Jewook
Kim†
a,
Hwanseok
Chang†
a,
Gwangmin
Bae
a,
Myungwoo
Choi
*b and
Seokwoo
Jeon
*a
aDepartment of Materials Science and Engineering, Korea University, Seoul 02841, Republic of Korea. E-mail: jeon39@korea.ac.kr
bNano Hybrid Technology Research Center, Korea Electrotechnology Research Institute (KERI), Changwon, 51543, Republic of Korea. E-mail: mwchoi@keri.re.kr
First published on 10th March 2025
Among sustainable energy-harvesting systems, thermoelectric technology has attracted considerable attention because of its ability to directly convert heat into electricity and diverse applications. Graphene, with its exceptional electrical conductivity and mechanical properties, is a promising candidate for thermoelectric materials. However, efficient thermoelectric applications require materials with a high Seebeck coefficient and low thermal conductivity—criteria that graphene does not inherently satisfy, owing to its gapless energy band structure and ballistic thermal conduction. This review examines the thermoelectric properties of graphene, optimization strategies, and the potential of graphene hybridization for thermoelectric applications. To overcome the intrinsic limitations of graphene for thermoelectric utilization, nanostructuring strategies based on its synthesis methods are discussed. Furthermore, strategies for graphene hybridization are introduced, with a focus on maximizing thermoelectric efficiency through interactions with nanostructured materials of various dimensions. Finally, the potential of graphene-based thermoelectric materials and future research directions are discussed.
Since its discovery in 2004, graphene has attracted considerable research interest owing to its high electrical conductivity,37 carrier mobility,38 mechanical strength,39 and flexibility.40 These outstanding electrical properties have led to its primary applications in energy storage systems—particularly as an anode material for lithium-ion batteries.41 Additionally, its large surface area and tunable surface chemistry have enabled its use as an absorbent material for environmental remediation.42 Beyond these primary applications, graphene has emerged as a promising candidate for thermoelectric materials, driven by its superior electrical properties, mechanical robustness, environmental sustainability, and cost-effectiveness. However, its intrinsic limitations, including a gapless band structure that suppresses the Seebeck coefficient and extremely high thermal conductivity arising from the long mean free path (MFP) of phonons, restrict its energy-conversion efficiency.13,31 Furthermore, the thermoelectric performance of graphene can vary significantly depending on the synthesis method,11,43 presenting additional challenges for optimization. To overcome these limitations, strategies such as nanostructuring,44,45 doping,46,47 grain-boundary tuning,31–33,48 and defect engineering13,49 have been explored to decouple and optimize the thermoelectric parameters of graphene. Despite significant research progress, the practical utilization of graphene in thermoelectric systems remains limited. These limitations arise from difficulties in establishing a temperature gradient across dimensions—particularly in the finite out-of-plane direction—due to its two-dimensional (2D) structure with a single atomic layer.50 Additionally, graphene exhibits a low thermoelectric conversion efficiency (5.42 × 10−3 at 350 K),13 even with the implementation of nanostructuring strategies.
As an alternative, a promising approach to harness the superior properties of graphene for thermoelectric applications is its selective use as a filler in hybrid composite systems. Hybridization with graphene allows the thermoelectric parameters to be decoupled by leveraging its ability to increase electrical conductivity,51,52 reduce thermal conductivity via phonon scattering,12,53 and induce energy-filtering effects.54 The enhancement of thermoelectric performance through graphene hybridization can be optimized by initially nanostructuring the thermoelectric material to increase its specific surface area, which facilitates graphene integration and interaction at the nanoscale. Notably, graphene hybridization with nanostructured thermoelectric materials has achieved significant improvements in zT values. For example, researchers improved the thermoelectric performance by incorporating partially reduced graphene oxide (rGO) into nanostructured TiO2.55 In this study, rGO flakes were strategically distributed near grain boundaries and defect sites within the TiO2 matrix, resulting in a synergistic effect. The highly conductive pathways provided by rGO, while activating phonon scattering at heterointerfaces, significantly reduced the thermal conductivity, causing a 3-fold improvement in the zT value at 325 K. This dual effect demonstrates the potential of graphene-based hybrid systems to overcome intrinsic limitations of conventional thermoelectric materials and achieve higher energy-conversion efficiencies.
While previous reviews56,57 have explored the thermoelectric properties of graphene and graphene-based composites, a focused analysis of the role of graphene in nanostructured hybrid thermoelectric systems remains absent. This review aims to fill that gap by analyzing theoretical and experimental findings related to the thermoelectric properties of graphene, including its intrinsic limitations and optimization strategies. The review further explores the integration of graphene with one-dimensional (1D), 2D, and three-dimensional (3D) nanostructured thermoelectric materials, clarifying the underlying mechanisms that maximize the filler effect of graphene, such as efficient carrier transport, phonon scattering, and energy filtering (Fig. 1). Finally, we discuss the practical potential and future directions of graphene-based hybrid thermoelectric materials for energy-harvesting applications.
![]() | ||
Fig. 1 Synthesis methods and hybridization strategies for utilizing graphene as a thermoelectric material. |
Nanostructuring strategies have been widely utilized to modify the transport properties of electrons and phonons for enhancing the thermoelectric performance of conventional materials. These approaches have been extended to graphene, and researchers have reported significant theoretical and experimental improvements in its thermoelectric properties. For example, Nam et al. fabricated bilayer-graphene without nanostructuring to compare its experimentally measured Seebeck coefficient and thermoelectric power with calculated values based on the Mott relation.70 Although the bandgap opening near the CNP under a back-gate voltage was expected to result in a Seebeck coefficient close to the theoretical prediction, the experimental values were significantly lower. This discrepancy was attributed to large fluctuations in the disorder potential, which likely hindered the formation of a bandgap. In contrast, nanostructured graphene has exhibited significant improvements in thermoelectric performance. H. Sevincli et al. investigated graphene nanoribbons (GNRs) with zigzag edges and calculated their thermoelectric parameters.44 The zigzag-edged graphene nanoribbons (ZGNRs) exhibited a high zT value of nearly 4, primarily owing to the suppressed phonon conductance at zigzag-edge defects. Additionally, the level of disorder within the ribbon was identified as a critical parameter for enhancing the thermoelectric performance of ZGNRs. These findings highlight the potential of nanostructuring for optimizing the thermoelectric properties of graphene—particularly through phonon-transport modulation.
To address the intrinsically low power factor of graphene, researchers have employed substrate engineering. Duan et al. constructed devices featuring single-layer graphene flakes on a hexagonal boron nitride (hBN) substrate (Fig. 2(a)).71 This configuration significantly enhanced the thermoelectric performance compared with devices on SiO2 substrates—particularly for active-cooling applications. By using the hBN substrate, the surface charge states and amount of impurities originating from the SiO2 substrate were reduced, increasing the carrier mobility and reducing the number of electron–hole puddles, which increased the Seebeck coefficient. Fig. 2(b) shows the gate-voltage (Vg) dependence of the Seebeck coefficient in the temperature range of 77–290 K. When Vg was applied to modulate the carrier concentration in graphene, the device exhibited a positive Seebeck coefficient in the hole-doped regime and a negative Seebeck coefficient in the electron-doped regime, with the highest Seebeck coefficient observed near the CNP. This led to a maximum power factor times temperature value of 10.35 W m−1 K−1 at room temperature, which exceeds those of other 2D materials and even some conventional thermoelectric materials.
![]() | ||
Fig. 2 (a) Schematic of graphene on an hBN device. (b) Back-gate dependence of S in the temperature range of 77–290 K. Reproduced with permission from ref. 71. |
While the theoretical enhancement of thermoelectric performance through the nanostructuring of graphene and the fabrication of passive devices via substrate engineering have been demonstrated, the practical utilization of graphene in sustainable thermoelectric energy-harvesting systems requires further investigation. Specifically, it is crucial to explore the applicability of nanostructured graphene in energy-conversion devices, with experimental validation of its thermoelectric performance.
![]() | ||
Fig. 3 (a) Test structure for the thermoelectric measurements of the GNR grown via CVD. (b) σ/k and ke/k ratios of the suspended GNR compared with reference data. (c) zT value of the suspended GNR compared with reference data. Reproduced with permission from ref. 45. (d) Scanning electron microscopy (SEM) images of different nucleation densities in the initial growth state depending on the synthesis temperature and pressure. (e) Thermal conductivity as a function of temperature for different grain sizes. Reproduced with permission from ref. 31. (f) zT value of the graphene with the optimized grain size. Reproduced with permission from ref. 48. (g) Growth mechanism of MHW-CVD graphene. (h) Concept of increasing the graphene zT value in the MHW-CVD system. (i) Comparison between the Seebeck coefficients of M-CVD and C-CVD graphene. (j) Comparison between the thermal conductivities of M-CVD and C-CVD graphene. (k) zT values of M-CVD and C-CVD graphene compared with reference data. Reproduced with permission from ref. 13. |
Another strategy for enhancing phonon scattering in graphene is grain-size reduction. By controlling the conditions during the CVD process for graphene synthesis, such as the operating pressure, temperature, heating-/cooling-time intervals, and precursor-gas concentration ratios, the average grain size of graphene is systematically controlled (Fig. 3(d)).31 As shown in Fig. 3(e), as the average grain size is reduced, the thermal conductivity tends to decrease, which results from the narrowing of the phonon dispersive range, limiting the contribution of long-wavelength phonons. For graphene with small grain sizes of 4.1 and 2.2 μm, the grain-size reduction narrows the phonon dispersive range, restricting the contribution of long-wavelength phonons and reducing the thermal conductivity. When the grain size is reduced to 500 nm, it becomes smaller than the phonon MFP of graphene, further reducing the thermal conductivity (340 W m−1 K−1). A subsequent study indicated that this reduction in grain size is more effective than reducing the power factor for reducing the thermal conductivity (Fig. 3(f)).48 Compared with a grain size of 4.1 μm, reductions to 2.2 and 0.5 μm resulted in electron MFP reductions of 23.3% and 49.8%, respectively, while the phonon MFPs decreased more substantially—by 24.5% and 68%, respectively. Accordingly, the highest thermoelectric zT value was achieved at a grain size of 0.5 μm, which was attributed to the effective phonon-boundary scattering. This indicates that controlling the grain size during CVD synthesis can increase the zT value.
Recent advancements in CVD have allowed more precise control over defects in graphene during its growth process. In particular, the mobile hot-wire CVD (MHW-CVD) system offers localized and sequential thermal treatment, allowing more refined control over graphene crystal growth compared with conventional CVD systems, which supply thermal energy globally.13,38,73–75 This system enables the synthesis of highly oriented graphene with only two crystal orientations, separated by high-angle tilt boundaries (HATBs) with a misorientation angle of approximately 30°. These HATBs significantly affect the thermoelectric properties of graphene.13 The strategy for enhancing the thermoelectric performance of graphene involves the formation of HATB domains that induce both energy filtering and phonon scattering. In crystallographically misaligned graphene, a charge-carrier transport gap (Eg) is formed at grain boundaries. As the misalignment angle increases, Eg increases. Additionally, phonon scattering is enhanced in regions with a higher degree of misalignment owing to disordered structures at the interfaces between neighboring graphene grains. Normally, graphene exhibits a zero bandgap and a low Seebeck coefficient, as observed for C-CVD graphene (conventional CVD-grown graphene) in Fig. 3(i). However, in M-CVD (MHW-CVD) graphene, the substantial Eg created at the HATB domains provides an energy-filtering function that selectively blocks low-energy carriers corresponding to Eg. Thus, M-CVD graphene is expected to have a relatively high Seebeck coefficient due to the selective transport of charge carriers. The Seebeck coefficient of M-CVD graphene is 50.1 μV K−1 at 350 K, representing a 138% increase compared with that of C-CVD graphene. As shown in Fig. 3(j), HATB domains also reduce the thermal conductivity of graphene. The thermal conductivity of M-CVD graphene at 350 K is approximately 382 W m−1 K−1, which is significantly lower than that of C-CVD graphene, indicating an increase in phonon scattering at the HATB domains. Fig. 3(k) shows the thermoelectric zT values for M-CVD and C-CVD graphene, which are derived from the experimental power factor and thermal conductivity data, along with other graphene-based thermoelectric materials. M-CVD graphene exhibits a maximum zT value of 3.78 × 10−3 at 350 K, which is nearly 7 times higher than that of C-CVD graphene. This approach can be used to enhance thermoelectric properties by simultaneously tuning grain boundaries and inducing nanostructuring effects through modifications in the CVD growth parameters, without the need for specific post-growth nanostructuring processes.
In top-down methods, minimizing defects during the exfoliation process is crucial for preserving the electrical properties of graphene flakes when they are assembled into films.76 The most common technique for producing graphene flakes is the modified Hummers’ method, which generates graphene oxide (GO).77Fig. 4(a) presents a schematic of the standard process for fabricating GO, which is followed by a post-treatment step to reduce GO into rGO. However, residual oxygenous groups and defects remain in the basal plane of rGO flakes even after the intense chemical process at an extremely high temperature (Fig. 4(b)).78 These oxidation-induced defects significantly degrade the thermoelectric performance of graphene, necessitating strategies to mitigate them for thermoelectric applications. Hu et al.79 systematically investigated the impact of the reduction temperature on GO synthesized via Hummers’ method, varying the reduction temperature from 500 to 3300 K.79 Their study revealed that as the reduction temperature increased, the degree of oxidation decreased, leading to the restoration of crystallinity and a substantial increase in the thermoelectric power factor, which reached ∼54.5 μW cm−1 K−2 at 3300 K. Raman spectroscopy supported these findings: a notable reduction in the D-to-G peak intensity ratio (ID/IG) from 1.11 to 0.08 was observed when the reduction temperature was increased from 1000 to 3300 K. Additionally, in contrast to GO reduced at 1000 K, the sample reduced at 3300 K exhibited a sharp 2D peak, indicating a more graphitic structure. The restoration of crystallinity at 3300 K resulted in a significant increase in its electrical conductivity, which reached approximately 4000 S cm−1 at 3000 K. Moreover, the increased Seebeck coefficient (∼150 μV K−1 at 1200 K) was attributed to the suppression of low-energy hole transport and controlled electron scattering. Furthermore, addressing oxidation-related defects is critical owing to their impact on the doping behavior of graphene. As illustrated in Fig. 4(c), the doping type of graphene can be tuned through annealing and selective functional group removal. However, rGO typically exhibits weak p-type thermoelectric behavior at room temperature.80 To address the fundamental challenges associated with the intrinsic properties of rGO, it is essential to develop fabrication methods that preserve the high structural quality while enabling scalable production of graphene flakes. Consequently, the need for low-oxidation graphene from top-down methods has drawn attention to synthesis techniques based on graphite intercalation compounds (GICs).81 Among the studies using GIC synthesis, flexible thermoelectric films produced from non-oxidized graphene flakes (NOGFs) were introduced. Fig. 4(d) presents schematics of the NOGF fabrication process, which contains before and after stage of expansion through intercalation. The resultant graphene was obtained via simple immersion method of graphite in potassium naphthalenide and tetrahydrofuran (THF) followed by mild sonication to form high-quality, large-scale flakes (Fig. 4(e)). The conductive film composed of NOGFs exhibits exceptional thermoelectric performance, achieving a room-temperature power factor of 673 μW m−1 K−2 (Fig. 4(f)).10 This impressive result is attributed to its high electrical conductivity of 3280 S cm−1 and Seebeck coefficient of −45.3 μV K−1, both of which originate from the minimal oxidation of the basal plane. Additionally, the films exhibit excellent mechanical stability, with <3% reduction in electrical conductivity at a bending radius of 1 mm and stability after >1000 bending cycles.
![]() | ||
Fig. 4 (a) Schematic of the conventional rGO fabrication process. (b) Transmission electron microscopy (TEM) image of rGO. Reproduced with permission from ref. 78. (c) Seebeck coefficient of graphene as the function of the annealing temperature. Reproduced with permission from ref. 80. (d) Schematic of the NOGF fabrication process. (e) TEM image of graphene. (f) Electrical conductivity (σ), Seebeck coefficient (S), and power factor of the NOGF-based flexible thin films. Reproduced with permission from ref. 10. (g) Schematic of the experimental process for fabricating complementary p- and n-type graphene films by varying the surfactant during exfoliation. (h) Electrical conductivity of NOGFs as a function of the temperature. (i) Power output as a function of the potential at different temperatures. Reproduced with permission from ref. 11. |
In a follow-up study, the separation of doping types in NOGFs was investigated by varying the surfactants used during processing. Fig. 4(g) presents the experimental approach, which was based on previous research (Fig. 4(d)), and highlights the two distinct doping types observed in the resulting TE devices.11 The process begins with the synthesis of a GIC by immersing raw graphite in a potassium naphthalenide solution, followed by gentle bath sonication for 30 min in DMSO with either polyvinylpyrrolidone (PVP) or pyrenebutyric acid (PBA) as a surfactant. By changing the adsorbed surfactant during the intercalation–exfoliation process (PVP for n-type, PBA for p-type), both extremely high electrical conductivity and high Seebeck coefficients of complementary n-type and p-type graphene can be achieved. Fig. 4(h) compares the in-plane thermoelectric properties of PVP-NOGF and PBA-NOGF films. Near room temperature (313 K), the PBA-NOGF film exhibits lower electrical conductivity (2330 S cm−1) than the PVP-NOGF film (3010 S cm−1), but both have relatively stable electrical conductivity across the temperature range. The power factor of the PBA-NOGF film (655 μW m−1 K−2) is similar to that of the PVP-NOGF film (621 μW m−1 K−2) under the same conditions. The power as a function of potential is presented in Fig. 4(i), showing the expected parabolic shape, with the height increasing as the temperature increases. The peak power output at 250 °C was 5.0 nW at 2.7 mV. The results of the study deviate from commonly used synthesis methods; rather, the researchers focused on enhancing the intrinsic thermoelectric properties of graphene. Additionally, distinct doping of graphene was achieved by adjusting the film's process variables, without additional post-treatment.
![]() | ||
Fig. 5 Hybridization of graphene with 1D materials for enhancing the thermoelectric performance. (a) SEM images of Si nanowire–graphitic shell composites. (b) Current–voltage (I–V) curves of Si/SiO2 nanowire–graphitic shell composites. (c) Dependence of the thermal conductivity on the thickness of the graphitic shell. Reproduced with permission from ref. 87. (d) TEM images of DTe nanowires and rGO hybridization. (e) Comparison of the Seebeck coefficient between GDTe nanowires and rGO. (f) Comparison of power factors between single constituents and hybrids. Reproduced with permission from ref. 88. |
Furthermore, Choi et al. utilized a ternary composite consisting of graphene, a polymer (PEDOT:PSS), and Te nanowires (TeNWs) to improve thermoelectric performance (Fig. 5(d)).88 The hybridization of graphene with TeNWs increased both the electrical conductivity and the Seebeck coefficient. The high charge-carrier mobility of graphene, combined with the flexibility of the polymer, optimized the electrical conduction pathways and efficiently filtered out low-energy charge carriers, allowing only high-energy carriers to contribute to electrical transport. In this study, a double carrier-filtering effect was induced at the heterojunctions, enhancing the thermoelectric properties without significantly increasing the thermal conductivity. The multiple interfaces formed by the combination of graphene, TeNWs, and PEDOT:PSS functioned as energy barriers for low-energy charge carriers, facilitating the selective transport of high-energy carriers. This selective filtering mechanism led to concurrent improvements in electrical conductivity and the Seebeck coefficient (Fig. 5(e)), increasing the power factor (Fig. 5(f)). These findings emphasize the significance of optimizing heterojunction characteristics to balance electrical conductivity and thermal conductivity in thermoelectric materials. The ability to precisely control carrier-transport properties through heterostructure design offers a promising route to further increase zT values.
Oh et al. fabricated a heterojunction network that combined GNRs with monolayer molybdenum disulfide (MoS2) to improve the thermoelectric performance (Fig. 6(a)).91 The heterojunction was designed to leverage the distinctive properties of each component: GNRs, known for their high electrical conductivity and band structure modulation capabilities, were paired with MoS2, which has semiconducting properties92 and a direct bandgap of approximately 1.8 eV. The GNR/MoS2 heterojunction network structure provided significant advantages for thermoelectric performance owing to its unique ability to form vertical carrier paths through a van der Waals interface. This structure allowed charge carriers to traverse between GNRs and MoS2 through the heterojunction, maximizing the interface's contribution to the overall electrical and thermoelectric properties. GNRs functioned as efficient conductive channels, while the MoS2 layers increased the Seebeck coefficient by providing additional bandgap tuning and acting as energy barriers for selective carrier transport. The introduction of a 1-nm-thick layer of benzyl viologen (BV) at the heterojunction further improved carrier transport by serving as a thin tunneling barrier and modulating the Fermi level of MoS2. By controlling the linewidth of the GNRs, the researchers demonstrated that it was possible to optimize the band structure and consequently the thermoelectric performance of the heterojunction network (Fig. 6(b)). GNRs with narrower linewidths exhibited higher quantum confinement, leading to an increased bandgap. This change increased the Seebeck coefficient and reduced the thermal conductivity. The GNR/MoS2 heterojunction network achieved an electrical conductivity of up to 1163 S cm−1 (Fig. 6(c)) and a power factor of 222 μW m−1 K−2 (Fig. 6(d)), representing significant improvements over pristine MoS2. The careful engineering of interfacial properties was key to the success of the heterojunction. By introducing the BV layer and tuning the GNR linewidth, the interfacial resistance was reduced, and band-to-band tunneling effects were enhanced, leading to more efficient charge transfer across the interface. This approach allowed optimization of the power factor without a corresponding increase in thermal conductivity, which is critical for achieving a high thermoelectric zT. Moreover, the multiple interfaces and the introduced energy-filtering effects were crucial in improving both the electrical conductivity and the Seebeck coefficient while keeping the thermal conductivity low Gautam et al. reported that the hybridization of MoS2 with rGO (partially reduced graphene oxide) significantly improved the thermoelectric performance.93 The primary challenge with MoS2 as a thermoelectric material lies in its low electrical conductivity94 despite its intrinsically low thermal conductivity. By forming a nanocomposite with rGO, they aimed to improve the thermoelectric zT by increasing the electrical conductivity, increasing the Seebeck coefficient, and reducing the thermal conductivity (Fig. 6(e)). The hybridization of rGO within the MoS2 matrix provided multiple conductive pathways, reducing the interface barrier height and enhancing the scattering of phonons. This reduced the thermal conductivity while creating more effective conductive channels for electron transport, increasing the electrical conductivity. The synergistic interaction between MoS2 and rGO was facilitated by their intimate contact, accelerating electron transfer across the nanocomposite. The presence of rGO effectively optimized the electron-transport properties of the MoS2 matrix, resulting in a higher electrical conductivity, which was approximately 171% higher than that of pristine MoS2 at room temperature (Fig. 6(f)). Moreover, the Seebeck coefficient of the MoS2/rGO nanocomposite was significantly increased (Fig. 6(g)). This was partly due to the reduction of interface barriers, which allowed more efficient transfer of high-energy charge carriers, increasing the thermoelectric potential. Additionally, the thermal conductivity was reduced through increased phonon scattering at the heterogeneous interfaces between MoS2 and rGO. The lattice part of the thermal conductivity was reduced by 61%, contributing to the overall improvement in thermoelectric performance (Fig. 6(h)). These combined improvements led to an impressive 60-fold enhancement in the thermoelectric zT of the composite compared with pristine MoS2.
![]() | ||
Fig. 6 Enhancing phonon scattering and energy filtering in the hybridization of graphene with 2D materials. (a) Schematic of the GNR–MoS2 heterojunction. (b) SEM images of GNR patterns with varying linewidths. (c) Electrical conductivity and (d) power factor of graphene hybridized with MoS2. Reproduced with permission from ref. 91. (e) TEM images of MoS2–RGO hybridization. Comparison of the (f) electrical conductivity, (g) Seebeck coefficient, and (h) Thermal conductivity between single MoS2 and hybridized MoS2–RGO. Reproduced with permission from ref. 93. |
Tang et al. explored this approach by incorporating graphene into a copper sulfide (Cu2−xS) matrix, which led to improved thermoelectric properties and thermal stability.84 The integration was performed through a combination of mechanical alloying and spark plasma sintering, creating a graphene–Cu2−xS composite with a 3D graphene network (Fig. 7(a)). The inclusion of graphene had multiple beneficial effects. First, the graphene network contributed to Fermi-level pinning at the Cu2−xS interfaces (Fig. 7(b)), ensuring that only high-energy carriers contributed to electrical transport, which significantly increased the electrical conductivity of the composite (Fig. 7(c)). This energy-filtering effect allowed high-energy carriers to dominate electrical transport while scattering low-energy carriers, which increased the Seebeck coefficient. This selective carrier transport significantly increased the composite's power factor, which reached 1197 μW m−1 K−2 at 873 K.
![]() | ||
Fig. 7 Hybridization of graphene with 3D bulk materials for enhancing the thermoelectric performance. (a) Schematic of the fabrication process for graphene–Cu2−xS. (b) Energy filtering in the depleted layer. (c) Seebeck coefficient as a function of the graphene concentration. (d) TEM images of graphene–Cu2−xS and the grain boundaries. (e) Thermal conductivity as a function of the graphene concentration. (f) Highest zT value of graphene–Cu2−xS. Reproduced with permission from ref. 84. |
Additionally, the interfacial engineering introduced by the graphene network enhanced phonon scattering, significantly reducing the thermal conductivity of the composite (Fig. 7(d)). The presence of graphene at the grain boundaries and within the copper sulfide lattice disrupted the continuity of the lattice and introduced additional scattering centers for phonons, reducing the lattice thermal conductivity. The porous 3D graphene structure also suppressed grain growth during sintering, leading to a high density of grain boundaries, which further contributed to phonon scattering. This resulted in a thermal conductivity as low as 0.67 W m−1 K−1 at 873 K, which was significantly lower than that of the pristine Cu2−xS (Fig. 7(e)). Overall, the high density of grain boundaries resulting from the hybridization of graphene with 3D Cu2−xS enhanced the energy filtering and phonon scattering, significantly increasing the thermoelectric zT, which reached 1.56 at 873 K—a value considerably higher than that of pristine Cu2−xS (Fig. 7(f)).
Graphene hybridization may be more effective when combined with the rational design of 3D nanostructured thermoelectric materials.95,96 Kim et al. fabricated highly ordered and large-scale periodic 3D ZnO nanostructures via proximity-field nanopatterning (PnP), achieving a practical zT (∼0.017 at 333 K and ∼0.024 at 423 K) for pure ZnO. This performance was attributed to the significant reduction in thermal conductivity due to extensive phonon scattering at the numerous periodic nanostructured interfaces.9 PnP—a 3D interference lithography-based nanofabrication method—has potential for the scalable production of nanostructured thermoelectric materials for device applications.97–105 Furthermore, structural design for thermoelectric materials can enhance the thermoelectric performance without the need for doping. Integrating graphene into these optimized 3D nanostructures can further improve the thermoelectric performance.12 For example, Choi et al. significantly improved the thermoelectric properties of ZnO through 3D nanostructuring and graphene hybridization.12 The 3D continuous, highly ordered porous ZnO matrix provided abundant active sites for the attachment of graphene nanoflakes, which were called GQDs. High-quality GQDs were synthesized from GICs in a process that involved minimal oxidation (Fig. 8(a)).106,107 The integration of low-oxidation GQDs into a 3D ZnO nanostructure led to significant enhancements in several key thermoelectric factors, such as the Seebeck coefficient, electrical conductivity, and thermal stability. The 5-nm GQD, with a distinct energy bandgap (∼3.2 eV) (Fig. 8(b)), exhibited a larger bandgap than ZnO (∼3 eV).108 This difference suggests the potential formation of an interfacial energy barrier, depending on the valence-band maximum of ZnO and the highest occupied molecular orbital of the GQD. Specifically, the interfacing of GQDs with the porous ZnO matrix resulted in the formation of well-controlled interfacial energy barriers approximately 0.63 eV in height, which functioned as electron filters. These energy barriers selectively allowed only high-energy electrons to contribute to electrical conduction, while scattering lower-energy carriers (Fig. 8(c)). This selective electron filtering contributed to a substantial improvement of 74% in the power factor compared with the pristine ZnO matrix, indicating the effectiveness of GQDs for enhancing thermoelectric performance (Fig. 8(f)). Moreover, the numerous heterogeneous interfaces between the GQDs and the ZnO matrix played a crucial role in inducing extensive phonon scattering (Fig. 8(d)), which significantly reduced the thermal conductivity from 1.49 to 0.785 W m−1 K−1 (Fig. 8(g)). This reduction in thermal conductivity, combined with the improved electrical properties, contributed to an increased thermoelectric zT for the composite. The zT reached 0.486 at 580 K, setting a new benchmark for ZnO-based thermoelectric materials (Fig. 8(h)).
![]() | ||
Fig. 8 Maximizing thermoelectric-performance enhancement through GQD interfacing in the nanostructured 3D matrix. (a) Schematic of the GQD fabrication process. (b) Bandgap differences between GQDs, rROQDs, and GOQDs. Reproduced with permission from ref. 106. (c) Field-emission SEM image and schematic of GQD-interfaced 3D ZnO (3D GQD@ZnO). (d) Sketch of phonon-scattering and energy-filtering effects at grain boundaries and GQD/ZnO interfaces. (e) TEM images of 3D GQD@ZnO (inset: high-resolution TEM image of 3D GQD@ZnO). (f)–(h) show the Seebeck coefficient, thermal conductivity, and zT of the 3D GQD@ZnO. Reproduced with permission from ref. 12. |
This research highlights the significant potential of utilizing nanosized graphene additives such as GQDs to fine-tune the electrical and thermal properties of thermoelectric materials. The combination of GQDs with a porous oxide matrix proved to be a viable strategy for improving the thermoelectric figure of merit by leveraging interface engineering and nanoscale material properties. The development of such hybrid materials not only provides a promising approach to address the limitations of traditional thermoelectric systems but also supports the advancement of sustainable energy-harvesting technologies capable of converting waste heat into useful electrical energy.
A promising strategy to overcome these limitations lies in the hybridization of graphene with thermoelectric matrices. Graphene, when employed as a filler, offers a unique advantage in decoupling the three critical thermoelectric parameters—electrical conductivity, Seebeck coefficient, and thermal conductivity—by facilitating charge-carrier transport, phonon scattering, and energy filtering. For example, 2D transition-metal dichalcogenides such as MoS2,93 Cu2Te,109 and PbSe110 can effectively modulate electron and phonon transport when integrated with graphene, enhancing thermoelectric performance. Additionally, polymer-based hybridization with materials such as PEDOT:PSS111 and PANI112 offers a lightweight, flexible, and cost-effective approach for fabricating scalable thermoelectric composites. Finally, the integration of graphene with oxide-based thermoelectric materials such as ZnO,12 Al-doped ZnO51 and lanthanum strontium titanium oxide,113 which are highly compatible with semiconductor processing, can significantly improve thermoelectric performance while maintaining practical manufacturability. The synergistic hybridization effect is further amplified when graphene is integrated with nanostructured or porous thermoelectric matrices. These structures increase the specific surface area, creating active sites for graphene interaction and improving electrical pathways while suppressing thermal conductivity through enhanced phonon scattering at the heterointerfaces. Fig. 9 shows the peak zT values in various temperature ranges before and after graphene hybridization.12,51,52,55,113–119 Graphene hybridization not only improves the performance of existing thermoelectric materials but also induces hybridization effects across the entire temperature range owing to the high thermal stability of graphene. This observation highlights the versatility of graphene as a thermoelectric filler capable of enhancing performance under both low- and high-temperature conditions. Furthermore, the integration of graphene into thermoelectric devices addresses many limitations of traditional thermoelectric materials and systems. Hybrid graphene-based thermoelectric materials exploiting the properties of graphene, such as non-toxicity, abundance, and cost-efficiency, play a pivotal role in sustainably advancing the development of energy-harvesting systems across diverse applications. Such advancements have significant implications for the development of eco-friendly and sustainable energy-harvesting technologies.
![]() | ||
Fig. 9 Peak ZT values of pristine materials and graphene composites: (1) Bi85Sb15/graphene,114 (2) SrTi0.9Nb0.1O3/rGO,115 (3) Bi2Te3/rGO,116 (4) ZnO/GQD,12 (5) CoSb3/graphene,117 (6) La0.067Sr0.9TiO3/rGO,113 (7) SrTiO3/rGO,52 (8) Sr0.93Ti0.90Nb0.10O3/rGO,118 (9) Nb–Tio2/GO,55 (10) ZnO/SLG,119 and (11) Zn0.98Al0.02O/GO.51 |
To accelerate the commercialization of graphene-based thermoelectric materials, several critical challenges must be addressed. First, the exploration of mass-production technologies for high-quality graphene using low-cost and scalable methods is essential.120 Methods such as CVD and solution-based exfoliation should be refined to produce graphene with controllable thickness and quality at an industrial scale. Second, future research should focus on developing robust and scalable synthesis processes for hybrid thermoelectric materials that effectively incorporate graphene while maintaining structural and functional integrity. In addition, further investigation into the mechanisms underlying graphene hybridization effects, such as phonon–electron interactions, energy filtering phenomena, and interfacial thermal resistance, will provide valuable insights for optimizing material design. Multiscale modeling and advanced characterization techniques should be employed to elucidate the structure–property–performance relationships of graphene-based thermoelectric materials. Another priority is the exploration of thermoelectric device technology tailored for graphene-based hybrid materials, e.g., designing and optimizing device architectures to maximize the energy-conversion efficiency of graphene-integrated systems.
In conclusion, graphene-based thermoelectric materials offer immense potential to revolutionize sustainable energy solutions. With the advancement of mass-production technologies, improvements in hybrid material design, and the development of scalable thermoelectric devices, graphene can play a pivotal role in next-generation energy-harvesting systems. Such efforts will be crucial in bridging the gap between fundamental research and real-world deployment, enabling the widespread use of graphene-based thermoelectric technologies in sustainable energy systems and beyond.
Footnote |
† These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |