This study demonstrates a new resistive switching material, F-doped TiO2 (F:TiO2), fabricated by atomic layer deposition (ALD) with an in-house fluorine source for resistive random access memory (RRAM) devices. Controlling oxygen vacancies is required since RRAM uses resistive switching (RS) characteristics by redistributing oxygen ions in oxide, and poor oxygen defect control has been shown to significantly reduce RRAM reliability. Therefore, this study designed an F based RRAM device using fluorine anions rather than oxygen defect for the main agent of RS behavior. We developed the F:TiO2 RRAM material using a novel in situ doping method in ALD and investigated its RS behaviors. The Pt/F:TiO2/Pt device exhibited forming-less bipolar RS and self-rectifying behavior by fluorine anion migration, effectively reducing the sneak current in crossbar array architecture RRAM. The doped fluorine passivated and reduced oxygen related defects in TiO2, confirmed by x-ray photoelectron spectroscopy analysis. Adopting the F-based RS material by ALD provides a viable candidate for high reliability RRAM.

Flash memory has shown unprecedented lithographic development, including double patterning, three-dimensional (3D) vertical processing, data processing for multi-bit operation, and error correction code (ECC) optimization.1–3 However, scaling limitations have also been demonstrated due to leakage current between word line (WL) and bit line (BL), coupling among adjacent cells, reduced saved charge as cell size decrease, etc. Since resistive random access memory (RRAM) exploits resistive switching (RS) behaviors relying on material inner ion movements, it can avoid flash memory problems that store data using charge traps.4,5 Therefore, the RRAM scaling limit is relatively low, promoting RRAM as an ideal candidate for the next-generation non-volatile memory (NVM).6–8 NVMs with a simple two-terminal structure are well-known for their high density using crossbar array architecture (CAA),9–11 such as spin-torque transfer magneto-resistive RAM, phase change RAM, ferroelectric tunnel junction, and RRAM. The CAA structure addresses and operates memory cells by WL and BL for reading and writing, respectively, since each memory cell does not have a selector. Thus, crosstalk issue is induced during operation due to sneak current caused by parallel CAA circuitry;6,12,13 hence, the sneak current increases with increasing array size. Previous studies have proposed various designs to manipulate sneak, including one selector–one resistor (1S–1R) structure14–17 with various selectors, e.g., diodes,6,18 Mott metal–insulator transition materials, potential barrier materials, and complementary resistive switches (CRSs).19,20 However, reliable selector fabrication and their layout complexity undermines the advantages from the compact CAA structure.

ALD enables low-temperature deposition to produce highly uniform thin films.21–23 This process aims to grow thin films by adopting self-limiting surface reactions, simplifying film thickness manipulation. Specifically, repetitive pulsing and purging composes chemical reactions with separately kept precursors while the deposition process continues. Thus, ALD produces thin films at low temperature with several advantages, such as high step coverage, uniformity, and precise film thickness control.24 Separated precursor dosing also limits gas phase reactions, providing not only exceptionally reactive precursor adoption but also extending the time to complete each reaction step. These advantages help improve the scalability and reliability of complex 3D electronic devices. Thus, there is considerable research interest in producing 3D RRAM structures, focusing strongly on reliable resistance material fabrication using ALD.25,26

However, reliable RRAM fabrication using the ALD process is not simple since RRAM RS behaviors require defect control. RRAM materials incorporate considerable oxygen vacancies to induce RS properties by oxygen migration. However, ALD exploits self-limiting surface reactions between reactants during film growth, which makes it fundamentally difficult to form oxygen defects within the film. Precursor dependency for ALD is also an obstacle to develop RS materials. On the other hand, oxygen defect control has been identified as a major mechanism underlying RRAM reliability.27–29 Therefore, this study proposed and fabricated an F based RRAM device using fluorine as an alternative to oxygen defect for the main RS behavior agent.

We formed F:TiO2 thin films by ALD using titanium tetra-isopropoxide (TTIP), deionized (DI) water, and an in-house fluorine source (HF dissolved in water) as precursors for Ti, O, and F reactants, respectively. Figure 1(a) shows the proposed ALD growth procedure for undoped TiO2 films: TTIP (0.3 s) → N2 purge (13 s) → H2O pulse (0.1 s) → N2 purge (13 s), where we pulsed water (H2O) and F source (H2O/HF) at a controlled y:x ratio to incorporate the F dopant into the films. Figure 1(b) shows one super-cycle for Ti–O:Ti–O/F (9:1) samples, i.e., one (x = 1) repetition of the ALD sequence [TTIP (0.3 s) → N2 purge (13 s) → H2O pulse (0.1 s) → N2 purge (13 s)] and nine (y = 9) ALD sequences {TTIP (0.3 s) → N2 purge (13 s) → H2O/HF pulse (0.1 s) → N2 purge (13 s) [Fig. 1(b)]}.

FIG. 1.

Proposed process to grow (a) undoped and (b) and (c) fluorine doped TiO2 thin films. O/F indicates fluorine single source [mixed H2O and dilute HF (48%–51% in H2O)]. In (b), x is the H2O source pulse ratio and y is the fluorine source (H2O/HF) pulse ratio in one super-cycle.

FIG. 1.

Proposed process to grow (a) undoped and (b) and (c) fluorine doped TiO2 thin films. O/F indicates fluorine single source [mixed H2O and dilute HF (48%–51% in H2O)]. In (b), x is the H2O source pulse ratio and y is the fluorine source (H2O/HF) pulse ratio in one super-cycle.

Close modal

Figure 1(c) shows the ALD-F:TiO2 film sequence with a single reactant (H2O/HF) [TTIP (0.3 s) → N2 purge (13 s) → H2O/HF pulse (0.1 s) → N2 purge (13 s)] to maximize the F concentration. The proposed technique provides a new protocol to synthesize oxygen related defects, which passivate TiO2 oxygen defects, with fluorine and induce self-rectifying capacity.

FIG. 2.

High-resolution XPS spectra for undoped TiO2 and F:TiO2 (1.2). (a) Ti 2p spectra (∼458.6 eV); (b) deconvolved of XPS O 1s spectra, OI and OII chemical states indicate the peaks originated from lattice oxygens (∼529.9 eV) and oxygen related defects (∼531.8 eV); and (c) F 1s spectra (∼684 eV).

FIG. 2.

High-resolution XPS spectra for undoped TiO2 and F:TiO2 (1.2). (a) Ti 2p spectra (∼458.6 eV); (b) deconvolved of XPS O 1s spectra, OI and OII chemical states indicate the peaks originated from lattice oxygens (∼529.9 eV) and oxygen related defects (∼531.8 eV); and (c) F 1s spectra (∼684 eV).

Close modal

Fluorine doped titanium oxide thin films were deposited on Pt(111)/Ti/SiO2/Si substrates using a Traveling Wave Lucida D100 system (NCD Technology, Inc. Korea) at 150 °C deposition temperature and ∼1 Torr operating pressure. Titanium tetra-isopropoxide [(Ti(OPri)4 or TTIP] (EG Chemical Co., Ltd., Korea), DI water, and hydrogen fluoride were used as Ti, O, and F precursors, respectively. The mixture of 50 ml DI water and 0.5 ml diluted hydrogen fluoride (HF 48%–51% diluted in water) were used to produce the homemade F source. For uniform dispersion of HF, the solution underwent five-minute stirring before being charged to a canister. Hydrofluoric acid and/or liquid hydrogen fluoride comprised the chemical state of this solution.30 We used a heater to maintain TTIP below 95 °C in high purity N2 (99.999%) atmosphere.

The ALD growth cycle for TiO2 film deposition progressed as TTIP (0.3 s) → N2 purge (13 s) → H2O pulse (0.1 s) → N2 purge (13 s) and for F:TiO2 film deposition as TTIP (0.3 s) → N2 purge (13 s) → HF pulse (0.1 s) → N2 purge (13 s).

After deposition, TiO2 and F:TiO2 films were post-annealed for 1 h at 500 °C. A 100 nm thick, 20 µm diameter Pt top electrode was formed by using an e-beam evaporator and shadow mask.

High-resolution x-ray diffraction (HR-XRD, SmartLab, Rigaku) was used to measure the crystal structure with Cu Kα1 radiation (λ = 1.5416 Å). F:TiO2 thin films were investigated using field emission scanning electron microscopy (FE-SEM, 7610f-plus, Hitachi) at 15 kV operating voltage with atomic force microscopy (AFM, MultiMode 8, Bruker, USA) to examine surface morphologies. Hall effect measurements (Ecopia HMS-3000) allowed calculating carrier concentration (n), resistivity (r), and Hall mobility (μH) using the van der Pauw method at room temperature. We used an ultraviolet–visible–near infrared (UV–vis–NIR) spectrophotometer (V-570, JASCO, 200–900 nm) to obtain optical transmittance spectra and Fourier transform infrared spectrophotometer (FTIR; Perkin Elmer 1760X spectrophotometer, 600–1000 cm−1) to measure Ti–F chemical bonding. X-ray photoelectron spectroscopy (XPS) from an Al Kα monochromator (1486.6 eV) with variable spot size (30–400 µm in 5 µm steps) was used to identify elemental presence on the F:TiO2 thin films (XPS, K-alpha, Thermo UK). RS characteristics for F:TiO2 based RRAM devices were measured at room temperature using a KEITHLEY 2635A Sourcemeter.

Density of states (DOS) was calculated from first principles using density functional theory (DFT) and the CASTEP code.31,32 We employed the Perdow–Burke–Ernzerhof exchange–correlation function with generalized gradient approximation (GGA) for geometrical optimization and GGA+U for DOS calculation.33,34 Pseudopotentials were constructed using the ultrasoft method with cutoff energy = 380 eV and self-consistence field tolerance = 1.0 × 10−6 eV/atom. Brillouin zone integration was performed using 2 × 2 × 2 k-points following the Monkhorst–Pack approximation.35 

Hydrogen fluoride in one super-cycle can be calculated from the HF content in the diluted fluoride source solution, i.e., 0.5 ml 48%–51% diluted HF in 50 ml DI water ≈0.5% HF in the source canister, where HF(%) = [y/(x + y)] × 0.5%.30 Both TiO2 and F:TiO2 (1.2) films were deposited at 150 °C and post-annealed at 500 °C, and then Ti 2p, O 1s, and F 1s were measured using XPS to examine fluorine presence and its relationship with TiO2 to assure surface oxygen vacancy passivation on TiO2.

Figures 2(a)2(c) show Ti 2p2/3, O 1s, and F 1s peaks at of 458.6, 529.9, and 684 eV binding energies, respectively, confirming F doping into TiO2.36,37Figure 2(a) compares the Ti 2p2/3 peaks for undoped TiO2 and F:TiO2 (1.2) films, exhibiting 0.2 eV shift attributed to the Fermi level receding from the conduction band due to oxygen vacancy passivation.36 

Figure S1 (supplementary material) shows XPS spectra of all F:TiO2 (0, 0.5, 1.0, and 1.2) films, with corresponding DOS for TiO2 (with oxygen vacancy) and F:TiO2 (passivated oxygen vacancy) in Fig. S2 (supplementary material). Valence and conduction bands for TiO2 and F:TiO2 mainly came from the O p-orbital and Ti d-orbital, respectively. However, the TiO2 Fermi level in the middle of the bandgap and the electron state located around the Fermi level can be attributed to oxygen defect states. The Fermi level moved to the top of the valence band for F:TiO2 compared with TiO2, the and oxygen vacancy defect energy level exhibited in TiO2 vanished in F:TiO2 due to oxygen vacancy passivation by F doping.

Figure 2(b) shows O 1s XPS and Gaussian fits to identify how oxygen and fluorine doping impacted oxygen vacancy passivation. Two constitutional O 1s peaks at ∼529.9 eV (OI) correspond to the Ti–O binding energy, whereas the peak at ∼531.8 eV (OII) corresponds to oxygen vacancies and non-lattice oxygen.37 Reduced OII peaks verify that fluorine doping passivated oxygen related defects. Figure S3 (supplementary material) shows that the peak area ratio rose from 5.9 to 15.17 eV after fluorine doping on the TiO2 latticeoxygen/defectoxygen. Thus, fluorine successfully passivated TiO2 oxygen related defects and non-lattice oxygen.

Figure 2(c) shows the F 1s peak for F:TiO2, confirming fluorine presence in F:TiO2 thin films. The peak at 684 eV corresponds to Ti–F binding energy. The F 1s binding energy intensity (peak at 684 eV) occurs above the 45 HF pulse portion. Table I shows that the F proportion in F:TiO2 thin films increased with increasing HF pulse portion, reaching 1.2 % in the TiO2 matrix. Fluorine proportion was calculated from XPS spectra for TiO2 and F:TiO2 films deposited on Pt substrates.

TABLE I.

Measured TiO2 thin film sample characteristics.

HF pulse portion inXPS (at. %)
SampleCycle ratio (Ti–F:Ti–O)Growth temp. (°C)1 super-cycle (%)TiOF
TiO2 0:100 (0:1) 150 29.26 ± 0.1 57.03 ± 0.1 
F:TiO2 (N1) 50:50 (1:1)  25 29.68 ± 0.1 57.68 ± 0.1 Not detected 
F:TiO2 (N2) 75:25 (3:1)  37.5 29.72 ± 0.1 57.91 ± 0.1 Not detected 
F:TiO2 (N3) 85:15 (6:1)  42.8 29.91 ± 0.1 58.17 ± 0.1 Not detected 
F:TiO2 (0.5) 90:10 (9:1)  45 30.32 ± 0.1 58.46 ± 0.1 0.5 ± 0.1 
F:TiO2 (0.7) 95:5 (19:1)  47.5 30.78 ± 0.1 58.93 ± 0.1 0.7 ± 0.1 
F:TiO2 (1.0) 98:2 (49:1)  49 31.01 ± 0.1 59.42 ± 0.1 1.0 ± 0.1 
F:TiO2 (1.2) 100:0 (100:0)  50 31.92 ± 0.1 62.61 ± 0.1 1.2 ± 0.1 
HF pulse portion inXPS (at. %)
SampleCycle ratio (Ti–F:Ti–O)Growth temp. (°C)1 super-cycle (%)TiOF
TiO2 0:100 (0:1) 150 29.26 ± 0.1 57.03 ± 0.1 
F:TiO2 (N1) 50:50 (1:1)  25 29.68 ± 0.1 57.68 ± 0.1 Not detected 
F:TiO2 (N2) 75:25 (3:1)  37.5 29.72 ± 0.1 57.91 ± 0.1 Not detected 
F:TiO2 (N3) 85:15 (6:1)  42.8 29.91 ± 0.1 58.17 ± 0.1 Not detected 
F:TiO2 (0.5) 90:10 (9:1)  45 30.32 ± 0.1 58.46 ± 0.1 0.5 ± 0.1 
F:TiO2 (0.7) 95:5 (19:1)  47.5 30.78 ± 0.1 58.93 ± 0.1 0.7 ± 0.1 
F:TiO2 (1.0) 98:2 (49:1)  49 31.01 ± 0.1 59.42 ± 0.1 1.0 ± 0.1 
F:TiO2 (1.2) 100:0 (100:0)  50 31.92 ± 0.1 62.61 ± 0.1 1.2 ± 0.1 

Figure S4 (supplementary material) shows FT-IR spectra for F:TiO2 at F = 0, 0.5, 1.0, and 1.2 %. The main absorption area 750–850 cm−1 corresponds to Ti–O bending, and the Ti–F peak at 890 cm−1 confirms the Ti–F bond presence in the TiO2 lattice.38,39

Figure 3 shows HR-XRD patterns and AFM images for F:TiO2 thin films for different fluorine concentrations. Anatase (JCPDS No. 21-1272) and rutile (JCPDS No. 21-1276) phases were found in all samples, with the main phase being anatase TiO2. Anatase/rutile phase ratios were calculated as40,41

(1)

and

(2)

respectively, where WA and WR represent anatase and rutile TiO2 mass portions, respectively; and IA and IR represent diffraction peak integral intensity for anatase and rutile phase TiO2, respectively. Anatase and rutile TiO2 comprised 60.18% and 39.81% undoped TiO2, respectively, and 59.22%, and 40.76% for F:TiO2 (1.2), respectively.

FIG. 3.

F:TiO2 films at different fluorine concentrations after post-annealing at 500 °C in atmosphere: (a) high-resolution x-ray diffraction patterns; and atomic force microscope images for (b) TiO2, (c) F:TiO2 (0.5), (d) F:TiO2 (1.0), and (e) F:TiO2 (1.2).

FIG. 3.

F:TiO2 films at different fluorine concentrations after post-annealing at 500 °C in atmosphere: (a) high-resolution x-ray diffraction patterns; and atomic force microscope images for (b) TiO2, (c) F:TiO2 (0.5), (d) F:TiO2 (1.0), and (e) F:TiO2 (1.2).

Close modal

The average grain size of F:TiO2 (0, 0.5, 1.0 and 1.2) films were calculated from the Scherrer equation using HR-XRD results,

(3)

where β is FWHM of the peak, K = 0.9 (i.e., the shape factor), and λ = 0.154 06 nm (Cu Kα wavelength).

Thus, TiO2 and F:TiO2 (1.2) film grain sizes increased from 15.19 to 19.62 nm with increasing F concentration.38,42Figures 3(b)3(e) show AFM images for F:TiO2 on Pt substrates. F:TiO2 root-mean-square (rms) roughness over 2 × 2 μm2 increased with increasing fluorine concentration, achieving 1.37, 1.50, 1.92, and 2.28 nm rms roughness for F at. % = 0, 0.5, 1.0, and 1.2, respectively. Increased rms values were attributed to the crystalline growth and different growth rates of two TiO2 phases of anatase and rutile. The proper surface roughness promoted nucleation and grain growth. Increasing rms value implies the TiO2 crystallinity and grain size enhancement with increasing fluorine doping ratio.42 Thus, the fact that crystallinity in F:TiO2 thin films improved as the F concentration increased can be confirmed based on the calculation of accurate grain size through XRD data and an increase in the rms value from the AFM images.

Figure S5 (supplementary material) shows that XPS peaks for F:TiO2 film C 1s reduced as F doping increased, confirming fluorine improved TTIP ligand reactivity, thereby contributing to increased TiO2 crystallinity. F:TiO2 films residual carbon concentration changes provide further evidence for improved F:TiO2 crystallinity. Figure 4(a) shows optical bandgaps with and without F:TiO2, exhibiting interband absorption edge shift in the short wavelength region with increasing fluorine doping. Thus, fluorine doping increased the TiO2 optical bandgap.43 The optical bandgap was evaluated based on the absorption edge using Tauc’s equation,44 

(4)

where hν is the photon energy, n = 2 for the direct bandgap in the materials, and C is a constant.

FIG. 4.

F:TiO2 films with different F concentrations: (a) Tauc plots for transmittance spectra to determine optical bandgap, (b) carrier concentration, and Hall mobility and resistivity.

FIG. 4.

F:TiO2 films with different F concentrations: (a) Tauc plots for transmittance spectra to determine optical bandgap, (b) carrier concentration, and Hall mobility and resistivity.

Close modal

Figure 4(a) shows optical bandgaps, estimated by extrapolating the linear portion of the curve to α = 0. The optical bandgap for undoped TiO2 increased slightly (3.29, 3.30, 3.33, and 3.34 eV for 0, 0.5, 1.0, and 1.2 F doping, respectively), attributed to improved TiO2 crystallinity due to F doping.43 A band tail of TiO2 refers to the localized electronic states located at the valence and edges of the conduction band. They were produced by disordered structures and defects (such as impurities or vacancies) of semiconductors. As F doping increases, the localized states of F:TiO2 decrease because of the crystallinity enhancement and oxygen defect passivation.45,46 Therefore, the valence band maximum decreased and the optical band of F:TiO2 slightly enhanced more than the actual bandgap of TiO2.

Figure 4(b) shows carrier concentration, mobility, and resistivity for TiO2 and F:TiO2 on glass substrates. Hall effects confirmed that F passivated vacancies and hence decreased carrier concentration. n-type conductivity was found in all films, and increasing fluorine doping concentration reduced the carrier concentration to 1.77 × 1014 (TiO2) and 1.81 × 1013 cm−3 [F:TiO2 (1.2)] due to the passivation of oxygen vacancies with fluorine ions. Unlike carrier concentration, mobility seems to increase in the F:TiO2 (0.5) film but tends to decrease according to the increase of F concentration. It is noteworthy that in the case of the F:TiO2 (0.5) film, the passivation effect reduced both the carrier concentration and grain boundary defect acting as the scattering center.47,48 The reduction contributed to the mobility enhancement of the F:TiO2 (0.5) film. However, for F:TiO2 (1.0) and F:TiO2 (1.2) films, the change of band tail state played a main role compared to the effect of scattering center mentioned above. The optical bandgap was slightly enhanced by F doping on TiO2. This enhancement attributed to the decrease of TiO2 band tail state due to F doping. The reduced band tail made it difficult for the electrons in F:TiO2 films to move from the valence band to conduction band, leading to the decrease in mobility.45 Thus, the mobility of F:TiO2 (1.0) and F:TiO2 (1.2) films decreased as the F concentration increased. As a result, the reduced charge carrier concentration and mobility contributed to the electrical resistivity increase to 774 (TiO2) and 3433 Ω cm [F:TiO2 (1.2)], respectively,

(5)

where ρ is the resistivity, σ is the conductivity, μ is the carrier mobility, and n is the carrier concentration.

Figure 5 shows that F:TiO2 thin films exhibit different rectifying characteristics in I–V curves. Devices exhibit forward characteristics under positive voltage because TiO2−x was formed locally around the bottom interface (in Pt/TiO2/Pt/Ti/SiO2) by Ti adhesion layer diffusion during post-annealing at 500 °C.49 Yang et al.50 previously showed Ti adhesion layer diffusion in the annealing process, where diffusion reduces TiO2−x at the bottom interface and induces diode characteristics. Figure S6 (supplementary material) shows the XPS depth profiles of TiO2/Pt/Ti/SiO2 samples before and after post-annealing at 500 °C. Before post-annealing, the Ti 2p signal was not detected, as shown in Fig. S6(a), while 0.7. % of the Ti 2p signal in the Pt substrate region was detected after post-annealing [Fig. S6(b)]. Figure 6(c) shows the calculated O/Ti atomic ratio before and after annealing based on the depth profile data (etching time from 25 to 400 s). The atomic ratio before annealing was 1.95–2 with 400 s etching time. After post-annealing, the ratio was not changed with 200 s etching time, while it ranged from 1.8–1.85 under 200–400 s etching time. These results suggest that post-annealing induced the diffusion of the Ti adhesion layer under the Pt bottom electrode, which led to the formation of the TiO2−x layer at the bottom interface and self-rectifying characteristic.

FIG. 5.

Resistive switching properties for Pt (100 nm)/F:TiO2 (30 nm)/Pt/Ti/SiO2 devices with various F doping concentrations: (a) TiO2, (b) F:TiO2 (0.5), (c) F:TiO2 (1.0), and (d) F:TiO2 (1.2).

FIG. 5.

Resistive switching properties for Pt (100 nm)/F:TiO2 (30 nm)/Pt/Ti/SiO2 devices with various F doping concentrations: (a) TiO2, (b) F:TiO2 (0.5), (c) F:TiO2 (1.0), and (d) F:TiO2 (1.2).

Close modal
FIG. 6.

(a) Rectification ratio of F:TiO2 films with varying F concentration (0, 0.5, 1.0, and 1.2) depending on resistance state, (b) endurance of Pt/F:TiO2/Pt at RT, and (c) retention of F Pt/F:TiO2/Pt at 85 °C.

FIG. 6.

(a) Rectification ratio of F:TiO2 films with varying F concentration (0, 0.5, 1.0, and 1.2) depending on resistance state, (b) endurance of Pt/F:TiO2/Pt at RT, and (c) retention of F Pt/F:TiO2/Pt at 85 °C.

Close modal

Although RS behaviors were exhibited by F:TiO2 films, TiO2 did not exhibit RS behaviors in Fig. 5. The blue arrows and numbers indicate sweep sequence (0 → +2.5 → 0 → −2.5 → 0 V, step size = 0.01). Pristine F:TiO2 was in the high resistance state (HRS) and exhibited forming-less resistive switching and set-first behavior. All F:TiO2 samples exhibited typical counterclockwise RS and asymmetric I–V curves.

The rectification ratio of F:TiO2 films is shown in Fig. 6(a). The TiO2 device exhibited a similar current level in both low resistance state (LRS) and HRS regardless of voltage polarity, whereas the F:TiO2 (1.2) device showed a higher current level in the positive bias compared to that in the negative bias. In the case of the TiO2 device, the rectification ratios of HRS and LRS were 1.4 and 1.6, respectively. On the other hand, for the F:TiO2 (1.2) device, the rectification ratios of HRS and LRS increased to 4.0 and 10, respectively. The self-rectifying properties were indicated by the difference in height of the Schottky barrier at the top interface. While the amount of F doping increased in F:TiO2, the Schottky barrier height at the top interface increased due to oxygen vacancy passivation. Therefore, the F:TiO2 (1.2) films showed the largest Schottky barrier height difference at both electrodes interfaces. Based on this difference, it can be expected that the rectification ratio of F:TiO2 (1.2) could be improved compared to TiO2. Figure 6(b) shows the endurance characteristic of the Pt/F:TiO2 (1.2)/Pt device measured at room temperature, which shows stable behavior of 250 cycles. Figure 6(c) shows the retention properties of LRS and HRS for Vread = 1 V at 85 °C over 24 h (time interval = 300 s) and the blue dash line indicates 10 years. Figure S7 (supplementary material) shows simulated CAA readout margin using the non-linear I–V curves in Fig. 5.

Figure 7(a) shows the Schottky barrier heights of both LRS and HRS in F:TiO2 (1.2) at ±0.5 V. The Schottky barrier heights were extracted from temperature-dependent (300–380 K) experimental I–V data shown in Fig. S8 (supplementary material).51,52 In HRS, the barrier heights were 0.24 eV (at +0.5 V) and 0.34 eV (at −0.5 V), and the Schottky barrier heights of LRS were 0.2 eV (+0.5 V) and 0.29 eV (−0.5 V), respectively. We found that Schottky barrier height changed in resistive switching behavior from the result.

FIG. 7.

(a) Extracted Schottky barrier heights from temperature-dependent (300–380 K) experiment for LRS and HRS; (b) F:TiO2 resistive switching mechanism.

FIG. 7.

(a) Extracted Schottky barrier heights from temperature-dependent (300–380 K) experiment for LRS and HRS; (b) F:TiO2 resistive switching mechanism.

Close modal

Figure 7(b) shows the RS mechanism for F:TiO2, inducing RS behaviors in F:TiO2 films. We could expect the conduction mechanism of F:TiO2 following an interface-limited conduction regime since the asymmetric I–V curve occurred. As shown in Figs. 2 and 5, despite the decrease in the amount of oxygen vacancies, the on/off ratio increased as the amount of F doping increased. Considering the change of on/off ratio with F doping concentration, it is suggested that the migration of the fluorine anion mainly contributed to the RS characteristics of F:TiO2 because fluorine having a small ionic radius could migrate more easily. Under positive voltage (set process), the passivated fluorine anion in bulk migrated to the oxygen site and/or interstitial site of the top interface by the electric field or electron scattering. The migrated F ions at the top interface act as the n-type dopant, which decreases the Schottky barrier height and depletion width and also could provide defect level assisting charge injection. This change switches the resistance state of F:TiO2 from HRS to LRS, and the applied negative voltage induces the reset as the opposite process.

This paper proposed and validated a simple ALD process for a new RS material F:TiO2 for RRAM devices. Fluorine concentration in F:TiO2 thin films can be controlled by engineering ALD precursor pulse portions for Ti–O/F and Ti–O. Increased fluorine doping concentration reduced oxygen defect in the fabricated TiO2 films. XPS, FT-IR, and Hall measurements confirmed that F anion presence passivated oxygen vacancies in the film. TiO2 crystallinity also increased because fluorine doping improved TTIP ligand reactivity, and the optical energy bandgap also slightly increased from 3.29 (TiO2) to 3.34 [F:TiO2 (1.2)] eV.

Thus, the proposed F:TiO2 exhibited rectifying properties, forming-less operation, set-first process, and anti-clockwise RS. These phenomena were based on the interface limited conduction regime. The migration of fluorine mainly contributed to the RS characteristics by controlling the Schottky barrier height of the top interface of F:TiO2. The on/off ratio and self-rectifying characteristics of F:TiO2 films can be improved by optimizing the F doping concentration with inserting a tunnel barrier.53,54 Enhanced asymmetric and RS properties due to fluorine presence in TiO2 make the proposed system attractive to achieve sneak current-free CAA memory devices.

The supplementary material includes the XPS spectra, simulated DOS, oxygen/defect oxygen area ratio, FT-IR spectrum, C 1s spectra of XPS spectrum, and readout margins.

This work was supported by the South Korean Ministry of Trade, Industry, and Energy (Grant No. 10068075), the National Research Foundation of Korea funded by the Korean Government (Grant No. 2019R1A2C2087604), and the Creative Materials Discovery Program through the National Research Foundation of Korea funded by the Ministry of Science and ICT (Grant No. 2018M3D1A1058536).

The authors have no conflicts to disclose.

The data that support the findings of this study are available within the article and its supplementary material and from the corresponding authors upon reasonable request.

1.
S.
Yu
and
P.-Y.
Chen
,
IEEE Solid-State Circuits Mag.
8
,
43
56
(
2016
).
2.
Y. V.
Pershin
and
M.
Di Ventra
,
Adv. Phys.
60
,
145
227
(
2011
).
3.
H.-S. P.
Wong
 et al.,
Proc. IEEE
100
,
1951
1970
(
2012
).
4.
M.
She
, in
Semiconductor Flash Memory Scaling Meeting
,
2003
.
5.
A. S.
Spinelli
,
C. M.
Compagnoni
, and
A. L.
Lacaita
,
Computers
6
,
16
(
2017
).
6.
M.-J.
Lee
 et al.,
Nat. Mater.
10
,
625
630
(
2011
).
7.
F.
Zhang
 et al.,
Nat. Mater.
18
,
55
61
(
2019
).
8.
R.
Waser
and
M.
Aono
,
A Collection of Reviews from Nature Journals
(
Rainer Waser and Masakazu Aono
,
2010
), pp.
158
165
.
9.
A.
Brataas
,
A. D.
Kent
, and
H.
Ohno
,
Nat. Mater.
11
,
372
381
(
2012
).
10.
M.
Gajek
 et al.,
Nat. Mater.
6
,
296
302
(
2007
).
11.
C.
Liu
 et al.,
Nat. Nanotechnol.
13
,
404
410
(
2018
).
12.
E.
Linn
,
R.
Rosezin
,
C.
Kügeler
, and
R.
Waser
,
Nat. Mater.
9
,
403
406
(
2010
).
13.
Y. C.
Bae
 et al.,
Adv. Funct. Mater.
22
,
709
716
(
2012
).
14.
B. J.
Choi
 et al.,
Adv. Mater.
28
,
356
362
(
2016
).
15.
L.
Zhang
 et al., in
IEEE 6th International Memory Workshop
(
IEEE
,
2014
), Vols. 1–4.
16.
S.
Kim
,
J.
Zhou
, and
W. D.
Lu
,
IEEE Trans. Electron Devices
61
,
2820
2826
(
2014
).
17.
J.
Zhou
,
K.-H.
Kim
, and
W.
Lu
,
IEEE Trans. Electron Devices
61
,
1369
1376
(
2014
).
18.
G. A.
Patterson
,
J.
Sune
, and
E.
Miranda
, “
Voltage-driven hysteresis model for resistive switching: SPICE modeling and circuit applications
,”
IEEE Trans. Comput.-Aided Des. Integr. Circuits Syst.
36
(
12
),
2044
2051
(
2017
).
19.
X.
Liu
 et al.,
IEEE Electron Device Lett.
34
,
235
237
(
2013
).
20.
F.
Nardi
,
S.
Balatti
,
S.
Larentis
, and
D. C.
Gilmer
, in
IEEE 2011 International Electron Devices Meeting
(
IEEE
,
2011
), pp.
31
34
.
21.
22.
H. Y.
Jeong
,
Y. I.
Kim
,
J. Y.
Lee
, and
S.-Y.
Choi
,
Nanotechnology
21
,
115203
(
2010
).
23.
J.
Shin
 et al.,
J. Appl. Phys.
109
,
033712
(
2011
).
24.
J. H.
Yoon
 et al.,
Adv. Mater.
25
,
1987
1992
(
2013
).
25.
K. J.
Yoon
 et al.,
Adv. Electron. Mater.
3
,
1700152
(
2017
).
26.
H.
Akinaga
and
H.
Shima
,
Proc. IEEE
98
,
2237
2251
(
2010
).
27.
S.-J.
Park
 et al.,
Nanotechnology
24
,
295202
(
2013
).
28.
M. Y.
Yang
 et al.,
Appl. Phys. Lett.
103
,
093504
(
2013
).
29.
Z.
Wei
 et al., in
IEEE International Interconnect Technology Conference
(
IEEE
,
2014
), pp.
349
352
.
30.
Y.-J.
Choi
and
H.-H.
Park
,
J. Mater. Chem. C
2
,
98
108
(
2014
).
31.
M. D.
Segall
 et al.,
J. Phys.: Condens. Matter
14
,
2717
(
2002
).
32.
M. C.
Payne
,
M. P.
Teter
,
D. C.
Allan
,
T. A.
Arias
, and
J. D.
Joannopoulos
,
Rev. Mod. Phys.
64
,
1045
(
1992
).
33.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
,
Phys. Rev. Lett.
77
,
3865
(
1996
).
34.
V. I.
Anisimov
,
J.
Zaanen
, and
O. K.
Andersen
,
Phys. Rev. B
44
,
943
(
1991
).
35.
H. J.
Monkhorst
and
J. D.
Pack
,
Phys. Rev. B
13
,
5188
(
1976
).
36.
H.
Huang
 et al.,
Appl. Surf. Sci.
544
,
148583
(
2021
).
37.
W.
Yu
 et al.,
Appl. Surf. Sci.
319
,
107
112
(
2014
).
38.
X.
Zhang
,
Y.
Wu
,
Y.
Huang
,
Z.
Zhou
, and
S.
Shen
,
J. Alloys Compd.
681
,
191
196
(
2016
).
39.
M.
Mansoorianfar
,
R.
Rahighi
,
A.
Hojjati-Najafabadi
,
C.
Mei
, and
D.
Li
,
Mater. Des.
198
,
109314
(
2021
).
40.
H.
Zhang
and
J. F.
Banfield
,
J. Phys. Chem. B
104
,
3481
3487
(
2000
).
41.
Y.-e.
Du
 et al.,
ChemistrySelect
1
,
6632
6640
(
2016
).
42.
D.
Haško
and
J.
Bruncko
,
Vacuum
84
(
1
),
166
169
(
2009
).
43.
Z.-M.
Liu
,
L.
Peng
, and
A.-W.
Tang
,
Chin. Chem. Lett.
27
,
1801
1804
(
2016
).
44.
J.
Tauc
,
R.
Grigorovici
, and
A.
Vancu
,
Phys. Status Solidi B
15
(
2
),
627
637
(
1966
).
45.
B.
Wang
,
G.
Biesold
,
M.
Zhang
, and
Z.
Lin
,
Chem. Soc. Rev.
50
,
13646
(
2021
).
46.
Y.-C.
Ho
,
M. N. F.
Hoque
,
E.
Stoneham
,
J.
Warzywoda
,
T.
Dallas
, and
Z.
Fan
,
J. Phys. Chem. C
121
(
43
),
23939
23946
(
2017
).
47.
H. J.
Avila-Paredes
,
K.
Choi
,
C.-T.
Chen
, and
S.
Kim
,
J. Mater. Chem.
19
(
27
),
4837
4842
(
2009
).
48.
K.-M.
Kang
and
H.-H.
Park
,
J. Phys. Chem. C
122
(
1
),
377
385
(
2018
).
49.
L.
Gao
,
B.
Hoskins
, and
D.
Strukov
,
Phys. Status Solidi RRL
10
,
426
430
(
2016
).
50.
J. J.
Yang
 et al.,
Adv. Mater.
22
,
4034
4038
(
2010
).
51.
K.
Jeon
,
J.
Kim
,
J. J.
Ryu
,
S.-J.
Yoo
,
C.
Song
,
M. K.
Yang
,
D. S.
Jeong
, and
G. H.
Kim
,
Nat. Commun.
12
(
1
),
1
15
(
2021
).
52.
S. E.
Kim
,
J. G.
Lee
,
L.
Ling
,
S. E.
Liu
,
H. K.
Lim
,
V. K.
Sangwan
,
M. C.
Hersam
, and
H. S.
Lee
,
Adv. Mater.
34
(
6
),
2106913
(
2021
).
53.
S.
Lee
,
J.
Woo
,
D.
Lee
,
E.
Cha
,
J.
Park
,
K.
Moon
,
J.
Song
,
Y.
Koo
, and
H.
Hwang
,
Appl. Phys. Lett.
104
(
5
),
052108
(
2014
).
54.
H.
Choi
,
J.
Yi
,
S.
Hwang
,
S.
Lee
,
S.
Song
,
S.
Lee
,
J.
Lee
,
D.
Son
,
J.
Park
, and
S.-J.
Kim
, in
3rd IEEE International Memory Workshop (IMW)
(
IEEE
,
2011
), pp.
1
4
.

Supplementary Material