Thermo-electrochemical cells for harvesting waste heat: History
Please note this is an old version of this entry, which may differ significantly from the current revision.
Subjects: Electrochemistry

Thermo-electrochemical cells (thermocells; TECs) represent a promising technology for harvesting and exploiting low-grade waste heat (<100–150 °C) ubiquitous in the modern environment. Based on temperature-dependent redox reactions and ion diffusion, emerging liquid-state thermocells convert waste heat energy into electrical energy, generating power at low costs, with minimal material consumption and negligible carbon footprint. New thermoelectric applications in the field include wearable and portable electronic devices in the health and performance monitoring sectors,  using body heat as a continuous source of energy etc. For waste heat conversion to partially replace fossil fuels as an alternative energy resource, power generation needs to be commercially viable, cost effective and sustainable.

  • Thermo-electrochemical cells
  • energy harvesters
  • low grade waste heat
  • wearable electronics
  • micro supercapacitors
  • thermoelectrics

[1]1. Introduction

The conversion of energy from primary energy sources to their final use is accompanied by several losses in the form of waste heat. Forman et al. [1] have estimated the waste heat potential in transport, commercial, industrial and residential sectors. Nearly 72% of the primary energy consumed was found to be lost as waste heat, of which ~63% of waste heat streams occurred at temperatures below 100 °C. In the USA, conventional industries and power plants are known to annually waste over 8000 TWh as low-grade heat [2]. This ubiquitous waste heat energy in the form of vehicle exhaust, industrial waste heat, geothermal heat, body heat, etc., is found distributed almost everywhere. However, vast amounts of low-grade heat is mostly discarded and rarely exploited commercially due to its intrinsic low temperatures, space-time variations and the lack of cost-effective and efficient energy-recovery technologies [3]. Renewable energies such as solar, wind or nuclear, which have a negligible carbon footprint compared to non-renewable fossil fuels, [4] are seeing a great resurgence. The recent UN Climate Change Conference, held at Glasgow, UK [5] was focussed on the threat of climate change through mitigation and conservation efforts, deforestation, methane abatement and reducing the carbon footprint and emissions from energy production. Due to the vast and renewable nature of waste heat energies and their huge energy potential, harvesting low-grade waste heat is beginning to attract a great deal of attention as a promising zero-carbon source of electricity.
The transformation of thermal gradients into electrode potentials for generating electricity has been investigated for a long time [6]. The generation of electric potential in the presence of a temperature gradient between different electrical conductors/or semiconductors was discovered by Thomas Johann Seebeck in 1821. Seebeck coefficients, representing the potential difference generated per unit temperature difference, are typically in the order of few µV/K for devices based on semiconductor materials [7]. Early research was mainly in the field of solid-state thermoelectric generators (TEGs) consisting of n- and p- type semiconductors connected in series as modules and then connecting a number of modules in parallel between the heat source and a cold sink. Under the influence of thermal gradients, mobile charge carriers ‘electrons/holes’ diffused from the hot electrode to the cold, building up charges and a small potential difference [8]. Solid-state semiconductor thermoelectrics have long been investigated for the conversion of thermal to electrical energy, and several exciting advances have been made [9][10][11]. Such devices are, however, not really suitable for low-grade heat harvesting, due to their relatively low efficiencies at ambient temperatures [12].
Early studies, which later led to thermoelectrochemical converters, were used for applying thermal corrections to electrochemical processes in the field of current sources and in the production of galvanic coatings. However, the gradual dissolution of anodes was found to limit their commercial application [13]. Landry [14] suggested the integration of low- and high-temperature processes to minimize heat losses. Wakao and Nozo [15] developed a process for recovering thermal energy from low-level heat generated during the mixing of nitric acid with water. The advent of nano-structured thermoelectrics has led to a great deal of interest in waste-heat conversion (WHC) to electricity [16]. Although early research was dominated by thermoelectrics, a number of articles have been published in leading journals on new types of WHC devices using a wide range of phenomena. These include thermomagnetic generators [17], ionic heat-to-electricity conversion [18], thermo-osmotic systems [2], liquid-state thermocells [19], high temperature pyroelectric systems [20][21] and organic Rankin cycles [22] for applications in industries, construction, transportation and energy sectors. These devices are used for harvesting energy from a heat source such as automotive exhaust systems [23], fuel cells [24] or hypersonic engines [25]. Such energy harvesting can reduce the charging requirements on batteries while eliminating wired power connections. Most R&D in this field is focussed on developing better n-type and p-type thermoelectric materials.
Thermoelectrochemical (TEC) cells or ‘thermocells’ are an alternative device design based on redox-active electrolytes which can produce much higher potential differences (mV/K vs. µV/K for solid-state semiconductor devices). A thermocell has two identical electrodes in contact with a redox couple electrolyte in the cell and an external connection [26]. Under a thermal gradient, the redox reaction causes the oxidation of the redox couple at the anode and reduction at the cathode. The reduced species are transported back to the anode through diffusion, convection and migration in the electrolyte, creating a continuous reaction and current flow. Within the degradation constraints of cell materials, thermocells are capable of continuously generating electricity without consuming materials or producing emissions. Chum and Osteryoung [27] have also discussed an alternative thermally regenerative electrochemical cycle (TREC) design where electrochemical reactions are used to convert heat into electricity within the constraints of Carnot-cycle efficiency.
A significant step in the development of thermo-electrochemical systems was the study of TEC cells based on inert electrodes, where the reactions at the electrodes occurred with the release of gaseous products (HNO3  NO, N2O4 and HNO2; Br2  Br); the released volatile products later condensed at the counter electrode as the reverse reaction [28][29][30]. Representative examples of such thermocells are shown in Figure 1Figure 1a shows a KBr/Br2-based TEC cell, where the intercalation of Br2 gas into graphite electrodes plays a key role in the cell operation. At the cold electrode (cathode), Br2 intercalates into graphite, picks up electrons and dissociates into Br ions. These ions diffuse through the bridge to the hot electrode (anode) and transform into Br2 gas, releasing 2e. These electrons travel from the anode to the cathode in the outer loop, completing the circuit. Figure 1b shows a cell based on an organic acetone-iso-propanol thermocouple. Reaction at the anode involves CH3-CH(OH)-CH3 ⟶ CH3-C(O)-CH3 + H2 + 2e, and the reverse reaction occurs at the cathode. Such processes allow the relatively long and continuous generation of potential difference and provide higher output power values. However, these systems are unstable, and their operation is accompanied by the release of toxic products such as bromine, hydrogen or nitrogen oxides.
Figure 1. Schematic representations: (a) KBr/Br2-based TEC cell, (b) acetone-isopropanol thermocell, (c) Cu2+/Cu system in CuSO4 solution with soluble electrodes, (d) a ferrite/ferrocyanide redox couple ([Fe (CN)6]3−/[Fe (CN)6]4−) with inert electrodes.

2. Redox Couples and Electrolytes

2.1. Redox Couples

There are two key research areas for increasing the current-generating efficiency of thermocells, namely increasing the potential difference in the cell and the current density at the electrodes. Cell potential can be enhanced by developing combinations of redoxcouples and electrolytes, increasing the maximum cell temperature through optimized design, solid electrolytes, separators, boiling electrolytes, etc. Current density can be improved using electrodes with large surface areas as well as through electrolytes with high rates of mass transfer. These aspects will be discussed in the following sections as potential strategies to improve the efficiency of thermocell devices.
The Seebeck coefficient of a redox couple, which determines the maximum potential difference generated in a thermocell, is a measure of the entropy change during oxidation or reduction. Structural changes in the redox species, solvent shells and solvent interactions are known to play key roles [31]. Although Seebeck coefficients have been measured/computed for a number of redox couples, their application in thermocells depends on additional factors such as thermal stability, electro-chemical reversibility and availability [19]. Migita et al. [32] have reported on thermo-electromotive forces and reaction entropies for a number of transition-metal redox couples in an amide-type room temperature ionic liquid using a non-isothermal cell. The highest Seebeck-coefficient values were reported for [Fe(CN)6]3−/4− as −1.49 mV/K and Fe(III/II) as 0.96 mV/K. Laux et al. [33] have reported on the iodide/triiodide (I/I3) redox couple in aqueous as well as non-aqueous solutions and in ionic liquids. The Seebeck coefficient in this system showed a strong dependence on the electrolyte concentration, changing threefold between 0.01 M and 2 M ethyl ammonium nitrate solutions; a maximum value of 0.97 mv/K was recorded at 0.01 M.
Abraham et al. [34] have reported on cobalt-based redox couples: Co2+/3+ (bpy)3(NTf2)2/3, (bpy = 2,2′-bipyridyl; NTf2 = bis(trifluoromethanesulfonyl)amide solutions, recording a maximum Seebeck coefficient of 2.19 mV/K. This high Seebeck coefficient was attributed to changes in the spin state of Co2+/3+, providing additional electronic contributions. Anari et al. [35] combined ferrocene and iodine results for enhancing thermoelectrochemical waste-heat harvesting abilities in terms of Seebeck coefficient and power output. These two couples were found to interact even in dilute solutions, and displayed a mixture of ferrocene, ferrocenium, iodine and triiodide ions; the highest Seebeck coefficient of 1.67 mV/K was observed by combining dibutanoylferrocene and iodine. Zinovyeya et al. [36] have reported on an organic thiolate/disulfide redox couple (McMT−/BMT), derived from 2-mercapto-5-methyl-1,3,4-thiadiazole (McMT); the thermoelectric power was found to depend strongly on the concentrations of the redox couple as well as the ionic liquid. 
In combined TEC systems, the unidirectional nature of mass and charge transfer in electrochemical systems creates difficulties in the design and manufacture of ready-made converters. Conductor arrangement can reduce the effective electrode surface area relative to the area of the TEC device by connecting cells in series or parallel. Kim et al. [37] showed that the Seebeck coefficient for Fe2+/Fe3+ transition can change signs for different aqueous systems. There are also reports of hypothetical Seebeck coefficients of up to 7 mV/K measured in electrolytes without a redox couple. The temperature-dependent potential in these systems was attributed to a combination of differences in Eastman entropies and charge-density changes at the electrode surfaces due to the Soret effect [38][39].

2.2. Electrolytes

Early thermocell research was primarily carried out on redox couples in aqueous electrolytes. High power output in aqueous ferri/ferrocyanide cells was attributed to the fast diffusion of ions due to the low viscosity of water [40]. As the low boiling point of water limits these cells to operating below 100 °C, research has been carried out on a range of ionic liquids, organic solvents for extending the thermal operating range and for water-insoluble or other redox couples with stability issues [41]. Ionic liquids are generally characterized by high boiling points, low vapor pressure, high ionic conductivity and low thermal conductivity, which in turn could lead to high Seebeck coefficients through higher entropy changes [42]. Detailed data on several redox couples and ionic liquids, TFSA(bis(trifluoromethylsulfonyl)amide, EMI+(1-ethyl-3-methylimidazolium), PP13+(1-methyl-1-propylpiperidinium, BMP+(1-butyl-1-methypyrrolidinium) and their blends, have been provided by these authors.
Redox reaction entropies in ionic liquids were found to depend strongly on Coulombic interactions between electrolyte ions and the redox couple. Abraham et al. [43] investigated the behaviour of an iodide/triiodide (I/I3) redox couple in 1-ethyl-3-methylimidazolium tetrafluoroborate; the temperature of the hot electrode was 130 °C. A Seebeck coefficient of 0.23 mV/K and power output of 29 mW/m2 were reported. Jiao et al. [44] have reported on the use of [Co(bpy)]+3/+2 redox couples in several ionic liquids (cation: 1-ethyl-3-methylimidazolium; anions: bis(trifluoromethylsulfonyl)imide, tetracynoborate, tris(pentafluoroethyl) trifluorophosphate etc.). The TEC response was found to depend linearly on ΔT (up to 140 °C) with Seebeck coefficients in the range 1.44 to 1.88 mV/K. Due to the generally high viscosity of ionic liquids, mass transport becomes a power-limiting factor, leading to slower rates of diffusion for redox ions. Kim et al. [45] have reported some of the highest Seebeck coefficients (2.9 mV/K) in mixed organic solvent electrolytes (20% methanol in water). The mass transport in electrolyte can also be improved by the addition of other species; the addition of MWCNTs to imidazolium-based ionic liquids lowered the ohmic resistance of the solution due to increased dissociation of ionic pairs and the formation of percolated networks [46]. However, the addition of CNTs to electrolytes containing Co2+/3+ (bpy)3 in [C2mim][NTf2] (see Dupont et al. [26] for nomenclature) showed reduced power with increases in viscosity outweighing the contributions from increases in conductivity [47]. Lithium-based ionic liquids, lithium bis(trifluoromethylsulfonyl)imide (Li[NTf2]) tetraglyme and a thiolate/disulfide redox organic couple in a mixture with 1-ethyl-3-methylimidazolium tetrafluoroborate/acetonitrile have also been tested to produce Seebeck coefficients of 1.4 and 0.6 mV/K, respectively [48][49].
While liquid electrolytes ensure fast ion diffusion, mass transport and high power density, possible leakage and low mechanical strength limit their application for wearable or flexible thermocells. Several quasi-solid-state electrolytes, as blends of polymeric solids with liquids, have been fabricated using a number of gelling agents such as polyvinyl alcohol, agar, poly(sodium acrylate) beads and cellulose [50]. Jin et al. [51] have reported on a quasi-solid-state electrolyte containing a cellulose-polymer matrix and a redox-active (ferri/ferrocyanide) aqueous phase. Organic gel electrolytes show excellent performance due to lower electronic conductivity, high ionic conductivity and good interfacial characteristics [52]. Yang et al. [53] fabricated two types of thermogalvanic gel electrolytes using two redox couples, aqueous ferri/ferrocyanide and ferric/ferrous chloride, with negative and positive Seebeck coefficients (−1.21 and 1.02 mV/K, respectively) gelled together using polyvinyl alcohol. This electrolyte was sandwiched between two polyamide electrodes and was connected in series to generate a larger potential difference; an output voltage up to 1 V could be generated in this system utilizing body heat.

3. Electrode Materials and Designs

Due to their non-reactive and thermally stable nature, platinum electrodes were used during early studies on thermocells. The electrolyte had to be stirred continuously to achieve Carnot efficiencies of 1.2% [54]. High power generation in thermocells requires large current densities, which may be achieved through high concentrations of redox mediators, increasing the number of sites for redox reactions and exposing electrodes to large thermal gradients [55]. Different types of thermocell electrodes and materials are detailed next.
Reversible metal electrodes: Thermocells using ‘reversible electrodes’ consist of soluble metal electrodes placed in solutions of corresponding salts at two different temperatures (Figure 1c). This cell can only work until a certain level of electrode consumption is reached; the direction of the temperature gradient is then reversed for the cell to work in the opposite direction. Most of the studies have reported on the Cu2+/Cu system due to its simple design and good reproducibility. Seebeck coefficients for this system could be modified from hypothetical values of 0.879 mV/K to 0.75–1.32 mV/K by changing experimental conditions and electrolyte additives [56]. Natural convection was found to have a significant influence on the efficiency of Cu/Cu2+-based thermogalvanic cells. Burmistrov et al. [57] have reported on the efficiency of thermocells based on copper, zinc and nickel metallic electrodes. While the efficiency parameters for Cu electrodes agreed well with the theoretical data and published results, the efficiency of zinc-based TEC cells was somewhat lower than theoretical expectations. The thermogalvanic system with nickel electrodes in an aqueous solution of nickel sulphate exhibited an unusual behaviour. The dependence of open-circuit voltage on temperature gradient exhibited an inflection point, with a Seebeck coefficient of 1.05 mV/K for ΔT < 30 °C and a Seebeck coefficient of 2.83 mV/K for ΔT = 30–60 °C. This result indicates the need for further research in the field towards improving thermocell efficiency.
Inert carbon electrodes: Carbon-based electrodes are becoming increasingly important as a promising and affordable alternative to platinum electrodes. Nanostructured carbon materials, such as single (SWNT) and multiwalled carbon nanotubes (MWNT) and graphene, have a large surface area, which helps in increasing the number of reaction sites. Depending on the number of nanotube walls, the specific surface area of CNTs lies in the range 50–1315 m2/g [58]. These materials also show fast electron-transfer kinetics for the ferrite/ferrocyanide redox couple. Both of these properties can increase the current density achieved with the thermocouple [59]. Most research in this area has been focused on composite materials involving nanostructured materials and additional treatment with additives or alloying [60]. The internal resistance of these electrodes fluctuates due to the porous structure; the highest power density (6.8 W/kg at a temperature difference of 20 °C) has been achieved by electrodes based on purified single-layer carbon nanotubes [55]. Similarly, carbon-nanotube (CNT) aerogel electrodes provide at their highest an efficiency of 3.95% in an aqueous ferri/ferrocyanide system. Hu et al. [61] have reported on a thermocell using 0.4 M ferri/ferrocyanide electrolyte, multiwalled nanotubes (MWNT) electrodes (10 nm dia., 0.5 cm2 area, 5 cm apart) and ΔT (61 °C), achieving an aerial power density of 1.36 W/m2. Composites of SWNTs with reduced graphene oxide in a 9:1 proportion, 0.45 µm thick, were able to generate an aerial power density of 0.46 W/m2
Hollow nickel microsphere electrodes: Burmistrov et al. [62] have shown that hollow Ni microspheres can be an effective electrode material for thermoelectrochemical cells and provide extremely high values of the hypothetical Seebeck coefficient and open circuit voltage. Electrodes were prepared by pressing tablets of nickel microspheres and reduced at 250 °C. The most effective composition, a KOH-based alkaline electrolyte, was chosen based on the influence of the alkali content on the hypothetical Seebeck coefficient value and specific power as determined by the number of charge carriers in the electrolyte and the intensity of reactions at the electrodes [63][64]. The basic electrode process in this system is described as
β-Ni(OH)2 + OH  β-NiOOH + H2O + e
This reaction proceeds on the hot electrode followed by electron release, resulting in the formation of a potential difference between the hot and cold electrodes (Figure 2). The accumulation of excess charge or heating on the nickel-oxide electrode results in the side reaction of β-NiOOH transforming into γ-NiOOH; the unstable γ-NiOOH gets hydrolysed to α-Ni(OH)2, which later transforms into β-Ni(OH)2. These reactions can lead to the absorption of electrons at the hot electrode and a lower thermocell potential. The Seebeck coefficient of these thermocells was found to reach 4.5 mV/K (ΔT up to 35 °C), one of the highest values reported for aqueous electrolytes.
Figure 2. (a) A schematic representation of KOH redox reactions mediated by Ni hollow microspheres, (b) and assembled cells with water heating/cooling and salt bridge, (c) polymer film body, (d) cell with adjustable distance between electrodes heating/cooling by Peltier elements, (e) structure view of the adjustable cell.
Additional electrode designs: The flexible nature of nano-carbons has led to new thermocell designs suitable for wrapping around curved or irregularly shaped surfaces such as vehicle exhaust pipes or cooling/heating lines in power plants. Hu et al. [61] used two layers of Nomex HT 4848 to separate two MWNT bucky paper electrodes at a distance of 2 mm; these scrolled electrodes increased the relative efficiency of the cell threefold to 1.4% relative to a Carnot engine. Thin coin-like thermocells have been developed using chemical vapor deposition by producing a catalyst with layers of Ti (30 nm), Al (10 nm) and Fe (2 nm) and using it to grow 100 µm tall MWNT forests on the thermocell casing. Nomex HT 4848 separators impregnated with 0.4 M ferri/ferrocyanide solution kept the electrodes apart; an aerial power density of 0.98 W/m2 was obtained for ΔT of 60 °C [30]. A stacked electrode configuration using SWNT and rGO (reduced graphene oxide) (in 9:1 proportion) consisting of layers of stainless-steel mesh as separators between 4.5 µm electrode films maintained a conductive path between individual films. A 10-stack configuration attained an efficiency of 2.63% relative to a Carnot engine. Marquardt et al. [65] have reported on a thermocell based on a proton-exchange membrane with hydrogen electrodes; the maximum open-circuit voltage with a power density of 45:3 mW/cm2 was observed at a humidifier temperature of 323 K. A Seebeck coefficient of 1.75 mV/K was observed for ΔT of 35 °C.

Additional electrode designs: The flexible nature of nano-carbons has led to new thermocell designs suitable for wrapping around curved or irregularly shaped surfaces such as vehicle exhaust pipes, cooling/heating lines in power plants. Hu et al. [81] used two layers of Nomex HT 4848 to separate two MWNT bucky paper electrodes at a distance of 2mm; these scrolled electrodes increased the relative efficiency of the cell threefold to 1.4% relative to a Carnot engine. Thin coin like thermocells have been developed using chemical vapor deposition by producing a catalyst with layers of Ti (30 nm), Al (10 nm) and Fe (2nm) and using it to grow 100µm tall MWNT forests on the thermocell casing. Nomex HT 4848 separators impregnated with 0.4 M ferri/ferro cyanide solution kept the electrodes apart; an aerial power density of 0.98 W/m2 was obtained for DT of 60ºC [30]. Stacked electrode configuration using SWNT and rGO (reduced graphene oxide) (in 9:1 proportion) consisting of layers of stainless-steel mesh as separators between 4.5µm electrode films maintaining a conductive path between individual films. A 10-stack configuration attained an efficiency of 2.63 % relative to a Carnot engine. Marquardt et al. [85] have reported on a thermocell based on a proton exchange membrane with hydrogen electrodes; maximum open circuit voltage with a power density of 45:3 mW/cm2 was observed at humidifier temperature of 323K. A Seebeck coefficient of 1.75 mV/K was observed for DT of 35ºC.

  1. Emerging applications

The conversion of waste heat to electrical energy is finding application in several fields. Some of the emerging applications are presented in this section. While some of these are stand alone on TE effect, others are in conjunction with other technologies.

Wearable and portable devices: Wearable electronic devices are gaining attention in the health and performance monitoring sectors towards long-term, continuous, self-powered operations using human body as a continuous supply of energy. Harvesting of body heat in wearable devices has been investigated extensively as an alternative to bulky batteries that require frequent charging or replacement [86,87] . As the core body temperature is regulated at 37°C, body heat can be a continuous source of energy; the total heat dissipated from the human body can range between 60 to 180W depending on the activity level [88]. As only a small fraction of body surface can be covered by TEG devices, overall efficiencies are likely to be quite small. Of the several approaches used for harvesting body heat such as electrostatic, piezoelectric, electromagnetic, pyroelectric and thermoelectric effects, we will focus our attention on thermoelectric mechanisms [89]. Wearable electronic devices find application in medical devices (real time monitoring, blood pressure sensors, ear-ware), smart watches, sportswear, wristbands, flexible devices for monitoring non-flat surfaces in industrial applications [90].

Among the well-known thermoelectric materials, alloys of bismuth telluride with antimony and selenium have been investigated extensively for room temperature operations. (BixSb1-x)Te3 alloy system is a p- type semiconductor legs with ZT ranging between 1-2; Bi2Te3-xSe x is an n-type semiconductor legs with ZT  around 1 [91–93]. An array of p- and n-type semiconductor legs are arranged in a thermoelectric module and are electrically connected in series, and then thermally connected in parallel towards greater overall efficiency [94]. A schematic representation of a typical TEG circuit in the form of a wrist band is shown in Fig. 3 for harvesting body heat. 

 

Fig. 3.   A schematic representation of a TEG wrist band for harvesting body heat

 

In rigid TEGs, n- and p-legs and interconnects are typically affixed to a thermally conductive ceramic substrate (alumina or aluminum nitride). In flexible TEGs, Gallium-Indium eutectic alloys have been used as flexible liquid-metal interconnects to ascertain integrity during operation; Kapton HN and polydimethylsiloxane are the two most common substrates [95]. Typical power outputs are found to range between 2 to 8.75µW/cm2 [96]. Liu et al [97] have presented several different designs for wearable devices for harvesting body heat in a range of applications.

Organic materials have also been investigated as flexible TEGs due to lower prices, light weight, low thermal conductivity, and high flexibility. While significant progress has been made, high performance organic TEGs still lag behind inorganic chalcogenides BiTe-based alloys. One of the strategies is to fabricate flexible TE devices from composites of organic and polymeric materials with micro/nanomaterials [98]. The TE properties of several p-type and n-type coordination polymers have been tabulated by Masoumi et al. [98]. π conjugated conducting polymers, are another group of organic TEs, with a capacity for doping with a wide variety of elements and adjustable doping levels [95]. Some of key conducting polymers include polyacetylene, polyvinylidene fluoride, polyphenylenevinylene, poly(3-hexylthiophene), polypyrrole etc. A power factor of 401 µW/ mK2 and a Seebeck coefficient of 43.5 µV/K has been achieved with some conducting polymers [99]. Other TE materials include carbon nanotubes, graphene, fullerenes, carbon nanodots, small molecules, organic composites with inorganic fillers etc. [94].

Energy storage devices: The integration of energy-harvesting and storage devices has been extensively investigated for emerging self-powered electronic devices [97]. Thermoelectric generators can be used to convert excess heat generated during the operation of electronic devices into electricity [93].  Using Soret effect, ionic thermoelectric supercapacitors utilize ionic electrolytes to produce charges and energy storage in a single device; the operation of these devices is however limited by long charging and discharging times and a rigid configuration [100]. Yang et al. [101] have reported on a TEG  device with n- and p- type modules consisting of Ag2Te and Ag2Se nanoparticle thin films. This device was directly linked with a planar micro supercapacitor. A Seebeck voltage of 82 mV was generated for DT of 15.8 K and a charging efficiency of 98%. Wu et al. [102] have reported on novel conjugated conducting polymer (PDAQ-BC) [DAQ: 2,6-diaminoanthraquinone; DBC: 3,6-dibromo-9-(4-bromophenyl)carbazole], which showed a specific capacitance of 180.5 F/g for a current of 1A/g. This polymer retained up to 95% coulombic efficiency and 85% capacitance after 5000 cycles of operation.

Park et al. [103] have reported an all-in-one energy system consisting of a TEG on one side of the substrate and a micro supercapacitor (MSC) on the other side. The TEG was constructed from screen printed p- and n- modules and a p- type TE film for alignment with electrodes. An MSC was fabricated on the other side using rGO/CNT electrodes and Au current collectors. Electrodes were positioned above and below the TEG-substrate-MSC system; the electrode on the TEG side was connected externally to the current collector of MSC. This system was able to generate 10.8V of electrical energy for thermal differences up to 10K and store it without loss.

Liu et al. [107] have reported on coupling thermoelectricity and electrocatalysis for hydrogen production via PbTe-PbS/TiO2 heterojunction involving a cathode, an anode and electrolyte operating at the hot-end and in-situ endothermic production of electrochemical hydrogen on the cold-end. At 70ºC and 1.0V bath voltage, hydrogen was generated at the rate of 6.1mL/cm2/h with energy and heat efficiencies reaching 88% and 49.9% respectively. Liu et al. [108] have reported on an advanced Li ion battery with charging based on a coupled thermoelectric approach. Richards et al. [109] have explored solid-state electrochemical heat engines (EHE) for generating electric power from available thermal energy based on reversible redox reactions. EHEs control and utilize the electrochemical potential of molecules undergoing redox reactions with temperature, composition and pressure playing an important role.

Liao et al. [110] have reported on a hybrid active/passive battery thermal management system in combination with thermoelectric elements (TEE) and phase change materials for the management of Li-ion batteries operating in extreme environments. TEEs were used to provide refrigeration at high temperatures and heating to preheat the batteries in cold environments. Nguyen and Shabani [111] have discussed about capturing the heat produced by proton exchange membrane fuel cells, heat recovery solutions and opportunities for integration with TEGs and thermally regenerative electrochemical cycles for power cogeneration applications.

 

  1. Conclusions and future perspectives

Development of thermoelectrochemical technologies for harvesting waste low-temperature heat opens the prospect of increasing the efficiency of various devices and mechanisms operating in exothermic mode or creating systems for generating electricity based on natural heat sources. Any process that can partially replace fossil fuels as a prime energy source will be used only if it is attractive to industry; an alternative energy source such as waste heat conversion (WHC) has to cost effective to becoming commercially viable. Geoffroy et al. [112] have shown that current WHC heat engines are not economically viable below 100ºC and require temperatures above 150ºC coupled with 100-1000 kW power outputs to be economically competitive. Studies in recent years have shown the possibility of significant increases in the power and conversion efficiency of TEC cells. Highest values of output power and cell potentials have been achieved for the redox ferri/ferrocyanide system and Co2+/3+, which offers great opportunities for further development and research in both aqueous and non-aqueous solvents.

Achieved results show the pathways to overcome key fundamental limitations of thermocell performance and set new tasks for fundamental research and further development of electrodes, electrolyte materials and cell design. One of the key tasks in thermocells development is to investigate mechanisms of entropy change via new redox couples and electrolytes. This will be related to increasing the hypothetical Seebeck coefficient, as well as improving the properties of electrodes and solvents to increase the mass transfer rate (diffusion capacity) towards increasing exchange currents and output power values.

One of the most promising thermoelectric power generation application involves vehicle waste heat recovery to improve fuel economy, wherein waste heat from the exhaust, is redirected to produce electricity. Other applications include harvesting industrial waste heat (incinerators, cement, steel mills etc.), geothermal, fuel oil-fired furnaces or gas water-heaters with this technology. Despite extensive research, WHC technology has yet to achieve significant market penetration. For WHCs to become be a serious contender, it has to compete with solar, wind, geothermal technologies in terms of capacity factors, capital costs, operational as well as maintenance costs. Significant developments are therefore required on multiple fronts towards achieving greater power density performances especially at higher temperatures. It is possible that TECs might have major advantages over solar cells and semiconductor thermoelectrics, in particular in the Wh/dollar efficiency [81]. Achieving industrial production volumes of MWCNTs and reducing their cost, according to the author, can make TEC a competitive device.

 

 

This entry is adapted from the peer-reviewed paper 10.3390/su14159483

References

  1. Forman, C.; Muritala, I. K.; Pardemann, R.; Meyer, B. Estimating the global waste heat potential. Renewable and Sustainable Energy Reviews 2016, 57, 1568–1579.
  2. Straub, A. P.; Yip, N. Y.; Lin, S.; Lee, J.; Elimelech, M. Harvesting low-grade heat energy using thermo-osmotic vapour transport through nanoporous membranes. Nature Energy 2016, 1, 1–6.
  3. Yang, Y.; Lee, S. W.; Ghasemi, H.; Loomis, J.; Li, X.; Kraemer, D.; Zheng, G.; Cui, Y.; Chen, G. Charging-free electrochemical system for harvesting low-grade thermal energy. Proc Natl Acad Sci U S A 2014, 111, 17011–17016.
  4. Evans, S. CarbonBrief/Energy, 2017. Solar, wind and nuclear have ‘amazingly low’ carbon footprints. Available online: https://www.carbonbrief.org/solar-wind-nuclear-amazingly-low-carbon-footprints/#:~:text=Simon%20Evans,-08.12.2017%20%7C%205&text=Building%20solar%2C%20wind%20or%20nuclear,of%20electricity%20out%20to%202050. (accessed on 10 Jan. 2022).
  5. COP26 Outcomes, 2021. Available online: https://ukcop26.org/the-conference/cop26-outcomes. (accessed on 10 Jan. 2022).
  6. Bouty, E. Phénomènes thermo-électriques et électro-thermiques au contact d’un métal et d’un liquide. Journal de Physique Théorique et Appliquée 1880, 9, 306–320.
  7. Snyder, G. J.; Toberer, E. S. Complex thermoelectric materials. Materials for Sustainable Energy: A Collection of Peer-Reviewed Research and Review Articles from Nature Publishing Group 2010, 101–110.
  8. Jangonda, C.; Patil, K.; Kinikar, A.; Bhokare, R.; Gavali, M. D. Review of various application of thermoelectric module. Intl. J. Innovative Research in Science, Engineering and Technology 2016, 5, 3393.
  9. Yang, X.; Wang, C.; Lu, R.; Shen, Y.; Zhao, H.; Li, J.; Zheng, X; Progress in Measurement of Thermoelectric Properties of Micro/Nano Thermoelectric Materials: A Critical Review. Nano Energy 2022, 107553.
  10. Chen, X. Q.; Fan, S. J.; Han, C.; Wu, T.; Wang, L. J.; Jiang, W.; Yang, J. P. Multiscale architectures boosting thermoelectric performance of copper sulfide compound. Rare metals 2021, 40(8), 2017-2025.
  11. Chen, X.; Zhang, H.; Zhao, Y.; Liu, W. D.; Dai, W.; Wu, T.; Yang, J. Carbon-encapsulated copper sulfide leading to enhanced thermoelectric properties. ACS applied materials & interfaces 2019, 11(25), 22457-22463.
  12. Vining, C. B. An inconvenient truth about thermoelectrics. Nature Materials 2009, 8, 83–85.
  13. deBethune, A. J.; Licht, T. S.; Swendeman, N. The Temperature Coefficients of Electrode Potentials. Journal of The Electrochemical Society 1959, 106, 616.
  14. Landry, B. A. Utilization of waste heat. Science 1953, 3, 3.
  15. Wakao, N.; Nojo, K. Nitric acid cycle process for extracting thermal energy from low-level heat sources. Nature 1978, 273, 25–27.
  16. Venkatasubramanian, R.; Siivola, E.; Colpitts, T.; O’Quinn, B. Thin-film thermoelectric devices with high room-temperature figures of merit. Nature 2001, 413, 597–602.
  17. Waske, A.; Dzekan, D.; Sellschopp, K.; Berger, D.; Stork, A.; Nielsch, K.; Fähler, S. Energy harvesting near room temperature using a thermomagnetic generator with a pretzel-like magnetic flux topology. Nature Energy 2018, 4, 68–74.
  18. Li, T.; Zhang, X.; Lacey, S. D.; Mi, R.; Zhao, X.; Jiang, F.; Song, J.; Liu, Z.; Chen, G.; Dai, J.; et al. Cellulose ionic conductors with high differential thermal voltage for low-grade heat harvesting. Nature Materials 2019, 18, 608–613.
  19. Yu, B.; Duan, J.; Cong, H.; Xie, W.; Liu, R.; Zhuang, X.; Wang, H.; Qi, B.; Xu, M.; Wang, Z. L.; et al. Thermosensitive crystallization-boosted liquid thermocells for low-grade heat harvesting. Science 2020, 370, 342–346.
  20. Pandya, S.; Wilbur, J.; Kim, J.; Gao, R.; Dasgupta, A.; Dames, C.; Martin, L. W. Pyroelectric energy conversion with large energy and power density in relaxor ferroelectric thin films. Nature Materials 2018, 17, 432–438.
  21. Thakre, A.; Kumar, A.; Song, H. C.; Jeong, D. Y.; Ryu, J. Pyroelectric Energy Conversion and Its Applications—Flexible Energy Harvesters and Sensors. Sensors 2019, Vol. 19, Page 2170 2019, 19, 2170.
  22. Quoilin, S.; Broek, M. van den; Declaye, S.; Dewallef, P.; Lemort, V. Techno-economic survey of Organic Rankine Cycle (ORC) systems. Renewable and Sustainable Energy Reviews 2013, 22, 168–186.
  23. Zhang, X.; Chau, K.T. An automotive thermoelectric photovoltaic hybrid energy system using maximum power point tracking. Energy Convers Manag 2011, 52(1), 641e7.
  24. Zhang, H.; W Kong, W.; Dong, F.; Xu, H.; Chen, B.; Ni, M. Application of cascading thermoelectric generator and cooler for waste heat recovery from solid oxide fuel cells. Energy Convers Manag 2017, 148, 1382-1390.
  25. Li, P.; Cai, L.; Zhai, P.; Tang, X.; Zhang, Q.; Nino, M. Design of a concentration solar thermoelectric generator. J Electron Mater 2010, 39(9), 1522-1530.
  26. Dupont, M. F.; MacFarlane, D. R.; Pringle, J. M. Thermo-electrochemical cells for waste heat harvesting – progress and perspectives. Chemical Communications 2017, 53, 6288–6302.
  27. Chum, H. L.; Osteryoung, R. A. Review of thermally regenerative electrochemical cells. Solar Energy Research Institute 1981.
  28. Lalancette, J.-M.; Roussel, R. Metals intercalated in graphite. V. A concentration cell with intercalated bromine. Canadian Journal of Chemistry 1976, 54, 3541–3544.
  29. Endo, M.; Yamagishi, Y.; Inagaki, M. Thermocell with graphite fiber-bromine intercalation compounds. Synthetic Metals 1983, 7, 203–209.
  30. Inagaki, M.; Matsumoto, A.; Sakai, M.; Maeda, Y. A cell of carbon-fibers and nitric acid with temperature difference. Nippon Kagaku Kaishi 1983, 309–311.
  31. Bonetti, M.; Nakamae, S.; Roger, M.; Guenoun, P. Huge Seebeck coefficients in nonaqueous electrolytes. The Journal of Chemical Physics 2011, 134, 114513.
  32. Burmistrov, I.; Gorshkov, N.; Kovyneva, N.; Kolesnikov, E.; Khaidarov, B.; Karunakaran, G.; Cho, E. B.; Kiselev, N.; Artyukhov, D.; Kuznetsov, D.; et al. High seebeck coefficient thermo-electrochemical cell using nickel hollow microspheres electrodes. Renewable Energy 2020, 157, 1–8.
  33. Im, H.; Kim, T.; Song, H.; Choi, J.; Park, J. S.; Ovalle-Robles, R.; Yang, H. D.; Kihm, K. D.; Baughman, R. H.; Lee, H. H.; et al. High-efficiency electrochemical thermal energy harvester using carbon nanotube aerogel sheet electrodes. Nature Communications 2016, 7.
  34. Artyukhov D.; Kiselev N.; Gorshkov N.; Kovyneva N.; Ganzha O.; Vikulova M.; Gorokhovsky A.; Offor P.; Boychenko E.; Burmistrov I. Harvesting Waste Thermal Energy Using a Surface-Modified Carbon Fiber-Based Thermo-Electrochemical Cell Sustainability 2021, 13(3), 1377.
  35. Inagaki, M.; Itoh, E.; Maeda, Y. Durable Performance of Thermocell with Carbon Cloth and Nitric Acid. TANSO 1985, 1985, 134–136.
  36. Battistel, A.; Peljo, P. Recent trends in thermoelectrochemical cells and thermally regenerative batteries. Current Opinion in Electrochemistry 2021, 30, 100853.
  37. Duan, J.; Yu, B.; Huang, L.; Hu, B.; Xu, M.; Feng, G.; Zhou, J. Liquid-state thermocells: Opportunities and challenges for low-grade heat harvesting. Joule 2021, 5, 768–779.
  38. Disalvo, F. J. Thermoelectric cooling and power generation. Science (1979) 1999, 285, 703–706.
  39. Abraham, T. J.; Macfarlane, D. R.; Baughman, R. H.; Jin, L.; Li, N.; Pringle, J. M. Towards ionic liquid-based thermoelectrochemical cells for the harvesting of thermal energy. Electrochimica Acta 2013, 113, 87–93.
  40. Romano, M. S.; Razal, J. M.; Antiohos, D.; Wallace, G.; Chen, J. Nano-carbon electrodes for thermal energy harvesting. Journal of Nanoscience and Nanotechnology 2015, 15, 1–14.
  41. He, J.; Al-Masri, D.; MacFarlane, D. R.; Pringle, J. M. Temperature dependence of the electrode potential of a cobalt-based redox couple in ionic liquid electrolytes for thermal energy harvesting. Faraday Discussions 2016, 190, 205–218.
  42. Kang, T. J.; Fang, S.; Kozlov, M. E.; Haines, C. S.; Li, N.; Kim, Y. H.; Chen, Y.; Baughman, R. H. Electrical Power From Nanotube and Graphene Electrochemical Thermal Energy Harvesters. Advanced Functional Materials 2012, 22, 477–489.
  43. Zhang, L.; Kim, T.; Li, N.; June Kang, T.; Chen, J.; Pringle, J. M.; Zhang, M.; Kazim, A. H.; Fang, S.; Haines, C.; et al. High Power Density Electrochemical Thermocells for Inexpensively Harvesting Low-Grade Thermal Energy. Advanced Materials 2017, 29, 1605652.
  44. Maeda, Y.; Kitamura, H.; Itoh, E.; Inagaki, M. A new carbon fiber and nitric acid cell with a temperature difference between electrodes. Synthetic Metals 1983, 7, 211–217.
  45. Mua, Y.; Quickenden, T. I. Power Conversion Efficiency, Electrode Separation, and Overpotential in the Ferricyanide/Ferrocyanide Thermogalvanic Cell. Journal of The Electrochemical Society 1996, 143, 2558–2564.
  46. Hinterleitner, B.; Knapp, I.; Poneder, M.; Shi, Y.; Müller, H.; Eguchi, G.; Eisenmenger-Sittner, C.; Stöger-Pollach, M.; Kakefuda, Y.; Kawamoto, N.; et al. Thermoelectric performance of a metastable thin-film Heusler alloy. Nature 2019 576:7785 2019, 576, 85–90.
  47. Li, D.; Sun, R. R.; Qin, X. Y. Improving thermoelectric properties of p-type Bi2Te3-based alloys by spark plasma sintering. Progress in Natural Science: Materials International 2011, 21, 336–340.
  48. Russ, B.; Glaudell, A.; Urban, J. J.; Chabinyc, M. L.; Segalman, R. A. Organic thermoelectric materials for energy harvesting and temperature control. Nature Reviews Materials 2016, 1, 1–14.
  49. Li, M.; Hong, M.; Dargusch, M.; Zou, J.; Chen, Z. G. High-efficiency thermocells driven by thermo-electrochemical processes. Trends in Chemistry 2021, 3, 561–574.
  50. Cho, C.; Stevens, B.; Hsu, J.-H.; Bureau, R.; Hagen, D. A.; Regev, O.; Yu, C.; Grunlan, J. C.; Cho, C.; Stevens, B.; et al. Completely Organic Multilayer Thin Film with Thermoelectric Power Factor Rivaling Inorganic Tellurides. Advanced Materials 2015, 27, 2996–3001.
  51. Wang, H.; Hsu, J.-H.; Yi, S.-I.; Lae Kim, S.; Choi, K.; Yang, G.; Yu, C.; Wang, H.; Yi, S.; Kim, S. L.; et al. Thermally Driven Large N-Type Voltage Responses from Hybrids of Carbon Nanotubes and Poly(3,4-ethylenedioxythiophene) with Tetrakis(dimethylamino)ethylene. Advanced Materials 2015, 27, 6855–6861.
  52. Wang, W.; Shu, G.; Tian, H.; Huo, D.; Zhu, X. A bimetallic thermally-regenerative ammonia-based flow battery for low-grade waste heat recovery. Journal of Power Sources 2019, 424, 184–192.
  53. Cheng, C.; Dai, Y.; Yu, J.; Liu, C.; Wang, S.; Feng, S. P.; Ni, M. Review of Liquid-Based Systems to Recover Low-Grade Waste Heat for Electrical Energy Generation. Energy and Fuels 2021, 35, 161–175.
  54. Hu, R.; Xu, D.; Luo, X. Liquid Thermocells Enable Low-Grade Heat Harvesting. Matter 2020, 3, 1400–1402.
  55. Black, J. J.; Murphy, T.; Atkin, R.; Dolan, A.; Aldous, L. The thermoelectrochemistry of lithium–glyme solvate ionic liquids: Towards waste heat harvesting. Physical Chemistry Chemical Physics 2016, 18, 20768–20777.
  56. Zhou, H.; Liu, P. High Seebeck Coefficient Electrochemical Thermocells for Efficient Waste Heat Recovery. ACS Applied Energy Materials 2018, 1, 1424–1428.
  57. Burmistrov, I.; Artyukhov, D.; Shindrov, A.; Gorshkov N.; Gorokhovsky, A. Thermo-Electrochemical Cells for Low-Grade Waste Heat Conversion. In Nanotech Middle East 2017 Conference and Exhibition. Dubai; Dubai, 2017; pp. 4–6.
  58. Sahami, S.; Weaver, M. J. Entropic and enthalpic contributions to the solvent dependence of the thermodynamics of transition-metal redox couples: Part II. Couples containing ammine and ethylenediamine ligands. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry 1981, 122, 171–181.
  59. Migita, T.; Tachikawa, N.; Katayama, Y.; Miura, T. Thermoelectromotive force of some redox couples in an amide-type room-temperature ionic liquid. Electrochemistry 2009, 77, 639–641.
  60. Laux, E.; Uhl, S.; Journot, T.; Brossard, J.; Jeandupeux, L.; Keppner, H. Aspects of Protonic Ionic Liquid as Electrolyte in Thermoelectric Generators. Journal of Electronic Materials 2016 45:7 2016, 45, 3383–3389.
  61. Anari, E. H. B.; Romano, M.; Teh, W. X.; Black, J. J.; Jiang, E.; Chen, J.; To, T. Q.; Panchompoo, J.; Aldous, L. Substituted ferrocenes and iodine as synergistic thermoelectrochemical heat harvesting redox couples in ionic liquids. Chemical communications 2016, 52, 745–748.
  62. Zinovyeva, V.; Nakamae, S.; Bonetti, M.; Roger, M. Enhanced Thermoelectric Power in Ionic Liquids. ChemElectroChem 2014, 1, 426–430.
  63. Kim, T.; Lee, J. S.; Lee, G.; Yoon, H.; Yoon, J.; Kang, T. J.; Kim, Y. H. High thermopower of ferri/ferrocyanide redox couple in organic-water solutions. Nano Energy 2017, 31, 160–167.
  64. Wang, H.; Zhao, D.; Ulla Khan, Z.; Puzinas, S.; Jonsson, M. P.; Berggren, M.; Crispin, X.; Wang, H.; Zhao, D.; Khan, Z. U.; et al. Ionic Thermoelectric Figure of Merit for Charging of Supercapacitors. Advanced Electronic Materials 2017, 3, 1700013.
  65. Krebs, S. Performance analysis of a Copper II Sulfate Pentahydrate based thermogalvanic cell. Electronic Theses and Dissertations 2015.
  66. Cabral, D. M.; Howlett, P. C.; MacFarlane, D. R. Electrochemistry of the tris(2,2‘-bipyridine) complex of iron(II) in ionic liquids and aprotic molecular solvents. Electrochimica Acta 2016, 220, 347–353.
  67. Yamato, Y.; Katayama, Y.; Miura, T. Effects of the Interaction between Ionic Liquids and Redox Couples on Their Reaction Entropies. Journal of The Electrochemical Society 2013, 160, H309–H314.
  68. Jiao, N.; Abraham, T. J.; MacFarlane, D. R.; Pringle, J. M. Ionic Liquid Electrolytes for Thermal Energy Harvesting Using a Cobalt Redox Couple. Journal of The Electrochemical Society 2014, 161, D3061–D3065.
  69. Salazar, P. F.; Stephens, S. T.; Kazim, A. H.; Pringle, J. M.; Cola, B. A. Enhanced thermo-electrochemical power using carbon nanotube additives in ionic liquid redox electrolytes. Journal of Materials Chemistry A 2014, 2, 20676–20682.
  70. Kazim, A. H.; Cola, B. A. Electrochemical Characterization of Carbon Nanotube and Poly (3,4-ethylenedioxythiophene)−Poly(styrenesulfonate) Composite Aqueous Electrolyte for Thermo-Electrochemical Cells. Journal of The Electrochemical Society 2016, 163, F867–F871.
  71. Wu, J.; Black, J. J.; Aldous, L. Thermoelectrochemistry using conventional and novel gelled electrolytes in heat-to-current thermocells. Electrochimica Acta 2017, 225, 482–492.
  72. Jin, L.; Greene, G. W.; MacFarlane, D. R.; Pringle, J. M. Redox-Active Quasi-Solid-State Electrolytes for Thermal Energy Harvesting. ACS Energy Letters 2016, 1, 654–658.
  73. Xiao, Y.; Zhong, X.; Guo, J.; Zhou, C.; Zuo, H.; Liu, Q.; Huang, Q.; Zhang, Q.; Diao, X. The role of interface between LiPON solid electrolyte and electrode in inorganic monolithic electrochromic devices. Electrochimica Acta 2018, 260, 254–263.
  74. Yang, P.; Liu, K.; Chen, Q.; Mo, X.; Zhou, Y.; Li, S.; Feng, G.; Zhou, J. Wearable thermocells based on gel electrolytes for the utilization of body heat. Angewandte Chemie 2016, 128, 12229–12232.
  75. Quickenden, T. I.; Mua, Y. A Review of Power Generation in Aqueous Thermogalvanic Cells. Journal of The Electrochemical Society 1995, 142, 3985–3994.
  76. Gunawan, A.; Li, H.; Lin, C. H.; Buttry, D. A.; Mujica, V.; Taylor, R. A.; Prasher, R. S.; Phelan, P. E. The amplifying effect of natural convection on power generation of thermogalvanic cells. International Journal of Heat and Mass Transfer 2014, 78, 423–434.
  77. Burmistrov, I.; Kovyneva, N.; Gorshkov, N.; Gorokhovsky, A.; Durakov, A.; Artyukhov, D.; Kiselev, N. Development of new electrode materials for thermo-electrochemical cells for waste heat harvesting. Renewable Energy Focus 2019, 29, 42–48.
  78. Koo, M. H.; Yoon, H. H. Fabrication of carbon nanotubes and charge transfer complex-based electrodes for a glucose/oxygen biofuel cell. Journal of Nanoscience and Nanotechnology 2013, 13, 7434–7438.
  79. Nugent, J. M.; Santhanam, K. S. V.; Rubio, A.; Ajayan, P. M. Fast Electron Transfer Kinetics on Multiwalled Carbon Nanotube Microbundle Electrodes. Nano Letters 2001, 1, 87–91.
  80. Hu, R.; A. Cola, B.; Haram, N.; N. Barisci, J.; Lee, S.; Stoughton, S.; Wallace, G.; Too, C.; Thomas, M.; Gestos, A.; et al. Harvesting Waste Thermal Energy Using a Carbon-Nanotube-Based Thermo-Electrochemical Cell. Nano Letters 2010, 10, 838–846.
  81. Kiselev, N.; Artyukhov, D.; Boychenko, E.; Gorshkov, N.; Glubokaya, A.; Burmistrov, I. Electrolyte concentration dependences of NiO based thermoelectrochemical cells performance. AIP Conference Proceedings 2022, 2456, 020005.
  82. Taganova, A. A.; Boychenko, E. A.; Kiselev, N. v.; Khaidarov, B. B.; Kolesnikov, E. A.; Yudin, A. G.; Vikulova, M. A.; Gorshkov, N. v.; Kuznetsov, D. v.; Burmistrov, I. N. Synthesis and Study of the Composition of Hollow Microspheres of NiO and NiO/Ni Composition for Thermoelectrochemical Energy Converters of Low-Potential Temperature Gradients of Thermal Units Into Electricity. Refractories and Industrial Ceramics 2021, 61, 715–719.
  83. Marquardt, T.; Kube, J.; Radici, P.; Kabelac, S. Experimental investigation of a thermocell with proton exchange membrane and hydrogen electrodes. International Journal of Hydrogen Energy 2020, 45, 12680–12690.
  84. Misra, V.; Bozkurt, A.; Calhoun, B.; Jackson, T.; Jur, J.; Lach, J.; Lee, B.; Muth, J.; Oralkan, O.; Ozturk, M.; et al. Flexible technologies for self-powered wearable health and environmental sensing. Proceedings of the IEEE 2015, 103, 665–681.
  85. Ando Junior, O. H.; Maran, A. L. O.; Henao, N. C. A review of the development and applications of thermoelectric microgenerators for energy harvesting. Renewable and Sustainable Energy Reviews 2018, 91, 376–393.
  86. Riemer, R.; Shapiro, A. Biomechanical energy harvesting from human motion: Theory, state of the art, design guidelines, and future directions. Journal of NeuroEngineering and Rehabilitation 2011, 8, 1–13.
  87. Invernizzi, F.; Dulio, S.; Patrini, M.; Guizzetti, G.; Mustarelli, P. Energy harvesting from human motion: materials and techniques. Chemical Society Reviews 2016, 45, 5455–5473.
  88. Llamas R. IDC Media Center. Worldwide Wearables Market to Nearly Double by 2021, According to IDC Internet Available online: https://www.idc.com/getdoc.jsp?containerId=prUS42818517 (accessed 22.11. 2021).
  89. Xu, Z.; Wu, H.; Zhu, T.; Fu, C.; Liu, X.; Hu, L.; He, J.; He, J.; Zhao, X. Attaining high mid-temperature performance in (Bi,Sb)2Te3 thermoelectric materials via synergistic optimization. NPG Asia Materials 2016 8:9 2016, 8, e302–e302.
  90. Nozariasbmarz, A. In-situ sintering decrystallization of thermoelectric materials using microwave radiation; North Carolina State University, 2017.
  91. Wang, S.; Tan, G.; Xie, W.; Zheng, G.; Li, H.; Yang, J.; Tang, X. Enhanced thermoelectric properties of Bi2(Te1−xSex)3-based compounds as n-type legs for low-temperature power generation. Journal of Materials Chemistry 2012, 22, 20943–20951.
  92. Leonov, V. Thermoelectric energy harvester on the heated human machine*. Journal of Micromechanics and Microengineering 2011, 21, 125013.
  93. Suarez, F.; Parekh, D. P.; Ladd, C.; Vashaee, D.; Dickey, M. D.; Öztürk, M. C. Flexible thermoelectric generator using bulk legs and liquid metal interconnects for wearable electronics. Applied Energy 2017, 202, 736–745.
  94. Park, H.; Lee, D.; Kim, D.; Cho, H.; Eom, Y.; Hwang, J.; Kim, H.; Kim, J.; Han, S.; Kim, W. High power output from body heat harvesting based on flexible thermoelectric system with low thermal contact resistance. Journal of Physics D: Applied Physics 2018, 51, 365501.
  95. Fan, Z.; Ouyang, J.; Fan, Z.; Ouyang, J. Thermoelectric Properties of PEDOT:PSS. Advanced Electronic Materials 2019, 5, 1800769.
  96. Liu, Y.; Wang, H.; Sherrell, P.C.; Liu, L.; Wang, Y.; Chen, J. Potentially Wearable Thermo-Electrochemical Cells for Body Heat Harvesting: From Mechanism, Materials, Strategies to Applications. Adv. Sci. 2021, 8, 2100669.
  97. Masoumi, S.; O’Shaughnessy, S.; Pakdel, A. Organic-based flexible thermoelectric generators: From materials to devices. Nano Energy 2022, 92, 106774.
  98. Peng, S.; Wang, D.; Lu, J.; He, M.; Xu, C.; Li, Y.; Zhu, S. A Review on Organic Polymer-Based Thermoelectric Materials. Journal of Polymers and the Environment 2017, 25, 1208–1218.
  99. Guan, X.; Cheng, H.; Ouyang, J. Significant enhancement in the Seebeck coefficient and power factor of thermoelectric polymers by the Soret effect of polyelectrolytes. Journal of Materials Chemistry A 2018, 6, 19347–19352.
  100. Yang, S.; Cho, K.; Park, Y.; Kim, S. Bendable thermoelectric generators composed of p- and n-type silver chalcogenide nanoparticle thin films. Nano Energy 2018, 49, 333–337.
  101. Kim, S. J.; Lee, H. E.; Choi, H.; Kim, Y.; We, J. H.; Shin, J. S.; Lee, K. J.; Cho, B. J. High-Performance Flexible Thermoelectric Power Generator Using Laser Multiscanning Lift-Off Process. ACS Nano 2016, 10, 10851–10857.
  102. Zhao, D.; Wang, H.; Khan, Z. U.; Chen, J. C.; Gabrielsson, R.; Jonsson, M. P.; Berggren, M.; Crispin, X. Ionic thermoelectric supercapacitors. Energy & Environmental Science 2016, 9, 1450–1457.
  103. Yang, K.; Cho, K.; Yang, S.; Park, Y.; Kim, S. A laterally designed all-in-one energy device using a thermoelectric generator-coupled micro supercapacitor. Nano Energy 2019, 60, 667–672.
  104. Wu, X.; Huang, B.; Wang, Q.; Wang, Y. Thermally chargeable supercapacitor using a conjugated conducting polymer: Insight into the mechanism of charge-discharge cycle. Chemical Engineering Journal 2019, 373, 493–500.
  105. Park, Y.; Cho, K.; Kim, S. Vertical all-in-one energy systems constructed with thermoelectric generators and microsupercapacitors. Journal of Power Sources 2021, 510, 230402.
  106. Liu, Z.; Cao, X.; Wang, B.; Xia, M.; Lin, S.; Guo, Z.; Zhang, X.; Gao, S. Coupling thermoelectricity and electrocatalysis for hydrogen production via PbTePbS/TiO2 heterojunction. Journal of Power Sources 2017, 342, 452–459.
  107. Liu, K.; Li, K.; Yang, Z.; Zhang, C.; Deng, J. An advanced Lithium-ion battery optimal charging strategy based on a coupled thermoelectric model. Electrochimica Acta 2017, 225, 330–344.
  108. Richards, G.; Gemmen, R. S.; Williams, M. C. Solid – state electrochemical heat engines. International Journal of Hydrogen Energy 2015, 40, 3719–3725.
  109. Liao, G.; Jiang, K.; Zhang, F.; E, J.; Liu, L.; Chen, J.; Leng, E. Thermal performance of battery thermal management system coupled with phase change material and thermoelectric elements. Journal of Energy Storage 2021, 43, 103217.
  110. Nguyen, H. Q.; Shabani, B. Proton exchange membrane fuel cells heat recovery opportunities for combined heating/cooling and power applications. Energy Conversion and Management 2020, 204, 112328.
  111. Geffroy, C.; Lilley, D.; Parez, P. S.; Prasher, R. Techno-economic analysis of waste-heat conversion. Joule 2021, 5, 3080–3096.
More
This entry is offline, you can click here to edit this entry!