Next Article in Journal
Laser Interference Lithography—A Method for the Fabrication of Controlled Periodic Structures
Previous Article in Journal
Nucleation and Stability of Toron Chains in Non-Centrosymmetric Magnetic Nanowires
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improved Thermal Anisotropy of Multi-Layer Tungsten Telluride on Silicon Substrate

1
College of Physical Science and Technology, Xiamen University, Xiamen 361005, China
2
College of Science, National University of Defense Technology, Changsha 410073, China
3
College of Advanced Interdisciplinary Studies, National University of Defense Technology, Changsha 410073, China
4
Jiujiang Research Institute of Xiamen University, Jiujiang 332105, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(12), 1817; https://doi.org/10.3390/nano13121817
Submission received: 18 May 2023 / Revised: 25 May 2023 / Accepted: 3 June 2023 / Published: 7 June 2023

Abstract

:
WTe2, a low-symmetry transition metal dichalcogenide, has broad prospects in functional device applications due to its excellent physical properties. When WTe2 flake is integrated into practical device structures, its anisotropic thermal transport could be affected greatly by the substrate, which matters a lot to the energy efficiency and functional performance of the device. To investigate the effect of SiO2/Si substrate, we carried out a comparative Raman thermometry study on a 50 nm-thick supported WTe2 flake (with κzigzag = 62.17 W·m−1·K−1 and κarmchair = 32.93 W·m−1·K−1), and a suspended WTe2 flake of similar thickness (with κzigzag = 4.45 W·m−1·K−1, κarmchair = 4.10 W·m−1·K−1). The results show that the thermal anisotropy ratio of supported WTe2 flake (κzigzagarmchair ≈ 1.89) is about 1.7 times that of suspended WTe2 flake (κzigzagarmchair ≈ 1.09). Based on the low symmetry nature of the WTe2 structure, it is speculated that the factors contributing to thermal conductivity (mechanical properties and anisotropic low-frequency phonons) may have affected the thermal conductivity of WTe2 flake in an uneven manner when supported on a substrate. Our findings could contribute to the 2D anisotropy physics and thermal transport study of functional devices based on WTe2 and other low-symmetry materials, which helps solve the heat dissipation problem and optimize thermal/thermoelectric performance for practical electronic devices.

1. Introduction

Low-symmetry materials exhibit anisotropic thermal [1,2,3,4], electrical [5,6,7,8], optical [5], optoelectronic [9], and thermoelectric [10] properties along different lattice orientations, which provides a brand-new opportunity to design high-performance devices. WTe2, a recently popular player in two-dimensional (2D) materials with low-symmetry lattice structures, has exhibited outstanding functional device applications [11,12] based on its heavy atomic mass, low-energy optical absorption [13], thickness-dependent anisotropy of Raman modes [14], and superconductivity properties [15,16]. It has been reported that the conductivity and the photoelectric anisotropy ratio of WTe2 film are about 103 and 300 [17], respectively, allowing it to be applied to anisotropic electric and photonic devices [18,19,20,21,22].
Considering that heat dissipation efficiency is crucial to device reliability and performance, while the thermal transport characteristics of WTe2 and other low-symmetry materials are quite sensitive to varying external conditions, it is necessary to explore the effect of external conditions before designing WTe2-based functional devices. The ambient temperature, mechanical strain, and substrate coupling will all lead to diverse thermal transport behaviors, which may affect thermal anisotropy and device performance [1,23,24]. In practical applications, especially low-symmetry materials are mostly integrated on substrates; therefore, understanding the effect of substrate is important for device design. For example, the thermal conductivity of suspended black phosphorene (BP) flake exhibits obvious anisotropy along different directions [1,2,10,25,26]. When supported on a solid substrate, the thermal anisotropy of BP flake will show obvious improvement compared to the suspended one, which is proved by molecular dynamics simulations [27].
As most 2D-integrated circuits are silicon-based, we chose an SiO2/Si substrate to explore the role of substrate coupling on multilayer WTe2. In this paper, the in-plane thermal conductivity of 50 nm-thick suspended and supported WTe2 samples along zigzag/armchair axes is investigated by laser-heating and electrical-heating Raman thermometry. The experimental results show that the in-plane thermal conductivity of supported WTe2 flake is κzigzag = 62.17 W·m−1·K−1, κarmchair = 32.93 W·m−1·K−1, and that of suspended WTe2 flake is κzigzag = 4.45 W·m−1·K−1, κarmchair = 4.10 W·m−1·K−1. Low-frequency phonons mainly contribute to the thermal conductivity along zigzag and armchair directions [28], and the thermal conductivity along the zigzag direction is greater for WTe2 flakes. In addition, it has been reported that the mechanical properties of materials will be affected when mechanical force is applied [29]. It is possible that the different contributions of low-frequency phonons and mechanical properties along zigzag and armchair directions caused by substrate have led to a significantly larger thermal anisotropy ratio for supported WTe2 flake (κzigzag/κarmchair ≈ 1.89) compared to the case for suspended WTe2 flake (κzigzag/κarmchair ≈ 1.09). Therefore, it is suggested that low-frequency phonons and mechanical properties are two of the possible reasons for the variation in the thermal anisotropy ratio of supported WTe2 flake. Our finding may be helpful to study the heat dissipation process in low-symmetry materials and offer guidance for efficient thermal management of low-symmetry material devices.

2. Materials and Methods

2.1. Preparation of WTe2 Samples

We first placed a small piece of polydimethylsiloxane (PDMS) on a clean slide and set it aside. A piece of bulk WTe2 (PDMS and bulk WTe2 were purchased from Shanghai Onway Technology Co., Ltd., Shanghai, China, http://www.onway-tec.com/, accessed on 7 September 2021) was placed on the scotch tape and folded repeatedly. The tape was then attached to the PDMS and pressed gently to make it fit closely to the PDMS. After ten seconds, the tape was quickly removed from the PDMS, and WTe2 flakes were transferred from the tape to the surface of the PDMS. We selected a uniform area of WTe2 flake on PDMS, then transferred it to the silicon substrate with periodic holes (with a diameter of 12 μm). The WTe2 flake covered the silicon holes to construct the suspending region of the WTe2 sample.
We took another piece of bulk WTe2 on a clean scotch tape and repeatedly folded the tape, then attached it to an SiO2 (285 nm)/Si substrate and pressed it gently so that the tape fit tightly to the substrate. After five minutes, the tape was slowly removed from the substrate, and WTe2 flakes with different optical contrasts covered the surface of the substrate, which could be observed through an optical microscope. We selected an area of WTe2 flake, then designed specific electrode patterns along zigzag and armchair directions of WTe2 flake by electron-beam lithography (Raith e-LINE Plus, Dortmund, Germany), and deposited Ti (with a thickness of 5 nm) and Au (with a thickness of 70 nm) on the surface of WTe2 flake by electron-beam evaporation (Kurt J. Lesker PVD75, Pittsburgh, PA, USA) to complete the preparation of the WTe2 device (supported WTe2 sample).

2.2. Characterizations of WTe2 Samples

Optical microscopy imaging was performed by the LV100D system (Nikon, Japan). The thickness and uniformity of samples were obtained by AFM (NT-MDT Company, Moscow, Russia). During the AFM measurement, there is an interaction force between the probe tip and the surface atoms of the sample. The interaction force is weaker when the distance between the tip and the sample surface is greater. When the tip is closer to the sample surface, the interaction force is greater. This change in interaction force causes a deformation of the cantilever beam, which is detected by the photosensitive detector and fed back to finally obtain the surface information of the sample.
Raman spectroscopy is a common tool to characterize the lattice structure, electrical, optical, and phonon properties of 2D materials [30,31,32]. Angle-dependent Raman spectroscopy was extensively used to study anisotropic lattice structures [33,34,35]. In this paper, we investigate the optical anisotropy of WTe2 samples by angle-dependent Raman spectroscopy (Renishaw, Wotton-under-Edge, UK) with a 532 nm excitation laser with 50× objective (NA = 0.55, and the laser spot size is about 616 nm), and the samples were rotated by a sample stage. Figure S1 shows the schematic diagram of the polarized-Raman configuration. During the characterization of angle-dependent Raman spectroscopy, we fixed the incident laser polarization (ei) parallel to the scattered laser polarization (es) (labeled as ei ‖ es, as shown in Figure S1) and set θ as the angle between the incident laser polarization (ei) and the zigzag axis. We placed the zigzag axis of the WTe2 sample parallel to the incident laser polarization (defined as 0°) [34,35], and the sample could be rotated clockwise from 0° to 360° by a sample stage. The optical anisotropy of the WTe2 sample is then investigated by analyzing the variation of Raman intensity with θ. The details of angle-dependent Raman spectroscopy are presented in the illustrations in Figure 1 and Figure S2.

2.3. Raman Thermometry Measurements of WTe2 Samples

Raman thermometry is an important method for investigating the thermal conductivity of 2D materials [36,37,38]. As the temperature increases, the Raman peak frequency will shift towards the lower frequency, which means that the Raman peak is redshifted [37]. The thermal conductivity can be obtained from the change in Raman frequency [39]. The thermal conductivity of WTe2 samples is investigated by Raman thermometry in the following two steps [28,40,41]. First, the in situ Raman spectra are obtained under a certain temperature range, from which the coefficient between Raman peak frequency and temperature can be extracted. The temperature-dependent in situ Raman spectra are performed with a temperature range from liquid nitrogen to room temperature to avoid damage to the sample [28,40,41]. Second, the sample is heated by increasing the laser power/bias voltage to analyze the Raman peak frequency shift against changing power. Laser power-dependent and bias voltage-dependent in situ Raman spectra are performed with low laser power and bias voltage range at room temperature [28,40]. Here, based on the different heating sources, Raman thermometry can be divided into laser-heating Raman thermometry (heating the sample with laser power) [41] and electrical-heating Raman thermometry (heating the sample with electrical power) [40].
The thermal conductivity of suspended WTe2 flake along zigzag and armchair directions is measured by laser-heating Raman thermometry with temperature-dependent (Renishaw, Wotton-under-Edge, UK) and laser-dependent in situ Raman spectra (Alpha 300R system, WITec Company, Ulm, Germany). The temperature is controlled by the A599 heating accessory holder (Bruker Corporation, Billerica, MA, USA). The in-plane thermal conductivity along different directions could be calculated by the equation κ zigzag = χ T ,   zigzag   ( 1 / 2 π h )   ( δ ω zigzag / δ P A ,   zigzag ) 1 and κ armchair = χ T ,   armchair   ( 1 / 2 π h )   ( δ ω armchair / δ P A ,   armchair ) 1 , where χ T is the first-order temperature coefficient, h is the thickness, PA is the absorbed laser power of suspended WTe2 flake (PA = AP, A is the absorptivity and P is the incident laser power), and δω is the Raman peak position shift caused by the increase in absorbed laser power δPA.
The thermal conductivity of supported WTe2 flake along zigzag and armchair directions is measured by electrical-heating Raman thermometry with temperature-dependent (Renishaw, Wotton-under-Edge, UK) and bias voltage-dependent in situ Raman spectra (Alpha 300R system, WITec Company, Ulm, Germany). The electrical measurement is measured by Keithley 4200. Based on the relationship between the temperature and the channel power density, we could calculate the in-plane thermal conductivity along different directions of supported WTe2 flake by the equation T zigzag   = T 0 ,   zigzag + P zigzag / κ and T armchair =   T 0 ,   armchair + P armchair / κ , where T is the temperature of supported WTe2 flake, κ is that along the direction perpendicular to current, P is the power density per channel length, and T0 is the temperature when the bias voltage is 0 V.

3. Results and Discussion

Figure 1a shows the low-symmetry crystal structure of WTe2. The tellurium-tungsten-tellurium atomic planes exhibit an orthorhombic lattice through vertical stacking along the z-axis by van der Waals force. The tungsten atoms form a W-W chain along the x-axis, and the upper tellurium layer is rotated 180° with respect to the bottom tellurium layer, showing a clear anisotropic structure. We define the x-axis as a zigzag direction and the y-axis as an armchair direction. Since the cleave energy in the zigzag direction is smaller than that in the armchair direction, bulk WTe2 will easily break along the zigzag direction during mechanical exfoliation [28] (it is observed that the edge along the zigzag direction is longer through the optical microscope).
We then explore the optical anisotropy of multilayer WTe2 by angle-dependent Raman spectroscopy, which contributes to the subsequent investigations of thermal transport properties along different lattice directions. The sample was then rotated clockwise from 0° to 360° during angle-dependent Raman measurements in a parallel-polarized configuration (labeled as e i e s ) , where the incident 532 nm laser polarization (ei) was parallel to the scattered laser polarization (es). Figure 1b shows the contour map of Raman intensity varying as a function of angle θ (the angle between laser polarization and zigzag direction) for three A1 modes located at about 133 cm−1 (4A1), 163 cm−1 (8A1), and 212 cm−1 (10A1).
According to the Placzek approximation [42,43], the variation of Raman intensity with θ for the A1 and A2 modes mentioned above can be understood. As θ ranges from 0° to 360°, the Raman intensity of A1 and A2 modes can be described by the following equation:
I e i · R ~ · e s 2
In a parallel-polarized configuration, the unitary vector of incident light is ei = (cos θ, sin θ, 0) and that of scattered light is es = (cos θ, sin θ, 0). R ~ is the Raman tensor. R ~ A 1 = a 0 0 0 b 0 0 0 c and R ~ A 2 = 0 d 0 d 0 0 0 0 0 . Then the Raman intensities of A1 and A2 modes become functions of θ, as is described [33]
I A 1 a 2 1 + b a 1 sin 2 θ 2 ,
I A 2 d 2 sin 2 2 θ ,
where a, b, and d are the elements of Raman tensors that determine the peak intensity. Thus, the Raman intensity of A1 mode exhibits two-fold symmetry, while that of A2 mode exhibits four-fold symmetry, in accordance with our experiment results as shown in Figure 1c–f. Equation (2) exhibits that the variation of Raman intensity of A1 mode with θ is related to the relationship between a and b. For a > b , the maximum Raman intensity value appeared at θ = 0° and θ = 180°, which means the incident light polarization is parallel to W-W chains (zigzag direction), such as in 4A1 and 8A1 modes. Conversely, the maximum Raman intensity is located at θ = 90° and θ = 270° (armchair direction), when the incident light polarization is perpendicular to W–W chains for a < b , namely, 10A1 mode.
After characterizing the anisotropic crystal structure and Raman modes of WTe2 flake, we then investigate the in-plane anisotropy of thermal transport. As WTe2 flakes are mostly integrated into silicon-based electronic devices, it is reasonable to investigate the thermal transport properties of WTe2 flakes supported on SiO2/Si substrates for actual applications. Here, we explored the anisotropic thermal transport of both suspended and supported WTe2 samples through laser-heating and electrical-heating Raman thermometry.
The height profile in Figure S3a clearly indicates that the thickness of suspended WTe2 flake is about 50 nm. After basic characterization, temperature- and laser power-dependent Raman spectra were used to study the anisotropic thermal transport of suspended WTe2 flake. Figure 2b shows the schematic diagram of the laser-heating Raman thermometry setup, where the laser (focused on the center of suspended WTe2 flake) is used to heat suspended WTe2 sample.
Figure S4a,b shows the peak positions of prominent Raman modes 8A1 (located at about 163 cm−1) and 10A1 (located at about 212 cm−1) as a function of temperature (ranging from 203 K to 273 K, and the step is 10 K). It was found that 8A1 and 10A1 modes exhibit apparent redshifts as the temperature increases due to anharmonic lattice vibrations and thermal expansion [44]. Where the frequencies of 8A1 Raman mode along the zigzag and armchair directions are redshifted by 0.9025 cm−1 and 0.748 cm−1, and the frequencies of 10A1 Raman mode along the zigzag and armchair directions are redshifted by 1.2535 cm−1and 0.821 cm−1, respectively. As reported in a previous study, 10A1 mode was more sensitive to temperature than 8A1 mode [14,28], so we chose 10A1 mode as the thermometer. Figure 2c indicates that the Raman peak position for 10A1 mode follows a linear tendency with increasing temperature along zigzag and armchair directions, as fitted by the equation [45]
ω T = ω 0 + χ T T ,
where ω 0 is the Raman peak position at 0 K, χ T is the first-order temperature coefficient, and T is the temperature. According to Figure 2c, the χ T of 10A1 mode can be extracted, χ T ,   zigzag = −0.0177 ± 0.0008 cm−1·K−1 and χ T ,   armchair = −0.0126 ± 0.0004 cm−1·K−1, which is within the same order of magnitude as the data of previous studies [28].
To calculate the thermal conductivity of suspended WTe2 flake, laser-dependent Raman spectra were further investigated in detail. The laser spot size (about 616 nm) is much smaller than the suspended area of WTe2 flake on silicon substrate (the diameter of the holes on silicon substrate is 12 μm), meaning the heat transfer to the substrate is negligible. The corresponding Raman spectra of suspended WTe2 flake as the incident laser power ranges from 106 μW to 325 μW are shown in Figure S4c,d, suggesting the redshifts of 8A1 and 10A1 modes. Where the frequencies of 10A1 Raman mode along the zigzag and armchair directions are redshifted by 1.402 cm−1 and 0.8485 cm−1. The linear relationship between the Raman peak position for 10A1 mode and increasing laser power is shown in Figure 2d, as described by the equation [37]
ω = ω P 2     ω P 1 = χ P P 2     P 1 =   χ P P ,
where χ P is the first-order laser-dependent coefficient and P is the laser power. For suspended WTe2 flake, χ P ,   zigzag = −0.0069 ± 0.0002 cm−1·μW−1 and χ P ,   armchair = −0.0053 ± 0.0001 cm−1·μW−1. The thermal conductivity κ of suspended WTe2 flake can be calculated from [36]
Q t = κ T · dS ,
where Q is the heat transferred along the cross-section area S (during the time t), and T is the temperature. Considering the radial heat dissipation from the center to the edge of the suspended area, Equation (6) is converted to κ = ( 1 / 2 π h )   ( P / T ) , where h is the thickness of suspended WTe2 flake. ΔT corresponds to the temperature change that is caused by ΔP. By analyzing the linear relationship between laser power and increasing temperature, the thermal conductivity can be calculated from
κ = χ T 1 2 π h δ ω δ P A 1 ,
where PA is the absorbed laser power of suspended WTe2 flake (PA = AP, A is the absorptivity and P is the incident laser power). δω is the Raman peak position shift caused by the increase in absorbed laser power δPA [36,37]. Figure S5 shows the angle-dependent absorbance plot of suspended WTe2 flake at room temperature (the measured wavelength is 532 nm). Considering the influence of CCD noise and dark current on the experimental data, the uncertainty of absorptivity is about 0.10%. As a result, the extracted thermal conductivity along the zigzag direction, κzigzag = 4.45 ± 0.20 W·m−1·K−1, is about 1.09 times that along the armchair direction, κarmchair = 4.10 ± 0.13 W·m−1·K−1, showing an obvious thermal anisotropy. Such anisotropy may be attributed to the phonons having different mean free paths along zigzag and armchair directions [28], which has also been observed in BP and Ta2NiS5 [3,41].
As a common substrate for practical applications, SiO2/Si substrate might impact the thermal transport of the WTe2 flake above. Hence, we investigate the thermal properties of multilayer WTe2 supported on SiO2/Si substrate by electrical-heating Raman thermometry. Figure S3b shows the similar thickness of supported and suspended WTe2 samples. Figure 3a,b shows the optical image of supported WTe2 flake and the schematic diagram of electrical-heating Raman thermometry. We also characterize the electrical transport properties of supported WTe2 flake with the bias voltage changing from −3 V to 3 V (Figure S6). First, the orientation-dependent output curves along zigzag and armchair directions reflect good ohmic contact between WTe2 flake and metal electrodes. Furthermore, the electrical power density P (power density per length) gradually increases with the voltage ranging from 0 V to 3 V. According to previous literature [40], P = F × IDS = (IDS × VDS)/L, where F means the bias voltage per channel length and IDS means source/drain current. Therefore, we can calculate the thermal conductivity of supported WTe2 flake by analyzing the in situ Raman spectra under different bias voltages.
Figure S7 shows the temperature-dependent and bias voltage-dependent Raman spectra of supported WTe2 flake. The curve in Figure 3c indicates that the Raman peak position for 10A1 mode exhibits a linear redshift with increasing temperature (ranging from 213 K to 293 K, and the step is 10 K). According to Equation (4), for the 10A1 mode of supported WTe2 flake, χ T , z i g z a g = −0.0153 ± 0.0002 cm−1·K−1 and χ T , a r m c h a i r = −0.0119 ± 0.0004 cm−1·K−1 as in Figure 3c. Then, we analyze the Raman spectra when electrical bias is applied along different directions of supported WTe2 flake. As the bias voltage ranges from 0 V·μm−1 to 0.2 V·μm−1 (with 0.02 V/μm step), the electrical power density reaches 3 mW·μm−1. Figure 3d shows the Raman peak position of 10A1 mode as a function of electrical power density, where the redshift along the zigzag direction is 0.6225 cm−1 and that along the armchair direction is 0.223 cm−1.
Unlike suspended WTe2 sample, the heat dissipation of supported WTe2 flake should be described with a different model [40]
κ T + P κ T T 0 = 0 ,
where κ is the in-plane thermal conductivity along the direction parallel to the current; κ is that along the direction perpendicular to the current; P is the power density per channel length; and T0 is the temperature when the bias voltage is 0 V. Since κ is a constant, the equation above can be simplified to T = T 0 + P / κ ( 1 1 / cos h ( L / L H ) ) , where LH is the thermal healing length and L is the channel length of the WTe2 device. Considering L»LH, it can be rewritten as the following equation [40]:
T = T 0 + P / κ .
The lattice mismatch between WTe2 flake and SiO2/Si substrate may diminish interfacial heat dissipation [40,46], and Chen et al. suggest that the heat dissipation towards substrate may not enhance the thermal anisotropy ratio for supported samples [27]. Therefore, in-plane thermal dissipation is considered to be dominant here. Figure S8 shows the schematic of in-plane thermal dissipation in the WTe2 device. Joule heat is generated when the current flows along the armchair direction of the WTe2 device and is then dissipated preferentially along the zigzag direction with greater thermal conductivity. For the in-plane heat dissipation of the WTe2 device, the effect of the contact resistance (Rc) between WTe2 flake and electrodes on the electric power needs to be considered. When the channel length of a WTe2 device is 10 μm, the power is dissipated mostly along the channel (Pch = Pt − I2Rc = 58%Pt [40], where Pch means channel power, Pt means total power). Therefore, the power density in Equation (9) is mainly the channel power density Pch. According to temperature-dependent and bias voltage-dependent Raman spectra, we can calculate the temperature variation caused by channel power density for supported WTe2 flake.
Figure 4a,b shows the temperature as a function of channel power density for supported WTe2 flake. The extracted in-plane thermal conductivity for supported WTe2 flake is κzigzag = 62.17 ± 0.36 W·m−1·K−1 and κarmchair = 32.93 ± 0.82 W·m−1·K−1 through Equation (9). Ma et al. found that the thermal conductivity of monolayer WTe2 sample along different directions was 20 and 9 W·m−1·K−1 [47], respectively, and it was previously reported that the thermal conductivity of 50 ± 2.8 nm-thick WTe2 sample supported on SiO2/Si substrate can reach 28 ± 11.41 W·m−1·K−1 [48], which are within the same order of magnitude as our results. Considering that the low-frequency phonons mainly contribute to the thermal conductivity [28] and that the thermal conductivity along the zigzag direction of WTe2 flake is larger. More importantly, the thermal anisotropy ratio of supported WTe2 flake (κzigzag/κarmchair ≈ 1.89) is about 1.7 times that of suspended WTe2 flake (κzigzag/κarmchair ≈ 1.09). It is inferred that the low-frequency phonons with anisotropy [47] may be affected when WTe2 flake is supported on silicon substrate, allowing the variation of the thermal anisotropy ratio. Furthermore, since the mechanical properties of materials will be affected when a mechanical force is applied [29], it is possible that the interaction between silicon substrate and WTe2 flake has affected the thermal transport properties in an uneven manner, and the examined thermal anisotropy has changed accordingly. Table S1 shows that the enhanced thermal anisotropy caused by substrate coupling can also be observed in BP sheets. The thermal anisotropy ratio of supported BP sheets has been improved and is almost unaffected by the heat shunting along the substrate [27], pointing out that substrate coupling can modulate the thermal anisotropy ratio of anisotropic 2D materials and inspire advanced device design.

4. Conclusions

In summary, the thermal transport behavior of 50 nm-thick suspended and supported WTe2 flakes along zigzag/armchair axes was investigated by Raman thermometry, demonstrating that substrate coupling could improve the thermal anisotropy of WTe2 flake to a great extent. The thermal conductivity (κ) of suspended WTe2 flake is κzigzag = 4.45 W·m−1·K−1 and κarmchair = 4.10 W·m−1·K−1. For WTe2 device supported on SiO2/Si substrate, the thermal conductivity changes to κzigzag = 62.17 W·m−1·K−1 and κarmchair = 32.93 W·m−1·K−1. When supported on a silicon substrate, the low-frequency phonons and mechanical properties that contribute to thermal dissipation may have anisotropic effects on the different lattice orientations of WTe2 flake, generating a higher thermal anisotropy ratio of supported WTe2 flake (about 1.7 times that of suspended WTe2 flake). Our work reveals the anisotropic thermal dissipation properties of multilayer WTe2 devices supported on silicon substrates, which may be useful for the application of WTe2 and other 2D anisotropic materials in nanodevices.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13121817/s1, Figure S1: Schematic diagram of polarized-Raman configuration. Figure S2: The angle-dependent Raman spectra of WTe2 flake supported on the SiO2/Si substrate; Figure S3: The AFM images of suspended and supported WTe2 samples; Figure S4: Temperature-dependent and laser power-dependent Raman spectra of suspended WTe2 flake. Figure S5: The angle-dependent absorptivity of suspended WTe2 flake; Figure S6: The output IV curves of supported WTe2 flake; Figure S7: Temperature-dependent and bias voltage-dependent Raman spectra of supported WTe2 flake; Figure S8: Schematic diagram of thermal dissipation in WTe2 device. Table S1: The enhanced thermal anisotropy ratio (κzigzagarmchair) of BP and WTe2 caused by substrate coupling.

Author Contributions

Conceptualization, C.D.; investigation, M.F., X.L. and J.L.; data curation, M.F., X.L. and Y.C.; writing—original draft, M.F.; methodology, Y.S. and Y.W.; formal analysis, Y.Z. and G.P.; project administration, W.C. and X.-A.Z.; writing—review and editing, C.D.; supervision, C.D. and X.-A.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the Natural Science Foundation of China (Grants No. 11874423 and No. 12174321), and Independent Scientific Research Projects of NUDT.

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 authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ong, Z.-Y.; Cai, Y.; Zhang, G.; Zhang, Y.-W. Strong thermal transport anisotropy and strain modulation in single-layer phosphorene. J. Phys. Chem. C 2014, 118, 25272–25277. [Google Scholar] [CrossRef] [Green Version]
  2. Jain, A.; McGaughey, A.J. Strongly anisotropic in-plane thermal transport in single-layer black phosphorene. Sci. Rep. 2015, 5, 8501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Luo, Z.; Maassen, J.; Deng, Y.; Du, Y.; Garrelts, R.P.; Lundstrom, M.S.; Ye, P.D.; Xu, X. Anisotropic in-plane thermal conductivity observed in few-layer black phosphorus. Nat. Commun. 2015, 6, 8572. [Google Scholar] [CrossRef] [Green Version]
  4. Xiao, P.; Chavez-Angel, E.; Chaitoglou, S.; Sledzinska, M.; Dimoulas, A.; Sotomayor Torres, C.M.; El Sachat, A. Anisotropic thermal conductivity of crystalline layered SnSe2. Nano Lett. 2021, 21, 9172–9179. [Google Scholar] [CrossRef]
  5. Qiu, G.; Du, Y.; Charnas, A.; Zhou, H.; Jin, S.; Luo, Z.; Zemlyanov, D.Y.; Xu, X.; Cheng, G.J.; Ye, P.D. Observation of optical and electrical in-plane anisotropy in high-mobility few-layer ZrTe5. Nano Lett. 2016, 16, 7364–7369. [Google Scholar] [CrossRef] [Green Version]
  6. Xu, X.; Song, Q.; Wang, H.; Li, P.; Zhang, K.; Wang, Y.; Yuan, K.; Yang, Z.; Ye, Y.; Dai, L. In-plane anisotropies of polarized raman response and electrical conductivity in layered tin selenide. ACS Appl. Mater. Interfaces 2017, 9, 12601–12607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Wang, H.; Chen, M.-L.; Zhu, M.; Wang, Y.; Dong, B.; Sun, X.; Zhang, X.; Cao, S.; Li, X.; Huang, J. Gate tunable giant anisotropic resistance in ultra-thin GaTe. Nat. Commun. 2019, 10, 2302. [Google Scholar] [CrossRef] [Green Version]
  8. Liu, E.; Fu, Y.; Wang, Y.; Feng, Y.; Liu, H.; Wan, X.; Zhou, W.; Wang, B.; Shao, L.; Ho, C.-H. Integrated digital inverters based on two-dimensional anisotropic ReS2 field-effect transistors. Nat. Commun. 2015, 6, 6991. [Google Scholar] [CrossRef] [Green Version]
  9. Zhang, E.; Wang, P.; Li, Z.; Wang, H.; Song, C.; Huang, C.; Chen, Z.-G.; Yang, L.; Zhang, K.; Lu, S. Tunable ambipolar polarization-sensitive photodetectors based on high-anisotropy ReSe2 nanosheets. ACS Nano 2016, 10, 8067–8077. [Google Scholar] [CrossRef]
  10. Fei, R.; Faghaninia, A.; Soklaski, R.; Yan, J.-A.; Lo, C.; Yang, L. Enhanced thermoelectric efficiency via orthogonal electrical and thermal conductances in phosphorene. Nano Lett. 2014, 14, 6393–6399. [Google Scholar] [CrossRef] [Green Version]
  11. Das, P.; Di Sante, D.; Cilento, F.; Bigi, C.; Kopic, D.; Soranzio, D.; Sterzi, A.; Krieger, J.; Vobornik, I.; Fujii, J. Electronic properties of candidate type-II Weyl semimetal WTe2. A review perspective. Electron. Struct. 2019, 1, 014003. [Google Scholar] [CrossRef] [Green Version]
  12. Li, Y.; Liu, J.; Zhang, P.; Zhang, J.; Xiao, N.; Yu, L.; Niu, P. Electrical transport properties of Weyl semimetal WTe2 under high pressure. J. Mater. Sci. 2020, 55, 14873–14882. [Google Scholar] [CrossRef]
  13. Homes, C.; Ali, M.; Cava, R.J. Optical properties of the perfectly compensated semimetal WTe2. Phys. Rev. B 2015, 92, 161109. [Google Scholar] [CrossRef] [Green Version]
  14. Chen, Y.; Deng, C.; Wei, Y.; Liu, J.; Su, Y.; Xie, S.; Cai, W.; Peng, G.; Huang, H.; Dai, M. Thickness dependent anisotropy of in-plane Raman modes under different temperatures in supported few-layer WTe2. Appl. Phys. Lett. 2021, 119, 063104. [Google Scholar] [CrossRef]
  15. Pan, X.-C.; Chen, X.; Liu, H.; Feng, Y.; Wei, Z.; Zhou, Y.; Chi, Z.; Pi, L.; Yen, F.; Song, F. Pressure-driven dome-shaped superconductivity and electronic structural evolution in tungsten ditelluride. Nat. Commun. 2015, 6, 7805. [Google Scholar] [CrossRef] [Green Version]
  16. Kang, D.; Zhou, Y.; Yi, W.; Yang, C.; Guo, J.; Shi, Y.; Zhang, S.; Wang, Z.; Zhang, C.; Jiang, S. Superconductivity emerging from a suppressed large magnetoresistant state in tungsten ditelluride. Nat. Commun. 2015, 6, 7804. [Google Scholar] [CrossRef] [Green Version]
  17. Zhang, Q.; Zhang, R.; Chen, J.; Shen, W.; An, C.; Hu, X.; Dong, M.; Liu, J.; Zhu, L. Remarkable electronic and optical anisotropy of layered 1T’-WTe2 2D materials. Beilstein J. Nanotechnol. 2019, 10, 1745–1753. [Google Scholar] [CrossRef] [Green Version]
  18. Wang, Q.; Yesilyurt, C.; Liu, F.; Siu, Z.B.; Cai, K.; Kumar, D.; Liu, Z.; Jalil, M.B.; Yang, H. Anomalous photothermoelectric transport due to anisotropic energy dispersion in WTe2. Nano Lett. 2019, 19, 2647–2652. [Google Scholar] [CrossRef]
  19. Jha, R.; Onishi, S.; Higashinaka, R.; Matsuda, T.D.; Ribeiro, R.A.; Aoki, Y. Anisotropy in the electronic transport properties of Weyl semimetal WTe2 single crystals. AIP Adv. 2018, 8, 101332. [Google Scholar] [CrossRef] [Green Version]
  20. Zhou, W.; Chen, J.; Gao, H.; Hu, T.; Ruan, S.; Stroppa, A.; Ren, W. Anomalous and Polarization-Sensitive Photoresponse of Td-WTe2 from Visible to Infrared Light. Adv. Mater. 2019, 31, 1804629. [Google Scholar] [CrossRef]
  21. Torun, E.; Sahin, H.; Cahangirov, S.; Rubio, A.; Peeters, F. Anisotropic electronic, mechanical, and optical properties of monolayer WTe2. J. Appl. Phys. 2016, 119, 074307. [Google Scholar] [CrossRef] [Green Version]
  22. Xu, Z.; Luo, B.; Chen, Y.; Li, X.; Chen, Z.; Yuan, Q.; Xiao, X. High sensitivity and anisotropic broadband photoresponse of Td-WTe2. Phys. Lett. A 2021, 389, 127093. [Google Scholar] [CrossRef]
  23. Chen, X.-K.; Hu, X.-Y.; Jia, P.; Xie, Z.-X.; Liu, J. Tunable anisotropic thermal transport in porous carbon foams: The role of phonon coupling. Int. J. Mech. Sci. 2021, 206, 106576. [Google Scholar] [CrossRef]
  24. Liu, X.; Zhang, G.; Zhang, Y.-W. Surface-engineered nanoscale diamond films enable remarkable enhancement in thermal conductivity and anisotropy. Carbon 2015, 94, 760–767. [Google Scholar] [CrossRef]
  25. Qin, G.; Yan, Q.-B.; Qin, Z.; Yue, S.-Y.; Hu, M.; Su, G. Anisotropic intrinsic lattice thermal conductivity of phosphorene from first principles. Phys. Chem. Chem. Phys. 2015, 17, 4854–4858. [Google Scholar] [CrossRef] [Green Version]
  26. Liu, T.-H.; Chang, C.-C. Anisotropic thermal transport in phosphorene: Effects of crystal orientation. Nanoscale 2015, 7, 10648–10654. [Google Scholar] [CrossRef]
  27. Chen, J.; Chen, S.; Gao, Y. Anisotropy enhancement of thermal energy transport in supported black phosphorene. J. Phys. Chem. Lett. 2016, 7, 2518–2523. [Google Scholar] [CrossRef] [PubMed]
  28. Chen, Y.; Peng, B.; Cong, C.; Shang, J.; Wu, L.; Yang, W.; Zhou, J.; Yu, P.; Zhang, H.; Wang, Y. In-Plane Anisotropic Thermal Conductivity of Few-Layered Transition Metal Dichalcogenide Td-WTe2. Adv. Mater. 2019, 31, 1804979. [Google Scholar] [CrossRef] [PubMed]
  29. Bormashenko, E.; Pogreb, R.; Sutovsky, S.; Lusternik, V.; Voronel, A. Mechanical and thermodynamic properties of infrared transparent low melting chalcogenide glass. Infrared Phys. Technol. 2002, 43, 397–399. [Google Scholar] [CrossRef]
  30. Zhang, X.; Tan, Q.-H.; Wu, J.-B.; Shi, W.; Tan, P.-H. Review on the Raman spectroscopy of different types of layered materials. Nanoscale 2016, 8, 6435–6450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Wu, J.B.; Zhao, H.; Li, Y.; Ohlberg, D.; Shi, W.; Wu, W.; Wang, H.; Tan, P.H. Monolayer molybdenum disulfide nanoribbons with high optical anisotropy. Adv. Opt. Mater. 2016, 4, 756–762. [Google Scholar] [CrossRef] [Green Version]
  32. Wu, J.-B.; Lin, M.-L.; Cong, X.; Liu, H.-N.; Tan, P.-H. Raman spectroscopy of graphene-based materials and its applications in related devices. Chem. Soc. Rev. 2018, 47, 1822–1873. [Google Scholar] [CrossRef] [Green Version]
  33. Song, Q.; Pan, X.; Wang, H.; Zhang, K.; Tan, Q.; Li, P.; Wan, Y.; Wang, Y.; Xu, X.; Lin, M. The in-plane anisotropy of WTe2 investigated by angle-dependent and polarized Raman spectroscopy. Sci. Rep. 2016, 6, 29254. [Google Scholar] [CrossRef]
  34. Zhou, Z.; Cui, Y.; Tan, P.-H.; Liu, X.; Wei, Z. Optical and electrical properties of two-dimensional anisotropic materials. J. Semicond. 2019, 40, 061001. [Google Scholar] [CrossRef]
  35. Ribeiro, H.B.; Pimenta, M.A.; De Matos, C.J.; Moreira, R.L.; Rodin, A.S.; Zapata, J.D.; De Souza, E.A.; Castro Neto, A.H. Unusual angular dependence of the Raman response in black phosphorus. ACS Nano 2015, 9, 4270–4276. [Google Scholar] [CrossRef] [PubMed]
  36. Balandin, A.A.; Ghosh, S.; Bao, W.; Calizo, I.; Teweldebrhan, D.; Miao, F.; Lau, C.N. Superior thermal conductivity of single-layer graphene. Nano Lett. 2008, 8, 902–907. [Google Scholar] [CrossRef] [PubMed]
  37. Yan, R.; Simpson, J.R.; Bertolazzi, S.; Brivio, J.; Watson, M.; Wu, X.; Kis, A.; Luo, T.; Hight Walker, A.R.; Xing, H.G. Thermal conductivity of monolayer molybdenum disulfide obtained from temperature-dependent Raman spectroscopy. ACS Nano 2014, 8, 986–993. [Google Scholar] [CrossRef] [PubMed]
  38. Peimyoo, N.; Shang, J.; Yang, W.; Wang, Y.; Cong, C.; Yu, T. Thermal conductivity determination of suspended mono-and bilayer WS2 by Raman spectroscopy. Nano Res. 2015, 8, 1210–1221. [Google Scholar] [CrossRef]
  39. Chen, C.-C.; Li, Z.; Shi, L.; Cronin, S.B. Thermal interface conductance across a graphene/hexagonal boron nitride heterojunction. Appl. Phys. Lett. 2014, 104, 081908. [Google Scholar] [CrossRef] [Green Version]
  40. Wei, Y.; Deng, C.; Zheng, X.; Chen, Y.; Zhang, X.; Luo, W.; Zhang, Y.; Peng, G.; Liu, J.; Huang, H. Anisotropic in-plane thermal conductivity for multi-layer WTe2. Nano Res. 2022, 15, 401–407. [Google Scholar] [CrossRef]
  41. Su, Y.; Deng, C.; Liu, J.; Zheng, X.; Wei, Y.; Chen, Y.; Yu, W.; Guo, X.; Cai, W.; Peng, G. Highly in-plane anisotropy of thermal transport in suspended ternary chalcogenide Ta2NiS5. Nano Res. 2022, 15, 6601–6606. [Google Scholar] [CrossRef]
  42. Liu, X.-L.; Zhang, X.; Lin, M.-L.; Tan, P.-H. Different angle-resolved polarization configurations of Raman spectroscopy: A case on the basal and edge plane of two-dimensional materials. Chin. Phys. B 2017, 26, 067802. [Google Scholar] [CrossRef]
  43. Loudon, R. The Raman effect in crystals. Adv. Phys. 2001, 50, 813–864. [Google Scholar] [CrossRef]
  44. Jana, M.K.; Singh, A.; Late, D.J.; Rajamathi, C.R.; Biswas, K.; Felser, C.; Waghmare, U.V.; Rao, C. A combined experimental and theoretical study of the structural, electronic and vibrational properties of bulk and few-layer Td-WTe2. J. Phys. Condens. Matter 2015, 27, 285401. [Google Scholar] [CrossRef] [PubMed]
  45. Calizo, I.; Balandin, A.; Bao, W.; Miao, F.; Lau, C. Temperature dependence of the Raman spectra of graphene and graphene multilayers. Nano Lett. 2007, 7, 2645–2649. [Google Scholar] [CrossRef] [PubMed]
  46. Vaziri, S.; Yalon, E.; Muñoz Rojo, M.; Suryavanshi, S.V.; Zhang, H.; McClellan, C.J.; Bailey, C.S.; Smithe, K.K.; Gabourie, A.J.; Chen, V. Ultrahigh thermal isolation across heterogeneously layered two-dimensional materials. Sci. Adv. 2019, 5, eaax1325. [Google Scholar] [CrossRef] [Green Version]
  47. Ma, J.; Chen, Y.; Han, Z.; Li, W. Strong anisotropic thermal conductivity of monolayer WTe2. 2D Mater. 2016, 3, 045010. [Google Scholar] [CrossRef]
  48. Laxmi, V.; Basu, N.; Nayak, P.K. Substrate dependent thermal conductivity in Td-WTe2 using micro-Raman spectroscopy. J. Raman Spectrosc. 2023, 54, 76–83. [Google Scholar] [CrossRef]
Figure 1. The crystal structure and angle-dependent Raman spectra of WTe2. (a) Top view and side view of the WTe2 crystal structure. (b) Contour map of angle-dependent Raman intensity for three A1 modes under parallel-polarized configuration. (cf) Polar plots and fit curves of angle-dependent Raman intensity for 3A2, 4A1, 8A1, and 10A1 peaks.
Figure 1. The crystal structure and angle-dependent Raman spectra of WTe2. (a) Top view and side view of the WTe2 crystal structure. (b) Contour map of angle-dependent Raman intensity for three A1 modes under parallel-polarized configuration. (cf) Polar plots and fit curves of angle-dependent Raman intensity for 3A2, 4A1, 8A1, and 10A1 peaks.
Nanomaterials 13 01817 g001
Figure 2. Temperature- and laser power-dependent in situ Raman spectra of suspended WTe2 sample. (a) Optical microscopy image of suspended WTe2 flake. (b) Schematic diagram of the experimental setup for Raman thermometry with laser-heating (the inset is the laser-heating process of suspended WTe2 flake). (c,d) The linear fitting of the 10A1 Raman peak position as a function of temperature (c) and laser power (d) along zigzag and armchair directions.
Figure 2. Temperature- and laser power-dependent in situ Raman spectra of suspended WTe2 sample. (a) Optical microscopy image of suspended WTe2 flake. (b) Schematic diagram of the experimental setup for Raman thermometry with laser-heating (the inset is the laser-heating process of suspended WTe2 flake). (c,d) The linear fitting of the 10A1 Raman peak position as a function of temperature (c) and laser power (d) along zigzag and armchair directions.
Nanomaterials 13 01817 g002
Figure 3. Temperature- and bias voltage-dependent in situ Raman spectra of supported WTe2 sample (WTe2 device supported on SiO2/Si substrate). (a) Optical image of supported WTe2 flake. (b) Schematic diagram of Raman thermometry with electrical heating (the inset depicts heat dissipation process in supported WTe2 flake). (c,d) The linear fitting of the 10A1 Raman peak position as a function of temperature (c) and electrical power density (d) along both zigzag and armchair directions.
Figure 3. Temperature- and bias voltage-dependent in situ Raman spectra of supported WTe2 sample (WTe2 device supported on SiO2/Si substrate). (a) Optical image of supported WTe2 flake. (b) Schematic diagram of Raman thermometry with electrical heating (the inset depicts heat dissipation process in supported WTe2 flake). (c,d) The linear fitting of the 10A1 Raman peak position as a function of temperature (c) and electrical power density (d) along both zigzag and armchair directions.
Nanomaterials 13 01817 g003
Figure 4. The effect of substrate coupling on thermal anisotropy. The fitting curve of temperature versus channel power density along zigzag (a) and armchair (b) directions for supported WTe2 device. (c) In-plane thermal conductivity of suspended and supported WTe2 samples along zigzag and armchair directions.
Figure 4. The effect of substrate coupling on thermal anisotropy. The fitting curve of temperature versus channel power density along zigzag (a) and armchair (b) directions for supported WTe2 device. (c) In-plane thermal conductivity of suspended and supported WTe2 samples along zigzag and armchair directions.
Nanomaterials 13 01817 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fang, M.; Liu, X.; Liu, J.; Chen, Y.; Su, Y.; Wei, Y.; Zhou, Y.; Peng, G.; Cai, W.; Deng, C.; et al. Improved Thermal Anisotropy of Multi-Layer Tungsten Telluride on Silicon Substrate. Nanomaterials 2023, 13, 1817. https://doi.org/10.3390/nano13121817

AMA Style

Fang M, Liu X, Liu J, Chen Y, Su Y, Wei Y, Zhou Y, Peng G, Cai W, Deng C, et al. Improved Thermal Anisotropy of Multi-Layer Tungsten Telluride on Silicon Substrate. Nanomaterials. 2023; 13(12):1817. https://doi.org/10.3390/nano13121817

Chicago/Turabian Style

Fang, Mengke, Xiao Liu, Jinxin Liu, Yangbo Chen, Yue Su, Yuehua Wei, Yuquan Zhou, Gang Peng, Weiwei Cai, Chuyun Deng, and et al. 2023. "Improved Thermal Anisotropy of Multi-Layer Tungsten Telluride on Silicon Substrate" Nanomaterials 13, no. 12: 1817. https://doi.org/10.3390/nano13121817

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