Alexander Brinkman’s research while affiliated with University of Twente and other places

What is this page?


This page lists works of an author who doesn't have a ResearchGate profile or hasn't added the works to their profile yet. It is automatically generated from public (personal) data to further our legitimate goal of comprehensive and accurate scientific recordkeeping. If you are this author and want this page removed, please let us know.

Publications (183)


Material characterization a) Symmetric (004) XRD ω‐2θ spectra of different Sn content GeSn:P structures. b) Asymmetric (224) RSM overlapped with a HR‐TEM micrograph of the Ge0.88Sn0.12 layer. The vertical and diagonal dotted lines indicate the pseudomorphic to the Ge lattice and cubic crystal lines, respectively. c) Carrier concentration depth profiles of the GeSn:P/GeSn structures. d) Calculated Γ‐ and L‐valley energies at 5K as a function of Sn concentration and lattice strain. The intersection line shows the conditions where the GeSn layers change from indirect to direct bandgap semiconductor. e) Doping dependent electron population of Γ‐valley for cubic GeSn layers with different Sn content. f) Top‐view SEM image of the Hall bar device and the measured MR (red) and Hall resistance (blue) at 2 K as a function of the applied magnetic field for undoped Ge0.86 Sn0.14 sample.
Phase coherence measurements. a) MR and c) Hall resistance as a function of the applied magnetic field at 2K for the GeSn:P samples with different Sn contents. Temperature dependence of the b) MR and d) Rxy for the Ge0.88Sn0.12 sample. Phase coherence lengths extracted from the WL fitting using the reduced HLN model e) as a function of Sn concentration and f) as a function of temperature for the Ge0.88Sn0.12 sample (HLN fitting in the inset).
SdH oscillations in n‐type GeSn. Isolated Shubnikov‐de Haas oscillations from the longitudinal resistance curves plotted as a function of 1/B for the a) 12 at.% and b) 14 at.% Sn samples, respectively. Extracted frequency of the oscillations using FFT is shown in the insets. c,d) Effective mass fitting from the Dingle equation and temperature dependence of oscillation peaks of a,b). Mass values are in free electron mass units (m0). e) Quantum scattering time fitting using the field‐dependent decay of oscillation amplitudes at different temperatures. f) Calculated effective mass at the Fermi level as a function of the n‐type doping concentration.
Phase‐Coherent Transport in GeSn Alloys on Si
  • Article
  • Full-text available

November 2024

·

45 Reads

Prateek Kaul

·

·

·

[...]

·

Germanium‐Tin (GeSn) is a novel semiconductor Group IV alloy that can be tuned from indirect to direct bandgap semiconductors by adjusting the Sn content. This property makes this alloy class attractive for integrated photonic applications and high‐mobility electronic devices. In this work, the GeSn alloy properties are investigated in the view of applications fields such as spintronics and quantum computing. Using low‐temperature magneto‐transport measurements, electron interference effects and deriving typical mesoscopic benchmark parameters such as the phase‐coherence length in GeSn‐based Hall bar structures for Sn concentrations up to 14 at.% is investigated. Furthermore, Shubnikov–de Haas oscillations provide direct access to the effective mass of the Γ‐valley electrons as well as the charge carrier mobility. This work provides a new insight into advanced group IV alloys desired for the study of spin dynamics and its quantum computing applications.

Download

Overview of normal transport in MBT flakes of thicknesses of 24.7 nm and 14.7 nm, depicted with red and blue lines respectively
a AFM images of the measured flakes. b Longitudinal resistance as a function of temperature at zero magnetic field and back gate voltage. c Longitudinal resistance as a function of applied back gate voltage. At zero back gate voltage, the flakes are in the hole regime, with increasing back gate the flakes can be gated to the electron regime. d The slope of the Hall Resistance in ∣B∣ < Bc for all back gate voltages. e Hall resistance as a function of perpendicular magnetic field for, from top to bottom, the electron, neutral, and hole regime. f Longitudinal resistance as a function of perpendicular magnetic field for, from top to bottom, the electron, neutral, and hole regime.
Overview of small current bias transport results at 50 mK in an Nb-MBT-Nb junction
a The inset shows an AFM image of the flake. b differential resistance for small current bias for different temperatures. c Differential resistance as function bias voltage and RF amplitude under irradiation of RF radiation at a frequency of 0.995 GHz, clearly indicating the presence of Josephson coupling. d Field dependence of the junction. At 3 T the MBT transition is observed from antiferromagnet to a ferromagnet. f Differential resistance map as function of magnetic field and current bias. Superconducting interference is observed, and e shows the corresponding line cuts.
Superconducting interference hysteresis
a and b show the zero bias voltage differential resistance for different values of the maximum field for two junctions. Curves are shifted for clarity. c The sum of the absolute square of the difference between the up and down sweep as a function of the magnetic field. d Schematic of the critical current relation as a function of magnetic field for a regular Fraunhofer pattern and a SQUID-like pattern. The stars indicate the magnetic fields where the supercurrent is suppressed in the case for a SQUID. The corresponding points are also indicated in a with stars.
Junction transport with large bias voltage
a Schematic of junction with large bias. The upper panel shows the concept of an induced gap in the MBT by the Nb with a possible interface indicated by the dashed lines. The lower panel schematically shows the voltage drop across the opaque (Z ≈ 0.5) and transparent (Z ≪ 0.5) interface. b Normalized differential conductance for larger biases estimated for a single interface up to 8 K. The data points show the experimental data. The solid lines show the BTK-fitted model for all temperatures. c The fitted superconducting gap Δ, quasi-particle lifetime Γ, interface parameter Z, and spin polarization P as a function of temperature. The error bars indicate the standard deviation for the fitted parameters.
Field dependence of gap feature
a Differential conductance for large bias for different magnetic fields. The data for finite magnetic fields are offset by −150 μS for clarity. b Difference between conductance at zero current and 1.5 μA bias, ΔG, as a function of the magnetic field. c Differential resistance map as function of magnetic field and voltage bias.
Josephson coupling across magnetic topological insulator MnBi2Te4

October 2024

·

116 Reads

Communications Materials

Topological superconductors hosting Majorana zero modes are of great interest for both fundamental physics and potential quantum computing applications. In this work, we investigate the transport properties of the intrinsic magnetic topological insulator MnBi2Te4 (MBT). In normal transport measurements, we observe the presence of chiral edge channels, though with deviations from perfect quantization due to factors such as non-uniform thickness, domain structures, and the presence of quasi-helical edge states. Subsequently, we fabricate superconducting junctions using niobium leads on MBT exfoliated flakes, which show an onset of supercurrent with clear Josephson coupling. The interference patterns in the superconducting junctions reveal interesting asymmetries, suggesting changes in the magnetic ordering of the MBT flakes under small applied magnetic fields. Moreover, the modulation of the critical current by magnetic field reveals a SQUID-like pattern, suggesting the presence of supercurrent through the quasi-helical edge states.



(a) Schematics of a Ge–Si core-shell nanowire Josephson device. (b) Scanning electron microscope image of a four-lead device. The junction is current biased either by I 2 − probe or I 4 − probe, and the voltage drop V is measured between the two inner leads, which have a separation of L = 100 nm. Note that the nanowire was broken due to electrostatic discharge only after the measurements.
Magnetic field dependence of the transport measured in both two-probe and four-probe configurations. Numerical differential resistance ∂ V SD / ∂ I bias vs I bias and (a) B out − of − plane for two-probe; (b) B in − plane for two-probe; (c) B out − of − plane for four-probe; and (d) B in − plane for four-probe. The in-plane field misaligns with the nanowire axis for about 25°. The black arrows indicate the current sweep direction.
Heat transport and model in a SNS Josephson junction. (a) Schematic of the two major cooling channels. P qp represents quasiparticle cooling, and P e − ph represents electron–phonon interaction induced cooling. Image is inspired by Ref. 22. (b) Quasiparticle cooling mechanism depicting the density of states of a superconductor at zero field (left) and finite field (right). (c) A heat-balanced I sw ( B ) model, including only electron–phonon cooling with parameter α, and (d) adding quasiparticle cooling with parameters β and B c bulk.
Temperature dependence of the switching current in device A in the two-probe configuration. (a) Numerical differential resistance ∂ V SD / ∂ I bias vs I bias and T bath at zero magnetic field and at (b) B in − plane = 200 mT.
Magnetic field enhanced critical current in Ge–Si nanowire Josephson junctions

July 2024

·

13 Reads

·

2 Citations

Anomalous critical current enhancement was observed with increasing magnetic field in Josephson junctions based on Ge–Si core-shell nanowires. Despite the predicted topological properties of these nanowires, which could potentially lead to non-trivial superconducting order parameter symmetries, our investigation unveils a more generalized, non-topological explanation for the observed critical current enhancement in these devices. Our findings suggest that the enhancement arises from a thermalization process induced by the magnetic field, wherein in-gap quasiparticles are generated. These quasiparticles play a crucial role in enhancing the cooling of the device, thereby lowering the effective temperature and resulting in an increased critical current. Furthermore, we elucidate how this thermalization effect varies with device geometry and measurement configuration.


(a) MnBi2Te4 and (b) VBi2Te4 have a unit cell structured as SLs separated by a VdW gap. The dashed boxes indicate the relative intercalated layers of MnTe and VTe in Bi2Te3. J|| is FM with either an (a) out-of-plane or (b) in-plane easy axis. J⊥ is AFM. (c) Bi2Te3 structured in QLs separated by a VdW gap. (d) VTe2.
(a) STEM image of a V-Bi-Te sample. The image is taken with a HAADF detector at 300 keV. A clear phase separation between bright and dark areas can be observed. (b) EDX scan of the V-Bi-Te sample. A strong separation between V-regions and Bi-regions can be observed. (c) STEM image of a smaller region on a V-Bi-Te sample. The bright areas (blue) show the QL structure of Bi2Te3 and the dark areas (orange) the VTe2 structure. (d) Histogram of the atomic distance in the x-direction. (e) 2θ-ω scans indicating (00l)-Bi2Te3 being dominant at low ϕV, while (00l)-VTe2 is dominant at high ϕV. The arrows indicate the disappearance of the respective phases as a function of ϕV. * indicates the Al2O3 substrate peak.
(a) RHEED pattern for ϕV = 1750 ∘C showing a double streak pattern related to the phases Bi2Te3 with a1= 4.31 Å and VTe2 with a2= 3.59 Å, indicated by the blue and white arrows, respectively. (b) Height distribution at the surface as a function of the ϕV. The insets show the surface morphology of the samples with ϕV = 1750 ∘C and ϕV = 1900 ∘C.
Phase Separation Prevents the Synthesis of VBi2Te4 by Molecular Beam Epitaxy

December 2023

·

58 Reads

Intrinsic magnetic topological insulators (IMTIs) have a non-trivial band topology in combination with magnetic order. This potentially leads to fascinating states of matter, such as quantum anomalous Hall (QAH) insulators and axion insulators. One of the theoretically predicted IMTIs is VBi2Te4, but experimental evidence of this material is lacking so far. Here, we report on our attempts to synthesise VBi2Te4 by molecular beam epitaxy (MBE). X-ray diffraction reveals that in the thermodynamic phase space reachable by MBE, there is no region where VBi2Te4 is stably synthesised. Moreover, scanning transmission electron microscopy shows a clear phase separation to Bi2Te3 and VTe2 instead of the formation of VBi2Te4. We suggest the phase instability to be due to either the large lattice mismatch between VTe2 and Bi2Te3 or the unfavourable valence state of vanadium.


Topological information device operating at the Landauer limit

December 2023

·

15 Reads

·

3 Citations

We propose and theoretically investigate a novel Maxwell's demon implementation based on the spin-momentum locking property of topological matter. We use nuclear spins as a memory resource which provides the advantage of scalability. We show that this topological information device can ideally operate at the Landauer limit; the heat dissipation required to erase one bit of information stored in the demon's memory approaches kBTln2. Furthermore, we demonstrate that all available energy, kBTln2 per one bit of information, can be extracted in the form of electrical work. Finally, we find that the current-voltage characteristic of topological information device satisfy the conditions of an ideal memristor.


Characteristics of the Cd3As2 nanoplate-based Josephson junctions
a Hinge states (green line) and skin modes (light blue) distributed on the surface of a Cd3As2 nanoplate with 112\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left(112\right)$$\end{document} surface orientation. Cd3As2 has a pair of Dirac points along kz\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${k}_{z}$$\end{document} direction, i.e., 001\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\left(001\right)$$\end{document} direction and the hinge states are the segments connecting the projection of the bulk Dirac points along the hinges. b Optical image of the device. The nanoplate is denoted by gray, while the Nb electrodes are denoted by green. Josephson junctions 1 and 2 are indicated, where the channel lengths are 800 and 600 nm, respectively. The four-terminal measurement method is adopted. Scale bar, 4 μm. c Color map of differential resistance dV/dI as a function of Idc\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${I}_{{dc}}$$\end{document} and Bz\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${B}_{z}$$\end{document} in junction 2 at the base temperature of 10 mK. The dark blue region represents the superconducting state, while its upper boundary denotes the critical current Ic\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${I}_{c}$$\end{document}. The orange curve shows the fitting results using the DF method. d dV/dI as a function of Idc\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${I}_{{dc}}$$\end{document} and Bz\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${B}_{z}$$\end{document} in junction 1 at the base temperature (10 mK). A SQUID-like supercurrent interference pattern is captured, where the width of the central lobe is close to others.
Broad and colossal edge supercurrent observed in junction 1
a The comparison of experimental data (scatters) extracted from Fig. 1d, Fraunhofer fit (red curve) and the edge-stepped supercurrent model fit (green curve). b The extracted supercurrent density profile Jcx\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${J}_{c}\left(x\right)$$\end{document} according to the experimental data in a. The blue dashed curve is obtained by the DF method, and its Gaussian fit is denoted by the red solid curve. Asymmetric supercurrent density peaks exist at two edges of the nanoplate. The full width of half maximum (FWHM) of the Gaussian peaks is 1.65\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$1.65$$\end{document} and 1.57μm\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$1.57\,{{{{{\rm{\mu }}}}}}{{{{{\rm{m}}}}}}$$\end{document}, respectively, which are considered as the width of the edge supercurrent channels.
Temperature-dependent SQUID-like interference patterns in junction 1
a Evolution of SQUID-like interference pattern with temperature. As increasing temperature, the central nodes of Ic\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${I}_{c}$$\end{document} oscillations go from nonzero to zero. b The extracted supercurrent density profile according to the IcBz\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${I}_{c}\left({B}_{z}\right)$$\end{document} data at 800 mK. The blue dashed curve is obtained by the DF method, and its Gaussian fit is denoted by the red solid curve. Symmetric Jc\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${J}_{c}$$\end{document} peaks exist at two edges of the nanoplate with almost vanishing bulk contribution. c The ratio of Jcleft peak\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${J}_{c}\left({{{{{\rm{left\; peak}}}}}}\right)$$\end{document}/Jcright peak\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\,{J}_{c}\left({{{{{\rm{right\; peak}}}}}}\right)$$\end{document} as a function of temperature. d The FWHM of Jc\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${J}_{c}$$\end{document} edge peaks as a function of temperature, revealing the variation of the width of the supercurrent edge channel versus temperature. e Temperature-dependent critical current IcT\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${I}_{c}\left(T\right)$$\end{document} at Bz=0mT\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${B}_{z}=0\ {{{{{\rm{mT}}}}}}$$\end{document}. The purple curve shows the fitting result based on the short ballistic junction model, and the parameters of interfacial transmission coefficient τ=0.4\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\tau=0.4$$\end{document} and proximity-induced superconducting gap Δ0=0.26meV\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${\Delta }_{0}=0.26\ {{{{{\rm{meV}}}}}}$$\end{document} are obtained.
Nonlocal measurements of supercurrent interference pattern at 10 mK
a Schematic of the nonlocal measurement configuration. The bias dc current Ib\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${I}_{b}$$\end{document} superimposed on an ac excitation ib\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${i}_{b}$$\end{document} is injected from terminal A to B, and the ac voltage drop Vnonlocal\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${V}_{{nonlocal}}$$\end{document} is monitored between terminals C and D. The electrodes A and B form junction 1 (L=800nm\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$L=800\,{{{{{\rm{nm}}}}}}$$\end{document}), while electrodes C and D correspond to junction 2 (L=600nm\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$L=600\,{{{{{\rm{nm}}}}}}$$\end{document}). b Schematic diagram of current distribution in the nonlocal measurement configuration, where the source-drain channel is in the normal state. The blue arrows represent the flowing direction of the edge current. Yellow and orange arrows denote the bulk and surface current flow, respectively. c The Vnonlocal\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${V}_{{nonlocal}}$$\end{document} as a function of bias current Ib\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${I}_{b}$$\end{document} and external field Bz\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${B}_{z}$$\end{document}. d The critical values of Ib\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${I}_{b}$$\end{document} and Inonlocal\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${I}_{{nonlocal}}$$\end{document} versus magnetic field Bz\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${B}_{z}$$\end{document}, which drives junction 2 into the normal state and triggers a nonzero nonlocal voltage. e The extracted nonlocal supercurrent density distribution in junction 2. The blue dashed curve is obtained from IcnonlocalBz\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${I}_{c}^{{nonlocal}}\left({B}_{z}\right)$$\end{document} in d, and the red solid curve corresponds to the Gaussian fitting results.
Broad and colossal edge supercurrent in Dirac semimetal Cd3As2 Josephson junctions

October 2023

·

226 Reads

·

4 Citations

Edge supercurrent has attracted great interest recently due to its crucial role in achieving and manipulating topological superconducting states. Proximity-induced superconductivity has been realized in quantum Hall and quantum spin Hall edge states, as well as in higher-order topological hinge states. Non-Hermitian skin effect, the aggregation of non-Bloch eigenstates at open boundaries, promises an abnormal edge channel. Here we report the observation of broad edge supercurrent in Dirac semimetal Cd3As2-based Josephson junctions. The as-grown Cd3As2 nanoplates are electron-doped by intrinsic defects, which enhance the non-Hermitian perturbations. The superconducting quantum interference indicates edge supercurrent with a width of ~1.6 μm and a magnitude of ~1 μA at 10 mK. The wide and large edge supercurrent is inaccessible for a conventional edge system and suggests the presence of non-Hermitian skin effect. A supercurrent nonlocality is also observed. The interplay between band topology and non-Hermiticity is beneficial for exploiting exotic topological matter.


Multiple Andreev reflections in topological Josephson junctions with chiral Majorana modes

May 2023

·

21 Reads

·

3 Citations

Andreev bound states (ABSs) occur in Josephson junctions when the total phase of the Andreev and normal reflections is a multiple of 2π. In ballistic junctions with an applied voltage bias, a quasiparticle undergoes multiple Andreev reflections before entering the leads, resulting in peaks in the current-voltage I(V) curve. Here we present a model for a two-dimensional S/TI/MTI/TI/S junction, where S is a superconductor, TI is a topological insulator, and MTI is a magnetic topological insulator barrier. We show that the interplay of broken time-reversal symmetry and topology results in an asymmetric I(V) curve. Such junctions are predicted to host chiral Majorana modes. We demonstrate that the peak positions in I(V) are directly linked to ABSs. We use this to show how the angle-resolved I(V) curve becomes a spectroscopic tool for the chirality and degeneracy of ABSs.


Enhancement of the Surface Morphology of (Bi0.4Sb0.6)2Te3 Thin Films by In Situ Thermal Annealing

February 2023

·

36 Reads

·

2 Citations

The study of the exotic properties of the surface states of topological insulators requires defect-free and smooth surfaces. This work aims to study the enhancement of the surface morphology of optimally doped, high-crystalline (Bi0.4Sb0.6)2Te3 films deposited by molecular beam epitaxy on Al2O3 (001) substrates. Atomic force microscopy shows that by employing an in situ thermal post anneal, the surface roughness is reduced significantly, and transmission electron microscopy reveals that structural defects are diminished substantially. Thence, these films provide a great platform for the research on the thickness-dependent properties of topological insulators.


FIG. 3. The angle-resolved asymmetric I(V ) curves for a S/TI/MTI/TI/S junction normalized by I∆ = De∆0/h. Each curve corresponds to a single incident angle θ ∈ (0, π/2), with corresponding bound state energy EABS(θ). The black dashed line is the angle average obtained in the lateral 2D junction limit. The parameters are µTI/mz ∼ 0.7, with mz/∆0 = 300, µS = µTI, µMTI = 0, the transparency ranges from D = 0.005 − 0.2, and the MTI barrier width is d = 1.5vF /mz. Inset: Nanowire limit. Each I(V ) curve corresponds to a single (normalized) quantized py = sin θ channel where the current is estimated by I(−θ) + I(θ). The graph is identical for ±eV /∆0.
Multiple Andreev reflections in two-dimensional Josephson junctions with broken time-reversal symmetry

January 2023

·

44 Reads

Andreev bound states (ABS) occur in Josephson junctions when the total phase of the Andreev and normal reflections is a multiple of 2π2\pi. In ballistic junctions with an applied voltage bias, a quasi-particle undergoes multiple Andreev reflections before entering the leads, resulting in peaks in the current-voltage I(V) curve. Here we present a general model for Josephson junctions with spin-active interlayers i.e., magnetic or topological materials with broken time-reversal symmetry. We investigate how ABS change the peak positions and shape of I(V), which becomes asymmetric for a single incident angle. We show how the angle-resolved I(V) curve becomes a spectroscopic tool for the chirality and degeneracy of ABS.


Citations (67)


... We observe that the visibility of the halfinteger Shapiro steps scales with the external magnetic field showing a very peculiar two-lobe structure. These findings provide a tool for proving the presence of topological phase [51], complementary to that discussed in Ref. [22], which could be promptly employed in, e.g., LAO/STO-type [19,21] or nanowire-based systems [20,52]. The appearance of a fractional Shapiro response marks a deviation from topologically-trivial to nontrivial junctions, potentially driven by the external magnetic field. ...

Reference:

Probing Topological Superconductivity of oxide nanojunctions using fractional Shapiro steps
Magnetic field enhanced critical current in Ge–Si nanowire Josephson junctions

... However, nonvanishing fluctuations in the energy exchange between resource and working substance are required for the demonic operation of the device in this case. A whole range of demonic devices spanning from the purely information-driven to nonthermal demons and beyond has been proposed [54][55][56][57][58][59], where these devices have in common that they seemingly violate the second law, see Ref. [60] for a detailed discussion. ...

Topological information device operating at the Landauer limit
  • Citing Article
  • December 2023

... The FWHM is observed to enhance with temperature, consistent with experimental findings in Cd 3 As 2 Josephson junctions. 73 Specifically, the FWHM follows a ffiffiffi ffi T p dependence with temperature, arising from the square root of T, which is characteristic of the Lorentzian width. 48,53 We now focus on the doped DSMs, where n e = 0, placing E F in the conduction band. ...

Broad and colossal edge supercurrent in Dirac semimetal Cd3As2 Josephson junctions

... In Josephson junctions characterized as "short" and with arbitrary barrier transparency, where the stripe length is significantly shorter than the superconducting coherence length, the phenomenon responsible for charge transfer is the multiple Andreev reflection (MAR) effect. [1][2][3] The essence of the MAR effect is that a quasiparticle undergoes several successive Andreev reflections from the "normal metal-superconductor (NS)" interfaces, accumulating energy until it exceeds E F + Δ, at which point the quasiparticle can escape into the conduction band of one of the superconducting sides (where E F is the Fermi energy and Δ is the order of a parameter). 4 In recent years, intensive research has focused on Josephson junctions based on low-dimensional structures (such as graphenelike materials, carbon nanotubes, some quasi-two-dimensional transition metal dichalcogenides, etc.). ...

Multiple Andreev reflections in topological Josephson junctions with chiral Majorana modes
  • Citing Article
  • May 2023

... The improvement in crystallinity resulting from the annealing process contributes to a decrement in disordered states and, consequently, lowers dislocation density. 31 Likewise, the value of N c demonstrates a decreasing trend during the annealing process. The overall change in structural parameters within the crystal can be attributed to crystallite size refinement and the elimination of dislocations within the films. ...

Enhancement of the Surface Morphology of (Bi0.4Sb0.6)2Te3 Thin Films by In Situ Thermal Annealing

... They are expected to appear at voltages of V = (2∆)/(en), where e is the elementary charge and n an integer [40,47]. The shape of the features are determined by the transport characteristics in the junction [48][49][50]. The positions of the voltage biases for possible multiple Andreev reflections are indicated by vertical dashed lines, where the first two are labeled 2∆ and ∆. ...

Anomalous temperature dependence of multiple Andreev reflections in a topological insulator Josephson junction

... This is naively because reducing the thickness down to a monolayer should eliminate the moderately strong interlayer interaction inherent to tellurides as compared to sulfides and selenides, and affect the stability of trimerizations. Indeed, the bottom-up synthetic approaches such as molecular beam epitaxy (MBE) and chemical vapor deposition (CVD) have successfully fabricated mono/few-layer MTe2 with different charge/spin orderings from the bulk counterparts (see Refs. [65][66][67][68][69][70][71][72][73][74][75][76] for VTe2, [77,78] for NbTe2, and [79][80][81][82] for TaTe2). On the other hand, we would also encourage more research using the top-down exfoliation method to assemble van der Waals heterostructures and pursue the relevant novel phenomena such as twistronics. ...

Thickness-Dependent Sign Change of the Magnetoresistance in VTe2 Thin Films

Solids

... A quartz crystal microbalance installed in the main chamber was used to calibrate the fluxes arising from the individual cells. To avoid Te vacancies, a flux ratio of (Bi + Sb):Te = 1:10 was provided with a growth rate of 0.07 QL/min [16]. The substrates were cleaned by means of ultrasonic immersion in acetone and ethanol. ...

Revisiting the van der Waals Epitaxy in the Case of (Bi0.4Sb0.6)2Te3 Thin Films on Dissimilar Substrates

... This is partially due to the observation of a topological state in BaBiO 3 [1][2][3][4] predicted by Yan et al. [5]. But the main interest in this material is caused by the fact that a unified theory for the insulating state in the pure BaBiO 3 and superconductivity in the doped phases is still missing [1,2,[6][7][8][9][10][11][12][13]. ...

Observing structural distortions in complex oxides by x-ray photoelectron diffraction
  • Citing Article
  • April 2022

Journal of Electron Spectroscopy and Related Phenomena

... This has allowed us to approximate water as non-dissociated molecules. We have used molecular dynamics (MD) simulations where H 2 O molecules are not dissociated to show that, if >0. 4 ML of water initially exists on the surface at 10 −5 torr, desorption rapidly occurs at room temperature at sufficiently long times without the need for heating, [22,23] in agreement with measurements [19,24]. As such, these surfaces do not support multi-layers of adsorbed water under the relevant low pressure conditions [11]. ...

Quantum oscillations in an optically-illuminated two-dimensional electron system at the LaAlO 3 /SrTiO 3 interface