Next Article in Journal
Sensor Input Type and Location Influence Outdoor Running Terrain Classification via Deep Learning Approaches
Previous Article in Journal
Analysis of P300 Evoked Potentials to Determine Pilot Cognitive States
Previous Article in Special Issue
An Attomolar-Level Biosensor Based on Polypyrrole and TiO2@Pt Nanocomposite for Electrochemical Detection of TCF3-PBX1 Oncogene in Acute Lymphoblastic Leukemia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

SWCNT/PEDOT:PSS/SA Composite Yarns with High Mechanical Strength and Flexibility via Wet Spinning for Thermoelectric Applications

Department of Materials Science, Tokai University, 4-1-1 Kitakaname, Hiratsuka 259-1292, Kanagawa, Japan
*
Author to whom correspondence should be addressed.
Sensors 2025, 25(19), 6202; https://doi.org/10.3390/s25196202
Submission received: 31 July 2025 / Revised: 20 September 2025 / Accepted: 23 September 2025 / Published: 7 October 2025
(This article belongs to the Special Issue Nanotechnology Applications in Sensors Development)

Abstract

To fabricate thermoelectric generators (TEGs) with high mechanical strength using single-walled carbon nanotubes (SWCNTs), we combined SWCNTs, poly(3, 4-ethylenedioxythiophene):poly(4-styrenesulfonate) (PEDOT:PSS), and sodium alginate (SA) to synthesize flexible SWCNT/PEDOT:PSS/SA composite yarns via wet spinning. The composite yarns were flexible and dense, with a diameter of approximately 290 µm. Their tensile strength and breaking strain were 151 MPa and 12.7%, respectively, which were approximately 10 and 4 times those of the SWCNT films. However, the thermoelectric properties of the composite yarns were inferior to those of the SWCNT films. The temperature distribution and output voltage of the fabricated TEG with composite yarns were measured at a heater temperature of 100 °C. The temperature difference generated by the TEG with composite yarns was approximately 75% of that generated by the TEG with SWCNT films because the composite yarn had a smaller specific surface area. The output voltage of the TEG with two composite yarns (0.21 mV) was lower than that of the TEG with two SWCNT films. However, arranging the composite yarns at a high density resulted in an output voltage exceeding that for the TEGs with SWCNT films. These findings are highly beneficial for yarn-based TEGs used in wearable sensors.

1. Introduction

As Internet of Things (IoT) technology progresses, its various potential applications are being explored [1,2]. Sensors that use IoT technology are expected to permeate almost every aspect of our daily lives, including smart cities, healthcare, smart agriculture, logistics, retail, and smart living. Among the various IoT sensors, wearable sensors offer the greatest growth potential for commercial development, as monitoring human activity has attracted significant interest owing to its various application areas, including sports, healthcare, and automotive driving [3,4,5,6]. One of the critical issues for wearable sensors is securing a power source to supply power to wearable sensors [7,8,9,10]. Because of the nature of wearable sensors, a wired power supply is inconvenient. In addition, because of the large number of wearable sensors used, it is difficult to replace batteries, and monitoring would be interrupted during replacement. Therefore, the most reliable approach is to use a self-powered system that employs energy harvesting [11,12,13,14,15].
Thermoelectric power generation, which converts thermal energy into electricity, is the most suitable self-powered system for wearable sensors such as humidity, position, and temperature sensors [16,17,18,19,20,21,22,23,24,25]. This is because it can generate electricity at a constant output, and body heat can be used as the heat source [26,27]. The requirements for thermoelectric generation in wearable sensors are that the generators be lightweight, compact, and flexible [28,29,30,31,32] and that the thermoelectric materials used must be low-cost, have a low environmental impact, and be nontoxic. In addition, thermoelectric materials can form thread-like structures that can be woven into fabrics. Considering the aforementioned conditions, the ideal thermoelectric materials for wearable sensors are single-walled carbon nanotubes (SWCNTs) and conductive organic polymers [33,34,35,36,37,38,39].
SWCNTs are composed of a single layer of hexagonally arranged carbon atoms (graphene) rolled into a cylindrical shape. They are extremely thin, ranging from 1 to 5 nm in diameter; a few micrometers long; and characterized by their light weight, high mechanical strength, and flexibility. Depending on their chirality, SWCNTs can exhibit metallic or semiconducting properties. However, only materials with semiconducting properties can be used as thermoelectric materials. In general, SWCNTs exhibit p-type semiconducting properties owing to the absorption of oxygen molecules [40,41,42,43,44], and the development of stable n-type SWCNTs is a major research topic [45,46,47,48,49]. SWCNT yarns have been fabricated via dry spinning from SWCNT forests, in which the nanotubes grow densely and vertically on a substrate [50,51,52]. SWCNT yarns have high strength, flexibility, and excellent thermoelectric properties [53,54,55,56,57]; however, SWCNT forests are only available through limited purchasing channels and require specialized equipment to be spun. In contrast, organic polymer yarns have high flexibility and can be easily fabricated using wet spinning without specialized equipment [58,59,60,61,62]. However, the strength, durability, and electrical conductivity of organic polymer yarns are inferior to those of SWCNT yarns. The most effective method for producing thermoelectric yarns for wearable sensor applications is combining SWCNTs and organic polymers through wet spinning [63,64,65,66,67]. Li et al., Kim et al., and Xu et al. fabricated carbon nanotube (CNT)/polyaniline, CNT/poly(vinylidene fluoride), and poly(3,4-ethylenedioxythiophene):poly(4-styrenesulfonate) (PEDOT:PSS) composite fibers, respectively, via wet spinning [68,69,70]. The main challenge is to fabricate SWCNT-based composite yarns with flexibility and high mechanical strength and to prepare high-performance TEGs using the composite yarns [71].
In this study, we combined SWCNTs, PEDOT:PSS, and sodium alginate (SA) to fabricate flexible SWCNT/PEDOT:PSS/SA composite yarns via wet spinning. The microstructures of the composite yarns were analyzed. The mechanical strength and thermoelectric properties of the composite yarns were measured at approximately 300 K and compared with those of SWCNT films fabricated using vacuum filtering. The tensile strength and breaking strain of the composite yarns were 151 MPa and 12.7%, respectively, which were approximately 10 and 4 times those of the SWCNT film. However, the thermoelectric properties of the composite yarns were inferior to those of the SWCNT films. TEGs were fabricated by attaching the yarns to fabric, and their performance was evaluated. The TEG with the two composite yarns had an output voltage of 0.21 mV. The results indicate that the synthesized TEGs composed of the SWCNT/PEDOT:PSS/SA composite yarns can be used as autonomous power supplies for wearable sensors.

2. Materials and Methods

Figure 1 illustrates the manufacturing process for the SWCNT/PEDOT:PSS/SA composite yarns. First, 0.08 g of SWCNT powder (ZEONANO SG101, purity > 90%, diameter 3–5 nm, ZEON, Tokyo, Japan), 0.32 g of PEDOT:PSS pellets (Sigma-Aldrich, St. Louis, MO, USA), and 40 mL of ion-exchanged water at a resistivity of 18.2 MΩ·cm were mixed in a 50-mL beaker. Subsequently, the mixture was ultrasonically dispersed to prepare the composite ink using an ultrasonic homogenizer (Branson, Ultrasonic Sonifier 250, Danbury, CT, USA) at a frequency of 20 kHz, ultrasonic amplitude of 104 μm, and ultrasonic horn-tip diameter of 12.7 mm. The ultrasonic dispersion amplitude was set to 60% (nominal value: 200 W) for 30 min in an ice bath. After ultrasonic dispersion, 0.4 g of SA (Fujifilm Wako Pure Chemical, Osaka, Japan) was added to the ink as a coagulant, and the composite ink was stirred for 15 min. The composition of the resulting ink was 0.2 wt% SWCNTs, 0.7 wt% PEDOT:PSS, and 1.0 wt% SA, with a solid concentration of 40 wt% PEDOT:PSS. The composite ink was filled into a syringe and spun into 600 mL of a 0.5 wt% CaCl2 aqueous solution using a syringe pump (MR-1A, AS ONE, Osaka, Japan) at a spinning speed of 0.36 mL/min. The immersion time was maintained at 9 min. Subsequently, the composite yarns were washed three times in a Petri dish with ion-exchanged water. It was then naturally dried for 24 h, and a weight was attached to one end to apply a light tensile load. The obtained composite yarns were thin and flexible. For comparison, using the same SWCNT powder, a 45-µm-thick SWCNT film was produced via vacuum filtration using a membrane filter (T100A090C, ADVANTEC®, pore size 1.0 µm, diameter 90 mm, film thickness 61 µm, Tokyo, Japan) at a temperature of approximately 300 K. The manufacturing process for the SWCNT films is presented in the Supplementary Materials (Figure S1) [72].
The microstructures of the composite yarns and SWCNT films were analyzed using field-emission scanning electron microscopy (FE-SEM; JSM-7100F, JEOL, Tokyo, Japan). The tensile strengths, breaking strains, and Young’s moduli of the composite yarns and reference SWCNT films were measured using tensile tests (MX-1000N-FA, IMADA, Toyohashi, Japan). Tensile test specimens were prepared by attaching 35-mm-long pieces of composite yarn to a 0.08-mm-thick piece of paper with an adhesive (Japanese Industrial Standards (JIS) R 7606). The tensile tests were performed at approximately 300 K at a testing speed of 10 mm/min. The Seebeck coefficients and electrical conductivities of the composite yarns and reference SWCNT films were measured at approximately 300 K using a thermoelectric measurement system (ZEM-3, ADVANCE RIKO, Yokohama, Japan).

3. Results and Discussion

3.1. Properties of Composite Yarns

The microstructures of the composite yarns and the reference SWCNT films are presented in Figure 2. The low-magnification SEM image of the composite yarn in Figure 2a reveals a relatively smooth surface with visible streaks along its length. The diameter of the composite yarn was approximately 290 µm. The SWCNT bundles were not visible, because of the low magnification. This is evident from the low-magnification SEM image of the SWCNT film presented in Figure 2b, in which the SWCNT bundles are also invisible. In the high-magnification SEM image of the composite yarn in Figure 2c, the surface unevenness is more apparent. However, SWCNT bundles were not observed in the composite yarn, even though the SWCNT film exhibited a bundle structure at the same magnification, as shown in Figure 2d. Here, SWCNT bundles with an average diameter of 100 nm were observed, and the pore size of the SWCNT film was approximately 500 nm or less. The SWCNT bundles were not observed in the composite yarn because both the PEDOT:PSS and SA were coated with SWCNT bundles.
The stress–strain curves of the composite yarn and a reference SWCNT film are shown in Figure 3. The stress–strain curve of the composite yarn was divided into an elastic region and a plastic region, whereas the SWCNT film did not clearly exhibit a boundary between them. Furthermore, the tensile stress and breaking strain of the composite yarn were significantly higher than those of the SWCNT film. For a detailed evaluation, the mechanical properties of the composite yarn and SWCNT film were compared, as shown in Table 1. The tensile strength and breaking strain of the composite yarn were 151 MPa and 12.7%, respectively, which were approximately 10 and 4 times those of the SWCNT film. The composite yarns had higher tensile strength because the addition of PEDOT:PSS and SA to the SWCNT bundles increased the contact area by forming a dense structure, as shown in Figure 2c. Meanwhile, the SWCNT film had lower tensile strength because each SWCNT bundle formed only point or line contacts, resulting in a porous structure, as shown in Figure 2d. The composite yarns exhibited high breaking strain owing to the significant plastic deformation of PEDOT:PSS—an organic material [73]. The Young’s modulus was calculated from the stress–strain curve in the elastic region, particularly in the low-strain region. The Young’s modulus of the composite yarn was 0.26 GPa, which was lower than that of the SWCNT film, indicating that the composite yarns had flexibility and elasticity. A photo of the composite yarn in a bent state is shown in the Supplementary Materials (Figure S2). The longitudinal sound velocity (vL) is expressed as v L =   E / ρ , where E and r represent the Young’s modulus and mass density, respectively. The composite yarn had a higher mass density (1.36 g/cm3) than the SWCNT film owing to its denser structure. Consequently, the longitudinal sound velocities of the composite yarn and SWCNT film were 437 and 1289 m/s, respectively. The composite film exhibited a lower vL value because of its low Young’s modulus and high mass density.
The thermoelectric properties of the composite yarn and SWCNT film are presented in Table 2. They exhibited p-type Seebeck coefficients (S) of 29.4 and 55.5 μV/K, respectively. The lower Seebeck coefficient of the composite yarn is attributed to the low Seebeck coefficient of the PEDOT:PSS film, which is approximately 15 μV/K owing to the high carrier concentration of the PEDOT:PSS [74]. The electrical conductivity (σ) of the composite yarn was approximately half that of the SWCNT film because the cross-linked structure of SA has almost no conductivity compared to its SWCNT and PEDOT:PSS [75]. In addition, the thermoelectric properties of the composite yarn were compared with those of the SWCNT/PEDOT composite [39]. The Seebeck coefficient of the composite yarn was comparable to that of the SWCNT/PEDOT composite; however, the electrical conductivity of the composite yarn was approximately 10 times lower than that of the SWCNT/PEDOT composite. The power factor (σS2) was calculated from the measured Seebeck coefficient and electrical conductivity. The composite yarn had a power factor of 1.3 µW/(m·K2), which was 14% of that of the SWCNT film. Therefore, the composite yarn exhibited higher tensile strength and breaking strain but inferior thermoelectric properties. In the future, the thermoelectric properties of composite yarns should be improved via doping and twisting while maintaining their mechanical strength [76,77,78].

3.2. Performance of Composite-Yarn TEGs

Figure 4a shows schematics of thermoelectric generators (TEGs) with the composite yarn and SWCNT films. We prepared two types of composite-yarn TEGs, including two and six composite yarns. For the TEG with two composite yarns, the CNT yarn was cut to a length of 20 mm. The cut yarns were aligned parallel at intervals of 13 mm, and both ends were bonded to polyester fabric (70 mm × 40 mm) using Ag paste. The ends of the yarns were connected in series using copper wire with a diameter of 50 μm. The other ends were connected to an external circuit using fine copper wire. For the TEG with six composite yarns, the length of the yarns was 20 mm. The cut yarns were aligned in parallel at intervals of 10 mm, and both ends were bonded to polyester fabric (size: 70 mm × 40 mm) using Ag paste. The method of connecting the copper wires was the same for both TEGs. For comparison, a TEG with SWCNT films was fabricated on polyester fabric (40 mm × 40 mm). The SWCNT films were 20 mm long and 5 mm wide and had an aligned interval of 10 mm. The methods for bonding to the fabric and creating copper-wire connections were the same as those used for the TEGs with composite yarns. A TEG with SWCNT films can be compared to a TEG with two yarns when the number of pieces is the same or to a TEG with six yarns when the volume of material used is the same. Photographs of the completed TEGs are presented in Figure 4b. The procedure for evaluating the TEG performance is shown in Figure 4c. The TEGs were suspended using a cotton thread directly above the heater (iStir HP 320; NEUATION, Gandhinagar, India), keeping the bottom of the TEG 15 mm from the heater. Photographs of the measurement setup are presented in Figure 4d. The copper fine wires from the TEGs were connected to a data logger (LR8432, HIOKI, Nagano, Japan), and the output voltage was measured for 10 min at a heater temperature of 100 °C. The temperature distribution of the TEGs was measured 5 and 10 min after the heater temperature reached 100 °C using a thermography camera (OPTXI40LTF20CFKT090, OPTRIS, Berlin, Germany) with spatial and temperature resolutions of 382 × 288 pixels and 0.08 °C, respectively. The temperature distribution of the TEGs with the two composite yarns and SWCNT films was analyzed using high-resolution close-up thermography images.
Figure 5 shows the thermal analyses of the TEGs under heating. To clarify the thermography images, photographs of the TEGs with composite yarns and SWCNT films are presented in Figure 5a,b, respectively, which show the same areas as the thermography images. Figure 5c,d show the thermography images of the TEGs with composite yarns and SWCNT films, respectively, after heating for 5 min. The temperature within the TEGs was measured at six locations, marked by squares. Both TEGs exhibited a temperature gradient from bottom to top. Despite the heater temperature being 100 °C, the temperature at the bottom of the TEGs was <40 °C. This was due to the absence of direct contact between the TEG and heater, where the TEG heating occurred via radiation and convection rather than heat transfer. The temperature difference between the bottom and top was 2.3 K for the TEG with composite yarns and 3.2 K for the TEG with SWCNT films. Even after 10 min of heating, the temperature differences created within the two TEGs hardly changed from their respective values after 5 min. The temperature difference was 2.3 K for the TEG with composite yarns and 3.1 K for the TEG with SWCNT films, as shown in Figure 5e,f. This indicated that the two TEGs reached equilibrium. Therefore, the temperature difference for the TEG with the composite yarn was approximately 75% of that generated by the TEG with SWCNT films. This phenomenon cannot be explained by the thermal conductivities of the composite yarn and SWCNT film: the composite yarn is expected to have lower thermal conductivity than the SWCNT film, as the thermal conductivities of the SWCNT film, PEDOT:PSS, and SA are 5.4, 0.3, and 0.6 W/(m·K), respectively [15,79,80]. One possible explanation is that the SWCNT film had a larger specific surface area owing to its porous structure, while the composite yarn had a smaller specific surface area because of its dense structure. The larger surface area enhances heat dissipation, resulting in a significant temperature gradient [81].
Figure 6 shows the output voltages of the TEGs under heating at 100 °C. In Figure 6a, the TEG with two composite yarns exhibits an average output voltage of 0.21 mV over the period of 400–800 s. By looking at the generated voltage for the TEGs and the specified temperature gradient, the Seebeck coefficient can be calculated. However, considering the number of TEGs, the calculated values do not match the values measured with ZEM-3. We believe the main cause of the mismatch is the temperature measurement taken with the thermographic camera. The camera displays the average temperature over an area larger than the CNT yarn. In Figure 6b, the TEG with two SWCNT films exhibits an average output voltage of 0.61 mV. Thus, for TEGs with the same number of yarns and films, the output voltage of the TEG with the SWCNT films was approximately three times that of the TEG with yarns. This is because the Seebeck coefficient of the SWCNT film exceeded that of the composite yarn, and the temperature difference of the TEG with the SWCNT film exceeded that of the TEG with the composite yarn. As shown in Figure 6c, the TEG with six composite yarns exhibited an average output voltage of 0.69 mV. Comparing TEGs with the same material volume of yarns and films revealed that the TEG with composite yarns had a higher output voltage than the TEG with SWCNT films. Therefore, although the TEG with composite yarns had a small temperature difference and a low Seebeck coefficient, increasing the yarn density could improve its output voltage. The results confirm the fabrication of TEGs with composite yarns that combine high strength and performance, which can be used as autonomous power supplies for wearable sensors. To incorporate CNT yarns into wearable devices, their thermoelectric properties must be further enhanced after treatments [82]. In addition, CNT yarns must be smaller in diameter while retaining their flexibility and strength.

4. Conclusions

To obtain SWCNT-TEGs with high mechanical strength and flexibility, we prepared composite yarns consisting of SWCNTs, PEDOT:PSS, and SA via wet spinning. For comparison, SWCNT films were prepared via vacuum filtration. The fabricated composite yarns were approximately 290 µm in diameter and exhibited flexibility, smooth surfaces, and dense structures. Tensile tests revealed that the tensile strength and breaking strain of the composite yarns were 151 MPa and 12.7%, respectively, which were approximately 10 and 4 times those of the SWCNT film. In contrast, the electrical conductivity and Seebeck coefficient of the composite yarns were lower than those of the SWCNT films because of the inferior thermoelectric properties of PEDOT:PSS. TEGs were fabricated by aligning multiple composite yarns on a polyester fabric. When the TEG with two composite yarns was heated at 100 °C, it generated an output voltage of 0.21 mV, which was lower than that of the TEG with two SWCNT films. This is because the composite yarn had a low Seebeck coefficient, and the TEG with the composite yarn created a smaller temperature gradient. However, because the composite yarns had a small diameter, the TEGs achieved high-density alignment. For the TEG with composite yarns having high-density alignment, the output voltage increased to 0.67 mV, which was higher than that of the TEG with SWCNT films when the volume of materials used was matched with the TEGs with composite yarns and SWCNT films. These findings are beneficial for the development of SWCNT-based TEGs, which have high mechanical strength and flexibility and can be used as autonomous power supplies for wearable sensors.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/s25196202/s1, Figure S1: Manufacturing process of the SWCNT films. Figure S2: Photo of the composite yarn in a bent state. Reference [72] is cited in the supplementary materials.

Author Contributions

Conceptualization, M.T.; methodology, K.U., Y.S., H.N., S.O. and Y.N.; investigation, K.U., Y.S., H.N., S.O. and Y.N.; writing—original draft, K.U., H.N., S.O., Y.N. and M.T.; supervision, M.T.; project administration, M.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

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.

Acknowledgments

All authors thank Zeon Corporation for providing the SG-CNT powders and M. Morikawa and Y. Oda at Tokai University for their experimental support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Nižetić, S.; Šolić, P.; López-de-Ipiña, D.; González-de-Artaza, D.; Patrono, L. Internet of Things (IoT): Opportunities, issues and challenges towards a smart and sustainable future. J. Clean. Prod. 2020, 274, 122877. [Google Scholar] [CrossRef] [PubMed]
  2. Ullo, S.L.; Sinha, G.R. Advances in smart environment monitoring systems using IoT and sensors. Sensors 2020, 20, 3113. [Google Scholar] [CrossRef]
  3. Mukhopadhyay, C.S.; Suryadevara, K.N.; Nag, A. Wearable sensors and systems in the IoT. Sensors 2021, 21, 7880. [Google Scholar] [CrossRef]
  4. Dian, F.J.; Reza, V.; Alireza, R. Wearables and the internet of things (IoT), applications, opportunities, and challenges: A survey. IEEE Access 2020, 8, 69200–69211. [Google Scholar] [CrossRef]
  5. Ciuti, G.; Ricotti, L.; Menciassi, A.; Dario, P. MEMS sensor technologies for human centred applications in healthcare, physical activities, safety and environmental sensing: A review on research activities in Italy. Sensors 2015, 15, 6441–6468. [Google Scholar] [CrossRef] [PubMed]
  6. Sun, W.; Guo, Z.; Yang, Z.; Wu, Y.; Lan, W.; Liao, Y.; Wu, X.; Liu, Y. A review of recent advances in vital signals monitoring of sports and health via flexible wearable sensors. Sensors 2022, 22, 7784. [Google Scholar] [CrossRef]
  7. Jiang, Q.; Antwi-Afari, M.F.; Fadaie, S.; Mi, H.; Anwer, S.; Liu, J. Self-powered wearable internet of things sensors for human-machine interfaces: A systematic literature review and science mapping analysis. Nano Energy 2024, 131, 110252. [Google Scholar] [CrossRef]
  8. Rong, G.; Zheng, Y.; Sawan, M. Energy solutions for wearable sensors: A review. Sensors 2021, 21, 3806. [Google Scholar] [CrossRef]
  9. Gao, M.; Wang, P.; Jiang, L.; Wang, B.; Yao, Y.; Liu, S.; Chu, D.; Cheng, W.; Lu, Y. Power generation for wearable systems. Energy Environ. Sci. 2021, 14, 2114–2157. [Google Scholar] [CrossRef]
  10. Chong, Y.W.; Ismail, W.; Ko, K.; Lee, C.Y. Energy harvesting for wearable devices: A review. IEEE Sens. J. 2019, 19, 9047–9062. [Google Scholar] [CrossRef]
  11. Xu, C.; Song, Y.; Han, M.; Zhang, H. Portable and wearable self-powered systems based on emerging energy harvesting technology. Microsyst. Nanoeng. 2021, 7, 25. [Google Scholar] [CrossRef]
  12. Dargusch, M.; Liu, W.D.; Chen, Z.G. Thermoelectric generators: Alternative power supply for wearable electrocardiographic systems. Adv. Sci. 2020, 7, 2001362. [Google Scholar] [CrossRef]
  13. Elahi, H.; Munir, K.; Eugeni, M.; Atek, S.; Gaudenzi, P. Energy harvesting towards self-powered IoT devices. Energies 2020, 13, 5528. [Google Scholar] [CrossRef]
  14. Wang, Z.L. Energy harvesting for self-powered nanosystems. Nano Res. 2008, 1, 1–8. [Google Scholar] [CrossRef]
  15. Chiba, T.; Amma, Y.; Takashiri, M. Heat source free water floating carbon nanotube thermoelectric generators. Sci. Rep. 2021, 11, 14707. [Google Scholar] [CrossRef]
  16. Yanfua, J.; 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, 2102990. [Google Scholar]
  17. Shi, Y.; Wang, Y.; Deng, Y.; Gao, H.; Lin, Z.; Zhu, W.; Ye, H. A novel self-powered wireless temperature sensor based on thermoelectric generators. Energy Convers. Manag. 2014, 80, 110–116. [Google Scholar] [CrossRef]
  18. Nakajima, T.; Hoshino, K.; Yamamoto, H.; Kaneko, K.; Okano, Y.; Takashiri, M. Stretchable and flexible painted thermoelectric generators on Japanese paper using inks dispersed with p- and n-type single-walled carbon nanotubes. Sensors 2024, 24, 2946. [Google Scholar] [CrossRef] [PubMed]
  19. Yamamoto, H.; Amezawa, T.; Okano, Y.; Hoshino, K.; Ochiai, S.; Sunaga, K.; Miyake, S.; Takashiri, M. High thermal durability and thermoelectric performance with ultra-low thermal conductivity in n-type single-walled carbon nanotube films by controlling dopant concentration with cationic surfactant. Appl. Phys. Lett. 2025, 126, 063902. [Google Scholar] [CrossRef]
  20. Liu, W.; Fang, Y.; Lyu, X.; Peng, X.; Zhong-Zhen Luo, Z.-Z.; Zou, Z. Construction and application of thermogalvanic hydrogels. Soft Sci. 2024, 4, 44. [Google Scholar] [CrossRef]
  21. Nakazawa, Y.; Ochiai, S.; Okano, Y.; Okutsu, R.; Amma, Y.; Takashiri, M. Enhancing heat-source free water-floating single-walled carbon nanotube thermoelectric generators with water-absorbing paper. J. Power Sources 2025, 642, 236966. [Google Scholar] [CrossRef]
  22. Kim, M.-K.; Kim, M.-S.; Lee, S.; Kim, C.; Kim, Y.-J. Wearable thermoelectric generator for harvesting human body heat energy. Smart Mater. Struct. 2014, 23, 105002. [Google Scholar] [CrossRef]
  23. 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]
  24. 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]
  25. 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]
  26. Sultana, A.; Alam, M.M.; Middya, T.R.; Mandal, D. A pyroelectric generator as a self-powered temperature sensor for sustainable thermal energy harvesting from waste heat and human body heat. Appl. Energy 2018, 221, 299–307. [Google Scholar] [CrossRef]
  27. Norimasa, O.; Tamai, R.; Nakayama, H.; Shinozaki, Y.; Takashiri, M. Self-generated temperature gradient under uniform heating in pin junction carbon nanotube thermoelectric generators. Sci. Rep. 2025, 15, 15956. [Google Scholar] [CrossRef] [PubMed]
  28. Yang, S.; Qiu, P.; Chen, L.; Shi, X. Recent developments in flexible thermoelectric devices. Small Sci. 2021, 1, 2100005. [Google Scholar] [CrossRef]
  29. Khan, S.; Kim, J.; Acharya, S.; Kim, W. Review on the operation of wearable sensors through body heat harvesting based on thermoelectric devices. Appl. Phys. Lett. 2021, 118, 200501. [Google Scholar] [CrossRef]
  30. Shi, X.L.; Chen, W.Y.; Zhang, T.; Zou, J.; Chen, Z.G. Fiber-based thermoelectrics for solid, portable, and wearable electronics. Energy Environ. Sci. 2021, 14, 729–764. [Google Scholar] [CrossRef]
  31. He, X.; Shi, J.; Hao, Y.; He, M.; Cai, J.; Qin, X.; Wang, L.; Yu, J. Highly stretchable, durable, and breathable thermoelectric fabrics for human body energy harvesting and sensing. Carbon Energy 2022, 4, 621–632. [Google Scholar] [CrossRef]
  32. Norimasa, O.; Chiba, T.; Hase, M.; Komori, T.; Takashiri, M. Improvement of thermoelectric properties of flexible Bi2Te3 thin films in bent states during sputtering deposition and post-thermal annealing. J. Alloys Compd. 2022, 898, 162889. [Google Scholar] [CrossRef]
  33. Iijima, S.; Ichihashi, T. Single-shell carbon nanotubes of 1-nm diameter. Nature 1993, 363, 603–605. [Google Scholar] [CrossRef]
  34. Blackburn, J.L.; Ferguson, A.J.; Cho, C.; Grunlan, J.C. Carbon-nanotube-based thermoelectric materials and devices. Adv. Mater. 2018, 30, 1704386. [Google Scholar] [CrossRef] [PubMed]
  35. Xiao, R.; Zhou, X.; Zhang, C.; Liu, X.; Han, S.; Che, C. Organic thermoelectric materials for wearable electronic devices. Sensors 2024, 24, 4600. [Google Scholar] [CrossRef] [PubMed]
  36. Chen, G.; Xu, W.; Zhu, D. Recent advances in organic polymer thermoelectric composites. J. Mater. Chem. C. 2017, 5, 4350–4360. [Google Scholar] [CrossRef]
  37. Wang, X.; Wang, H.; Liu, B. Carbon nanotube-based organic thermoelectric materials for energy harvesting. Polymers 2018, 10, 1196. [Google Scholar] [CrossRef]
  38. Chiba, T.; Yabuki, H.; Takashiri, M. High thermoelectric performance of flexible nanocomposite films based on Bi2Te3 nanoplates and carbon nanotubes selected using ultracentrifugation. Sci. Rep. 2023, 13, 3010. [Google Scholar] [CrossRef]
  39. Seki, Y.; Takashiri, M. Freestanding bilayers of drop-cast single-walled carbon nanotubes and electropolymerized poly(3,4-ethylenedioxythiophene) for thermoelectric energy harvesting. Org. Electron. 2018, 76, 105478. [Google Scholar] [CrossRef]
  40. Valentini, L.; Armentano, I.; Lozzi, L.; Santucci, S.; Kenny, J.M. Interaction of methane with carbon nanotube thin films: Role of defects and oxygen adsorption. Mater. Sci. Eng. C 2004, 4, 527–533. [Google Scholar] [CrossRef]
  41. Kang, D.; Park, N.; Ko, J.; Bae, E.; Park, W. Oxygen-induced p-type doping of a long individual single-walled carbon nanotube. Nanotechnology 2005, 16, 1048. [Google Scholar] [CrossRef]
  42. Wang, Y.; Chen, Z.; Huang, H.; Wang, D.; Liu, D.; Wang, L. Organic radical compound and carbon nanotube composites with enhanced electrical conductivity towards high-performance p-type and n-type thermoelectric materials. J. Mater. Chem. A 2020, 8, 24675–24684. [Google Scholar] [CrossRef]
  43. Yonezawa, S.; Chiba, T.; Seki, Y.; Takashiri, M. Origin of n-type properties in single wall carbon nanotube films with anionic surfactants investigated by experimental and theoretical analyses. Sci. Rep. 2021, 11, 5758. [Google Scholar] [CrossRef] [PubMed]
  44. Seki, Y.; Nagata, K.; Takashiri, M. Facile preparation of air-stable n-type thermoelectric single-wall carbon nanotube films with anionic surfactants. Sci. Rep. 2020, 10, 8104. [Google Scholar] [CrossRef] [PubMed]
  45. Cheng, X.; Wang, X.; Chen, G. A convenient and highly-tunable way to n-type carbon nanotube thermoelectric composite film using common alkylammonium cationic surfactant. J. Mater. Chem. A 2018, 6, 19030–19037. [Google Scholar] [CrossRef]
  46. Brownlie, L.; Shapter, J. Advances in carbon nanotube n-type doping: Methods, analysis and applications. Carbon 2018, 126, 257–270. [Google Scholar] [CrossRef]
  47. Khongthong, J.; Raginov, I.N.; Khabushev, M.E.; Goldt, E.A.; Kondrashov, A.V.; Russakov, M.D.; Shandakov, D.S.; Krasnikov, V.D.; Nasibulin, G.A. Aerosol doping of SWCNT films with p- and n-type dopants for optimizing thermoelectric performance. Carbon 2024, 218, 118670. [Google Scholar] [CrossRef]
  48. Amma, Y.; Miura, K.; Nagata, S.; Nishi, T.; Miyake, S.; Miyazaki, K.; Takashiri, M. Ultra-long air-stability of n-type carbon nanotube films with low thermal conductivity and all-carbon thermoelectric generators. Sci. Rep. 2022, 12, 21603. [Google Scholar] [CrossRef]
  49. Yamamoto, H.; Okano, Y.; Uchida, K.; Kageshima, M.; Kuzumaki, T.; Miyake, S.; Takashiri, M. Phonon transports in single-walled carbon nanotube films with different structures determined by tensile tests and thermal conductivity measurements. Carbon Trends 2024, 17, 100435. [Google Scholar] [CrossRef]
  50. Miao, M. Yarn spun from carbon nanotube forests: Production, structure, properties and applications. Particuology 2013, 11, 378–393. [Google Scholar] [CrossRef]
  51. Dariyal, P.; Arya, K.A.; Singh, P.B.; Dhakate, R.S. A review on conducting carbon nanotube fibers spun via direct spinning technique. J. Mater. Sci. 2021, 56, 1087–1115. [Google Scholar] [CrossRef]
  52. Chen, G.; Davis, C.R.; Futaba, N.D.; Sakurai, S.; Kobashi, K.; Yumura, M.; Hata, K. A sweet spot for highly efficient growth of vertically aligned single-walled carbon nanotube forests enabling their unique structures and properties. Nanoscale 2016, 8, 162–171. [Google Scholar] [CrossRef]
  53. Nakamura, M.; Yamashita, I. Thermoelectric cloths using carbon nanotube yarn for wearable electronics. Jpn. J. Appl. Phys. 2024, 63, 010803. [Google Scholar] [CrossRef]
  54. Zheng, Y.; Zhang, Q.; Jin, W.; Jing, Y.; Chen, X.; Han, X.; Bao, Q.; Liu, Y.; Wang, X.; Wang, S.; et al. Carbon nanotube yarn based thermoelectric textiles for harvesting thermal energy and powering electronics. J. Mater. Chem. A 2020, 8, 2984–2994. [Google Scholar] [CrossRef]
  55. Tzounis, L.; Gravalidis, C.; Vassiliadou, S.; Logothetidis, S. Fiber yarns/CNT hierarchical structures as thermoelectric generators. Mater. Today Proc. 2017, 4, 7070–7075. [Google Scholar] [CrossRef]
  56. Ali, K.M.; Okamoto, N.; Abe, R.; Pandey, M.; Moneim, A.A.; Nakamura, M. Gas phase doping of pre-fabricated CNT yarns for enhanced thermoelectric properties. Synth. Met. 2021, 280, 116874. [Google Scholar] [CrossRef]
  57. Culebras, M.; Ren, G.; O’Connell, S.; Vilatela, J.J.; Collins, N.M. Lignin doped carbon nanotube yarns for improved thermoelectric efficiency. Adv. Sustain. Syst. 2020, 4, 2000147. [Google Scholar] [CrossRef]
  58. Liu, L.; Chen, J.; Liang, L.; Deng, L.; Chen, G. A PEDOT:PSS thermoelectric fiber generator. Nano Energy 2022, 102, 107678. [Google Scholar] [CrossRef]
  59. Kim, J.-Y.; Lee, W.; Kang, Y.H.; Cho, S.Y.; Jang, K.-S. Wet-spinning and post-treatment of CNT/PEDOT:PSS composites for use in organic fiber-based thermoelectric generators. Carbon 2018, 133, 293–299. [Google Scholar] [CrossRef]
  60. Sarabia-Riquelme, R.; Noble, L.E.; Espejo, P.A.; Ke, Z.; Graham, K.R.; Mei, J.; Paterson, A.F.; Weisenberger, M.C. Highly conductive n-type polymer fibers from the wet-spinning of n-doped PBDF and their application in thermoelectric textiles. Adv. Funct. Mater. 2023, 34, 2311379. [Google Scholar] [CrossRef]
  61. Pope, J.; Lekakou, C. Thermoelectric polymer composite yarns and an energy harvesting wearable textile. Smart Mater. Struct. 2019, 28, 095006. [Google Scholar] [CrossRef]
  62. Kim, Y.; Lund, A.; Noh, H.; Hofmann, A.I.; Craighero, M.; Darabi, S.; Zokaei, S.; Park, J.I.; Yoon, M.-H.; Müller, C. Robust PEDOT:PSS wet-spun fibers for thermoelectric textiles. Macromol. Mater. Eng. 2020, 305, 1900749. [Google Scholar] [CrossRef]
  63. Yang, X.; Zhang, K. Direct wet-spun single-walled carbon nanotubes-based p-n segmented filaments toward wearable thermoelectric textiles. ACS Appl. Mater. Interfaces 2022, 14, 44704–44712. [Google Scholar] [CrossRef]
  64. Wen, N.; Zeo, X.; Zhou, J.; Sun, C.; Chen, C.; Jiang, D.; Xu, H.; Wang, W.; Pan, L.; Fan, Z. Boosting thermoelectric performance of wet-spun PEDOT:PSS-based organic/inorganic composite fibers via a dual-interfacial engineering approach. Small 2025, 21, 2500866. [Google Scholar] [CrossRef] [PubMed]
  65. Zhang, C.; Liu, Y.; Li, H.; Liu, S.; Li, P.; Zhang, H.; He, C. Enhanced thermoelectric properties of carbon nanotubes/polyaniline fibers through engineering doping level and orientation. Compos. Sci. Technol. 2024, 253, 110660. [Google Scholar] [CrossRef]
  66. Zhang, C.; Zhang, Q.; Zhang, D.; Wang, M.; Bo, Y.; Fan, X.; Li, F.; Liang, J.; Huang, Y.; Ma, R.; et al. Highly stretchable carbon nanotubes/polymer thermoelectric fibers. Nano Lett. 2021, 21, 1047–1055. [Google Scholar] [CrossRef]
  67. Zhang, M.; Cao, X.; Wen, M.; Chen, C.; Wen, Q.; Fu, Q.; Deng, H. Highly electrical conductive PEDOT:PSS/SWCNT flexible thermoelectric films fabricated by a high-velocity non-solvent turbulent secondary doping approach. ACS Appl. Mater. Interfaces 2023, 15, 10947–10957. [Google Scholar] [CrossRef]
  68. Li, H.; Liu, Y.; Liu, S.; Li, P.; Zhang, C.; He, C. Wet-spun flexible carbon nanotubes/polyaniline fibers for wearable thermoelectric energy harvesting. Compos. Part A 2023, 166, 107386. [Google Scholar] [CrossRef]
  69. Kim, J.-Y.; Mo, J.-H.; Kang, Y.H.; Cho, S.Y.; Jang, K.-S. Thermoelectric fibers from well-dispersed carbon nanotube/poly(vinylidene fluoride) pastes for fiber-based thermoelectric generators. Nanoscale 2018, 10, 19766–19773. [Google Scholar] [CrossRef] [PubMed]
  70. Xu, C.; Yang, S.; Li, P.; Wang, H.; Li, H.; Liu, Z. Wet-spun PEDOT:PSS/CNT composite fibers for wearable thermoelectric energy harvesting. Compos. Commun. 2022, 32, 101179. [Google Scholar] [CrossRef]
  71. Liu, L.; Guo, X.; Liu, W.; Lee, C. Recent progress in the energy harvesting technology—From self-powered sensors to self-sustained IoT, and new applications. Nanomaterials 2021, 11, 2975. [Google Scholar] [CrossRef] [PubMed]
  72. Okano, Y.; Ochiai, S.; Nakayama, H.; Nagai, K.; Takashiri, M. Optimization of Ultrasonic Dispersion of Single-Walled SWCNT Inks for Improvement of Thermoelectric Performance in SWCNT Films Using Heat Source-Free Water-Floating SWCNT Thermoelectric Generators. Materials 2025, 18, 3339. [Google Scholar] [CrossRef]
  73. Yang, Y.; Deng, H.; Fu, Q. Recent progress on PEDOT:PSS based polymer blends and composites for flexible electronics and thermoelectric devices. Mater. Chem. Front. 2020, 4, 3130–3152. [Google Scholar] [CrossRef]
  74. Yemata, A.T.; Zheng, Y.; Kyaw, K.K.A.; Wang, X.; Song, J.; Chin, S.W.; Xu, J. Modulation of the doping level of PEDOT:PSS film by treatment with hydrazine to improve the Seebeck coefficient. RSC Adv. 2020, 10, 1786–1792. [Google Scholar] [CrossRef]
  75. Wu, W.; Li, Y.; Yang, L.; Ma, Y.; Yan, X. Preparation and characterization of coaxial multiwalled carbon nanotubes/polyaniline tubular nanocomposites for electrochemical energy storage in the presence of sodium alginate. Synth. Met. 2014, 193, 48–57. [Google Scholar] [CrossRef]
  76. Wu, T.; Shi, X.-L.; Liu, W.-D.; Li, M.; Yue, F.; Huang, P.; Liu, Q.; Chen, Z.-G. High thermoelectric performance and flexibility in rationally treated PEDOT:PSS fiber bundles. Adv. Fiber Mater. 2024, 6, 607–618. [Google Scholar] [CrossRef]
  77. Behabtu, N.; Young, C.C.; Tsentalovich, D.E.; Kleinerman, O.; Wang, X.; Ma, A.W.K.; Bengio, E.A.; Ter Waarbeek, R.F.; De Jong, J.J.; Hoogerwerf, R.E.; et al. Strong, light, multifunctional fibers of carbon nanotubes with ultrahigh conductivity. Science 2013, 339, 182–186. [Google Scholar] [CrossRef] [PubMed]
  78. Zhao, X.; Kong, D.; Tao, J.; Kong, N.; Mota-Santiago, P.; Lynch, P.A.; Shao, Y.; Zhang, J. Wet twisting in spinning for rapid and cost-effective fabrication of superior carbon nanotube yarns. Adv. Funct. Mater. 2024, 34, 2400197. [Google Scholar] [CrossRef]
  79. Liu, J.; Wang, X.; Coates, N.E.; Segalman, R.A.; Cahill, D.G. Thermal conductivity and elastic constants of PEDOT:PSS with high electrical conductivity. Macromolecules 2015, 48, 585–591. [Google Scholar] [CrossRef]
  80. Hatami, M.; Ganji, D.D. Natural convection of sodium alginate (SA) non-Newtonian nanofluid flow between two vertical flat plates by analytical and numerical methods. Case Stud. Therm. Eng. 2014, 2, 14–22. [Google Scholar] [CrossRef]
  81. Miura, K.; Amezawa, T.; Tanaka, S.; Takashiri, M. Improved heat dissipation of dip-coated SWCNT/mesh sheets with high flexibility and free-standing strength for thermoelectric generators. Coatings 2024, 14, 126. [Google Scholar] [CrossRef]
  82. Zeng, H.; Gao, C.; Yu, Y.; Jiang, M.; Deng, T.; Zhu, J. Wet spinning enabled advanced PEDOT:PSS composite fibers for smart devices. Acc. Mater. Res. 2025, 6, 952–963. [Google Scholar] [CrossRef]
Figure 1. Manufacturing process for the SWCNT/PEDOT:PSS/SA composite yarns.
Figure 1. Manufacturing process for the SWCNT/PEDOT:PSS/SA composite yarns.
Sensors 25 06202 g001
Figure 2. Microstructure and surface morphology of the composite yarn and SWCNT films. Low-magnification images of the (a) composite yarn and (b) SWCNT film; high-magnification images of the (c) composite yarn and (d) SWCNT film.
Figure 2. Microstructure and surface morphology of the composite yarn and SWCNT films. Low-magnification images of the (a) composite yarn and (b) SWCNT film; high-magnification images of the (c) composite yarn and (d) SWCNT film.
Sensors 25 06202 g002
Figure 3. Stress–strain curves of the composite yarn and SWCNT film.
Figure 3. Stress–strain curves of the composite yarn and SWCNT film.
Sensors 25 06202 g003
Figure 4. (a) Manufacturing process for the TEG with composite yarns; (b) photographs of completed TEGs; (c) measurement procedure for the TEG; (d) photographs of the measurement procedure.
Figure 4. (a) Manufacturing process for the TEG with composite yarns; (b) photographs of completed TEGs; (c) measurement procedure for the TEG; (d) photographs of the measurement procedure.
Sensors 25 06202 g004
Figure 5. Photographs of the (a) TEG with composite yarns and (b) TEG with SWCNT films. Thermography images of the (c) TEG with composite yarns after 5 min of heating, (d) TEG with SWCNT films after 5 min of heating, (e) TEG with composite yarns after 10 min of heating, and (f) TEG with SWCNT films after 10 min of heating.
Figure 5. Photographs of the (a) TEG with composite yarns and (b) TEG with SWCNT films. Thermography images of the (c) TEG with composite yarns after 5 min of heating, (d) TEG with SWCNT films after 5 min of heating, (e) TEG with composite yarns after 10 min of heating, and (f) TEG with SWCNT films after 10 min of heating.
Sensors 25 06202 g005
Figure 6. Time dependence of the output voltage for TEGs heated at 100 °C: (a) TEG with two composite yarns; (b) TEG with two SWCNT films; (c) TEG with six composite yarns.
Figure 6. Time dependence of the output voltage for TEGs heated at 100 °C: (a) TEG with two composite yarns; (b) TEG with two SWCNT films; (c) TEG with six composite yarns.
Sensors 25 06202 g006
Table 1. Mechanical properties of the composite yarn and reference SWCNT film.
Table 1. Mechanical properties of the composite yarn and reference SWCNT film.
SampleTensile StrengthBreaking StrainYoung’s ModulusMass DensityLongitudinal Sound Velocity
Composite yarn151 MPa12.7%0.26 GPa1.36 g/cm3437 m/s
SWCNT film16 MPa3.6%0.70 GPa0.42 g/cm31289 m/s
Table 2. Thermoelectric properties of the composite yarn and reference SWCNT film.
Table 2. Thermoelectric properties of the composite yarn and reference SWCNT film.
SampleSeebeck CoefficientElectrical ConductivityPower Factor
Composite yarn29.4 μV/K15.3 S/cm1.3 μW/(m·K2)
SWCNT film55.5 μV/K29.0 S/cm8.9 μW/(m·K2)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Uchida, K.; Shinozaki, Y.; Nakayama, H.; Ochiai, S.; Nakazawa, Y.; Takashiri, M. SWCNT/PEDOT:PSS/SA Composite Yarns with High Mechanical Strength and Flexibility via Wet Spinning for Thermoelectric Applications. Sensors 2025, 25, 6202. https://doi.org/10.3390/s25196202

AMA Style

Uchida K, Shinozaki Y, Nakayama H, Ochiai S, Nakazawa Y, Takashiri M. SWCNT/PEDOT:PSS/SA Composite Yarns with High Mechanical Strength and Flexibility via Wet Spinning for Thermoelectric Applications. Sensors. 2025; 25(19):6202. https://doi.org/10.3390/s25196202

Chicago/Turabian Style

Uchida, Keisuke, Yoshiyuki Shinozaki, Hiroto Nakayama, Shuya Ochiai, Yuto Nakazawa, and Masayuki Takashiri. 2025. "SWCNT/PEDOT:PSS/SA Composite Yarns with High Mechanical Strength and Flexibility via Wet Spinning for Thermoelectric Applications" Sensors 25, no. 19: 6202. https://doi.org/10.3390/s25196202

APA Style

Uchida, K., Shinozaki, Y., Nakayama, H., Ochiai, S., Nakazawa, Y., & Takashiri, M. (2025). SWCNT/PEDOT:PSS/SA Composite Yarns with High Mechanical Strength and Flexibility via Wet Spinning for Thermoelectric Applications. Sensors, 25(19), 6202. https://doi.org/10.3390/s25196202

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