Keito Uchida‡
a,
Shunya Sakane‡*b,
Takashi Shimizua,
Akito Ayukawab,
Haruhiko Udono
b and
Hideki Tanaka
*a
aFaculty of Science and Engineering, Chuo University, 1-13-27, Kasuga, Bunkyo-ku, Tokyo, 112-8551, Japan. E-mail: htanaka@kc.chuo-u.ac.jp
bGraduate School of Science and Engineering, Ibaraki University, 4-12-1, Nakanarusawa-cho, Hitachi, Ibaraki 316-8511, Japan. E-mail: shunya.sakane.sz12@vc.ibaraki.ac.jp
First published on 5th June 2025
Graphene-based materials are anticipated to be used as thermoelectric conversion devices due to their flexibility and low toxicity, in addition to their high thermoelectric performance. In this study, we demonstrated an enhancement in the thermoelectric power factor of graphene thin films with agglomerated PEDOT:PSS through π–π interactions. The graphene/PEDOT:PSS thin films were prepared by a spin-coating method. Atomic force microscopy, X-ray diffraction and Raman spectroscopy revealed that PEDOT:PSS agglomerated on graphene thin films through π–π interactions. The fabricated sample exhibited a 1.6 times higher power factor compared to graphene single-phase thin films. The local π–π interactions with PEDOT:PSS contribute to the electron transfer from graphene to PEDOT and the enhanced crystallinity of graphene throughout the thin film, resulting in a high power factor. This study contributes to the development of graphene-based thermoelectric materials.
On the other hand, carbon-based materials such as diamond, carbon nanotubes, and graphene attract attention because they are composed of carbon, which is abundant in nature, and have excellent mechanical properties that make them widely applicable as wearable materials.18–24 In particular, graphene, with its two-dimensional structure of carbon atoms arranged in a honeycomb lattice, has two fundamental unit vectors in reciprocal space with its unit cell. The two points at the corners of the Brillouin zone, corresponding to these unit vectors, are known as Dirac points. Near these Dirac points, graphene exhibits a unique electronic structure called the Dirac cone,23,25–27 which contributes to its high mobility of 60000 cm2 V−1 S−1.28 In contrast, the Seebeck coefficient of graphene monolayers is extremely low because the Fermi level is located near Dirac points and both electrons and holes contribute to electrical conduction.29,30 However, controlling the Fermi level by applying an electric field to the graphene monolayer allows for the selective control of the carrier conduction type and a consequent improvement of S. As a result, high S2σ values of 360 μW cm−1 K−2 (p-type) and 190 μW cm−1 K−2 (n-type) have been reported for a micron-scale graphene monolayer.31–33 Graphene multilayers prepared by chemical vapor deposition (CVD) are capable of centimeter-scale thin film fabrication and have been reported to exhibit high S2σ values of 69.3 μW cm−1 K−2 (p-type) and 32.9 μW cm−1 K−2 (n-type).34
For practical applications, a high S2σ is required without the application of an electric field. In a previous report, graphene thin films can be prepared through spin-coating using graphene ink, which has also demonstrated a relatively high S of 41 μV K−1 without the application of an electric field, resulting in an S2σ of 0.187 μW cm−1 K−2 (p-type).35 To improve the thermoelectric properties of graphene thin films, compositing them with aromatic ring compounds such as poly(3,4-ethylenedioxythiophene):poly(styrene sulfonate) (PEDOT:PSS) is effective because it improves crystallinity and enables electron transfer through π–π interactions. In fact, the π–π interaction has been demonstrated in various graphene composites such as rGO and graphene quantum dots (GQDs)/PEDOT:PSS.36–42
In this study, graphene/PEDOT:PSS thin films were fabricated by a spin-coating method with the aim of obtaining high S2σ through π–π interactions. We experimentally demonstrated a high S2σ derived from π–π interactions and thoroughly investigated the S2σ enhancement mechanism. This study would contribute to the development of thermoelectric performance enhancement in graphene-based materials.
The structures of these thin films were evaluated by X-ray diffraction (XRD) using an X-ray diffractometer (Rigaku, SmartLab), scanning electron microscopy (SEM)–energy dispersive X-ray spectroscopy (EDX) using a field emission SEM (Hitachi High-Technologies, S-5500), atomic force microscopy (AFM; Agilent Technologies), and Raman spectroscopy (Lambda Vision, Micro-RAM300/CHK-532-785-Gr1). For Raman spectroscopy, a 532 nm (30 mW) light source and a 50× objective lens were used. The electronic states were evaluated by X-ray photoelectron spectroscopy (XPS) with the characteristic X-ray source of Al Kα radiation (JEOL, JPS-9010).
σ was calculated from the sheet resistance measured by the van der Pauw method and film thickness measured using the cross-sectional SEM images of cut samples. S (=ΔV/ΔT) was calculated from the voltage difference (ΔV) and temperature difference (ΔT) measured with a self-made system and corrected with ZEM-3 (ADVANCE RIKO). Carrier concentration and mobility were measured by Hall effect measurements. Local electrical conductivity was measured by a 4-probe method with a probe distance of 100 μm.
The inset of Fig. 2(a) shows a photograph of the thin film surface in the GP3 sample, showing an agglomerated PEDOT:PSS area and an exposed graphene region on the thin film. The area around the agglomerated PEDOT:PSS was observed by surface SEM (Fig. 2(a)), showing that the agglomerated PEDOT:PSS area was less uneven, while the exposed graphene region was more uneven. Elemental mapping by SEM–EDX revealed that the agglomerated PEDOT:PSS contained a large amount of S (Fig. 2(b)), which is attributed to PEDOT:PSS. From the above results, PEDOT:PSS agglomerated on graphene layers.
The structures of the agglomerated PEDOT:PSS region and the exposed graphene region mentioned above were analyzed in more detail. Fig. 2(c) and (d) show the Raman spectra of the PEDOT:PSS and graphene regions of the GP3 sample, respectively. In the PEDOT:PSS region of the GP3 sample (Fig. 2(c)), a typical Raman band at 1449 cm−1 corresponding to the CC symmetric stretching vibration of PEDOT was observed, where the peak position shifted to the higher wavenumber side compared to the peak position of 1447 cm−1 in P3 (Fig. S2, ESI†). This is due to the π–π interaction between graphene and PEDOT.36 Raman bands were also observed around 1520 cm−1 and 1600 cm−1 originating from the C
C asymmetric and antisymmetric stretching vibrations of PEDOT,43 respectively, which are also observed in P3. On the other hand, in the graphene region (Fig. 2(d)), three typical Raman bands were observed: a G band at approximately 1578 cm−1 originating from the in-plane vibrational motion of sp2 hybridized carbon, and D and 2D bands at around 1346 cm−1 and 2683 cm−1, respectively, associated with overtones from the stretching motion of sp3 hybridized carbon.44,45 These Raman bands were also observed at similar wavenumbers in G3 (Fig. S2, ESI†).
Fig. 2(e) and (f) show the AFM images of the PEDOT:PSS and graphene regions of the GP3 sample, respectively. The PEDOT:PSS region of the GP3 sample shows fibrous morphology, while P3 exhibits a random granular structure (Fig. S2, ESI†). This is because PEDOT:PSS typically has a coiled structure and random molecular arrangement, whereas due to the π–π interaction with graphene, PEDOT grows along the surface of graphene nanosheets, transforming the coiled structure into a linear one and its random molecular arrangement into an ordered arrangement.46 In the AFM images of both G3 and the graphene region of the GP3 sample, multiple sheet structures were observed.47 The RMS roughness of the graphene region of the GP3 sample was approximately 15 nm, which is comparable to that of G3 (∼20 nm).
Fig. 3(a) shows the XRD 2θ–ω scans of the GPn and G3 samples. For all samples, the glass substrate-derived peak at 21.7° and the 002 and 004 diffraction peaks of graphene at 26.4° and 54.5°, respectively, were observed. As shown in Fig. 3(a), the 002 diffraction peak intensity increased with the number of graphene spin-coating cycles. The plot of the 002 diffraction peak intensity, normalized by the glass substrate peak, against the number of graphene spin-coating cycles was approximated with a straight line through the origin (Fig. S3, ESI†). This indicates that the amount of graphene in the thin films increased proportionally with the number of spin coating cycles, which is consistent with the results shown in Fig. 1(b). Fig. 3(b) shows the full width at half maximum (FWHM) of the 002 diffraction peak of graphene. All GPn samples exhibited a reduced FWHM compared to G3. This suggests a smaller variation in the interlayer spacing of graphene, leading to high crystallinity, likely due to π–π interactions with PEDOT.44
![]() | ||
Fig. 3 (a) XRD 2θ–ω scans for each thin film and (b) the dependence of the FWHM of the 002 graphene peak on the number of graphene spin coating cycles (2n). |
Fig. 4 shows the S (2p) XPS spectra of the PEDOT:PSS region in GP3 and P3 samples. Two peaks were observed at 166–170 eV and 162–166 eV attributed to sulfur in PSS and PEDOT, respectively. Notably, the PSS-derived peak intensity was lower for the GP3 sample compared to P3, indicating a reduction in the amount of PSS present at or near the surface detectable by XPS. This reduction can be caused by a reorganization within the PEDOT:PSS complex facilitated by the π–π interactions between graphene and PEDOT. The PEDOT-derived peak (162–166 eV) was fitted with two mixed Gaussian/Lorentzian functions, revealing peaks centered at 163.6 eV (neutral PEDOT) and 164.6 eV (oxidized PEDOT).48 These fitting results are summarized in Table 1. The GP3 sample exhibited a higher ratio of the neutral PEDOT component compared to P3, potentially due to the electron transfer from graphene to PEDOT:PSS. Furthermore, the PEDOT-derived peaks showed a slight shift towards lower binding energies in the GP3 sample, indicating the reduction of PEDOT, which is consistent with previous reports.49 Additionally, the peak corresponding to sulfur atoms in ionic PSS (component 3) decreased for the GP3 sample compared with P3. This is likely because PSS interacts less strongly with the PEDOT chains due to stronger PEDOT–graphene interactions.
![]() | ||
Fig. 4 S (2p) XPS spectra of (a) GP3 and (b) P3 samples. The blue lines represent the sum of curves fitted with four components (dashed curves). |
Sample | Component | Center (eV) | Area ratio (%) |
---|---|---|---|
GP3 | 1 | 163.6 | 18.8 |
2 | 164.6 | 44.4 | |
3 | 167.5 | 2.6 | |
4 | 168.3 | 34.2 | |
P3 | 1 | 163.7 | 12.4 |
2 | 164.7 | 46.0 | |
3 | 167.8 | 18.9 | |
4 | 168.9 | 22.7 |
The S, σ and S2σ values of the fabricated thin films plotted against the number of graphene spin-coating cycles are shown in Fig. 5(a)–(c), respectively. Fig. 5(a) shows that the GPn samples exhibited S values similar to those of graphene, except for GP1. In contrast, the σ values of the GPn samples were higher than those of G3. Notably, the GP3 sample exhibited a σ value approximately 1.7 times higher than that of G3. Additionally, GPn samples also exhibited higher S and σ than those of the PEDOT:PSS single-phase thin film (S ≈ 4 μV K−1, σ ≈ 46 S cm−1). Consequently, the S2σ value increased by about 1.6 times compared with G3, as shown in Fig. 5(c). To investigate the origin of this high S2σ, carrier concentration (p) and Hall mobility (μ) were measured by Hall effect measurements. As shown in Fig. 5(d), the GP3 sample exhibited higher p in contrast to its lower Hall mobility when compared to G3. This observation can be attributed to the electron transfer from graphene to PEDOT within the GP3 sample. This electron transfer is facilitated by the strong π–π interactions between the graphene and the conjugated polymer of PEDOT.36,39 However, a drop in σ was observed for the GP4 sample as shown in Fig. 5(b), which is due to a reduced effective interaction ratio between the PEDOT:PSS and the graphene layers with the increase of the number of graphene layers.
![]() | ||
Fig. 5 The dependences of (a) S, (b) σ, (c) S2σ, and (d) p (left axis) and μ (right axis) on the number of graphene spin coating cycles (2n) for each thin film. |
To investigate the carrier transport in detail, we measured the temperature dependence of the electrical properties of GP3 and G3. As shown in Fig. 6(a), the σ of both samples slightly increased as the temperature increased, indicating their semiconductor-like properties. In some previous reports, the carrier transport mechanism in reduced graphene oxide has been discussed using the Mott-variable range hopping (VRH) model and thermal activation.50 In the case of the three-dimensional Mott-VRH model, σ exhibits a T dependence given by σ ∝ exp(−(T0/T)−1/4), where T0 is the Mott characteristic temperature. However, the VRH model does not provide a good fit to the σ of GP3 and G3 because our samples show a nonlinear T−1/4 dependence of the logarithm of σ, as shown in Fig. 6(b).
Fig. 6(c) shows the T dependence of p. The p was fitted with the following equation based on a simple two-band model of bulk graphite:35,51
![]() | (1) |
The Hall mobilities of GP3 and G3 as a function of T are shown in Fig. 6(d). The μ values of both GP3 and G3 decreased as T increased. Considering that the dependence is similar for both samples, GP3 likely exhibits similar carrier scattering mechanisms to G3. This decrease in dependence might be considered to arise from acoustic phonon scattering. Although, in general, the large contribution of acoustic phonon scattering tends to be proportional to T−3/2 (dashed line), the experimental value exhibits a weaker dependence than T−1/2 (dotted line). Therefore, the carrier scattering mechanism must be extremely complicated because it might contain some other carrier scattering processes such as those caused by impurities, defects, interfaces, etc. Based on the above discussion, the high S2σ of GP3 originating from high σ is attributed to the enhancement of p by doping from PEDOT due to the π–π interactions.
Despite the agglomeration of PEDOT:PSS on graphene, the abovementioned thermoelectric properties were measured over the entire region. To understand the relationship between the electrical properties and the local structures, it is essential to measure local electrical properties. Here, the local electrical conductivity (σL) of the GP3 sample was measured using a 4-probe method. Fig. 7(a) shows the σL at various positions in the graphene region of the GP3 sample; in the PEDOT:PSS region, the contact resistance was too high to measure σL. Dotted and dashed lines indicate the average values of GP3 and G3, respectively. The σL of the graphene region in the GP3 sample was obviously higher than that of G3 in most areas. This indicates that the σL is high not only for graphene in contact with PEDOT:PSS, but also for the entire thin film, including the graphene region spatially distant from PEDOT:PSS. These results suggest that the local π–π interactions with PEDOT:PSS contribute to the electron transfer from graphene to PEDOT and the enhanced crystallinity of graphene throughout the thin film, resulting in high S2σ (Fig. 7(b)).
![]() | ||
Fig. 7 (a) Local electrical conductivity (σL) in various graphene regions in GP3 and (b) schematic diagram of graphene doping from PEDOT:PSS via local π–π interactions. |
Finally, the S2σ of the GP3 sample fabricated in this study (1.2 μW cm−1 K2) was compared with those of previously reported PEDOT:PSS/graphene-based thin films (Table 2). The GP3 sample prepared in this study exhibited high S2σ comparable to a previous report (1.50 μW cm−1 K2) for PEDOT:PSS/rGO(20 wt%) prepared by the pad-dry-cure method.40 In this study, through the formation of agglomerated PEDOT:PSS, a relatively high S2σ was obtained despite the simplicity of the spin-coating method. Although the S2σ values of GPn samples were relatively low compared with other inorganic thermoelectric materials, we believe that our fabrication approach can contribute to the development of graphene-based thermoelectric materials.
Sample | σ (S cm−1) | S (μV K−1) | S2σ (μW cm−1 K−2) | Ref. |
---|---|---|---|---|
This work | 497 | 48 | 1.2 | — |
Graphene | 81 | 41 | 0.187 | 35 |
PEDOT:PSS/graphene (70 wt%) | 96 | 17 | 0.029 | 41 |
PEDOT:PSS/graphene (RTCVD) | 193 | 54 | 0.054 | 42 |
PEDOT:PSS/rGO (2 wt%) | 32 | 59 | 0.11 | 38 |
PEDOT:PSS/rGO (3 wt%) | 637 | 27 | 0.46 | 37 |
PEDOT:PSS/rGO (16 wt%) | 51 | 61 | 0.050 | 36 |
PEDOT:PSS/rGO (20 wt%) | 410 | 61 | 1.5 | 40 |
PEDOT:PSS/GQD (13 wt%) | 72 | 15 | 0.015 | 39 |
Footnotes |
† Electronic supplementary information (ESI) available: Cross-sectional SEM images of GP3, Raman spectra and AFM images of graphene and PEDOT:PSS single-phase thin films are included. In addition, the intensity of the 002 XRD patterns of GPn and G3 samples is shown. See DOI: https://doi.org/10.1039/d5ma00454c |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |