1. Introduction
There is currently interest in realizing synthetic topological quantum matter with topologically protected quasiparticles at its edges [
1,
2,
3], with potential application in topological quantum computation [
4,
5,
6,
7,
8,
9]. Haldane fractional spin quasiparticles in a spin one chain and Majorana Fermions in topological superconductors are good examples [
9,
10,
11]. To realize Majorana Fermions Kitaev proposed [
11,
12] a chain of quantum dots on a p-wave superconductor that carries such non-local zero energy Majorana Fermions localized on its two ends, the Majorana zero modes (MZMs). Since then there have been numerous proposals to realize the Kitaev chain [
13,
14,
15,
16,
17,
18,
19]. In all cases, experimental confirmation of the presence of the MZMs has proved to be a non-trivial and challenging task [
20,
21,
22,
23,
24,
25,
26,
27,
28].
Recent progress in semiconductor quantum dots in nanowires [
29,
30,
31,
32,
33,
34,
35,
36] opens the possibility of realizing Kitaev chains and optical detection of their Majorana zero modes. In this work, we consider such an array of InAsP quantum dots embedded in an InP nanowire as the material system [
29,
30,
31,
32,
33,
34,
35,
36] for realization of MZM, and study its signature in light-matter interaction. As the schematic in
Figure 1(a) shows, we combine a semiconductor nanowire with p-wave superconductor [
37,
38,
39,
40,
41,
42]. The p-wave pairing in this system is introduced by proximity effect among electrons that are spin-polarized by an external magnetic field, making sure that Cooper pairs can only form between electrons in the conduction band (CB) of adjacent dots. We will show that one can tune the system parameters into topological regime, where two MZMs appear at the two ends of the chain. With semiconductor quantum dots, light can generate a hole in the valence band (VB) and an electron in the conduction band. The electron adds to an existing gas of Majorana Fermions while the hole then interacts with all the quasiparticles of the Kitaev chain, including MZMs, to form composite objects similar to excitons and trions in the Fermi Edge Singularity problem [
43,
44,
45,
46]. This leads to a structure in the absorption spectrum of the chain as a function of photon energy. Here we present a theory for the signatures of the MZMs in the optical spectra of the semiconductor nanowire.
After describing the model in
Section 2, in
Section 3 we introduce the exact diagonalization (ED) method, and introduce Majorana and bond Fermion representation of the Kitaev Hamiltonian. Next, in
Section 4 we describe exciton-Majorana Fermion complexes and predict the absorption spectrum. We focus discussion on the optical signature of the MZM in the absorption spectrum. Finally, in
Section 5 we conclude by summarising our results and discuss potential experiments detecting Majorana Fermions in a semiconductor Kitaev chain.
2. Kitaev chain in a semiconductor nanowire
Figure 1(a) shows a schematic representation of the Kitaev chain we are considering. It consists of a hexagonal InP nanowire with an array of embedded InAsP quantum dots in the proximity of a p-wave superconductor [
37,
38,
39,
40,
41,
42], in the presence of applied external magnetic field. Such arrays have been extensively investigated [
29,
30,
31,
32,
33,
34,
35,
36], including their excitonic complexes [
29,
36]. As
Figure 1(b) shows, in our model we include the lowest conduction spin level of each dot and the highest spin valence band level, which are effectively both spin-polarized due to the external magnetic field. Consequently, in the presence of superconductivity, only the electrons from the adjacent conduction levels can pair up, as there is only one conduction level available in each dot. The Kitaev Hamiltonian
in Equation (
1a) describes the hopping and pairing of electrons in conduction band levels. The chemical potential is tuned to bring the chain to near half filling in the absence of superconductivity. Therefore, in the equilibrium the valence levels are full and the system is described by the Kitaev Hamiltonian. However, when a photon with energy close to the band gap of InAsP illuminates the dots, it generates a hole in VB and an electron in CB. The electron becomes one of the electrons in CB and decomposes into quasiparticles of the superconducting state. The hole then forms a bound state with the quasiparticles of the electronic system that is in a collective superconducting state. These possible bound states generate peaks in the absorption spectrum of the system, among which there is the signature of MZM, as we shall show below.
The hole is described by a simple tight binding Hamiltonian
in Equation (1b). We also consider electron-hole interaction,
in Equation (1c), which is strongest when both conduction electron and valence band hole are on the same quantum dot. Hence, we write the full Hamiltonian of the system as
where
is the normal Fermionic creation operator of an electron(hole) in dot
i,
is hopping between adjacent conduction(valence) levels,
is pairing energy between adjacent conduction levels,
is the chemical potential measured from the conduction energy level, and
is the CB to VB energy gap in each dot. In the interaction term Equation (1c),
V is the Coulomb attraction energy between electrons and holes, where we also introduced
, the electron(hole) number operator in dot
i.
Figure 1(b) schematically shows different terms of Equation (1) between two adjacent dots.
Next, before describing the absorption experiment, we start with a brief discussion of Kitaev Hamiltonian.
3. Majorana and bond Fermions in Kitaev Hamiltonian
The Kitaev Hamiltonian
in Equation (
1a), originally introduced in Ref. [
11], supports two MZMs localized on the two ends of the chain, when the Hamiltonian is in topological regime. For a finite chain, the topological region is centered on parameters
and
, which is our focus throughout this work. Here, after describing the exact diagonalization (ED) method for normal Fermions, following Kitaev [
11], we show how using Majorana Fermions reveals the usefulness of a new set of Fermions we refer to as
bond Fermions. Next, after matching energy spectra obtained by ED in both normal and bond Fermion bases, we shall use the bond Fermion basis for the rest of the paper.
3.1. Exact diagonalization in normal Fermion basis
We start off by introducing the exact diagonalization method (ED) for finding the energy spectrum of the Kitaev Hamiltonian. In ED we span the Hilbert space of the system by configuration basis [
47]. For our electronic system being made of
N spinless orbitals, there are
possible configurations, which we construct as
where
is the vacuum of electrons,
, corresponds to having(1) or not having(0) electron in orbital
i.
For a given number of electrons
M we generate electron configurations
. But as Kitaev Hamiltonian, being a Hamiltonian for a superconductor, does not conserve particle number, its eigenstates are coherent linear combinations of electronic configurations with different electron numbers as
where we are populating
N sites with
electrons. To solve for coefficients
, we apply the Hamiltonian on this state, and by using the orthogonality of the configurations we obtain the eigenvalue equation
However, since the Kitaev Hamiltonian
in Equation (
1a) only changes particle number in pairs, the matrix element
is non-zero only if
M and
have the same parity, i.e. if they are both even or odd. This parity symmetry allows us to break the Hilbert space into two decoupled subspaces of even and odd configurations. In
Appendix A we explicitly show the configurations and the Hamiltonian matrix
in each of these subspaces, for the case of
.
3.2. Bond Fermions
We now express the Kitaev Hamiltonian in Equation (
1a) in terms of Majorana and bond Fermions. First, as schematically shown in
Figure 2, we write each electron operator ,
c and
in terms of two Majorana Fermion operators
and
as
where the
’s are Majorana Fermion operators. Majorana Fermions satisfy a slightly different anti-commutation relation than the ordinary Fermions,
.
Using Equation (
5) and Majorana anti-commutation relations, the Hamiltonian
can be written in terms of Majorana Fermions as
The form in Equation (
6) shows the pairing between Majoranas of different types in adjacent sites. But it also shows that the Hamiltonian is not diagonal in Majorana Fermions, they are not quasiparticles of the Kitaev Hamiltonian. Following Kitaev [
11], as shown in
Figure 2, we define a new set of Fermionic operators, bond Fermions, which are made of two Majoranas of different types from adjacent sites as
where we also defined
, to which we refer as the zero mode, out of the two unpaired Majoranas at the two ends of the chain, as shown in
Figure 2. Then, the Hamiltonian in terms of bond Fermion operators is
where in the second and the third sum one should identify
. Note that in topological regime, when
and
, the bond Fermions diagonalize the Hamiltonian in Equation (
8) and reduce it to
which implies a set of
quasiparticles with energy
, and one non-local quasiparticle
with zero energy, hence the name zero mode. In this case, since bond Fermions are the quasiparticles of Kitaev Hamiltonian, their configurations are the eigenstates of the system.
In this spirit, we also use bond Fermion configurations for exact diagonalization of Kitaev Hamiltonian. In the same fashion as in Equation (
2) we define bond Fermion configurations as
where
is the vacuum of bond Fermions, and we used the
overline to distinguish these configurations from the normal Fermion configurations. Next, an equation similar to the equation in Equation (
3) can be written for the eigenstates of the Hamiltonian in terms of bond Fermion configurations, where now
would represent the
configuration of having
M bond Fermions. And similar to the case of normal Fermions, since
in Equation (
8) conserves the parity of bond Fermion numbers too, we can split the Hilbert space into even and odd subspaces. In
Appendix A we explicitly show the bond Fermion configurations and the Hamiltonian matrix of Equation (
8) in each of these subspaces, for the case of
.
3.3. Energy spectrum
To demonstrate the usefulness of bond Fermion basis, we now describe the energy spectrum of a chain of quantum dots, obtained both in the normal and bond Fermion basis.
Figure 3 shows the energy spectrum for the case of
. Throughout the work, we consider
, as it is the case for conduction bands hopping integrals. As we mentioned above, in this case, the configurations of bond Fermions are also the eigenstates of the system. Being in topological regime, with these parameters the system has a doubly degenerate ground state, one in the odd subspace
with all bond Fermions, and the other in the even subspace
, which is missing the zero energy bond Fermion
. Next, we have the singly excited states, missing one non-zero bond Fermion, with excitation energy
, from which we have two in each subspace,
and
in the even subspace, and
and
in the odd subspace. Finally, in each subspace, there is one doubly excited state, missing two non-zero bond Fermion with excitation energy
,
in the even subspace, and
in the odd subspace.
Table 1 summarizes the description of the spectrum in terms of bond Fermions.
4. Kitaev chain and a light induced valence hole
Absorption of a photon injects an electron-hole pair into the system. Therefore, the relevant optically excited states live in the subspace of all configurations with one hole. Here, after studying the energy spectrum of the full Hamiltonian in Equation (1) with one hole in the configuration space of bond Fermions, we discuss the absorption spectrum of the chain and the optical signature of the MZM.
4.1. Exact diagonalization of electron-hole system
Having demonstrated the benefit of bond Fermion basis, we now study the Hamiltonian with one hole in the configuration basis of bond Fermions and one hole. Using
to refer to
M bond Fermions being in their
configuration, and the hole being at site
m, we can find the spectrum by solving an equation similar to Equation (
4), but considering the full Hamiltonian
H in Equation (1) rather than
, and in the configuration basis of bond Fermions and one hole.
For instance, for
dots, following the convention we introduced in
Section 3.3 and
Table 1, we can list these configurations as
In this subspace, the hole Hamiltonian
in Equation (1b) amounts to a constant
and mixes states with the same electronic configurations and different locations of the hole by hopping matrix element
. Therefore, with the ordering in
Table 2, the full Hamiltonian with one hole for the example of
dots has a structure like
where each block is a
matrix,
is the identity matrix times
, and the diagonal blocks are given by the matrix elements of
in Equation (1) over the configurations in
Table 2. The interaction term
in Equation (1c) for each of the diagonal blocks
is
, and it mixes up different bond Fermion configurations as we have
which implies that when the hole is not at the two ends of the chain then the interaction mixes up two non-zero bond Fermions, and when it is at one of the two ends, the interaction mixes the zero mode with one of the non-zero ones. For instance, for the operator
and
in the even configuration basis in
Table 2 we have
while
, in a similar fashion to
, mixes
with
, and
with
.
4.2. Energy spectrum of the electron-hole system
Figure 4 shows the energy spectrum of a chain of length
dots in the even subspace and for
, and
, as the electron-hole interaction
V increases; for a localized hole (
) on the left panel, and for a mobile hole with
on the right panel. Both cases show branching into two groups, pertaining to bonding and antibonding pairs of states, mixed by the interaction
V.
The case of localized hole allows us to understand the spectrum better. There are four states associated with each dot, and since the two end dots are geometrically the same, the spectrum always shows four pairs of doubly degenerate states. As we show in the
Appendix B, and it can be seen from Equation (
13), for the two end dots two of these four states are mixtures of
and
that give us visible peaks at
described in
Section 4.3.1, and also indicated on the plot. The two other pairs of degenerate levels are mixtures of
and
, which do not get excited by absorbing a photon. For the middle dot, as can be seen from Equation (
13), one pair of states are mixture of
and
, where only the bonded state gets excited by absorbing a photon (see
Appendix B), resulting in the peak
; also described in
Section 4.3.1 and shown on the plot. Finally, the last pair of states, which also do not get excited by absorbing a photon, are mixtures of
and
.
As can be seen in the right panel of
Figure 4, for a mobile hole, when
, we still have four pairs of doubly degenerate states as a result of the chain’s spatial symmetry. At
for the case of localized hole there is an extra triple degeneracy because of the non-dispersive nature of the localized hole band. But for a mobile hole, it can be seen on the right panel of
Figure 4, that at
the degenerate levels split into sets of triples, corresponding to the three propagating modes of the hole band. More importantly, in this case, since the three dot subspaces are connected by hole hopping (see Equation (
11)), the above described pairs of states mix up by
, and the ones that are closer in energy mix more. As a result of this mixture, more peaks arise in the absorption spectrum, as we discuss in the next section.
4.3. Absorption spectrum
As the schematic in
Figure 1(a) shows, in absorption experiment a photon probes the chain along the nanowire. InAsP dots having significantly smaller bandgap than InP bulk of the nanowire [
48], guaranties that the photon can only be absorbed by the dots. For calculating the absorption spectrum of the chain, we assume that the photon creates an electron-hole pair with uniform probability along the nanowire, and so define the polarization operator as
where we also introduced the local electron-hole pair operator
.
We are assuming that one can also setup the system to create the electron-hole pair on a chosen specific dot
i [
35,
49,
50,
51], i.e. acting with the operator
on the chain, rather than
P. As we discuss, having access to such spatially resolved spectrum is important in detecting the optical signature of the MZM.
The polarization operator
– we use this notation to simultaneously refer to
P and
– takes the ground state of the system to an excited state with one hole and one electron in CB. Since the ground state can be degenerate, as it is when
and
, the absorption spectrum has an even and an odd part pertaining to each ground state
where
are the eigenstates of the one hole subspace and the corresponding electron parity, and we used the notation
to simultaneously refer to the regular absorption spectrum
A, and
the spatially resolved absorption spectrum coming from dot
i.
4.3.1. Analytic result for localized hole
If
and the hole is localized the full Hamiltonian becomes block diagonal (see Equation (
11)), i.e. the subspaces of having the hole in each of the dots decouple. Consequently, we have
At the heart of topological regime when
and
, as depicted graphically in
Figure 5, and expressed in Equation (A6) , an electron created at site
i by
, decomposes into a superposition of creation and annihilation operators of the two bond Fermions on its two sides,
and
. If the electron is created at one of the two ends, one of the bond Fermions is the zero mode
. On the other hand, when the hole is at site
i, the interaction
mixes the two bond Fermions, as shown in
Figure 5, and expressed in Equations (
A8) and (
A12). As we show in the
Appendix B, combining these two mechanisms, one can find an analytic expression for the spatially resolved absorption spectrum for a chain of arbitrary length
N if the hole is created on site
i as
where
and then the full absorption spectrum is given by the simple sum in Equation (
16).
The bottom row of
Figure 6 shows the results in Equations (
17) and (18) for the case of
. The peak
is only present in the middle, while the peaks
are present at the two ends of the chain. As we show in the
Appendix B, the two peaks
have a mixture of zero mode in them, while
is purely made of non-zero bond Fermions. At
,
is purely made of zero mode while
is purely made of non-zero bond Fermions. As we increase
V,
acquires more zero mode contribution while
mixes more with a non-zero bond Fermions. At the same time, by increasing
V, the peak at
diminishes, as can be seen from Equation (18c). If not too weak,
peak is a better resolved optical signature for the MZM than
, as it is separated from the rest of the spectrum by
V, and we expect to have
. This presents an advantage over scanning tunnelling microscopy approach for detecting MZM [
21]. Moreover, if one can perform spatially resolved absorption spectroscopy on the chain, the presence of the zero mode can be determined by the presence of a visible peak at high energy near
when probing the end dots, and its absence when probing other dots.
4.3.2. Absorption for mobile hole
When the hole is mobile, there are
N itinerant hole states with different energies. Therefore, one would expect
N different transitions to each electronic state. More importantly, as can be seen in Equation (
11), hopping hole mixes up different subspaces of having the hole in different dots. As a result more transitions become allowed leading to the emergence of more peaks in the absorption spectrum.
In
Figure 6 we compare the absorption spectrum of a mobile hole with
(top row), and the analytic result of Equation (
17) for localized hole (bottom row), for the case of
. It is evident how more peaks are visible for the case of mobile hole, while the major peaks are still close to the location of
and
. Moreover, note how in the full spectrum
(top left) there is only one visible peak at high energy near
, and how the same peak is large in
(top middle) and faint in
(top right), pertaining to the localized nature of the MZM that
carries. In plotting
Figure 6 we used
, as for a mobile hole the even and odd parts of the absorption spectrum are not the same. But since there is no preference between the two ground states, one would expect to observe an average of the two.
The same logic is valid for a chain of any length, as the analytic result in Equation (
17) is for general
N.
Figure 7 shows the absorption spectrum of a chain of length
. Here, we set
while changing
. When
we approach the idealized case of localized hole, where the subspaces of having the hole on each dot are decoupled. Growing
mixes up the modes of different dots. Consequently, the zero mode starts leaking out of the two ends of the chain. On the first panel of
Figure 7 we can see that at high energy there is still only one visible peak near
, until about
a faint peak appears to the right of it. This makes
a very robust signature for a relatively large range of hole hopping. Having access to spatially resolved spectrum we can further confirm that the peak is indeed coming from the two ends. It can be seen from the third panel of
Figure 7 that the there is no visible high energy peak on the site next to the end dot (
) until around
. In contrast, we can observe in
that
, which also contains a large share of zero mode and is a stronger peak, starts leaking out of the end dot very quickly for small
’s.
5. Conclusion
We present here a theory of Majorana excitons, photo-excited conduction electron-valence band hole pairs, interacting with Majorana Fermions in a Kitaev chain of semiconductor quantum dots embedded in a nanowire. Using exact diagonalization techniques and Majorana and Bond Fermions we compute the energy spectra of the system. We confirm the existence of nonlocal bond Fermion, a superposition of Majorana Fermions at the two ends of the chain, with zero energy. We introduce a valence band hole and describe its interaction with Majorana fermions. We predict interband absorption spectra and discuss the signature of Majorana Zero Modes in the absorption spectra. We demonstrate how spatially resolved absorption spectrum can be used to confirm the localized character of the MZMs.
We hope this preliminary work motivates future theoretical and experimental work on hybrid nanowire semiconductor quantum dots /superconductor systems for the demonstration of Majorana Fermions.
Author Contributions
Conceptualization, P.H. and M.K.; methodology, P.H.; software, M.M., D.M., and H.A.; validation, M.M., D.M., H.A., and D.G.; formal analysis, H.A., M.M., and D.G.; investigation, P.H., M.M., and H.A.; resources, P.H., D.M.; data curation, H.A. and M.M.; writing—original draft preparation, H.A. and M.M; writing—review and editing, P.H., H.A., M.M., and D.M.; visualization, M.M. and H.A.; supervision, P.H.; project administration, P.H.; funding acquisition, P.H. and M.K.. All authors have read and agreed to the published version of the manuscript.
Funding
This research was supported by the Quantum Sensors and Applied Quantum Computing Challenge Programs at the National Research Council of Canada, by NSERC Discovery Grant No. RGPIN- 2019-05714, and University of Ottawa Research Chair in Quantum Theory of Quantum Materials, Nanostructures, and Devices.
Data Availability Statement
The codes used to produce data presented in this study are openly available in "Kitaev Exciton" at reference [
52].
Conflicts of Interest
The authors declare no conflict of interest.
Appendix A. Exact diagonalization for chain of length three
Following Equation (
2), for a chain of length
, the configurations of normal Fermions in the even and odd subspaces are
Computing the matrix elements of all terms in the Kitaev Hamiltonian
in Equation (
1a) between every pairs of configurations in Equation (
A1a) and Equation (A1b) we can explicitly write the matrix
in Equation (
4), in each of the subspaces as
for which we used the same ordering as in Equation (A1).
Then, for ED in the bond Fermion basis, following Equation (
10), the configurations are
and one can use Equation (
8) to find the corresponding Kitaev Hamiltonian matrices in each of the subspaces as
for which we used the same ordering as in Equation (A3). Notice how the two Hamiltonian matrices in Equation (A4) become diagonal when
and
.
Appendix B. Analytic calculation of absorption spectrum for localized hole
Following the notation in
Table 1, for a chain of arbitrary length
N, and for
and
, we express the two degenerate ground states of the system as
Here, we derive the result in Equations (
17) and (18) using
in Equation (
A5a), and the procedure is the same for
.
To start, first note that from Equation (7) we have
which means for
we have
where we used the same notation as in
Table 2 for the excited states. On the other hand, from Equation (
12a) we have
Recalling that both
and
are eigenstates of
with excitation energy
, then in the basis
, the full Hamiltonian
is
with the following two eigenstates
Now using Equation (
A7) we have
, and
, which give us the result in Equation (
17) for the case of
.
Next, we show the second case for
, and the procedure is the same for
. In this case, using Equation (A6b) we have
where
is identical to the other ground state
up to a global phase, hence its excitation energy is zero. Then from Equation (12b) we have
Therefore, considering that
and
are eigenstates of
with excitation energy
and zero, respectively, the full Hamiltonian
in the basis
is
and its two eigenstates are given by
with
Now using Equation (
A11) we have
, and
, where we used Equation (
A15) to express these matrix elements in terms of
in Equation (18c). From this follows the result in Equation (
17) for the case of
.
References
- Gyongyosi, L.; Imre, S. A survey on quantum computing technology. Computer Science Review 2019, 31, 51–71. doi:10.1016/j.cosrev.2018.11.002. [CrossRef]
- Field, B.; Simula, T. Introduction to topological quantum computation with non-Abelian anyons. Quantum Science and Technology 2018, 3, 045004. doi:10.1088/2058-9565/aacad2. [CrossRef]
- Campbell, E.T.; Terhal, B.M.; Vuillot, C. Roads towards fault-tolerant universal quantum computation. Nature 2017, 549, 172–179. doi:10.1038/nature23460. [CrossRef]
- Stern, A.; Lindner, N.H. Topological quantum computation—from basic concepts to first experiments. Science 2013, 339, 1179–1184. doi:10.1126/science.1231473. [CrossRef]
- Nayak, C.; Simon, S.H.; Stern, A.; Freedman, M.; Sarma, S.D. Non-Abelian anyons and topological quantum computation. Reviews of Modern Physics 2008, 80, 1083. doi:10.1103/revmodphys.80.1083. [CrossRef]
- Sarma, S.D.; Freedman, M.; Nayak, C. Majorana zero modes and topological quantum computation. npj Quantum Information 2015, 1, 1–13. doi:10.1038/npjqi.2015.1. [CrossRef]
- Das Sarma, S.; Freedman, M.; Nayak, C. Topological quantum computation. Physics Today 2006, 59, 32–38. doi:10.1063/1.2337825. [CrossRef]
- Freedman, M.; Kitaev, A.; Larsen, M.; Wang, Z. Topological quantum computation. Bulletin of the American Mathematical Society, 40, 31–38. doi:10.48550/arXiv.quant-ph/0101025. [CrossRef]
- Jaworowski, B.; Hawrylak, P. Quantum bits with macroscopic topologically protected states in semiconductor devices. Applied Sciences 2019, 9, 474. doi:10.3390/app9030474. [CrossRef]
- Haldane, F.D.M. Nonlinear field theory of large-spin Heisenberg antiferromagnets: semiclassically quantized solitons of the one-dimensional easy-axis Néel state. Physical Review Letters 1983, 50, 1153. doi:10.1103/PhysRevLett.50.1153. [CrossRef]
- Kitaev, A.Y. Unpaired Majorana fermions in quantum wires. Physics-Uspekhi 2001, 44, 131. doi:10.1070/1063-7869/44/10s/s29. [CrossRef]
- Kitaev, A.Y. Fault-tolerant quantum computation by anyons. Annals of Physics 2003, 303, 2–30. doi:10.1016/s0003-4916(02)00018-0. [CrossRef]
- Lutchyn, R.M.; Sau, J.D.; Sarma, S.D. Majorana fermions and a topological phase transition in semiconductor-superconductor heterostructures. Physical Review Letters 2010, 105, 077001. doi:10.1103/PhysRevLett.105.077001. [CrossRef]
- Mourik, V.; Zuo, K.; Frolov, S.M.; Plissard, S.; Bakkers, E.P.; Kouwenhoven, L.P. Signatures of Majorana fermions in hybrid superconductor-semiconductor nanowire devices. Science 2012, 336, 1003–1007. doi:10.1126/science.1222360. [CrossRef]
- Sau, J.D.; Sarma, S.D. Realizing a robust practical Majorana chain in a quantum-dot-superconductor linear array. Nature Communications 2012, 3, 964. doi:10.1038/ncomms1966. [CrossRef]
- Leijnse, M.; Flensberg, K. Parity qubits and poor man’s Majorana bound states in double quantum dots. Physical Review B 2012, 86, 134528. doi:10.1103/PhysRevB.86.134528. [CrossRef]
- Dvir, T.; Wang, G.; van Loo, N.; Liu, C.X.; Mazur, G.P.; Bordin, A.; Ten Haaf, S.L.; Wang, J.Y.; van Driel, D.; Zatelli, F.; others. Realization of a minimal Kitaev chain in coupled quantum dots. Nature 2023, 614, 445–450. doi:10.1038/s41586-022-05585-1. [CrossRef]
- Nadj-Perge, S.; Drozdov, I.K.; Li, J.; Chen, H.; Jeon, S.; Seo, J.; MacDonald, A.H.; Bernevig, B.A.; Yazdani, A. Observation of Majorana fermions in ferromagnetic atomic chains on a superconductor. Science 2014, 346, 602–607. doi:10.1126/science.1259327. [CrossRef]
- Sun, H.H.; Jia, J.F. Detection of Majorana zero mode in the vortex. npj Quantum Materials 2017, 2, 34. doi:10.1038/s41535-017-0037-4. [CrossRef]
- Liu, D.E.; Baranger, H.U. Detecting a Majorana-fermion zero mode using a quantum dot. Physical Review B 2011, 84, 201308. doi:10.1103/physrevb.84.201308. [CrossRef]
- Jäck, B.; Xie, Y.; Yazdani, A. Detecting and distinguishing Majorana zero modes with the scanning tunnelling microscope. Nature Reviews Physics 2021, 3, 541–554. doi:10.1038/s42254-021-00328-z. [CrossRef]
- Pientka, F.; Romito, A.; Duckheim, M.; Oreg, Y.; von Oppen, F. Signatures of topological phase transitions in mesoscopic superconducting rings. New Journal of Physics 2013, 15, 025001. doi:10.1088/1367-2630/15/2/025001. [CrossRef]
- Pikulin, D.I.; van Heck, B.; Karzig, T.; Martinez, E.A.; Nijholt, B.; Laeven, T.; Winkler, G.W.; Watson, J.D.; Heedt, S.; Temurhan, M.; others. Protocol to identify a topological superconducting phase in a three-terminal device. arXiv preprint arXiv:2103.12217 2021. doi:10.48550/arXiv.2103.12217. [CrossRef]
- Liu, J.; Potter, A.C.; Law, K.T.; Lee, P.A. Zero-bias peaks in the tunneling conductance of spin-orbit-coupled superconducting wires with and without Majorana end-states. Physical Review Letters 2012, 109, 267002. doi:10.1103/PhysRevLett.109.267002. [CrossRef]
- Sarma, S.D.; Pan, H. Disorder-induced zero-bias peaks in Majorana nanowires. Physical Review B 2021, 103, 195158. doi:10.1103/PhysRevB.103.195158. [CrossRef]
- Rubbert, S.; Akhmerov, A. Detecting Majorana nonlocality using strongly coupled Majorana bound states. Physical Review B 2016, 94, 115430. doi:10.1103/physrevb.94.115430. [CrossRef]
- Aghaee, M.; Akkala, A.; Alam, Z.; Ali, R.; Ramirez, A.A.; Andrzejczuk, M.; Antipov, A.E.; Astafev, M.; Bauer, B.; Becker, J.; others. InAs-Al hybrid devices passing the topological gap protocol. arXiv preprint arXiv:2207.02472. doi:10.48550/arXiv.2207.02472. [CrossRef]
- Baldelli, N.; Bhattacharya, U.; González-Cuadra, D.; Lewenstein, M.; Graß, T. Detecting Majorana zero modes via strong field dynamics. ACS Omega 2022. doi:10.1021/acsomega.2c07169. [CrossRef]
- Cygorek, M.; Korkusinski, M.; Hawrylak, P. Atomistic theory of electronic and optical properties of InAsP/InP nanowire quantum dots. Physical Review B 2020, 101, 075307. doi:10.1103/physrevb.101.075307. [CrossRef]
- Manalo, J.; Cygorek, M.; Altintas, A.; Hawrylak, P. Electronic and magnetic properties of many-electron complexes in charged InAsxP1-x quantum dots in InP nanowires. Physical Review B 2021, 104. doi:10.1103/PhysRevB.104.125402. [CrossRef]
- Koong, Z.X.; Ballesteros-Garcia, G.; Proux, R.; Dalacu, D.; Poole, P.J.; Gerardot, B.D. Multiplexed single photons from deterministically positioned nanowire quantum dots. Physical Review Applied 2020, 14, 034011. doi:10.1103/PhysRevApplied.14.034011. [CrossRef]
- Dalacu, D.; Mnaymneh, K.; Lapointe, J.; Wu, X.; Poole, P.J.; Bulgarini, G.; Zwiller, V.; Reimer, M.E. Ultraclean emission from InAsP quantum dots in defect-free wurtzite InP nanowires. Nano Letters 2012, 12, 5919–5923. doi:10.1021/nl303327h. [CrossRef]
- Jaworowski, B.; Rogers, N.; Grabowski, M.; Hawrylak, P. Macroscopic singlet-triplet qubit in synthetic spin-one chain in semiconductor nanowires. Scientific Reports 2017, 7, 5529. doi:10.1038/s41598-017-05655-9. [CrossRef]
- Phoenix, J.; Korkusinski, M.; Dalacu, D.; Poole, P.J.; Zawadzki, P.; Studenikin, S.; Williams, R.L.; Sachrajda, A.S.; Gaudreau, L. Magnetic tuning of tunnel coupling between InAsP double quantum dots in InP nanowires. Scientific Reports 2022, 12, 5100. doi:10.1038/s41598-022-08548-8. [CrossRef]
- Northeast, D.B.; Weber, J.F.; Dalacu, D.; Phoenix, J.; Poole, P.J.; Aers, G.; Lapointe, J.; Williams, R.L. Optical fibre-based (plug-and-play) single photon source using InAsP quantum dot nanowires and gradient-index lens collection. arXiv preprint arXiv:2104.11197. doi:10.1038/s41598-021-02287-y. [CrossRef]
- Laferrière, P.; Yeung, E.; Korkusinski, M.; Poole, P.J.; Williams, R.L.; Dalacu, D.; Manalo, J.; Cygorek, M.; Altintas, A.; Hawrylak, P. Systematic study of the emission spectra of nanowire quantum dots. Applied Physics Letters 2021, 118, 161107. doi:10.1063/5.0045880. [CrossRef]
- Talantsev, E.; Iida, K.; Ohmura, T.; Matsumoto, T.; Crump, W.; Strickland, N.; Wimbush, S.; Ikuta, H. p-wave superconductivity in iron-based superconductors. Scientific Reports 2019, 9, 14245. doi:10.1038/s41598-019-50687-y. [CrossRef]
- Wang, W.S.; Zhang, C.C.; Zhang, F.C.; Wang, Q.H. Theory of chiral p-wave superconductivity with near nodes for Sr2RuO4. Physical Review Letters 2019, 122, 027002. doi:10.1103/PhysRevLett.122.027002. [CrossRef]
- Yuan, N.F.; Mak, K.F.; Law, K. Possible topological superconducting phases of MoS2. Physical Review Letters 2014, 113, 097001. doi:10.1103/PhysRevLett.113.097001. [CrossRef]
- Frigeri, P.; Agterberg, D.; Koga, A.; Sigrist, M. Superconductivity without Inversion Symmetry: MnSi versus CePt3Si. Physical Review Letters 2004, 92, 097001. doi:10.1103/PhysRevLett.92.097001. [CrossRef]
- Hardy, F.; Huxley, A. p-wave superconductivity in the ferromagnetic superconductor URhGe. Physical Review Letters 2005, 94, 247006. doi:10.1103/PhysRevLett.94.247006. [CrossRef]
- Ishida, K.; Mukuda, H.; Kitaoka, Y.; Asayama, K.; Mao, Z.; Mori, Y.; Maeno, Y. Spin-triplet superconductivity in Sr2RuO4 identified by 17O Knight shift. Nature 1998, 396, 658–660. doi:10.1038/25315. [CrossRef]
- Mahan, G. Many-Particle Physics; Physics of Solids and Liquids, 2012. doi:10.1007/978-1-4757-5714-9. [CrossRef]
- Wojs, A.; Hawrylak, P. Negatively charged magnetoexcitons in quantum dots. Physical Review B 1995, 51, 10880–10885. doi:10.1103/PhysRevB.51.10880. [CrossRef]
- Hawrylak, P. Excitonic effects in optical spectra of a quasi-one-dimensional electron gas. Solid State Communications 1992, 81, 525–527. doi:https://doi.org/10.1016/0038-1098(92)90605-9. [CrossRef]
- Hawrylak, P. Optical properties of a two-dimensional electron gas: Evolution of spectra from excitons to Fermi-edge singularities. Physical Review B 1991, 44, 3821–3828. doi:10.1103/PhysRevB.44.3821. [CrossRef]
- Weiße, A.; Fehske, H., Exact Diagonalization Techniques. In Computational Many-Particle Physics; Fehske, H.; Schneider, R.; Weiße, A., Eds.; Springer Berlin Heidelberg: Berlin, Heidelberg, 2008; pp. 529–544. doi:10.1007/978-3-540-74686-7_18. [CrossRef]
- Adachi, S., Energy-Band Structure: Energy-Band Gaps. In Properties of Group-IV, III-V and II-VI Semiconductors; John Wiley & Sons, Ltd, 2005; chapter 6, p. 116. doi:10.1002/0470090340.ch6. [CrossRef]
- Dalacu, D.; Poole, P.J.; Williams, R.L. Tailoring the geometry of bottom-up nanowires: application to high efficiency single photon sources. Nanomaterials 2021, 11. doi:10.3390/nano11051201. [CrossRef]
- Laferriere, P.; Yeung, E.; Giner, L.; Haffouz, S.; Lapointe, J.; Aers, G.C.; Poole, P.J.; Williams, R.L.; Dalacu, D. Multiplexed single-photon source based on multiple quantum dots embedded within a single nanowire. Nano Letters 2020, 20, 3688–3693. doi:10.1021/acs.nanolett.0c00607. [CrossRef]
- Laferriére, P.; Haffouz, S.; Northeast, D.B.; Poole, P.J.; Williams, R.L.; Dalacu, D. Position-controlled telecom single photon emitters operating at elevated temperatures. Nano Letters 2023, 23, 962–968. doi:10.1021/acs.nanolett.2c04375. [CrossRef]
- Allami, H. Kitaev Exciton. Available at https://github.com/hassan-allami/KitaevExciton.git, 2023.
|
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2023 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).