Introduction

The continuous scaling and requirements for low-power operation of field-effect transistors (FETs) limit the range of threshold voltage (Vth), which is the gate-source voltage (VGS) at which a FET turns on. In modern FETs, the gates are made of metal and consequently, the threshold voltage cannot easily be adjusted because it is fixed by the work function of the metal. This problem is even more exacerbated in metal-semiconductor FETs (MESFETs) and high-electron-mobility transistors (HEMTs), which are used in high-speed communication systems1. In such FETs, there is an additional constraint on the threshold voltage as prohibitively large gate leakage current (IG) is obtained if Vth was not properly adjusted. This is because the channel carrier density in these FETs is usually controlled by a Schottky barrier between the metal gate and semiconductor2. The gate current is therefore the current of the Schottky junction which must stay in the off state to keep the gate leakage negligible. The inability to tune the work function of the gate metal, on which the Schottky barrier height (SBH) and consequently the turn-on voltage of the Schottky junction critically depend, limits possible applications of MESFETs and HEMTs.

A better control of the threshold voltage of FETs could be obtained if the gate metal was replaced by a material of adjustable work function. Prior to the 45 nm technology node, poly-Si gates were used to set the threshold voltage in metal-oxide-semiconductor FETs during their fabrication3,4. However, poly-Si cannot be used in MESFETs and HEMTs and its work function cannot be adjusted after fabrication, i.e., during device operation. Such adjustment is possible in graphene because its Fermi level (and therefore the work function) can be controlled by an external electric field. The field effect in graphene has mainly been exploited in the applications in electronics in which graphene was used as a transistor channel of graphene field-effect transistors (GFETs). However, graphene also forms a Schottky junction with most of the semiconductors5,6, which has previously been used to realize photodetectors7,8,9,10, solar cells11,12,13, sensors14,15, optical modulators16, mixers17, and MESFETs with a graphene gate18,19,20,21,22. The ability of graphene to form a Schottky junction with semiconductors and consequently act as a gate in a MESFET, opens up a perspective of the control of the threshold voltage of FETs. The electrostatic control of the SBH of graphene-semiconductor Schottky junctions has been used in the past to realize barristors23, vertical heterostructures24, and strain sensors25.

Here we demonstrate a semiconductor MESFET with a graphene gate, i.e., a graphene-semiconductor FET (GESFET), in which the carrier density in the GaAs transistor channel was set by a monolayer graphene gate. An additional Al/AlOx control gate was fabricated on top of the graphene gate to externally control the threshold voltage of the transistor. The control gate sets the Fermi level of the graphene gate which adjusts the SBH at the graphene/GaAs interface and therefore the threshold voltage. By changing the voltage between the control and graphene gate, it was possible to change the threshold voltage of the GESFET by up to ~1.4 V. This allowed to change the sign of Vth (e.g., to shift Vth from −0.8 to 0.6 V) in GESFETs with a thin (~250 nm) GaAs channel, i.e., to turn their operation from depletion to enhancement mode after fabrication, which cannot be obtained in conventional MESFETs and HEMTs. The ability to externally control the FET mode was exploited to realize the logic gates with a positive switching threshold, i.e., the logic gates which could be cascaded. The demonstrated GESFETs can also be independently operated by each of the gates as dual-gate transistors. The dual-gate operation allowed the realization of novel functionalities, e.g., frequency up or down-conversion mixers without input resistors and therefore input power dissipation.

Results

GESFET with a single gate

Figure 1a shows the schematic of a GESFET in which the transistor channel between the source (S) and drain (D) was made of n-type GaAs. The gate (G) was made of monolayer graphene grown by chemical vapor deposition (CVD), which was transferred on top of the GaAs channel. The graphene gate was externally accessed through an Au electrode which forms an Ohmic contact with graphene. The depletion layer of the Schottky junction, formed between the graphene gate and GaAs channel, extends into the channel and reduces the thickness of the conductive (i.e., not depleted) part of the channel. This thickness was controlled by the Schottky junction voltage (VGS), which modulates the depletion layer width and therefore the drain current (ID) between the source and drain. For instance, the decrease of VGS increases the depletion layer width and therefore decreases the drain current. The largest modulation of the drain current was obtained when the Schottky junction was reversely biased (VGS < 0 V) due to the weak modulation of the depletion layer width in forward bias (VGS > 0 V)2. The use of the forward bias was also limited by the forward current of the Schottky junction which represents the gate leakage current in this FET. The electrical characteristics of the graphene/GaAs Schottky junction are shown in Supplementary Fig. 1.

Fig. 1: Standard (single-gate) GESFET.
figure 1

a A schematic of a standard GESFET fabricated on an intrinsic GaAs/AlGaAs substrate (orange/yellow/gray) epitaxially grown by molecular-beam epitaxy (MBE). The graphene gate (gray with hexagonal pattern) was transferred on top of the n-GaAs transistor channel (yellow) and was externally accessed through the Au gate pad (transparent yellow). The source and drain contacts (lavender) were evaporated on an n+-GaAs layer (orange) to create an ohmic contact with the n-GaAs channel. b The equilibrium band diagram of a GESFET in the vertical direction, i.e., from the graphene gate (G) to surface states (orange bar) and n-GaAs channel (S). The Fermi level in graphene (EFG) is equal to the Fermi level in the channel (EFS) at VGS = 0 V. The work function of the graphene gate and channel are ΦG and ΦS, respectively. The bottom of the conduction band and the top of the valence band in GaAs are denoted by Ec and Ev, respectively. Vss is the voltage drop on the surface states and ΦB0 is the SBH. c The same band diagram at VGS < 0 V, i.e., when the gate-source voltage is used to deplete the n-GaAs channel. The electrons from the depletion region fill the surface and graphene states (red). The thick dashed line shows the position of the Fermi level in each of the materials. The SBH is reduced to ΦB1 < ΦB0.

At VGS = 0 V, the conductivity of the GESFET channel depends on its thickness. We found that thin channels (<50 nm) were completely depleted at VGS = 0 V. Such GESFETs could only be turned on at positive VGS (provided that the channel is not too thin), i.e., they had Vth > 0 V and operated in the enhancement mode. However, such FETs exhibited poor performance due to the weak modulation of the depletion layer width and large gate leakage current of the Schottky junction under forward bias. The thicker channels were conductive at VGS = 0 V, requiring negative VGS to turn off, i.e., they had Vth < 0 V and operated in the depletion mode.

Figure 1 b shows the band structure of the unbiased (VDS = 0 V) GESFET in the vertical direction at VGS = 0 V. The Fermi level at the interface between graphene and GaAs is to a large extent pinned by the surface states of GaAs due to their large density of states (DOS). The Fermi-level pinning reduces the SBH by eVss to ΦB0, where e is the elementary charge and Vss is the voltage drop at the interface (the formation of the Schottky junction is illustrated in Supplementary Fig. 2). The Schottky barrier depletes the electrons in the n-type GaAs close to the interface.

The electron density in the GaAs channel is controlled by applying VGS < 0 V which reversely biases the Schottky junction and further depletes the channel, as shown in Fig. 1c. Most of the electrons, originating from the depletion region of GaAs, fill the surface states which slightly raises the Fermi level at the surface due to the large DOS of the surface states. The remaining smaller number of electrons transfer to graphene, significantly raising the Fermi level in graphene6,26 (equivalent to the reduction of the work function of graphene) due to its low DOS. This decreases the SBH from the initial value ΦB0 to ΦB1, reducing the carrier depletion in the channel. The reduction of the SBH is therefore a consequence of the limited DOS of graphene and negatively affects transistor operation.

Dual-gate GESFET

To control the SBH (i.e., the work function of graphene), a second control Al/AlOx gate stack is fabricated on top of graphene, resulting in a dual-gate GESFET shown in Fig. 2a. The graphene gate and control Al gate (C) form a capacitor in which the induced charge depends on the capacitor voltage VGC rather than VGS. This prevents the influence of VGS on the carrier density in graphene, i.e., it prevents the filling of the electronic states in graphene by the electrons from the GaAs channel when the Schottky junction is reversely biased (VGS < 0 V). As a consequence, the electrons originating from the depletion region in GaAs fill only the surface states, as illustrated in Fig. 2b for VGC = 0 V. As the surface states have a very large DOS, the SBH is only slightly reduced from ΦB0 to ΦB ≈ ΦB0, i.e., the control gate prevents the lowering of the work function of graphene and deterioration of the transistor properties. In this case, the dual-gate GESFET operates as efficiently as the MESFET with a standard metal gate, in which the SBH also does not depend on VGS. The mean SBH was estimated to ΦB ≈ 0.75 eV, as shown in Supplementary Figs. 35. Scanning electron microscopy (SEM) images of one of the fabricated dual-gate GESFETs are shown in Fig. 2d, e and its proposed electrical symbol in Fig. 2f.

Fig. 2: Dual-gate GESFET.
figure 2

a A schematic of a dual-gate GESFET which differs from the standard (single-gate) GESFET in having a control gate (C; white) fabricated on top of the graphene gate. b The equilibrium band diagram of a dual-gate GESFET in the vertical direction, i.e., from the control-gate stack, comprising gate metal (C) on top of gate oxide (Ox), to the graphene gate (G), surface states (orange bar) and n-GaAs channel (S). The diagram is drawn for VGC = 0 V (i.e., when the control gate is connected to the graphene gate) and VGS < 0 V (i.e., when the channel is depleted). The SBH is ΦB. c The same band diagram for VGC > 0 V, i.e., when the voltage between the graphene and control gate is used to increase the SBH to ΦB + eVq, where Vq is the voltage drop between graphene and its metal contact. Vox is the voltage drop on the gate oxide, i.e., VGC = Vox + Vq. d A SEM image of a dual-gate GESFET. The top surface of the source and drain contacts is n+-GaAs layer which became rough after annealing. e The enlarged part of the previous image within the enclosed white dashed rectangle. f The proposed circuit symbol of a dual-gate GESFET.

Figure 3a shows the transfer characteristics of a dual-gate GESFET at different drain voltages VDS when the control gate is absent (dashed lines) or connected to the graphene gate (solid lines). At positive gate voltages (VGS > 0 V), the n-GaAs channel is fully undepleted and there is no difference in the drain current between the GESFET without and with the control gate. In the absence of the control gate, the reduction of the SBH, as the gate voltage is decreased, impedes the channel depletion. This results in a large subthreshold swing (Sth = 161 mV dec−1) and very negative threshold voltage (Vth = −2.1 V). However, when the control gate is connected to the graphene gate, SBH is mainly unaffected by the decrease of VGS. This depletes the channel faster compared to the GESFET without the control gate, leading to the smaller subthreshold swing (Sth = 86 mV dec−1) and less negative threshold voltage (Vth = −0.92 V). In both cases (with or without the control gate), the transfer curves are quadratic at large drain voltages (VDS > VGS − Vth) and mainly independent of VDS because the GESFET is in the saturation region. At low drain voltages (e.g., VDS = 0.1 V), the GESFET is in the ohmic region and the drain current is much smaller. The maximum drain current ID = 72 A m−1 was obtained at VGS = 0.6 V which is smaller than that of commercial GaAs MESFETs with the same channel length27. Similarly, the extrinsic carrier mobility was only μext = 162 cm2 V−1 s−1. This is a consequence of the non-optimized metal contacts with the GaAs channel resulting in a large contact resistance Rc = 6.7 kΩ μm corresponding to the contact resistivity ρc ~ 3 × 10−3 Ω cm2, which is much larger than that of commercial devices28. The transconductance, gate leakage currents, and contact resistance in the fabricated GESFETs are discussed in Supplementary Figs. 6 and 7.

Fig. 3: The electrical characteristics of a dual-gate GESFET without the control gate (dashed lines) and with the control gate connected to the graphene gate (solid lines).
figure 3

The channel length L = 1.5 μm and width W = 4.8 μm. a The transfer curves of the GESFET where the drain current ID (normalized by the channel width W) is measured as a function of VGS at VDS ranging from 0.1 to 1.6 V with a step of 0.5 V. The inset shows the same transfer curves in the semi-log scale and the corresponding subthreshold swings. b The output curves of the same dual-gate GESFET with ID measured as a function of VDS at VGS ranging from −1.5 to 0.6 V with a step of 0.3 V.

Figure 3b shows the output characteristics of the same GESFET at different gate voltages VGS. Similar to the transfer curves, the control gate does not make much difference if the channel is not depleted, i.e., for VGS > 0 V. This is because the Fermi level of the graphene gate is inside the valence band in equilibrium, as shown in Fig. 1b. At positive gate voltages, the Fermi level in graphene can only shift deeper into the valence band where the DOS of graphene is large enough to suppress such shift and the increase of the SBH in the absence of the control gate. However, at negative gate voltages, the Fermi level shifts toward the Dirac point at which there is the minimum (zero) of the DOS of graphene. Therefore, as the gate voltage becomes more negative, the Fermi level noticeably shifts toward smaller DOS which decreases the SBH in the absence of the control gate. For this reason, as VGS decreases, the drain current decreases faster in the presence of the control gate, as indicated by the arrows in Fig. 3b. Once the Fermi level crosses the Dirac point (which happens at VGS ~ −0.6 V in Fig. 3b) and enters the conduction band, the further decrease of VGS shifts the Fermi level toward larger DOS, which suppresses the Fermi level shift and the reduction of the SBH in the absence of the control gate. This suppresses further split between the drain currents in the absence and presence of the control gate. However, this effect cannot be seen in Fig. 3b because the channel was depleted by reducing VGS below −0.6 V in the presence of the control gate.

The requirement to connect the control gate to the graphene gate in order to operate a GESFET as efficiently as a conventional MESFET is necessary only in the GESFETs in which the Fermi level is close to the Dirac point (e.g., as in the GESFET in Fig. 3). However, if the Fermi level is deep into the valence band, the DOS of graphene is large enough to suppress the changes of the SBH with VGS. Supplementary Figures 8 and 9 show the electrical characteristics of such a GESFET. It exhibits almost the same subthreshold swing as the conventional MESFET with a Ti/Au gate and turns off before the decrease of VGS shifts the Fermi level close to the Dirac point of graphene. Such GESFETs typically exhibit the smallest subthreshold swing (Sth = 65 mV dec−1 was obtained in this case).

In contrast to a standard MESFET, the work function of a graphene gate in a GESFET can be controlled by the control gate, which is illustrated in Fig. 2c and experimentally confirmed in Supplementary Fig. 4b. When a voltage between the control gate and graphene gate is applied, e.g., VGC > 0 V as in Fig. 2c, the most of this voltage drops on the oxide of the control gate. However, due to the limited DOS of graphene, there is also a voltage drop Vq between the metal contact of the graphene gate and graphene (usually modeled by the quantum capacitance). This lowers the Fermi level in graphene by eVq and consequently increases the SBH by the same amount to ΦB + eVq. Therefore, the increase of VGC (i.e., the decrease of the control gate potential VC below the potential of the graphene gate VG), increases the threshold voltage Vth because it increases the depletion of carriers in the channel and decreases the drain current.

The influence of the control gate on the threshold voltage is illustrated in Fig. 4 which shows that the increase of VGC shifts the transfer curve, i.e., the threshold voltage Vth, to larger gate voltages. By increasing VGC from 0 to 1.8 V, the threshold voltage was increased from −0.76 to 0.56 V, as shown in the inset of Fig. 4 (the corresponding leakage currents are shown in Supplementary Fig. 10). This means that the control gate of the GESFET can be used to change the transistor operation from the depletion to the enhancement mode, representing an advantage with respect to the conventional MESFETs. The latter can only operate in one mode after the fabrication, typically the depletion mode. In this case, VGS of the opposite sign of VDS is required to fully deplete the channel (e.g., in an n-type MESFET, VGS < 0 V is required at VDS > 0 V). That limits the application of such MESFETs in digital electronics because the opposite signs of VGS and VDS prevent the direct cascading of transistors which is required in logic gates.

Fig. 4: The transfer curves of a dual-gate GESFET at different VGC and VDS = 1 V.
figure 4

The leftmost curve (black) corresponds to a GESFET in which the control gate is disconnected. Starting from the next curve (purple) and going to the right, VGC increases from 0 to 1.8 V in the steps of 0.2 V. The inset shows the threshold voltage Vth of the GESFET (obtained from the transfer curves), which increases as VGC increases and becomes positive for VGC ≥ 1.2 V. Approximately, Vth = 0.76VGC − 0.83 V.

Logic gates with dual-gate GESFETs

The presented depletion mode GESFETs can overcome the inability of conventional depletion mode MESFETs and HEMTs to be cascaded in logic gates. This is illustrated in Fig. 5 which shows the static voltage transfer characteristics of a depletion-load inverter in which the threshold voltage of the driver GESFET was controlled by the voltage of its control gate. Both GESFETs have a negative threshold voltage when the control gate is connected to the gate (VGC = 0 V). A negative threshold voltage is required for the load transistor to keep it in the on state when its gate is connected to its source. However, a negative threshold voltage of the driver transistor results in a negative switching threshold (VM = −0.62 V) of the inverter at a positive voltage supply VDD = 2.5 V, as shown in Fig. 5. In this case, the logic zero corresponds to a negative voltage at the input (VIN < VM) even though the output voltage is always positive (0 V < VOUT < VDD). In order to shift the switching threshold to positive voltages, VGC > 0 V is required, as discussed in Fig. 4. The increase of the switching threshold VM with the increase of VGC is demonstrated in Fig. 5b. For example, the switching threshold was shifted to VM = 0.3 V at VGC = 2 V.

Fig. 5: A depletion-load inverter with an externally controlled switching threshold.
figure 5

a A schematic of a depletion-load inverter realized with two dual-gate GESFETs. The top GESFET (load) is kept in the on state by having both gates connected to the source. The bottom GESFET (driver) has the control gate connected to a negative power supply voltage −VCC so that its gate (i.e., the input voltage VIN) is always at the higher potential. This shifts the threshold voltage of the driver to positive values if VCC is sufficiently large. b The static voltage transfer characteristics of the inverter for different VGC (the inverter shown in Supplementary Fig. 11) and VCC (the present inverter). The switching threshold VM of the inverter is the input voltage at which VOUT = VDD/2. Logic 0 at the input corresponds to VIN < VM and logic 1 to VIN > VM. The corresponding voltage gain is shown in Supplementary Fig. 12.

In realistic applications, setting VGC > 0 V requires a voltage source to be connected between the gate and the control gate of the driver transistor (see Supplementary Fig. 11), which is impractical because this would require each driver GESFET to be connected to a separate voltage source in an integrated circuit (IC). This problem could be overcome simply by connecting the control gates of all driver GESFETs in the IC to a negative power supply voltage −VCC < 0 V, as shown in Fig. 5a, thereby ensuring VGC = VIN + VCC > 0 V for a large VCC. This is demonstrated in Fig. 5b in which the switching threshold VM was increased with the increase of VCC. At low VCC, the switching threshold was more negative compared to VGC = 0 V (e.g., VM = −1 V at VCC = 0.5 V) because a negative input voltage makes VGC < 0 V when VCC is small. However, as VCC was increased, VM increased too and it reached VM = 0.7 V at VCC = 2 V.

The use of a negative supply −VCC allows to shift the switching threshold of the investigated logic gates to positive voltages which is required in practical applications. However, larger shifts are obtained at larger VCC, which leads to the larger voltage drop VGC = VIN + VCC on the oxide of the control gate. This could result in the oxide breakdown when VIN = VDD (logic 1 at the input). For example, if the present inverters were to be cascaded, VM = VDD/2 = 1.25 V, i.e., VCC > 2 V at VDD = 2.5 V would be required. This problem could be mitigated by simply reducing VDD. However, this was not possible here due to the limited transconductance gm of the GESFETs (typically, gm/W < 100 S m−1) because of the large contact resistance of the used non-optimized contacts. As an alternative, a gate oxide of a similar thickness but larger relative dielectric constant (e.g., HfO2) could be used to increase the transconductance. Nevertheless, all investigated inverters exhibit an over-unity voltage gain, as shown in Supplementary Fig. 12.

Frequency mixers with dual-gate GESFETs

A control gate could also be used as the transistor gate in its own right. Supplementary Figure 13 shows the electrical characteristics of a dual-gate GESFET in which the control gate voltage VCS is used to modulate the drain current at a fixed VGS. In this case, the depletion of the GaAs channel is solely controlled by the change of the SBH. A transistor with two equally functioning gates could be used to simplify more complex electronic circuits by reducing their component count and therefore the space they take in an IC. For example, a frequency mixer typically consists of a nonlinear device (a diode or transistor) whose nonlinearity is exploited to mix the signals at the two inputs of the mixer1. In an ambipolar resistive mixer based on GFETs, the inputs are often connected to the gate of a GFET via resistors29. The resistors take a valuable space in an IC and also lead to an unwanted input power dissipation (because the resistors create a current path between the inputs). Such frequency mixers can be simplified by a dual-gate GESFET in which the inputs are directly connected to its gates thereby completely eliminating the input resistors and the corresponding power dissipation.

Figure 6 shows a simple upconversion mixer that mixes the baseband data signal at the intermediate frequency (IF) with the high-frequency signal of the local oscillator (LO). The output radio-frequency (RF) signal consists of the IF signal shifted to the LO frequency. The load resistor RL was chosen to fit the resistance of the GESFET at the DC operating point, so that the DC voltage at the output is VDD/2. The conversion loss of the mixer is 17.5 dB, which is comparable to that of the best graphene mixers at such low LO power30, even though the present mixer was measured in a DC probe station with an externally connected load resistor. The same mixer can also downconvert the RF signal if the RF and IF ports are swapped, see Supplementary Fig. 15. Further investigations of GESFET-based mixers should be performed to reach the performance of the conventional single-gate mixers.

Fig. 6: An upconverting frequency mixer based on a dual-gate GESFET.
figure 6

a A schematic of the mixer. The output RF signal was obtained by mixing the input IF and LO signals. b The power spectrum of the input IF and output RF signals. The input IF signal is represented with a sine wave at a frequency fIF = 1 kHz, while the LO frequency is fLO = 100 kHz. The output RF signal comprises a signal at fLO, lower sideband at fRF = fLO − fIF = 99 kHz, higher sideband at fRF = fLO + fIF = 101 kHz, and other harmonics of much smaller power. The conversion loss of 17.5 dB represents a power difference between the IF and RF sideband signals. The corresponding waveforms are shown in Supplementary Fig. 14.

Discussion

We demonstrated single and dual-gate GaAs FETs with a monolayer graphene gate. In both types of GESFET, the conductivity of the GaAs channel was modulated by changing the width of the depletion layer of the Schottky junction formed between the graphene gate and GaAs channel. Dual-gate GESFETs had an additional control gate which was fabricated on top of the graphene gate and used to control the threshold voltage of the GESFETs. The control gate sets the Fermi level of the graphene gate and therefore the SBH, which in turn sets the threshold voltage of the GESFETs. This allowed to change the threshold voltage of the dual-gate GESFETs from ~ −0.8 V (depletion mode) to ~0.6 V (enhancement mode) by changing the control gate voltage by 1.8 V.

The control of the threshold voltage can be performed independently on each dual-gate GESFET in a circuit. This was demonstrated by a depletion-load inverter in which the driver GESFET was operated in the enhancement mode, resulting in a positive switching threshold of the inverter. The control gate can also be connected to the graphene gate to reduce the subthreshold swing, or operated independently to provide an additional functionality. The latter was demonstrated by realizing simple frequency mixers in which two different signals were applied to each of the gates.

The presented technology can be applied to any other semiconductor material (e.g., Si, InGaAs, and InP) which makes a Schottky junction with graphene31. The figures of merit of the GESFETs are expected to be mainly affected by the choice of the channel material and their dual-gate configuration. For example, the operating voltage and power dissipation are directly related to the bandgap of the channel material. However, slightly larger power dissipation is expected in GESFETs (with respect to MESFETs) because they have the additional control gate resulting in an additional leakage current at large operating voltages. Similarly, the device speed (or bandwidth) is directly related to the carrier mobility in the channel, but we expect that the ultimate speed will be limited by the resistance of the graphene gate at very short gate lengths (L < 100 nm).

The demonstrated dual-gate GESFETs with a graphene gate pave the way for the development of FETs in which the threshold voltage can be varied almost arbitrarily after fabrication (i.e., during device operation) thus satisfying the demand of different range of applications.

Methods

Fabrication of the transistor channel

A stack of semi insulating AlGaAs, lightly n-doped GaAs (~1017 cm−3), and n+-doped GaAs (>1018 cm−3) was grown by MBE on an intrinsic (100) GaAs wafer. Electron-beam (e-beam) lithography (Raith eLINE at 10 kV) was used to define the mesa structure. A solution of H3PO4 was used to remove n+-GaAs above the transistor channel. Source and drain contacts were patterned by e-beam lithography. The contacts comprising a stack of Au/Ge/Ni/Au (20/75/17/100 nm) were deposited by e-beam evaporation (at a base pressure of 10−6 mbar) followed by annealing in an Ar/H2 atmosphere (for 20 s at 325 C).

Fabrication of the graphene gate

Graphene grown by CVD32,33 was directly transferred on the chip and then patterned by e-beam lithography to define the gate Schottky junction (with a channel length 1 μm < L < 15 μm). O2 plasma was used to remove the excess graphene. The gate contact was patterned by e-beam lithography on graphene outside the channel area. It consisted of pure Au (100 nm) deposited by e-beam evaporation.

Fabrication of the control gate

The dual-gate GESFETs were fabricated from standard (single-gate) GESFETs after they were electrically characterized. The control gate was patterned by e-beam lithography on graphene covering the channel. It consisted of Al (100 nm) evaporated by e-beam lithography. Al oxidized in air forming a thin (~4 nm) native layer of AlOx on all surfaces of Al, including the surface in contact with graphene. This formed a control Al/AlOx gate stack on top of graphene with a gate capacitance of Cox = 1.4 μF cm−2. 34,35To demonstrate that the observed effects are not related to a possible direct contact between Al and the GaAs channel, we also fabricated reference devices without the graphene gate, see Supplementary Fig. 16.

Device characterization

The electrical measurements were performed in FormFactor probe stations Summit 11000 and EP6 in air ambient. Keithley 2600B series source-measure units, a function generator Tektronix AFG 3022B, and an oscilloscope Keysight DS09064A were used in electrical characterizations of the realized GESFETs, inverters, and mixers. The capacitance–voltage (CV) measurements were performed using Keithley 4200A-SCS parameter analyzer with a CV unit 4215-CVU. The SEM characterization was performed in Raith eLINE at 10 kV.