Scalable van der Waals Heterojunctions for High ... - ACS Publications

Sep 25, 2017 - Scalable van der Waals Heterojunctions for High-Performance. Photodetectors. Chao-Hui Yeh,. †. Zheng-Yong Liang,. †. Yung-Chang Lin...
0 downloads 0 Views 5MB Size
Research Article www.acsami.org

Cite This: ACS Appl. Mater. Interfaces 2017, 9, 36181-36188

Scalable van der Waals Heterojunctions for High-Performance Photodetectors Chao-Hui Yeh,† Zheng-Yong Liang,† Yung-Chang Lin,‡ Tien-Lin Wu,† Ta Fan,† Yu-Cheng Chu,† Chun-Hao Ma,†,§ Yu-Chen Liu,† Ying-Hao Chu,§ Kazutomo Suenaga,‡ and Po-Wen Chiu*,†,∥ †

Department of Electrical Engineering, National Tsing Hua University, Hsinchu 30013, Taiwan National Institute of Advanced Industrial Science and Technology (AIST), Tsukuba 305-8565, Japan § Department of Materials Science and Engineering, National Chiao Tung University, Hsinchu 30010, Taiwan ∥ Institute of Atomic and Molecular Sciences, Academia Sinica, Taipei 10617, Taiwan ‡

S Supporting Information *

ABSTRACT: Atomically thin two-dimensional (2D) materials have attracted increasing attention for optoelectronic applications in view of their compact, ultrathin, flexible, and superior photosensing characteristics. Yet, scalable growth of 2D heterostructures and the fabrication of integrable optoelectronic devices remain unaddressed. Here, we show a scalable formation of 2D stacks and the fabrication of phototransistor arrays, with each photosensing element made of a graphene−WS2 vertical heterojunction and individually addressable by a local top gate. The constituent layers in the heterojunction are grown using chemical vapor deposition in combination with sulfurization, providing a clean junction interface and processing scalability. The aluminum top gate possesses a self-limiting oxide around the gate structure, allowing for a self-aligned deposition of drain/source contacts to reduce the access (ungated) channel regions and to boost the device performance. The generated photocurrent, inherently restricted by the limited optical absorption cross section of 2D materials, can be enhanced by 2 orders of magnitude by top gating. The resulting photoresponsivity can reach 4.0 A/W under an illumination power density of 0.5 mW/cm2, and the dark current can be minimized to few picoamperes, yielding a low noise-equivalent power of 2.5 × 10−16 W/Hz1/2. Tailoring 2D heterostacks as well as the device architecture moves the applications of 2D-based optoelectronic devices one big step forward. KEYWORDS: WS2, graphene, two-dimensional materials, heterostructure, photoresponsivity



INTRODUCTION

To date, vertical 2D heterostructures are constructed via a sequential exfoliation and transfer. This technique gains an advantage over its ease of handling and provides 2D flakes with high quality but is limited in its scalability and control over the thickness. Routes that allow up-scaling and produce 2D heterostacks with desired functionalities are much needed. Of the various routes available for producing 2D materials, chemical vapor deposition (CVD) stands out as the most useful for electronic applications because of its scalability, reliability, and integrability with silicon technology.2,13 Postsulfurization or -selenization of transition-metal oxides, on the other hand, has the advantage of site-controllable manufacturing.14,15 Recently, in-plane and vertical MoS2−WS2 heterojunctions have been grown using CVD and postsulfurization, respectively. However, nontransfer stacking of graphene with different transition-metal dichalcogenides (TMDs) remains unaddressed. Here, we bridge this gap and show scalable fabrication of graphene−WS2 (Gr−WS2) heterojunc-

Two-dimensional (2D) layered materials have been a subject of intensive research for more than a decade because of their wide range of fascinating properties. Of particular interest are their applications in nanoelectronics1−4 and optoelectronics5−7 considering their unique features, which are absent in conventional semiconductor analogues such as their excellent mechanical flexibility and optical transparency. Versatile van der Waals stacks of 2D materials have delivered enormous opportunities to allow electronic functionalities by design.8 Examples are recently reported vertical field-effect transistors (FETs),9 barristors,10 and photovoltaic cells.11,12 The prototypical vertical FETs made of transferred graphene−MoS2 heterojunctions showed a remarkable on-current density, 2−5 orders of magnitude larger than that of conventional FETs.9 A gate-controllable photocurrent with a responsivity of 0.22 A/W and an external quantum efficiency of 55% has also been realized using the same structure, rendering the 2D layered heterojunctions an exciting material system for electronics and optoelectronics with novel functionalities or better performance. © 2017 American Chemical Society

Received: July 24, 2017 Accepted: September 25, 2017 Published: September 25, 2017 36181

DOI: 10.1021/acsami.7b10892 ACS Appl. Mater. Interfaces 2017, 9, 36181−36188

Research Article

ACS Applied Materials & Interfaces

Figure 1. Raman and STEM spectral characterizations of synthesized TMDs. (a−d) Raman spectra and characteristic maps acquired for four different TMDs: WS2, MoS2, WSe2, and MoSe2. Scale bars: 50 μm. (e−h) STEM images taken from the four materials shown in parts a−d. Scale bars: 2 nm. Inset: Corresponding SAED pattern.

tion arrays for high-performance photodetectors. The graphene layers are grown directly on transparent substrates by plasmaenhanced CVD, followed by patterned stacking of WS2 on top using postsulfurization, with the thickness controlled by the predeposited metal oxide WOx. Through modulation by an aluminum top gate, the generated photocurrent can be enhanced by more than 2 orders of magnitude while shining monochromatic light. The gated Gr−WS2 junctions can deliver a photoresponsibility as high as 4.0 A/W, 2−4 orders of magnitude larger than that of devices solely comprised of graphene or WS2.16 More importantly, the dark current of the Gr−WS2 junction, which determines the signal-to-noise ratio and is difficult to eliminate by signal processing through an external circuit, can be suppressed to a few picoamperes, giving a low noise-equivalent power of 2.5 × 10−16 W/Hz1/2. Moreover, we also transferred the Gr−WS2 junction arrays to a flexible substrate and fabricated top-gated Gr−WS2 phototransistors. Photodetection under bending was explored and discussed.



MO3 − x +

7−x 3−x X* → MX 2 + XO2 2 2

where H* and X* are hydrogen and chalcogen radicals, respectively. Here, we show the reactions with M = Mo and W and X = S and Se. It has been known that it is difficult to directly react pure metal M and stoichiometric oxide MO3 with chalcogens.17 Here, hydrogen radicals serve as reducing agents and are used to facilitate the conversion of MO3 to an intermediate MO3−x, which can react much more effectively with chalcogens. It should be noted that a higher level of hydrogen content will cause an excessive reduction of MO3−x to M, making the chalcogenization reaction incomplete. As a result, the amount of hydrogen used in the reaction determines the layer quality, the number of nucleation sites, and the grain size accordingly. We characterize the resulting graphene and TMD films using Raman spectroscopy, X-ray photoelectron spectroscopy (XPS), atomic force microscopy (AFM), and scanning transmission electron microscopy (STEM). Parts a−d of Figure 1 show the Raman properties of WS2, MoS2, WSe2, and MoSe2 under a laser excitation wavelength of 532 nm. For WS2 and MoS2, the most prominent Raman peaks, corresponding to in-plane (E12g) and out-of-plane (A1g) optical modes, can be explicitly identified. Note that the frequency difference Δω between the E12g and A1g modes depends sensitively upon the number of layers due to the stronger dielectric screening of long-range Coulomb interactions between the effective charges in thicker samples.18,19 The measured Δω values are 63.5 and 23.7 cm−1 for WS2 and MoS2, respectively. These frequency differences correspond to a thickness of 2−3 layers, in good agreement

RESULTS AND DISCUSSION

A horizontal 3-in. tube furnace equipped with a home-designed remote plasma system and a multicoiled radio-frequency generator was used to grow graphene on sapphire and to chalcogenize transition-metal oxides deposited by thermal evaporation. The chalcogenization takes place with the following two sequential reactions: MO3 + 2x H* → MO3 − x + x H 2O 36182

DOI: 10.1021/acsami.7b10892 ACS Appl. Mater. Interfaces 2017, 9, 36181−36188

Research Article

ACS Applied Materials & Interfaces

Figure 2. Cartoon illustration of the fabrication process of a phototransistor based on the vertical Gr−TMD junction. (a) Direct CVD growth of graphene and strip patterning. (b and c) Patterned deposition of the transition-metal oxide and chalcogenization. (d) Formation of the Al/AlOx topgated electrode and self-alignment metals.

Figure 3. Raman and TEM spectral characterizations of the Gr−WS2 vertical junction. (a and b) High-resolution cross-sectional TEM images of Gr−WS2 junction. (c) Raman spectra of the multilayer WS2 (taken on the red spot of the optical image in the inset) and crossed junction area (taken on the green spot of the optical image in the inset). Scale bar: 100 μm. (d) Spatial Raman mapping of the G band (1598 cm−1) and E12g mode (351 cm−1) at the Gr−WS2 junction area. Scale bar: 40 μm.

Information). The selected-area electron diffraction (SAED) patterns (insets to Figure 1e−h), which were taken with a 1μm-diameter aperture, show diffraction spots arranged in multiple hexagons, indicating polycrystalline structures with grain sizes smaller than 1 μm. Other characterizations of the TMD films, including the chemical composition using XPS, topology using AFM, and layer-dependent photoluminescence, can be found in Figures S2−S4 in the Supporting Information. As learned from bulk semiconductors, the performance of electronic devices is intimately related to the functional interface or heterostructures within the devices. Exquisite control over the composition and perfection of interfaces is required for the successful fabrication of high-performance planar devices. As such, we use the nontransfer method to make 2D heterojunctions, with a focus on the Gr−WS2 stacks in the current study. Figure 2 depicts the fabrication process flow of individually addressable phototransistor arrays based on a Gr−

with the number of layers measured using AFM and STEM (Figures S1 and S2 in the Supporting Information). For WSe2 and MoSe2, on the other hand, they are featured in a single prominent vibrational mode, appearing at 250.5 and 241.7 cm−1 for WSe2 E12g and MoSe2 A1g, respectively. The B2g mode, which is found at 307 cm−1 for WSe2, indicates a multilayer structure.20 What are also shown are the spatially resolved Raman maps of the most intense mode for WS2, MoS2, WSe2, and MoSe2. The areas of Raman responses match well with the defined structures and are uniform in intensity at a fixed peak position. Parts e−h of Figure 1 show the atomic structure of the TMD films. The contrast of the annular dark-field images varies with the power of the atomic number ∼ Z1.7, sensitively allowing us to differentiate the transition metal with the chalcogen atom and to assess the crystalline quality according to the defect density of chalcogens (Figure S2 in the Supporting 36183

DOI: 10.1021/acsami.7b10892 ACS Appl. Mater. Interfaces 2017, 9, 36181−36188

Research Article

ACS Applied Materials & Interfaces

Figure 4. Device geometry, energy band diagram, and electrical properties of Gr−WS2 photodetectors on a rigid substrate. (a) Optical image of Gr− WS2 junction photodetectors on sapphire, with a close-up of the junction area at the lower panel and a schematic of light illumination through the device at the right panel. (b and c) Output characteristics of the device. The inset shows the direction of current flow through the Gr−WS2 junction. (d) Gating response (Ids−Vgs) of the Gr−WS2 junction photodetector in dark and illuminated states, acquired at a bias voltage of Vds = −1 V. The illumination power is 30 mW/cm2. Inset: Time-resolved photoresponse of the device, recorded at fixed biases of Vds = −1 V and Vgs = 1 V. The modulation frequency of the chopper is 0.1 Hz. (e) Band diagram of the top-gated Gr−WS2 junction photodetector. An electrostatic gating changes the band slope of WS2, determining the electron density in the n-type WS2 as illuminated with light. (f) Photoresponsivity and photocurrent as a function of the incident laser power.

WS2 junction. In brief, a few-layer graphene film was first grown on a transparent sapphire wafer using a remote-catalyzed CVD technique,21 followed by patterning into strips with 250 μm width and 1000 μm length using reactive oxygen plasma etching (Figure 2a). The graphene surface was cleaned prior to the deposition of metal oxide using the method described in our previous study.22 Tungsten oxide strips deposited perpendicular to the graphene strips was then defined by a hard mask (Figure S5 in the Supporting Information). Sulfurization was carried out in a plasma-enhanced CVD system at 700 °C for 20 min (Figure 2b,c). The Gr−WS2 junction is formed at the cross of the two strips, each side of which is contacted with Cr (1 nm)/Au (50 nm) as an electrical contact. In the last step, an Al/AlOx top gate was made on top of the junction, followed by a self-aligned deposition of drain/ source contacts (Figures 2d and S6 in the Supporting Information).23 Parts a and b of Figure 3 show the typical high-resolution TEM (HRTEM) images of the Gr−WS2 junction in the cross section. To ensure continuity of the film, the average thickness of graphene and WS2 was grown to be 2−3 layers. A proper increase of the thickness gains the advantage of high light absorption, generating a higher photocurrent. However, growing multilayer graphene using plasma-enhanced CVD on dielectric substrates is not straightforward. It usually leads to a vertical outgrowth of graphene nanowalls or nanopedals because of the defective growth mechanism and fast deposition rate.24 Likewise, sulfurization of a metal oxide often occurs on the top surface and leads to an incomplete reaction in the bottom. Parts a and b of Figure 3 show thin and thick Gr−WS2 junctions, respectively. It can be seen that each constituent layer is well spaced: ∼3.4 Å for graphene layers and ∼6.3 Å for

WS2. More importantly, the interface between graphene and WS2 is shaped and clean. This characteristic is fundamentally critical in the performance of the optoelectronic devices thus constructed. Parts c and d of Figure 3 show the Raman spectra and spatial mapping of the Gr−WS2 junction. The characteristic Raman peaks of graphene and WS2 in the noncross region are also taken for comparison. The G peak of graphene appears at 1598 cm−1 in the noncross region, with a narrow full width at halfmaximum of 28 cm−1, indicative of decent crystallinity. For WS2, the characteristic E12g and A1g modes peak at 353.5 and 417 cm−1 in the noncross region, respectively. No remarkable peak shift is found for both graphene and WS2 in the crossed junction region, in accordance with other reported vertical TMD−TMD or Gr−TMD heterojunctions.25,26 Only a slight A1g upshift of ∼1 cm−1 is noticed, presumably due to a weak charge transfer or strain induced by the underlying graphene layers.27−29 A similar A1g upshift is also found in the Gr−MoS2 junctions (Figure S7 in the Supporting Information). The spatially resolved Raman maps depict the uniformity of the graphene and WS2 strips as well as the Gr−WS2 junction, as shown in Figure 3d. WS2 has received considerable attention for optoelectronic applications because of its semiconducting properties with a direct band gap in the visible spectrum,30 large exciton binding energy, large absorption coefficient,31 and high external quantum efficiency.32 Figure 4a shows a perspective view schematic and an optical microscopy top-view image of our photodetector comprised of perpendicular graphene and WS2 strips. The crossed junction area is 250 μm (Gr) × 80 μm (WS2) in dimension, with WS2 lying above graphene and in intimate contact with an Al/AlOx top gate. The thicknesses of 36184

DOI: 10.1021/acsami.7b10892 ACS Appl. Mater. Interfaces 2017, 9, 36181−36188

Research Article

ACS Applied Materials & Interfaces

layer heterojunctions. At Vgs > 0, the band bends downward at the WS2−gate interface, driving the photoexcited holes to the graphene side and the electrons to WS2. Under a forward bias, the current increases exponentially because of the lowering of both contact and Gr−WS2 interjunction barriers. On the other hand, a negative Vgs causes an upward bending of the WS2 band, depleting both graphene and WS2 as the photocarriers are generated. In this case, the photocurrent is suppressed. The photoresponsivity, external quantum efficiency, and photogain are important figures of merit in evaluating the performance of photodetection. The photoresponsivity is a measure of a device’s electrical response to incident light and is defined as R = Iph/P, where Iph and P stand for the measured photocurrent and illumination power of the laser, respectively. Figure 4f depicts the photocurrent and photoresponsivity as a function of the incident power at biases of Vds = −1 V and Vgs = 1 V. Under this condition, the transistor is tuned to the ON state and the photocurrent monotonically increases with an increase of the incident power. On the contrary, the photoresponsivity monotonically decreases with an increase of the incident power from 4 A/W at P = 0.15 mW/cm2 to 0.3 A/W at P = 60 mW/cm2. This behavior is consistent with other photodetectors made of solely TMDs and attributed to a shortening of the exciton recommendation time caused by the increasing number of electron−hole pairs and the influence of trap states between the gate oxide and semiconductor.31 The external quantum efficiency and photogain of our Gr−WS2 photodetectors are given in Figure S8 in the Supporting Information. In addition to the operation on rigid substrates, we also transferred the Gr−WS2 junctions onto poly(ethylene terephthalate) (PET). The same top-gated structure was fabricated after transfer, forming bendable Gr−WS2 phototransistor arrays. Parts a−c of Figure 5 show the optical images of the Gr−WS2 junctions on PET, with an individually addressable top gate for each junction. In general, the transistor behaviors of the devices on PET are qualitatively similar to those on rigid

the graphene and WS2 strips were determined by TEM to be ∼3 and ∼4 nm, respectively. To measure the photoconductivity, the device was illuminated with a focused laser beam from the backside of the transparent substrate. Figure 4b shows the Ids−Vds curves for the WS2 strip under dark conditions at different gate voltages. The current increases with increasing positive Vgs, indicating the n-type nature of WS2. The low current density is attributed partly to an appreciable Schottky barrier formed at the metal−WS2 contacts and partly to a parasitic resistance in the access (ungated) regions. For charge transport going through the Gr−WS2 junction, the Ids− Vds curves become asymmetric (Figure 4c). The rectification stems from the two asymmetric Schottky barriers at the Au− Gr−WS2 and Au−WS2 junctions and allows current to pass through the Gr−WS2 junction only when WS2 is negatively biased. Here, graphene plays a critical role in Fermi-level unpinning, which yields a low barrier at the Au−Gr−WS2 junction.33 Because the device is illuminated with a focused laser beam (wavelength = 514 nm and power = 50 μW), the current density going across the Gr−WS2 junction is enhanced by a factor of 150 in comparison to that going along the WS2 strip at Vgs = 1 V. This remarkable enhancement is also seen in the dependence of the photocurrent generation upon the application of a top-gated voltage, shown in Figure 4d. In the dark state, the Gr−WS2 junction functions as an n-type vertical FET; i.e., the channel only conducts at Vgs > 0. When we illuminated the device biased at Vds = −1 V, the channel current remains blocked at Vgs < 0 but remarkably increased at Vgs > 0. The room-temperature ON/OFF ratio of the Gr−WS2 vertical FET reaches 104, higher than that of the similar device made of Gr−MoS2.30 This result indicates that massive photoexcited carriers can be generated in the atomic thin WS2 layers and effectively modulated by an external electric field. Figure 4d inset shows the time-resolved photoresponse of the device. Under repeated ON/OFF cycles of light illumination, the timeresolved photoresponse of the Gr−WS2 junction at fixed biases of Vds = −1 V and Vgs = 1 V yields alternating high and low impedance states, with an ON/OFF ratio (defined as Iph/Idark) of ∼50, rendering the device a high-quality photosensitive switch. To determine the sensitivity of our phototransistor, we measure the noise in the dark current. Given that shot noise is the dominant source of the total noise in our device (higher than the thermal noise and 1/f noise), the dark-current shot noise is written as 2eIdark ∼ 1 × 10−15 A/Hz1/2 for Idark = 3 pA, where e stands for electronic charge. The noise-equivalent power, which is a measure of the sensitivity of a photodetector and equal to the shot noise divided by the responsivity, is obtained as ∼2.5 × 10−16 W/Hz1/2, 2 orders of magnitude smaller than that of a commercial silicon avalanche photodiode based on p−n junctions.34 For our Gr−WS2 junction, it increases at Vgs > 0. The observed gate modulation of the photocurrent in our devices can be understood from the junction band diagram (Figure 4e). Because the vertical channel length of WS2 (3−5 nm) is much shorter than the depletion length of the Gr−WS2 junction, the band bending at the Gr−WS2 interface dominates the band slope of the entire WS2 layers. As such, the application of an external electric field provides an effective means of modulating this monotonic band slope, capable of controlling the separation and transport of photocarriers in vertically stacked

Figure 5. Optical image and photoresponse of Gr−WS2 photodetectors on a flexible PET substrate. (a−c) Optical images of Gr− WS2 photodetectors on PET. Scale bar: 50 μm. (d) Time-resolved photoresponse of the device recorded at fixed biases of Vds = −1 V and Vgs = 1 V. Photocurrents at illumination powers of 1 and 30 mW/cm2 were acquired. (e) Photocurrent as a function of the bending cycle at an illumination power of 30 mW/cm2. 36185

DOI: 10.1021/acsami.7b10892 ACS Appl. Mater. Interfaces 2017, 9, 36181−36188

Research Article

ACS Applied Materials & Interfaces

∼20 min. The samples were rapidly cooled to room temperature as the reaction was completed. Device Fabrication. Drain/source metal contacts, Cr (1 nm)/Au (50 nm), were first deposited on each side of the graphene and WS2 strips. Then, standard photolithography was applied to define a topgate window (500 × 250 μm2) covering the crossed Gr−WS2 junction area. A thin aluminum film with a thickness of 7 nm was deposited, followed by an oxidation process in a sealed chamber filled with highpurity oxygen (>99.999%) at a pressure of 3 kgf/cm2 overnight. Then, the aluminum top gate (100 nm) was deposited and staked on the surface oxidation layer (AlOx). Afterward, self-aligned deposition of an 8-nm-thick gold film was applied to reduce the access resistance of the device. Optoelectronic Measurements. The Gr−WS2 junction photodetectors were made on sapphire, which was glued on a glass substrate with a hole of 5 mmdiameter for laser illumination. A solid-state laser was used to provide a light source with a wavelength of 532 nm. The output power is controllable in the range from 100 μW to 80 mW and coupled with a chopper to define the switching ON/OFF intervals. The output power was calibrated with a hand-held continuous-wave laser power meter. For I−V measurements, the drain and gate biases were provided through Keithley 2400 sourcemeters, and the generated photocurrent was detected via a Keithley 2000 multimeter connected to a current preamplifier.

substrates, except that the photocurrent on PET is lowered by a factor of ∼7. The degraded performance in the photoresponse might be attributed to the rough surface and the trapped water layer between the substrate and graphene. Figure 5d shows the time-resolved photocurrent for Gr−WS2 junctions on PET under alternating ON/OFF light illuminations at Vds = −1 V in an ambient environment. A rise and decay response times of 800 ms and 3 s are measured, respectively. No clear persistent photocurrent is seen in the repeated ON/OFF switch of light illumination. Nevertheless, the Gr−WS2 photodetectors on PET take longer to saturate and decay in comparison to the same structure lying on sapphire. Figure 5e shows the ON-state photocurrent as a function of the bending cycles. A bending cycle is referred to as the completion of an ON/OFF light illumination under a bending radius of 10 mm and the bias conditions of Vds = −1 V and Vgs = 1 V. The Gr−WS2 junction maintains its structural integrity and exhibits a stable photocurrent within the tested cycles, indicative of its decent flexibility and reliable durability for applications in wearable optoelectronic devices.





CONCLUSION In summary, we show all-CVD growth of vertical Gr−WS2 heterojunctions. An individually addressable Al/AlOx top gate was fabricated on top of the Gr−WS2 junction, forming phototransistors on either sapphire or PET substrates. The Al/ AlOx gate architecture allows for a self-aligned deposition of drain/source contacts, reducing the access regions and lowering the parasitic resistance accordingly. The resulting Gr−WS2 photodetectors exhibit gate-tunable high photocurrents. The photoresponsivity reaches 4.0 A/W. The low dark current yields an equivalent noise power as low as 2.5 × 10−16 W/Hz1/2. Further, the self-passivated interface oxidation layer has shown its potential to act as a robust gate insulator on our Gr−WS2 photodetectors and is also anticipated to extend its unique features to other 2D materials for widespread applications.35,36 An ongoing research is to optimize the number of constituent layers in the Gr−TMD junctions and to explore other TMD− TMD junctions for a superior performance of photodetection. Their better control over the thickness, clean interjunction interface, scalability, reliability, and integrability with silicon technology make the 2D heterojunctions an ideal candidate for future optoelectronic applications.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.7b10892. Additional AFM, XPS, Raman, PL, HRTEM, and STEM characterizations of as-grown TMD films, electrical properties of as-grown graphene, contact improvement via access resistance reduction, information on a homemade hard mask kit as well as the chalcogenization procedure, and the external quantum efficiency and photocurrent gain of Gr-WS2 phototransistors. (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Chao-Hui Yeh: 0000-0002-9437-055X Yung-Chang Lin: 0000-0002-3968-7239 Tien-Lin Wu: 0000-0001-9061-6279 Ying-Hao Chu: 0000-0002-3435-9084

EXPERIMENTAL SECTION

Author Contributions

Growth of Vertical Heterojunctions. Few-layer graphene was grown using our previously reported method involving the direct formation of graphene on oxide with remote catalyzation. The graphene film was then patterned as 250 × 1000 μm2 stripes. The Raman characterization and van der Pauw measurements of our asgrown graphene film are given in Figure S9 in the Supporting Information. For TMDs, a patterned transition-metal oxide seed layer was deposited by thermal evaporation using tungsten(VI) oxide with 99.995% purity or molybdenum(VI) oxide with 99.97% purity. Subsequently, sulfurization or selenization was carried out in a 3-in. quartz tube placed in a dual-walled tube furnace equipped with radiofrequency plasma. The samples were placed at the center of the tube furnace, and the chalcogen precursors (sulfur or selenium powders with 99.998% purity) were located at the upstream of the gas flow, set up with an independent heating source. The samples and chalcogen powders were heated to 700−800 and 140 °C, respectively, depending on the kind of transition metal. Vaporization of the chalcogen precursor was carried by argon (200 sccm) and hydrogen (5−40 sccm) through a plasma region. The chalcogenization reaction takes

C.-H.Y. and Z.-Y.L. contributed equally to the work, and both contributors are considered first authors. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We are thankful to the Best Champion Technology Co., Ltd., for technical support. C.-H.Y. acknowledges C.-C. Andow Liao for valuable discussion. P.-W.C. appreciates the project support of Taiwan Ministry of Science and Technology: Grants MOST 103-2628-M-007-004-MY3 and MOST 103-2119-M-007-008MY3.



REFERENCES

(1) Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-Layer MoS2 Transistors. Nat. Nanotechnol. 2011, 6, 147−150.

36186

DOI: 10.1021/acsami.7b10892 ACS Appl. Mater. Interfaces 2017, 9, 36181−36188

Research Article

ACS Applied Materials & Interfaces

Sapphire and Imaging Its Grain Boundary. ACS Nano 2013, 7, 8963− 8971. (19) Mouri, S.; Miyauchi, Y.; Matsuda, K. Tunable Photoluminescence of Monolayer MoS2 via Chemical Doping. Nano Lett. 2013, 13, 5944−5948. (20) Tonndorf, P.; Schmidt, R.; Bottger, P.; Zhang, X.; Borner, J.; Liebig, A.; Albrecht, M.; Kloc, C.; Gordan, O.; Zahn, D. R.; Michaelis de Vasconcellos, S.; Bratschitsch, R. Photoluminescence Emission and Raman Response of Monolayer MoS2, MoSe2, and WSe2. Opt. Express 2013, 21, 4908−4916. (21) Teng, P.-Y.; Lu, C.-C.; Akiyama-Hasegawa, K.; Lin, Y.-C.; Yeh, C.-H.; Suenaga, K.; Chiu, P.-W. Remote Catalyzation for Direct Formation of Graphene Layers on Oxides. Nano Lett. 2012, 12, 1379− 1384. (22) Lin, Y. C.; Lu, C. C.; Yeh, C. H.; Jin, C. H.; Suenaga, K.; Chiu, P. W. Graphene Annealing: How Clean Can It Be? Nano Lett. 2012, 12, 414−419. (23) Lu, C. C.; Lin, Y. C.; Yeh, C. H.; Huang, J. C.; Chiu, P. W. High Mobility Flexible Graphene Field-Effect Transistors with Self-Healing Gate Dielectrics. ACS Nano 2012, 6, 4469−4474. (24) Tu, C.-H.; Chen, W.; Fang, H.-C.; Tzeng, Y.; Liu, C.-P. Heteroepitaxial Nucleation and Growth of Graphene Nanowalls on Silicon. Carbon 2013, 54, 234−240. (25) Tongay, S.; Fan, W.; Kang, J.; Park, J.; Koldemir, U.; Suh, J.; Narang, D. S.; Liu, K.; Ji, J.; Li, J. B.; Sinclair, R.; Wu, J. Q. Tuning Interlayer Coupling in Large-Area Heterostructures with CVD-Grown MoS2 and WS2 Monolayers. Nano Lett. 2014, 14, 3185−3190. (26) Zhang, X.; Meng, F.; Christianson, J. R.; Arroyo-Torres, C.; Lukowski, M. A.; Liang, D.; Schmidt, J. R.; Jin, S. Vertical Heterostructures of Layered Metal Chalcogenides by Van Der Waals Epitaxy. Nano Lett. 2014, 14, 3047−3054. (27) Pierucci, D.; Henck, H.; Naylor, C. H.; Sediri, H.; Lhuillier, E.; Balan, A.; Rault, J. E.; Dappe, Y. J.; Bertran, F.; Le Fevre, P.; Johnson, A. T. C.; Ouerghi, A. Large Area Molybdenum Disulphide-Epitaxial Graphene Vertical Van Der Waals Heterostructures. Sci. Rep. 2016, 6, 26656. (28) Zhang, W. J.; Chuu, C. P.; Huang, J. K.; Chen, C. H.; Tsai, M. L.; Chang, Y. H.; Liang, C. T.; Chen, Y. Z.; Chueh, Y. L.; He, J. H.; Chou, M. Y.; Li, L. J. Ultrahigh-Gain Photodetectors Based on Atomically Thin Graphene-MoS2 Heterostructures. Sci. Rep. 2015, 4, 3826. (29) Pierucci, D.; Henck, H.; Avila, J.; Balan, A.; Naylor, C. H.; Patriarche, G.; Dappe, Y. J.; Silly, M. G.; Sirotti, F.; Johnson, A. T. C.; Asensio, M. C.; Ouerghi, A. Band Alignment and Minigaps in Monolayer MoS2-Graphene Van Der Waals Heterostructures. Nano Lett. 2016, 16, 4054−4061. (30) 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. (31) Ganatra, R.; Zhang, Q. Few-Layer MoS2: A Promising Layered Semiconductor. ACS Nano 2014, 8, 4074−4099. (32) Tan, H.; Fan, Y.; Zhou, Y.; Chen, Q.; Xu, W.; Warner, J. H. Ultrathin 2D Photodetectors Utilizing Chemical Vapor Deposition Grown WS2 with Graphene Electrodes. ACS Nano 2016, 10, 7866− 7873. (33) Guimarães, M. H. D.; Gao, H.; Han, Y. M.; Kang, K.; Xie, S.; Kim, C. J.; Muller, D. A.; Ralph, D. C.; Park, J. Atomically Thin Ohmic Edge Contacts Between Two-Dimensional Materials. ACS Nano 2016, 10, 6392−6399. (34) Krainak, M. A.; Sun, X. L.; Yang, G. N.; Lu, W. Comparison of Linear-Mode Avalanche Photodiode Lidar Receivers for Use at One Micron Wavelength. Proc. SPIE 2010, 7681. (35) Carey, B. J.; Ou, J. Z.; Clark, R. M.; Berean, K. J.; Zavabeti, A.; Chesman, A. S.; Russo, S. P.; Lau, D. W.; Xu, Z. Q.; Bao, Q.; Kevehei, O.; Gibson, B. C.; Dickey, M. D.; Kaner, R. B.; Daeneke, T.; KalantarZadeh, K. Wafer-Scale Two-Dimensional Semiconductors from Printed Oxide Skin of Liquid Metals. Nat. Commun. 2017, 8, 14482.

(2) Yu, L.; El-Damak, D.; Radhakrishna, U.; Ling, X.; Zubair, A.; Lin, Y.; Zhang, Y.; Chuang, M. H.; Lee, Y. H.; Antoniadis, D.; Kong, J.; Chandrakasan, A.; Palacios, T. Design, Modeling, and Fabrication of Chemical Vapor Deposition Grown MoS2 Circuits with E-Mode FETs for Large-Area Electronics. Nano Lett. 2016, 16, 6349−6356. (3) Tosun, M.; Chuang, S.; Fang, H.; Sachid, A. B.; Hettick, M.; Lin, Y. J.; Zeng, Y. P.; Javey, A. High-Gain Inverters Based on WSe2 Complementary Field-Effect Transistors. ACS Nano 2014, 8, 4948− 4953. (4) Ho, P. H.; Chang, Y. R.; Chu, Y. C.; Li, M. K.; Tsai, C. A.; Wang, W. H.; Ho, C. H.; Chen, C. W.; Chiu, P. W. High-Mobility InSe Transistors: The Role of Surface Oxides. ACS Nano 2017, 11, 7362− 7370. (5) Massicotte, M.; Schmidt, P.; Vialla, F.; Schadler, K. G.; ReserbatPlantey, A.; Watanabe, K.; Taniguchi, T.; Tielrooij, K. J.; Koppens, F. H. L. Picosecond Photoresponse in Van Der Waals Heterostructures. Nat. Nanotechnol. 2015, 11, 42−46. (6) Cheng, R.; Li, D. H.; Zhou, H. L.; Wang, C.; Yin, A. X.; Jiang, S.; Liu, Y.; Chen, Y.; Huang, Y.; Duan, X. F. Electroluminescence and Photocurrent Generation from Atomically Sharp WSe 2 /MoS 2 Heterojunction p-n Diodes. Nano Lett. 2014, 14, 5590−5597. (7) Mak, K. F.; Shan, J. Photonics and Optoelectronics of 2D Semiconductor Transition Metal Dichalcogenides. Nat. Photonics 2016, 10, 216−226. (8) Sarkar, D.; Xie, X.; Liu, W.; Cao, W.; Kang, J.; Gong, Y.; Kraemer, S.; Ajayan, P. M.; Banerjee, K. A Subthermionic Tunnel Field-Effect Transistor with an Atomically Thin Channel. Nature 2015, 526, 91− 95. (9) Yu, W. J.; Li, Z.; Zhou, H. L.; Chen, Y.; Wang, Y.; Huang, Y.; Duan, X. F. Vertically Stacked Multi-Heterostructures of Layered Materials for Logic Transistors and Complementary Inverters. Nat. Mater. 2012, 12, 246−252. (10) Yang, H.; Heo, J.; Park, S.; Song, H. J.; Seo, D. H.; Byun, K.-E.; Kim, P.; Yoo, I.; Chung, H.-J.; Kim, K. Graphene Barristor, a Triode Device with a Gate-Controlled Schottky Barrier. Science 2012, 336, 1140. (11) Yu, W. J.; Liu, Y.; Zhou, H. L.; Yin, A. X.; Li, Z.; Huang, Y.; Duan, X. F. Highly Efficient Gate-Tunable Photocurrent Generation in Vertical Heterostructures of Layered Materials. Nat. Nanotechnol. 2013, 8, 952−958. (12) Lopez-Sanchez, O.; Lembke, D.; Kayci, M.; Radenovic, A.; Kis, A. Ultrasensitive Photodetectors Based on Monolayer MoS2. Nat. Nanotechnol. 2013, 8, 497−501. (13) Wang, H.; Yu, L.; Lee, Y. H.; Shi, Y.; Hsu, A.; Chin, M. L.; Li, L. J.; Dubey, M.; Kong, J.; Palacios, T. Integrated Circuits Based on Bilayer MoS2 Transistors. Nano Lett. 2012, 12, 4674−4680. (14) Woods, J. M.; Jung, Y.; Xie, Y.; Liu, W.; Liu, Y.; Wang, H.; Cha, J. J. One-Step Synthesis of MoS2/WS2 Layered Heterostructures and Catalytic Activity of Defective Transition Metal Dichalcogenide Films. ACS Nano 2016, 10, 2004−2009. (15) Xue, Y. Z.; Zhang, Y. P.; Liu, Y.; Liu, H. T.; Song, J. C.; Sophia, J.; Liu, J. Y.; Xu, Z. Q.; Xu, Q. Y.; Wang, Z. Y.; Zheng, J. L.; Liu, Y. Q.; Li, S. J.; Bao, Q. L. Scalable Production of a Few-Layer MoS2/WS2 Vertical Heterojunction Array and Its Application for Photodetectors. ACS Nano 2016, 10, 573−580. (16) Britnell, L.; Gorbachev, R. V.; Jalil, R.; Belle, B. D.; Schedin, F.; Mishchenko, A.; Georgiou, T.; Katsnelson, M. I.; Eaves, L.; Morozov, S. V.; Peres, N. M.; Leist, J.; Geim, A. K.; Novoselov, K. S.; Ponomarenko, L. A. Field-Effect Tunneling Transistor Based on Vertical Graphene Heterostructures. Science 2012, 335, 947−950. (17) Feng, Y.; Zhang, K.; Wang, F.; Liu, Z.; Fang, M.; Cao, R.; Miao, Y.; Yang, Z.; Mi, W.; Han, Y.; Song, Z.; Wong, H. S. Synthesis of Large-Area Highly Crystalline Monolayer Molybdenum Disulfide with Tunable Grain Size in a H2 Atmosphere. ACS Appl. Mater. Interfaces 2015, 7, 22587−22593. (18) Zhang, Y.; Zhang, Y.; Ji, Q.; Ju, J.; Yuan, H.; Shi, J.; Gao, T.; Ma, D.; Liu, M.; Chen, Y.; Song, X.; Hwang, H. Y.; Cui, Y.; Liu, Z. Controlled Growth of High-Quality Monolayer WS2 Layers on 36187

DOI: 10.1021/acsami.7b10892 ACS Appl. Mater. Interfaces 2017, 9, 36181−36188

Research Article

ACS Applied Materials & Interfaces (36) Kalantar-zadeh, K.; Ou, J. Z.; Daeneke, T.; Mitchell, A.; Sasaki, T.; Fuhrer, M. S. Two Dimensional and Layered Transition Metal Oxides. Applied Materials Today 2016, 5, 73−89.

36188

DOI: 10.1021/acsami.7b10892 ACS Appl. Mater. Interfaces 2017, 9, 36181−36188