Janina Maultzsch’s research while affiliated with Friedrich-Alexander-University Erlangen-Nürnberg 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 (295)


a) 2 fabricating patterned TMD/G heterostructures on a Si/SiO2 substrate. The gray and yellow plates represent CVD‐grown graphene and the deposited fun‐TMD layer, respectively, while the green block pattern indicates the photochemically modified domain created by direct laser writing. b) Illustration of major chemical processes in (a): Chemical exfoliation of bulk MoS₂ powder produces monolayer‐rich TMD nanosheets (ce‐TMDs) dispersed in water. Covalent functionalization of ce‐TMDs with diazonium salts forms functionalized TMDs (fun‐TMDs), with 4‐iodophenyl groups tethered to the basal plane. Upon laser irradiation of fun‐TMD/graphene stacks, photochemical reactions such as deiodination and phenyl radical formation may occur. These radicals can then undergo a series of reactions with surrounding materials or chemical species (red dashed arrows), leading to photochemically modified heterostructure domains exclusively in the irradiated areas. Atoms are displayed as follows: white for H, gray for C, red for O, yellow for S/Se, cyan for Mo/W, and purple for I.
a) Raman spectra (λex = 532 nm) of ce‐MoS2, fun‐MoS2, pristine G, and fun‐MoS2/G (measured at 17 and 2 mW). The Raman intensities of ce‐MoS2, fun‐MoS2, and fun‐MoS2/G are normalized to the intensity of the E¹2g mode. Orange, blue, and red dots mark MoS2, graphene, and functional group‐related peaks. The asterisk labels the Raman peaks that correspond to the silicon substrate. b) I(2D)/I(G) ratio of fun‐MoS2/G as a function of laser power, with a dashed trendline for guidance. c) Raman I(2D)/I(G) map of patterned fun‐MoS2/G (λex = 532 nm, laser power = 7 mW), showing photochemically modified regions with lower I(2D)/I(G) ratios.
a) AFM image of patterned MoS2/G heterostructures. The blue dashed block indicates the location of highly irradiated region. b) KPFM image of patterned MoS2/G heterostructures.
a) XPS survey spectra and b) C 1s core level spectra of non‐irradiated (bottom) and irradiated (top) regions of fun‐MoS2/G heterostructures. The C 1s spectra are deconvoluted using a Lorentzian function. c) Optical image of the patterned MoS2/G heterostructures, where fun‐MoS2 appears as the black layer, and the transparent graphene beneath shows a slight blue reflection. Green areas indicate laser‐written patterns, while the red dashed box highlights the ToF‐SIMS mapping area, comparing the non‐irradiated (upper) and irradiated (lower) regions. d–f) High‐resolution ToF‐SIMS intensity maps (negative mode) of MoS2⁻, O⁻, and OH⁻ secondary ions on the surface of fun‐MoS2/G, with O⁻ and OH⁻ signals predominantly detected in the irradiated region. The color scales in (d–f) represent intensity in counts.
Relaxed structures of I‐Ph‐2MoS2 a), BPh‐2MoS2 b), OH‐Ph‐2MoS2 c), and Ph‐2MoS2 d). The atoms are displayed as follows: white for H, black for C, red for O, yellow for S, gray for Mo, and purple for I. For motifs (a,c), the functional groups must be positioned as far apart as possible; otherwise, functionalization on sulfur atoms with the same coordinates within the MoS2 plane would necessitate a significant increase in the interlayer distance.
Patterned Assembly of Transition Metal Dichalcogenide/Graphene Heterostructures via Direct Laser Writing
  • Article
  • Full-text available

May 2025

·

23 Reads

Xin Chen

·

Stefan Wolff

·

Sofiia Zuieva

·

[...]

·

Andreas Hirsch

Connecting two‐dimensional (2D) material layers via interface linkers represents a new avenue for fabricating 2D heterostructures. Utilizing light to remotely modulate this interface function allows for seamless assembly and patterning in a single run. Here, an efficient method for fabricating patterned 2D heterostructures using direct laser writing is demonstrated, drawing a conceptual parallel to laser printing. In the approach, functionalized transition metal dichalcogenide (TMD) dispersions serve as inks, graphene as the substrate, and a Raman laser as the patterning tool. Unlike laser printing's electrostatic interactions, the method achieves patterned assembly through covalent bonding between TMDs and graphene. Selective Raman laser irradiation of functionalized TMD/graphene heterostructures triggers localized reactions, forming chemically modified domains exclusively in the laser‐irradiated regions, as confirmed by Raman spectroscopy, Kelvin probe force microscopy (KPFM), and time‐of‐flight secondary ion mass spectrometry (ToF‐SIMS). Experimental and theoretical analyses of the interface composition and structure provide new insights into laser‐induced chemistry. The work demonstrates the potential for high‐throughput assembly of customizable 2D heterostructures, with enhanced compatibility for subsequent patterning through photolabile linkers and photoinduced coupling. Additionally, the results provide deeper insights into chemistry within confined 2D spaces, offering a novel approach to nanoscale heterostructure engineering.

Download

Semimetallic 2D Defective Graphene Networks with Periodic 4–8 Defect Lines

April 2025

·

6 Reads

physica status solidi (b)

Theoretical simulations of the electronic properties of graphene‐like 2D carbon networks with a periodic arrangement of defect lines formed by alternating four‐ and eight‐membered rings are presented. These networks can be seen as arrays of armchair‐edged nanoribbons (AGNRs), which are covalently connected at the edges. Using a combination of density functional theory and a simple tight binding model, it is shown that the electronic properties of these networks can be understood to arise from the family behavior of the constituting AGNRs, plus a rigid shift due to an “inter‐ribbon” coupling across the defect lines. As a result, one class of zero‐bandgap semiconducting 2D networks and two classes of semimetallic networks with quasilinear band close to the Fermi energy are found. The formation of closed‐ring electron‐ and hole‐like Fermi surfaces due to hybridization across the defect lines offers interesting perspectives of using such defective 2D networks for transport applications or the realization of carbon‐based nodal‐line semimetals.


Semimetallic two-dimensional defective graphene networks with periodic 4-8 defect lines

April 2025

·

9 Reads

We present theoretical simulations of the electronic properties of graphene-like two-dimensional (2D) carbon networks with a periodic arrangement of defect lines formed by alternating four- and eight-membered rings. These networks can be seen as arrays of armchair-edged nanoribbons (AGNRs), which are covalently connected at the edges. Using a combination of density functional theory and a simple tight binding model, we show that the electronic properties of these networks can be understood to arise from the family behaviour of the constituting AGNRs, plus a rigid shift due to an 'inter-ribbon' coupling across the defect lines. As a result, we find one class of zero-band-gap semiconducting 2D networks, and two classes of semimetallic networks with quasi-linear band close to the Fermi energy. The formation of closed-ring electron- and hole-like Fermi surfaces due to hybridization across the defect lines offers interesting perspectives of using such defective 2D networks for transport applications or the realization of carbon-based nodal line semimetals.



Figure 3. Spectroscopic characterizations of (4,2)-CcGNR and (6,2)-CcGNR. (a) MALDI-TOF MS analysis of polymer P1 and P2 (matrix: DCTB, linear mode). (b) FTIR spectra of P1, P2 (4,2)-CcGNR and (6,2)-CcGNR. (c) Raman spectra of (4,2)-CcGNR and (6,2)-CcGNR measured at 532 nm excitation wavelength. (d) UV-vis absorption spectra of model compounds 1 and 2 in CH2Cl2 (10 -5 M), (4,2)-CcGNR and (6,2)-CcGNR in NMP.
Figure 4. (a) Time-resolved complex terahertz photoconductivity of both (4,2)-CcGNR and (6,2)-CcGNR. (b) Frequency-resolved terahertz conductivity measured at ∼1.5 ps after photoexcitation. The solid lines are fits to the Drude−Smith model.
Scheme 1. Synthetic Route toward Model Compounds 1 and 2. Regents and conditions: (a) (Triisopropylsilyl)acetylene, CuI, PdCl2(PPh3)2, THF/TEA, r.t., 12 h, 95%; (b) n-BuLi, I2, THF, -78 °C, 16 h, 97%; (c) i. (3-bromonaphthalen-2-yl)boronic acid, Pd(PPh3)4, K2CO3, THF/EtOH/H2O, 60 o C, 36 h; ii. TBAF, THF, r.t., 20 min, 79%; (d) NBS, AgNO3, acetone, r.t., 1 h, 66%; (e) InCl3, toluene, 95 o C, 24 h, 87%; (f) CuI, piperidine, toluene, Air, r.t., 6 h, 83%; (g) PtCl2, toluene, 90 o C, 24 h, 93%; (h) 2-chloroiodobenzene, Pd(PPh3)4, Na2CO3, nBu4NBr, THF/EtOH/H2O, 60 o C, 10 h, 85%; (i) n-BuLi, triisopropyl borate, THF, -78 °C, 16 h, 61%; (j) Pd(PPh3)4, K2CO3, toluene/EtOH/H2O, 95 o C, 48 h, 55% for 14 and 60% for 15; (k) DDQ, TfOH, DCM, 0 o C, 45 min, 62% for 1 and 65% for 2.
Cove-edged Chiral Graphene Nanoribbons with Chirality-Dependent Bandgap and Carrier Mobility

February 2025

·

56 Reads

Graphene nanoribbons (GNRs) have garnered significant interest due to their highly customizable physicochemical properties and potential utility in nanoelectronics. Besides controlling widths and edge structures, the inclusion of chirality in GNRs brings another dimension for fine-tuning their optoelectronic properties, but related studies remain elusive owing to the absence of feasible synthetic strategies. Here, we demonstrate a novel class of cove-edged chiral GNRs (CcGNRs) with a tunable chiral vector (n,m). Notably, the bandgap and effective mass of (n,2)- CcGNR show a distinct positive correlation with the increasing value of n, as indicated by theory. Within this GNR family, two representative members, namely, (4,2)- CcGNR and (6,2)-CcGNR, are successfully synthesized. Both CcGNRs exhibit prominently curved geometries arising from the incorporated [4]helicene motifs along their peripheries, as also evidenced by the single-crystal structures of the two respective model compounds (1 and 2). The chemical identities and optoelectronic properties of (4,2)- and (6,2)-CcGNRs are comprehensively investigated via a combination of IR, Raman, solid-state NMR, UV-vis, and THz spectroscopies as well as theoretical calculations. In line with theoretical expectation, the obtained (6,2)-CcGNR possesses a low optical bandgap of 1.37 eV along with charge carrier mobility of 8 cm2/Vs, whereas (4,2)-CcGNR exhibits a narrower bandgap of 1.26 eV with increased mobility of 14 cm2/Vs. This work opens up a new avenue to precisely engineer the bandgap and carrier mobility of GNRs by manipulating their chiral vector.


Figure 1. Illustration of the step-by-step approach to prepare and transfer bilayer MoS2 flake onto a gold-coated TEM grid. (a and b) MoS2 flakes were exfoliated using PDMS gel and transferred onto a SiO2/Si substrate (steps 1 and 2). (c) The gold-coated TEM grid was carefully positioned in relation to the target flakes under the light microscope. The target flake was then transferred to the grid by sequentially applying IPA and KOH solution. Finally, the MoS2 flake was cleaned using water and acetone bath, followed by heating on a hotplate in air.
Figure 2. Crystal structure and Burgers vector analysis on different types of dislocation, implying the presence of 2H and 3R polytypes in the exfoliated bilayer MoS2 flake. (a) and (d) Top and side views of crystalline structures for bilayer 2H and 3R MoS2. The red and blue arrows in the top views represent the corresponding Burgers vectors of type (1/3)〈2110〉 and (1/3)〈0110〉. (b) and (c) Series of {1100} and {2110} DF TEM images acquired from the
Figure 3. Correlative determination of 2H and 3R polytypes within the bilayer MoS2 flake using 3D ED, Raman and PL spectroscopy. (a) and (b) Rocking curves of (1100) (red dots) and (1120) (black dots) reflections, extracted from 3D ED datasets (insets), which were acquired from dislocation-free areas of bilayer MoS2. Solid lines are the simulated rocking curves. The rocking curves confirm the layer numbers as well as the different atomic staking orders (i.e., polymorphism). (c) and (e) Low-frequency Raman and PL spectra of the 2H and 3R bilayer MoS2; the peaks were fit by Lorentzians (dashed lines). (d) The amplitude ratio and frequency difference between LB and S modes in 2H and 3R bilayer MoS2. The inset in (e) highlights the average value of the A exciton and B exciton energy differences between 2H and 3R bilayer MoS2. The error bars in (d) and inset in (e) represent the standard deviation derived from three independent Raman and PL spectra measured from three different sample regions (holes on the TEM grid).
Identification of Polytypism and Their Dislocations in Bilayer MoS2 Using Correlative Transmission Electron Microscopy and Raman Spectroscopy

February 2025

·

12 Reads

Stacking orders and topological defects substantially influence the physical properties of 2D van der Waals (vdW) materials. However, the inherent features of 2D materials challenge the effectiveness of single characterization techniques in identifying stacking sequences, necessitating correlative approaches. Using bilayer MoS2 as a benchmark, we differentiate its polytypism and specific dislocations through transmission electron microscopy (TEM) and Raman spectroscopy. Perfect and partial dislocations were revealed in TEM, which are closely linked to the stacking sequences, thus indirectly indicating the 2H and 3R polytypes. 3D electron diffraction reconstruction on relrods and low-frequency Raman spectroscopy further validated these polytypes owing to their reliance on crystal symmetry. Surprisingly, we unexpectedly resolved both polytypes despite starting with 2H bulk crystal, pointing to a possible phase transition during mechanical exfoliation. The correlative TEM-Raman approach can be extended to other 2D materials, paving the way for property alteration via stacking and defect engineering.


(a) Brillouin zone of twisted bilayer graphene with corresponding mini Brillouin zone and rotational wavevector q1 ${{\bf{q}}_1 }$ . (b) Molecular structure of 1,4,5,8,9,11‐hexaazatriphenylenehexacarbonitrile (HATCN).
(a) Calculated phonon dispersion of single‐layer graphene. The green arrow indicates the rotational wavevector q1 ${{\bf{q}}_1 }$ , which represents the difference between the two reciprocal lattice vectors of the two layers in tBLG. (b) Phonon dispersion as a function of twist angle ϑ ${\vartheta }$ . Red circles indicate the moiré phonons for a given q1 ${{\bf{q}}_1 }$ [same as in (a)].(c) Raman spectra ( λ=532.17 ${\lambda = 532.17}$ nm) of a ∼6∘ ${ \sim 6^\circ }$ twisted bilayer graphene sample with corresponding mode, before (grey) and after (purple) HATCN functionalization. The characteristic mode of the HATCN molecule is indicated.
(a) Raman map of the 2D‐mode frequency after non‐covalent functionalization. The 2D mode was fitted by a single Lorentzian for all graphene layers. Areas with different layer numbers are marked in the map. The area framed by the blue line shows the twisted bilayer (2 L) area. (b) Raman map of the R′‐mode frequency and corresponding twist angle after non‐covalent functionalization. The green framed area shows no R′ mode because the twist angle is smaller than ≲ 5.3°. Red arrows point to areas with higher R′‐mode frequency and therefore larger twist angle. There is no significant change in the twist angle map before and after chemical functionalization. (c) Raman map showing the intensity of the HATCN peak. The green (high HATCN intensity) and cyan (low HATCN intensity) crosses mark the position of the spectra in Figure 3 (d). (d) exemplary Raman spectra of tBLG with non‐covalent HATCN functionalization, emphasizing the intensity difference of the HATCN peak and the change in frequency of the R′ mode at different positions on the sample.
(a) and (b); moiré lattice of 5° and 7° twisted bilayer graphene, respectively. Schematic views of possible arrangements of HATCN molecules (red) on AB‐stacked regions, showing different densities of the HATCN molecules for the two twist angles. In the schematic view of the 7° moiré lattice, the nitrogen atoms of the CN groups are too close to each other. As a result, the HATCN molecules cannot attach to each AB‐stacked region. For additional arrangements see Supporting Information Figure S9.
Comparsion of the HATCN peak intensity distribution at 50 °C (left) and 100 °C (right) of the same tBLG sample shown in Figure 3. A clear increase of the HATCN peak intensity is observed in the upper left part of the twisted bilayer area (green circle) at 100 °C.
Moiré Lattice of Twisted Bilayer Graphene as Template for Non‐Covalent Functionalization

December 2024

·

17 Reads

We present a novel approach to achieve spatial variations in the degree of non‐covalent functionalization of twisted bilayer graphene (tBLG). The tBLG with twist angles varying between ~5° and 7° was non‐covalently functionalized with 1,4,5,8,9,11‐hexaazatriphenylenehexacarbonitrile (HATCN) molecules. Our results show a correlation between the degree of functionalization and the twist angle of tBLG. This correlation was determined through Raman spectroscopy, where areas with larger twist angles exhibited a lower HATCN peak intensity compared to areas with smaller twist angles. We suggest that the HATCN adsorption follows the moiré pattern of tBLG by avoiding AA‐stacked areas and attach predominantly to areas with a local AB‐stacking order of tBLG, forming an overall ABA‐stacking configuration. This is supported by density functional theory (DFT) calculations. Our work highlights the role of the moiré lattice in controlling the non‐covalent functionalization of tBLG. Our approach can be generalized for designing nanoscale patterns on two‐dimensional (2D) materials using moiré structures as a template. This could facilitate the fabrication of nanoscale devices with locally controlled varying chemical functionality.


Moiré Lattice of Twisted Bilayer Graphene as Template for Non‐Covalent Functionalization

November 2024

·

6 Reads

Angewandte Chemie

We present a novel approach to achieve spatial variations in the degree of non‐covalent functionalization of twisted bilayer graphene (tBLG). The tBLG with twist angles varying between ~5° and 7° was non‐covalently functionalized with 1,4,5,8,9,11‐hexaazatriphenylenehexacarbonitrile (HATCN) molecules. Our results show a correlation between the degree of functionalization and the twist angle of tBLG. This correlation was determined through Raman spectroscopy, where areas with larger twist angles exhibited a lower HATCN peak intensity compared to areas with smaller twist angles. We suggest that the HATCN adsorption follows the moiré pattern of tBLG by avoiding AA‐stacked areas and attach predominantly to areas with a local AB‐stacking order of tBLG, forming an overall ABA‐stacking configuration. This is supported by density functional theory (DFT) calculations. Our work highlights the role of the moiré lattice in controlling the non‐covalent functionalization of tBLG. Our approach can be generalized for designing nanoscale patterns on two‐dimensional (2D) materials using moiré structures as a template. This could facilitate the fabrication of nanoscale devices with locally controlled varying chemical functionality.


Production of Ultrathin and High-Quality Nanosheet Networks via Layer-by-Layer Assembly at Liquid-Liquid Interfaces

November 2024

·

67 Reads

·

4 Citations

ACS Nano

Solution-processable 2D materials are promising candidates for a range of printed electronics applications. Yet maximizing their potential requires solution-phase processing of nanosheets into high-quality networks with carrier mobility (μNet) as close as possible to that of individual nanosheets (μNS). In practice, the presence of internanosheet junctions generally limits electronic conduction, such that the ratio of junction resistance (RJ) to nanosheet resistance (RNS), determines the network mobility via μNS/μNet ≈ RJ/RNS + 1. Hence, achieving RJ/RNS < 1 is a crucial step for implementation of 2D materials in printed electronics applications. In this work, we utilize an advanced liquid-interface deposition process to maximize nanosheet alignment and network uniformity, thus reducing RJ. We demonstrate the approach using graphene and MoS2 as model materials, achieving low RJ/RNS values of 0.5 and 0.2, respectively. The resultant graphene networks show a high conductivity of σNet = 5 × 10⁴ S/m while our semiconducting MoS2 networks demonstrate record mobility of μNet = 30 cm²/(V s), both at extremely low network thickness (tNet < 10 nm). Finally, we show that the deposition process is compatible with nonlayered quasi-2D materials such as silver nanosheets (AgNS), achieving network conductivity close to bulk silver for networks <100 nm-thick.


Fig 4. Optoelectronic properties of MoS2 stacked monolayer networks. A. UV-visible absorbance spectra for MoS2 networks 1 to 6 layers thick. B. The PL signal from a 1L MoS2 film (gray circles). This spectrum has been fitted to two Lorentzian components (dashed lines). The overall fit is shown by the solid black line. C. Plot of absorbance measured at a photon energy of 2eV vs thickness for MoS2 films. The red line is a linear fit to the data and the absorbance coefficient at 2eV, α2eV is indicated in the plot. D. Transmission scanner histograms showing local network thickness for MoS2 networks from 1L to 6L. Inset: Ratio of standard deviation of network thickness to mean network thickness plotted versus mean network thickness. E. Contact resistance, expressed as RCw, where w is the channel width (2mm), plotted vs thickness. F. Conductivity versus thickness for 1L to 6L MoS2 networks.
Fig 6. Beyond layered 2D materials. A. Reflectance spectra for 1L to 3L AgNS networks. B.
Production of Ultra-Thin and High-Quality Nanosheet Networks via Layer-by-Layer Assembly at Liquid-Liquid Interfaces

October 2024

·

94 Reads

Solution-processable 2D materials are promising candidates for a range of printed electronics applications. Yet maximising their potential requires solution-phase processing of nanosheets into high-quality networks with carrier mobility ({\mu}Net) as close as possible to that of individual nanosheets ({\mu}NS). In practise, the presence of inter-nanosheet junctions generally limits electronic conduction, such that the ratio of junction resistance (RJ) to nanosheet resistance (RNS), determines the network mobility via . Hence, achieving RJ/RNS<1 is a crucial step for implementation of 2D materials in printed electronics applications. In this work, we utilise an advanced liquid-interface deposition process to maximise nanosheet alignment and network uniformity, thus reducing RJ. We demonstrate the approach using graphene and MoS2 as model materials, achieving low RJ/RNS values of 0.5 and 0.2, respectively. The resultant graphene networks show a high conductivity of {\sigma}Net = 5 \times 104 S/m while our semiconducting MoS2 networks demonstrate record mobility of {\mu}Net = 30 cm2/Vs, both at extremely low network thickness (tNet <10 nm). Finally, we show that the deposition process is compatible with non-layered quasi-2D materials such as silver nanosheets (AgNS), achieving network conductivity close to bulk silver for networks <100 nm thick. We believe this work is the first to report nanosheet networks with RJ/RNS<1 and serves to guide future work in 2D materials-based printed electronics.


Citations (56)


... [14] Particularly promising has been the printing of MoS 2 -based transistors with state-ofthe-art mobility (up to 30 cm 2 /Vs). [3][4][5]15,16] Such devices will be important in applications (e.g., wearable devices) where they will be routinely exposed to mechanical deformations. The resultant-induced strains can affect the electronic properties of the materials comprising the device, leading to strain-induced changes in device operation. ...

Reference:

Using Electrical Impedance Spectroscopy to Separately Quantify the Effect of Strain on Nanosheet and Junction Resistance in Printed Nanosheet Networks
Production of Ultrathin and High-Quality Nanosheet Networks via Layer-by-Layer Assembly at Liquid-Liquid Interfaces
  • Citing Article
  • November 2024

ACS Nano

... The peak positions are characteristic of multilayered 2H-WS2 [55]. The secondorder 2LA mode is activated by defects, in particular sulfur vacancies [56]. No obvious differences were observed between the Raman spectrum of the WS2-20s film, where the particles have an almost horizontal orientation, and the spectra of the WS2-30s and WS2-90s films with mixed horizontal and vertical morphologies. ...

TEM-processed defect densities in single-layer TMDCs and their substrate-dependent signature in PL and Raman spectroscopy

... When light-coupling is favored, the interlayer moiré excitons can be brightened through hybridization with a nearly resonant intralayer exciton. In heterobilayers, this is made possible under various compound choices that have nearly aligned conduction or valence bands 8,[19][20][21] (Fig. 1a). In moiré traps of different local stacking registries, the C 3 rotational symmetry dictates different center-of-mass envelope forms of the intralayer component hybridized to a tightly trapped interlayer exciton wavepacket (Fig. 1c), whereas an s-wave envelope leads to brightening. ...

Hybrid Moiré Excitons and Trions in Twisted MoTe2-MoSe2 Heterobilayers
  • Citing Article
  • April 2024

Nano Letters

... In the latter class, recently proposed materials include T-carbon [7], T-graphene [8,9] and PHOTHgraphene [10]. In a previous work [11], we studied the electronic properties of graphene nanoribbons with bisecting 4-8 ring defect lines. We found that, akin to defect-free armchair-edged nanoribbons [12,13] and carbon nanotubes, the defective ribbons can be classified into two families of semiconducting nanoribbons with bandgaps increasing due to quantum confinement for decreasing ribbon width, and a third family of nanoribbons featuring linear crossing bands at the Fermi energy. ...

Family behavior and Dirac bands in armchair nanoribbons with 4–8 defect lines

... The different ribbon widths or edge topologies for the specific group also have an influence on the electronic properties, for example, the AGNRs are predicted to be semiconductors while the band gap is controlled by adjusting the ribbon width [28][29][30][31][32]. In recent years, a series of CGNRs with different cove topologies have been synthesized in experiments [33,34]. It is revealed that the band gaps of CGNRs can be tailored by controlling the size of coves or the distance of the adjacent cove units [22,35]. ...

Cove-Edged Chiral Graphene Nanoribbons with Chirality-Dependent Bandgap and Carrier Mobility
  • Citing Article
  • December 2023

Journal of the American Chemical Society

... An alternative method for size separation using commonly available benchtop centrifuges is liquid cascade centrifugation (LCC). 16 This widely used method [17][18][19][20][21][22][23][24] has been discussed in recent reviews for graphene and metal chalcogenides, and reported for a range of materials including MXenes. 17,25 By sequentially increasing the centrifuge speed (or time) over multiple centrifugation steps and extracting the sediment from each step, a series of samples are obtained. ...

Sonication-assisted liquid phase exfoliation of two-dimensional CrTe3 under inert conditions

Ultrasonics Sonochemistry

... Third, by analyzing the charge fluctuations we can, as we will show below, address all four rates in one go. To achieve this, the analysis of waiting times [28][29][30][31] as well as the full counting statistics [32][33][34][35][36] in terms of so-called factorial cumulants [37][38][39][40][41][42][43][44][45][46][47][48][49][50] will be used. This paper is organized as follows. ...

Independent and coherent transitions between antiferromagnetic states of few-molecule systems

... The proposed importance of the local minimum at Q for polaron formation in bulk MoS 2 poses the question whether polarons can actually form in single-layer (SL) MoS 2 , since the energy difference between K and Q is much larger in n-doped SLs [17][18][19]. In addition, a recent theoretical study postulated that stable polarons do not form in SL MoS 2 [20], although only monopolarons were considered in these calculations. ...

Robustness of Momentum-Indirect Interlayer Excitons in MoS 2 /WSe 2 Heterostructure against Charge Carrier Doping
  • Citing Article
  • March 2023

ACS Photonics

... [30] Heterojunction engineering to suppress recombination in Sb 2 S 3 absorber has been extensively explored, with a variety of oxide semiconductors used in combination with Sb 2 S 3 , such as TiO 2 , [31] Al 2 O 3 , [32,33] ZnO, [34,35] and SnO 2 . [36,37] SnO 2 is an ideal inorganic electron extraction material, extensively investigated in the field of halide perovskites solar cells (PSC) as electron-transport layer (ETL). [38][39][40] The choice of SnO 2 (bandgap of 3.6-4.1 eV) as ETL in a photoanode configuration for OER, owns numerous advantages: 1) favorable conduction band states that accelerates the electron extraction; 2) high electron mobility for enhancing the electron extraction and transport, decreasing the carrier recombination; and 3) low crystallization temperature with reduced fabrication cost; 4) excellent lattice match with conductive glasses as FTO or ITO. ...

Continuous, crystalline Sb2S3 ultrathin light absorber coatings in solar cells based on photonic concentric p-i-n heterojunctions

Nano Energy

... Additionally, it may offer further benefits for long-term environmental stability, as the lower intercalation competitiveness of H 2 O and O 2 is expected to intrinsically limit their ability to interfere with the intercalation process or passivate vacancy sites. Such behavior would be generally consistent with the well-documented hydrophobic nature of MoS 2 basal planes [110]. ...

Oxidation and phase transition in covalently functionalized MoS2