Next Article in Journal
A Novel Solution for Solving the Frequency Regulation Problem of Renewable Interlinked Power System Using Fusion of AI
Next Article in Special Issue
Influence of Thermoelectric Properties and Parasitic Effects on the Electrical Power of Thermoelectric Micro-Generators
Previous Article in Journal
Renewable Electricity Generation in Small Island Developing States: The Effect of Importing Ammonia
Previous Article in Special Issue
Additive Manufacturing of Bulk Thermoelectric Architectures: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Review on Wearable Thermoelectric Generators: From Devices to Applications

1
Guangxi Key Laboratory of Information Material, School of Material Science and Engineering, Guilin University of Electronic Technology, Guilin 541004, China
2
Guangxi Key Laboratory for Relativistic Astrophysics, School of Physical Science and Technology, Guangxi University, Nanning 530004, China
3
Engineering Research Center of Electronic Information Materials and Devices, Ministry of Education, Guangxi Key Laboratory of Precision Navigation Technology and Application, Guilin University of Electronic Technology, Guilin 541004, China
4
Nagoya Industrial Science Research Institute, Nagoya 464-0819, Japan
5
Center of Nanotechnology, King Abdulaziz University, Jeddah 21589, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Energies 2022, 15(9), 3375; https://doi.org/10.3390/en15093375
Submission received: 1 March 2022 / Revised: 28 April 2022 / Accepted: 28 April 2022 / Published: 5 May 2022
(This article belongs to the Special Issue Advanced Thermoelectric Generation Technologies 2022)

Abstract

:
Wearable thermoelectric generators (WTEGs) can incessantly convert body heat into electricity to power electronics. However, the low efficiency of thermoelectric materials, tiny terminal temperature difference, rigidity, and neglecting optimization of lateral heat transfer preclude WTEGs from broad utilization. In this review, we aim to comprehensively summarize the state-of-the-art strategies for the realization of flexibility and high normalized power density in thermoelectric generators by establishing the links among materials, TE performance, and advanced design of WTEGs (structure, heatsinks, thermal regulation, thermal analysis, etc.) based on inorganic bulk TE materials. Each section starts with a concise summary of its fundamentals and carefully selected examples. In the end, we point out the controversies, challenges, and outlooks toward the future development of wearable thermoelectric devices and potential applications. Overall, this review will serve to help materials scientists, electronic engineers, particularly students and young researchers, in selecting suitable thermoelectric devices and potential applications.

1. Introduction

In the past few decades, with the advancement of nanomaterials and microelectronics technology, small-sized, low-power-consuming, and multifunctional electronic devices, especially wearable devices, have experienced rapid development. Examples include electronic skin and wristbands that can monitor human physiological data. In addition, the integration of these devices with the Internet of Things (IoT) has also attracted a lot of attention, especially for communication systems, environmental monitoring, and biomedical devices [1,2,3,4,5]. This integration allows devices to be equipped with wireless connectivity to directly transmit data from various sensors to a cloud diagnostic server for further analysis. For example, a portable electrocardiogram (ECG) system can monitor the heart data of patients with heart disease and remind medical staff to take action once arrhythmia occurs [6,7]. Smart wearable devices, such as intelligent clothes that monitor temperature and air parameters, can alert personnel who are exposed to toxic or substandard air conditions, thus enabling swift evacuation. However, these devices require laborious, non-stop measuring of parameters. Lately, soft and durable batteries, such as environmentally friendly water-zinc batteries [8] and lithium sulfide batteries [9,10], have also been developed as power sources for electronic devices. However, most batteries either need to be replaced regularly or need to be recharged. The limited lifespan and the need for frequent replacement further weaken the functionality of wearables. For example, in 24-h health monitoring applications, rechargeable batteries are undesirable because the operation may be interrupted during charging. This can lead to a severe lack of data availability and hamper medical treatment. Targeting this issue, self-powered devices that facilitate continuous monitoring of data over longer periods of time have become a popular direction of research in recent years. An adult can generate more than 100 W through activities such as breathing, heating, blood transport, and walking [11]. Therefore, merely converting 1% of the energy produced by the human body into electricity is sufficient to support the operation of most portable electronic devices [12,13,14]. Various energy harvesting technologies, such as triboelectric generators (TENGs) [15,16,17], piezoelectric generators [18], and thermoelectric generators (TEGs) [19,20,21], have been developed to convert human biomechanical energy (from human movement and body temperature) into electricity. However, triboelectric generators (TENGs) and piezoelectric generators are unstable in terms of energy harvesting from the human body. While body heat is the main energy source of human metabolism and is a constant heat source, thermoelectric generators (TEGs) can provide permanent DC power for electronic devices by collecting and converting body heat without noise and maintenance. Nevertheless, achieving an efficiency of 1% is more challenging than expected, as TEGs only cover a small part of the human body, so wearable thermoelectric generators are required to maximize efficiency. The maximization requires the precise design of TEGs at the material, device, and system levels. The most direct and effective way of enhancing efficiency is to improve the ZT of TE materials. At present, inorganic TE materials represented by bismuth telluride can reach a ZT of 2.2 at room temperature [22]. Organic materials represented by PEDOT:PSS [23,24,25,26] also appear as promising candidates because of their good flexibility. However, the highest ZT of the most advanced organic materials could not exceed 0.75 [22]. In addition, thin-film materials/devices exhibit high ZT and power density [27,28], but their large-area heat collection is unsatisfactory. The WTEGs based on inorganic bulk have the most potential as power sources for wearable electronic devices because of their high output and are discussed in this paper. Figure 1 shows the overview of WTEGs in this review. We introduce the trends in wearable technology and the basic concepts of WTEG in Section 1 and Section 2. The optimization strategies based on inorganic bulk WTEGs, including TE materials and thermal management, are summarized in Section 3. Section 4 discusses the markets of WTEG and various WTEGs-driven devices related to wearable health monitoring sensors and environmental monitoring. Section 5 describes existing challenges while predicting future trends of WTEG.

2. Thermoelectric Generators

2.1. Basic Concept of Thermoelectric Generators

Commonly, the thermoelectric effect first includes the Seebeck effect, which was discovered in 1821 by Estonian physicist Thomas Johann Seebeck, and the Peltier effect and Thomson effect, discovered by Peltier in 1834 and Thomson in 1851, respectively. Different metals (or semiconductors) have different free electron densities. When two different metals (or semiconductors) are in contact with each other (Figure 2A), the electrons on the contact surface will diffuse to eliminate the difference in electron density. A stable voltage is generated at the other end of the metals (or semiconductors), and the diffusion rate of electrons is proportional to the temperature of the contact area. The same metal (or semiconductor) at different temperatures also has different free electron densities, so as long as the temperature difference between the two ends of the metal is maintained, the electrons can continue to diffuse, and an open-circuit voltage is generated at both ends of the metal (or semiconductor) called Seebeck voltage (V), as given in Equation (1) [29,30].
V = S × Δ T            
where S and ΔT are the Seebeck coefficient and temperature difference, respectively. The practical TE conversion efficiency of a TEG is given by the following Equation (2):
η = P Q h            
where P is the power output to the load (output energy), and Qh is the heat energy absorbed at the hot side (input energy). Theoretically, the maximum energy conversion efficiency depends only on the temperature difference of the device and the intrinsic ZT of the TE material, as Equations (3) and (4) show [31]:
η m a x = T h T c T h     1 + Z T a v g 1 1 + Z T a v g + T c T h                                  
T a v g = T h + T c 2                
where Th is the temperature of the hot side of the TEG, Tc is that of its cold side, and (Th-Tc/Th) is the maximum efficiency of the Carnot cycle. Tavg is the average temperature. ZTavg is evaluated by the following equation, Equation (5) [31]:
Z T a v g = S 2 σ κ × T a v g            
where σ and κ represent the electrical conductivity and thermal conductivity, respectively. To obtain high TE conversion efficiency, high ZT is required. Meanwhile, to elevate ZTavg, the σ and S values should be large, and the κ value ought to be small. However, σ, S, and κ are strongly coupled through the carrier concentration (n) [32]. In other words, it is challenging to tune σ, S, and κ independently, hence realizing high ZTavg is not within easy reach.
A typical TEG module consists of pairs of π-type TE legs connected in series, as shown in Figure 2B. Based on the Seebeck effect, each pair of π-type TE legs consists of the p-type and n-type semiconductors fixed in series between two insulating substrates. When a heat source of temperature Th is applied onto one side of the TEG, and another side of the TEG is exposed to a low temperature (Tc) environment, a temperature difference (ΔT) is created between the TE pair. Therefore, most of the charge carriers generated at p-type (holes) TE legs and n-type (electrons) TE legs move from the higher to the lower temperature side, which generates an open-circuit voltage (Voc) across these TE pairs [33,34]. The electrical energy generated by a module depends on the number of thermocouples, the thermocouple geometry, TE properties of the thermocouple material, the thermal and electrical properties of the contact layers, electrodes, substrates, and the temperature difference. To simplify the evaluation, we assume that the absolute value of the Seebeck coefficient and the thermal and electrical conductivities of the p-type and n-type TE materials are equal.
The produced power (P) on the load resistance (RL) connected to a TEG module is given by:
P = V 2 R T E G + R L 2 × R L        
where RTEG is the internal resistance of TEG. When the load resistance (RL) is equal to RTEG, the maximum power (Pmax) is obtained:
P m a x = V o c 2 4 × R T E G                
The Voc produced by a TEG module and the RTEG can be evaluated by Equations (8) and (9) [35]:
V o c = N S T h T c 1 + 2 r t h L c / L                
R T E G = 2 N ρ r e + L A        
where N is the number of TE couples, and A and L are the cross-sectional area and length of TE legs, respectively, and L c is the thickness of the contact layer. Next, there is r e = ρ c / ρ and r t h = λ c / λ (where ρ c is the electrical contact resistivity, λ c is the thermal contact conductivity, and ρ and λ are the electrical resistivity and thermal conductivity of TE materials, respectively). r e and r t h are usually referred to as the electrical and thermal contact parameters, respectively. For commercially TEG modules, appropriate values are r e ~0.1 and r t h ~0.2.
The maximum power (Pmax) is obtained as [35]:
P m a x = S 2 2 ρ × A N T h T c 2 r e + L 1 + 2 r t h L c L 2                          
The maximum energy conversion efficiency ( η m a x ) is obtained as [35]:
η m a x = T H T C T H 1 + 2 r t h L c L 2 2 1 2 T H T C T H + 4 Z T H L + r e L + 2 r t h L c                  

2.2. Heat Transfer Model of Human Body as Heat Source

The TEG uses the temperature gradient as the driving force, thus enabling its application in aerospace, vehicle, and other fields in which temperature gradients are generated. The TEG module consisting of 36 pairs of bismuth telluride (Bi2Te3) was first commercialized by General Electric in 1959 [36]. Because the human body can act as a stable heat source, wearable thermoelectric generators have been developed to drive wearable electronic devices. When the TEG is worn on the human body, the thermal system becomes complicated due to the thermal contact resistance and different environments (Air velocity, ambient temperature). Figure 3A shows a schematic for the heat transfer model of a wearable thermoelectric generator. The heat flows from the skin through the TEG to the air. Note that the temperature difference between the two ends of the TEG (ΔTint = Th − Tc) is not the temperature difference between the human skin and the environment (ΔText = Ts − Ta) but is estimated by the following equation, Equation (12) [37]:
Δ T i n t = ( T h T c ) = ( T s T a ) × R t h , g R t h , h + R t h , g + R t h , c              
where Ta is the ambient temperature, Ts is the temperature of the human skin, and Rth,h, Rth,g, and Rth,c are the thermal resistances of skin contact, TEG, and air convection, respectively. The practical thermal resistance of a TEG is given by Equation (13) [37]:
R t h , g = 1 κ     L F F × A T E G                                        
where ATEG is the area of the TEG, FF is the fill factor (the area of TE legs/the area of the WTEGs), κ is the thermal conductivity of the TE material, and L is the height of the thermoelectric legs.
The core temperature of a human body is maintained between 36.5 and 37.5 °C, through thermoregulation [38]. The thermal conductivity of skin has been reported to between 0.26–0.3 W/mK, which is a complex function of specific physiological parameters, such as body weight, age, body fat, and gender [39]. Figure 3B shows the variation of skin temperature relative to air temperature in different parts of the human body [40].
When a TEG is placed on human skin, it is affected by the skin’s thermal resistance in addition to the thermal contact resistance at the TEG-skin interface [41,42]. Since the skin itself is a rough surface, air gaps between the TEG and the skin can significantly hinder the heat transfer performance of the TEG. Additionally, since the skin is soft and elastic with pressure, the contact resistance at the skin/TEG interface is highly dependent on the amount of pressure applied. Figure 3C shows the estimated heat transfer coefficients for different locations on the body as a function of contact pressure varying from 0 to 1 kPa. Generally, the pressure of tight clothing corresponds to 0.5 to 0.8 kPa [43].
Figure 3D shows the calculated heat transfer coefficients for a typical small heat sink consisting of 11 parallel fins, 0.5 mm thick, and 5 mm high. This calculation assumes that the heat sink base area is 1 × 1 cm and is a function of air velocities and thermal conductivity of the material. The heat transfer coefficient of the heat sink can be used to estimate the air thermal resistance of the interface between the TEG and the air interface [44].

2.3. Different Types of Wearable TEGs Based on Inorganic Bulk Materials

The purpose of WTEGs is to collect more heat for higher output power and be more comfortable for wearing, so the requirements for WTEGs include not only high output power but also flexibility and/or stretchability. The WTEGs are divided into different types and are discussed in detail below.

2.3.1. Rigid Wearable Thermoelectric Generators (r-WTEGs)

Generally, a rigid WTEG consists of a number of pairs of π-type thermoelectric legs connected in series by metal and packaged by two rigid substrates with good thermal conductivity. Rigid-shaped devices are firstly used in commercial products because of their superior heat collection and heat dissipation abilities, which can provide devices with large temperature differences and output power. Seiko Thermic watch (Figure 4A) was the first commercial body-heat-powered wristwatch, which was manufactured in 1998 [45]. The thermal to electrical energy conversion efficiency of this watch was about 0.1%; its open-circuit voltage was 300 mV, and the output power was approximately 25 µW over 1.5 °C temperature difference across the thermoelectric modules when the watch was worn. It can be equipped with heat sinks according to different application environments. To make the TEG work indoors and on a seated person (without forced air convection), a 38 × 34 mm heat sink with small pins and a TEG were assembled in a “five-to-seven” design. The “five to seven” design contains 32 thermopiles and a pin heat sink. The metal protection grid above the TEG protects the heat sink from being touched. The approximate average power generation at daytime of 25 0µW corresponds to 20 µW/cm2, which exceeds solar cells in many indoor situations, especially considering that TEG power is also available at nighttime [46]. In addition, Settaluri et al. added a copper sheet as the high-efficiency heat collector that can fit the human arm and a heat sink to reduce the air thermal resistance for the r-WTEGs, as shown in Figure 4B [47]. The TEG module contains 256 Bi2Te3 thermocouples (2 mm thick, 1 mm2 area) for power generation and two ceramic plates (0.7 mm thick, 10 cm2 area) for electrical insulation. A 0.1 mm thick flexible copper plate with an area of 14.6 × 6.4 cm was employed to make direct contact with the skin and effectively increase heat flow. Infrared images show that the heat sink can increase the temperature gradient for the TEG. Moreover, the heat sink surface can be designed with flat, groove, and checkerboard patterns to increase heat exchange with the air, showing steady-state open-circuit voltages of 85, 94, and 108 mV, and power outputs of 17.6, 21.6, and 28.5 µW/cm2, respectively. On the other hand, Nozariasbmarz et al. used nanostructured TE materials with a low fill factor (FF) module [48]. The fabricated r-WTEGs exhibited about 80% power output enhancement compared to commercial designs, resulting in a high power density of 35 μW/cm2 (Figure 4C). Conventional bulk TE materials have several weaknesses, including the complexity of processing, heavy-weight, and rigidity, all of which limit their applicability, especially when in contact with curved heat sources. In this regard, flexible thermoelectric generators are promising for wearable applications because their conformability enables effective contact with curved heat sources to maximize heat harvesting.

2.3.2. Flexible Thermoelectric Generators (FTEG)

Since most human body parts are curved or even dynamic, only a small part of the human body can be used for rigid wearable TEG. In order to obtain body heat from a larger area, flexible wearable devices have been proposed. The FTEG can be divided into two types, one is the non-stretchable FTEG that can be worn on curved surfaces such as human arms, and the other is the stretchable FTEG that can be worn on dynamic surfaces such as joints.

Non-Stretchable Flexible Thermoelectric Generators (ns-FTEGs)

Non-stretchable Flexible Thermoelectric Generators (ns-FTEGs) have achieved great development in recent years due to their wide applicability and high efficiency. A common and simple way to fabricate ns-FTEGs is to connect the TE legs through flexible electrodes, as shown in Figure 5A. The device was fabricated by Yuan et al., using FPCB as the connecting electrode and supporting substrate among TE elements, and considering the multi-objective optimization of power density, material consumption, and power matching with wearable electronic devices [49]. The results show that the ns-FTEGs exhibit excellent flexibility and high output, with power densities of 3.5 and 12.3µW/g, respectively, under no-wind conditions. In addition, a self-powered wearable bracelet was also fabricated. The strategy was to combine FTEG with a booster, which can be worn on the wrist with an output voltage of 2.8–3.3 V to drive multi-sensory systems, tailor-made intelligent power management, and LCD data displays. The system, which was fully man-powered, could continuously and simultaneously monitor the temperature, humidity, and activity of the human body. ns-FTEGs can be used as a continuous green energy supply for wearable monitoring systems. However, the TE legs of those modules are easily oxidized and fall off the WTEG during application. In order to enhance the reliability of ns-FTEGs during the application, an effective way to encapsulate the TE legs and flexible electrodes together into a flexible encapsulation material is needed. As shown in Figure 5B, Kim et al. used the nickel lift-off method to embed the screen-printed Cu thin film into Polydimethylsiloxane (PDMS) [50]. The design solved the problem of poor adhesion between the Cu electrode and the PDMS, and also prevented the TE legs from being oxidized. The TE module generated an open-circuit output voltage of 2.9 mV and an output power of 3 µW at an ambient temperature of 15 °C. Nonetheless, while the encapsulation makes the device more reliable, it inevitably causes a part of the heat to be shunted from the TE legs, thus rendering the device too thin to build a large temperature difference. In 2018, Park et al., proposed a method in which the TE legs were encapsulated by bakelite individually and connected in series by a flexible wire (Figure 5C) [51]. Thus, the TEG possessed good flexibility and reliability. The TE module could generate an output voltage of 8.8 mV and power of 138.67 μW or a power density of 5.60 μW/cm2 at room temperature. Compared with the two aforementioned structures, the assembly process of this type of TEG is more complicated, leading to difficulty in mass production.

Stretchable Flexible Thermoelectric Generators (s-FTEGs)

Compared with r-WTEGs and ns-FTEGs, stretchable flexible thermoelectric generators (s-FTEGs) can be applied to complex and even dynamic surfaces to minimize thermal energy loss at the interfaces.
There are various types of stretchable FTEG structures based on organic materials or thin films, but the method of fabrication based on inorganic bulk materials is mainly to use stretchable electrodes to connect thermoelectric legs and encapsulate them with elastic organics. A common material for making stretchable electrodes is liquid metal (LM). As shown in Figure 6A, the device is composed of top and bottom LM-embedded elastomer (LMEEs) layers that act as a heat spreader and form electrical interconnections between the TE legs by LM [52]. The results show that the s-FTEG can perform well after more than 1000 cycles at 30% stretch. At a temperature difference of 10 °C, it could generate an open-circuit voltage of about 60 mV. When the temperature difference is raised to 60 °C, it could generate an open-circuit voltage of about 320 mV and a power density of 0.087 mW/cm2. Snake-shaped copper foil is another form of flexible electrode, as shown in Figure 6B [53]. These hot-pressed p-type (Sb2Te3) and n-type (Bi2Te3) TE legs are interconnected by stretchable snake-shaped copper electrodes and embedded in Ecoflex elastomers. When the s-FTEG is stretched, the flexible snake-shaped copper electrode is elongated to maintain the thermoelectric connection between the TE elements. The results show that its 10 × 10 array device has an internal resistance of 22 Ω, and its output power is about 0.15 mW/cm2 at ΔT = 19 K.
s-FTEG can be applied to special parts of the human body, such as the joints of limbs, because of its superior flexibility. Among the three wearable TEGs, s-FTEG has the lowest output power but can drive some low-power sensors.
Table 1 shows the performance of the three types of WTEGs. Rigid WTEG has the highest output power, followed by ns-FTEG, and s-FTEG comes at last. Meanwhile, ns-FTEG has attracted the attention of scholars in recent years due to its high output power and its ability to fit most areas of the human body.

3. Optimization of Wearable Thermoelectric Generators

3.1. Materials for Wearable Thermoelectric Generators

Thermoelectric legs are the basic elements of thermoelectric power generation devices and thermoelectric heating and cooling devices. The improvement of their performance plays an indispensable role in the performance of the entire thermoelectric device. At present, the research of inorganic thermoelectric materials mainly focuses on metals and semiconductor alloys. There are three main categories of thermoelectric materials, namely high-temperature thermoelectric materials, such as silicon-germanium, intermediate-temperature thermoelectric materials, such as lead-telluride, and room-temperature thermoelectric materials, such as bismuth-telluride [77,78,79]. In general, wearable thermoelectric generators work around room temperature, and inorganic bulk thermoelectric materials with good thermoelectric performance at room temperature are discussed below.

3.1.1. Bismuth Telluride Material

Bi2Te3 is the best candidate for thermoelectric materials working at around room temperature. It was firstly reported in the 1950s. Bi2Te3 is a narrow gap-layered semiconductor material with a bandgap energy of about 0.15 eV; its crystalline structure belongs to the rhombohedral crystal system and space group R 3 ¯ m [55,80,81,82]. Bi2Te3 crystals have a layered structure, consisting of one kind of Bi site and two kinds of Te site, which form a five-layered crystal structure. Two adjacent Te layers are bound by van der Waals force, which causes easy cleavage [83].
In practical applications, the solid solutions of Bi2Te3 and Sb2Te3 are commonly used to prepare p-type materials (BixSb1-x) 2Te3. The regulation of the hole carrier of Bi2Te3 can be achieved by combining a Sb2Te3 solid solution with Bi2Te3, which also introduces a large number of point defects to significantly reduce the lattice thermal conductivity of the material. The n-type Bi2Te3-xSex can be obtained from solid solutions of Bi2Te3 and Bi2Se3. Both components are appropriate for body heat harvesting thanks to their high thermoelectric performance [38].
For half a century, many methods have been attempted to improve the thermoelectric properties of Bi2Te3, including the synthesis of low-dimensional nanomaterials, the optimization of material composition through doping, and the development of novel preparation processes [84]. Typically, the zone melting method is a traditional method for preparing Bi2Te3, through which samples with a ZT value of about 1 can be obtained. However, because Bi2Te3 is prone to cleavage and belongs to the direction of crystal growth, it is fragile and possesses poor mechanical properties. Hence it is difficult to process and not suitable for wearable devices. Therefore, to overcome the mechanical and thermoelectric challenges, numerous studies have been carried out on the preparation of dense polycrystalline bulk material by powder sintering process, including hot pressing and spark plasma sintering. Zheng et al., prepared a p-type Bi0.5Sb1.5Te3 compound by combining melt spinning with SPS. The bending strength and compressive strength of this sample were 5 times and 3 times higher than that of the zone melting one, respectively, and the ZT value also reached a maximum value of 1.22 at 340 K [85]. In addition to the sintering process, the introduction of nanostructures also greatly improves the thermoelectric properties of materials. Through effectively scattering phonons by decreasing grain size, introducing point defects, and dislocation arrays to the grain boundary, the lattice thermal conductivity can reduce rapidly, which improves the thermoelectric properties [22]. For bulk materials, ball milling and melt spinning are the most common ways of making grain size nanoscale. Kim et al. successfully prepared Bi0.5Sb1.5Te3 by using melt spinning together with Te composite, reducing the lattice thermal conductivity to 0.33 Wm−1K−1 and obtaining the highest ZT value of 1.86 at 320 K [86]. Poudel et al. prepared p-type BixSb2-xTe3 nanocomposites with ZT values of 1.4 at 100 °C by ball milling combined with hot-pressing [87]. Like the introduction of nanostructures, modulation doping is still an important method to improve the thermoelectric performance of Bi2Te3. For example, the thick graphene nanoplatelets (GNps) at the grain boundaries of Bi2Te3 polycrystal can boost the ZT value to 1.5, which increases the device output power by 35% [82]. Apart from the above two methods, there are several novel processes employed to achieve the control of defects and nanostructures. For instance, Nozariasbmarz et al. investigated the fact that combining microwave processing, glass inclusions, and annealing could enhance the ZT value of p-type (BixSb1-x)2Te3 to 1.2 at room temperature by introducing defects, which promoted a significant reduction in thermal conductivity [88].
In recent years, low-dimensional nanoscale-related research has further refined the thermoelectric properties of materials. In 2022, Chen et al., reported a flexible film of n-type Bi2Te3/single-walled carbon nanotube (SWCNT) hybrids with a cement-rebar nanoarchitecture constructed by a new solution-processed method. Benefiting from the novel structure, an extremely low thermal conductivity of ~0.33 Wm−1K−1 was realized at room temperature. Furthermore, the TE performance of this film is independent of mechanical bending, which makes it applicable to practical devices [28]. Different from traditional alloys, superlattice material is a kind of thermoelectric material that uses acoustic mismatch between superlattice components to reduce lattice thermal conductivity and improve the ZT value. The ZT value of the superlattice material (Bi2Te3/Sb2Te3) can reach 2.4 at 300 K [89]. Compared with the bulk material, the superlattice film greatly improves the material properties of Bi2Te3. If this type of thin-film material possesses excellent flexibility and lower manufacturing costs, it will be more suitable for wearable devices.

3.1.2. Other Room Temperature TE Materials

Bi2Te3-based thermoelectric materials have been widely used in the industry due to their excellent performance and decent price. However, tellurium compounds are considered mildly toxic and have poor mechanical properties. At present, the TE materials with better performance at room temperature are Ag2Se [90], MgAgSb [91,92,93], SnSe [94,95,96,97], Ag2Te [98,99,100], PbTe [101], etc. PbTe and Ag2Te are tellurides, which we will not discuss. It is worth noting that the MgAgSb-based material is also a non-toxic and better-performing material with good mechanical properties that is expected to replace Bi2Te3. There are three different crystal structures in MgAgSb: a half-heusler γ phase at high temperatures (630–700 K), a Cu2Sb-related β phase at about 560–630 K, and a tetragonal structure α phase with a = 9.1761 Å and c = 12.6960 Å at low temperatures (under 560 K) [F4]. MgAgSb can improve its TE performance by adjusting the percentage of Ag and Sb. Zhao et al. used ball milling and obtained an alloy with a ZT value of 0.7 at room temperature [102]. With the addition of multi-walled carbon nanotubes (MWCNTs), the electrical conductivity of the MgAgSb increases dramatically. This is because the addition of MWCNTs improves both phonon scattering and electrical transportation. At room temperature, the ZT of MgAg0.97Sb0.99 with 0.1 wt% MWCNTs is 0.83 [103]. Furthermore, the performance of MgAgSb can be optimized by Zn-doping and heat-treating. Zheng et al. obtained a method for preparing MgAgSb with excellent properties. The average ZT is about 1.3 in the range from 323 to 548 K in the sample Mg0.97Zn0.03Ag0.9Sb0.95 with heat-treating for 10 days, which is one of the highest values among all reported p-type MgAgSb-based thermoelectric materials. Tin selenide (SnSe) is one of the most efficient thermoelectric materials because it has low thermal conductivity as well as reasonable electrical conductivity. It has two stable phases: low-temperature α-SnSe (T<~800 K) and high-temperature β-SnSe (~800 K<T<1134 K). The mechanical properties of polycrystalline SnSe are comparable to other advanced thermoelectric materials [95]. For low-temperature α-SnSe, Zhao et al. obtained a high ZT of about 1.34, ranging from 0.7 to 2.0 at 300 to 773 K, realized in hole-doped tin selenide (SnSe) crystals [97]. As Su et al. reported, they found an average ZT of about 1.7 at 300 to 773 K in chlorine-doped and lead-alloyed tin selenide crystals by phonon-electron decoupling [96]. For high temperature use, Zhou et al., used hole-doped SnSe polycrystalline samples and with tin oxides, removed an exhibit of ZT of roughly 3.1 at 783 K. The key to performance enhancement is to remove the deleterious, thermally conductive oxides from the surface of SnSe grains [94]. Ag2Se-based TE materials are also well-studied. They are less toxic and can be prepared by costless synthetic methods. Silver and selenium powder can be mixed at room temperature, but the price of silver powder is a little high. It has been reported that the ZT value of Ag2Se obtained by doping carbon nanotubes with 0.5 wt% can reach 0.88 [104]. Day et al., calculated and predicted Ag2Se with a ZT around 1.0 using a single parabolic band model [105], and it was experimentally proven to show ZT ~1.0 near room temperature through optimizing the composition and nanostructure control [106]. With the development of Ag2Se-based thermoelectric materials and the requirement of removing Te from industrial materials, Ag2Se is expected to replace Bi2Te3-based materials as a room temperature thermoelectric material. Nanostructure TE materials are promising in WTEG and other fields. It has a larger Seebeck coefficient, and a large number of nanometer features in nanomaterials can effectively scatter phonons and reduce thermal conductivity, which could increase the ZT of the TE materials. The progress of nanostructured bismuth telluride has also stimulated the development of other thermoelectric materials in the nanometer direction to some extent. Jiang et al. have tried to obtain a high-performance Ag2Se film with a power density of 22.0 W/m2. It was realized by the synthesis of multi-sized Ag2Se nanostructures [107]. In addition, nanostructured monoclinic Cu2Se with a low carrier concentration has been synthesized by a wet mechanical alloying process combined with spark plasma sintering. It has obtained a ZT of 0.72 at 380 K by Chen et al. [108].

3.2. Device Design Optimization for Wearable Thermoelectric Generators

3.2.1. Geometric Parameters

The performance of a WTEG is determined by its temperature difference and the heat flow through the TE legs, while the temperature difference of TEG is mainly determined by its thermal resistance in a given ambient condition. The thermal resistance of TEG can be increased by reducing the fill factor, i.e., reducing the number of TE legs. In addition, increasing the aspect ratio of the legs and/or reducing the thermal conductivity of the legs can also play the same role [44,109,110], but this also affects the amount of heat flowing through the TE leg. Wherefore the optimal matching of several parameters must be considered in improving the output of WTEG. Leonov et al. reported a thermal resistance-matching mechanism to optimize the performance of WTEG and the thermal resistance of TEG should satisfy Equation (14) [111]:
R t h , g , o p t i m a l = R t h , F R t h , h + R t h , c 2 R t h , h + R t h , c + R t h , F            
where Rth,F is the parallel parasitic thermal resistance in a TEG upon the imaginary removal of the TE material, Rth,h is the thermal resistance between the body core and the surface of the skin, and Rth,c is the thermal resistance of the ambient air. However, the study by Leonov et al., cannot directly predict the output of the TEG from the specific geometric parameters of the TE legs.
In order to express the influence of the geometric parameters of WTEG more intuitively and multi-dimensionally, Lee et al. analyzed the influence of the length, width, and FF of the thermoelectric leg on the output performance of a device under different convection conditions by a numerical analysis method and the results are summarized in Figure 7A [37]. The results clearly show that the length and FF of the TE legs affect the power output of the WTEG, and the width and FF affect the voltage output of the WTEG. In 2020, Zhu et al., proposed a simplified system considering external thermal resistance, output power, and voltage matching [112]. It takes the engineering length (L*), that is, the ratio of L and FF, as the key parameter for determining the power density of the device and the aspect ratio (r = L/A) as the key parameter for the voltage density. The optimizations above are all based on numerical analysis, which cannot be applied for lateral heat transfer. This type of method is reliable for rigid TEG or in the case of good heat transfer. However, the results are not reliable for most flexible TEGs due to their poor heat transfer. Zhu and Peng et al. introduced the influence of the width of the thermoelectric legs on the output power of the flexible TEG using finite element analysis [64]. They proposed that in the case of poor heat transfer, tall and thin thermoelectric legs are more suitable than thick legs. By contrast, in the case of poor heat transfer, the influence of the width of the thermoelectric leg on the thermoelectric power output can be ignored (Figure 7B).

3.2.2. Heat Sink

Based on the thermoelectric principle, the output power of WTEG is proportional to the square of the temperature difference (ΔT), so the optimization of temperature difference must be considered [113,114]. The temperature difference depends on the thermal design of the WTEG system, and adding a heat sink to the WTEG is an effective method of increasing the temperature gradient. It goes without saying that, for the WTEG system, not only temperature difference but also flexibility is required for the heat sink. Various designs of heat sinks for WTEG are discussed in detail below.
Figure 8A shows the temperature distribution in a typical wearable TEG system at room temperature and in a cold environment. The results show that more than 40% of the heat is lost at the interface between the TEG and the air, while 20–30% of the heat is lost at the interface between the TEG and the human body, and only less than 30% of the heat can be used for power generation in the TEG [44]. The contact thermal resistance between the skin and TEG can only be optimized slightly, because it is affected by pressure and skin roughness. Although it can be reduced by increasing the pressure, doing so will make it less comfortable to wear. The heat sink is an effective means to reduce the thermal resistance of the TEG and air interface and has accumulated attention in recent years. For example, Park et al. added a plate fin heat sink to each pair of π-type thermocouples in the chain f-TEG to obtain a large temperature difference while maintaining its flexibility (Figure 8B) [67]. When worn on the wrist and during the “walking” condition, it successfully generated an open-circuit voltage of 12.1 mV with a power output density of 6.97 µW/cm2. However, high fin heat sinks can compromise wearing comfort. Hyland et al., added and optimized a Cu heat spreader with an area of 16 cm2 and a thickness of 250 µm on the cold side of the rigid WTEG (Figure 8C) [54]. The results show that the device equipped with the heat spreader successfully generated a power output density of 6 µW/cm2 when the ambient temperature was 18.3 °C. A spreader effectively improves the lateral heat transfer but has limited ability to transfer heat to the external environment. In order to enhance the external heat transfer, Kim et al., fabricated a water-storage heat sink based on superabsorbent polymer (SAP) material for FTEG [71]. The thermoelectric power density was increased to 13 µW/cm2 and could be maintained for 22 h through the strategy of water evaporation, which took away the heat from the cold side of the TEG (Figure 8D). Unlike the evaporation of water for the heat sink, which requires continuous water replenishment, Lee et al. used the heat sink to enhance the output of TEG based on solid-liquid phase change material (octadecane), as shown in Figure 8E [74]. While the flexible heat sinks can be reused after solidification at room temperature, the results show that the power generated by the flexible TEG remained at 20 µW/cm2 only for 33 min. Phase change materials based on octadecane have not been able to satisfy the long-term energy supply of electronic systems. In the future, high phase change latent heat materials will also be the way to improve TEG output. Compared with other thick and heavy heat sinks, radiative cooling can be realized by thin films to enhance the flexibility of TEGs, which has attracted the extensive attention of scholars in recent years. Khan et al. proposed a TEG integrated with a flexible, micrometer-thin poly (vinylidene fluoride-co-hexafluoropropylene) radiative cooling heat sink (Figure 8F) [65]. The radiative cooling heat sink can radiate heat to the cosmic space through the atmospheric window, and the fabricated TEG achieves a power density of 12.48 µW/cm2 (outdoors). This TEG module is not only small but also provides high output power compared to TEGs with plate fin heat sinks.

3.2.3. Encapsulation Material

The package is to stabilize and protect the π-type thermoelectric structure. For different TEG applications, encapsulation has various requirements for thermal conductivity, oxidation resistance, and heat insulation. Here, we divide the packages into two categories: substrates for encapsulating electrodes and filler material for encapsulating thermoelectric legs. Different types of encapsulations are discussed in detail in the sections below.

Substrate

The function of the substrate for rigid WTEG is to fix and enhance the lateral heat transfer, and its materials are generally aluminum oxide and aluminum nitride with high hardness and high thermal conductivity and insulation. However, the substrate material of flexible or stretched WTEG is mainly elastomer (PDMS), which can stabilize and protect the electrode from oxidation. Because of its low thermal conductivity, PDMS resists the lateral heat transfer between the hot and cold ends of WTEG, so the modification to make PDMS with higher thermal conductivity has become a hot spot in recent years. For example, Hong et al. doped AlN in Ecoflex to enhance the thermal conductivity [61]. This successfully increased the thermal conductivity of Ecoflex from 0.16 to 0.76 W/mK. The thermoelectric device with an area of 5 × 5 cm is composed of 72 pairs of thermocouples, which can not only be used to manage human body temperature but also generate a power density of 4.5 μW/cm2 at room temperature. Coincidentally, Sargolzaeiaval et al. also added eutectic gallium indium (EGaIn) to PDMS to improve the thermal conductivity of the substrate for WTEG based on liquid metal electrodes (Figure 9). Worn on the wrist, it achieved a power density of 5.2 μW/cm2 under natural convection and over 30 μW/cm2 at an air velocity of 1.2 m/s [62].

Filler Material

In contrast to the substrate, the fill material requires thermal insulation to reduce thermal shunting through the TE legs. Figure 10A shows the results of PDMS and melamine foam (MF) simulated by Zhu et al. [64]. The low thermal conductivity of MF can not only increase the output power but also reduce the fill factor and enhance the flexibility of the device. Figure 10B shows a schematic diagram of its optimized strategy using low thermal conductivity MF as the filler material. A power density of 7 μW/cm2 was achieved in a natural environment. Using a strategy similar to that adopted by Zhu for low thermal conductivity MF, Jung et al. prepared porous PDMS (Figure 10C) [70]. The power density reached 9.7 μW/cm2 when equipped with a SAP based heat sink. Obviously, the method of low-cost and high-efficiency filler material fabrication will continue to be a hot issue in the future.

3.2.4. Fabricating Process of Thermoelectric Modules

At present, the manufacturing technology based on rigid thermoelectric devices for refrigeration has matured. A low-cost and high-efficiency fabrication process for the commercialization of flexible wearable thermoelectric generators is the key. In 2016, Kim et al. reported FTEGs produced by screen printing technique (SPT) and laser multi-scanning (LMS) lift-off process. Screen-printed TEGs were fabricated on SiO2/a-Si/quartz substrates by the SPT process, and then rigid quartz substrates were completely separated from pristine TEGs (a-Si) by the LMS process [115]. The FTEG fabricated by this technique exhibits high flexibility (no significant change in device performance after 8000 repeated bending cycles) and excellent output performance (4.78 mW/cm2 and 20.8 mW/g at ΔT = 25 °C). However, the fabrication process is complicated. In 2018, Shi et al., proposed a low-cost and efficient FTEG fabrication method based on the commercialized flexible printed circuit (FPC), as shown in Figure 11A [58]. It was first arranged and fixed on the TE legs through the sieve tray and then welded onto the customized FPC. This semi-automatic assembly method is simple and low-cost, but the assembly and welding of electrodes are still complicated and inefficient. Zhu et al. simplified this step through etching techniques [64]. They first printed the circuit on transfer paper and then transferred it to copper foil with PI on the back, and finally etched the electrodes. As to the assembling technology of s-FTEG, Li et al. proposed a highly automated fabrication process by doping silver-nickel (Ag-Ni) particles into PDMS to improve the thermal conductivity of the substrate, as shown in Figure 11B [76]. Ag-Ni particles were first mixed with PDMS, and then silver nanowires (Ag-NWs) were printed as soft electrodes onto the mixture through a mask. Afterward, the mixture of PDMS and Ag-Ni particles was magnetically aligned along the direction of the legs to facilitate heat transfer from the skin to the legs. Next, 220 couples of Bi2Te3-based TE legs were automatically arranged onto the Ag-NW electrodes via vacuum suction nozzles, and then the gaps between them were filled with PDMS. The electrodes of the upper substrate are prepared in the same way and mounted on top of the device. The highly automated fabrication process greatly improves fabrication efficiency. The results show that when the flexible TEG with an area of 39 × 43 mm2 is worn on the wrist, the highest power density generated under a 10 °C temperature gradient is 6.96 μW/cm2.

4. Thermoelectric and Wearable Applications

In recent years, major breakthroughs have been witnessed in low-temperature thermoelectric materials in terms of performance and flexibility. More efficient, lightweight, and safe thermoelectric devices have been reported one after another. Wearable devices have great potential applications in healthcare, environmental detection, and other fields. This paper summarizes the application and market of thermoelectric wearable devices in the following two aspects.

4.1. Market Research for Wearable Thermoelectric

The global thermoelectric generator market was valued at 472.5 million USD in 2020 and is projected to reach 1443.3 million USD by 2030, growing at a compound annual growth rate (CAGR) of 11.8% from 2021 to 2030. The thermoelectric generators market is segmented into low power (<10 W), medium power (10 W–1 kW), and high power (>1 kW) based on wattage. The power produced by thermoelectric generators is highly dependent upon the temperature applied to their plates. The low power (<10 W) segment of the thermoelectric generators market is projected to grow at the highest CAGR during the forecast period, with the increasing use of wearables and handheld consumer electronic devices being the major. In the context of the global COVID-19 pandemic, wearables devices are rapidly moving into the preventative care, diagnostics, and urgent care segment with an anticipated market of 9.3 B USD by 2024, driving improvement of the WTEG market [116].

4.1.1. Biomedical Devices

The demand for autonomous monitoring and diagnosis is large in the healthcare industry since the increasing healthcare costs and the rapid aging of the global population in developed countries. By 2030, for example, nearly 25% percent of US residents are expected to be aged 65 or older. Among them, heart rate, pulse, and other health indicators need to be as continuous as possible, and TE materials can provide energy for wearable devices to avoid the potential danger of battery replacement to the greatest extent. In some relatively simple applications proposed earlier, it is confirmed that this wearable detection device based on WTEG can be realized and applied in the medical field. Tom et al. used a small watch-type thermoelectric generator to build a wireless pulse oxymeter powered by body heat [117]. The system, shown in Figure 12A, is an integration of WTEG and wireless sensing, has an output of about 30 µW per square centimeter when the watch is close to the skin, and the watch is large enough to provide a steady stream of power to the sensor. However, this kind of watch-type thermoelectric device has poor flexibility and should develop in the direction of higher comfort and flexibility under the premise of meeting the power supply demand. In Figure 12B, Kim et al. printed the conventional thermoelectric power generation and regeneration circuit in the flexible PCB, giving the device higher flexibility and adding the flexible heat sink to enable sufficient temperature difference, successfully establishing the self-powered wearable ECG system at the arm site [71]. This flexible design ensures the energy of the ECG system while greatly improving the comfort experience. With the continuous improvement of the flexibility and efficiency of thermoelectric wearables, more and more previously undesirable applications have been reported, in addition to routine tests, such as blood pressure, pulse and heartbeat. A self-powered hearing aid based on thermoelectric power generation devices was proposed and designed by Ekuakille et al. [118]. As shown in Figure 12C, the hearing aid is mainly composed of an amplifier circuit, which has extremely low power consumption. In daily life, the power supply can be achieved entirely by the body heat-driven flexible thermoelectric device, a design that truly makes the external hearing aid an “artificial ear”.

4.1.2. Human Activity Monitoring Device

Flexible thermoelectric systems that monitor human motion states can provide fully continuous detection in clinical diagnosis, behavioral analysis, and information communication. Shi et al. developed a wearable thermoelectric system that can monitor the motion state of a person in real-time [58]. The system is powered by the body heat of the wrist and provides continuous detection based on inertial and strain sensors. It can determine the wearer’s current movement state only through the feedback waveform. Although it can assess significant state changes, such as running, jumping, and walking, it is subtle, such as speed discrimination (Figure 13A). More precise sensors are helpful for detecting movement status. Yang et al., proposed a hand state detection device self-powered by WTEG [53]. The system uses the stress tensile sensor to detect more sensitive factors, and subtle changes in hand movements can be seen in Figure 13B. This wearable application based on a highly sensitive stretch sensor is also widely used in laryngeal language discrimination and ocular visual sensing, but motor state recognition devices in which more flexible thermoelectric systems collect energy from the dynamic surface of the human skin have not been reported.

4.1.3. Environmental Monitoring Device

Signals from the surrounding environment (such as ultraviolet light, air composition, temperature) are also important compared with health care. Wearable devices can help the wearer visualize unknown signals from the environment and reduce/avoid health hazards from the environment, and even predict potential hazards. The focus is on people in hazardous operations. Lee et al. designed a wearable temperature measurement glove driven by an FTEG. It can accurately measure the temperature of touching objects with a temperature-sensitive sensor and warn users with LEDs to avoid burns caused by touching unknown objects [76]. In addition to the above temperature detection thermoelectric gloves, more practical, lighter, and more fully functional hand wearables have been reported (Figure 14A). In 2020, Yuan et al., introduced a TE system that can help the wearer perceive more external information [119]. The system has the functions of monitoring temperature, humidity, wind speed, identifying materials, and acceleration, and can visualize the measurement data on an LCD display (Figure 14B). The device is also known as “electronic skin” because of its lightweight and is highly flexible, and its energy is entirely provided by a hand-shaped WTEG. The wearable application of environment detection focuses on protecting the health and safety of workers in dangerous places. Hand wearable devices are limited by the contact area of the thermoelectric devices and the skin and the volume of the wearable devices themselves, so the wearable applications in other parts have also had great development. For example, Fazio et al. have developed a model to detect the wearer’s surrounding environment and its own biophysical parameters [120]. The system integrates heart rate ECG, accelerometer, temperature, and electrochemical gas sensors, using a flexible thermoelectric and solar power-combined power supply (Figure 14C). In addition, smart clothing is also equipped with a wireless communication port, which can realize information visualization and risk early warning by uploading the collected information to the mobile terminal or the cloud. The addition of the mobile IoT make wearables respond more quickly to collect information and further guarantees the health and safety of the wearer. In addition to the apps described above, MATRIX Co., Ltd. reported a man-powered smartwatch with multiple sensors built for movement tracking and health state measurement in 2020. It was able to interact with a phone and the cloud to support IOS and Android systems and is currently the only thermoelectric wearable product commercially available.
Energy collection of wearable thermoelectrics is of great practical significance to the power supply of portable electronics for renewable energy when requiring light and sustainable work in fields such as medical, health, and environmental monitoring. In recent years, with the continuous improvement of the basic research on TE materials performance, the continuous realization of lower energy consumption and the flexibility of PCB, wearable thermoelectric generators have shown great potential in their application. In the future, as these technologies become more mature, combining the diversification of the IoT and the cloud, lighter, comfortable, and highly sensitive wearable devices will appear in our vision and life.

5. Conclusions and Challenges

Wearable thermoelectric generators (WTEGs) can incessantly convert body heat into electricity to power electronics. This review provides an overview of the technology for wearable thermoelectric generators based on inorganic bulk thermoelectric materials, from materials to device design and finally to applications. First, we explained the basic concepts and types of WTEGs. Because of WTEG’s small-size, lightweight, no-noise, and reliability, it was used in the human body as a self-powered system, which is irreplaceable by other generators. Then, we introduced the popular inorganic bulk thermoelectric materials for WTEGs this year. Currently, most WTEGs use inorganic bulk thermocouples based on Bi2Te3 with high ZT to power wearable electronic devices. In addition, we also show the strategy of optimizing the output performance of TEG in recent years, the core content of which is to improve the temperature difference by optimizing the topology structure (geometry, connection, and heat dissipation) and fabricating process of the wearable thermoelectric generator while taking into account high reliability and flexibility. Besides r-WTEG and ns-FTEG, stretchable FTEGs have attracted a lot of attention in recent years because of their excellent flexibility and use in many special occasions. In applications, WTEG was used to power watches in the early days, but it has now been developed for various fields, such as biology and environment.
With the increasing market of wearable electronic devices, the development prospect and application market of efficient wearable thermoelectric generators are broad. As a relatively new field, wearable thermoelectric generators have made great progress in the past ten years, but nowadays, various wearable electronic devices with multi-function and high energy consumption require necessary properties of WTEG, such as high efficiency, durability, biological compatibility, and flexibility. In thermoelectric materials, high-efficiency thermoelectric generators require high ZT materials. Current thermoelectric materials and thermoelectric devices have not achieved satisfactory performance, so a lot of work is still needed in this research area. In general, Bi2Te3 exhibits superior TE performance at room temperature and can be mass-produced. Unfortunately, the development of wearable thermoelectric generators for human energy harvesting has been hindered by the inherent rigidity, toxicity, and danger of Bi2Te3. Although Ag2Se, MgAgSb, SnSe, and other non-toxic materials exhibit good performance at room temperature and show the potential to replace Bi2Te3, there are still many challenges, such as TE properties, mechanical properties, cost, mass production, and reliability. In addition, heat sinks are a good strategy to improve device efficiency, but most of the flexible heat sinks designed for wearable thermoelectric generators are composed of thick hydrogels and copper plates, which do not take into account flexibility. Therefore, with the development of modern heat dissipation materials, such as graphene and radiant cooling, flexible, lightweight, and efficient wearable thermoelectric generators are expected. In order to solve the insufficient energy supply of self-powered wearable devices, the current trend is to combine a variety of wearable generators, such as solar power generators combined with thermoelectric and triboelectric nanogenerators. How to combine multiple wearable generators more effectively is also an urgent problem to be solved. In the future, the rapid development of new materials and technologies will bring greater application spaces and opportunities for wearable thermoelectric generators.

Author Contributions

Investigation, methodology, and writing—original draft preparation, S.Z.; investigation, Z.F., R.S., B.F. and Z.J.; conceptualization and supervision, Y.P. and J.G.; conceptualization, methodology, writing—review and editing, and supervision, L.M. and K.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (grant no. U21A2054, 51961011, 52061009), the National Key Research and Development Program of China (no. 2017YFE9128000), and the Natural Science Foundation of Guangxi, China (grant no., 2021AC19206 and AD19110020).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to confidentiality.

Acknowledgments

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, Y.; Shi, G.; Qin, J.; Lowe, S.E.; Zhang, S.; Zhao, H.; Zhong, Y.L. Recent Progress of Direct Ink Writing of Electronic Components for Advanced Wearable Devices. ACS Appl. Electron. Mater. 2019, 1, 1718–1734. [Google Scholar] [CrossRef]
  2. Hu, X.; Li, F.; Song, Y. Wearable Power Source: A Newfangled Feasibility for Perovskite Photovoltaics. ACS Energy Lett. 2019, 4, 1065–1072. [Google Scholar] [CrossRef]
  3. Jia, Y.; Jiang, Q.; Sun, H.; Liu, P.; Hu, D.; Pei, Y.; Liu, W.; Crispin, X.; Fabiano, S.; Ma, Y.; et al. Wearable Thermoelectric Materials and Devices for Self-Powered Electronic Systems. Adv. Mater. 2021, 33, e2102990. [Google Scholar] [CrossRef] [PubMed]
  4. Yan, J.; Liao, X.; Yan, D.; Chen, Y. Review of Micro Thermoelectric Generator. J. Microelectromech. Syst. 2018, 27, 1–18. [Google Scholar] [CrossRef]
  5. Haras, M.; Skotnicki, T. Thermoelectricity for IoT—A review. Nano Energy 2018, 54, 461–476. [Google Scholar] [CrossRef]
  6. Gibson, J.S.; Liu, X.; Georgakopoulos, S.V.; Wie, J.J.; Ware, T.H.; White, T.J. Reconfigurable Antennas Based on Self-Morphing Liquid Crystalline Elastomers. IEEE Access 2016, 4, 2340–2348. [Google Scholar] [CrossRef]
  7. Yang, Z.; Zhou, Q.; Lei, L.; Zheng, K.; Xiang, W. An IoT-cloud Based Wearable ECG Monitoring System for Smart Healthcare. J. Med. Syst. 2016, 40, 286. [Google Scholar] [CrossRef]
  8. Guo, Z.; Ma, Y.; Dong, X.; Huang, J.; Wang, Y.; Xia, Y. An Environmentally Friendly and Flexible Aqueous Zinc Battery Using an Organic Cathode. Angew. Chem. Int. Ed. Engl. 2018, 57, 11737–11741. [Google Scholar] [CrossRef]
  9. Zhang, S.S. Sulfurized Carbon: A Class of Cathode Materials for High Performance Lithium/Sulfur Batteries. Front. Energy Res. 2013, 1, 10. [Google Scholar] [CrossRef] [Green Version]
  10. He, J.; Chen, Y.; Lv, W.; Wen, K.; Li, P.; Qi, F.; Wang, Z.; Zhang, W.; Li, Y.; Qin, W.; et al. Highly-flexible 3D Li2S/graphene cathode for high-performance lithium sulfur batteries. J. Power Sources 2016, 327, 474–480. [Google Scholar] [CrossRef]
  11. Starner, T. Human-powered wearable computing. IEEE Access 1996, 35, 618–629. [Google Scholar] [CrossRef]
  12. Dieffenderfer, J.; Goodell, H.; Mills, S.; McKnight, M.; Yao, S.; Lin, F.; Beppler, E.; Bent, B.; Lee, B.; Misra, V.; et al. Low-Power Wearable Systems for Continuous Monitoring of Environment and Health for Chronic Respiratory Disease. IEEE J. Biomed. Health Inform. 2016, 20, 1251–1264. [Google Scholar] [CrossRef] [PubMed]
  13. García Núñez, C.; Manjakkal, L.; Dahiya, R. Energy autonomous electronic skin. npj Flex. Electron. 2019, 3, 1. [Google Scholar] [CrossRef]
  14. Huang, H.; Li, X.; Liu, S.; Hu, S.; Sun, Y. TriboMotion: A Self-Powered Triboelectric Motion Sensor in Wearable Internet of Things for Human Activity Recognition and Energy Harvesting. IEEE Internet Things J. 2018, 5, 4441–4453. [Google Scholar] [CrossRef]
  15. Ha, M.; Park, J.; Lee, Y.; Ko, H. Triboelectric Generators and Sensors for Self-Powered Wearable Electronics. ACS Nano 2015, 9, 3421–3427. [Google Scholar] [CrossRef]
  16. Wang, Z.L. On Maxwell’s displacement current for energy and sensors: The origin of nanogenerators. Mater. Today 2017, 20, 74–82. [Google Scholar] [CrossRef]
  17. Dong, K.; Peng, X.; An, J.; Wang, A.C.; Luo, J.; Sun, B.; Wang, J.; Wang, Z.L. Shape adaptable and highly resilient 3D braided triboelectric nanogenerators as e-textiles for power and sensing. Nat. Commun. 2020, 11, 2868. [Google Scholar] [CrossRef]
  18. Chen, Z.; Wang, Z.; Li, X.; Lin, Y.; Luo, N.; Long, M.; Zhao, N.; Xu, J.B. Flexible Piezoelectric-Induced Pressure Sensors for Static Measurements Based on Nanowires/Graphene Heterostructures. ACS Nano 2017, 11, 4507–4513. [Google Scholar] [CrossRef]
  19. Freer, R.; Powell, A.V. Realising the potential of thermoelectric technology: A Roadmap. J. Mater. Chem. C 2020, 8, 441–463. [Google Scholar] [CrossRef]
  20. Soleimani, Z.; Zoras, S.; Ceranic, B.; Cui, Y.; Shahzad, S. A comprehensive review on the output voltage/power of wearable thermoelectric generators concerning their geometry and thermoelectric materials. Nano Energy 2021, 89, 106325. [Google Scholar] [CrossRef]
  21. Zeng, W.; Tao, X.-M.; Lin, S.; Lee, C.; Shi, D.; Lam, K.-H.; Huang, B.; Wang, Q.; Zhao, Y. Defect-engineered reduced graphene oxide sheets with high electric conductivity and controlled thermal conductivity for soft and flexible wearable thermoelectric generators. Nano Energy 2018, 54, 163–174. [Google Scholar] [CrossRef]
  22. Hasan, M.N.; Nafea, M.; Nayan, N.; Mohamed Ali, M.S. Thermoelectric Generator: Materials and Applications in Wearable Health Monitoring Sensors and Internet of Things Devices. Adv. Mater. Technol. 2021, 2101203. [Google Scholar] [CrossRef]
  23. Chen, G.; Xu, W.; Zhu, D. Recent advances in organic polymer thermoelectric composites. J. Mater. Chem. C 2017, 5, 4350–4360. [Google Scholar] [CrossRef]
  24. Elmoughni, H.M.; Menon, A.K.; Wolfe, R.M.W.; Yee, S.K. A Textile-Integrated Polymer Thermoelectric Generator for Body Heat Harvesting. Adv. Mater. Technol. 2019, 4, 1800708. [Google Scholar] [CrossRef]
  25. Wang, H.; Yu, C. Organic Thermoelectrics: Materials Preparation, Performance Optimization, and Device Integration. Joule 2019, 3, 53–80. [Google Scholar] [CrossRef] [Green Version]
  26. Cui, Y.J.; Wang, B.L.; Wang, K.F. Energy conversion performance optimization and strength evaluation of a wearable thermoelectric generator made of a thermoelectric layer on a flexible substrate. Energy 2021, 229, 120694. [Google Scholar] [CrossRef]
  27. Bharti, M.; Singh, A.; Singh, B.P.; Dhakate, S.R.; Saini, G.; Bhattacharya, S.; Debnath, A.K.; Muthe, K.P.; Aswal, D.K. Free-standing flexible multiwalled carbon nanotubes paper for wearable thermoelectric power generator. J. Power Sources 2020, 449, 227493. [Google Scholar] [CrossRef]
  28. Chen, Z.; Lv, H.; Zhang, Q.; Wang, H.; Chen, G. Construction of a cement–rebar nanoarchitecture for a solution-processed and flexible film of a Bi2Te3/CNT hybrid toward low thermal conductivity and high thermoelectric performance. Carbon Energy 2021, 4, 115–128. [Google Scholar] [CrossRef]
  29. DiSalvo, F.J. Thermoelectric cooling and power generation. Science 1999, 285, 703–706. [Google Scholar] [CrossRef]
  30. Majumdar, A. Materials science. Thermoelectricity in semiconductor nanostructures. Science 2004, 303, 777–778. [Google Scholar] [CrossRef]
  31. Toberer, G.J.S.E.S. Complex thermoelectric materials. Nat. Mater. 2008, 7, 105–114. [Google Scholar]
  32. Rosi, F.D. Thermoelectricity and thermoelectric power generation. Solid-State Electron. 1968, 11, 833–868. [Google Scholar] [CrossRef]
  33. Selvan, K.V.; Hasan, M.N.; Mohamed Ali, M.S. State-of-the-Art Reviews and Analyses of Emerging Research Findings and Achievements of Thermoelectric Materials over the Past Years. J. Electron. Mater. 2018, 48, 745–777. [Google Scholar] [CrossRef]
  34. Selvan, K.V.; Hasan, M.N.; Mohamed Ali, M.S. Methodological reviews and analyses on the emerging research trends and progresses of thermoelectric generators. Int. J. Energy Res. 2018, 43, 113–140. [Google Scholar] [CrossRef] [Green Version]
  35. Rowe, D.M. Thermoelectrics Handbook; CRC Press: Boca Raton, FL, USA, 2006. [Google Scholar]
  36. Zoui, M.A.; Bentouba, S.; Stocholm, J.G.; Bourouis, M. A Review on Thermoelectric Generators: Progress and Applications. Energies 2020, 13, 3606. [Google Scholar] [CrossRef]
  37. Lee, Y.G.; Kim, J.; Kang, M.-S.; Baek, S.-H.; Kim, S.K.; Lee, S.-M.; Lee, J.; Hyun, D.-B.; Ju, B.-K.; Moon, S.E.; et al. Design and Experimental Investigation of Thermoelectric Generators for Wearable Applications. Adv. Mater. Technol. 2017, 2, 1600292. [Google Scholar] [CrossRef]
  38. Nozariasbmarz, A.; Collins, H.; Dsouza, K.; Polash, M.H.; Hosseini, M.; Hyland, M.; Liu, J.; Malhotra, A.; Ortiz, F.M.; Mohaddes, F.; et al. Review of wearable thermoelectric energy harvesting: From body temperature to electronic systems. Appl. Energy 2020, 258, 114069. [Google Scholar] [CrossRef]
  39. Leonov, V.; Vullers, R.J.M. Thermoelectric Generators on Living Beings. In Proceedings of the 5th European Conference on Thermoelectrics (ECT 2007), Odessa, Ukraine, 10–12 September 2007. [Google Scholar]
  40. Webb, P. Temperatures of skin, subcutaneous tissue, muscle and core in resting men in cold, comfortable and hot conditions. Appl. Physiol. 1992, 64, 471–476. [Google Scholar] [CrossRef]
  41. Saggin, B.; Tarabini, M.; Lanfranchi, G. A Device for the Skin–Contact Thermal Resistance Measurement. IEEE Trans. Instrum. Meas. 2012, 61, 489–495. [Google Scholar] [CrossRef]
  42. Ho, H.N.; Jones, L.A. Thermal model for hand-object interactions. In Proceedings of the 2006 14th Symposium on Haptic Interfaces for Virtual Environment and Teleoperator Systems, Alexandria, VA, USA, 25–26 March 2006; pp. 461–467. [Google Scholar]
  43. Mori, Y.; Kioka, E.; Tokura, H. Effects of pressure on the skin exerted by clothing on responses of urinary catecholamines and cortisol, heart rate and nocturnal urinary melatonin in humans. Int. J. Biometeorol. 2002, 47, 1–5. [Google Scholar] [CrossRef]
  44. Suarez, F.; Nozariasbmarz, A.; Vashaee, D.; Öztürk, M.C. Designing thermoelectric generators for self-powered wearable electronics. Energy Environ. Sci. 2016, 9, 2099–2113. [Google Scholar] [CrossRef]
  45. Qi, Y.; McAlpine, M.C. Nanotechnology-enabled flexible and biocompatible energy harvesting. Energy Environ. Sci. 2010, 3, 1275–1285. [Google Scholar] [CrossRef]
  46. Leonov, V.; Torfs, T.; Fiorini, P.; Van Hoof, C. Thermoelectric Converters of Human Warmth for Self-Powered Wireless Sensor Nodes. IEEE Sens. J. 2007, 7, 650–657. [Google Scholar] [CrossRef]
  47. Settaluri, K.T.; Lo, H.; Ram, R.J. Thin Thermoelectric Generator System for Body Energy Harvesting. J. Electron. Mater. 2011, 41, 984–988. [Google Scholar] [CrossRef]
  48. Nozariasbmarz, A.; Kishore, R.A.; Poudel, B.; Saparamadu, U.; Li, W.; Cruz, R.; Priya, S. High Power Density Body Heat Energy Harvesting. ACS Appl. Mater. Interfaces 2019, 11, 40107–40113. [Google Scholar] [CrossRef]
  49. Yuan, J.; Zhu, R. A fully self-powered wearable monitoring system with systematically optimized flexible thermoelectric generator. Appl. Energy 2020, 271, 115250. [Google Scholar] [CrossRef]
  50. Kim, S.J.; We, J.H.; Cho, B.J. A wearable thermoelectric generator fabricated on a glass fabric. Energy Environ. Sci. 2014, 7, 1959–1965. [Google Scholar] [CrossRef]
  51. 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. J. Phys. D Appl. Phys. 2018, 51, 365501. [Google Scholar] [CrossRef]
  52. Zadan, M.; Malakooti, M.H.; Majidi, C. Soft and Stretchable Thermoelectric Generators Enabled by Liquid Metal Elastomer Composites. ACS Appl. Mater. Interfaces 2020, 12, 17921–17928. [Google Scholar] [CrossRef]
  53. Yang, Y.; Hu, H.; Chen, Z.; Wang, Z.; Jiang, L.; Lu, G.; Li, X.; Chen, R.; Jin, J.; Kang, H.; et al. Stretchable Nanolayered Thermoelectric Energy Harvester on Complex and Dynamic Surfaces. Nano Lett. 2020, 20, 4445–4453. [Google Scholar] [CrossRef]
  54. Hyland, M.; Hunter, H.; Liu, J.; Veety, E.; Vashaee, D. Wearable thermoelectric generators for human body heat harvesting. Appl. Energy 2016, 182, 518–524. [Google Scholar] [CrossRef] [Green Version]
  55. Nozariasbmarz, A.; Suarez, F.; Dycus, J.H.; Cabral, M.J.; LeBeau, J.M.; Öztürk, M.C.; Vashaee, D. Thermoelectric generators for wearable body heat harvesting: Material and device concurrent optimization. Nano Energy 2020, 67, 104265. [Google Scholar] [CrossRef]
  56. Jo, S.E.; Kim, M.K.; Kim, M.S.; Kim, Y.J. Flexible thermoelectric generator for human body heat energy harvesting. Electron. Lett. 2012, 48, 1015–1017. [Google Scholar] [CrossRef]
  57. Shi, Y.; Wang, Y.; Mei, D.; Feng, B.; Chen, Z. Design and Fabrication of Wearable Thermoelectric Generator Device for Heat Harvesting. IEEE Rob. Autom. Lett. 2018, 3, 373–378. [Google Scholar] [CrossRef]
  58. Wang, Y.; Shi, Y.; Mei, D.; Chen, Z. Wearable thermoelectric generator to harvest body heat for powering a miniaturized accelerometer. Appl. Energy 2018, 215, 690–698. [Google Scholar] [CrossRef]
  59. Kim, C.S.; Lee, G.S.; Choi, H.; Kim, Y.J.; Yang, H.M.; Lim, S.H.; Lee, S.-G.; Cho, B.J. Structural design of a flexible thermoelectric power generator for wearable applications. Appl. Energy 2018, 214, 131–138. [Google Scholar] [CrossRef]
  60. 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. Appl. Energy 2017, 202, 736–745. [Google Scholar] [CrossRef]
  61. Hong, S.; Gu, Y.; Seo, J.K.; Wang, J.; Liu, P.; Meng, Y.S.; Xu, S.; Chen, R. Wearable thermoelectrics for personalized thermoregulation. Sci. Adv. 2019, 5, eaaw0536. [Google Scholar] [CrossRef] [Green Version]
  62. Sargolzaeiaval, Y.; Padmanabhan Ramesh, V.; Neumann, T.V.; Misra, V.; Vashaee, D.; Dickey, M.D.; Öztürk, M.C. Flexible thermoelectric generators for body heat harvesting—Enhanced device performance using high thermal conductivity elastomer encapsulation on liquid metal interconnects. Appl. Energy 2020, 262, 114370. [Google Scholar] [CrossRef]
  63. Ramesh, V.P.; Sargolzaeiaval, Y.; Neumann, T.; Misra, V.; Vashaee, D.; Dickey, M.D.; Ozturk, M.C. Flexible thermoelectric generator with liquid metal interconnects and low thermal conductivity silicone filler. npj Flex. Electron. 2021, 5, 5. [Google Scholar] [CrossRef]
  64. Zhu, S.; Peng, Y.; Gao, J.; Miao, L.; Lai, H.; Liu, C.; Zhang, J.; Zhang, Y.; Zhou, S.; Koumoto, K.; et al. Simultaneous Realization of Flexibility and Ultrahigh Normalized Power Density in a Heatsink-Free Thermoelectric Generator via Fine Thermal Regulation. ACS Appl. Mater. Interfaces 2022, 14, 1045–1055. [Google Scholar] [CrossRef]
  65. Lee, D.; Park, H.; Park, G.; Kim, J.; Kim, H.; Cho, H.; Han, S.; Kim, W. Liquid-metal-electrode-based compact, flexible, and high-power thermoelectric device. Energy 2019, 188, 116019. [Google Scholar] [CrossRef]
  66. Park, H.; Kim, D.; Eom, Y.; Wijethunge, D.; Hwang, J.; Kim, H.; Kim, W. Mat-like flexible thermoelectric system based on rigid inorganic bulk materials. J. Phys. D Appl. Phys. 2017, 50, 494006. [Google Scholar] [CrossRef]
  67. Eom, Y.; Wijethunge, D.; Park, H.; Park, S.H.; Kim, W. Flexible thermoelectric power generation system based on rigid inorganic bulk materials. Appl. Energy 2017, 206, 649–656. [Google Scholar] [CrossRef]
  68. Xu, Q.; Deng, B.; Zhang, L.; Lin, S.; Han, Z.; Zhou, Q.; Li, J.; Zhu, Y.; Jiang, F.; Li, Q.; et al. High-performance, flexible thermoelectric generator based on bulk materials. Cell Rep. Phys. Sci. 2022, 3, 100780. [Google Scholar] [CrossRef]
  69. Kim, J.; Khan, S.; Wu, P.; Park, S.; Park, H.; Yu, C.; Kim, W. Self-charging wearables for continuous health monitoring. Nano Energy 2021, 79, 105419. [Google Scholar] [CrossRef]
  70. Jung, S.-J.; Shin, J.; Lim, S.-S.; Kwon, B.; Baek, S.-H.; Kim, S.K.; Park, H.-H.; Kim, J.-S. Porous organic filler for high efficiency of flexible thermoelectric generator. Nano Energy 2021, 81, 105604. [Google Scholar] [CrossRef]
  71. Kim, C.S.; Yang, H.M.; Lee, J.; Lee, G.S.; Choi, H.; Kim, Y.J.; Lim, S.H.; Cho, S.H.; Cho, B.J. Self-Powered Wearable Electrocardiography Using a Wearable Thermoelectric Power Generator. ACS Energy Lett. 2018, 3, 501–507. [Google Scholar] [CrossRef]
  72. Khan, S.; Kim, J.; Roh, K.; Park, G.; Kim, W. High power density of radiative-cooled compact thermoelectric generator based on body heat harvesting. Nano Energy 2021, 87, 10618. [Google Scholar] [CrossRef]
  73. Liu, Y.; Yin, L.; Zhang, W.; Wang, J.; Hou, S.; Wu, Z.; Zhang, Z.; Chen, C.; Li, X.; Ji, H.; et al. A wearable real-time power supply with a Mg3Bi2-based thermoelectric module. Cell Rep. Phys. Sci. 2021, 2, 100412. [Google Scholar] [CrossRef]
  74. Lee, G.; Kim, C.S.; Kim, S.; Kim, Y.J.; Choi, H.; Cho, B.J. Flexible heatsink based on a phase-change material for a wearable thermoelectric generator. Energy 2019, 179, 12–18. [Google Scholar] [CrossRef]
  75. Hou, Y.; Yang, Y.; Wang, Z.; Li, Z.; Zhang, X.; Bethers, B.; Xiong, R.; Guo, H.; Yu, H. Whole Fabric-Assisted Thermoelectric Devices for Wearable Electronics. Adv. Sci. 2022, 9, 2103574. [Google Scholar] [CrossRef] [PubMed]
  76. Lee, B.; Cho, H.; Park, K.T.; Kim, J.S.; Park, M.; Kim, H.; Hong, Y.; Chung, S. High-performance compliant thermoelectric generators with magnetically self-assembled soft heat conductors for self-powered wearable electronics. Nat. Commun. 2020, 11, 5948. [Google Scholar] [CrossRef] [PubMed]
  77. Peng, Y.; Miao, L.; Liu, C.; Song, H.; Kurosawa, M.; Nakatsuka, O.; Back, S.Y.; Rhyee, J.S.; Murata, M.; Tanemura, S.; et al. Constructed Ge Quantum Dots and Sn Precipitate SiGeSn Hybrid Film with High Thermoelectric Performance at Low Temperature Region. Adv. Energy Mater. 2021, 12, 2103191. [Google Scholar] [CrossRef]
  78. Gorai, P.; Stevanović, V.; Toberer, E.S. Computationally guided discovery of thermoelectric materials. Nat. Rev. Mater. 2017, 2, 17053. [Google Scholar] [CrossRef]
  79. Wu, Z.; Zhang, S.; Liu, Z.; Mu, E.; Hu, Z. Thermoelectric converter: Strategies from materials to device application. Nano Energy 2022, 91, 106692. [Google Scholar] [CrossRef]
  80. Hong, M.; Chasapis, T.C.; Chen, Z.G.; Yang, L.; Kanatzidis, M.G.; Snyder, G.J.; Zou, J. n-Type Bi2Te3-xSex Nanoplates with Enhanced Thermoelectric Efficiency Driven by Wide-Frequency Phonon Scatterings and Synergistic Carrier Scatterings. ACS Nano 2016, 10, 4719–4727. [Google Scholar] [CrossRef]
  81. Hong, M.; Chen, Z.-G.; Zou, J. Fundamental and progress of Bi2Te3-based thermoelectric materials. Chin. Phys. B 2018, 27, 048403. [Google Scholar] [CrossRef] [Green Version]
  82. Parás-Hernández, F.U.; Fabián-Mijangos, A.; Cardona-Castro, M.A.; Alvarez-Quintana, J. Enhanced performance nanostructured thermoelectric converter for self-powering health sensors. Nano Energy 2020, 74, 104854. [Google Scholar] [CrossRef]
  83. Liu, C.-J.; Lai, H.C.; Liu, Y.-L.; Chen, L.-R. High thermoelectric figure-of-merit in p-type nanostructured (Bi,Sb)2Te3 fabricated via hydrothermal synthesis and evacuated-and-encapsulated sintering. J. Mater. Chem. 2012, 22, 4825–4831. [Google Scholar] [CrossRef]
  84. Jun, P. Preparation and Characterization of Telluride Based Thermoeletric Materials and Device. Ph.D. Thesis, University of Science and Technology Beijing, Beijing, China, 2019. [Google Scholar]
  85. Zhen, Y. The Fabrication and Mechanical Performance of p-Type Bismuth Telluride-Based Thermoelectric Materials. Ph.D. Thesis, Wuhan University of Technology, Wuhan, China, 2015. [Google Scholar]
  86. Kim, S.I.; Lee, K.H.; Mun, H.A.; Kim, H.S.; Hwang, S.W.; Roh, J.W.; Yang, D.J.; Shin, W.H.; Li, X.S.; Lee, Y.H.; et al. Dense dislocation arrays embedded in grain boundaries for high-performance bulk thermoelectrics. Science 2015, 348, 109–114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Poudel, B.; Hao, Q.; Ma, Y.; Lan, Y.; Minnich, A.; Yu, B.; Yan, X.; Wang, D.; Muto, A.; Vashaee, D. High-Thermoelectric Performance of Nanostructured Bismuth Antimony Telluride Bulk Alloys. Science 2008, 320, 634–638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Nozariasbmarz, A.; Vashaee, D. Effect of Microwave Processing and Glass Inclusions on Thermoelectric Properties of P-Type Bismuth Antimony Telluride Alloys for Wearable Applications. Energies 2020, 13, 4524. [Google Scholar] [CrossRef]
  89. 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. [Google Scholar] [CrossRef]
  90. Lu, Y.; Qiu, Y.; Cai, K.; Ding, Y.; Wang, M.; Jiang, C.; Yao, Q.; Huang, C.; Chen, L.; He, J. Ultrahigh power factor and flexible silver selenide-based composite film for thermoelectric devices. Energy Environ. Sci. 2020, 13, 1240–1249. [Google Scholar] [CrossRef]
  91. Gao, W.; Yi, X.; Cui, B.; Wang, Z.; Huang, J.; Sui, J.; Liu, Z. The critical role of boron doping in the thermoelectric and mechanical properties of nanostructured α-MgAgSb. J. Mater. Chem. C 2018, 6, 9821–9827. [Google Scholar] [CrossRef]
  92. Zhang, T.; Dong, B.; Wang, X. Optimization of the thermoelectric performance of α-MgAgSb-based materials by Zn-doping. J. Mater. Sci. 2021, 56, 13715–13722. [Google Scholar] [CrossRef]
  93. Zheng, Y.; Liu, C.; Miao, L.; Li, C.; Huang, R.; Gao, J.; Wang, X.; Chen, J.; Zhou, Y.; Nishibori, E. Extraordinary thermoelectric performance in MgAgSb alloy with ultralow thermal conductivity. Nano Energy 2019, 59, 311–320. [Google Scholar] [CrossRef]
  94. Zhou, C.; Lee, Y.K.; Yu, Y.; Byun, S.; Luo, Z.Z.; Lee, H.; Ge, B.; Lee, Y.L.; Chen, X.; Lee, J.Y.; et al. Polycrystalline SnSe with a thermoelectric figure of merit greater than the single crystal. Nat. Mater. 2021, 20, 1378–1384. [Google Scholar] [CrossRef]
  95. Chen, Z.-G.; Shi, X.; Zhao, L.-D.; Zou, J. High-performance SnSe thermoelectric materials: Progress and future challenge. Prog. Mater. Sci. 2018, 97, 283–346. [Google Scholar] [CrossRef] [Green Version]
  96. Su, L.; Wang, D.; Wang, S.; Qin, B.; Wang, Y.; Qin, Y.; Jin, Y.; Chang, C.; Zhao, L.-D. High thermoelectric performance realized through manipulating layered phonon-electron decoupling. Science 2022, 375, 1385–1389. [Google Scholar] [CrossRef] [PubMed]
  97. Zhao, L.-D.; Tan, G.; Hao, S.; He, J.; Pei, Y.; Chi, H.; Wang, H.; Gong, S.; Xu, H.; Dravid, V.P.; et al. Ultrahigh power factor and thermoelectric performance in hole-doped single-crystal SnSe. Science 2015, 351, 141–144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Deng, S.; Jiang, X.; Chen, L.; Qi, N.; Tang, X.; Chen, Z. Ultralow Thermal Conductivity and High Thermoelectric Performance in AgCuTe1-xSex through Isoelectronic Substitution. ACS Appl. Mater. Interfaces 2021, 13, 868–877. [Google Scholar] [CrossRef] [PubMed]
  99. Jiang, J.; Zhu, H.; Niu, Y.; Zhu, Q.; Song, S.; Zhou, T.; Wang, C.; Ren, Z. Achieving high room-temperature thermoelectric performance in cubic AgCuTe. J. Mater. Chem. A 2020, 8, 4790–4799. [Google Scholar] [CrossRef]
  100. Gao, J.; Miao, L.; Liu, C.; Wang, X.; Peng, Y.; Wei, X.; Zhou, J.; Chen, Y.; Hashimoto, R.; Asaka, T.; et al. A novel glass-fiber-aided cold-press method for fabrication of n-type Ag2Te nanowires thermoelectric film on flexible copy-paper substrate. J. Mater. Chem. A 2017, 5, 24740–24748. [Google Scholar] [CrossRef]
  101. Zhong, Y.; Tang, J.; Liu, H.; Chen, Z.; Lin, L.; Ren, D.; Liu, B.; Ang, R. Optimized Strategies for Advancing n-Type PbTe Thermoelectrics: A Review. ACS Appl. Mater. Interfaces 2020, 12, 49323–49334. [Google Scholar] [CrossRef]
  102. Zhao, H.; Sui, J.; Tang, Z.; Lan, Y.; Jie, Q.; Kraemer, D.; McEnaney, K.; Guloy, A.; Chen, G.; Ren, Z. High thermoelectric performance of MgAgSb-based materials. Nano Energy 2014, 7, 97–103. [Google Scholar] [CrossRef]
  103. Lei, J.; Zhang, D.; Guan, W.; Ma, Z.; Cheng, Z.; Wang, C.; Wang, Y. Enhancement of thermoelectric figure of merit by the insertion of multi-walled carbon nanotubes in α-MgAgSb. Appl. Phys. Lett. 2018, 113, 083901. [Google Scholar] [CrossRef] [Green Version]
  104. Wang, H.; Liu, X.; Zhou, Z.; Wu, H.; Chen, Y.; Zhang, B.; Wang, G.; Zhou, X.; Han, G. Constructing n-type Ag2Se/CNTs composites toward synergistically enhanced thermoelectric and mechanical performance. Acta Mater. 2022, 223, 117502. [Google Scholar] [CrossRef]
  105. Day, T.; Drymiotis, F.; Zhang, T.; Rhodes, D.; Shi, X.; Chen, L.; Snyder, G.J. Evaluating the potential for high thermoelectric efficiency of silver selenide. J. Mater. Chem. C 2013, 1, 7568–7573. [Google Scholar] [CrossRef] [Green Version]
  106. Jood, P.; Chetty, R.; Ohta, M. Structural stability enables high thermoelectric performance in room temperature Ag2Se. J. Mater. Chem. A 2020, 8, 13024–13037. [Google Scholar] [CrossRef]
  107. Jiang, C.; Ding, Y.; Cai, K.; Tong, L.; Lu, Y.; Zhao, W.; Wei, P. Ultrahigh Performance of n-Type Ag2Se Films for Flexible Thermoelectric Power Generators. ACS Appl. Mater. Interfaces 2020, 12, 9646–9655. [Google Scholar] [CrossRef] [PubMed]
  108. Chen, J.; Liu, T.; Bao, D.; Zhang, B.; Han, G.; Liu, C.; Tang, J.; Zhou, D.; Yang, L.; Chen, Z.G. Nanostructured monoclinic Cu2Se as a near-room-temperature thermoelectric material. Nanoscale 2020, 12, 20536–20542. [Google Scholar] [CrossRef] [PubMed]
  109. Liu, Y.; Shi, Y.; Li, J.; Guo, X.; Wang, Y.; Xiang, Q.; Guo, S.; Ze, R.; Zeng, J.; Xiang, Y.; et al. Comprehensive performance prediction and power promotion for wearable thermoelectric generator with flexible encapsulation in practical application. Energy Convers. Manag. 2020, 220, 113080. [Google Scholar] [CrossRef]
  110. Tanwar, A.; Lal, S.; Razeeb, K. Structural Design Optimization of Micro-Thermoelectric Generator for Wearable Biomedical Devices. Energies 2021, 14, 2339. [Google Scholar] [CrossRef]
  111. Leonov, V.; Vullers, R.J.M. Wearable Thermoelectric Generators for Body-Powered Devices. J. Electron. Mater. 2009, 38, 1491–1498. [Google Scholar] [CrossRef]
  112. Zhu, K.; Deng, B.; Zhang, P.; Kim, H.S.; Jiang, P.; Liu, W. System efficiency and power: The bridge between the device and system of a thermoelectric power generator. Energy Environ. Sci. 2020, 13, 3514–3526. [Google Scholar] [CrossRef]
  113. Pietrzyk, K.; Soares, J.; Ohara, B.; Lee, H. Power generation modeling for a wearable thermoelectric energy harvester with practical limitations. Appl. Energy 2016, 183, 218–228. [Google Scholar] [CrossRef]
  114. Elghool, A.; Basrawi, F.; Ibrahim, T.K.; Habib, K.; Ibrahim, H.; Idris, D.M.N.D. A review on heat sink for thermo-electric power generation: Classifications and parameters affecting performance. Energy Convers. Manag. 2017, 134, 260–277. [Google Scholar] [CrossRef]
  115. 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. [Google Scholar] [CrossRef]
  116. Thermoelectric Generator Market. Thermoelectric Generator Market by Material, Application, and End-use Industry: Global Opportunity Analysis and Industry Forecast, 2021–2030. Available online: https://www.researchandmarkets.com/reports/5521526/thermoelectric-generator-market-by-material# (accessed on 5 February 2022).
  117. Tom Torfs, V.L. Ruud Vullers, Pulse Oximeter Fully Powered by Human Body Heat. Sens. Transducers 2007, 80, 1230–1238. [Google Scholar]
  118. Lay-Ekuakille, A.; Vendramin, G.; Trotta, A.; Mazzotta, G. Thermoelectric Generator Design Based on Power from Body Heat for Biomedical Autonomous Devices. In International Workshop on Medical Measurements and Applications; IEEE: Cetraro, Italy, 2009. [Google Scholar]
  119. Yuan, J.; Zhu, R.; Li, G. Self-Powered Electronic Skin with Multisensory Functions Based on Thermoelectric Conversion. Adv. Mater. Technol. 2020, 2000419. [Google Scholar] [CrossRef]
  120. de Fazio, R.; Cafagna, D.; Marcuccio, G.; Minerba, A.; Visconti, P. A Multi-Source Harvesting System Applied to Sensor-Based Smart Garments for Monitoring Workers’ Bio-Physical Parameters in Harsh Environments. Energies 2020, 13, 2161. [Google Scholar] [CrossRef]
Figure 1. A schematic diagram showing the structure, optimization strategy, and applications of WTEG.
Figure 1. A schematic diagram showing the structure, optimization strategy, and applications of WTEG.
Energies 15 03375 g001
Figure 2. Schematic diagrams of the (A) Seebeck effect [22]. (B) TEG module [22].
Figure 2. Schematic diagrams of the (A) Seebeck effect [22]. (B) TEG module [22].
Energies 15 03375 g002
Figure 3. (A) Schematic for the heat transfer model of a wearable thermoelectric generator [37]. (B) Reported human skin temperatures for different points on the body at varying ambient temperatures. (C) Heat transfer coefficient of the TEG/skin interface as a function of pressure for various locations on the body [44]. (D) Heat transfer coefficient of a small heat sink plotted as a function of the thermal conductivity of the heat sink material for different air velocities ranging from almost 0 to 1.4 m/s [44].
Figure 3. (A) Schematic for the heat transfer model of a wearable thermoelectric generator [37]. (B) Reported human skin temperatures for different points on the body at varying ambient temperatures. (C) Heat transfer coefficient of the TEG/skin interface as a function of pressure for various locations on the body [44]. (D) Heat transfer coefficient of a small heat sink plotted as a function of the thermal conductivity of the heat sink material for different air velocities ranging from almost 0 to 1.4 m/s [44].
Energies 15 03375 g003
Figure 4. (A) The schematic of the Seiko Thermic watch that could generate about 25 µW from body heat to power the watch [45]. (B) Schematic diagram of a low-profile r-WTEGs and infrared image of copper with a high emissivity coating showing effective cooling using the large surface area available in a wristband [47]. (C) The schematic of a low FF TEG module and experimental setup used to measure the TEG module on the human body [48].
Figure 4. (A) The schematic of the Seiko Thermic watch that could generate about 25 µW from body heat to power the watch [45]. (B) Schematic diagram of a low-profile r-WTEGs and infrared image of copper with a high emissivity coating showing effective cooling using the large surface area available in a wristband [47]. (C) The schematic of a low FF TEG module and experimental setup used to measure the TEG module on the human body [48].
Energies 15 03375 g004
Figure 5. (A) Structural design and fabricated images of ns-FTEGs and FTEGs-powered wearable multi-sensory bracelet [49]. (B) Schematic illustration of the flexibility of the embedded Cu film (Cu foil vs. the screen-printed Cu film) in PDMS and demonstration of band-type flexible TE generator for harvesting thermal energy from human skin at an air temperature of 15 °C (scale bar, 1 cm) [50]. (C) Schematic design and structure of ns-FTEGs encapsulated by bakelite [51].
Figure 5. (A) Structural design and fabricated images of ns-FTEGs and FTEGs-powered wearable multi-sensory bracelet [49]. (B) Schematic illustration of the flexibility of the embedded Cu film (Cu foil vs. the screen-printed Cu film) in PDMS and demonstration of band-type flexible TE generator for harvesting thermal energy from human skin at an air temperature of 15 °C (scale bar, 1 cm) [50]. (C) Schematic design and structure of ns-FTEGs encapsulated by bakelite [51].
Energies 15 03375 g005
Figure 6. (A) s-FTEGs with stretchable liquid metal electrodes [52]. (B) s-WTEGs with snake-shaped copper electrodes [53].
Figure 6. (A) s-FTEGs with stretchable liquid metal electrodes [52]. (B) s-WTEGs with snake-shaped copper electrodes [53].
Energies 15 03375 g006
Figure 7. (A) Results of TE geometric parameters obtained by numerical analysis [37]. (B) Results of TE geometric parameters obtained by Finite Element Simulation [64].
Figure 7. (A) Results of TE geometric parameters obtained by numerical analysis [37]. (B) Results of TE geometric parameters obtained by Finite Element Simulation [64].
Energies 15 03375 g007
Figure 8. (A) Temperature difference from core body temperature to the ambient air [44]. (B) Schematic structure of FTEG with a copper heat sink and the fabricated device worn on the wrist [67]. (C) Schematic of the r-WTEG with a heat spreader and the experimental setup of the TEG on the wrist [54]. (D) Schematic of the FTEG with a flexible heat sink base on SAP [71] (E) with a flexible PCM heat sink [74] and (F) with a radiative cooling heat sink [72].
Figure 8. (A) Temperature difference from core body temperature to the ambient air [44]. (B) Schematic structure of FTEG with a copper heat sink and the fabricated device worn on the wrist [67]. (C) Schematic of the r-WTEG with a heat spreader and the experimental setup of the TEG on the wrist [54]. (D) Schematic of the FTEG with a flexible heat sink base on SAP [71] (E) with a flexible PCM heat sink [74] and (F) with a radiative cooling heat sink [72].
Energies 15 03375 g008
Figure 9. Schematic of the FTEG with flexible high thermal-conductivity substrates based on EGaIn/PDMS [62].
Figure 9. Schematic of the FTEG with flexible high thermal-conductivity substrates based on EGaIn/PDMS [62].
Energies 15 03375 g009
Figure 10. (A)The internal temperature distributions of the TEG filled with PDMS and MF [64]. (B) Schematic of the WTEG filled with MF [64]. (C) Schematic of the WTEG filled with porous and thermal conductivity of PDMS and porous PDMS [70].
Figure 10. (A)The internal temperature distributions of the TEG filled with PDMS and MF [64]. (B) Schematic of the WTEG filled with MF [64]. (C) Schematic of the WTEG filled with porous and thermal conductivity of PDMS and porous PDMS [70].
Energies 15 03375 g010
Figure 11. (A) Semi-automatic fabrication process for the wearable TEG based on FPCB and the fabricated device [58]. (B) Highly automated fabrication process for the wearable TEG based on AgNW electrode and the fabricated device [76].
Figure 11. (A) Semi-automatic fabrication process for the wearable TEG based on FPCB and the fabricated device [58]. (B) Highly automated fabrication process for the wearable TEG based on AgNW electrode and the fabricated device [76].
Energies 15 03375 g011
Figure 12. (A) Watch-type thermoelectric and oximeter [117]. (B) Flexible thermoelectric ECG detector [71]. (C) Thermal and power hearing aid [118].
Figure 12. (A) Watch-type thermoelectric and oximeter [117]. (B) Flexible thermoelectric ECG detector [71]. (C) Thermal and power hearing aid [118].
Energies 15 03375 g012
Figure 13. (A) Human movement state detection [58]. (B) Detection of hand activity [53].
Figure 13. (A) Human movement state detection [58]. (B) Detection of hand activity [53].
Energies 15 03375 g013
Figure 14. (A) Temperature detection device based on s-FTEG [76]. (B) A multifunctional detection system based on a palm-type thermoelectric generator [119]. (C) Intelligent clothing system [120].
Figure 14. (A) Temperature detection device based on s-FTEG [76]. (B) A multifunctional detection system based on a palm-type thermoelectric generator [119]. (C) Intelligent clothing system [120].
Energies 15 03375 g014
Table 1. Performances and features summary for different types of WTEG.
Table 1. Performances and features summary for different types of WTEG.
Types of WTEGsMaterials of
TE Legs
ZT or
Efficiency
EncapsulationElectrodeAmbient TemperatureHeat SinkPower Density
(μW/cm2)
Ref.
r-WTEGBismuth Telluride0.8PDMS/Alumina Copper18.3 °CHeat spreader6[54]
r-WTEGCommercial Bi2Te3/Alumina Copper22 °CPlate fin heat sink20[46]
r-WTEGCommercial Bi2Te3/CeramicCopper23 °CPlate fin heat sink28.5[47]
r-WTEGBi2Te30.14%AlN Copper17 °CPlate fin heat sink35[48]
r-WTEGBi2Te3/AlNCopper25 °CHeat spreader44[55]
ns-FTEGCommercial Bi2Te3/PDMSFPC13 °CNo0.084[56]
ns-FTEGBi2Te3(Glass fabric)/PDMSCopper15 °CNo0.75[50]
ns-FTEGCommercial Bi2Te3/NoFPC19 °CNo0.4[57]
ns-FTEGCommercial Bi2Te3/PDMSFPC25 °CNo0.526[58]
ns-FTEGCommercial Bi2Te3/PolymerCopper25 °CNo2.28[59]
ns-FTEGCommercial Bi2Te30.35PDMSLM5 °CNo2.5[60]
ns-FTEGCommercial Bi2Te30.85NoFPC13 °CNo3.5[49]
ns-FTEGCommercial Bi2Te30.71PDMSCopper24 °CNo4.5[61]
ns-FTEGCommercial Bi2Te3/PDMSLM24 °CNo5.2[62]
ns-FTEGCommercial Bi2Te3/Aerogel/PDMSLM24 °CNo5.4[63]
ns-FTEGCommercial Bi2Te3/MFCopper24 °CHeat spreader7[64]
ns-FTEGBi2Te3/BakeliteCopper20 °CPlate fin heat sink1.6[65]
ns-FTEGBi2Te3/BakeliteCopper24 °CFlexible heat sink4.78[51]
ns-FTEGBi2Te3/BakeliteCopper24 °CFlexible heat sink5.6[66]
ns-FTEGBi2Te3/BakeliteCopper24 °CPlate fin heat sink5.6[67]
ns-FTEGBi2Te30.75FabricCopper24 °CAir velocity: 0.2m/s6.3[68]
ns-FTEGBi2Te3/NoCopper24 °CPlate fin heat sink8[69]
ns-FTEGBi2Te3/PDMSLM&Copper24 °CHydrogels8.3[65]
ns-FTEG Commercial Bi2Te30.75Porous PDMSCopper24 °CSAP heat sink9.7[70]
ns-FTEGCommercial Bi2Te30.68PolymerCopper24 °CSAP heat sink13[71]
ns-FTEGBi2Te3/NoFPC24 °CRadiant cooling12.5[72]
ns-FTEGMg3Bi2/Bi2Te3/Porous PUCopper16 °CAir velocity: 1.1m/s20.6[73]
ns-FTEGCommercial Bi2Te30.64PolymerCopper13 °CPCM heat sink20[74]
s-FTEGBi2Te3 (PEDOT:PSS)/PDMSCopperΔT=33KNo0.87[53]
s-FTEGBi2Te3/Elastic fabricPolyester Fiber ΔT=33KNo1 [75]
s-FTEGCommercial Bi2Te33/PDMSLMΔT=10KNo4[52]
s-FTEGBi2Te3/PDMS/Ag-NiAgNW10 °CNo6.97[76]
LM: liquid metal. SAP: absorbent polymer. MF: Melamine foam. PU: Polyurethane. AgNW: Silver nanowires.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhu, S.; Fan, Z.; Feng, B.; Shi, R.; Jiang, Z.; Peng, Y.; Gao, J.; Miao, L.; Koumoto, K. Review on Wearable Thermoelectric Generators: From Devices to Applications. Energies 2022, 15, 3375. https://doi.org/10.3390/en15093375

AMA Style

Zhu S, Fan Z, Feng B, Shi R, Jiang Z, Peng Y, Gao J, Miao L, Koumoto K. Review on Wearable Thermoelectric Generators: From Devices to Applications. Energies. 2022; 15(9):3375. https://doi.org/10.3390/en15093375

Chicago/Turabian Style

Zhu, Sijing, Zheng Fan, Baoquan Feng, Runze Shi, Zexin Jiang, Ying Peng, Jie Gao, Lei Miao, and Kunihito Koumoto. 2022. "Review on Wearable Thermoelectric Generators: From Devices to Applications" Energies 15, no. 9: 3375. https://doi.org/10.3390/en15093375

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop