Skip to main content

REVIEW article

Front. Energy Res., 01 June 2023
Sec. Process and Energy Systems Engineering
Volume 11 - 2023 | https://doi.org/10.3389/fenrg.2023.1124288

Functional materials for solar thermophotovoltaic devices in energy conversion applications: a review

www.frontiersin.orgModupeola Dada* www.frontiersin.orgPatricia Popoola www.frontiersin.orgAlice Alao www.frontiersin.orgFolasayo Olalere www.frontiersin.orgEvlly Mtileni www.frontiersin.orgNtanzi Lindokuhle www.frontiersin.orgMakinita Shamaine
  • Chemical, Metallurgical and Materials Engineering, Tshwane University of Technology, Pretoria, South Africa

Fossil fuels are now used to meet over 80% of the world’s energy demands, but they have the disadvantages of being unsustainable economically and polluting the environment. Solar energy is also one of the most desired alternative forms of renewable energy due to the quantity of direct sunlight among these sources. Due to the difficulties with solar cells, less than 1% of this energy is harvested and transformed into electricity. Notably, solar thermal and photovoltaic systems are the traditional methods for converting solar energy into electricity. It can be challenging to turn the solar energy captured by these systems into power. In contrast to conventional conversion methods, which involve converting solar energy directly into electricity, this article conducts a thorough investigation of solar thermophotovoltaic devices and the high-tech materials used in solar thermophotovoltaic systems as a solution to the conversion challenges.

Introduction

In 2019, the global energy consumption growth rate declined by +0.6% compared to its ever-increasing trend, contributing to slow economic growth. Consumption in Algeria and Indonesia was dynamic; however, consumption in South Africa, Saudi Arabia, and Nigeria continued to increase (Xu G. et al., 2020; Zhou et al., 2020). Generally, energy consumption is a socio-economic human need, and the energy demand is met by fossil fuels such as carbon and hydrogen compounds, from which petroleum, natural gas, and coal are derived (Sharma and Ghoshal, 2015; Pareek et al., 2020). Coal was one of the first fossil fuels used for steam engines, transportation, and the production of steel, while petroleum was used for fuel in combustion engines and lighting paraffin lamps, and natural gas was used for cooking and to generate electricity (Kalair et al., 2021; Welsby et al., 2021). Nonetheless, the adverse effects of these fossil fuels include, but are not limited to, the emission of nitrous oxide (N2O) and carbon monoxide (CO). Inhalation of CO causes dizziness and headaches that may lead to death (Vohra et al., 2021; Abbasi et al., 2022). N2O, on the other hand, generates ground-level ozone, which is harmful to crops and the respiratory system. Moreover, oil and coal contain sulfur, which, in contact with moisture, forms sulfuric acid, resulting in very damaging acid rain (Guenet et al., 2021; Qasim et al., 2021). Greenhouse gases are also one of the most destructive emissions of fossil fuels, causing global warming and disrupting the surface temperature of the Earth from sustaining life; consequently, alternative energy sources have been developed (Malhotra, 2020; Shen et al., 2020). This study reviews the innovation in renewable energy sources that are cleaner, more accessible, and derived from natural sources. These sources include wind, geothermal, hydrogen, hydroelectric, biomass, ocean, and solar energy.

Solar energy

The primary source of solar energy, which travels at 3.0 × 108 m per second, is the Sun. The Sun, made up of helium gas and hydrogen, makes this energy in its core through fusion (Kalogirou, 2013; Gong et al., 2019). Fusion involves hydrogen isotopes, and with the transformation of matter, it comes together to form helium atoms, and this transformed matter is given off as radiant energy by the Sun (Okutsu et al., 2021; Kenjo et al., 2022). Radiant energy emitted from the Sun reaches the Earth surface in tiny portions, approximately 1.7 × 1018 W, and these rations are enough to supply the energy needed on Earth (Sisay, 2022). The Sun supplies energy to different parts of the Earth in small fractions per time, making it necessary to capture the solar energy through solar collectors before transforming it into electricity, as shown in Figure 1 (Panwar et al., 2011; Zhang L. et al., 2021).

FIGURE 1
www.frontiersin.org

FIGURE 1. Solar energy.

Conversion of solar energy into electricity

A solar cell built from semiconductor materials is one device that is electronically a collector to convert solar energy into electricity (Fukuda et al., 2020; Kim J. Y. et al., 2020). The material used for solar cells must absorb sunlight, which raises electrons in the light to higher-energy states, and then the high-energy electron moves from the cell to an external circuit (Fonash, 2012; Burlingame et al., 2020). Urbanization is constantly increasing global electricity consumption, and like other types of energy, the need to reduce the price of electricity increases its supply, performance, and storage, making solar cell devices a fundamental solution (Asongu et al., 2020; Zhang et al., 2022). Solar thermal systems and photovoltaics are two methods of converting solar energy into electricity (Pasupathi et al., 2020; Rashidi et al., 2020). Solar thermal systems comprise concentrated solar power, which uses solar energy to generate electricity (Javadi et al., 2020; Osorio et al., 2022). The process involves using a solar collector with a mirrored surface to direct sunlight into a standby receiver, which, in turn, heats a liquid. The heated liquid produces steam, which produces electricity (Peuser et al., 2013; Ndukwu et al., 2021). The photovoltaic process of generating electricity involves the use of solar cells made up of silicon, which supplies electricity when the radiant energy from sunlight strikes the cell, triggering the electrons in the cell to move, and this movement of electrons jerks an electric current, switching from solar energy to electricity (Grätzel, 2005; Prabhu and ValanArasu, 2020).

Limitations to the conversion methods

There is an absolute theoretical Shockley–Queisser (SQ) limitation to the efficiency of conventional solar cells (Shockley and Queisser, 1961; Markvart, 2022). The conversion of solar energy to electricity, as shown in Figure 2 by solar cells is established by the photoelectric effect, which is an interaction between the transformed matter and the electromagnetic wave (Guillemoles et al., 2019; Ehrler et al., 2020). During their study, Shockley and Queisser (1961) realized there was a mismatch between the emission angles and absorption; therefore, they proposed that the Sun and solar cells act as black bodies. To this effect, a single layer of solar cells consisting of silicon was detailed through emission angle restrictions, photon recycling, and optical concentration to having an upper limit of a little above 32% for a 1.1eV gap (Lu et al., 2021; Chen et al., 2022).

FIGURE 2
www.frontiersin.org

FIGURE 2. Schematic representation of the thermophotovoltaic process.

The SQ limit defines 32%–33.6% as the maximum solar energy conversion efficiency achievable for any solar cell material (Xiang et al., 2019; Kim S. et al., 2020). This limitation, which was developed in 1961, is applicable to the principle of detailed balancing, which equates the photon flux that goes into the solar cell device with the electron or photon flux that goes out of it at different open-circuit conditions (Rühle, 2016; Green and Ho-Baillie, 2019).

Exploring solutions to the theoretical limitations

In recent times, possible ways of increasing the efficiency of solar cells above the absolute limit have been found, namely, by adding multiple layers of solar cells, which increases the incident intensity, the current density, and the voltage (Kim J. Y. et al., 2020; Park, 2020). Angle restriction filters can also be used to reduce the existing recombination current; multiple semiconductors with several bandgaps can also be used to decrease thermal losses and increase efficiency (Beard et al., 2013; Tennyson et al., 2019). Axelevitch (2018) reviewed the ways of improving the efficiency of single-junction solar cells, with specific attention given to solar cells enhanced with the plasmon. The author described using multi-junction solar cells, down-conversion solar cells, up-conversion solar cells, multiple exciton generation solar cells, solar cells with intermediate bands, and hot carrier solar cells as enhancement mechanisms of solar cells from the SQ limitations (Krügener et al., 2021; Yao and Hou, 2022). The possibility of using nano-structures made up of gold or silver nanoparticles was also discussed, concluding that the combination of an up-converter and a plasmon is a promising solution to the SQ limitation (Chen et al., 2021; Singh and Jen, 2021). The plasmon with extreme energy photons will generate multiple charged carriers under the absorption of one photon, while the up-converter uses the wavelength photons to increase the efficiency of solar cells (Gerislioglu et al., 2019; Huang et al., 2021). Nonetheless, a preferred alternative for exceeding the SQ limitations is the conversion of solar energy to heat first before generating electrical power through solar thermophotovoltaic devices, as shown in Figure 3.

FIGURE 3
www.frontiersin.org

FIGURE 3. Electric power through solar thermophotovoltaic devices.

Solar thermophotovoltaics and their devices

There are several technological options for converting primary energy into electricity. A few of them may be converted directly (for example, PV and fuel cells), but the vast majority require the intermediary creation of heat, which is then converted into electricity using a heat engine (Snyder and Toberer, 2008; Shi et al., 2020). Consequently, many other types of heat engines have been invented, but only those based on solid state devices (thermoelectrics and thermionics) have been widely employed, particularly in the energy and aerospace sectors (Mahan and Sales, 1997; Chen et al., 2003), while the dynamic systems (Rankine, Stirling, and Brayton) are still under development until they demonstrate high levels of reliability.

In thermoelectrics, thermal energy is directly converted into electrical energy via thermoelectric modules, which are solid state devices (Snyder and Ursell, 2003; Pei et al., 2012). The “Seebeck effect,” which is the appearance of an electrical voltage induced by a temperature gradient across a material, lies at the heart of the conversion process. The inverse of this is the “Peltier effect,” which causes a temperature gradient to form when voltage is applied (Tritt, 2011; Pourkiaei et al., 2019). As a result, the performance of the thermoelectric (TE) device is directly dependent on the temperature gradient (DT), the thermoelectric figure of merit (ZT), and the material parameter (Jia et al., 2021; Mao et al., 2021). The thermoelectric efficiency is defined for power generation by combining the Carnot efficiency (DT/Thot) (Zou et al., 2020; Zhang Z. et al., 2021). To increase this efficiency, high ZT values and a significant temperature differential across the thermoelectric material are required (Yang et al., 2021). Mahan and Sofo (1996) studied the electronic structure required to provide a high figure of merit in thermoelectrics, and it was discovered that a delta-shaped transport distribution maximizes thermoelectric properties. Their result indicates that for maximum thermoelectric efficiency, a narrow distribution of the energy of the electrons involved in the transport process is required. Nonetheless, according to Sootsman et al. (2009), current thermoelectric devices have a ZT of 0.8 and function at only around 5%–6% efficiency. By raising ZT by a factor of 4 and depending on DT, the estimated efficiency rises to 30%. However, the difficulty in developing high-ZT thermoelectric materials is attaining high electronic conductivity (s), high thermoelectric power (S), and low thermal conductivity (k) in the same solid. These characteristics are governed by the specifics of the electronic structure and the dispersion of charge carriers (electrons or holes) and so cannot be controlled independently. Dresselhaus et al. (2007) discovered a simultaneous increase in power factor and decrease in thermal conductivity using nanocomposites when compared to alloy samples of the same chemical makeup; Nandihalli et al. (2020) studied polymer-based thermoelectric nanocomposites; the level of material performance for output power factor PF = σS2 and energy conversion efficiency was determined through the dimensionless figure of merit ZT = σS2T/k (σ,S,T,andK are the electrical conductivity, Seebeck coefficient, temperature, and thermal conductivity, respectively). The authors concluded that the trade-off relationships between conductivity and Seebeck coefficient in polymer-based materials, as well as in inorganic thermoelectric (TE) materials, limit the ability to improve TE performance. These issues can be solved, however, by modifying the interfaces between the polymer and inorganic or organic additives. As a result, a suitable manufacturing process is required, in which interfacial density and defects at the interface of nanocomposites may be regulated in order to enhance electrical conductivity and the Seebeck coefficient at the same time. Therefore, significant interest from researchers in developing advanced thermoelectric properties through organic and inorganic nanomaterial-based hybrid nanocomposites has gained popularity among researchers (Bisht et al., 2021).

Compared to thermoelectrics, thermionic energy converters are power generators and their thermal management process is shown in Figure 4. The thermionic converter uses heat as its source of energy and transfers energy through mechanical work at no point during its operation. As a result, it is classified as a fuel cell or a photovoltaic cell (Herring and Nichols, 1949; Schwede et al., 2010).

FIGURE 4
www.frontiersin.org

FIGURE 4. Thermal management process.

Thermophotovoltaics (TPV) is a solid state alternative to thermoelectric and thermionic converters, which is very efficient. The two most prevalent methods for harnessing solar energy are photovoltaic, in which sunlight directly excites electron–hole pairs in a semiconductor, and the solar-thermal technique, in which sunlight powers a mechanical heat engine (Daneshvar et al., 2015; Burger et al., 2020). Photovoltaic systems directly convert solar radiation into electricity and are used to generate power from solar energy. Another approach uses a solar collector to convert Sun photons to thermal energy, which is then used in a thermal engine to generate electricity (Joshi et al., 2009; Lupangu and Bansal, 2017).

Photovoltaic power generation is intermittent and only efficiently utilizes a fraction of the Sun spectrum, but the inherent irreversibilities of tiny heat engines make the solar-thermal technique best suited for utility-scale power facilities. As a result, hybrid solutions for solar power generation are becoming increasingly important. Solar thermophotovoltaics (STPV) can also be referred to as systems that collect and re-emit solar light as heat radiation before directing it to photovoltaic cells (Hosenuzzaman et al., 2015; Das et al., 2018). With the STPV system, optimization may involve selecting an emitter spectrum and reflecting the unused portion of the radiation from the receiver back to the emitter surface. STPV systems’ primary features are modularity, portability, the absence of moving parts, pollution-free operation, great efficiency, and high-power densities. Typically, the goal is to enhance photovoltaic conversion efficiency by matching the spectrum of the light to the bandgap of the cell (Davies and Luque, 1994; Wang et al., 2019). Solar thermophotovoltaics promise to leverage the benefits of both optimal approaches by converting sunlight into thermal emission tuned to energies that are directly above the photovoltaic bandgap. This is achieved by using a hot absorber–emitter with high efficiency and harnessing the entire solar spectrum, scalability and compactness due to their solid state nature, and dispatchability due to the ability to store energy using thermal or chemical means (Kohiyama et al., 2016; Chen et al., 2020).

STPV systems have a variety of possible applications, including electrical and thermal energy supply, grid-independent storage appliances, waste-heat recovery, and space and aerospace power applications. However, at high working temperatures, effective sunlight collection in the absorber and spectrum control in the emitter are extremely difficult (Datas and Martí, 2017; Gupta et al., 2018); because of this limitation, earlier experimental demonstrations have had a conversion efficiency of approximately or below 1%. Hence, a number of factors influence solar thermophotovoltaic efficiency, including the sunlight concentration ratio, absorber/emitter temperature/efficiency, photon recycling efficiency, and TPV cell characteristics (Bitnar et al., 2013; Zhou et al., 2016).

In a study of STPV systems, a newly developed Fresnel lens was used, and the calculated system efficiency with a tungsten emitter achieved 35%. The system comprised a solar concentrator, an absorber emitter, a spectrum filter, PV cell arrays, a bottom reflector, and a cooling system. A parabolic reflector was utilized as the concentrator in this case. The emitter might be composed of grayscale or spectrum-selective material. Forty GaSb cells were linked in series and positioned on the cooling system’s inner surface. The bottom surface totally reflected heat radiation, ensuring that it was absorbed completely by the cells (Zenker et al., 2001; Khvostikov et al., 2007). The solar collector collects the Sun’s rays, which subsequently radiate to the absorber surface area. The absorber was a critical component of the STPV system, acting as a spectrum-selective surface that transforms solar light into thermal energy. This surface increases solar energy absorption in the visible and near-infrared ranges while minimizing heat emission in the infrared region (Chaudhuri, 1992; Nam et al., 2014).

Hence, the absorber surface must function at elevated temperatures with a high energy density to achieve thermal stability. Another selective surface that radiates heat energy toward the PV cell is the emitter. According to the SQ limit, a single-junction solar PV has a significant limitation in utilizing solar energy because photons with less than the required energy to generate electron–hole pairs are practically useless, while photons with more energy than the bandgap result in excess energy at the PV surface and reduce efficiency (Mojiri et al., 2013; Abbas et al., 2022). As a result, by customizing the bandgap, emitters give the most efficient spectrum for the PV cell.

Another innovative STPV system offers electrical energy that is passed via the solid oxide electrolyzer cell (SOEC) to create hydrogen. The ultimate purpose of an electrolyzer cell is to create hydrogen from the intake water. Heat and power are required for the electrolyzer cell to perform this feat. Thus, a steam electrolyzer that runs at high temperatures is an effective way to significantly cut energy use (Charache et al., 1999; Dashiell et al., 2006). Consequently, the efficiency of numerous intermediary energy conversion processes is critical to the performance of STPV. The absorber converts optically focused sunlight into heat, the absorber temperature increases, heat conducts to the emitter, and the hot emitter thermally radiates toward the photovoltaic cell, where radiation is eventually harnessed to excite charge carriers and create electricity (Kohiyama et al., 2020; Hou et al., 2023).

The overall efficiency η stpv can be expressed as a product of the optical efficiency of concentrating sunlight (η o), the thermal efficiency of converting and delivering sunlight as heat to the emitter (η t), and the efficiency of generating electrical power from the thermal emission (η tpv) (Tervo et al., 2020):

ηstpv=ηoηtηtpv.

The TPV efficiency η tpv hinges on the spectral properties and temperature of the emitter. A spectrally selective emitter should have a high emittance for energies above the photovoltaic bandgap (Eg) and a low emittance for energies below the bandgap. To excite enough thermal modes for substantial emission above the bandgap, the emitter temperature should ideally be high enough that Planck’s blackbody peak coincides with the bandgap; in other words, by Wien’s displacement law (Tian et al., 2021)

Teopt2336KeV1·Eg.

The emitter’s high-temperature operation faces two major obstacles to a successful STPV power conversion: efficiently collecting sunlight to meet Teopt and preserving spectral selectivity at raised temperatures. For the absorber, one approach to effectively enhance the intrinsic solar absorptivity of materials is to use macrocavity geometries. Because of the high aspect ratio of the cavity needed to enhance absorption, this approach typically requires high levels of optical concentration to reach Teopt. A high optical concentration necessitates sophisticated systems with poor optical efficiencies (η o ≈ 65%). Tungsten has poor inherent spectrum selectivity as an emitter at Teopt because its emissivity at low photon energies (<Eg) increases with temperature, accompanied by an increase in electrical resistance. Ultimately, depending on the intrinsic spectral properties of materials for their absorber–emitter performance is limited with experimental STPV conversion efficiencies. Hence, to improve the performance of the absorber–emitter, the design of structured surfaces with spectral properties approaching those of ideal STPV components is achieved with simulation studies using realistic nanophotonic surfaces, which predict STPV efficiencies exceeding 40% (Bhatt et al., 2020).

Solar thermal generators, which offer very high energetically dense thermal storage, are an eventual alternative to solar PV because an STPV system requires a broadband absorber capable of absorbing the entire solar spectrum, a narrowband emitter capable of converting the absorber’s absorbed energy in the form of heat to photons in a narrowband spectrum just above the bandgap of the PV cell, and a low bandgap PV cell capable of effectively generating electron–hole pairs while avoiding thermalization losses (Xu Y. et al., 2020).

The broadband absorber and narrow band emitter create an intermediate structure that may be built with metamaterials and their two-dimensional metasurface equivalents (Liu et al., 2019). Metasurfaces have the special benefit of being compact and having a high absorptance. To produce strong absorptance across a wide range of incidence angles, both the absorber and the emitter should be angle-insensitive (Anggraini et al., 2022). The major goal in both circumstances is to maximize the absorptance profile since, according to Kirchhoff’s Law, absorptance equals emittance in thermodynamic equilibrium; furthermore, because the STPV system must function under high-temperature settings, the materials used to construct the absorbers and emitters are carefully chosen from a limited number of refractory materials (Azzali et al., 2021; Bendelala and Cheknane, 2022). Nevertheless, oxidation is a significant issue when employing refractory materials (Rana et al., 2021). Most refractory metals with high melting points oxidize at significantly lower temperatures and pressures. Tungsten (W), rhenium (Re), tantalum (Ta), and titanium nitride (TiN) are refractory metals with extremely high melting temperatures of 3695.15 K, 3458.15 K, 3290.15 K, and 3203.15 K, respectively. Ongoing research on absorbers, emitters, and STPV systems in general is hampered by non-planar three-dimensional (3D) or multilayer designs and reduced efficiency (Tong, 2018). Fabrication of 3D designs is inherently complex for the primary lens (or mirror), the absorber, the PV cell, and a photon recuperator mechanism, which are the key components of TPV systems (Palfinger, 2006).

Hence, the TPV efficiency is optimized by optimizing three parameters: absorber, PV cell temperatures, and cell voltage. When photons of above-bandgap energy released by the heat emitter are absorbed by the PV cell, the photovoltaic effect produces electron-hole pairs (EHPs) (i.e., photogeneration) (Zou et al., 2020; Ren et al., 2021; Zheng et al., 2021). When EHPs in a PV cell are recombined, photons with energies larger than the bandgap are released (i.e., radiative recombination). The above-bandgap spectral radiation released by the PV cell quantum efficiency (Qe) is mostly the result of radiative recombination (Prentice, 1999). Although Qe comprises a free carrier and lattice emission in the PV cell, its contributions are unimportant to photogeneration or radiative recombination. The photocurrent flows through the external circuit as EHPs produced by photogeneration and radiative recombination undergo charge separation and migrate to electrodes (Sahoo and Mishra, 2018). Charge carriers that are lost due to recombination (non-radiative, radiative, and surface recombination) are unable to contribute to photocurrent production. Internal quantum efficiency is defined as the ratio of the number of created EHPs that may be employed for photocurrent production to the number of absorbed photons (Goodnick and Honsberg, 2022).

Advanced NF-TPV device ideas, particularly hybrid devices with NF-TPV integrated with a thermionic device or light-emitting diode and NF-TPV devices with multi-junction PV cells, are gaining research interest (Song et al., 2022). The trade-off between resistance and shading losses in the frontside is mitigated by serial integration of thermionic and thermophotovoltaic devices as a means of resolving the difficulties in an NF-TPV device. The near-field thermophotovoltaic (nTiPV) device with a thermionic cathode [i.e., LaB6] and an anode [i.e., BaF2] is positioned on the emitter and PV cell sides of the NF-TPV device, respectively (Datas and Vaillon, 2021). Electrons are emitted from the heated cathode and transmitted to the anode through the thermionic effect in nTiPV devices, along with photons. Electrons passing through the anode and reaching the top side of the PV cell negate the accumulated holes caused by the photovoltaic effect (Qiu et al., 2022). As a result, an electrical connection is created via the vacuum, eliminating the necessity for a frontside contact electrode. Additionally, electrons arriving at the PV cell’s backside electrode are given to the emitter cathode, releasing the extra-potential of thermionic and photovoltaic effects. The elimination of the trade-off between series resistance and shading losses opens the door to significantly improved performance, even for scaled NF-TPV devices (Song et al., 2022). The space charge effect creates an electrostatic potential within the vacuum gap, acting as a potential barrier to thermionic transfer. When the vacuum gap approaches the near-field regime, not only is photon transport accelerated but also thermionic electron transport since the electric potential barrier is reduced (Khalid et al., 2016; Xiao et al., 2017).

As a result, with no frontside contact electrode, increased photocurrent from evanescent modes may be successfully transported to the emitter side cathode. At a realizable vacuum gap of 100 nm and an emitter temperature of 1000 K, the power density of the nTiPV device was 3.7 times and 10.7 times higher than that of the NF-TPV device in the same operation condition with ‘ideal (i.e., series resistance is 0 mΩ)' and ‘realistic (i.e., series resistance is 10 mΩ)' conditions, respectively. With a 1-cm2 macro-scale device area with the same vacuum gap and emitter temperature parameters, the nTiPV device had a power density of 6.73 W/cm2 and 18% conversion efficiency (Jang et al., 2016). As a result of the trade-off between resistive and shading losses, the nTiPV device may be promising for the macro-size application of NF-TPV conversion. Fabrication of the anode and cathode with sufficiently low work-functions, which is required for improved thermionic emission in the near-field domain, still remains a problem for the specified application (Song et al., 2022).

In contrast to PV cells, a light-emitting diode (LED) produces luminescence by radiative recombination of bias-driven injected charge carriers. If the TPV device’s passive emitter is replaced with an active emitter, such as an LED, the PV cells’ above-bandgap absorption can be boosted due to the electroluminescence created by the LED. A PV-LED hybrid device of this type is known as a thermophotonics (TPX) device. If the nearfield idea is implemented in the TPX device, it is evident that the evanescent mode will improve both PV and LED performances, resulting in a better power density. This gadget works by driving the LED with a portion of the electricity provided by the PV cell (Sadi et al., 2020).

A near-field TPV (nTPX) device used AlGaAs ternary semiconductors in both the PV cell and the LED. Using rigorous balancing analysis, they simulated the photon flow that conveys chemical potential and the non-radiative recombination rate. When the LED and PV cell temperatures were 600 and 300 K, respectively, at the 10-nm vacuum gap, their nTPX device could create 9.6 W/cm2 of electrical power with a conversion efficiency of 9.8%. With the same temperature difference, the suggested nTPX device exceeded the record-high power output density and conversion efficiency of a conduction-based thermoelectric generator. A TPX device’s performance is generally more sensitive to emitter side temperature (i.e., LED temperature) than a TPV device. First and foremost, bandgap alignment between the LED and PV cells is required for high-performance TPX devices. Since the bandgap energy depends on the temperature, the performance of TPX devices could be degraded even if the temperature of the LED slightly deviates from the optimum point. In addition, because parameters of the non-radiative recombination rate, such as Auger and SRH recombination lifetimes (𝜏Auger and 𝜏SRH) and intrinsic carrier concentration (𝑛𝑖), are also functions of the temperature, the increase in LED temperature can result in detrimental effects on the TPX performance; in other words, as the LED temperature increases, 𝜏Auger and 𝜏SRH decrease and 𝑛𝑖 increases (Legendre and Chapuis, 2022).

High-tech materials for solar thermophotovoltaic devices

Conventional photovoltaic materials convert solar energy directly to electricity; hence, they undergo theoretical limitations. However, ways of increasing the overall efficiency are to convert solar energy first to heat and then use the heat to generate electrical energy (Nevala et al., 2019, Hassan et al., 2020). Table 1 shows the solar thermophotovoltaic devices and the devices used for this application are referred to as solar thermophotovolataics (STPV). Devices used for this application are referred to as solar thermophotovoltaics (STPV), with the flowchart shown in Figure 5. STPV pairs low-efficiency conventional solar cells with an added layer of high-tech material that helps in multiplying the SQ limitation, making it possible for the cells to generate more power (Alam and Khan, 2019; Wong et al., 2020). The device works on the principle of dispersing waste solar energy as heat inside the solar cell, a by-product of the nuclear/chemical reactions or mechanical work (Jayawardena et al., 2020; Gong et al., 2021). The heat is then absorbed by the transitional component at temperatures that will allow this component to produce thermal radiation. The configurations of the cell and the high-tech materials used in the devices are fine-tuned to the right wavelengths for the cell to capture light, which improves its efficiency (Guillemoles et al., 2019).

TABLE 1
www.frontiersin.org

TABLE 1. Literature review on solar thermophotovoltaic devices.

FIGURE 5
www.frontiersin.org

FIGURE 5. STPV and its energy flow flowchart.

According to Jayawardena et al. (2020), perovskite cells mixed with lead–tin as an absorber can achieve a fill factor above 80% by post-treating the absorber with guanidinium bromide. The authors showed that the post-treatments were favorable in aligning the cathode and anode interfaces, thus enabling a bipolar extraction, which resulted in the device having an area of 0.43 cm2, a fill factor above 80% and 14.4% power conversion efficiency. Rau et al. (2005) proposed using fluorescent collectors with photonic structures, which act as an omnidirectional spectral band stop-filter, to enhance the efficiency of photovoltaic solar cells, and they concluded that the combination of fluorescent collectors with photonic structures can close the theoretical SQ limitation while saving about 99% of the solar cell material. Thus, the authors recommended more research should be focused on the potential of two- to three-dimensional photonic structures used with fluorescent collectors. Briggs et al. (2013) developed an up-converter solar cell using thermodynamics to exceed the SQ limitation. The results showed that the efficiency of the solar cell increased from 28% to 34% with an increase in the up-converter quantum yield and capacity. Jia et al. (2016) reported the use of silicon nanocrystals built into the dielectric matrix as a promising high-tech material for solar cells. Conversely, the material did not exceed the SQ limitation; therefore, the authors investigated the maximum efficiency of the material, and they stated that the practical limit of the solar cell’s efficiency was 32%. Based on these results, they made suggestions for further studies to figure out the reason for the limitation and the proposed solution for improving the performance of the cell.

Trupke et al. (2002) tried generating a multiple electron–hole pair, a down-conversion high-energy photon, in enhancing the efficiency of solar cells. The authors detailed that there was an increment in the efficiency of the solar cell from 30.9% to 39.63%. They investigated the reason why intermediate-band solar cells could not exceed the SQ limitation, and they realized that the restriction was attributed to the radiative recombination through the intermediate band, yet they stated that suppressing the radiative recombination by introducing a quantum ratchet to the intermediate band can enhance the solar cell efficiency; therefore, the authors recommended using quantum ratchets as a more effective alternative to single-gap solar cells. On the other hand, Wang et al. (2013) used carrier transport and photon recycling simulation to figure out the reasons why thin-film GaAs solar cells did not exceed the SQ limitation by studying a single-junction thin-film solar cell and the influence of the design parameters. The authors concluded that an increment in efficiency will occur via enhancements on the backside mirror reflectivity above 95%, thus naming the series resistance and the back mirror reflectivity as the two important factors to focus on when creating high-efficiency thin-film solar cells, while Schaller et al. (2006) showed that charge carriers in large portions can enhance the performance of solar cells by increasing the photon to exciton conversion by 700%.

Xu et al. (2015) used nanostructured solar cells as photovoltaic devices, which under a 1.5 solar illumination showed a maximum efficiency of about 42%. However, they reported that the device did not exceed the theoretical limit for planar devices with optical concentrations, even though it exceeded the SQ limitations for non-planar devices. The authors attributed the failure to reach the SQ limitations to the principle of detailed balance with good knowledge of the absorption in the device structure. They recommended that nanostructured devices should be developed with limited absorption at wavelengths and angles very different from the incident illumination. More so, the improvement of the devices should come from the open-circuit voltage with non-radiative recombination and good-quality surface passivation. Mann et al. (2016) showed that large absorption of the cross sections is not responsible for the enhancements of solar cells using nanophotonic materials for photovoltaics; however, increasing the directivity bounds, which consist of the nanoscale concentrations in macroscopic solar cells, and the maintenance of high short-circuit currents are the significant voltage enhancement factors. Bierman et al. (2016) developed a high-tech nanophotonic crystal that was used to emit the desired wavelengths of light when heated while integrated into a system vertically aligned with carbon nanotubes, which serve as absorbers while operating at about 1,000 . When the crystal was heated, it continued to give out light that matches the band of wavelengths that the solar cell uses to convert to electric current. The carbon nanotube enables all the energy of the photons to get converted to heat, which, in turn, re-emits light that matches the solar cell’s peak efficiency through the nanophotonic crystal. The authors, using an absorber, solar stimulator, photovoltaic receiver, and filter all in one device, argue that a device coupled with a thermal storage system could provide continuous on-demand electrical power. They recommended further studies on increasing the current laboratory size of the device for commercial purposes (Chan et al., 2013; Davids et al., 2020).

Future recommendations

The literature has shown that converting solar energy into heat and then from heat into electricity is an efficient way of exceeding the SQ limitation. This knowledge has sparked significant interest in using solar thermophotovoltaic devices, where photons from a hot emitter are used to transverse a vacuum gap that is absorbed by the solar cell and used to generate electricity. Nonetheless, recent reports have shown that the temperature of the emitter is still too low to give off a photon flux sufficient for the photovoltaic cell, thus limiting the service life of these devices. New approaches use thermophotovoltaic energy conversion mechanisms such as photon-assisted tunneling with a bipolar grating-coupled complementary metal–oxide–silicon tunnel diode or a micro thermophotovoltaic generator, and these are recommended solutions to increase the efficiency. Solar thermophotovoltaic devices lack suitable structural designs that overcome the thermal losses experienced with current fabrication techniques, which can be improved. Thus, further studies need to be conducted to enhance current designs. The real implementation of an NF-TPV device is extremely challenging because it requires maintaining a high-temperature difference in the sub-micron gap between the low-bandgap PV cell and the emitter. In addition, the temperature of the PV cell should be kept near room temperature to prevent performance degradation. Therefore, a thermally isolated emitter would be desirable for a real system. Furthermore, innovative materials that can be used as absorbers, and emitters should be explored for long-term thermal stability. Ways to reduce the cost of setting up the existing thermophotovoltaic systems are few, and cheaper alternatives should also be investigated.

Conclusion

In this paper, we discussed how the world’s demand for energy consumption led to the development of fossil fuels; however, economical sustainability and environmental pollution have created a need for cleaner energy sources. Solar energy was outlined as a preferred alternative source of renewable energy attributed to its availability and sustainability; however, the efficiency of this source of energy in its conversion mechanism to electrical power is limited. Therefore, we presented a literature review on different solar thermophotovoltaic devices, materials, and methods researchers have used in exceeding the theoretical limitations, and future recommendations and solutions were shown. Surveying the studies presented, it is clear that nanomaterials as advanced functional materials can enhance the efficiency of solar thermophotovoltaic devices. In general, solar energy is a significant source for fulfilling the required energy demands.

Author contributions

MD: conceptualization, writing, and editing of the manuscript. PP: supervision, writing, and editing of the manuscript. EM: writing and editing of the manuscript. NL: writing and editing of the manuscript. MS: writing and editing of the manuscript. AA: writing and editing of the manuscript. FO: review, writing, and editing of the manuscript. All authors contributed to the article and approved the submitted version.

Acknowledgments

The authors would like to thank the Surface Engineering Research Laboratory, Durban University of Technology, and the Tshwane University of Technology, Pretoria, South Africa, for their scientific and technical support.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Abbas, M. A., Kim, J., Rana, A. S., Kim, I., Rehman, B., Ahmad, Z., et al. (2022). Nanostructured chromium-based broadband absorbers and emitters to realize thermally stable solar thermophotovoltaic systems. Nanoscale 14 (17), 6425–6436. doi:10.1039/d1nr08400c

PubMed Abstract | CrossRef Full Text | Google Scholar

Abbasi, K. R., Shahbaz, M., Zhang, J., Irfan, M., and Alvarado, R. (2022). Analyze the environmental sustainability factors of China: The role of fossil fuel energy and renewable energy. Renew. Energy 187, 390–402. doi:10.1016/j.renene.2022.01.066

CrossRef Full Text | Google Scholar

Alam, M. A., and Khan, M. R. (2019). Shockley–Queisser triangle predicts the thermodynamic efficiency limits of arbitrarily complex multijunction bifacial solar cells. Proc. Natl. Acad. Sci. 116 (48), 23966–23971. doi:10.1073/pnas.1910745116

PubMed Abstract | CrossRef Full Text | Google Scholar

Anggraini, L., Tarigan, H. J., and Simatupang, J. W. (2022). Metamaterial and metasurface based emitters for solar thermal photovoltaic applications: Analytical review. Bull. Electr. Eng. Inf. 11 (6), 3249–3257. doi:10.11591/eei.v11i6.3588

CrossRef Full Text | Google Scholar

Asongu, S. A., Agboola, M. O., Alola, A. A., and Bekun, F. V. (2020). The criticality of growth, urbanization, electricity and fossil fuel consumption to environment sustainability in Africa. Sci. Total Environ. 712, 136376. doi:10.1016/j.scitotenv.2019.136376

PubMed Abstract | CrossRef Full Text | Google Scholar

Axelevitch, A. (2018). Photovoltaic efficiency improvement: Limits and possibilities. Sci. Revs Chem. Commun. 8 (1), 115.

Google Scholar

Azzali, N., Meucci, M., Di Rosa, D., Mercatelli, L., Silvestroni, L., Sciti, D., et al. (2021). Spectral emittance of ceramics for high temperature solar receivers. Sol. Energy 222, 74–83. doi:10.1016/j.solener.2021.05.019

CrossRef Full Text | Google Scholar

Beard, H., Cholleti, A., Pearlman, D., Sherman, W., and Loving, K. A. (2013). Applying physics-based scoring to calculate free energies of binding for single amino acid mutations in protein-protein complexes. PloS one 8 (12), e82849.

PubMed Abstract | CrossRef Full Text | Google Scholar

Bendelala, F., and Cheknane, A. (2022). A transparent metasurface absorber/emitter with high solar thermal transfer efficiency for combined solar/thermal conversion application. Plasmonics 17 (3), 921–929. doi:10.1007/s11468-021-01580-w

CrossRef Full Text | Google Scholar

Bhatt, R., Kravchenko, I., and Gupta, M. (2020). High-efficiency solar thermophotovoltaic system using a nanostructure-based selective emitter. Sol. Energy 197, 538–545. doi:10.1016/j.solener.2020.01.029

CrossRef Full Text | Google Scholar

Bierman, D. M., Lenert, A., Chan, W. R., Bhatia, B., Celanović, I., Soljačić, M., et al. (2016). Enhanced photovoltaic energy conversion using thermally based spectral shaping. Nat. Energy 1 (6), 16068–16077. doi:10.1038/nenergy.2016.68

CrossRef Full Text | Google Scholar

Bisht, N., More, P., Khanna, P. K., Abolhassani, R., Mishra, Y. K., and Madsen, M. (2021). Progress of hybrid nanocomposite materials for thermoelectric applications. Mater. Adv. 2 (6), 1927–1956. doi:10.1039/d0ma01030h

CrossRef Full Text | Google Scholar

Bitnar, B., Durisch, W., and Holzner, R. (2013). Thermophotovoltaics on the move to applications. Appl. Energy 105, 430–438. doi:10.1016/j.apenergy.2012.12.067

CrossRef Full Text | Google Scholar

Briggs, J. A., Atre, A. C., and Dionne, J. A. (2013). Narrow-bandwidth solar upconversion: Case studies of existing systems and generalized fundamental limits. J. Appl. Phys. 113 (12), 124509. doi:10.1063/1.4796092

CrossRef Full Text | Google Scholar

Burger, T., Sempere, C., Roy-Layinde, B., and Lenert, A. (2020). Present efficiencies and future opportunities in thermophotovoltaics. Joule 4 (8), 1660–1680. doi:10.1016/j.joule.2020.06.021

CrossRef Full Text | Google Scholar

Burlingame, Q., Ball, M., and Loo, Y. L. (2020). It’s time to focus on organic solar cell stability. Nat. Energy 5 (12), 947–949. doi:10.1038/s41560-020-00732-2

CrossRef Full Text | Google Scholar

Chan, W. R., Bermel, P., Pilawa-Podgurski, R. C., Marton, C. H., Jensen, K. F., Senkevich, J. J., et al. (2013). Toward high-energy-density, high-efficiency, and moderate-temperature chip-scale thermophotovoltaics. Proc. Natl. Acad. Sci. 110 (14), 5309–5314. doi:10.1073/pnas.1301004110

PubMed Abstract | CrossRef Full Text | Google Scholar

Charache, G. W., Baldasaro, P. F., Danielson, L. R., DePoy, D. M., Freeman, M. J., Wang, C. A., et al. (1999). InGaAsSb thermophotovoltaic diode: Physics evaluation. J. Appl. Phys. 85 (4), 2247–2252. doi:10.1063/1.369533

CrossRef Full Text | Google Scholar

Chaudhuri, T. K. (1992). A solar thermophotovoltaic converter using Pbs photovoltaic cells. Int. J. Energy Res. 16 (6), 481–487. doi:10.1002/er.4440160605

CrossRef Full Text | Google Scholar

Chen, G., Dresselhaus, M. S., Dresselhaus, G., Fleurial, J. P., and Caillat, T. (2003). Recent developments in thermoelectric materials. Int. Mater. Rev. 48 (1), 45–66. doi:10.1179/095066003225010182

CrossRef Full Text | Google Scholar

Chen, J., Jia, D., Johansson, E. M., Hagfeldt, A., and Zhang, X. (2021). Emerging perovskite quantum dot solar cells: Feasible approaches to boost performance. Energy & Environ. Sci. 14 (1), 224–261. doi:10.1039/d0ee02900a

CrossRef Full Text | Google Scholar

Chen, M., Yan, H., Zhou, P., and Chen, X. (2020). Performance analysis of solar thermophotovoltaic system with selective absorber/emitter. J. Quantitative Spectrosc. Radiat. Transf. 253, 107163. doi:10.1016/j.jqsrt.2020.107163

CrossRef Full Text | Google Scholar

Chen, Y., Jia, B., Guan, X., Han, L., Wu, L., Guan, P., et al. (2022). Design and analysis of III-V two-dimensional van der Waals heterostructures for ultra-thin solar cells. Appl. Surf. Sci. 586, 152799. doi:10.1016/j.apsusc.2022.152799

CrossRef Full Text | Google Scholar

Daneshpour, R., and Mehrpooya, M. (2018). Design and optimization of a combined solar thermophotovoltaic power generation and solid oxide electrolyser for hydrogen production. Energy Convers. Manag. 176, 274–286. doi:10.1016/j.enconman.2018.09.033

CrossRef Full Text | Google Scholar

Daneshvar, H., Prinja, R., and Kherani, N. P. (2015). Thermophotovoltaics: Fundamentals, challenges and prospects. Appl. Energy 159, 560–575. doi:10.1016/j.apenergy.2015.08.064

CrossRef Full Text | Google Scholar

Das, U. K., Tey, K. S., Seyedmahmoudian, M., Mekhilef, S., Idris, M. Y. I., Van Deventer, W., et al. (2018). Forecasting of photovoltaic power generation and model optimization: A review. Renew. Sustain. Energy Rev. 81, 912–928. doi:10.1016/j.rser.2017.08.017

CrossRef Full Text | Google Scholar

Dashiell, M. W., Beausang, J. F., Ehsani, H., Nichols, G. J., Depoy, D. M., Danielson, L. R., et al. (2006). Quaternary InGaAsSb thermophotovoltaic diodes. IEEE Trans. Electron Devices 53 (12), 2879–2891. doi:10.1109/ted.2006.885087

CrossRef Full Text | Google Scholar

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

CrossRef Full Text | Google Scholar

Datas, A., and Martí, A. (2017). Thermophotovoltaic energy in space applications: Review and future potential. Sol. Energy Mater. Sol. Cells 161, 285–296. doi:10.1016/j.solmat.2016.12.007

CrossRef Full Text | Google Scholar

Datas, A., and Vaillon, R. (2021). “Thermophotovoltaic energy conversion,” in Ultra-high temperature thermal energy storage, transfer and conversion (Sawston, Cambridge: Woodhead Publishing), 285–308.

CrossRef Full Text | Google Scholar

Davids, P. S., Kirsch, J., Starbuck, A., Jarecki, R., Shank, J., and Peters, D. (2020). Electrical power generation from moderate-temperature radiative thermal sources. Science 367 (6484), 1341–1345. doi:10.1126/science.aba2089

PubMed Abstract | CrossRef Full Text | Google Scholar

Davies, P. A., and Luque, A. (1994). Solar thermophotovoltaics: Brief review and a new look. Sol. energy Mater. Sol. cells 33 (1), 11–22. doi:10.1016/0927-0248(94)90284-4

CrossRef Full Text | Google Scholar

Dresselhaus, M. S., Chen, G., Tang, M. Y., Yang, R. G., Lee, H., Wang, D. Z., et al. (2007). New directions for low dimensional thermoelectric materials. Adv. Mater. 19 (8), 1043–1053. doi:10.1002/chin.200726202

CrossRef Full Text | Google Scholar

Ehrler, B., Alarcón-Lladó, E., Tabernig, S. W., Veeken, T., Garnett, E. C., and Polman, A. (2020). Photovoltaics reaching for the shockley–queisser limit. Washington, D.C: ACS Publishing.

Google Scholar

Ferrero, D., Lanzini, A., Santarelli, M., and Leone, P. (2013). A comparative assessment on hydrogen production from low-and high-temperature electrolysis. Int. J. hydrogen energy 38 (9), 3523–3536. doi:10.1016/j.ijhydene.2013.01.065

CrossRef Full Text | Google Scholar

Fink, Y., Winn, J. N., Fan, S., Chen, C., Michel, J., Joannopoulos, J. D., et al. (1998). A dielectric omnidirectional reflector. Science 282 (5394), 1679–1682. doi:10.1126/science.282.5394.1679

PubMed Abstract | CrossRef Full Text | Google Scholar

Fonash, S. (2012). Solar cell device physics. New York: Academic press.

Google Scholar

Fukuda, K., Yu, K., and Someya, T. (2020). The future of flexible organic solar cells. Adv. Energy Mater. 10 (25), 2000765. doi:10.1002/aenm.202000765

CrossRef Full Text | Google Scholar

Gerislioglu, B., Ahmadivand, A., and Adam, J. (2019). Infrared plasmonic photodetectors: The emergence of high photon yield toroidal metadevices. Mater. Today Chem. 14, 100206. doi:10.1016/j.mtchem.2019.100206

CrossRef Full Text | Google Scholar

Gong, J., Li, C., and Wasielewski, M. R. (2019). Advances in solar energy conversion. Chem. Soc. Rev. 48 (7), 1862–1864. doi:10.1039/c9cs90020a

PubMed Abstract | CrossRef Full Text | Google Scholar

Gong, Y., Qiu, R., Niu, C., Fu, J., Jedlicka, E., Giridharagopal, R., et al. (2021). Ag incorporation with controlled grain growth enables 12.5% efficient kesterite solar cell with open circuit voltage reached 64.2% Shockley–Queisser limit. Adv. Funct. Mater. 31 (24), 2101927. doi:10.1002/adfm.202101927

CrossRef Full Text | Google Scholar

Goodnick, S. M., and Honsberg, C. (2022). “Solar cells,” in Springer handbook of semiconductor devices. Nature Switzerland, Springer, Cham publishing, 699–745.

CrossRef Full Text | Google Scholar

Grätzel, M. (2005). Solar energy conversion by dye-sensitized photovoltaic cells. Inorg. Chem. 44 (20), 6841–6851. doi:10.1021/ic0508371

PubMed Abstract | CrossRef Full Text | Google Scholar

Green, M. A., and Ho-Baillie, A. W. (2019). Pushing to the limit: Radiative efficiencies of recent mainstream and emerging solar cells. ACS Energy Lett. 4 (7), 1639–1644. doi:10.1021/acsenergylett.9b01128

CrossRef Full Text | Google Scholar

Guenet, B., Gabrielle, B., Chenu, C., Arrouays, D., Balesdent, J., Bernoux, M., et al. (2021). Can N2O emissions offset the benefits from soil organic carbon storage. Glob. Change Biol. 27 (2), 237–256. doi:10.1111/gcb.15342

CrossRef Full Text | Google Scholar

Guillemoles, J. F., Kirchartz, T., Cahen, D., and Rau, U. (2019). Guide for the perplexed to the Shockley–Queisser model for solar cells. Nat. photonics 13 (8), 501–505. doi:10.1038/s41566-019-0479-2

CrossRef Full Text | Google Scholar

Gupta, M. C., Ungaro, C., Foley IV, J. J., and Gray, S. K. (2018). Optical nanostructures design, fabrication, and applications for solar/thermal energy conversion. Sol. Energy 165, 100–114. doi:10.1016/j.solener.2018.01.010

CrossRef Full Text | Google Scholar

Hassan, H., Yousef, M. S., Fathy, M., and Ahmed, M. S. (2020). Impact of condenser heat transfer on energy and exergy performance of active single slope solar still under hot climate conditions. Solar Energy 204, 79–89.

CrossRef Full Text | Google Scholar

Herring, C., and Nichols, M. H. (1949). Thermionic emission. Rev. Mod. Phys. 21 (2), 185–270. doi:10.1103/revmodphys.21.185

CrossRef Full Text | Google Scholar

Hosenuzzaman, M., Rahim, N. A., Selvaraj, J., Hasanuzzaman, M., Malek, A. A., and Nahar, A. (2015). Global prospects, progress, policies, and environmental impact of solar photovoltaic power generation. Renew. Sustain. energy Rev. 41, 284–297. doi:10.1016/j.rser.2014.08.046

CrossRef Full Text | Google Scholar

Hou, G., Lin, Z., Wang, Q., Zhu, Y., Xu, J., and Chen, K. (2023). Integrated silicon-based spectral reshaping intermediate structures for high performance solar thermophotovoltaics. Sol. Energy 249, 227–232. doi:10.1016/j.solener.2022.11.026

CrossRef Full Text | Google Scholar

Huang, J., Ojambati, O. S., Chikkaraddy, R., Sokołowski, K., Wan, Q., Durkan, C., et al. (2021). Plasmon-induced trap state emission from single quantum dots. Phys. Rev. Lett. 126 (4), 047402. doi:10.1103/physrevlett.126.047402

PubMed Abstract | CrossRef Full Text | Google Scholar

Jang, Y. H., Jang, Y. J., Kim, S., Quan, L. N., Chung, K., and Kim, D. H. (2016). Plasmonic solar cells: From rational design to mechanism overview. Chem. Rev. 116 (24), 14982–15034. doi:10.1021/acs.chemrev.6b00302

PubMed Abstract | CrossRef Full Text | Google Scholar

Javadi, F. S., Metselaar, H. S. C., and Ganesan, P. (2020). Performance improvement of solar thermal systems integrated with phase change materials (PCM), a review. Sol. Energy 206, 330–352. doi:10.1016/j.solener.2020.05.106

CrossRef Full Text | Google Scholar

Jayawardena, K. D. G. I., Bandara, R. M. I., Monti, M., Butler-Caddle, E., Pichler, T., Shiozawa, H., et al. (2020). Approaching the Shockley–Queisser limit for fill factors in lead–tin mixed perovskite photovoltaics. J. Mater. Chem. A 8 (2), 693–705. doi:10.1039/c9ta10543c

CrossRef Full Text | Google Scholar

Jia, N., Cao, J., Tan, X. Y., Dong, J., Liu, H., Tan, C. K. I., et al. (2021). Thermoelectric materials and transport physics. Mater. Today Phys. 21, 100519. doi:10.1016/j.mtphys.2021.100519

CrossRef Full Text | Google Scholar

Jia, X., Puthen-Veettil, B., Xia, H., Yang, T. C. J., Lin, Z., Zhang, T., et al. (2016). All-silicon tandem solar cells: Practical limits for energy conversion and possible routes for improvement. J. Appl. Phys. 119 (23), 233102. doi:10.1063/1.4954003

CrossRef Full Text | Google Scholar

Joshi, A. S., Dincer, I., and Reddy, B. V. (2009). Performance analysis of photovoltaic systems: A review. Renew. Sustain. Energy Rev. 13 (8), 1884–1897. doi:10.1016/j.rser.2009.01.009

CrossRef Full Text | Google Scholar

Kalair, A., Abas, N., Saleem, M. S., Kalair, A. R., and Khan, N. (2021). Role of energy storage systems in energy transition from fossil fuels to renewables. Energy Storage 3 (1), e135. doi:10.1002/est2.135

CrossRef Full Text | Google Scholar

Kalogirou, S. A. (2013). Solar energy engineering: Processes and systems. Cambridge, Massachusetts: Academic Press.

Google Scholar

Kenjo, S., Ogino, Y., Mukai, K., Bakr, M., Yagi, J., and Konishi, S. (2022). Employing of ZrCo as a fuel source in a discharge-type fusion neutron source operated in self-sufficient mode. Int. J. Hydrogen Energy 47 (5), 3054–3062. doi:10.1016/j.ijhydene.2021.10.250

CrossRef Full Text | Google Scholar

Khalid, K. A. A., Leong, T. J., and Mohamed, K. (2016). Review on thermionic energy converters. IEEE Trans. electron devices 63 (6), 2231–2241. doi:10.1109/ted.2016.2556751

CrossRef Full Text | Google Scholar

Khvostikov, V. P., Gazaryan, P. Y., Khvostikova, O. A., Potapovich, N. S., Sorokina, S. V., Malevskaya, A. V., et al. (2007). GaSb Applications for solar thermophotovoltaic conversion. AIP Conf. Proc. 890 (1), 139–148.

CrossRef Full Text | Google Scholar

Kim, J. Y., Lee, J. W., Jung, H. S., Shin, H., and Park, N. G. (2020). High-efficiency perovskite solar cells. Chem. Rev. 120 (15), 7867–7918. doi:10.1021/acs.chemrev.0c00107

PubMed Abstract | CrossRef Full Text | Google Scholar

Kim, S., Márquez, J. A., Unold, T., and Walsh, A. (2020). Upper limit to the photovoltaic efficiency of imperfect crystals from first principles. Energy & Environ. Sci. 13 (5), 1481–1491. doi:10.1039/d0ee00291g

CrossRef Full Text | Google Scholar

Kohiyama, A., Shimizu, M., Konno, K., Furuhashi, T., and Yugami, H. (2020). Effective photon recycling in solar thermophotovoltaics using a confined cuboid emitter. Opt. Express 28 (26), 38567–38578. doi:10.1364/oe.412764

PubMed Abstract | CrossRef Full Text | Google Scholar

Kohiyama, A., Shimizu, M., and Yugami, H. (2016). Unidirectional radiative heat transfer with a spectrally selective planar absorber/emitter for high-efficiency solar thermophotovoltaic systems. Appl. Phys. Express 9 (11), 112302. doi:10.7567/apex.9.112302

CrossRef Full Text | Google Scholar

Krügener, J., Rienäcker, M., Schäfer, S., Sanchez, M., Wolter, S., Brendel, R., et al. (2021). Photonic crystals for highly efficient silicon single junction solar cells. Sol. Energy Mater. Sol. Cells 233, 111337. doi:10.1016/j.solmat.2021.111337

CrossRef Full Text | Google Scholar

Legendre, J., and Chapuis, P. O. (2022). GaAs-based near-field thermophotonic devices: Approaching the idealized case with one-dimensional PN junctions. Sol. Energy Mater. Sol. Cells 238, 111594. doi:10.1016/j.solmat.2022.111594

CrossRef Full Text | Google Scholar

Lenert, A., Bierman, D. M., Nam, Y., Chan, W. R., Celanović, I., Soljačić, M., et al. (2014). A nanophotonic solar thermophotovoltaic device. Nat. Nanotechnol. 9 (2), 126–130. doi:10.1038/nnano.2013.286

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, G., Liu, X., Chen, J., Li, Y., Shi, L., Fu, G., et al. (2019). Near-unity, full-spectrum, nanoscale solar absorbers and near-perfect blackbody emitters. Sol. Energy Mater. Sol. Cells 190, 20–29. doi:10.1016/j.solmat.2018.10.011

CrossRef Full Text | Google Scholar

Lu, Y., Li, K., Yang, X., Lu, S., Li, S., Zheng, J., et al. (2021). HTL-free Sb2 (S, Se) 3 solar cells with an optimal detailed balance band gap. ACS Appl. Mater. Interfaces 13 (39), 46858–46865. doi:10.1021/acsami.1c10758

PubMed Abstract | CrossRef Full Text | Google Scholar

Lupangu, C., and Bansal, R. C. (2017). A review of technical issues on the development of solar photovoltaic systems. Renew. Sustain. Energy Rev. 73, 950–965. doi:10.1016/j.rser.2017.02.003

CrossRef Full Text | Google Scholar

Mahan, G. D., and Sofo, J. O. (1996). The best thermoelectric. Proc. Natl. Acad. Sci. 93 (15), 7436–7439. doi:10.1073/pnas.93.15.7436

PubMed Abstract | CrossRef Full Text | Google Scholar

Mahan, G., Sales, B., and Sharp, J. (1997). Thermoelectric materials: New approaches to an old problem. Phys. Today 50 (3), 42–47. doi:10.1063/1.881752

CrossRef Full Text | Google Scholar

Malhotra, R. (2020). Fossil energy. New York: Springer.

Google Scholar

Mann, S. A., Grote, R. R., Osgood, R. M., Alù, A., and Garnett, E. C. (2016). Opportunities and limitations for nanophotonic structures to exceed the Shockley–Queisser limit. ACS Nano 10 (9), 8620–8631. doi:10.1021/acsnano.6b03950

PubMed Abstract | CrossRef Full Text | Google Scholar

Mao, J., Chen, G., and Ren, Z. (2021). Thermoelectric cooling materials. Nat. Mater. 20 (4), 454–461. doi:10.1038/s41563-020-00852-w

PubMed Abstract | CrossRef Full Text | Google Scholar

Markvart, T. (2022). Shockley: Queisser detailed balance limit after 60 years. Wiley Interdiscip. Rev. Energy Environ. 11 (4), e430. doi:10.1002/wene.430

CrossRef Full Text | Google Scholar

Mojiri, A., Taylor, R., Thomsen, E., and Rosengarten, G. (2013). Spectral beam splitting for efficient conversion of solar energy—a review. Renew. Sustain. Energy Rev. 28, 654–663. doi:10.1016/j.rser.2013.08.026

CrossRef Full Text | Google Scholar

Nam, Y., Yeng, Y. X., Lenert, A., Bermel, P., Celanovic, I., Soljačić, M., et al. (2014). Solar thermophotovoltaic energy conversion systems with two-dimensional tantalum photonic crystal absorbers and emitters. Sol. Energy Mater. Sol. Cells 122, 287–296. doi:10.1016/j.solmat.2013.12.012

CrossRef Full Text | Google Scholar

Nandihalli, N., Liu, C. J., and Mori, T. (2020). Polymer based thermoelectric nanocomposite materials and devices: Fabrication and characteristics. Nano Energy 78, 105186. doi:10.1016/j.nanoen.2020.105186

CrossRef Full Text | Google Scholar

Ndukwu, M. C., Bennamoun, L., and Simo-Tagne, M. (2021). Reviewing the exergy analysis of solar thermal systems integrated with phase change materials. Energies 14 (3), 724. doi:10.3390/en14030724

CrossRef Full Text | Google Scholar

Nevala, S. M., Hamuyuni, J., Junnila, T., Sirviö, T., Eisert, S., Wilson, B. P., et al. (2019). Electro-hydraulic fragmentation vs conventional crushing of photovoltaic panels–Impact on recycling. Waste Manag. 87, 43–50. doi:10.1016/j.wasman.2019.01.039

PubMed Abstract | CrossRef Full Text | Google Scholar

Okutsu, K., Yamashita, T., Kino, Y., Nakashima, R., Miyashita, K., Yasuda, K., et al. (2021). Design for detecting recycling muon after muon-catalyzed fusion reaction in solid hydrogen isotope target. Fusion Eng. Des. 170, 112712. doi:10.1016/j.fusengdes.2021.112712

CrossRef Full Text | Google Scholar

Osorio, J. D., Wang, Z., Karniadakis, G., Cai, S., Chryssostomidis, C., Panwar, M., et al. (2022). Forecasting solar-thermal systems performance under transient operation using a data-driven machine learning approach based on the deep operator network architecture. Energy Convers. Manag. 252, 115063. doi:10.1016/j.enconman.2021.115063

CrossRef Full Text | Google Scholar

Palfinger, G. (2006). Low dimensional Si/SiGe structures deposited by UHV-CVD for thermophotovoltaics (Doctoral dissertation). Switzerland.

Google Scholar

Panwar, N. L., Kaushik, S. C., and Kothari, S. (2011). Role of renewable energy sources in environmental protection: A review. Renew. Sustain. energy Rev. 15 (3), 1513–1524. doi:10.1016/j.rser.2010.11.037

CrossRef Full Text | Google Scholar

Pareek, A., Dom, R., Gupta, J., Chandran, J., Adepu, V., and Borse, P. H. (2020). Insights into renewable hydrogen energy: Recent advances and prospects. Mater. Sci. Energy Technol. 3, 319–327. doi:10.1016/j.mset.2019.12.002

CrossRef Full Text | Google Scholar

Park, N. G. (2020). Research direction toward scalable, stable, and high efficiency perovskite solar cells. Adv. Energy Mater. 10 (13), 1903106. doi:10.1002/aenm.201903106

CrossRef Full Text | Google Scholar

Pasupathi, M. K., Alagar, K., Mm, M., and Aritra, G. (2020). Characterization of hybrid-nano/paraffin organic phase change material for thermal energy storage applications in solar thermal systems. Energies 13 (19), 5079. doi:10.3390/en13195079

CrossRef Full Text | Google Scholar

Pei, Y., Wang, H., and Snyder, G. J. (2012). Band engineering of thermoelectric materials. Adv. Mater. 24 (46), 6125–6135. doi:10.1002/adma.201202919

PubMed Abstract | CrossRef Full Text | Google Scholar

Peuser, F. A., Remmers, K. H., and Schnauss, M. (2013). Solar thermal systems: Successful planning and construction. Beuth Verlag Germany: Routledge.

Google Scholar

Pourkiaei, S. M., Ahmadi, M. H., Sadeghzadeh, M., Moosavi, S., Pourfayaz, F., Chen, L., et al. (2019). Thermoelectric cooler and thermoelectric generator devices: A review of present and potential applications, modeling and materials. Energy 186, 115849. doi:10.1016/j.energy.2019.07.179

CrossRef Full Text | Google Scholar

Prabhu, B., and ValanArasu, A. (2020). Stability analysis of TiO2-Ag nanocomposite particles dispersed paraffin wax as energy storage material for solar thermal systems. Renew. Energy 152, 358–367. doi:10.1016/j.renene.2020.01.043

CrossRef Full Text | Google Scholar

Prentice, J. S. C. (1999). Optical generation rate of electron-hole pairs in multilayer thin-film photovoltaic cells. J. Phys. D Appl. Phys. 32 (17), 2146–2150. doi:10.1088/0022-3727/32/17/302

CrossRef Full Text | Google Scholar

Qasim, W., Xia, L., Lin, S., Wan, L., Zhao, Y., and Butterbach-Bahl, K. (2021). Global greenhouse vegetable production systems are hotspots of soil N2O emissions and nitrogen leaching: A meta-analysis. Environ. Pollut. 272, 116372. doi:10.1016/j.envpol.2020.116372

PubMed Abstract | CrossRef Full Text | Google Scholar

Qiu, H., Xu, H., Ni, M., and Xiao, G. (2022). Photo-thermo-electric modeling of photon-enhanced thermionic emission with concentrated solar power. Sol. Energy Mater. Sol. Cells 246, 111922. doi:10.1016/j.solmat.2022.111922

CrossRef Full Text | Google Scholar

Rana, A. S., Zubair, M., Danner, A., and Mehmood, M. Q. (2021). Revisiting tantalum based nanostructures for efficient harvesting of solar radiation in STPV systems. Nano Energy 80, 105520. doi:10.1016/j.nanoen.2020.105520

CrossRef Full Text | Google Scholar

Rashidi, S., Yang, L., Khoosh-Ahang, A., Jing, D., and Mahian, O. (2020). Entropy generation analysis of different solar thermal systems. Environ. Sci. Pollut. Res. 27, 20699–20724. doi:10.1007/s11356-020-08472-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Rau, U., Einsele, F., and Glaeser, G. C. (2005). Efficiency limits of photovoltaic fluorescent collectors. Appl. Phys. Lett. 87 (17), 171101. doi:10.1063/1.2112196

CrossRef Full Text | Google Scholar

Ren, L., Yu, A., Wang, W., Guo, D., Jia, M., Guo, P., et al. (2021). Pn junction based direct-current triboelectric nanogenerator by conjunction of tribovoltaic effect and photovoltaic effect. Nano Lett. 21 (23), 10099–10106. doi:10.1021/acs.nanolett.1c03922

PubMed Abstract | CrossRef Full Text | Google Scholar

Rühle, S. (2016). Tabulated values of the Shockley–Queisser limit for single junction solar cells. Sol. energy 130, 139–147. doi:10.1016/j.solener.2016.02.015

CrossRef Full Text | Google Scholar

Sadi, T., Radevici, I., and Oksanen, J. (2020). Thermophotonic cooling with light-emitting diodes. Nat. Photonics 14 (4), 205–214. doi:10.1038/s41566-020-0600-6

CrossRef Full Text | Google Scholar

Sahoo, G. S., and Mishra, G. P. (2018). Design and modeling of an SJ infrared solar cell approaching upper limit of theoretical efficiency. Int. J. Mod. Phys. B 32 (02), 1850014. doi:10.1142/s0217979218500145

CrossRef Full Text | Google Scholar

Schaller, R. D., Sykora, M., Pietryga, J. M., and Klimov, V. I. (2006). Seven excitons at a cost of one: Redefining the limits for conversion efficiency of photons into charge carriers. Nano Lett. 6 (3), 424–429. doi:10.1021/nl052276g

PubMed Abstract | CrossRef Full Text | Google Scholar

Schwede, J. W., Bargatin, I., Riley, D. C., Hardin, B. E., Rosenthal, S. J., Sun, Y., et al. (2010). Photon-enhanced thermionic emission for solar concentrator systems. Nat. Mater. 9 (9), 762–767. doi:10.1038/nmat2814

PubMed Abstract | CrossRef Full Text | Google Scholar

Sharma, S., and Ghoshal, S. K. (2015). Hydrogen the future transportation fuel: From production to applications. Renew. Sustain. energy Rev. 43, 1151–1158. doi:10.1016/j.rser.2014.11.093

CrossRef Full Text | Google Scholar

Shen, M., Huang, W., Chen, M., Song, B., Zeng, G., and Zhang, Y. (2020). (Micro) plastic crisis: Un-ignorable contribution to global greenhouse gas emissions and climate change. J. Clean. Prod. 254, 120138. doi:10.1016/j.jclepro.2020.120138

CrossRef Full Text | Google Scholar

Shi, X. L., Zou, J., and Chen, Z. G. (2020). Advanced thermoelectric design: From materials and structures to devices. Chem. Rev. 120 (15), 7399–7515. doi:10.1021/acs.chemrev.0c00026

PubMed Abstract | CrossRef Full Text | Google Scholar

Shockley, W., and Queisser, H. J. (1961). Detailed balance limit of efficiency of p-n junction solar cells. J. Appl. Phys. 32 (3), 510–519. doi:10.1063/1.1736034

CrossRef Full Text | Google Scholar

Singh, A. K., and Jen, T. C. (2021). “Introduction to nanomaterials and their applications in optoelectronics,” in Nanomaterials for optoelectronic applications (Canada: Apple Academic Press), 1–75.

CrossRef Full Text | Google Scholar

Sisay, T. (2022). Solar irradiance modeling using feed-forward neural network and multiple linear regression over bahir dar. Doctoral dissertation. Debreberhan, Ethiopia: Debreberhan University.

Google Scholar

Snyder, G. J., and Toberer, E. S. (2008). Complex thermoelectric materials. Nat. Mater. 7 (2), 105–114. doi:10.1038/nmat2090

PubMed Abstract | CrossRef Full Text | Google Scholar

Snyder, G. J., and Ursell, T. S. (2003). Thermoelectric efficiency and compatibility. Phys. Rev. Lett. 91 (14), 148301. doi:10.1103/physrevlett.91.148301

PubMed Abstract | CrossRef Full Text | Google Scholar

Song, J., Han, J., Choi, M., and Lee, B. J. (2022). Modeling and experiments of near-field thermophotovoltaic conversion: A review. Sol. Energy Mater. Sol. Cells 238, 111556. doi:10.1016/j.solmat.2021.111556

CrossRef Full Text | Google Scholar

Sootsman, J. R., Chung, D. Y., and Kanatzidis, M. G. (2009). New and old concepts in thermoelectric materials. Angew. Chem. Int. Ed. 48 (46), 8616–8639. doi:10.1002/anie.200900598

PubMed Abstract | CrossRef Full Text | Google Scholar

Tennyson, E. M., Doherty, T. A., and Stranks, S. D. (2019). Heterogeneity at multiple length scales in halide perovskite semiconductors. Nat. Rev. Mater. 4 (9), 573–587. doi:10.1038/s41578-019-0125-0

CrossRef Full Text | Google Scholar

Tervo, E. J., Callahan, W. A., Toberer, E. S., Steiner, M. A., and Ferguson, A. J. (2020). Solar thermoradiative-photovoltaic energy conversion. Cell Rep. Phys. Sci. 1 (12), 100258. doi:10.1016/j.xcrp.2020.100258

CrossRef Full Text | Google Scholar

Tian, Y., Liu, X., Ghanekar, A., and Zheng, Y. (2021). Scalable-manufactured metal–insulator–metal based selective solar absorbers with excellent high-temperature insensitivity. Appl. Energy 281, 116055. doi:10.1016/j.apenergy.2020.116055

CrossRef Full Text | Google Scholar

Tong, X. C. (2018). Functional metamaterials and metadevices, 110. Bolingbrook, IL: Springer.

Google Scholar

Tritt, T. M. (2011). Thermoelectric phenomena, materials, and applications. Annu. Rev. Mater. Res. 41, 433–448. doi:10.1146/annurev-matsci-062910-100453

CrossRef Full Text | Google Scholar

Trupke, T., Green, M. A., and Würfel, P. (2002). Improving solar cell efficiencies by down-conversion of high-energy photons. J. Appl. Phys. 92 (3), 1668–1674. doi:10.1063/1.1492021

CrossRef Full Text | Google Scholar

Ungaro, C., Gray, S. K., and Gupta, M. C. (2015). Solar thermophotovoltaic system using nanostructures. Opt. express 23 (19), A1149–A1156. doi:10.1364/oe.23.0a1149

PubMed Abstract | CrossRef Full Text | Google Scholar

Vohra, K., Vodonos, A., Schwartz, J., Marais, E. A., Sulprizio, M. P., and Mickley, L. J. (2021). Global mortality from outdoor fine particle pollution generated by fossil fuel combustion: Results from GEOS-Chem. GEOS-Chem. Environ. Res. 195, 110754. doi:10.1016/j.envres.2021.110754

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, X., Khan, M. R., Gray, J. L., Alam, M. A., and Lundstrom, M. S. (2013). Design of GaAs solar cells operating close to the Shockley–Queisser limit. IEEE J. Photovoltaics 3 (2), 737–744. doi:10.1109/jphotov.2013.2241594

CrossRef Full Text | Google Scholar

Wang, Y., Liu, H., and Zhu, J. (2019). Solar thermophotovoltaics: Progress, challenges, and opportunities. Apl. Mater. 7 (8), 080906. doi:10.1063/1.5114829

CrossRef Full Text | Google Scholar

Welsby, D., Price, J., Pye, S., and Ekins, P. (2021). Unextractable fossil fuels in a 1.5° C world. Nature 597 (7875), 230–234. doi:10.1038/s41586-021-03821-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Wong, J., Omelchenko, S. T., and Atwater, H. A. (2020). Impact of semiconductor band tails and band filling on photovoltaic efficiency limits. ACS Energy Lett. 6 (1), 52–57. doi:10.1021/acsenergylett.0c02362

CrossRef Full Text | Google Scholar

Wu, F., Liu, T., and Xiao, S. (2023). Polarization-sensitive photonic bandgaps in hybrid one-dimensional photonic crystals composed of all-dielectric elliptical metamaterials and isotropic dielectrics. Appl. Opt. 62 (3), 706–713. doi:10.1364/ao.480083

PubMed Abstract | CrossRef Full Text | Google Scholar

Wu, F., Wu, X., Xiao, S., Liu, G., and Li, H. (2021). Broadband wide-angle multilayer absorber based on a broadband omnidirectional optical Tamm state. Opt. Express 29 (15), 23976–23987. doi:10.1364/oe.434181

PubMed Abstract | CrossRef Full Text | Google Scholar

Xiang, C., Zhao, X., Tan, L., Ye, J., Wu, S., Zhang, S., et al. (2019). A solar tube: Efficiently converting sunlight into electricity and heat. Nano Energy 55, 269–276. doi:10.1016/j.nanoen.2018.10.077

CrossRef Full Text | Google Scholar

Xiao, G., Zheng, G., Qiu, M., Li, Q., Li, D., and Ni, M. (2017). Thermionic energy conversion for concentrating solar power. Appl. Energy 208, 1318–1342. doi:10.1016/j.apenergy.2017.09.021

CrossRef Full Text | Google Scholar

Xu, G., Schwarz, P., and Yang, H. (2020). Adjusting energy consumption structure to achieve China's CO2 emissions peak. Renew. Sustain. Energy Rev. 122, 109737. doi:10.1016/j.rser.2020.109737

CrossRef Full Text | Google Scholar

Xu, Y., Gong, T., and Munday, J. N. (2015). The generalized Shockley-Queisser limit for nanostructured solar cells. Sci. Rep. 5 (1), 13536–13539. doi:10.1038/srep13536

PubMed Abstract | CrossRef Full Text | Google Scholar

Xu, Y., Wang, J., Yu, F., Guo, Z., Cheng, H., Yin, J., et al. (2020). Flexible and efficient solar thermal generators based on polypyrrole coated natural latex foam for multimedia purification. ACS Sustain. Chem. Eng. 8 (32), 12053–12062. doi:10.1021/acssuschemeng.0c03164

CrossRef Full Text | Google Scholar

Yang, S., Qiu, P., Chen, L., and Shi, X. (2021). Recent developments in flexible thermoelectric devices. Small Sci. 1 (7), 2100005. doi:10.1002/smsc.202100005

CrossRef Full Text | Google Scholar

Yao, H., and Hou, J. (2022). Recent advances in single-junction organic solar cells. Angew. Chem. 134 (37), e202209021. doi:10.1002/anie.202209021

CrossRef Full Text | Google Scholar

Zenker, M., Heinzel, A., Stollwerck, G., Ferber, J., and Luther, J. (2001). Efficiency and power density potential of combustion-driven thermophotovoltaic systems using GaSb photovoltaic cells. IEEE Trans. Electron Devices 48 (2), 367–376. doi:10.1109/16.902740

CrossRef Full Text | Google Scholar

Zhang, L., Shi, X. L., Yang, Y. L., and Chen, Z. G. (2021). Flexible thermoelectric materials and devices: From materials to applications. Mater. Today 46, 62–108. doi:10.1016/j.mattod.2021.02.016

CrossRef Full Text | Google Scholar

Zhang, X., Han, L., Wei, H., Tan, X., Zhou, W., Li, W., et al. (2022). Linking urbanization and air quality together: A review and a perspective on the future sustainable urban development. J. Clean. Prod. 346, 130988. doi:10.1016/j.jclepro.2022.130988

CrossRef Full Text | Google Scholar

Zhang, Z., Ding, T., Zhou, Q., Sun, Y., Qu, M., Zeng, Z., et al. (2021). A review of technologies and applications on versatile energy storage systems. Renew. Sustain. Energy Rev. 148, 111263. doi:10.1016/j.rser.2021.111263

CrossRef Full Text | Google Scholar

Zheng, M., Lin, S., Tang, Z., Feng, Y., and Wang, Z. L. (2021). Photovoltaic effect and tribovoltaic effect at liquid-semiconductor interface. Nano Energy 83, 105810. doi:10.1016/j.nanoen.2021.105810

CrossRef Full Text | Google Scholar

Zhou, W., Chen, Q., Luo, D., Jiang, R., and Chen, J. (2020). Global energy consumption analysis based on the three-dimensional network model. IEEE Access 8, 76313–76332. doi:10.1109/access.2020.2989186

CrossRef Full Text | Google Scholar

Zhou, Z., Sakr, E., Sun, Y., and Bermel, P. (2016). Solar thermophotovoltaics: Reshaping the solar spectrum. Nanophotonics 5 (1), 1–21. doi:10.1515/nanoph-2016-0011

CrossRef Full Text | Google Scholar

Zou, H., Dai, G., Wang, A. C., Li, X., Zhang, S. L., Ding, W., et al. (2020). Alternating current photovoltaic effect. Adv. Mater. 32 (11), 2001532. doi:10.1002/adma.202001532

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: solar energy, thermophotovoltaic devices, energy conversion applications, renewable energy sources, functional material

Citation: Dada M, Popoola P, Alao A, Olalere F, Mtileni E, Lindokuhle N and Shamaine M (2023) Functional materials for solar thermophotovoltaic devices in energy conversion applications: a review. Front. Energy Res. 11:1124288. doi: 10.3389/fenrg.2023.1124288

Received: 15 December 2022; Accepted: 04 May 2023;
Published: 01 June 2023.

Edited by:

Muhammad Wakil Shahzad, Northumbria University, United Kingdom

Reviewed by:

Feng Wu, Guangdong Polytechnic Normal University, China
Tolutope Oluwasegun Siyanbola, Covenant University, Nigeria

Copyright © 2023 Dada, Popoola, Alao, Olalere, Mtileni, Lindokuhle and Shamaine. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Modupeola Dada, dadadupeola@gmail.com

Download