Next Article in Journal
Impact of NaHCO3/Na2CO3 Buffer Reagent on Mitigating the Corrosion of C110 Steel in Water-Based Annulus Protection Fluid at Ultrahigh Temperature
Previous Article in Journal
Optimization of Hybrid Composite–Metal Joints: Single Pin
Previous Article in Special Issue
Observation of Boson Peak of Fragile Baltic Amber Glass by Terahertz Time-Domain Spectroscopy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Interactions of Terahertz Photons with Phonons of Two-Dimensional van der Waals MoS2/WSe2/MoS2 Heterostructures and Thermal Responses

1
State Key Laboratory of Optoelectronic Materials and Technologies, Guangdong Province Key Laboratory of Display Material and Technology, School of Electronics and Information Technology, Sun Yat-sen University, Guangzhou 510275, China
2
State Key Laboratory of Integrated Chips and Systems, Frontier Institute of Chip and System, Fudan University, Shanghai 200433, China
3
State Key Laboratory of Integrated Chips and Systems, School of Information Science and Technology, Fudan University, Shanghai 200433, China
4
Guangdong Province Key Laboratory of Display Material and Technology, Center for Neutron Science and Technology, School of Physics, Sun Yat-sen University, Guangzhou 510275, China
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(7), 1665; https://doi.org/10.3390/ma18071665
Submission received: 10 March 2025 / Revised: 31 March 2025 / Accepted: 2 April 2025 / Published: 4 April 2025
(This article belongs to the Special Issue Terahertz Vibrational Spectroscopy in Advanced Materials)

Abstract

:
The interaction between terahertz (THz) photons and phonons of materials is crucial for the development of THz photonics. In this work, typical two-dimensional (2D) van der Waals (vdW) transition metal chalcogenide (TMD) layers and heterostructures are used in THz time-domain spectroscopy (TDS) measurements, low-wavenumber Raman spectroscopy measurements, calculation of 2D materials’ phonon spectra, and theoretical analysis of thermal responses. The TDS results reveal strong absorption of THz photons in the frequency range of 2.5–10 THz. The low-wavenumber Raman spectra show the phonon vibration characteristics and are used to establish phonon energy bands. We also set up a computational simulation model for thermal responses. The temperature increases and distributions in the individual layers and their heterostructures are calculated, showing that THz photon absorption results in significant increases in temperature and differences in the heterostructures. These give rise to interesting photothermal effects, including the Seebeck effect, resulting in voltages across the heterostructures. These findings provide valuable guidance for the potential optoelectronic application of the 2D vdW heterostructures.

1. Introduction

Terahertz waves are electromagnetic radiation with a frequency range of 0.1–10 THz and a wavelength of 0.03–3 mm. Due to their frequency band lying between the radio wave region and the optical region, they possess dual characteristics of both macroscopic classical theory and microscopic quantum theory, thus exhibiting unique application value [1,2]. The unique frequency range of THz spectroscopy (covering characteristic frequencies of molecular vibrations and lattice vibrations) makes it a powerful tool for characterizing the physical properties of materials, such as measuring complex permittivity, dynamic conductivity, and studying microscopic processes like low-energy electron transitions and carrier dynamics. Furthermore, THz technology demonstrates broad application prospects in multiple cutting-edge fields, including biomedical imaging, security inspection, 6G communications, and materials science [3,4,5,6,7]. Its potential for high-frequency communications is especially regarded as one of the core technologies to break through the communication rate bottleneck in the future [8,9,10].
The main challenges and development directions of THz technology lie in achieving high spatial resolution, broadband spectral coverage, high signal-to-noise ratio, high efficiency, and system integrability. In this context, exploring emerging materials, such as 2D materials, heterostructures, and integration methods, is attracting increasing attention [11,12].
Two-dimensional materials have the advantage of being easily integrated with traditional optoelectronic materials, systems, and devices, bringing novel phenomena and potential applications in the THz field [13,14]. In particular, the vdW heterostructures are stacked from different layered materials, with layers bound together by vdW forces, free from the constraints of lattice matching between the material and the substrate. At the same time, since electrons in atomically thin layers are directly exposed, different quantum states between layers can interact and couple in ways that are not possible in other systems, enabling the ultrafast transfer of electrons and holes across layers within an atomically spaced distribution. Compared to individual 2D atomic crystal materials, 2D vdW heterostructures exhibit significantly enhanced performance advantages over monolayer materials [15,16,17].
The research directions regarding the coupling between THz photons and phonons in 2D materials, such as black phosphorus, MoS2, and WSe2, involve several aspects [18,19,20]. (1) Phonon polaritons: the coupling of THz photons with phonons to form quasiparticles. (2) Thermal effects and energy conversion: THz wave energy dissipation due to the conversion of electromagnetic wave energy into lattice vibrations, i.e., the phonons. (3) Nonlinear optical response: nonlinear effects of phonons induced by strong-field THz pulses. Phonons, as the quantized modes of atomic oscillations within a lattice, play a crucial role in the thermal and electrical properties of materials. Optical phonons can be generated through optical excitation, and manipulating phonons in semiconductors allows for the control of material properties at the quantum level. Traditional methods for phonon excitation typically rely on the illumination of the materials using infrared or visible light sources. However, despite the unique characteristics of THz radiation, the direct excitation of coherent optical phonons has faced the challenge of energy mismatch.
Currently, there is limited research on the interactions between photons and phonons of transition metal dichalcogenide (TMD) materials and their heterostructures. It is mainly because monolayer TMD materials, such as MoS2, exhibit extremely high transmittance to THz waves, and as the incident wave frequency increases, the absorption rate of THz waves gradually decreases [21,22,23,24]. Zhu et al. [25] fabricated a broadband THz wave absorber based on MoS2. Through phase modulation, broadband coherent absorption can be achieved, with the coherent absorption rate being adjustable from 1.57% to 99.97%. This research result indicates that with certain scientific modulation methods, TMD materials can also exhibit excellent absorption performance for THz wave radiation. H. Buss et al. [26] studied the broadband transient conductivity and the low-energy excitations in multilayer MoS2. When the sample was subjected to radiation with a frequency exceeding 2 THz at low temperatures, a significant enhancement in conductivity was observed. This temperature-dependent behavior suggests that multilayer MoS2 possesses strong high-frequency optoelectronic responses in the THz regime, particularly at frequencies above 2 THz.
In this work, 2D vdW MoS2/WSe2/MoS2 heterostructures with high-quality atomic-level crystalline interfaces were fabricated using an optical micromechanical arm alignment technique. THz time-domain spectroscopy measurements revealed significant absorption of different structurally layered materials in the range of 0.5 to 15 THz, attributed to intraband absorption by excess free electrons in few-layer MoS2, WSe2, and their heterostructures. Furthermore, based on low-wavenumber Raman spectroscopy results, we propose a THz photon-to-phonon transition mechanism. When the few-layer MoS2 and WSe2 in the 2D heterostructure absorb photons with energies that are integer multiples of their own phonon energies, they release phonons corresponding to those energy multiples. These newly released phonons participate in phonon scattering, leading to thermal dissipation and subsequently causing an increase in the temperature of the heterostructure. Such a mechanism is corroborated by theoretical calculations of the phonon spectrum of MoS2 and WSe2. Additionally, by establishing a thermal conduction model for the heterostructure, we systematically investigated the thermodynamic changes in the heterostructure under continuous wave irradiation from 0.5 to 15 THz. The simulation results indicate that the overall temperature of the heterostructure increases and reveal the presence of the Seebeck effect arising from the temperature gradient.

2. Experiments and Methods

2.1. MoS2/WSe2/MoS2 Heterosruacture Fabrication

First, the bottom n-type MoS2 few layers were mechanically exfoliated using 3M adhesive tape on the (300 nm/500 μm) SiO2/Si substrates. Then, with the assistance of polydimethylsiloxane (PDMS) for transfer, the p-type WSe2 and another n-type MoS2 few-layers were exfoliated and sequentially transferred onto the bottom MoS2 flake using optical micromanipulator alignment technology. To avoid a short circuit, the top MoS2 layer should not come into direct contact with the bottom MoS2 layer.

2.2. Analytical Methods

The optical images were recorded for the fabrication of the vertically stacked vdW MoS2/WSe2/MoS2 heterostructure using an optical microscope (Olympus: BX51, Tokyo, Japan) equipped with a 50× objective lens (N.A. = 0.50) and a quartz-tungsten-halogen lamp (100 W), operating in reflection mode.
Raman and PL spectra were obtained at room temperature using a Raman spectrometer (Renishaw: InVia Reflex, Wotton-under-Edge, UK) equipped with a CCD detector and a 50× objective (Leica, Wetzler, Germany, model 11566074, N.A. = 0.75). The measurements were performed in a backscattering configuration. The laser used was a 532 nm solid-state laser (second harmonic from an Nd:YAG laser, model: RL532C, Renishaw), with a power of 52 mW. The laser was vertically incident on the sample surface. Both Raman and PL signals were then collected from the same side through reflection. The spectrometer was equipped with a diffraction grating of 2400 l/mm, providing high spectral resolution. The spectrometer covered a broad spectral range from 5 cm−1 to 3200 cm−1, ensuring the detection of both low- and high-frequency Raman modes. Rayleigh scattering was effectively suppressed using a Renishaw ultra-narrow-band notch filter (supplied by Renishaw, model: VBG) integrated into the spectrometer system, which facilitated the reliable acquisition of low-wavenumber Raman spectra.
The peak selection criteria were based on signal-to-noise (SNR) analysis:
SNR = I peak I baseline σ noise
where Ipeak represents the intensity of the low-wavenumber Raman peak, Ibaseline is the baseline intensity, and σnoise is the standard deviation of the noise background. In our study, when the SNR exceeded 3, we claimed that the signal was a real Raman mode rather than noise.
The atomic force microscope study (AFM) for structural and morphological measurements was conducted using a Veeco Dimension-Icon system scanning probe microscope (Bruker Inc.: Dimension fastscan bio, Billerica, MA, USA) with SiN probes (Scanasyst-Air, calibrated spring constants of 0.3–0.5 N m−1 and a nominal tip radius of 2–5 nm). Topographic characterization was carried out in Peak Force Tapping mode under ambient conditions.
Time-resolved THz spectroscopy characterizations were conducted using a fiber-coupled THz time-domain spectroscopy (THz-TDS) system (TeraFlash pro, Toptica, Pittsford, NY, USA). In order to analyze the THz absorption properties of the few-layer MoS2, WSe2, and their vertically stacked heterostructures prepared on a high-resistance Si substrate (500-μm thickness covered by a 300-nm SiO2 layer, ρ~20,000 Ω·cm), THz-TDS spectroscopy measurements were conducted at room temperature. The large-area MoS2 and WSe2 flakes were purchased from the manufacturer (Shenzhen SixCarbon Technology Co., Ltd., Shenzhen, China) and were grown using the CVD method. The large-area MoS2/WSe2/MoS2 heterostructure was prepared using a wet chemical etching process (the detailed method is shown in the Supporting Information Section S1). The experimental samples for the TDS measurements should be at least 1.5 cm × 1.5 cm in area. For a specific TDS measurement, the sample was attached to a hollow iron plate for testing, and the THz wave focused on the sample with a spot size of 3 mm. The test chamber was filled with high-purity N2 to ensure that the air humidity was less than 10%. The TDS spectrum for the sample was recorded using a time constant of 300 ms, an integration time of 900 ms, a step size of 1.5 µm for distance changes between the excitation and probe beams, and a scan point number of 700.
The density functional theory (DFT) calculations were performed using the Vienna Ab Initio Simulation package (VASP) [27,28]. Specifically, in order to unveil the origins of the phonons, we calculated the electronic structure and force constant matrices using DFT implemented in the VASP and phonopy [29,30] package. The DFT calculations used the generalized gradient approximation (GGA) [31] of Perdew–Burke–Ernzerhof (PBE) [32] as the exchange-correlation function, and the plane wave cutoff energy was set to 520 eV, and the k-point mesh was a Gamma-centered 6 × 6 × 1 Monkhorst–Pack grid. Rigorous convergence tests were conducted to determine the computational parameters, as we had previously done in calculations of the dielectric functions of vdW polar crystals [33]. Accordingly, the plane wave cutoff energy was set to 520 eV, and the k-point mesh was a Gamma-centered 6 × 6 × 1 Monkhorst–Pack grid. The energy convergence accuracy was set to 1 × 10−10 eV, while the force convergence criterion for each atom was set to 1 × 10−8 eV/Å. With these parameters, we calculated the Born effective charge tensor [34] using the DFT linear response scheme combined with the iterative Green’s function method of density functional perturbation theory and then obtained the phonon spectrum in a 3 × 3 × 2 supercell combined with the second-order force constants calculated using the phonopy package.
The finite-element method based on triangular prismatic elements was conducted using an in-house-developed solver. This method demonstrates a robust capability in addressing the multiscale characteristics inherent to geometric models. The transient heat conduction equation was solved via an implicit backward differentiation formula with a stabilized time-stepping strategy (Δt: 0.001–1 s). Material properties include thermal conductivity, density, and specific heat capacity. Boundary conditions enforced natural convective heat transfer at the air and large-area silicon substrate interfaces (h = 5–25 W·m−2·K−1), with prescribed thermal source constraints additionally applied. Temperature-time curves at the heterostructure centers were extracted through linear interpolation in the course of post-processing.

3. Results and Discussion

In our study, we focused on the 2D vdW MoS2/WSe2/MoS2 heterostructure, which is uniquely suited for optoelectronic applications due to its synergistic material properties and interfacial advantages. Both materials are TMDs with complementary electronic characteristics: MoS2 provides high electron mobility (~200 cm2/V·s), while WSe2 exhibits superior hole mobility (>100 cm2/V·s), enabling balanced charge transport. Their bandgaps (1.2–1.9 eV for MoS2, 1.2–1.7 eV for WSe2) form a type-II staggered alignment, enhancing photocarrier separation and suppressing recombination—critical for photodetectors and THz applications. The type-II band alignment in MoS2/WSe2 outperforms alternatives like MoS2/WS2 (type-I) or MoSe2/WSe2 (type-I), which suffer from inefficient charge separation. Additionally, the small lattice mismatch (3.8%) minimizes interfacial strain and defects, ensuring high-quality heterostructures with efficient phonon transport. In contrast, other TMD pairs (e.g., MoS2/MoTe2, WSe2/WS2) exhibit larger mismatches (>4%), degrading device performance. Experimentally, MoS2 and WSe2 are air-stable and mechanically robust, enabling reliable exfoliation and precise stacking. This contrasts with air-sensitive materials like black phosphorus or ReS2, which complicate fabrication. The trilayer n-p-n configuration further enhances optoelectronic performance, facilitating strong in-plane p-n junctions and efficient carrier extraction. Crucially, the heterostructure exhibits strong THz photon-phonon coupling, as evidenced by low-wavenumber Raman modes and phonon simulations. THz irradiation excites interlayer vibrations, converting photon energy into lattice heating—a key mechanism for THz detection. This, combined with its thermal and electronic properties, makes MoS2/WSe2/MoS2 an optimal platform for advanced optoelectronics.
The schematic illustration of the MoS2/WSe2/MoS2 heterostructures is depicted in Figure 1a–f,i. The space-cell structure model of MoS2 and WSe2 is shown in Figure 1a,b. The diagrams of the preparation processes of the heterostructure are displayed in Figure 1d–f. The cross-sectional view and top view of the vertical MoS2/WSe2/MoS2 heterostructure are shown in Figure 1c,i. The corresponding optical microscope images captured after each transfer procedure during the whole fabrication procedure are given in Figure 1j–l. The Raman spectra of the MoS2, WSe2, MoS2/WSe2, and MoS2/WSe2/MoS2 heterostructures are presented in Figure 1g. Distinct Raman peaks are observed at ~383.3 cm−1 and 408.5 cm−1, corresponding to the vibration modes E 2 g 1 (in-plane) and A1g (out-of-plane) of MoS2, respectively. The wavenumber difference Δd of these two vibration modes can simply reflect the layer number of the MoS2 [35]. The Δd values of the bottom and top MoS2 are 25.2 cm−1 and 23.6 cm−1, respectively. Both values suggest more than five layers of the two MoS2 layers. As for the Raman peaks of WSe2, a strong peak at 250.2 cm−1 is related to the A1g mode, and a small signature at 306.8 cm−1 may be assigned to the normally inactive B2g mode. Notably, the Raman peaks of MoS2 in the overlapped region exhibit slight broadening and shifting, indicating charge transfer across the interfaces. Consequently, a depletion layer was formed at the MoS2/WSe2/MoS2 vdW heterostructure [36].
Figure 1h depicts the photoluminescence (PL) spectra of the vertically stacked heterostructure. Two peaks of the top MoS2 layer are observed at 1.84 eV (673.9 nm) and 2.0 eV (620 nm), respectively. They correspond to the A1 and B1 direct excitonic transitions with the energy splitting due to valence band spin-orbital coupling [37]. The A1 and B1 peaks of the bottom MoS2 layers appear to be slightly shifted to 1.83 eV (677.6 nm) and 1.99 eV (623.1 nm), respectively. This suggests that the bottom and top MoS2 layers have different layer numbers. The middle WSe2 layer exhibits a PL peak at 1.62 eV (765.4 nm), corresponding to its indirect band gap [38]. However, the PL intensity is strongly suppressed in the overlapped region of the three layers (Figure 1h, green line), which can be attributed to the radiative recombination of the spatially separated electron-hole pairs in different layers.
AFM images recorded at three different regions (marked with colored boxes in Figure 1l) are shown in Figure 1m–o. They clearly display the atomically flat surfaces of the mechanically exfoliated MoS2 and WSe2 layers. AFM height profiles extracted from each layer of MoS2/WSe2/MoS2 are shown in the insets. The results display the heights of the three layers, i.e., 4.85 nm, 3.12 nm, and 4.25 nm, respectively, for the bottom MoS2, middle WSe2, and top MoS2. These thicknesses are consistent with the Raman results shown in Figure 1g. AFM results demonstrate that the atoms within the layers are uniformly arranged, and the interfaces of the stacked structure are flat without any dislocations or defects, possessing a high-quality atomic-level crystalline interface [39].
Figure 2a–d depicts the optical micrographs of large-area MoS2, WSe2, MoS2/WSe2, and MoS2/WSe2/MoS2, respectively. A schematic illustration of the TDS spectroscopy is shown in Figure 2e. Figure 2f displays THz-TDS results from the blank substrate, MoS2, WSe2, MoS2/WSe2, and MoS2/WSe2/MoS2. The time-domain peak and valley of the THz wave transmitting through large-area vdW MoS2, WSe2, and their heterostructures on high-resistance Si are ~5 ps. Figure 2i displays THz-TDS results from the N2 atmosphere. The time-domain peak position of the THz signal transmitting through N2 absorption is ~1 ps.
Figure 2g,h presents a detailed analysis of the time-domain spectra (shown around the 5 ps region in Figure 2f) for different 2D vdW MoS2, WSe2, and their heterostructures. The strength of THz absorption is represented by the difference between the adjacent peaks (or valleys) of the time-domain signal. As the THz signal passes through WSe2, MoS2, MoS2/WSe2, and MoS2/WSe2/MoS2, the signal intensity gradually decreases. The THz electric field transmitting through the blank substrate, MoS2/substrate, WSe2/substrate, MoS2/WSe2/substrate, and MoS2/WSe2/MoS2/substrate is defined as Esub(τ), EM(τ), EW(τ), EM/W(τ), and EM/W/M(τ), respectively. The maximum values of EM(τ), EW(τ), EM/W(τ), and EM/W/M(τ) respectively show reductions of 3.03%, 2.5%, 3.9%, and 5.1% compared to Esub(τ). This attenuation of the THz transmission arises from absorption in the few-layer MoS2 and WSe2, as well as their heterostructures, either due to the intraband transition of excess free electrons [40] or phonon excitation, as discussed below.
The frequency-domain spectrum is obtained by conducting a Fourier transform of the TDS spectrum. In the range of 0.5–20 THz, clear spectral lines in the frequency domain can be observed, as shown in Figure 2j. The enlarged frequency domain spectrum from 1.5 to 4.5 THz (regions marked by blue dotted boxes in Figure 2j) is presented in Figure 2k. The THz wave amplitude in the frequency domain exhibits a gradual decrease across the material sequence: WSe2, MoS2, and MoS2/WSe2 to MoS2/WSe2/MoS2.
The transmittance coefficient of the material can be calculated according to [41]:
Transmittance = E sample ( τ ) 2 E substrate ( τ ) 2
where Esample (τ) is the amplitude of the sample signal, and Esubstrate (τ) is the amplitude of the substrate signal. The extinction coefficient of the material can be calculated according to
Extinction   coefficient = 1 Transmittance
The extinction coefficient varies across the frequency ranges of 0.5–15 THz, 0.5–10 THz, and 0.5–5 THz, as shown in Figure 2l–n. In the THz frequency range of 0.5–2 THz, the extinction coefficient of few-layer MoS2 and WSe2 is higher than that of the heterostructure. In the range of 2–5 THz, the extinction of THz signals of MoS2/WSe2 and MoS2/WSe2/MoS2 heterostructures gradually increases, surpassing the values observed in few-layer MoS2 and WSe2. Significant signal fluctuations emerge for the frequency range above 5 THz. In this spectral range, the extinction coefficient of few-layer MoS2, WSe2, and MoS2/WSe2 increases gradually, whereas that of MoS2/WSe2/MoS2 declines. According to the data in Figure 2n, the extinction coefficient of WSe2, MoS2, MoS2/WSe2, and MoS2/WSe2/MoS2 was 30%, 31.8%, 32.1%, and 32.6% at 2.52 THz.
In the traditional photon detection mechanism, the photon energy must exceed the semiconductor band gap energy to enable electron transitions from the valence band to the conduction band, thereby altering the conductivity [42]. However, the THz band from 0.1 to 10 THz corresponds to a photon energy of 0.414 to 41.4 meV, which is much smaller than the band gap of MoS2 (1.83 eV) and WSe2 (1.62 eV). The absorption of the THz wave in the vdW crystals and heterostructures may, therefore, arise from phonon excitations by the THz wave. For THz waves with frequencies less than 1 THz, the corresponding wavenumber (the reciprocal of the wavelength) is less than 33 cm−1. This suggests that it is possible to explore the coupling between THz photons and phonons in TMDs and their heterostructures using low-wavenumber Raman spectroscopy, which can reveal the phonon modes in the range of 0–100 cm−1.
A 532-nm laser is used to obtain the low-wavenumber Raman spectra of MoS2, WSe2, MoS2/WSe2, and MoS2/WSe2/MoS2. According to the SNR of the spectral bumps, different Raman peaks can be assigned (Tables S1 and S2 in the Supporting Information). Figure 3a shows the Raman spectra of a few layers of MoS2. Two peaks at 10.23 cm−1 and 28.01 cm−1 can be observed, which respectively correspond to frequencies of 0.31 THz and 0.84 THz. Figure 3b presents spectra obtained from few-layer WSe2. Peaks at 13.98 cm−1, 20.83 cm−1, 29.03 cm−1, 35.85 cm−1, 61.9 cm−1, 75.32 cm−1, and 97.01 cm−1, corresponding to frequencies of 0.42 THz, 0.64 THz, 0.87 THz, 1.08 THz, 1.86 THz, 2.26 THz, and 2.91 THz, respectively, can be recorded. For the MoS2/WSe2 heterostructure, five peaks, located at 9.88 cm−1 (0.3 THz), 13.98 cm−1 (0.42 THz), 23.56 cm−1 (0.71 THz), 27.66 cm−1 (0.83 THz), and 35.85 cm−1 (1.08 THz), can be recorded (Figure 3c). Figure 3d depicts the low-wavenumber Raman spectra of the MoS2/WSe2/MoS2 heterostructure, with peaks at 9.88 cm−1, 13.99 cm−1, 20.83 cm−1, and 27.84 cm−1, corresponding to frequencies of 0.3 THz, 0.42 THz, 0.63 THz, and 0.84 THz, respectively. Table 1 provides a summary of the above results.
Based on the results of the above low-wavenumber Raman spectroscopy, we discover several typical phonon modes of MoS2 and WSe2, corresponding to the following frequencies:
MoS2: 10.23 cm−1 to 0.31 THz (ν1), 28.01 cm−1 to 0.84 THz (ν2).
WSe2: 13.98 cm−1 to 0.42 THz (ν3), 20.83 cm−1 to 0.64 THz (ν4).
For a 2.52-THz photon, these frequencies correspond respectively to ~8 (ν1), 3 (ν2), 6 (ν3), and ~4 (ν4). According to these observations, we propose a mechanism for THz absorption in the vdW crystals and heterostructures. Specifically, when a THz wave irradiates the vdW crystals, the vdW lattices will absorb the photons by emitting corresponding optical phonons, provided the photon energy equals an integer multiple of the optical phonon energies in the crystals. Based on this mechanism and the phonon energies extracted from the low-wavenumber Raman spectroscopy measurements, we assume that the TMDs used in our study may strongly absorb THz photons with energies that are integer multiples of ν1 to ν4. Figure 3e illustrates this photon-to-phonon transition process in our heterostructure. Specifically, when the MoS2 lattice absorbs THz photons at 2.52 THz, eight phonons (ν1) or three phonons (ν2) can be released. Similarly, when the WSe2 lattice absorbs a 2.52-THz photon, six phonons (ν3) and four phonons (ν4) can be released. These newly released phonons enhance the lattice vibration, thus leading to an increment in the crystal temperature. This is the main mechanism governing the THz wave absorption in our 2D MoS2/WSe2/MoS2 heterostructures. It should be noted that bulk MoS2 and WSe2 are centrosymmetric crystals in principle. In such systems, Raman-active phonon modes are infrared-inactive, meaning they cannot be directly excited by THz photons. However, the samples in our experiment consist of few-layer structures. When the layer number is odd, the crystal loses centrosymmetry, rendering most phonon modes both Raman- and infrared-active. In our study, the layer numbers are 5 for the top MoS2 layer, 7 for the bottom MoS2 layer, and 5 for the middle WSe2 layer. Consequently, THz photon absorption in these samples is still governed by the aforementioned multi-phonon excitation processes.
Through theoretical calculations, we plotted the phonon dispersion of monolayer and bulk MoS2 and WSe2 in Figure 4. Combined with Tables S3 and S4 in the Supporting Information, the corresponding phonon frequencies of MoS2 and WSe2 with different thicknesses at the Γ points are shown. There are three acoustic branches, whose frequencies are zero at the Γ point in the phonon dispersion of each crystal. In particular, for the bulk crystals, three branches are close to the acoustic ones in the dispersion curves. The vibrations of atoms in each of these three branches are in phase within the same layer but are oppositely phased in adjacent layers. These are new optical phonons that are absent in monolayer crystals.
The values of the new optical phonon branches in bulk MoS2 (Figure 4c,d) at the Γ point are about 0.42 THz and 0.71 THz, respectively. These agree with the results for low-wavenumber Raman experiments of MoS2 (0.31 THz and 0.84 THz). The values of the new optical phonon branches in bulk WSe2 at the Γ point are 0.38 THz and 0.61 THz, respectively (Figure 4e,f). These also agree with the results for low-wavenumber Raman experiments of WSe2 (0.42 THz and 0.64 THz).
Based on the previous TDS results, few-layer MoS2, WSe2, and their heterostructures all have strong absorption of THz waves. After absorption, they will release phonons and lead to an increase in crystalline temperature. We then simulate the thermal responses of different vdW crystals, accounting for heat dissipation from the large-area silicon substrate and air. We employ an in-house thermal simulation solver, utilizing a finite element method based on triangular prismatic elements to establish a thermal conduction model for the vdW MoS2/WSe2/MoS2 heterostructures. The THz frequency range used in our simulations was 0.5–15 THz, and the physical parameters of the vdW heterostructure were derived from experimentally prepared samples, as shown in Figure S1 and Table S5. A series of simulation results related to the temperature changes in the vdW heterostructure is then readily obtained.
First, we investigated the temperature changes, ∆T, in the stacked region of the heterostructure upon THz irradiation from 0.5 THz to 15 THz at different power densities. The irradiation duration, Δt, is 120 s, consisting of 60-s radiation intervals alternating with 60-s recovery periods. As shown in Figure 5a, the heterostructure exhibits a temperature rise after irradiation in the 0.5–15 THz frequency range. In particular, at the same power density, the ∆T increases against the frequency for 0.5 to 5 THz, then decreases as the irradiation frequency is reduced. According to the aforementioned TDS results, the extinction ratio of the heterostructure shows an increasing trend within the 0.5–5 THz range, while it gradually decreases above 5 THz. This frequency-dependent temperature rise, therefore, strongly correlates with the extinction coefficient trends observed in the TDS extinction. In the range of the 0.5–15 THz frequency band, we calculated the ΔT in the bottom MoS2 layer, the middle WSe2 layer, and the top MoS2 layer under continuous irradiation by three THz pulses with different power densities. The results are shown in Figures S2–S4.
Additionally, we compared the temperature changes ΔT of individual MoS2, WSe2, and the MoS2/WSe2/MoS2 heterostructure irradiated by different THz frequencies at a surface power density of 2798.73 mW/cm2, as presented in Figure 5b. The results reveal that within the ranges of 0.5–2.5 THz and 5–15 THz, the temperature rise of individual MoS2 and WSe2 is greater than that of the heterostructure. However, within the range of 2.5–5 THz, the trend is reversed, with the heterostructure exhibiting a higher temperature increase compared to the individual materials. More attention should be given to the frequency ranges below 2.5 THz and above 10 THz. In the frequency range below 2.5 THz, the thermal simulation results indicate that ΔT in the heterojunction is significantly influenced by interlayer thermal conductivity differences. The high thermal conductivity of the top MoS2 layer (100 W/(m·K)) accelerates heat diffusion to the substrate (SiO2/Si), while the low thermal conductivity of the middle WSe2 layer (40 W/(m·K)) partially hinders heat transfer. However, the overall thermal resistance distribution remains dominated by rapid lateral heat diffusion. The competition between these heat diffusion pathways results in a significant reduction in the temperature gradient within the heterojunction at lower frequencies, with the temperature rise being lower than that of monolayer materials (Figure 5b). These results predict that, in low-frequency applications, the thermal response of the heterojunction is likely more dependent on the thermal dissipation capacity of the substrate rather than interlayer cooperative effects. In the frequency range above 10 THz, the simulation results suggest that the thermal response of the heterojunction is likely limited by the thermal diffusion rate at the surface and interlayer thermal conduction, rather than the heat generation capability of the material. The reduced absorption depth of high-frequency THz energy in the heterojunction likely causes heat to become more concentrated in the top MoS2 layer, while the high thermal conductivity of the top layer further accelerates heat dissipation to the environment (natural convection coefficient h = 5−25 W·m−2·K−1). Furthermore, the heat distribution at high frequencies is likely more significantly affected by interlayer thermal conduction. Due to the thermal conductivity differences between the layers, heat transfer between layers involves scattering and uneven energy distribution. These factors make the conduction path of heat within the material more complex, increasing energy dissipation and limiting the accumulation of overall temperature rise.
We then simulated the time-resolved temperature rise profiles of the heterostructure under THz irradiation with different power densities, as shown in Figure 5c. It can be observed that following three consecutive THz pulses at 2.52 THz, the heterostructure exhibits nonlinear thermal responses. Specifically, during each pulse cycle, the temperature rises rapidly before transitioning to a sublinear growth phase without saturation. Upon pulse termination, an initial rapid cooling phase is followed by gradual thermal equilibration. Simultaneously, within a single pulse cycle, the heat of the heterostructure accumulated during the first 60 s of irradiation is mostly dissipated through thermal diffusion and convection in the subsequent 60 s. However, after continuous irradiation by three THz pulse cycles, the heterostructure still retains a portion of heat accumulation, manifesting as a temperature increase background. Under the maximum power density of 2798.7 mW/cm2, the final temperature rise of the heterostructure is approximately 2.5 mK.
To better understand the rapid temperature change process, we extracted the temperature rise curves of the heterostructure within the first 100 ms of THz radiation under different power densities from Figure 5c and plotted them in Figure 5d. Clearly, the ΔT of the heterostructure exhibits a nonlinear increase within the first 50 ms of irradiation at 2.52 THz waves. Beyond 50 ms, the ΔT demonstrates a linear growth trend over time, and the slope of the temperature curve shows a positive correlation with the power density of the THz wave irradiation on the sample.
We continue to investigate the ΔT in the stacked region of the heterostructure under continuous irradiation at 2.52 THz by three THz pulses of varying power densities. The results are displayed in Figure 5e. The temperature of the stacked structure region of the heterostructure increased overall after being irradiated by three consecutive THz pulses. Moreover, the ΔT increases linearly against the power density. This observation indicates that as the THz power density increases, the energy radiated onto the surface of the stacked structure per unit time also increases. Consequently, the material absorbs more THz photons, which are then converted into additional lattice thermal energy, leading to a rise in the material temperature. These findings indicate that the THz response mechanism of the heterostructure can be attributed to the photothermal effect [43]. Additionally, we also calculated the temperature changes in each layer of the heterostructure after THz irradiation. The simulation results are shown in Figure S5. The results indicate that the bottom MoS2 layer, the middle WSe2 layer, and the top MoS2 layer of the heterostructure all exhibited similar trends of temperature rise.
Photothermoelectric effect (PTE) describes photon-to-electron conversion due to the concentration or temperature gradient of photogenerated hot carriers in materials under light excitation. The gradient can drive the directional flow of electrons, thereby generating open-circuit-voltage or short-circuit-current photoelectric responses [44]. In the case of our 2D vdW heterostructures, after the crystalline lattice absorbs THz photons, the localized temperature rises within the material, forming a temperature gradient. According to the Seebeck effect [45], the temperature difference between the two terminals of the material can generate a potential difference (thermal voltage) as follows:
Δ V = S × Δ T
where S denotes the Seebeck coefficient and ΔT represents the temperature gradient. For the vdW MoS2/WSe2/MoS2 heterostructure, the total Seebeck coefficient SM/W/M can be calculated by the following formula [46]:
S M / W / M = σ BM × S BM + σ MW × S MW + σ TM × S TM σ BM + σ MW + σ TM
where σ is the conductivity of the material. The subscript “M/W/W” represents the MoS2/WSe2/MoS2, the subscript “BM” represents bottom MoS2, the subscript “MW” represents middle WSe2, and the subscript “TM” represents top MoS2. According to the relevant literature on the thermoelectric effect of MoS2 and WSe2 of different thicknesses [47,48], the Seebeck coefficient SBM and the conductivity σBM of bottom MoS2 are ~−507 μV/W, 1670 S/m, respectively. The SW and σW of middle WSe2 in the middle are about 50 μV/W and 34.5 S/m, respectively. For the top MoS2, the STM and σTM are approximately −494 μV/W and 1120 S/m, respectively. Combined with Equation (4), the total Seebeck coefficient SM/W/M is calculated as −495.04 μV/W. Based on the variation of ΔT in heterostructure irradiated by different THz power densities (Figure 5e), the electric potential difference ΔV of the overall heterostructure can be calculated as a function of the THz power densities, as shown in Figure 5h. The ΔV of the heterostructure is linearly dependent on the THz power density, with a maximum value of 1.2 μV at the highest irradiation intensity.
Next, we explored the ΔT in each layer of the heterostructure along the direction perpendicular to the sample surface, as depicted in Figure 5f. From the bottom layer to the top layer, the temperature of each material gradually rises with distance from the silicon substrate. The bottom MoS2 layer, closest to the silicon substrate, exhibits the smallest temperature change after THz pulse irradiation. Although the middle WSe2 layer is relatively thin (with a thickness of 1.86 nm), the temperature difference between its upper and lower surfaces becomes the largest under the same conditions due to its lower thermal conductivity (40 W/(m·K)) compared to MoS2 (100 W/(m·K)).
The Seebeck coefficient in the vertical (out-of-plane) direction of the material shows distinct characteristics compared to that in the in-plane direction. According to the results reported in the relevant literature [49,50], the out-of-plane Seebeck coefficient of the MoS2 and WSe2 are approximately −115 μV/W and +129 μV/W, respectively. According to Equation (4) and Figure 5f, we calculated the open-circuit voltage VOC distribution of each layer of the heterostructure in the direction perpendicular to the sample surface, as presented in Figure 5i. It can be seen that the VOC distribution of each layer in the heterostructure is consistent with the temperature change of each layer. The above results further confirm the existence of the PTE effect.
Finally, we also explored the temperature distribution across the lateral region of the whole heterostructure surface, with the results presented in Figure 5g. From one end of the sample to the other, the temperature gradually increases from the region of the bottom MoS2 layer, reaching its maximum in the stacked region of the three crystals, and then gradually decreases in the region of the top MoS2 layer on the right side. Based on the temperature gradient across the lateral region in the whole heterostructure, we calculated the VOC distribution using Equations (4) and (5). The results indicate that the VOC distribution exhibits spatial consistency with the temperature variation profile, thereby providing further evidence for the PTE effect observed.
It is noted that all of the above simulations are conducted under room temperature (~25 °C) and low humidity (<10%) conditions, without directly investigating the effects of temperature and humidity variations on the aforementioned processes. Based on the current findings, temperature and humidity may regulate the photo-thermal-electric effect of the heterostructure through several mechanisms. First, our thermal simulations assumed fixed thermal conductivities (MoS2: 100 W/(m·K); WSe2: 40 W/(m·K)), but practical scenarios suggest thermal conductivity (κ) decreases with rising temperature. For example, κ MoS 2 may decline below 80 W/(m·K) at temperatures exceeding 100 °C. Such reductions would elevate thermal resistance and amplify the temperature gradient within the heterostructure. While simulations indicate a linear dependence of ΔT on THz power density (Figure 5e in our original manuscript), the temperature dependence of thermal conductivity κ(T) at elevated temperatures may disrupt this linear relationship. This deviation would arise primarily from reduced thermal diffusion rates, particularly at multilayer interfaces. Furthermore, the temperature-sensitive thermal diffusion properties of the SiO2/Si substrate may also disrupt overall thermal equilibrium [51]. Secondly, the Seebeck voltage (ΔV = S × ΔT) is influenced by the temperature dependence of the Seebeck coefficients in individual layers. If the Seebeck coefficients vary with temperature, the observed linear relationship between ΔV and power density (Figure 5h in our original manuscript) may deviate. Thirdly, under high humidity, adsorbed water films at heterojunction interfaces may form low layers, exacerbating temperature gradients and impairing thermal conduction. Humidity fluctuations could also modulate material surface states and interfacial bonding, potentially altering Seebeck coefficients and their thermal response. Although current experiments are restricted to low humidity (<10%), future studies will systematically probe the role of humidity in thermo-electrical conversion. These issues will be explored systematically in our future study.

4. Conclusions

In the current study, we employed optical micromechanical arm alignment technology to successfully fabricate 2D vdW MoS2/WSe2/MoS2 few-layer structures and their heterostructures with a high-quality crystalline interface. THz-TDS measurements show obvious absorption in the heterostructure in the range of 0.5–15 THz. The few-layer WSe2, MoS2, MoS2/WSe2, and MoS2/WSe2/MoS2 exhibit an extinction coefficient of about 30%, 31.8%, 32.1%, and 32.6% at 2.52 THz, respectively. Several strong phonon modes are observed in the low-wavenumber Raman spectra of few-layer MoS2 and WSe2, corresponding to frequencies of 0.31 THz and 0.84 THz for MoS2, and 0.64 THz and 0.42 THz for WSe2. The energy of an incident THz photon may be equal to that of a few such phonons. Therefore, we propose a THz photo-to-phonon transition mechanism. In the present heterostructure, absorption of THz photons by the few-layer MoS2 and WSe2 lattices results in the emission of an integer number of their respective optical phonons. These newly emitted phonons lead to the rise of the temperature of the heterostructure.
Theoretical simulations of the phonon dispersion show degenerate states in the acoustic phonon branches of monolayer MoS2 and WSe2, with these states lifting in few-layer structures. This further validates our proposed THz photon-to-phonon transition mechanism.
The thermal response simulation results show the overall temperature increase of the heterostructure subjected to the THz pulses. Additionally, the temperature change ΔT and the PTE voltage exhibit a linear increase with the THz power density. Because photon-phonon couplings are ubiquitous and important in THz photonics, the results of this work are expected to gain more applications in 6G communication, terahertz imaging, macromolecular detection, astronomy, etc.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma18071665/s1, Table S1. The SNR calculations of MoS2 and WSe2 Raman peaks; Table S2. The SNR calculations of MoS2/WSe2 and MoS2/WSe2/MoS2 heterostructure Raman peaks; Table S3. The corresponding frequency values of monolayer and few-layer MoS2 materials with different thicknesses at the Γ point; Table S4. The corresponding frequency values of monolayer and few-layer WSe2 materials at the Γ point; Table S5. The physical parameters of different materials for simulation modeling; Figure S1. The geometric morphology, size, and thickness parameters of the materials. (a) Bottom layer MoS2. (b) Middle layer WSe2. (c) Top layer MoS2. (d) MoS2/WSe2. (e) MoS2/WSe2/MoS2. (f) The side view of the MoS2/WSe2/MoS2 model on a large–area silicon substrate. (g) Top view of the model; Figure S2. The temperature change ΔT of the top MoS2 after irradiation by terahertz waves of 0.5–15 THz with THz power density variation; Figure S3. The temperature increase ΔT of the middle WSe2 after irradiation by terahertz waves of 0.5–15 THz with THz power density variation; Figure S4. The temperature change ΔT of the bottom MoS2 after irradiation by terahertz waves of 0.5–15 THz with THz power density variation; Figure S5. The simulation of temperature changes ΔT in each layer of the MoS2/WSe2/MoS2 heterojunction after irradiation by a single terahertz pulse (pulse duration Δt = 120 s). (a) Bottom MoS2, (b) Middle WSe2, (c) Top MoS2, (d) MoS2/WSe2/MoS2 heterostructure.

Author Contributions

S.D., N.X. and H.C. conceived the idea and initiated the present study. J.H., S.D. and N.X. carried out the experiments and analyzed the data. Y.W. and X.R. provided the theoretical simulation and calculation of thermal responses. Y.F. and W.W. performed a phonon spectra theoretical simulation. H.C. and H.Z. offered suggestions for analyzing the data. J.H., H.C., S.D. and N.X. co-wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Key Basic Research Program of China (Grant Nos. 2019YFA0210200, 2019YFA0210201, and 2019YFA0210203), the National Natural Science Foundation of China (Grant Nos. 51290171 and 62488101), the Science and Technology Department of Guangdong Province (Grant Nos. 2020B1212060030 and 2023B1212060025), the Key Program of the National Natural Science Foundation of China (Grant Nos. 92373201), and the Fundamental Research Funds for the Central Universities.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ferguson, B.; Zhang, X.-C. Materials for terahertz science and technology. Nat. Mater. 2002, 1, 26–33. [Google Scholar] [CrossRef] [PubMed]
  2. Jiang, Y.; Li, G.; Ge, H.; Wang, F.; Li, L.; Chen, X.; Lu, M.; Zhang, Y. Machine Learning and Application in Terahertz Technology: A Review on Achievements and Future Challenges. IEEE Access 2022, 10, 53761–53776. [Google Scholar]
  3. Chen, Z.; Han, C.; Wu, Y.; Li, L.; Huang, C.; Zhang, Z.; Wang, G.; Tong, W. Terahertz Wireless Communications for 2030 and Beyond: A Cutting-Edge Frontier. IEEE Commun. Mag. 2021, 59, 66–72. [Google Scholar]
  4. Li, X.; Li, J.; Li, Y.; Ozcan, A.; Jarrahi, M. High-throughput terahertz imaging: Progress and challenges. Light Sci. Appl. 2023, 12, 233. [Google Scholar]
  5. Wang, C.; Shi, F.; Zhao, M.; Ao, J.; Jia, C.; Chen, S. Convolutional Neural Network-Based Terahertz Spectral Classification of Liquid Contraband for Security Inspection. IEEE Sens. J. 2021, 21, 18955–18963. [Google Scholar]
  6. Dang, S.; Amin, O.; Shihada, B.; Alouini, M.-S. What should 6G be? Nat. Electron. 2020, 3, 20–29. [Google Scholar]
  7. Xu, Z.; Li, Y.; Han, B.; Wang, Y.; Yuan, Q.; Li, Y.; He, W.; Hao, J.; Wu, L.; Yao, J. All-Silicon Polarization-Insensitive Metamaterial Absorber in the Terahertz Range. Materials 2024, 17, 2098. [Google Scholar] [CrossRef]
  8. Jia, S.; Lo, M.C.; Zhang, L.; Ozolins, O.; Udalcovs, A.; Kong, D.; Pang, X.; Guzman, R.; Yu, X.; Xiao, S.; et al. Integrated dual-laser photonic chip for high-purity carrier generation enabling ultrafast terahertz wireless communications. Nat. Commun. 2022, 13, 1388. [Google Scholar]
  9. Aqlan, B.; Himdi, M.; Vettikalladi, H.; Le-Coq, L. Experimental Realization of Sub-THz Circularly Polarized Antenna Based on Metasurface Superstrate at 300 GHz. Materials 2021, 14, 4796. [Google Scholar] [CrossRef]
  10. Huang, Y.; Shen, Y.; Wang, J. From Terahertz Imaging to Terahertz Wireless Communications. Engineering 2023, 22, 106–124. [Google Scholar]
  11. Mahdouani, M.; Bourguiga, R.; Gardelis, S. Auger Recombination and Carrier-Surface Optical Phonon Interaction in Van Der Waals Heterostructures Composed of Graphene and 2D Transition Metal Chalcogenides. Materials 2025, 18, 720. [Google Scholar] [CrossRef] [PubMed]
  12. Zhao, T.; Xie, P.; Wan, H.; Ding, T.; Liu, M.; Xie, J.; Li, E.; Chen, X.; Wang, T.; Zhang, Q.; et al. Ultrathin MXene assemblies approach the intrinsic absorption limit in the 0.5–10 THz band. Nat. Photonics 2023, 17, 622–628. [Google Scholar]
  13. Regan, E.C.; Wang, D.; Paik, E.Y.; Zeng, Y.; Zhang, L.; Zhu, J.; MacDonald, A.H.; Deng, H.; Wang, F. Emerging exciton physics in transition metal dichalcogenide heterobilayers. Nat. Rev. Mater. 2022, 7, 778–795. [Google Scholar]
  14. Yang, X.; Li, J.; Song, R.; Zhao, B.; Tang, J.; Kong, L.; Huang, H.; Zhang, Z.; Liao, L.; Liu, Y.; et al. Highly reproducible van der Waals integration of two-dimensional electronics on the wafer scale. Nat. Nanotechnol. 2023, 18, 471–478. [Google Scholar] [PubMed]
  15. Liu, L.; Cai, Z.; Xue, S.; Huang, H.; Chen, S.; Gou, S.; Zhang, Z.; Guo, Y.; Yao, Y.; Bao, W.; et al. A mass transfer technology for high-density two-dimensional device integration. Nat. Electron. 2025, 8, 135–146. [Google Scholar]
  16. Tan, C.; Yin, S.; Chen, J.; Lu, Y.; Wei, W.; Du, H.; Liu, K.; Wang, F.; Zhai, T.; Li, L. Broken-Gap PtS2/WSe2 van der Waals Heterojunction with Ultrahigh Reverse Rectification and Fast Photoresponse. ACS Nano 2021, 15, 8328–8337. [Google Scholar]
  17. Wang, Y.; Gu, Y.; Cui, A.; Li, Q.; He, T.; Zhang, K.; Wang, Z.; Li, Z.; Zhang, Z.; Wu, P.; et al. Fast Uncooled Mid-Wavelength Infrared Photodetectors with Heterostructures of van der Waals on Epitaxial HgCdTe. Adv. Mater. 2022, 34, e2107772. [Google Scholar]
  18. Luo, T.; Ilyas, B.; von Hoegen, A.; Lee, Y.; Park, J.; Park, J.-G.; Gedik, N. Time-of-flight detection of terahertz phonon-polariton. Nat. Commun. 2024, 15, 2276. [Google Scholar]
  19. Yoon, Y.; Lu, Z.; Uzundal, C.; Qi, R.; Zhao, W.; Chen, S.; Feng, Q.; Kim, W.; Naik, M.H.; Watanabe, K.; et al. Terahertz phonon engineering with van der Waals heterostructures. Nature 2024, 631, 771–776. [Google Scholar]
  20. Kim, S.J.; Kang, B.J.; Puc, U.; Kim, W.T.; Jazbinsek, M.; Rotermund, F.; Kwon, O.P. Highly Nonlinear Optical Organic Crystals for Efficient Terahertz Wave Generation, Detection, and Applications. Adv. Opt. Mater. 2021, 9, 2101019. [Google Scholar]
  21. Wu, D.; Ma, Y.; Niu, Y.; Liu, Q.; Dong, T.; Zhang, S.; Niu, J.; Zhou, H.; Wei, J.; Wang, Y.; et al. Ultrabroadband photosensitivity from visible to terahertz at room temperature. Sci. Adv. 2018, 4, eaao3057. [Google Scholar] [CrossRef] [PubMed]
  22. Yan, X.; Zhu, L.; Zhou, Y.; E, Y.; Wang, L.; Xu, X. Dielectric property of MoS2 crystal in terahertz and visible regions. Appl. Opt. 2015, 54, 6732–6736. [Google Scholar] [CrossRef] [PubMed]
  23. Deng, X.-Y.; Deng, X.-H.; Su, F.-H.; Liu, N.-H.; Liu, J.-T. Broadband ultra-high transmission of terahertz radiation through monolayer MoS2. J. Appl. Phys. 2015, 118, 224304. [Google Scholar] [CrossRef]
  24. Arcos, D.; Gabriel, D.; Dumcenco, D.; Kis, A.; Ferrer-Anglada, N. THz time-domain spectroscopy and IR spectroscopy on MoS2. Phys. Status Solidi (b) 2016, 253, 2499–2504. [Google Scholar] [CrossRef]
  25. Zhu, W.; Xiao, F.; Kang, M.; Sikdar, D.; Liang, X.; Geng, J.; Premaratne, M.; Jin, R. MoS2 Broadband Coherent Perfect Absorber for Terahertz Waves. IEEE Photonics J. 2016, 8, 5502207. [Google Scholar] [CrossRef]
  26. Betz, M.; Elezzabi, A.Y.; Tsen, K.-T.; Buss, J.H.; Smith, R.P.; Coslovich, G.; Kaindl, R.A. Broadband transient THz conductivity of the transition-metal dichalcogenide MoS2. In Proceedings of the Ultrafast Phenomena and Nanophotonics XIX, San Francisco, CA, USA, 7–12 February 2015. [Google Scholar]
  27. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  28. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758. [Google Scholar] [CrossRef]
  29. Togo, A. First-principles Phonon Calculations with Phonopy and Phono3py. J. Phys. Soc. Jpn. 2023, 92, 012001. [Google Scholar] [CrossRef]
  30. Togo, A.; Chaput, L.; Tadano, T.; Tanaka, I. Implementation strategies in phonopy and phono3py. J. Phys. Condens. Matter 2023, 35, 353001. [Google Scholar] [CrossRef]
  31. Langreth, D.C.; Mehl, M.J. Beyond the local-density approximation in calculations of ground-state electronic properties. Phys. Rev. B 1983, 28, 1809–1834. [Google Scholar] [CrossRef]
  32. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar]
  33. Fang, Y.; Chen, H.; Li, Z.; Wang, W. Hyperbolic Phonon Polaritons and Wave Vector Direction-Dependent Dielectric Tensors in Anisotropic Crystals. J. Phys. Chem. C 2024, 128, 7359–7369. [Google Scholar]
  34. Gonze, X.; Lee, C. Dynamical matrices, Born effective charges, dielectric permittivity tensors, and interatomic force constants from density-functional perturbation theory. Phys. Rev. B 1997, 55, 10355. [Google Scholar]
  35. Tonndorf, P.; Schmidt, R.; Bottger, P.; Zhang, X.; Borner, J.; Liebig, A.; Albrecht, M.; Kloc, C.; Gordan, O.; Zahn, D.R.; et al. Photoluminescence emission and Raman response of monolayer MoS2, MoSe2, and WSe2. Opt. Express 2013, 21, 4908–4916. [Google Scholar]
  36. Liu, L.; Xu, N.; Zhang, Y.; Zhao, P.; Chen, H.; Deng, S. Van der Waals Bipolar Junction Transistor Using Vertically Stacked Two-Dimensional Atomic Crystals. Adv. Funct. Mater. 2019, 29, 1807893. [Google Scholar]
  37. Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.Y.; Galli, G.; Wang, F. Emerging photoluminescence in monolayer MoS2. Nano Lett. 2010, 10, 1271–1275. [Google Scholar]
  38. Steinhoff, A.; Kim, J.H.; Jahnke, F.; Rosner, M.; Kim, D.S.; Lee, C.; Han, G.H.; Jeong, M.S.; Wehling, T.O.; Gies, C. Efficient Excitonic Photoluminescence in Direct and Indirect Band Gap Monolayer MoS2. Nano Lett. 2015, 15, 6841–6847. [Google Scholar]
  39. Liu, L.; Xu, N.S.; Ke, Y.; Chen, H.; Zhang, Y.; Deng, S. Sensing by Surface Work Function Modulation: High Performance Gas Sensing using van der Waals Stacked Bipolar Junction Transistor. Sens. Actuators B Chem. 2021, 345, 130340. [Google Scholar]
  40. Koch, M.; Mittleman, D.M.; Ornik, J.; Castro-Camus, E. Terahertz time-domain spectroscopy. Nat. Rev. Methods Primers 2023, 3, 48. [Google Scholar]
  41. Lui, C.; Frenzel, A.; Pilon, D.; Lee, Y.-H.; Ling, X.; Akselrod, G.; Kong, J.; Gedik, N. Trion-induced negative photoconductivity in monolayer MoS2. Phys. Rev. Lett. 2014, 113, 166801. [Google Scholar]
  42. Balkanski, M.; Wallis, R.F. Semiconductor Physics and Applications; Oxford University Press: Oxford, UK, 2000; Volume 8. [Google Scholar]
  43. He, R.; Chen, Z.; Lai, H.; Zhang, T.; Wen, J.; Chen, H.; Xie, F.; Yue, S.; Liu, P.; Chen, J.; et al. van der Waals Transition-Metal Oxide for Vis-MIR Broadband Photodetection via Intercalation Strategy. ACS Appl. Mater. Interfaces 2019, 11, 15741–15747. [Google Scholar] [PubMed]
  44. Long, M.; Wang, P.; Fang, H.; Hu, W. Progress, Challenges, and Opportunities for 2D Material Based Photodetectors. Adv. Funct. Mater. 2018, 29, 1803807. [Google Scholar]
  45. Chen, H.; Wang, X.; Liu, S.; Cao, Z.; Li, J.; Zhu, H.; Li, S.; Xu, N.; Deng, S. Monolithic Multiparameter Terahertz Nano/Microdetector Based on Plasmon Polariton Atomic Cavity. Adv. Mater. 2025, 37, e2410946. [Google Scholar]
  46. Hicks, L.D.; Dresselhaus, M.S. Effect of quantum-well structures on the thermoelectric figure of merit. Phys. Rev. B Condens. Matter 1993, 47, 12727–12731. [Google Scholar]
  47. Kayyalha, M.; Maassen, J.; Lundstrom, M.; Shi, L.; Chen, Y.P. Gate-tunable and thickness-dependent electronic and thermoelectric transport in few-layer MoS2. J. Appl. Phys. 2016, 120, 134305. [Google Scholar]
  48. Kim, W.; Kim, H.; Hallam, T.; McEvoy, N.; Gatensby, R.; Nerl, H.C.; O’Neill, K.; Siris, R.; Kim, G.T.; Duesberg, G.S. Field-Dependent Electrical and Thermal Transport in Polycrystalline WSe2. Adv. Mater. Interfaces 2018, 5, 1701161. [Google Scholar] [CrossRef]
  49. Kang, M.-S.; Kang, S.-Y.; Lee, W.-Y.; Park, N.-W.; Kown, K.C.; Choi, S.; Kim, G.-S.; Nam, J.; Kim, K.S.; Saitoh, E.; et al. Large-scale MoS2 thin films with a chemically formed holey structure for enhanced Seebeck thermopower and their anisotropic properties. J. Mater. Chem. A 2020, 8, 8669–8677. [Google Scholar]
  50. Lee, W.-Y.; Kang, M.-S.; Park, N.-W.; Kim, G.-S.; Nguyen, A.D.; Choi, J.W.; Yoon, Y.-G.; Kim, Y.S.; Jang, H.W.; Saitoh, E.; et al. Layer dependence of out-of-plane electrical conductivity and Seebeck coefficient in continuous mono- to multilayer MoS2 films. J. Mater. Chem. A 2021, 9, 26896–26903. [Google Scholar] [CrossRef]
  51. Sahoo, S.; Gaur, A.P.; Ahmadi, M.; Guinel, M.J.F.; Katiyar, R.S. Temperature-dependent Raman studies and thermal conductivity of few-layer MoS2. J. Phys. Chem. C 2013, 117, 9042–9047. [Google Scholar]
Figure 1. vdW MoS2/WSe2/MoS2 heterostructure illustrations and morphological characterizations. (af) Schematic illustration of heterostructure fabrication procedures. (a,b) The MoS2 and WSe2 space-cell structure model. (c) The cross−sectional view of the vertical MoS2/WSe2/MoS2 heterostructure. (d) The few−layer MoS2 on SiO2/Si substrate. (e) The vertical MoS2/WSe2 heterostructure. (f) The vertical MoS2/WSe2/MoS2 heterostructure. (g,h) Raman (g) and photoluminescence (PL) (h) spectra of the bottom MoS2 (purple line), middle WSe2 (red line), top MoS2 (blue line), and overlapping regions of the three flakes (green line). (i) The top view of the heterostructure. (jl) The optical microscope images of the vertically stacked structure; scale bar: 20 μm. (mo) AFM images of the three−box area shown in (l). Insets: the AFM height profiles of bottom MoS2 (4.85 nm), middle WSe2 (3.12 nm), and top MoS2 (4.25 nm), respectively.
Figure 1. vdW MoS2/WSe2/MoS2 heterostructure illustrations and morphological characterizations. (af) Schematic illustration of heterostructure fabrication procedures. (a,b) The MoS2 and WSe2 space-cell structure model. (c) The cross−sectional view of the vertical MoS2/WSe2/MoS2 heterostructure. (d) The few−layer MoS2 on SiO2/Si substrate. (e) The vertical MoS2/WSe2 heterostructure. (f) The vertical MoS2/WSe2/MoS2 heterostructure. (g,h) Raman (g) and photoluminescence (PL) (h) spectra of the bottom MoS2 (purple line), middle WSe2 (red line), top MoS2 (blue line), and overlapping regions of the three flakes (green line). (i) The top view of the heterostructure. (jl) The optical microscope images of the vertically stacked structure; scale bar: 20 μm. (mo) AFM images of the three−box area shown in (l). Insets: the AFM height profiles of bottom MoS2 (4.85 nm), middle WSe2 (3.12 nm), and top MoS2 (4.25 nm), respectively.
Materials 18 01665 g001
Figure 2. THz time−domain spectral characterization of large−area 2D vdW MoS2, WSe2 crystals and their heterostructures. (ad) Optical micrographs of large−area MoS2, WSe2, MoS2/WSe2, and MoS2/WSe2/MoS2, respectively; scale bar: 20 μm. (e) Schematic illustration of the THz−TDS system used for the experiment. (f) Time−domain spectra for blank substrate, MoS2 on the substrate, WSe2 on the substrate, MoS2/WSe2 on the substrate, and MoS2/WSe2/MoS2 on the substrate. (gh) Enlarged spectra of the pink and orange dotted boxes in (f). (i) Time−domain spectrum of the N2 atmosphere. (j) Frequency−domain spectra obtained by Fourier transform of the time−domain spectra of N2 atmosphere, blank substrate, MoS2, WSe2, MoS2/WSe2, and MoS2/WSe2/MoS2. The black box: the amplitude variations of different materials at frequencies ranging from 1.5 to 4.5 THz. (k) Enlarged spectra of the black dotted boxes in (j). (ln) The calculated extinction coefficient curves of the frequency−domain spectra of MoS2, WSe2, MoS2/WSe2, and MoS2/WSe2/MoS2: (l) 0.5–15 THz, (m) 0.5–10 THz, (n) 0.5–5 THz.
Figure 2. THz time−domain spectral characterization of large−area 2D vdW MoS2, WSe2 crystals and their heterostructures. (ad) Optical micrographs of large−area MoS2, WSe2, MoS2/WSe2, and MoS2/WSe2/MoS2, respectively; scale bar: 20 μm. (e) Schematic illustration of the THz−TDS system used for the experiment. (f) Time−domain spectra for blank substrate, MoS2 on the substrate, WSe2 on the substrate, MoS2/WSe2 on the substrate, and MoS2/WSe2/MoS2 on the substrate. (gh) Enlarged spectra of the pink and orange dotted boxes in (f). (i) Time−domain spectrum of the N2 atmosphere. (j) Frequency−domain spectra obtained by Fourier transform of the time−domain spectra of N2 atmosphere, blank substrate, MoS2, WSe2, MoS2/WSe2, and MoS2/WSe2/MoS2. The black box: the amplitude variations of different materials at frequencies ranging from 1.5 to 4.5 THz. (k) Enlarged spectra of the black dotted boxes in (j). (ln) The calculated extinction coefficient curves of the frequency−domain spectra of MoS2, WSe2, MoS2/WSe2, and MoS2/WSe2/MoS2: (l) 0.5–15 THz, (m) 0.5–10 THz, (n) 0.5–5 THz.
Materials 18 01665 g002
Figure 3. The low-wavenumber Raman analysis for vdW MoS2 and WSe2 and their heterostructures. (ad) Low−wavenumber Raman spectra of MoS2 (a), WSe2 (b), MoS2/WSe2 (c), and MoS2/WSe2/MoS2 (d), respectively. The atomic vibrations shown in (a,b) correspond to E 2 g 2 modes, which are associated with the Raman peaks at 10.23 cm−1/0.31 THz (a) and 13.98 cm−1/0.42 THz (b), respectively. (e) Schematic illustration of the THz photon–phonon transition mechanism.
Figure 3. The low-wavenumber Raman analysis for vdW MoS2 and WSe2 and their heterostructures. (ad) Low−wavenumber Raman spectra of MoS2 (a), WSe2 (b), MoS2/WSe2 (c), and MoS2/WSe2/MoS2 (d), respectively. The atomic vibrations shown in (a,b) correspond to E 2 g 2 modes, which are associated with the Raman peaks at 10.23 cm−1/0.31 THz (a) and 13.98 cm−1/0.42 THz (b), respectively. (e) Schematic illustration of the THz photon–phonon transition mechanism.
Materials 18 01665 g003
Figure 4. The phonon dispersion analysis for the monolayer and bulk MoS2 and WSe2 (theoretical results). (a) Monolayer MoS2. (b) Monolayer WSe2. (c) Bulk MoS2. The purple dashed box: the calculation results of the phonon dispersion at the Γ–M point. (d) Amplifying the phonon dispersion of bulk MoS2 from (c) in the blue-dashed box at the Γ–M point. (e) Bulk WSe2. The purple dashed box: the calculation results of the phonon dispersion at the Γ–M point. (f) Enlarged phonon dispersion of bulk WSe2 from (e) in the blue−dashed box at the Γ–M point.
Figure 4. The phonon dispersion analysis for the monolayer and bulk MoS2 and WSe2 (theoretical results). (a) Monolayer MoS2. (b) Monolayer WSe2. (c) Bulk MoS2. The purple dashed box: the calculation results of the phonon dispersion at the Γ–M point. (d) Amplifying the phonon dispersion of bulk MoS2 from (c) in the blue-dashed box at the Γ–M point. (e) Bulk WSe2. The purple dashed box: the calculation results of the phonon dispersion at the Γ–M point. (f) Enlarged phonon dispersion of bulk WSe2 from (e) in the blue−dashed box at the Γ–M point.
Materials 18 01665 g004
Figure 5. Simulations of thermal responses under THz irradiation of the MoS2/WSe2/MoS2 heterostructure. (a) Temperature changes, ΔT, of the heterostructure upon irradiation by THz waves of 0.5–15 THz with varied power densities. (b) Temperature changes of MoS2, WSe2, and the MoS2/WSe2/MoS2 heterostructure as a function of the irradiation frequency. The power density is kept at 2798.73 mW/cm2. (c) Time−resolved temperature changes of the MoS2/WSe2/MoS2 heterostructure upon irradiation with different THz power densities. The THz wave frequency is kept at 2.52 THz. (d) The temperature rise of the heterostructure over time within 100 ms of THz radiation under different power densities. (e) The temperature increases ΔT of the heterostructure as a function of the THz power density. (f) The variation of ΔT of each layer in the heterostructure along the direction perpendicular to the sample surface. (g) The temperature distribution along the red solid line. (h) The thermal voltage, ΔV, across the heterostructure as a function of the THz power density. (i) The open-circuit voltage VOC of each layer in the heterostructure along the direction perpendicular to the sample surface. (j) The variation of VOC along the red solid line.
Figure 5. Simulations of thermal responses under THz irradiation of the MoS2/WSe2/MoS2 heterostructure. (a) Temperature changes, ΔT, of the heterostructure upon irradiation by THz waves of 0.5–15 THz with varied power densities. (b) Temperature changes of MoS2, WSe2, and the MoS2/WSe2/MoS2 heterostructure as a function of the irradiation frequency. The power density is kept at 2798.73 mW/cm2. (c) Time−resolved temperature changes of the MoS2/WSe2/MoS2 heterostructure upon irradiation with different THz power densities. The THz wave frequency is kept at 2.52 THz. (d) The temperature rise of the heterostructure over time within 100 ms of THz radiation under different power densities. (e) The temperature increases ΔT of the heterostructure as a function of the THz power density. (f) The variation of ΔT of each layer in the heterostructure along the direction perpendicular to the sample surface. (g) The temperature distribution along the red solid line. (h) The thermal voltage, ΔV, across the heterostructure as a function of the THz power density. (i) The open-circuit voltage VOC of each layer in the heterostructure along the direction perpendicular to the sample surface. (j) The variation of VOC along the red solid line.
Materials 18 01665 g005
Table 1. The characteristics of low−wavenumber Raman modes and corresponding wavelengths, frequencies, and phonon energies of different TMDs and their heterostructures.
Table 1. The characteristics of low−wavenumber Raman modes and corresponding wavelengths, frequencies, and phonon energies of different TMDs and their heterostructures.
MaterialsRaman Shift
(cm−1)
Wavelength
(cm)
Frequency
(THz)
Phonon Energy
(meV)
MoS210.230.0970.311.28
28.010.0360.843.48
WSe2 13.98 0.071 0.42 1.74
20.83 0.048 0.62 2.57
MoS2/WSe2 9.88 0.101 0.3 1.23
13.98 0.071 0.42 1.74
23.56 0.042 0.71 2.94
27.66 0.036 0.83 3.44
MoS2/WSe2/WSe2 9.88 0.101 0.30 1.24
13.99 0.072 0.42 1.74
20.83 0.048 0.63 2.61
27.84 0.036 0.84 3.48
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

Huang, J.; Xu, N.; Wu, Y.; Ran, X.; Fang, Y.; Zhu, H.; Wang, W.; Chen, H.; Deng, S. Interactions of Terahertz Photons with Phonons of Two-Dimensional van der Waals MoS2/WSe2/MoS2 Heterostructures and Thermal Responses. Materials 2025, 18, 1665. https://doi.org/10.3390/ma18071665

AMA Style

Huang J, Xu N, Wu Y, Ran X, Fang Y, Zhu H, Wang W, Chen H, Deng S. Interactions of Terahertz Photons with Phonons of Two-Dimensional van der Waals MoS2/WSe2/MoS2 Heterostructures and Thermal Responses. Materials. 2025; 18(7):1665. https://doi.org/10.3390/ma18071665

Chicago/Turabian Style

Huang, Jingwen, Ningsheng Xu, Yumao Wu, Xue Ran, Yue Fang, Hongjia Zhu, Weiliang Wang, Huanjun Chen, and Shaozhi Deng. 2025. "Interactions of Terahertz Photons with Phonons of Two-Dimensional van der Waals MoS2/WSe2/MoS2 Heterostructures and Thermal Responses" Materials 18, no. 7: 1665. https://doi.org/10.3390/ma18071665

APA Style

Huang, J., Xu, N., Wu, Y., Ran, X., Fang, Y., Zhu, H., Wang, W., Chen, H., & Deng, S. (2025). Interactions of Terahertz Photons with Phonons of Two-Dimensional van der Waals MoS2/WSe2/MoS2 Heterostructures and Thermal Responses. Materials, 18(7), 1665. https://doi.org/10.3390/ma18071665

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