Abstract
Graphene, with superior electrical tunabilities, has arisen as a multifunctional insertion layer in vertically stacked devices. Although the role of graphene inserted in metal-semiconductor junctions has been well investigated in quasi-static charge transport regime, the implication of graphene insertion at ultrahigh frequencies has rarely been considered. Here, we demonstrate the diode operation of vertical Pt/n-MoSe2/graphene/Au assemblies at ~200-GHz cutoff frequency (fC). The electric charge modulation by the inserted graphene becomes essentially frozen above a few GHz frequencies due to graphene quantum capacitance–induced delay, so that the Ohmic graphene/MoSe2 junction may be transformed to a pinning-free Schottky junction. Our diodes exhibit much lower total capacitance than devices without graphene insertion, deriving an order of magnitude higher fC, which clearly demonstrates the merit of graphene at high frequencies.
Graphene insertion paves the way to effortlessly achieve 6G sub-THz communication devices.
INTRODUCTION
Two-dimensional (2D) van der Waals (vdW) semiconductors have been extensively studied as active components for advanced electronic devices (1–6). The high charge carrier mobilities and associated physical properties along the in-plane direction have been the main subjects of 2D crystals. In this regard, one of the main challenges in 2D crystal-based electronic device applications has been the high contact resistance (RC) (7, 8). Previous reports have shown various ways to mitigate this issue, which includes identification of new Ohmic contact materials (7, 9), development of new strategy to deposit metals on 2D semiconductor (10, 11), and the adaptation of a vertical device geometry with larger contact area (12). Owing to the recent advancement of RC engineering, a few high-frequency devices based on 2D semiconductors have been recently reported (11–15). Still, high cutoff frequency over 100 GHz seems not possible to achieve from RC improvement alone.
Graphene, the most extensively studied 2D crystal with large electrical tunability and chemical stability (16), can be inserted as a multifunctional layer in various vdW heterostructures (17–19). For example, inserted graphene has been previously used as a layer for barristor device (17). Moreover, the in-plane charge transport behavior of graphene itself has also been studied well at DC and high frequencies (20–23). However, vertical charge transport across inserted graphene and its implication for device performance at high frequencies have been largely neglected. In particular, considering the capacitance is one of the key parameters for high-frequency devices (11), the role of graphene insertion regarding device capacitance should be properly addressed to understand the fundamental limitation on the high-frequency performance of vdW 2D devices.
Here, we demonstrate the operation of vertical Schottky diodes with high cutoff frequencies (fC) of ~200 GHz at maximum via graphene insertion, which covers millimeter wave (mm-Wave) range for signal reception and energy harvesting (24, 25). We fabricate Pt/n-MoSe2 vertical Schottky diodes with various Ohmic contacts and elucidate the role of graphene insertion regarding the modification of device capacitance at high frequency. We find that the electrical charge modulation by graphene is essentially frozen above a few gigahertz frequencies due to the delay associated with the quantum capacitance of graphene (CGR), while charge modulation needs a minimum time longer than the delay. Ohmic Au/graphene/n-MoSe2 junction transforms to an ideal pinning-free Schottky junction above a few gigahertz frequencies, which leads to lower total capacitance and order-of-magnitude higher cutoff frequencies than devices without graphene insertion. In addition, 28-GHz mixer applications are demonstrated by using two graphene-inserted Schottky diodes, which clearly displays the merit of graphene-insertion at high frequencies. Our work elucidates the role of graphene insertion at high frequencies and paves the way for improving the high-frequency performance of vertically stacked vdW devices.
RESULTS
Vertical Schottky diodes incorporating layered 2D semiconductor with various kinds of electrodes were fabricated. Figure 1A displays a 3D schematic of the vertical Schottky diode with Pt/n-MoSe2/graphene/Au. Top Pt electrode with relatively thick (~160 nm) 2H-MoSe2 flake serves as almost an ideal metal-semiconductor Schottky junction. Raman spectroscopy and atomic force microscopy (AFM) results for MoSe2 flakes are shown in figs. S1A and S2, respectively. Au (with and without graphene insertion) was chosen as Ohmic contact as the other side (bottom-contact) of contact, so that Pt/n-MoSe2/graphene/Au may form a vertical diode. Similar Ohmic contacts were observed with different graphene thickness for insertion, from monolayer (1L) to thick graphite or multilayer graphene (ML). Fabrication details are described in Methods. We also note that Ti contact (of Au/Ti bilayer form) can support better Ohmic behavior than Au, but here, Au as a bottom electrode was mainly discussed because of the suitable comparison between control groups and its superior air stability to Ti. Figure S3 shows the effects of source/drain top contact metals (Pt, Ti, and Au) on MoSe2 channel field effect transistor and also displays the effects of top and bottom Au on a vertical MoSe2 diode, where bottom Au contact clearly shows quiet an Ohmic behavior.
Fig. 1. Structure and DC properties of vertical Schottky diodes.
(A) Schematic of the Pt/n-MoSe2/graphene/Au vertical Schottky diode. (B) OM image and 2D schematic (inset) of Pt/n-MoSe2/graphene/Au diode. (C) OM image and 2D schematic (inset) of Pt/n-MoSe2/Au diode. (D) TEM images of Au/monolayer graphene/MoSe2, Au/4 ~ five-layer graphene/n-MoSe2, Au/graphite/n-MoSe2, and Pt/n-MoSe2 interfaces (from left to right). Pt/n-MoSe2 image is obtained by scanning TEM mode. (E and F) DC I-V characteristics of Pt/n-MoSe2/graphene/Au and Pt/n-MoSe2/Au diodes. (G) Current responsivity and nonlinearity (inset) of Pt/n-MoSe2/graphene/Au diode. (H) 1/C2-V curve with calculated carrier concentration and built-in potential.
Schottky diodes with graphene insertion of various thickness (and without graphene insertion) were fabricated in a coplanar waveguide (CPW) device pattern for radiofrequency (RF) probing. Figure 1B (with graphene) and Fig. 1C (without graphene) show optical microscopy (OM) images of the diodes in a CPW pattern. The yellow dashed line in Fig. 1B traces the patterned chemical vapor deposited (CVD) graphene on the bottom Au electrode, and the vertical contact area is 10 × 10 μm2 as patterned (red dashed line). Figure 1D shows cross-sectional high-resolution transmission electron microscopy (HRTEM) images of the MoSe2/graphene/Au Ohmic junction with various graphene thickness, including the image of Pt/MoSe2 junction. HRTEM images confirm the thickness of graphene insertion and show the clean interfaces without noticeable process-related residues at the junction.
The DC current-voltage (I-V) characteristics of the vertical diodes with (and without) graphene insertion are shown in Fig. 1E (and Fig. 1F). The ideality factors from each semilog-scale I-V plot were estimated to be 1.15 (n-MoSe2/graphene/Au) and 1.20 (n-MoSe2/Au). (The extraction and fitting of ideality factors are introduced in Supplementary Text 1). Low-temperature I-V characteristics were also measured in the range of 80 to 300 K, as depicted in fig. S4, where a Schottky barrier of ~0.17 eV at the Pt-MoSe2 junction was consistently measured. The Schottky barrier height (qΦB,Pt) is quite small, and we attribute it to the existence of Fermi pinning effects between the Pt contact and n-MoSe2 (26) as the origin of such a small qΦB.
The responsivity and the nonlinearity [i.e., the ratio of the rectified DC current and the RF power delivered to nonlinear devices and a measure of the deviation from a linear device, respectively (27)] are the important figures of merits (FOMs) in RF rectifier applications as estimated from the I-V characteristics. Figure 1G displays the responsivity and the nonlinearity (inset) plots of the graphene/Au-contact device, where their peak values are calculated as 22.7 A/W and 5.29, respectively. These numbers are regarded as high compared to those of reported Schottky devices (11, 28). These values were also extracted for Au-contact device, as plotted in fig. S5. In general, a diode rectifier should satisfy the conditions of certain responsivity (> 7 A/W) and nonlinearity (> 3) for RF applications (27), which means that our Schottky devices have strong potential for RF rectifiers. For more information on the Schottky junction characteristics of the Pt/n-MoSe2/Au diode, a Mott-Schottky plot was (29) obtained from the capacitance-voltage (C-V) measurement. As shown in the 1/C2-V plot of Fig. 1H, we extracted the built-in potential (ϕi,Pt = 0.146 V) at the Pt/n-MoSe2 junction and carrier concentration of MoSe2 (n = 7.13 × 1016 cm−3).
With the promising DC diode performance, high-frequency performance including fC measurements were conducted using a network analyzer. Figure 2 (A and B, respectively) shows the circuit diagram and CPW device scheme with the ground–signal–ground (G-S-G) configuration (30) for the one-port S-parameter measurement. A small-signal equivalent circuit of the Schottky diode is included as the device under test (red box) in the diagram of Fig. 2A. One-port scattering parameters (S11) for each diode were achieved from 50 MHz to 40 GHz using a network analyzer and subsidiary circuitry. S11 in the one-port network equals the voltage reflection coefficient, Γ = (ZL – Z0)/(ZL + Z0) (30), where ZL and Z0 are the impedances of each diode and the transmission line (= 50 ohm in this case), respectively. Thus, the impedance of each device can be extracted from its S11 parameter chart, the Smith chart in Fig. 2C. Figure 2D shows the series resistance (real part, solid) and capacitive reactance (imaginary part, dot) in log-scale impedance plots, as extracted.
Fig. 2. S-parameter and RF characteristics of vertical Schottky diodes.
(A) Circuit diagram of the test setup for S-parameter measurement including the small-signal equivalent circuit of the Schottky diode. Cp indicates the parasitic capacitance. (B) 3D schematic of the G-S-G patterned device. (C) The Smith chart on S11 parameters of Pt/n-MoSe2/1L graphene/Au (red), Pt/n-MoSe2/3L graphene/Au (green), and Pt/n-MoSe2/Au (blue) diodes. (D) Impedance-frequency plot of of Pt/n-MoSe2/1L graphene/Au (red), Pt/n-MoSe2/3L graphene/Au (green), and Pt/n-MoSe2/Au (blue) diodes with extracted cutoff frequencies (inset). (E) Impedance-frequency plot of Pt/60-nm n-MoSe2/1L graphene/Au diode as compared with Pt/160-nm n-MoSe2/1L graphene/Au diode. Inset shows the cutoff frequencies extracted from the devices with 60-, 128-, 160-nm-thick MoSe2. The bars in the inset graphs for 1L Gr.-inserted device with 160-nm MoSe2 indicates all three measured fC (110, 145, and 220 GHz).
The impedances for three Schottky diodes of Pt/MoSe2/1L graphene/Au (red), Pt/MoSe2/3L graphene/Au (green), and Pt/MoSe2/Au (blue) are displayed separately in Fig. 2D. The cutoff frequency of devices can be determined as the intersection of resistance and capacitive reactance in the impedance plot (31) and 1L graphene–inserted device shows an order of magnitude higher fC of 220 GHz than those of Au-contact (25.9 GHz) and 3L graphene–inserted (33.5 GHz) device. This ultrahigh fc can be attributed to two factors: low series resistance RS and low junction capacitance Cj. According to the small-signal equivalent circuit, the junction resistance (Rj) and Cj originate from the depletion region at the Schottky junction, whereas RS is composed of the contact resistance (RC) and undepleted region resistance (RI). The total impedance, ZL, can be expressed as
(1) |
and ZL ≅ Rs − j∣Xc∣ = Rs − j∣1/2πfCj∣ at f ≫ 1/2πRjCj. According to simple mathematics to justify the approximation in Supplementary Text 2, |Im(ZL)| maximum (in the imaginary part of ZL) should appear at the frequency of , which is experimentally found to be ~100 MHz or less as shown in fig. S6. Hence, the criteria f ≫ 1/2πRjCjfor our approximation could be well satisfied at gigahertz frequencies; the real and imaginary parts of impedance at the frequencies higher than a few gigahertz can be interpreted as the series resistance (RS) and the capacitive reactance (XC), respectively. In this respect, the capacitance of 1L graphene–inserted device must be smaller than that of Au- or 3L graphene–inserted device at a same high frequency (Fig. 2D). Furthermore, at their cutoff frequencies, 1L graphene–inserted device shows quite lower RS (~22 ohm) than those (40 to 55 ohm) of Au contact and 3L graphene–inserted devices as indicated by dashed arrows.
The capacitances of various devices with the same 160-nm-thick MoSe2 are compared. According to the Smith chart in Fig. 2C, the three clockwise circles end to meet different capacitive reactance (−2.0j versus ~ −0.6j) at the same 40 GHz. This clearly indicates that the capacitance of 1L graphene–inserted device must be about three times lower than those of Au- and 3L graphene–inserted diodes, although MoSe2 thickness in each device is almost the same, to be ~160 nm. The impedance feature of 3L graphene–inserted device seems similar to that of Au-contact device, resulting in a fC of 33.5 GHz. This behavior from 3L graphene appears quite similar to those of another device with a few nanometer-thin graphite (fC ~ 20 GHz for 4L and 10L graphene, fig. S7). We have also fabricated a few more 1L graphene–inserted devices with ~160-nm-thick MoSe2, which consistently exhibit high fC of 110 to 220 GHz (see fig. S8). The inset graph of Fig. 2D summarizes fC results from the diodes with 1L ~ ML graphene and from Au-contact diode.
We also studied the effect of MoSe2 thickness in our vertical devices. We found that, even with 1L graphene insertion, the fC decreases below 100 GHz if MoSe2 thickness is not thick enough. According to Fig. 2E, fC of the device with 60-nm-thin MoSe2 is limited to 37.3 GHz, resulting from smaller RS and XC than those of our main device with 160-nm-thick MoSe2. RS comprises undepleted thickness-induced RI and RC (constant), and thinner MoSe2 thus brings relatively smaller RS (due to smaller RI) in favor of high fC. However, thinner MoSe2 also causes smaller XC with higher Cj, which would be critically opposed to increasing fC. Likewise, fC of the device with 128-nm-thin MoSe2 was ~67 GHz (fig. S9), which is higher than 37.3 GHz but lower than 110 GHz (the smallest fC of L graphene–inserted diode with 160-nm-thick MoSe2). The inset graph of Fig. 2E summarizes measured fC results according to MoSe2 thickness. Regarding an optimum MoSe2 thickness for maximum fC (from smallest RS and Cj), further study would be necessary on the basis of any precise information on the carrier concentration in working MoSe2 and its dielectric constant.
Apart from above device performances, another important feature of the vertical diodes to note is their ageing stability and reproducibility; even if they are kept in ambient air (45% relative humidity at room temperature) for more than a week, they show no difference in terms of the RF impedance feature and the fC behavior (see figs. S8, S10, and S11 for more details including reproducibility). We attribute such a stability to encapsulation effects probably achieved from the top Pt electrode covering the thick MoSe2. Last, Ti-contact device with a different architecture (Pt/MoSe2/Ti) was also fabricated and its fC (24.4 GHz) appears very comparable to that of Au-contact device (25.9 GHz). According to fig. S12, those two devices with different Ohmic contact seem to have almost identical impedance characteristics in high-frequency regime, where the Pt/MoSe2 junction capacitance must be the control factor for impedance.
DISCUSSION
To elucidate the role of graphene (Gr) insertion in vdW Schottky diode, the junction capacitance values of each diode were extracted from the imaginary part of impedances as shown in Fig. 3A. Although the three diodes with 160-nm-thick MoSe2 initially show almost the same capacitance (Cj ~ 110 nF/cm2), each becomes to display different behavior as frequency increases. Schottky diode without graphene insertion seems to show a constant capacitance regardless of frequency. Because the diode capacitance mostly originates from the depletion thickness in MoSe2, the observed constant capacitance could be attributed to the fixed depletion thickness at Pt/n-MoSe2 Schottky junction. On the other hand, the capacitance of 1L graphene–inserted diode largely decreases with the input frequency reaching to ~43 nF/cm2 at 20-GHz input frequency, and it is almost three times smaller value than that of Au-contact case. The capacitance of ML (or 3L) graphene–inserted diode appears to moderately decrease to ~88 nF/cm2. The observed frequency-dependent capacitance from graphene-inserted devices strongly suggests that the charge depletion at the MoSe2/Gr junction arises under high-frequency operation; reduction in total capacitance takes place because of series capacitance effects. According to the detailed calculation in fig. S13 and Supplementary Text 3, the total capacitance [Cj,tot, or Cj,high = Cj,PtCj,Au-Gr/(Cj,Pt + Cj,Au-Gr)] at high frequencies (> 40 GHz) becomes ~4 times smaller than that (Cj,low = Cj,Pt) of the Pt/n-MoSe2/Au diode if the MoSe2 flake is sufficiently thick. Cj,Pt and Cj,Au-Gr indicate the junction capacitances due to depleted MoSe2 near Pt and Au/Gr, respectively. These explanations are consistent with the result of Figs. 3A and 2D. Along with the decrease in Cj,tot, RS (= RC + RI) simultaneously decreases, because RI decreases with the increase of depletion thickness at high frequencies. RS thus becomes small with the frequency increase, resulting in a markedly high fC as combined with the effects of small Cj,tot.
Fig. 3. Capacitance lowering at high-frequency regime and its explanation.
(A) Capacitance-frequency plot that is extracted from the imaginary part of the impedance of each diode in Fig. 2. (B) Band diagram of Pt/n-MoSe2/1L-graphene/Au and Pt/n-MoSe2/ML-graphene/Au diode at DC and high-frequency AC. Trapping-induced modulation is frozen at high AC frequencies. (C) Plot of MoSe2 depletion thickness (xd,Au-Gr) as a function of frequency (D) Plots of built-in potential ϕi,Au−Gr between MoSe2 and graphene as a function of frequency. Max. ϕi,Au−Gr between MoSe2 and graphene appears to be ~1.3 V which well matches with the theoretical work function difference (qWAu-qWMoSe2 ≃ 1.3 eV) between Au and n-MoSe2. It means that graphene’s Fermi level decreases down to EO,Gr at high frequency although initial maximum of EF,Gr-EO,Gr was 1.3 eV, the same as work function difference between Au and MoSe2. (E) Accumulated total electron charge density, N versus frequency plot, which unintentionally follows exponential form. Inset logarithmic plot results in fO of ~4.18 GHz, to be a characteristic frequency.
The rise of depletion thickness in MoSe2 near Gr could be attributed to the graphene’s failure in its charge modulation at high frequencies. According to previous studies on DC operation (32), 1L graphene between two materials acts as a tunneling layer that is nearly transparent to the vertically transported electrons, while such vertical tunneling originates from momentum mismatch. Moreover, graphene can modulate the potential difference (ΔV) between the two materials due to the graphene-captured charges (Qe) and quantum capacitance (CGR); ΔV = Qe/CGR. Under DC voltage sweep, the Schottky barrier at the graphene/MoSe2 junction is effectively nullified, so that the Fermi level of each material may be aligned maintaining an equilibrium Ohmic junction at a certain temperature (Fig. 3B, upper band diagram). However, at much high frequencies, the electrical behavior of the Au/1L graphene/MoSe2 junction seems not Ohmic anymore but rather ideal Schottky, as shown in the band diagrams (Fig. 3B, middle). It is probably because tunneling electrons may not be captured by the graphene layer in such a short duration. As a result, charge modulation is prevented and 1L graphene only plays as a tunneling layer. It must be also considered that CGR-induced delay (RSCGR) exists in the AC behavior. Charge trapping in graphene might need a minimum time (scattering-induced mean free time, τ) which would be longer than RSCGR delay. At such a high frequency (f ≫ 1/ RSCGR), charge modulation may largely be frozen.
Following analysis may support aforementioned statements. The high frequency–induced MoSe2 depletion thickness (xd,Au-Gr) of Fig. 3B (middle) at 1L graphene/MoSe2 junction can be extracted out by considering blue and red plots for capacitance in Fig. 3A which reflect upper and middle band diagrams of Fig. 3B, respectively. Calculation details found in fig. S14, which include a formula (1/Cj,Au-Gr = xd,Au-Gr/ɛMoSe2, ɛMoSe2 is dielectric permittivity). The calculation results in frequency-dependent plot of xd,Au-Gr as seen in Fig. 3C, where at 40 GHz the thickness saturates to ~137 nm. Since we know ɛMoSe2 (~ 8.5ɛo) and Nd (7.13 × 1016 cm−3) in the inset equation (xd,Au-Gr = ), the built-in potential ϕi,Au−Gr between MoSe2 and graphene can be plotted as a function of frequency in Fig. 3D, where 1.3 V appears to be a maximum. A voltage of 1.3 V well matches with the theoretical work function difference (qWAu − qWMoSe2 = ~1.3 eV) (29, 33) between Au and n-MoSe2. This result indicates that pinning-free Schottky barrier certainly arises at the high frequency due to the failure of graphene’s charge modulation and the maximum attains to 1.3 eV. Likewise, the frequency-dependent Schottky barrier comes from frequency-limited charge trapping or modulation limit in inserted graphene layer, which means that graphene’s Fermi level (EF,Gr) decreases as a function of frequency as described with EF,Gr- EO,Gr where EO,Gr is constant as a minimum Fermi level. Maximum of EF,Gr- EO,Gr becomes 1.3 eV, the same as work function difference between Au and MoSe2. Now, frequency-dependent EF,Gr- EO,Gr is plotted along in Fig. 3D, and in Fig. 3E it is converted to accumulated (total) electron charge density, N which can be obtained by integrating the density of states (dN/dE) with energy via a relationship (dN/dE ∝ √N). Theoretical analysis and calculation details are in Supplementary Text 4. According to Fig. 3E, N is accidentally well fitted to an exponential form as frequency varies, and the slope of inset logarithmic plot shows a characteristic frequency fO, to be ~4.18 GHz which is for 36% charge accumulation in the Dirac cone. τ (= 1/fO) becomes ~0.24 ns. On the basis of τ (= RSCGR) value, approximate CGR can be worked out to be ~1.14 μF/cm2 (RS = ~210 ohm at 4.18 GHz, and diode area = ~100 μm2), which is a quite comparable to number values of undoped graphene’s quantum capacitance (34, 35).
For ML graphene–inserted devices, the RC delay associated with the quantum capacitance of graphene is still effective and the depletion thickness should have the frequency dependence. However, graphite will replace the position of Au as shown in another band diagram of Fig. 3B (bottom). Graphite has much smaller work function than Au, which results in a limited depletion thickness of MoSe2 and moderate decrease of Cj,tot. Consequently, ML (over 3L) graphene–inserted diodes behave similar to Au-contact device at high frequencies. More analysis details including the plots analogous to Fig. 3 (C to E) are found in fig. S15.
We designed another vertical vdW devices to more directly confirm the graphene’s role under high-frequency operations. Here, we fabricated devices with Ti/n-MoSe2/1L graphene/Au junction as schematically presented in Fig. 4A, which should exhibit Ohmic characteristics under DC operation. We found that the device indeed exhibits quasi-Ohmic characteristics as shown in Fig. 4B, which confirms that both junctions at graphene/MoSe2 and Ti/MoSe2 interface have ignorable barrier under DC condition. Under RF conditions, however, the device exhibits totally different results as seen in the equivalent circuit of Fig. 4C. One-port S-parameter measurement was conducted with a similar setup to Fig. 2A, but this time, graphene/Au electrode is forwardly biased with DC voltage ranging from 0 to 1 V as depicted in Fig. 4 (A and C). As a result, the Smith chart of Fig. 4D shows diode-like behavior with circular contour in lower plane, which implies the existence of junction capacitance under RF input signal. In addition, the circular contour varies with applied DC voltage. According to more detailed analysis in Fig. 4E, Rj values (real part in impedance) decrease with the forward bias voltage at low 50 MHz while total capacitance increases with the voltage at high 20 GHz (because depletion thickness decreases as described in Fig. 4G). On the basis of the Eq. 1 and the circuit of Fig. 4C, it is expected that the impedance Z under low-frequency AC input approaches to junction resistance, Rj (≅ RS + Rj) because Rj is much larger than RS in this low-frequency regime. As the frequency increases, Z approaches to RS – j/2πfCj by ignoring Rj. At gigahertz high frequencies, imaginary part (−j/2πfCj) becomes much more important than RS, so that Cj is extracted. Likewise, Fig. 4 (D and E) directly indicate a diode-like behavior of our new device under forward DC bias conditions. The underlying principle of this behavior is presented graphically in Fig. 4G where the depletion thickness is modulated by forward DC bias. The impedance plots in Fig. 4F for Ti/n-MoSe2/1 L graphene show fC of 33 GHz resembling the behavior of Pt/n-MoSe2/1 L graphene diode.
Fig. 4. Experiment for clarifying the formation of barrier at Au/1L graphene/n-MoSe2 interface.
(A) 2D schematic and (B) DC I-V characteristic on the cross-section of the device which has Ohmic contacts at each terminal, one of which is Au/1L graphene contact. (C) Circuit diagram of the measurement including the small-signal equivalent circuit of the device. (D) S11 parameter of the device on the Smith chart in which characteristic semicircle contour of diode can be seen. (E) Extracted resistance under 50 MHz and junction capacitance under 20 GHz AC input signal with different DC forward bias from 0 to 1 V. (F) Impedance-frequency plot under zero DC bias. (G) Band diagram which shows the change of Schottky barrier and junction capacitance at Au/1L graphene/n-MoSe2 junction under forward bias.
As our last work, we have fabricated a high-frequency mixer using two Pt/n-MoSe2/graphene/Au diodes, to more demonstrate the validity of our Schottky diode with inserted graphene. Figure S16 (A to F) displays 28-GHz mixer applications using two graphene-inserted Schottky diodes, which clearly displays the merit of graphene-insertion at high frequencies.
In summary, we have elucidated the implication of graphene insertion in vdW layered vertical devices at ultrahigh frequencies. The Pt/n-MoSe2 Schottky diode with graphene/Au contact shows an extremely high fC of approximately 220 GHz in maximum, which would be applied to high-frequency mixer and mm-Wave energy harvesting. The mechanism of such a high fC is attributed to the fact that quantum capacitance–driven charge modulation by the inserted graphene becomes essentially frozen above gigahertz frequencies, so that the Ohmic graphene/MoSe2 junction may be transformed to a pinning-free Schottky junction. By integrating the diodes within a peripheral circuit, we successfully demonstrate an RF mixer for 5G communication. Although high cutoff frequencies over ~100 GHz have been reported for devices with single-crystalline Si (28, 36, 37) and GaAs (38, 39), achieving 5G-compatible frequencies (> 28 GHz) using thin film–based devices is challenging (40, 41) (see fig. S17 for comparison.) We believe that our vertical device geometry is essentially compatible with the conventional fabrication process and has great potential in high-frequency applications. Equipped with month-long ambient stability and fabrication ease, we conclude that our vertical diode with inserted graphene enlightens a new possibility toward future RF devices with 2D layered semiconductors.
MATERIALS AND METHODS
Device fabrication
For all diodes in this study, a glass substrate (Eagle XG) was sequentially cleaned with acetone, ethanol, and deionized water using an ultrasonicator. Then, 10-nm Al2O3 was deposited using an atomic layer deposition (ALD) system as a buffer layer. For the Pt/n-MoSe2/1 L graphene/Au diode, a 5 nm Ti/10 nm Au bottom electrode was patterned by conventional photolithography and deposited using a DC magnetron sputter. For the wet-transfer, polystyrene (PS) was initially spin-coated onto CVD monolayer graphene on the 285 nm SiO2/p+-Si substrate (Graphene Supermarket). Then, the monolayer graphene film was separated from the SiO2/p+-Si substrate using deionized water to be wet-transferred onto the bottom electrode. After wet-transfer, PS was removed by toluene. A 2H-MoSe2 flake (HQ Graphene) was then mechanically exfoliated and dry-transferred onto the bottom electrode. As a last step, a 150-nm Pt top electrode was patterned by the same method used for the bottom electrode. For the case of the Pt/n-MoSe2/Au diode, all the processes were same but the wet transfer process of graphene sheet that was omitted.
Last, for the mixer application, two Pt/n-MoSe2/graphene/Au diodes with similar characteristics were selected, and the glass substrate was diced except for the area containing the device. The device was then connected by silver paste to a mixer circuit board. All Schottky diodes were fabricated according to the guidelines for the CPW pattern (I-V, C-V, S11 parameter, and power spectrum measurements) or mixer pattern.
Material characterization
Noncontact mode AFM (NX-10, Park Systems) was used to characterize the surface topography as well as the thickness of 2H-MoSe2 flakes. Raman spectroscopy was conducted by using a 532-nm laser source with a Raman spectrometer (LabRam Aramis, and Horriba). TEM samples were prepared using a focused ion beam (Crossbeam 540, ZEISS). TEM and STEM images were acquired with double Cs-aberration corrected JEOL ARM-200F, which operated at 200 kV.
DC electrical measurements
DC current-voltage (I-V) characteristic measurements were carried out by using a semiconductor parameter analyzer (HP4155C, Agilent Technologies). Low-temperature measurements were performed in the dark under vacuum (~1.5 mtorr). DC capacitance-voltage (C-V) measurements were conducted by using a precision LCR meter (HP4284A, Agilent Technologies) under a small signal frequency of 1 MHz.
Nonlinearity and responsivity FOM from DC I-V characteristic
The nonlinearity FOM of a diode is the ratio of the differential conductance to the conductance, i.e.
(2) |
and is a measure of the deviation from a linear resistor (27).
On the other hand, the responsivity FOM is defined as the ratio of the rectified DC current and the RF input power that is delivered to device (11) i.e.
(3) |
For a nonlinear device, current as a function of voltage can be expanded in a Taylor series at V = V0
(4) |
If an external voltage V = V0 + vcosωt is applied across the device, the current can be expressed as
(5) |
Therefore, if the amplitude of applied AC input signal v is small enough, the rectified current due to AC signal is approximately
(6) |
And for small signal AC input v(t) = vcosωt, time-averaged power Pin is calculated as
(7) |
Then, the responsivity FOM can be expressed as
(8) |
The nonlinearities the responsivities in fig. S5 and Fig. 1 are calculated from I-V characteristics in Fig. 1 (E and F).
S-parameter and fC measurement
A 1-port network was constructed by using a reference G-S-G CPW pattern with a characteristic impedance of 50 ohm. Scattering parameter S11 measurements were performed by using a vector network analyzer (MS4647B, Anritsu) after calibration. The RF input signal was set to sweep from 50 MHz to 40 GHz.
Power spectrum and mixer measurement
For output power spectrum measurements using two terminals, an RF signal generator (MG3695A, Anritsu) was connected to the input terminal of devices. A spectrum analyzer (MS2850A-046, Anritsu) was then connected to the output terminal of the devices so that the output signal could be analyzed. For mixer demonstration, two RF signal generators were used along with two Schottky diodes, where the generators were connected to the RF and LO terminals.
Acknowledgments
We thank K. Lee and J. Park for the preparation of CVD graphene and critical comments on this work.
Funding: This work was supported by National Research Foundation of Korea grant 2017R1A5A1014862 (S.H., S.L., M.J., Y.L., L.J.W., S.P., K.K., Y.-W.S., and S.I.) and National Research Foundation of Korea grant 2021R1A6A3A13044763 (S.H.).
Author contributions: Conceptualization: S.H. and S.I. Methodology: S.H., C.-U.H., S.P., M.J., S.L., Y.L., K.K., L.J.W., C.J., and J.-G.Y. Investigation: S.H., C.-U.H., C.J., S.L., and Y.L. Visualization: S.H. Funding acquisition: S.H. and S.I. Project administration: S.H., J.-G.Y., and S.I. Supervision: JKY and S.I. Writing—original draft: S.H. and S.I. Writing—review and editing: S.H., C.-U.H., S.L., M.J., C.J., Y.L., L.J.W., S.P., K.K., Y.-W.S., J.-G.Y., and S.I.
Competing interests: The authors declare that they have no competing interests.
Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials.
Supplementary Materials
This PDF file includes:
Supplementary Texts 1 to 4
Figs. S1 to S17
References
REFERENCES AND NOTES
- 1.S. Lukman, L. Ding, L. Xu, Y. Tao, A. C. Riis-Jensen, G. Zhang, O. Y. S. Wu, M. Yang, S. Luo, C. Hsu, L. Yao, G. Liang, H. Lin, Y.-W. Zhang, K. S. Thygesen, Q. J. Wang, Y. Feng, J. Teng, High oscillator strength interlayer excitons in two-dimensional heterostructures for mid-infrared photodetection. Nat. Nanotechnol. 15, 675–682 (2020). [DOI] [PubMed] [Google Scholar]
- 2.W. Choi, S. Hong, Y. Jeong, Y. Cho, H. G. Shin, J. H. Park, Y. Yi, S. Im, Dynamic oscillation via negative differential resistance in type III junction organic/two-dimensional and oxide/two-dimensional transition metal dichalcogenide diodes. Adv. Funct. Mater. 31, 2009436 (2020). [Google Scholar]
- 3.H. G. Shin, D. Kang, Y. Jeong, K. Kim, Y. Cho, J. Park, S. Hong, Y. Yi, S. Im, High-performance van der Waals junction field-effect transistors utilizing organic molecule/transition metal dichalcogenide interface. ACS Nano 14, 15646–15653 (2020). [DOI] [PubMed] [Google Scholar]
- 4.J. Y. Lim, A. Pezeshki, S. Oh, J. S. Kim, Y. T. Lee, S. Yu, D. K. Hwang, G.-H. Lee, H. J. Choi, S. Im, Homogeneous 2D MoTe2p-n junctions and CMOS inverters formed by atomic-layer-deposition-induced doping. Adv. Mater. 29, 1701798 (2017). [DOI] [PubMed] [Google Scholar]
- 5.D. Sarkar, X. Xie, W. Liu, W. Cao, J. Kang, Y. Gong, S. Kraemer, P. M. Ajayan, K. Banerjee, A subthermionic tunnel field-effect transistor with an atomically thin channel. Nature 526, 91–95 (2015). [DOI] [PubMed] [Google Scholar]
- 6.J. Wu, H.-Y. Chen, N. Yang, J. Cao, X. Yan, F. Liu, Q. Sun, X. Ling, J. Guo, H. Wang, High tunnelling electroresistance in a ferroelectric van der Waals heterojunction via giant barrier height modulation. Nat. Electron. 3, 466–472 (2020). [Google Scholar]
- 7.Y. Wang, J. H. Kim, R. J. Wu, J. Martinez, X. Song, J. Yang, F. Zhao, A. Mkhoyan, H. Y. Jeong, M. Chhowalla, Van der Waals contacts between three-dimensional metals and two-dimensional semiconductors. Nature 568, 70–74 (2019). [DOI] [PubMed] [Google Scholar]
- 8.M. Farmanbar, G. Brocks, Ohmic contacts to 2D semiconductors through van der Waals bonding. Adv. Electron. Mater. 2, 1500405 (2016). [Google Scholar]
- 9.P.-C. Shen, C. Su, Y. Lin, A.-S. Chou, C.-C. Cheng, J.-H. Park, M.-H. Chiu, A.-Y. Lu, H.-L. Tang, M. M. Tavakoli, G. Pitner, X. Ji, Z. Cai, N. Mao, J. Wang, V. Tung, J. Li, J. Boker, A. Zettl, C.-I. Wu, T. Palacios, L.-J. Li, J. Kong, Ultralow contact resistance between semimetal and monolayer semiconductors. Nature 593, 211–217 (2021). [DOI] [PubMed] [Google Scholar]
- 10.H. Cho, D. Kang, Y. Lee, H. Bae, S. Hong, Y. Cho, K. Kim, Y. Yi, J. H. Park, S. Im, Dramatic reduction of contact resistance via ultrathin LiF in two-dimensional MoS2 field effect transistors. Nano Lett. 21, 3503–3510 (2021). [DOI] [PubMed] [Google Scholar]
- 11.X. Zhang, J. J. Grajal, J. L. Vazquez-Roy, U. Radhakrishna, X. Wang, W. Chern, L. Zhou, Y. Lin, P.-C. Shen, X. Ji, X. Ling, A. Zubair, Y. Zhang, H. Wang, M. Dubey, J. Kong, M. Dresselhaus, T. Palacios, Two-dimensional MoS2-enabled flexible rectenna for Wi-Fi-band wireless energy harvesting. Nature 566, 368–372 (2019). [DOI] [PubMed] [Google Scholar]
- 12.S. J. Yang, K.-T. Park, J. Im, S. Hong, Y. Lee, B.-W. Min, K. Kim, S. Im, Ultrafast 27 GHz cutoff frequency in vertical WSe2 Schottky diodes with extremely low contact resistance. Nat. Commun. 11, 1574 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.D. Krasnozhon, D. Lembke, C. Nyffeler, Y. Leblebici, A. Kis, MoS2 transistors operating at gigahertz frequencies. Nano Lett. 14, 5905–5911 (2014). [DOI] [PubMed] [Google Scholar]
- 14.A. Sanne, R. Ghosh, A. Rai, M. N. Yogeesh, S. H. Shin, A. Sharma, K. Jarvis, L. Mathew, R. Rao, D. Akinwande, S. Banerjee, Radio frequency transistors and circuits based on CVD MoS2. Nano Lett. 15, 5039–5045 (2015). [DOI] [PubMed] [Google Scholar]
- 15.R. Cheng, S. Jiang, Y. Chen, Y. Liu, N. Weiss, H.-C. Cheng, H. Wu, Y. Huang, X. Duan, Few-layer molybdenum disulfide transistors and circuits for high-speed flexible electronics. Nat. Commun. 5, 5143 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.L. Ju, B. Geng, J. Horng, C. Girit, M. Martin, Z. Hao, H. A. Bechtel, X. Liang, A. Zettl, Y. R. Shen, F. Wang, Graphene plasmonics for tunable terahertz metamaterials. Nat. Nanotechnol. 6, 630–634 (2011). [DOI] [PubMed] [Google Scholar]
- 17.J. Shim, H. S. Kim, Y. S. Shim, D.-H. Kang, H.-Y. Park, J. Lee, J. Jeon, S. J. Jung, Y. J. Song, W.-S. Jung, H. Lee, S. Park, J. Kim, S. Lee, Y.-H. Kim, J.-H. Park, Extremely Large Gate Modulation in Vertical Graphene/WSe2 Heterojunction Barristor Based on a Novel Transport Mechanism. Adv. Mater. 28, 5293–5299 (2016). [DOI] [PubMed] [Google Scholar]
- 18.D. Seo, D. Y. Lee, J. Kwon, J. J. Lee, T. Taniguchi, K. Watanabe, G.-H. Lee, K. S. Kim, J. Hone, Y. D. Kim, H.-J. Choi, High-performance monolayer MoS2 field-effect transistor with large-scale nitrogen-doped graphene electrodes for Ohmic contact. Appl. Phys. Lett. 115, 012104 (2019). [Google Scholar]
- 19.K.-E. Byun, H.-J. Chung, J. Lee, H. Yang, H. J. Song, J. Heo, D. H. Seo, S. Park, S. W. Hwang, I. Yoo, K. Kim, Graphene for true Ohmic contact at metal-semiconductor junctions. Nano Lett. 13, 4001–4005 (2013). [DOI] [PubMed] [Google Scholar]
- 20.S. Kim, J. Nah, I. Jo, D. Shahrjerdi, L. Colombo, Z. Yao, E. Tutuc, S. K. Banerjee, Realization of a high mobility dual-gated graphene field-effect transistor with Al2O3 dielectric. Appl. Phys. Lett. 94, 062107 (2009). [Google Scholar]
- 21.D. De Fazio, D. G. Purdie, A. K. Ott, P. Braeuninger-Weimer, T. Khodkov, S. Goossens, T. Taniguchi, K. Watanabe, P. Livreri, F. H. L. Koppens, S. Hofmann, I. Goykhman, A. C. Ferrari, A. Lombardo, High-mobility, wet-transferred graphene grown by chemical vapor deposition. ACS Nano 13, 8926–8935 (2019). [DOI] [PubMed] [Google Scholar]
- 22.Y. Wu, K. A. Jenkins, A. Valdes-Garcia, D. B. Farmer, Y. Zhu, A. A. Bol, C. Dimitrakopoulos, W. Zhu, F. Xia, P. Avouris, Y.-M. Lin, State-of-the-art graphene high-frequency electronics. Nano Lett. 12, 3062–3067 (2012). [DOI] [PubMed] [Google Scholar]
- 23.N. Petrone, I. Meric, T. Chari, K. L. Shepard, J. Hone, Graphene field-effect transistors for radio-frequency flexible electronics. 3, IEEE J. Electron Devices Soc., 44–48 (2015).
- 24.S. Sankaran, O. Kenneth, Schottky barrier diodes for millimeter wave detection in a foundry CMOS process. IEEE Electron Device Lett. 26, 492–494 (2005). [Google Scholar]
- 25.G. Hueber, M. A. Niknejad, Millimeter-Wave Circuits for 5G and Radar (Cambridge Univ. Press, 2019).
- 26.K. Sotthewes, R. van Bremen, E. Dollekamp, T. Boulogne, K. Nowakowski, D. Kas, H. J. W. Zandvliet, P. Bampoulis, Universal fermi-level pinning in transition-metal dichalcogenides. J. Phys. Chem. C. Nanomater. Interfaces 123, 5411–5420 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.E. Donchev, J. S. Pang, P. M. Gammon, A. Centeno, F. Xie, P. K. Petrov, J. D. Breeze, M. P. Ryan, D. J. Riley, N. McN, The rectenna device: from theory to practice (a review). MRS Energy Sustain. 1, 1 (2014). [Google Scholar]
- 28.R. Han, Y. Zhang, D. Coquillat, H. Videlier, W. Knap, E. Brown, K. K. O, A 280-GHz Schottky diode detector in 130-nm digital CMOS. IEEE J. Solid State Circuits 46, 2602–2612 (2011). [Google Scholar]
- 29.R. Muller, T. Kamins, M. Chan, Device Electronics for Integrated Circuits (John Wiley & Sons Inc., 2003). [Google Scholar]
- 30.D. Pozar, Microwave Engineering (John Wiley & Sons Inc., 2011). [Google Scholar]
- 31.A. Chasin, S. Steudel, F. Vanaverbeke, K. Myny, M. Nag, T. Ke, S. Schols, G. Gielen, J. Genoe, P. Heremans, UHF IGZO Schottky diode. In 2012 International Electron Devices Meeting 12.14.11–12.14.14 (IEEE, 2012). [Google Scholar]
- 32.X. Zhu, S. Lei, S.-H. Tsai, X. Zhang, J. Liu, G. Yin, M. Tang, C. M. Torres Jr., A. Navabi, Z. Jin, S.-P. Tsai, H. Qasem, Y. Wang, R. Vajtai, R. K. Lake, P. M. Ajayan, K. L. Wang, A study of vertical transport through graphene toward control of quantum tunneling. Nano Lett. 18, 682–688 (2018). [DOI] [PubMed] [Google Scholar]
- 33.L. L. Zhang, X. Zhao, H. Ji, M. D. Stoller, L. Lai, S. Murali, S. Mcdonnell, B. Cleveger, R. M. Wallace, R. S. Ruoff, Nitrogen doping of graphene and its effect on quantum capacitance, and a new insight on the enhanced capacitance of N-doped carbon. Energ. Environ. Sci. 5, 9618–9625 (2012). [Google Scholar]
- 34.Q. Zhang, S. Zhang, B. A. Sperling, N. V. Nguyen, Band offset and electron affinity of monolayer MoSe2 by internal photoemission. J. Electron. Mater. 48, 6446–6450 (2019). [Google Scholar]
- 35.J. Xia, F. Chen, J. Li, N. Tao, Measurement of the quantum capacitance of graphene. Nat. Nanotechnol. 4, 505–509 (2009). [DOI] [PubMed] [Google Scholar]
- 36.K. M. Strohm, J. Buechler, E. Kasper, SIMMWIC rectennas on high-resistivity silicon and CMOS compatibility. IEEE Trans. Microw. Theory Tech. 46, 669–676 (1998). [Google Scholar]
- 37.C. Mishra, U. Pfeiffer, R. Rassel, S. Reynolds, Silicon Schottky diode power converters beyond 100 GHz. In 2007 IEEE Radio Frequency Integrated Circuits (RFIC) Symposium (IEEE, 2007), pp. 547–550. [Google Scholar]
- 38.H. Zhao, V. Drakinskiy, P. Sobis, J. Hanning, T. Bryllert, A.-Y. Tang, J. Stake, Development of a 557 GHz GaAs monolithic membrane-diode mixer. In 2012 International Conference on Indium Phosphide and Related Materials (2012), pp. 102–105. [Google Scholar]
- 39.J. Rollin, G. I. Chance, S. R. Davies, A membrane planar diode for 200 GHz mixing applications. In Infrared and Millimeter Waves, Conference Digest of the 2004 Joint 29th International Conference on 2004 and 12th International Conference on Terahertz Electronics (IEEE, 2004), pp. 205–206. [Google Scholar]
- 40.J. Oh, S. Moon, D. S. Kang, S.-D. Kim, High-performance 94-GHz single-balanced diode mixer using disk-shaped GaAs Schottky diodes. IEEE Electron Device Lett. 30, 206–208 (2009). [Google Scholar]
- 41.D. G. Georgiadou, J. Semple, A. A. Sagade, H. Forsten, P. Rantakari, Y.-H. Lin, F. Alkhalil, A. Seitkhan, K. Loganathan, H. Faber, T. D. Anthopoulos, 100 GHz zinc oxide Schottky diodes processed from solution on a wafer scale. Nat. Electron. 3, 718–725 (2020). [Google Scholar]
- 42.B. W. H. Baugher, H. O. H. Churchill, Y. Yang, P. Jarillo-Herrero, Optoelectronic devices based on electrically tunable p–n diodes in a monolayer dichalcogenide. Nat. Nanotechnol. 9, 262–267 (2014). [DOI] [PubMed] [Google Scholar]
- 43.A. Ortiz-Conde, F. J. G. Sánchez, J. Muci, Exact analytical solutions of the forward non-ideal diode equation with series and shunt parasitic resistances. Solid State Electron. 44, 1861–1864 (2000). [Google Scholar]
- 44.J. Zhang, Y. Li, B. Zhang, H. Wang, Q. Xin, A. Song, Flexible indium-gallium-zinc-oxide Schottky diode operating beyond 2.45 GHz. Nat. Commun. 6, 7561 (2015). [DOI] [PubMed] [Google Scholar]
- 45.A. Chasin, M. Nag, A. Bhoolokam, K. Myny, S. Steudel, S. Schols, J. Genoe, G. Gielen, P. Heremans, Gigahertz operation of a-IGZO schottky diodes. IEEE Transactions on Electron Devices 60, 3407–3412 (2013). [Google Scholar]
- 46.J.-P. Ao, K. Takahashi, N. Shinohara, N. Niwa, T. Fujiwara, Y. Ohno, S-Parameter Analysis of GaN Schottky Diodes for Microwave Power Rectification. 2010 IEEE Compound Semiconductor Integrated Circuit Symposium (2010), pp. 1–4.
- 47.J. E. Lee, G. Ahn, J. Shim, Y. S. Lee, S. Ryu, Optical separation of mechanical strain from charge doping in graphene. Nat. Commun. 3, 1024 (2012). [DOI] [PubMed] [Google Scholar]
- 48.G. Nazir, H. Kim, J. Kim, K. S. Kim, D. H. Shin, M. F. Khan, D. S. Lee, J. Y. Hwang, C. Hwang, J. Suh, J. Eom, S. Jung, Ultimate limit in size and performance of WSe2 vertical diodes. Nat. Commun. 9, 5371 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.C.-M. Kang, J. Wade, S. Yun, J. Lim, H. Cho, J. Roh, H. Lee, S. Nam, D. D. C. Bradley, J.-S. Kim, C. Lee, 1 GHz pentacene diode rectifiers enabled by controlled film deposition on SAM-treated Au anodes. Adv. Electron. Mater. 2, 1500282 (2016). [Google Scholar]
- 50.D. Im, H. Moon, M. Shin, J. Kim, S. Yoo, Towards gigahertz operation: Ultrafast low turn-on organic diodes and rectifiers based on C60 and tungsten oxide. Adv. Mater. 23, 644–648 (2011). [DOI] [PubMed] [Google Scholar]
- 51.S. Sankaran, K. K. O, Schottky diode with cutoff frequency of 400 GHz fabricated in 0.18 μm CMOS. Electron. Lett. 41, 506–508 (2005). [Google Scholar]
- 52.B. T. Bulcha, J. L. Hesler, V. Drakinskiy, J. Stake, A. Valavanis, P. Dean, L. H. Li, N. S. Barker, Design and characterization of 1.8-3.2 THz Schottky-based harmonic mixers. IEEE Trans. Terahertz Sci. Technol. 6, 737–746 (2016). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Supplementary Texts 1 to 4
Figs. S1 to S17
References