thanks: These two authors contributed equally.thanks: These two authors contributed equally.

Origin of Interstitial Do** Induced Coercive Field Reduction in Ferroelectric Hafnia

Tianyuan Zhu Department of Physics, School of Science, Westlake University, Hangzhou, Zhejiang 310030, China Institute of Natural Sciences, Westlake Institute for Advanced Study, Hangzhou, Zhejiang 310024, China    Liyang Ma Department of Physics, School of Science, Westlake University, Hangzhou, Zhejiang 310030, China    Xu Duan School of Science, Zhejiang University of Science and Technology, Hangzhou, Zhejiang 310023, China    Shi Liu [email protected] Department of Physics, School of Science, Westlake University, Hangzhou, Zhejiang 310030, China Institute of Natural Sciences, Westlake Institute for Advanced Study, Hangzhou, Zhejiang 310024, China
Abstract

Hafnia-based ferroelectrics hold promise for nonvolatile ferroelectric memory devices. However, the high coercive field required for polarization switching remains a prime obstacle to their practical applications. A notable reduction in coercive field has been achieved in ferroelectric Hf(Zr)1+xO2 films with interstitial Hf(Zr) dopants [Science 381, 558 (2023)], suggesting a less-explored strategy for coercive field optimization. Supported by density functional theory calculations, we demonstrate the Pca21𝑃𝑐𝑎subscript21Pca2_{1}italic_P italic_c italic_a 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT phase, with a moderate concentration of interstitial Hf dopants, serves as a minimal model to explain the experimental observations, rather than the originally assumed rhombohedral phase. Large-scale deep potential molecular dynamics simulations suggest that interstitial defects promote the polarization reversal by facilitating Pbcn𝑃𝑏𝑐𝑛Pbcnitalic_P italic_b italic_c italic_n-like mobile 180°°\degree° domain walls. A simple pre-poling treatment could reduce the switching field to less than 1 MV/cm and enable switching on a subnanosecond timescale. High-throughput calculations reveal a negative correlation between the switching barrier and dopant size and identify a few promising interstitial dopants for coercive field reduction.

Ferroelectric hafnia (HfO2) has emerged as a promising candidate for integrating ferroelectric functionalities into integrated circuits, enabled by its robust nanoscale ferroelectricity and exceptional silicon compatibility [1, 2, 3]. However, the high coercive field (csubscriptc\mathcal{E}_{\rm c}caligraphic_E start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT) required to switch the polarization in HfO2-based thin films remains a critical issue, impeding the commercialization of this silicon-compatible ferroelectric [3, 4]. Typically, polycrystalline thin films of hafnia fabricated through atomic-layer deposition display an csubscriptc\mathcal{E}_{\rm c}caligraphic_E start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT exceeding 1 MV/cm [5]. High-quality epitaxial thin films, obtained by pulsed laser deposition, can show even higher csubscriptc\mathcal{E}_{\rm c}caligraphic_E start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT values of \approx 2–5 MV/cm [6, 7, 8, 9]. The need for applying high electric fields, close to the material’s breakdown strength, for polarization reversal seriously limits its field cycling endurance. Reducing csubscriptc\mathcal{E}_{\rm c}caligraphic_E start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT without sacrificing the nanoscale ferroelectricity of hafnia is a pressing challenge for the wider adoption of HfO2-based ferroelectrics.

Various strategies such as do** [10, 11], strain [12, 13], and superlattices [14] have been theoretically explored to lower the barrier for polarization switching, aiming to reduce csubscriptc\mathcal{E}_{\rm c}caligraphic_E start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT. Investigations based on density functional theory (DFT) revealed that the polarization switching process in the ferroelectric othorhombic (O𝑂Oitalic_O) Pca21𝑃𝑐𝑎subscript21Pca2_{1}italic_P italic_c italic_a 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT phase often involves an intermediate tetragonal (T𝑇Titalic_T) P42/nmc𝑃subscript42𝑛𝑚𝑐P4_{2}/nmcitalic_P 4 start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / italic_n italic_m italic_c phase, and the energy difference between these two phases could serve as a measure of the switching barrier [11, 15]. Substitutional Si has been suggested as an effective dopant for reducing csubscriptc\mathcal{E}_{\rm c}caligraphic_E start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT due to its intrinsic sp3𝑠superscript𝑝3sp^{3}italic_s italic_p start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bonding with oxygen, which helps stabilize the intermediate T𝑇Titalic_T phase [11, 16]. Conversely, most other substitutional dopants have shown limited effectiveness in influencing csubscriptc\mathcal{E}_{\rm c}caligraphic_E start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT based on DFT calculations [11]. Recently, a substantial reduction in csubscriptc\mathcal{E}_{\rm c}caligraphic_E start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT, down to \approx0.65 MV/cm, has been achieved in thin films of Hf(Zr)1+xO2, a composition rich in hafnium-zirconium [Hf(Zr)] [17]. This reduction was attributed to the interstitial do** of Hf and Zr atoms into a polar rhombohedral (R𝑅Ritalic_R) R3m𝑅3𝑚R3mitalic_R 3 italic_m phase. DFT calculations show that a 7%-strained R𝑅Ritalic_R-phase Hf1.08O2 exhibits a switching barrier of 7.6 meV/atom, significantly lower than that of 20 meV/atom in stoichiometric R𝑅Ritalic_R-HfO2.

However, assuming a polar R𝑅Ritalic_R phase in Hf1.08O2 films grown via magnetron sputtering leads to notable discrepancies between theory and experiment [18, 19]. The unstrained R𝑅Ritalic_R phase is actually a nonpolar cubic P4¯3m𝑃¯43𝑚P\bar{4}3mitalic_P over¯ start_ARG 4 end_ARG 3 italic_m phase [3, 20]; the polar R3m𝑅3𝑚R3mitalic_R 3 italic_m symmetry only emerges under an equibiaxial compressive strain within the (111) crystallographic plane. Although the relative stability between the R𝑅Ritalic_R and O𝑂Oitalic_O phases can be reversed with 8% interstitial do** [17], highly-doped R𝑅Ritalic_R-Hf1.08O2 without strain remains nonpolar; a giant compressive strain of 7% is nevertheless required to induce a polarization of \approx15 μ𝜇\muitalic_μC/cm2, still below the experimental value of 22 μ𝜇\muitalic_μC/cm2 in polycrystalline films [17]. Moreover, X-ray diffraction (XRD) patterns of Hf(Zr)1+xO2 thin films show only a slight offset from both unstrained R𝑅Ritalic_R and O𝑂Oitalic_O phases [21], indicating the absence of high strains.

In this work, we combine DFT and deep potential molecular dynamics (DPMD) simulations [22] to investigate how interstitial Hf do** reduces the coercive field in ferroelectric HfO2. We find that interstitial Hf dopants diminish the energy difference between the polar orthorhombic Pca21𝑃𝑐𝑎subscript21Pca2_{1}italic_P italic_c italic_a 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT phase and the intermediate tetragonal P42/nmc𝑃subscript42𝑛𝑚𝑐P4_{2}/nmcitalic_P 4 start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / italic_n italic_m italic_c phase, lowering the switching barrier and coercive field. Unlike the R𝑅Ritalic_R phase, which demands additional compressive strains for polarization, the Pca21𝑃𝑐𝑎subscript21Pca2_{1}italic_P italic_c italic_a 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT phase of Hf1.03O2, with a moderate do** concentration of 3%, already exhibits a low switching barrier and a high polarization. We propose that the Pca21𝑃𝑐𝑎subscript21Pca2_{1}italic_P italic_c italic_a 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT phase of HfO2 with interstitial Hf defects better explains experimental observations. Large-scale MD simulations at finite temperatures with a DFT-derived machine learning force field demonstrate that 0.5% interstitial do** in Pca21𝑃𝑐𝑎subscript21Pca2_{1}italic_P italic_c italic_a 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT HfO2 is enough to reduce the switching field significantly. Moreover, after a pre-poling treatment during which interstitial dopants induce the formation of mobile domain walls, the switching field can drop below 1 MV/cm, comparable to experimental results. Further DFT calculations covering various interstitial dopants reveal a negative correlation between the switching barrier and dopant size.

All DFT calculations are carried out using the Vienna ab initio simulation package (VASP) [23] with the projector augmented-wave (PAW) method [24, 25] and the Perdew-Burke-Ernzerhof (PBE) exchange correlation functional [26]. The cut-off energy for the plane-wave basis is set to 600 eV. The Brillouin zones of the 12-atom pseudocubic unit cell and the 2×\times×2×\times×2 supercell for doped systems are sampled by ΓΓ\Gammaroman_Γ-centered 4×\times×4×\times×4 and 2×\times×2×\times×2 Monkhorst-Pack [27] k𝑘kitalic_k-point meshes, respectively. All the structures are fully optimized until the atomic forces converge to 0.01 eV/Å. The polarization values are determined using the Berry phase method [28, 29]. The polarization switching pathway is determined using the nudged elastic band (NEB) method [30]. The bond order analysis is performed with the crystal orbital Hamilton population (COHP) method [31] implemented in LOBSTER [32]. To investigate the ferroelectric switching in Hf1+xO2 at finite temperatures, we perform isobaric-isothermal ensemble (NPT𝑁𝑃𝑇NPTitalic_N italic_P italic_T) MD simulations using a deep neural network-based force field [33, 34], with model accuracy verification and sample inputs available via an online notebook [35]. The electric fields are included in MD simulations using the “force method” [36, 37, 15], where an additional force isubscript𝑖\mathcal{F}_{i}caligraphic_F start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is added on the ion i𝑖iitalic_i according to i=Zisubscript𝑖superscriptsubscript𝑍𝑖\mathcal{F}_{i}=Z_{i}^{*}\cdot\mathcal{E}caligraphic_F start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ⋅ caligraphic_E, with Zisuperscriptsubscript𝑍𝑖Z_{i}^{*}italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT being the Born effective charge (BEC) tensor of ion i𝑖iitalic_i.

We start with an analysis of the phase stability of Hf1+xO2 with interstitial Hf dopants (Hfi), considering the monoclinic (M𝑀Mitalic_M) P21/c𝑃subscript21𝑐P2_{1}/citalic_P 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_c, polar orthorhombic (O𝑂Oitalic_O) Pca21𝑃𝑐𝑎subscript21Pca2_{1}italic_P italic_c italic_a 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, tetragonal (T𝑇Titalic_T) P42/nmc𝑃subscript42𝑛𝑚𝑐P4_{2}/nmcitalic_P 4 start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / italic_n italic_m italic_c, rhombohedral (R𝑅Ritalic_R) R3m𝑅3𝑚R3mitalic_R 3 italic_m, and cubic (C𝐶Citalic_C) Fm3¯m𝐹𝑚¯3𝑚Fm\bar{3}mitalic_F italic_m over¯ start_ARG 3 end_ARG italic_m phases. Two do** concentrations, 3.125% and 6.25% (abbreviated as 3% and 6%), are modeled by intercalating one and two excess Hf atom within the hollow sites in a 96-atom 2×\times×2×\times×2 supercell, respectively. As shown in Fig. 1(a), pristine HfO2 polymorphs exhibit an inverse correlation between the thermodynamic stability and crystal symmetry.With increasing Hfi concentration, the energies of these phases become closer, suggesting higher Hfi formation energy in lower-symmetry M𝑀Mitalic_M and O𝑂Oitalic_O phases but lower in higher-symmetry C𝐶Citalic_C and R𝑅Ritalic_R phases. This dependence is due to different effective void sizes in each phase [see the inset of Fig. 1(a)]. Specifically, the M𝑀Mitalic_M and O𝑂Oitalic_O phases have smaller voids due to alternating fourfold-coordinated (4C) and threefold-coordinated (3C) oxygen atoms, hindering Hfi integration and resulting in higher formation energy.

Consistent with previous DFT results [17], our calculations show that, compared to the T𝑇Titalic_T phase, Hfi do** destabilizes the lower-symmetry M𝑀Mitalic_M and O𝑂Oitalic_O phases while imposing minimal impact on the relative stability of the R𝑅Ritalic_R phase. This suggests that the R𝑅Ritalic_R phase would become energetically more favorable than the O𝑂Oitalic_O phase above a critical Hfi concentration, if it remains stable. However, the Hfi-doped R𝑅Ritalic_R phase is nonpolar, requiring a (111) in-plane compressive strain to achieve a polarized R𝑅Ritalic_R phase. Using a 36-atom hexagonal supercell with a 7% in-plane compressive strain and an excess Hf atom, Wang et al. reported a polarized R𝑅Ritalic_R-Hf1.08O2 with a low switching barrier of 7.6 meV/atom [17]. Notably, such a high strain state should correspond to a large out-of-plane interplanar spacing, d111=3.11subscript𝑑1113.11d_{111}=3.11italic_d start_POSTSUBSCRIPT 111 end_POSTSUBSCRIPT = 3.11 Å. This does not align with the experimental XRD data of Hf1+xO2 films, which shows a (111) peak at 2θ=30.1°2𝜃30.1°2\theta=30.1\degree2 italic_θ = 30.1 °, indicating only a 0.1°°\degree° shift compared to the stoichiometric O𝑂Oitalic_O phase (with d111=subscript𝑑111absentd_{111}=italic_d start_POSTSUBSCRIPT 111 end_POSTSUBSCRIPT =2.96 Å, see Fig. S1).

In contrast to the R𝑅Ritalic_R phase, the O𝑂Oitalic_O phase is inherently polar and does not depend on applied strain for polarization. The ferroelectric switching barrier in the O𝑂Oitalic_O phase can be estimated from the energy difference between the T𝑇Titalic_T phase and O𝑂Oitalic_O phase (ΔE=ETEOΔ𝐸superscript𝐸𝑇superscript𝐸𝑂\Delta E=E^{T}-E^{O}roman_Δ italic_E = italic_E start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT - italic_E start_POSTSUPERSCRIPT italic_O end_POSTSUPERSCRIPT[11]. As illustrated in Fig. 1(a), interstitial do** elevates the energy of the O𝑂Oitalic_O phase relative to the T𝑇Titalic_T phase, thereby reducing ΔEΔ𝐸\Delta Eroman_Δ italic_E. Figure 1(b) plots the NEB energy variation along the switching pathway in the O𝑂Oitalic_O phase for different do** concentrations. The switching barrier decreases drastically from 28 meV/atom in pristine HfO2 to 12.3 meV/atom in Hf1.03O2, and further drops to 7.5 meV/atom in Hf1.06O2.

As summarized in Fig. 1(c), for the O𝑂Oitalic_O phase, both the switching barrier and polarization decrease with increasing Hfi concentration. This reduction in polarization is attributed to the diminished local displacements of polar oxygen atoms surrounding Hfi. Further increasing Hfi concentration destabilizes the O𝑂Oitalic_O phase, causing it to transform into a Pbcn𝑃𝑏𝑐𝑛Pbcnitalic_P italic_b italic_c italic_n-like structure during DFT geometry relaxation (see Fig. S2). Importantly, O𝑂Oitalic_O-Hf1.06O2, with a lower Hfi concentration, exhibits a switching barrier (7.5 meV/atom) comparable to that of 7%-strained R𝑅Ritalic_R-Hf1.08O2. It also features a higher spontaneous polarization along the [111] direction, matching closely to the experimental value. Although large local strains in polycrystalline thin films cannot be ruled out, the Pca21𝑃𝑐𝑎subscript21Pca2_{1}italic_P italic_c italic_a 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT phase with moderate interstitial Hf do** serves as a minimal model to explain experimental observations. We note that Hfi exhibits a negative BEC of 2e2𝑒-2e- 2 italic_e, indicating that its slight movement induces a change in polarization equivalent to the displacement of two negative elementary charges. This counterintuitive result is likely due to the crowding effect of the confining cage around Hfi. Nevertheless, since the Hfi is strongly confined, it does not change the polarization switching mechanism identified with DFT-based NEB calculations.

Limited by computational cost, our DFT investigations focused on small supercells. This raises a legitimate question: how does the reduced energy difference between the O𝑂Oitalic_O and T𝑇Titalic_T phases, seen in zero-Kelvin DFT calculations, manifest in larger supercells at elevated temperatures that better represent the experimental sample conditions? To address this, we perform large-scale MD simulations at finite temperatures for an O𝑂Oitalic_O-Hf1.005O2 supercell of 20,772 atoms, considering different Hfi dopant distributions. The force field is a deep neural network-based model potential [15], which accurately reproduces various properties including the energy variation along the switching pathway in Hf1+xO2 (see Fig. S3). All simulations are conducted at 400 K to facilitate the switching process on (sub)nanosecond time scales, thereby mitigating computational cost. We gauge the ease of switching by determining the lowest field strength (ssubscripts\mathcal{E}_{\rm s}caligraphic_E start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT) that triggers polarization switching within 200 ps in MD simulations. As shown in Fig. 2(a), the switching field is 5.3 MV/cm for pristine single-domain HfO2. Introducing uniformly distributed 0.5% Hfi dopants decreases ssubscripts\mathcal{E}_{\rm s}caligraphic_E start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT to 4.4 MV/cm. The local enrichment of Hfi dopants further reduce ssubscripts\mathcal{E}_{\rm s}caligraphic_E start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT: the cluster-like and two-dimensional (2D) distributions of Hfi lead to ssubscripts\mathcal{E}_{\rm s}caligraphic_E start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT of 2.8 and 3.1 MV/cm, respectively (see dopant distributions in Fig. S4).

We discover that a simple pre-poling treatment can bring ssubscripts\mathcal{E}_{\rm s}caligraphic_E start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT to an even lower value of \approx0.8 MV/cm. The switching process extracted from MD simulations is presented in Fig. 2(c-f), with the color scheme explained in Fig. 2(b). The starting configuration has 3C oxygen atoms displaced upward (P-) relative to 4C oxygen atom (NP). The 2D-confined Hfi dopants are distributed randomly within the xz𝑥𝑧xzitalic_x italic_z plane between P- and NP layers. A main impact of Hfi is the slight downward displacement of 4C atoms [left inset of Fig. 2(c)]. Upon poling under an electric field of =3.13.1\mathcal{E}=3.1caligraphic_E = 3.1 MV/cm, the Hfi enriched layer facilitates the nucleation of 180°°\degree° domain walls characterized by oppositely polarized 3C atoms, called type-P walls [see Fig. 2(d)]. Locally, the type-P wall resembles a Pbcn𝑃𝑏𝑐𝑛Pbcnitalic_P italic_b italic_c italic_n unit cell, consistent with DFT calculations showing a spontaneous transformation of the O𝑂Oitalic_O phase to a Pbcn𝑃𝑏𝑐𝑛Pbcnitalic_P italic_b italic_c italic_n-like structure at high Hfi concentration (see Fig. S2). We confirm that Hfi dopants reduce the type-P domain-wall energy (see Fig. S5). Consequently, polarization switching is driven by the motion of a type-P wall, while the other type-P wall is pinned by Hfi dopants. When the external field is turned off, the mobile type-P wall remains stable. After pre-poling, a much lower electric field of 0.8 MV/cm is sufficient to drive the domain-wall motion again.

Above results show that the value of ΔEΔ𝐸\Delta Eroman_Δ italic_E, easily accessible through DFT calculations at the unit cell level, can indeed serve as a useful descriptor for csubscriptc\mathcal{E}_{\rm c}caligraphic_E start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT. To identify more types of dopants for regulating csubscriptc\mathcal{E}_{\rm c}caligraphic_E start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT in ferroelectric HfO2, we perform high-throughput DFT calculations covering diverse dopant atoms, X=𝑋absentX=italic_X = Hf, Zr, Ti, La, Ta, Y, Nb, Sn, Ge, Si, Al, Ga, In, Sb, Mg, Ca, Cu, Zn, with a fixed concentration of 3%. As shown in Fig. 3(a), ΔEΔ𝐸\Delta Eroman_Δ italic_E between the T𝑇Titalic_T and O𝑂Oitalic_O phases with interstitial dopants (Xisubscript𝑋iX_{\rm i}italic_X start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT) is plotted against the dopant atomic radius, revealing a clear negative correlation. Interstitial dopants with large atomic radii, such as Hf, Zr, and Ti, tend to reduce the ΔEΔ𝐸\Delta Eroman_Δ italic_E, while smaller interstitial atoms such as Si, Ge, and Sn tend to increase ΔEΔ𝐸\Delta Eroman_Δ italic_E, potentially leading to higher csubscriptc\mathcal{E}_{\rm c}caligraphic_E start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT.

This negative correlation between ΔEΔ𝐸\Delta Eroman_Δ italic_E and dopant size for interstitial do** contrasts sharply with substitutional do**, where a volcano-like dependence (mostly a positive correlation) of ΔEΔ𝐸\Delta Eroman_Δ italic_E on the ionic radius of the substitutional dopant was observed [11]. We further examine substitutional do** with atom X𝑋Xitalic_X (XHfsubscript𝑋HfX_{\rm Hf}italic_X start_POSTSUBSCRIPT roman_Hf end_POSTSUBSCRIPT) and plot ΔEΔ𝐸\Delta Eroman_Δ italic_E as a function of the ionic radius of X𝑋Xitalic_X [see Fig.3(b)]. Different from interstitial do**, the ΔEΔ𝐸\Delta Eroman_Δ italic_E of HfO2 under substitutional do** increases with dopant size, consistent with previous DFT results [11]. Notably, only a few small substitutional dopants, like SiHf, effectively lower ΔEΔ𝐸\Delta Eroman_Δ italic_E. This has been attributed to the formation of stable sp3𝑠superscript𝑝3sp^{3}italic_s italic_p start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT bonds between the substitutional Si dopant and neighboring oxygen atoms in the T𝑇Titalic_T phase [11].

To comprehend the diametrically opposing effects of Hfi and Sii on ΔEΔ𝐸\Delta Eroman_Δ italic_E and csubscriptc\mathcal{E}_{\rm c}caligraphic_E start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT, we analyze the bonding of interstitial atoms in the T𝑇Titalic_T and O𝑂Oitalic_O phases of HfO2. As depicted in Fig. 4, the intercalated atom is surrounded by eight oxygen atoms. We quantify the bonding strength of an interstitial dopant based on the bond-valence conservation principle [38, 39]. According to this principle, each atom i𝑖iitalic_i has a preferred atomic valence V0,isubscript𝑉0𝑖V_{0,i}italic_V start_POSTSUBSCRIPT 0 , italic_i end_POSTSUBSCRIPT, which is often the nominal oxidation state of the atom (e. g., +44+4+ 4 for Hf). The actual atomic valence, Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, is obtained by summing the individual bond valence (Vijsubscript𝑉𝑖𝑗V_{ij}italic_V start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT) for bonds between the atom and its neighbors (j𝑗jitalic_j), Vi=jVijsubscript𝑉𝑖subscript𝑗subscript𝑉𝑖𝑗V_{i}=\sum_{j}V_{ij}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT. A smaller deviation of Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT from V0,isubscript𝑉0𝑖V_{0,i}italic_V start_POSTSUBSCRIPT 0 , italic_i end_POSTSUBSCRIPT reflects stronger overall bonding strength. The bond valence for each oxygen-dopant pair is calculated as Vij=exp[(RoRij)/B]subscript𝑉𝑖𝑗subscript𝑅osubscript𝑅𝑖𝑗𝐵V_{ij}=\exp{[(R_{\rm o}-R_{ij})/B]}italic_V start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = roman_exp [ ( italic_R start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT - italic_R start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ) / italic_B ], where Rijsubscript𝑅𝑖𝑗R_{ij}italic_R start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT is the bond length, Rosubscript𝑅oR_{\rm o}italic_R start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT and B𝐵Bitalic_B are Brown’s empirical parameters [40]. The calculated Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT for Hfi is 3.19 in the T𝑇Titalic_T phase and 3.02 in the O𝑂Oitalic_O phase, indicating stronger bonding of Hfi in the T𝑇Titalic_T phase. This is further supported by the projected COHP analysis. As shown in the right panel of Fig. 4, the total integration of COHP (ICOHP) for Hfi--O pairs is --20.0 and --19.0 eV in the T𝑇Titalic_T and O𝑂Oitalic_O phases, respectively. The higher magnitude of ICOHP in the T𝑇Titalic_T phase confirms stronger bonding between Hfi and oxygen atoms.

The difference between the energy of the O𝑂Oitalic_O and T𝑇Titalic_T phases containing Hfi (EO[Hfi]superscript𝐸𝑂delimited-[]subscriptHfiE^{O}[{\rm Hf_{i}}]italic_E start_POSTSUPERSCRIPT italic_O end_POSTSUPERSCRIPT [ roman_Hf start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT ] and ET[Hfi]superscript𝐸𝑇delimited-[]subscriptHfiE^{T}[{\rm Hf_{i}}]italic_E start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT [ roman_Hf start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT ]) can be directly related to the differences in the formation energy of Hfi (Ef[Hfi]subscript𝐸𝑓delimited-[]subscriptHfiE_{f}[{\rm Hf_{i}}]italic_E start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT [ roman_Hf start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT ]) as follows:

ΔEΔ𝐸\displaystyle\Delta Eroman_Δ italic_E =ET[Hfi]EO[Hfi]absentsuperscript𝐸𝑇delimited-[]subscriptHfisuperscript𝐸𝑂delimited-[]subscriptHfi\displaystyle=E^{T}[{\rm Hf_{i}}]-E^{O}[{\rm Hf_{i}}]= italic_E start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT [ roman_Hf start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT ] - italic_E start_POSTSUPERSCRIPT italic_O end_POSTSUPERSCRIPT [ roman_Hf start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT ]
=(ET[Hfi]E0TμHf)(EO[Hfi]E0OμHf)+E0TE0Oabsentsuperscript𝐸𝑇delimited-[]subscriptHfisubscriptsuperscript𝐸𝑇0subscript𝜇Hfsuperscript𝐸𝑂delimited-[]subscriptHfisubscriptsuperscript𝐸𝑂0subscript𝜇Hfsubscriptsuperscript𝐸𝑇0subscriptsuperscript𝐸𝑂0\displaystyle=(E^{T}[{\rm Hf_{i}}]-E^{T}_{0}-\mu_{\rm Hf})-(E^{O}[{\rm Hf_{i}}% ]-E^{O}_{0}-\mu_{\rm Hf})+E^{T}_{0}-E^{O}_{0}= ( italic_E start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT [ roman_Hf start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT ] - italic_E start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_μ start_POSTSUBSCRIPT roman_Hf end_POSTSUBSCRIPT ) - ( italic_E start_POSTSUPERSCRIPT italic_O end_POSTSUPERSCRIPT [ roman_Hf start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT ] - italic_E start_POSTSUPERSCRIPT italic_O end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_μ start_POSTSUBSCRIPT roman_Hf end_POSTSUBSCRIPT ) + italic_E start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E start_POSTSUPERSCRIPT italic_O end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT
=EfT[Hfi]EfO[Hfi]+ΔE0=ΔEf[Hfi]+ΔE0,absentsubscriptsuperscript𝐸𝑇𝑓delimited-[]subscriptHfisubscriptsuperscript𝐸𝑂𝑓delimited-[]subscriptHfiΔsubscript𝐸0Δsubscript𝐸𝑓delimited-[]subscriptHfiΔsubscript𝐸0\displaystyle=E^{T}_{f}[{\rm Hf_{i}}]-E^{O}_{f}[{\rm Hf_{i}}]+\Delta E_{0}=% \Delta E_{f}[{\rm Hf_{i}}]+\Delta E_{0},= italic_E start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT [ roman_Hf start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT ] - italic_E start_POSTSUPERSCRIPT italic_O end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT [ roman_Hf start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT ] + roman_Δ italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = roman_Δ italic_E start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT [ roman_Hf start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT ] + roman_Δ italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ,

where μHfsubscript𝜇Hf\mu_{\rm Hf}italic_μ start_POSTSUBSCRIPT roman_Hf end_POSTSUBSCRIPT is the chemical potential of Hf and E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the energy of the pristine crystal. The relatively stronger Hfi--O bonding in the T𝑇Titalic_T phase suggests a lower formation energy of Hfi than that in the O𝑂Oitalic_O phase (EfT[Hfi]<EfO[Hfi]subscriptsuperscript𝐸𝑇𝑓delimited-[]subscriptHfisubscriptsuperscript𝐸𝑂𝑓delimited-[]subscriptHfiE^{T}_{f}[{\rm Hf_{i}}]<E^{O}_{f}[{\rm Hf_{i}}]italic_E start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT [ roman_Hf start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT ] < italic_E start_POSTSUPERSCRIPT italic_O end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT [ roman_Hf start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT ][41], which is responsible for the reduced ΔEΔ𝐸\Delta Eroman_Δ italic_E compared to the undoped value (ΔE0Δsubscript𝐸0\Delta E_{0}roman_Δ italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT). In the case of Sii, the pCOHP curve reveals a stronger anti-bonding character, which is corroborated by the small values of Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Importantly, both the magnitudes of ICOHP and Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are larger in the O𝑂Oitalic_O phase than those in the T𝑇Titalic_T phase. This points to a lower formation energy of Sii in the low-energy O𝑂Oitalic_O phase, which will further increases the energy difference between T𝑇Titalic_T and O𝑂Oitalic_O phases.

In summary, our investigations combining both zero-Kelvin DFT calculations and large-scale MD simulations at finite temperatures establish the link between interstitial do** and coercive field in ferroelectric Pca21𝑃𝑐𝑎subscript21Pca2_{1}italic_P italic_c italic_a 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT HfO2. Unit-cell-level DFT calculations act as a mean-field-like analysis, demonstrating that the coercive field reduction from interstitial Hf do** is due to the lower defect formation energy in the intermediate P42/nmc𝑃subscript42𝑛𝑚𝑐P4_{2}/nmcitalic_P 4 start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / italic_n italic_m italic_c phase. MD simulations, better representing experimental conditions, suggest that interstitial Hf dopants likely promote switching by forming mobile Pbcn𝑃𝑏𝑐𝑛Pbcnitalic_P italic_b italic_c italic_n-like domain walls. With pre-poling, a switching field of <<<1.0 MV/cm can drive polarization reversal within subnanoseconds. Finally, high-throughput DFT calculations reveal a negative correlation between the switching barrier and the size of the interstitial dopant. The comprehensive understanding of interstitial do** effects in hafnia offers useful guidelines for optimizing the coercive field in this silicon-compatible ferroelectric oxide.

Acknowledgements.
T.Z., L.M. and S.L. acknowledge the supports from National Key R&D Program of China (2021YFA1202100), National Natural Science Foundation of China (12361141821, 12074319), and Westlake Education Foundation. The computational resource is provided by Westlake HPC Center.
Refer to caption
Figure 1: Phase stability and ferroelectric switching of Hf1+xO2. (a) Energy landscape of monoclinic (M𝑀Mitalic_M) P21/c𝑃subscript21𝑐P2_{1}/citalic_P 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_c, orthorhombic (O𝑂Oitalic_O) Pca21𝑃𝑐𝑎subscript21Pca2_{1}italic_P italic_c italic_a 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, tetragonal (T𝑇Titalic_T) P42/nmc𝑃subscript42𝑛𝑚𝑐P4_{2}/nmcitalic_P 4 start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / italic_n italic_m italic_c, rhombohedral (R𝑅Ritalic_R) R3m𝑅3𝑚R3mitalic_R 3 italic_m, and cubic (C𝐶Citalic_C) Fm3¯m𝐹𝑚¯3𝑚Fm\bar{3}mitalic_F italic_m over¯ start_ARG 3 end_ARG italic_m phases under interstitial do** of different concentrations (3% and 6% in short of 3.125% and 6.25%, respectively). The energy of T𝑇Titalic_T phase is set to zero as a reference. The inset schematically shows a interstitial Hf atom intercalated within a void (purple polyhedron) formed by surrounding 6 Hf and 8 O atoms. (b) Energy variation along the polarization switching pathway of the O𝑂Oitalic_O phase under different do** concentrations. (c) Polarization projected along the [111] crystallographic orientation and switching barrier of the O𝑂Oitalic_O and R𝑅Ritalic_R phases. The results of 7%-strained R𝑅Ritalic_R phase (hollow yellow circle) and experimental remanent polarization (dashed line) are extracted from Ref. [17].
Refer to caption
Figure 2: Interstitial Hf dopants promoted polarization reversal at elevated temperatures. (a) Switching fields (ssubscripts\mathcal{E}_{\rm s}caligraphic_E start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT) of pristine and differently distributed interstitial-doped HfO2. (b) Alternately arranged nonpolar (NP) and polar (P-/P+) oxygen atoms in a Pca21𝑃𝑐𝑎subscript21Pca2_{1}italic_P italic_c italic_a 2 start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT unit cell. The P- oxygen atoms exhibit a positive local displacement (δ𝛿\deltaitalic_δ) of 0.54 Å. Distributions of δ𝛿\deltaitalic_δ in a 12×\times×12×\times×12 supercell in the presence of two-dimensional-confined interstitial Hf dopants (denoted by hollow circles) between P- and NP layers without an applied electric field (c), under =3.13.1\mathcal{E}=3.1caligraphic_E = 3.1 MV/cm poling (d), with the electric field turned off (e), and under subsequently applied =0.80.8\mathcal{E}=0.8caligraphic_E = 0.8 MV/cm (f). The left inset of (c) displays the local Pbcn𝑃𝑏𝑐𝑛Pbcnitalic_P italic_b italic_c italic_n-like structure surrounding an interstitial Hf. The top inset of (e) displays the atomic structure of a mobile type-P domain wall [42].
Refer to caption
Figure 3: Energy difference (ΔE=ETEOΔ𝐸superscript𝐸𝑇superscript𝐸𝑂\Delta E=E^{T}-E^{O}roman_Δ italic_E = italic_E start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT - italic_E start_POSTSUPERSCRIPT italic_O end_POSTSUPERSCRIPT) between T𝑇Titalic_T-phase and O𝑂Oitalic_O-phase HfO2 with different dopants. (a) ΔEΔ𝐸\Delta Eroman_Δ italic_E as a function of atomic radius for interstitial Xisubscript𝑋iX_{\rm i}italic_X start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT do**. (b) ΔEΔ𝐸\Delta Eroman_Δ italic_E as a function of ionic radius for substitutional XHfsubscript𝑋HfX_{\rm Hf}italic_X start_POSTSUBSCRIPT roman_Hf end_POSTSUBSCRIPT do**. The do** concentration is set to 3%. The value for pristine HfO2 is denoted by a dashed line as a reference.
Refer to caption
Figure 4: Bonding analysis for interstitial Hf and Si with their neighboring oxygen atoms. (a) Hfi in T𝑇Titalic_T phase. (b) Hfi in O𝑂Oitalic_O phase. (c) Sii in T𝑇Titalic_T phase. (d) Sii in O𝑂Oitalic_O phase. The left panel shows the local atomic structure surrounding the interstitial atom, with the bond valence sum Vi=jexp[(RoRij)/B]subscript𝑉𝑖subscript𝑗subscript𝑅osubscript𝑅𝑖𝑗𝐵V_{i}=\sum_{j}\exp{[(R_{\rm o}-R_{ij})/B]}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT roman_exp [ ( italic_R start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT - italic_R start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ) / italic_B ] highlighted at the bottom. The Brown’s empirical parameters Ro=1.923subscript𝑅o1.923R_{\rm o}=1.923italic_R start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT = 1.923 Å and B=0.375𝐵0.375B=0.375italic_B = 0.375 Å for Hf, and Ro=1.624subscript𝑅o1.624R_{\rm o}=1.624italic_R start_POSTSUBSCRIPT roman_o end_POSTSUBSCRIPT = 1.624 Å and B=0.389𝐵0.389B=0.389italic_B = 0.389 Å for Si [40] are used in the bond valence calculation. The right panel shows the projected crystal orbital Hamilton population (pCOHP) averaged over eight interstitial–oxygen pairs, labeled with the total integration of COHP (ICOHP).

References

  • Böscke et al. [2011] T. S. Böscke, J. Müller, D. Bräuhaus, U. Schröder, and U. Böttger, Ferroelectricity in hafnium oxide thin films, Appl. Phys. Lett. 99, 102903 (2011).
  • Cheema et al. [2020] S. S. Cheema, D. Kwon, N. Shanker, R. dos Reis, S.-L. Hsu, J. Xiao, H. Zhang, R. Wagner, A. Datar, M. R. McCarter, C. R. Serrao, A. K. Yadav, G. Karbasian, C.-H. Hsu, A. J. Tan, L.-C. Wang, V. Thakare, X. Zhang, A. Mehta, E. Karapetrova, R. V. Chopdekar, P. Shafer, E. Arenholz, C. Hu, R. Proksch, R. Ramesh, J. Ciston, and S. Salahuddin, Enhanced ferroelectricity in ultrathin films grown directly on silicon, Naure 580, 478 (2020).
  • Schroeder et al. [2022] U. Schroeder, M. H. Park, T. Mikolajick, and C. S. Hwang, The fundamentals and applications of ferroelectric HfO2Nat. Rev. Mater. 7, 653 (2022).
  • Noheda et al. [2023] B. Noheda, P. Nukala, and M. Acuautla, Lessons from hafnium dioxide-based ferroelectrics, Nat. Mater. 22, 562 (2023).
  • Schroeder et al. [2014] U. Schroeder, E. Yurchuk, J. Müller, D. Martin, T. Schenk, P. Polakowski, C. Adelmann, M. I. Popovici, S. V. Kalinin, and T. Mikolajick, Impact of different dopants on the switching properties of ferroelectric hafnium oxide, Jpn. J. Appl. Phys. 53, 08LE02 (2014).
  • Wei et al. [2018] Y. Wei, P. Nukala, M. Salverda, S. Matzen, H. J. Zhao, J. Momand, A. S. Everhardt, G. Agnus, G. R. Blake, P. Lecoeur, B. J. Kooi, J. Íñiguez, B. Dkhil, and B. Noheda, A rhombohedral ferroelectric phase in epitaxially strained Hf0.5Zr0.5O2 thin films, Nat. Mater. 17, 1095 (2018).
  • Lyu et al. [2019] J. Lyu, I. Fina, R. Solanas, J. Fontcuberta, and F. Sánchez, Growth window of ferroelectric epitaxial Hf0.5Zr0.5O2 thin films, ACS Appl. Electron. Mater. 1, 220 (2019).
  • Estandía et al. [2019] S. Estandía, N. Dix, J. Gazquez, I. Fina, J. Lyu, M. F. Chisholm, J. Fontcuberta, and F. Sánchez, Engineering ferroelectric Hf0.5Zr0.5O2 thin films by epitaxial stress, ACS Appl. Electron. Mater. 1, 1449 (2019).
  • Song et al. [2020] T. Song, R. Bachelet, G. Saint-Girons, R. Solanas, I. Fina, and F. Sánchez, Epitaxial ferroelectric La-doped Hf0.5Zr0.5O2 thin films, ACS Appl. Electron. Mater. 2, 3221 (2020).
  • Qi and Rabe [2020] Y. Qi and K. M. Rabe, Phase competition in with applied electric field from first principles, Phys. Rev. B 102, 214108 (2020).
  • Yang et al. [2020] H. Yang, H.-J. Lee, J. Jo, C. H. Kim, and J. H. Lee, Role of Si do** in reducing coercive fields for ferroelectric switching in HfO2Phys. Rev. Appl. 14, 064012 (2020).
  • Wei et al. [2022] W. Wei, G. Zhao, X. Zhan, W. Zhang, P. Sang, Q. Wang, L. Tai, Q. Luo, Y. Li, C. Li, and J. Chen, Switching pathway-dependent strain-effects on the ferroelectric properties and structural deformations in orthorhombic HfO2J. Appl. Phys. 131, 154101 (2022).
  • Zhou et al. [2022] S. Zhou, J. Zhang, and A. M. Rappe, Strain-induced antipolar phase in hafnia stabilizes robust thin-film ferroelectricity, Sci. Adv. 8, eadd5953 (2022).
  • Zhao et al. [2024] H. J. Zhao, Y. Fu, L. Yu, Y. Wang, Y. Yang, L. Bellaiche, and Y. Ma, Creating ferroelectricity in monoclinic (HfO2)1/(CeO2)1 superlattices, Phys. Rev. Lett. 132, 256801 (2024).
  • Ma et al. [2023] L. Ma, J. Wu, T. Zhu, Y. Huang, Q. Lu, and S. Liu, Ultrahigh oxygen ion mobility in ferroelectric hafnia, Phys. Rev. Lett. 131, 256801 (2023).
  • Falkowski et al. [2018] M. Falkowski, C. Künneth, R. Materlik, and A. Kersch, Unexpectedly large energy variations from dopant interactions in ferroelectric HfO2 from high-throughput ab initio calculations, npj Comput. Mater. 4, 73 (2018).
  • Wang et al. [2023] Y. Wang, L. Tao, R. Guzman, Q. Luo, W. Zhou, Y. Yang, Y. Wei, Y. Liu, P. Jiang, Y. Chen, S. Lv, Y. Ding, W. Wei, T. Gong, Y. Wang, Q. Liu, S. Du, and M. Liu, A stable rhombohedral phase in ferroelectric Hf(Zr)1+xO2 capacitor with ultralow coercive field, Science 381, 558 (2023).
  • Fina and Sánchez [2021] I. Fina and F. Sánchez, Epitaxial ferroelectric HfO2 films: Growth, properties, and devices, ACS Appl. Electron. Mater. 3, 1530 (2021).
  • Wei et al. [2023] J. Wei, G. Xu, M. Huang, Y. Wu, W. Li, H. Xiang, and S. Chen, Intrinsic instability of rhombohedral (Hf,Zr)O2 phases and endurance crisis of Hf-based ferroelectric devices, Preprint at Research Square 10.21203/rs.3.rs-2634489/v1 (2023).
  • Zhu et al. [2023] T. Zhu, S. Deng, and S. Liu, Epitaxial ferroelectric hafnia stabilized by symmetry constraints, Phys. Rev. B 108, L060102 (2023).
  • Qi et al. [2020] Y. Qi, S. Singh, C. Lau, F.-T. Huang, X. Xu, F. J. Walker, C. H. Ahn, S.-W. Cheong, and K. M. Rabe, Stabilization of competing ferroelectric phases of HfO2 under epitaxial strain, Phys. Rev. Lett. 125, 257603 (2020).
  • Zhang et al. [2018] L. Zhang, J. Han, H. Wang, R. Car, and W. E, Deep potential molecular dynamics: A scalable model with the accuracy of quantum mechanics, Phys. Rev. Lett. 120, 143001 (2018).
  • Kresse and Furthmüller [1996] G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B 54, 11169 (1996).
  • Blöchl [1994] P. E. Blöchl, Projector augmented-wave method, Phys. Rev. B 50, 17953 (1994).
  • Kresse and Joubert [1999] G. Kresse and D. Joubert, From ultrasoft pseudopotentials to the projector augmented-wave method, Phys. Rev. B 59, 1758 (1999).
  • Perdew et al. [1996] J. P. Perdew, K. Burke, and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett. 77, 3865 (1996).
  • Monkhorst and Pack [1976] H. J. Monkhorst and J. D. Pack, Special points for Brillouin-zone integrations, Phys. Rev. B 13, 5188 (1976).
  • King-Smith and Vanderbilt [1993] R. D. King-Smith and D. Vanderbilt, Theory of polarization of crystalline solids, Phys. Rev. B 47, 1651 (1993).
  • Vanderbilt and King-Smith [1993] D. Vanderbilt and R. D. King-Smith, Electric polarization as a bulk quantity and its relation to surface charge, Phys. Rev. B 48, 4442 (1993).
  • Sheppard et al. [2012] D. Sheppard, P. Xiao, W. Chemelewski, D. D. Johnson, and G. Henkelman, A generalized solid-state nudged elastic band method, J. Chem. Phys. 136, 074103 (2012).
  • Dronskowski and Bloechl [1993] R. Dronskowski and P. E. Bloechl, Crystal orbital Hamilton populations (COHP): Energy-resolved visualization of chemical bonding in solids based on density-functional calculations, J. Phys. Chem. 97, 8617 (1993).
  • Maintz et al. [2016] S. Maintz, V. Deringer, A. Tchougréeff, and R. Dronskowski, Lobster: A tool to extract chemical bonding from plane-wave based DFT, J. Comput. Chem. 37, 1030 (2016).
  • Wu et al. [2021] J. Wu, Y. Zhang, L. Zhang, and S. Liu, Deep learning of accurate force field of ferroelectric HfO2Phys. Rev. B 103, 024108 (2021).
  • Wu et al. [2023] J. Wu, J. Yang, L. Ma, L. Zhang, and S. Liu, Modular development of deep potential for complex solid solutions, Phys. Rev. B 107, 144102 (2023).
  • [35] Ultrahigh oxygen ion mobility in ferroelectric hafnia, https://nb.bohrium.dp.tech/detail/1034100470.
  • Umari and Pasquarello [2002] P. Umari and A. Pasquarello, Ab initio molecular dynamics in a finite homogeneous electric field, Phys. Rev. Lett. 89, 157602 (2002).
  • Liu et al. [2016] S. Liu, I. Grinberg, and A. M. Rappe, Intrinsic ferroelectric switching from first principles, Nature 534, 360 (2016).
  • Brown [2009] I. D. Brown, Recent developments in the methods and applications of the bond valence model, Chem. Rev. 109, 6858 (2009).
  • Liu et al. [2013] S. Liu, I. Grinberg, H. Takenaka, and A. M. Rappe, Reinterpretation of the bond-valence model with bond-order formalism: An improved bond-valence-based interatomic potential for PbTiO3Phys. Rev. B 88, 104102 (2013).
  • Gagné and Hawthorne [2015] O. C. Gagné and F. C. Hawthorne, Comprehensive derivation of bond-valence parameters for ion pairs involving oxygen, Acta Cryst. B71, 562 (2015).
  • Wei et al. [2021] J. Wei, L. Jiang, M. Huang, Y. Wu, and S. Chen, Intrinsic defect limit to the growth of orthorhombic HfO2 and (Hf, Zr)O2 with strong ferroelectricity: First-principles insights, Adv. Funct. Mater. 31, 2104913 (2021).
  • Wu et al. [2024] J. Wu, J. Yang, Y. Sun, and S. Liu, Low-coercive-field ferroelectric hafnia with mobile domain walls, Preprint at arXiv 10.48550/arXiv.2403.09426 (2024).