We report the effect of remote surface optical (RSO) phonon scattering on carrier mobility in monolayer graphene gated by ferroelectric oxide. We fabricate monolayer graphene transistors back-gated by epitaxial (001) Ba0.6Sr0.4TiO3 films, with field effect mobility up to 23 000 cm2 V−1 s−1 achieved. Switching ferroelectric polarization induces nonvolatile modulation of resistance and quantum Hall effect in graphene at low temperatures. Ellipsometry spectroscopy studies reveal four pairs of optical phonon modes in Ba0.6Sr0.4TiO3, from which we extract RSO phonon frequencies. The temperature dependence of resistivity in graphene can be well accounted for by considering the scattering from the intrinsic longitudinal acoustic phonon and the RSO phonon, with the latter dominated by the mode at 35.8 meV. Our study reveals the room temperature mobility limit of ferroelectric-gated graphene transistors imposed by RSO phonon scattering.
I. INTRODUCTION
Leveraging its high mobility, superb mechanical strength, and optical transparency, extensive research has been carried out on graphene-based field effect transistor (FET) devices1,2 for developing radio-frequency transistors,3,4 nonvolatile memory,5 flexible electronics,6 and optoelectronics.7,8 As the electronic properties of graphene are highly susceptible to the interfacial dielectric environment due to its two-dimensional (2D) nature, the choice of the gate and substrate materials for graphene FETs can have a significant impact on the device performance.9 For example, the gate can be a major source of charged impurities10 as well as providing dielectric screening.11–13 The interfacial charge dynamics can induce undesired switching hysteresis and compromise the device retention.9,14 One important phenomenon is the remote surface optical (RSO) phonon from the dielectric layer, which is the major mechanism that limits room temperature mobility15–17 and saturation current9,18,19 in the graphene channel.
Recently, the ferroelectric/2D van der Waals heterostructure has emerged as a promising platform for developing high-performance logic, memory, and optical applications.14,20 Ferroelectrics possess nonvolatile switchable polarization with high doping capacity21 and exhibit second harmonic generation22 as well as negative capacitance effect,23 which can be utilized to design novel functionalities in the 2D devices, such as electron collimation,24 nonlinear optical filtering.,25 steep slope switching,26 and neuromorphic computing.27 The widely explored ferroelectric systems include ferroelectric oxides,6,8,28–30 polymers,5,31,32 and 2D semiconductors.20 Among them, the ferroelectric perovskite oxides have the distinct advantages of large bandgap, large polarization, high-κ dielectric constant, high endurance, low coercive field, and fast switching time.14 For electronic applications, it is important to understand the effect of interfacial ferroelectric layer on the channel mobility. The high-κ nature of ferroelectric oxides implies the presence of soft optical phonon modes, which can significantly affect the mobility of the 2D channel at room temperature.12,33 This effect is, especially, prominent in monolayer graphene (1LG) due to its linear dispersion.9,34 Despite the rapid progress in developing ferroelectric/graphene-based applications, the effect of RSO phonon on the electronic and thermal transport of the device has not been examined to date.
In this work, we report a comprehensive magnetotransport study of graphene FETs back-gated by ferroelectric Ba0.6Sr0.4TiO3 (BSTO) thin films. The devices exhibit high field effect mobility up to 23 000 cm2 V−1 s−1 and nonvolatile field effect modulation of quantum Hall effect. Four RSO phonon modes have been identified by considering the optical phonon modes in BSTO and the dielectric boundary condition. The temperature dependence of resistivity in graphene points to the dominant effect of the 35.8 meV RSO phonon at high temperature. Our study reveals the room temperature mobility limit in graphene imposed by the ferroelectric oxide gate, providing important material information for designing high-performance ferroelectric/graphene FETs for nanoelectronic applications.
II. SAMPLE PREPARATION AND CHARACTERIZATION
We work with epitaxial 100 and 300 nm (001) BSTO thin films deposited on Nb-doped SrTiO3 (Nb:STO) substrates using off-axis RF magnetron sputtering. The films are deposited in 25 mTorr process gas of Ar and O2 (ratio 2:1) at 600 °C. X-ray diffraction (XRD) studies show that these films are single crystalline with a c-axis lattice constant of 3.99 Å [Fig. 1(a)]. Atomic force microscopy (AFM) images reveal smooth surface morphology with a typical root-mean-square roughness of 3–4 Å [Fig. 1(a) inset]. Thin graphite flakes (kish graphite from Sigma-Aldrich®) are mechanically exfoliated on BSTO. The monolayer flakes are identified optically and characterized via Raman spectroscopy. Figure 1(b) compares the Raman spectra of graphene on BSTO and bare BSTO. There is no appreciable D band observed in the 1LG sample, indicating a lack of atomic defects. From the Lorentz fits to each peak, we deduce a large 2D to G band intensity ratio , which is comparable to those observed in high-mobility graphene sandwiched between SiO2 substrates and HfO2 top-layers,17 confirming the high quality of our samples.35 Selected flakes are fabricated into Hall bar devices via electron-beam lithography followed by electron-beam evaporation of 5 nm Cr/25 nm Au as the electrodes [Fig. 1(c)]. The conductive Nb:STO substrate serves as the back-gate electrode for field effect studies [Fig. 1(d)]. We perform variable temperature magnetotransport measurements in a Quantum Design PPMS using standard lock-in technique (SR830) at an excitation current of 50 nA. The results are based on three graphene samples (denoted as devices A, B, and C).
(a) XRD θ–2θ scan of a 100 nm BSTO film on Nb:STO. Inset: AFM topography image. (b) Raman spectra of 1LG on BSTO normalized to the G band intensity and bare BSTO. (c) Optical image of device A, with the graphene edge outlined (dotted line). (d) Device schematic. (e) vs 1/B at 10 K with the background resistivity subtracted, and (f) the corresponding semi-log plot of vs 1/B. The dashed lines are fits to Eq. (1).
(a) XRD θ–2θ scan of a 100 nm BSTO film on Nb:STO. Inset: AFM topography image. (b) Raman spectra of 1LG on BSTO normalized to the G band intensity and bare BSTO. (c) Optical image of device A, with the graphene edge outlined (dotted line). (d) Device schematic. (e) vs 1/B at 10 K with the background resistivity subtracted, and (f) the corresponding semi-log plot of vs 1/B. The dashed lines are fits to Eq. (1).
The BSTO-gated graphene FETs show high mobility compared with those gated by SiO2. Figure 1(e) shows the Shubnikov–de Haas oscillation of the longitudinal resistivity taken on device A at 10 K and the back-gate voltage of Vg = −1 V. The oscillation period corresponds to a charge density of n = 5.36 × 1012 cm−2. The oscillatory amplitude is given by10
where is the zero-field resistivity, is the thermal factor, is the cyclotron frequency with e being the elementary charge, and is the quantum scattering time. Here, is the effective mass in monolayer graphene, with being the Fermi velocity. Fitting vs 1/B reveals [Fig. 1(f)], while the extracted Hall mobility , corresponding to a transport scattering time of . These results are comparable to previous reports for graphene on BSTO29 and SiO2.10 The large ratio between the transport and quantum scattering times indicates that the mobility is dominated by small-angle scattering events, e.g., from charged impurities residing within the BSTO substrate.10 This ratio is larger than those reported for graphene gated by HfO2 and SiO29 and agrees well with the theoretical value for long-ranged scatterers considering the dielectric screening of BSTO.14
Figure 2(a) shows of device A measured at 2 K. We observe a hysteresis between the up-sweep and down-sweep curves, which corresponds to ferroelectric polarization switching.14 The Dirac points locate at Vg = 2.4 V for the up-sweep state and Vg = 0.9 V for the down-sweep state. To determine the carrier density n, we characterize the Hall resistivity [Fig. 2(b) inset] at different back-gate voltages and deduce the Hall coefficient , with . We denote the carrier density in electron- and hole-doped regions as and , respectively. Figure 2(b) plots vs Vg, where we identify three distinct regions for the up-sweep branch associated with different polarization states of BSTO. When BSTO is uniformly polarized in Pup (region III) and Pdown (region I) states, it behaves as a normal high-κ dielectric. In the intermediate Vg-range (region II), we observe a gradual change of carrier density with a steeper slope in n(Vg). As BSTO is a relaxor, the polarization switching is associated with the alignment of polar nanoregions,36 in contrast to the abrupt switching in canonical ferroelectrics such as Pb(Zr,Ti)O3. The induced polarization corresponds to the enhanced doping efficiency. Similar behaviors are also observed in the down-sweep branch. Around the Dirac point, we fit the conductivity by , where μFE is the field effect mobility and ρshort is the resistivity due to short-ranged scatterers. Figure 2(a) inset shows the fits to the up-sweep branch, which yields μFE of 4700 cm2 V−1 s−1 for holes and 23 000 cm2 V−1 s−1 for electrons. For the down-sweep branch, we extract μFE of 2700 cm2 V−1 s−1 for holes and 10 300 cm2 V−1 s−1 for electrons. Figure 2(c) shows of this device at 8.9 T, which exhibits well-developed quantum Hall states in both branches. The filling factors’ sequence corresponds to 4(n + ½), which is the signature behavior of monolayer graphene.
Magnetotransport studies of device A. (a) ρxx(Vg) at 2 K. The arrows label the Vg sweeping direction. Inset: σ vs with fits (dashed lines), with nh and ne denoting the hole- and electron-doped regions, respectively. (b) 1/eRH vs Vg at 2 K. Inset: ρXy vs B at Vg = −1 V for up-sweep with a linear fit (dashed line). (c) ρxx(Vg) at 10 K and 8.9 T. The filling factors for the down-sweep branch are labeled.
Magnetotransport studies of device A. (a) ρxx(Vg) at 2 K. The arrows label the Vg sweeping direction. Inset: σ vs with fits (dashed lines), with nh and ne denoting the hole- and electron-doped regions, respectively. (b) 1/eRH vs Vg at 2 K. Inset: ρXy vs B at Vg = −1 V for up-sweep with a linear fit (dashed line). (c) ρxx(Vg) at 10 K and 8.9 T. The filling factors for the down-sweep branch are labeled.
III. TEMPERATURE DEPENDENCE OF RESISTIVITY
To investigate the temperature dependence of resistivity in BSTO-gated graphene, we conduct Hall measurements at different temperatures to convert Vg to n. Here, we choose to work with the hole-doped region for the up-sweep branch, where we have access to the largest density range. Figure 3(a) shows n vs Vg at 80 K. BSTO exhibits a linear n(Vg) relation, , with being the gating efficiency. We extract the dielectric constant of BSTO , where d is the BSTO film thickness and is the vacuum permittivity. The temperature dependence of [Fig. 3(b)] is consistent with previous report.29 We then shift the charge neutral point at different temperatures to and convert the resulting [Fig. 3(c) inset] to ρxx(n) based on the gating efficiency [Fig. 3(c)]. At low n, ρxx increases with decreasing temperature, which can be attributed to thermally activated charge carriers in the electron–hole puddle region.37 At high doping level, the sample exhibits metallic T-dependence associated with phonon scattering.15,17
(a) n(Vg) at 80 K with a linear fit (dashed line). (b) vs T for 100 nm (solid dots) and 300 nm (open circles) BSTO. (c) ρxx vs −n at 20–200 K with 20 K intervals. Inset: color map of .
(a) n(Vg) at 80 K with a linear fit (dashed line). (b) vs T for 100 nm (solid dots) and 300 nm (open circles) BSTO. (c) ρxx vs −n at 20–200 K with 20 K intervals. Inset: color map of .
In graphene, the temperature dependence of resistivity can be modeled as15
Here, is the residual resistivity due to impurity scattering, which is temperature-independent; is associated with the longitudinal acoustic (LA) phonon scattering intrinsic to graphene; and is associated with RSO phonon scattering from the interfacial dielectric layer.
A. Effect of LA phonon
In the nondegenerate equipartition acoustic phonon system,38 the LA phonon contribution depends linearly on temperature,15,28
where DA is the acoustic deformation potential, is the areal mass density of graphene, and is the sound velocity for LA phonons in graphene. Equation (3) can well describe the low temperatures data of our samples, where ρxx exhibits a linear T-dependence that is independent of n. From the slopes of low temperature ρxx(T) for devices A–C, we obtain DA = 20 ± 6 eV, which is within the range of previous reports (10–30 eV).9
B. Effect of RSO phonon
With increasing temperature, ρ(T) becomes nonlinear and highly dependent on n, which can be attributed to the onset of RSO phonon contribution . The RSO phonon effect has previously been studied in graphene interfaced with SiO2,15 Al2O3,16 and HfO2.17 Under relaxation time approximation,12,34,39 can be expressed as the sum of independent contributions from individual RSO phonon modes,
where is the contribution from ith RSO phonon mode . Here, is the matrix element for scattering between electron (k) and phonon (q) states; is the corresponding electron-phonon coupling strength,17 with being the ith intermediate dielectric constant depending on the frequency-dependent dielectric function of BSTO. At , we can assume and are temperature-independent and, thus, decouple the density and temperature dependences of resistivity as , where captures the density dependence of resistivity and is the unitless magnitude of .
We first consider the temperature dependence of . Phenomenologically, can be expressed by the generalized Lyddane–Sachs–Teller (LST) relation,
where is the optical permittivity and and are the frequencies of the ith transverse optical (TO) and longitudinal optical (LO) phonon modes, respectively. The independent broadening parameters, and , account for anharmonic lattice interactions. From Eq. (5), in the case of zero phonon broadening, the dielectric function diverges to infinity at the TO modes and approaches zero at the LO modes. The intermediate dielectric constants are defined by rewriting the real part of the LST relation [Eq. (5)] into the form of lossless Lorentzian oscillator approximation,
where is defined as . From the deduced , we can calculate the electron–phonon coupling strength . For graphene sandwiched between BSTO and vacuum, the RSO modes can be determined by matching the dielectric boundary condition .33,39
To obtain and , we carry out spectroscopic ellipsometry to study the frequencies of the optical phonon modes and dielectric properties of BSTO. Two ellipsometer apparatuses are used for measurements performed at room temperature. A commercial variable angle of an incidence spectroscopic ellipsometer (IR-VASE Mark-II; J.A. Woollam Co., Inc.) is employed for the infrared spectral range (650–3000 cm−1). An in-house built instrument is used for the far-infrared spectral range (50–650 cm−1).40 Ellipsometry data are obtained in the notation, and the measurements are taken at multiple angles of incidence.41 A bare SrTiO3 substrate is measured in addition for comparison with model calculations using literature values for phonon mode parameters.42 The ellipsometry data are analyzed for the phonon mode properties of the epitaxial thin film. A three-phase (substrate-film-ambient) model is established, where the film is modeled by the BSTO layer thickness. The far-infrared and infrared dielectric function is modeled using Eq. (5), where the LO phonon mode frequency can be directly obtained from the best-match model analysis. The broadening parameters in the best-match model calculations account for finite absorption losses or phonon damping near TO and LO modes. A detailed discussion of the phonon mode analysis can be found in Refs. 43–45. The BSTO dielectric function is modeled with four phonon mode pairs, . The calculated and data using the above model description are compared against the measured ellipsometry data, while the model parameters are varied until reaching a best match between the calculated and measured data. Parameters for variation are BSTO optical permittivity and film thickness, TO and LO phonon frequencies, and their broadening parameters.
Figure 4(a) shows the frequency dependence of the real part of the complex dielectric function of BSTO . This function is calculated from Eq. (5) by setting all broadening parameters to zero and using the TO and LO mode parameters obtained from the best-match model calculation using the measured ellipsometry data as target values. In the frequency range of interest, we identify four pairs of TO and LO phonon modes: , ; , ; , ; , and . By solving the dielectric boundary condition, we deduce four RSO phonon modes, with each one located between a pair of TO and LO modes: , , , and . The corresponding coupling strength is calculated based on the intermediate dielectric constants deduced from Eq. (6): , , , and . Figure 4(b) plots the individual contribution of each RSO phonon scattering to the temperature dependence of resistivity in graphene: . At T > 50 K, it is clear that is dominated by the mode. The lowest phonon mode has a negligibly small , as revealed by the dielectric function spectrum so it does not couple strongly to the graphene channel. As a result, even though it has the highest excited phonon population, its contribution to resistivity is insignificant.
(a) Real function of for BSTO simulated based on the ellipsometry data. (b) Simulated for four RSO phonon modes.
(a) Real function of for BSTO simulated based on the ellipsometry data. (b) Simulated for four RSO phonon modes.
Figure 5 shows ρxx(T) of device A at various carrier densities beyond the electron–hole puddle region and the corresponding fitting curves using Eq. (2). We first consider the contributions from all four RSO phonons to models. Based on the RSO phonon frequencies and their corresponding coupling strength , Eq. (4) can be rewritten as
ρxx(T) at selected hole densities taken on device A with fits to Eq. (2), with given by (a) Eq. (7) (solid lines) and (b) Eq. (8) (dashed lines).
As shown in Fig. 5(a), this model well captures the temperature dependence of resistivity at all densities. We then consider fitting ρ(T) using a single effective Bose–Einstein distribution with the dominant phonon frequency of 35.8 meV [Fig. 5(b)],
where is the fitted density dependent resistivity. This model also yields excellent fit to the experimental results for devices A [Fig. 5(b)], B, and C. These results further confirm the dominating role of the 35.8 meV RSO phonon mode in determining the high temperature resistivity of graphene on BSTO, consistent with the simulation result in Fig. 4(b).
Figure 6(a) shows the density dependence of the coefficient C(n) extracted from these three devices. All samples exhibit a power-law density dependence , which can be attributed to the long-range nature of RSO phonon scattering. The exponent β ranges from 0.9 to 1.3, consistent with the values reported in graphene interfaced with SiO215 and HfO2.17 This value is higher than the predicted β = 1/2 for Thomas–Fermi approximation for electron screening.34 The stronger n-dependence has been attributed to the finite-q corrections to the scattering matrix element |Hkk′|2.33,34
(a) Resistivity coefficient C(n) vs carrier density n on devices A–C with fits to n−β (dashed lines). Here, β = 1.3 (device A), 0.9 (device B), and 1.2 (device C). (b) Mobility limit at 300 K from scattering imposed by LA and RSO phonon modes in device A.
(a) Resistivity coefficient C(n) vs carrier density n on devices A–C with fits to n−β (dashed lines). Here, β = 1.3 (device A), 0.9 (device B), and 1.2 (device C). (b) Mobility limit at 300 K from scattering imposed by LA and RSO phonon modes in device A.
C. Modeling of room temperature mobility
Based on the fitting results, we calculate the mobility limit of graphene on BSTO imposed by the LA phonon and various RSO phonon modes using . Figure 6(b) summarizes the mobilities at 300 K from the individual phonon scattering mechanisms using the parameters extracted from device A. The LA phonon limited mobility has dependence on the density, which gives at the doping level of interest. Among all RSO phonons, the 35.8 meV mode is the dominating scattering source at 300 K, yielding . Combining all phonon contributions, the overall mobility is around and exhibits very weak density dependence. This result is about twice of the room temperature mobility limit for HfO2-gated graphene,17 despite the significantly higher doping capacity of BSTO.
IV. CONCLUSIONS
In conclusion, we have investigated the effect of RSO phonon scattering in graphene FET back-gated by ferroelectric BSTO. The high-temperature resistance is dominated by the 35.8 meV RSO phonon mode, which limits the room temperature mobility of graphene FET to about . The BSTO gate provides efficient dielectric screening, high doping capacity, and the promise for local doping control via nanoscale domain patterning, while its impact on the room temperature mobility in graphene FET is highly competitive compared with high-κ dielectrics such as HfO2. Our study sheds light on the technological potential of ferroelectric perovskite oxide as the gate material for graphene-based nanoelectronics.
ACKNOWLEDGMENTS
We thank Dawei Li for valuable discussions. This work was primarily supported by the U.S. Department of Energy (DOE), Office of Science, Basic Energy Sciences (BES), under Award No. DE-SC0016153 (graphene FET preparation and characterization). Additional support was provided by NSF through Grant Nos. DMR-1710461 (preparation of BSTO), and EPSCoR RII Track-1: Emergent Quantum Materials and Technologies (EQUATE), Award No. OIA-2044049 (ellipsometry studies and data modeling). The research was performed, in part, in the Nebraska Nanoscale Facility: National Nanotechnology Coordinated Infrastructure, the Nebraska Center for Materials and Nanoscience, which are supported by NSF ECCS: 2025298, and the Nebraska Research Initiative.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose
Author Contributions
Hanying Chen: Data curation (lead); Formal analysis (equal); Validation (lead); Writing – original draft (lead). Tianlin Li: Formal analysis (supporting); Validation (supporting); Writing – review & editing (equal). Yifei Hao: Data curation (supporting); Formal analysis (supporting); Writing – review & editing (supporting). Anil Rajapitamahuni: Data curation (equal); Writing – review & editing (supporting). Zhiyong Xiao: Formal analysis (equal); Writing – review & editing (supporting). Stefan Schoeche: Data curation (equal); Writing – review & editing (equal). Mathias Schubert: Funding acquisition (supporting); Supervision (equal); Writing – review & editing (supporting). Xia Hong: Conceptualization (lead); Formal analysis (equal); Funding acquisition (lead); Supervision (lead); Writing – original draft (equal); Writing – review & editing (equal).
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.