Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Indication of thermal roughening in the

retrieved mean inner potential across a


5 grain boundary in SrTiO3 annealed at
different temperatures
Piu Rajak, Sung Bo Lee & Somnath
Bhattacharyya

Journal of Materials Science


Full Set - Includes `Journal of Materials
Science Letters'
ISSN 0022-2461
J Mater Sci
DOI 10.1007/s10853-015-9468-0

1 23

Your article is protected by copyright and all


rights are held exclusively by Springer Science
+Business Media New York. This e-offprint is
for personal use only and shall not be selfarchived in electronic repositories. If you wish
to self-archive your article, please use the
accepted manuscript version for posting on
your own website. You may further deposit
the accepted manuscript version in any
repository, provided it is only made publicly
available 12 months after official publication
or later and provided acknowledgement is
given to the original source of publication
and a link is inserted to the published article
on Springer's website. The link must be
accompanied by the following text: "The final
publication is available at link.springer.com.

1 23

Author's personal copy


J Mater Sci
DOI 10.1007/s10853-015-9468-0

Indication of thermal roughening in the retrieved mean inner


potential across a R5 grain boundary in SrTiO3 annealed
at different temperatures
Piu Rajak1 Sung Bo Lee2 Somnath Bhattacharyya1

Received: 7 July 2015 / Accepted: 28 September 2015


Springer Science+Business Media New York 2015

Abstract We have measured the mean inner potential


depth at a R5 grain boundary in a SrTiO3 bicrystal by
reconstructing the exit-face wave function from an image
focal series collected by transmission electron microscopy.
We find that, as the annealing temperature increases, the
potential depth at the grain boundary exponentially
increases. We interpret the temperature dependence of the
potential depth as the signature of a grain-boundary thermal roughening transition.

Introduction
Atomically smooth crystalline surfaces correspond to singular points such as cusps in the plot of the surface energy
against the surface normal direction (c-plot) [1]. As is well
accepted, an atomically smooth surface undergoes a thermal roughening transition for simple thermodynamic reasons relating to the free energy required to form a step on
the crystalline smooth surface (step free energy) [2, 3]. As
temperature increases, thermal fluctuations increase the
step entropy per unit length, thus decreasing the step free
energy. At a critical temperature, it goes to zero. The

& Sung Bo Lee


bolee@snu.ac.kr
& Somnath Bhattacharyya
somnath.tem@gmail.com
1

Department of Metallurgical and Materials Engineering,


Indian Institute of Technology (IIT) Madras,
Chennai 600036, India

Department of Materials Science and Engineering and


Research Institute of Advanced Materials (RIAM), Seoul
National University, Seoul 08826, South Korea

critical temperature is called the thermal roughening transition temperature, TR. A thermal roughening transition
corresponds to the disappearance of the cusp in the c-plot
[5, 6]. A thermal roughening transition of a smooth surface
for a cusp orientation is accompanied by a facetingdefaceting transition of a vicinal surface inclined from the
cusp orientation [7].
The concepts of thermal roughening and facetingdefaceting transitions [26] hold for grain boundaries (GBs)
[810]. If the GB energy varies strongly with the GB plane
normal direction (or the inclination angle) and there are
thus cusps in the polar plot of the GB energy (c) against the
normal direction, then most GBs, which do not correspond
to the cusp orientations, are expected to be faceted [8] with
a hill-and-valley shape. As for crystalline surfaces, GBs are
indeed observed to undergo a thermal roughening transition and a correlated defaceting transition [9, 10].
Hsieh and Balluffi [9] examined a facetingdefaceting
transition in a R3 asymmetric h111i tilt GB in Al and Au,
and a R11 asymmetric boundaries in Al, which were
observed to be reversible with temperature. (The quantity R
means a degree of fit between the structures of the two
adjoining grains and is described by the reciprocal of the
ratio of coincidence site lattice (CSL) points to the total
number of site points [11, 12].) Lee et al. [10] showed that
a GB with R5 [001] misorientation relationship in a SrTiO3
bicrystal underwent a defaceting transition. Grain boundaries in the specimens as-received and annealed in air at
1100, 1300, and 1500 C showed faceting into long, flat
symmetric (310) planes and short asymmetric components
[(100)//(430)], whereas they became defaceted after
annealing at 1600 C. As temperature decreased, a faceted
structure appeared again, indicating that the facetingdefaceting is reversible [10], as observed by Hsieh and Balluffi [9]. The GBs examined were not observed to contain

123

Author's personal copy


J Mater Sci

any amorphous phase [10]. The defaceting transition


observed in the GB with increasing annealing temperature
was interpreted as being caused by the thermal roughening
transitions of both the symmetric and asymmetric GB
planes [10]. Furthermore, using impedance spectroscopy,
they showed that the GB defaceting transition in the
SrTiO3 R5 bicrystalline GB causes a sudden change in the
charge-carrier transport across the GB between 1500 and
1600 C, no GB impedance in the lower temperature range,
but a GB impedance appearing at 1600 C [10], which
clearly indicates that there is a change in GB potential
between the two temperatures.
It is well accepted that GBs often produce a detrimental
effect on the performance of electrical devices, such as
polycrystalline Si thin-film transistors (TFT) and solar
cells. This is because GBs in these materials can generate
defect states within the band gap acting as preferential sites
for electronhole recombination, thus forming potential
barriers against carrier transportation at GBs and reducing
the electrical conductivity. The GB potential has been
measured by electron-beam-induced current (EBIC) techniques and Kelvin probe force microscopy (KPFM) [13,
14]. Tsurekawa et al. [13] measured the potential barrier
heights at GBs in polycrystalline Si by Kelvin probe force
microscopy (KPFM) and found that the potential barrier
heights were measured to be lower at CSL GBs with low R.
Rizk and Nouet [14] examined the electrical activity of
GBs in Si bicrystals (R9, R13, and R25) uncontaminated
and intentionally contaminated with Cu, Ni, or Fe by deeplevel transient spectroscopy (DLTS), electron-beam-induced current (EBIC) measurements, and thermally stimulated capacitance (TSCAP). They [10] found that the
grain-boundary barrier effect was observed for the
R25 GB. However, they did not clarify why the R9 and
R13 GBs did not exhibit barrier effects. They concluded
that the formation of the GB potential is generally correlated with the presence of precipitates at the GBs. Interestingly, they also detected a barrier effect at the
uncontaminated R25 GB, which was attributed to a structural disorder at the GB. However, they did not examine
the possible variation of the GB potential with increasing
annealing temperature.
The GB potential can also be measured by transmission
electron microscopy (TEM)-related techniques. Incoming
electron that reaches a specimen experiences an electrostatic potential due to inherent electronic nature of the
specimen. The volume average of the electrostatic potential
within the material is termed as mean inner potential [15].
Mean inner potential across GB of SrTiO3 was determined using off-axis electron holography and was correlated with space charge accumulation across GB [16]. This
potential profile across an interface was also determined
using Fresnel contrast analysis [17]. Both the above-

123

mentioned methods were used to determine the change in


the mean inner potential across SiC interfaces, which was
correlated with the spatial charge distribution across a
double Schottky barrier [18, 19]. The mean inner potential
profile was determined using through focal series TEM
images of varying defocus, e.g., across intergranular glassy
films and interfaces between grains and glassy pockets in
polycrystalline Si3N4 [20] to reveal the existence of space
charge layer, across the interfaces of a Ge quantum well
embedded within rare-earth oxide layers [21] to determine
structural abruptness, and within intercalated unit cell of
topological superconductor Cu0.12Be2Se3 [22] to determine
stoichiometry. However, despite strong indication of effect
of thermal fluctuations at the GB on the GB potential [9,
10], relevant studies have been missing.
In this paper, the effect of increasing annealing temperatures on projected mean inner potential depth at the GB
(hereafter GB potential depth, for short) across a R5
SrTiO3 bicrystalline GB (used in the previous study [10])
has been examined. The phase retrieval method we applied
in this present study measures the relative phase change
within an imaged region. Therefore, the measured GB
potential depth here is the difference in mean inner
potential between the adjacent crystal and the GB. The GB
potential depth was determined by reconstructing exit wave
function using a set of defocus images. The change in GB
potential depth is interpreted as corresponding to the
decreasing behavior of the GB step free energy with
increasing temperature.

Experimental methods
We examined nominally undoped SrTiO3 bicrystals containing a 36.8 [001] tilt (R5) GB (Shinkosha Co., Ltd.,
Japan) annealed at various temperatures, which are the
same as used in the previous study [10]. The dimensions of
the bicrystal were 10 9 5 9 0.5 mm3 with the central GB
between the 5 9 5 9 0.5 mm3 single crystals. The
bicrystals were annealed in air at 1100 C for 3 days, at
1300 C for 3 days, at 1500 C for 1 day, and at 1600 C
for 1 day. They were heated at a rate of 10 C/min, and
after annealing, were cooled in the furnace to room temperature. (See Ref. [10] for more details.) For TEM, disks
(3 mm in diameter) were cut from the bicrystal. The disks
were mechanically polished, dimpled, and finally ion milled to perforation by a GATAN DuoMill at an angle of
1213 at 3 keV.
We measured the GB potential depth using the Zeiss
Libra 120 operated at 120 keV, which is equipped with an
in-column Omega-type energy filter (Carl Zeiss AG, Germany) [23]. For the present analysis, the GBs were always
tilted to be edge-on (parallel to the electron beam

Author's personal copy


J Mater Sci

direction). To retrieve exit-face wave function, a set of


bright field images (with an objective aperture of size
-1) at a different defocus were taken using energy
0.1248 A
slit around zero-loss peak (slit width 10 eV). Therefore,
only elastically scattered coherent electrons were used to
form an image. For each specimen, seven images were
taken from 1000 to 4000 nm with a step size of 500 nm.
Therefore, here we mainly retrieved the low-spatial frequency components of the phase. To reconstruct exit-face
wave function, we used the full-resolution wave reconstruction (FRWR) software (http://elim.physik.uni-ulm.de),
which was developed by Koch [24]. Using the two images,
one with an energy window of 10 eV around zero-loss peak
and the other without energy window (with the whole
spectrum), the specimen thickness along the electron beam
was determined using Eq. (22) of our previous report [20].
Theoretical background
When an incoming electron reaches a specimen it experiences an electrostatic potential due to the inherent electronic nature of the specimen. Unit-cell volume averaged
value of this potential is termed as mean inner potential.
The mean inner potential of an assembly of neutral atom is
calculated as [25],
Vr

NX
atoms
h0
j
f s 0
2pm0 jejX j1 el

where X is the unit-cell volume, m0 is the rest mass of the


electron, e is the charge of an electron, h0 is Plancks
P atoms j
constant, and Nj1
fel s 0 is the electron scattering
factor of an assembly of neutral atoms [25] at zero scattering angle.
For transmission electron microscopic (TEM) studies,
we reconstruct mean inner potential distribution of an
imaged region in 2D. When the imaged region is uniform
from all aspects (structure, chemistry, ionicity, etc.), the
projected potential does not have any variation. The
introduction of non-uniformity within the imaged region
produces mean inner potential variation. As non-uniformity
increases, the potential variation increases. Non-uniformity
can be caused by various reasons, such as local changes in
crystal tilt, local changes of chemistry and/or atomic density, and accumulation of space charge. To conclude, a
combination of the aforementioned local changes induces
the variation in the mean inner potential at the GB.
Incoming plane electron wave experiences phase shift
due to the variation in mean inner potential and specimen
thickness (along the electron beam) within an imaged
region. The exit-face wave function can be described as
wr ei/r ;

where phase (/) can be written as


/r rVrt
Vr

/r
rt

where V(r) is the projected mean inner potential, t is the


specimen thickness along the electron beam, and r is the
relativistic electron interaction constant. (For the present
work, the value was 8.631 9 10-3 V-1 nm-1.) We can
write V(r) as [26]
V r V 0 r iV 00 r V 000 r ;
where V0 (r) is the electrostatic potential due to the elastic
interaction of electrons with the sample, V00 (r) is the
electron loss due to the inelastic scattering, and V 000 r is
the elastic scattering (and quasi-elastic thermal diffuse
scattering) outside the objective aperture which can be
described as an objective aperture-dependent pseudo-absorptive scattering potential [27]. Thickness can be
determined using two images, zero-loss filtered and unfiltered [27].
In the previous study [10], the GB assumed a faceted
morphology with long flat symmetric (310) planes separated by short asymmetric components [(100)//(430)] in the
specimens as-received and annealed at 1100, 1300, and
1500 C. However, the specimen annealed at 1600 C
became wavy and defaceted. For the faceted GBs shown in
the specimens as-received and annealed at 1100, 1300, and
1500 C, the potential measurements were made in long
symmetric regions. For the defaceted morphology in the
specimen annealed at 1600 C, suitable GB regions were
selected for the measurements.
The faceted and defaceted structures are not expected to
change during air cooling from the high temperatures used.
This expectation is corroborated by observations in our
previous study [10]. In the report [10], the SrTiO3 bicrystal
was first annealed at 1600 C and then cooled down and
annealed at 1100 C. The specimen was cooled in the
furnace to room temperature. The GB in the step-annealed
specimen was faceted, while the GB in the specimen only
annealed at 1600 C was observed to be defaceted. The
observations indicate that the structures obtained at the
high temperatures were effectively trapped to room
temperature.

Results and discussion


Figure 1 shows reconstructed amplitude, phase, and surface plots (normalized with the minimum value of the
imaged region) of the phase of the specimens as-received,
and annealed at the various temperatures (please refer to

123

Author's personal copy


J Mater Sci

Fig. 1 Reconstructed amplitude, phase, and surface plots of the


phase of the specimens as-received (ac) and annealed at 1100 C (d
f), 1300 C (gi), 1500 C (jl), and 1600 C (mo). Due to the
coherence criterion, keeping GB in the middle of the imaged regions,
we are only concerned about the phase variation within a distance
13.4 nm in reconstructed phases (presented in b, e, h, k, and n) and in
corresponding surface plots. Larger regions were shown for presentation purpose only

Ref. [24] to know details about reconstructed phase and


amplitude).
The phase relationship between two points within the
imaged region depends on coherence width (Xc). Within
the distance, the illuminating radiation can be treated as
perfectly coherent [28]. Coherence width can be measured
using the equation Xc = k/2phc, where k is the electron

123

wavelength and hc is the illumination semi-angle (the beam


convergence). For our experiments, the illumination semiangle was kept at 0.25 mrad and, using the above equation,
the coherence width was calculated to be 2.13 nm. Like
coherence width, there is another criterion called incoherence width (Xi) [28], beyond which the illumination is
completely incoherent. Since we have used partially
coherent electron beam, so incoherence width is also
important, which is, in our experiments (Xi = k/hc),
13.4 nm. Over the distances between Xc and Xi, the illumination on the object is partially coherent. Beyond the
distances of Xi (13.4 nm in the present work), the illumination becomes fully incoherent. Therefore, keeping GB in
the middle of the imaged regions, we were only concerned
about the phase variation within a distance 13.4 nm in
reconstructed phases (Fig. 1b, e, h, k, and n) and in corresponding surface plots. Regions at both sides of the GB
of SrTiO3 shown in Fig. 1 are only for presentation
purpose.
Variation in mean inner potential occurs due to the
introduction of non-uniformity within the imaged region.
As mentioned above, non-uniformity is caused by various
irregularities at the GB, such as local changes in crystal tilt,
local changes in chemistry and/or atomic density, and
accumulation of space charge, its increase leading to the
increase in potential variation. Grain-boundary grooving
arising from TEM specimen preparation can be also a
cause of potential variation.
We always selected the region where t was almost
uniform along the GB to determine projected mean inner
potential of the GB and subtracted this value from the
potential of the adjacent region which had the same t to
measure the GB potential depth. For each of the specimens, each measurement of the GB potential depth was
averaged over 2 nm distance along the GB and such a
measurement was repeated over around 100 nm along the
length of GB to have a good statistics. When measuring
GB potential depth, we always considered the straight
projected region of the GB. Table 1 shows the average
values of the measurements (*50 sets for each of the
specimens) with their standard deviations. For the GB
width, we took a line scan across the reconstructed
amplitude of the GB (averaged over 2 nm distance along
the GB). Its full width at half maximum (FWHM) was
measured as the GB width. This measurement was done
along the GB over 100 nm distance, and the average
values with their standard deviations are specified in
Table 1.
Plotting the GB potential depth as a function of temperature (Fig. 2), it can be found that the GB potential
depth remains nearly the same for the specimens as-received and annealed at 1100 C, but that it starts to
abruptly increase between 1100 and 1300 C, showing a

Author's personal copy


J Mater Sci
Table 1 Summary of changes
in specimen thickness, GB
width, and GB potential depth
with increasing annealing
temperature

Annealing temperature

Specimen thickness (nm)

As-received

15.5 6.3

1.65 0.3

0.8 0.05

13.7 4.1

2.24 0.4

0.75 0.03
1.45 0.07

1300 C

11 3.5

2.77 0.4

1500 C

14.6 3.2

3.83 0.3

2.9 0.15

1600 C

12.7 2.4

3.37 0.4

3.8 0.2

Potential depth (V)

3.5
3.0
2.5
2.0
1.5
1.0

As-rcv

1100

1200

GB potential depth (V)

1100 C

4.0

0.5

GB width (FWHM) (nm)

1300

1400

1500

1600

Annealing temperature ( C)
Fig. 2 Plot of the change in GB potential depth with increasing
annealing temperature. As-rcv in the x-axis label means the asreceived specimen. The error bar is the standard deviation of the
measurements for each temperature

kind of an exponential profile. The deviation values in the


determined GB potential depth (error bar) also exhibits a
similar tendency, revealing larger values for the specimens
annealed at 1500 and 1600 C.
Our previous report [10] showed that the GBs
annealed at 1100, 1300, and 1500 C were observed to
be faceted, but in contrast, the GB became defaceted in
that annealed at 1600 C. As suggested by Cahn [8] and
Hsieh and Balluffi [9], it is explained by a faceting
defaceting transition accompanying thermal roughening
transitions of cusp-oriented GB planes, which are the
(310) symmetric and (100)//(430) asymmetric planes for
the present study. There is no dispute that a roughening
transition occurs thermodynamically. Thus, in the present study, we relate the dependence of the potential
depth at the (310) symmetric grain-boundary plane on
annealing temperature to its thermal roughening transition behavior.
Certainly, even below the thermal roughening transition
temperature, TR, of a GB plane, the roughness of the grainboundary plane is likely to increase with increasing
annealing temperature to TR. As noted in the introduction,
this is due to the increase in step entropy (i.e., the

decrease in step free energy) with increasing annealing


temperature.
As mentioned above, the GB potential would be
affected by local changes in crystal tilt, local changes of
chemistry and/or atomic density, and accumulation of
space charge. However, it should be noted that the GB is
expected to undergo a thermal roughening transition
thermodynamically with increasing temperature and that
the effect of the roughening transition is likely to be
revealed in the change in GB potential depth with
temperature.
As noted above, the thermal roughening transition is
explained in terms of the disappearance of the GB step free
energy with increasing temperature. As temperature
increases, the step free energy decreases very slowly at
relatively low temperatures, but abruptly at higher temperatures, following KosterlitzThouless type [29].
The values of the GB potential depth and its deviation
measured in the present study indicate the degree of disorder of the GB. As noted in the previous work [10], the
facetingdefaceting transition examined is correlated with
the thermal roughening transition of the long, flat symmetric (310) plane, which is a subject of the present
analyses. If the temperature dependence of the degree of
disorder (Fig. 2) can be defined as a course to a thermal
roughening transition, it should be also explained by the
decreasing behavior of the GB step free energy with
increasing temperature. Put it another way, the decreasing
behavior of the GB step free energy should be reflected in
the temperature dependence of GB potential. The
decreasing behavior of the GB step free energy is expected
to cause the degree of disorder at the GB to change in the
following way. As mentioned above, at relatively low
temperatures, the step free energy gradually decreases,
which corresponds to a gradual increase in the degree of
disorder. The abrupt decease in step free energy at higher
temperatures will be represented as an abrupt increase in
the degree of disorder. Such an increasing tendency is
found in the profile of the GB potential depth (Fig. 2),
which certainly suggests that the increases in GB potential
depth and in its deviation with increasing temperature are
characterized by a course to thermal roughening. As shown
in Table 1, the GB width tended to increase linearly with
increasing annealing temperature. Although indicating a

123

Author's personal copy


J Mater Sci

structural disordering with increasing annealing temperature, the change in GB width does not seem to well reflect
the thermal roughening behavior, as compared with the
case of the GB potential depth (Fig. 2).
For the specimen annealed at 1600 C, since the GB was
observed to be fully defaceted, the GB potential depth was
measured in suitable regions but not the (310) symmetric
planes. However, the average orientation of the GB was
nominally close to the symmetric (310) plane [10].
Therefore, it is reasonable to assume that the GB potential
measured from the specimen annealed at 1600 C
approximates to that of a thermally roughened (310) plane.

Concluding remarks
To sum up, we measured the GB potential depth in a
SrTiO3 bicrystal, which is the same as used in the previous
study [10], by reconstructing exit-face wave function using
a set of defocused images. As noted above, the magnitude
of the GB potential depth and its deviation indicates the
degree of disorder of the GB, which is expected to increase
with increasing annealing temperature in terms of a thermal
roughening transition. Actually, the potential depth at the
GB increases with increasing annealing temperature. The
tendency is consistent with the decreasing behavior of the
GB step free energy with increasing temperature (i.e.,
thermal roughening transition).
Acknowledgements This research was supported by the Basic
Science Research Program through the National Research Foundation
of Korea (NRF) funded by the Ministry of Education (NRF2013R1A1A2005181) (RIAM).

References
1. Herring C (1951) Some theorems on the free energies of crystal
surfaces. Phys Rev 82:8793
2. Burton WK, Cabrera N, Frank FC (1951) The growth of crystals
and the equilibrium structure of their surfaces. Phil Trans R Soc
Lond A 243:299358
3. van Beijeren H (1977) Exactly solvable model for the roughening
transition of a crystal surface. Phys Rev Lett 38:993996
4. Avron JE, Balfour LS, Kuper CG, Landau J, Lipson SG, Schulman LS (1980) Roughening transition in the 4He solid-superfluid
interface. Phys Rev Lett 45:814817
5. Rottman C, Wortis M (1984) Statistical mechanics of equilibrium
crystal shapes: interfacial phase diagrams and phase transitions.
Phys Rep 103:5979
6. Wortis M (1988) Equilibrium crystal shapes and interfacial phase
transitions. In: Vanselow R, Howe RF (eds) Chemistry and
physics of solid surfaces VII. Springer, Berlin, pp 367405
7. Robinson IK, Vlieg E, Hornis H, Conrad EH (1991) Surface
morphology of Ag(110) close to its roughening transition. Phys
Rev Lett 67:18901893
8. Cahn J (1982) Transitions and phase equilibria among grain
boundary structures. J Phys Colloque 43:C6-199C6-213

123

9. Hsieh TE, Balluffi RW (1989) Observations of roughening/defaceting phase transitions in grain boundaries. Acta Metall
37:21332139
10. Lee SB, Lee JH, Cho PS, Kim DY, Sigle W, Phillipp F (2007)
High-temperature resistance anomaly at a strontium titanate grain
boundary and its correlation with grain-boundary faceting-defaceting transition. Adv Mater 19:391395
11. Kronberg ML, Wilson FH (1949) Secondary recrystallization in
copper. Metall Trans 185:501514
12. Brandon DG, Ralph B, Ranganathan S, Wald MS (1964) A field
ion microscope study of atomic configuration at grain boundaries.
Acta Metall 12:813821
13. Rizk R, Nouet G (1997) Electrical activity and structural configuration of transition metals (Cu, Ni, Fe) in silicon bicrystals.
Interface Sci 4:303316
14. Tsurekawa S, Kido K, Watanabe T (2005) Measurements of
potential barrier height of grain boundaries in polycrystalline
silicon by Kelvin probe force microscopy. Philos Mag Lett
85:4149
15. Ross FM, Stobbs WM (1991) Computer modelling for Fresnel
contrast analysis. Philos Mag A 63:3770
16. Ravikumar V, Rodrigues RP, Dravid VP (1997) Space-charge
distribution across internal interfaces in electroceramics using
electron holography: I, pristine grain boundaries. J Am Ceram
Soc 80:11171130
17. Dunin-Borkowski RE (2000) The development of Fresnel contrast analysis, and the interpretation of mean inner potential
profiles at interfaces. Ultramicroscopy 83:193216
18. Kleebe HJ, Siegelin F (2003) Schottky barrier formation in liquid-phase-sintered silicon carbide. Z Metallkd 94:211217
19. Siegelin F, Kleebe HJ, Sigl LS (2003) Interface characteristics
affecting electrical properties of Y-doped SiC. J Mater Res
18:26082617
20. Bhattacharyya S, Koch CT, Ruhle M (2006) Projected potential
profiles across intergranular glassy films. J Ceram Soc Jpn
114:10051012
21. Das T, Bhattacharyya S (2011) Structure and chemistry across
interfaces at nanoscale of a Ge quantum well embedded within
rare earth oxide layers. Microsc Microanal 17:759765
22. Das T, Bhattacharyya S, Joshi BP, Thamizhavel A, Ramakrishnan S (2013) Direct evidence of intercalation in a topological
insulator turned superconductor. Mater Lett 93:370373
23. Bhattacharyya S, Jinschek JR (2008) Retrieval of absorptive
potential variation in electron beam sensitive specimens using a
single energy-filtered bright-field TEM image. Microsc Res Tech
71:669675
24. Koch CT (2008) A flux-preserving non-linear inline holography
reconstruction algorithm for partially coherent electrons.
Ultramicroscopy 108:141150
25. Peng LM (1999) Electron atomic scattering factors and scattering
potentials of crystals. Micron 30:625648
26. Bhattacharyya S, Subramaniam A, Koch CT, Cannon RM, Ruhle
M (2006) The evolution of amorphous grain boundaries during
in situ heating experiments in Lu-Mg doped Si3N4. Mater Sci Eng
A 422:92101
27. Bhattacharyya S, Koch CT, Ruhle M (2006) Projected potential
profiles across interfaces obtained by reconstructing the exit face
wave function from through focal series. Ultramicroscopy
106:525538
28. Spence JCH (1981) Experimental high resolution electron
microscopy. Clarendon Press, Oxford, p 93
29. Kosterlitz JM, Thouless DJ (1973) Ordering, metastability and
phase transitions in two-dimensional systems. J Phys C
6:11811203

You might also like