Evidence for chiral superconductivity on a silicon surface F. Ming1X. Wu2 3C. Chen2K. D. Wang2P. Mai4T. A. Maier4J. Strockoz5 6 J. W. F. Venderbos5 6C. Gonzalez7 8J. Ortega9S. Johnston10 11and H. H. Weitering10 11

2025-04-26 0 0 8.2MB 25 页 10玖币
侵权投诉
Evidence for chiral superconductivity on a silicon surface
F. Ming,1X. Wu,2, 3 C. Chen,2K. D. Wang,2P. Mai,4T. A. Maier,4J. Strockoz,5, 6
J. W. F. Venderbos,5, 6 C. Gonzalez,7, 8 J. Ortega,9S. Johnston,10, 11 and H. H. Weitering10, 11
1State Key Laboratory of Optoelectronic Materials and Technologies, School of Electronics and Information Technology
and Guangdong Province Key Laboratory of Display Material, Sun Yat-sen University, Guangzhou 510275, China
2Department of Physics, Southern University of Science and Technology,
Shenzhen, Guangdong 518055, China
3School of Physical Sciences, Great Bay University, Dongguan, Guangdong 523000, China
4Computational Sciences and Engineering Division, Oak Ridge National Laboratory, Oak Ridge, TN, 37831-6494, USA
5Department of Physics, Drexel University, Philadelphia, PA 19104, USA
6Department of Materials Science and Engineering, Drexel University, Philadelphia, PA 19104, USA
7Departamento de Física de Materiales, Universidad Complutense de Madrid, 28040 Madrid, Spain
8Instituto de Magnetismo Aplicado UCM-ADIF, Vía de Servicio A-6,
900, E-28232 Las Rozas de Madrid, Spain
9Departamento de Física Teórica de la Materia Condensada and Condensed Matter Physics Center (IFIMAC),
Facultad de Ciencias, Universidad Autónoma de Madrid, 28049 Madrid, Spain
10Department of Physics and Astronomy,
The University of Tennessee, Knoxville, TN 37996, USA
11Institute of Advanced Materials and Manufacturing, The University of Tennessee, Knoxville, TN 37996, USA
Sn adatoms on a Si(111) substrate with 1/3 monolayer coverage form a two-dimensional trian-
gular adatom lattice with one unpaired electron per site and an antiferromagnetic Mott insulating
state. The Sn layers can be modulation hole-doped and metallized using heavily-doped p-type
Si(111) substrates, and become superconducting at low temperatures. While the pairing symme-
try of the superconducting state is currently unknown, the combination of repulsive interactions
and frustration inherent to the triangular adatom lattice opens up the possibility for a chiral or-
der parameter. Here, we study the superconducting state of Sn/Si(111) using scanning tunneling
microscopy/spectroscopy and quasi-particle interference imaging. We find evidence for a doping-
dependent Tcwith a fully gapped order parameter, the presence of time-reversal symmetry breaking,
and a strong enhancement of the zero-bias conductance near the edges of the superconducting do-
mains. While each individual piece of evidence could have a more mundane interpretation, our
combined results suggest the tantalizing possibility that Sn/Si(111) is an unconventional chiral d-
wave superconductor.
Superconductivity – dissipationless electrical conduc-
tivity in conjunction with perfect diamagnetism – is a
profound manifestation of a macroscopic quantum phe-
nomenon. Microscopically, supercurrents are carried by
Cooper pairs whose pair wave functions become phase
locked as they condense, like bosons, into a coherent
macroscopic quantum state [1]. In conventional su-
perconductors, electron pairing is mediated by virtual
phonon exchange. In this case, the relatively slow motion
of the ions provides a time-retarded effective attractive
interaction that allows the electrons to overcome their
mutual repulsion resulting in Cooper pairs with s-wave
symmetry, where the composite spin and orbital angu-
lar momenta of the electrons are zero. Higher angular
momentum states are typically driven by repulsive in-
teractions [2,3] as is the case for e.g. high-Tccuprate
superconductors [3,4]. Here, electron repulsion is min-
imized by imposing a nodal structure with correspond-
ing sign change in the superconducting wave function.
More recent emphasis on topological materials systems
have raised expectations for the discovery of novel multi-
component order parameters that are topologically dis-
tinct from those of ordinary Cooper pair condensates [5
15]. Besides the microscopic nature of the pairing inter-
actions, the physics of these systems is dictated by bro-
ken symmetries such as crystal, spin rotation, and time-
reversal symmetries, though experimental validation of
intrinsically topological order parameters remains scant.
Superconductivity has recently been discovered in a
system comprised of one-third monolayer of Sn deposited
on degenerately doped p-type Si(111) substrates [17]. Its
pairing symmetry, however, remains undetermined. This
system is of particular interest because the undoped Sn
monolayer is an antiferromagnetic single-band Mott in-
sulator [16,18] that becomes superconducting upon hole
doping, drawing interesting comparisons with the high-
Tccuprates [3,19] with d-wave order parameters. The Sn
layer, however, has triangular lattice symmetry imposed
by the Si(111) substrate. This geometry naturally allows
for the existence of a chiral order parameter with topolog-
ical edge states [7,12,20], if repulsive interactions domi-
nate the pairing. The appearance of such an exotic order
parameter is expected to furthermore depend on the elec-
tron correlation strength, shape of the Fermi surface, and
the doping level [7,20,21]. In particular, recent renor-
malization group calculations for the Sn/Si(111) system
indicated a competition between chiral d- and f-wave and
triplet p-wave instabilities, depending on the doping level
arXiv:2210.06273v1 [cond-mat.supr-con] 12 Oct 2022
2
FIG. 1. Structure, topography and spectral properties of the superconducting (3×3)-Sn surface on Si(111).
a, Structure model with the three outermost surface layers and one Sn adatom per (3×3) unit cell. b, Noninteracting
dispersion of the dangling-bond surface state according to Ref. 16. The inset shows the Fermi surface in the hexagonal surface
Brillouin zone. The high symmetry points are indicated. The Γpoint is located at the center of the hexagon. c, Comparison of
differential conductance spectra of two Si(111)(3×3)-Sn surfaces. The spectrum for the undoped surface is from Ref. 17,
acquired at 77 K, showing the upper and lower Hubbard bands (UHB/LHB) and a small gap around the Fermi level; the other
spectrum is from a doped surface (p= 0.08) obtained at 0.5 K, showing an extra quasiparticle peak (QPP) near the Fermi level,
consistent with Ref. 17.d, Topographic STM image (Vs=0.1V, It= 0.1nA) showing a near perfect Sn adatom lattice, along
with a substitutional Si defect, vacancy, and other defect, labelled with a triangle, square and a circle, respectively (p= 0.08).
e, Normalized STS spectra for three different hole concentrations, revealing a clear doping dependence of the superconducting
gap. f, A set of raw dI/dV spectra taken at equidistant locations along the dotted line shown in panel d, starting on the left.
(p= 0.08). g, Normalized dI/dV spectra as a function of temperature (p= 0.08). The spectra in panels fand gare offset
vertically for clarity.
and value of the nearest-neighbor Hubbard repulsion [21].
At the same time, electron-phonon interactions, partic-
ularly to interfacial Si modes [22], could drive a conven-
tional s-wave pairing [17].
Here we study the superconducting state of the
Sn/Si(111) interface using scanning tunneling microscopy
and spectroscopy (STM/STS) and quasiparticle interfer-
ence (QPI) imaging. Our observations reveal a strong
doping dependence of the superconducting Tc, a fully
gapped order parameter, the presence of time-reversal
symmetry breaking, and a strong enhancement of the
zero-bias conductance near the edges of the supercon-
ducting domains. While each of these observations may
have a mundane explanation, we discuss why we believe
that a chiral d-wave scenario offers the most consistent
interpretation of the measurements and theoretical mod-
eling. Final confirmation, however, awaits experimental
validation concerning the topological nature of the edge-
state conductance.
At 1/3 monolayer coverage, the Sn adatoms form
a (3×3) superlattice on the Si(111) surface with
one half-filled dangling-bond orbital per site and a Sn-
Sn distance of 6.65 Å; see Fig.1a. All other chemical
bonds in the system are passivated. The non-interacting
dangling-bond surface state has a bandwidth W0.5eV
(Fig. 1b), which is comparable to the on-site Hubbard in-
teraction U0.66 eV of the dangling bond orbitals [16].
As such, the system is a Mott insulator with an upper
and a lower Hubbard band (UHB/LHB) straddling the
Fermi level (Fig. 1c).
Figure 1dshows an STM image of the triangular Sn
adatom lattice. The Sn atoms are clearly resolved and
well ordered. The dark point defects correspond to sub-
stitutional Si adatoms (most prevalent) and Sn adatom
vacancies. Holes are introduced via modulation doping,
using boron-doped Si substrates with different doping
levels [17,18]. (For a discussion on dopant segregation,
see Supplementary Note 1 and Supplementary Figure 1).
The hole concentration in the dangling-bond surface state
is estimated from the spectral weight transfer in the tun-
neling spectra, associated with the introduction of holes
and formation of a quasiparticle peak in the Mott gap
(see Fig. 1band Extended Data Fig. 1) [18].
Fig. 1eshows the normalized dI/dV tunneling spec-
tra for excess hole concentrations of p= 0.06,0.08, and
0.10, recorded at T= 0.5K. (The p= 0.10 data were
reported in Ref. 17.) These spectra are representative of
the superconducting density of states (DOS). Here, we
3
FIG. 2. Low-temperature differential conductance and
QPI spectra of the p= 0.08 (3×3)-Sn surface on
Si(111).a, Best fits of the low-energy normalized dI/dV
spectra at T= 0.5K, for different pairing symmetries. The
inset zooms in on the zero bias region. b,c, QPI images ac-
quired at Vs=±5mV, beyond the superconducting gap. d,
Real space conductance map g(r, V ), obtained at zero bias.
The bright six-leaf features are scattering features from sur-
face defects. e, Corresponding QPI spectrum g(q, V )ob-
tained from the conductance map in panel dvia Fourier trans-
formation. The six dark blue crosses in panels b,c, and ein-
dicate the Bragg peaks, while the colored contours highlight
characteristic features in each QPI image.
divided the raw tunneling spectra by the normal state
dI/dV spectrum obtained in a perpendicular 15 Tesla
magnetic field. This field is large enough to completely
suppress the superconductivity for the p= 0.08 and 0.10
samples, which have upper critical field values of Hc2(0.5)
of 3 Tesla [17] and 13 Tesla, respectively (see Extended
Data Fig. 2). This procedure is a bit problematic for
the p= 0.06 sample, where the upper critical field ex-
ceeds the magnetic-field capability of our instrument (15
Tesla).
The Sn adatom lattice in Figure 1ais highly ordered,
while the boron dopants are located in the silicon bulk.
The superconducting DOS is, therefore, spatially uniform
away from localized point defects and the edges of the
(3×3) domains. Fig. 1fshows a series of spectra
recorded at T=0.5K along the line segment in Fig. 1d.
This level of homogeneity distinguishes the Sn on Si(111)
system from e.g. complex oxides, which exhibit consider-
able electronic inhomogeneity, often in conjunction with
various competing orders [23,24].
The normalized dI/dV spectra of the p= 0.08 sam-
ple are plotted as a function of temperature in Fig. 1g.
The gap feature persists up to about 8K. Detailed Dynes
fits [25] of the spectra assuming s-wave and dx2y2±idxy
order parameter symmetries, as well as zero bias con-
ductance measurements as a function of temperature,
consistently produce a Tcof about 7.6±0.2K with
some evidence of superconducting fluctuations above Tc
(see Ref. 17, Extended Data Fig. 3, and Fig. 2a). A
similar procedure for the p= 0.10 sample produces Tc
= 4.7±0.3K [17], while the Tcof the p= 0.06 sample
was difficult to ascertain because the spectra cannot be
properly normalized [Hc2(0.5 K) >15 T]. We conserva-
tively estimate its Tcto be around 9K (see Extended
Data Fig. 2).
Fitting the p= 0.08 dI/dV spectra with an s-wave
gap produces a reasonable fit but with notable discrep-
ancies near zero-bias. Turning to potential chiral order
parameters, we find that a chiral d-wave fit also agrees
well with the data and even improves the fit at low en-
ergies. A chiral p-wave gap clearly fails to describe the
spectra, particularly at low-energy (Fig. 2a). This fail-
ure occurs because any p-wave gap function must vanish
at the M-point by symmetry. This point corresponds
to the van Hove singularity and lays close to the Fermi
surface [17,18]. It therefore affects the gap significantly,
producing pronounced shoulders in the DOS that are not
observed experimentally. Other parameter symmetries
such as extended chiral p-wave and nematic d-wave sym-
metries (i.e., dx2y2and dxy ) do not fit the spectra either,
see Supplementary Note 2. (A multigap order parameter
can also be ruled out since this is a single-band system.)
Only s- and chiral d-wave symmetries produce good re-
sults and it is not possible to conclusively discriminate
between the two based on fitting alone.
Important details about the Fermi surface and or-
der parameter symmetry can be obtained from spectro-
scopic STM imaging [26]. Here, one acquires a spatial
map of the differential tunneling conductance g(r, V ) =
dI(r, V )/dV . Such dI/dV maps typically reveal the
presence of electronic standing waves as quasi-particles
are scattered elastically by defects on the surface. The
power spectrum of the differential conductance map –
the QPI spectrum – then identifies the dominant scatter-
ing processes contributing to the standing wave pattern.
In itinerant systems, these typically correspond to scat-
tering wavevectors connecting different k-points on the
constant energy contours (corresponding to the imaging
bias) q= 2k±G, where Gis a reciprocal lattice vector
of the (3×3) adatom lattice.
Figs. 2b,cshow the T= 0.5K QPI spectra taken
at ±5meV bias (p= 0.08). Both spectra reveal the
4
FIG. 3. Comparison of the measured QPI spectra with
theory.a,b, Experimental QPI images g(q, V )obtained at
zero bias on the p= 0.10 surface above (panel a) and below
Tc(panel b). c,d, Simulated QPI images for a superconduc-
tor with a chiral d-wave (panel c) and s-wave (panel d) order
parameter, assuming non-magnetic defects. In each image,
the six dark blue crosses indicate the locations of the Bragg
points, while the colored contours highlight characteristic fea-
tures in each QPI spectrum.
warped hexagonal Fermi contour of the normal state
(G=0), highlighted in magenta, along with several
scattering replica’s (G6=0) as indicated by the light
blue dumbbell-shaped contours [18]. These spectra agree
very well with previous calculations for the spectra in
the normal state, and are fully consistent with the band
structure for the Sn surface state [18]. (The presence of
the quasiparticle band and its dispersion is inconsistent
with an interpretation in terms of impurity band physics;
see Supplementary Note 3.) Real space differential con-
ductance maps at zero bias (Fig. 2d), i.e. deep inside the
superconducting gap, reveal the existence of very strong
star-like scattering features centered at the various sur-
face defects. The corresponding Fourier map (Fig. 2e)
now reveals the presence of a flower-like feature centered
at q=0and with six “petals” pointing towards the Bragg
points of the (3×3) lattice, as outlined by the red con-
tour. Meanwhile, the Fermi contour seen at ±5meV is
suppressed. This flower feature appears to be intimately
related to the superconductivity; it only exists when the
sample is in the superconducting state (Fig. 3a,b) and
when the tunneling bias is within the superconducting
gap (Fig. 2b,c,e). Hence, they are unique features of
the superconducting state (see Extended Data Fig. 4for
additional QPI results).
To elucidate the origin of the flower-like features, we
first simulated the QPI patterns for the s-wave and chi-
ral d-wave order parameters using the T-matrix formal-
ism and assuming nonmagnetic scattering (see Meth-
ods). Fig. 3shows the experimental QPI spectra of the
p= 0.10 sample, along with the simulated spectra for
the dx2y2±idxy and s-wave state. The experimental
pattern in Fig. 3bis well reproduced in the calculations
for the d-wave pairing channel (Fig. 3c). (The theoret-
ical features exhibit much more curvature because the
calculations are based on the non-interacting band dis-
persion whereas the experimental band dispersion has
correlation-driven band renormalizations [18].) Impor-
tantly, the flower is absent for the s-wave pairing channel,
as shown in Fig. 3d.
Our simulations reveal that the flower features only
appear when time reversal symmetry is broken. Such
would be the case for non-magnetic scattering in a chiral
superconductor, as simulated above, but it could also be
due to magnetic scattering in an s-wave superconductor
(see Extended Data Fig. 5). In particular, the star-like
scattering features in the real-space QPI maps are very
similar to those observed for magnetic point scatterers
in s-wave systems, and have been attributed to a focus-
ing effect of magnetic bound states or Yu–Shiba–Rusinov
(YSR) states due to Fermi surface anisotropy [2731]. To
discriminate between the s-wave and d-wave scenarios, it
is essential to establish the nature of the defects on the
surface.
The most prevalent scattering defect on the surface is
the substitutional Si adatom (replacing a Sn adatom). It
shows up as a dark void in filled-state STM images and
as a depressed adatom in the empty state images (see Ex-
tended Data Fig. 6). This observation indicates that the
spz-like dangling bond orbital of the Si atom is empty,
and because the adatom forms three covalent backbonds
with the Si substrate, the Si adatom is expected to be
non-magnetic. This is confirmed via first-principles Den-
sity Functional Theory (DFT) total-energy calculations
(which show that the Si adatoms are placed 0.6 Å be-
low the Sn adatoms) and by STM image simulations (see
Methods and Extended Data Fig. 6). In addition, spin-
polarized DFT calculations (see Methods) confirm the
nonmagnetic nature of this defect.
It is not possible to ascertain the nature of all native
defects on the surface (see Extended Data Fig. 7a) and
thus rule out any magnetic scattering contribution to the
QPI pattern. We therefore created a new type of defect
by depositing a tiny excess amount of (nonmagnetic) Sn
atoms at 120 K. STM images indicate that additional
Sn adatoms are located at three-fold symmetric inter-
stitial adatom sites, surrounded by three Sn adatoms of
the 2D host lattice (see Extended Data Fig. 6). The
excess Sn atoms easily move under the STM tip at tun-
neling biases in excess of ±0.8V, indicating that they
are weakly bound to the surface. The interstitial adatom
5
FIG. 4. Defect states and edge states on the (3×3)-Sn surface (p=0.08).a, STS point spectra (0.5 K) for each
Sn atom along the blue-dotted line in the topographic image (Vs= 0.5 V, It= 1 nA) at the top with a substitutional Si
defect in the middle. The bottom spectrum corresponds to the left end of the line. The spectrum recorded right on top of the
defect is indicated by a triangle. Spectra near the defect site exhibit two gap states at EB=±0.6meV. b, A (3×3)-Sn
superconducting domain (Vs= 0.8V, It= 0.1nA) next to a semiconducting Si(111)(23×23)R30-Sn domain on the far
left side of the image (bright strip) . c-e, Registry-aligned real-space conductance maps g(r, V = 0 mV) of the (3×3)-Sn
domain measured at different temperatures. In panels b-e, the vertical dashed lines label the domain boundary; the dashed
circles label the locations of the same defect. f, Averaged zero-bias conductance as a function of distance from the domain
boundary. Each line is obtained from a conductance map g(r, V = 0 mV), recorded at the indicated temperature. The first six
curves are fitted with an exponential decay. g, Fitted decay lengths as a function of temperature. The dashed line is a guide
to the eye. h, STS spectra taken at 0.5 K along the dotted line in panel b, starting at the domain boundary on the left. i,
The 15 bottommost spectra from panel hafter subtracting the dI/dV spectrum recorded deep inside the (3×3) domain.
These spectra highlight the edge state contribution (indicated by shading) to the measured dI/dV spectra. j, Simulated DOS
of a chiral d+ idsuperconductor, approaching the open edge of a cylinder (see Methods). As with panel i, a spectrum from the
center of the cylinder is subtracted to highlight the contribution from the edge state to the total DOS inside the superconducting
gap. Spectra in panel h,iand jare shifted vertically for clarity.
location is validated by DFT total energy minimization
and STM image simulations. In particular, notice the
excellent agreement between the experimental and theo-
retical STM images, as shown in Extended Data Fig. 6.
This level of agreement gives us confidence that the cal-
culations capture the local electronic structure correctly.
Importantly, spin-polarized DFT indicates that these de-
fect centers are also nonmagnetic. The latter can be un-
derstood from the fact that the interstitial Sn atom and
its three nearest neighbors have negligible contribution
to the DOS at the Fermi level (Extended Data Fig. 6),
thus pre-empting a potential magnetic instability.
Figure 4aplots the STS spectra across a substitutional
Si defect, which reveals the existence of two in-gap states
(p= 0.08). These states appear in pairs located symmet-
rically about the Fermi level reminiscent of YSR bound
states though the substitutional defect is nonmagnetic.
All defects on the surface we have checked produce these
YSR-like in gap states albeit at different energies, in-
cluding the native vacancy defects, interstitial Si, excess
Sn adatoms, as well as various other defects (see Ex-
tended Data Fig. 7). This is to be expected in a chiral
d-wave scenario because potential and magnetic defects
are both pair breaking [32,33]. In the s-wave scenario,
on the other hand, one would have to assume that these
defects are all magnetic [34], which seems unlikely (see
Extended Data Fig. 5). Interestingly, the nonmagnetic
interstitial Sn adatoms produce the strongest star-like
scattering features in real-space QPI maps and the most
intense flower features in the corresponding power spec-
trum (see Extended Data Fig. 7e-g). These enhanced
scattering features near the excess Sn adatoms suggest
that the s-wave scenario can be dismissed because time-
reversal symmetry should not be broken in such a case.
Alternatively, one might suggest that nonmagnetic impu-
rities are present in a magnetically ordered background,
摘要:

EvidenceforchiralsuperconductivityonasiliconsurfaceF.Ming,1X.Wu,2,3C.Chen,2K.D.Wang,2P.Mai,4T.A.Maier,4J.Strockoz,5,6J.W.F.Venderbos,5,6C.Gonzalez,7,8J.Ortega,9S.Johnston,10,11andH.H.Weitering10,111StateKeyLaboratoryofOptoelectronicMaterialsandTechnologies,SchoolofElectronicsandInformationTechnology...

展开>> 收起<<
Evidence for chiral superconductivity on a silicon surface F. Ming1X. Wu2 3C. Chen2K. D. Wang2P. Mai4T. A. Maier4J. Strockoz5 6 J. W. F. Venderbos5 6C. Gonzalez7 8J. Ortega9S. Johnston10 11and H. H. Weitering10 11.pdf

共25页,预览5页

还剩页未读, 继续阅读

声明:本站为文档C2C交易模式,即用户上传的文档直接被用户下载,本站只是中间服务平台,本站所有文档下载所得的收益归上传人(含作者)所有。玖贝云文库仅提供信息存储空间,仅对用户上传内容的表现方式做保护处理,对上载内容本身不做任何修改或编辑。若文档所含内容侵犯了您的版权或隐私,请立即通知玖贝云文库,我们立即给予删除!
分类:图书资源 价格:10玖币 属性:25 页 大小:8.2MB 格式:PDF 时间:2025-04-26

开通VIP享超值会员特权

  • 多端同步记录
  • 高速下载文档
  • 免费文档工具
  • 分享文档赚钱
  • 每日登录抽奖
  • 优质衍生服务
/ 25
客服
关注