ArticlePDF Available

Imaging the Spatial Distribution of Electronic States in Graphene Using Electron Energy-Loss Spectroscopy: Prospect of Orbital Mapping

Authors:

Abstract

The spatial distributions of antibonding π Ã and σ Ã states in epitaxial graphene multilayers are mapped using electron energy-loss spectroscopy in a scanning transmission electron microscope. Inelastic channeling simulations validate the interpretation of the spatially resolved signals in terms of electronic orbitals, and demonstrate the crucial effect of the material thickness on the experimental capability to resolve the distribution of unoccupied states. This work illustrates the current potential of core-level electron energy-loss spectroscopy towards the direct visualization of electronic orbitals in a wide range of materials, of huge interest to better understand chemical bonding among many other properties at interfaces and defects in solids.
Imaging the Spatial Distribution of Electronic States in Graphene Using
Electron Energy-Loss Spectroscopy: Prospect of Orbital Mapping
M. Bugnet ,1,2,3,* M. Ederer ,4V. K. Lazarov ,5L. Li ,6Q. M. Ramasse ,1,2,7 S. Löffler ,4,and D. M. Kepaptsoglou 1,5,
1SuperSTEM Laboratory, SciTech Daresbury Campus, Daresbury WA4 4AD, United Kingdom
2School of Chemical and Process Engineering, University of Leeds, Leeds LS2 9JT, United Kingdom
3Univ Lyon, CNRS, INSA Lyon, UCBL, MATEIS, UMR 5510, 69621 Villeurbanne, France
4University Service Centre for Transmission Electron Microscopy, TU Wien, Wiedner Hauptstraße 8-10/E057-02, 1040 Wien, Austria
5Department of Physics, University of York, York YO10 5DD, United Kingdom
6Department of Physics and Astronomy, University of West Virginia, Morgantown, West Virginia 26506, USA
7School of Physics and Astronomy, University of Leeds, Leeds LS2 9JT, United Kingdom
(Received 4 June 2021; revised 23 December 2021; accepted 25 January 2022; published 14 March 2022)
The spatial distributions of antibonding πand σstates in epitaxial graphene multilayers are mapped
using electron energy-loss spectroscopy in a scanning transmission electron microscope. Inelastic
channeling simulations validate the interpretation of the spatially resolved signals in terms of electronic
orbitals, and demonstrate the crucial effect of the material thickness on the experimental capability to
resolve the distribution of unoccupied states. This work illustrates the current potential of core-level
electron energy-loss spectroscopy towards the direct visualization of electronic orbitals in a wide range of
materials, of huge interest to better understand chemical bonding among many other properties at interfaces
and defects in solids.
DOI: 10.1103/PhysRevLett.128.116401
The vast majority of physical and chemical properties of
crystalline materials originates from electronic states gov-
erning chemical bonding. In addition, defects, interfaces,
and surfaces have a direct influence on macroscopic
material properties. Imaging electronic states, such as
chemical bonds at crystal imperfections and discontinuities
in real space, is thus of fundamental and technological
interest to enable the development of new materials with
novel functionalities. While total electronic charge den-
sities have been reconstructed using either electron dif-
fraction [1,2] or high-resolution imaging [3] in the
transmission electron microscope, and more recently
imaged with atomic-scale resolution using four-dimensional
scanning transmission electron microscopy (STEM) [46],
the direct observation ofindividual electronic states has been
achieved primarily using scanning tunneling microscopy
[710], albeit with surface sensitivity only. Electron energy-
loss spectroscopy (EELS) in an electron microscope is a
spectroscopy technique probing site- and momentum-
projected empty states in the conduction band [11].
Following the development of aberration correctors and
high stability electron optics, atomic resolution EELS in the
scanning transmission electron microscope has become
routinely available, leading to elemental (chemical) mapping
[1214], and providing real-space atomic scale localization
of electronic states [1521] using the energy-loss near-edge
structure (ELNES) of the spectroscopic signal.
The ELNES, or spectrum fine structure, arising from
core-level excitation provides a wealth of information on
chemical bonding between atoms, and can be interpreted
by first-principles calculations in favorable cases.
However, a quantitative interpretation of ELNES maps
at atomic resolution requires to also take into account the
channeling characteristics of the swift electron beam
before and after the inelastic event [15,2225],and
resulting EELS signal mixing. The appropriate description
and/or deconvolution of the electron beam propagation
allows for the precise determination of the origin of
spatially resolved variations in fine structures arising from
orbital orientation [26] and localization [27]. It has been
theoretically predicted that aberration-corrected STEM-
EELS should allow for the mapping of electronic orbitals
[24]. A first experimental proof of principle was reported
through real-space mapping of electronic transitions to Ti
dorbitals in bulk rutile TiO2[27], but thus far, mapping
electronic orbitals in real space remains extremely chal-
lenging and elusive, be it in bulk crystals or at crystal
imperfections and discontinuities.
Graphene, a flagship two-dimensional material with
exceptional physical and mechanical properties, has
received tremendous scientific interest for potential
Published by the American Physical Society under the terms of
the Creative Commons Attribution 4.0 International license.
Further distribution of this work must maintain attribution to
the author(s) and the published articles title, journal citation,
and DOI.
PHYSICAL REVIEW LETTERS 128, 116401 (2022)
0031-9007=22=128(11)=116401(6) 116401-1 Published by the American Physical Society
electronic applications [28,29]. The atomic scale analysis
of individual graphene flakes is almost exclusively
achieved in the top surface view, thus enabling a path to
probe single atom chemical bonding [3032] and phononic
response [33]. The chemical bonding in graphene can be
described as in-C-plane (σ) and orthogonal out-of-C-plane
(π) covalent bonds. The ELNES of the C-K edge therefore
represents the excitation of core states probing in-C-plane
σ(1s2px;y) orbitals and out-of-C-plane π(1s2pz)
orbitals, as illustrated schematically in Fig. S1 [49]. While
πstate distributions around nitrogen and boron dopants
in monolayer graphene have been evidenced from the
ELNES [34], the prospect of mapping orbitals at vacancies
and nitrogen dopants in a single graphene sheet has
been explored theoretically only a few years ago [35].
Nevertheless, even if this is intuitively the appropriate
direction to observe individual in-C-plane σbonds, inelas-
tic channeling computations show that the STEM-EELS
mapping of σorbitals in top surface view in pristine
graphene is not possible due to symmetry constraints
[35,36]. The observation of graphene layers in side view,
however, provides a pathway to directly visualizing the
distribution of πstates at the atomic scale using STEM-
EELS. While the description of the atomic scale distribu-
tion of out-of-C-plane πand in-C-plane σstates may
appear simple enough from a chemical bonding perspec-
tive, experimental evidence using STEM-EELS is lacking.
Moreover, considering the aforementioned subtle effects
associated with the localization of the EELS signal, due for
instance to channeling of the incident electron beam, the
interpretation of energy-filtered real-space maps can be
very complex and must be validated through careful
numerical work.
In this Letter, real-space maps of πand σstates in
epitaxial graphene multilayers are recorded in side view,
combining state-of-the-art high spatial and energy resolu-
tion STEM-EELS with inelastic channeling calculations.
The interpretation of the spatial distribution of orbital
signals, based on the excellent agreement between com-
puted and experimental data, highlights the successful
direct mapping of the πstate distribution at atomic reso-
lution in the transmission electron microscope. The theo-
retical approach provides a powerful platform to determine
the origin of the energy-filtered signal.
The epitaxial graphene/SiC specimen was synthesized
by thermal decomposition of SiC at 1300 °C in ultrahigh
vacuum (UHV). For completeness, and as shown in
Fig. 1(a), we note that a thin capping film of Bi2Se3
was additionally deposited on top of the graphene layers by
molecular beam epitaxy at 275325 °C [37]. This specimen
was selected due to the convenient cross-section geometry
of the graphene layers; the properties and electronic
structure of interfaces with the 6HSiC substrate and
the Bi2Se3capping film are the subject of separate studies
and not discussed here. This results in a structure
comprising of a so-called graphene buffer layer(BL)
in contact with the underlying SiC substrate, capped with a
number of layers of epitaxialgraphene [here five such
layers are seen in Fig. 1(a)], whose macroscopic properties
are known to be nigh-on identical to those of free-standing
graphene [38,39].
The cross-section STEM lamellae were prepared by
focused ion beam milling. The thickness of the specimen
in the regions of investigation was evaluated to 25 nm by
Fourier-Log deconvolution of low-loss EELS spectra [11].
The STEM-EELS experiments were carried out using a
Nion HERMES microscope, equipped with a high-energy-
resolution monochromator, a Csaberration corrector up to
the fifth order, a Gatan Enfinium spectrometer, and oper-
ated at 60 kV. The convergence and collection semiangles
were 30 and 66 mrad, respectively. The specimen was
oriented in the ½10 ¯
10zone axis of SiC, corresponding to
the ½21 ¯
30zone axis of graphene (see Fig. S1 [49]). The
C-K edge was acquired with a 1.1 Å probe size and a step of
0.3Å, providing high spatial sampling while preserving
(a)
(b)
FIG. 1. (a) High resolution STEM-HAADF image of a six-layer
epitaxial graphene assembly, grown on 6HSiC and topped
with Bi2Se3, simultaneously acquired with core-loss EELS data.
(b) C-K edge spectra corresponding to the probe positionned in-
C-plane (solid orange line) and between layers (solid blue line),
as indicated in (a). Spectra are integrated over the width of the
whole image, presented after background subtraction, and shifted
vertically for visualization.
PHYSICAL REVIEW LETTERS 128, 116401 (2022)
116401-2
the specimen from electron beam damage. The monochro-
mator slit width was adjusted to provide an effective energy
resolution of 100 meV, as measured at the zero-loss
peak full width at half maximum. While not the highest
achievable resolution on the instrument, these conditions
provided a good compromise of beam current (given the
chosen probe size) while still being narrower than expected
spectral features. The presented STEM-EELS dataset was
acquired from a region of 4.8×3.9nm2, with a sampling
of 110 ×88 pixels2. Subpixel scanning (16 ×16)was
employed, hence leading to a 1760 ×1408 pixels2simul-
taneously acquired HAADF image in Fig. 1(a). The
experimental EELS maps and HAADF image in Fig. 2
are directly cropped from a 88 ×88 pixels2region in the
original dataset. The dwell time was 0.2 s, at a dispersion of
0.05 eV=pixel. The experimental maps were obtained after
background extraction (modeled with a power-law func-
tion), and energy filtering with a 2 eV window for πand σ
states.
The spatial variations of the C-K ELNES are highlighted
in Fig. 1(b), where spectra corresponding to in-C-plane
(solid orange line) and out-of-C-plane (solid blue line)
probe positions are displayed. It is noteworthy that the
instrumental broadening of the electron source is narrower
than the intrinsic linewidth of the fine structures. This is
expected to facilitate orbital mapping since the spectral
features are not limited by the energy resolution of the
electron source but by physical phenomena linked to, e.g.,
the excited state lifetime broadening, core-hole screening,
or other multielectronic interactions. The πand σfine
structures are in good agreement with existing work on
free-standing graphene layers [39,40], with a sharp exci-
tonic feature visible around 294.5 eV. Although the edges
overall look comparable for the in- and out-of-C plane
probe positions, the πintensity increases noticeably
between the epitaxial graphene layers at out-of-C-plane
positions. This behavior was systematically observed and is
characteristic of all C-K near-edge structures between the
epitaxial graphene layers Gr2Gr6 (see Fig. S2 [49]). The
ELNES of the BL and between the graphene BL and Gr2
are influenced by significant covalent bonding between the
graphene BL and SiC [39], and thus are not considered
here. In a first approximation, the spectral variations
observed for the Gr2Gr6 graphene layers can be related
to the simple picture of out-of-C-plane delocalization of π
states, in contrast to the in-C-plane nature of σstates.
Indeed, while the πbonding takes place between C
neighboring atoms of a single graphene layer, the lobes
of the antibonding πorbitals are delocalized around the C
planes, as shown schematically in Fig. S1 of the
Supplemental Material [49]. On the contrary, the σorbitals
are contained essentially within the graphene planes.
Nevertheless, the magnitude of this out-of-C-plane delo-
calization measured by fine structure mapping, and the
ability to spatially distinguish πfrom σstates using a
convergent electron-probe in STEM-EELS are nontrivial.
In order to rationalize experimental findings, we carried
out extensive numerical calculations of the fine structure
maps. The effect of the graphene-SiC interface, partially
influenced by covalent bonding, and of the graphene-
Bi2Se3interface on orbital mapping are beyond the scope
of this Letter, therefore a structure made exclusively of
graphene layers was considered for inelastic channeling
calculations. For simulating the elastic electron propagation
both before and after the inelastic scattering events, the
multislice algorithm [41,42] was used. For the inelastic
interaction between the probe beam and the sample
electrons, we calculate the mixed dynamic form factor
[24,43], based on density functional theory data obtained
with WIEN2k [44]. All simulated STEM-EELS maps were
(a) (b) (c) (d) (i)
(e) (f) (g) (h)
FIG. 2. (a),(b),(c),(d) Experimental π,σ,π=σmaps, and HAADF image, respectively. (e),(f),(g),(h) Theoretical π,σ,π=σmaps
with shot noise, and ADF image, respectively. The position of atomic planes from the HAADF signal is indicated with green circles.
(i) π=σprofiles from (c),(g), and HAADF intensity integrated in the range indicated by the vertical orange, blue, and red bars in (c),(g),
and (d). All scale bars indicate 1 nm.
PHYSICAL REVIEW LETTERS 128, 116401 (2022)
116401-3
calculated with the same parameters (acceleration voltage,
convergence or collection angle, orientation, sampling,
etc.) as used in the experiments. For Fig. 2, the simulated
ideal maps were blurred using a Gaussian filter with a
standard deviation of 1.1 Å to mimic instrumental broad-
ening due to partial coherence of the electron source [45].
Subsequently, shot noise was added based on the exper-
imental noise characteristics, which were evaluated from
the electron intensity in the experimental maps; π:
31676.4, σ:40026.5e=nm2.
The experimental πand σmaps, shown in Figs. 2(a)
and 2(b), respectively, both display higher intensity where
the Cplanes are located. The localization of the σstates on
the Cplanes is expected. For the πstates (which one might
expect to be stronger around the Cplanes), the apparent,
counterintuitive localization on the planes can be explained
by channeling effects of the electron beam. These obser-
vations are confirmed in the computed maps obtained by
inelastic channeling simulations in Figs. 2(e) and 2(f).
Rather than analyzing the absolute intensities, we inves-
tigate the ratio between the πand the σintensities as
shown in Figs. 2(c) and 2(g), as a way to normalize the π
intensity variations. The ratio is maximized between the C
planes in these maps, as exemplified by the vertical π=σ
profiles plotted versus the HAADF intensity in Fig. 2(i).
HAADF intensity minima coincide with π=σintensity
profile maxima, which are almost exactly equidistant
from two graphene layers. The visual agreement between
the calculated and experimental π,σ, and π=σmaps is
supported by the remarkable overlap of the calculated
and experimental π=σline profiles. This successful
reproduction of the experimental data underlines the
robustness of the inelastic channeling calculations per-
formed in this work to interpret the experimental spectral
data. Most importantly, this result provides an undeniable
proof that the contrast obtained from πand σreal-space
fine-structure maps at high resolution does match the
expected localization of corresponding unoccupied elec-
tronic orbitals. It also highlights that beyond the atomic site
where core-level excitation takes place, the localization of
the πand σorbitals in two-dimensional maps is inti-
mately linked to the channeling of the electron beam, and is
thus strongly affected by the specimen projected thick-
ness [46].
While the channeling of the swift electron beam pri-
marily depends on the alignment of the electron beam path
with the atomic columns, the projected thickness also
strongly modifies the atomic-scale contrast in fine structure
maps. To evaluate the influence of the projected thickness
on the expected πand σorbital contrast, we performed
inelastic channeling calculations using the same simulation
parameters as in Fig. 2but considering specimens with
different thicknesses: 0.43 (a single graphene unit cell),
12.8, and 25.6 nm. The latter corresponds to the estimated
thickness of the TEM lamella considered experimentally.
The πand σmaps corresponding to these projected
thicknesses under ideal conditions (no noise, no instru-
mental broadening, etc.) are presented in Figs. 3(c) and
3(d), respectively. These maps differ from those displayed
in Figs. 2(e),2(f), which contain noise and instrumental
broadening. The πmap of the thinnest specimen displays
lobes outside the Cplanes, in agreement with the πcharge
(a) (c)
(b) (d)
FIG. 3. (a) Charge density for the energy interval between 0.73 and 3.46 eVabove the Fermi level. (b) Projected density of states in
graphite. The zaxis corresponds to the crystallographic caxis of graphite, perpendicular to the carbon layers. (c) πmaps for projected
thicknesses of 0.43 (left), 12.8 (middle), and 25.6 nm (right). The position of atomic planes is indicated with green circles. (d) σmaps
for the same thicknesses. All maps are shown without noise and instrumental broadening. All scale bars indicate 5 Å.
PHYSICAL REVIEW LETTERS 128, 116401 (2022)
116401-4
density in Fig. 3(a). Additional intensity is also visible on
the Ccolumns, and becomes more prominent for larger and
more realistic projected thicknesses. At a thickness of
25.6 nm, the intensity of the πstates on the Ccolumns
is stronger than outside the Cplanes, in agreement with the
experimental πmaps in Fig. 2. For all thicknesses, it is
noteworthy that the intensity in the πmaps is expected to
fade out beyond 1Å away from the Cplanes. The
intensity in the σmaps is, as expected, exclusively
contained within the Cplanes, and peaked on the C
columns. In addition, it is noteworthy that the atomic
resolution contrast is smoothed out with increasing thick-
ness. It should be noted that the σmaps also contain some
intensity from pzstates, i.e., states with πsymmetry, as
shown in the PDOS in Fig. 3(b).
These simulated fine structure maps, in which the elastic
channeling conditions of the electron beam were taken into
account, highlight the fact that the specimen thickness must
be considered carefully to interpret STEM-EELS orbital
mapping experiments successfully. Halving the projected
thickness down to 12.8 nm is expected to lead to a result
similar to the current experimental thickness of 25.6 nm.
The direct comparison of πorbital maps with the πcharge
density plot in Fig. 3(a) is not reasonable for the exper-
imental thickness considered, nor even for 12.8 nm, but
only for an unrealistically small thickness of the order of
0.43 nm. Therefore, it is suggested that smaller projected
thickness will only be meaningful below few nm to provide
better visualization of electronic orbitals using STEM-
EELS in the present case. The noise level is also a major
hurdle to overcome, and is clearly visible when comparing
the πand σmaps in Figs. 2(e) (shot noise added) and 3(c)
(no shot noise), and Figs. 2(f) (shot noise added) and 3(d)
(no shot noise), respectively. It is expected that orbital
mapping in STEM-EELS might benefit from a new gen-
eration of detectors with improved sensitivity and lower
noise level [47,48].
In conclusion, the spatial distribution of antibonding π
and σorbitals in epitaxial graphene multilayers was
mapped successfully by electron energy-loss spectroscopy
in the aberration-corrected scanning transmission electron
microscope. Inelastic channeling calculations unambigu-
ously reproduce the experimental πand σorbital maps
with high level of accuracy, and demonstrate the decisive
effect of the specimen thickness on the orbital mapping
capabilities in graphene. The real-space visualization, at
atomic resolution, of unoccupied electronic states with
different symmetry defines a pathway to better understand
chemical bonding at interfaces and defects in solids. This is
particularly relevant to foster defect engineering and tune
the properties of solids for a wide range of promising
applications where physical and chemical phenomena
occur at surfaces (e.g., photocatalysis) or interfaces (e.g.,
spintronics). This work further illustrates the potentiality of
orbital mapping using STEM-EELS.
The electron microscopy work was supported by the
EPSRC (UK). SuperSTEM Laboratory is the EPSRC
National Research Facility for Advanced Electron
Microscopy. The authors would like to thank Hitachi
High-Tech Corporation (UK and Japan), Orsay Physics
and Tescan for the preparation of FIB lamellae. M.B. is
grateful to the SuperSTEM Laboratory for microscope
access, and to the School of Chemical and Process
Engineering at the University of Leeds for a visiting
associate professorship and financial support. M. E. and
S. L. acknowledge funding from the Austrian Science Fund
(FWF) under Grant No. I4309-N36. L. L. acknowledges
funding from U.S. National Science Foundation under
Grant No. EFMA-1741673.
M. B. and M. E. contributed equally to this work.
*mbugnet@superstem.org
stefan.loeffler@tuwien.ac.at
dmkepap@superstem.org
[1] J. M. Zuo, M. Kim, M. Okeeffe, and J. C. H. Spence,
Nature (London) 401, 49 (1999).
[2] P. N. H. Nakashima, A. E. Smith, J. Etheridge, and B. C.
Muddle, Science 331, 1583 (2011).
[3] J. C. Meyer, S. Kurasch, H. J. Park, V. Skakalova, D.
Künzel, A. Groß, A. Chuvilin, G. Algara-Siller, S. Roth,
T. Iwasaki, U. Starke, J.H. Smet, and U. Kaiser, Nat. Mater.
10, 209 (2011).
[4] K. Müller, F. F. Krause, A. B´ech´e, M. Schowalter, V.
Galioit, S. Löffler, J. Verbeeck, J. Zweck, P.
Schattschneider, and A. Rosenauer, Nat. Commun. 5,
5653 (2014).
[5] W. Gao, C. Addiego, H. Wang, X. Yan, Y. Hou, D. Ji, C.
Heikes, Y. Zhang, L. Li, H. Huyan, T. Blum, T. Aoki, Y. Nie,
D. G. Schlom, R. Wu, and X. Pan, Nature (London) 575,
480 (2019).
[6] G. Sánchez-Santolino, N. R. Lugg, T. Seki, R. Ishikawa,
S. D. Findlay, Y. Kohno, Y. Kanitani, S. Tanaka, S. Tomiya,
Y. Ikuhara, and N. Shibata, ACS Nano 12, 8875
(2018).
[7] J. Repp, G. Meyer, S. M. Stojković, A. Gourdon, and C.
Joachim, Phys. Rev. Lett. 94, 026803 (2005).
[8] J. Repp, G. Meyer, S. Paavilainen, F. E. Olsson, and M.
Persson, Science 312, 1196 (2006).
[9] W. Wang, X. Shi, C. Lin, R. Q. Zhang, C. Minot, M. A. Van
Hove, Y. Hong, B. Z. Tang, and N. Lin, Phys. Rev. Lett. 105,
126801 (2010).
[10] L. Gross, N. Moll, F. Mohn, A. Curioni, G. Meyer, F.
Hanke, and M. Persson, Phys. Rev. Lett. 107, 086101
(2011).
[11] R. F. Egerton, Electron Energy-Loss Spectroscopy in the
Electron Microscope (Springer Science & Business Media,
New York, 2011).
[12] M. Bosman, V. J. Keast, J. L. Garcia-Munoz, A. J.
DAlfonso, S. D. Findlay, and L. J. Allen, Phys. Rev. Lett.
99, 086102 (2007).
[13] K. Kimoto, T. Asaka, T. Nagai, M. Saito, Y. Matsui, and K.
Ishizuka, Nature (London) 450, 702 (2007).
PHYSICAL REVIEW LETTERS 128, 116401 (2022)
116401-5
[14] D. A. Muller, L. F. Kourkoutis, M. Murfitt, J. H. Song, H. Y.
Hwang, J. Silcox, N. Dellby, and O. L. Krivanek, Science
319, 1073 (2008).
[15] M. Bugnet, S. Löffler, D. Hawthorn, H. A. Dabkowska,
G. M. Luke, P. Schattschneider, G. A. Sawatzky, G. Radtke,
and G. A. Botton, Sci. Adv. 2, e1501652 (2016).
[16] R. F. Klie, Q. Qiao, T. Paulauskas, A. Gulec, A. Rebola, S.
Öğüt, M. P. Prange, J. C. Idrobo, S. T. Pantelides, S.
Kolesnik, B. Dabrowski, M. Ozdemir, C. Boyraz, D.
Mazumdar, and A. Gupta, Phys. Rev. Lett. 108, 196601
(2012).
[17] N. Gauquelin, E. Benckiser, M. K. Kinyanjui, M. Wu, Y. Lu,
G. Christiani, G. Logvenov, H.-U. Habermeier, U. Kaiser,
B. Keimer, and G. A. Botton, Phys. Rev. B 90, 195140
(2014).
[18] M. Haruta, Y. Fujiyoshi, T. Nemoto, A. Ishizuka, K.
Ishizuka, and H. Kurata, Phys. Rev. B 97, 205139 (2018).
[19] Y. Wang, M. R. S. Huang, U. Salzberger, K. Hahn, W. Sigle,
and P. A. van Aken, Ultramicroscopy 184, 98 (2018).
[20] A. Teurtrie, E. Popova, I. Koita, E. Chikoidze, N. Keller, A.
Gloter, and L. Bocher, Adv. Funct. Mater. 29, 1904958
(2019).
[21] A. Gloter, V. Badjeck, L. Bocher, N. Brun, K. March, M.
Marinova, M. Tenc´e, M. Walls, A. Zobelli, O. St´ephan, and
C. Colliex, Mater. Sci. Semicond. Process. 65, 2 (2017).
[22] H. Tan, S. Turner, E. Yücelen, J. Verbeeck, and G. Van
Tendeloo, Phys. Rev. Lett. 107, 107602 (2011).
[23] N. R. Lugg, M. Haruta, M. J. Neish, S. D. Findlay, T.
Mizoguchi, K. Kimoto, and L. J. Allen, Appl. Phys. Lett.
101, 183112 (2012).
[24] S. Löffler, V. Motsch, and P. Schattschneider, Ultramicros-
copy 131, 39 (2013).
[25] M. J. Neish, M. P. Oxley, J. Guo, B. C. Sales, L. J. Allen,
and M. F. Chisholm, Phys. Rev. Lett. 114, 106101 (2015).
[26] M. J. Neish, N. R. Lugg, S. D. Findlay, M. Haruta, K.
Kimoto, and L. J. Allen, Phys. Rev. B 88, 115120 (2013).
[27] S. Löffler, M. Bugnet, N. Gauquelin, S. Lazar, E. Assmann,
K. Held, G. A. Botton, and P. Schattschneider, Ultramicros-
copy 177, 26 (2017).
[28] K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y.
Zhang, S. V. Dubonos, I. V. Grigorieva, and A. A. Firsov,
Science 306, 666 (2004).
[29] A. K. Geim and K. S. Novoselov, Nat. Mater. 6, 183 (2007).
[30] K. Suenaga and M. Koshino, Nature (London) 468, 1088
(2010).
[31] W. Zhou, M. D. Kapetanakis, M. P. Prange, S. T. Pantelides,
S. J. Pennycook, and J.-C. Idrobo, Phys. Rev. Lett. 109,
206803 (2012).
[32] Q. M. Ramasse, C. R. Seabourne, D.-M. Kepaptsoglou, R.
Zan, U. Bangert, and A. J. Scott, Nano Lett. 13, 4989
(2013).
[33] F.S. Hage, G. Radtke, D. M. Kepaptsoglou, M. Lazzeri, and
Q. M. Ramasse, Science 367, 1124 (2020).
[34] D. M. Kepaptsoglou, T. P. Hardcastle, C. R. Seabourne, U.
Bangert, R. Zan, J. A. Amani, H. Hofsäss, R. J. Nicholls,
R. M. D. Brydson, A. J. Scott, and Q. M. Ramasse, ACS
Nano 9, 11398 (2015).
[35] L. Pardini, S. Löffler, G. Biddau, R. Hambach, U. Kaiser, C.
Draxl, and P. Schattschneider, Phys. Rev. Lett. 117, 036801
(2016).
[36] A. H. Tavabi, P. Rosi, E. Rotunno, A. Roncaglia, L. Belsito,
S. Frabboni, G. Pozzi, G. C. Gazzadi, P.-H. Lu, R. Nijland,
M. Ghosh, P. Tiemeijer, E. Karimi, R. E. Dunin-Borkowski,
and V. Grillo, Phys. Rev. Lett. 126, 094802 (2021).
[37] Y. Liu, Y. Y. Li, S. Rajput, D. Gilks, L. Lari, P. L. Galindo,
M. Weinert, V. K. Lazarov, and L. Li, Nat. Phys. 10, 294
(2014).
[38] C. Berger, Z. Song, X. Li, X. Wu, N. Brown, C. Naud, D.
Mayou, T. Li, J. Hass, A. N. Marchenkov, E. H. Conrad,
P. N. First, and W. A. de Heer, Science 312, 1191 (2006).
[39] G. Nicotra, Q. M. Ramasse, I. Deretzis, A. La Magna,
C. Spinella, and F. Giannazzo, ACS Nano 7, 3045 (2013).
[40] I. Palacio, A. Celis, M. N. Nair, A. Gloter, A. Zobelli, M.
Sicot, D. Malterre, M. S. Nevius, W. A. de Heer, C. Berger,
E. H. Conrad, A. Taleb-Ibrahimi, and A. Tejeda, Nano Lett.
15, 182 (2015).
[41] E. J. Kirkland, Advanced Computing in Electron Micros-
copy (Plenum Press, New York, 1998).
[42] J. M. Cowley and A. F. Moodie, Acta Crystallogr. 10, 609
(1957).
[43] S. Löffler, Study of real space wave functions with electron
energy loss spectrometry, Ph.D. thesis, Vienna University of
Technology, 2013.
[44] P. Blaha, K. Schwarz, F. Tran, R. Laskowski, G. K. H.
Madsen, and L. D. Marks, J. Chem. Phys. 152, 074101
(2020).
[45] P. Schattschneider, M. Stöger-Pollach, S. Löffler, A. Steiger-
Thirsfeld, J. Hell, and J. Verbeeck, Ultramicroscopy 115,21
(2012).
[46] R. Hovden, H. L. Xin, and D. A. Muller, Phys. Rev. B 86,
195415 (2012).
[47] B. Plotkin-Swing, G. J. Corbin, S. De Carlo, N. Dellby, C.
Hoermann, M. V. Hoffman, T. C. Lovejoy, C. E. Meyer, A.
Mittelberger, R. Pantelic, L. Piazza, and O. L. Krivanek,
Ultramicroscopy 217, 113067 (2020).
[48] S. Cheng, A. Pofelski, P. Longo, R. D. Twesten, Y. Zhu, and
G. A. Botton, Ultramicroscopy 212, 112942 (2020).
[49] See Supplemental Material at http://link.aps.org/supplemental/
10.1103/PhysRevLett.128.116401 for a structural model of
the samples used for inelastic channeling simulations, and
further spectral and mapping data used to produce the figures
in the main manuscript.
PHYSICAL REVIEW LETTERS 128, 116401 (2022)
116401-6
... Valence EELS (VEELS) is limited by the delocalization of the excitation on the nanometer scale, much larger than the size of the valence orbitals themselves 17 . Although recent VEELS work has shown atomic-scale contrast in certain energy ranges in graphene, the contrast is a function of inelastic scattering cross sections between different orbitals and sample thickness, making it non-trivial to isolate valence electron charge densities 18,19 . Valence electron densities are commonly measured using scanning tunneling microscopy (STM) 20 , but these are limited to surfaces and energy ranges typically only a few eV below the Fermi level 21 . ...
... A monolayer of MoS 2 is a twodimensional direct bandgap semiconductor in its 2H phase where the Mo atoms are sandwiched between two S atoms (Fig. 1b). The semiconducting nature and direct band gap are useful for optoelectronics and catalysis applications 18,19 . Fig. 1c shows an ADF-STEM image of a super-cell of MoS 2 . ...
Article
Full-text available
Four-dimensional scanning transmission electron microscopy (4D-STEM) has recently gained widespread attention for its ability to image atomic electric fields with sub-Ångstrom spatial resolution. These electric field maps represent the integrated effect of the nucleus, core electrons and valence electrons, and separating their contributions is non-trivial. In this paper, we utilized simultaneously acquired 4D-STEM center of mass (CoM) images and annular dark field (ADF) images to determine the projected electron charge density in monolayer MoS2. We evaluate the contributions of both the core electrons and the valence electrons to the derived electron charge density; however, due to blurring by the probe shape, the valence electron contribution forms a nearly featureless background while most of the spatial modulation comes from the core electrons. Our findings highlight the importance of probe shape in interpreting charge densities derived from 4D-STEM and the need for smaller electron probes.
... With orbital mapping [8,9,10], however, we strive to fly even closer to the sun. Contrary to elemental mapping, a tiny energy window of often less than 2 eV is necessary. ...
... In our case, however, this direction would result only in rotationally symmetric maps with negligible differences between different energy loss windows for pristine graphite [9,38]. Thus, a side view is required and we opt to choose the [1 0 1 0] zone axis, the same as in [10]. Two peaks dominate the DOS above the Fermi energy and, likewise, the energy loss near edge structure (ELNES) of the graphite K-edge. ...
Preprint
Full-text available
A new material characterization technique is emerging for the transmission electron microscope (TEM). Using electron energy-loss spectroscopy, real space mappings of the underlying electronic transitions in the sample, so called orbital maps, can be produced. Thus, unprecedented insight into the electronic orbitals responsible for most of the electrical, magnetic and optical properties of bulk materials can be gained. However, the incredibly demanding requirements on spatial as well as spectral resolution paired with the low signal-to-noise ratio severely limits the day-to-day use of this new technique. With the use of simulations, we strive to alleviate these challenges as much as possible by identifying optimal experimental parameters. In this manner, we investigate representative examples of a transition metal oxide, a material consisting entirely of light elements, and an interface between two different materials to find and compare acceptable ranges for sample thickness, acceleration voltage and electron dose for a scanning probe as well as for parallel illumination.
... Previous efforts to reveal electron orbital information using (S) TEM-based techniques have not gone unrewarded. A variety of electron energy loss spectroscopy (EELS) experiments have successfully been used to reconstruct electron orbitals or tease out bonding information from carefully crafted experiments [9][10][11] . Likewise, convergent beam electron diffraction (CBED) techniques have similarly been used to extract electron orbital information 12 . ...
Article
Full-text available
Recent studies of secondary electron (SE) emission in scanning transmission electron microscopes suggest that material’s properties such as electrical conductivity, connectivity, and work function can be probed with atomic scale resolution using a technique known as secondary electron e-beam-induced current (SEEBIC). Here, we apply the SEEBIC imaging technique to a stacked 2D heterostructure device to reveal the spatially resolved electron density of an encapsulated WSe2 layer. We find that the double Se lattice site shows higher emission than the W site, which is at odds with first-principles modelling of valence ionization of an isolated WSe2 cluster. These results illustrate that atomic level SEEBIC contrast within a single material is possible and that an enhanced understanding of atomic scale SE emission is required to account for the observed contrast. In turn, this suggests that, in the future, subtle information about interlayer bonding and the effect on electron orbitals could be directly revealed with this technique.
Article
The concept of electronic orbitals has enabled the understanding of a wide range of physical and chemical properties of solids through the definition of, for example, chemical bonding between atoms. In the transmission electron microscope, which is one of the most used and powerful analytical tools for high‐spatial‐resolution analysis of solids, the accessible quantity is the local distribution of electronic states. However, the interpretation of electronic state maps at atomic resolution in terms of electronic orbitals is far from obvious, not always possible, and often remains a major hurdle preventing a better understanding of the properties of the system of interest. In this review, the current state of the art of the experimental aspects for electronic state mapping and its interpretation as electronic orbitals is presented, considering approaches that rely on elastic and inelastic scattering, in real and reciprocal spaces. This work goes beyond resolving spectral variations between adjacent atomic columns, as it aims at providing deeper information about, for example, the spatial or momentum distributions of the states involved. The advantages and disadvantages of existing experimental approaches are discussed, while the challenges to overcome and future perspectives are explored in an effort to establish the current state of knowledge in this field. The aims of this review are also to foster the interest of the scientific community and to trigger a global effort to further enhance the current analytical capabilities of transmission electron microscopy for chemical bonding and electronic structure analysis.
Article
Hydrogen donor doping of correlated electron systems such as vanadium dioxide (VO2) profoundly modifies the ground state properties. The electrical behavior of HxVO2 is strongly dependent on the hydrogen concentration; hence, atomic scale control of the doping process is necessary. It is however a nontrivial problem to quantitatively probe the hydrogen distribution in a solid matrix. As hydrogen transfers its sole electron to the material, the ionization mechanism is suppressed. In this study, a methodology mapping the doping distribution at subnanometer length scale is demonstrated across a HxVO2 thin film focusing on the oxygen–hydrogen bonds using electron energy loss spectroscopy (EELS) coupled with first-principles EELS calculations. The hydrogen distribution was revealed to be nonuniform along the growth direction and between different VO2 grains, calling for intricate hydrogenation mechanisms. Our study points to a powerful approach to quantitatively map dopant distribution in quantum materials relevant to energy and information sciences.
Article
Strong elastic deformation resistance and an intrinsic stable energetic state contribute to the mega pressure hysteresis in Ce-rich CaCu 5 -type metal hydrides.
Article
Visualization of individual electronic states ascribed to specific unoccupied orbitals at the atomic scale can reveal fundamental information about chemical bonding, but it is challenging since bonding often results in only subtle variations in the whole density of states. Here, we utilize atomic-resolution energy-loss near-edge fine structure (ELNES) spectroscopy to map out the electronic states attributed to specific unoccupied pz orbital around a fourfold coordinated silicon point defect in graphene, which is further supported by theoretical calculations. Our results illustrate the power of atomic-resolution ELNES towards the probing of defect-site-specific electronic orbitals in monolayer crystals, providing insights into understanding the effect of chemical bonding on the local properties of defects in solids.
Article
Recent advances in scanning transmission electron microscopy have enabled atomic-scale focused, coherent, and monochromatic electron probes, achieving nanoscale spatial resolution, meV energy resolution, sufficient momentum resolution, and a wide energy detection range in electron energy-loss spectroscopy (EELS). A four-dimensional EELS (4D-EELS) dataset can be recorded with a slot aperture selecting the specific momentum direction in the diffraction plane and the beam scanning in two spatial dimensions. In this paper, the basic principle of the 4D-EELS technique and a few examples of its application are presented. In addition to parallelly acquired dispersion with energy down to a lattice vibration scale, it can map the real space variation of any EELS spectrum features with a specific momentum transfer and energy loss to study various locally inhomogeneous scattering processes. Furthermore, simple mathematical combinations associating the spectra at different momenta are feasible from the 4D dataset, e.g., the efficient acquisition of a reliable electron magnetic circular dichroism (EMCD) signal is demonstrated. This 4D-EELS technique provides new opportunities to probe the local dispersion and related physical properties at the nanoscale.
Article
Scanning transmission electron microscopy (STEM) is one of the most powerful characterization tools in materials science research. Due to instrumentation developments such as highly coherent electron sources, aberration correctors, and direct electron detectors, STEM experiments can examine the structure and properties of materials at length scales of functional devices and materials down to single atoms. STEM encompasses a wide array of flexible operating modes, including imaging, diffraction, spectroscopy, and 3D tomography experiments. This review outlines many common STEM experimental methods with a focus on quantitative data analysis and simulation methods, especially those enabled by open source software. The hope is to introduce both classic and new experimental methods to materials scientists and summarize recent progress in STEM characterization. The review also discusses the strengths and weaknesses of the various STEM methodologies and briefly considers promising future directions for quantitative STEM research. Expected final online publication date for the Annual Review of Materials Research, Volume 53 is July 2023. Please see http://www.annualreviews.org/page/journal/pubdates for revised estimates.
Article
Full-text available
The component of orbital angular momentum (OAM) in the propagation direction is one of the fundamental quantities of an electron wave function that describes its rotational symmetry and spatial chirality. Here, we demonstrate experimentally an electrostatic sorter that can be used to analyze the OAM states of electron beams in a transmission electron microscope. The device achieves postselection or sorting of OAM states after electron-material interactions, thereby allowing the study of new material properties such as the magnetic states of atoms. The required electron-optical configuration is achieved by using microelectromechanical systems technology and focused ion beam milling to control the electron phase electrostatically with a lateral resolution of 50 nm. An OAM resolution of 1.5ℏ is realized in tests on controlled electron vortex beams, with the perspective of reaching an optimal OAM resolution of 1ℏ in the near future.
Article
Full-text available
We characterize a hybrid pixel direct detector and demonstrate its suitability for electron energy loss spectroscopy (EELS). The detector has a large dynamic range, narrow point spread function, detective quantum efficiency ≥ 0.8 even without single electron arrival discrimination, and it is resilient to radiation damage. It is capable of detecting ∼5 × 10⁶ electrons/pixel/second, allowing it to accommodate up to 0.8 pA per pixel and hence a >100 pA EELS zero-loss peak (ZLP) without saturation, if the ZLP is spread over >125 pixels (in the non-dispersion direction). At the same time, it can reliably detect isolated single electrons in the high loss region of the spectrum. The detector uses a selectable threshold to exclude low energy events, and this results in essentially zero dark current and readout noise. Its maximum frame readout rate at 16-bit digitization is 2250 full frames per second, allowing for fast spectrum imaging. We show applications including EELS of boron nitride in which an unsaturated zero loss peak is recorded at the same time as inner shell loss edges, elemental mapping of an STO/BTO/LMSO multilayer, and efficient parallel acquisition of angle-resolved EEL spectra (S(q, ω)) of boron nitride.
Article
Full-text available
The WIEN2k program is based on the augmented plane wave plus local orbitals (APW+lo) method to solve the Kohn–Sham equations of density functional theory. The APW+lo method, which considers all electrons (core and valence) self-consistently in a full-potential treatment, is implemented very efficiently in WIEN2k, since various types of parallelization are available and many optimized numerical libraries can be used. Many properties can be calculated, ranging from the basic ones, such as the electronic band structure or the optimized atomic structure, to more specialized ones such as the nuclear magnetic resonance shielding tensor or the electric polarization. After a brief presentation of the APW+lo method, we review the usage, capabilities, and features of WIEN2k (version 19) in detail. The various options, properties, and available approximations for the exchange-correlation functional, as well as the external libraries or programs that can be used with WIEN2k, are mentioned. References to relevant applications and some examples are also given.
Article
Full-text available
The charge-density distribution in materials dictates their chemical bonding, electronic transport, and optical and mechanical properties. Indirectly measuring the charge density of bulk materials is possible through X-ray or electron-diffraction techniques by fitting their structure factors1–3, assuming that the sample is perfectly homogeneous within the beam illuminated area. The recent development of scanning tunnelling microscopy and atomic force microscopy enables us to see chemical bonds, but only on surfaces4–6. It therefore remains a challenge to resolve charge density in nanostructures and functional materials with imperfect crystalline structures, for example those with defects, interfaces or boundaries where new physics emerges. Here we describe the development of a real-space imaging technique that can directly map the local charge density of crystalline materials, using scanning transmission electron microscopy alongside an angle-resolved pixelated fast electron detector. Using this technique, we imaged the interfacial charge distribution and ferroelectric polarization in a SrTiO3/BiFeO3 heterojunction and discovered charge accumulation at the interface that was induced by the penetration of the polarization field of BiFeO3, which is validated through side-by-side comparison with density functional theory calculations. The charge-density imaging method established in this work advances electron microscopy from detecting atoms to imaging electrons, paving a new route towards study local bonding in crystalline solids.
Article
Full-text available
Bismuth iron garnet Bi3Fe5O12 (BIG) is a multifunctional insulating oxide exhibiting remarkably the largest known Faraday rotation and linear magnetoelectric coupling. Enhancing the electrical conductivity in BIG while preserving its magnetic properties would further widen its range of potential applications in oxitronic devices. Here, a site‐selective codoping strategy in which Ca²⁺ and Y³⁺ substitute for Bi³⁺ is applied. The resulting p‐ and n‐type doped BIG films combine state‐of‐the‐art magneto‐optical properties and semiconducting behaviors above room temperature with rather low resistivity: 40 Ω cm at 450 K is achieved in an n‐type Y‐doped BIG; this is ten orders of magnitude lower than that of Y3Fe5O12. High‐resolution electron spectromicroscopy unveils the complete dopant solubility and the charge compensation mechanisms at the local scale in p‐ and n‐type systems. Oxygen vacancies as intrinsic donors play a key role in the conduction mechanisms of these doped BIG films. On the other hand, a self‐compensation of Ca²⁺ with oxygen vacancies tends to limit the conduction in p‐type Ca/Y‐doped BIG. These results highlight the possibility of integrating n‐type and p‐type doped BIG films in spintronic structures as well as their potential use in gas sensing applications.
Article
Seeing single silicon atom vibrations Vibrational spectroscopy can achieve high energy resolution, but spatial resolution of unperturbed vibrations is more difficult to realize. Hage et al. show that a single-atom impurity in a solid (a silicon atom in graphene) can give rise to distinctive localized vibrational signatures. They used high-resolution electron energy-loss spectroscopy in a scanning transmission electron microscope to detect this signal. An experimental geometry was chosen that reduced the relative elastic scattering contribution, and repeated scanning near the silicon impurity enhanced the signal. The experimental vibration frequencies are in agreement with ab initio calculations. Science , this issue p. 1124
Article
Direct electron detectors (DeDs) have been widely used for imaging studies because of their higher beam sensitivity, lower noise, improved pixel resolution, etc. However, there have been limited studies related to the performance in spectroscopic applications for the direct electron detection. Hereby, taking the advantage of the DeD installed on a high-performance electron energy-loss spectrometer, we systematically studied the performance of a DeD (Gatan's K2 IS) fitted on an aberration-corrected transmission electron microscope (TEM) equipped with an electron monochromator. Using SrTiO3 as the model system, the point spread function in the zero-loss region of the spectrum and the performance for core loss spectroscopy have been investigated under both 200 kV and 80 kV operating conditions. We demonstrate that the K2 detector can achieve an overall better performance at 200 kV than a charge coupled device (CCD) detector. At 80 kV, the K2 DeD is still better than a CCD, except for the relative broad tails of the zero-loss peak. The signal-to-noise ratio is very close for DeD and CCD under 80 kV. Based on our data obtained at different operating voltages, it is clear that DeD will benefit the microscopy community and boost the development of cutting-edge materials science studies by pushing the frontiers in electron energy-loss spectroscopy.
Article
Probing the charge density distributions in materials at atomic scale remains an extremely demanding task, particularly in real space. However, recent advances in differential phase contrast-scanning transmission electron microscopy (DPC-STEM) bring this possibility closer by directly visualizing the atomic electric field. DPC-STEM at atomic resolutions measures how a sub-angstrom electron probe passing through a material is affected by the atomic electric field, the field between the nucleus and the surrounding electrons. Here, we perform a fully quantitative analysis which allows us to probe the charge density distributions inside atoms, including both the positive nuclear and the screening electronic charges, with subatomic resolution and in real space. By combining state-of-the-art DPC-STEM experiments with advanced electron scattering simulations we are able to map the spatial distribution of the electron cloud within individual atomic columns. This work constitutes a crucial step toward the direct atomic scale determination of the local charge redistributions and modulations taking place in materials systems.
Article
One of the key factors in understanding high-Tc cuprate superconductors is the spatial distribution of holes in the sample. Since x-ray absorption spectroscopy (XAS) and electron-energy-loss spectroscopy (EELS) can directly measure the unoccupied 2p states of the oxygen, many experiments have already been performed for a wide range of doping conditions. While polarization-dependent x-ray fluorescence yield absorption cannot spatially resolve different oxygen sites, EELS combined with scanning transmission electron microscopy has the ability to resolve them with atomic resolution. However, since cuprate superconductors are extremely sensitive to electron irradiation, it has not been possible, to date, to characterize them with atomic resolution. Here, we succeeded in atomic-resolution two-dimensional mapping of holes in La2−xSrxCuO4±δ by overcoming the problem of irradiation damage using an advanced integration technique. In addition, we have demonstrated the anisotropic chemical bond related to the difference between px and py orbitals was observed with atomic resolution. The present approach enables atomic-resolution anisotropic spectroscopy and is expected to give information complementary to the polarization-dependent XAS.
Article
Electron energy-loss spectroscopy and energy-dispersive X-ray spectroscopy are two of the most common means for chemical analysis in the scanning transmission electron microscope. The marked progress of the instrumentation hardware has made chemical analysis at atomic resolution readily possible nowadays. However, the acquisition and interpretation of atomically resolved spectra can still be problematic due to image distortions and poor signal-to-noise ratio of the spectra, especially for investigation of energy-loss near-edge fine structures. By combining multi-frame spectrum imaging and automatic energy-offset correction, we developed a spectrum imaging technique implemented into customized DigitalMicrograph scripts for suppressing image distortions and improving the signal-to-noise ratio. With practical examples, i.e. SrTiO3 bulk material and Sr-doped La2CuO4 superlattices, we demonstrate the improvement of elemental mapping and the EELS spectrum quality, which opens up new possibilities for atomically resolved EELS fine structure mapping.