Ferroelectric domain wall (DW) based nano-electronics is an emerging new field of research. It is only recently with advancements in electron and atomic force microscopy instrumentation that the complex nature of these 2D entities can be probed. In this Research Update, the advances in aberration corrected scanning transmission electron microscopy applied to ferroelectric topological defects are summarized. We discuss sub-atomic imaging and diffraction techniques used to observe changes in polarization, chemical composition, charge density, and strain at DWs and vortices. We further highlight the current achievements in mapping the 3D nature of ferroelectric polar skyrmions and in situ biasing. This Review will focus on both the fundamental physics of DW and polar vortex formation and their dynamics. Finally, we discuss how electron spectroscopy can be used to relate the quantified structural distortions of polar topological entities to changes in their oxidation state and band structure.

In the past 20 years, there has been a significant shift in interest within the ferroelectric community from domains to domain walls (DWs) as the active element for device applications.1 At almost the exact same time, commercially available aberration corrected scanning transmission electron microscopy (STEM) was achieved.2–4 This significant advancement in STEM enabled researchers to start analyzing the sub-atomic shifts at ferroelectric domain walls and polar interfaces,5 thus probing the fundamental physics at the spatial resolution of the DW itself. Previous theories regarding DW width, structure, and nucleation mechanisms are now regularly investigated experimentally via STEM.6,7 With further advances in STEM detectors, computing power, and post-processing, the previously suggested ideas such as complex polarization rotation at ferroelectric domain walls8 as seen in ferromagnetic DWs and room temperature ferroelectric skyrmions9 can now finally be proven experimentally.10 

The results from STEM characterization have made many impactful studies toward developing domain wall nanoelectronics. More specifically, this type of characterization is vital for optimizing thin film growth in terms of strain, thickness, and purity levels for specific DW type based devices. Substitutional doping in perovskite and manganite unit cells has been used to tune material properties such as magnetism11 and conductivity.12 Additional control of the domain formation is possible through epitaxial strain.

Atomic-level DW studies have understandably concentrated on the properties of unusual DWs in well-known ferroelectrics. Often these materials have been previously commercialized and thus have established growth processes, ground-state domain structures, and compatible dopants. We wish to highlight the opportunity of finding non-classical behavior and emergent phenomena by searching for new, improper ferroelectrics, which intrinsically contain charged domain walls. Such uncommon ferroelectrics are more typically found as bulk single crystals rather than thin films. While there are relatively fixed and predictable DWs in thin films, samples cut from bulk single crystals can vary widely in the domain structure.

Characterizing DW properties requires crossover of expertise between high-resolution electron microscopy techniques and ferroelectric properties. These properties offer challenges and opportunities for DW experiments. For example, the stability under the electron beam varies not only between different materials but also the type of DW and polar topological feature. Furthermore, while the crystal symmetry is the overriding energy consideration for DW orientation in bulk crystals, DW or vortex orientation in nanoscale samples such as TEM lamellae can vary significantly. These factors are important as the DW contrast is often the most accessible clue for understanding the overall domain structure and polarity. Thus, a careful understanding of TEM contrast mechanisms and the applicability of different polarization mapping techniques is crucial for DW studies.

In this Review, we discuss how aberration corrected STEM techniques can be used to obtain reliable maps of the polarization, electric fields, and charge distribution within ferroelectric DWs and other polar topologies such as vortices, and how to avoid the most common sources of error and artifacts attributed to these techniques. As detailed in Tables I and II, we have structured the review by physical scale and the associated electron microscopy techniques possible. We begin the review by describing how to identify the domain patterns in the bulk sample via scanning electron microscopy (SEM) and the most appropriate electron microscopy techniques to use with increasing magnification, leading to pico-meter characterization. As one of the most exciting aspects of ferroelectric topological entities is their mobility, we also detail the in situ biasing options to investigate their dynamics at the different magnifications. Finally, we highlight some of the most recent advances in STEM characterization methods for ferroelectrics such as visualizing electric charge density at sub-angstrom resolution13 and the benefits of coupling polarization characterization with electron energy loss spectroscopy (EELS) band structure analysis.14 

TABLE I.

Summary of the review sections.

Microscale: Domain configuration SEM, EBSD, FIB, TEM, 2 beam DF TEM, holography 
Nanoscale: Domain analysis STEM ADF, detector DPC, 4DSTEM (DPC, strain, phase) 
Picoscale: Domain wall/vortex Segmented detector DPC/iDPC, aberration corrected STEM ADF, aberration corrected STEM ABF, 
 4DSTEM (ptychography, electric field, electrostatic potential) 
Complementary techniques DW dynamics induced by the electron beam, in situ TEM holders, EELS 
Microscale: Domain configuration SEM, EBSD, FIB, TEM, 2 beam DF TEM, holography 
Nanoscale: Domain analysis STEM ADF, detector DPC, 4DSTEM (DPC, strain, phase) 
Picoscale: Domain wall/vortex Segmented detector DPC/iDPC, aberration corrected STEM ADF, aberration corrected STEM ABF, 
 4DSTEM (ptychography, electric field, electrostatic potential) 
Complementary techniques DW dynamics induced by the electron beam, in situ TEM holders, EELS 
TABLE II.

Electron microscopy techniques to use at different magnifications.

TechniqueExperienceCommon artifactsAdvantagesDisadvantages
Microscale: Domain configuration SEM Beginner Surface charging causes contrast reversal Large area analysis, ease-of-use Requires contrast interpretation, not all DWs visible, surface only, limited polarization information 
 EBSD Intermediate Scanning distortions when using long dwell times and large areas Large area analysis, crystal orientation information Limited polarization information 
 Low magnification TEM Beginner Dislocations, and bending contours appear similar to DWs Quick, large area analysis Limited polarization information, DWs must cause strain to show diffraction contrast 
 2-beam DF TEM Intermediate Diffraction contrast, tilt sensitive Large area analysis, identifying the domain configuration Time consuming, tilt sensitive, multiple images, and prior knowledge required for polarization interpretation 
 Holography Advanced Diffraction contrast Large area analysis Requires specific hardware, difficult post-processing 
Nanoscale: Domain analysis STEM DPC segmented detector Intermediate De-scan imperfection causing a diffraction pattern shift, diffraction contrast Quick, real-time, detailed polarization information Requires microscope re-alignment to transition from domain-scale analysis to DW analysis, segmented detector required 
 Low magnification STEM ADF Beginner Image becomes coherent at too-low collection angles, contrast reversal DWs exhibit strain-related contrast, can quickly transition from DW identification to DW analysis Very little quantitative knowledge gained at low magnification 
 GPA/Fourier masking Beginner Noise related artifacts Quick, requires only lattice resolution images, can give specific polarization information Prior knowledge of unit cells for each domain needed 
 Microprobe 4DSTEM Advanced Processing artifacts, de-scan imperfection Detailed structural information and polarization direction can be captured over a large area Post-processing required, acquisition time can be minutes long 
 
TechniqueExperienceCommon artifactsAdvantagesDisadvantages
Microscale: Domain configuration SEM Beginner Surface charging causes contrast reversal Large area analysis, ease-of-use Requires contrast interpretation, not all DWs visible, surface only, limited polarization information 
 EBSD Intermediate Scanning distortions when using long dwell times and large areas Large area analysis, crystal orientation information Limited polarization information 
 Low magnification TEM Beginner Dislocations, and bending contours appear similar to DWs Quick, large area analysis Limited polarization information, DWs must cause strain to show diffraction contrast 
 2-beam DF TEM Intermediate Diffraction contrast, tilt sensitive Large area analysis, identifying the domain configuration Time consuming, tilt sensitive, multiple images, and prior knowledge required for polarization interpretation 
 Holography Advanced Diffraction contrast Large area analysis Requires specific hardware, difficult post-processing 
Nanoscale: Domain analysis STEM DPC segmented detector Intermediate De-scan imperfection causing a diffraction pattern shift, diffraction contrast Quick, real-time, detailed polarization information Requires microscope re-alignment to transition from domain-scale analysis to DW analysis, segmented detector required 
 Low magnification STEM ADF Beginner Image becomes coherent at too-low collection angles, contrast reversal DWs exhibit strain-related contrast, can quickly transition from DW identification to DW analysis Very little quantitative knowledge gained at low magnification 
 GPA/Fourier masking Beginner Noise related artifacts Quick, requires only lattice resolution images, can give specific polarization information Prior knowledge of unit cells for each domain needed 
 Microprobe 4DSTEM Advanced Processing artifacts, de-scan imperfection Detailed structural information and polarization direction can be captured over a large area Post-processing required, acquisition time can be minutes long 
 

SEM provides a platform for selective domain contrast imaging depending on the chosen detector via a non-destructive characterization technique of the entire bulk sample. With a secondary electron detector, c+ (polarization pointing out of the surface) and c− (polarization pointing into the surface) domains exhibit darker and brighter contrast, respectively, than the a domains.15–19 This is due to positively and negatively charged domain surfaces repelling and absorbing SEs, respectively, but is true only either (1) at an equilibrium voltage where the beam does not charge the sample or (2) during the first few seconds of imaging, before the beam has charged the sample.20–22 It is known as voltage contrast. At lower voltages, the sample is positively charged and thus bright, while at higher voltages the sample is negatively charged and dark. Varying the voltage allows the equilibrium to be found for each material. Voltage contrast also applies to the a domains where it has been used to study point defects for a-type head-to-head or tail-to-tail domain walls.23 

The lattice rotation across the DWs and peak stress can be estimated by using the SEM based technique, electron backscatter diffraction (EBSD).24,25 Rather than precisely interpreting contrast, the knowledge of crystal orientation relative to the observed domain walls can often be sufficient to determine at least the polarization axes. For example, in tetragonal crystals, a DW along [110]pc means an a–a 90° DW and [100] means an a–c 90° DW angled 45° to the surface and wobbly or undefined curtain-like DWs mean c+/c− 180° DWs.26 If no DW contrast can be seen, ferroelastic domains can be identified by using EBSD to create a topology image similar to those often obtained by piezoresponse force microscopy. Mapping the approximate polarization in the SEM allows one to then target specific regions of interest for STEM characterization via combined (dual beam) SEM and focused ion beam (FIB) sample preparation.

SEM imaging techniques listed above can also be used to analyze the material within the FIB lamella and thus confirm if there are changes in the DW pattern after thinning. Piezoresponse force microscopy (PFM) and c-AFM27 mapping of the FIB lamella can also be done to quantify changes in polarization and conduction and relate these measurements back to the SEM imaging and diffraction. In general, epitaxial films and especially thin films are most likely to retain their domain structure because the domain formation is governed by nanoscale strain. Single crystals, on the other hand, are freer to adjust their domain structure to minimize their free energy during the lift-out and thinning of a lamella. This factor should be kept in mind when choosing the FIB lamella orientation, shape, thickness, and whether to include extra thin “windows.” Altering the FIB preparation techniques can result in drastically different strain throughout the lamella and thus DW pattern confinement effects.28 In this way, the local strain state of the lamella can be designed to preserve existing DWs or create new ones.29–31 Final thinning by argon ion milling requires additional consideration for ferroelectrics.32 The surface damage induced by gallium ion milling provides a polarizable “dead layer” on the surface, which makes c domains more easily stabilized but negates the need to form 180° c+/c− DWs for electrostatic screening. Thus, removing the dead layer with argon polishing may induce 180° c+/c− DWs, especially in “Z-cut” uniaxial crystals.

More recently, studies have shown that the SEM probe can also be used to induce domains, switch their polarization direction, and even move the DWs.33–36 Kianirad et al. detailed the motion of the DW can be controlled via the dose rate, total dose, and scan direction,37 as shown in Fig. 1. Sections IV A and IV B will discuss the TEM/STEM based in situ biasing DW dynamic techniques.

FIG. 1.

SEM characterization of DW motion in LiTaO3 for Ps↑ domains surrounded by Ps↓ domains (a)–(c). Domain wall motion for Ps↑ surrounding Ps↓ domains (d)–(f). The electron beam was scanned horizontally and from top to bottom in the images. The acceleration voltage was 0.3 kV, and the scanned area was 0.12 × 0.09 mm2.37 

FIG. 1.

SEM characterization of DW motion in LiTaO3 for Ps↑ domains surrounded by Ps↓ domains (a)–(c). Domain wall motion for Ps↑ surrounding Ps↓ domains (d)–(f). The electron beam was scanned horizontally and from top to bottom in the images. The acceleration voltage was 0.3 kV, and the scanned area was 0.12 × 0.09 mm2.37 

Close modal

Before high resolution (HR)-STEM analysis of a DW can be carried out, the desired DW must first be found and the polarization in the surrounding domains identified. Diffraction contrast in low-magnification TEM can be used initially to see the domain structure of the lamella as a whole. If the aim is to monitor domains over a large area, then specific modes of TEM can be used. Tilting to a two-beam dark field condition can isolate intensity from a specific diffraction spot. In ferroelectric crystals, some diffraction spots break Friedel’s law and can be used to directly relate the image intensity to the polarization direction (Fig. 2).38–40 The drawback is that multiple images must be carefully acquired under different conditions to map the polarity of the entire domain structure (Fig. 2). More simply, an off-centered objective aperture will enhance diffraction contrast at low-medium magnification TEM to detect any subtle DWs present, which is particularly useful for in situ studies.41 

FIG. 2.

G-vector dependence of contrast in the dark field images for domain I in the FM phase, together with a model of domain I. These images were taken from the x = 0.47 sample at room temperature using the (a) 002 A, (b) 002 B, (c) 110 B, and (d) ¯ 110 B reflections, where the subscripts A and B indicate two original tetragonal variants. In (e), the resultant polarization-vector directions produced by a sum of [001] and ⟨110⟩ components are denoted by the heavy arrows.38 

FIG. 2.

G-vector dependence of contrast in the dark field images for domain I in the FM phase, together with a model of domain I. These images were taken from the x = 0.47 sample at room temperature using the (a) 002 A, (b) 002 B, (c) 110 B, and (d) ¯ 110 B reflections, where the subscripts A and B indicate two original tetragonal variants. In (e), the resultant polarization-vector directions produced by a sum of [001] and ⟨110⟩ components are denoted by the heavy arrows.38 

Close modal

STEM imaging relies on collecting electrons that have passed through a thin sample by a scanned beam. This imaging technique has two main differences in comparison to conventional TEM: (1) The detector is placed in the backfocal plane where the diffraction space information is held and (2) the intensity of each pixel in a STEM image is collected separately, allowing highly specific, localized information to be isolated from adjacent areas. STEM annual dark field (ADF) imaging uses a doughnut-shaped annular detector. The high angle annular dark field (HAADF) collects highly scattered electrons, thus providing useful images of heavier atoms differentiated by intensity. The annular bright field (ABF) collects weakly scattered electrons and is particularly useful for imaging lighter elements.

In comparison to TEM, STEM imaging is less impacted by diffraction contrast, and thus, it can be difficult to identify domains with small spontaneous strain differences. To maximize diffraction contrast and identify domains in the sample, microprobe STEM can be used at the cost of atomic resolution capabilities. Alternatively, it is recommended to use an inner collection angle of the annular detector very close to the convergence angle, low angle-annular dark field (LAADF).29,42 Using LAADF gives the added advantage of quickly transitioning from low to high resolution STEM without realignment. In this way, much of the polarity for individual domains can be identified based on DW orientations relative to the crystal structure. However, when investigating unconventional DW configurations or novel materials, interpreting the DW contrast to make a “most-plausible domain polarization” map can be overly simplistic,43 and more in-depth techniques are required.

Geometric phase analysis (GPA) allows quantitative strain maps to be determined for any lattice resolution image.44,45 Conventional TEM atomic resolution images are sensitive to thickness changes and subject to contrast reversal, making large-area strain mapping sometimes problematic.46 Conversely, STEM is relatively insensitive to surface defects, thickness changes, and sample tilts. STEM allows fields-of-view of over 100 nm to be strain-mapped, narrowing down the possible polarization for each domain to two opposite directions. The direction of lattice rotation across PbTiO3 DWs, for example, can be used to fingerprint the exact polarization in each domain and, by extension, the entire domain structure.47 Thus, GPA strain and rotation maps have been used extensively to map the domain structure in various investigations of PbTiO3.48–51 GPA fingerprinting of polarization should be possible for other ferroelectrics exhibiting lattice rotation at DWs, e.g., LiNbO352 and BiFO3.

By masking fast Fourier transforms (FFTs), domain mapping can be further simplified.4,53 Rather than displaying relative changes in spacing/strain as GPA does, FFT masking displays the location in the image of specific spacings/strain states. Thus, while it cannot be used to identify small strain changes or produce rotation maps, FFT masking allows specific “superlattice” or unique structural spacings that characterize individual domains or DWs to be identified in a real-space image. These unique atomic spacings are especially prevalent in non-perovskite ferroelectrics with complex unit cells. Boracites,54,55 for example, have a unique superlattice reflection related to the polarization axis in the domain. FFT masking is further useful for cryogenic investigations such as those on charge-ordered manganites,56 where the sample drift is an issue because it is subject to the same flexible image conditions as GPA.

As the domain and DW formation depends on nanoscale strain,57 it is vital to investigate the coupling of polarization to spontaneous strain around DWs and within the 2D features itself. FFT53 and GPA51 can be used to map out the strain states of the domains. However, the quantification of strain states by these methods is only possible if the region in the field of view is highly strained. The electron diffraction patterns collected from 4DSTEM datasets58 allow researchers to measure very low changes in strain percentages using patterned probes59 and even weakly scattering 2D materials using direct electron detectors.60 Another advantage of using 4DSTEM strain mapping for ferroelectric materials is the flexibility in the possible magnifications while retaining high precision. This is vital for investigating the internal strain measurements of individual ferroelectric DWs and vortices at the picoscale while being able to relate these measurements back to the bulk domain configuration of the TEM lamella. It is important to note that there is a slight trade-off between the resolution in real and reciprocal space; however, with advancements in beam shaping, beam tilting (precision electron diffraction), probe patterning, camera speeds, and improved data processing, this limitation is becoming less of an issue.61–65 

STEM differential phase contrast (DPC) and center of mass (CoM) imaging have become the tool of choice for visualizing internal electric and magnetic fields. Beam deflections due to internal fields were initially detected by single detectors using Fresnel and Foucault imaging, but it was a delicate, time-consuming, and difficult-to-control task to produce interpretable image contrast.66 In contrast, the development of quadrant annular detectors allowed multidirectional beam shifts to be interpreted from a single acquisition.67 Images detecting shifts in a STEM BF disk partially overlapping an annular quadrant detector are generally referred to as DPC.68 When studies identified that redistribution of intensity within the BF diffraction disk could also contribute to the contrast in DPC images,69 it became recognized that the contrast is finally determined by the CoM of the BF disk. Thus, the terminology of CoM imaging was introduced. However, while the CoM can be approximated from quadrant detectors, it can be more accurately measured by recording the entire BF disk, as in 4DSTEM.70 Hence, polarization maps constructed from 4DSTEM datasets are also referred to as CoM imaging.13 

When mapping polarization in a ferroelectric domain via DPC, one is trying to measure the effect of net electric field (E) in the sample. In the absence of any external biasing, this means the spontaneous polarization (Ps) minus the depolarization (D) caused by the screening charge at DWs, as explained by the following equation:
(1)
and the change in D, E, and Ps moving across a DW is related to the screening charge (ρc) by
(2)
As detailed by MacLaren et al.,69 when the DW is fully screened,
and
2Ps for a 180° DW, where the DW is perpendicular to Ps

In most examples of ferroelectric DPC studies, there is a measured electric field, and thus, the polarization is not fully screened. Equation (2) also means that taking the differential of DPC maps to map the charge density measures the net of (bound charge due to polarization) – (screening charge). It is important to clarify that maps of εrε0∇ · E represent the unscreened portion of the bound charge at the DW, not the local free charge density (ρc). However, ρc at the DW can be calculated by adding the differential of the DPC map to ∇ · Ps [from Eq. (2)].69 ∇ · Ps may need to be calculated using phase-field simulations for complex DWs.14,71,72

When the electric field gradient is greater than the probe size, as is the case for nanoscale electric fields across domains vs an electron probe, the effect is purely a beam deflection.73,74 It is recommended to use microprobe STEM with a small convergence angle α of ∼ 1 mrad to increase sensitivity to small deflections. Under the kinematical (single scattering) approximation, the deflection angle (Θ) is expressed as75,76
(3)
where CHT=e(E0+Ee)Ee(2E0+Ee) is the interaction constant that corrects for relativity.

Ee is the electron energy, e is the elementary charge, λ is the electron wavelength, E0 is the electron rest energy, Ui is the mean inner potential,77 Ue is the (additional) electrostatic potential, t is the thickness, and ∇rt is the local thickness gradient. When there is no electrostatic potential, Ue = 0, Eq. (3) can be used to measure the mean inner potential or the thickness gradient of a sample.75 More importantly, for ferroelectrics, when there is a uniform thickness, Eq. (3) can quantify the net electric field (E = ∇rUe) from the measured beam deflection θ.

Typical electric field strengths are 3–4 orders of magnitude smaller than atomic potential gradients at 0.1 V/nm–1 V/nm. Thus, increasing the relative sample thickness either physically or by using a lower accelerating voltage is recommended to increase the deflection angle and thus the sensitivity.73 The deflection angle can be measured directly via 4DSTEM, but DPC maps require normalization by the sum of the quadrants and calibration for quantitative interpretation.75 Some of the sources of error are channelling effects, strong diffracting conditions exciting asymmetric Friedel’s pairs, and smaller potential gradients (defects or interfaces), causing asymmetry within the BF disk. Each of these can cause artifacts in a CoM-derived electric field map from DPC or 4DSTEM data. Normalization and tilting away from zone axis to achieve quasi-kinematical conditions can mitigate these effects, and a vacuum dataset can correct for scanning-induced beam shifts if de-scan settings are imperfect.78 

The best practice for measuring E-fields is to use microprobe STEM with an objective aperture blocking all diffracted beams and place the BF disk just inside the inner detector radius (DPC) or inner virtual detector radius (4DSTEM).74,79,80 Thus, there is no signal/overlap of the BF disk in the absence of a long-range potential gradient and the CoM can be ignored to focus on the differential signal intensity. Such artifact-free ferroelectric DPC maps should then resemble magnetic field maps with DWs appearing as black boundaries between uniformly colored domains.43 

For researchers investigating multiferroic domain walls and needing to correlate the electric and magnetic fields in a single multiferroic, the sample should be achievable by changing the accelerating voltage between, e.g., 60 kV and 300 kV. At 60 kV, the sample appears approximately seven times thicker to the electron beam, increasing the deflection due to the electric field. Meanwhile, the electron velocity is reduced by ∼43%, decreasing the deflection due to the magnetic field proportionally. The electric field DPC map can be acquired at 60 kV, and then, the voltage changed to 300 kV, changing the contributions to the beam deflection by an order of magnitude, to measure a DPC map of the (perpendicular) magnetic field. Thus, it should be possible to map out both the magnetic and electric field transitions across magnetoelectric domain walls.

Aberration corrected TEM was first used by Jia et al.5 to image the change in investigation of the cation–oxygen dipoles per unit cell at ferroelectric DWs. By using the negative spherical-aberration imaging,2,81,82 they were able to experimentally quantify the large difference in atomic displacement at charged vs uncharged DWs. This type of imaging technique has been used to confirm non-classical ferroelectric allowed polarization theories such as dipole continuous rotation in vortex structures,83 unique cycloidal polarization order,84 and Néel-like polarization in ferroelectric DWs,85 as shown in Fig. 3. However, atomic resolution TEM imaging is hindered by the inability to map out the lighter elements in more complex unit cells of other materials, and the atomic resolution imaging is sensitive to thickness changes and subject to contrast reversal. In this Review, we will focus on advances in aberration corrected atomic resolution STEM imaging and other STEM techniques such as 4DSTEM.

FIG. 3.

The panoramic view of nesting tetragonal and monoclinic domains viewed along the [110]T direction. The yellow dotted lines trace the boundaries between the tetragonal and monoclinic phases, and the white dotted lines trace the domain walls in the monoclinic phase. The phase and domain boundaries are differentiated by mapping the relative displacements of O2 columns (red arrows) with respect to centers of the nearest neighboring Zr/Ti columns over the whole image. Scale bar: 1 nm. (b)–(d) Enlarged view of the relative displacements of the O2 columns in domains M-I, T-I, and M-II in (a). The colored circles overlapped on the images denote different column types: Pb/O1—yellow, Zr/Ti—blue, and O2—red. The vertical dashed and solid lines indicate positions of the Zr/Ti columns and centers of their nearest neighbors. The relative displacements of O2 columns with respect to these centers can be directly identified in this way.85 

FIG. 3.

The panoramic view of nesting tetragonal and monoclinic domains viewed along the [110]T direction. The yellow dotted lines trace the boundaries between the tetragonal and monoclinic phases, and the white dotted lines trace the domain walls in the monoclinic phase. The phase and domain boundaries are differentiated by mapping the relative displacements of O2 columns (red arrows) with respect to centers of the nearest neighboring Zr/Ti columns over the whole image. Scale bar: 1 nm. (b)–(d) Enlarged view of the relative displacements of the O2 columns in domains M-I, T-I, and M-II in (a). The colored circles overlapped on the images denote different column types: Pb/O1—yellow, Zr/Ti—blue, and O2—red. The vertical dashed and solid lines indicate positions of the Zr/Ti columns and centers of their nearest neighbors. The relative displacements of O2 columns with respect to these centers can be directly identified in this way.85 

Close modal

With the development of aberration correctors, STEM imaging modes can be used to identify atomic column positions down to a precision of 3 pm–5 pm.86–89 Thus, for ABO3 perovskite ferroelectrics, HAADF can identify the heavier “A” and “B” site atoms, while ABF can be used to see the displacements or rotations of oxygen octahedra (Fig. 4).90–96 These complimentary aspects of high resolution (HR) STEM are crucial as the positions of all atoms in the unit cell must be measured to determine the displacements with respect to the paraelectric phase and thus the net polarity. For some more complex unit cells, STEM ABF is not able to resolve all the positions of the lighter elements. More recently, a segmented ADF detector integrated differential phase contrast (iDPC) imaging technique97,98 can be used to locate the positions of the required lighter elements for polarization direction determination.89 The technique iDPC will be discussed in more detail later in Sec. III A 1. Once the position of the lighter elements relative to heavier elements for the polarization direction has been established, HAADF alone can measure the displacement of the heavier atoms, while the position of invisible lighter atoms can be assumed. For example, in the well characterized PbTiO3, O and Ti columns move in the same direction away from the neutral position at the center of the unit cell. However, O2− moves further than Ti4+99 so that the polarization, pointing negative to positive, forms exactly opposite to the direction of Ti4+ displacement. This phenomenon allows the “reverse B-site displacement vector” method of using only HR-STEM HAADF images to place arrows denoting polarization vectors on each unit cell of PbTiO3 or BiFeO3.100–104 

FIG. 4.

Simulated annular bright-field scanning transmission electron microscopy (ABF-STEM) images along the zone axis of [11 ¯ 0] for a−a−a− (a) and [101] for a−b+a− (b) OOR patterns. (c) and (g) show the ABF-STEM images of the 2.6 nm-thick CaTiO3 (111) film along the zone axis of [11 ¯ 0], measured using 80-keV (c) and 200-keV (g) electron kinetic energy. (d) and (h) display the maps of OOR patterns, identified by deep neural network analysis, in the same regions as in (c) and (g), respectively. The color indicates the probability of each OOR pattern. (e) and (i) present the polarization vectors for each unit cell of the same regions as in (c) and (g), respectively. The arrows denote the polarization direction; the stronger the polarization, the darker the arrow color. The strength of polarization is also expressed as a color map, ranging from white (weak) to red (strong). (f) illustrates a schematic free energy landscape, showing the relaxation of R3c state (a−a−a−) into Pnma state (a−b+a−).90 

FIG. 4.

Simulated annular bright-field scanning transmission electron microscopy (ABF-STEM) images along the zone axis of [11 ¯ 0] for a−a−a− (a) and [101] for a−b+a− (b) OOR patterns. (c) and (g) show the ABF-STEM images of the 2.6 nm-thick CaTiO3 (111) film along the zone axis of [11 ¯ 0], measured using 80-keV (c) and 200-keV (g) electron kinetic energy. (d) and (h) display the maps of OOR patterns, identified by deep neural network analysis, in the same regions as in (c) and (g), respectively. The color indicates the probability of each OOR pattern. (e) and (i) present the polarization vectors for each unit cell of the same regions as in (c) and (g), respectively. The arrows denote the polarization direction; the stronger the polarization, the darker the arrow color. The strength of polarization is also expressed as a color map, ranging from white (weak) to red (strong). (f) illustrates a schematic free energy landscape, showing the relaxation of R3c state (a−a−a−) into Pnma state (a−b+a−).90 

Close modal

1. Integrated differential phase contrast imaging

While iDPC was proposed by Rose,105 its commercialization with a segmented annular detector has led to widespread adoption.106 iDPC has a contrast transfer function closely proportional to Z1, as opposed to Z1.7 for HAADF. Thus, iDPC allows light elements to be efficiently imaged alongside heavier elements.89,98,107,108 Atomic resolution CoM-DPC derived from annular quadrants or 4DSTEM data74 requires a large convergence angle α of ∼ 30 mrad. In contrast to the deflection described for nanoscale E fields above, the effect of (atomic) potentials smaller than the probe size is a redistribution of intensity in the ∼30 mrad BF disk. CoM represents the expectation value of the electron beam’s quantum mechanical probability current flow through the sample (Fig. 5).109,110

FIG. 5.

Structural characterization of the PbTiO(n)/SrTiO(10) multilayer. Superposition of the iDPC-STEM image and a structure model of PbTiO3, showing the atom displacements around the vortex structure. The red, green, and yellow dots denote the positions of Pb2+, Ti4+, and O2− columns, respectively. The arrows denote the polarization direction. The dashed lines indicate the interface between PbTiO3 and SrTiO3. The zoom-in images attached to this iDPC-STEM image exhibit the unit cell of PbTiO3 along [010]pc in which positions of Ti, O, and Pb are clearly shown. Scale bar: 1 nm.111 

FIG. 5.

Structural characterization of the PbTiO(n)/SrTiO(10) multilayer. Superposition of the iDPC-STEM image and a structure model of PbTiO3, showing the atom displacements around the vortex structure. The red, green, and yellow dots denote the positions of Pb2+, Ti4+, and O2− columns, respectively. The arrows denote the polarization direction. The dashed lines indicate the interface between PbTiO3 and SrTiO3. The zoom-in images attached to this iDPC-STEM image exhibit the unit cell of PbTiO3 along [010]pc in which positions of Ti, O, and Pb are clearly shown. Scale bar: 1 nm.111 

Close modal

2. Ptychography

Ptychography is a well established technique in light112 and x-ray113–116 based characterization; however, it is only recently the technique has been exploited in STEM research.117–120 The main benefit of potentially using pytchography for atomic resolution polarization mapping at ferroelectric DWs and polar vortices would be the ability to image clearly the lighter elements within even complex unit cells. The main benefit of using ptychography over segmented detector iDPC imaging is that much lower doses can be used, achieving the same potential atomic column resolution. Additionally, as described by Yang et al.,118 the phase information extracted from a 4DSTEM dataset collected on a direct electron detector while simultaneously collecting a z-contrast image on a HAADF detector. The main drawback of ptychography currently is the time taken to process the data, whereas iDPC can be done live via commercial software from most microscope vendors.

In order to maximize the information that can be extracted from atomic resolution TEM and STEM images, there are a number of steps that can be taken during and after acquisition. Live drift corrected frame integration121 helps signal-to-noise while allowing a lower dose through a smaller beam current and thus a smaller electron probe and better resolution. Furthermore, summing images with orthogonal scan directions can correct for scanning distortions,122 especially with the use of non-rigid registration software.123–125 Post-acquisition, Fourier space filters such as double Gaussian can be used to further reduce noise,126 and statistical reconstruction127,128 can be used to further clean up noisy data.129–131 This approach has been used, in particular, for samples like hexagonal rare-earth manganites where spacing between atomic columns is smaller within the unit cell compared to PbTiO3,132 and a more complex atomic displacement corresponding to their polarization,133,134 and beam sensitive ferroelectric materials like LiNbO3.52 

Following image filtering, mapping the atomic displacements and thus assigning polarization per unit cell can typically be automated using 2D Gaussian fitting python based programs such as Atomap135 and TEMUL toolkit,136 Pycroscopy,137 and Matlab based scripts (Fig. 6).5,101,138,139

FIG. 6.

(a) Sketch of a planar-view HAADF-STEM image overlaid with a cross-sectional strain map, displaying an overview of the PbTiO3 film. (b) A reversed Ti-displacement vector map (front) based on an atomic-resolved cross-sectional HAADF-STEM image (back), showing a trapezoidal domain with the convergent polarization configuration. In the back HAADF-STEM image, the yellow and red circles denote the Pb and Ti atom columns, respectively, and the yellow arrows denote the directions of −δTi vectors. (c) A reversed Ti-displacement vector map based on the atomic-resolved planar-view HAADF-STEM image of stripe domains, showing the convergent Meron arrays (black circles) at h-t-h domain walls (black solid lines). (d) A reversed Ti-displacement vector map (top) of a magnified polar Meron array with the convergent polarization configuration based on an atomic-resolved planar-view HAADF-STEM image (bottom).139 

FIG. 6.

(a) Sketch of a planar-view HAADF-STEM image overlaid with a cross-sectional strain map, displaying an overview of the PbTiO3 film. (b) A reversed Ti-displacement vector map (front) based on an atomic-resolved cross-sectional HAADF-STEM image (back), showing a trapezoidal domain with the convergent polarization configuration. In the back HAADF-STEM image, the yellow and red circles denote the Pb and Ti atom columns, respectively, and the yellow arrows denote the directions of −δTi vectors. (c) A reversed Ti-displacement vector map based on the atomic-resolved planar-view HAADF-STEM image of stripe domains, showing the convergent Meron arrays (black circles) at h-t-h domain walls (black solid lines). (d) A reversed Ti-displacement vector map (top) of a magnified polar Meron array with the convergent polarization configuration based on an atomic-resolved planar-view HAADF-STEM image (bottom).139 

Close modal

While mapping the domain polarity provides important information about the likely macroscopic properties and behavior of a DW, the true power of aberration corrected STEM lies in probing the interior of the DW itself. Using specific alignment conditions, 4DSTEM datasets can also be used to measure the internal electric fields and electrostatic potentials of ferroelectric topological features.140 Particularly, effective studies embracing this approach include the studies of Yadav et al.141 and Gao et al.13 Both studies measured the polarization vectors from atomic imaging and the electric field/charge density from 4DSTEM. Yadav et al. used the independent measurements to estimate the free energy gradient near polar vortexes and thus confirmed the existence of local areas of negative capacitance. Gao et al. demonstrated a spatial difference between the polarization and the charge density and persuasive evidence that the 4DSTEM-derived electric field/charge density was an independent measurement to that of the atomic structure. When the samples are thicker than 7 nm and beam broadening cannot be neglected, atomic resolution CoM maps should not be interpreted as the gradient of the atomic potential, but the gradient of the probe (Fig. 7).73,74,109,142 As extremely thin samples are required, the ideal samples would be 2D ferroelectric materials such as CuInP2S6.143–145 

FIG. 7.

(a) Sketch of a ferroelectric crystal slab when no external field is applied. The small black arrows give the microscopic dipoles of the BiFeO3 pseudocubic unit cells, while the large gray arrow depicts the mesoscopic field due to the whole crystal polarization. (b) Effect of a uniform ferroelectric polarization on the potential slices for multislice simulations. An additive linear term induces disruption of the periodicity of the potential slices. (c) Modified multislice approach including the interaction of the electrons with a transverse electric field in the free-space propagation. After interacting with the potential slice Vi, the electron wavefunction is propagated in the free space until the next potential slice Vi + 1, applying a fractional beam tilt corresponding to the effect induced by the ferroelectric polarization of one slice.142 

FIG. 7.

(a) Sketch of a ferroelectric crystal slab when no external field is applied. The small black arrows give the microscopic dipoles of the BiFeO3 pseudocubic unit cells, while the large gray arrow depicts the mesoscopic field due to the whole crystal polarization. (b) Effect of a uniform ferroelectric polarization on the potential slices for multislice simulations. An additive linear term induces disruption of the periodicity of the potential slices. (c) Modified multislice approach including the interaction of the electrons with a transverse electric field in the free-space propagation. After interacting with the potential slice Vi, the electron wavefunction is propagated in the free space until the next potential slice Vi + 1, applying a fractional beam tilt corresponding to the effect induced by the ferroelectric polarization of one slice.142 

Close modal

Yadav et al.141 took a different approach to mapping out the polar skyrmions, more akin to the description above for electric field mapping in domains. Polar skyrmions are effectively merged/continuous ferroelectric domains with a more gradual potential gradient than atomic fields. Thus, Yadav et al. used a smaller convergence angle to separate the diffraction disks and measured exclusively the shift of the BF disk on the pixelated detector. In this way, the longer range electric fields are effectively decoupled from atomic potential gradients and can be compared to the polarization vectors.

By combining aberration corrected TEM imaging and fast cameras with in situ biasing holders, experimentation not possible by any other method at the same spatial and time resolution is now accessible to the ferroelectric research community. Understanding the relationship between the DW and vortex formation during phase changes within ferroelectric materials is essential for future nanoelectronics based on these polar entities. In 2019, Linze et al.146 published a thorough review on the progress of in situ biasing TEM holder based microscopic processes of ferroelectric domain switching. There has been extensive work to date on live imaging during the in situ biasing via lower magnification TEM to investigate domain nucleation, domain switching, relaxation, and interplay between different types of domains (Fig. 8).28,111,147–155 These in situ biasing experiments have improved our understanding of the underlying physics governing ferroelectric domain dynamics. However, to date, real-time biasing quantifiable sub-atomic imaging has not been achieved. With improvements in TEM cameras and in situ biasing holders, this is the next hurdle to overcome for the TEM ferroelectric community.

FIG. 8.

(a) Schematic and TEM image of the experimental setup: A thin cross-sectional Pb(Zr0.2Ti0.8)O3 film with a concentration of ferroelastic 90° domains (a-domains) was grown on SrRuO3/DyScO3 and SrRuO3/SrTiO3. A mobile tungsten tip acts as one electrode for electrical switching with the SRO layer being grounded, whereas a diamond indenter is used for mechanical switching. (b) High-resolution HAADF STEM image of c/a/c-domains overlaid with vectors describing the head-to-tail polarization arrangement. Scale bar: 1 nm. (c) Image sequence showing a clamped 90° domain is stable at applied negative voltages from 0 → (−18) V. At −19 V, the domain is eventually erased and, simultaneously, damage to the film occurs due to the high strength of the electric field. Scale bar: 100 nm.150 

FIG. 8.

(a) Schematic and TEM image of the experimental setup: A thin cross-sectional Pb(Zr0.2Ti0.8)O3 film with a concentration of ferroelastic 90° domains (a-domains) was grown on SrRuO3/DyScO3 and SrRuO3/SrTiO3. A mobile tungsten tip acts as one electrode for electrical switching with the SRO layer being grounded, whereas a diamond indenter is used for mechanical switching. (b) High-resolution HAADF STEM image of c/a/c-domains overlaid with vectors describing the head-to-tail polarization arrangement. Scale bar: 1 nm. (c) Image sequence showing a clamped 90° domain is stable at applied negative voltages from 0 → (−18) V. At −19 V, the domain is eventually erased and, simultaneously, damage to the film occurs due to the high strength of the electric field. Scale bar: 100 nm.150 

Close modal

As described earlier, the electron beam of an SEM can be used to induce and move DWs while imaging. Hart et al.156 was the first to show that the TEM beam could be used to induce ferroelectric domain nucleation in PbTiO3 (Fig. 9), and Barzilay157 used atomic resolution TEM to form 2 nm domain periodicity in BaTiO3. Recently, there has been a surge in interest in atomic scale manipulation of 2D materials using an aberration controlled STEM probe.158–162 Conroy et al.55 showed that ferroelectric DWs can also be controllably moved via the STEM probe for the boracite material while monitoring sub-atomic shifts and thus polarization changes during movement. Using the STEM probe as an applied electric field allows one to study the dynamics of ferroelectric domain walls under vast different experimental conditions such as probe scan direction, speed of scan, dose, and probe size without complicated in situ experimental preparation. However, this may not be possible for all ferroelectric materials and most likely is restricted to materials with low coercive fields.

FIG. 9.

(a) Dark field TEM images showing electron beam-induced domain motion after 1 h of uniform irradiation. The dark arrows indicate domain polarization. All panels show the same domain. The top panel images show where the domain intersects the top edge, and the bottom panel images show where the domain intersects the bottom edge. (b) Observation of multiple c− domain nucleation sites along the bottom edge after uniform irradiation. (c) A c− domain that retracted 1 µm from the top edge after 1 h of uniform irradiation.156 

FIG. 9.

(a) Dark field TEM images showing electron beam-induced domain motion after 1 h of uniform irradiation. The dark arrows indicate domain polarization. All panels show the same domain. The top panel images show where the domain intersects the top edge, and the bottom panel images show where the domain intersects the bottom edge. (b) Observation of multiple c− domain nucleation sites along the bottom edge after uniform irradiation. (c) A c− domain that retracted 1 µm from the top edge after 1 h of uniform irradiation.156 

Close modal

Atomic scale core loss EELS has been used extensively in ferroelectric oxide thin film hetero-interfaces, relating changes in polarization to chemical bonding and oxidation states.163,164 More recently, with improved STEM and spectrometer designs, the energy resolution now possible has allowed researchers to measure the core loss fine structure changes within the DWs of homogeneous materials. Mundy et al.165 showed the clear Mn valence change at charged DWs in manganites, and Rojac et al.166 measured Fe4+ cations and bismuth vacancies at DWs, revealing p-type conduction at domain walls caused by the presence of electron holes associated with Fe4+.

Monochromated high-energy resolution low loss EELS has been used to probe changes in the electronic structure at quantum wells and 2DEGs for other material systems. More recently, Zhang et al.14 has shown the clear change in bandgap at conducting domain walls in BiFeO3, and Conroy et al.167 has shown the change in bandgap and also the signal enhancement of the energy loss intensity in the 2 eV–3.5 eV regime in conducting LiNbO3 DWs. With advances in spectrometer design, low loss EELS can now also be used to map out changes in phonon modes at the atomic scale in 2D materials.168 We propose this technique would be ideal to map out the phonon modes in ferroelectric DWs at the spatial resolution needed to understand the fundamental physics governing the ferroelastic ferroelectric coupling.

Aberration corrected STEM is the only experimental technique that allows one to probe the internal physical and electronic structure of ferroelectric DWs and polar vortices at the picoscale. Advances in our knowledge of the fundamental physics that govern the formation of 3D polar topological features such as room temperature skyrmions have only been made possible by pushing the limits of STEM techniques.169 We can now directly relate theoretical density functional theory (DFT) studies with experimental results, hence increasing the speed of material research and development design. With the end goal of using these polar topological entities for devices based on their dynamics, in situ STEM allows the research community to investigate exactly how the atomic displacements occur during movement. Thus, informing us of how to improve our materials to lower the energy required to move these ferroelectric polar entities. In this Review, we have laid out how to identify ferroelectric regions of interest via electron microscopy from the bulk sample scale and then step by step details of what types of electron microscopy characterization techniques can be used with increasing magnification finalizing in aberration corrected picoscale STEM, as shown in Table II.

There are various ways in which phase effects and 4DSTEM datasets promise to improve atomic scale electric field measurements. Recently, there has been work on re-shaping or even splitting the STEM beam170 to use the electron dose more efficiently, thus obtaining the same polarization data more efficiently. These methods can also make use of interference effects generated by multi-beam scanning to gain further phase sensitivity. There has been a major focus on controlled beam sculpting into vortex patterns,171–174 thus opening up further types of available characterization techniques for 3D polar ferroelectric topologies, as seen in ferromagnetic materials. Multi-beam electron diffraction is another fascinating development, which may enable efficient phase sensitivity.175 Time resolution at the attosecond timescale is also possible via microscopes with laser pumped sources.176 As described in the STEM perspective by Idrobo,177 there has been a renaissance of using cryogenic temperatures (liquid nitrogen or helium) for physical materials. El Baggari et al.56 have shown how aberration corrected cryogenic STEM can be used to quantify the atomic displacement within the unit cell between phases and thus investigate the broken symmetry states. As certain ferroelectric materials have a ferromagnetic phase at lower temperatures, cryogenic STEM would also allow researchers to study the charge to spin relationship of these multiferroic phases. The STEM is a powerful one-stop laboratory to characterize ferroelectric topological defects and their dynamics by combining polarization mapping, oxidation mapping, and band structure changes all at DW or polar entity resolution via EELS and in situ biasing.

All authors contributed to the discussions and manuscript preparation.

The data that support the findings of this study are available within the article.

This research was supported by the US-Ireland R&D Partnership Programme (Grant No. USI 120) and the Science Foundation Ireland (Grant No. 16/US/3344). M.C. acknowledges funding from the SFI Industry Fellowship (Grant No. 18/IF/6282) and Royal Society University Research Fellowship (URF∖R1∖201318).

4DSTEM

four-dimensional scanning transmission electron microscopy

CoM

center of mass

DPC

differential phase contrast

DW

domain wall

EELS

electron energy loss spectroscopy

FFT

fast Fourier transform

FIB

focused ion beam

GPA

geometric phase analysis

SEM

scanning electron microscopy

STEM

scanning transmission electron microscopy

TEM

transmission electron microscopy

4DSTEM

four-dimensional scanning transmission electron microscopy

CoM

center of mass

DPC

differential phase contrast

DW

domain wall

EELS

electron energy loss spectroscopy

FFT

fast Fourier transform

FIB

focused ion beam

GPA

geometric phase analysis

SEM

scanning electron microscopy

STEM

scanning transmission electron microscopy

TEM

transmission electron microscopy

1.
G.
Catalan
,
J.
Seidel
,
R.
Ramesh
, and
J. F.
Scott
, “
Domain wall nanoelectronics
,”
Rev. Mod. Phys.
84
,
119
156
(
2012
).
2.
M.
Haider
et al, “
Electron microscopy image enhanced
,”
Nature
392
,
768
769
(
1998
).
3.
In recognition of aberration-corrected TEM
,”
Nat. Nanotechnol.
15
,
417
(
2020
).
4.
P. E.
Batson
,
N.
Dellby
, and
O. L.
Krivanek
, “
Sub-ångstrom resolution using aberration corrected electron optics
,”
Nature
418
,
617
620
(
2002
).
5.
C.-L.
Jia
et al, “
Atomic-scale study of electric dipoles near charged and uncharged domain walls in ferroelectric films
,”
Nat. Mater.
7
,
57
61
(
2008
).
6.
R. C.
Miller
and
G.
Weinreich
, “
Mechanism for the sidewise motion of 180° domain walls in barium titanate
,”
Phys. Rev.
117
,
1460
1466
(
1960
).
7.
D.
Berlincourt
and
H. H. A.
Krueger
, “
Domain processes in lead titanate zirconate and barium titanate ceramics
,”
J. Appl. Phys.
30
,
1804
1810
(
1959
).
8.
T.
Malis
and
H.
Gleiter
, “
Investigation of the structure of ferroelectric domain boundaries by transmission electron microscopy
,”
J. Appl. Phys.
47
,
5195
5200
(
1976
).
9.
M. A.
Pereira Gonçalves
,
C.
Escorihuela-Sayalero
,
P.
Garca-Fernández
,
J.
Junquera
, and
J.
Íñiguez
, “
Theoretical guidelines to create and tune electric skyrmion bubbles
,”
Sci. Adv.
5
,
eaau7023
(
2019
).
10.
S.
Das
et al, “
Observation of room-temperature polar skyrmions
,”
Nature
568
,
368
372
(
2019
).
11.
L.
Keeney
et al,
Nanoscale Ferroelectrics and Multiferroics: Key Processing and Characterization Issues, and Nanoscale Effects
(
Wiley
,
2016
), pp.
789
829
.
12.
E.
Salje
and
H.
Zhang
, “
Domain boundary engineering
,”
Phase Transitions
82
,
452
469
(
2009
).
13.
W.
Gao
et al, “
Real-space charge-density imaging with sub-Ångström resolution by four-dimensional electron microscopy
,”
Nature
575
,
480
484
(
2019
).
14.
Y.
Zhang
et al, “
Intrinsic conductance of domain walls in BiFeO3
,”
Adv. Mater.
31
,
1902099
(
2019
).
15.
D.
Grüner
and
Z.
Shen
, “
Direct scanning electron microscopy imaging of ferroelectric domains after ion milling
,”
J. Am. Ceram. Soc.
93
,
48
50
(
2010
).
16.
G.
Rosenman
et al, “
Observation of ferroelectric domain structures by secondary-electron microscopy in as-grown KTiOPO4 crystals
,”
Phys. Rev. B
54
,
6222
6226
(
1996
).
17.
S.
Zhu
and
W.
Cao
, “
Direct observation of ferroelectric domains in LiTaO3 using environmental scanning electron microscopy
,”
Phys. Rev. Lett.
79
,
2558
2561
(
1997
).
18.
P.
Potnis
,
N.-T.
Tsou
, and
J.
Huber
, “
A review of domain modelling and domain imaging techniques in ferroelectric crystals
,”
Materials
4
,
417
447
(
2011
).
19.
A.
Sogr
,
A.
Maslovskaya
, and
I.
Kopylova
, “
Advanced modes of imaging of ferroelectric domains in SEM
,”
Ferroelectrics
341
,
29
37
(
2006
).
20.
V. V.
Aristov
,
L. S.
Kokhanchik
, and
Y. I.
Voronovskii
, “
Voltage contrast of ferroelectric domains of lithium niobate in SEM
,”
Phys. Status Solidi A
86
,
133
141
(
1984
).
21.
S.
Zhu
and
W.
Cao
, “
Imaging of 180° ferroelectric domains in LiTaO3 by means of scanning electron microscopy
,”
Phys. Status Solidi A
173
,
495
502
(
1999
).
22.
K.
Tanaka
and
T.
Suhara
, “
Non-destructive observation of ferroelectric domain inverted structures in MgO:LiNbO3 by scanning electron microscope
,”
Electron. Lett.
51
,
923
925
(
2015
).
23.
L. S.
Kokhanchik
, “
The use of surface charging in the SEM for lithium niobate domain structure investigation
,”
Micron
40
,
41
45
(
2009
).
24.
T. L.
Burnett
et al, “
Electron backscatter diffraction as a domain analysis technique in BiFeO3-PbTiO3 single crystals
,”
IEEE Trans. Ultrason. Ferroelectrics, Freq. Control
55
,
957
962
(
2008
).
25.
M. U.
Farooq
et al, “
Electron backscatter diffraction mapping of herringbone domain structures in tetragonal piezoelectrics
,”
J. Appl. Phys.
104
,
024111
(
2008
).
26.
R.
Le Bihan
, “
Study of ferroelectric and ferroelastic domain structures by scanning electron microscopy
,”
Ferroelectrics
97
,
19
46
(
1989
).
27.
A. B.
Mosberg
et al, “
FIB lift-out of conducting ferroelectric domain walls in hexagonal manganites
,”
Appl. Phys. Lett.
115
,
122901
(
2019
).
28.
Z.
Chen
et al, “
Giant tuning of ferroelectricity in single crystals by thickness engineering
,”
Sci. Adv.
6
,
eabc7156
(
2020
).
29.
A.
Schilling
,
R. M.
Bowman
,
G.
Catalan
,
J. F.
Scott
, and
J. M.
Gregg
, “
Morphological control of polar orientation in single-crystal ferroelectric nanowires
,”
Nano Lett.
7
,
3787
3791
(
2007
).
30.
L. J.
McGilly
and
J. M.
Gregg
, “
Polarization closure in PbZr(0.42)Ti(0.58)O3 nanodots
,”
Nano Lett.
11
,
4490
4495
(
2011
).
31.
J. R.
Whyte
et al, “
Ferroelectric domain wall injection
,”
Adv. Mater.
26
,
293
298
(
2014
).
32.
S. R.
Burns
,
J. M.
Gregg
, and
V.
Nagarajan
, “
Nanostructuring ferroelectrics via focused ion beam methodologies
,”
Adv. Funct. Mater.
26
,
8367
8381
(
2016
).
33.
E.
Vlasov
,
D.
Chezganov
,
M.
Chuvakova
, and
V. Y.
Shur
, “
The ferroelectric domain structures induced by electron beam scanning in lithium niobate
,”
Scanning
2018
,
7809826
.
34.
X.
Li
,
K.
Terabe
,
H.
Hatano
, and
K.
Kitamura
, “
Electron-beam domain writing in stoichiometric LiTaO3 single crystal by utilizing resist layer
,”
Jpn. J. Appl. Phys., Part 2
45
,
L399
L402
(
2006
).
35.
X.
Li
,
K.
Terabe
,
H.
Hatano
, and
K.
Kitamura
, “
Domain patterning in LiNbO3 and LiTaO3 by focused electron beam
,”
J. Cryst. Growth
292
,
324
327
(
2006
).
36.
V. Y.
Shur
,
D. S.
Chezganov
,
A. R.
Akhmatkhanov
, and
D. K.
Kuznetsov
, “
Domain patterning by electron beam of MgO doped lithium niobate covered by resist
,”
Appl. Phys. Lett.
106
,
232902
(
2015
).
37.
H.
Kianirad
,
F.
Laurell
, and
C.
Canalias
, “
Domain wall motion in stoichiometric LiTaO3 induced by low-energy electron beam
,”
Appl. Phys. Lett.
115
,
052901
(
2019
).
38.
T.
Asada
and
Y.
Koyama
, “
Ferroelectric domain structures around the morphotropic phase boundary of the piezoelectric material Pb Zr1−xTixO3
,”
Phys. Rev. B
75
,
214111
(
2007
).
39.
K.
Aoyagi
et al, “
Diffraction contrast analysis of 90° and 180° ferroelectric domain structures of PbTiO3 thin films
,”
Sci. Technol. Adv. Mater.
12
,
034403
(
2011
).
40.
M.
Tanaka
and
G.
Honjo
, “
Electron optical studies of barium titanate single crystal films
,”
J. Phys. Soc. Jpn.
19
,
954
970
(
1964
).
41.
M.-G.
Han
et al, “
Interface-induced nonswitchable domains in ferroelectric thin films
,”
Nat. Commun.
5
,
4693
(
2014
).
42.
H.
Lu
et al, “
Electrical tunability of domain wall conductivity in LiNbO3 thin films
,”
Adv. Mater.
31
,
1902890
(
2019
).
43.
T.
Matsumoto
, in
Nanoscale Ferroelectric-Multiferroic Materials for Energy Harvesting Applications
, edited by
H.
Kimura
,
Z.
Cheng
, and
T.
Jia
(
Elsevier
,
2019
),
109
135
.
44.
M. J.
Hÿtch
,
E.
Snoeck
, and
R.
Kilaas
, “
Quantitative measurement of displacement and strain fields from HREM micrographs
,”
Ultramicroscopy
74
,
131
146
(
1998
).
45.
V.
Grillo
and
F.
Rossi
, “
STEM_CELL: A software tool for electron microscopy. Part 2 analysis of crystalline materials
,”
Ultramicroscopy
125
,
112
129
(
2013
).
46.
D. B.
Williams
and
C. B.
Carter
,
Transmission Electron Microscopy
(
Springer
,
2009
), pp.
483
509
.
47.
Y. L.
Tang
,
Y. L.
Zhu
, and
X. L.
Ma
, “
On the benefit of aberration-corrected HAADF-STEM for strain determination and its application to tailoring ferroelectric domain patterns
,”
Ultramicroscopy
160
,
57
63
(
2016
).
48.
Y. L.
Tang
et al, “
Observation of a periodic array of flux-closure quadrants in strained ferroelectric PbTiO3 films
,”
Science
348
,
547
(
2015
).
49.
Y.
Liu
,
Y.-L.
Tang
,
Y.-L.
Zhu
,
W.-Y.
Wang
, and
X.-L.
Ma
, “
Spatial coupling of ferroelectric domain walls and crystallographic defects in the PbTiO3 films
,”
Adv. Mater. Interfaces
3
,
1600342
(
2016
).
50.
Y.
Sun
et al, “
Subunit cell–level measurement of polarization in an individual polar vortex
,”
Sci. Adv.
5
,
eaav4355
(
2019
).
51.
M.
Hadjimichael
,
Y.
Li
,
E.
Zatterin
,
G. A.
Chahine
,
M.
Conroy
,
K.
Moore
,
E. N. O.
Connell
,
P.
Ondrejkovic
,
P.
Marton
,
J.
Hlinka
,
U.
Bangert
,
S.
Leake
, and
P.
Zubko
, “
Metal ferroelectric supercrystals with periodically curved metallic layers
,”
Nat. Mater.
(published online
2021
).
52.
J.
Gonnissen
et al, “
Direct observation of ferroelectric domain walls in LiNbO3: Wall-meanders, kinks, and local electric charges
,”
Adv. Funct. Mater.
26
,
7599
7604
(
2016
).
53.
K.
Moore
,
M.
Conroy
, and
U.
Bangert
, “
Rapid polarization mapping in ferroelectrics using fourier masking
,”
J. Microsc.
279
,
222
228
(
2020
).
54.
R. G. P.
McQuaid
,
M. P.
Campbell
,
R. W.
Whatmore
,
A.
Kumar
, and
J. M.
Gregg
, “
Injection and controlled motion of conducting domain walls in improper ferroelectric Cu–Cl boracite
,”
Nat. Commun.
8
,
15105
(
2017
).
55.
M.
Conroy
et al, “
Probing the dynamics of topologically protected charged ferroelectric domain walls with the electron beam at the atomic scale
,”
Microsc. Microanal.
26
,
3030
3032
(
2020
).
56.
I.
El Baggari
et al, “
Nature and evolution of incommensurate charge order in manganites visualized with cryogenic scanning transmission electron microscopy
,”
Proc. Natl. Acad. Sci. U. S. A.
115
,
1445
(
2018
).
57.
W. T.
Lee
,
E. K. H.
Salje
, and
U.
Bismayer
,
J. Phys.: Condens. Matter
15
,
1353
(
2003
).
58.
C.
Ophus
, “
Four-dimensional scanning transmission electron microscopy (4D-STEM): From scanning nanodiffraction to ptychography and beyond
,”
Microsc. Microanal.
25
,
563
582
(
2019
).
59.
S. E.
Zeltmann
et al, “
Patterned probes for high precision 4D-STEM Bragg measurements
,”
Ultramicroscopy
209
,
112890
(
2020
).
60.
Y.
Han
et al, “
Strain mapping of two-dimensional heterostructures with subpicometer precision
,”
Nano Lett.
18
,
3746
3751
(
2018
).
61.
V. B.
Ozdol
et al, “
Strain mapping at nanometer resolution using advanced nano-beam electron diffraction
,”
Appl. Phys. Lett.
106
,
253107
(
2015
).
62.
K.
Müller
et al, “
Scanning transmission electron microscopy strain measurement from millisecond frames of a direct electron charge coupled device
,”
Appl. Phys. Lett.
101
,
212110
(
2012
).
63.
T. C.
Pekin
,
C.
Gammer
,
J.
Ciston
,
A. M.
Minor
, and
C.
Ophus
, “
Optimizing disk registration algorithms for nanobeam electron diffraction strain mapping
,”
Ultramicroscopy
176
,
170
(
2017
).
64.
G.
Guzzinati
et al, “
Electron Bessel beam diffraction for precise and accurate nanoscale strain mapping
,”
Appl. Phys. Lett.
114
,
243501
(
2019
).
65.
K.
Müller-Caspary
,
A.
Oelsner
, and
P.
Potapov
, “
Two-dimensional strain mapping in semiconductors by nano-beam electron diffraction employing a delay-line detector
,”
Appl. Phys. Lett.
107
,
072110
(
2015
).
66.
J.
Zweck
, “
Imaging of magnetic and electric fields by electron microscopy
,”
J. Phys.: Condens. Matter
28
,
403001
(
2016
).
67.
J. N.
Chapman
,
I. R.
McFadyen
, and
S.
McVitie
, “
Modified differential phase contrast Lorentz microscopy for improved imaging of magnetic structures
,”
IEEE Trans. Magn.
26
,
1506
(
1990
).
68.
T.
Matsumoto
,
M.
Koguchi
, and
Y.
Takahashi
, in
EMC 2008 14th European Microscopy Congress 1–5 September 2008, Aachen, Germany
, edited by
S.
Richter
and
A.
Schwedt
(
Springer Berlin Heidelberg
,
2008
), pp.
549
550
.
69.
I.
MacLaren
et al, “
On the origin of differential phase contrast at a locally charged and globally charge-compensated domain boundary in a polar-ordered material
,”
Ultramicroscopy
154
,
57
(
2015
).
70.
S. R.
Spurgeon
et al, “
Towards data-driven next-generation transmission electron microscopy
,”
Nat. Mater.
(published online
2020
).
71.
Y. L.
Li
,
S. Y.
Hu
,
Z. K.
Liu
, and
L. Q.
Chen
, “
Effect of electrical boundary conditions on ferroelectric domain structures in thin films
,”
Appl. Phys. Lett.
81
,
427
429
(
2002
).
72.
F.
Xue
,
Y.
Gu
,
L.
Liang
,
Y.
Wang
, and
L.-Q.
Chen
, “
Orientations of low-energy domain walls in perovskites with oxygen octahedral tilts
,”
Phys. Rev. B
90
,
220101
(
2014
).
73.
K.
Müller-Caspary
et al, “
Measurement of atomic electric fields and charge densities from average momentum transfers using scanning transmission electron microscopy
,”
Ultramicroscopy
178
,
62
(
2017
).
74.
M. C.
Cao
et al, “
Theory and practice of electron diffraction from single atoms and extended objects using an EMPAD
,”
Microscopy
67
,
i150
i161
(
2018
).
75.
M.
Wu
and
E.
Spiecker
, “
Correlative micro-diffraction and differential phase contrast study of mean inner potential and subtle beam-specimen interaction
,”
Ultramicroscopy
176
,
233
(
2017
).
76.
M. R.
McCartney
and
D. J.
Smith
, “
Electron holography: Phase imaging with nanometer resolution
,”
Annu. Rev. Mater. Res.
37
,
729
767
(
2007
).
77.
M.
O’Keeffe
and
J. C. H.
Spence
, “
On the average Coulomb potential (Σ0) and constraints on the electron density in crystals
,”
Acta Crystallogr., Sect. A
50
,
33
45
(
1994
).
78.
N.
Shibata
et al, “
Direct visualization of local electromagnetic field structures by scanning transmission electron microscopy
,”
Acc. Chem. Res.
50
,
1502
1512
(
2017
).
79.
Y. P.
Ivanov
,
A.
Chuvilin
,
S.
Lopatin
, and
J.
Kosel
, “
Modulated magnetic nanowires for controlling domain wall motion: Toward 3D magnetic memories
,”
ACS Nano
10
,
5326
5332
(
2016
).
80.
S.
Lopatin
,
Y. P.
Ivanov
,
J.
Kosel
, and
A.
Chuvilin
, “
Multiscale differential phase contrast analysis with a unitary detector
,”
Ultramicroscopy
162
,
74
(
2016
).
81.
C. L.
Jia
,
M.
Lentzen
, and
K.
Urban
, “
Atomic-resolution imaging of oxygen in perovskite ceramics
,”
Science
299
,
870
873
(
2003
).
82.
C. L.
Jia
and
K.
Urban
, “
Atomic-resolution measurement of oxygen concentration in oxide materials
,”
Science
303
,
2001
2004
(
2004
).
83.
C.-L.
Jia
,
K. W.
Urban
,
M.
Alexe
,
D.
Hesse
, and
I.
Vrejoiu
, “
Direct observation of continuous electric dipole rotation in flux-closure domains in ferroelectric Pb(Zr,Ti)O3
,”
Science
331
,
1420
1423
(
2011
).
84.
X. K.
Wei
et al, “
An unconventional transient phase with cycloidal order of polarization in energy-storage antiferroelectric PbZrO3
,”
Adv. Mater.
32
,
1907208
(
2020
).
85.
X.-K.
Wei
et al, “
Néel-like domain walls in ferroelectric Pb(Zr,Ti)O3 single crystals
,”
Nat. Commun.
7
,
12385
(
2016
).
86.
Y.
Wang
,
U.
Salzberger
,
W.
Sigle
,
Y.
Eren Suyolcu
, and
P. A.
van Aken
, “
Oxygen octahedra picker: A software tool to extract quantitative information from STEM images
,”
Ultramicroscopy
168
,
46
52
(
2016
).
87.
Y.-M.
Kim
,
S. J.
Pennycook
, and
A. Y.
Borisevich
, “
Quantitative comparison of bright field and annular bright field imaging modes for characterization of oxygen octahedral tilts
,”
Ultramicroscopy
181
,
1
7
(
2017
).
88.
P.
Gao
et al, “
Picometer-scale atom position analysis in annular bright-field STEM imaging
,”
Ultramicroscopy
184
,
177
187
(
2018
).
89.
A.
Kumar
et al, “
Atomic-resolution electron microscopy of nanoscale local structure in lead-based relaxor ferroelectrics
,”
Nat. Mater.
20
,
62
67
(
2020
).
90.
J. R.
Kim
et al, “
Stabilizing hidden room-temperature ferroelectricity via a metastable atomic distortion pattern
,”
Nat. Commun.
11
,
4944
(
2020
).
91.
Q.
He
et al, “
Towards 3D mapping of BO6 octahedron rotations at perovskite heterointerfaces, unit cell by unit cell
,”
ACS Nano
9
,
8412
8419
(
2015
).
92.
G.
Sanchez-Santolino
et al, “
Oxygen octahedral distortions in LaMO3/SrTiO3 superlattices
,”
Microsc. Microanal.
20
,
825
(
2014
).
93.
R.
Aso
,
D.
Kan
,
Y.
Shimakawa
, and
H.
Kurata
, “
Atomic level observation of octahedral distortions at the perovskite oxide heterointerface
,”
Sci. Rep.
3
,
2214
(
2013
).
94.
C.
Chen
et al, “
Atomic and electronic structure of the SrNbO3/SrNbO3.4 interface
,”
Appl. Phys. Lett.
105
,
221602
(
2014
).
95.
L.
Wang
et al, “
Hole-trapping-induced stabilization of Ni4+ in SrNiO3/LaFeO3 superlattices
,”
Adv. Mater.
32
,
2005003
(
2020
).
96.
Y.
Liu
et al, “
Controlled growth and atomic-scale mapping of charged heterointerfaces in PbTiO3/BiFeO3 bilayers
,”
ACS Appl. Mater. Interfaces
9
,
25578
25586
(
2017
).
97.
E.
Yücelen
,
I.
Lazić
, and
E. G. T.
Bosch
, “
Phase contrast scanning transmission electron microscopy imaging of light and heavy atoms at the limit of contrast and resolution
,”
Sci. Rep.
8
,
2676
(
2018
).
98.
S.
de Graaf
,
J.
Momand
,
C.
Mitterbauer
,
S.
Lazar
, and
B. J.
Kooi
, “
Resolving hydrogen atoms at metal-metal hydride interfaces
,”
Sci. Adv.
6
,
eaay4312
(
2020
).
99.
G.
Shirane
,
J. D.
Axe
,
J.
Harada
, and
J. P.
Remeika
, “
Soft ferroelectric modes in lead titanate
,”
Phys. Rev. B
2
,
155
159
(
1970
).
100.
A. K.
Yadav
et al, “
Observation of polar vortices in oxide superlattices
,”
Nature
530
,
198
201
(
2016
).
101.
C. T.
Nelson
et al, “
Spontaneous vortex nanodomain arrays at ferroelectric heterointerfaces
,”
Nano Lett.
11
,
828
(
2011
).
102.
Y. L.
Tang
et al, “
Atomic-scale mapping of dipole frustration at 90° charged domain walls in ferroelectric PbTiO3 films
,”
Sci. Rep.
4
,
4115
(
2014
).
103.
J.
Ma
et al, “
Modulation of charged a1/a2 domains and piezoresponses of tensile strained PbTiO3 films by the cooling rate
,”
RSC Adv.
9
,
13981
13990
(
2019
).
104.
W. Y.
Wang
et al, “
Large scale arrays of four-state vortex domains in BiFeO3 thin film
,”
Appl. Phys. Lett.
109
,
202904
(
2016
).
105.
H.
Rose
, “
Electron microscopy
,”
Optik
39
,
416
436
(
1974
).
106.
I.
Lazić
,
E. G. T.
Bosch
, and
S.
Lazar
, “
Phase contrast STEM for thin samples: Integrated differential phase contrast
,”
Ultramicroscopy
160
,
265
280
(
2016
).
107.
N.
Shibata
et al, “
Differential phase-contrast microscopy at atomic resolution
,”
Nat. Phys.
8
,
611
(
2012
).
108.
M.
Campanini
et al, “
Imaging and quantification of charged domain walls in BiFeO3
,”
Nanoscale
12
,
9186
9193
(
2020
).
109.
K.
Müller
et al, “
Atomic electric fields revealed by a quantum mechanical approach to electron picodiffraction
,”
Nat. Commun.
5
,
5653
(
2014
).
110.
A.
Lubk
and
J.
Zweck
, “
Differential phase contrast: An integral perspective
,”
Phys. Rev. A
91
,
023805
(
2015
).
111.
K.
Du
et al, “
Manipulating topological transformations of polar structures through real-time observation of the dynamic polarization evolution
,”
Nat. Commun.
10
,
4864
(
2019
).
112.
A. M.
Maiden
,
J. M.
Rodenburg
, and
M. J.
Humphry
, “
Optical ptychography: A practical implementation with useful resolution
,”
Opt. Lett.
35
,
2585
2587
(
2010
).
113.
P.
Thibault
et al, “
High-resolution scanning x-ray diffraction microscopy
,”
Science
321
,
379
382
(
2008
).
114.
A.
Maiden
,
G.
Morrison
,
B.
Kaulich
,
A.
Gianoncelli
, and
J.
Rodenburg
, “
Soft X-ray spectromicroscopy using ptychography with randomly phased illumination
,”
Nat. Commun.
4
,
1669
(
2013
).
115.
M.
Dierolf
et al, “
Ptychographic X-ray computed tomography at the nanoscale
,”
Nature
467
,
436
439
(
2010
).
116.
F.
Pfeiffer
, “
X-ray ptychography
,”
Nat. Photonics
12
,
9
17
(
2018
).
117.
Y.
Jiang
et al, “
Electron ptychography of 2D materials to deep sub-ångström resolution
,”
Nature
559
,
343
349
(
2018
).
118.
H.
Yang
et al, “
Simultaneous atomic-resolution electron ptychography and Z-contrast imaging of light and heavy elements in complex nanostructures
,”
Nat. Commun.
7
,
12532
(
2016
).
119.
W.
Gao
,
Z. D.
Hood
, and
M.
Chi
, “
Interfaces in heterogeneous catalysts: Advancing mechanistic understanding through atomic-scale measurements
,”
Acc. Chem. Res.
50
,
787
795
(
2017
).
120.
H.
Yang
et al, “
Electron ptychographic phase imaging of light elements in crystalline materials using Wigner distribution deconvolution
,”
Ultramicroscopy
180
,
173
179
(
2017
).
122.
C.
Ophus
,
J.
Ciston
, and
C. T.
Nelson
, “
Correcting nonlinear drift distortion of scanning probe and scanning transmission electron microscopies from image pairs with orthogonal scan directions
,”
Ultramicroscopy
162
,
1
9
(
2016
).
123.
B.
Berkels
and
C. H.
Liebscher
, “
Joint non-rigid image registration and reconstruction for quantitative atomic resolution scanning transmission electron microscopy
,”
Ultramicroscopy
198
,
49
57
(
2019
).
124.
L.
Jones
et al, “
Optimising multi-frame ADF-STEM for high-precision atomic-resolution strain mapping
,”
Ultramicroscopy
179
,
57
62
(
2017
).
125.
L.
Jones
et al, “
Smart align—a new tool for robust non-rigid registration of scanning microscope data
,”
Adv. Struct. Chem. Imaging
1
,
8
(
2015
).
126.
O. L.
Krivanek
et al, “
Atom-by-atom structural and chemical analysis by annular dark-field electron microscopy
,”
Nature
464
,
571
574
(
2010
).
127.
A.
Borisevich
et al, “
Mapping octahedral tilts and polarization across a domain wall in BiFeO3 from Z-contrast scanning transmission electron microscopy image atomic column shape analysis
,”
ACS Nano
4
,
6071
6079
(
2010
).
128.
M. C.
Sarahan
,
M.
Chi
,
D. J.
Masiel
, and
N. D.
Browning
, “
Point defect characterization in HAADF-STEM images using multivariate statistical analysis
,”
Ultramicroscopy
111
,
251
257
(
2011
).
129.
L.
Houben
,
A.
Thust
, and
K.
Urban
, “
Atomic-precision determination of the reconstruction of a 90° tilt boundary in YBa2Cu3O7-δ by aberration corrected HRTEM
,”
Ultramicroscopy
106
,
200
214
(
2006
).
130.
I.
MacLaren
,
R.
Villaurrutia
,
B.
Schaffer
,
L.
Houben
, and
A.
Peláiz-Barranco
, “
Atomic-scale imaging and quantification of electrical polarisation in incommensurate antiferroelectric lanthanum-doped lead zirconate titanate
,”
Adv. Funct. Mater.
22
,
261
266
(
2012
).
131.
B. D. A.
Levin
,
E. L.
Lawrence
, and
P. A.
Crozier
, “
Tracking the picoscale spatial motion of atomic columns during dynamic structural change
,”
Ultramicroscopy
213
,
112978
(
2020
).
132.
K.
Moore
et al, “
Highly charged 180 degree head-to-head domain walls in lead titanate
,”
Commun. Phys.
3
,
231
(
2020
).
133.
Q. H.
Zhang
et al, “
Direct observation of interlocked domain walls in hexagonal RMnO3 (R = Tm, Lu)
,”
Phys. Rev. B
85
,
020102
(
2012
).
134.
T.
Matsumoto
et al, “
Multivariate statistical characterization of charged and uncharged domain walls in multiferroic hexagonal YMnO3 single crystal visualized by a spherical aberration-corrected STEM
,”
Nano Lett.
13
,
4594
4601
(
2013
).
135.
M.
Nord
,
P. E.
Vullum
,
I.
MacLaren
,
T.
Tybell
, and
R.
Holmestad
, “
Atomap: A new software tool for the automated analysis of atomic resolution images using two-dimensional Gaussian fitting
,”
Adv. Struct. Chem. Imaging
3
,
9
(
2017
).
136.
M. H.
Eoghan O’Connell
, TEMUL Toolkit, https://temul-toolkit.readthedocs.io/en/latest/,
2020
.
137.
S.
Somnath
,
C. R.
Smith
,
N.
Laanait
,
R. K.
Vasudevan
, and
S.
Jesse
, “
USID and pycroscopyopen source frameworks for storing and analyzing imaging and spectroscopy data
,”
Microsc. Microanal.
25
,
220
221
(
2019
).
138.
S. M.
Anthony
and
S.
Granick
, “
Image analysis with rapid and accurate two-dimensional Gaussian fitting
,”
Langmuir
25
,
8152
(
2009
).
139.
G. F.
Nataf
et al, “
Domain-wall engineering and topological defects in ferroelectric and ferroelastic materials
,”
Nat. Rev. Phys.
2
,
634
648
(
2020
).
140.
J. A.
Hachtel
,
J. C.
Idrobo
, and
M.
Chi
, “
Sub-Ångstrom electric field measurements on a universal detector in a scanning transmission electron microscope
,”
Adv. Struct. Chem. Imaging
4
,
10
(
2018
).
141.
A. K.
Yadav
et al, “
Spatially resolved steady-state negative capacitance
,”
Nature
565
,
468
471
(
2019
).
142.
M.
Campanini
et al, “
Atomic-resolution differential phase contrast STEM on ferroelectric materials: A mean-field approach
,”
Phys. Rev. B
101
,
184116
(
2020
).
143.
J. A.
Brehm
et al, “
Tunable quadruple-well ferroelectric van der Waals crystals
,”
Nat. Mater.
19
,
43
48
(
2020
).
144.
Z.
Guan
et al, “
Recent progress in two-dimensional ferroelectric materials
,”
Adv. Electron. Mater.
6
,
1900818
(
2020
).
145.
K.
Müller-Caspary
et al, “
Atomic-scale quantification of charge densities in two-dimensional materials
,”
Phys. Rev. B
98
,
121408
(
2018
).
146.
L.
Li
,
L.
Xie
, and
X.
Pan
, “
Real-time studies of ferroelectric domain switching: A review
,”
Rep. Prog. Phys.
82
,
126502
(
2019
).
147.
C. T.
Nelson
et al, “
Domain dynamics during ferroelectric switching
,”
Science
334
,
968
971
(
2011
).
148.
P.
Gao
et al, “
Atomic-scale mechanisms of ferroelastic domain-wall-mediated ferroelectric switching
,”
Nat. Commun.
4
,
2791
(
2013
).
149.
H.
Chang
et al, “
Watching domains grow: In situ studies of polarization switching by combined scanning probe and scanning transmission electron microscopy
,”
J. Appl. Phys.
110
,
052014
(
2011
).
150.
P.
Gao
et al, “
Ferroelastic domain switching dynamics under electrical and mechanical excitations
,”
Nat. Commun.
5
,
3801
(
2014
).
151.
J. K.
Lee
et al, “
Direct observation of asymmetric domain wall motion in a ferroelectric capacitor
,”
Acta Mater.
61
,
6765
6777
(
2013
).
152.
M.-G.
Han
et al, “
Ferroelectric switching dynamics of topological vortex domains in a hexagonal manganite
,”
Adv. Mater.
25
,
2415
2421
(
2013
).
153.
L.
Li
et al, “
Giant resistive switching via control of ferroelectric charged domain walls
,”
Adv. Mater.
28
,
6574
6580
(
2016
).
154.
P.
Gao
et al, “
Direct observations of retention failure in ferroelectric memories
,”
Adv. Mater.
24
,
1106
(
2012
).
155.
J. X.
Zhang
et al, “
Large field-induced strains in a lead-free piezoelectric material
,”
Nat. Nanotechnol.
6
,
98
102
(
2011
).
156.
J. L.
Hart
et al, “
Electron-beam-induced ferroelectric domain behavior in the transmission electron microscope: Toward deterministic domain patterning
,”
Phys. Rev. B
94
,
174104
(
2016
).
157.
M.
Barzilay
and
Y.
Ivry
, “
Formation and manipulation of domain walls with 2 nm domain periodicity in BaTiO3 without contact electrodes
,”
Nanoscale
12
,
11136
11142
(
2020
).
158.
F. S.
Hage
,
M. D.
Kapetanakis
,
J. C.
Idrobo
,
Q. M.
Ramasse
, and
D.
Kepaptsoglou
, “
Atomic-scale spectroscopic imaging of the extreme-UV optical response of B- and N-doped graphene
,”
Adv. Funct. Mater.
29
,
1901819
(
2019
).
159.
R.
Mishra
,
R.
Ishikawa
,
A. R.
Lupini
, and
S. J.
Pennycook
, “
Single-atom dynamics in scanning transmission electron microscopy
,”
MRS Bull.
42
,
644
652
(
2017
).
160.
O.
Dyck
et al, “
Building structures atom by atom via electron beam manipulation
,”
Small
14
,
1801771
(
2018
).
161.
A.
Maksov
et al, “
Deep learning analysis of defect and phase evolution during electron beam-induced transformations in WS2
,”
npj Comput. Mater.
5
,
12
(
2019
).
162.
O.
Dyck
et al, “
Doping of Cr in graphene using electron beam manipulation for functional defect engineering
,”
ACS Appl. Nano Mater.
3
,
10855
10863
(
2020
).
163.
Y.-M.
Kim
et al, “
Direct observation of ferroelectric field effect and vacancy-controlled screening at the BiFeO3/LaxSr1−xMnO3 interface
,”
Nat. Mater.
13
,
1019
1025
(
2014
).
164.
Y.
Zhang
et al, “
Anisotropic polarization-induced conductance at a ferroelectric–insulator interface
,”
Nat. Nanotechnol.
13
,
1132
1136
(
2018
).
165.
J. A.
Mundy
et al, “
Functional electronic inversion layers at ferroelectric domain walls
,”
Nat. Mater.
16
,
622
627
(
2017
).
166.
T.
Rojac
et al, “
Domain-wall conduction in ferroelectric BiFeO3 controlled by accumulation of charged defects
,”
Nat. Mater.
16
,
322
327
(
2017
).
167.
M.
Conroy
et al, “
Atomic-scale characterization of ferro-electric domains in lithium niobate-revealing the electronic properties of domain walls
,”
Microsc. Microanal.
25
,
576
577
(
2019
).
168.
F. S.
Hage
,
G.
Radtke
,
D. M.
Kepaptsoglou
,
M.
Lazzeri
, and
Q. M.
Ramasse
, “
Single-atom vibrational spectroscopy in the scanning transmission electron microscope
,”
Science
367
,
1124
1127
(
2020
).
169.
S.
Das
et al, “
Local negative permittivity and topological phase transition in polar skyrmions
,”
Nat. Mater.
(published online
2020
).
170.
C.
Ophus
et al, “
Efficient linear phase contrast in scanning transmission electron microscopy with matched illumination and detector interferometry
,”
Nat. Commun.
7
,
10719
(
2016
).
171.
J.
Verbeeck
et al, “
Shaping electron beams for the generation of innovative measurements in the (S)TEM
,”
C. R. Phys.
15
,
190
199
(
2014
).
172.
K. Y.
Bliokh
et al, “
Theory and applications of free-electron vortex states
,”
Phys. Rep.
690
,
1
70
(
2017
).
173.
L.
Clark
et al, “
Exploiting lens aberrations to create electron-vortex beams
,”
Phys. Rev. Lett.
111
,
064801
(
2013
).
174.
R.
Shiloh
et al, “
Nanostructuring of electron beams
,”
Phys. Scr.
94
,
034004
(
2019
).
175.
X.
Hong
et al, “
Multibeam electron diffraction
,” arXiv:2009.09134 (
2020
).
176.
Y.
Morimoto
and
P.
Baum
, “
Diffraction and microscopy with attosecond electron pulse trains
,”
Nat. Phys.
14
,
252
256
(
2018
).
177.
J. C.
Idrobo
, “
A new resolution quest in electron microscopy
,”
Nat. Rev. Mater.
(published online
2020
).