David A. Strubbe's research while affiliated with University of California and other places

What is this page?


This page lists the scientific contributions of an author, who either does not have a ResearchGate profile, or has not yet added these contributions to their profile.

It was automatically created by ResearchGate to create a record of this author's body of work. We create such pages to advance our goal of creating and maintaining the most comprehensive scientific repository possible. In doing so, we process publicly available (personal) data relating to the author as a member of the scientific community.

If you're a ResearchGate member, you can follow this page to keep up with this author's work.

If you are this author, and you don't want us to display this page anymore, please let us know.

Publications (27)


A computational materials science paradigm for a Course-based Undergraduate Research Experience (CURE)
  • Preprint

June 2024

·

3 Reads

David A. Strubbe

Course-based Undergraduate Research Experiences (CUREs) bring the excitement of research into the classroom to improve learning and the sense of belonging in the field. They can reach more students, earlier in their studies, than typical undergraduate research. Key aspects are: students learn and use research methods, give input into the project, generate new research data, and analyze it to draw conclusions that are not known beforehand. CUREs are common in other fields but have been rare in materials science and engineering. I propose a paradigm for computational material science CUREs, enabled by web-based simulation tools from nanoHUB.org that require minimal computational skills. After preparatory exercises, students each calculate part of a set of closely related materials, following a defined protocol to contribute to a novel class dataset which they analyze, and also calculate an additional property of their choice. This approach has been used successfully in several class projects.

Share

The computed values of ⟨S^2⟩ for the high-spin reference state and the N, T, V, and Z states of ethylene within SF-BSE. T is a triplet while the others are singlets. The spin of the many-body states is treated with a symmetric (‘ S ’) or antisymmetric (‘ A ’) combination of the individual spins.
The frontier orbitals of ethylene under zero torsion, in the high-spin reference state (up-spin orbitals occupied), plotted with XCrysDen [27]. The down-spin π¯∗ orbital in this illustration has a reversal of phase relative to the up-spin orbital.
The computed ⟨S^2⟩0 for ethylene for torsion angles from 0° to 90°. The High-Spin Reference state is a triplet, so the values are reported as deviations from the expected S(S+1)=2 .
The four cases (as in equation (18)) to consider for calculating matrix elements of the nested commutators of creation/annihilation operators, labelled as O in the figure, necessary for the calculation of Δ⟨S^2⟩ . The indices i and j represent the newly unoccupied state upon the SF excitation, for the bra and ket, respectively, and a¯ and b¯ represent the newly occupied state for the bra and ket, respectively. (a) The single-particle transitions for the bra and ket involve the same up-spin orbital (‘i = j’) and same down-spin orbital (‘ a¯=b¯ ’), (b) different up-spin orbitals but the same down-spin orbitals (‘i ≠ j’ but ‘ a¯=b¯ ’ ), (c) the same up-spin orbitals but different down-spin orbitals (‘i = j’ but ‘ a¯≠b¯ ’ ), and (d) different up-spin and down-spin orbitals (‘i ≠ j’ and ‘ a¯≠b¯ ’).
The electron configurations and values of ⟨S^2⟩ as calculated in SF-BSE for (a) the high-spin reference state, (b) the 3A2 (with MS=0 ) ground state, and (c) the ³ E excited state for the NV⁻ center in diamond. Values of ⟨S^2⟩ are presented for both the spin-restricted and spin-unrestricted cases.

+1

Computation of the expectation value of the spin operator S^2 for the spin-flip Bethe–Salpeter equation
  • Article
  • Publisher preview available

May 2024

·

18 Reads

Electronic Structure

Electronic Structure

Spin-flip (SF) methods applied to excited-state approaches like the Bethe–Salpeter equation allow access to the excitation energies of open-shell systems, such as molecules and defects in solids. The eigenstates of these solutions, however, are generally not eigenstates of the spin operator S^2 . Even for simple cases where the excitation vector is expected to be, for example, a triplet state, the value of ⟨S^2⟩ may be found to differ from 2.00; this difference is called ‘spin contamination’. The expectation values ⟨S^2⟩ must be computed for each excitation vector, to assist with the characterization of the particular excitation and to determine the amount of spin contamination of the state. Our aim is to provide for the first time in the SF methods literature a comprehensive resource on the derivation of the formulas for ⟨S^2⟩ as well as its computational implementation. After a brief discussion of the theory of the SF Bethe–Salpeter equation (BSE) and some examples further illustrating the need for calculating ⟨S^2⟩ , we present the derivation for the general equation for computing ⟨S^2⟩ with the eigenvectors from an SF-BSE calculation, how it is implemented in a Python script, and timing information on how this calculation scales with the size of the SF-BSE Hamiltonian.

View access options

Two-level quantum defect screening in WS2
a Two defect configurations that are considered in this work: substitution on the W site (MW, red) is shown on the left and on S site (MS, red) is depicted on the right. W atoms are colored green and S atoms are displayed in yellow. b Transition dipole moment vs. single-particle excitation energy at the single-shot PBE0. The marker and color scheme stand for the defect structure and whether the ground state is singlet or not. Each point stands for a charge defect that is thermodynamically stable within a certain Fermi level (EF) range in the band gap, and with electronic structures that possess two localized defect levels within the band gap, as shown in the inset. Below the conduction band (CB, light green) and above the valence band (VB, light blue), a filled state (orange arrow) to empty state (white arrow) transition is shown.
Molecular orbital trend within the 3d transition metal series for MS and MW defects
a The molecular orbital diagram shows the splitting between anti-bonding and bonding state (ΔAB) as well as the splitting with d orbitals (Δd) for a typical MW and MS defect. b A schematic of the bonding and anti-bonding state for different 3d transition metals in MW (blue) and MS (yellow) positions. The conduction band minimum (CBM) and valence band maximum (VBM) are drawn as black lines.
Thermodynamic charge transition levels and electronic structure of CoS
a Formation energy of CoS as a function of Fermi level for the neutral and the two charged states. The charge transition levels, i.e., (+/0) and (0/−), are referenced to the band-edge positions of pristine WS2 as obtained with PBE0 incorporating 22% of Fock exchange PBE0(0.22). b Orbital diagram of the localized defect states for neutral CoS. Resonant states within the valence band and conduction band manifolds are not depicted. The characters of the localized states are indicated by the specific d orbitals of the Co atom [e.g., Co(dz2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${d}_{{z}^{2}}$$\end{document}) as the highest occupied state in the minority channel] and, if any, the d orbitals of the W atoms in the nearest neighbor (WNN). The occupied (unoccupied) states are shown by the filled (empty) rectangles, the height of which indicates the degree of dispersion. The localized electrons in the majority (minority) channel are indicated by the arrows pointing up (down). The band-edge positions (in horizontal dashed lines) refer to those of the pristine WS2 obtained with PBE0 (0.22). Energies are referenced to the vacuum level. SOC is not taken into account for the localized defect states. c Top view of the charge density (in blue) for the three CoS0\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${}_{{{{{{{{\rm{S}}}}}}}}}^{0}$$\end{document} defect states as indicated in (b). The W and S atoms are represented by the gray and yellow spheres, respectively. The isovalue is 0.001 e/ų.
CoS defect formation and characterization
a The process of forming a high density of VS, (b) low-temperature deposition of Co atoms in situ, and (c) subsequent placement into a sulfur vacancy (VS) with the assistance of the STM probe that is used to selectively manipulate atoms at voltage ranges below −1.3 V is shown schematically. Corresponding scanning tunneling micrographs that capture WS2/Gr/SiC(0001) (d) after defect introduction via Ar⁺ bombardment and e post Co-deposition are plotted (Itunnel = 30 pA, Vsample = 1.2 V). Scale bars, 2 nm. STM images (f) before a voltage excitation and (g) after Co substitution within an identified VS are also shown (Itunnel = 30 pA, Vsample = 1.2 V, Vexcitation = −2.1 V). Scale bars, 2 nm. Itunnel is the tunneling current, Vsample is the sample bias voltage, and Vexcitation is the applied excitation voltage. h The apparent height difference of CoS compared to adsorption atop as-grown WS2 is measured to be 0.15 nm, taken from line scans across both (f) (maxima shown with blue dashed line) and (g) (maxima shown with magenta dashed line) red highlighted regions.
Experimental and simulated CoS scanning tunneling spectroscopy
a Scanning tunneling spectra (STS) recorded on a CoS defect and the as-grown WS2 monolayer on graphene (Vmodulation = 5 mV), where defect resonances, VBM and CBM onsets, in-gap states, and the shift between neutral (white background) to an anionic charge state (gray background) are labeled. Vmodulation is the bias modulation. b In-gap states identified are located at peak maxima of 0.36 eV and 0.47 eV, each with a full-width half maximum near 0.045 eV. Peak widths are broadened due to vibronic excitations (black lines). Differential conductance (dI/dV) imaging maps over the defect are depicted at (c) −0.9 eV (vertical black dashed line in a), (d) 0.373 eV (vertical green dashed lines in a,b), and (e) 0.486 eV (vertical red dashed lines in a,b) (Vmodulation = 5 mV), showing CoS orbital geometries. Scale bars, 0.25 nm. f–h Simulated STS maps using PBE0 over CoS orbitals identifying energy range densities near experimentally measured values. Scale bars, 0.25 nm. Isocontour value, 7 × 10⁻⁶ Å⁻³. A charging peak is identified in (a), where the (i) lowest unoccupied CoS0\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${}_{{{{{{{{\rm{S}}}}}}}}}^{0}$$\end{document} state becomes (j) resonant with the EF of the substrate and an electron is donated to the lowest unoccupied state (LUS) at sufficient Vsample (or equivalent tip potential, μtip) produce the CoS−1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${}_{{{{{{{{\rm{S}}}}}}}}}^{-1}$$\end{document} defect. Both (c) and (f) are representative of the CoS−1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${}_{{{{{{{{\rm{S}}}}}}}}}^{-1}$$\end{document} orbital densities collected at the specified energy (the charging ring onset in (c) is removed for clarity).
A substitutional quantum defect in WS2 discovered by high-throughput computational screening and fabricated by site-selective STM manipulation

April 2024

·

78 Reads

·

3 Citations

Nature Communications

John C. Thomas

·

·

·

[...]

·

Point defects in two-dimensional materials are of key interest for quantum information science. However, the parameter space of possible defects is immense, making the identification of high-performance quantum defects very challenging. Here, we perform high-throughput (HT) first-principles computational screening to search for promising quantum defects within WS2, which present localized levels in the band gap that can lead to bright optical transitions in the visible or telecom regime. Our computed database spans more than 700 charged defects formed through substitution on the tungsten or sulfur site. We found that sulfur substitutions enable the most promising quantum defects. We computationally identify the neutral cobalt substitution to sulfur (CoS0\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${}_{{{{{{{{\rm{S}}}}}}}}}^{0}$$\end{document}) and fabricate it with scanning tunneling microscopy (STM). The CoS0\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${}_{{{{{{{{\rm{S}}}}}}}}}^{0}$$\end{document} electronic structure measured by STM agrees with first principles and showcases an attractive quantum defect. Our work shows how HT computational screening and nanoscale synthesis routes can be combined to design promising quantum defects.


Figure 1. (a) Schematic illustration of the catalyst-free growth of Bi 2 Te 3 /F 4 -TCNQ core−shell nanoribbons. (i) Growth of Bi 2 Te 3 cores on a SiO 2 -coated silicon substrate in high vacuum and (ii) in situ surface coating with highly electronegative molecules, achieved by reorienting the quartz tube inside the furnace without environmental exposure. (b) Schematic illustration of a nanoribbon integrated with a thermoelectric characterization device. (Inset) Top-view scanning electron microscope image of the integrated device.
Figure 2. TEM analysis of a Bi 2 Te 3 /F 4 -TCNQ core−shell interface shows (a) a thin F 4 -TCNQ shell layer fully covering the Bi 2 Te 3 nanoribbons and (b) indication of an oxide-free Bi 2 Te 3 /F 4 -TCNQ interface in the region indicated by the red box in panel a. (c) FFT of panel b. (d) Electron energy loss spectroscopy at the O K-edge for the uncoated (blue lines) and coated (tan lines) nanoribbons.
Figure 3. Quantitative electron diffraction analysis of the Bi 2 Te 3 /F 4 -TCNQ core−shell nanoribbon on a thermoelectric characterization device. (a) Double-exposure TEM image, indicating the diffraction collection region. (b) Selected area electron diffraction pattern. (c) Weighted radial histogram of the peak positions before (blue dash-dot line) and after (red solid line) elliptical correction; for comparison, the expected histogram for bulk Bi 2 Te 3 is included (vertical solid lines; 28 † indicates X-ray-forbidden reflections derived from the (300) reflection). (c, inset) Peak location map aligned to the [1010] reflection, indicating the peaks used in the elliptical correction algorithm (red circles) to correct for inherent slight misalignments in the condenser, objective, and projector lens systems.
Enabling Oxidation Protection and Carrier-Type Switching for Bismuth Telluride Nanoribbons via in Situ Organic Molecule Coating

December 2023

·

14 Reads

·

1 Citation

Nano Letters

Thermoelectric materials with high electrical conductivity and low thermal conductivity (e.g., Bi2Te3) can efficiently convert waste heat into electricity; however, in spite of favorable theoretical predictions, individual Bi2Te3 nanostructures tend to perform less efficiently than bulk Bi2Te3. We report a greater-than-order-of-magnitude enhancement in the thermoelectric properties of suspended Bi2Te3 nanoribbons, coated in situ to form a Bi2Te3/F4-TCNQ core–shell nanoribbon without oxidizing the core–shell interface. The shell serves as an oxidation barrier but also directly functions as a strong electron acceptor and p-type carrier donor, switching the majority carriers from a dominant n-type carrier concentration (∼10²¹ cm–3) to a dominant p-type carrier concentration (∼10²⁰ cm–3). Compared to uncoated Bi2Te3 nanoribbons, our Bi2Te3/F4-TCNQ core–shell nanoribbon demonstrates an effective chemical potential dramatically shifted toward the valence band (by 300–640 meV), robustly increased Seebeck coefficient (∼6× at 250 K), and improved thermoelectric performance (10–20× at 250 K).


Fig. 1: Two-Level Quantum Defect Screening in WS 2 . a Two defect configurations that are considered in this work: substitution on W site (M W ) and on S site (M S ). b Transition dipole moment vs. single-
Fig. 2: Molecular orbital trend within the 3d transition metal series for M S and M W defects. a The molecular orbital diagram shows the splitting between anti-bonding and bonding state (∆AB) as well as the splitting with d orbitals (∆d) for a typical M W and M S defect. b A schematic of the bonding and anti-bonding state for different 3d transition metals in M W (blue) and M S (yellow) positions.
Fig. 3: Thermodynamic charge transition levels and electronic structure of Co S . a Formation energy of Co S as a function of Fermi level. The charge transition levels are referenced to the band-edge positions of pristine WS 2 as obtained with PBE0 incorporating 22% of Fock exchange PBE0(0.22). b Orbital diagram of the localized defect states for neutral Co S . Resonant states within the valence band and conduction band manifolds are not depicted. The occupied (unoccupied) states are shown by the filled (empty) rectangles, the height of which indicates the degree of dispersion. The band-edge positions refer to those of the pristine WS 2 obtained with PBE0(0.22). Energies are referenced to the vacuum level. SOC is not taken into account for the localized defect states. c Top view of the charge density (in blue) for the three Co 0 S defect states as indicated in b. The isovalue is 0.001 e/Å 3 .
Fig. 5: Experimental and Simulated Co S Scanning Tunneling Spectroscopy. a STS spectra recorded on a Co S defect and the as-grown WS 2 monolayer on graphene (V modulation = 5 mV). b In-gap states identified are located at peak maxima of 0.36 eV and 0.47 eV, each with a full-width half maximum near 0.045 eV. Differential conductance (dI/dV) imaging maps over the defect are depicted at c −0.9 eV, d 0.373 eV, and e 0.486 eV (V modulation = 5 mV), showing Co S orbital geometries. Scale bars, 0.25 nm. f-h
A substitutional quantum defect in WS2 discovered by high-throughput computational screening and fabricated by site-selective STM manipulation

October 2023

·

87 Reads

·

1 Citation

Point defects in two-dimensional materials are of key interest for quantum information science. However, the space of possible defects is immense, making the identification of high-performance quantum defects extremely challenging. Here, we perform high-throughput (HT) first-principles computational screening to search for promising quantum defects within WS 2 , which present localized levels in the band gap that can lead to bright optical transitions in the visible or telecom regime. Our computed database spans more than 700 charged defects formed through substitution on the tungsten or sulfur site. We found that sulfur substitutions enable the most promising quantum defects. We computationally identify the neutral cobalt substitution to sulfur (Co$_{\rm S}^{0}$) as very promising and fabricate it with scanning tunneling microscopy (STM). The Co$_{\rm S}^{0}$ electronic structure measured by STM agrees with first principles and showcases an attractive new quantum defect. Our work shows how HT computational screening and novel defect synthesis routes can be combined to design new quantum defects.


Pausing ultrafast melting by multiple femtosecond-laser pulses

June 2023

·

49 Reads

An intense femtosecond-laser excitation of a solid induces highly nonthermal conditions. In materials like silicon, laser-induced bond-softening leads to a highly incoherent ionic motion and eventually nonthermal melting. But is this outcome an inevitable consequence, or can it be controlled? Here, we performed ab initio molecular dynamics simulations of crystalline silicon after multiple femtosecond-laser pulse excitations with total energy above the nonthermal melting threshold. Our results demonstrate an excitation mechanism, which can be generalized to other materials, that pauses nonthermal melting and creates a metastable state instead, which has an electronic structure similar to the ground state. Our findings open the way for a search for metastable structural and/or electronic transitions in the high-excitation regime. In addition, our approach could be used to switch off nonthermal contributions in experiments, which could be used to obtain reliable electron-phonon coupling constants more easily.


Figure 1: Side views of the training set structures, illustrating the four possible locations of the Ni dopant within MoS 2 : (a) Mo-substituted, (b) S-substituted, (c) octahedral intercalation, and (d) tetrahedral intercalation. Sphere colors indicate S (yellow), Mo (green), and Ni (red).
Figure 9: Relaxation of a dopant/vacancy structure as calculated by DFT: (a) tetrahedrally intercalated Ni adjacent to an S vacancy; (b) Ni has moved into the S site, forming Ssubstituted MoS 2 .
Figure 10: Side-view snapshots of the model system and radial distribution function (I) before deposition, (II) at the end of the deposition, and (III) at the end of the annealing stage. A clear transition from amorphous to crystalline is observed (II) → (III). Spheres represent S (yellow), Mo (green), and Ni (red) atoms.
Comparison of structural parameters as obtained from DFT and ReaxFF for bulk Ni-doped MoS 2
Development and Demonstration of a ReaxFF Reactive Force Field for Ni-Doped MoS$_2

February 2023

·

88 Reads

The properties of $\mathrm{MoS_2}$ can be tuned or optimized through doping. In particular, Ni doping has been shown to improve the performance of $\mathrm{MoS_2}$ for various applications, including catalysis and tribology. To enable investigation of Ni-doped $\mathrm{MoS_2}$ with reactive molecular dynamics simulations, we developed a new ReaxFF force field to describe this material. The force field parameters were optimized to match a large set of density-functional theory (DFT) calculations of 2H-$\mathrm{MoS_2}$ doped with Ni, at four different sites (Mo-substituted, S-substituted, octahedral intercalation, and tetrahedral intercalation), under uniaxial, biaxial, triaxial, and shear strain. The force field was evaluated by comparing ReaxFF- and DFT-relaxed structural parameters and the tetrahedral/octahedral energy difference in doped 2H, energies of doped 1H and 1T monolayers, and doped 2H structures with vacancies. We demonstrated the force field with reactive simulations of sputtering deposition and annealing of Ni-doped MoS$_2$ films. Results show that the developed force field can accurately model the phase transition of Ni-doped $\mathrm{MoS_2}$ from amorphous to crystalline. The newly developed force field can be used in subsequent investigations to study the properties and behavior of Ni-doped $\mathrm{MoS_2}$ using reactive molecular dynamics simulations.


Quantifying Hidden Symmetry in the Tetragonal CH$_3$NH$_3$PbI$_3$ Perovskite

January 2023

·

11 Reads

The assignment of an exact space group to the tetragonal CH$_3$NH$_3$PbI$_3$ perovskite structure is experimentally challenging and controversial in the literature. Average orientation of the methylammonium ion that gives symmetry to the experimental measurement is not captured in a static density functional theory calculation, although the quasi-I4cm and quasi-I4/mcm structures are commonly used in calculations. In this work we have developed a methodology to quantify the hidden symmetry of these structures using group theory, to enable use of symmetries in understanding spectroscopy and other properties. We study the approximate symmetry of vibrational modes, including analysis of degenerate representations, as well as the dielectric, elastic, electro-optic, Born effective charge, and Raman tensors and the dynamical matrix. Comparing to each subgroup of the full tetragonal D$_{4h}$, our results show that the quasi-I4cm is best described by the expected corresponding point group C$_{4v}$, whereas the quasi-I4/mcm (despite corresponding to point group D$_{4h}$) is best described by the lower symmetry C$_{2v}$. Our methodology can be useful generally for analysis of other soft hybrid materials or any approximately symmetric material.


Surface Effects on Anisotropic Photoluminescence in One-Dimensional Organic Metal Halide Hybrids

December 2022

·

178 Reads

One-dimensional (1D) organic metal halide hybrids exhibit strongly anisotropic optical properties, highly efficient light emission, and large Stokes shift, holding promises for novel photodetection and lighting applications. However, the fundamental mechanisms governing their unique optical properties and in particular the impacts of surface effects are not understood. Here, we investigate 1D C4N2H14PbBr4 by polarization-dependent time-averaged and time-resolved photoluminescence (TRPL) spectroscopy, as a function of photoexcitation energy. Surprisingly, we find that the emission under photoexcitation polarized parallel to the 1D metal halide chains can be either stronger or weaker than that under perpendicular polarization, depending on the excitation energy. We attribute the excitation-energy-dependent anisotropic emission to fast surface recombination, supported by first-principles calculations of optical absorption in this material. The fast surface recombination is directly confirmed by TRPL measurements, when the excitation is polarized parallel to the chains. Our comprehensive studies provide a more complete picture for a deeper understanding of the optical anisotropy in 1D organic metal halide hybrids.


Bulk Assembly of Organic Metal Halide Nanoribbons

November 2022

·

23 Reads

Organic metal halide hybrids with low-dimensional structures at the molecular level have received great attention recently for their exceptional structural tunability and unique photophysical properties. Here we report for the first time the synthesis and characterization of a one-dimensional (1D) organic metal halide hybrid material, which contains metal halide nanoribbons with a width of three octahedral units. It is found that this material with a chemical formula C$_8$H$_{28}$N$_5$Pb$_3$Cl$_{11}$ shows a dual emission with a photoluminescence quantum efficiency (PLQE) of around 25% under ultraviolet (UV) light irradiation. Photophysical studies and density functional theory (DFT) calculations suggest the coexisting of delocalized free excitons and localized self-trapped excitons in metal halide nanoribbons leading to the dual emission. This work shows once again the exceptional tunability of organic metal halide hybrids that bridge between molecular systems with localized states and crystalline ones with electronic bands.


Citations (8)


... Defects control the properties of many functional materials and devices 1 , like solar cells 2,3 , batteries 4,5 , catalysts [6][7][8] , and quantum computers [9][10][11][12] . To discover better materials for these applications it is thus necessary to predict how their defects behave. ...

Reference:

Machine-learning structural reconstructions for accelerated point defect calculations
A substitutional quantum defect in WS2 discovered by high-throughput computational screening and fabricated by site-selective STM manipulation

Nature Communications

... These functionals are therefore denoted as non-decomposable. The potentials based on inverting the KS equation [21][22][23][24][25] also represent a bottom-up construction subject to the condition that the admissible densities are considered. 26 The bottom-up approach has been further used in orbital-free DFT methods (OF-DFT) by Chai and Weeks. ...

Nuclear cusps and singularities in the nonadditive kinetic potential bifunctional from analytical inversion
  • Citing Article
  • October 2022

Physical Review A

... Particularly, recent studies demonstrate accurate description of ground electronic states for several materials with strong local Coulomb repulsions and SOCs such as transition metal oxides [13], pyrochlore iridates [14], americium monochalcogenides [15] and osmium-based double perovskites [16]. We also note that recent developments of fully relativistic GW approximations (FR-GWA) [17][18][19][20][21][22][23][24][25][26] obtained spin-orbit split energy bands of semiconductors in good agreements with experiments. ...

Spinor G W /Bethe-Salpeter calculations in BerkeleyGW: Implementation, symmetries, benchmarking, and performance
  • Citing Article
  • September 2022

... At high temperatures, the random spinning of the MA + ion within the cage makes the structure pseudo-cubic, and is not even close to any symmetry, complicating theoretical analyses. 15 For the room-temperature tetragonal structure, the average over space and time of this random spinning makes this structure quasi-I4cm or quasi-I4/mcm. So, the tetragonal MAPI does not have any exact symmetry, but is considered to have approximate symmetry. ...

Stress Effects on Vibrational Spectra of a Cubic Hybrid Perovskite: A Probe of Local Strain
  • Citing Article
  • December 2020

The Journal of Physical Chemistry C

... N(λ) of the film in its region of medium and strong absorption. Some amorphous materials can contain voids [93,94], and the voids volume fraction (with respect to the entire volume of the material) has been approximated as follows: ...

Computational generation of voids in a -Si and a -Si:H by cavitation at low density
  • Citing Article
  • February 2020

Physical Review Materials

... regard, this work shares parallels with attempts to learn the U parameter for DFT+U functionals [28][29][30][31] and the dielectric screening when solving the Bethe-Salpeter equation [32] (albeit in the latter case the dielectric screening is a physical observable, but for the purposes of that work it was used as an ingredient for subsequent calculations of optical spectra). This work also differs from ML methods that seek to relate structural information directly to observable quantities on a practical level [33][34][35][36][37][38][39][40][41][42], because in our case first-principles calculations will not be bypassed completely. ...

Deep learning and density-functional theory
  • Citing Article
  • August 2019

Physical Review A

... Solid lubrication is an environmentally friendly manufacture technology with properties of high temperature resistance, high vacuum, non-volatility, low coefficient of friction, etc. [1][2][3][4]. The solid particles also include materials with layered structures, soft metals with multiple slip surfaces, metal oxides, metal fluorides, and multi-phase solid lubricating coatings [5][6][7]. Graphite and molybdenum disulfide have proven suitable for mechanical engineering applications under extreme conditions, including aerospace and high-temperature molding [8,9]. In terms of experimental research, Chen et al. [10] prepared a large area graphite layer with low friction coefficient, excellent wear resistance, durability and high load capacity by a self-assembly technique. ...

Solid Lubrication with MoS 2 : A Review

Lubricants

... For these calculations, we again use the Octopus code [120][121][122][123][132][133][134][135] and the LDA functional. The quasi-minimal residual (QMR) method 136 (qmr_dotp) with the final tolerance of 10 −6 is applied to solve linear equations. ...

Orbital magneto-optical response of periodic insulators from first principles

npj Computational Materials