoltaics,” J. Opt., 16 (3), 035102 (2014). https://doi.org/10.1088/2040-8978/16/3/035102 Google Scholar

85. 

B. Zhao et al., “Thermophotovoltaic emitters based on a two-dimensional grating/thin-film nanostructure,” Int. J. Heat Mass Transf., 67 637 –645 (2013). https://doi.org/10.1016/j.ijheatmasstransfer.2013.08.047 Google Scholar

86. 

C. Wu et al., “Metamaterial-based integrated plasmonic absorber/emitter for solar thermo-photovoltaic systems,” J. Opt., 14 024005 (2012). https://doi.org/10.1088/2040-8978/14/2/024005 Google Scholar

87. 

T. Cao et al., “Mid-infrared tunable polarization-independent perfect absorber using a phase-change metamaterial,” J. Opt. Soc. Am., B30 1580 –1585 (2013). https://doi.org/10.1364/JOSAB.30.001580 JOSAAH 0030-3941 Google Scholar

88. 

J. Bossard and D. Werner, “Metamaterials with angle selective emissivity in the near-infrared,” Opt. Express, 21 (5), 5215 –5225 (2013). https://doi.org/10.1364/OE.21.005215 OPEXFF 1094-4087 Google Scholar

89. 

L. Mo et al., “High-efficiency plasmonic metamaterial selective emitter based on an optimized spherical core-shell nanostructure for planar solar thermophotovoltaics,” Plasmonics, 10 529 –538 (2015). https://doi.org/10.1007/s11468-014-9837-6 1557-1955 Google Scholar

90. 

E. Sakr and P. Bermel, “Thermophotovoltaics with spectral and angular selective doped-oxide thermal emitters,” Opt. Express, 25 (20), A880 –A895 (2017). https://doi.org/10.1364/OE.25.00A880 OPEXFF 1094-4087 Google Scholar

91. 

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

92. 

I. Čelanović et al., “Design and optimization of one-dimensional photonic crystals for thermophotovoltaic applications,” Opt. Lett., 29 (8), 863 –865 (2004). https://doi.org/10.1364/OL.29.000863 OPLEDP 0146-9592 Google Scholar

93. 

M. Modest, Radiative Heat Transfer, 3rd ed.Academic, Oxford (2013). Google Scholar

94. 

Y. Yeng, “Photonic crystals for high temperature applications,” (2014). Google Scholar

95. 

V. Stelmakh et al., “Performance of tantalum-tungsten alloy selective emitters in thermophotovoltaic systems,” Proc. SPIE, 9115 911504 (2014). https://doi.org/10.1117/12.2043696 PSISDG 0277-786X Google Scholar

96. 

C. Schlemmer et al., “Thermal stability of micro-structured selective tungsten emitters,” AIP Conf. Proc., 653 164 –173 (2003). https://doi.org/10.1063/1.1539374 Google Scholar

97. 

N. R. Denny et al., “Effects of thermal processes on the structure of monolithic tungsten and tungsten alloy photonic crystals,” Chem. Mater., 19 4563 –4569 (2007). https://doi.org/10.1021/cm071228p CMATEX 0897-4756 Google Scholar

98. 

N. R. Denny et al., “In situ high temperature TEM analysis of sintering in nanostructured tungsten and tungsten molybdenum alloy photonic crystals,” J. Mater. Chem., 20 (8), 1538 –1545 (2010). https://doi.org/10.1039/B918423F JMACEP 0959-9428 Google Scholar

99. 

S. G. Rudisill, Z. Wang and A. Stein, “Maintaining the structure of templated porous materials for reactive and high-temperature applications,” Langmuir, 28 7310 –7324 (2012). https://doi.org/10.1021/la300517g LANGD5 0743-7463 Google Scholar

100. 

H. J. Lee et al., “Hafnia-plugged microcavities for thermal stability of selective emitters,” Appl. Phys. Lett., 102 (24), 241904 (2013). https://doi.org/10.1063/1.4811703 APPLAB 0003-6951 Google Scholar

101. 

U. Guler, A. Boltasseva and V. Shalaev, “Refractory plasmonics,” Science, 344 263 –264 (2014). https://doi.org/10.1126/science.1252722 SCIEAS 0036-8075 Google Scholar

102. 

V. Rinnerbauer et al., “High-temperature stability and selective thermal emission of polycrystalline tantalum photonic crystals,” Opt. Express, 21 (9), 11482 –11491 (2013). https://doi.org/10.1364/OE.21.011482 OPEXFF 1094-4087 Google Scholar

103. 

V. Stelmakh et al., “Thick sputtered tantalum coatings for high-temperature energy conversion applications,” J. Vac. Sci. Technol. A, 33 061204 (2015). https://doi.org/10.1116/1.4935156 JVSTAL 0022-5355 Google Scholar

104. 

V. Stelmakh et al., “High-temperature tantalum tungsten alloy photonic crystals: stability, optical properties, and fabrication,” Appl. Phys. Lett., 103 123903 (2013). https://doi.org/10.1063/1.4821586 APPLAB 0003-6951 Google Scholar

105. 

P. Nagpal et al., “Fabrication of carbon/refractory metal nanocomposites as thermally stable metallic photonic crystals,” J. Mater. Chem., 21 10836 –10843 (2011). https://doi.org/10.1039/c1jm10997a JMACEP 0959-9428 Google Scholar

106. 

D. Peykov et al., “Effects of surface diffusion on high temperature selective emitters,” Opt. Express, 23 (8), 9979 –9993 (2015). https://doi.org/10.1364/OE.23.009979 OPEXFF 1094-4087 Google Scholar

107. 

M. Shimizu, A. Kohiyama and H. Yugami, “Evaluation of thermal stability in spectrally selective few-layer metallo-dielectric structures for solar thermophotovoltaics,” J. Quant. Spectrosc. Radiat. Transfer, 212 45 –49 (2018). https://doi.org/10.1016/j.jqsrt.2018.02.037 JQSRAE 0022-4073 Google Scholar

108. 

R. Nelson, “A brief history of thermophotovoltaic development,” Semicond. Sci. Technol., 18 S141 –S143 (2003). https://doi.org/10.1088/0268-1242/18/5/301 SSTEET 0268-1242 Google Scholar

109. 

B. Bitnar, W. Durisch and R. Holzner, “Thermophotovoltaics on the move to applications,” Applied Energy, 105 430 –438 (2013). https://doi.org/10.1016/j.apenergy.2012.12.067 APENDX 0306-2619 Google Scholar

110. 

“Lithium-ion battery maintenance guidelines,” Google Scholar

Biographies of the authors are not available.

© 2019 Society of Photo-Optical Instrumentation Engineers (SPIE)
Reyu Sakakibara, Veronika Stelmakh, Walker R. Chan, Michael Ghebrebrhan, John D. Joannopoulos, Marin Soljacic, and Ivan Čelanović "Practical emitters for thermophotovoltaics: a review," Journal of Photonics for Energy 9(3), 032713 (26 February 2019). https://doi.org/10.1117/1.JPE.9.032713
Received: 16 October 2018; Accepted: 5 February 2019; Published: 26 February 2019
Lens.org Logo
CITATIONS
Cited by 96 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Silicon carbide

Photonic crystals

Solar cells

Silicon

Aluminum

Hybrid fiber optics

Prototyping


CHORUS Article. This article was made freely available starting 26 February 2020

Back to Top