ACS Publications. Most Trusted. Most Cited. Most Read
Locally Phase-Engineered MoTe2 for Near-Infrared Photodetectors
My Activity
  • Open Access
Article

Locally Phase-Engineered MoTe2 for Near-Infrared Photodetectors
Click to copy article linkArticle link copied!

  • Jan Hidding
    Jan Hidding
    Zernike Institute for Advanced Materials, University of Groningen, 9747 AG Groningen, The Netherlands
    More by Jan Hidding
  • Cédric A. Cordero-Silis
    Cédric A. Cordero-Silis
    Zernike Institute for Advanced Materials, University of Groningen, 9747 AG Groningen, The Netherlands
  • Daniel Vaquero
    Daniel Vaquero
    Nanotechnology Group, USAL─Nanolab, Universidad de Salamanca, E-37008 Salamanca, Spain
  • Konstantinos P. Rompotis
    Konstantinos P. Rompotis
    Zernike Institute for Advanced Materials, University of Groningen, 9747 AG Groningen, The Netherlands
  • Jorge Quereda
    Jorge Quereda
    Departamento de Física de Materiales, GISC, Universidad Complutense de Madrid, E-28040 Madrid, Spain
  • Marcos H. D. Guimarães*
    Marcos H. D. Guimarães
    Zernike Institute for Advanced Materials, University of Groningen, 9747 AG Groningen, The Netherlands
    *Email: [email protected]
Open PDFSupporting Information (1)

ACS Photonics

Cite this: ACS Photonics 2024, 11, 10, 4083–4089
Click to copy citationCitation copied!
https://doi.org/10.1021/acsphotonics.4c00896
Published September 16, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

Transition-metal dichalcogenides (TMDs) are ideal systems for two-dimensional (2D) optoelectronic applications owing to their strong light-matter interaction and various band gap energies. New techniques to modify the crystallographic phase of TMDs have recently been discovered, allowing the creation of lateral heterostructures and the design of all-2D circuitry. Thus, far, the potential benefits of phase-engineered TMD devices for optoelectronic applications are still largely unexplored. The dominant mechanisms involved in photocurrent generation in these systems remain unclear, hindering further development of new all-2D optoelectronic devices. Here, we fabricate locally phase-engineered MoTe2 optoelectronic devices, creating a metal (1T′) semiconductor (2H) lateral junction and unveil the main mechanisms at play for photocurrent generation. We find that the photocurrent originates from the 1T′–2H junction, with a maximum at the 2H MoTe2 side of the junction. This observation, together with the nonlinear IV-curve, indicates that the photovoltaic effect plays a major role in the photon-to-charge current conversion in these systems. Additionally, the 1T′–2H MoTe2 heterojunction device exhibits a fast optoelectronic response over a wavelength range of 700–1100 nm, with a rise and fall times of 113 and 110 μs, respectively, 2 orders of magnitude faster when compared to a directly contacted 2H MoTe2 device. These results show the potential of local phase-engineering for all-2D optoelectronic circuitry.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2024 The Authors. Published by American Chemical Society

Special Issue

Published as part of ACS Photonics special issue “Rising Stars in Photonics”.

Introduction

Click to copy section linkSection link copied!

Historically, low-dimensional systems have been proposed, fabricated, and studied for enhanced photodetection (1,2) and potential on-chip photonic applications. (3−5) Particularly, two-dimensional (2D) materials beyond graphene have shown high interest due to their strong light-matter interaction and compatibility with silicon photonic new generation devices. (6−8) From the large family of 2D crystals, the group of materials called the transition-metal dichalcogenides (TMDs) have gained much attention in the last two decades due to their versatility, mechanical strength, atomically flat interfaces, and strong absorption at the monolayer limit, making them promising candidates for future (opto)electronic and (opto)spintronic applications. (9) The most commonly studied crystal structure of the TMD family is the hexagonal (2H) phase, for which most TMDs are semiconducting and possess a thickness-dependent band gap. (10,11) Apart from the 2H phase, however, TMDs can present a multitude of different crystallographic phases, such as the semiconducting 3R phase, or the semimetallic 1T, 1T′, and 1Td phases, which possess different symmetries and (opto)electronic properties. To benefit from these different properties in a single device, researchers have recently focused on gaining control of the crystallographic phase of TMDs, allowing them to transform the phase of single TMD crystals at will. (12,13) This new and emerging field is now termed the field of phase-engineering and opens the door to creating on-chip 2D circuitry with 2D metals and semiconductors.
In the literature, multiple methods are used to induce a 2H to 1T′ phase transformation for different TMDs, such as crystal deformation, (14−17) electrostatic doping, (18) chemical doping, (19,20) laser heating, (21) and so forth. In particular, MoTe2 gained much attention as the energy barrier between the 2H and 1T′ phase is the smallest (∼40 meV). (22) This low-energy barrier allows, for moderate laser power, to induce a phase change in MoTe2. It was demonstrated, using X-ray and electron microscopy techniques, that the 2H phase evolves into a 1T′ by thermodynamically driven processes, changing the system into a 2D lateral junction. (21,23,24) It was shown electrically that the Schottky barrier, present when directly contacting the 2H TMD with metallic contacts, is significantly reduced when contacting a 2H TMD via a phase-transformed 1T′ region, (19,21,25,26) with these two-dimensional lateral junctions, approaching the quantum limit for the contact resistance. (27) The possibility of fabricating high-quality contacts for a 2D semiconductor is essential to increase the speed of optoelectronic devices based on these materials.
Apart from electrical characterization, only a few reports explored the benefits of local phase engineering on the optoelectronic performance of TMD devices. (28−30) Lin et al. report an increased responsivity for 2H MoTe2 devices using 1T′ interlayer contacts. (29) However, no scanning photocurrent measurements are performed, which makes it difficult to disentangle the possible microscopic mechanism involved in the photocurrent generation to either the photovoltaic effect (PVE), due to the build in electric field at the Schottky barriers, or the photothermal effect (PTE), due to the different Seebeck coefficients of the 2H and 1T′ regions. (9,31,32)
Here, we perform scanning photocurrent measurements on 1T′–2H MoTe2 heterojunction devices, which allow us to spatially resolve the areas involved in the photocurrent generation, giving insights into the underlying mechanisms involved. First, we phase-transform the sides of an exfoliated 2H MoTe2 crystal to a 1T′ phase using local heating by laser irradiation, which allows us to contact 2H MoTe2 via the semimetallic 1T′ regions. We find a clear nonlinear behavior for the 1T′–contacted 2H region, indicative of a Schottky barrier between the 1T′ and 2H regions. Additionally, using the scanning photocurrent measurements, we clearly observe that the photocurrents are generated at the 1T′–2H junction rather than at the Ti/Au electrodes or the 1T′ region. More specifically, we find that the peak of the photocurrent is generated at the 2H side of the junction, which suggests that the observed photocurrents in the 1T′–2H junction can be attributed to the PVE rather than the PTE. Lastly, we characterize the optoelectronic performance of the MoTe2 photodetector by performing time-resolved and laser power-dependent photocurrent measurements. We find fast rise and fall times of 113 and 110 μs, respectively, over a broad spectral range of 700–1100 nm. By comparing our 1T′–2H MoTe2 photodetector to a 2H MoTe2 diode where the electrodes are directly deposited on the 2H MoTe2 crystal, we are able to show that the temporal response of 1T′–contacted 2H MoTe2 is 2 orders of magnitude faster. This indicates that phase-engineering can be considered another tool for improving the performance of TMD-based optoelectronic devices.

Results and Discussion

Click to copy section linkSection link copied!

Raman Spectroscopy

The device used to perform the optoelectronic measurements is depicted in Figure 1a. The green region is the untreated 2H MoTe2, while the dark green regions, indicated by the dashed white line, were irradiated with a laser to induce the phase transformation from 2H to 1T′ (details of the device fabrication can be found in the Methods section). To confirm the phase transformation, we performed Raman spectroscopy measurements, as depicted in Figure 1b. Before laser irradiation, we observe the in-plane E2g mode at 235 cm–1 and an out-of-plane Ag mode near 174 cm–1, indicative of the 2H MoTe2 phase. After laser irradiation, these peaks are strongly suppressed, and we observe two new peaks at 124 and 138 cm–1, corresponding to the Ag mode of 1T′ MoTe2. This significant change in the Raman spectrum indicates the successful phase transformation of the irradiated regions. To quantify the degree of phase change in our devices, we calculated the spectral weight of each peak in the phase-changed device (see Supporting Information Figure S5 and Table S1), finding values of 0.01 for the peak at ∼232 cm–1 and of 2.14 for the peak at ∼124 cm–1, corresponding to the 2H and the 1T′ phases, respectively. Our results are in agreement with previous reports, which extensively characterized the laser-induced phase change with several techniques additional to Raman spectroscopy, including X-ray diffraction and electron microscopy. (21,23,24)

Figure 1

Figure 1. (a) Optical micrograph of a phase-changed MoTe2 device, where the phase-changed regions are outlined with the white dashed line, while the bright green part is the unaltered 2H MoTe2 region. (b) Raman spectra obtained before (green) and after (purple) the phase transformation, which clearly indicate a successful phase transformation. The spectra before the phase change is multiplied by 3 for clarity. (c) IdsVds measurements, as indicated in (a), with Vg ranging from 0 to 50 V, taken at 78 K. The nonlinear IV characteristics show the Schottky behavior. The IV measurement for the two 1T′ regions are depicted in the inset, which clearly show Ohmic behavior. (d) Transfer curve measured with a Vds of 3 V, taken at 78 K, shows a clear n-type behavior.

Electrical Characterization

After fabricating electrical contacts to the phase-changed region, we electrically characterized the device in a cryostat at 78 K. We swept the drain-source voltage (Vds) and measured the drain-source current (Ids) for the different regions (both 1T′ and the 1T′–2H junction). In Figure 1c, the 2-probe IdsVds measurements are shown for the 1T′–2H–1T′ junction at different gate voltages, ranging from 0 to 50 V. We observe a clear nonlinear behavior for the Ids as a function of the Vds, indicative of a Schottky barrier present in our device, which could either be between the Ti/Au contact and the 1T′–MoTe2, or the 1T′–2H junction. To determine this, we performed the IdsVds measurement on the phase-transformed 1T′ region only and observe a clear linear behavior showing Ohmic contact between the Ti/Au contacts and 1T′ region, as depicted in the inset of Figure 1c. Therefore, we expect the nonlinear behavior observed in Figure 1c to originate from a Schottky barrier between the 1T′–2H junction. Figure 1d displays the gate transfer curve of the 1T′–2H–1T′ device at Vds = 3 V, showing a clear n-type transistor behavior and a threshold voltage of Vth = 40 V, confirming the semiconducting nature of the 2H phase. For the 1T′ regions, we find a two-probe resistance of 8 and 18 kΩ, which again indicate the successful transformation from the semiconducting 2H MoTe2 phase to the semimetallic 1T′ phase.

Optoelectrical Characterization

To observe the optoelectronic response of the 1T′–2H–1T′ sample, we perform both scanning photocurrent measurements and time-resolved photocurrent measurements, as described in the Methods section. Unless otherwise stated, the optoelectrical measurements were performed at room temperature and high vacuum (1 × 10–6 mbar) conditions. First, the scanning photocurrent measurements are presented, which give more insights into the origin of the photocurrent, after which the time-resolved photocurrent measurements are discussed.
The scanning photocurrent measurements enable us to spatially identify where the photocurrent is generated and thus allow us to check whether the photocurrent originates from the Ti/Au contacts or from the MoTe2 flake itself. When performing the scanning photocurrent measurements, the laser beam is focused and scanned across the sample in a raster-like fashion while recording both the reflection and the generated photocurrent. Figure 2a shows the recorded reflection map with an illumination wavelength of 700 nm, a power of 1 μW, and a full width at half-maximum (FHWM) spot size of 0.70 ± 0.02 μm (see Supporting Information). The contours of the flake and the Ti/Au contacts are clearly visible and highlighted with the white outlines for clarity. It should be noted that the different phases of MoTe2 can also be clearly distinguished, and their junctions are highlighted with the white dashed line. The corresponding photocurrent maps with a Vds of −2, 0, and 2 V are depicted in Figure 2b–d, respectively. We clearly observe that the photocurrent originates locally from the 1T′–2H junction, rather than the Ti/Au contacts, which is the case when directly contacting the 2H MoTe2. This, again, confirms a low Schottky barrier between the Ti/Au and 1T′ MoTe2, indicating a successful phase transformation.

Figure 2

Figure 2. (a) Reflectivity map of the scanning photocurrent measurement of the device depicted in Figure 1a with the corresponding photocurrent map in (b), (c), and (d), taken at RT. The white outlines indicate the position of the flake and Ti/Au contacts, while the white dashed lines indicate the 1T′–2H junctions for clarity. The photocurrent maps are obtained with λ = 700 nm, P = 1 μW, and a Vds of (b) −2, (c) 0, and (d) 2 V. From the photocurrent maps, we can clearly see that the induced photocurrent originates from the 1T′–2H junction rather than from the Ti/Au contacts.

We observe photocurrents with opposite signs at the two 1T′-2H junctions for Vds = 0 V. This is in line with the expected behavior for two possible mechanisms: photovoltaic effect (PVE) due to a Schottky barrier at the 1T′–2H junction or the photothermoelectric effect (PTE) due to a different Seebeck coefficient of the two MoTe2 phases. A more detailed analysis of the influence of the Vds and the Vg on the photocurrent is shown in the Supporting Information. The two mechanisms are schematically depicted in Figure 3. While both the PVE and the PTE can give rise to photocurrent in the 2H–1T′ interface, we conclude that the PTE contribution should be negligible. For the PTE, local heating due to the laser irradiation causes a temperature gradient (ΔT) which is converted into a voltage difference (VPTE) due to a difference in the Seebeck coefficient between the 1T′ (S1T) and 2H (S2H) phases. (9,33) By using the maximum-generated photocurrent in our measurements, we calculate a range of unrealistic temperature gradients above 7000 K, indicating that the PTE cannot solely explain the observed photocurrent. A more detailed explanation can be found in the Supporting Information.

Figure 3

Figure 3. (a) Schematic of the photothermoelectric (PTE) and photovoltaic effect (PVE). For the PTE, the laser locally heats the device, which creates a temperature gradient, which via the Seebeck effect causes an induced photocurrent (IPTE). For the PVE, the localized electric field at the Schottky barrier causes a separation of the photoinduced carriers, resulting in IPVE. It should be noted that the two effects produce a photocurrent with the same direction and that the induced photocurrent is opposite on both junctions. (b) Line trace of the photocurrent mapping, as indicated in red in Figure 2c, showing the reflection (white) and negative (purple) and positive (green) photocurrent peaks along the line trace. Both the maximum and minimum photocurrents are obtained in the 2H region, as expected for a localized electric field from a Schottky barrier. The phases are indicated by the purple (1T′) and green (2H) backgrounds.

On the other hand, for the PVE-driven photocurrent, the localized electric field at the 1T′–2H interface causes the photoinduced electron–hole pairs to separate, resulting in a photocurrent. (33) For our n-type MoTe2, the band alignment is depicted in Figure 3a. The electric field from the Schottky barrier is positioned in the 2H region. By taking a line scan of the reflection map and photocurrent map, indicated by the red line in Figure 2c, we can more accurately determine the position of the photocurrent peak with respect to the 1T′–2H junction. Here, we find that the peak of the photocurrent arises in the 2H region rather than at the 1T′–2H junction, which is in line with the expectation for the PVE. (32) Additionally, we can estimate the depletion region due to the Schottky barrier by using the following equation:
W=2ϵ0ϵrϕbieNd
(1)
where W is the width of the Schottky barrier, e is the electron charge, and ϵ0 is the permittivity of free space. By assuming a donor density Nd of 1011 cm2, (34) a barrier height ϕbi of 60 meV, (29) and a relative ϵr of 12, we estimate W to be ∼2 μm, which corresponds well to the FWHM of 1.5 ± 0.2 μm we find by fitting the positive photocurrent peak with a Gaussian.
Buscema et al. observed a similar photocurrent sign in their scanning photocurrent measurements on an n-type MoS2 photodiode. (35) However, they attribute the induced photocurrent to the PTE, as they observe clear photocurrent generation in the center of their Ti/Au contacts and see a linear IdsVds behavior with no indication of a Schottky barrier. In contrast, for our 1T′-2H MoTe2 junctions, we observe a clear nonlinear IV-curve, indicating that the Schottky barrier plays a more important role in our devices. Furthermore, they see a pronounced photocurrent even when exciting below the bandgap of MoS2. Unfortunately, our optoelectronic setup only allows for excitation up to 1100 nm, which is still above the bandgap of MoTe2 (∼1.1 eV ∝ ∼1127 nm for bulk). (36,37) Therefore, we suggest further research to be performed on below-band gap excitation to determine to what extent the PTE is contributing to the observed photocurrent.
To characterize the optoelectronic performance of our MoTe2 photodetector, we perform time-resolved and power-dependent photocurrent measurements, as depicted in Figure 4. By measuring the induced photocurrent versus time, using a chopper to modulate the light on and off (see Figure 4a), we are able to extract the rise (τr) and fall times (τf) of the device, which are defined as the time required for the photocurrent to increase from 10 to 90%, and decrease from 90 to 10% of the maximum photocurrent, respectively. By performing these measurements over a range of different excitation wavelengths, we find that we get short rise and fall times of ∼113 and ∼110 μs, respectively, independent of the wavelength, as depicted in Figure 4b. These response times correspond to a 3 dB frequency of 0.35/τr = 3 kHz, (38) which are close to the performance of graphene/MoTe2/graphene photodetectors. (39) In contrast, when directly contacting the 2H MoTe2 with Ti/Au electrodes, we find a much slower response (over 100-fold slower), as shown in Figure S2c in the Supporting Information. Here, we observe a waveform similar to a “sawtooth” response, indicative of capacitive behavior, at 20 Hz. This is a result of the highly resistive contact with the 2H region of the device, leading to photogating and slow charging of the device. This shows that using the 1T′ regions to contact the 2H MoTe2 increases the temporal response of our MoTe2 photodetector by more than 2 orders of magnitude.

Figure 4

Figure 4. (a) Time-resolved photocurrent, taken at RT, where the photocurrent (purple) in the device is plotted together with the chopper signal (gray) versus time, shows the fast response of our MoTe2 photodetector. The dashed gray rectangles indicate the region used to determine the rise and fall times, as depicted in the inset of (b). (b) Extracted rise (purple) and fall times (green) indicate no wavelength dependence on the fast response for wavelengths ranging from 700 to 1100 nm. The inset shows the rise (purple) and fall (green) curves of the photocurrent from which the rise and fall times are extracted. (c) Power-dependent measurements for different Vds, ranging from −2 to 2 V, with a maximum responsivity of 4.5 × 10–8 A/μW. Here, the responsivity (R) of the device is plotted as a function of the laser excitation power (P) and fitted at high laser excitation power to a power law: RPα–1. The measured R for 700 and 1068 nm are indicated by the filled circles and unfilled squares, respectively. (d) Extracted index of the power law (α) from the fitting in (c) versus Vds for the wavelengths 700 nm (purple) and 1068 nm (green).

The response dynamics displayed at the 1T′–2H junction are fast compared to other TMD-based photodetectors. (38,40) More specifically, compared to other MoTe2-based photodetectors, they are 1 order of magnitude faster than the report of Huang et al. on directly contacted 2H MoTe2, (41) and similar to the ones found by Lin et al. in 1T′–contacted 2H MoTe2. (29) On other TMD-based devices, a variety of different response dynamics are reported, with the fastest responses reported for deep UV and mid-IR detectors on graphene/MoTe2/black phosphorus devices, which reach bandwidths of 2.1 MHz. (42)
To determine the responsivity of our devices, we vary the excitation power at a fixed excitation wavelength (700 and 1068 nm). From these power-dependent measurements, we are able to determine the responsivity by R = IPC/P, where IPC is the induced photocurrent, and P is the power of the laser, (9,33) and find a maximum R of 4.5 × 10–8 A/μW with a wavelength of 700 nm and a 2 V bias. The value we find is comparable to other reports on TMD-based photodetectors, ranging approximately from 7.25 × 10–11 A/μW to 9.708 × 10–3 A/μW. (40,41,43,44) For the same wavelength and bias voltage, we also calculate a maximum external quantum efficiency (EQE) of ∼8%. Given that there is commonly a trade-off between fast response and high responsivity in these devices, (38) it is not unexpected that both the EQE and the responsivity of our devices fall within the middle or lower half of reported values. (30,39,45) Additionally, the responsivity of our device is measured with a focused laser spot rather than illuminating the entire device. This could lead to an underestimation of the generated photocurrent, as only a small fraction of the photodetector area is used to generate the photocurrent. Moreover, the capacitance added by the global back-gate can lead to an increase in the response times of our device.
Figure 4c clearly shows a decrease of responsivity with incident excitation power for P > 1.9 μW, which is commonly observed in TMD photodetectors. (46,47) It can be associated with a reduced number of photogenerated carriers available for extraction under high photon flux due to the saturation of recombination/trap states that influence the lifetime of the generated carriers. (48) The responsivity versus laser power can be expressed by a power law RPdα–1 for P > 1.9 μW, where P is the laser power, and α is the index of the power law. (41,49) From the fit, we are able to extract α for the two different wavelengths at different Vds values, as shown in Figure 4d. The deviation from the ideal slope of α = 1, where the responsivity does not depend on the laser power, can be attributed to complex processes in the carrier generation, trapping, and electron–hole recombination in MoTe2. (50,51) For the PTE, a value of α ∼ 0.8 is expected, while we find a value of α ∼ 0.25, which indicates again that the PTE is not primarily responsible for the generated photocurrent in our device. (31)

Conclusions

Click to copy section linkSection link copied!

In conclusion, laser-induced phase transformation is a simple, scalable, and reliable methodology to engineer MoTe2 optoelectronic devices. Our results indicate that contacting the 2H region of MoTe2 via a phase-transformed 1T′ region is beneficial for the temporal optoelectronic response of MoTe2-based photodetectors and does not require complex heterostructure fabrication. Our scanning photocurrent measurements and nonlinear IV curves clearly show that the origin of the photocurrent in our devices can be ascribed to the Schottky barrier between the 1T′ and 2H junction, rather than the photothermoelectric effect or Schottky barriers at the Ti/Au electrode-TMD interface. Contacting MoTe2 via the phase-transformed 1T′ region, therefore, allows one to study the intrinsic properties of the TMD rather than the electrode-TMD interactions, beneficial for fundamental research. Additionally, an increase of 2 orders of magnitude in the optoelectronic temporal response is observed when contacting the 2H MoTe2 via the 1T′ regions. This shows that tailoring the crystallographic phase of TMDs locally and altering their optoelectronic response at will can have a profitable effect on the optoelectronic operation. Our results, in combination with the wide variety of phase-engineering techniques and different TMDs available, could lead to a further improved performance of TMD-based optoelectronic devices, leading to more sensitive, faster, and flexible photodetectors.

Methods

Click to copy section linkSection link copied!

Device Fabrication

The 2H MoTe2 flakes are obtained by mechanical exfoliation (bulk crystal supplied by HQ graphene) and transferred onto a Si/SiO2 (285 nm) substrate in a nitrogen environment. Using an optical microscope, the MoTe2 flakes are selected based on their size, thickness, and homogeneous surface. Details of the selection criteria, as well as Figure S4, a discussion on the impact of the flake thickness to the laser-induced phase change, can be found in the Supporting Information. Next, the Raman spectra are obtained with an inVia Raman Renishaw microscope using a linearly polarized laser in backscattering geometry. The excitation wavelength and grating used were λ = 532 nm and 2400 l/mm, respectively. The laser power was ∼100 μW with a diffraction-limited spot of ∼1 μm. Using the same system, the 2H–1T′ phase transformation is performed by selectively illuminating parts of the MoTe2 flake with the 532 nm laser beam in a raster-like fashion, using steps of 500 nm and 0.1 s illumination. We find that a laser power of ≥3.25 mW (laser spot size around 500 nm) is needed to initiate the phase transformation. Finally, using standard lithography techniques, the Ti/Au (5/55 nm) contacts are fabricated on top of the flake by means of electron beam lithography and electron beam evaporation.

Optoelectronic Measurements

The electrical characterization (i.e., IV-sweeps, transfer curves) is performed using Keithleys 2400 and 2450 source measure units at 78 K. For the optoelectronic measurements, a supercontinuum white light laser (NKT Photonics SuperK EXTREME) is used as the illumination source, and the measurements are taken at room temperature. The induced photocurrent is measured in a short-circuit configuration using a Stanford Research Systems SR830 lock-in amplifier, which is referenced to the frequency of the optical chopper. The photocurrents are either measured directly by the lock-in amplifier or converted to a voltage using a home build current preamplifier, which is subsequently measured by the lock-in amplifier. The time-resolved photoresponse of the device, as depicted in Figure 3a, is measured using a chopper and an oscilloscope (Keysight DSOX1204A) at room temperature.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsphotonics.4c00896.

  • Details over the photothermoelectric temperature gradient calculation; laser spot size calculation and data for 700 and 1064 nm for the reported measurements; equation and additional details for charge carrier mobility and further details of another phase-engineered device, including responsivity and transfer curves; drain-source and gate voltage dependence of the photocurrent; normalized Raman spectra and spectral weight contribution of the 1T′ phase characteristic peaks; laser irradiation power; and flake thickness dependence for phase change (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
  • Authors
    • Jan Hidding - Zernike Institute for Advanced Materials, University of Groningen, 9747 AG Groningen, The Netherlands
    • Cédric A. Cordero-Silis - Zernike Institute for Advanced Materials, University of Groningen, 9747 AG Groningen, The NetherlandsOrcidhttps://orcid.org/0000-0002-0842-3169
    • Daniel Vaquero - Nanotechnology Group, USAL─Nanolab, Universidad de Salamanca, E-37008 Salamanca, Spain
    • Konstantinos P. Rompotis - Zernike Institute for Advanced Materials, University of Groningen, 9747 AG Groningen, The Netherlands
    • Jorge Quereda - Departamento de Física de Materiales, GISC, Universidad Complutense de Madrid, E-28040 Madrid, Spain
  • Author Contributions

    J.H. and C.A.C.-S. contributed equally. J.H. and C.A.C.-S. fabricated the samples, and together with D.V. performed both the electrical and optical measurements under the supervision of M.H.D.G.; K.R. joined to perform measurements and sample characterization on additional samples also under the supervision of M.H.D.G.; J.H. performed the data analysis and, together with C.A.C.-S and M.H.D.G., wrote the paper with comments from all authors.

  • Funding

    This work was supported by the Dutch Research Council (NWO-STU.019.014), the European Union (ERC, 2D-OPTOSPIN, 101076932), the “Materials for the Quantum Age”(QuMat) program (Registration No. 024.005.006) which is part of the Gravitation program financed by the Dutch Ministry of Education, Culture and Science (OCW), the Zernike Institute for Advanced Materials, and the innovation program under grant agreement No. 881603 (Graphene Flagship).

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

The authors acknowledge Prof. M. A. Loi and E. K. Tekelenburg for their help with the Raman measurements and thank J. G. Holstein, H. Adema, H. de Vries, A. Joshua, and F. H. van der Velde for their technical support. Sample fabrication was performed using NanoLabNL facilities.

References

Click to copy section linkSection link copied!

This article references 52 other publications.

  1. 1
    Miao, J.; Hu, W.; Guo, N.; Lu, Z.; Zou, X.; Liao, L.; Shi, S.; Chen, P.; Fan, Z.; Ho, J. C.; Li, T.-X.; Chen, X. S.; Lu, W. Single InAs Nanowire Room-Temperature Near-Infrared Photodetectors. ACS Nano 2014, 8, 36283635,  DOI: 10.1021/nn500201g
  2. 2
    Huang, W.; Xing, C.; Wang, Y.; Li, Z.; Wu, L.; Ma, D.; Dai, X.; Xiang, Y.; Li, J.; Fan, D.; Zhang, H. Facile Fabrication and Characterization of Two-Dimensional Bismuth(III) Sulfide Nanosheets for High-Performance Photodetector Applications Under Ambient Conditions. Nanoscale 2018, 10, 24042412,  DOI: 10.1039/C7NR09046C
  3. 3
    Ilyas, N.; Li, D.; Song, Y.; Zhong, H.; Jiang, Y.; Li, W. Low-Dimensional Materials and State-of-the-Art Architectures for Infrared Photodetection. Sensors 2018, 18, 4163,  DOI: 10.3390/s18124163
  4. 4
    Huang, W.; Zhang, Y.; You, Q.; Huang, P.; Wang, Y.; Huang, Z. N.; Ge, Y.; Wu, L.; Dong, Z.; Dai, X.; Xiang, Y.; Li, J.; Zhang, X.; Zhang, H. Enhanced Photodetection Properties of Tellurium@Selenium Roll-to-Roll Nanotube Heterojunctions. Small 2019, 15, 1900902  DOI: 10.1002/smll.201900902
  5. 5
    Rojas-Lopez, R. R.; Brant, J. C.; Ramos, M. S. O.; Castro, T. H. L. G.; Guimarães, M. H. D.; Neves, B. R. A.; Guimarães, P. S. S. Photoluminescence and Charge Transfer in the Prototypical 2D/3D Semiconductor Heterostructure MoS2/GaAs. Appl. Phys. Lett. 2021, 119, 233101,  DOI: 10.1063/5.0068548
  6. 6
    Zhai, X.-P.; Ma, B.; Wang, Q.; Zhang, H.-L. 2D Materials Towards Ultrafast Photonic Applications. Phys. Chem. Chem. Phys. 2020, 22, 2214022156,  DOI: 10.1039/D0CP02841J
  7. 7
    An, J.; Zhao, X.; Zhang, Y.; Liu, M.; Yuan, J.; Sun, X.; Zhang, Z.; Wang, B.; Li, S.; Li, D. Perspectives of 2D Materials for Optoelectronic Integration. Adv. Funct. Mater. 2022, 32, 2110119  DOI: 10.1002/adfm.202110119
  8. 8
    Liang, G.; Yu, X.; Hu, X.; Qiang, B.; Wang, C.; Wang, Q. J. Mid-Infrared Photonics and Optoelectronics in 2D Materials. Mater. Today 2021, 51, 294316,  DOI: 10.1016/j.mattod.2021.09.021
  9. 9
    Buscema, M.; Island, J. O.; Groenendijk, D. J.; Blanter, S. I.; Steele, G. A.; Van Der Zant, H. S.; Castellanos-Gomez, A. Photocurrent Generation with Two-Dimensional van der Waals Semiconductors. Chem. Soc. Rev. 2015, 44, 36913718,  DOI: 10.1039/C5CS00106D
  10. 10
    Mak, K. F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T. F. Atomically Thin MoS2: A New Direct-Gap Semiconductor. Phys. Rev. Lett. 2010, 105, 136805  DOI: 10.1103/PhysRevLett.105.136805
  11. 11
    Manzeli, S.; Ovchinnikov, D.; Pasquier, D.; Yazyev, O. V.; Kis, A. 2D Transition Metal Dichalcogenides. Nat. Rev. Mater. 2017, 2, 17033,  DOI: 10.1038/natrevmats.2017.33
  12. 12
    Aftab, S.; Shehzad, M. A.; Salman Ajmal, H. M.; Kabir, F.; Iqbal, M. Z.; Al-Kahtani, A. A. Bulk Photovoltaic Effect in Two-Dimensional Distorted MoTe2. ACS Nano 2023, 17, 1788417896,  DOI: 10.1021/acsnano.3c03593
  13. 13
    Hu, X.; Zhang, F.; Hu, Z.; He, P.; Tao, L.; Zheng, Z.; Zhao, Y.; Yang, Y.; He, J. Preparation of 1T’- and 2H–MoTe2 Films and Investigation of their Photoelectric Properties and Ultrafast Photocarrier Dynamics. Opt. Mater. 2023, 136, 113467  DOI: 10.1016/j.optmat.2023.113467
  14. 14
    Dave, M.; Vaidya, R.; Patel, S. G.; Jani, A. R. High Pressure Effect on MoS2 and MoSe2 Single Crystals Grown by CVT Method. Bulletin of Materials Science 2004, 27, 213216,  DOI: 10.1007/BF02708507
  15. 15
    Song, S.; Keum, D. H.; Cho, S.; Perello, D.; Kim, Y.; Lee, Y. H. Room Temperature Semiconductor-Metal Transition of MoTe2 Thin Films Engineered by Strain. Nano Lett. 2016, 16, 188193,  DOI: 10.1021/acs.nanolett.5b03481
  16. 16
    Lin, Y.-C.; Dumcenco, D. O.; Huang, Y.-S.; Suenaga, K. Atomic Mechanism of the Semiconducting-to-Metallic phase Transition in Single-Layered MoS2. Nat. Nanotechnol. 2014, 9, 391396,  DOI: 10.1038/nnano.2014.64
  17. 17
    Shang, B.; Cui, X.; Jiao, L.; Qi, K.; Wang, Y.; Fan, J.; Yue, Y.; Wang, H.; Bao, Q.; Fan, X.; Wei, S.; Song, W.; Cheng, Z.; Guo, S.; Zheng, W. Lattice-Mismatch-Induced Ultrastable 1T-Phase MoS2-Pd/Au for Plasmon-Enhanced Hydrogen Evolution. Nano Lett. 2019, 19, 27582764,  DOI: 10.1021/acs.nanolett.8b04104
  18. 18
    Wang, Y.; Xiao, J.; Zhu, H.; Li, Y.; Alsaid, Y.; Fong, K. Y.; Zhou, Y.; Wang, S.; Shi, W.; Wang, Y.; Zettl, A.; Reed, E. J.; Zhang, X. Structural Phase Transition in Monolayer MoTe2 Driven by Electrostatic Doping. Nature 2017, 550, 487491,  DOI: 10.1038/nature24043
  19. 19
    Kappera, R.; Voiry, D.; Yalcin, S. E.; Branch, B.; Gupta, G.; Mohite, A. D.; Chhowalla, M. Phase-Engineered Low-Resistance Contacts for Ultrathin MoS2 transistors. Nat. Mater. 2014, 13, 11281134,  DOI: 10.1038/nmat4080
  20. 20
    Ma, Y.; Liu, B.; Zhang, A.; Chen, L.; Fathi, M.; Shen, C.; Abbas, A. N.; Ge, M.; Mecklenburg, M.; Zhou, C. Reversible Semiconducting-to-Metallic Phase Transition in Chemical Vapor Deposition Grown Monolayer WSe2 and Applications for Devices. ACS Nano 2015, 9, 73837391,  DOI: 10.1021/acsnano.5b02399
  21. 21
    Cho, S.; Kim, S.; Kim, J. H.; Zhao, J.; Seok, J.; Keum, D. H.; Baik, J.; Choe, D.-H.; Chang, K. J.; Suenaga, K.; Kim, S. W.; Lee, Y. H.; Yang, H. Phase Patterning for Ohmic Homojunction Contact in MoTe2. Science 2015, 349, 625628,  DOI: 10.1126/science.aab3175
  22. 22
    Duerloo, K.-A. N.; Li, Y.; Reed, E. J. Structural Phase Transitions in Two-Dimensional Mo- and W-Dichalcogenide Monolayers. Nat. Commun. 2014, 5, 4214,  DOI: 10.1038/ncomms5214
  23. 23
    Tan, Y.; Luo, F.; Zhu, M.; Xu, X.; Ye, Y.; Li, B.; Wang, G.; Luo, W.; Zheng, X.; Wu, N.; Yu, Y.; Qin, S.; Zhang, X.-A. Controllable 2H-to-1T’ Phase Transition in Few-Layer MoTe2. Nanoscale 2018, 10, 1996419971,  DOI: 10.1039/C8NR06115G
  24. 24
    Kang, S.; Won, D.; Yang, H.; Lin, C.-H.; Ku, C.-S.; Chiang, C.-Y.; Kim, S.; Cho, S. Phase-controllable laser thinning in MoTe2. Appl. Surf. Sci. 2021, 563, 150282  DOI: 10.1016/j.apsusc.2021.150282
  25. 25
    Zhang, X. Low Contact Barrier in 2H/1T’ MoTe2 In-Plane Heterostructure Synthesized by Chemical Vapor Deposition. ACS Appl. Mater. Interfaces 2019, 11, 1277712785,  DOI: 10.1021/acsami.9b00306
  26. 26
    Bae, G. Y.; Kim, J.; Kim, J.; Lee, S.; Lee, E. MoTe2 Field-Effect Transistors with Low Contact Resistance through Phase Tuning by Laser Irradiation. Nanomaterials 2021, 11, 2805,  DOI: 10.3390/nano11112805
  27. 27
    Wang, Y.; Chhowalla, M. Making Clean Electrical Contacts on 2D Transition Metal Dichalcogenides. Nature Reviews Physics 2022, 4, 101112,  DOI: 10.1038/s42254-021-00389-0
  28. 28
    Yamaguchi, H.; Blancon, J. C.; Kappera, R.; Lei, S.; Najmaei, S.; Mangum, B. D.; Gupta, G.; Ajayan, P. M.; Lou, J.; Chhowalla, M.; Crochet, J. J.; Mohite, A. D. Spatially Resolved Photoexcited Charge-Carrier Dynamics in Phase-Engineered Monolayer MoS2. ACS Nano 2015, 9, 840849,  DOI: 10.1021/nn506469v
  29. 29
    Lin, D.-Y.; Hsu, H.-P.; Liu, G.-H.; Dai, T.-Z.; Shih, Y.-T. Enhanced Photoresponsivity of 2H-MoTe2 by Inserting 1T-MoTe2 Interlayer Contact for Photodetector Applications. Crystals 2021, 11, 964,  DOI: 10.3390/cryst11080964
  30. 30
    Ding, Y.; Qi, R.; Wang, C.; Wu, Q.; Zhang, H.; Zhang, X.; Lin, L.; Cai, Z.; Xiao, S.; Gu, X.; Nan, H. Broad-Band Photodetector Based on a Lateral MoTe2 1T-2H-1T Homojunction. J. Phys. Chem. C 2023, 127, 2007220081,  DOI: 10.1021/acs.jpcc.3c05592
  31. 31
    Xu, X.; Gabor, N. M.; Alden, J. S.; Van Der Zande, A. M.; McEuen, P. L. Photo-thermoelectric Effect at a Graphene Interface Junction. Nano Lett. 2010, 10, 562566,  DOI: 10.1021/nl903451y
  32. 32
    Zhang, Y.; Li, H.; Wang, L.; Wang, H.; Xie, X.; Zhang, S. L.; Liu, R.; Qiu, Z. J. Photothermoelectric and Photovoltaic Effects Both Present in MoS2. Sci. Rep. 2015, 5, 7938,  DOI: 10.1038/srep07938
  33. 33
    Huo, N.; Konstantatos, G. Recent Progress and Future Prospects of 2D-Based Photodetectors. Adv. Mater. 2018, 30, 1801164  DOI: 10.1002/adma.201801164
  34. 34
    HQgraphene., MoTe2 (2H Molybdenum Ditelluride). https://www.hqgraphene.com/MoTe2.php (accessed May 23 2023).
  35. 35
    Buscema, M.; Barkelid, M.; Zwiller, V.; van der Zant, H. S. J.; Steele, G. A.; Castellanos-Gomez, A. Large and Tunable Photothermoelectric Effect in Single-Layer MoS2. Nano Lett. 2013, 13, 358363,  DOI: 10.1021/nl303321g
  36. 36
    Ruppert, C.; Aslan, O. B.; Heinz, T. F. Optical Properties and Band Gap of Single- and Few-Layer MoTe2 Crystals. Nano Lett. 2014, 14, 62316236,  DOI: 10.1021/nl502557g
  37. 37
    Lezama, I. G.; Arora, A.; Ubaldini, A.; Barreteau, C.; Giannini, E.; Potemski, M.; Morpurgo, A. F. Indirect-to-Direct Band Gap Crossover in Few-Layer MoTe2. Nano Lett. 2015, 15, 23362342,  DOI: 10.1021/nl5045007
  38. 38
    Yadav, P.; Dewan, S.; Mishra, R.; Das, S. Review of Recent Progress, Challenges, and Prospects of 2D Materials-Based Short Wavelength Infrared Photodetectors. J. Phys. D: Appl. Phys. 2022, 55, 313001,  DOI: 10.1088/1361-6463/ac6635
  39. 39
    Zhang, K.; Fang, X.; Wang, Y.; Wan, Y.; Song, Q.; Zhai, W.; Li, Y.; Ran, G.; Ye, Y.; Dai, L. Ultrasensitive Near-Infrared Photodetectors Based on a Graphene-MoTe2-Graphene Vertical van der Waals Heterostructure. ACS Appl. Mater. Interfaces 2017, 9, 53925398,  DOI: 10.1021/acsami.6b14483
  40. 40
    Xiao, H.; Lin, L.; Zhu, J.; Guo, J.; Ke, Y.; Mao, L.; Gong, T.; Cheng, H.; Huang, W.; Zhang, X. Highly Sensitive and Broadband Photodetectors Based on WSe2/MoS2 Heterostructures with van der Waals Contact Electrodes. Appl. Phys. Lett. 2022, 121, 023504  DOI: 10.1063/5.0100191
  41. 41
    Huang, H. Highly Sensitive Visible to Infrared MoTe2 Photodetectors Enhanced by the Photogating Effect. Nanotechnology 2016, 27, 445201  DOI: 10.1088/0957-4484/27/44/445201
  42. 42
    Shen, D. High-Performance Mid-IR to Deep-UV van der Waals Photodetectors Capable of Local Spectroscopy at Room Temperature. Nano Lett. 2022, 22, 34253432,  DOI: 10.1021/acs.nanolett.2c00741
  43. 43
    Yu, W.; Li, S.; Zhang, Y.; Ma, W.; Sun, T.; Yuan, J.; Fu, K.; Bao, Q. Near-Infrared Photodetectors Based on MoTe2/Graphene Heterostructure with High Responsivity and Flexibility. Small 2017, 13, 1700268  DOI: 10.1002/smll.201700268
  44. 44
    Yin, L.; Zhan, X.; Xu, K.; Wang, F.; Wang, Z.; Huang, Y.; Wang, Q.; Jiang, C.; He, J. Ultrahigh Sensitive MoTe2 Phototransistors Driven by Carrier Tunneling. Appl. Phys. Lett. 2016, 108, 043503  DOI: 10.1063/1.4941001
  45. 45
    Xie, Y.; Wu, E.; Zhang, J.; Hu, X.; Zhang, D.; Liu, J. Gate-Tunable Photodetection/Voltaic Device Based on BP/MoTe2 Heterostructure. ACS Appl. Mater. Interfaces 2019, 11, 1421514221,  DOI: 10.1021/acsami.8b21315
  46. 46
    Lee, H. S.; Min, S.-W.; Chang, Y.-G.; Park, M. K.; Nam, T.; Kim, H.; Kim, J. H.; Ryu, S.; Im, S. MoS2 Nanosheet Phototransistors with Thickness-Modulated Optical Energy Gap. Nano Lett. 2012, 12, 36953700,  DOI: 10.1021/nl301485q
  47. 47
    Lopez-Sanchez, O.; Lembke, D.; Kayci, M.; Radenovic, A.; Kis, A. Ultrasensitive Photodetectors Based on Monolayer MoS2. Nat. Nanotechnol. 2013, 8, 497501,  DOI: 10.1038/nnano.2013.100
  48. 48
    Buscema, M.; Groenendijk, D. J.; Blanter, S. I.; Steele, G. A.; Van Der Zant, H. S. J.; Castellanos-Gomez, A. Fast and Broadband Photoresponse of Few-Layer Black Phosphorus Field-Effect Transistors. Nano Lett. 2014, 14, 33473352,  DOI: 10.1021/nl5008085
  49. 49
    Zhang, X.; Liu, B.; Yang, W.; Jia, W.; Li, J.; Jiang, C.; Jiang, X. 3D-Branched Hierarchical 3C-SiC/ZnO Heterostructures for High-Performance photodetectors. Nanoscale 2016, 8, 1757317580,  DOI: 10.1039/C6NR06236A
  50. 50
    Kind, H.; Yan, H.; Messer, B.; Law, M.; Yang, P. Nanowire Ultraviolet Photodetectors and Optical Switches. Adv. Mater. 2002, 14, 158160,  DOI: 10.1002/1521-4095(20020116)14:2<158::AID-ADMA158>3.0.CO;2-W
  51. 51
    Liu, F.; Shimotani, H.; Shang, H.; Kanagasekaran, T.; Zólyomi, V.; Drummond, N.; Fal’ko, V. I.; Tanigaki, K. High-Sensitivity Photodetectors Based on Multilayer GaTe Flakes. ACS Nano 2014, 8, 752760,  DOI: 10.1021/nn4054039
  52. 52
    Hidding, J.; Cordero-Silis, C. A.; Vaquero, D.; Panagiotis Rompotis, K.; Quereda, J.; Guimarães, M. H. D. Locally Phase-Engineered MoTe2 for Near-Infrared Photodetectors . 2024, 2406.01376, arXiv, https://arxiv.org/abs/2406.01376 (accessed Jun 03, 2024).

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 2 publications.

  1. Lingfeng Peng, Jianwei Zou, Jinlin Liu, Hao Liu, Jing Sun, Fan Dang, Ningjiu Zhao, Yongjun Hu, Hailong Chen. Unveiling Defect-Dependent Nonradiative Carrier Dynamics in Few-Layer 2H-MoTe2 by Femtosecond Mid-Infrared Spectroscopy. The Journal of Physical Chemistry C 2025, Article ASAP.
  2. Yuan Gao, Haiyan Nan, Huilin Zuo, Renxian Qi, Zijian Wang, Jialing Jian, Zhengjin Weng, Wenhui Wang, Shaoqing Xiao, Xiaofeng Gu. Enhancing photodetection efficiency in MoTe2/MoS2 van der Waals heterojunctions by modulating the phase regions. Surfaces and Interfaces 2025, 65 , 106511. https://doi.org/10.1016/j.surfin.2025.106511

ACS Photonics

Cite this: ACS Photonics 2024, 11, 10, 4083–4089
Click to copy citationCitation copied!
https://doi.org/10.1021/acsphotonics.4c00896
Published September 16, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

1171

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. (a) Optical micrograph of a phase-changed MoTe2 device, where the phase-changed regions are outlined with the white dashed line, while the bright green part is the unaltered 2H MoTe2 region. (b) Raman spectra obtained before (green) and after (purple) the phase transformation, which clearly indicate a successful phase transformation. The spectra before the phase change is multiplied by 3 for clarity. (c) IdsVds measurements, as indicated in (a), with Vg ranging from 0 to 50 V, taken at 78 K. The nonlinear IV characteristics show the Schottky behavior. The IV measurement for the two 1T′ regions are depicted in the inset, which clearly show Ohmic behavior. (d) Transfer curve measured with a Vds of 3 V, taken at 78 K, shows a clear n-type behavior.

    Figure 2

    Figure 2. (a) Reflectivity map of the scanning photocurrent measurement of the device depicted in Figure 1a with the corresponding photocurrent map in (b), (c), and (d), taken at RT. The white outlines indicate the position of the flake and Ti/Au contacts, while the white dashed lines indicate the 1T′–2H junctions for clarity. The photocurrent maps are obtained with λ = 700 nm, P = 1 μW, and a Vds of (b) −2, (c) 0, and (d) 2 V. From the photocurrent maps, we can clearly see that the induced photocurrent originates from the 1T′–2H junction rather than from the Ti/Au contacts.

    Figure 3

    Figure 3. (a) Schematic of the photothermoelectric (PTE) and photovoltaic effect (PVE). For the PTE, the laser locally heats the device, which creates a temperature gradient, which via the Seebeck effect causes an induced photocurrent (IPTE). For the PVE, the localized electric field at the Schottky barrier causes a separation of the photoinduced carriers, resulting in IPVE. It should be noted that the two effects produce a photocurrent with the same direction and that the induced photocurrent is opposite on both junctions. (b) Line trace of the photocurrent mapping, as indicated in red in Figure 2c, showing the reflection (white) and negative (purple) and positive (green) photocurrent peaks along the line trace. Both the maximum and minimum photocurrents are obtained in the 2H region, as expected for a localized electric field from a Schottky barrier. The phases are indicated by the purple (1T′) and green (2H) backgrounds.

    Figure 4

    Figure 4. (a) Time-resolved photocurrent, taken at RT, where the photocurrent (purple) in the device is plotted together with the chopper signal (gray) versus time, shows the fast response of our MoTe2 photodetector. The dashed gray rectangles indicate the region used to determine the rise and fall times, as depicted in the inset of (b). (b) Extracted rise (purple) and fall times (green) indicate no wavelength dependence on the fast response for wavelengths ranging from 700 to 1100 nm. The inset shows the rise (purple) and fall (green) curves of the photocurrent from which the rise and fall times are extracted. (c) Power-dependent measurements for different Vds, ranging from −2 to 2 V, with a maximum responsivity of 4.5 × 10–8 A/μW. Here, the responsivity (R) of the device is plotted as a function of the laser excitation power (P) and fitted at high laser excitation power to a power law: RPα–1. The measured R for 700 and 1068 nm are indicated by the filled circles and unfilled squares, respectively. (d) Extracted index of the power law (α) from the fitting in (c) versus Vds for the wavelengths 700 nm (purple) and 1068 nm (green).

  • References


    This article references 52 other publications.

    1. 1
      Miao, J.; Hu, W.; Guo, N.; Lu, Z.; Zou, X.; Liao, L.; Shi, S.; Chen, P.; Fan, Z.; Ho, J. C.; Li, T.-X.; Chen, X. S.; Lu, W. Single InAs Nanowire Room-Temperature Near-Infrared Photodetectors. ACS Nano 2014, 8, 36283635,  DOI: 10.1021/nn500201g
    2. 2
      Huang, W.; Xing, C.; Wang, Y.; Li, Z.; Wu, L.; Ma, D.; Dai, X.; Xiang, Y.; Li, J.; Fan, D.; Zhang, H. Facile Fabrication and Characterization of Two-Dimensional Bismuth(III) Sulfide Nanosheets for High-Performance Photodetector Applications Under Ambient Conditions. Nanoscale 2018, 10, 24042412,  DOI: 10.1039/C7NR09046C
    3. 3
      Ilyas, N.; Li, D.; Song, Y.; Zhong, H.; Jiang, Y.; Li, W. Low-Dimensional Materials and State-of-the-Art Architectures for Infrared Photodetection. Sensors 2018, 18, 4163,  DOI: 10.3390/s18124163
    4. 4
      Huang, W.; Zhang, Y.; You, Q.; Huang, P.; Wang, Y.; Huang, Z. N.; Ge, Y.; Wu, L.; Dong, Z.; Dai, X.; Xiang, Y.; Li, J.; Zhang, X.; Zhang, H. Enhanced Photodetection Properties of Tellurium@Selenium Roll-to-Roll Nanotube Heterojunctions. Small 2019, 15, 1900902  DOI: 10.1002/smll.201900902
    5. 5
      Rojas-Lopez, R. R.; Brant, J. C.; Ramos, M. S. O.; Castro, T. H. L. G.; Guimarães, M. H. D.; Neves, B. R. A.; Guimarães, P. S. S. Photoluminescence and Charge Transfer in the Prototypical 2D/3D Semiconductor Heterostructure MoS2/GaAs. Appl. Phys. Lett. 2021, 119, 233101,  DOI: 10.1063/5.0068548
    6. 6
      Zhai, X.-P.; Ma, B.; Wang, Q.; Zhang, H.-L. 2D Materials Towards Ultrafast Photonic Applications. Phys. Chem. Chem. Phys. 2020, 22, 2214022156,  DOI: 10.1039/D0CP02841J
    7. 7
      An, J.; Zhao, X.; Zhang, Y.; Liu, M.; Yuan, J.; Sun, X.; Zhang, Z.; Wang, B.; Li, S.; Li, D. Perspectives of 2D Materials for Optoelectronic Integration. Adv. Funct. Mater. 2022, 32, 2110119  DOI: 10.1002/adfm.202110119
    8. 8
      Liang, G.; Yu, X.; Hu, X.; Qiang, B.; Wang, C.; Wang, Q. J. Mid-Infrared Photonics and Optoelectronics in 2D Materials. Mater. Today 2021, 51, 294316,  DOI: 10.1016/j.mattod.2021.09.021
    9. 9
      Buscema, M.; Island, J. O.; Groenendijk, D. J.; Blanter, S. I.; Steele, G. A.; Van Der Zant, H. S.; Castellanos-Gomez, A. Photocurrent Generation with Two-Dimensional van der Waals Semiconductors. Chem. Soc. Rev. 2015, 44, 36913718,  DOI: 10.1039/C5CS00106D
    10. 10
      Mak, K. F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T. F. Atomically Thin MoS2: A New Direct-Gap Semiconductor. Phys. Rev. Lett. 2010, 105, 136805  DOI: 10.1103/PhysRevLett.105.136805
    11. 11
      Manzeli, S.; Ovchinnikov, D.; Pasquier, D.; Yazyev, O. V.; Kis, A. 2D Transition Metal Dichalcogenides. Nat. Rev. Mater. 2017, 2, 17033,  DOI: 10.1038/natrevmats.2017.33
    12. 12
      Aftab, S.; Shehzad, M. A.; Salman Ajmal, H. M.; Kabir, F.; Iqbal, M. Z.; Al-Kahtani, A. A. Bulk Photovoltaic Effect in Two-Dimensional Distorted MoTe2. ACS Nano 2023, 17, 1788417896,  DOI: 10.1021/acsnano.3c03593
    13. 13
      Hu, X.; Zhang, F.; Hu, Z.; He, P.; Tao, L.; Zheng, Z.; Zhao, Y.; Yang, Y.; He, J. Preparation of 1T’- and 2H–MoTe2 Films and Investigation of their Photoelectric Properties and Ultrafast Photocarrier Dynamics. Opt. Mater. 2023, 136, 113467  DOI: 10.1016/j.optmat.2023.113467
    14. 14
      Dave, M.; Vaidya, R.; Patel, S. G.; Jani, A. R. High Pressure Effect on MoS2 and MoSe2 Single Crystals Grown by CVT Method. Bulletin of Materials Science 2004, 27, 213216,  DOI: 10.1007/BF02708507
    15. 15
      Song, S.; Keum, D. H.; Cho, S.; Perello, D.; Kim, Y.; Lee, Y. H. Room Temperature Semiconductor-Metal Transition of MoTe2 Thin Films Engineered by Strain. Nano Lett. 2016, 16, 188193,  DOI: 10.1021/acs.nanolett.5b03481
    16. 16
      Lin, Y.-C.; Dumcenco, D. O.; Huang, Y.-S.; Suenaga, K. Atomic Mechanism of the Semiconducting-to-Metallic phase Transition in Single-Layered MoS2. Nat. Nanotechnol. 2014, 9, 391396,  DOI: 10.1038/nnano.2014.64
    17. 17
      Shang, B.; Cui, X.; Jiao, L.; Qi, K.; Wang, Y.; Fan, J.; Yue, Y.; Wang, H.; Bao, Q.; Fan, X.; Wei, S.; Song, W.; Cheng, Z.; Guo, S.; Zheng, W. Lattice-Mismatch-Induced Ultrastable 1T-Phase MoS2-Pd/Au for Plasmon-Enhanced Hydrogen Evolution. Nano Lett. 2019, 19, 27582764,  DOI: 10.1021/acs.nanolett.8b04104
    18. 18
      Wang, Y.; Xiao, J.; Zhu, H.; Li, Y.; Alsaid, Y.; Fong, K. Y.; Zhou, Y.; Wang, S.; Shi, W.; Wang, Y.; Zettl, A.; Reed, E. J.; Zhang, X. Structural Phase Transition in Monolayer MoTe2 Driven by Electrostatic Doping. Nature 2017, 550, 487491,  DOI: 10.1038/nature24043
    19. 19
      Kappera, R.; Voiry, D.; Yalcin, S. E.; Branch, B.; Gupta, G.; Mohite, A. D.; Chhowalla, M. Phase-Engineered Low-Resistance Contacts for Ultrathin MoS2 transistors. Nat. Mater. 2014, 13, 11281134,  DOI: 10.1038/nmat4080
    20. 20
      Ma, Y.; Liu, B.; Zhang, A.; Chen, L.; Fathi, M.; Shen, C.; Abbas, A. N.; Ge, M.; Mecklenburg, M.; Zhou, C. Reversible Semiconducting-to-Metallic Phase Transition in Chemical Vapor Deposition Grown Monolayer WSe2 and Applications for Devices. ACS Nano 2015, 9, 73837391,  DOI: 10.1021/acsnano.5b02399
    21. 21
      Cho, S.; Kim, S.; Kim, J. H.; Zhao, J.; Seok, J.; Keum, D. H.; Baik, J.; Choe, D.-H.; Chang, K. J.; Suenaga, K.; Kim, S. W.; Lee, Y. H.; Yang, H. Phase Patterning for Ohmic Homojunction Contact in MoTe2. Science 2015, 349, 625628,  DOI: 10.1126/science.aab3175
    22. 22
      Duerloo, K.-A. N.; Li, Y.; Reed, E. J. Structural Phase Transitions in Two-Dimensional Mo- and W-Dichalcogenide Monolayers. Nat. Commun. 2014, 5, 4214,  DOI: 10.1038/ncomms5214
    23. 23
      Tan, Y.; Luo, F.; Zhu, M.; Xu, X.; Ye, Y.; Li, B.; Wang, G.; Luo, W.; Zheng, X.; Wu, N.; Yu, Y.; Qin, S.; Zhang, X.-A. Controllable 2H-to-1T’ Phase Transition in Few-Layer MoTe2. Nanoscale 2018, 10, 1996419971,  DOI: 10.1039/C8NR06115G
    24. 24
      Kang, S.; Won, D.; Yang, H.; Lin, C.-H.; Ku, C.-S.; Chiang, C.-Y.; Kim, S.; Cho, S. Phase-controllable laser thinning in MoTe2. Appl. Surf. Sci. 2021, 563, 150282  DOI: 10.1016/j.apsusc.2021.150282
    25. 25
      Zhang, X. Low Contact Barrier in 2H/1T’ MoTe2 In-Plane Heterostructure Synthesized by Chemical Vapor Deposition. ACS Appl. Mater. Interfaces 2019, 11, 1277712785,  DOI: 10.1021/acsami.9b00306
    26. 26
      Bae, G. Y.; Kim, J.; Kim, J.; Lee, S.; Lee, E. MoTe2 Field-Effect Transistors with Low Contact Resistance through Phase Tuning by Laser Irradiation. Nanomaterials 2021, 11, 2805,  DOI: 10.3390/nano11112805
    27. 27
      Wang, Y.; Chhowalla, M. Making Clean Electrical Contacts on 2D Transition Metal Dichalcogenides. Nature Reviews Physics 2022, 4, 101112,  DOI: 10.1038/s42254-021-00389-0
    28. 28
      Yamaguchi, H.; Blancon, J. C.; Kappera, R.; Lei, S.; Najmaei, S.; Mangum, B. D.; Gupta, G.; Ajayan, P. M.; Lou, J.; Chhowalla, M.; Crochet, J. J.; Mohite, A. D. Spatially Resolved Photoexcited Charge-Carrier Dynamics in Phase-Engineered Monolayer MoS2. ACS Nano 2015, 9, 840849,  DOI: 10.1021/nn506469v
    29. 29
      Lin, D.-Y.; Hsu, H.-P.; Liu, G.-H.; Dai, T.-Z.; Shih, Y.-T. Enhanced Photoresponsivity of 2H-MoTe2 by Inserting 1T-MoTe2 Interlayer Contact for Photodetector Applications. Crystals 2021, 11, 964,  DOI: 10.3390/cryst11080964
    30. 30
      Ding, Y.; Qi, R.; Wang, C.; Wu, Q.; Zhang, H.; Zhang, X.; Lin, L.; Cai, Z.; Xiao, S.; Gu, X.; Nan, H. Broad-Band Photodetector Based on a Lateral MoTe2 1T-2H-1T Homojunction. J. Phys. Chem. C 2023, 127, 2007220081,  DOI: 10.1021/acs.jpcc.3c05592
    31. 31
      Xu, X.; Gabor, N. M.; Alden, J. S.; Van Der Zande, A. M.; McEuen, P. L. Photo-thermoelectric Effect at a Graphene Interface Junction. Nano Lett. 2010, 10, 562566,  DOI: 10.1021/nl903451y
    32. 32
      Zhang, Y.; Li, H.; Wang, L.; Wang, H.; Xie, X.; Zhang, S. L.; Liu, R.; Qiu, Z. J. Photothermoelectric and Photovoltaic Effects Both Present in MoS2. Sci. Rep. 2015, 5, 7938,  DOI: 10.1038/srep07938
    33. 33
      Huo, N.; Konstantatos, G. Recent Progress and Future Prospects of 2D-Based Photodetectors. Adv. Mater. 2018, 30, 1801164  DOI: 10.1002/adma.201801164
    34. 34
      HQgraphene., MoTe2 (2H Molybdenum Ditelluride). https://www.hqgraphene.com/MoTe2.php (accessed May 23 2023).
    35. 35
      Buscema, M.; Barkelid, M.; Zwiller, V.; van der Zant, H. S. J.; Steele, G. A.; Castellanos-Gomez, A. Large and Tunable Photothermoelectric Effect in Single-Layer MoS2. Nano Lett. 2013, 13, 358363,  DOI: 10.1021/nl303321g
    36. 36
      Ruppert, C.; Aslan, O. B.; Heinz, T. F. Optical Properties and Band Gap of Single- and Few-Layer MoTe2 Crystals. Nano Lett. 2014, 14, 62316236,  DOI: 10.1021/nl502557g
    37. 37
      Lezama, I. G.; Arora, A.; Ubaldini, A.; Barreteau, C.; Giannini, E.; Potemski, M.; Morpurgo, A. F. Indirect-to-Direct Band Gap Crossover in Few-Layer MoTe2. Nano Lett. 2015, 15, 23362342,  DOI: 10.1021/nl5045007
    38. 38
      Yadav, P.; Dewan, S.; Mishra, R.; Das, S. Review of Recent Progress, Challenges, and Prospects of 2D Materials-Based Short Wavelength Infrared Photodetectors. J. Phys. D: Appl. Phys. 2022, 55, 313001,  DOI: 10.1088/1361-6463/ac6635
    39. 39
      Zhang, K.; Fang, X.; Wang, Y.; Wan, Y.; Song, Q.; Zhai, W.; Li, Y.; Ran, G.; Ye, Y.; Dai, L. Ultrasensitive Near-Infrared Photodetectors Based on a Graphene-MoTe2-Graphene Vertical van der Waals Heterostructure. ACS Appl. Mater. Interfaces 2017, 9, 53925398,  DOI: 10.1021/acsami.6b14483
    40. 40
      Xiao, H.; Lin, L.; Zhu, J.; Guo, J.; Ke, Y.; Mao, L.; Gong, T.; Cheng, H.; Huang, W.; Zhang, X. Highly Sensitive and Broadband Photodetectors Based on WSe2/MoS2 Heterostructures with van der Waals Contact Electrodes. Appl. Phys. Lett. 2022, 121, 023504  DOI: 10.1063/5.0100191
    41. 41
      Huang, H. Highly Sensitive Visible to Infrared MoTe2 Photodetectors Enhanced by the Photogating Effect. Nanotechnology 2016, 27, 445201  DOI: 10.1088/0957-4484/27/44/445201
    42. 42
      Shen, D. High-Performance Mid-IR to Deep-UV van der Waals Photodetectors Capable of Local Spectroscopy at Room Temperature. Nano Lett. 2022, 22, 34253432,  DOI: 10.1021/acs.nanolett.2c00741
    43. 43
      Yu, W.; Li, S.; Zhang, Y.; Ma, W.; Sun, T.; Yuan, J.; Fu, K.; Bao, Q. Near-Infrared Photodetectors Based on MoTe2/Graphene Heterostructure with High Responsivity and Flexibility. Small 2017, 13, 1700268  DOI: 10.1002/smll.201700268
    44. 44
      Yin, L.; Zhan, X.; Xu, K.; Wang, F.; Wang, Z.; Huang, Y.; Wang, Q.; Jiang, C.; He, J. Ultrahigh Sensitive MoTe2 Phototransistors Driven by Carrier Tunneling. Appl. Phys. Lett. 2016, 108, 043503  DOI: 10.1063/1.4941001
    45. 45
      Xie, Y.; Wu, E.; Zhang, J.; Hu, X.; Zhang, D.; Liu, J. Gate-Tunable Photodetection/Voltaic Device Based on BP/MoTe2 Heterostructure. ACS Appl. Mater. Interfaces 2019, 11, 1421514221,  DOI: 10.1021/acsami.8b21315
    46. 46
      Lee, H. S.; Min, S.-W.; Chang, Y.-G.; Park, M. K.; Nam, T.; Kim, H.; Kim, J. H.; Ryu, S.; Im, S. MoS2 Nanosheet Phototransistors with Thickness-Modulated Optical Energy Gap. Nano Lett. 2012, 12, 36953700,  DOI: 10.1021/nl301485q
    47. 47
      Lopez-Sanchez, O.; Lembke, D.; Kayci, M.; Radenovic, A.; Kis, A. Ultrasensitive Photodetectors Based on Monolayer MoS2. Nat. Nanotechnol. 2013, 8, 497501,  DOI: 10.1038/nnano.2013.100
    48. 48
      Buscema, M.; Groenendijk, D. J.; Blanter, S. I.; Steele, G. A.; Van Der Zant, H. S. J.; Castellanos-Gomez, A. Fast and Broadband Photoresponse of Few-Layer Black Phosphorus Field-Effect Transistors. Nano Lett. 2014, 14, 33473352,  DOI: 10.1021/nl5008085
    49. 49
      Zhang, X.; Liu, B.; Yang, W.; Jia, W.; Li, J.; Jiang, C.; Jiang, X. 3D-Branched Hierarchical 3C-SiC/ZnO Heterostructures for High-Performance photodetectors. Nanoscale 2016, 8, 1757317580,  DOI: 10.1039/C6NR06236A
    50. 50
      Kind, H.; Yan, H.; Messer, B.; Law, M.; Yang, P. Nanowire Ultraviolet Photodetectors and Optical Switches. Adv. Mater. 2002, 14, 158160,  DOI: 10.1002/1521-4095(20020116)14:2<158::AID-ADMA158>3.0.CO;2-W
    51. 51
      Liu, F.; Shimotani, H.; Shang, H.; Kanagasekaran, T.; Zólyomi, V.; Drummond, N.; Fal’ko, V. I.; Tanigaki, K. High-Sensitivity Photodetectors Based on Multilayer GaTe Flakes. ACS Nano 2014, 8, 752760,  DOI: 10.1021/nn4054039
    52. 52
      Hidding, J.; Cordero-Silis, C. A.; Vaquero, D.; Panagiotis Rompotis, K.; Quereda, J.; Guimarães, M. H. D. Locally Phase-Engineered MoTe2 for Near-Infrared Photodetectors . 2024, 2406.01376, arXiv, https://arxiv.org/abs/2406.01376 (accessed Jun 03, 2024).
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsphotonics.4c00896.

    • Details over the photothermoelectric temperature gradient calculation; laser spot size calculation and data for 700 and 1064 nm for the reported measurements; equation and additional details for charge carrier mobility and further details of another phase-engineered device, including responsivity and transfer curves; drain-source and gate voltage dependence of the photocurrent; normalized Raman spectra and spectral weight contribution of the 1T′ phase characteristic peaks; laser irradiation power; and flake thickness dependence for phase change (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.