Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe₁–xTex
The iron-chalcogenide superconductor FeSe₁–xTex (0.5 < x < 1) was investigated by scanning-tunneling microscopy/ spectroscopy (STM/STS) and break-junction techniques. In the STM topography of the samples, randomly distributed Te and Se surface atomic structure patterns correlate well with th...
Gespeichert in:
Datum: | 2013 |
---|---|
Hauptverfasser: | , , |
Format: | Artikel |
Sprache: | English |
Veröffentlicht: |
Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України
2013
|
Schriftenreihe: | Физика низких температур |
Schlagworte: | |
Online Zugang: | http://dspace.nbuv.gov.ua/handle/123456789/118228 |
Tags: |
Tag hinzufügen
Keine Tags, Fügen Sie den ersten Tag hinzu!
|
Назва журналу: | Digital Library of Periodicals of National Academy of Sciences of Ukraine |
Zitieren: | Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe₁–xTex / T. Ekino, A. Sugimoto, A.M. Gabovich // Физика низких температур. — 2013. — Т. 39, № 3. — С. 343–353. — Бібліогр.: 50 назв. — англ. |
Institution
Digital Library of Periodicals of National Academy of Sciences of Ukraineid |
irk-123456789-118228 |
---|---|
record_format |
dspace |
spelling |
irk-123456789-1182282017-05-30T03:02:48Z Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe₁–xTex Ekino, T. Sugimoto, A. Gabovich, A.M. К 75-летию со дня рождения И. К. Янсона The iron-chalcogenide superconductor FeSe₁–xTex (0.5 < x < 1) was investigated by scanning-tunneling microscopy/ spectroscopy (STM/STS) and break-junction techniques. In the STM topography of the samples, randomly distributed Te and Se surface atomic structure patterns correlate well with the bulk composition, demonstrating that nanoscale surface features directly reflect bulk properties. The high-bias STS measurements clarified the gap-like structure at ≈ 100–300 meV, which is consistent with the break-junction data. These highenergy structures were also found in sulfur substituted FeS₀.₁Te₀.₉. Possible origin of such spectral peculiarities is discussed. The superconducting gap 2Δ ≈ 3.4 ± 0.2 meV at temperature T = 4.2 K was found in the break junction of FeSe₁–xTex with the critical temperature Tc ≈ 10 K. The corresponding characteristic gap to Tc ratio 2Δ/kBTc ≈ 4 ± 0.2 indicates moderate superconducting coupling (kB is the Boltzmann constant). 2013 Article Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe₁–xTex / T. Ekino, A. Sugimoto, A.M. Gabovich // Физика низких температур. — 2013. — Т. 39, № 3. — С. 343–353. — Бібліогр.: 50 назв. — англ. 0132-6414 PACS: 74.50.+r, 74.55.+v, 74.70.–b, 74.70.Xa http://dspace.nbuv.gov.ua/handle/123456789/118228 en Физика низких температур Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України |
institution |
Digital Library of Periodicals of National Academy of Sciences of Ukraine |
collection |
DSpace DC |
language |
English |
topic |
К 75-летию со дня рождения И. К. Янсона К 75-летию со дня рождения И. К. Янсона |
spellingShingle |
К 75-летию со дня рождения И. К. Янсона К 75-летию со дня рождения И. К. Янсона Ekino, T. Sugimoto, A. Gabovich, A.M. Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe₁–xTex Физика низких температур |
description |
The iron-chalcogenide superconductor FeSe₁–xTex (0.5 < x < 1) was investigated by scanning-tunneling microscopy/
spectroscopy (STM/STS) and break-junction techniques. In the STM topography of the samples, randomly
distributed Te and Se surface atomic structure patterns correlate well with the bulk composition, demonstrating
that nanoscale surface features directly reflect bulk properties. The high-bias STS measurements
clarified the gap-like structure at ≈ 100–300 meV, which is consistent with the break-junction data. These highenergy
structures were also found in sulfur substituted FeS₀.₁Te₀.₉. Possible origin of such spectral peculiarities
is discussed. The superconducting gap 2Δ ≈ 3.4 ± 0.2 meV at temperature T = 4.2 K was found in the break junction
of FeSe₁–xTex with the critical temperature Tc ≈ 10 K. The corresponding characteristic gap to Tc ratio
2Δ/kBTc ≈ 4 ± 0.2 indicates moderate superconducting coupling (kB is the Boltzmann constant). |
format |
Article |
author |
Ekino, T. Sugimoto, A. Gabovich, A.M. |
author_facet |
Ekino, T. Sugimoto, A. Gabovich, A.M. |
author_sort |
Ekino, T. |
title |
Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe₁–xTex |
title_short |
Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe₁–xTex |
title_full |
Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe₁–xTex |
title_fullStr |
Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe₁–xTex |
title_full_unstemmed |
Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe₁–xTex |
title_sort |
scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of fese₁–xtex |
publisher |
Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України |
publishDate |
2013 |
topic_facet |
К 75-летию со дня рождения И. К. Янсона |
url |
http://dspace.nbuv.gov.ua/handle/123456789/118228 |
citation_txt |
Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe₁–xTex / T. Ekino, A. Sugimoto, A.M. Gabovich // Физика низких температур. — 2013. — Т. 39, № 3. — С. 343–353. — Бібліогр.: 50 назв. — англ. |
series |
Физика низких температур |
work_keys_str_mv |
AT ekinot scanningtunnelingmicroscopyspectroscopyandbreakjunctiontunnelingspectroscopyoffese1xtex AT sugimotoa scanningtunnelingmicroscopyspectroscopyandbreakjunctiontunnelingspectroscopyoffese1xtex AT gabovicham scanningtunnelingmicroscopyspectroscopyandbreakjunctiontunnelingspectroscopyoffese1xtex |
first_indexed |
2025-07-08T13:35:32Z |
last_indexed |
2025-07-08T13:35:32Z |
_version_ |
1837085999112388608 |
fulltext |
© T. Ekino, A. Sugimoto, and A.M. Gabovich, 2013
Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3, pp. 343–353
Scanning-tunneling microscopy/spectroscopy and
break-junction tunneling spectroscopy of FeSe1–xTex
T. Ekino and A. Sugimoto
Hiroshima University, Graduate School of Integrated Arts and Sciences, Higashi-Hiroshima 739-8521, Japan
E-mail: ekino@hiroshima-u.ac.jp
A.M. Gabovich
Institute of Physics, National Academy of Sciences of Ukraine, 46 Nauka Ave., Kyiv 03028, Ukraine
Received November 12, 2012
The iron-chalcogenide superconductor FeSe1–xTex (0.5 < x < 1) was investigated by scanning-tunneling mi-
croscopy/spectroscopy (STM/STS) and break-junction techniques. In the STM topography of the samples, ran-
domly distributed Te and Se surface atomic structure patterns correlate well with the bulk composition, demon-
strating that nanoscale surface features directly reflect bulk properties. The high-bias STS measurements
clarified the gap-like structure at ≈ 100–300 meV, which is consistent with the break-junction data. These high-
energy structures were also found in sulfur substituted FeS0.1Te0.9. Possible origin of such spectral peculiarities
is discussed. The superconducting gap 2Δ ≈ 3.4 ± 0.2 meV at temperature T = 4.2 K was found in the break junc-
tion of FeSe1–xTex with the critical temperature Tc ≈ 10 K. The corresponding characteristic gap to Tc ratio
2Δ/kBTc ≈ 4 ± 0.2 indicates moderate superconducting coupling (kB is the Boltzmann constant).
PACS: 74.50.+r Tunneling phenomena; Josephson effects;
74.55.+v Tunneling phenomena: single particle tunneling and STM;
74.70.–b Superconducting materials other than cuprates;
74.70.Xa Pnictides and chalcogenides.
Keywords: tunneling spectroscopy, break junction, scanning-tunneling microscopy/spectroscopy, energy gap,
iron-based superconductors, FeSe1–xTex.
1. Introduction
The discovery of iron-arsenide superconductor
LaFeAsO1–xFx exhibiting Tc = 26 K in 2008 [1] stimu-
lated the subsequent synthesis of novel iron-based super-
conductors, e.g., LiFeAs [2], BaFe2As2 [3], and Fe(Se,Te)
[4]. In particular, SmFeAsO1–xFx compound has the high-
est Tc 55 K [5–8], which is a record among non-cuprate
superconductors. As for the microscopic Cooper pairing
mechanism supporting such high Tc’s, several possible
candidates were proposed [9,10] and until now it is unclear
which of them is a true one.
Among iron-based superconductors there is a PbO type
β-FeSe with the simplest crystal structure. Its specific fea-
ture, especially beneficial for surface studies, is a good
cleavability over the ab-plane. The sample surface is be-
lieved to be so stable that it is not expected to be recon-
structed. The superconductivity of FeSe maintains against
substitution of the chalcogenide ions such as Fe(Se,M)
(M = Si, Sb, S, Te) [11,12], and Fe(Te,S) [13]. For in-
stance, Tc of Te substituted compound Fe(Se,Te) can be
easily manipulated by changing the Se/Te composition
ratio. In particular, Tc of FeSe is enhanced up to ~15 K by
replacing Se by Te, which has the larger ion radius, while
Tc of FeSe is only 8 K. Furthermore, Tc rises up to 37 K
under the high pressure of 8.9 GPa [14], which is especial-
ly remarkable because such a high critical temperature
occurs in a pure binary compound system. Therefore, it
seems very important to investigate gradual composition
changes and their influence on superconductivity micro-
scopically, e.g., as a function of the Se/Te ratio.
In this paper, single crystals of Fe1.01Se1–xTex (x = 0.5–1)
were investigated by means of the scanning tunneling mi-
croscopy (STM) and spectroscopy (STS), and the results of
the nanoscale surface measurements within the large Te
range were compared and discussed. The tunneling spec-
troscopy measurements of the superconducting gapping
were also carried out by the break-junction (BJ) method
which is extremely sensitive to the electronic spectrum
variations in the superconducting state.
T. Ekino, A. Sugimoto, and A.M. Gabovich
344 Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3
2. Experimental
Fe1.01Se1–xTex single crystals were synthesized by a
standard process [15]. The mixed powder of Fe, Se, and Te
pressed into pellet was double-sealed in an evacuated
quartz tube, which was held at 1000 °C for 36 hours to
cool down to 400 °C at a rate of –3 °C/h, followed by fur-
nace cooling to room temperature, T. The pristine samples
thus obtained were annealed at 400 °C for 100 hours. The
electron probe micro analyzer (EPMA) was employed to
determine the actual composition of the crystal. The
resistivity measurements were done by a standard dc four
probe method.
The STM apparatus used in this experiment was an
Omicron LT-UHV-STM system, which has been modified
to further reduce external disturbance of the sample
[16,17]. The sample was cleaved along the layer in situ at
T = 77 K in an ultra-high vacuum (UHV) sample prepara-
tion chamber of ~10–8 Pa to avoid any contamination or
migration of atoms on the crystal surface. The Pt/Ir wire
was used as the tunneling tip, which was cleaned by a
high-voltage field emission process with Au single crystal
target before the scanning operation. The STM measure-
ments were carried out at T = 4.9–77 K under the UHV
condition of ~ 10–8 Pa evacuated by the ion pump. A con-
stant current mode was adopted to obtain the STM images.
The BJ tunneling spectra were measured using an ac mod-
ulation technique with lock-in amplifier. By this method,
fresh and clean superconductor–insulator–superconductor
(SIS) junction interface can be obtained along the crack of
the thin platelet single crystal at T = 4.2 K [18,19].
3. Results and discussion
Prior to the STM/STS measurements, we determined
the composition ratio x from EPMA in Fe1.01Se1–xTex
with 0.5 < x < 1 (hereafter, we denote it as FeSe1–xTex).
The results showed that the analyzed compositions of Te
and Se were in a good agreement with nominal x. The
found Te content exceeded x in the range 0.2 < x < 0.5.
The analyzed Fe content 1.02 was slightly larger than the
nominal value of 1.01. Single crystals of nominal x < 0.2
were not obtained in the present synthesis procedures.
From the temperature dependence of resistivity, the maxi-
mum Tc was determined: Tc (onset) = 15.9 K, Tc(0) = 14.5 K
for x = 0.5. The critical temperature decreases with either
increasing or decreasing x from 0.5. The lowest onset Tc
was found as ~ 11 K for x = 0.9. In the T-dependence of
resistivity for the annealed FeTe sample (x = 1), a noticea-
ble drop in the resistivity, which is related to the tetragon-
al-to-orthorhombic structural phase transition, was ob-
served at around 72 K [20–23]. As the Te content x
decreases, the resistivity-drop temperature also goes down
accompanying the increase of resistivity. Such a feature
was not observed for x = 0.9 before the annealing process.
Before the STM measurements, we confirmed the sur-
face electronic cleanliness by measuring the dependence
of tunneling current I on the tip-sample distance z, I(z).
Figure 1 shows that I(z) curve is exponential. From the
perfect linearity of the logI-z dependence, the local work
function, φ = 5.2 eV, is determined. The spatial resolution
of the tip was ascertained by monitoring the atomic struc-
tures of gold single crystal during the tip preparation
process. For the STM measurements, the sample bias
voltage V = ± 0.01–0.8 V and the tunneling current It =
= 0.3–0.4 nA were adopted as the feedback conditions, but
the results did not substantially depend on these parame-
ters. The STM images did not depend on the annealing
process. Clear surface spots of atomic arrangements form-
ing the square-lattice structures were always observed. The
period of lattice spot structures is ~ 0.37–0.38 nm for all
compositions x, thus no distinguishable differences were
found among them within the resolution.
The STM images for FeSe1–xTex with various x meas-
ured at 4.9 K are shown in Fig. 2. Since the crystal struc-
ture consists of the stacking of edge-sharing FeSe4 tetrahe-
dral layers, the cleaved surface is always the same. This is
very advantageous for the investigation of the ab-plane
properties in this compound. They exhibit the coexistence
of the bright and dark spots for all the topographies. In
Fig. 2(a), bright and dark spots are shown to be randomly
distributed over the wide area of the atomic image. Such
coexistence is realized only for the limited range of com-
positions x, where bright and dark spots correspond to Te
and Se atoms, respectively, as was discussed earlier [15].
Similar STM images were also reported in Refs. 24–27.
Fig. 1. Dependence of the tunneling current I on the tip-sample
distance z between PtIr tip and FeSe1–xTex single crystal. The
upper-right inset shows the optical micrograph of the single crys-
tal, while the lower left inset displays a histogram of the local
work function (the barrier height).
1 mm
0 2 4 6 8
Local work function, eV
F
re
q
u
en
cy
4
2
0,1
8
6
4
2
0,1
8
6
4
0.1 0.2
z, nm
I,
n
A
Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe1–xTex
Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 345
The dynamic range of the topographic contrast slightly
depends on the frames. The bright spots are located just
on the atomic grid points of the square lattice structure for
x = 0.5–0.9 (Figs. 2(a)–(d)), while the bright shining spots
shown for x = 1, Fig. 2(e), seem to be slightly away from
the formal atomic positions and extend to a few atoms.
Therefore, the origin of these spots in Fig. 2(e) is probably
different from that of the bright spots in Figs. 2(a)–(d). By
assuming these spots in Fig. 2(e) as excess Fe atoms, the
clusters are counted in the larger area of topography, and
the ratio of ~ 0.05 was obtained. This is consistent with the
data from EPMA. The atomic spots in Fig. 2(e) except the
discussed Fe spots are homogeneous, which are naturally
attributed to Te atoms. Interestingly, Te atoms as the brigh-
ter spots in Figs. 2(a)–(d) turn into the darker spots in
Fig. 2(e), but it is understood if we consider the difference
in the atomic sizes (Te atom having larger size than Se
one) for Figs. 2(a)–(d) or the height of the excess Fe atoms
on the surface for Fig. 2(e), respectively, which results in
relative change in the contrast.
A direct evidence of distinguishing atoms in line with
the discussion is given in Fig. 3, where the STM topogra-
phy and its cross section along the indicated line is pre-
sented for x = 0.7 and x = 0.9. The ripple patterns reflect-
ing the sample-tip distance being kept constant are readily
seen in the cross sectional profiles. The positions of spiky
dips among the regular shallow ripples in the height pro-
files in Figs. 3(a) and (b) correspond to the darker spots in
the topography. The depth of the spikes is ~ 0.02–0.03 nm
for both profiles. Since the ionic radii of Te and Se are 0.221
and 0.198 nm, respectively, the difference in the height z
reflects just the difference of ionic radii 0.023 nm be-
tween Te and Se. Therefore, the difference in ion sizes be-
tween them is now visualized. Figure 3(c) shows the dark-
spot ratio in the STM images versus the Se content (1–x)
determined from EPMA. The dark-spot ratio was defined
by counting the numbers of dark nd and bright nb spots
within the area of typically 30×30 nm and then normalized
as nd/(nb + nd). The bright shinning spots discussed above
were sometimes observed in the whole x ranges, but they
were omitted from the counting. Almost perfect linear rela-
tionship was obtained between the dark-spot ratio on the
STM image and the Se content taken from EPMA. From
these results, we identify the origin of the bright and dark
spots in the whole investigated range of 0.5 < x <0.9. The
results testify that the bulk characteristics directly correlate
with the nanoscale local features of the freshly cleaved
crystal surface. These convincing topographic studies with
atomic resolution are ensured by the high quality of the
local tunnel barriers manifested by the constancy of the
measured work function.
In spite of such clear observations of the surface atomic
arrangements of FeSe1-xTex with precise identification of
Te and Se atoms on the crystal surface, we observed by
STS no clear local superconducting gaps. The electronic
surface inhomogeneity might be a possible primary origin
Fig. 2. STM topography of FeSe1–xTex for 0.5 < x < 1. T = 4.9 K, I = 0.3 nA, V = –0.4 V.
T. Ekino, A. Sugimoto, and A.M. Gabovich
346 Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3
of the obscure superconducting gap structures. To confirm
such a nanoscale electronic irregularity we built STS con-
ductance mappings.
Figure 4 demonstrates the STM topography and the
mapping of the differential conductance dI/dV at zero bias
and V = 30 mV for FeSe0.3Te0.7 in the superconducting
state at 4.9 K. The patch-like patterns in the conductance
map of 5×5 nm roughly resemble those of STM topogra-
phy in the same region, which is due to Te/Se distributions
as indicated in Fig. 2. However, a closer look indicates
slight discrepancies between the contrast patterns of the
topography and conductance mapping. This can be ex-
plained by the manifestation of the local density of states
integrated over the energy revealed in the topography
while density of states itself determines the conductance
mapping. To clarify this point further, the line profiles of
the conductance are shown for the line cuts from three dif-
ferent locations. The conductance magnitude becomes re-
markably inhomogeneous when the bias voltage exceeds
+20–30 mV while it is fairly homogeneous for the negative
bias. This indicates the existence of impurity states in the
empty band. Since the energy scale of the inhomogeneity
starting at 20–30 meV is much higher than Tc ~ 10 K
(~ 0.9 meV) of FeSe0.3Te0.7, it does not seem to be directly
related to the superconducting properties. The Bardeen-
Cooper-Schrieffer (BCS) pairing energy is about 2 meV,
but no trace of a superconducting gap is seen in any of local
conductance profiles in the low bias regions below 10 mV.
It stems from the figure that there are no serious local na-
noscale inhomogeneities at zero bias which would be able to
smear superconducting-like spectra with their small energy
gaps. (This is in contrast to the pattern at V > 20–30 meV
where the substantial irregularity does exist.) The gap-
related features might have been washed out by surface
variations of the electronic properties emerging despite
low-T and UHV cleaving conditions. Nevertheless, it is
impossible now to indicate any sound reason of such varia-
tions. One might also consider the absence of gapping be-
ing due to bad resolution of the STS chamber. However,
measurements with the same STM/STS apparatus demon-
strated conspicuous atomic arrangements and the super-
conducting gap features in the other layered superconduc-
tor β-ZrNCl having a similar Tc = 13–15 K [28].
The existence of the superconducting gap could be
found by examination of the zero-bias conductance in the
superconducting state. Corresponding attempts were made
and the results obtained are shown in Fig. 4, in which the
dark-contrast region near zero-bias corresponding to the
depression of the electronic density of states might be due
to the manifestation of the superconductivity gap, because
such deep dark contrast becomes generally weak upon
warming. Anyway, there are no traces of the apparent gap
structure in the observed line profiles.
Hence, another effective method of BJ tunneling spec-
troscopy (BJTS) [18] was applied to reveal superconduct-
ing gaps in the materials concerned. In this technique, the
symmetric geometry of electrodes is realized so that the
overwhelming majority of junctions are of the SIS nature.
Since the crack section is perpendicular to the plate the
tunneling current passes mainly through the ab crystal plane.
Quite a number of iron-based superconductors are be-
lieved to manifest multiple or anisotropic gaps [29] as a
consequence of the multiple electronic band structure [10].
Fig. 3. Line profiles of the STM topography for FeSe0.3Te0.7 (a)
and FeSe0.1Te0.9 (b). Dark spot ratio from STM versus Se content
(1–x) from the electron-probe micro analyzer (EPMA) (c).
Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe1–xTex
Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 347
In particular, two-gap behavior was demonstrated in
FeSe1–x [30]. We sought for this phenomenon in our sam-
ples of FeSe1–xTex but succeeded in the observation of the
maximum-gap features only. This major gap, which is of
the order of Tc if the gaps are relatively weakly coupled, is
a parameter that at least approximately reflects the strength
of the pairing interaction [31].
Figure 5(a) shows three representative BJ tunneling con-
ductances measured at T = 4.2 K for FeSe0.5Te0.5, which
possesses the highest bulk Tc = 15 K among the series
FeSe1–xTex. The top curve shows a conspicuous depletion of
the electron density of states as well as inexpressive conduc-
tance peaks. The pattern can be associated with the presence
of the superconducting gap. The gap smearing correlates
with a large observed zero-bias leakage conductance as
compared to the standard BCS density of states.
At the same time, the middle and bottom curves in
Fig. 5(a) reveal more subtle gap-like features demonstrat-
ing hints of multigapness. The inner coherent peaks of the
double-gap structure in the middle curve correspond to the
gap peaks of the top curve positioned at ± 2–3 mV, while
the outer ones correlate with the rather strong peaks of the
bottom curve positioned at ± 5–6 mV. Since the latter bias
voltages are twice as much as the former ones, a plausible
conclusion may be made that the outer and inner peaks are
generated by tunneling through SIS (the bottom curve) and
SIN (the top curve, N stands for a normal metal) junctions,
respectively rather than by random scatter of multiple gap
values due to the sample inhomogeneity. Indeed, SIN junc-
tions are often formed instead of SIS counterparts even in
symmetric break junctions. This can be understood as the
consequence of the crack occurrence at a crystal defect or
grain boundary [32].
Fig. 4. The STM topography, the dI/dV mapping, and the dI/dV line profiles for FeSe0.3Te0.7 at 4.9 K.
T. Ekino, A. Sugimoto, and A.M. Gabovich
348 Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3
The above arguments were checked by the conductance
fitting using the broadened BCS density of states with the
account of thermal broadening. In the fitting, the well-
known approximation for the density of states, N(E, Γ) =
= |Re{(E – iΓ)/[(E – iΓ)2 – Δ2]1/2}|, was used, where Γ and
Δ are the phenomenological broadening and gap parame-
ters, respectively [33]. The fitted results are depicted as
dashed lines in Figs. 5(b) and (c). The overall features in-
cluding the broadened gap peaks and especially the sub-
stantial leakage conductance at zero bias are fairly well
reproduced. On the other hand, the calculated conductance
peaks are much shaper than the experimental data. The
best fitting parameters at T = 4.2 K are Δ = 1.6–1.8 meV
and Γ = 0.2–0.3 meV, respectively. As for the microscopic
origin of Γ, there exist several possible factors such as gap
anisotropy, inhomogeneity, etc., although there are no
quantitative estimations. On the other hand, relatively large
but rather common value obtained here implies the extrin-
sic interface influence. Such a large Γ value probably re-
flects the depression of the superconducting gap features in
STS. From the fitting results in Figs. 5(b) and (c), in which
the correspondence between experimental data and calcu-
lated curves is rather good in the zero bias region, the fea-
ture of sub-gap conductance seems to have nothing in
common with any non-conventional anisotropic pairing.
Figure 6 shows the T-dependence of the conductance in
the gap region corresponding to Fig. 5(b). The gap struc-
ture already broadened at 4.2 K is further broadened upon
warming and smeared out at a local Tc = 10 K in the junc-
tion. This is much lower than the bulk Tc = 15 K from the
resistivity measurements. The figure also indicates the SIN
junction formation in BJ as is indicated in Fig. 5(b), where
the gap structure is gradually smeared out but the gap-edge
peaks do not shift to zero at T → Tc, which is in contrast to
the behavior of SIS junctions [32]. The low-T gap value
2Δ = 3.2–3.6 meV and the junction local Tc = 10 K give
the characteristic gap to Tc ratio 2Δ/kBTc = 3.7–4.2. This is
slightly larger than the BCS value ≈ 3.5, indicating a mod-
erate or strong-coupling. Similar gap values were reported
in the earlier STS studies [24,25,27].
Fig. 6. Temperature evolution of dI/dV for FeSe0.5Te0.5. The
conductance is shifted up for the clarity.
Fig. 5. The dI/dV curves of FeSe0.5Te0.5 (a) at 4.2 K from break-
junction tunneling spectroscopy (BJTS). The dI/dV fittings using
the broadened BCS density of states (dotted curves) (b) and (c).
Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe1–xTex
Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 349
In the BJ measurements, substantial variation of both
gaps was found. Most probably, it is because BJTS is sen-
sitive to local changes in stoichiometry as described in our
previous measurements [34]. These BJ studies reveal an
upper limit of the superconducting gap value. The solid
curve in Fig. 7(a) shows the maximum gap size observed
in this series of measurements, where the peak-to-peak
separation of Vp–p ≈ 13–14 mV was obtained in the gap
structure with moderately pronounced peaks. This gap
may be attributed to the maximum Tc = 15 K in this com-
pound series. The representative outer gap structure in
Fig. 5(a) is also shown for the comparison. We can recog-
nize that the larger gap size possesses the larger conduc-
tance leakage (probably reflecting the low superconduct-
ing volume fraction). This is consistent with the fact that
we often obtained immature gap size and Tc, as shown in
Figs. 5 and 6. This set of thermodynamic parameters cor-
relates well with Vp–p ≈ 40 mV and Tc ≈ 48 K of the su-
perconducting NdFeAsO0.9F0.1 found in our BJTS mea-
surements [35], thereby supporting the idea of a common
mechanism of superconductivity among the iron-based
compounds [10,36].
In fact, one-unit-cell film of FeSe fabricated by molecu-
lar beam epitaxy was found to exhibit superconductivity
above 50 K and reveals the gap values of ≈ 20 mV and
≈ 40 mV in STS investigations [37,38]. Notwithstanding
the difference in Tc, the coherent peak positions in Ref. 37
correlate with our results of Vp–p ≈ 40 mV for the symme-
tric BJ junction. In Fig. 7(b), superconducting zero-T gap
versus Tc plots for several iron-based superconductors are
shown together with those of some copper oxides as well
as of previously investigated by us layered nitrochloride
superconductors [28,39]. All the measurements were done
by the fixed design of BJTS. The plots demonstrate a re-
markable difference in the slope reflecting the difference in
the coupling strength and, possibly, in the underlying me-
chanism of Cooper pairing. Namely, the slope is about the
s-wave BCS weak-coupling value in iron-based supercon-
ductors, while it is more than twice that in the copper-
oxide and nitrochloride superconductors. Therefore, it is
clear that cuprates and nitrochlorides possess unusually
large energy scale of the pairing interaction, the origin of
which remains to be clarified [39,40]. As for cuprates, it
might be a consequence of the charge-density-wave
(CDW) influence on superconductivity [41].
We have extended the bias voltage range to investigate
the background normal-state electronic states. Figure 8(a)
shows representative tunneling conductances obtained in
both STS (A and B) and BJTS (C and D) measurements.
The STS conductance is obtained by averaging 4096 spec-
tra in a 10×10 nm region, while the BJTS conductance is a
single trace from a junction. The revealed broadened gap-
like structure in STS spectra appears to be asymmetric with
respect to zero bias, but the bias polarity of the enhanced
peak in the asymmetric conductance depends on the mea-
surement run as shown in two characteristic curves of
Fig. 8(a). The asymmetric gap structure of curve A pos-
sesses a moderately enhanced peak appearing at the negative
bias of ~ –300 mV, while a subtle peak at the positive bias
≈ 100 mV becomes distinct only after normalizing by the
background value. Note that the bias polarity is defined with
respect to the sample. The high-energy gap structures found
in STS persist even at 77 K, showing the peaks at –300 mV
and +100mV, as displayed in Fig. 8(b).
The sub-gap conductance depends linearly on the bias
showing a very weak bend at –100–150 mV, which becomes
visible after the normalization. This bias is symmetric with
respect to the smaller peak occurring at +100 mV in the
normalized conductance, which is distinctly manifested in
the raw conductance data of B only in the positive bias.
Curves C and D in Fig. 8(a) describe the high-energy
structure of the BJTS conductance similar to that observed
by STS. The conductance is reasonably symmetric in ac-
cordance with the SIS nature of the junction. The well-
distinguished gap-edge structures are seen at ≈ ±300 mV
for C with small bends at ±100mV, accompanying an in-
Fig. 7. The maximum gap structure observed in the measure-
ments of FeSe1–xTex (a). Peak-to-peak value of the superconduct-
ing gap 2Δp–p versus Tc(K) for several superconductors
[35,38,39] (b).
FeSe Te1–x x
La Sr CuO1.85 0.15 4
Nd Ce CuO1.85 0.15 4
Bi Sr CuO2 2 6
�-HfNCl
�-ZrNCl
(a)
BJTS
= 4.2 KT
8
7
6
5
0.17
0.16
0.15
–10 –5 0 5 10
Voltage, mV
d
I/
d
V
,
m
S
15–15
(b)
NdFeAsO F0.9 0.1
40
30
20
10
0
10 20 30 40 50
Tc, K
2
,
m
eV
�
p
–
p
T. Ekino, A. Sugimoto, and A.M. Gabovich
350 Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3
tensive broad zero-bias peak due to a superconducting
weak-link behavior of the junction. The broadened double-
peak structure is seen in D at ±400 mV and ±150 mV, pos-
sessing the slightly larger energy than in C. These gap fea-
tures are in agreement with the STS spectra.
The lack of superconductivity manifestations in spec-
trum D is probably due to break of the sample in the non-
superconducting region or the gap averaging out by inho-
mogeneity. This explanation correlates with relatively large
conductance leakage. It might also happen that the absence
of superconductivity is due to the insufficient resolution of
the ac modulation voltage (> 1 mV) in this case.
The line profiles of the conductance mapping are shown
in Figs. 9(a) and (b) from the STS measurements at 4.9 K
within the length scale of ≈ 10 nm. The broadened peaks at
≈ – 300 mV (Fig. 9(a)) and +100 mV (Fig. 9(b)) are evi-
dent anywhere in the spatial region, which demonstrate
that the conductance below |V| = 400 mV is fairly homoge-
neous at least in this range with small accidental variations.
The broad peak structure at –300 mV is consistent with
the binding energy of 300 meV in the density of states ori-
ginating from iron d-state as observed in the valence-band
photoemission spectroscopy [42]. On the other hand, the
conductance peak positions which change with the bias
polarity (–300 mV and +100 mV) could be due to the
tunneling from different sections of the Fermi surface. In
fact, the angle-resolved photoemission spectroscopy along
the high-symmetry line of the Brillouin zone showed a
broad peak feature at ≈ 300 meV around the Γ point and a
nondispersive band at ≈ 100 meV around the M point, res-
pectivey [43]. The energies of asymmetric peaks with
strongly asymmetric conductance background observed in
STS could be explained by the difference in the tunneling
probability along the different directions in the Brillouin
zone. Since the negative bias corresponds to the electron
tunneling from the sample into the tip, the observed peak at
–300 mV might reflect the bottom of the electron band,
while the +100 mV peak reflects the empty states of the top
of the hole like band, both of which are located at the Γ
point. The conductance features like this exist in the broad
composition range, showing similar nanoscale homogeneity.
The remarkable segmental dependence of the conduc-
tance background occurring intensively in the particular
surface regions, that is, stronger positive or negative bias
dependences in Figs.9(a) or (b), respectively, could hide
the peak in the opposite bias. Such features could happen
when the shape of the potential barrier plays an important
role. In the vacuum tunneling using an STM tip, local en-
hancement of the electric field modifying the barrier yields
a strong bias dependence.
The density-wave formation in the layered compound
may be another plausible driving force of the conductance
asymmetry as predicted by the theory [44]. As for the gap
size observed here, it is typical to those of the charge den-
sity wave compounds [45,46]. Specifically, the asymmetry
in the gap structure could be a manifestation of the influ-
ence of the CDW order parameter phase. This is a possible
reason of the pseudogap (dip-hump) peak existence on one
bias polarity branch only [47]. The existence of the asymme-
tric gap feature in the STS conductance could be a smoking
gun of the CDW appearance in FeSe1–xTex, which might
exist in iron-based compounds along with spin density
waves (SDWs) [36]. At the same time, SDW can also pro-
mote asymmetry of the tunneling conductance [44].
Assuming the energy scale of the conventional CDW
compounds, the present gap might disappear at ≈ 100–250 K.
This is in accordance with Fig. 8(b), where the gap-like
structures are observed even at 77 K. In fact, the structural
phase transition near 90 K has been reported in FeSe [48],
and confirmed in our own resistivity measurements [49].
These characteristic features can be associated with high
bias gap-like structures observed here.
The high-energy gap like structure is reproduced in the
sulfur-substituted compound. Figure 10 shows the STS
conductance profile for Fe1.02S0.1Te0.9 where composi-
tions were determined by EPMA. Rather homogeneous
line profile in the spatial range of 10 nm is similar to that
in the Se-substituted compound. The pronounced peak at
Fig. 8. The dI/dV curves for FeSe0.5Te0.5 obtained from STS (A,
B) and BJTS (C, D) at T = 4–5 K (a), and STS at 77 K (b).
–1000 0 1000
Voltage, mV
1
0
B
JT
S
,
m
S
S
T
S
,
n
S
d
I/
d
V 2
0
FeSe Te0.5 0.5(a)
C (BJTS)
D (BJTS)
A (STS)
B (STS)
4.2 K (BJTS) – 4.9 K (STS)
–500 0 500
(b)
1
0
–500 0 500
STS
= 77 KT
d
I/
d
V
,
n
S
Voltage, mV
Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe1–xTex
Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 351
≈ –400 mV and a weak structure at +100–200 mV are
slightly larger than in the Se compound, but essentially in
the same energy range. The difference is probably due to
modifications of electronic states upon the change of the
ion sizes, since the S ion is 7–9% smaller than the Se one.
In order to study the high-energy feature in connection
with the magnetic phase transition at 72 K [20–23], the
STS measurements have been carried out for the non-
superconducting Fe1.01Te. The results showing the shallow
zero-bias depression of conductance and no apparent high-
energy gap structure are in contrast to Fig. 8, although both
conductance spectra display noticeable asymmetry with
respect to the bias voltage. This difference of the observed
electronic structures in a normal metal FeTe and a super-
conductor FeSe1–xTex is consistent with calculations [50].
4. Summary
We have synthesized and measured the surface and
electronic properties of Fe1.01Se1–xTex (0.5 < x < 1) single
crystals by means of STM/STS and BJ tunneling spectros-
copy. The STM topographies distinguish Se and Te atoms
in the whole x range, indicating that surface local events
directly reflect the bulk properties. The STS and break
junction measurements clarified the asymmetric gap-like
structure at ≈ ±100 mV and –300 mV. These features seem
to be consistent with the photoemission spectra, while the
distinct asymmetric gap-like feature can be explained by a
possible CDW formation. Such structures are also ob-
served in sulfur substituted Fe1.02S0.1Te0.9, but not in
Fe1.01Te. The superconducting gap structure is probed by
BJ tunneling. A smeared BCS density of states with
2Δ (4 K) ≈ 3.4 ± 0.2 meV for Tc ~ 10 K was observed. The
characteristic gap ratio 2Δ(0)/kBTc = 3.7–4.2 indicates an
intermediate superconducting coupling strength.
Acknowledgements
We thank the Natural Science Center for Basic Research
and Development, Hiroshima University for EPMA analysis
and supplying liquid helium. This work was supported by a
Fig. 9. Line profiles of dI/dV for FeSe0.5Te0.5 measured by STS at 4.9 K, showing the peak at –300 mV (a) and +100 mV (b). The pro-
files of (a) and (b) were taken from different samples.
(a) (b)
x,
n
m
T = 4.9 K T = 4.9 K
4
2
0
–2
–4
–600–400–200 0 200 400 600
4
2
0
–400 –200 0 200 400
4
2
0
2
1
0
–1
–2
x,
n
m
Voltage, mV Voltage, mV
d
I/
d
V
,
n
S
d
I/
d
V
,
n
S
Fig. 10. Line profile of dI/dV from STS for FeS0.1Te0.9 at 4.9 K.
T. Ekino, A. Sugimoto, and A.M. Gabovich
352 Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3
Grand-in-Aid for Scientific Research (245403770) of the
Ministry of Education, Culture, Sports, Science, and Tech-
nology (MEXT) of Japan. The work was partially supported
by the Project N 8 of the 2012-2014 Scientific Cooperation
Agreement between Poland and Ukraine.
1. Y. Kamihara, T. Watanabe, M. Hirono, and H. Hosono, J.
Am. Chem. Soc. 130, 3296 (2008).
2. X.C. Wang, Q.Q. Liu, Y.X. Lv, W.B. Gao, L.X. Yang, R.C. Yu,
F.Y. Li, and C.Q. Jin, Solid State Commun. 148, 538 (2008).
3. M. Rotter, M. Tegel, and D. Johrendt, Phys. Rev. Lett. 101,
107006 (2008).
4. F.C. Hsu, J.Y. Luo, K.W. The, T.K. Chen, T.W. Huang,
P.M. Wu, Y.C. Lee, Y.L. Huang, Y.Y. Chu, D.C. Yan, and
M.K. Wu, Proc. Natl. Acad. Sci. 105, 14262 (2008).
5. X.H. Chen, T. Wu, G. Wu, R.H. Liu, H. Chen, and D.F.
Fang, Nature 453, 761 (2008).
6. G.F. Chen, Z. Li, D. Wu, G. Li, W.Z. Hu, J. Dong, P. Zheng,
J.L. Luo, and N.L. Wang, Phys. Rev. Lett. 100, 247002
(2008).
7. H.H. Wen, G. Mu, L. Fang, H. Yang, and X. Zhu, Europhys.
Lett. 82, 17009 (2008).
8. Z.-A. Ren, J. Yang, W. Lu, X.-L. Shen, Z. Cai Li, G.-C. Che,
X.-L. Dong, L.-L. Sun, F. Zhou, and Z.-X. Zhao, Europhys.
Lett. 82, 57002 (2008).
9. K. Kuroki, R. Arita, and H. Aoki, Butsuri 64, 627 (2009) (in
Japanese).
10. A.A. Kordyuk, Fiz. Nizk. Temp. 38, 1119 (2012) [Low Temp.
Phys. 38, 888 (2012)].
11. S. Pandya, S. Sherif, L.S.S. Chandra, and V. Ganesan,
Supercond. Sci. Tech. 24, 045011 (2011).
12. Y. Mizuguchi, F. Tomioka, S. Tsuda, T. Yamaguchi, and Y.
Takano, J. Phys. Soc. Jpn. 78, 074712 (2009).
13. K. Deguchi, Y. Mizuguchi, Y. Kawasaki, T. Ozaki, S. Tsuda,
T. Yamaguchi, and Y. Takano, Supercond. Sci. Tech. 24,
055008 (2011).
14. S. Medvedev, T.M. McQueen, I.A. Troyan, T. Palasyuk,
M.I. Eremets, R.J. Cava, S. Naghavi, F. Casper, V. Kseno-
fontov, G. Wortmann, and C. Felser, Nature Mater. 8, 630
(2009).
15. R. Ukita, A. Sugimoto, and T. Ekino, Physica C 471, S622
(2011).
16. A. Sugimoto, T. Ekino, and H. Eisaki, J. Phys. Soc. Jpn. 77,
043705 (2008).
17. A. Sugimoto, K. Shohara, T. Ekino, Y. Watanabe, Y.
Harada, S. Mikusu, K. Tokiwa, and T. Watanabe, Physica C
469, 1020 (2009).
18. T. Ekino, T. Takabatake, H. Tanaka, and H. Fujii, Phys. Rev.
Lett. 75, 4262 (1995).
19. T. Ekino, A. Sugimoto, H. Okabe, K. Shohara, R. Ukita, J.
Akimitsu, and A.M. Gabovich, Physica C 470, 358 (2010).
20. B.C. Sales, A.S. Sefat, M.A. McGuire, R.Y. Jin, and D. Mand-
rus, Phys. Rev. B 79, 094521 (2009).
21. Y. Mizuguchi, F. Tomioka, S. Tsuda, T. Yamaguchi, and Y.
Takano. Appl. Phys. Lett. 94, 012503 (2009).
22. K.W. Yeh, T.W. Huang, Y.L. Huang, T.K. Chen, F.C. Hsu,
P.M. Wu, Y.C. Lee, Y.Y. Chu, C.L. Chen, J.Y. Luo, D.C.
Yan, and M.K. Wu, Europhys. Lett. 84, 37002 (2008).
23. W. Bao, Y. Qiu, Q. Huang, M.A. Green, P. Zajdel, M.R.
Fitzsimmons, M. Zhernenkov, S. Chang, M. Fang, B. Qian,
E.K. Vehstedt, J. Yang, H.M. Pham, L. Spinu, and Z.Q.
Mao, Phys. Rev. Lett. 102, 247001 (2009).
24. T. Kato, Y. Mizuguchi, H. Nakamura, T. Machida, H.
Sakata, and Y. Takano, Phys. Rev. B 80, 180507 (2009).
25. T. Hanaguri, S. Niitaka, K. Kuroki, and H. Takagi, Science
328, 5977 (2010).
26. F. Massee, S. de Jong, Y. Huang, J. Kaas, E. van Heumen,
J.B. Goedkoop, and M.S. Golden, Phys. Rev. B 80, 140507
(2009).
27. C-L. Song, Y-L. Wang, P. Cheng, Y-P. Jiang, W. Li, T.
Zhang, Z. Li, K. He, L. Wang, J-F. Jia, H-H. Hung, C. Wu,
X. Ma, X. Chen, and Q-K. Xue, Science 332, 1410 (2011).
28. A. Sugimoto, K. Shohara, T. Ekino, Z. Zheng, and S.
Yamanaka, Phys. Rev. B 85, 144517 (2012).
29. G. R. Stewart, Rev. Mod. Phys. 83, 1589 (2011).
30. R. Khasanov, M. Bendele, A. Amato, K. Coder, H. Keller,
H.-H. Klaus, H. Luetkens, and E. Pomjakushima, Phys. Rev.
Lett. 104, 087004 (2010).
31. V.Z. Kresin and S.A. Wolf, Physica C 169, 476 (1990).
32. T. Ekino, H. Fuji, T. Nakama, and R. Yagasaki, Phys. Rev. B
56, 7851 (1997).
33. R.C. Dynes, V. Narayanamurti, and J.P. Garno, Phys. Rev.
Lett. 41, 1509 (1978).
34. T. Ekino, H. Fujii, M. Kosugi, Y. Zenitani, and J. Akimitsu,
Phys. Rev. B 53, 5640 (1996).
35. T. Ekino, A. Sugimoto, H. Okabe, K. Shohara, R. Ukita, J.
Akimitsu, and A.M. Gabovich, Physica C 470, S358 (2010).
36. A. Chubukov, Annu. Rev. Condens. Matter Phys. 3, 57
(2012).
37. Q.-Y. Wang, Z. Li, W.-H. Zhang, Z.-C. Zhang, J.-S. Zhang,
W. Li, H. Ding, Y.-B. Ou, P. Deng, K. Chang, J. Wen, C.-L.
Song, K. He Ke, J.-F. Jia, S.-H. Ji, Y.-Y. Wang, L.-L. Wang,
X. Chen, X.-C. Ma, and Q.-K. Xue, Chin. Phys. Lett. 29,
037402 (2012).
38. S. He, J. He, W. Zhang, L. Zhao, D. Liu, X. Liu, D. Mou,
Y.-B. Ou, Q.-Y. Wang, Z. Li, L. Wang, Y. Peng, Y. Liu, C.
Chen, L. Yu, G. Liu, X. Dong, J. Zhang, C. Chen, Z. Xu, X.
Chen, X. Ma, Q. Xue, and X.J. Zhou, arXiv:1207.6823.
39. T. Takasaki, T. Ekino, A. Sugimoto, K. Shohara, S.
Yamanaka, and A.M. Gabovich, Euro. Phys. J. B 73, 471
(2010).
40. T. Ekino, T. Doukan, and H. Fujii, J. Low Temp. Phys. 105,
563 (1996).
41. T. Ekino, A.M. Gabovich, Mai Suan Li, M. Pękała, H.
Szymczak, and A.I. Voitenko, J. Phys. Condens. Matter 23,
385701 (2011).
42. R. Yoshida, T. Wakita, H. Okazaki, Y. Mizuguchi, S. Tsuda,
Y. Takano, H. Takeya, K. Hirata, T. Muro, M. Okawa, K.
Ishizaka, S. Shin, H. Harima, M. Hirai, Y. Muraoka, and T.
Yokota, J. Phys. Soc. Jpn. 78, 034708 (2009).
Scanning-tunneling microscopy/spectroscopy and break-junction tunneling spectroscopy of FeSe1–xTex
Low Temperature Physics/Fizika Nizkikh Temperatur, 2013, v. 39, No. 3 353
43. K. Nakayama, T. Sato, P. Richard, T. Kawahara, Y. Sekiba,
T. Qian, G.F. Chen, J.L. Luo, N.L. Wang, H. Ding, and T.
Takahashi, Phys. Rev. Lett. 105, 197001 (2010).
44. A.M. Gabovich and A.I. Voitenko, Phys. Rev. B 56, 7785
(1997).
45. G. Gruner, Density Wave in Solids, Addison-Wesley,
Reading, Massachusetts (1994).
46. A.M. Gabovich, A.I. Voitenko, and M. Ausloos, Phys. Rep.
367, 583 (2002).
47. A.M. Gabovich and A.I. Voitenko, Phys. Rev. B 75, 064516
(2007).
48. T.M. McQueen, Q. Huang, V. Ksenofontov, C. Felser, Q.
Xu, H. Zandbergen, Y.S. Hor, J. Allred, A.J. Williams, D.
Qu, J. Checkelsky, N.P. Ong, and R.J. Cava, Phys. Rev. B
79, 014522 (2009).
49. R. Ukita, A. Sugimoto, and T. Ekino, Physica C 471, 622
(2011).
50. A. Subedi, L. Zhang, D.J. Singh, and M.H. Du, Phys. Rev. B
78, 134514 (2008).
|