S. A. Crooker’s research while affiliated with Los Alamos National Laboratory 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 (320)


Quantum oscillations of holes in GaN
  • Preprint

January 2025

·

17 Reads

Chuan F. C. Chang

·

Joseph E. Dill

·

·

[...]

·

GaN has emerged to be a major semiconductor akin to silicon due to its revolutionary impacts in solid state lighting, critically enabled by p-type doping, and high-performance radio-frequency and power electronics. Suffering from inefficient hole doping and low hole mobility, quantum oscillations in p-type GaN have not been observed, hindering fundamental studies of valence bands and hole transport in GaN. Here, we present the first observation of quantum oscillations of holes in GaN. Shubnikov-de Haas (SdH) oscillations in hole resistivity are observed in a quantum-confined two-dimensional hole gas at a GaN/AlN interface, where polarization-induced doping overcomes thermal freeze-out, and a sharp and clean interface boosts the hole mobility enough to unmask the quantum oscillations. These holes degenerately occupy the light and heavy hole bands of GaN and have record-high mobilities of ~1900 cm2/Vs and ~400 cm2/Vs at 3K, respectively. We use magnetic fields up to 72 T to resolve SdH oscillations of holes from both valence bands to extract their respective sheet densities, quantum scattering times, and the effective masses of light holes (0.5-0.7 m0) and heavy holes (1.9 m0). SdH oscillations of heavy and light holes in GaN constitute a direct metrology of valence bands and open new venues for quantum engineering in this technologically important semiconductor. Like strained silicon transistors, strain-engineering of the valence bands of GaN is predicted to dramatically improve hole mobilities by reducing the hole effective mass, a proposal that can now be explored experimentally, particularly in a fully fabricated transistor, using quantum oscillations. Furthermore, the findings of this work suggest a blueprint to create 2D hole gases and observe quantum oscillations of holes in related wide bandgap semiconductors such as SiC and ZnO in which such techniques are not yet possible.


Structural and magnetization characterizations and MR measurements of the Cr2Te3 thin film. a) A schematic of the atomic structure of Cr2Te3, viewed along [100] zone axis, where CrI, CrII, and CrIII are three inequivalent Cr sites. b) A cross‐sectional HAADF‐STEM image of the Cr2Te3 thin film taken along the [100] axis with iDPC technique, consistent with the atomic structure model. The blue circles mark the atomic columns with a weak contrast, indicating partially occupied vacancy sites due to chemical disorder. c) A schematic of the atomic structure (top) and simulated diffraction pattern (bottom) obtained from the atomic model of Cr2Te3, viewed along [210] zone axis. d) A cross‐sectional HAADF‐STEM image of Cr2Te3 thin film taken along the [210] zone axis. Inset: (d1–d3) corresponding FFT patterns of the square‐colored areas indicated in the HAADF‐STEM image in (d), matching the simulated electron diffraction pattern in (c). e) Out‐of‐plane magnetic hysteresis loops measured at different temperatures. (From bottom to top, measurement temperatures are 25, 45, 65, 85, 105, 125, and 145 K, respectively.) f) Out‐of‐plane MCD hysteresis loops measured at different temperatures (From bottom to top, measurement temperatures are 10, 30, 50, 75, and 95 K, respectively) using 700 nm light. g) Temperature‐dependence of longitudinal resistivity ρxx as a function of the out‐of‐plane magnetic field. (From bottom to top, measurement temperatures are 5, 25, 45, 65, 75, 85, 105, 125, and 145 K, respectively.) h) HC as a function of temperature extracted from magnetic (circle), MCD (square), and MR (triangle) hysteresis measurements.
Hall effect measurements and two interpretations of the observed unconventional AHE. a) A schematic of the Hall bar device used for Hall effect measurements. b) Shown from bottom to top are the magnetic field‐dependent Hall resistivity ρyx measured at temperatures of 5, 25, 45, 65, 75, 85, 105, 125, and 145 K, respectively. The solid arrows indicate the looping directions, showing a change of polarity of the measured Hall resistivity with changing temperature. c) A representative ρAHE loop measured at 25 K, obtained by subtracting the OHE contribution. d) Fitting by an AHE loop and a THE loop, and e) fitting by two AHE loops to reproduce the observed humps and dips in the AHE loop in (c).
MFM images measured at representative temperatures of 45 and 10 K with different magnetic fields. Evolution of magnetic domains of the sample measured at 45 K with field values of a) 0.3 T, b) 0.4 T, c) 0.5 T, and d) 0.6 T; and at 10 K with field values of e) 0.6 T, f) 0.7 T, and g) 0.8 T and h) 0.9 T.
Minor ρAHE loops and fitting results of the temperature‐dependent ρAHE loops for the unannealed samples. Minor ρAHE loops measured at 25 K for different stopping fields a) −0.11 T, b) −0.41 T, c) −0.51 T, and d) −1 T (Gray lines are corresponding full AHE loops for comparison). e–i) ρAHE1 and j–n) ρAHE2 loops with opposite polarities that are used to fit ρAHE loops, respectively. o–s) Corresponding fitting results (red) and measured (black) ρAHE loops. From left to right, the measuring temperature decreases from 75 to 5 K.
Temperature‐dependent Hall resistivity for the annealed samples. a–e) Magenta and f–j) blue curves represent ρAHE1 and ρAHE2 loops with different polarities that are used in the fittings, respectively. k–o) The corresponding fitting results (red) for the measured (black) total AHE resistivity ρAHE loops at different temperatures. From left to right, the temperature decreases from 55 to 5 K.

+1

Unconventional Anomalous Hall Effect Driven by Self‐Intercalation in Covalent 2D Magnet Cr2Te3
  • Article
  • Full-text available

November 2024

·

83 Reads

·

1 Citation

Covalent 2D magnets such as Cr2Te3, which feature self‐intercalated magnetic cations located between monolayers of transition‐metal dichalcogenide material, offer a unique platform for controlling magnetic order and spin texture, enabling new potential applications for spintronic devices. Here, it is demonstrated that the unconventional anomalous Hall effect (AHE) in Cr2Te3, characterized by additional humps and dips near the coercive field in AHE hysteresis, originates from an intrinsic mechanism dictated by the self‐intercalation. This mechanism is distinctly different from previously proposed mechanisms such as topological Hall effect, or two‐channel AHE arising from spatial inhomogeneities. Crucially, multiple Weyl‐like nodes emerge in the electronic band structure due to strong spin‐orbit coupling, whose positions relative to the Fermi level is sensitively modulated by the canting angles of the self‐intercalated Cr cations. These nodes contribute strongly to the Berry curvature and AHE conductivity. This component competes with the contribution from bands that are less affected by the self‐intercalation, resulting in a sign change in AHE with temperature and the emergence of additional humps and dips. The findings provide compelling evidence for the intrinsic origin of the unconventional AHE in Cr2Te3 and further establish self‐intercalation as a control knob for engineering AHE in complex magnets.

Download

Energy shifts and broadening of excitonic resonances in electrostatically-doped semiconductors

November 2024

·

32 Reads

Tuning the density of resident electrons or holes in semiconductors provides crucial insight into the composition of excitonic complexes that are observed as absorption or photoluminescence resonances in optical studies. Moreover, we can change the way these resonances shift and broaden in energy by controlling the quantum numbers of the resident carriers with magnetic fields and doping levels, and by selecting the quantum numbers of the photoexcited or recombining electron-hole (e-h) pair through optical polarization. We discuss the roles of distinguishability and optimality of excitonic complexes, showing them to be key ingredients that determine the energy shifts and broadening of optical resonances in charge-tunable semiconductors. A distinguishable e-h pair means that the electron and hole undergoing photoexcitation or recombination have quantum numbers that are not shared by any of the resident carriers. An optimal excitonic complex refers to a complex whose particles come with all available quantum numbers of the resident carriers. All optical resonances may be classified as either distinct or indistinct depending on the distinguishability of the e-h pair, and the underlying excitonic complex can be classified as either optimal or suboptimal. The universality of these classifications, inherited from the fundamental Pauli exclusion principle, allows us to understand how optical resonances shift in energy and whether they should broaden as doping is increased. This understanding is supported by conclusive evidence that the decay of optical resonances cannot be simply attributed to enhanced screening when resident carriers are added to a semiconductor. Finally, applying the classification scheme in either monolayer or moire heterobilayer systems, we relate the energy shift and amplitude of the neutral exciton resonance to the compressibility of the resident carrier gas.



FIG. 1. Temperature dependence of the resistivity ρ xx in samples A-D at zero magnetic field. The inset shows the superconducting transitions in greater detail over a more limited temperature range and the determination of T c,onset . The sheet resistance (R S ) is given in terms of the resistance quantum of a Cooper pair h/4e 2 = 6.45 k.
FIG. 2. Cross-sectional ADF-TEM images of the lowest and highest T c samples, A and D, respectively. Sample D clearly demonstrates a 3 unit-cell FeSe (001) epitaxial layer on STO (001), along with the expected crystalline FeTe and amorphous Te capping layers, while clear images of the grown layers in sample A could not be achieved.
FIG. 3. MR of sample A at various temperatures in the range 1.2 K T 15 K with magnetic field (a) in-plane (H ab) and (b) out-of-plane (H c).
FIG. 4. Temperature dependence of the upper critical field in the parallel field configuration H ab of sample A using the ρ 10 n , ρ 50 n , ρ 90 n , and ρ onset n criteria. The overall qualitative behavior of each curve is similar. Lines are a guide to the eye.
FIG. 6. Temperature dependence of the upper critical field anisotropy parameter γ = H ab c2 (T )/H c c2 (T ). Filled data points are obtained from actual experimental results where H ab c2 and H c c2 were obtained at the same temperature. Low temperature, open (unfilled) points for sample C were calculated using WHH-predicted values of H ab c2 (T ) and experimentally measured values of H c c2 (T ). Open points for sample D were calculated using experimental (predicted) values of H ab c2 (T ) [H c c2 (T )] above 20 K and predicted (experimental) values at and below 20 K. Lines are a guide to the eye.
Temperature dependence and limiting mechanisms of the upper critical field of FeSe thin films

March 2024

·

91 Reads

·

3 Citations

Physical Review B

We use magnetoresistance measurements at high magnetic field (μ0H≤65 T) and low temperature (T≥500 mK) to gain fresh insights into the behavior of the upper critical field Hc2 in superconducting ultrathin FeSe films of varying degrees of disorder, grown by molecular beam epitaxy on SrTiO3. Measurements of Hc2 across samples with a widely varying superconducting critical temperature (1.2 K ≤Tc≤21 K) generically show similar qualitative temperature dependence. We analyze the temperature dependence of Hc2 in the context of Werthamer-Helfand-Hohenberg (WHH) theory. The analysis yields parameters that indicate a strong Pauli paramagnetic pair-breaking mechanism which is also reflected by pseudoisotropic superconductivity in the limit of zero temperature. In the lower Tc samples, we observe a spin-orbit scattering-driven enhancement of Hc2 above the strongly-coupled Pauli paramagnetic limit. We also observe clear deviations from WHH theory at low temperature, regardless of Tc. We attribute this to the multiband superconductivity of FeSe and possibly to the emergence of a low-temperature, high-field superconducting phase.


Emergence of composite many-body exciton states in WS 2 and MoSe 2 monolayers

January 2024

·

64 Reads

·

5 Citations

When doped with a high density of mobile charge carriers, monolayer transition-metal dichalcogenide (TMD) semiconductors can host new types of composite many-particle exciton states that do not exist in conventional semiconductors. Such multiparticle bound states arise when a photoexcited electron-hole pair couples not to just a single Fermi sea that is quantum-mechanically distinguishable (as in the case of conventional charged excitons or trions), but rather couples simultaneously to multiple Fermi seas, each having distinct spin and valley quantum numbers. Composite six-particle “hexciton” states were recently identified in electron-doped WSe2 monolayers, but under suitable conditions they should also form in all other members of the monolayer TMD family. Here we present spectroscopic evidence demonstrating the emergence of many-body hexcitons in charge-tunable WS2 monolayers (at the A-exciton) and MoSe2 monolayers (at the B-exciton). The roles of distinguishability and carrier screening on the stability of hexcitons are discussed.


Optical properties of lead halide perovskite crystals with various band gaps at a temperature of 1.6 K. The colored lines show the photoluminescence spectra for continuous wave excitation at Eexc = 3.06 eV, using the laser power density of 5 mW/cm². For MAPb(Br0.05Cl0.95)3 the excitation energy is Eexc = 3.23 eV. The black lines show the PLE spectrum for FA0.9Cs0.1PbI2.8Br0.2 and the reflectivity spectra for the other samples.
Magneto‐optical properties of excitons, as well as resident electrons and holes in a FAPbBr3 crystal. a) Reflectivity spectra measured in σ⁺ (red line) and σ⁻ (blue line) polarization in the longitudinal magnetic field BF = 7 T at T = 1.6 K. b) Exciton Zeeman splitting as function of BF measured in magneto‐reflectivity. Slope of the linear fit gives gF,X = +2.7. c) Time‐resolved Kerr ellipticity signal (blue) measured at BV = 0.5 T, T = 6 K, using the laser photon energy of 2.188 eV. The electron (black) and hole (green) components are obtained from decomposing the signal. d) Dependence of the Zeeman splitting of the hole (green), the electron (black), and their sum (blue) on BV.
a) Anisotropy of the electron (black circles) and hole (green circles) g‐factors measured by SFRS for CsPbBr3. B = 5 T and T = 1.6 K. Red circles present the exciton g‐factor. The black and green lines are calculated with Equation (2), using the parameters from Table 1. The blue crosses are experimental data of ge + gh, the blue line is the sum of the fits shown by the black and green lines. b) Magnetic field dependence of the exciton Zeeman splitting measured from magneto‐reflectivity in the Faraday geometry for CsPbBr3. T = 1.6 K. The symbols are experimental data and the line is a linear fit.
Dependence of the exciton g‐factor measured by magneto‐reflectivity (closed red circles are our data and open red circles are from Refs. [14, 16]) and of the sum of carrier g‐factors evaluated from TRKE and SFRS (crosses) on the band gap energy. The data are taken at cryogenic temperatures of 1.6 − 10 K. The dashed lines are model calculations[¹⁰] with Equations (3) and (4) for the hole (green) and the electron (black) g‐factors, which closely match the experimental data given in Table 1. The blue line is the sum of their contributions calculated with Equation (5). Size of the symbols includes the error.
Weak Dispersion of Exciton Landé Factor with Band Gap Energy in Lead Halide Perovskites: Approximate Compensation of the Electron and Hole Dependences

November 2023

·

74 Reads

·

20 Citations

The optical properties of lead halide perovskite semiconductors in vicinity of the bandgap are controlled by excitons, so that investigation of their fundamental properties is of critical importance. The exciton Landé or g‐factor gX is the key parameter, determining the exciton Zeeman spin splitting in magnetic fields. The exciton, electron, and hole carrier g‐factors provide information on the band structure, including its anisotropy, and the parameters contributing to the electron and hole effective masses. Here, gX is measured by reflectivity in magnetic fields up to 60 T for lead halide perovskite crystals. The materials band gap energies at a liquid helium temperature vary widely across the visible spectral range from 1.520 up to 3.213 eV in hybrid organic–inorganic and fully inorganic perovskites with different cations and halogens: FA0.9Cs0.1PbI2.8Br0.2, MAPbI3, FAPbBr3, CsPbBr3, and MAPb(Br0.05Cl0.95)3. The exciton g‐factors are found to be nearly constant, ranging from +2.3 to +2.7. Thus, the strong dependences of the electron and hole g‐factors on the bandgap roughly compensate each other when combining to the exciton g‐factor. The same is true for the anisotropies of the carrier g‐factors, resulting in a nearly isotropic exciton g‐factor. The experimental data are compared favorably with model calculation results.


Exciton‐Polaritons in CsPbBr 3 Crystals Revealed by Optical Reflectivity in High Magnetic Fields and Two‐Photon Spectroscopy

November 2023

·

46 Reads

·

10 Citations

physica status solidi (RRL) - Rapid Research Letters

Cesium lead bromide (CsPbBr 3 ) is a representative material of the emerging class of lead halide perovskite semiconductors that possess remarkable optoelectronic properties. Its optical properties in the vicinity of the band gap energy are greatly contributed by excitons, which form exciton‐polaritons due to strong light‐matter interactions. We examine exciton‐polaritons in solution‐grown CsPbBr ³ crystals by means of circularly‐polarized reflection spectroscopy measured in high magnetic fields up to 60 T. The excited 2P exciton state is measured by two‐photon absorption. Comprehensive modelling and analysis provides detailed quantitative information about the exciton‐polariton parameters: exciton binding energy of 32.5 meV, oscillator strength characterized by longitudinal‐transverse splitting of 5.3 meV, damping of 6.7 meV, reduced exciton mass of 0.18 m 0 , exciton diamagnetic shift of 1.6 μ eV/T ² , and exciton Landé factor g X =+2.35. We show that the exciton states can be well described within a hydrogen‐like model with an effective dielectric constant of 8.7. From the measured exciton longitudinal‐transverse splitting we evaluate the Kane energy of E p =15 eV, which is in reasonable agreement with values of 11.8−12.5 eV derived from the carrier effective masses. This article is protected by copyright. All rights reserved.


Deconstructing magnetization noise: Degeneracies, phases, and mobile fractionalized excitations in tetris artificial spin ice

October 2023

·

52 Reads

·

1 Citation

Proceedings of the National Academy of Sciences

Direct detection of spontaneous spin fluctuations, or “magnetization noise,” is emerging as a powerful means of revealing and studying magnetic excitations in both natural and artificial frustrated magnets. Depending on the lattice and nature of the frustration, these excitations can often be described as fractionalized quasiparticles possessing an effective magnetic charge. Here, by combining ultrasensitive optical detection of thermodynamic magnetization noise with Monte Carlo simulations, we reveal emergent regimes of magnetic excitations in artificial “tetris ice.” A marked increase of the intrinsic noise at certain applied magnetic fields heralds the spontaneous proliferation of fractionalized excitations, which can diffuse independently, without cost in energy, along specific quasi-1D spin chains in the tetris ice lattice.


Strain-engineered WSe2/NiPS3 heterostructures host QEs displaying sharp, localized PL peaks with a strong degree of spontaneous circular polarization
a,b, Optical (a) and PL (b) images of a WSe2/NiPS3 heterostructure. The portion of the WSe2 monolayer that does not overlap with NiPS3 emits bright PL (region R1), whereas PL is quenched when the WSe2 is coupled to the NiPS3 (region R2). Strong PL was restored by indentations marked by the white square. c, Schematic of the sample structure and an atomic force microscopy topography image and cross section of a representative indentation (inset). d–f, Representative σ⁺-resolved (blue) and σ–-resolved (red) PL spectra from different individual nanoindentations acquired under linearly polarized laser excitation at 4 K. Although the peaks marked as a and b are σ⁺ polarized with DCP values of 0.36 and 0.89, peaks marked c–f are σ– polarized with DCP values of –0.40, –0.29, –0.33 and –0.37, respectively. Some of the localized emission peaks ride on top of a broad PL background that displays little polarization, suggesting that the DCP of localized emission peaks could even be higher if the broad PL background is subtracted.
Source data
Demonstration of quantum light emission from strain-engineered WSe2/NiPS3 heterostructures
a,b, σ⁺-polarized (blue) and σ–-polarized (red) PL spectra of an indented WSe2/NiPS3 heterostructure acquired under σ⁺-polarized (a) and σ–-polarized (b) laser excitation. c–e, PL intensity time trace (c), PL decay curve (d) and second-order photon correlation trace (e) measured from the PL peak marked by the blue column in a and b acquired under 514 nm, 40 MHz pulsed laser excitation.
Source data
Correlation of chiral quantum light emission with local magnetization
a, Zig-zag AFM order overlaid with the atomic NiPS3 structure where Ni electronic spins are aligned parallel (blue arrows) and antiparallel (red arrows) to the crystallographic a axis. b, Schematic of an indentation region cross section where the local strain creates exciton-capture centres in WSe2. c,d, Scanning NV-centre mapping of the surface magnetization associated with the morphology of a nanoindentation (c) is strongly correlated with the magnetic field measured in the indentation region (d). e, NV spectra taken both off (grey) and on (red) the indented region show a shift in the detected magnetic signal of Bindent = 26 ± 2 µT. A high-resolution atomic force microscopy image of the nanoindentation is provided (inset of c) as the NV scanning probes have reduced spatial resolution. The scanning NV centre has a standoff distance of 70 ± 20 nm, meaning that the measured magnetic field is reduced by spatial averaging and inverse distance scaling. kcps, kilo counts per second.
Source data
Magneto-PL studies of chiral QEs
a, Low-temperature, polarization-resolved PL spectra acquired as a function of external B field applied perpendicular to the sample plane (Faraday geometry). An expanded view of a localized PL peak (1.59 eV; grey bar) displays a strong DCP of –0.6 (left). b, Energy splitting (left axis) between the σ⁺ and σ– PL peaks of a localized exciton at 1.59 eV (blue data points) and a 2D exciton at 1.75 eV (red data points). The error bars represent the uncertainty in the peak positions determined by the Gaussian fit. The DCP (right axis) of the 1.59 eV peak versus B field is plotted as black triangles. The blue, red and black dashed lines are guides to the eye. c, Low-temperature, polarization-resolved PL spectra of an indented region with a PL peak at ~1.70 eV, exhibiting a strong DCP of ~0.7 before the temperature- and field-dependent measurement (top row). Low-temperature, polarization-resolved PL spectra of the same location acquired after heating the sample to 180 K, raising the field to 6 T and cooling the sample back down to 4 K while maintaining B at 6 T (second row). Low-temperature, polarization-resolved PL spectra acquired as the field is reduced back to 0 T (two bottom rows). d, The experiment in c was repeated by applying an external field of –6 T at 180 K. e, DCP of the 1.69 eV peak plotted against an external B field. The green arrows originate from the DCP value originally observed in the top-most panels of c and d and indicate the process involving heating the sample to 180 K and cooling back down to 4 K while maintaining the B field at 6 T (–6 T). The red (blue) arrows and data points represent the process and DCP observed during the reduction in B from 6 T (–6 T) to 0 T at 4 K immediately after cooling down. Since the DCP values are calculated from the peak intensities of the σ+/–-polarized spectra, they are contributed not only by the QE emission but also by the broad PL background that increases as the field is ramped from +6 to –6 T. The grey-shaded region represents variations in the DCP measured at 0 T.
Source data
Proximity-induced chiral quantum light generation in strain-engineered WSe2/NiPS3 heterostructures

August 2023

·

202 Reads

·

42 Citations

Nature Materials

Quantum light emitters capable of generating single photons with circular polarization and non-classical statistics could enable non-reciprocal single-photon devices and deterministic spin–photon interfaces for quantum networks. To date, the emission of such chiral quantum light relies on the application of intense external magnetic fields, electrical/optical injection of spin-polarized carriers/excitons or coupling with complex photonic metastructures. Here we report the creation of free-space chiral quantum light emitters via the nanoindentation of monolayer WSe2/NiPS3 heterostructures at zero external magnetic field. These quantum light emitters emit with a high degree of circular polarization (0.89) and single-photon purity (95%), independent of pump laser polarization. Scanning diamond nitrogen-vacancy microscopy and temperature-dependent magneto-photoluminescence studies reveal that the chiral quantum light emission arises from magnetic proximity interactions between localized excitons in the WSe2 monolayer and the out-of-plane magnetization of defects in the antiferromagnetic order of NiPS3, both of which are co-localized by strain fields associated with the nanoscale indentations.


Citations (54)


... Determining the upper critical field H c2 and its anisotropy γ = H ab c2 /H c c2 is critical for elucidating the superconducting mechanism and assessing application potential [26,27]. The temperature dependence of H c2 reflects the electronic structure governing superconductivity and provides insights into pairbreaking mechanisms, which are crucial for both fundamental understanding and applications [9,28]. FeSe 0.4 Te 0.6 single crystals exhibit pronounced curvature in H c2 , with exceptionally high upper critical fields (H ab c2 ≈ H c c2 ≈ 48 T) [29] and a superconducting anisotropy γ decreasing from 3 near T c to 0.99 at 3.8 K [29]. ...

Reference:

Upper critical field and critical current density of Zn and Mn substituted FeSe0.4Te0.6 single crystals
Temperature dependence and limiting mechanisms of the upper critical field of FeSe thin films

Physical Review B

... In monolayer transition metal dichalcogenides (TMDs), the range of excitonic states observed includes neutral excitons, or bound electron-hole pairs [4,5], their excited states [6], charged excitons [7], and biexcitons [8]. Under specific conditions, even six-and eight-particle complexes can be observed [9,10]. In stacked bilayers, this landscape is further enriched by the possibility of forming dipolar and quadrupolar interlayer excitons [11][12][13] in addition to the emergence of moiré excitons under suitable conditions [1,3]. ...

Emergence of composite many-body exciton states in WS 2 and MoSe 2 monolayers
  • Citing Article
  • January 2024

... The Landé g-factor of charge carriers and excitons determines their Zeeman splitting in magnetic eld and thus is a key parameter in spin physics. We found recently that in bulk lead halide perovskites the electron and hole as well as exciton g-factors follow universal dependences on the band gap energy [13,27]. In NCs, the carrier connement induces mixing of the electronic ground states with higher bands, which results in a con-siderable renormalization of the electron g-factor (g e ), as predicted theoretically and found experimentally in CsPbI 3 NCs in glass [14,28]. ...

Weak Dispersion of Exciton Landé Factor with Band Gap Energy in Lead Halide Perovskites: Approximate Compensation of the Electron and Hole Dependences

... It is known that the prominence of an excitonic peak at the absorption edge increases from iodide to bromide and chloride since materials with a wider band gap show a more pronounced excitonic peak, with exciton binding energies larger than the roomtemperature thermal energy k B T = 0.026 eV. 40,41 This means that for CsPbBr 3 and CsPb(Br/Cl) 3 the various excited-state species have a different relative importance in the pathway leading to NIR emission and that, although the NIR emission spectra of CsPbBr 3 and CsPb(Br/Cl) 3 are the same, the exciton and charge carrier dynamics associated with the NIR luminescence may be different. This can be expected based on the difference in exciton binding energy between the pure CsPbBr 3 and mixed-halide CsPb(Br/Cl) 3 , with excitons being energetically more stable in the latter, leading to a larger separation in energy between the exciton and band gap absorption. ...

Exciton‐Polaritons in CsPbBr 3 Crystals Revealed by Optical Reflectivity in High Magnetic Fields and Two‐Photon Spectroscopy
  • Citing Article
  • November 2023

physica status solidi (RRL) - Rapid Research Letters

... [37,38] Heterostructures composed of TMDs and NiPS 3 have recently exhibited novel phenomena driven by interactions between excitons and magnetic order via proximity effects. [39] Although NiPS 3 exhibits relatively weak interlayer coupling (as evidenced by its minimal thickness dependence of the Néel temperature), strain exerts a significant influence on its magnetic ordering. [40] While moiré superlattices are crucial for modulating quantum states at the nanoscale, experimental studies exploring the influence of magnetic materials on moiré excitons remain limited. ...

Proximity-induced chiral quantum light generation in strain-engineered WSe2/NiPS3 heterostructures

Nature Materials

... For S-TMD MLs, the g-factors are approximately equal to −4 [13,19,21,23,25,27,[44][45][46][47][48] for the A excitons. It was also shown that a significant variation in the ML dielectric environment does not affect the g-factor [49,50] contrary to other physical quantities such as the exciton binding energies and therefore the diamagnetic shift of excitons [49,[51][52][53][54]. Moreover, previous studies have found that the g-factor of a neutral exciton investigated in WSSe ML with almost 50/50 S/Se ratio is also close to −4 [55], similarly to the parent compounds WS 2 and WSe 2 . ...

Proximity-enhanced valley Zeeman splitting at the WS2/graphene interface

... The above argument implies that any bulk antiferromagnet (AFM) which has a nonzero ME response (at linear or higher order) has a corresponding roughnessrobust net magnetic dipole moment per unit area ("surface magnetization"), despite the vanishing bulk magnetization [5] (Fig. 1(a)). Such surface magnetization has utility as a directly detectable probe of the underlying AFM bulk Néel vector [4,6], even on surfaces parallel to atomic planes which are magnetically compensated in the bulk ( Fig. 1(a)) [4][5][6][7][8][9]. We showed in Ref. 5 that such "induced" surface magnetization in ME AFMs occurs when higher-order local bulk magnetic multipoles centered on the magnetic sites order ferroically in the unit cell ( Fig. 1(a), second and fourth panels). ...

Topological surface magnetism and Néel vector control in a magnetoelectric antiferromagnet

npj Quantum Materials

... The magnetoresistance in our model has a B/T scaling collapse, in alignment with experimental observations, albeit with a slight deviation from the proposed quadrature scaling ρ(B, T ) ∝ (µ B B) 2 + (k B T ) 2 in Refs. 4 and 5. Further, the magnetoresistance in our model for ω c ≳ ω * c scales as ρ(B, T ) ∝ (α − γ)k B T +μ B B, in accordance with the measured scaling 0.5 k B T +µ B B in recent experiments on nano-patterned YBCO [3]. Finally, we estimate a crossover field scale B * /T = 0.5 Tesla per Kelvin for moderate disorder strength k F ξ ≈ 10, and m f = 4m e [39], in reasonable agreement with experiments on cuprates and pnictides [1,4]. ...

Observation of cyclotron resonance and measurement of the hole mass in optimally doped La 2 − x Sr x CuO 4
  • Citing Article
  • April 2021

... Information on its optical and spin properties in low magnetic fields can be found in refs. [44][45][46]. Here we focus on exciton-polariton parameters and their modification in strong magnetic fields. ...

Weak dispersion of exciton Land\'e factor with band gap energy in lead halide perovskites: Approximate compensation of the electron and hole dependences

... With a small twisting angle in (nearly)commensurate bilayer TMDs an applied gate voltage may also brighten interlayer R M h 2p excitons, where the effective Hamiltonian can be written in the form of H BL in Eq. (15) [2,43,44]. It would be instructive to consider resonant mixing in the recently found valley-dependent magnetic proximity effects [45], leading to an intriguing possibility of the topological effects and gap closing at only one (K or K') valley [46] and the resulting change of the helicity of the emitted light. ...

Asymmetric magnetic proximity interactions in MoSe2/CrBr3 van der Waals heterostructures

Nature Materials