Next Article in Journal
Effect of UV-Irradiation and ZnO Nanoparticles on Nonlinear Optical Response of Specific Photochromic Polymers
Next Article in Special Issue
Correlation between Optical Localization-State and Electrical Deep-Level State in In0.52Al0.48As/In0.53Ga0.47As Quantum Well Structure
Previous Article in Journal
Scanning Thermal Microscopy of Ultrathin Films: Numerical Studies Regarding Cantilever Displacement, Thermal Contact Areas, Heat Fluxes, and Heat Distribution
Previous Article in Special Issue
Phonons in Short-Period GaN/AlN Superlattices: Group-Theoretical Analysis, Ab initio Calculations, and Raman Spectra
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fractional Charge States in the Magneto-Photoluminescence Spectra of Single-Electron InP/GaInP2 Quantum Dots

1
Ioffe Physical-Technical Institute of the Russian Academy of Sciences, 194021 St. Petersburg, Russia
2
Electrical Engineering, University of Notre Dame, Notre Dame, IN 46556, USA
3
Department of Physics and Engineering, ITMO University, 197101 St. Petersburg, Russia
4
SyMMES, IRIG, University Grenoble Alpes, CEA, CNRS, F-38000 Grenoble, France
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(2), 493; https://doi.org/10.3390/nano11020493
Submission received: 24 January 2021 / Revised: 9 February 2021 / Accepted: 13 February 2021 / Published: 16 February 2021

Abstract

:
We used photoluminescence spectra of single electron quasi-two-dimensional InP/GaInP2 islands having Wigner-Seitz radius ~4 to measure the magnetic-field dispersion of the lowest s, p, and d single-particle states in the range 0–10 T. The measured dispersion revealed up to a nine-fold reduction of the cyclotron frequency, indicating the formation of nano-superconducting anyon or magneto-electron (em) states, in which the corresponding number of magnetic-flux-quanta vortexes and fractional charge were self-generated. We observed a linear increase in the number of vortexes versus the island size, which corresponded to a critical vortex radius equal to the Bohr radius and closed-packed topological vortex arrangements. Our observation explains the microscopic mechanism of vortex attachment in composite fermion theory of the fractional quantum Hall effect, allows its description in terms of self-localization of ems and represents progress towards the goal of engineering anyon properties for fault-tolerant topological quantum gates.

1. Introduction

The integer (IQHE) [1] and fractional (FQHE) [2] quantum Hall effects, which reveal a skipping-type dissipationless edge transport in two-dimensional (2D) electron (e) gas semiconductor structures in a strong perpendicular magnetic field, are two of the most important discoveries in the physics of the last decades of the 20th century. The IQHE provides a precise method of measuring the fundamental fine-structure constant and gave the first example of edge/surface topological states that are robust to external perturbations [3,4]. The FQHE yields very rich physics involving fractionally charged anyon states [5] and Majorana zero-energy modes [6]. These open up the possibility of building topologically protected quantum gates (TQG) [6,7], which promise to dramatically improve the operation of quantum computers and accelerate the creation of their compact hardware—one of the challenges of science and technology in the 21st century.
The IQHE is well understood in terms of noninteracting electrons and the quantization of their Hall conductance by integer Landau-level fillings, ν [8]. The FQHE—representing quantization by fractional fillings ν = p/k [9], where k and p are the rational integers—is instead considered to be a property of strongly interacting electrons condensing into incompressible liquids of anyon quasiparticles (QPs) having charge 1/k, as was suggested by Laughlin [10]. This liquid state is described by Laughlin’s wavefunction with k zeros (k-WZs) localized at the positions of each electron keeping them separated, thus reducing the interaction energy. The FQHE can also be described using the concept of composite fermion (CFs) particles consisting of an electron with 2p magnetic-flux-quanta (MFQ) attached, as was shown further by Jain [11]. In this approach, the FQHE states are the IQHE states of weakly interacting CFs and the CF wave-function of k Landau level state has 2kp-WZs, i.e., ν = k/(2pk + 1). Thus, both approaches are equivalent. While these theoretical approaches have described a rich set of experimental data since the discovery of the FQHE nearly 40 years ago, a microscopic description thereof remains elusive [12]. This results from the fact that they do not clarify the connection between WZs and MFQ, i.e., the microscopic mechanism of MFQ attachment, and its dependence on the localization of the electrons. This creates uncertainty in the interpretation of FQHE magneto-conductivity experiments related to the realization of topological quantum gates (TQG) [13,14,15].
Theoretical proposals [16,17] and a methodology [18] for topological quantum computing are based on Laughlin’s QP or Jain’s CF description of the FQHE, and recommend using a set of localized QPs/CFs pulled from the appropriate host liquids. Thus, experiments demonstrating the possibility of the creation of single localized fractionally charged QPs/CFs are a necessary step for the realization of TQGs, and are critical for our understanding of the microscopic nature of the FQHE. So far, however, the existence of localized QP/CF puddles/molecules has only been observed using high-spatial-resolution electrometry of 2D quantum Hall structures [19,20] or near-field photoluminescence measurements of quantum dots (QDs) [21], while the localization of single fractionally charged QP/CF has not been reported.
Here, we demonstrate such localization and report the formation of fractionally charged states of single electrons, which gives insights into the mechanism of its coupling to MFQ. This clarifies the unresolved issues in the microscopic description of the FQHE and gives new perspectives for the realization of TQG. We use measurements of magneto-photoluminescence (magneto-PL) spectra of quasi-2D islands represented by self-organized semiconductor InP/GaInP2 QDs for the detection of the charge of single-electron states for a Wigner-Seitz radius rs~4 in a range of magnetic field 0–10 T. These QDs are large (~150 nm), providing a strong Wigner localization regime, together with in situ electron doping and a built-in magnetic field [22], and, thus, have unique properties compared to other QD material systems [23,24]. Using these, we measured a fractional charge in the range 1/3–1/9, which decreased linearly with an increase in the size of the QD and the angular momentum. This implies the formation of MFO vortexes and their fixed radius, which is limited by the Bohr radius. Thus, we revealed a critical role of localization in the coupling of the electron and MFO-vortexes, resulting in the formation of a magneto-electron (em) particle. Its formation can be explained by the nano-superconducting properties of the electrons occupying quantized states in the Wigner localization regime and the self-generation of vortexes. The fractional charge of such em appears to be due to vortex screening of the electric field. Our observation shows that Laughlin’s k-WZs are related to close-packed topological vortex structures of em, and provides an explanation for FQHE magneto-conductivity plateaus in terms of self-confinement of em’s in a quasi-ordered potential modulation, which, in turn, opens up new possibilities of deterministic engineering of anyon particle properties for use in TQGs.

2. Single-Electron InP/GaInP2 QDs

The measured InP/GaInP2 QDs (see Appendix A.1 for growth details) had a flat lens shape, a lateral size D~50–180 nm, up to 20 electrons and a Wigner-Seitz radius of up to rs~5, as described in [22,25]. The dots revealed a small elongation and asymmetry, which, in most cases, could be described as a combination of a ~5% elliptical distortion (D||/D⊥ = 1.05) and a 10% change of R⊥. The dots had up to ~0.2 content of Ga (xGa) due to intermixing during growth, which resulted in an increase of the emission energy of the 00 transition E0 (see below) on ~150 meV.
The InP QDs are surrounded by atomically ordered GaInP2 20–50 nm size domains (AODs) having a CuPtB-type crystal structure, which is a [111]B-oriented GaP/InP monolayer superlattice [22]. Thus, they represent core-shell QD-AOD composites. Relaxation of the bonds of the domain was observed for some dots and resulted in an 80 meV decrease of E0. The domains generates a strong [111]B-oriented piezo-electric field, which produces electron doping of QDs and induces a built-in magnetic field (Bbi) [21,26].
The structures studied had a relatively large dot density (~109 cm−2) and we used an optical fiber probe cryo-magnetic near-field scanning optical microscope (NSOM) with a spatial resolution of 20-300 nm to measure the PL spectra of a single QD (see Appendix A.2 for details of NSOM technique). Since the QDs had a wide range of sizes, electron populations and Bbi, the measurements presented certain technical challenges (including luck), the main one being the tedious scanning process to find the appropriate dot for a specific study, and separating its spectrum from neighboring dots. In our experiments, QDs with different electron populations were selected using analysis of the number of spectral features in PL spectra, as described in detail in [25]. The spectra were measured at 10 K and an external magnetic field Be = 0–10 T.
In our NSOM measurements of InP/GaInP2 structures, we found nearly ten single-electron QDs from a set of ~200 dots. Here, we present the data for eight dots, denoted S00, S02, S03, S04, S09, S10, S11 and S12. Table 1 contains a list of these dots, their type (see below) and parameters, such as, quantum confinement (ħω0), D, rs, E0, xGa, the separation of the electrons in a photo-excited state dWM, and Bbi (see [25]). The type and parameters were determined using magneto-PL spectra (see Appendix B.1 for spectra processing).
From this set, the dot S00 was the only one having weak Wigner localization (WL) (rs~1) and the smallest size (60 nm), for which a singly charged exciton (trion) formed in the photo-excited state. The magneto-PL spectra of this dot are discussed in the Supplementary Materials (S1). The spectra were similar to those of the 2e excitonic (tetron) dot discussed in detail in [27], with a paramagnetic (negative) B-dispersion.
For the rest of the dots, which had stronger Wigner localization (rs > 2), the trion becomes unstable decomposing into a 2e Wigner molecule (WM) weakly bound to the photo-excited hole. Two of the dots, S02 and S09 had rs~2.5 (D < 100 nm), and were thus in the regime of the onset of WL; the rest had rs~4 (D > 100 nm) and were thus in the regime of strong WL.
The change of D was from 85 nm (for S02) to 150 nm (for S04), which corresponded to a change in ħω0 from 2.5 to 0.6 meV. The xGa varied from 0.1 for S11 and S12 to 0.2 for S10 with xGa~0.15 for the five remaining QDs. This corresponded to a change of emission energy E0 from ~1.7 to 1.8 eV.

3. Magneto-PL Spectra: Experiment

Figure 1a compares the spectra of all seven WM QDs at Be = 0 T plotted in Stokes shift energy units. As can be seen from Figure 1a, a zero-energy peak, denoted 00, and a few Stokes peaks, denoted 0N, which are the signature of 2e-WM emission [27], were observed for all dots. For dots S02, S09 and S10, only a single 01 Stokes component (SC) peak was clearly observed in the spectra, while N > 1 SCs had an order-of-magnitude smaller intensity. We denoted these single SC dots as S- type. For the rest QDs up to four SCs were observed; their intensities were larger or the same as that of 00 component. We denoted these multiple peak SC dots as S*-type. Note that the dots S11 and S03, which had the same ħω0, had nearly the same intensity distribution of 0N components, in which 02 and 03 components were dominant. They differed only by the emission energy, i.e., xGa (see Table 1).
Figure 1b presents magneto-PL spectra of the S11 and S12 QDs measured in the Be range 0–10 T. It shows that the S11 and S12 QDs were observed at the same NSOM spectra, i.e., probe position, and thus, neighbored one another, being separated by~200 nm. Their spectra partially overlapped (at 1.696 eV for Be = 0 T) and were separated using Lorentzian decomposition. At Be = 0 T, the spectrum of the S11 dot also partially overlapped at 1.688 eV with the spectrum of another QD (D01m) with eight electrons and ħω0 = 3.5 meV. This dot will be discussed elsewhere.
The intensity of the spectra of the S11/S12 dots versus Be strongly increased/decreased, most probably due to the lateral shift of the NSOM probe toward the S11 dot.
Thus, for Be > 7 T, the spectral shape of S12 QD could hardly be analyzed, but the position of the center of gravity of the spectra could be clearly traced.
When Be increased, the S*-type peaks of the S11 dot merged into two at 4 T and then into a single peak at 6 T. We denote these types of spectra as D and P, respectively. A further increase in Be led to a splitting of the P-type spectrum and its transformation into the S-type spectrum at Be = 10 T. For the S12 dot, a P-type single peak shape seemed to appear at Be = 4 T and 8 T, revealing some intermediary fine structure. For the QD S04 (see [27] and Figure 5a below), the P- and D-type spectra seemed to appear at 1 T and 4 T, respectively, but no S-type spectrum was observed over the entire range of Be 0–10 T.
The key observation from the comparison of Figure 1a,b is that, in the spectrum of the QD S11 at Be = 10 T and in the that of the QD10 at Be = 0 T, the peak separation and their intensity distribution were nearly the same. The only difference was a slightly larger full-width-at-half-maximum (1.6 versus 1.2 meV), which may have been due to the larger value of xGa in S10 (see Table 1). These features, which indicate that these are nearly identical dots, are clearly seen in Figure 1c, in which we compare their spectra. In Figure 1d, we compare an NSOM image of the S10 dot (see [25]), dWM estimated from the experimental spectrum (see below) and the calculated electron density for this dot for Be = 0. The measured data and the calculation are in excellent agreement. This indicates that QD11 had zero internal magnetic field B at Be = 10 T. This corresponded to Bbi directed opposite to Be and equaled −10 T. Thus, we can conclude that the S*-type spectrum is a signature of Bbi and the fact that we did not observe S-type spectra for QD S04 corresponds to Bbi > 10 T.
The total internal field in the QD is B = Be + Bbi, and a negative (positive) shift of the spectral components versus Be corresponded to a positive (negative) shift versus B. This shift, outlined in Figure 1b by dashed lines connecting the peak maxima, revealed a weak positive B-dispersion in the S*-range and a negative B-dispersion in the S-, P- and D-ranges (see also the more detailed analysis in Figure 5a–c below). The average value of the dispersion of the P-spectra was a few times stronger than that of the S and S*, namely ~0.1 meV/T. For S04, a similar B-dispersion was observed [27]. The dispersion showed a dependence on the dot size, which may clearly be seen in Figure 1b by comparing the slope of the dashed lines of the P-type spectra of the S11 and S12 dots. The slope corresponds to a dispersion ~0.5 meV/T for S11 (D = 130 nm) and ~1 meV/T for S12 (D = 110 nm), thus increasing with a decrease in the dot size.

4. Magneto-PL Spectra: Phenomenology Description

4.1. Intensity Distribution of the Stokes Components

In Figure 2a, we show a cartoon of the confinement potential and energy levels, together with a cross-section of the corresponding wave-functions ψN(r,θ) and electron positions in the initial (IS) and photo-excited (PS) states of a single-electron QD having rs = 3.8 (analog of S11). In the left and lower inserts, we show the calculated electron density distributions for the PS and the six lowest ISs (see S2). Note that in Figure 2a and elsewhere, for PL transitions (see vertical arrows in the right), we use the notations 0N, where two numbers are radial quantum numbers (see Appendix B.2) of the PS and IS states, respectively, and, for single-electron states having the same angular momentum, the notations Ns, Np and Nd, where N’ = N + 1, to number the states. It can be seen from Figure 2a that in the dot, a 2e-WM with dWM = 70 nm formed by photon absorption generating an electron-hole (e–h) pair. An e-h recombination, resulting in photon emission corresponding to a transition between the lowest PS-IS 1s-states, which have total angular momentum Lz = 0, left the residual 1s electron in a position shifted from that of the IS by a dWM/2. This shift broke the selection rule ΔN = 0 for an optical transition, and thus generated 0N (N > 0) PL SCs with an energy separation 2ħω0. The relative intensity of 0N PL transitions was equal to the ratio of the square of the overlap integrals O I N ( d ) = ψ 0 ( r + d , θ ) ψ N ( r , θ ) d r d θ for d= dWM/2. In Figure 2b,c we present numerically calculated ratios [ O I N ( d ) / O I 0 ( d ) ] 2 for B = 0 and B~7 T (see below) and corresponding PL spectra for d = dWM/2, respectively. These ratios are also shown as bars in Figure 1a. The comparison shows good agreement with the experiment. The enhanced intensity of higher N components in S*-dots was therefore due to Bbi and a corresponding increase of electron localization and Lz (see below).
For larger Lz, Npy and N’dy states appeared in the PL spectra. For these states, only zero Stokes energy components 1py and 1dy were expected, owing to the spatial matching between their and the 1s2 PS maxima of the electron density distributions (see inserts in Figure 2a). The 1px and 1dx states were expected to have nearly zero intensity, owing to the opposite phase of the upper and lower maxima, leading to a zero overlap integral. As we will show below, a single-peak P-type spectrum can be attributed to transitions involving the 1dy component of Lz = 1 PS and a second peak in the D-type spectra to the 1py component Lz = 3 of PS. The S*-type spectra can be attributed to transitions involving the 1s component of the Lz = 3 or 5 states.

4.2. Magic Lz Numbers of the Photo-Excited State

The increase of Lz in a few-electron 2D WMs versus B occurred via certain values known as “magic” numbers LzMN, which depend on the WM configuration [28]. LzMN appeared because of the necessity of matching the spatial structure of the electron wave-functions and the spatial symmetry of the electron arrangement, and resulted from the competition between the Coulomb interaction and quantum and magnetic field confinement [29]. For 2e-WM, LzMN are odd, i.e., Lz = 1, 3, and so on. They correspond to single electron PS configurations 1s + 1px, 1py + 1dx, 1dy + 1fx, … for Lz = 1, 1px + 1dx, 1s + 1fx, … for Lz = 3 and so on.
In Figure 3, we present the results of our calculations (see S3 for details) of the average orbital angular momentum 〈Lz〉 = Lz of the 2e-WM for an InP/GaInP2 QD having D = 150 nm (rs = 4.3), corresponding to S04, versus B for B = 0-5 T. The inserts show the electron density distribution for B = 0, 1.0, 2.25, 3.0, 4.25 and 5 T. Results for smaller rs are presented in S3.
Broadened plateaus at values of LzMN =1 and 3 centered at B1~0.5 T and B3~1.25 T are seen in Figure 3. These plateaus were washed out for LzMN = 7 and higher, after which Lz showed a near-linear dependence on B. The plateaus shifted to higher B as D (rs) decreased, and for D = 110 nm (rs =3.4), they were centered at fields that were two times larger. Assigning the P-, D- and S*-type spectra in Figure 1b to Lz = 1, 3 and 5, the experimental BLz fields (1.5, 4 and 8 T) seemed to be two times larger than the calculated ones (~0.7, 2 and 3 T). This may have indicated some mechanism for reducing the effective field acting on an electron, related to vortex formation (see below).
The electron density distributions in Figure 3 show a strong decrease of the size of single-electron maxima De versus B, which was nearly by a factor of two in the range from 0 to 5 T (~50 nm versus 30 nm). At the same time, dWM did not change. Thus, the enhancement of higher-order SCs in the PL spectrum in S*-dots was related to a decrease in De. Such a decrease can be formally described as an increase of the effective mass, since De~ 1/m*2. In Figure 2b, we present calculated OIN2 ratios using a nine-fold increase of m* corresponding to B~7 T. The calculations yielded an S*-type order-of-magnitude increase of SCs having N > 1, confirming that these dots had strong Bbi (see also calculated spectra in Figure 2c).

4.3. Magnetic Field Shifts

For the analysis of the B-dispersion of PL spectral components, we used the following expression for the emission energy of the ground state of a 2e-WM weakly bound to a hole (2eWM-h) in a parabolic potential [27]:
E2eWM-h(B) = E02eWM-h(0) + EC,2e(B) + Er.m.,2e(B) +ħω0h(B)
where the first term is the “free” WM energy, the second term is the energy of the Coulomb interaction between the two electrons, the third term is the energy of the relative motion of the two electrons and the fourth term is the cyclotron energy of the hole (see Appendix B.2). In expression (1), we take advantage of the fact that cyclotron energy of center-of-mass c.m. motion does not depend on the number of electrons [30], i.e., Ec.m.,2e(B) = ħω0e(B) + const. We also use the fact that mh* > >me*, and we neglect the energy of the relative motion of electron and hole. We further assume, for simplicity, that the energy of the electron-hole interaction does not depend on B. The second term in (1) represents the difference between the energy of the 2e state of interacting and noninteracting electrons and it is approximately given by EC,2e(B) = 0.1 ω c m * / ( ω c 2 +   4 ω 0 2 4 ) ) (see [31]). The third term gives LzMN transitions discussed above and makes much smaller contribution (see S3). Thus, the B-dispersion of the 00 transition (i.e., 1s state) is
E00(B) = E2eWM-h(B)-E02eWM-h(0)≈ħω0h(B),
accounting for the fact that ħω0h(B) ≈ 2EC,2e(B). For Ns, px,y and dx,y transitions, we find from the Fock-Darwin spectrum (see Appendix B.2)
E0N(B) ≈ ħω0h(B) − (2N + 1)ħω0e(B),
Epx,py(B) ≈ ħω0h(B) − 2ħω0e(B) ± 1/2ħωce (B),
Edx,dy(B) ≈ ħω0h(B) − 3ħω0e(B) ± 1ħωce(B),
In Figure 4 we present a calculation of the energies given by Equation (3) for an electron charge e* = 1 and 1/7, together with the experimental points for S* N’s peaks of S11 dot at B = 8 T, at which positive dispersion begins. The energies show a strong negative B-dispersion (~1 meV/T) for e* = 1 resulting, from the small value of m*e and the fact that m*e<<m*h, i.e., a large ωce and ωce >>ωch. Such large dispersion also have 1p and 1d states presented in Figure 4 too. For e* = 1/7, a negative dispersion decreases down to −0.1 meV/T, owing to the decrease of ωce to values close to ωch.
The positive dispersion of the S*-region indicates ωce < ωch and for e* = 1/7 results in a more than three-fold decrease of m*h (see example in Figure 4 for B> 8 T), which can indicate heavy-light hole mixing. A qualitative comparison of B-dispersions presented in Figure 4 with Figure 1b for B > 7 T shows that matching the peaks of the S- and S*-type spectra is achieved only for e* = 1/7, thus indicating the observation of the fractional charge of single-electron states. The appearance of the S*-type spectra implies a recombination involving e-h states having lz = LzMN = 3 or 5, leaving the residual electron in the 1s state. Thus, positive dispersion is a signature for the mixing of the holes of the states having large angular momentum.

5. Magneto-Electrons

5.1. Size-Dependence of Fractional Charge

A quantitative comparison of B-dispersions and a fit using Equation (3) are shown for dots S04, S11 and S12 in Figure 5a–c, respectively. This gives fractional charge values 1/k, where k is the number of vortexes, versus D (see Table 2). In the figure, the PL spectra are shown as contour plots, on which peak positions (circles) and fitted curves were superimposed. Note that the fit includes Bbi, which is -10 T for the S11 and S12 and −15 T for the S04 QDs.
The fit makes it possible to consistently match all observed peaks to single-electron ISs N’s, 1py and 1dy in these dots. We found that the fitting procedure was robust (in terms of self-consistency) against variations of level sequence, 1/k and Bbi.
The most straightforward fitting was obtained for S11 (see Figure 5b), which had a complete data set, in terms of the measured B-range, peak resolution and intensity. For this dot, the two low-field oscillations of the 1s-peak, which had an amplitude/period 0.3 meV/2T, could be attributed to switching between 1s and 1px/1py states; formation of P-type spectra could be attributed to the contribution of the 1dy state and its crossing with the 2s state at B = 3 T; the formation of the D-type spectra could be attributed to the contribution of the 1py at B > 5 T; and the formation of S*-type spectra can be attributed to a positive dispersion at B > 7 T.
Almost all these features were resolved in the S04 dot (see Figure 5a), accounting for the limited B-range due to the large Bbi. For the S04 dot, however, the 2s state was not resolved, probably because of an overlap with the 1dx state, and S*-type peaks were smeared out/crossed because of the smaller energy separation.
For the S12 dot, which had a limited set of identified peaks due to low PL intensity, it was possible to resolve (see Figure 5c) the following: the crossing between the 2s and 1dy states, manifesting the onset of P-type spectra at B = 3 T; the transition between D- and S*-type spectra at B = 6 T; and the positive B-dispersion of the 3s and 4s peaks in the S*-range.
The measured fractional charge (see Table 2) was 1/9, 1/7 and 1/5 for the N’s states, 1/7 and 1/5 for 1p states and 1/5, ¼ and 1/3 for 1d states. This decreased with D and angular momentum. The number of vortexes versus D is shown in Figure 6 and reveals a linear dependence in the measured range. Also, for D = 150 nm, it shows a linear decrease versus lz.
Note that the fit included a matching between N’s states in S- and S*-type spectra and revealed a suppression of the 1s (00) component in the S*-type spectra, which means that the peaks denoted 0N at Be = 0 T in Figure 1b were actually 0(N + 1) peaks, as shown in Figure 5a,b. This corresponded to a much smaller “effective” De (see lower plot in Figure 1b) due to vortexes and related WZs (see inserts in Figure 6 below).

5.2. Magneto-Electrons and Fractional Quantum Hall Effect

Our results presented in Figure 5a–c and Figure 6 show self-generation of MFQ vortexes by a single electron in the strong Wigner-localization regime and imply a nano-superconducting characteristic of its quantum states. We denote such states as magneto-electron states or simply as magneto-electrons – emks. Note that for lz > 0, vortexes of opposite directions were generated for the positive and negative parts of wave-functions, which was thus related to Majorana zero energy modes.
The k(D) dependence implies a critical size D0 for which fractional charge can be generated, corresponding to k = 2. This, in turn, implies hard-wall vortexes, i.e., the existence of a fixed vortex size dv and close-packed spatial structures. Assuming conservation of linear k(D) dependence for D < 100 nm and an effective size of the electron density generating vortexes equal to Dv = D/2, we could estimate dvD0/4 = 20 nm or a radius rv~10 nm for the s-state. This value was close to the Bohr radius aB* of InP/GaInP2 QDs (~8 nm see Appendix B.1 and S1). Possible close-packed structures of aB*-size-vortexes for different D and single-electron states are shown in the inserts in Figure 6. The observed decrease of k versus lz may have been due to the reduction of Dv due the spatial “splitting” of electron density.
The vortexes of ems represented current loops and could be considered as a “bound” charge part of the electron wave function. Thus, a fractional charge 1/k appeared, owing to the vortex screening the electric field. In InP/GaInP2 QDs, self-formation of ems generating vortexes and Bbi may have been due to the dissipationless electron motion in a lateral electric field produced by the AOD GaInP2 shell. Thus, one can assume that for the S-type S10 dot, which had nearly the same D as the S*-type S11 and S03 dots, the lateral field was absent and Bbi was not generated. This may have been due to the different GaInP2-shell structure, in which AODs were weakly ordered because of large xGa.
We should point out that different vortex-type single-electron eigenstates were considered theoretically in billiard models of square and related shape geometries in magnetic fields [32]. In this case, these states are excited states and vortex solutions arise because of the matching of the spatial geometry of the eigenstates to the shape of the confining potential. These vortex eigenstates appeared to be similar to the vortexes in mesoscopic superconductors of similar shape [33], which link these systems to em and give rise to the notion of the superconducting 2D “droplet” formed by a single-electron quantum state.
For rs~4, the FQHE regime begins at B~0.2 T, corresponding to ν = 1, where ν = 4h/(πD2)/e0/B is effective Landau level filling factor, assuming uniform charge distribution within QD. The formation of emk has thresholds B~0.2k T corresponding to ν = 1/k and Dk~kaB*. This makes it possible to describe the FQHE, i.e., the appearance of corresponding conductance plateaus in quantum Hall (QH) bar structures, using a localization of emks in the disorder-induced potential fluctuations. The key effect in such a description is a quasi-ordering of the fluctuations [34], which is analogous to spontaneous symmetry breaking induced by quantum monitoring [35]. This results in a spatial modulation having a half-period equal to the spatial size kaB*. This implies a self-confinement of emks at the corresponding ν. Therefore, localized/delocalized emks exist in the range between 1/k and 1/(k +1). In this range, delocalized emks provide edge hopping conductivity. For IQHE, Wigner-crystal-type network patterns of the localized electron states, observed in a scanning electrometer probe experiment for ν = 1 [36], naturally emerge from such self-confinement.
Within this picture, the alternation of ¼ and ½ charge versus size of the resonator, i.e., the number of electrons, observed in Aharonov-Bohm phase interferometry measurements of quantum Hall (QH) bars for ν = 5/2 [14], can be explained by alternation of the size of em by the Majorana modes.

5.3. Topological Quantum Computing

While in our experiment, the effective density of electrons was ~1010 cm−2, the density used in QH bars for experiments aiming to realize TQGs [16] was nearly twenty times larger [14]. At such high density, the anyon emks are hard to control locally, making it difficult to measure their statistics [17] and perform elementary topological quantum computing operations, i.e., anyon interchange (braiding) and pairing (fusion) [16]. The same control seems more manageable at low densities; we have already demonstrated optically induced pairing and the interchange of ems in the same InP/GaInP2 QDs, but with six emks [21]. Moreover, owing to Bbi, these were done at zero magnetic field and at relatively high temperature (10 K), which was more practical. Also, in InP/GaInP2 QDs, the properties of emk anyons could be controlled by Ne, D, QD shape and Bbi. For InP/GaInP2 QDs, TQG could be realized using adjacent nano-electrodes attached to single electron transistors [20]. Such a chip made it possible to move emks around each other and to control the local charge to measure the initialization and manipulation of q-bits.

6. Conclusions

In conclusion, using measurements of the magneto-PL spectra of single-electron InP/GaInP2 quantum dots having Wigner-Seitz radius ~4, we demonstrated the existence of a magneto-electron state having a fractional charge. This is the anyon state, in which a few magnetic flux quanta vortexes are self-generated by a single electron in Wigner localization regime, providing nano-superconducting properties of quantum states. The emergence of the magneto-electron corresponds to the three- to nine-fold reduction of the cyclotron frequency of single electron Fork-Darwin states and its dependence from quantum dot size measured. These imply a close-packed arrangements of the magnetic-flux vortexes and make it possible to describe the fractional quantum Hall effect in terms of magneto-electron self-confinement in the disorder-induced potential fluctuations. Such self-confinement gives rise to the possibility of engineering of anyon properties (statistics) of magneto-electrons, and could be used to build anyon-based fault-tolerant topological quantum gates.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/11/2/493/s1. S1: Trion versus exciton emission in S00 QD. S2: Electron states of InP/GaInP2 QDs versus size. S3: Theoretical modeling of electronic structure in magnetic field.

Author Contributions

A.M., D.L. and A.V. conceived the magneto-PL NSOM experiment. A.B. and S.R. made calculations of the overlap integrals and S.B. made calculations of electronic structure in magnetic field. The manuscript was written A.M. with the input from all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Russian Science Foundation grant number #19-19-00246.

Data Availability Statement

The data that supports the findings of this study are available within the article and its Supplementary Material.

Acknowledgments

We wish to thank G. Snider and A. Orlov for discussions of using single electron transistors for topological quantum gate proposal.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Appendix A.1. Structure Growth

The samples were grown on a [001] oriented GaAs substrate by a metal-organic chemical vapor deposition under 100 mbar pressure at a temperature of 725 °C [25]. The dots are formed by deposition of 3–7 monolayers of InP on 40–500 nm thick Ga0.52In0.48P (or simply GaInP2) lattice matched to the GaAs and capped by a 40–60 nm thick GaInP2 layer. Their detailed characterization including optical (steady-state and time-resolved photoluminescence, photo-reflection, Raman spectroscopy, conventional and nano-indentation near-field scanning optical microscopy) together with and structural (cross-section and plan-view transmission electron microscopy) is presented in Refs. [25,26,37,38].

Appendix A.2. Measurement Technique

A single dot magneto-PL spectra were measured using a near-field scanning optical microscope (NSOM) operating at 10 K and magnetic fields of up to 10 T. We used tapered fiber probes coated with Al, having an aperture size of 50–300 nm in a collection-illumination mode. The spectra were excited by the 514.5 nm Ar-laser line and measured using a CCD (multi-channel) detector together with a 280 mm focal length monochromator. The excitation power measured before the fiber coupler was ~5 μW, which provided a power density of ~0.5 W/cm2. The spectral resolution of the system is 0.2–0.4 meV.

Appendix B

Appendix B.1. Spectra Processing

For the Lorenz contour deconvolution of the spectral shape, we used a multi-peak fitting procedure from Origin 8.0 graphic software.
The values of ħω0 were estimated from the splitting of the Stokes peaks in the PL spectra, and D from charge-density distributions calculated for specific ħω0 using the configuration-interaction approach (see [21]) or from the NSOM images. This allows one to calculate rs = 4N/πD2aB*, where aB* is the Bohr radius of In(Ga)P; or in effective atomic units, Bohr radius aB * = ħ2(4πεε0)/m*e2, where ε and m* are the dielectric constant of the material and the effective mass of the electron, respectively. For our dots it is 8.7 nm for xGa = 0 (see also S1).

Appendix B.2. Fock-Darwin States

We used the Fock-Darwin spectrum [30] to fit observed shifts of PL peaks versus intra dot magnetic field B using Equation (3):
ε i   ( n , m ) = ħ ω i ( B ) ( n + 1 ) 1 2 ħ ω c i m ,
where n = 0, 1, … is the principal quantum number; m = −n, −n + 2, …, n − 2, n is the azimuthal quantum number, which is equal to angular momentum (lz); N = (n-|m|)/2 is the radial quantum number; ωci = e*B/m*i, ω0i(B) = (ω0i2 + ωci2/4)0.5 and i marks electron (e) or hole (h) having corresponding charge e* and effective mass m*e/m*h =0.08me/0.56me.
In S2 we present electron density distributions of six N’m single electron states having N’ = N+1 = 1, 2 for D = 110, 130 and 150 nm (ħω0 = 0.6, 0.8 and 1.25 meV). These states are the 1s, 1px, 1py and 1dx, 1dy and 2s shells (levels), having |lz| = 0, 1, 1, 2, 2 and 0, respectively. Owing to the circular asymmetry, the 1px, 1py, 1dx, 1dy and 2s levels are split on ~0.25ħω0, and the 1dy and 2s states are mixed.

References

  1. Klitzing, A.K.; Dorda, G.; Pepper, M. New Method for High-Accuracy Determination of the Fine-Structure Constant Based on Quantized Hall Resistance. Phys. Rev. Lett. 1980, 45, 494–497. [Google Scholar] [CrossRef] [Green Version]
  2. Tsui, D.C.; Stormer, H.L.; Gossard, A.C. Two-Dimensional Magnetotransport in the Extreme Quantum Limit. Phys. Rev. Lett. 1982, 48, 1559–1562. [Google Scholar] [CrossRef] [Green Version]
  3. Kosterlitz, M.; Thouless, D.J. Long range order and metastability in two dimensional solids and superfluids. (Application of dislocation theory). J. Phys. C 1972, 5, L124–L126. [Google Scholar] [CrossRef]
  4. Thouless, D.J.; Kohmoto, M.; Nightingale, M.P.; den Nijs, M. Quantized Hall Conductance in a Two-Dimensional Periodic Potential. Phys. Rev. Lett. 1982, 49, 405–408. [Google Scholar] [CrossRef] [Green Version]
  5. Wilczek, F. Magnetic Flux, Angular Momentum, and Statistics. Phys. Rev. Lett. 1982, 48, 1144–1146. [Google Scholar] [CrossRef]
  6. Sarma, S.D.; Freedman, M.; Nayak, C. Majorana zero modes and topological quantum computation. NPJ Quantum Inf. 2015, 1, 15001. [Google Scholar] [CrossRef] [Green Version]
  7. Kitaev, A. Fault-tolerant quantum computation by anyons. Ann. Phys. 2003, 303, 2–30. [Google Scholar] [CrossRef] [Green Version]
  8. Laughlin, R.B. Quantized Hall conductivity in two dimensions. Phys. Rev. B 1981, 23, 5632–5633. [Google Scholar] [CrossRef]
  9. Haldane, F.D. Fractional Quantization of the Hall Effect: A Hierarchy of Incompressible Quantum Fluid State. Phys. Rev. Lett. 1983, 51, 605–608. [Google Scholar] [CrossRef]
  10. Laughlin, R.B. Anomalous Quantum Hall Effect: An Incompressible Quantum Fluid with Fractionally Charged Excitations. Phys. Rev. Lett. 1983, 50, 1395–1398. [Google Scholar] [CrossRef] [Green Version]
  11. Jain, J.K. Composite-fermion approach for the fractional quantum Hall effect. Phys. Rev. Lett. 1989, 63, 199–202. [Google Scholar] [CrossRef]
  12. Rotondo, P.; Molinari, L.G.; Ratti, P.; Gherardi, M. Devil’s Staircase Phase Diagram of the Fractional Quantum Hall Effect in the Thin-Torus Limit. Phys. Rev. Lett. 2016, 116, 256803. [Google Scholar] [CrossRef] [Green Version]
  13. Dolev, M.; Gross, Y.; Chung, Y.C.; Heiblum, M.; Umansky, V.; Mahalu, D. Dependence of the tunneling quasiparticle charge determined via shot noise measurements on the tunneling barrier and energetics. Phys. Rev. B 2010, 81, 161303. [Google Scholar] [CrossRef] [Green Version]
  14. Willett, R.L.; Pfeiffer, L.N.; West, K.W. Alternation and interchange of e/4 and e/2 period interference oscillations consistent with filling factor 5/2 non-Abelian quasiparticles. Phys. Rev. B 2010, 82, 205301. [Google Scholar] [CrossRef]
  15. Fu, H.; Wang, P.; Shan, P.; Xiong, L.; Pfeiffer, L.N.; West, K.; Kastner, M.A.; Lin, X. Competing ν = 5/2 fractional quantum Hall states in confined geometry. Proc. Natl. Acad. Sci. USA 2016, 113, 12386–12390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Bonesteel, N.E.; Hormozi, L.; Zikos, G.; Simon, S.H. Braid Topologies for Quantum Computation. Phys. Rev. Lett. 2005, 95, 140503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Sarma, S.D.; Freedman, M.; Nayak, C. Topologically Protected Qubits from a Possible Non-Abelian Fractional Quantum Hall State. Phys. Rev. Lett. 2005, 94, 166802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Macaluso, E.; Comparin, T.; Mazza, L.; Carusotto, I. Fusion Channels of Non-Abelian Anyons from Angular-Momentum and Density-Profile Measurements. Phys. Rev. Lett. 2019, 123, 226801. [Google Scholar] [CrossRef] [Green Version]
  19. Martin, J.; Ilani, S.; Verdene, B.; Smet, J.; Umansky, V.; Mahalu, D.; Schuh, D.; Abstreiter, G.; Yacoby, A. Localization of Fractionally Charged Quasi-Particles. Science 2004, 305, 980–983. [Google Scholar] [CrossRef] [Green Version]
  20. Venkatachalam, V.; Yacoby, A.; Pfeiffer, L.; West, K. Local charge of the ν=5/2 fractional quantum Hall state. Nature 2011, 469, 185–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Mintairov, A.M. Molecular States of Composite Fermions in Self-Organized InP/GaInP Quantum Dots in Zero Magnetic Field. Semiconductors 2020, 54, 138–143. [Google Scholar] [CrossRef]
  22. Mintairov, A.M.; Lebedev, D.V. Atomic ordering and bond relaxation in optical spectra of self-organized InP/GaInP2 Wigner molecule structures. Appl. Phys. Lett. 2019, 115, 202104. [Google Scholar] [CrossRef]
  23. Michler, E. Single Semiconductor Quantum Dots; Springer: Berlin/Heidelberg, Germany, 2009; p. 390. [Google Scholar]
  24. Self-Assembled Quantum Dots; Wang, Z.M. (Ed.) Springer: Berlin/Heidelberg, Germany, 2008; p. 468. [Google Scholar]
  25. Mintairov, A.M.; Kapaldo, J.; Merz, J.L. Control of Wigner localization and electron cavity effects in near-field emission spectra of In(Ga)P/GaInP quantum-dot structures. Phys. Rev. B 2018, 97, 195443–195449. [Google Scholar] [CrossRef]
  26. Mintairov, A.M.; Merz, J.L.; Blundell, S. Molecular States of Electrons: Emission of Single Molecules in Self-Organized InP/GaInP Quantum Dots. In Fingerprints in the Optical and Transport Properties of Quantum Dots; IntechOpen Limited: London, UK, 2012; pp. 125–152. [Google Scholar]
  27. Mintairov, A.M.; Kapaldo, J.; Merz, J.L.; Vlasov, A.S.; Blundell, S.A. Wigner molecules and charged excitons in near-field magnetophotoluminescence spectra of self-organized InP/GaInP quantum dots. Phys. Rev. B 2017, 95, 11544–11610. [Google Scholar] [CrossRef]
  28. Ruan, W.Y.; Liu, Y.Y.; Bao, C.G.; Zhang, Z.Q. Origin of magic angular momenta in few-electron quantum dots. Phys. Rev. B 1995, 51, 7942. [Google Scholar] [CrossRef] [Green Version]
  29. Phannkuche, D.; Gernhardts, R.R.; Maksym, P.A.; Gudmundson, V. Theory of quantum dot helium. Phys. B 1993, 189, 6–15. [Google Scholar] [CrossRef]
  30. Jacak, L.; Hawrylak, P.; Wojs, A. Quantum Dots; Springer: Berlin/Heidelberg, Germany, 1998; p. 176. [Google Scholar]
  31. Kyriakidis, J.; Pioro-Ladriere, M.; Ciorga, M.; Sachrajda, A.S.; Hawrylak, P. Voltage-tunable singlet-triplet transition in lateral quantum dots. Phys. Rev. B 2002, 66, 035320. [Google Scholar] [CrossRef] [Green Version]
  32. Chibotaru, L.F.; Ceulemans, A.; Lorenzini, M.; Moshchalkov, V.V. Vorticity quantum numbers for confined electrons. Europhys. Lett. 2003, 63, 159–165. [Google Scholar] [CrossRef]
  33. Chibotaru, L.F.; Ceulemans, A.; Bruyndoncx, V.; Moshchalkov, V.V. Symmetry induced formation of antvortexes in mesoscopic superconductors. Nature 2000, 408, 833–835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Mintairov, A.M.; He, Y.; Merz, J.L.; Jin, Y.; Goldman, R.S.; Kudrawiec, R.; Misiewicz, J.; Akimov, I.A.; Yakovlev, D.R.; Bayer, M. Quasi-ordering of composition fluctuations and their interaction with lattice imperfections in an optical spectra of dilute nitride alloys. Semicond. Sci. Technol. 2016, 31, 095012. [Google Scholar] [CrossRef]
  35. García-Pintos, L.P.; Tielas, D.; Campo, A. Spontaneous Symmetry Breaking Induced by Quantum Monitoring. Phys. Rev. Lett. 2019, 123, 090403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Zhitenev, N.B.; Fulton, T.A.; Yacoby, A.; Hess, H.F.; Pfeiffer, L.N.; West, K.W. Imaging of localized electronic states in the quantum Hall regime. Nature 2000, 404, 473–476. [Google Scholar] [CrossRef] [PubMed]
  37. Mintairov, A.M.; Sun, K.; Merz, J.L.; Li, C.; Vlasov, A.S.; Vinokurov, D.A.; Kovalenkov, O.V.; Tokranov, V.; Oktyabrsky, S. Nanoindentation and near-field spectroscopy of single semiconductor quantum dots. Phys. Rev. B 2004, 69, 155306. [Google Scholar] [CrossRef]
  38. Kapaldo, J.; Rouvimov, S.; Merz, J.L.; Bert, N. Ga–In intermixing, intrinsic doping, and Wigner localization in the emission spectra of self-organized InP/GaInP quantum dots. J. Phys. D Appl. Phys. 2016, 49, 475301. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) Comparison of PL spectra (thick lines) S and S* 1e InP/GaInP2 QDs (see Table 1) at Be = 0 T. Thin/dashed lines are Lorentzian contour decomposition of the spectra and bars are calculated 0N transition intensities (red and green for B = 0 and ~7 T, respectively, see text). The star symbols denote the neighboring dots. (b) PL spectra of S11and S12 InP/GaInP2 QDs measured at Be = 0, 0.5, 1.0, … and 10 T. Dashed lines connect the positions of the peak maxima. (c) Comparison of PL spectra of S11 (upper) and S10 (lower). QDs measured at Be = 10 and 0 T, respectively. Dashed lines are the spectrum and its constituents obtained from Lorentzian contour decomposition. The star symbols denote the neighboring dot. (d) Measured NSOM PL intensity distribution (left) and calculated 2e-WM electron density (right) for the S10 QD. The dumbbell in the left image corresponds to dWM, as-measured from the experimental PL spectrum.
Figure 1. (a) Comparison of PL spectra (thick lines) S and S* 1e InP/GaInP2 QDs (see Table 1) at Be = 0 T. Thin/dashed lines are Lorentzian contour decomposition of the spectra and bars are calculated 0N transition intensities (red and green for B = 0 and ~7 T, respectively, see text). The star symbols denote the neighboring dots. (b) PL spectra of S11and S12 InP/GaInP2 QDs measured at Be = 0, 0.5, 1.0, … and 10 T. Dashed lines connect the positions of the peak maxima. (c) Comparison of PL spectra of S11 (upper) and S10 (lower). QDs measured at Be = 10 and 0 T, respectively. Dashed lines are the spectrum and its constituents obtained from Lorentzian contour decomposition. The star symbols denote the neighboring dot. (d) Measured NSOM PL intensity distribution (left) and calculated 2e-WM electron density (right) for the S10 QD. The dumbbell in the left image corresponds to dWM, as-measured from the experimental PL spectrum.
Nanomaterials 11 00493 g001
Figure 2. (a) Cross-sections of the parabolic confinement potential and energy levels of the initial (IS) and photo-excited (PS) state of a 1e-QD in the Wigner localization regime (rs = 3.8). The curves are the cross-section of the wave-functions of the corresponding energy levels and vertical arrows show a set of PL transitions involving Ns, 1px,y and 1dx,y states, the electron-density distribution of which are shown on the left and lower inserts (in the 200 × 200 nm2 frame). The thick arrow outlines a radiative recombination process and connects the photo-excited hole (smaller circle) with one of the electron (larger circles). (b) Square of the ratio of the overlap integrals OIN (see text) for B = 0 T (upper) and 7 T (lower). The dashed line shows the value of dWM/2 in (a). (c) PL spectra calculated using (b) for B = 0 T (upper) and 7 T (lower).
Figure 2. (a) Cross-sections of the parabolic confinement potential and energy levels of the initial (IS) and photo-excited (PS) state of a 1e-QD in the Wigner localization regime (rs = 3.8). The curves are the cross-section of the wave-functions of the corresponding energy levels and vertical arrows show a set of PL transitions involving Ns, 1px,y and 1dx,y states, the electron-density distribution of which are shown on the left and lower inserts (in the 200 × 200 nm2 frame). The thick arrow outlines a radiative recombination process and connects the photo-excited hole (smaller circle) with one of the electron (larger circles). (b) Square of the ratio of the overlap integrals OIN (see text) for B = 0 T (upper) and 7 T (lower). The dashed line shows the value of dWM/2 in (a). (c) PL spectra calculated using (b) for B = 0 T (upper) and 7 T (lower).
Nanomaterials 11 00493 g002
Figure 3. Average orbital angular momentum L z calculated for the ground state of an InP/GaInP2 QD having rs = 4.3. The inserts are densities for B = 0 T, 1 T, and up to 5 T in the 200 × 200 nm2 frame. The contours indicate the deformed confining potential.
Figure 3. Average orbital angular momentum L z calculated for the ground state of an InP/GaInP2 QD having rs = 4.3. The inserts are densities for B = 0 T, 1 T, and up to 5 T in the 200 × 200 nm2 frame. The contours indicate the deformed confining potential.
Nanomaterials 11 00493 g003
Figure 4. Stokes shift of N’s, 1px,y and 1dx,y states versus magnetic field – thick(thin) curve for electron charge of 1(1/7), balls—experiment for S*. Transition to positive dispersion shown for S*-range corresponds to mh*/3.
Figure 4. Stokes shift of N’s, 1px,y and 1dx,y states versus magnetic field – thick(thin) curve for electron charge of 1(1/7), balls—experiment for S*. Transition to positive dispersion shown for S*-range corresponds to mh*/3.
Nanomaterials 11 00493 g004
Figure 5. Contour plots of PL spectra of S04 (a), S11(b) and S12 QDs (c), peak positions (circles) and their Fock-Darwin spectra fit using Equation (3) versus internal magnetic field B = Be + Bbi.
Figure 5. Contour plots of PL spectra of S04 (a), S11(b) and S12 QDs (c), peak positions (circles) and their Fock-Darwin spectra fit using Equation (3) versus internal magnetic field B = Be + Bbi.
Nanomaterials 11 00493 g005
Figure 6. Number of vortexes (circles) versus dot size. Inserts show electron densities and vortex arrangement for 1s (upper, D = 110, 130 and 150 nm), 2s (lower left, D = 150 nm), 1p and 1d (right, D = 130 nm). Solid lines outline linear approximation of k(D), dashed lines show k = 2 (horizontal) and critical D for vortex formation (vertical).
Figure 6. Number of vortexes (circles) versus dot size. Inserts show electron densities and vortex arrangement for 1s (upper, D = 110, 130 and 150 nm), 2s (lower left, D = 150 nm), 1p and 1d (right, D = 130 nm). Solid lines outline linear approximation of k(D), dashed lines show k = 2 (horizontal) and critical D for vortex formation (vertical).
Nanomaterials 11 00493 g006
Table 1. Parameters of electronic states of single electron InP/GaInP2 QDs measured in magneto- PL spectra.
Table 1. Parameters of electronic states of single electron InP/GaInP2 QDs measured in magneto- PL spectra.
NoTypeħω0
(me)
D(a)
(nm)
rsE0
(eV)
xGadWM
(nm)
Bbi
T
S00trion4.060~11.7200.15-0
S02 [25]S2.5852.51.7240.15350
S09 [25]S2.0902.61.7120.15400
S10 [25]S0.71404.11.7980.2750
S12S*1.251103.41.7050.16510
S03 [25]S*0.81303.81.7650.157010
S11S*0.81303.81.6950.17010
S04 [27]S*0.61504.31.7430.158015
(a) Averaged value (between D|| and D) of the size of the area containing 96% of calculated electron density.
Table 2. Measured fractional charge of single electron states.
Table 2. Measured fractional charge of single electron states.
Quantum DotDot Size
D (nm)
Fractional Charge
1/k
N’s1p1d
S041101/91/71/5
S111301/71/51/4
S121501/5-1/3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mintairov, A.; Lebedev, D.; Vlasov, A.; Bogdanov, A.; Ramezanpour, S.; Blundell, S. Fractional Charge States in the Magneto-Photoluminescence Spectra of Single-Electron InP/GaInP2 Quantum Dots. Nanomaterials 2021, 11, 493. https://doi.org/10.3390/nano11020493

AMA Style

Mintairov A, Lebedev D, Vlasov A, Bogdanov A, Ramezanpour S, Blundell S. Fractional Charge States in the Magneto-Photoluminescence Spectra of Single-Electron InP/GaInP2 Quantum Dots. Nanomaterials. 2021; 11(2):493. https://doi.org/10.3390/nano11020493

Chicago/Turabian Style

Mintairov, Alexander, Dmitrii Lebedev, Alexei Vlasov, Andrey Bogdanov, Shahab Ramezanpour, and Steven Blundell. 2021. "Fractional Charge States in the Magneto-Photoluminescence Spectra of Single-Electron InP/GaInP2 Quantum Dots" Nanomaterials 11, no. 2: 493. https://doi.org/10.3390/nano11020493

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop