ArticlePDF Available

Spin-polarized Majorana zero modes in proximitized superconducting penta-silicene nanoribbons

Authors:

Abstract and Figures

We theoretically propose penta-silicene nanoribbons (p-SiNRs) with induced p-wave superconductivity as a platform for the emergence of spin-polarized Majorana zero-modes (MZMs). The model explicitly considers the key ingredients of well-known Majorana hybrid nanowire setups: Rashba spin-orbit coupling, magnetic field perpendicular to the nanoribbon plane, and first nearest neighbor hopping with p-wave superconducting pairing. The energy spectrum of the system, as a function of chemical potential, reveals the existence of MZMs with a well-defined spin orientation localized at the opposite ends of both the top and bottom chains of the p-SiNR, associated with well-localized and nonoverlapping wave function profiles. Well-established experimental techniques enable the fabrication of highly ordered p-SiNRs, complemented by a thin lead film on top, responsible for inducing p-wave superconductivity through proximity effect. Moreover, the emergence of MZMs with explicit opposite spin orientations for some set of model parameters opens a new avenue for exploring quantum computing operations, which accounts for both MZMs and spin properties, as well as for new MZMs probe devices based on spin-polarized electronic transport mechanisms.
Spinful case—magnetic field up: (a–e) Bulk energy dispersion of the superconducting p-SiNRs for the spinful situation, as a function of kx\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$k_{x}$$\end{document}, for μ=-2.7t\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\mu =-2.7t$$\end{document}, -2.35t\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$-2.35t$$\end{document}, 1.1t, 2.09t and 2.2t, respectively. (f) Energy spectrum as a function of the chemical potential. Vertical lines indicate the chosen values of chemical potential shown on top panels. (g)–(k) Zero-energy states spectrum. (l)–(q) Probability density per lattice site |ψ|2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$|\psi |^2$$\end{document}, associated with zero-energy states on the real axis of the spin-polarized Kitaev, top or bottom chains.
… 
This content is subject to copyright. Terms and conditions apply.
1
Vol.:(0123456789)
Scientic Reports | (2023) 13:17965 | https://doi.org/10.1038/s41598-023-44739-7
www.nature.com/scientificreports
Spin‑polarized Majorana
zero modes in proximitized
superconducting penta‑silicene
nanoribbons
R. C. Bento Ribeiro
1, J. H. Correa
2,3, L. S. Ricco
4, I. A. Shelykh
4,5, Mucio A. Continentino
1,
A. C. Seridonio
6, M. Minissale
7, G. Le Lay
7 & M. S. Figueira
8*
We theoretically propose penta‑silicene nanoribbons (p‑SiNRs) with induced p‑wave superconductivity
as a platform for the emergence of spin‑polarized Majorana zero‑modes (MZMs). The model
explicitly considers the key ingredients of well‑known Majorana hybrid nanowire setups: Rashba
spin‑orbit coupling, magnetic eld perpendicular to the nanoribbon plane, and rst nearest neighbor
hopping with p‑wave superconducting pairing. The energy spectrum of the system, as a function of
chemical potential, reveals the existence of MZMs with a well‑dened spin orientation localized at
the opposite ends of both the top and bottom chains of the p‑SiNR, associated with well‑localized
and nonoverlapping wave function proles. Well‑established experimental techniques enable the
fabrication of highly ordered p‑SiNRs, complemented by a thin lead lm on top, responsible for
inducing p‑wave superconductivity through proximity eect. Moreover, the emergence of MZMs with
explicit opposite spin orientations for some set of model parameters opens a new avenue for exploring
quantum computing operations, which accounts for both MZMs and spin properties, as well as for new
MZMs probe devices based on spin‑polarized electronic transport mechanisms.
Ultra-scaling of nanoelectronic devices, beyond Moores law, still using the ubiquitous silicon technology, could
come from silicene13, the rst silicon-based graphene-like articial two-dimensional (2D) quantum material,
which further engendered the Xenes family4, and which was used to fabricate an atom-thin channel in a eld
eect transistor5, 6. Moreover, topological silicon nanowires hosting Majorana fermions could be a materials
platform for a quantum computer7. However, like other nanowire candidates, even proximitized ones based
on heavier constituents with larger spin-orbit coupling, until now, no conclusive experimental measurements
guarantee incontrovertibly the existence of topologically protected Majorana zero modes (MZMs) for the pos-
sible realization of qubits8, 9.
Since the appearance of the generic Kitaev model10, several platforms were proposed to realize it, both from
theoretical1117, and experimental points of view1824. A helpful review of the experimental state-of-the-art on
this subject can be found in Refs.9, 25, 26. is model considers p-wave superconductor pairing between electrons
in dierent sites of a one-dimensional chain (Kitaev chain) and predicts the existence of unpaired MZMs at
opposite ends of a nite Kitaev chain. However, until now, there are no conclusive experimental measurements
that guarantee without doubt the existence of topologically protected MZMs2630. e experimental detection
of MZMs remains an elusive problem, and they were not really observed until now. Per se, this situation justies
the search for new platforms.
One possible alternative platform is the one-dimensional honeycomb nanoribbons (HNRs) that have been
receiving growing attention in the literature3134. Nevertheless, the mono-elemental 2D graphene-like materials
coined Xenes, where X represents elements from group IIIA to group VIA of the periodic table, could constitute
OPEN
1Centro Brasileiro de Pesquisas Físicas, Rua Dr. Xavier Sigaud, 150, Urca, Rio de Janeiro, RJ 22290-180,
Brazil. 2Universidad Tecnológica del Perú, Nathalio Sánchez, 125, 15046 Lima, Peru. 3AGH University of Krakow,
Academic Centre for Materials and Nanotechnology, al. A. Mickiewicza 30, 30-059 Kraków, Poland. 4Science
Institute, University of Iceland, Dunhagi-3, 107 Reykjavik, Iceland. 5Russian Quantum Center, Skolkovo IC, Bolshoy
Bulvar 30 bld. 1, Moscow 121205, Russia. 6School of Engineering, Department of Physics and Chemistry, São Paulo
State University (UNESP), Ilha Solteira, SP 15385-000, Brazil. 7Aix-Marseille Université, CNRS, PIIM UMR 7345,
13397 Marseille Cedex, France. 8Instituto de Física, Universidade Federal Fluminense, Av. Litorânea s/N, Niterói,
RJ CEP: 24210-340, Brazil. *email: gueira7255@gmail.com
Content courtesy of Springer Nature, terms of use apply. Rights reserved
2
Vol:.(1234567890)
Scientic Reports | (2023) 13:17965 | https://doi.org/10.1038/s41598-023-44739-7
www.nature.com/scientificreports/
possible candidates to build HNRs with the ability to harbor MZMs at their ends3538. Penta-Silicene (X=Si) is
an up-and-coming candidate in this family for obtaining a p-SiNR geometry that can host MZMs3941.
A paradigmatic breakthrough would be the experimental implementation of the generic Kitaev toy model
with a silicon platform10. In a previous work34, we addressed the problem of Majorana spin discrimination
employing a double-spin Kitaev zigzag honeycomb nanoribbons (KzHNR), which mimics two parallel Kitaev
chains connected by the hopping t (see gure1 of34). Since such KzHNRs have not been realized in experiments,
we look instead in the present paper at the possibility of obtaining MZMs in p-SiNRs, harboring Dirac fermions,
which have been epitaxially grown on Ag(110) surfaces39, 4244. Typically, highly perfect, atom thin, massively
aligned single strand p-SiNRs, 0.8nm in width, and with lengths extending to tens of nanometers were obtained
by molecular beam epitaxy upon insitu Si deposition onto Ag(110) surfaces held at room temperature, as shown
in Fig.1a. In scanning tunneling microscopy (STM) and high-resolution nc-AFM images, these p-SiNRs appear
as two shied lines of protrusions along the [110] direction as shown in Fig.1b,c and are separated by twice the
nearest neighbor Ag-Ag distance, i.e., 0.577nm. eir hidden internal atomic structure was initially uncovered
employing thorough density functional theory (DFT) calculations and simulations of the STM images39, point-
ing to an arrangement of pure Si pentagonal building blocks, as displayed in Fig.1d, which denes the missing
pentagonal row (P-MR) model employed in the Supplemental information of reference39 to optimize the angles
and the distance between the silicon atoms in the pentagonal arrangement. is unique atomic geometry was later
directly visualized by high-resolution non-contact atomic force microscopy (Fig.1c from40). We will theoretically
demonstrate that these p-SiNRs could constitute a tantalizing disruptive new Kitaev platform.
eoretically, the most well-established example of a topological superconductor hosting MZMs is a spinless
chain with p-wave superconducting pairing between neighboring sites, as proposed by Kitaev10. is type of
superconductivity seems to be extremely rare in nature45. However, spinless orbital p-wave superconductivity can
be engineered from a conventional spin-singlet s-wave superconductor when associated with a material exhibiting
helical bands46. is was rst achieved experimentally by Mourik etal.19, using a semiconducting nanowire with
strong Rashba spin-orbit interaction in proximity to an s-wave superconductor subject to a magnetic eld aligned
along the axis of the nanowire. Despite signicant advances in sample fabrication, measurement techniques, and
theoretical understanding, a denitive signature of MZMs detected through electronic transport measurements
in such hybrid systems remains challenging47. e main reason stems from inevitable inherent disorder, which
generates trivial zero-energy states, mimicking the MZMs signatures48.
We propose a new experimental platform to circumvent this disorder issue, still within the same conceptual
approach. Compared to previous proposals of Majorana nanowire devices, the paradigmatic dierence is the
replacement of the semiconducting nanowire by a new, highly ordered material, with transport measurements
replaced by scanning tunneling spectroscopy (STS). Specically, our proposal relies on the implementation of
long and atomically precise p-SiNRs, grown insitu under ultra-high-vacuum (UHV) in the cleanest conditions
on the Ag(110) surface, all aligned along the [110] direction3941.
e localized zero energy states associated with MZMs at the ends of p-SINRs can be detected by low-tem-
perature STS via an STM immersed in a strong parallel magnetic eld, following the methodology of Yazdani
and co-workers49. Since silver is not a superconductor, we will proximitize the p-SiNRs insitu with lead islands.
As mentioned before, lead is a conventional BCS superconductor with a relatively high critical temperature that
can be easily grown at Ag(110) surfaces50, which is known to interact only very weakly with the Si nanoribbons
while preserving its structural and electronic properties51, 52. To this end, a thin lead lm will be evaporated
insitu on top of the p-SiNRs and annealed at temperature
200C
, as already mentioned in4. Hence, at variance
with the classical fabrication and measurement procedures, all the steps in the experiments and measurements
will be performed in perfectly controlled and highly cleaned conditions in a UHV system, which will comprise
a surface science chamber (with all analytical tools and the two Si and Pb evaporators) directly linked to the
STM/STS chamber.
Moreover, within the present proposal, we characterize the topological phase transitions (TPTs) employing
the spinless version of the model and the inclusion of the p-wave superconducting pairing and the magnetic
eld reveal the emergence of topologically protected MZMs with the spin discriminated at opposite ends of the
p-SiNRs; this result constitutes the main nding of the work. We also calculate the wave function of the MZMs
at the ends of the p-SiNR, showing its topological signature.
Figure1. p-SiNR on Ag(110) surface. (a,b) Experimental STM images (uncorrected dri), (c) High-resolution
nc-AFM image. (d) Top and cross view of the arrangement of the Si pentagonal building blocks. (a,b) Courtesy
Eric Salomon, (c) Reprinted with permission from40. Copyright 2023 American Chemical Society. (d) From
Cerda etal.39.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
3
Vol.:(0123456789)
Scientic Reports | (2023) 13:17965 | https://doi.org/10.1038/s41598-023-44739-7
www.nature.com/scientificreports/
The model
Lattice transformations
In Fig.2a, to reduce the geometry complexity of the p-SiNR and facilitate the tight-binding calculations, we rede-
ne its structure using square-shaped pentagons. In the geometry of the pentagons that constitute the p-SiNRs
of Fig.2b, four silicon atoms are located on the missing silver row, and only one exhibits a buckling structure
(pink atoms). We neglect the buckling structure of these atoms and employ a planar conguration composed
of square-shaped pentagons. As the distance between the silicon atoms that constitute the pentagons are close,
we consider them equal to
a0
and identify it as the lattice parameter of the p-SiNR. We also dene the nearest
neighbor hopping as equal to t, which is considered the energy unit in all the calculations.
L2Na0
, is the length
of the p-SiNR, and N is the number of sites of the corresponding Kitaev chain (top or bottom), employed in the
calculation, as indicated in Fig.2c, that exhibits the shape of the p-SiNR and the unit cell composed of six atoms
inside the dashed rectangle employed in the calculations. We expect these simplications will not change the
results once we keep the lattice.
Eective Hamiltonian—spinless case
e total Hamiltonian, which describes the spinless p-SiNR of Fig.3 is given by
with
(1)
H=Ht+H,
Figure2. (a) Penta-silicene (p-SiNRs) lattice transformation adopted. (b) Penta-silicene angles. (c) Sketch of
nonequivalent Si atoms placed at the vertices of the “square” pentagonal lattice. e dashed rectangle depicts the
unit cell of the system.
Figure3. Sketch of the p-SiNRs: e penta-silicene system comprises two Kitaev chains, one on the top and
the second on the bottom, hybridized via hopping t. e ellipses represent the superconducting p-wave pairing
between silicon atoms belonging to the same chain, represented by pink (top) and yellow (bottom,) respectively
(in the real material, these atoms correspond to the buckled one). e arrows only indicate the spin polarization
needed to dene a Kitaev chain.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
4
Vol:.(1234567890)
Scientic Reports | (2023) 13:17965 | https://doi.org/10.1038/s41598-023-44739-7
www.nature.com/scientificreports/
where
µ
is the chemical potential, the index (−) and (
+
) dierentiate the creation and annihilation operators for
electrons and holes, respectively, and H.c. is the Hermitian conjugate. e system Hamiltonian of Eq.(2) was
built according to the unit cell of nonequivalent Si atoms (a, b, c, d, e, f) shown in Fig.2c.
e p-SiNRs are grown on Ag(110) surfaces in the setup proposed here. However, silver is not a superconduc-
tor, and to generate a p-wave pairing
on the pink and yellow atoms of Fig.3, we evaporate insitu a thin lead
lm over the Ag(110) surface in such a way that the buckled silicon atoms enter in contact with the lead atoms.
Under the presence of a strong Rashba spin orbit coupling (RSOC) arising from the Pb atoms and an applied
magnetic eld, the s-wave Cooper pairs of the Pb lm can enter into the p-SiNR region via proximity eect
(Andreev reections)12, giving rise to a p-wave-induced pairing in the double p-SiNRs structure. By following
the same procedure done in our previous work34 and based on the Kitaev model10, we introduce a spinless p-wave
superconducting pairing
between the “external” pink and yellow atoms of the same type as shown in Fig.3.
e Hamiltonian, which describes such a pairing, reads
Eective Hamiltonian—spinful case
In order to properly account for the spin degree of freedom in the superconducting p-SiNRs, we follow our
previous work34. Considering also the spin degree of freedom on both
and
H
,
We introduce a Zeeman eect due to the application of an external magnetic eld perpendicular to the p-SiNRs
plane. e Hamiltonian, which accounts for the Zeeman eect, reads:
wherein
Z
is the eective strength of the external Zeeman magnetic eld
B
, and
σ=↑,
is the spin index for
each operator.
e extrinsic RSOC induced in the p-SiNRs can be modulated by the action of an external electric eld
E
applied perpendicularly to the nanoribbon plane5356. Its corresponding general Hamiltonian reads
(2)
H
t=−
N
i=1
µa
i,+ai,+a
i,ai,+b
i,+bi,+b
i,bi,+c
i,+ci,+c
i,ci,
+d
i,+di,+d
i,di,+e
i,+ei,+e
i,ei,+f
i,+fi,+f
i,fi,
N
i=1
ta
i,+bi,+bi,a
i,+b
i,+ci,+ci,b
i,
+c
i,+di,+di,c
i,+d
i,+ei,+ei,d
i,+e
i,+fi,+fi,e
i,
N1
i=1
ta
i+1,+fi,+fi,a
i+1,+a
i+1,+ci,+ci,a
i+1,+d
i+1,+fi,+fi,d
i+1,+
H.c.,
(3)
H
=
N1
i=1
b
i,+b
i+1,b
i+1,b
i,++e
i,+e
i+1,e
i+1,e
i,++
H.c..
(4)
H
t=−
N
i=1,σ
µa
i,+,σai,+,σa
i,,σai,,σ+b
i,+,σbi,+,σb
i,,σbi,,σ+c
i,+,σci,+,σc
i,,σci,,
σ
+d
i,+,σdi,+,σd
i,,σdi,,σ+e
i,+,σei,+,σe
i,,σei,,σ+f
i,+,σfi,+,σf
i,,σfi,,σ
N
i=1,σ
ta
i,+,σbi,+,σbi,,σa
i,,σ+b
i,+,σci,+,σci,,σb
i,,σ+c
i,+,σdi,+,σ
di,,σc
i,,σ+d
i,+,σei,+,σei,,σd
i,,σ+e
i,+,σfi,+,σfi,,σe
i,,σ
N1
i=1,σ
ta
i+1,+,σfi,+,σfi,,σa
i+1,,σ+a
i+1,+,σci,+,σci,,σa
i+1,,σ
+
d
i
+
1,
+
,σfi,
+
,σ
fi,
,σd
i
+
1,
,σ
+
H.c.
(5)
H
=
N1
i=1,σ
b
i,+,σb
i+1,,σb
i+1,,σb
i,+,σ+e
i,+,σe
i+1,,σe
i+1,,σe
i,+,σ+
H.c..
(6)
H
z=
N
i=1,σ
Z sgn )a
i,σai,σ+b
i,σbi,σ+c
i,σci,σ+d
i,σdi,σ+e
i,σei,σ+f
i,σfi,σ+
H.c.,
Content courtesy of Springer Nature, terms of use apply. Rights reserved
5
Vol.:(0123456789)
Scientic Reports | (2023) 13:17965 | https://doi.org/10.1038/s41598-023-44739-7
www.nature.com/scientificreports/
where
u
i,j
=−
R
a0
ˆ
k
×
δi,
j
, with
R
being the extrinsic RSOC parameter,
δi
,
j
is the vector that connects the adjacent
lattice sites i and j, and
γ
the Pauli matrices. e index
¯σ
indicates the opposite spin direction of
σ
. e Eq.(7)
turns into
where
γ
1
=
1
2+
i
3
2
,
γ
2
=
1
2
i
3
2
,
γ
3
=
1
2
i
3
2
and
γ
4
=
1
2+
i
3
2
.
Notice that from Eqs.(6) and(7), we are assuming the external Zeeman magnetic eld
B
perfectly perpen-
dicular to the RSOC, i.e,
BB�= 0
and
B=0
. In Rashba nanowires setups, this condition is responsible for
the vanishing of the induced superconducting gap at zero momentum (inner gap) and the opening of a constant
gap at nite momentum (outer gap), which characterizes the topological phase transition and the concomitant
emergence of MZMs protected by the outer gap12.
However, from the experimental perspective, ensuring that the magnetic eld is applied only in the perpen-
dicular direction of the RSOC eld can be challenging. en, it is natural to consider also the eects of
B= 0
.
In this situation, we have both components of the Zeeman eld, and the critical magnetic eld condition for the
topological phase transition remains the same. However, the behavior of the outer gap is not constant anymore,
which aects the topological protection of the MZMs towards fault-tolerant quantum computing operations.
e eect of
B
in the outer gap is not so detrimental if the RSOC is strong.
It is worth noticing that the opposite cases of
BB= 0
and
B=0
can lead to the vanishing of the outer
gap, hence preventing the topological phase and emergence of MZMs. erefore, since our system is qualitatively
described by the similar underlying physics of Rashba nanowires, it is appropriate to experimentally ensure the
dominance of the magnetic eld component perpendicular to the Rashba eld.
We now can dene the total system Hamiltonian as
which can be written in the corresponding Bogolyubov–de Gennes (BdG) form in k-space as
Htotal(k)=T
H
BdG(k)�
, with
where
Hσ,σ(±k)
and
H,σ,σ(±k)
represent the matrix elements for dierent spin directions and the matrix
elements corresponding to the part of the matrix where superconducting couplings
appear, respectively. e
spinor
was constructed with the fermionic operators in the following order:
e spin alignment for each situation in the next section is computed numerically. We calculate the mean value
of the Pauli matrix in
ˆz
direction
ˆ
Sz
, i.e.,
ˆ
Sz=|ˆ
Sz|
, w here
|
are the eigenvectors of the total Hamiltonian
given by Eq.(9).
In hybrid semiconducting-superconducting nanowires, sometimes called Majorana nanowires, the following
features strongly suggest the emergence of MZMs at the nanowire ends12:
(a) Closing and subsequent reopening of the superconducting gap in the bulk relation dispersion as the chemi-
cal potential
µ
changes, indicating a TPT;
(b) Emergence of persistent zero-modes for specic system parameter values associated with nonoverlapping
wave functions localized at the opposite ends of the nanowire.
To obtain the TPTs present in the p-SiNRs, we will consider the innite case given by the Hamiltonian of Eq.(1).
We calculate the bulk band structure, discussed in detail in the supplemental material (SM). To investigate the
existence of MZMs in the p-SiNRs, we will analyze the spinless p-SiNRs with nite size
N=100
and calculate
(7)
H
R=
N
i
,
j=
1,σ
ic
i,σ(ui,j.γ)cj,(¯σ) +
H.c.,
(8)
H
R=
N
i=1,σ
γ1(a
i,σbi+1/2, ¯σ)+γ2(b
i+1/2,σai,¯σ)
+(a
i,σci1, ¯σ)(c
i1,σai,¯σ)+(i)(a
i,σfi,¯σ)
+(i)(f
i,σai,¯σ)+γ3(b
i+1/2,σci+1, ¯σ)+γ4(c
i+1,σbi+1/2, ¯σ
)
+(i)(c
i+1,σdi+1, ¯σ)+(i)(d
i+1,σci+1, ¯σ)(d
i+1,σfi,¯σ)
+(f
i,σdi+1, ¯σ)+γ3(d
i+1,σei+3/2, ¯σ)+γ4(e
i+3/2,σdi+1, ¯σ
)
+
γ1(e
i
+
3/2,σfi
+
2,
¯
σ)
+
γ2(f
i
+
2,σei
+
3/2,
¯
σ)
+
H.c.,
(9)
Htotal =Ht+HZ+HR+H,
(10)
H
BdG(k)=
H
,
(k)H
,
(k)H
,,
(k)H
,,
(k)
H,(k)H,(k)H,↓↑(k)H,,(k)
H
,,(k)H
,,(k)H
,(k)H
,(k)
H
,,
(
k)H
,,
(
k)H
,
(
k)H
,
(
k)
,
(11)
T
=(ak,,bk,,ck,,dk,,ek,,fk,,ak,,bk,,ck,,dk,,ek,,fk,,
a
k
,
,b
k
,
,c
k
,
,d
k
,
,e
k
,
,f
k
,
,a
k
,
,b
k
,
,c
k
,
,d
k
,
,e
k
,
,f
k
,
)
.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
6
Vol:.(1234567890)
Scientic Reports | (2023) 13:17965 | https://doi.org/10.1038/s41598-023-44739-7
www.nature.com/scientificreports/
the energy spectrum as a function of the chemical potential
µ
and the probability density function
|ψ|2
associ-
ated with the zero-energy states which arise on the real axis of the energy spectrum.
Both the energies
En
and eigenvectors
ψn
per site are obtained by numerically solving the Schrödinger equa-
tion
Hψn=Enψn
for the Hamiltonian of Eq.(1). To evaluate the position dependence of the wave functions
associated with zero energy states, we numerically calculate the eigenvector
ψn
when
En=0
, which allows
obtaining the probability density per lattice site according to
Results and discussion
Finite spinless p‑SiNRs
We employed the following parameter set in all the calculations:
=0.5t
,
Z=0.1t
,
R=0.05t
and
N=100
.
e top panels of Fig.4 show the bulk energy dispersion of the p-SiNRs, in the presence of the superconduct-
ing p-wave pairing, described by Eqs.(13), along the
kx
direction, for three representative values of chemical
potential
µ
[vertical lines in panel (d)]. Figure4a depicts the closing of the superconductor (SC) gap at
kx=0
for
µ=0.0t
. As the value of
µ
enhances, the SC gap opens as shown in panels (b) for
µ=0.4t
and closes again
at
kx=0
for
µ=0.7t
as shown in panel (c). is closing and reopening of the SC gap with the tuning of
µ
char-
acterize a topological phase transition. e bulk-boundary correspondence principle57 ensures the topologically
protected MZMs at the ends of the p-SiNRs.
To verify the emergence of MZMs associated with the TPTs depicted in Fig.4a–c, we plot the p-SiNRs
energy spectrum as a function of
µ
in Fig.4d. ere are no zero-energy modes for the values of
µ
where the gap
closes (red and magenta vertical lines). However, for values of
µ
inside the topological gap, for example, when
µ=0.4t
(green vertical line), two zero-energy states appear on the real axis, indicating the presence of MZMs
at the opposite ends of the p-SiNRs, topologically protected by the eective p-wave SC gap (Fig.4b). is nd-
ing is similar to what was obtained in our previous work34, wherein the MZMs emerge at the opposite ends of
a nite double zHNR.
Figure4f shows isolated zero-energy modes for
µ=0.4t
, which are associated with a nonoverlapping wave
function, well-localized at the ends of the p-SiNRs, as depicted in Fig.4j; which together with the topological
phase transition (Fig.4a–c), is a piece of strong evidence that topologically protected MZMs emerge at the
opposite ends of the spinless p-SiNRs. In the Supplemental Material, we developed an extensive analysis of the
topological and trivial phases of the spinless p-wave superconducting p-SiNR, that can be distinguished by the
Zak number topological invariant58. However, we cannot aord to do the same study for the spinful case due to
the extreme mathematical complexity.
Although there are zero-energy modes for other values of
µ
(Fig.4e,g), they are not associated with wave func-
tions well-localized at the ends of the p-SiNR, as can be seen in Fig.4i,k, for
µ=0.0t
and
µ=0.7t
, respect ively.
is implies that a zero mode cannot be exclusively attributed to MZMs. In this context, the genuine nature of
the zero-modes shown in Fig.4e,f, for instance, can be experimentally distinguished by combining STM and
atomic force microscopic (AFM) measurements49, 59. is approach allows associating the zero-bias conductance
(12)
|
ψ
n|2=
ψ
n
ψ
n.
Figure4. Spinless case: (ac) Bulk energy dispersion for the spinless p-SiNRs as a
kx
function. e colors red,
green, and magenta used in the panels correspond to
µ=0.0t
, 0.4t, and 0.7t, respectively. (d) Energy spectrum
as a function of the chemical potential. (eg) Zero-energy states spectrum. (ik) Probability density per lattice
site
|ψ|2
, associated with zero-energy states on the real axis of the Kitaev, top or bottom chains.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
7
Vol.:(0123456789)
Scientic Reports | (2023) 13:17965 | https://doi.org/10.1038/s41598-023-44739-7
www.nature.com/scientificreports/
peaks at the p-SINR ends (Fig.4f) with their well-localized wave functions through spatially resolved conduct-
ance maps, cf. Fig.4j.
Furthermore, we highlight that we analyze only one region of all energy spectrum shown in Fig.4d, which
presents other ranges of chemical potentials wherein a zero-energy state, associated with the emergence of MZMs,
arises. A more detailed study of the energy spectra can be found in the SM. We can also observe that, unlike the
system of our previous work34, the energy spectrum of Fig.4d is asymmetric at about
µ=0
.
Finite spinful p‑SiNRs
Now we will analyze how the spinless scenario shown in Fig.4 is aected by the presence of both Zeeman eld
(Eq.6) and extrinsic RSOC (Eq.7) coupling within the spinful description (Eq.9).
Figure5a–e exhibit the energy dispersion of the p-SiNRs given by the eigenenergies of BdG Hamiltonian
(Eq.10) as a function of
kx
, for distinct values of the chemical potential
µ
, indicated by vertical lines in Fig.5f.
e spin polarization is indicated by the vertical color bar, in which the red color represents the spin
↑= 1
, w hile
the blue color stands for spin
↓= 1
, and the light shades of colors mean the spin is neither up nor down. As
µ
is tuned, we can see the opening and closing of the superconducting gap, thus indicating a TPT, as previously
veried in the spinless situation (Fig.4a–c). However, here we can notice that each TPT associated with a specic
value of
µ
has a preferential spin orientation, except Fig.5c, where the system exhibits a conventional band gap.
e spin-polarized TPTs in Fig.5a,b,d,e lead to the appearance of spin-polarized zero-modes in Fig.5f, which
shows the system energy spectrum as a function of
µ
. ese zero-modes indicate the emergence of spin-polarized
MZMs at the ends of the p-SiNRs as
µ
is changed, similar to those found in34.
e panels g–k of Fig.5 depict the corresponding energy levels sorted in ascending order. e dierent values
of
µ
used to calculate the MZMs are indicated by vertical black lines in Fig.5f. For
µ=−2.7t
(Fig.5g), there are
two zero modes on the real axis of spin up (red points), associated with nonoverlapping wave functions shown
in Fig.5l.
For
µ=−2.35t
(Figs.5h and 6h), there are two energy-states in the spin-up direction and the other two
with spin-down, associated with degenerate (blue and red) nonoverlapping wave functions shown in Figs.5m
and6m, respectively. It is worth mentioning that this situation is absent from our previous paper34 and constitutes
a pivotal dierence exclusively related to the p-SiNRs. Here, pairs of MZMs at opposite edges split into top and
bottom chains of the nanoribbon with opposite spins. ese MZMs, acting as an eective two-level electronic
system, would allow the recovery of the spin degree of freedom as a good quantum number for purposes of
quantum computing, as well as to dene the intrinsic spins of the regular fermions built up by these top and
bottom pairs of MZMs edge states, respectively. By considering two p-SiNRs as source and drain reservoirs with
an immersed quantum dot, a type of single electron transistor (SET) could be set15, 60, 61, allowing to detect a
spin-polarized zero-bias conductance provided by one of the spins associated with a given pair of MZMs placed
at one particular nanoribbon chain.
e chemical potential
µ=1.1t
(Fig.5i) represents a non-topological region, there are four spin-down energy
states outside the real axis, there are no MZMs, and the wave functions completely overlap along the ribbon.
e system exhibits a particular “semiconductor” phase with the gap controlled by the
µ
variation and two delta
peaks generated by the zero modes outside the real axis, inside the gap.
Figure5. Spinful case—magnetic eld up: (ae) Bulk energy dispersion of the superconducting p-SiNRs for
the spinful situation, as a function of
kx
, for
µ=−2.7t
,
2.35t
, 1.1t, 2.09t and 2.2t, respectively. (f) Energy
spectrum as a function of the chemical potential. Vertical lines indicate the chosen values of chemical potential
shown on top panels. (g)–(k) Zero-energy states spectrum. (l)–(q) Probability density per lattice site
|ψ|2
,
associated with zero-energy states on the real axis of the spin-polarized Kitaev, top or bottom chains.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
8
Vol:.(1234567890)
Scientic Reports | (2023) 13:17965 | https://doi.org/10.1038/s41598-023-44739-7
www.nature.com/scientificreports/
For
µ=2.09t
(Fig.5j), there are two zero modes on the real axis of spin up (red points). Finally, for
µ=2.2t
(Fig.5k), there are four MZMs with spin-down energy states on the real axis. is situation happens because,
at
µ=2.09t
, a TPT occurs for spin-up, the gap closes at
kπ
, and for
µ>2.09t
the gap denes a trivial
band insulator for this spin orientation and MZMs with spin-up are not available anymore. ese well-localized
probability densities describing wave functions centered at the opposite ends of the superconducting p-SiNRs,
associated with zero-energy edge states, indicate the emergence of MZMs in the same way previously found for
the spinless system.
Figure6 represents the same situation as Fig.5 but with the magnetic eld pointing in the opposite direc-
tion. e net eect on the p-SiNRs is to change the MZMs, for all
µ
values, in spin up to down and vice versa.
erefore, it is possible to select the spin polarization of the MZMs by changing the chemical potential
µ
or the
magnetic eld orientation.
In Fig.7, we mainly analyze the dispersion relation, energy spectrum, and nature of the zero-modes at
µ=0
of Fig.5, with the magnetic eld pointing in the up direction. Figure7a depicts E(k) as a function of
kx
, showing
that there is a nite topological superconducting gap only for the spin-down orientation (blue line), while the
spin-up (red line) remains gapless. is behavior suggests a spin-polarized TPT at zero chemical potential, imply-
ing that only the system’s spin-down component is within the topological regime. At the same time, the spin-up
belongs to a metallic phase. Fig.7b,c represent two MZMs of spin-down with its correspondent nonoverlapped
wave function, respectively, and Fig.7d, shows detail at around
µ=0
region.
Additionally, we investigate how the energy spectrum as a function of
µ
is aected by the length of the
p-SiNRs. Fig.8 exhibits the superconducting p-SiNRs’ energy spectrum for increasing nanoribbon length values
N. From the smallest system considered (
N=10
, Fig.8a) to the largest one (
N=100
, Fig.8e), it can be noticed
a decrease of the amplitude of oscillations at around the real axis (
E=0
), and at the same time the denition of
the MZMs on the real axis improves as N increases, and for
N=100
the MZMs are well dened in all the real
axis. It should be mentioned that these oscillations around zero energy are expected for short Majorana nanowires
due to the overlap between Majorana wavefunctions of opposite ends. erefore, such oscillations are expected
to decrease as the system becomes larger. e same behavior was veried in the work34.
Figure6. Spinful case—magnetic eld down: e same situation of Fig.5 but with the magnetic eld pointing
in the opposite direction.
Figure7. Analysis in detail of the
µ=0
case of Fig.5 with the magnetic eld pointing in the up direction.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
9
Vol.:(0123456789)
Scientic Reports | (2023) 13:17965 | https://doi.org/10.1038/s41598-023-44739-7
www.nature.com/scientificreports/
Conclusions and perspectives
is paper demonstrates the emergence of topologically protected MZMs at opposite ends of spinless and spinful
p-SiNRs with p-wave superconducting pairing. ese MZMs exhibit spin discrimination, and their polarization
can be controlled by adjusting the nanoribbon chemical potential or the external magnetic eld. To implement
our ndings experimentally, we propose a material engineering of p-SiNRs grown over an Ag(110) surface (cf.
Fig.1a), with a thin Pb lm deposited on top4, 50. In this device, the proximity eect will enable the penetration
of Cooper pairs from the Pb s-wave superconductor into the p-SiNRs12, and in combination with an external
magnetic eld and the extrinsic RSOC modulated by the action of an external electric eld
E
applied perpen-
dicularly to the nanoribbon plane5356, it will induce p-wave pairing in the buckled atoms of the double p-SiNRs
structure (cf. Fig.1d).
We should highlight the potential applications driven by the spin-polarized MZMs presented in this work,
notably demonstrated in the results of Fig.7, with the down spin component associated with MZMs, while the
up component displays metallic features, resulting in a half-metallic behavior for the system62, 63. is property
could be harnessed to design a single Majorana transistor (SMT) built from a quantum dot (QD) sandwiched by
nite p-SiNR leads6, 64, 65. is setup resembles the conventional single electron transistor (SET)66. e SMT can
be a valuable tool for discerning between MZMs and trivial Andreev bound states15, 17. Particularly, the leakage
of MZMs through the QD67, along with both local and crossed Andreev reections induced by a specic spin
orientation within the p-SiNR-QD-p-SiNR SMT structure, is expected to generate distinct electronic transport
signatures, enabling the identication of MZMs.
In addition to the spin-polarization of MZMs, our proposal also features the emergence of two MZMs located
at opposite ends of the p-SiNR top chain, while another two with opposed spins are at the bottom, as illustrated
in Figs.5h and 6h. ese MZMs acting as an eective two-level electronic system would allow the recovery of the
spin degree of freedom as a good quantum number for purposes of quantum computing implementation, as well
as to dene the intrinsic spins of the regular fermions built-up by these top and bottom couples of edge MZMs,
respectively. It is crucial for implementing quantum computing operations between two qubits, as it requires the
presence of two fermionic sites, i.e., four MZMs68, 69. erefore, our proposal is a promising candidate for realizing
hybrid quantum computing operations70, 71 between conventional qubits and spin-polarized Majorana-based
qubits and paves the way for dening quantum computing operations using Majorana spintronics72.
Data availability
e data that support the ndings of this study are available from the corresponding author upon reasonable
request.
Received: 6 July 2023; Accepted: 11 October 2023
References
1. Vogt, P. et al. Silicene: Compelling experimental evidence for graphene like two-dimensional silicon. Phys. Rev. Lett. 108, 155501
(2012).
2. Fleurence, A. et al. Experimental evidence for epitaxial silicene on diboride thin lms. Phys. Rev. Lett. 108, 245501. https:// doi.
org/ 10. 1103/ PhysR evLett. 108. 245501 (2012).
Figure8. Energy spectrum as a function of the chemical potential
µ
, for distinct lengths of superconducting
p-SiNRs, namely, for
N=10
(a),
N=20
(b),
N=40
(c),
N=60
(d) and
N=100
(e).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
10
Vol:.(1234567890)
Scientic Reports | (2023) 13:17965 | https://doi.org/10.1038/s41598-023-44739-7
www.nature.com/scientificreports/
3. Feng, B. et al. Evidence of silicene in honeycomb structures of silicon on Ag(111). Nano Lett. 12, 3507. https:// doi. org/ 10. 1021/
nl301 047g (2012).
4. Dávila, M. & Le Lay, G. Silicene: Genesis, remarkable discoveries, and legacy. Mater. Today Adv. 16, 100312 (2022).
5. Tao, L. et al. Silicene eld-eect transistors operating at room temperature. Nat. Nanotechnol. 10, 227–231 (2015).
6. Le Lay, G. Silicene transistors. Nat. Nanotechnol. 10, 202 (2015).
7. Frolov, S. M., Plissard, S. R., Nadj-Perge, S., Kouwenhoven, L. P. & Bakkers, E. P. Quantum computing based on semiconductor
nanowires. MRS Bull. 38, 809 (2013).
8. Rančić, M. J. Exactly solving the Kitaev chain and generating Majorana-zero-modes out of noisy qubits. Sci. Rep. 12, 19882 (2022).
9. Flensberg, K., von Oppen, F. & Stern, A. Engineered platforms for topological superconductivity and Majorana zero modes. Nat.
Rev. Mater.https:// doi. org/ 10. 1038/ s41578- 021- 00336-6 (2021).
10. Kitaev, A.Y. Unpaired Majorana fermions in quantum wires. Phys. Usp. 44, 131. http:// stacks. iop. org/ 1063- 7869/ 44/i= 10S/a= S29
(2001)
11. Oreg, Y., Refael, G. & von Oppen, F. Helical liquids and Majorana bound states in quantum wires. Phys. Rev. Lett. 105, 177002.
https:// doi. org/ 10. 1103/ PhysR evLett. 105. 177002 (2010).
12. Aguado, R. Majorana quasiparticles in condensed matter. Riv Nuovo Cimento 40, 523. https:// doi. org/ 10. 1393/ ncr/ i2017- 10141-9
(2017).
13. Schuray, A., Weithofer, L. & Recher, P. Fano resonances in Majorana bound states-quantum dot hybrid systems. Phys. Rev. B 96,
085417. https:// doi. org/ 10. 1103/ PhysR evB. 96. 085417 (2017).
14. Prada, E., Aguado, R. & San-Jose, P. Measuring Majorana nonlocality and spin structure with a quantum dot. Phys. Rev. B 96,
085418. https:// doi. org/ 10. 1103/ PhysR evB. 96. 085418 (2017).
15. Ricco, L. S., de Souza, M., Figueira, M. S., Shelykh, I. A. & Seridonio, A. C. Spin-dependent zero-bias peak in a hybrid nanowire-
quantum dot system: Distinguishing isolated Majorana fermions from Andreev bound states. Phys. Rev. B 99, 155159. https:// doi.
org/ 10. 1103/ PhysR evB. 99. 155159 (2019).
16. Zhang, H., Liu, D. E., Wimmer, M. & Kouwenhoven, L. P. Next steps of quantum transport in Majorana nanowire devices. Nat.
Commun. 10, 5128. https:// doi. org/ 10. 1038/ s41467- 019- 13133-1 (2019).
17. Prada, E. et al. From Andreev to Majorana bound states in hybrid superconductor-semiconductor nanowires. Nat. Rev. Phys. 2,
575. https:// doi. org/ 10. 1038/ s42254- 020- 0228-y (2020).
18. Das, A. et al. Zero-bias peaks and splitting in an Al-InAs nanowire topological superconductor as a signature of Majorana fermions.
Nat. Phys. 8, 887 EP. https:// doi. org/ 10. 1038/ nphys 2479 (2012).
19. Mourik, V. et al. Signatures of Majorana fermions in hybrid superconductor–semiconductor nanowire devices. Science 336, 1003.
https:// doi. org/ 10. 1126/ scien ce. 12223 60 (2012).
20. Nadj-Perge, S. et al. Observation of Majorana fermions in ferromagnetic atomic chains on a superconductor. Science 346, 602.
https:// doi. org/ 10. 1126/ scien ce. 12593 27 (2014).
21. Krogstrup, P. et al. Epitaxy of semiconductor-superconductor nanowires. Nat. Mater. 14, 400. https:// doi. org/ 10. 1038/ nmat4 176
(2015).
22. Jeon, S. et al. Distinguishing a Majorana zero mode using spin-resolved measurements. Science 358, 772. https:// doi. org/ 10. 1126/
scien ce. aan36 70 (2017).
23. Clarke, D. J. Experimentally accessible topological quality factor for wires with zero energy modes. Phys. Rev. B 96, 201109(R).
https:// doi. org/ 10. 1103/ PhysR evB. 96. 201109 (2017).
24. Gül, Ö. et al. Ballistic Majorana nanowire devices. Nat. Nanotechnol. 13, 192. https:// doi. org/ 10. 1038/ s41565- 017- 0032-8 (2018).
25. Lutchyn, R. M. et al. Majorana zero modes in superconductor-semiconductor heterostructures. Nat. Rev. Mater. 3, 52. h ttps:// doi.
org/ 10. 1038/ s41578- 018- 0003-1 (2018).
26. Jäck, B., Xie, Y. & Yazdani, A. Detecting and distinguishing Majorana zero modes with the scanning tunnelling microscope. Nat.
Rev. Phys. 3, 541 (2021).
27. Pan, H. & Das Sarma, S. Physical mechanisms for zero-bias conductance peaks in Majorana nanowires. Phys. Rev. Res. 2, 013377.
https:// doi. org/ 10. 1103/ PhysR evRes earch.2. 013377 (2020).
28. Pan, H., Cole, W. S., Sau, J. D. & DasSarma, S. Generic quantized zero-bias conductance peaks in superconductor-semiconductor
hybrid structures. Phys. Rev. B 101, 024506. https:// doi. org/ 10. 1103/ PhysR evB. 101. 024506 (2020).
29. Pan, H., Liu, C.-X., Wimmer, M. & DasSarma, S. Quantized and unquantized zero-bias tunneling conductance peaks in Majorana
nanowires: Conductance below and above
2
e
2/h
. Phys. Rev. B 103, 214502. htt ps: // doi. org/ 10. 1103/ PhysR evB. 103. 214502 (2021).
30. Kim, H. et al. Toward tailoring Majorana bound states in articially constructed magnetic atom chains on elemental superconduc-
tors. Sci. Adv. 4, eaar5251. https:// doi. org/ 10. 1126/ sciadv. aar52 51 (2018).
31. Klinovaja, J., Ferreira, G. J. & Loss, D. Helical states in curved bilayer graphene. Phys. Rev. B 86, 235416. https:// doi. org/ 10. 1103/
PhysR evB. 86. 235416 (2012).
32. Klinovaja, J. & Loss, D. Giant spin-orbit interaction due to rotating magnetic elds in graphene nanoribbons. Phys. Rev. X 3, 011008.
https:// doi. org/ 10. 1103/ PhysR evX.3. 011008 (2013).
33. Maiellaro, A., Romeo, F. & Citro, R. Topological phase diagram of a Kitaev ladder. Eur. Phys. J. Spec. Top. 227, 1397. https:// doi.
org/ 10. 1140/ epjst/ e2018- 800090-y (2018).
34. Ribeiro, R. C. B., Correa, J. H., Ricco, L. S., Seridonio, A. C. & Figueira, M. S. Spin-polarized Majorana zero modes in double zigzag
honeycomb nanoribbons. Phys. Rev. B 105, 205115. https:// doi. org/ 10. 1103/ PhysR evB. 105. 205115 (2022).
35. Dutreix, C., Guigou, M., Chevallier, D. & Bena, C. Majorana fermions in honeycomb lattices. Eur. Phys. J. B 87, 296. https:// doi.
org/ 10. 1140/ epjb/ e2014- 50243-9 (2014).
36. Ma, T., Yang, F., Huang, Z. & Lin, H.-Q. Triplet p-wave pairing correlation in low-doped zigzag graphene nanoribbons. Sci. Rep.
7, 42262 EP. https:// doi. org/ 10. 1038/ srep4 2262 (2017).
37. Zhaoa, A. & Wang, B. Two-dimensional graphene-like xenes as potential topological materials. APL Mater. 8, 030701. https:// doi.
org/ 10. 1063/1. 51359 84 (2020).
38. Grazianetti, C. & Martella, C. e rise of the xenes: From the synthesis to the integration processes for electronics and photonics.
Materials (Basel) 14, 4170 (2021).
39. Cerdá, J. I. et al. Unveiling the pentagonal nature of perfectly aligned single-and double-strand Si nano-ribbons on Ag(110). Nat.
Commun. 7, 13076 (2016).
40. Sheng, S. et al. e pentagonal nature of self-assembled silicon chains and magic clusters on Ag(110). Nano Lett. 18, 2937 (2018).
41. Yue, S. et al. Observation of one-dimensional Dirac fermions in silicon nanoribbons. Nano Lett. 22, 695. https:// doi. org/ 10. 1021/
acs. nanol ett. 1c038 62 (2022).
42. De Padova, P., Quaresima, C., Ottaviani, C., Sheverdyaeva, P.M., Moras, P., Carbone, C., Topwal, D., Olivieri, B., Kara, A., Oughad-
dou, H., Aufray, B. & LeLay, G. Evidence of graphene-like electronic signature in silicene nanoribbons. Appl. Phys. Lett. 96. https://
doi. org/ 10. 1063/1. 34591 43 (2010).
43. De Padova, P.et al. 1d graphene-like silicon systems: Silicene nano-ribbons. J. Phys. Condens. Matter 24, 223001 (2012).
44. Pawlak, R. et al. Quantitative determination of atomic buckling of silicene by atomic force microscopy. Proc. Natl. Acad. Sci. 117,
228 (2020).
45. Das Sarma, S. In search of Majorana. Nat. Phys. 19, 165. https:// doi. org/ 10. 1038/ s41567- 022- 01900-9 (2023).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
11
Vol.:(0123456789)
Scientic Reports | (2023) 13:17965 | https://doi.org/10.1038/s41598-023-44739-7
www.nature.com/scientificreports/
46. Laubscher, K. & Klinovaja, J. Majorana bound states in semiconducting nanostructures. J. Appl. Phys. 130, 081101. https:// doi.
org/ 10. 1063/5. 00559 97 (2021).
47. Frolov, S. Quantum computing’s reproducibility crisis: Majorana fermions. Nature 592, 350. https:// doi. org/ 10. 1038/ d41586- 021-
00954-8 (2021).
48. Souto, R.S., Tsintzis, A., Leijnse, M. & Danon, J. Probing Majorana localization in minimal kitaev chains through a quantum dot
(2023). arXiv: 2308. 14751 [cond-mat.mes-hall]
49. Jäck, B., Xie, Y. & Yazdani, A. Detecting and distinguishing Majorana zero modes with the scanning tunnelling microscope. Nat.
Rev. Phys. 3, 541. https:// doi. org/ 10. 1038/ s42254- 021- 00328-z (2021).
50. Tsud, N. et al. Interfacial reconstruction in the system Pb/Ag(110). Surf. Sci. 542, 112. https:// doi. org/ 10. 1016/ S0039- 6028(03)
00977-4 (2003).
51. Podsiadły-Paszkowska, A. & Krawiec, M. Dirac fermions in silicene on Pb(111) surface. Phys. Chem. Chem. Phys. 17, 2246. https://
doi. org/ 10. 1039/ C4CP0 5104A (2015).
52. Stepniak-Dybala, A. & Krawiec, M. Formation of silicene on ultrathin Pb(111) lms. J. Phys. Chem. C 123, 17019. https:// doi. org/
10. 1021/ acs. jpcc. 9b043 43 (2019).
53. Min, H. et al. Intrinsic and Rashba spin-orbit interactions in graphene sheets. Phys. Rev. B 74, 165310. https:// doi. org/ 10. 1103/
PhysR evB. 74. 165310 (2006).
54. Zarea, M. & Sandler, N. Rashba spin-orbit interaction in graphene and zigzag nanoribbons. Phys. Rev. B 79, 165442. https:// doi.
org/ 10. 1103/ PhysR evB. 79. 165442 (2009).
55. Ezawa, M. Valley-polarized metals and quantum anomalous hall eect in silicene. Phys. Rev. Lett. 109, 055502. https:// doi. org/ 10.
1103/ PhysR evLett. 109. 055502 (2012).
56. Jiao, Z., Yao, Q. & Zandvliet, H. Tailoring and probing the quantum states of matter of
2d
Dirac materials with a buckled honeycomb
structure. Physica E: Low-dimens. Syst. Nanostruct. 121, 114113. https:// doi. org/ 10. 1016/j. physe. 2020. 114113 (2020).
57. Al ase, A. Boundary Physics and Bulk-Boundary Correspondence in Topological Phases of Matter 1st edn. (Springer eses, Springer
Nature Switzerland AG, 2019).
58. Zak, J. Berry’s phase for energy bands in solids. Phys. Rev. Lett. 62, 2747. https:// doi. org/ 10. 1103/ PhysR evLett. 62. 2747 (1989).
59. Pawlak, R. et al. Probing atomic structure and Majorana wavefunctions in mono-atomic Fe chains on superconducting Pb surface.
npj Quantum Inf. 2, 16035. https:// doi. org/ 10. 1038/ npjqi. 2016. 35 (2016).
60. Campo, V. L., Ricco, L. S. & Seridonio, A. C. Isolating Majorana fermions with nite Kitaev nanowires and temperature: Universality
of the zero-bias conductance. Phys. Rev. B 96, 045135. https:// doi. org/ 10. 1103/ PhysR evB. 96. 045135 (2017).
61. Ricco, L. S., Campo, V. L., Shelykh, I. A. & Seridonio, A. C. Majorana oscillations modulated by fano interference and degree of
nonlocality in a topological superconducting-nanowire-quantum-dot system. Phys. Rev. B 98, 075142. https:// doi. org/ 10. 1103/
PhysR evB. 98. 075142 (2018).
62. Tsai, W.-F. et al. Gated silicene as a tunable source of nearly 100% spin-polarized electrons. Nat. Commun. 4, 1500. https:// doi. org/
10. 1038/ ncomm s2525 (2013).
63. Jiang, P. et al. Robust generation of half-metallic transport and pure spin current with photogalvanic eect in zigzag silicene
nanoribbons. J. Phys. Condens. Matter 31, 495701. https:// doi. org/ 10. 1088/ 1361- 648x/ ab3dd6 (2019).
64. Ihn, T. et al. Graphene single-electron transistors. Mater. Today 13, 44. https:// doi. org/ 10. 1016/ S1369- 7021(10) 70033-X (2010).
65. Niu, W. et al. Exceptionally clean single-electron transistors from solutions of molecular graphene nanoribbons. Nat. Mater. 22,
180. https:// doi. org/ 10. 1038/ s41563- 022- 01460-6 (2023).
66. Goldhaber-Gordon, D. et al. Kondo eect in a single-electron transistor. Nature 391, 156. https:// doi. org/ 10. 1038/ 34373 (1998).
67. Vernek, E., Penteado, P. H., Seridonio, A. C. & Egues, J. C. Subtle leakage of a Majorana mode into a quantum dot. Phys. Rev. B 89,
165314. https:// doi. org/ 10. 1103/ PhysR evB. 89. 165314 (2014).
68. Karzig, T. et al. Scalable designs for quasiparticle-poisoning-protected topological quantum computation with Majorana zero
modes. Phys. Rev. B 95, 235305. https:// doi. org/ 10. 1103/ PhysR evB. 95. 235305 (2017).
69. Steiner, J. F. & von Oppen, F. Readout of Majorana qubits. Phys. Rev. Res. 2, 033255. https:// doi. org/ 10. 1103/ PhysR evRes earch.2.
033255 (2020).
70. Leijnse, M. & Flensberg, K. Quantum information transfer between topological and spin qubit systems. Phys. Rev. Lett. 107, 210502.
https:// doi. org/ 10. 1103/ PhysR evLett. 107. 210502 (2011).
71. Leijnse, M. & Flensberg, K. Hybrid topological-spin qubit systems for two-qubit-spin gates. Phys. Rev. B 86, 104511. https:// doi.
org/ 10. 1103/ PhysR evB. 86. 104511 (2012).
72. Liu, X., Li, X., Deng, D.-L., Liu, X.-J. & Das Sarma, S. Majorana spintronics. Phys. Rev. B 94, 014511. https:// doi. org/ 10. 1103/ PhysR
evB. 94. 014511 (2016).
Acknowledgements
M.S.F., M.A.C., and A.C.S. acknowledge nancial support from the National Council for Scientic and Tech-
nological Development (CNPq) grant numbers 311980/2021-0, 305810/2020-0, 308695/2021-6, respectively.
M.S.F. acknowledges the Foundation for Support of Research in the State of Rio de Janeiro (FAPERJ) processes
number 210 355/2018 and 211.605/2021. M.A.C. acknowledges nancial support to the Foundation for Support
of Research in the State of Rio de Janeiro (FAPERJ) for the fellowship of the Programa Cientistas do Nosso Estado,
E-26/201.223/2021. L.S.R. and I.A.S. acknowledge the Icelandic Research Fund (Rannis), grant No. 163082-051.
Author contributions
All authors participate in the scientic discussion of the work. All authors reviewed the paper. M.S.F., R.C.B.R.,
M.M., G.L.L., L.S.R., A.C.S., and J.H.C. edit the paper. R.C.B.R. performed the numerical calculations. R.C.B.R.,
and L.S.R. performed analytical calculations.
Competing interests
e authors declare no competing interests.
Additional information
Supplementary Information e online version contains supplementary material available at https:// doi. org/
10. 1038/ s41598- 023- 44739-7.
Correspondence and requests for materials should be addressed to M.S.F.
Reprints and permissions information is available at www.nature.com/reprints.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
12
Vol:.(1234567890)
Scientic Reports | (2023) 13:17965 | https://doi.org/10.1038/s41598-023-44739-7
www.nature.com/scientificreports/
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Open Access is article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the
Creative Commons licence, and indicate if changes were made. e images or other third party material in this
article are included in the articles Creative Commons licence, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder. To view a copy of this licence, visit http:// creat iveco mmons. org/ licen ses/ by/4. 0/.
© e Author(s) 2023
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... In the same vein, we previously designed another scalable platform based on the straight, highly perfect, narrow (0.8 nm in width), single-stranded penta-silicene NRs (p-SiNRs) spontaneously self-aligned at room temperature along the [110] direction of a silver (110) template [44]. Similar conditions, as described above for the zHNRs, cause the structure of the p-SiNRs to turn into a double spin-polarized Kitaev chain, with each of the edge displaying spin up or down orientations. ...
... The band structure of the penta-silicene reveals several topological phase transitions. For the corresponding finite case, the energy spectrum as a function of chemical potential displays different configurations of MZMs, as indicated in figure 5 of our previous article [44]. These p-SiNRs should host MZMs at their ends, which could be detected in situ by tunneling spectroscopy. ...
Article
Full-text available
Majorana fermions are a fascinating class of particles with unique and intriguing properties: they are their own antiparticles, as first theorized by the Italian physicist Ettore Majorana in 1937. In recent decades, research in condensed matter physics shows theoretically that in certain exotic states of matter, such as topological superconductors, pairs of Majorana fermions can emerge as bound states at defects or interfaces, known as Majorana Zero Modes (MZMs). They behave like non-local anyons and could be used as decoherence-protected qubits. After the seminal work of Kitaev (2001), one-dimensional artificial setups have been developed in line with the concept of the Kitaev chain to implement MZMs. As no definite proof has yet been widely accepted by the community, improvements in the architectures and setups have been realized, and different platforms have been devised, which could be kinds of ‘DNA’ in this rapidly evolving vivid ecosystem. Here, we sequence these ‘DNAs’ and draw perspectives for topological quantum computation.
Article
Full-text available
Artificial Kitaev chains, formed by quantum dots coupled via superconductors, have emerged as a promising platform for realizing Majorana bound states. Even a minimal Kitaev chain (a quantum dot–superconductor–quantum dot setup) can host Majorana states at discrete sweet spots. However, unambiguously identifying Majorana sweet spots in such a system is still challenging. In this work, we propose an additional dot coupled to one side of the chain as a tool to identify good sweet spots in minimal Kitaev chains. When the two Majorana states in the chain overlap, the extra dot couples to both and thus splits an even-odd ground-state degeneracy when its level is on resonance. In contrast, a ground-state degeneracy will persist for well-separated Majorana states. This difference can be used to identify points in parameter space with spatially separated Majorana states, using tunneling spectroscopy measurements. We perform a systematic analysis of different relevant situations. We show that the additional dot can help distinguish between Majorana sweet spots and other trivial zero-energy crossings. We also characterize the different conductance patterns, which can serve as a guide for future experiments aiming to study Majorana states in minimal Kitaev chains.
Article
Full-text available
Only single-electron transistors with a certain level of cleanliness, where all states can be properly accessed, can be used for quantum experiments. To reveal their exceptional properties, carbon nanomaterials need to be stripped down to a single element: graphene has been exfoliated into a single sheet, and carbon nanotubes can reveal their vibrational, spin and quantum coherence properties only after being suspended across trenches1–3. Molecular graphene nanoribbons4–6 now provide carbon nanostructures with single-atom precision but suffer from poor solubility, similar to carbon nanotubes. Here we demonstrate the massive enhancement of the solubility of graphene nanoribbons by edge functionalization, to yield ultra-clean transport devices with sharp single-electron features. Strong electron–vibron coupling leads to a prominent Franck–Condon blockade, and the atomic definition of the edges allows identifying the associated transverse bending mode. These results demonstrate how molecular graphene can yield exceptionally clean electronic devices directly from solution. The sharpness of the electronic features opens a path to the exploitation of spin and vibrational properties in atomically precise graphene nanostructures.
Article
Full-text available
Majorana-zero-modes (MZMs) were predicted to exist as edge states of a physical system called the Kitaev chain. MZMs should host particles that are their own antiparticles and could be used as a basis for a qubit which is robust-to-noise. However, all attempts to prove their existence gave inconclusive results. Here, the Kitaev chain is exactly solved with a quantum computing methodology and properties of MZMs are probed by generating eigenstates of the Kitev Hamiltonian on 3 noisy qubits of a publicly available quantum computer. After an ontological elaboration I show that two eigenstates of the Kitaev Hamiltonian exhibit eight signatures attributed to MZMs. The results presented here are a most comprehensive set of validations of MZMs ever conducted in an actual physical system. Furthermore, the findings of this manuscript are easily reproducible for any user of publicly available quantum computers, solving another important problem of research with MZMs—the result reproducibility crisis.
Article
Full-text available
The recent outcomes related to the Xenes, the two-dimensional (2D) monoelemental graphene-like materials, in three interdisciplinary fields such as electronics, photonics and processing are here reviewed by focusing on peculiar growth and device integration aspects. In contrast with forerunner 2D materials such as graphene and transition metal dichalcogenides, the Xenes pose new and intriguing challenges for their synthesis and exploitation because of their artificial nature and stabilization issues. This effort is however rewarded by a fascinating and versatile scenario where the manipulation of the matter properties at the atomic scale paves the way to potential applications never reported to date. The current state-of-the-art about electronic integration of the Xenes, their optical and photonics properties, and the developed processing methodologies are summarized, whereas future challenges and critical aspects are tentatively outlined.
Article
Majorana particles are the same as their antiparticle, and their analogues in condensed matter may be a platform for quantum computing. Here I describe the search for these modes in semiconductor heterostructures and how disorder is a limiting factor. Majorana zero modes are emergent excitations in topological superconductors. This Perspective introduces the physics of these modes, recaps the recent history of the experimental search for them and discusses the future prognosis for success.
Article
In the realm of two-dimensional (2D) materials, besides the ones initially peeled from lamellar crystals, the artificial emerging elemental ones, called Xenes, appear as strong contenders to graphene in a booming new field. The very first synthetic Xene was silicene, created in 2012. On the occasion of its tenth anniversary, this concise review, describes the birth of silicene, in situ, under ultra-high vacuum, and surveys its most tantalizing properties: its Dirac features, its 2D topological insulator character, its easy functionalization, its insertion as atom-thin channel in Field Effect Transistors operating at room temperature. Silicene has striking variants and amazing doubles in the quantum world; these lookalikes are briefly described and their origins discussed. We owe to silicene the legacy of all its descendants from borophene to tellurene, and fascinating prospects for spintronics, the emergence and control of Majorana fermions, possibly for quantum computing.
Article
We study the emergence of Majorana zero modes (MZMs) at the ends of a finite double zigzag honeycomb nanoribbon (zHNR), considering a minimal model composed of the first nearest neighbor hopping term, Rashba spin-orbit coupling (RSOC), p-wave superconducting pairing, and an applied external magnetic field (EMF). The energy spectrum profiles reveal regions with either spin up or down MZMs belonging to distinct topological phase transitions, which are characterized by their corresponding winding numbers and can be accessed by tuning the chemical potential of the nanoribbons. Hybrid systems constituted by zHNRs deposited on conventional s-wave superconductors are potential candidates for experimentally realizing our proposal. The spin discrimination of MZMs suggests a possible route for performing topological-conventional qubit operations using Majorana spintronics.
Article
In this Tutorial, we give a pedagogical introduction to Majorana bound states (MBSs) arising in semiconducting nanostructures. We start by briefly reviewing the well-known Kitaev chain toy model in order to introduce some of the basic properties of MBSs before proceeding to describe more experimentally relevant platforms. Here, our focus lies on simple “minimal” models where the Majorana wave functions can be obtained explicitly by standard methods. In the first part, we review the paradigmatic model of a Rashba nanowire with strong spin–orbit interaction (SOI) placed in a magnetic field and proximitized by a conventional s-wave superconductor. We identify the topological phase transition separating the trivial phase from the topological phase and demonstrate how the explicit Majorana wave functions can be obtained in the limit of strong SOI. In the second part, we discuss MBSs engineered from proximitized edge states of two-dimensional (2D) topological insulators. We introduce the Jackiw–Rebbi mechanism leading to the emergence of bound states at mass domain walls and show how this mechanism can be exploited to construct MBSs. Due to their recent interest, we also include a discussion of Majorana corner states in 2D second-order topological superconductors. This Tutorial is mainly aimed at graduate students—both theorists and experimentalists—seeking to familiarize themselves with some of the basic concepts in the field.
Article
Among the major avenues that are being pursued for realizing quantum bits, the Majorana-based approach has been the most recent to be launched. It attempts to realize qubits that store quantum information in a topologically protected manner. The quantum information is protected by non-local storage in localized and well-separated Majorana zero modes, and manipulated by exploiting their non-abelian quantum statistics. Realizing these topological qubits is experimentally challenging, requiring superconductivity, helical electrons (created by spin–orbit coupling) and breaking of time-reversal symmetry to all cooperate in an uncomfortable alliance. Over the past decade, several candidate materials systems for realizing Majorana-based topological qubits have been explored, and there is accumulating, though still debated, evidence that zero modes are indeed being realized. This Review surveys the basic physical principles on which these approaches are based, the materials systems that are being developed and the current state of the field. We highlight both the progress that has been made and the challenges that still need to be overcome.