Next Article in Journal
Perfect Photon Indistinguishability from a Set of Dissipative Quantum Emitters
Next Article in Special Issue
Thermal Transport in 2D Materials
Previous Article in Journal
Green, Sustainable Synthesis of γ-Fe2O3/MWCNT/Ag Nano-Composites Using the Viscum album Leaf Extract and Waste Car Tire for Removal of Sulfamethazine and Bacteria from Wastewater Streams
Previous Article in Special Issue
Review of Photothermal Technique for Thermal Measurement of Micro-/Nanomaterials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Temperature Dependence of Thermal Conductivity of Giant-Scale Supported Monolayer Graphene

1
College of New Materials and New Energies, Shenzhen Technology University, Shenzhen 518116, China
2
School of Mechanical and Automotive Engineering, Shanghai University of Engineering Science, Shanghai 201620, China
3
College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen 518055, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(16), 2799; https://doi.org/10.3390/nano12162799
Submission received: 16 July 2022 / Revised: 9 August 2022 / Accepted: 11 August 2022 / Published: 15 August 2022
(This article belongs to the Special Issue Energy Transport at the Micro/Nanoscale)

Abstract

:
Past work has focused on the thermal properties of microscale/nanoscale suspended/supported graphene. However, for the thermal design of graphene-based devices, the thermal properties of giant-scale (~mm) graphene, which reflects the effect of grains, must also be investigated and are critical. In this work, the thermal conductivity variation with temperature of giant-scale chemical vapor decomposition (CVD) graphene supported by poly(methyl methacrylate) (PMMA) is characterized using the differential transient electrothermal technique (diff-TET). Compared to the commonly used optothermal Raman technique, diff-TET employs joule heating as the heating source, a situation under which the temperature difference between optical phonons and acoustic phonons is eased. The thermal conductivity of single-layer graphene (SLG) supported by PMMA was measured as 743 ± 167 W/(m·K) and 287 ± 63 W/(m·K) at 296 K and 125 K, respectively. As temperature decreased from 296 K to 275 K, the thermal conductivity of graphene was decreased by 36.5%, which can be partly explained by compressive strain buildup in graphene due to the thermal expansion mismatch.

1. Introduction

The discovery of graphene created explosive growth in research on the various properties of graphene [1,2,3,4]. Due to its high thermal conductivity (k) [5], ultrahigh electron mobility [6], and remarkable mechanical strength [7], graphene is regarded to have broad and promising applications including flexible electronics [8,9], photovoltaic devices [10], batteries [11], and so on [12]. With the increasing intense heat of electronics, graphene with super-high k has potential applications for the thermal management of microelectronics. A massive amount of work has focused on the experimental and theoretical calculation of the k of suspended and supported graphene, and great progress has been achieved [13,14,15,16,17]. The effect of various parameters such as system size [17], defects [18,19,20], substrates [21,22], and grain size [23,24,25] on the k of graphene has been deeply studied. Last but not least, since temperature affects the phonon propagation, the k of graphene shows strong temperature dependence. It has been reported that the k of supported graphene increases with increasing temperature, with the peak k appearing around 300 K [14]. When the temperature is lower than 300 K, the phonon-substrate scattering dominates the phonon scattering [14].
Thermal property measurement is critical to the application of graphene in thermal management. Since graphene is extremely thin, the commonly used thermal conductivity measurement techniques of graphene are optothermal Raman spectroscopy [26,27,28] and the micro-bridge method [14]. For optothermal Raman spectroscopy, the accuracy of the measured k depends on the absorbed laser power and the temperature coefficient of the Raman shift of the G peak [29]. Li et al. developed the laser-flash optothermal Raman technique, which eases the laser absorption coefficient uncertainty in the measurement of thermal conductivity [29,30]. It is assumed that optical phonons and acoustic phonons are in thermal equilibrium under intense laser irradiation. However, Ruan et al. first reported that the phonon branches were in strong thermal non-equilibrium by employing density functional perturbation theory. The temperature of out-of-plane acoustic phonons (ZA) is 14.8% lower than that of transverse optical phonons (TO) [31]. Under this situation, the k of graphene will be overestimated. Wang et al. first detected the lumped temperature difference between optical and acoustic phonons by nanosecond energy transport state-resolved Raman spectroscopy [32]. Seol et al. developed the micro-bridge method to measure the k of supported graphene. The micro-bridge method uses joule heating as the heating source. In the experiment, the graphene supported by SiO2 acts as the thermal bridge between the heating part and the temperature-sensing part [14]. In principle, the micro-bridge method is feasible, but its measurement accuracy is guaranteed when the thermal resistance of the sample is comparable to that of the resistance thermometer in the experiment setup. In addition, it is technically challenging and time consuming to fabricate the whole measurement device [33].
So far, most thermal conductivity measurements of graphene focus on microscale samples. It is reported that the k of suspended graphene is size/geometry dependent. The k of graphene has a ~log L (L: sample length) dependence which was confirmed by experiments (up to 9 μm) [17] and theoretical simulations [34]. Although the k of supported graphene is less sensitive to the length size due to substrate coupling [21,35], the k of macroscale supported graphene is critical, since the edge-phonon scattering usually leads to a great reduction in k. Furthermore, graphene has important applications in the field of flexible electronics, which calls for the thermal properties of graphene supported by flexible materials. To date, seldom work has been conducted on the temperature dependence of the thermal conductivity of millimeter-scale graphene on soft substrates, yet such knowledge is essential to graphene-based device design and optimization.
In this work, the temperature dependence of the thermal conductivity of millimeter-scale SLG supported by poly(methyl methacrylate) (PMMA) is characterized by a differential transient thermoelectrical technique (diff-TET) [36]. The diff-TET employs self-heating as the heating source, which ensures the thermal equilibrium among phonon branches. The sample size is approximately 2–3 mm in width and 1–3 mm in length, which significantly suppresses the effect of size on thermal conductivity. The k of PMMA is much lower than that of graphene. Choosing extremely thin (~457 nm) PMMA as the substrate can ensure the contribution of graphene to the overall thermal conductivity is comparable to that of PMMA. Thus, it is feasible to measure the k of graphene with high accuracy. The k of supported SLG is reported to be 743 ± 167 W/(m·K) at 296 K, which is 23.8% higher than that of micro-scale SiO2-supported graphene measured using the micro-bridge method [14]. The k of supported SLG is reduced to 287 ± 63 W/(m·K) when the temperature is 125 K, partly due to the strain-induced rips in the sample.

2. Sample Preparation and Characterization

2.1. Sample Preparation

SLG supported by PMMA characterized in this work was obtained from Jiangsu Xianfeng nanoscale materials company. The samples used in this experiment were fabricated using the chemical vapor deposition (CVD) method. The graphene is grown on a copper (Cu) in a controlled chamber pressure CVD system. At first, a clean Cu foil was annealed at 1077 °C with a H2 flow rate of 500 sccm. Then, the H2 flow rate and chamber pressure were adjusted to 70 sccm and 108 Torr, respectively. By introducing 0.15 sccm CH4 into the chamber, the graphene starts to grow. The copper was etched off after PMMA was coated on the graphene. Finally, graphene supported by PMMA was transferred to a filter substrate. The layer number of graphene was provided in the product technical data.
The as-received graphene was cut into the desired size by scissors. Then, the experimental sample was suspended between two electrodes, as shown in Figure 1a. The effective k of these samples will be measured by diff-TET without any further processing. The thickness of PMMA (δp) is necessary to determine the k of graphene when subtracting the effect of PMMA. Atomic force microscopy (AFM) was used to determine the thickness of PMMA. The supported graphene was transferred to a silicon substrate; then, the thickness of PMMA was determined with AFM. Figure 1b shows the AFM image of the as-received supported graphene. The thickness of PMMA was determined to be 457 nm.

2.2. Structure Study Based on Raman Spectroscopy

Even though the layer number of the as-received sample was given in the technical data sheet, the layer number of the as-received graphene on PMMA was verified with confocal Raman spectroscopy (Horiba, LabRam Odyssey). In total, 30 random spots were tested. In the Raman spectroscopy test, a 532 nm laser with ~0.45 mW was focused on the graphene under 50× objective. The integration time was 10 s. Spectra of the as-received sample are shown in Figure 2. No D peak (~1340 cm−1) can be observed in the Raman spectra, indicating that the supported graphene has rare D band-related defects. Peaks at 1588.3 cm−1 (G band) and 2681.5 cm−1 (2D band) can be observed. We used a ratio (IG/I2D) of the intensity of the G band to the intensity of 2D band to evaluate the layer number of the as-received sample [37]. It was found that the layer number of each tested spot is 1. Thus, we can verify that the layer number of graphene in the whole sample is 1, and the graphene distributes on the PMMA uniformly. Since the to-be-measured sample was cut from the as-received sample, we can confirm that the graphene layer number of each to-be-measured sample is 1. The PMMA thickness, graphene layer number, length, and width for the four samples are summarized in Table 1.

3. Thermal Transport Characterization of Giant-Scale Graphene Supported by PMMA

The thermal diffusivity of graphene supported by PMMA was characterized by diff-TET [36]. As shown in Figure 1a, the sample was suspended between two electrodes. Silver paste was used to eliminate the thermal contact resistance and electric contact resistance between the sample and the electrodes. The whole sample was placed in a vacuum chamber (Janis CCS-100/204N) to eliminate the convective effect on the measurement of the thermal diffusivity. During the experiment, a step current provided by Keithley 6221 was fed through the sample to induce joule heating. The rise of the temperature led to a change in resistance and voltage. The voltage evolution of the sample was recorded by an oscilloscope (Tektronix MDO32). The normalized temperature rise obtained from the experiment is determined as T * = ( V V 0 ) / ( V 1 V 0 ) . Here, V0 and V1 are the initial voltage and steady-state voltage over the sample, respectively. The environmental temperature was controlled by the temperature controller which ranges from 7 K to 300 K. Since most of the samples were broken at 125 K with a sudden jump resistance to a very high value, we did the experiments from room temperature (RT) down to 125 K in 25 K steps. During the joule heating, the thermal transport in the sample can be regarded as one-dimensional, thus the energy governing equation can be expressed as:
( ρ c p T ) t = k 2 T x 2 + q ˙
where ρ, cp, and k are the density, specific heat capacity of the sample, respectively. q ˙ is the heating power per unit volume, which equals to I 2 R s / A L . Here, A and L are the cross-sectional area and length of the sample, respectively. I and Rs are the fed-in current and the sample resistance, respectively. Since the two electrodes are much larger than the sample, the temperature of the two electrodes can be regarded as environmental temperature. Thus, the boundary conditions can be depicted as: T(x = 0, x = L) = T0 (T0: room temperature). The theoretical normalized temperature rise is defined as: T * ( t ) = [ T ( t ) T 0 ] / [ T ( t ) T 0 ] . The theoretical solution to Equation (1) is solved as:
T * = 48 π 4 m = 1 1 exp [ ( 2 m 1 ) 4 π 2 α e f f t / L 2 ] ( 2 m 1 ) 4
A Matlab program created based on Equation (2) was used to fit the normalized temperature rise obtained by the experiments. The trial value of αeff which gives the best fit of the experiment data was taken as the sample’s effective thermal diffusivity. Figure 3b shows the normalized voltage evolutions and the fitting curves for S1 at 275, 200, and 125 K. Magnificent fitting was obtained. The TET technique has been rigorously proven to be a quick and effective method to measure the thermal diffusivity of various conductive and non-conductive micro/nanoscale samples. More details can be found in references [13,36].
The effective thermal diffusivity (αeff) includes the radiation effect (αrad) which can be expressed as:
α r a d = 8 ε k B T 3 L 2 / ( π 2 δ ρ c p )
where ε, kB, and δ are emissivity, the Stefan–Boltzmann constant, and thickness of the sample, respectively. The emissivity of the sample is taken from [36] as 0.08. After obtaining the real thermal diffusivity (αreal), the effective thermal conductivity (keff) can be obtained as keff = αreal·(ρcp)p. The volumetric heat capacity of PMMA was used for the whole sample with high accuracy due to the extremely low mass proportion of graphene. In part 2, the Raman spectroscopy study of graphene showed that the graphene is distributed on the PMMA uniformly and all the graphene in samples is found to be a single layer. The relationship between effective thermal conductivity and thermal conductivity of PMMA and graphene can be depicted as:
k e f f = k g δ g + k p δ p δ g + δ p
Here, δp and δg are the thickness of PMMA and graphene, and are taken as 457 nm and 0.335 nm, respectively. kp and kg are thre thermal conductivities of PMMA and graphene, respectively. The temperature-dependent volumetric heat capacity and thermal conductivity of PMMA are taken from references [38,39]. The test methods of the heat capacity and k of PMMA are the vacuum calorimeter and the 3ω method, respectively [38,39]. The thermal conductivity of SLG supported by PMMA is obtained from the average graphene thermal conductivities of the four samples. Two critical points should be explained here. In thermal characterization, the graphene is first heated by the electrical current, then the thermal energy is transferred to the PMMA. Work by Liu et al., has proven that graphene and PMMA reached thermal equilibrium in the cross-sectional direction during thermal characterization [36]. Additionally, the interface thermal resistance (~10−2 K/W) between graphene and PMMA is neglectable in comparison to the thermal resistance (~106 K/W) of PMMA. More details can be found in [36].

4. Results and Discussion

4.1. Abnormal Temperature Coefficient of Resistance for Graphene Supported by PMMA

The normalized resistance (R* = R/R0, R0: resistance at RT) variation against the temperatures of S1, S2, S3, and S4 is presented in Figure 4a. The resistances of the four samples jump to very high values at 100 K, which indicates severe fractures in the sample. Thus, the normalized resistance is shown to be between RT and 125 K. It has been reported that the resistivity (ρe) of graphene is linearly proportional to T, as temperature is higher than the Bloch–Gruneisen temperature (ΘBG) of graphene and is proportional to T4 in the opposite case (T < ΘBG) [40]. However, the R* of the four samples decreases slightly as temperature decreases, then increases as temperature is decreased from RT to 125 K. The turning point is in the range of 250–275 K. The R* of S2 and S4 is increased by 35% and 23%, while the R* of S1 and S3 is decreased by 0.012% and 0.0075% as temperature changes from RT to 125 K. To obtain more details on the resistance of supported graphene, the temperature dependence of R* of S5 was measured every 10–25 K for three rounds. Figure 4b shows the normalized resistance of three rounds for S5. As temperature decreases from RT to 150 K, the R* trends of S5 are similar to that of S1 and S3. As temperature continues to decrease from 150 K, the R* of S5_r1 and S5_r2 first decreases and then increases with temperature, while the R* of S5_r3 continues to increase. Hinnefeld et al. studied the resistance of graphene supported by PDMS under strain and it was found that the tensile strain in the supported graphene can lead to rips in the graphene. These rips cause the resistance of the supported graphene to increase significantly with increased strain [41]. As temperature decreases, the graphene will expand and PMMA will contract. However, as the graphene is adhered to PMMA by van der Waals forces, the graphene will contract. The Raman spectra study of the experimental sample in the following part also confirms that the compressive strain builds in graphene as the temperature decreases from RT to 125.
The temperature coefficient of resistance (TCR) of graphene supported by PMMA is further discussed based on the data of S5_r3. The TCR of S5_r3 is shown in Figure 4c. The inset gives a close up to the TCR as temperature is in the range of 7–180 K. Apart from the intrinsic TCR of relaxed graphene, the overall TCR of the sample is affected by the thermal expansion of PMMA and graphene. The overall TCR of the whole sample can be described by the following equation: T C R = T C R i + ( β p β g ) γ . Here, T C R i and γ are the intrinsic TCR of relaxed graphene and a positive constant, respectively. βp and βg are the thermal expansion coefficients (β) of PMMA and graphene, respectively. Figure 4d shows the TEC of graphene [40] and PMMA [39]. As graphene and PMMA are held together by van der Waals forces, the compressive strain/stress will build in graphene. As shown in Figure 4c, βp decreases with decreasing temperature and remains positive when the temperature is above 0 K. βg remains negative when temperature is lower than RT. In the entire temperature range, T C R i remains negative [42,43]. When the temperature is close to RT, βp is relatively large, which helps maintain a positive TCR in supported graphene. As temperature is decreased, the difference between βp and βg becomes smaller in comparison to the effect of T C R i , causing the TCR of supported graphene to be negative.

4.2. Thermal Properties of SLG Supported by PMMA

The real thermal diffusivity without the effect of radiation of four samples is shown in Figure 5a. As temperature is decreased from RT, the αreal of S1, S2 and S4 decreases first, then the αreal of S2 remains constant while the αreal of S1 and S4 increases slightly. The change in the αreal of S3 is relatively small when compared to the other three samples. The k of SLG supported by PMMA is presented in Figure 5b. For comparison, the k of SLG supported by SiO2 [14,27] and bilayer graphene with PMMA residue [42] are shown in Figure 5b. At RT, the k of supported SLG is 743 ± 167 W/(m·K), which is smaller than that of SiO2-supported SLG (840 W/(m·K)) [27]. In [27], the k of supported graphene is measured by a comprehensive Raman optothermal method which eliminated the laser absorption uncertainty [27]. However, this work did not take the thermal non-equilibrium between different phonon branches into account. In this situation, the k of supported graphene is overestimated. The k in this work is larger than two literature values (~600 W/(m·K) for SiO2-supported SLG [14] and 560 W/(m·K) for bilayer graphene with PMMA [42]). The differences in the k of our work and above two differences may be attributed to the different interface coupling strength.
As temperature decreases from RT to 275 K, the k of supported SLG first decreases by 36.5%, which is much larger than the k reduction in [14,27]. As discussed in Section 4.1, there is compressive strain that builds in the graphene as temperature decreases from RT. There is limited work about the strain effect on the k of graphene supported by a flexible substrate. As such, we turn our attention to the effect on the k of suspended graphene. As reported in the literature, compressive strain in suspended graphene suppresses thermal conductivity [43]. This is attributed to the fluctuant structure created in graphene under compressive strain, under which the phonons can be scattered significantly more often [43,44]. With −0.05% strain in suspended graphene, it has been reported that the k can be reduced by 26% [43]. We speculate that the compressive strain induced by the TEC mismatch between graphene and PMMA results in a large k reduction when temperature decreases from RT to 275 K. Thus, Raman spectroscopy is used to investigate the G peak under different temperatures. The graphene sample suspended between two electrodes was placed in a cryogenic cell where the temperature can be controlled between 77 and 300 K. A confocal Raman is used to collect the Raman spectra of graphene under different temperatures.
Figure 6a shows the frequency shift of the G peak (ΔωG) of SLG on PMMA as a function of temperature. The G peak blueshifts as temperature goes down from RT. The temperature dependent frequency shift of the G peak is attributed to three parts: thermal expansion of the lattice ( Δ ω G E ); anharmonic effect ( Δ ω G A ), changing the phonon energy; and compressive strain ( Δ ω G S ), induced by the TEC mismatch between graphene and PMMA [45]. Thus, ΔωG can be given by [45]:
Δ ω G ( T ) = Δ ω G E ( T ) + Δ ω G A ( T ) + Δ ω G S ( T )
The G-band frequency shift due to the strain induced by the TEC mismatch can be expressed as [45]:
Δ ω G S ( T ) = β ε ( T ) .
where β is the biaxial strain coefficient of the G band. The biaxial strain coefficient of the G band has been determined to be around −70 cm−1/% at RT [46]. Since the temperature dependence of the G-band frequency is more reliable than the TEC of graphene and PMMA, the G-band frequency shift due to temperature variation ( Δ ω G E + Δ ω G A ) obtained from theoretical calculations [47] are subtracted from ΔωG to obtain Δ ω G S ( T ) . Strain buildup in graphene is calculated according to Equation (6). Figure 6b shows the temperature dependence of the strain buildup in graphene between graphene and PMMA. At 275 K, the strain is −0.047%, which indicates that compressive strain was built in graphene [46]. Li et al. reported that the k of suspended graphene can be reduced by 26% with −0.05% strain in graphene [43]. However, whether the −0.047% strain is sufficient to explain the large k reduction at 275 K is still doubtful. It also has been reported that the k of graphene nanoribbons are insensitive to compressive strain [48]. For graphene supported by flexible substrate, the effect of compressive strain on k still needs more theoretical calculations. As temperature continues decreasing from 275 K, the variation of k is not obvious, especially in the range of 150–225 K. We interpreted this trend from the structure of the sample. The SEM figure of S3 after cryogenic test is obtained and shown in Figure 7. A lot of micro-scale rips are observed. We speculate that there are micro-rips in the graphene when temperature is lower than 225 K. The rips suppress the k of graphene as compared to the k of supported graphene without any rips [14,42]. At lower temperatures, the effect of rips on thermal conductivity dominates, which keeps the changes in the k of graphene small.

5. Conclusions

In summary, we first reported the resistance–temperature relationship for supported graphene since this relationship is critical for explaining the behavior of graphene in thermal characterization. The TCR of our sample reduced from a positive value at RT to a negative value at low temperatures, while the free-standing graphene has a negative TCR throughout the entire temperature range. The abnormal R–T relationship is due to the TEC mismatch between graphene and PMMA and rips in graphene appearing at lower temperature. Using the diff-TET technique, the temperature dependence of the thermal conductivity of graphene supported by a flexible substrate is obtained. Through monitoring the blueshift of the Raman peak, compressive strain is confirmed to be built up in the graphene. The large reduction in k as temperature decreases from RT to 275 K is partly due to the compressive strain built up in graphene. This work can provide some guidance on the design of graphene-based electrothermal devices.

Author Contributions

Conceptualization, J.L. and L.M.; methodology, J.L., S.X. and Y.X.; validation, J.L., P.L.; investigation, P.L. and Q.W.; data curation, Y.X.; writing—original draft preparation, J.L.; writing—review and editing, J.L. and Y.X.; visualization, P.L.; supervision, J.L. and L.M.; project administration, J.L.; funding acquisition, J.L. and L.M. All authors have read and agreed to the published version of the manuscript.

Funding

Natural Science Foundation of Top Talent of SZTU 2019209; Guangdong Basic and Applied Basic Research Foundation 2020A1515110389; Project of Youth Innovative Talents in Higher Education Institutions of Guangdong 2020KQNCX069.

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

Support of this work by the Natural Science Foundation of Top Talent of SZTU (2019209, to J.L.), the Guangdong Basic and Applied Basic Research Foundation (2020A1515110389, to J.L.), and the Project of Youth Innovative Talents in Higher Education Institutions of Guangdong (2020KQNCX069, to L.M.) is gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ghosh, S.; Calizo, I.; Teweldebrhan, D.; Pokatilov, E.P.; Nika, D.L.; Balandin, A.A.; Bao, W.; Miao, F.; Lau, C.N. Extremely high thermal conductivity of graphene: Prospects for thermal management applications in nanoelectronic circuits. Appl. Phys. Lett. 2008, 92, 151911. [Google Scholar] [CrossRef]
  2. Geim, A.K. Graphene: Status and Prospects. Science 2009, 324, 1530–1534. [Google Scholar] [CrossRef] [PubMed]
  3. Ni, G.X.; Yang, H.Z.; Ji, W.; Baeck, S.J.; Toh, C.T.; Ahn, J.H.; Pereira, V.M.; Ozyilmaz, B. Tuning Optical Conductivity of Large-Scale CVD Graphene by Strain Engineering. Adv. Mater. 2014, 26, 1081–1086. [Google Scholar] [CrossRef] [PubMed]
  4. Novoselov, K.S.; Fal’ko, V.I.; Colombo, L.; Gellert, P.R.; Schwab, M.G.; Kim, K. A roadmap for graphene. Nature 2012, 490, 192–200. [Google Scholar] [CrossRef]
  5. Gu, X.K.; Wei, Y.J.; Yin, X.B.; Li, B.W.; Yang, R.G. Colloquium: Phononic thermal properties of two-dimensional materials. Rev. Mod. Phys. 2018, 90, 041002. [Google Scholar] [CrossRef]
  6. Bolotin, K.I.; Sikes, K.J.; Jiang, Z.; Klima, M.; Fudenberg, G.; Hone, J.; Kim, P.; Stormer, H.L. Ultrahigh electron mobility in suspended graphene. Solid. State. Commun. 2008, 146, 351–355. [Google Scholar] [CrossRef]
  7. Lee, C.; Wei, X.D.; Kysar, J.W.; Hone, J. Measurement of the elastic properties and intrinsic strength of monolayer graphene. Science 2008, 321, 385–388. [Google Scholar] [CrossRef]
  8. You, R.; Liu, Y.Q.; Hao, Y.L.; Han, D.D.; Zhang, Y.L.; You, Z. Laser Fabrication of Graphene-Based Flexible Electronics. Adv. Mater. 2020, 32, 1901981. [Google Scholar] [CrossRef]
  9. Secor, E.B.; Lim, S.; Zhang, H.; Frisbie, C.D.; Francis, L.F.; Hersam, M.C. Gravure Printing of Graphene for Large-Area Flexible Electronics. Adv. Mater. 2014, 26, 1401052. [Google Scholar] [CrossRef] [PubMed]
  10. Freitag, M.; Low, T.; Avouris, P. Increased Responsivity of Suspended Graphene Photodetectors. Nano Lett. 2013, 13, 1644–1648. [Google Scholar] [CrossRef] [PubMed]
  11. Cai, X.Y.; Lai, L.F.; Shen, Z.X.; Lin, J.Y. Graphene and graphene-based composites as Li-ion battery electrode materials and their application in full cells. J. Mater. Chem. A 2017, 5, 15423–15446. [Google Scholar] [CrossRef]
  12. Cai, C.L.; Wang, T.; Zhang, Y.X.; He, N.Y. Facile fabrication of ultra-large graphene film with high photothermal effect and thermal conductivity. Appl. Surf. Sci. 2021, 563, 150354. [Google Scholar] [CrossRef]
  13. Cai, W.W.; Moore, A.L.; Zhu, Y.W.; Li, X.S.; Chen, S.S.; Shi, L.; Ruoff, R.S. Thermal Transport in Suspended and Supported Monolayer Graphene Grown by Chemical Vapor Deposition. Nano Lett. 2010, 10, 1645–1651. [Google Scholar] [CrossRef]
  14. Seol, J.H.; Jo, I.; Moore, A.L.; Lindsay, L.; Aitken, Z.H.; Pettes, M.T.; Li, X.S.; Yao, Z.; Huang, R.; Broido, D.; et al. Two-Dimensional Phonon Transport in Supported Graphene. Science 2010, 328, 213–216. [Google Scholar] [CrossRef] [PubMed]
  15. Ong, Z.Y.; Pop, E. Effect of substrate modes on thermal transport in supported graphene. Phys. Rev. B 2011, 84, 075471. [Google Scholar] [CrossRef]
  16. Wei, Z.Y.; Yang, J.K.; Bi, K.D.; Chen, Y.F. Mode dependent lattice thermal conductivity of single layer graphene. J. Appl. Phys. 2014, 116, 153503. [Google Scholar] [CrossRef]
  17. Xu, X.F.; Pereira, L.F.C.; Wang, Y.; Wu, J.; Zhang, K.W.; Zhao, X.M.; Bae, S.; Bui, C.T.; Xie, R.G.; Thong, J.T.L.; et al. Length-dependent thermal conductivity in suspended single-layer graphene. Nat. Commun. 2014, 5, 3689. [Google Scholar] [CrossRef] [PubMed]
  18. Malekpour, H.; Ramnani, P.; Srinivasan, S.; Balasubramanian, G.; Nika, D.L.; Mulchandani, A.; Lake, R.K.; Balandin, A.A. Thermal conductivity of graphene with defects induced by electron beam irradiation. Nanoscale 2016, 8, 14608–14616. [Google Scholar] [CrossRef] [PubMed]
  19. Hu, S.Q.; Chen, J.; Yang, N.; Li, B.W. Thermal transport transport in graphene with defect and doping: Phonon modes analysis. Carbon 2017, 116, 139–144. [Google Scholar] [CrossRef]
  20. Li, M.Y.; Deng, T.Z.X.; Zheng, B.; Zhang, Y.; Liao, Y.G.; Zhou, H.M. Effect of Defects on the Mechanical and Thermal Properties of Graphene. Nanomaterials 2019, 9, 347. [Google Scholar] [CrossRef] [PubMed]
  21. Chen, J.; Zhang, G.; Li, B.W. Substrate coupling suppresses size dependence of thermal conductivity in supported graphene. Nanoscale 2013, 5, 532–536. [Google Scholar] [CrossRef] [PubMed]
  22. Li, M.; Zhang, J.C.; Hu, X.J.; Yue, Y. Thermal transport across graphene/SiC interface: Effects of atomic bond and crystallinity of substrate. Appl. Phys. A 2015, 119, 415–424. [Google Scholar] [CrossRef]
  23. Liu, H.K.; Lin, Y.; Luo, S.N. Grain Boundary Energy and Grain Size Dependences of Thermal Conductivity of Polycrystalline Graphene. J. Phys. Chem. C 2014, 118, 24797–24802. [Google Scholar] [CrossRef]
  24. Limbu, T.B.; Hahn, K.R.; Mendoza, F.; Sahoo, S.; Razink, J.J.; Katiyar, R.S.; Weiner, B.R.; Morell, G. Grain size-dependent thermal conductivity of polycrystalline twisted bilayer graphene. Carbon 2017, 117, 367–375. [Google Scholar] [CrossRef]
  25. Lee, W.; Kihm, K.D.; Kim, H.G.; Shin, S.; Lee, C.; Park, J.S.; Cheon, S.; Kwon, O.M.; Lim, G.; Lee, W. In-Plane Thermal Conductivity of Polycrystalline Chemical Vapor Deposition Graphene with Controlled Grain Sizes. Nano Lett. 2017, 17, 2361–2366. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, S.; Li, Q.; Zhang, Q.; Qu, Y.; Ji, H.; Ruoff, R.S.; Cai, W. Thermal conductivity measurements of suspended graphene with and without wrinkles by microRaman mapping. Nanotechnology 2012, 23, 365701. [Google Scholar] [CrossRef] [PubMed]
  27. Li, Q.Y.; Xia, K.L.; Zhang, J.; Zhang, Y.Y.; Li, Q.Y.; Takahashi, K.; Zhang, X. Measurement of specific heat and thermal conductivity of supported and suspended graphene by a comprehensive Raman optothermal method. Nanoscale 2017, 9, 10784–10793. [Google Scholar] [CrossRef]
  28. Wang, R.D.; Wang, T.Y.; Zobeiri, H.; Li, D.C.; Wang, X.W. Energy and Charge Transport in 2D Atomic Layer Materials: Raman-Based Characterization. Nanomaterials 2020, 10, 1807. [Google Scholar] [CrossRef]
  29. Li, Q.Y.; Katakami, K.; Ikuta, T.; Kohno, M.; Zhang, X.; Takahashi, K. Measurement of thermal contact resistance between individual carbon fibers using a laser-flash Raman mapping method. Carbon 2019, 141, 92–98. [Google Scholar] [CrossRef]
  30. Li, Q.Y.; Zhang, X.; Takahashi, K. Variable-spot-size laser-flash Raman method to measure in-plane and interfacial thermal properties of 2D van der Waals heterostructures. Int. J. Heat Mass Trans. 2018, 125, 1230–1239. [Google Scholar] [CrossRef]
  31. Vallabhaneni, A.K.; Singh, D.; Bao, H.; Murthy, J.; Ruan, X.L. Reliability of Raman measurements of thermal conductivity of single-layer graphene due to selective electron-phonon coupling: A first-principles study. Phys. Rev. B 2016, 93, 125432. [Google Scholar] [CrossRef]
  32. Wang, R.D.; Zobeiri, H.; Xie, Y.S.; Wang, X.W.; Zhang, X.; Yue, Y.N. Distinguishing Optical and Acoustic Phonon Temperatures and Their Energy Coupling Factor under Photon Excitation in nm 2D Materials. Adv. Sci. 2020, 7, 2000097. [Google Scholar] [CrossRef]
  33. Liu, J.; Li, P.; Zheng, H. Review on Techniques for Thermal Characterization of Graphene and Related 2D Materials. Nanomaterials 2021, 11, 2787. [Google Scholar] [CrossRef]
  34. Wang, L.; Hu, B.B.; Li, B.W. Logarithmic divergent thermal conductivity in two-dimensional nonlinear lattices. Phys. Rev. E 2012, 86, 040101. [Google Scholar] [CrossRef] [PubMed]
  35. Su, R.X.; Zhang, X. Size effect of thermal conductivity in monolayer graphene. Appl. Therm. Eng. 2018, 144, 488–494. [Google Scholar] [CrossRef]
  36. Liu, J.; Xu, Z.L.; Cheng, Z.; Xu, S.; Wang, X.W. Thermal Conductivity of Ultrahigh Molecular Weight Polyethylene Crystal: Defect Effect Uncovered by 0 K Limit Phonon Diffusion. ACS Appl. Mater. Inter. 2015, 7, 27279–27288. [Google Scholar] [CrossRef]
  37. Graf, D.; Molitor, F.; Ensslin, K.; Stampfer, C.; Jungen, A.; Hierold, C.; Wirtz, L. Spatially resolved raman spectroscopy of single- and few-layer graphene. Nano Lett. 2007, 7, 238–242. [Google Scholar] [CrossRef] [PubMed]
  38. Putnam, S.A.; Cahill, D.G.; Ash, B.J.; Schadler, L.S. High-precision thermal conductivity measurements as a probe of polymer/nanoparticle interfaces. J. Appl. Phys. 2003, 94, 6785. [Google Scholar] [CrossRef]
  39. Esposito, M.; Buontempo, S.; Petriccione, A.; Zarrelli, M.; Breglio, G.; Saccomanno, A.; Szillasi, Z.; Makovec, A.; Cusano, A.; Chhiuchiolo, A.; et al. and Giordano, M. Fiber Bragg Grating sensors to measure the coefficient of thermal expansion of polymers at cryogenic temperatures. Sens. Actuators 2013, 189, 195–203. [Google Scholar] [CrossRef]
  40. D’Souza, R.; Mukherjee, S. First-principles study of the electrical and lattice thermal transport in monolayer and bilayer graphene. Phys. Rev. B 2017, 95, 085435. [Google Scholar] [CrossRef]
  41. Hinnefeld, J.H.; Gill, S.T.; Zhu, S.; Swanson, W.J.; Li, T.; Mason, N. Reversible Mechanical and Electrical Properties of Ripped Graphene. Phys. Rev. Appl. 2015, 3, 014010. [Google Scholar] [CrossRef]
  42. Pettes, M.T.; Jo, I.S.; Yao, Z.; Shi, L. Influence of Polymeric Residue on the Thermal Conductivity of Suspended Bilayer Graphene. Nano Lett. 2011, 11, 1195–1200. [Google Scholar] [CrossRef] [PubMed]
  43. Li, X.B.; Maute, K.; Dunn, M.L.; Yang, R.G. Strain effects on the thermal conductivity of nanostructures. Phys. Rev. B 2010, 81, 245318. [Google Scholar] [CrossRef]
  44. Guo, Z.X.; Zhang, D.; Gong, X.G. Thermal conductivity of graphene nanoribbons. Appl. Phys. Lett. 2009, 95, 163103. [Google Scholar] [CrossRef]
  45. Yoon, D.; Son, Y.W.; Cheong, H. Negative Thermal Expansion Coefficient of Graphene Measured by Raman Spectroscopy. Nano Lett. 2011, 11, 3227–3231. [Google Scholar] [CrossRef]
  46. Yoon, D.; Son, Y.W.; Cheong, H. Strain-Dependent Splitting of the Double-Resonance Raman Scattering Band in Graphene. Phys. Rev. Lett. 2011, 106, 155502. [Google Scholar] [CrossRef]
  47. Bonini, N.; Lazzeri, M.; Marzari, N.; Mauri, F. Phonon anharmonicities in graphite and graphene. Phys. Rev. Lett. 2007, 99, 176802. [Google Scholar] [CrossRef]
  48. Wei, N.; Xu, L.Q.; Wang, H.Q.; Zheng, J.C. Strain engineering of thermal conductivity in graphene sheets and nanoribbons: A demonstration of magic flexibility. Nanotechnology 2011, 22, 105705. [Google Scholar] [CrossRef]
Figure 1. (a) Topology of S3 under microscope; (b) AFM image of the as-received SLG supported by PMMA.
Figure 1. (a) Topology of S3 under microscope; (b) AFM image of the as-received SLG supported by PMMA.
Nanomaterials 12 02799 g001
Figure 2. (a,b) Raman spectra of as-received sample.
Figure 2. (a,b) Raman spectra of as-received sample.
Nanomaterials 12 02799 g002
Figure 3. (a) Schematic setup of the TET measurement (not to scale); (b) Normalized voltage evolution and the fitting curve for S1 at 275, 200, and 125 K.
Figure 3. (a) Schematic setup of the TET measurement (not to scale); (b) Normalized voltage evolution and the fitting curve for S1 at 275, 200, and 125 K.
Nanomaterials 12 02799 g003
Figure 4. (a) Normalized resistance of S1, S2, S3, and S4. (b) Normalized resistance of S1, S3, S5_r1, S5_r2. and S5_r3. (c) Temperature coefficient of resistance of S5. The TCR was obtained based on the data of S5_r3. (d) Thermal expansion coefficient of PMMA (Reprinted/adapted with permission from Ref. [39]. Copyright 2013, Elsevier) and suspended graphene (Reprinted/adapted with permission from Ref. [40]. Copyright 2011, American Chemical Society). The data in blue square are obtained through experiment while the data shown in black square are estimated values.
Figure 4. (a) Normalized resistance of S1, S2, S3, and S4. (b) Normalized resistance of S1, S3, S5_r1, S5_r2. and S5_r3. (c) Temperature coefficient of resistance of S5. The TCR was obtained based on the data of S5_r3. (d) Thermal expansion coefficient of PMMA (Reprinted/adapted with permission from Ref. [39]. Copyright 2013, Elsevier) and suspended graphene (Reprinted/adapted with permission from Ref. [40]. Copyright 2011, American Chemical Society). The data in blue square are obtained through experiment while the data shown in black square are estimated values.
Nanomaterials 12 02799 g004
Figure 5. (a) Real thermal diffusivity without radiation effect for S1, S2, S3, and S4; (b) Thermal conductivity variation against temperature of SLG supported by PMMA. For comparison, the thermal conductivities of SLG supported by SiO2 (Reprinted/adapted with permission from Ref. [14]. Copyright 2010, The American Association for the Advancement of Science) and bilayer graphene with PMMA residue (Reprinted/adapted with permission from Ref. [42]. Copyright 2011, American Chemical Society).
Figure 5. (a) Real thermal diffusivity without radiation effect for S1, S2, S3, and S4; (b) Thermal conductivity variation against temperature of SLG supported by PMMA. For comparison, the thermal conductivities of SLG supported by SiO2 (Reprinted/adapted with permission from Ref. [14]. Copyright 2010, The American Association for the Advancement of Science) and bilayer graphene with PMMA residue (Reprinted/adapted with permission from Ref. [42]. Copyright 2011, American Chemical Society).
Nanomaterials 12 02799 g005
Figure 6. (a) Frequency shift of the G band of SLG supported by PMMA. (b) Temperature dependence of the strain buildup in graphene due to the TEC mismatch between graphene and PMMA.
Figure 6. (a) Frequency shift of the G band of SLG supported by PMMA. (b) Temperature dependence of the strain buildup in graphene due to the TEC mismatch between graphene and PMMA.
Nanomaterials 12 02799 g006
Figure 7. SEM figure of S3 after the cryogenic test. Many micro-scale rips can be observed.
Figure 7. SEM figure of S3 after the cryogenic test. Many micro-scale rips can be observed.
Nanomaterials 12 02799 g007
Table 1. PMMA thickness, graphene layer number, length, and width of S1, S2, S3, S4, and S5.
Table 1. PMMA thickness, graphene layer number, length, and width of S1, S2, S3, S4, and S5.
SampleS1S2S3S4S5
PMMA thickness(nm)457457457457457
Graphene layer number11111
Length (mm)1.731.532.352.252.09
Width (mm)1.921.853.202.800.74
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, J.; Li, P.; Xu, S.; Xie, Y.; Wang, Q.; Ma, L. Temperature Dependence of Thermal Conductivity of Giant-Scale Supported Monolayer Graphene. Nanomaterials 2022, 12, 2799. https://doi.org/10.3390/nano12162799

AMA Style

Liu J, Li P, Xu S, Xie Y, Wang Q, Ma L. Temperature Dependence of Thermal Conductivity of Giant-Scale Supported Monolayer Graphene. Nanomaterials. 2022; 12(16):2799. https://doi.org/10.3390/nano12162799

Chicago/Turabian Style

Liu, Jing, Pei Li, Shen Xu, Yangsu Xie, Qin Wang, and Lei Ma. 2022. "Temperature Dependence of Thermal Conductivity of Giant-Scale Supported Monolayer Graphene" Nanomaterials 12, no. 16: 2799. https://doi.org/10.3390/nano12162799

APA Style

Liu, J., Li, P., Xu, S., Xie, Y., Wang, Q., & Ma, L. (2022). Temperature Dependence of Thermal Conductivity of Giant-Scale Supported Monolayer Graphene. Nanomaterials, 12(16), 2799. https://doi.org/10.3390/nano12162799

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