Nanoscale View Article Online

PAPER

Cite this: Nanoscale, 2014, 6, 433

View Journal | View Issue

Low-frequency noise in multilayer MoS2 fieldeffect transistors: the effect of high-k passivation†

Published on 09 October 2013. Downloaded on 24/10/2014 22:14:30.

Junhong Na,a Min-Kyu Joo,ab Minju Shin,ab Junghwan Huh,c Jae-Sung Kim,a Mingxing Piao,a Jun-Eon Jin,a Ho-Kyun Jang,a Hyung Jong Choi,d Joon Hyung Shimd and Gyu-Tae Kim*a Diagnosing of the interface quality and the interactions between insulators and semiconductors is significant to achieve the high performance of nanodevices. Herein, low-frequency noise (LFN) in mechanically exfoliated multilayer molybdenum disulfide (MoS2) (11.3 nm-thick) field-effect transistors with back-gate control was characterized with and without an Al2O3 high-k passivation layer. The carrier number fluctuation (CNF) model associated with trapping/detrapping the charge carriers at the interface nicely described the noise behavior in the strong accumulation regime both with and without the Al2O3 passivation layer. The interface trap density at the MoS2–SiO2 interface was extracted from the LFN analysis, and estimated to be Nit  1010 eV1 cm2 without and with the passivation layer. This suggested that the accumulation channel induced by the back-gate was not significantly influenced by the passivation layer. The Hooge mobility fluctuation (HMF) model implying the bulk conduction was found to describe the drain current fluctuations in the subthreshold regime, which is rarely observed in other nanodevices, attributed to those extremely thin channel sizes. In the case of the thick-MoS2 (40 nmthick) without the passivation, the HMF model was clearly observed all over the operation regime, Received 9th August 2013 Accepted 27th September 2013

ensuring the existence of the bulk conduction in multilayer MoS2. With the Al2O3 passivation layer, the

DOI: 10.1039/c3nr04218a

change in the noise behavior was explained from the point of formation of the additional top channel in the MoS2 because of the fixed charges in the Al2O3. The interface trap density from the additional CNF

www.rsc.org/nanoscale

model was Nit ¼ 1.8  1012 eV1 cm2 at the MoS2–Al2O3 interface.

Introduction Two-dimensional (2-D) nanostructures have attracted considerable interest because of their unique properties and the easiness in fabricating the complex structures compared with three- or one-dimensional nanostructures. Among those 2-D nanostructures, graphene has drawn particular interest in recent years owing to extremely large charge carrier mobility.1 However, because of the absence of a band gap, device operation suitable for logic applications is not straightforward.2–4 Layered transition-metal dichalcogenides (TMDs) such as MoS2 have nite band gaps, and so are more appropriate for logic devices.5

a

School of Electrical Engineering, Korea University, Seoul 136-713, South Korea. E-mail: [email protected]

b

IMEP-LAHC, Grenoble INP-MINATEC, 3 Parvis Louis Neel, 38016 Grenoble, France

c

Department of Electronics and Telecommunications, Norwegian University of Science and Technology, NO-7491 Trondheim, Norway d

School of Mechanical Engineering, Korea University, Seoul 136-713, South Korea

† Electronic supplementary information (ESI) available: The effects of thermal annealing, the change of the off-current, the numerically calculated gate capacitances, the raw LFN data, and specic band diagrams are described in detail here. See DOI: 10.1039/c3nr04218a

This journal is © The Royal Society of Chemistry 2014

An electron mobility of 217 cm2 V1 s1 has been reported in monolayer MoS2 passivated with a high-k dielectric.6 This is a substantial improvement over the value of 3 cm2 V1 s1 that was reported previously.7 The large mobility of monolayer MoS2 may be an overestimate due to the wrong choice of the gate capacitance used in the dual-gate measurement system.8,9 Signicant mobility improvement due to the effect of the high-k dielectrics on the MoS2 may not be as large as the initial report suggested; however, further works are required to explain the effects of the high-k dielectric on the charge carrier mobility, as well as the device characteristics of MoS2 transistors.9 Multilayer MoS2 has a comparable band gap (Eg ¼ 1.2 eV) to that of silicon (where Eg ¼ 1.1 eV), and a lower band gap than monolayer MoS2 (where Eg ¼ 1.8 eV). There is therefore greater potential for multilayer MoS2 applications in wide spectral response photodetectors compared with monolayer MoS2.10 In addition, it has a relatively high electron mobility compared with either amorphous Si, low-temperature polycrystalline Si, or amorphous metal oxide semiconductors in view of thin-lm transistors.10,11 The lm thickness can be controlled, and very thin lms, including monolayers, can be obtained from the topdown processes,12–14 which are more exible for customizing the nanodevices.

Nanoscale, 2014, 6, 433–441 | 433

View Article Online

Published on 09 October 2013. Downloaded on 24/10/2014 22:14:30.

Nanoscale

High-k materials and semiconductors are inseparable in the nanodevice applications, particularly eld-effect transistors (FETs). The high-k materials must be used for the gate dielectrics to achieve the ultimate eld effect and the lower gate leakage current. Due to the xed charges and the interface states from the high-k dielectrics, it is difficult to use the high-k for the gate insulators in the FETs. It is important not only to reduce those aws, but also to diagnose them exactly. However, there are few works for diagnosing the interactions between the high-k and the MoS2.11,15 Therefore, in order to apply the multilayer MoS2 to the nanodevices, the effects of a high-k dielectric to the MoS2 should be understood. The low-frequency noise (LFN) characteristics of FETs have been used to provide information relating to the interface quality between the Si channel and the gate dielectric in Si metal–oxide–semiconductor eld-effect transistors (MOSFETs).16–19 As the channel length of MOSFETs continues to be reduced by the scaling of device dimensions, uctuations in the drain current become increasingly signicant, and so the importance of LFN analysis has also grown. However, very little LFN analysis of MoS2 FETs has been carried out to date. In this article, passivation of MoS2 using high-k dielectrics such as Al2O3 and the effects on the electrical properties of multilayer MoS2 were characterized using LFN measurements.

Experimental Bulk MoS2 crystals (SPI Supplies, USA) were mechanically exfoliated directly onto a p++ Si substrate, which was thermally grown using a 300 nm-thick SiO2 layer. Fig. 1(a) (le image) shows the as-exfoliated multilayer MoS2 on the 300 nm-thick SiO2 substrate. We can estimate the thickness of MoS2 from the optical image.20,21 The channel length (0.28 mm) and width (2.3 mm) were dened using electron beam lithography. A 10 nm-thick Ti adhesion layer followed by a 100 nm-thick Au layer was deposited using electron beam evaporation, and electric contacts were dened using a li-off process with acetone. Fig. 1(a) (right image) shows an optical image aer metallization. In Fig. 1(b), atomic force microscopy (AFM) measurements conrmed that the thickness of the multilayer MoS2 was 11.3 nm. MoS2 is a multilayered structure with an interlayer distance of 0.65 nm, which is consistent with the images shown in Fig. 1(c). We can therefore determine the number of layers in the structure: an 11.3 nm-thick MoS2 lm contains approximately 17 layers. A schematic of the multilayer MoS2 FET is shown in Fig. 1(d) before depositing the Al2O3 passivation layer. The electrical properties of the FET were characterized using a Keithley 4200 semiconductor analyzer. LFN measurements were carried out using our custom LFN characterization system,22 which was congured using a data acquisition (DAQ) module. All direct current (DC) and alternating current (AC) electrical measurements were carried out in ambient air at room temperature. A 30 nm-thick Al2O3 passivation layer was deposited on the MoS2 at 220  C using atomic layer deposition (ALD) (Atomic Classic, CN-1 Corporation). The duration of the growth was ˚ per cycle). 10 hours (300 cycles with a growth rate of 1 A

434 | Nanoscale, 2014, 6, 433–441

Paper

Trimethylaluminum (TMA) (UP Chemical Co., Ltd.) and deionized (DI) water were used as precursors. Ultra-pure N2 (99.999%) was used as a carrier and purging gas. The sequence of the ALD cycle was H2O (1 s)/N2 (60 s)/TMA (1 s)/N2 (60 s). The initial pressure of the growth chamber was 5  102 Torr, and it was maintained at 1 Torr during the deposition. In order to ensure that thermal annealing during the ALD process did not affect the comparison between devices with and without an Al2O3 layer, a thermal annealing process with the same conditions as the ALD growth process (i.e. 220  C, 10 hours, 7  102 Torr, but with no precursors) was applied to those devices without an Al2O3 passivation layer (see the ESI for further details†). From Fig. S1,† the as-fabricated MoS2 transistor showed Schottky-like behavior in its output characteristics; however, following the thermal annealing process, ohmic-like behavior was observed. Even these relatively lowtemperature thermal anneals appear to be sufficient to achieve ohmic-like behavior to MoS2 devices. Typically, a signicant increase in drain current, which was observed here following the thermal anneal, is attributed to a decrease in the contact resistance. There may be a number of reasons for this. First, remnant polymer residues, which may come from the scotch tape or the electron beam resist between the channel and the contact, may be removed during the thermal annealing process to improve the effective junction area.23 These residues can also act like an interfacial layer between the MoS2 and the metal layer and be removed by the thermal annealing.24 Another possible reason for the increase in the current is that gas molecules such as oxygen and water can be desorbed from the MoS2 surface during the annealing process. Typically, oxygen and water adsorbed onto the surface can capture electrons from the MoS2 and decrease the carrier concentration.25–28 If those molecules are desorbed from the surface, especially near the contact, it can result in a decrease of the depletion layer width in the MoS2–metal Schottky junction, followed by a decrease of the contact resistance. In order to do the precise static and LFN analysis by obtaining the linear I–V shape, therefore, the thermal annealing process should be required.

Results and discussion Fig. 2(a) shows the typical transfer characteristics of the multilayer MoS2 transistors with and without the Al2O3 passivation. We can conrm that the MoS2 FET operates as an n-channel device. Important electrical parameters such as the eld-effect mobility, mFE, the threshold voltage, Vth, and the on/off ratio are shown in Fig. 2(b) with and without the Al2O3 passivation layer. Interestingly, the on/off ratio decreased from  105 to  103 when the Al2O3 passivation layer was included, resulting from an increase in the off-current from 1013 A to 1010 A. In Fig. S2,† all other MoS2 transistors that we measured showed the same results, which is the off-current increase aer the Al2O3 passivation. This can be explained by considering that an additional channel is induced by the Al2O3 passivation (as discussed later in the paper). If an additional channel is induced near the top of the multilayer MoS2 in parallel with the MoS2 channel, it will be difficult to deplete the additional channel by

This journal is © The Royal Society of Chemistry 2014

View Article Online

Published on 09 October 2013. Downloaded on 24/10/2014 22:14:30.

Paper

Nanoscale

Fig. 1 (a) Optical images before and after the formation of source/drain metal contacts on the multilayer MoS2. The scale bar is 5 mm long. (b) AFM images of the multilayer MoS2 with the source/drain metal contacts. The inset shows that the thickness of MoS2 layer was approximately 11.3 nm. The scale bar is 2 mm long. (c) Schematic image of multilayer MoS2. The gap between monolayers is 0.65 nm. (d) Schematic image of a backgated multilayer MoS2 FET. The source/drain contacts are Ti/Au. The back-gate is p++ Si and the gate dielectric is SiO2.

controlling the back-gate voltage because of the geometry of the gate (i.e. there is poor gate-coupling). This off-current increase due to the Al2O3 deposition is not good news for electronic device applications, but we believe that these phenomena can be reduced and removed by staking the top-gate metal on the Al2O3 high-k dielectric.29 Choosing the work function of the gate

metal can naturally deplete the electrons in the channel due to the work function difference.30 Also, by controlling the top-gate bias, the electrons in the additional top channel of MoS2 can be easily modulated due to the enhancement of the gate-coupling. The threshold voltage shied from 11.1 V to 20.5 V by including the Al2O3 passivation. Although most of the gas

Fig. 2 Changes in the DC electrical characteristics of the multilayer MoS2 FETs with and without the Al2O3 passivation. (a) Transfer characteristics without (black line) and with (red line) the Al2O3 passivation. The inset shows a logarithmic plot of the transfer characteristics. (b) On and offcurrent with and without the Al2O3 passivation layer, as well as the on/off ratios. The table in the inset shows the threshold voltages and mFE.

This journal is © The Royal Society of Chemistry 2014

Nanoscale, 2014, 6, 433–441 | 435

View Article Online

Published on 09 October 2013. Downloaded on 24/10/2014 22:14:30.

Nanoscale

molecules on the MoS2 surface – such as oxygen and water – were removed during the thermal annealing process, whatever molecules remain may be adsorbed on the MoS2 surface and can be removed during the ALD process. In addition, the overall effect of doping due to the Al2O3 layer may explain why Vth shied. A slight increase in mFE from 8.4 to 9.9 cm2 V1 s1 was observed with the Al2O3 passivation layer; mFE was 17.9% larger with the high-k dielectric. These charge carrier mobilities were extracted without considering the fringing elds and the dielectric constant of the Al2O3. When these factors are taken into consideration, the capacitance was 19.2% larger for the device with the Al2O3 passivation layer (see the ESI for details of the numerical calculation of the gate capacitance†). This variation between the gate capacitance values can be compensated between the mFE values for devices without and with the Al2O3 passivation. The effects of the high-k dielectric and the thermal annealing should be separated, especially when extracting mFE. The eld-effect mobility includes the effects of the contact resistance between the metal contact layer and the semiconductor channel. The suppression of Coulomb scattering in the high-k environment is one possible mechanism for an increase of mFE of MoS2,6,9,15,31 however, it is important to consider that the decrease of the contact resistance following the ALD process may occur.8,9 We have already established that the thermal annealing effects of the ALD result in a reduction in the contact resistance. An anomalous effect of the high-k dielectric on mFE was not observed in our devices. High mobility of MoS2 FETs, however, is absolutely necessary to the device applications. Using a low work function metal such as scandium for the contacts could raise the mobility value,32 and decreasing of the substrate interactions might be a way of getting the higher mobility.33 The selection of the proper number of layers might be another way to improve the mobility value.34 Fig. 3(a) and (b) show the normalized drain current spectral density (SI/Ids2) as a function of frequency (f) at different gate voltages with and without the Al2O3 passivation, respectively. In both cases, a 1/f dependence of the noise was observed. The normalized drain current spectral densities decreased as the gate voltages increased, which can be explained by considering that the noise sources were almost constant; however, as Vgs increased, the drain current also increased, resulting in a larger signal-to-noise ratio. It appears that these multilayer MoS2 transistors follow the conventional 1/f noise characteristics that have been reported for numerous nanoscale devices, including graphene,35–37 carbon nanotubes,38,39 ZnO nanowires,40,41 SnO2 nanowires,42–44 Bi2Se3 thin lms.45 In order to directly compare the noise characteristics at a given same drain current, we plot the normalized drain current spectral density, SI/Ids2, as a function of the drain current, Ids. These data are shown in Fig. 3(c) both with and without the Al2O3 passivation. At relatively large currents, in the strong accumulation regime, the noise levels were very similar. However, at relatively low current levels, i.e. in the subthreshold regime, the noise level was higher with the Al2O3 passivation layer. It follows that the Al2O3 layer resulted in an additional noise source that could be observed at the low currents.

436 | Nanoscale, 2014, 6, 433–441

Paper

Fig. 3 LFN characteristics of the multilayer MoS2 FETs. The normalized drain current spectral density as a function of the frequency at various gate voltages is shown (a) without and (b) with the Al2O3 passivation layer. The dashed lines in the curves show the 1/f dependence. (c) The mean normalized drain current spectral density as a function of the drain current for frequencies of 10, 12, 14, 16, 18, and 20 Hz, both without (black) and with (red) the Al2O3 passivation layer.

Two models exist to describe the LFN in conventional Si MOSFETs.16 The Hooge mobility uctuation (HMF) model describes uctuations of the carrier mobility due to uctuations of the mean free-path of electrons, which leads to uctuations in the charge carrier mobility. The carrier number uctuation (CNF) model describes uctuations in the charge carrier density due to charges localized at the trap sites near the channel–oxide interface. To investigate which noise model best describes the observed behavior of the multilayer MoS2 FETs, we plotted SI/ Ids2 as a function of the drain current, Ids, without (Fig. 4(a)) and with (Fig. 4(b)) the Al2O3 passivation layer. According to the CNF

This journal is © The Royal Society of Chemistry 2014

View Article Online

Paper

Nanoscale

model, the normalized drain current spectral density can be expressed as:16  2 SI gm ¼ SVfb (1) Ids Ids 2 where gm is the gate transconductance and SV is the at band voltage spectral density, which can be expressed as:

Published on 09 October 2013. Downloaded on 24/10/2014 22:14:30.

SVfb ¼

q2 kB TNit WLCox 2 f

(2)

where q is the electronic charge, kB is the Boltzmann constant, T is the temperature in Kelvin, W and L are the width and the length of the channel, respectively, Cox is the gate capacitance per unit area, and Nit is the interface trap density. The CNF model ts well to the SI/Ids2  Ids data as in Fig. 4(a) and (b) in the strong accumulation regime, both with and without Al2O3 passivation. (The strong accumulation regimes are represented

in the transfer characteristics and in the band diagram in Fig. 4(c) and (d), respectively.) This result is plausible because the accumulation channel in the MoS2 is probably induced near the MoS2–SiO2 interface by modulating the back-gate voltage, regardless of the existence of an Al2O3 passivation layer; therefore, trap sites near the MoS2–SiO2 interface can be expected to be the dominant noise source. Fig. 4(d) shows band diagrams in which a strong accumulation regime occurs when the back-gate voltage is sufficiently large (and positive) to accumulate electrons near the SiO2 gate dielectric, regardless of the presence of the Al2O3 passivation layer. The interface trap density calculated using eqn (1) and (2) was Nit ¼ 7.2  1010 cm2 eV1 without the Al2O3 passivation layer and Nit ¼ 5.5  1010 cm2 eV1 with the Al2O3 passivation layer. Those two values are similar, and so the surface channel near the SiO2 was not strongly affected by the top passivation layer within the strong accumulation regime. In addition, Nit was one order of magnitude larger compared with

Fig. 4 LFN, transfer characteristics, and band diagrams with and without the Al2O3 passivation layer. Normalized drain current spectral density (NSI) from the noise spectra as a function of the drain current (a) without and (b) with the Al2O3 passivation layer. The HMF and CNF models were fitted to the raw data. CNF_top and CNF_bottom refer to the fitted carrier number fluctuation models for the top (near the Al2O3) and the bottom (near the SiO2) surfaces of the channel. The fitting data for the summation of the HMF and CNF models are denoted by “SUM”. (c) Drain current as a function of the gate overdrive voltage, Vov ¼ Vgs  Vth, both with and without the Al2O3 passivation layer. The subthreshold and the strong accumulation regimes are indicated by the gray region. (d) Simplified band diagrams without (left) and with (right) the Al2O3 passivation. The band diagrams at the top, middle, and bottom indicate when the device is as fabricated, with an applied negative gate voltage, and with an applied positive gate voltage, respectively. The conduction and valence band edges of SiO2 and Al2O3 have been omitted for clarity.

This journal is © The Royal Society of Chemistry 2014

Nanoscale, 2014, 6, 433–441 | 437

View Article Online

Published on 09 October 2013. Downloaded on 24/10/2014 22:14:30.

Nanoscale

Paper

Nit  109 cm2 eV1 at the Si–SiO2 interface in Si MOSFETs.17 This one order of difference between the Nit values can be further improved through the optimization of the fabrication process. At small currents, i.e. the subthreshold regime, there were large discrepancies between the CNF model and the measured SI/Ids2 data shown in Fig. 4(a) and (b). The subthreshold regimes are represented as the transfer characteristics in Fig. 4(c), and as band diagrams in Fig. 4(d). Considering that multilayer MoS2 FETs operate as accumulation-mode transistors, since MoS2 is a natural n-type semiconductor, and that the devices are junctionless transistors, i.e. they have the same doping concentrations in the source, drain, and channel regions, we can expect that the threshold voltages should be less than or equal to the at-band voltages.46 The enlarged section of the transfer characteristic curve below the threshold voltage without the Al2O3 passivation layer, shown in Fig. 4(c), is consistent with FETs that operate in accumulation-mode and are junctionless. These kinds of “stretched-out” transfer characteristics are typically observed in accumulation-mode and junctionless transistors, and are regarded as the bulk conduction portions.47,48 However, the threshold voltages extracted from the transfer characteristics in this study were larger than the at band voltages.32 Due to the very small contribution of bulk conduction to the total current, this stretched-out part of the transfer characteristics can be ignored, and the surface conduction channel is important in determining the threshold voltages. If the doping density or the effective channel size of the multilayer MoS2 is small enough, the bulk conduction channel is hard to observe in the transfer characteristics. In this study, the transfer curve shown in Fig. 4(c) suggests a relatively small bulk conduction portion. When the bulk conduction is the dominant current pathway (in the subthreshold regime), there is a possibility that the CNF model may not be appropriate, and that instead, the HMF model may better describe the noise characteristics of the subthreshold regime. According to the HMF model, the normalized drain current spectral density can be expressed as:16 SI qaH ¼ 2 fWLQacc Ids

(3)

where aH is the Hooge parameter and Qacc is the accumulated charge. If the accumulated charge follows the form of Qacc ¼ Cox(Vgs  Vth) and the accumulated charge is approximated to the bulk channel charge, the following relationship holds: SI qaH mVds 1 ¼ $ 2 Ids fL2 Ids

(4)

where m is the charge carrier mobility and Vds is the drain voltage. By tting the HMF model to the measured SI/Ids2 without the Al2O3 passivation layer (Fig. 4(a)), the behavior in the subthreshold regime is well described. This HMF model is not observed even in the Si junctionless transistors that have the high portion of bulk conduction.49 This might be the distinguished properties of layered nanomaterials which will be discussed below. The Hooge parameters extracted from eqn (3) and (4) with and without the Al2O3 passivation were

438 | Nanoscale, 2014, 6, 433–441

almost identical, at aH ¼ 2.5  103. This Hooge parameter is similar to or smaller than previous reports for graphenes, carbon nanotubes, amorphous Si, and metal oxide semiconductors.35,36,39,43,50 In many reports of noise analyses, it is unclear whether the noise characteristics follow the HMF or CNF model. The HMF model has rarely been introduced to explain the noise characteristics, even in Si MOSFETs. Particularly, the HMF model has been used to describe the noise characteristics of p-MOS buriedchannel structures,51 and more recently, Persson et al.52 showed that the HMF and the CNF models could be used to describe the noise properties in the vertical wrap-gated Sn-doped InAs nanowire FETs with the core channel and the surface channel, respectively. In the case of the multilayer MoS2 FET, however, the HMF model is relatively easily observed without the buriedchannel structures or any intentional doping and the gate-allaround structures. Even in the case of the relatively thickmultilayer (40 nm) MoS2 FET, we conrmed that only the HMF model was tted all over the operation regime as in Fig. 5. This implies that the inuence of the traps in the gate insulator to the conduction is very weak. Recently, Das et al.53 suggested the resistor network model to describe the current distribution in multilayer MoS2 FET based on Thomas–Fermi charge screening and interlayer coupling. According to those results, accumulation charges induced by the back gate bias can be distributed deeply along to the multilayer MoS2 due to the high T–F charge screening length (l  7 nm) and the interlayer resistance (Rint) has high value, so the dominant current ow may not occur though the lowest layer of the multilayer MoS2. Once the dominant conduction is far from the gate insulator, the probability arising from the carrier trapping/detrapping in the oxide traps in the gate insulator can be reduced, which is already conrmed by other studies that we have referred to above.51,52 This might be a reason why the HMF model is well tted instead of the CNF model in the thick-multilayer (40 nm) MoS2 FETs in the entire regime. In the case of the 11.3 nm thick-multilayer MoS2 FET, the CNF model could be introduced to describe the noise properties in Fig. 4(a) because the thickness of the MoS2 is relatively small, although the dominant current ow does not occur through the lowest layer. If then, there is still a possibility to introduce the CNF model in the strong accumulation regime. Our results can describe Das et al.’s work,53 but more specic theory about the current ow in 2-D nanomaterials should be studied. Also, using the relatively well-described Si accumulation-mode or junctionless transistor theory to describe the electrical properties of MoS2 FETs should be accompanied. In the device with the Al2O3 passivation layer (Fig. 4(b)), the HMF model does not adequately t the SI/Ids2 behavior. In the subthreshold regime, the surface channel near the SiO2 was not the dominant pathway, and instead, bulk conduction and other surface current pathways induced by the Al2O3 passivation were dominant in the noise characteristics. Fig. 4(d) shows the band diagrams with the Al2O3 layer; this indicates the possibility of additional surface channels near the MoS2–Al2O3 interface. By introducing the additional CNF model to t the data in the subthreshold regime, a good t was obtained, as shown in

This journal is © The Royal Society of Chemistry 2014

View Article Online

Published on 09 October 2013. Downloaded on 24/10/2014 22:14:30.

Paper

Fig. 5 (a) Optical images after the formation of source/drain metal contact on the thick-multilayer MoS2. ‘D’ and ‘S’ represent the drain and the source contact, respectively. Other electrodes were intended for the four probe measurements but were not used in the noise measurements. The scale bar is 5 mm long. The inset shows that the thickness of the MoS2 layer was approximately 40 nm by AFM. (b) Normalized drain current spectral density (NSI) from the noise spectra as a function of the drain current without the Al2O3 passivation layer. Only the HMF model (HMF) was fitted to the raw data.

Fig. 4(b). This distinct difference between the devices with and without the Al2O3 passivation layer is consistent with the bump in the curve shown in Fig. 3(c), which was attributed to additional surface channels induced by the Al2O3 passivation layer. Considering the high value of the interlayer resistance (Rint),53 the existence of the additional top surface channel induced by the Al2O3 passivation is very convincing. When both of the charges induced by Al2O3 on the topmost layer of the MoS2 and the Rint are high, the conditions are sufficiently satised so that the dominant current ow can occur through the topmost layer in the deep subthreshold regime of the multilayer MoS2 FET with the Al2O3 passivation. If Cox at the top surface channel caused by the Al2O3 layer can be approximated to that of the surface channel induced by the back-gate, the interface trap density for the Al2O3–MoS2 interface can be extracted as Nit ¼ 1.8  1012 cm2 eV1 from the additional CNF model using eqn (1) and (2). If the Cox is considered to be overestimated owing to the geometrical gate

This journal is © The Royal Society of Chemistry 2014

Nanoscale

Schematic diagrams of the cross-section of the device (a) without and (b) with the Al2O3 passivation layer. In ambient air, gas molecules (mainly oxygen and water) are absorbed on to the MoS2 surface and induce a depletion region. Following the Al2O3 passivation, the gas molecules in ambient air can no longer deplete the multilayer MoS2; however, positive fixed charges in the Al2O3 layer can induce an additional conduction pathway via electrostatic doping at the top of the MoS2 layer. Fig. 6

coupling effect, then Nit will be lower according to the Cox2 dependence as in eqn (2); therefore, we can expect Nit  1011 cm2 eV1. This is consistent with the interface trap density (Dit) ¼ 2.4  1012 cm2 eV1, and Dit ¼ 2.6  1011 cm2 eV1 for the MoS2–Al2O3 interface in top-gate15 and back-gate11 geometries, calculated using the subthreshold swing, respectively. In addition, Dit in high-k materials has been reported to lie in the range 1011 < Dit < 1012 cm2 eV1.18,19,54 If the electrical parameters such as Cox or mobility can be extracted more precisely by using the analytical or numerical solutions, we believe that the noise analyses of the MoS2 FETs can be a more powerful tool to diagnose the interface quality between the high-k dielectric and MoS2.

Nanoscale, 2014, 6, 433–441 | 439

View Article Online

Published on 09 October 2013. Downloaded on 24/10/2014 22:14:30.

Nanoscale

High-k dielectrics, including Al2O3, HfO2, and ZrO2, are known to retain more xed charges than SiO2.18,19,39,54 To ensure stable device operation, the effects of these xed charges on the device characteristics should be studied. Carbon nanotube transistors have been observed to change from p-type to n-type following Al2O3 passivation; this was attributed to xed charges, with the positive polarity in the Al2O3 leading to electrostatic doping.55,56 Shin et al.57 suggested that there may be a nonuniform distribution of xed charges in Al2O3. In InGaAs–Al2O3 structures, negative xed charges may be located near the interface between the Al2O3 and the InGaAs, with positive xed charges in the Al2O3 away from the interface. This suggests that the total electric eld in the MoS2 in our samples may be the sum of the potential due to localized charge distribution near the MoS2–Al2O3 interface and in the bulk part of the Al2O3 layer. Then, the dominant xed charges can be positive in our case, as shown in Fig. 6, because the bulk part of the Al2O3 layer that we deposited is relatively as large as the 30 nm-thick. It is clear that an additional top surface channel in the multilayer MoS2 transistors is induced by the Al2O3 passivation. The increase in the off-current and the noise characteristics is evidence of this. If those phenomena result from xed charges in the Al2O3 layer, the electrostatic doping effect may be positive in the case of the multilayer MoS2 transistors reported here.

Conclusions We have performed LFN characterization of multilayer MoS2 FETs (11.3 nm) with and without an Al2O3 passivation layer. In the strong accumulation regime, the carrier number uctuation (CNF) model associated with trapping/detrapping the charge carriers at the interface was used to describe the noise behavior, both with and without the passivation. The interface trap density at the MoS2–SiO2 interface extracted from the LFN measurements was Nit ¼ 7.2  1010 eV1 cm2 without the Al2O3 passivation layer and Nit ¼ 5.5  1010 eV1 cm2 with the passivation layer. Those two similar Nit values represented that the accumulation channel induced by the back-gate was not signicantly inuenced by the passivation layer. The Hooge mobility uctuation (HMF) model was introduced to describe the electrical properties in terms of bulk conduction in the subthreshold regime. For the thick-MoS2 (40 nm) FET, the HMF model was tted all over the operation regime, ensuring the existence of the bulk conduction in multilayer MoS2. With the addition of the Al2O3 passivation, an additional top surface channel was induced, which led to an increase of the offcurrent. This change is also apparent in the noise behavior so the additional (CNF) model was successfully introduced when the Al2O3 passivation layer was placed. The trap density at the MoS2–Al2O3 interface was Nit ¼ 1.8  1012 eV1 cm2. These results can be a great help to understand the interactions between MoS2 and high-k materials.

Acknowledgements This work was supported by the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT and

440 | Nanoscale, 2014, 6, 433–441

Paper

Future Planning (Converging Research Center Program, 2013K000175 and Global Frontier Research Program, No. 20110031638).

Notes and references 1 A. K. Geim, Science, 2009, 324, 1530–1534. ¨ 2 M. Y. Han, B. Ozyilmaz, Y. Zhang and P. Kim, Phys. Rev. Lett., 2007, 98, 206805. 3 D. S. L. Abergel, V. Apalkov, J. Berashevich, K. Ziegler and T. Chakraborty, Adv. Phys., 2010, 59, 261–482. 4 M. F. Craciun, S. Russo, M. Yamamoto and S. Tarucha, Nano Today, 2011, 6, 42–60. 5 Nat. Nanotechnol., 2012, 7, 683–683. 6 B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti and A. Kis, Nat. Nanotechnol., 2011, 6, 147–150. 7 K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V. Khotkevich, S. V. Morozov and A. K. Geim, Proc. Natl. Acad. Sci. U. S. A., 2005, 102, 10451–10453. 8 M. S. Fuhrer and J. Hone, Nat. Nanotechnol., 2013, 8, 146– 147. 9 B. Radisavljevic and A. Kis, Nat. Nanotechnol., 2013, 8, 147– 148. 10 W. Choi, M. Y. Cho, A. Konar, J. H. Lee, G.-B. Cha, S. C. Hong, S. Kim, J. Kim, D. Jena, J. Joo and S. Kim, Adv. Mater., 2012, 24, 5832–5836. 11 S. Kim, A. Konar, W. S. Hwang, J. H. Lee, J. Lee, J. Yang, C. Jung, H. Kim, J. B. Yoo, J. Y. Choi, Y. W. Jin, S. Y. Lee, D. Jena, W. Choi and K. Kim, Nat. Commun., 2012, 3, 1011. 12 A. Castellanos-Gomez, M. Barkelid, A. M. Goossens, V. E. Calado, H. S. J. van der Zant and G. A. Steele, Nano Lett., 2012, 12, 3187–3192. 13 Y. Liu, H. Nan, X. Wu, W. Pan, W. Wang, J. Bai, W. Zhao, L. Sun, X. Wang and Z. Ni, ACS Nano, 2013, 7, 4202–4209. 14 X. Lu, M. I. B. Utama, J. Zhang, Y. Zhao and Q. Xiong, Nanoscale, 2013, 5, 8904–8908. 15 L. Han and P. D. Ye, IEEE Electron Device Lett., 2012, 33, 546– 548. 16 G. Ghibaudo, O. Roux, C. Nguyen-Duc, F. Balestra and J. Brini, Phys. Status Solidi A, 1991, 124, 571–581. 17 T. Boutchacha, G. Ghibaudo, G. Gu´ egan and T. Skotnicki, Microelectron. Reliab., 1997, 37, 1599–1602. 18 M. Bigang, S. P. Devireddy, Z. Celik-Butler, W. Fang, A. Zlotnicka, T. Hsing-Huang and P. J. Tobin, IEEE Trans. Electron Devices, 2004, 51, 1679–1687. 19 C. Claeys, E. Simoen, A. Mercha, L. Pantisano and E. Young, J. Electrochem. Soc., 2005, 152, F115–F123. 20 H. Li, G. Lu, Z. Yin, Q. He, H. Li, Q. Zhang and H. Zhang, Small, 2012, 8, 682–686. 21 W. Ying Ying, G. Ren Xi, N. Zhen Hua, H. Hui, G. Shu Peng, Y. Huan Ping, C. Chun Xiao and Y. Ting, Nanotechnology, 2012, 23, 495713. 22 M.-K. Joo, P. Kang, Y. Kim, G.-T. Kim and S. Kim, Rev. Sci. Instrum., 2011, 82, 034702. 23 O. L. Jeong, C. Park, K. Ju-Jin, K. Jinhee, P. Jong Wan and Y. Kyung-Hwa, J. Phys. D: Appl. Phys., 2000, 33, 1953.

This journal is © The Royal Society of Chemistry 2014

View Article Online

Published on 09 October 2013. Downloaded on 24/10/2014 22:14:30.

Paper

24 W. Yunsung, S. D. Georg and R. Siegmar, Nanotechnology, 2007, 18, 095203. 25 D. J. Late, B. Liu, H. S. S. R. Matte, V. P. Dravid and C. N. R. Rao, ACS Nano, 2012, 6, 5635–5641. 26 H. Qiu, L. Pan, Z. Yao, J. Li, Y. Shi and X. Wang, Appl. Phys. Lett., 2012, 100, 123104. 27 F. K. Perkins, A. L. Friedman, E. Cobas, P. M. Campbell, G. G. Jernigan and B. T. Jonker, Nano Lett., 2013, 13, 668– 673. 28 P. Woanseo, P. Juhun, J. Jingon, L. Hyungwoo, J. Hyunhak, C. Kyungjune, H. Seunghun and L. Takhee, Nanotechnology, 2013, 24, 095202. 29 C.-W. Lee, I. Ferain, A. Afzalian, R. Yan, N. D. Akhavan, P. Razavi and J.-P. Colinge, Solid-State Electron., 2010, 54, 97–103. 30 H. Wang, L. Yu, Y.-H. Lee, Y. Shi, A. Hsu, M. L. Chin, L.-J. Li, M. Dubey, J. Kong and T. Palacios, Nano Lett., 2012, 12, 4674–4680. 31 D. Jena and A. Konar, Phys. Rev. Lett., 2007, 98, 136805. 32 S. Das, H.-Y. Chen, A. V. Penumatcha and J. Appenzeller, Nano Lett., 2012, 13, 100–105. 33 K. Nagashio, T. Nishimura, K. Kita and A. Toriumi, Appl. Phys. Express, 2009, 2, 025003. 34 S.-L. Li, K. Wakabayashi, Y. Xu, S. Nakaharai, K. Komatsu, W.-W. Li, Y.-F. Lin, A. Aparecido-Ferreira and K. Tsukagoshi, Nano Lett., 2013, 13, 3546–3552. 35 G. Liu, W. Stillman, S. Rumyantsev, Q. Shao, M. Shur and A. A. Balandin, Appl. Phys. Lett., 2009, 95, 033103. 36 A. N. Pal and A. Ghosh, Appl. Phys. Lett., 2009, 95, 082105. 37 S. Qinghui, L. Guanxiong, D. Teweldebrhan, A. A. Balandin, S. Rumyantsev, M. S. Shur and Y. Dong, IEEE Electron Device Lett., 2009, 30, 288–290. 38 Y.-M. Lin, J. Appenzeller, J. Knoch, Z. Chen and P. Avouris, Nano Lett., 2006, 6, 930–936. 39 U. J. Kim, K. H. Kim, K. T. Kim, Y.-S. Min and W. Park, Nanotechnology, 2008, 19, 285705. 40 W. Wang, H. D. Xiong, M. D. Edelstein, D. Gundlach, J. S. Suehle, C. A. Richter, W.-K. Hong and T. Lee, J. Appl. Phys., 2007, 101, 044313.

This journal is © The Royal Society of Chemistry 2014

Nanoscale

41 J. Lee, B.-Y. Yu, C. H. Lee, G.-C. Yi, S. H. Son, G.-T. Kim and G. Ghibaudo, Phys. E., 2008, 40, 2147–2149. 42 S. Ju, P. Chen, C. Zhou, Y.-g. Ha, A. Facchetti, T. J. Marks, S. K. Kim, S. Mohammadi and D. B. Janes, Appl. Phys. Lett., 2008, 92, 243120. 43 J. Huh, M.-K. Joo, D. Jang, J.-H. Lee and G. T. Kim, J. Mater. Chem., 2012, 22, 24012–24016. 44 M.-K. Joo, J. Huh, M. Mouis, S. J. Park, D.-Y. Jeon, D. Jang, J.-H. Lee, G.-T. Kim and G. Ghibaudo, Appl. Phys. Lett., 2013, 102, 053114. 45 M. Z. Hossain, S. L. Rumyantsev, K. M. F. Shahil, D. Teweldebrhan, M. Shur and A. A. Balandin, ACS Nano, 2011, 5, 2657–2663. 46 J. P. Colinge, A. Kranti, R. Yan, C. W. Lee, I. Ferain, R. Yu, N. Dehdashti Akhavan and P. Razavi, Solid-State Electron., 2011, 65–66, 33–37. 47 J. P. Colinge, IEEE Trans. Electron Devices, 1990, 37, 718– 723. 48 D.-Y. Jeon, S. J. Park, M. Mouis, S. Barraud, G.-T. Kim and G. Ghibaudo, Solid-State Electron., 2013, 80, 135– 141. 49 D.-Y. Jeon, S. J. Park, M. Mouis, S. Barraud, G.-T. Kim and G. Ghibaudo, Solid-State Electron., 2013, 81, 101–104. 50 S. Liang-Yu, L. Huang-Kai, H. Chia-Chin and H. JianJang, J. Disp. Technol., 2012, 8, 695–698. 51 G. Ghibaudo and J. Chroboczek, Solid-State Electron., 2002, 46, 393–398. 52 K.-M. Persson, B. G. Malm and L.-E. Wernersson, Appl. Phys. Lett., 2013, 103, 033508. 53 S. Das and J. Appenzeller, Nano Lett., 2013, 13, 3396–3402. 54 G.-W. Lee, J.-H. Lee, H.-W. Lee, M.-K. Park, D.-G. Kang and H.-K. Youn, Appl. Phys. Lett., 2002, 81, 2050–2052. 55 N. Moriyama, Y. Ohno, T. Kitamura, S. Kishimoto and T. Mizutani, Nanotechnology, 2010, 21, 165201. 56 U. J. Kim, H. B. Son, E. H. Lee, J. M. Kim, S. C. Min and W. Park, Appl. Phys. Lett., 2010, 97, 032117. 57 B. Shin, J. R. Weber, R. D. Long, P. K. Hurley, C. G. V. d. Walle and P. C. McIntyre, Appl. Phys. Lett., 2010, 96, 152908.

Nanoscale, 2014, 6, 433–441 | 441

Low-frequency noise in multilayer MoS2 field-effect transistors: the effect of high-k passivation.

Diagnosing of the interface quality and the interactions between insulators and semiconductors is significant to achieve the high performance of nanod...
983KB Sizes 0 Downloads 0 Views