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 MoS
2/WSe
2/MoS
2 heterostructures is depicted in
Figure 1a–f,i. The space-cell structure model of MoS
2 and WSe
2 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 MoS
2/WSe
2/MoS
2 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 MoS
2, WSe
2, MoS
2/WSe
2, and MoS
2/WSe
2/MoS
2 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
(in-plane) and A
1g (out-of-plane) of MoS
2, respectively. The wavenumber difference Δd of these two vibration modes can simply reflect the layer number of the MoS
2 [
35]. The Δd values of the bottom and top MoS
2 are 25.2 cm
−1 and 23.6 cm
−1, respectively. Both values suggest more than five layers of the two MoS
2 layers. As for the Raman peaks of WSe
2, a strong peak at 250.2 cm
−1 is related to the A
1g mode, and a small signature at 306.8 cm
−1 may be assigned to the normally inactive B
2g mode. Notably, the Raman peaks of MoS
2 in the overlapped region exhibit slight broadening and shifting, indicating charge transfer across the interfaces. Consequently, a depletion layer was formed at the MoS
2/WSe
2/MoS
2 vdW heterostructure [
36].
Figure 1h depicts the photoluminescence (PL) spectra of the vertically stacked heterostructure. Two peaks of the top MoS
2 layer are observed at 1.84 eV (673.9 nm) and 2.0 eV (620 nm), respectively. They correspond to the A
1 and B
1 direct excitonic transitions with the energy splitting due to valence band spin-orbital coupling [
37]. The A
1 and B
1 peaks of the bottom MoS
2 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 MoS
2 layers have different layer numbers. The middle WSe
2 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 MoS
2 and WSe
2 layers. AFM height profiles extracted from each layer of MoS
2/WSe
2/MoS
2 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 MoS
2, middle WSe
2, and top MoS
2. 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 MoS
2, WSe
2, MoS
2/WSe
2, and MoS
2/WSe
2/MoS
2, respectively. A schematic illustration of the TDS spectroscopy is shown in
Figure 2e.
Figure 2f displays THz-TDS results from the blank substrate, MoS
2, WSe
2, MoS
2/WSe
2, and MoS
2/WSe
2/MoS
2. The time-domain peak and valley of the THz wave transmitting through large-area vdW MoS
2, WSe
2, and their heterostructures on high-resistance Si are ~5 ps.
Figure 2i displays THz-TDS results from the N
2 atmosphere. The time-domain peak position of the THz signal transmitting through N
2 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 MoS
2, WSe
2, 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 WSe
2, MoS
2, MoS
2/WSe
2, and MoS
2/WSe
2/MoS
2, the signal intensity gradually decreases. The THz electric field transmitting through the blank substrate, MoS
2/substrate, WSe
2/substrate, MoS
2/WSe
2/substrate, and MoS
2/WSe
2/MoS
2/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 MoS
2 and WSe
2, 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: WSe
2, MoS
2, and MoS
2/WSe
2 to MoS
2/WSe
2/MoS
2.
The transmittance coefficient of the material can be calculated according to [
41]:
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
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 MoS
2 and WSe
2 is higher than that of the heterostructure. In the range of 2–5 THz, the extinction of THz signals of MoS
2/WSe
2 and MoS
2/WSe
2/MoS
2 heterostructures gradually increases, surpassing the values observed in few-layer MoS
2 and WSe
2. Significant signal fluctuations emerge for the frequency range above 5 THz. In this spectral range, the extinction coefficient of few-layer MoS
2, WSe
2, and MoS
2/WSe
2 increases gradually, whereas that of MoS
2/WSe
2/MoS
2 declines. According to the data in
Figure 2n, the extinction coefficient of WSe
2, MoS
2, MoS
2/WSe
2, and MoS
2/WSe
2/MoS
2 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 MoS
2 (1.83 eV) and WSe
2 (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 MoS
2, WSe
2, MoS
2/WSe
2, and MoS
2/WSe
2/MoS
2. 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 MoS
2. 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 WSe
2. 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 MoS
2/WSe
2 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 MoS
2/WSe
2/MoS
2 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 MoS
2 lattice absorbs THz photons at 2.52 THz, eight phonons (
ν1) or three phonons (
ν2) can be released. Similarly, when the WSe
2 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 MoS
2/WSe
2/MoS
2 heterostructures. It should be noted that bulk MoS
2 and WSe
2 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 MoS
2 layer, 7 for the bottom MoS
2 layer, and 5 for the middle WSe
2 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 MoS
2 and WSe
2 in
Figure 4. Combined with
Tables S3 and S4 in the Supporting Information, the corresponding phonon frequencies of MoS
2 and WSe
2 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 MoS
2 (
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 MoS
2 (0.31 THz and 0.84 THz). The values of the new optical phonon branches in bulk WSe
2 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 WSe
2 (0.42 THz and 0.64 THz).
Based on the previous TDS results, few-layer MoS
2, WSe
2, 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 MoS
2/WSe
2/MoS
2 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, Δ
t, 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 Δ
T in the bottom MoS
2 layer, the middle WSe
2 layer, and the top MoS
2 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 Δ
T of individual MoS
2, WSe
2, and the MoS
2/WSe
2/MoS
2 heterostructure irradiated by different THz frequencies at a surface power density of 2798.73 mW/cm
2, 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 MoS
2 and WSe
2 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 Δ
T in the heterojunction is significantly influenced by interlayer thermal conductivity differences. The high thermal conductivity of the top MoS
2 layer (100 W/(m·K)) accelerates heat diffusion to the substrate (SiO
2/Si), while the low thermal conductivity of the middle WSe
2 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 MoS
2 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/cm
2, 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 Δ
T of the heterostructure exhibits a nonlinear increase within the first 50 ms of irradiation at 2.52 THz waves. Beyond 50 ms, the Δ
T 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 Δ
T 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 Δ
T 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 MoS
2 layer, the middle WSe
2 layer, and the top MoS
2 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:
where S denotes the Seebeck coefficient and Δ
T represents the temperature gradient. For the vdW MoS
2/WSe
2/MoS
2 heterostructure, the total Seebeck coefficient S
M/W/M can be calculated by the following formula [
46]:
where
σ is the conductivity of the material. The subscript “M/W/W” represents the MoS
2/WSe
2/MoS
2, the subscript “BM” represents bottom MoS
2, the subscript “MW” represents middle WSe
2, and the subscript “TM” represents top MoS
2. According to the relevant literature on the thermoelectric effect of MoS
2 and WSe
2 of different thicknesses [
47,
48], the Seebeck coefficient S
BM and the conductivity
σBM of bottom MoS
2 are ~−507 μV/W, 1670 S/m, respectively. The S
W and
σW of middle WSe
2 in the middle are about 50 μV/W and 34.5 S/m, respectively. For the top MoS
2, the S
TM and
σTM are approximately −494 μV/W and 1120 S/m, respectively. Combined with Equation (4), the total Seebeck coefficient S
M/W/M is calculated as −495.04 μV/W. Based on the variation of Δ
T in heterostructure irradiated by different THz power densities (
Figure 5e), the electric potential difference Δ
V of the overall heterostructure can be calculated as a function of the THz power densities, as shown in
Figure 5h. The Δ
V 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 Δ
T 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 MoS
2 layer, closest to the silicon substrate, exhibits the smallest temperature change after THz pulse irradiation. Although the middle WSe
2 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 MoS
2 (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 MoS
2 and WSe
2 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 MoS
2 layer, reaching its maximum in the stacked region of the three crystals, and then gradually decreases in the region of the top MoS
2 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 (MoS
2: 100 W/(m·K); WSe
2: 40 W/(m·K)), but practical scenarios suggest thermal conductivity (
κ) decreases with rising temperature. For example,
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 Δ
T 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 SiO
2/Si substrate may also disrupt overall thermal equilibrium [
51]. Secondly, the Seebeck voltage (Δ
V = S × Δ
T) 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 Δ
V and power density (
Figure 5h in our original manuscript) may deviate. Thirdly, under high humidity, adsorbed water films at heterojunction interfaces may form low
-κ 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.