Atomic-Monolayer MoS Band-to-Band Tunneling Field...

8
www.MaterialsViews.com 1 © 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.small-journal.com Atomic-Monolayer MoS 2 Band-to-Band Tunneling Field-Effect Transistor Yann-Wen Lan,* Carlos M. Torres Jr.,* Shin-Hung Tsai, Xiaodan Zhu, Yumeng Shi, Ming-Yang Li, Lain-Jong Li, Wen-Kuan Yeh, and Kang L. Wang* The 20th century ushered in the Information Age with the silicon electronics industry’s aggressive and persistent scaling down of the complementary metal-oxide-semiconductor field-effect transistors’ (MOSFET) dimensions. However, we now face the impending stall of Moore’s Law and must find alternative materials and novel device concepts which may augment commercial silicon technology for enabling high performance and low power consumption. As a consequence, a tunneling field-effect transistor (TFET) has been imple- mented in various material systems, [1–5] Although the TFET features a similar device structure as the MOSFET, its fun- damental switching mechanism differs from the thermionic emission process by relying upon the modulation of quantum mechanical band-to-band tunneling (BTBT) through a tunnel barrier in order to inject carriers into its device channel. Fur- thermore, since the TFET relies on quantum mechanical DOI: 10.1002/smll.201601310 Transistors Dr. Y.-W. Lan National Nano Device Laboratories (NDL) National Applied Research Laboratories Hsinchu 30078, Taiwan E-mail: [email protected] Dr. Y.-W. Lan, Dr. C. M. Torres Jr., [+] S.-H. Tsai, X. Zhu, Prof. K. L. Wang Department of Electrical Engineering University of California at Los Angeles Los Angeles, CA 90095, USA E-mail: [email protected]; [email protected] Dr. Y. Shi, Dr. M.-Y. Li, Prof. L.-J. Li Physical Sciences and Engineering Division King Abdullah University of Science and Technology (KAUST) Thuwal 23955-6900, Kingdom of Saudi Arabia Dr. M.-Y. Li Research Center for Applied Sciences Academia Sinica Taipei 10617, Taiwan Prof. W.-K. Yeh National Nano Device Laboratories (NDL) National Applied Research Laboratories Hsinchu 30078, Taiwan Prof. W.-K. Yeh Department of Electrical Engineering National University of Kaohsiung Kaohsiung 811, Taiwan [+] Present address: Space and Naval Warfare (SPAWAR) Systems Center Pacific, San Diego, CA 92152, USA tunneling through a tunnel barrier, both its off-state leakage current as well as its subthreshold swing (SS) can actually be much lower than that of the MOSFET. Thus, the TFET is a promising contender for low-energy electronics albeit there are several obstacles it must first circumvent in order to truly find practical applications. Some of these obstacles in bulk materials include its very low on-state current, the spatial inhomogeneity of the semiconductor channel’s thickness and doping which ultimately smears out its band edges (and limits the steepness of the SS), the limited magnitude of the electric field across the tunnel barrier, and the efficacy of the electro- static gate control over the device channel. Furthermore, the dangling bonds in bulk materials can lead to surface states and traps, which are detrimental to the SS. Recently, the advent of atomically thin 2D crystals has sparked a paradigm shift in nanotechnology. [6–13] Since the last decade, we are now able to truly investigate and imple- ment novel device concepts at the ultimate physical limit in addition to having at our disposal a myriad of 2D materials, each of which exhibits unique electronic and optical proper- ties. [14–17] There exist many 2D materials with energy band- gaps in the range between 1 and 2 eV for scaled low-power devices, in which they all possess a fully terminated surface (free of dangling bonds), which is beneficial for lowering the SS [18] and these bandgap values can provide the thermal barriers needed to achieve low off-state currents. Further- more, because abrupt p–n junctions are limited by the solid solubility of the dopant atoms in bulk semiconductors, this restricts the magnitudes of the internal electric fields and lowers the tunneling current. By the use of an atomically thin layer, 2D materials may give much higher tunneling current densities. In fact, some theoretical works predict that layered transition metal dichalcogenide (TMD) materials support direct BTBT in their monolayer sheet forms in addition to providing high on-state currents due to the atomically thin tunneling barrier and low SS. [19–23] However, there currently only exist a few experimental reports on TFETs made from few-layered TMD materials. [24–26] Accordingly, an experi- mental demonstration of BTBT in a monolayer TMD-based TFET is critical to determine the utility of 2D materials for future applications. In this letter, we demonstrate a novel vertical tunneling field-effect transistor with a monolayer of MoS 2 as the channel (MoS 2 -TFET). To the best of our knowledge, this is the first experimental demonstration of utilizing a monolayer TMD material as a channel for the small 2016, DOI: 10.1002/smll.201601310

Transcript of Atomic-Monolayer MoS Band-to-Band Tunneling Field...

www.MaterialsViews.com

1© 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.small-journal.com

Atomic-Monolayer MoS2 Band-to-Band Tunneling Field-Effect Transistor

Yann-Wen Lan,* Carlos M. Torres Jr.,* Shin-Hung Tsai, Xiaodan Zhu, Yumeng Shi, Ming-Yang Li, Lain-Jong Li, Wen-Kuan Yeh, and Kang L. Wang*

The 20th century ushered in the Information Age with the silicon electronics industry’s aggressive and persistent scaling down of the complementary metal-oxide-semiconductor field-effect transistors’ (MOSFET) dimensions. However, we now face the impending stall of Moore’s Law and must find alternative materials and novel device concepts which may augment commercial silicon technology for enabling high performance and low power consumption. As a consequence, a tunneling field-effect transistor (TFET) has been imple-mented in various material systems,[1–5] Although the TFET features a similar device structure as the MOSFET, its fun-damental switching mechanism differs from the thermionic emission process by relying upon the modulation of quantum mechanical band-to-band tunneling (BTBT) through a tunnel barrier in order to inject carriers into its device channel. Fur-thermore, since the TFET relies on quantum mechanical

DOI: 10.1002/smll.201601310

Transistors

Dr. Y.-W. LanNational Nano Device Laboratories (NDL)National Applied Research LaboratoriesHsinchu 30078, TaiwanE-mail: [email protected]

Dr. Y.-W. Lan, Dr. C. M. Torres Jr.,[+] S.-H. Tsai, X. Zhu, Prof. K. L. WangDepartment of Electrical EngineeringUniversity of California at Los AngelesLos Angeles, CA 90095, USAE-mail: [email protected]; [email protected]

Dr. Y. Shi, Dr. M.-Y. Li, Prof. L.-J. LiPhysical Sciences and Engineering DivisionKing Abdullah University of Science and Technology (KAUST)Thuwal 23955-6900, Kingdom of Saudi Arabia

Dr. M.-Y. LiResearch Center for Applied SciencesAcademia SinicaTaipei 10617, Taiwan

Prof. W.-K. YehNational Nano Device Laboratories (NDL)National Applied Research LaboratoriesHsinchu 30078, Taiwan

Prof. W.-K. YehDepartment of Electrical EngineeringNational University of KaohsiungKaohsiung 811, Taiwan[+]Present address: Space and Naval Warfare (SPAWAR) Systems Center Pacific, San Diego, CA 92152, USA

tunneling through a tunnel barrier, both its off-state leakage current as well as its subthreshold swing (SS) can actually be much lower than that of the MOSFET. Thus, the TFET is a promising contender for low-energy electronics albeit there are several obstacles it must first circumvent in order to truly find practical applications. Some of these obstacles in bulk materials include its very low on-state current, the spatial inhomogeneity of the semiconductor channel’s thickness and doping which ultimately smears out its band edges (and limits the steepness of the SS), the limited magnitude of the electric field across the tunnel barrier, and the efficacy of the electro-static gate control over the device channel. Furthermore, the dangling bonds in bulk materials can lead to surface states and traps, which are detrimental to the SS.

Recently, the advent of atomically thin 2D crystals has sparked a paradigm shift in nanotechnology.[6–13] Since the last decade, we are now able to truly investigate and imple-ment novel device concepts at the ultimate physical limit in addition to having at our disposal a myriad of 2D materials, each of which exhibits unique electronic and optical proper-ties.[14–17] There exist many 2D materials with energy band-gaps in the range between 1 and 2 eV for scaled low-power devices, in which they all possess a fully terminated surface (free of dangling bonds), which is beneficial for lowering the SS[18] and these bandgap values can provide the thermal barriers needed to achieve low off-state currents. Further-more, because abrupt p–n junctions are limited by the solid solubility of the dopant atoms in bulk semiconductors, this restricts the magnitudes of the internal electric fields and lowers the tunneling current. By the use of an atomically thin layer, 2D materials may give much higher tunneling current densities. In fact, some theoretical works predict that layered transition metal dichalcogenide (TMD) materials support direct BTBT in their monolayer sheet forms in addition to providing high on-state currents due to the atomically thin tunneling barrier and low SS.[19–23] However, there currently only exist a few experimental reports on TFETs made from few-layered TMD materials.[24–26] Accordingly, an experi-mental demonstration of BTBT in a monolayer TMD-based TFET is critical to determine the utility of 2D materials for future applications. In this letter, we demonstrate a novel vertical tunneling field-effect transistor with a monolayer of MoS2 as the channel (MoS2-TFET). To the best of our knowledge, this is the first experimental demonstration of utilizing a monolayer TMD material as a channel for the

small 2016, DOI: 10.1002/smll.201601310

communicationswww.MaterialsViews.com

2 www.small-journal.com © 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

study of BTBT in a p–n and n+–n junction, in which these diode junctions are electrostatically induced by a local top-gate electrode. Furthermore, by applying either positive or negative top-gate voltages, a strong gate-induced doping of the monolayer MoS2 channel is achieved and the presence of large internal electric fields results in the direct manifestation of a BTBT current and an ambipolar transport.

The device schematic of the vertical monolayer MoS2 tunneling field-effect transistor (MoS2-TFET) and its optical micrograph are shown in Figure 1. The cross-sectional view of the MoS2 device structure is shown in Figure 1a. The three-terminal device consists of a degenerately doped n++ silicon substrate (ND ≈ 1019 cm−3) as the drain (D), a monolayer of chemical vapor deposition (CVD) grown MoS2 as the channel with source electrodes (S) contacted on top of the MoS2, and sputtered indium tin oxide (ITO) (≈45 nm) as the top-gate (G). A thermally grown thin SiO2 tunnel barrier separates the n++ silicon and monolayer MoS2 channel, whereas an atomic-layer deposited (≈55 nm) HfO2 serves as the top-gate dielectric in between the monolayer MoS2 channel and the ITO top-gate electrode. In this fabrication process, arrays of vertical monolayer MoS2-TFETs are isolated from each other via a thick (≈300 nm) SiO2 field oxide. The underlying silicon surface which is not covered by the thick SiO2 field oxide, but rather surrounded by it, serves as the current confinement region of the device for current injection. Thus, the fabrica-tion process was designed to be compatible with silicon com-plementary metal oxide semiconductor (CMOS) technology. A top-view optical micrograph of an actual MoS2-TFET is presented in Figure 1b. The detailed fabrication process is described in the Experimental Section. In this particular study, a common-source configuration was employed during the electrical measurements and a schematic biasing circuit

for the MoS2 device is shown in Figure 1a. Both of the source contacts on top of the monolayer MoS2 are grounded during the electrical measurements in order to achieve a uniform potential distribution across the monolayer MoS2 region. Accordingly, an electric current can vertically tunnel from the underlying n++ silicon substrate to the monolayer MoS2 that serves as a constant current source, followed by in-plane/lat-eral transport along the monolayer MoS2 channel toward the source electrodes. The measurements were performed in air and at room temperature. It should be noted that the con-tact between MoS2 and the n++ silicon substrate is dominated by a tunneling behavior while the contact between MoS2 and metal is Ohmic-like behavior. Finally, we note that the Fermi level of the MoS2 is modulated only in the area directly underneath the top-gate electrode as illustrated in Figure 1c and that the remaining MoS2 regions which connect this cen-tral region to the source electrodes act as parasitic series resistors to ground. In the next section, we shall proceed to discuss the energy band diagrams in regards to the BTBT tunneling current.

In order to clearly understand the physics, we first focus on describing the two modes of operation for the gate induced p–n junction using energy band diagrams. There are two possible cases for the gate induced p–n junction of the MoS2 vertical devices depending upon the polarity of the applied top-gate voltage Vg. The first case, shown in Figure 1c, describes positive charges induced on the left side of the monolayer MoS2 which occurs when Vg < 0. In this scenario, an in-plane/lateral p–n junction is formed due to the local electrostatically induced holes via the negative voltage on the top-gate on the left side of the MoS2. The corresponding energy band diagram is schematically shown in Figure 2a. The second case, shown in Figure 1d, describes negative charges

small 2016, DOI: 10.1002/smll.201601310

Figure 1. Schematic illustration of a vertical MoS2-TFET structure and device optical microscope image. a) Cross-sectional view of the device structure with a monolayer MoS2 as the channel. The capital letters S, D, and G represent the source, drain, and gate, respectively. b) Optical micrograph (top-view) of an actual MoS2-TFET device. The scale bar is 100 μm. The dashed circle outlines the MoS2 region. c,d) Schematic plots for the electrostatically induced p–n junction and n+–n (similar to n–p) junction in the monolayer MoS2 channel.

www.MaterialsViews.com

3© 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.small-journal.com

accumulated on the left side of the monolayer MoS2 and occurs when Vg > 0. Accordingly, in this scenario, an in-plane n+–n junction is produced due to the local electrostatically induced electrons via the positive voltage on the top-gate as compared to the lower electron density that exists in the intrinsically n-type monolayer MoS2 in the region not cov-ered by the top-gate. Thus, the transport behavior is expected to be qualitatively similar to an n–p junction. The corre-sponding energy band diagram is schematically presented in Figure 2b. This local electrostatic control of the p–n and the n+–n junction behavior with the top-gate is experimentally observed in the electrical measurements. Figure 2c shows the Ids − Vds curves for three different top-gate voltages (Vg = −10, 0, and +4 V), in which it exhibits linear characteristics at Vg = 0 V, meaning an Ohmic-like behavior exists across both contacts (n++ silicon/MoS2 and MoS2/source electrode). The reason for the nonsymmetric Ids − Vds curves at Vg = −10 V and Vg = +4 V is due to the fact that the electrical transport is affected by the local electrostatic control of the p–n and the n+–n junction behavior with the top-gate. Upon the applica-tion of a negative top-gate voltage, a p–n junction is induced along the MoS2 planar channel. On the other hand, an n+–n junction is induced along the MoS2 planar channel with the application of a positive top-gate voltage. Accordingly, the measured drain current increases (e.g., diode-like charac-teristics) in both the forward-biased region with the appli-cation of a negative top-gate voltage (Vg = −10 V) and the reverse-biased region with the application of a positive top-gate voltage (Vg = +4 V). We noted that the leakage current

across the HfO2 top-gate dielectric is typically around 50 pA for the highest applied gate voltages of Vg = ±10 V. Thus, the magnitude of this leakage current across the HfO2 top-gate dielectric is far less than the change in Ids due to the modulation of the MoS2 layer via the top-gate voltage in our MoS2-TFETs. Figure 2d shows a 2D contour plot of the current intensity as a function of both the bias voltage (Vds) and the top-gate voltage (Vg). This 2D plot clearly illustrates the existence of a field-effect on the measured drain current via different applied top-gate voltages in either the forward-biased (Vg < 0) or reverse-biased (Vg > 0) regions.

For this specific device (A25), the threshold top-gate voltage for the formation of a p–n junction and an n+–n junc-tion is roughly found to be −6 and 0 V, respectively. At top-gate voltages between −6 and 0 V, both the current amplitude and the Ids − Vds curves almost remain the same. This serves as a transition region which varies among the devices. Conse-quently, either a p–n junction or an n+–n junction forms along the monolayer MoS2 channel with the application of a local negative or positive top-gate voltage. More interestingly, we find that the drain current in the reverse-biased region of the p–n junction (Vds < 0) and the forward-biased region of the n+–n junction (Vds > 0) also increase for certain applied top-gate voltages (see Figure 3). This increase in drain current is attributed to BTBT in the local electrostatically induced p–n junction across the monolayer MoS2 channel, in which electrons may tunnel from the valence band to the conduc-tion band, or vice versa. The required conditions for BTBT are listed as follows: (1) There exist occupied energy states

small 2016, DOI: 10.1002/smll.201601310

Figure 2. Energy band diagrams and electrical characterization for the p–n and n+–n junction of the monolayer MoS2-TFETs. a) Band diagram of the p–n junction induced along the MoS2 channel by the application of a negative top-gate voltage. b) Band diagram of the n+–n junction induced along the MoS2 channel by the application of a positive gate voltage. c) Drain current versus bias voltages for three top-gate voltages of −10, 0, and +4 V. d) Current intensity plotted as a function of both the bias voltages and the top-gate voltages. Either a p–n junction or an n+–n junction forms along the monolayer MoS2 channel with the application of a local negative or positive top-gate voltage.

communicationswww.MaterialsViews.com

4 www.small-journal.com © 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

in the region from which the electrons tunnel. (2) There exist unoccupied or available energy states at the same energy level in the region to which the electrons can tunnel into. (3) The tunneling barrier height and width is small enough so that a finite tunneling probability exists. (4) The transverse momentum is conserved in the tunneling process. On the basis of the above conditions, a BTBT current could exist in our MoS2-TFETs.

In order to better display the increase in the measured drain current arising from the BTBT process, a 3D plot (drain current as a function of both the bias voltages and the top-gate voltages) in another monolayer MoS2-TFET (B21) is presented in Figure 3. This 3D plot clearly shows a bipolar transport behavior. The substantial increase in the drain current upon reverse biasing the p–n junction (Vds < 0) and forward biasing the n+–n junction (Vds > 0) is attributed to electron tunneling from the valence band into the con-duction band of the MoS2. This situation is illustrated in the two graphs of Figure 3, in which electron tunneling from the valence band into the conduction band occurs when a large gate voltage is negatively applied to the p–n junction (Figure 3b) and is positively applied to the n+–n junction (Figure 3a), respectively. The first BTBT operational mode occurs when a large negative top-gate voltage induces a lat-eral (e.g., in-plane) p–n junction along the monolayer MoS2 channel. In this case, BTBT occurs at reverse bias, which is similar to the traditional TFET. The second BTBT opera-tional mode occurs when a large positive top-gate voltage induces a lateral (e.g., in-plane) n+–n junction along the mono-layer MoS2 channel. In this case, BTBT occurs at forward bias, which is a novel effect and unique to our MoS2-TFET devices. Furthermore, the fact that the tunneling cur-rent increases with increasing gate voltages for both cases shows that the tunneling window (ΔE) depends on the gate

voltages. Once the gate voltages exceed the threshold voltage for BTBT, more electrons can tunnel due to the enhanced tunneling probability. It is noted that if we apply a forward Vds bias with a large negative gate voltage, the energy band on the left side of the MoS2 is dragged downward, and then BTBT will not occur because there is no tunneling window between the left side of the valence band and right side of the conduction band along the planar MoS2 channel. However, if the reverse Vds bias is applied, the tunneling window is open due to the superposition effect from the reverse Vds bias and then the band-to-band tunneling phenomenon will appear in this circumstance. Similarly, for a large positive gate voltage, the application of a forward Vds bias will further increase the tunneling window, as shown in Figure 3a. Therefore, we believe that the superposition effect from the applied gate and drain voltages is one of the key factors, which enables the realization of the band-to-band tunneling in our device system. Additionally, the BTBT process can be either direct or indirect. For direct tunneling to occur, the conduction band minimum and the valence band maximum must have the same momentum. In our case, direct tunneling is the most likely process since a monolayer MoS2 semiconductor with direct bandgap satisfies this condition. Accordingly, the electrons can tunnel from the vicinity of the valence band maximum to the vicinity of the conduction band minimum without a change of momentum in our monolayer MoS2-TFETs (see Figure 3a,b). Nevertheless, the contribution of phonons will definitely exist in any system with temperature higher than 0 K. In our devices, scattering processes involving spontaneous phonon emission is possible and may result in noise as well as a slightly lowered effective conductivity.

The most important figures of merit for a tunneling FET are the subthreshold swing and the ON/OFF ratio. In our first batch of MoS2-TFETs, the average subthreshold slope

small 2016, DOI: 10.1002/smll.201601310

Figure 3. BTBT in a local top-gate voltage induced p–n (or n+–n) junction along the MoS2 channel for a MoS2-TFET. 3D plot of the measured drain current as a function of both the bias voltages and the top-gate voltages. a) Schematic band diagram of the BTBT that occurs when large positive top-gate voltages induce an n+–n junction along the monolayer MoS2 channel. b) Schematic band diagram of the BTBT that occurs when large negative top-gate voltages induce a p–n junction along the monolayer MoS2 channel.

www.MaterialsViews.com

5© 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.small-journal.com

is about 3.1 V per decade with an ON/OFF ratio of about two orders magnitude as shown in Figure 4a. Although these values are still far away from the minimum subthreshold swing of a conventional device (≈60 mV per decade), we believe it could be improved by reducing the parasitics upon optimizing our prototype device structure. In our opinion, the implementation of other 2D atomic-layered materials featuring both direct and smaller bandgaps (e.g., black phos-phorous) than monolayer MoS2 may further improve the ON-state current in our TFETs. Other possibilities[27] include the use of staggered and broken-gap heterojunctions in order to shorten the tunneling distance, implementing layered 2D van der Waals heterojunctions, passivating the defect states which may exist at the surface of transition metal dichalco-genides, use of strain engineering, implementing materials with low tunneling effective masses, and exploiting materials (e.g., bilayer graphene) which feature van Hove singularities (e.g., peaks in the density of state) near the band edges.[28] However, it is interesting that p-type conduction in MoS2 is efficiently achieved in our vertical MoS2-TFET device struc-ture upon application of negative top-gate voltages, since it is not easy to achieve hole-dominated transport in MoS2 due to the fact that MoS2 is an intrinsically n-type material. This feature reasserts that a lateral p–n junction is formed due to the local electrostatically induced holes via the applied nega-tive top-gate voltage on the left side of the MoS2. To confirm the existence of band-to-band tunneling behavior, the tem-perature dependence (77–300 K) of the tunnel current is performed at a large positive top-gate voltage (Vg = +10 V), which enables BTBT to occur in the forward bias voltages whereas a thermionic emission mechanism dominates the electrical transport in the reverse Vds bias. As is evident in Figure 4b, the measured drain current exhibits a weak tem-perature dependence in the forward bias regime which is indicative of BTBT current, whereas the measured drain current exhibits a temperature dependence greater than an order of magnitude in the reverse bias regime due to the cur-rent being dominated by a thermionic emission mechanism (e.g., BTBT does not dominate in this case).

In order to further evaluate the observed BTBT current in the local field-induced lateral p–n junctions across the mono-layer MoS2, we have performed numerical calculations based on the following analytical model. Using the Wentzel–Kramers–

Brillouin approximation, the direct tunneling probability for the BTBT process in monolayer MoS2 can be approximated as

exp 2 ( )d expBTBT0

1g

Te

k E EE

∫ξξξ= −

≡ −

(1)

where e is the electronic charge, ξ is the electric field, E = 0 is the valence band edge, E = Eg is the conduction band edge, k(E) is the magnitude of the imaginary wave vector with “least action for tunneling” within the forbidden gap, and ξI = 2/e ( )d

0

g

k E EE

∫( ) is essentially the “least action integral” which is an intrinsic property of the MoS2 and was calcu-lated to be 1.02 V nm−1 in theory.[22] According to theoretical works,[21] the tunneling current in the local field-induced lat-eral p–n junctions across the monolayer MoS2 can be approx-imately described by

In 12 1 coshBTBT

S VB

BBTBTI

g g eh k T E

K T T( )= + ∆

(2)

where h is Planck’s constant, gs and gv are the spin and valley degeneracies, respectively, kB is the Boltzmann con-stant, T is the temperature, and ΔE is the tunneling energy window between the conduction band and the valence band for the case of the n+–n junction (see Figure 3a) and for the case of the p–n junction (see Figure 3b). Next, we combine these two equations and define the following parameters. The fitting parameter a is defined as ( / )S V Bg g e h k T and is related to the current amplitude. The fitting parameter b is defined as Eg + ΔE, which corresponds to the built-in voltage Vbi. The fitting parameter c is a variable constant ( / )D 1 dsc W e eV bξ= + , while the electric field ξ is defined as the built-in voltage divided by the depletion width WD, where

(2 / )( ( / ))((1/ ) (1/ ))D 0 r ds A DW e V b e N Nε ε= + + ; ε0 is the per-mittivity of free space, εr is relative dielectric constant, NA is the acceptor-doping concentration, and ND is the donor-doping concentration. Additionally, since the applied bias, Vds, also alters the tunneling window, ΔE, and the electric field, ξ, the BTBT current equation is therefore modified as

In 12 1 cosh

( )exp

( )ds g

B

ds

dsI a

eV E bk T

c eV beV b( )= + − −

− ++

(3)

small 2016, DOI: 10.1002/smll.201601310

Figure 4. a) Drain current versus top-gate voltages for the MoS2-TFET at Vds = +2 and +3 V. b) Temperature dependence of the BTBT current in a monolayer MoS2-TFET. Drain current as a function of the bias voltages for a top-gate voltage of Vg = +10 V at different temperatures.

communicationswww.MaterialsViews.com

6 www.small-journal.com © 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Accordingly, the measured tunneling current was fit by the above equation with the appropriate fitting parameters a, b, and c. Note that the BTBT current is due to the gate voltage induced p–n or n+–n junctions, so that the experi-mental data used for fitting must be the measured current, in which we properly subtracted the background current. This background current is the measured current at a certain gate voltage in which the current does not vividly increase. This threshold voltage is slightly different for each meas-ured device. In all of the data fitting plots, we indicate this critical point of background subtraction in the ordinate axis. For instance, the threshold voltage for background subtrac-tion in the n+–n junction and the p–n junction in device B21 are 0 and −5 V, respectively, which are labeled in the ordinate axis in Figure 5. Figure 5 shows the measured BTBT current compared with the calculated tunneling current in device B21. The n+–n junction that is induced along the lateral mono layer MoS2 channel by the applied positive top-gate voltages is shown in Figure 5a, whereas the p–n junction that is induced by the applied negative top-gate voltages is shown in Figure 5b. Both cases indicate that the experimental data are well matched with the fitting model and we present sim-ilar results for another two devices in the Supporting Infor-mation (Figure S1, Supporting Information).

The fitting parameters for the three devices are summa-rized and listed in Table 1, in which the built-in voltage is directly obtained from the fitting parameter b and the deple-tion width WD is acquired from ( / )D 1 dsW c e eV bξ= + . It is noted that the average built-in voltage is larger than 1.8 V (corresponding to the bandgap of monolayer MoS2), which is

reasonable because in the presence of BTBT, the minimum energy must be at least around or larger than this value. Fur-thermore, by fitting the experimental data with this model, the depletion widths are found to be about a few nanometers (average value ≈ 3 nm) in all of our monolayer MoS2-TFETs. These extracted values of depletion width corroborate the predicted values (≈2.5 nm) in the theoretical work.[29] Lastly, the average internal electric fields induced by the top-gate electrode are estimated to be around 6 MV cm−1, which is comparable or even higher than the corresponding electric fields induced in Si-based TFETs.[27] After optimizing the device structure design, the magnitudes of the internal elec-tric fields could be further enhanced and thus higher on-state tunneling current densities are possible to achieve which would benefit the practical applications of low-power dis-sipation MoS2-TFETs. Here we also provide Table 2 which summarizes the experimental on-state tunneling currents in various TFETs found in the literature.

In summary, we have demonstrated experiments and numerical simulation to confirm the BTBT in a novel vertical tunneling field-effect transistor with a monolayer MoS2 as the channel (MoS2-TFET). The application of either positive or negative local top-gate voltages leads to the formation of in-plane/lateral n+–n or p–n junctions, respectively. Increasing the gate voltages results in a strong gate-induced doping of the monolayer MoS2 channel as well as the presence of large internal electric fields. Accordingly, this results in the direct manifestation of a BTBT current and an ambipolar trans-port in our MoS2-TFETs. The merit of these MoS2-TFETs lies in their suitability for the development of low-power

small 2016, DOI: 10.1002/smll.201601310

Figure 5. Experimental and simulation results of the BTBT current in a monolayer MoS2-TFET. a) For the case of the n+–n junction that is formed along the lateral monolayer MoS2 with the application of positive top-gate voltages. b) For the case of the p–n junction that is formed along the lateral monolayer MoS2 with the application of negative top-gate voltages.

Table 1. The summary of the fitting parameters for the three measured MoS2-TFETs.

Device Gate voltage [V]

a [pA]

b [eV]

c [(eV)1/2]

Built-in voltage [V] (at Vds = 0)

Depletion width, [nm] (at Vds = 0)

B21 11 10.9 ± 0.04 1.77 ± 0.03 1.85 ± 0.04 1.77 ± 0.03 2.41 ± 0.08

−9 12.6 ± 0.01 1.97 ± 0.01 2.84 ± 0.01 1.97 ± 0.01 3.91 ± 0.02

B43 6 69.9 ± 1.33 1.81 ± 0.03 4.29 ± 0.05 1.81 ± 0.03 5.66 ± 0.10

−10 51.0 ± 0.98 1.81 ± 0.03 4.46 ± 0.04 1.81 ± 0.03 5.88 ± 0.10

A25 10 5.54 ± 0.06 2.31 ± 0.03 0.64 ± 0.02 2.31 ± 0.03 0.95 ± 0.04

−10 1.16 ± 0.03 2.31 ± 0.05 0.65 ± 0.05 2.31 ± 0.05 0.97 ± 0.08

www.MaterialsViews.com

7© 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.small-journal.com

electronics. Furthermore, this work constitutes the first experimental demonstration of utilizing a monolayer TMD material as the channel for the study of BTBT in lateral p–n and n+–n junctions, in which these diode junctions are elec-trostatically induced by a local top-gate electrode. The MoS2 was deposited using chemical vapor deposition, a potentially large-scale and practical fabrication technique. The concept and understanding presented here is easily transferable to a myriad of van der Waals materials, TMDs in particular, with Eg in the range of 1–2 eV, thereby providing us with an extra degree of freedom in tuning the threshold voltage and sub-threshold leakage current for various potential applications. While this work clearly demonstrates the feasibility of cur-rent switching via BTBT in a monolayer of MoS2, simulta-neous improvements in the tunnel barrier profiles as well as the contact between van der Waals materials and metal electrodes can further increase the current densities. This will enable low-power electronic applications through their direct integration with commercial silicon-based CMOS technology for achieving augmented functionalities.

Experimental Section

In this work, we commenced the fabrication process with a 100 mm degenerately doped n++ (ND ≈ 1 × 1019 cm−3) silicon wafer and performed a standard local oxidation of silicon proce-dure in order to define arrays of active areas for the MoS2-TFETs, which were isolated from each other by a 300 nm thick SiO2 field oxide. Afterward, we transferred large-area CVD monolayer MoS2 on top of the substrate (e.g., SiO2 field oxide) so that it covered several arrays of active areas of the MoS2-TFETs. The single crys-talline MoS2 within one triangular domain was grown using a CVD method.[37,38] These triangles assemble into the large-area mon-olayer MoS2 sheet and were transferred onto the substrate using a poly-methyl methacrylate (PMMA) transfer method. Subsequently, a photolithography step was performed to mask circular regions of the MoS2 covering the active areas. The MoS2 outside of the active regions was etched in order to isolate the various devices and its monolayer nature was confirmed by Raman measurement (see Figure S2, Supporting Information). More sample characterization results such as atomic force microscope (AFM) height measure-ment, photo-luminescence (PL) mapping, and Raman mapping are included in the Supporting Information (Figures S3 and S4, Sup-porting Information). Since the MoS2 shows a strong PL emission,

we believe our MoS2 samples are lightly doped n-type semicon-ductors. The MoS2 area of each device is about 8 × 104 μm2. A second photolithography step was performed in order to pattern and deposit the base contacts (20 nm thick Ti/100 nm thick Au) on the MoS2. A 1 nm thick Ti seed layer was evaporated on top of MoS2 and naturally oxidized in air, followed by atomic layer dep-osition of a ≈55 nm thick HfO2 as the dielectric material. A third photolithography step was performed in order to define a central circular top-gate region which encompasses the entire active area of the MoS2-TFETs. We then radio-frequency (RF) sputtered ≈45 nm of ITO at room temperature into this circular region followed by lift-off. Finally, a fourth photolithography step was performed in order to pattern and deposit the side electrodes (150 nm thick Al/50 nm thick Au) on top of the HfO2. These metallic side electrodes inti-mately contact the central ITO top-gate region and allow for easy probing and biasing of the MoS2-TFETs.

Supporting Information

Supporting Information is available from the Wiley Online Library or from the author.

Acknowledgements

Y.-W.L. and C.M.T. contributed equally to this work. This work was in part supported by the National Science Foundation (NSF) under Award No. NSF-EFRI-1433541. C.M.T.Jr. thanks the Department of Defense SMART (Science, Mathematics, and Research for Trans-formation) Scholarship for graduate scholarship funding. This research was funded in part by the National Science Council of Taiwan under Contract No. NSC 103-2917-I-564-017. The authors would like to acknowledge the collaboration of this research with King Abdul-Aziz City for Science and Technology (KACST) via The Center of Excellence for Green Nanotechnologies (CEGN). K.L.W., Y.-W.L., and C.M.T.Jr. conceived the idea and designed the experi-ments; Y.-W.L. and C.M.T.Jr. performed the electrical measure-ments; L.-J.L., M.Y.L., and Y.S. synthesized and contributed the materials; C.M.T.Jr. fabricated the devices; Y.-W.L., X. Z., W.-K.Y. and S.-H.T. analyzed the data. All of the authors discussed the results and wrote the paper together.

[1] A. M. Ionescu, H. Riel, Nature 2011, 479, 329.[2] A. C. Seabaugh, Q. Zhang, Proc. IEEE 2010, 98, 2095.[3] Q. Zhang, A. C. Seabaugh, IEEE Electron Device Lett. 2006, 27,

297.[4] J. Appenzeller, Y.-M. Lin, J. Knoch, Ph. Avouris, Phys. Rev. Lett.

2004, 93, 196805.[5] K. K. Bhuwalka, S. Sedlmaier, A. K. Ludsteck, C. Tolksdorf,

J. Schulze, E. Ignaz, IEEE Trans. Electron Devices 2004, 51, 279.[6] B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, A. Kis, Nat.

Nanotechnol. 2011, 6, 147.

small 2016, DOI: 10.1002/smll.201601310

Table 2. The summary of the measured on-state tunneling currents in various TFETs.

Reference Material Type Ids/W [μA μm−1]

[30] Si/Ge N 0.4

[31] Si N 0.01

[32] Si NW N 0.0002

[33] Si P 1.2

[34] GeSn P 26

[35] InAs/Si NW P 2.4

[36] Si NW P 0.00035

This work Monolayer MoS2 N 0.00004

communicationswww.MaterialsViews.com

8 www.small-journal.com © 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

[7] L. Britnell, R. V. Gorbachev, R. Jalil, B. D. Belle, F. Schedin, A. Mishchenko, T. Georgiou, M. I. Katsnelson, L. Eaves, S. C. Morozov, N. M. R. Peres, J. Leist, A. K. Geim, K. S. Novoselov, L. A. Ponomarenko, Science 2012, 335, 947.

[8] C.-H. Lee, G.-H. Lee, A. M. van der Zande, W. Chen, Y. Li, M. Han, X. Cui, G. Arefe, C. Nuckolls, T. F. Heinz, J. Guo, J. Hone, P. Kim, Nat. Nanotechnol. 2014, 9, 676.

[9] A. K. Geim, I. V. Grigorieva, Nature 2013, 499, 419.[10] S. Najmaei, Z. Liu, W. Zhou, X. L. Zou, G. Shi, S. D. Lei,

B. I. Yakobson, J. C. Idrobo, P. M Ajayan, J. Lou, Nat. Mater. 2013, 12, 754.

[11] H. Li, X. Duan, X. Wu, X. Zhuang, H. Zhou, Q. Zhang, X. Zhu, W. Hu, P. Ren, P. Guo, L. Ma, X. Fan, X. Wang, J. Xu, A. Pan, X. Duan, J. Am. Chem. Soc. 2014, 136, 3756.

[12] H. Li, Q. Zhang, X. Duan, X. Wu, X. Fan, X. Zhu, X. Zhuang, W. Hu, H. Zhou, A. Pan, X. Duan. J. Am. Chem. Soc. 2015, 137, 5284.

[13] X. Duan, C. Wang, J. C. Shaw, R. Cheng, Y. Chen, H. Li, C. Wu, Y. Tang, Q. Zhang, A. Pan, J. Jiang, R. Yu, Y. Huang, X. Duan, Nat. Nanotechnol. 2014, 9, 1024.

[14] Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, M. S. Strano, Nat. Nanotechnol. 2012, 7, 699.

[15] Y. J. Zhang, T. Oka, R. Suzuki, J. T. Ye, Y. Iwasa, Science 2014, 344, 725.

[16] L. N. Nguyen, Y. W. Lan, J. H. Chen, T. R. Chang, Y. L. Zhong, H. T. Jeng, L. J. Li, C. D. Chen, Nano Lett. 2014, 14, 2381.

[17] M. Buscema, D. J. Groenendijk, G. A. Steele, H. S. J. van der Zant, A. Castellanos-Gomez, Nat. Commun. 2014, 5, 4651.

[18] G. Fiori, F. Bonaccorso, G. Lannaccone, T. Palacios, D. Neumaier, A. Seabaugh, S. K. Banerjee, L. Colombo, Nat. Nanotechnol. 2014, 9, 768.

[19] D. Jena, Proc. IEEE 2013, 101, 1585.[20] X.-W. Jiang, J. Gong, N. Xu, S.-S. Li, J. Zhang, Y. Hao, L.-W. Wang,

Appl. Phys. Lett. 2014, 104, 023512.[21] R. K. Ghosh, S. Mahapatra, IEEE Trans. Electron Devices 2012, 60,

274.[22] R. K. Ghosh, S. Mahapatra, IEEE J. Electron Devices Soc. 2013, 1, 175.[23] M. Li, D. Esseni, J. J. Nahas, D. Jena, H. G. Xing, IEEE Trans. Elec-

tron Devices 2015, 3, 200.

[24] T. Roy, M. Tosun, X. Cao, H. Fang, D.-H. Lien, P. Zhao, Y.-Z. Chen, Y.-L. Chueh, J. Guo, A. Javey, ACS Nano 2015, 9, 2071.

[25] S. Das, A. Prakash, R. Salazar, J. Appenzeller, ACS Nano 2014, 8, 1681.

[26] D. Sarkar, X. Xie, W. Liu, W. Cao, J. Kang, Y. Gong, S. Kraemer, P. M. Ajayan, K. Banerjee, Nature 2015, 526, 91.

[27] H. Lu, A. Seabaugh, IEEE J. Electron Devices Soc. 2014, 2, 44.[28] G. Alymov, V. Vyurkov, V. Rhyzii, D. Svintsov, Sci. Rep. 2016, 6,

24654.[29] H Ilatikhameneh, Y. Tan, B Novakovic, G. Klimeck, R. Rahman,

J. Appenzeller, 2015, arXiv:1502.01760.[30] S. H. Kim, H. Kam, C. Hu, T.-J. K. Liu, in Proc. VLSI Symp. Tech. Dig.

Honolulu, HI, USA, 2009, pp. 178–179.[31] Q. Huang, R. Huang, Z. Zhan, Y. Qiu, in Proc. IEDM Tech. Dig.San

Francisco, CA, USA, 2012, pp. 8.5.1–8.5.4.[32] R. Gandhi, Z. Chen, N. Singh, K. Banerjee, S. Lee, IEEE Electron

Device Lett. 2011, 32, 437.[33] K. Jeon, W.-Y. Loh, P. Patel, C. Y. Kang, J. Oh, A. Bowonder, C. Park,

C. S. Park, C. Smith, P. Majhi, H.-H. Tseng, R. Jammy, T.-J. K. Liu, C. Hu, in Proc. VLSI Symp. Tech. Dig. Honolulu, HI, USA, 2010, p. 121.

[34] Y. Yang, S. Su, P. Guo, W. Wang, X. Gong, L. Wang, K. L. Low, G. Zhang, C. Xue, B. Cheng, G. Han, Y.-C. Yeo, in Proc. IEDM Tech. Dig. San Francisco, CA, USA, 2012, pp. 16.3.1–16.3.4.

[35] H. Riel, K. E. Moselund, C. Bessire, M. T. Bjork, A. Schenk, H. Ghoneim, H. Schmid, in Proc. IEDM Tech. Dig. San Francisco, CA, USA, 2012, pp. 16.6.1–16.6.4.

[36] R. Gandhi, Z. Chen, N. Singh, K. Banerjee, S. Lee, IEEE Electron Device Lett. 2011, 32, 1504.

[37] Y.-H. Lee, L. Yu, H. Wang, W. Fang, X. Ling, Y. Shi, C.-T. Lin, J.-K. Huang, M.-T. Chang, C.-S. Chang, M. Dresselhaus, T. Palacios, L.-J. Li, J. Kong, Nano Lett. 2013, 13, 1852.

[38] W.-T. Hsu, Z.-A. Zhao, L.-J. Li, C.-C. Chen, M.-H. Chiu, P.-S. Chang, Y.-C. Chou, W.-H. Chang, ACS Nano 2014, 83, 2951.

Received: April 16, 2016Revised: July 20, 2016Published online:

small 2016, DOI: 10.1002/smll.201601310