You are on page 1of 29

pubs.acs.

org/JPCC Article

Large Exciton-Driven Linear and Nonlinear Optical Processes in


Monolayers of Nitrogen Arsenide and Nitrogen Antimonide
Miroslav Kolos, Rekha Verma, František Karlický,* and Sitangshu Bhattacharya*
Cite This: J. Phys. Chem. C 2022, 126, 14931−14959 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: We demonstrate that atomically thin nitrogen-based binary


group V wide band gap indirect semiconductors (β-NX (X = As, Sb))
embody 3-fold valley degenerated band structures and strong linear and
nonlinear optical activities. Contrary to NAs, the fundamental optical
Downloaded via 223.189.238.161 on September 26, 2023 at 10:47:07 (UTC).

absorption in NSb is quite resilient toward the temperature variations


between 0 and 450 K. In the absence of lattice vibrations, exciton in NSb
is, however, less strongly bound (1.30 eV) compared to the tighter case in
NAs (1.62 eV), leading to a more delocalized Mott−Wannier-type texture.
The inhomogeneous excitonic line widths for both monolayers within
these temperatures are within the 100−400 meV range. The most
significant nonlinear second harmonic optical coefficients (2ω reso-
nances) in NAs and NSb monolayers are obtained as ∼636 and 270 pm/V
at 3.2 eV (387 nm) and 2.6 eV (477 nm), respectively. We also present a
detailed analysis of in-plane biaxial strain on these structures and found
that tensile strain dynamically stabilizes structures with no loss in absorbance spectra (and does not influence exciton binding energy
in NAs). The nonlinear coefficient also significantly improves at the two most important 810 and 1560 nm wavelengths compared to
the cases offered by the monolayer transitional metal dichalcogenides. Our analyses originate from a fully ab initio G0W0+Bethe-
Salpeter excited-state theory. The temperature-dependent linear absorption spectra are evaluated by including the electron−phonon
self-energies, whereas the nonlinear spectra are treated using the modern theory of polarization within the same perturbative
approach. Our reference findings provide important advanced characteristics for applications of two-dimensional β-NX (X = As, Sb)
materials and should motivate further theoretical and experimental investigations of these interesting materials.

■ INTRODUCTION
Excitons are the essential ingredients responsible for optical
eigenvalues that the standard DFT rigorously struggles for.
Nowadays, these methods have become a standard routine task
excitation in materials. Fundamentally pictured as a quasi- to understand the excitonic-driven optical properties in bulk to
particle in the framework of many-body interactions, an atomically thin layers,5−12 including comparison with the
exciton is an electron−hole pair bound together by a quantum Monte Carlo calculations.13
Coulombic force of attraction. The pair can be generated When the intensity of the light becomes very high, the
when the electron is kicked out by, for example, photons from semiconductor optical properties tune to the nonlinear regime.
the ground-state orbital. How long this pair will hold together One can then observe frequency overtones like second, third,
depends on its binding energy. The exciton so formed and other higher harmonics.14 For example, suppose the crystal
resembles an atomic hydrogen, and likewise, distinct optical structure is such that both the inversion and time-reversal
spectra are therefore expected. Similar to the hydrogen symmetries are broken. In that case, one can then show that
electronic energy quantization, the exciton Coulombic energy the induced time-varying electric dipoles do not cancel out on
lowers the width of the direct band gap of the material. In an average, instead of leading to a minimum detected second
principle, the one-electron picture obtained from the standard harmonic signal as a nonlinear one.15 In order to capture these
generalized gradient approximation (GGA) density functional nonlinear dynamics at an atomic scale in a purely ab initio way,
theory (DFT) does not account for such interactions. Instead,
excitons are a two-body dynamically correlated system, and
thus, to understand their energy and lifetime, many-body Received: May 30, 2022
methods must be used, such as the Bethe−Salpeter equation Revised: August 5, 2022
(BSE) over the single-shot GW corrections.1−4 The solution of Published: August 25, 2022
the former can directly access the information about the
optical absorption spectrum and the bound-state electron−
hole energies, while the latter fixes the missing energy
© 2022 The Authors. Published by
American Chemical Society https://doi.org/10.1021/acs.jpcc.2c03708
14931 J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

a time-dependent DFT (TD-DFT) is therefore needed. This and defines the momentum forbidden dark and bright states.
TD-DFT formalism was initially developed by Runge and For example, the monolayer transitional metal dichalcogenide
Gross,16 who numerically integrated the time-dependent (TMDC) WSe2 possesses a time-reversal symmetry and thus
Kohn−Sham equation directly in the real-time domain. The has a large spin−orbit splitting in its valence band at K point.
accuracy of such calculations, however, depends on the level of The bright (dark) exciton is thus formed when an intravalley
approximation used in the exchange-correlation kernel of the (intervalley) electron pairs with a hole having a parallel
time-dependent Hamiltonian.3,17 The kernel hierarchy starts (antiparallel) spin.8,30 In recent years, there have been many
from an independent particle approximation (only the Kohn− benchmark strategic experiments to push the binding energies
Sham eigenvalues) to a time-dependent Hartree (arising due to and nonlinear coefficients to their limits (for example, see
local field effects resulting in system inhomogeneities) to an Table 4 below). The excitons in two-dimensional semi-
adiabatic local density approximation (quasi-statically space conductors possess rather larger binding energies and oscillator
and time-varying density) and to the most advanced time- strengths due to an in-plane quantum confinement than their
dependent screen exchange (electron−hole correlation is bulk structures. In the latter case, the electric field is properly
screen exchanged). This last kernel particularly contains the screened out by the surrounding charges, while in monolayers,
excitonic dynamics and is also known as the time-dependent it is not. Therefore, an intense Coulombic interaction develops,
BSE (TD-BSE) or the real-time BSE.18 The modern theory of leading to a strongly bounded pair with enhanced optical
polarization developed by Kingsmith and Vanderbilt 19 absorption. There is, however, another justification as to why
prescribes a way to obtain the coupling between the time- such an atomic layer should absorb a large amount of light.
varying electric field and the Bloch electrons in an extended Castro Neto et al.31 showed that the strong peaks in the optical
system that is required in the time-dependent Hamiltonian. conductivity could also be justified from the ground-state
The Berry’s phase change around the complete Brillouin zone electronic band structure. They observed that the peaks
(BZ) leads to the macroscopic time-varying polarization, and correspond to those points in the BZ where the transition
therefore, the TD-DFT equation of motion with such various bands run parallel to each other. This is known as band-
kernel hierarchy can be solved. The nonlinear coefficients can nesting. It was recently demonstrated that this band-nesting
then be obtained by imposing a suitable cutoff in the could also justify the nonlinear second harmonic generation
polarization Fourier series.20 (SHG) signals in uniaxially tensile strained TMDCs.32
There is, however, a fundamental part missing which needs Nevertheless, shifting the focus from the traditional planar
to be addressed. The BSE so developed was accounted only for monolayers, Taheri et al.33 demonstrated that group V
the frozen atom condition; i.e., atoms were assumed to be fixed monolayers based on nitrogen (NX, where X = P, As, and
in space or, in other words, in the absence of lattice vibrations. Sb) in their β-phase exhibit thermodynamically stable buckled
As a result, the pole of the dielectric function (i.e., it is structures. Using a purely ab initio technique (along with a
imaginary part) is required to be manually tuned so that the temperature dependent BSE calculation), it was recently
broadening in the absorption spectrum matches with that of shown27 that this β-phase nitrogen phosphide (NP) monolayer
the experimental spectrum. The role of this broadening carries can exhibit a very high excitonic binding energy and strong
significant importance since it directly confirms Heisenberg’s absorbance spectra in the visible region. The nonlinear
uncertainty principle. The absence of the broadening would coefficients SHG along with the third harmonic generation
simply mean that the exciton lives forever in that state, (THG) were also found to be very high compared to the
characterized by a sharp delta-like spectrum. One should note traditional TMDs and monolayer transitional metal mono-
that, in principle, this thermal broadening should enter as an chalcogenides (TMMCs). In this work, we extend the
exciton−phonon coupling matrix in the BSE. However, the roadmap of exciton-driven optical activities in monolayers of
current theoretical strategy addresses this formidable challenge β-phase nitrogen arsenide (NAs) and nitrogen antimony
by adopting its entry as an electron−phonon coupling matrix (NSb). We ask the following: (1) How an electron taking part
into the BSE Hamiltonian.21 This first-cut approximation in the formation of an exciton relaxes from the conduction
approach quintessentially removes the manual tuning of the band in the presence of lattice vibrations, (2) how temperature
broadening parameter but is also quite successful in justifying controls the excitonic nonradiative line widths via exciton−
the temperature-dependent absorption spectra, the underlying phonon coupling, (3) can band-nesting in these cases justify
excitonic binding energies, and lifetimes in a variety of the strong absorption, (4) are the excitons of the Mott−
dimensional systems like monolayer MoS2,22 WSe2,23 h-BN Wannier or Frenkel type, (5) what are the SHG coefficients,
(both bulk24,25 and monolayer26), monolayer β-phase NP,27 and (6) can an in-plane biaxial strain improve the binding
bulk zinc-blende GaN,28 carbon-nanotube,29 etc., demonstrat- energies and SHG coefficients? We respond to these questions
ing a remarkable accuracy with experimental observations. The in a purely ab initio way. We use the many-body perturbation
temperature-dependent analyses of the optical properties are theory at the level of BSE to justify the excitonic-driven linear
therefore seems more plausible on one hand; however, on the spectra. All temperature corrections are computed using
other hand, it makes the entire computation an expensive affair density functional perturbation theory (DFPT) on the top of
with rigorous cascaded process-flow. Due to these trade-offs, the ground-state DFT eigenvalues. The nonlinear SHG
such calculations are rarely reported in the literature. coefficient is computed by solving the TD-BSE in the real-
Two quantities are greatly required in order to observe a time domain. Berry’s phase polarization is used to compute the
strong exciton-driven optical properties in semiconductors: a couplings between Bloch electrons and the time-dependent
weak dielectric screening and crystalline noncentrosymmetry electric field. In a hope to improve the optical properties, we
(i.e., lack of inversion symmetry). Whether or not spin−orbit extend our analyses in the presence of an in-plane biaxial strain
coupling (SoC) will originate depends on the availability of an also, and we report the optimized thermodynamic stability,
additional crystalline time-reversal symmetry. The latter leads energy gaps, binding energies, absorption and SHG spectra for
to the coupling between spin and momentum in the valleys both compressive and tensile strains. All strain-dependent
14932 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Table 1. Electronic Ab Initio Energy Comparisons in β-NX (X = P, As, Sb)a,b


Material (ML) a (Å) z (Å) Indirect gap (eV) Direct gap (eV) Δ0 (meV) ref
β-NP 2.73 0.87 1.79 (PBE), 4.15 (G0W0) 4.95 (G0W0, M) Kolos et al.27
2.78 1.48 (PBE), 2.64 (HSE06) Yu et al.43
2.73 0.86 Taheri et al.33
β-NAs 2.79 1.00 2.36 (PBE, K Γ ↔ M) 3.06 (PBE, M Γ) 44 (PBE, Γ) this work
4.51 (G0W0, K Γ ↔ M) 5.00 (G0W0, M Γ) this work
3.03 1.70 (PBE), 2.76 (HSE06) Yu et al.43
2.98 0.97 Taheri et al.33
β-NSb 3.00 1.13 2.06 (PBE, K Γ ↔ M) 2.52 (PBE, K) 57 (PBE, Γ) this work
3.70 (G0W0, Γ ↔ K) 4.80 (G0W0, K) this work
3.58 (PBE, Γ) this work
4.45 (G0W0, Γ) this work
2.88 (PBE, M Γ) this work
4.77 (G0W0, M Γ) this work
3.33 1.35 (PBE), 2.26 (HSE06) Yu et al.43
3.28 1.02 Taheri et al.33
a
Blank spaces denotes values not reported. ba and z are the in-plane lattice parameter and buckling heights, respectively. Δ0 is the spin-orbit
splitting energy. K Γ and M Γ symbols mean specific k-points of gap location on the K-Γ and M-Γ paths, respectively.

calculations follow the same extensive process flow as Table 3. DFT and Zero-Point Gap (Using the Dynamical
discussed above. Our results on the electronic and optical HAC Theory) in Various Materials Including This Work
properties of β-phase NAs and NSb are summarized in Tables (ZB = zinc-blende)
1, 2, and 6. A sizable comparison with other monolayers and
Materials DFT (eV) Zero-point gap (meV)

Table 2. Ab Initio Lattice Vibration and Formation Energy Bulk BP 0.53 (HSE06)62 20 (PBE)63
Comparisons in β-NX (X = P, As, Sb)a ML WSe2 1.26 (PBE)23 31 (PBE)23
trans-polyacetylene 0.6345 (LDA)60 40 (LDA)60
ωLO − Formation ML MoS2 1.69 (LDA)64 75 (LDA)22
Material ωTO|Γ ωZO|Γ energy (eV/ dEgidT
(ML) (cm−1) (cm−1) atom) (meVK−1) ref ML β-NSba 2.06 (PBE) 80 (PBE)
ML β-NAsa 2.36 (PBE) 105 (PBE)
β-NP 566 672 6.67 Kolos et
al.27 Si 0.46 (LDA)65 123 (LDA)66
480 570 6.23 Yu et ML β-NP 1.79 (PBE)27 200 (PBE)27
al.43 SiC 2.54 (LDA)67 223 (LDA)66
500, 534 600 Taheri et Bulk-ZB BN 6.4 ± 0.5c,68 366 (LDA)36
al.33 Polyethylene 5.78 (LDA)60 280 (LDA)60
β-NAs 566 672 10.7 −0.18 this work Diamond 5.5 (LDA)69 622 (LDA)59
480 570 5.58 Yu et Bulk-ZB LiF 10.9b,70 718 (LDA)36
al.43
a
500, 534 600 Taheri et This work. bWith Hartree−Fock corrections. cExperimental value.
al.33
β-NSb 523 596 9.3 −0.47 this work
460 470 5.31 Yu et down the Coulombic interactions due to periodic images along
al.43 the out-of-plane direction, a vacuum-slab-vacuum structure
480, 500 517 Taheri et was created with a vacuum separation extending to 20 Å on
al.33 either side. Figure 1 below demonstrates the schematic view of
a
ωLO and ωTO are the zone-centre longitudinal and transverse optical such a monolayer. The optimized in-plane and out-of-plane
in-plane frequencies, while ωZO is the optical out-of plane frequency. lattice constants were found to be (2.79, 1.00)Å for NAs and
dEig (3.00, 1.13)Å for NSb, respectively. These optimizations were
shows the temperature dependent indirect band-gap.
dT
done by using the open-source Quantum Espresso DFT
package.34 Fully relativistic norm-conserving pseudopotentials
bulks are provided in Tables 3, 4, and 5. In addition, we also were first generated35 which also included nonlinear core-
put the convergence and supporting figures in the Supporting corrections in N (core states: [He], valence states 2s and 2p),
Information. What follows, in the Methodologies section, we As (core states [Ar], valence states 3d, 4s and 4p), and Sb
provide a lucid picture of our computational details for the (core states [Kr], valence states 4d, 5s and 5p). A Perdew−
ground-state, excited-state, and nonlinear analysis. In the Burke−Ernzerhof (PBE) exchange-correlation functional was
Results and Discussion, we justify our results together with used rather than a local density approximation (LDA) since
summarizing our outcomes.


the latter is known to underestimate the electron−phonon
couplings by 30%.36 A kinetic cutoff energy of 80 Ry for all the
METHODOLOGIES atomic species was found to be sufficient (see Figure S1 in the
Ground-State Calculations. β-Phase NAs and NSb Supporting Information) to achieve energy convergences. A
monolayers are thermodynamically stable in their non-co- plane-wave basis set was then used, and the cell was finally
planar structures. A unit cell composed of two atoms was thus allowed to undergo energy minimization at Γ-centered 12 × 12
made, leading to a C3v 3m point group 3-fold symmetry. To cut × 1 Monkhorst−Pack grid constraining all forces and energy
14933 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Table 4. Comparison of Experimental Quasi-Particle Direct Band-Gap, Fundamental Exciton Binding Energy (BE), Line
Width (LW), and Nonlinear SHG Coefficient (at Excitation Wavelength in nm) in Single and Few Atomic Layers, including
This Work (rt = room temperature 300 K)a
Material Substrate rt Band-gap (eV) rt BE (eV) rt LW (meV) |χ(2)|(λnm) (pm/V)
72 72 72
BP (4L) PDMS 0.84 0.139 20
BP (29L) SiO2/Si
ML WSe2 SiO2/Si(e) 2.0273 0.3773 5074 16.5 (1560)75
Sapphire 1.89b,76 0.240b,76
Au 1.75b,76 0.140b,76
SiO2/Si(e) 2.63(4−300 K)77 0.887(4−300 K)77
hBN encap (comm)78 3.9−4.2(4 K)
32−34(292 K)
Quartz(e) ∼13 (160 K)79
Si/SiO2/(e) 1 × 104 (816)80
ML MoS2 Fused silica(e) 2.17 ± 0.181
0.31 ± 0.04 81
160 (810)82
SiO2/Si(e) 1 × 105 (810)83
Sapphireb 2.1176 0.24076
Si/SiO2(c) 684
Si/SiO2(c) 5 × 103 (816)83
amorphous quartz(e) 30−100 (680−1080)85
Au 1.9b,76 0.090b,76
hBN/Fused silica(e) 2.47 ± 0.0886 0.44 ± 0.08f,86
isolatedc 4422
HOPG(c) 2.15 ± 0.06 (77 K) 87
0.2 (77 K) 87

HOPG(c) 2.15 ± 0.1 (77 K)88 ∼0.3 (77 K)88


hBN encap (comm)78 2.0−4.5 (4 K)
44−46 (292 K)
hBN encap (CPT)78 3.9−5.0 (4 K)
44−49 (292 K)
ML MoSe2 Si/SiO2(e) 0.1 (5 K)89 4090 37(1560)75
91
6H-SiC(0001)(m) 1.58 (40 K)
hBN encap (VPT)78 2.4−4.9 (4 K)
34−36 (292 K)
hBN encap(m) 6.6 (4 K)92
Si/SiO2(c) 50 (1200−1800)93
ML MoTe2 Si/SiO2(e) 0.540 (10 K)94
ML WS2 suspended and Si/SiO2(c)b 9 × 103(832)95
SiO2/Si(e) 2.14 ± 0.04 (5 K) 0.32 ± 0.04 (5 K)
Si/SiO2(e) 16.2 (1560)75
SiO2(c) 2.31−2.5396 0.26−0.4896
SiO2/Si(e) 2.7397 0.71 ± 0.0197
SiO2/Si(e) 3.01 (4−300 K)77 0.929 (4−300 K)77
Fused silica(e) 2.33 ± 0.0586 0.32 ± 0.0586
Fused quartz(e) 2.38 ± 0.0681 0.360 ± 0.06081
Si/SiO2(e) 2490
hBN encap (comm)78 4.3−4.8 (4 K)
23−25 (292 K)
Sapphire(c) 5.96 (10 K)98
ML InSe hBN(e) 2.9 (100 K)99
hBN encap(e) 223 ± 138 (810)100
at 300 K
ML GaSe Si/SiO2(c) 2.4 × 103101
ML hBN Fused silica(e) 10 (810)82
Graphite(m) 6.1102 1.9102 32 (10 K)−38 (300 K)102
isolatedc26 7.36 1.81 97 (0 K)−260 (300 K)
hBN (110 ± 2L) SiO2(e) 42 (810)103
ML β-NPc27 isolated 4.92 2e
151 800 (364)
ML β-NAscd isolated 5.21 (M Γ) 1.63e 278 636 (387)
52 (810)
14 (1560)
ML β-NSbcd isolated 4.45 (Γ) 258 270 (477)
4.92 (M Γ) 1.74e 31 (810)
21 (1560)

14934 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Table 4. continued

a
Symbols: c = chemical vapor deposition, e = mechanical exfoliation, m = molecular beam epitaxy, L = layer, PDMS = polydimethylsiloxane, rt =
room temperature (300 K). b= temperature not mentioned. c= ab initio. d=this work, comm = commercial, encap = encapsulation, CV(P)T =
chemical vapor(pressure) transport and blank space = not reported. Data at other than 300 K are explicitly mentioned. eThe rt-binding energies are
obtained after adding the respective G0W0 corrections to the temperature dependent PBE direct gaps. fObtained by analyzing the B-excitonic
transition state.

Table 5. Experimental Excitonic Line Width Parameters in Monolayer NAs and NSb together with Conventional TMDCsa
Material (ML) Substrate Γ0(0K) (meV) γAC (μeV/K) γOP (meV) ωOP (meV) temp range (K)
90
WS2 SiO2/Si(e) 9.2 28 6.5 20 0−300
MoSe2 SiO2/Si(e)90 4.3 91 15.6 30 0−300
WSe2 Sapphire(c)106 1.6 ± 0.3 60 5−50
Quartz(e)79 7.3 4 39 31 40
MoS2 hBN/ML MoS2/hBN78 4 ± 0.2 70 ± 5 42.6 ± 1.2 24.2 4−300
quartz/suspended107 4.5 45 5−40
β-NP†27 isolated 151 33 90 74.4 0−1000
β-NAs†§ isolated 105 0.4 36 71.9 0−450
β-NSb†§ isolated 129 0.18 52 76.9 0−450
a
The notations are c = chemical vapor deposition, e = mechanical exfoliation, † = ab initio, §=this work and blank space = not mentioned.

Table 6. Optical Phonon Mode Comparisons (in cm−1) at Γ in β-NX (X = As, Sb) at Various Strain Levelsa
Strain Material Modes 2% 4% 6% 8% 10%
NAs E 519.4 476.6 436.8 399.5 365.5
Tensile A1 640.4 612.1 585.9 561.9 539.9
NSb E 477.6 435.4 396.5 360.3 326.5
A1 563.3 533.9 507.2 482.5 459.9
NAs E 617.2 672.9 733.1 799.3 871.8
Compressive A1 706.9 745.0 786.6 832.9 884.1
NSb E 573.4 627.4 685.9 749.8 815.9
A1 632.0 671.6 715.1 763.4 818.6
a
Both E and A1 modes are infra-red and Raman active in all cases.

Using the same energy threshold, the dynamical matrices and


the perturbed potentials on these grids were then computed by
mapping the self-consistent charge densities on the regular
grids. A non self-consistent calculation was then carried out on
these random grids, which led to the electron−phonon
corrected electronic states along the BZ. Using this last step,
the electron−phonon matrix elements were finally evaluated
after constructing the initial states.
Linear Spectra G0W0+BSE Calculations. The excited-
state corrections were computed using the MBPT open-source
code package Yambo.37 The G0W0 corrections were applied to
the five electronic bands on either side of the valence band
maximum, which we also find to be most crucial in capturing
Figure 1. Optimized in-plane and out-of-plane crystal structure of β- the optical transitions. A total of 200 bands (lowest 20
phase NX (X = As, N). Both these structures possess a trigonal occupied bands and highest 180 unoccupied bands) were used
symmetry and belong to space group 156 and point group 3m. to sum up the irreducible polarization response function. Local
field effects were introduced in this linear response sum by
cut-offs below 10−5 Ry/Bohr and 10−5 Ry, respectively. The switching on the random-phase approximation (RPA) kernel.
self-consistent charge densities and states were computed using An energy cutoff of about 10 Ry was found sufficient to
a two-spinor wave function and switching on the noncollinear converge this sum describing the system inhomogenity. The
spin−orbit interactions. microscopic dielectric function is then constructed by
Electron−Phonon Coupling Calculations. The lattice convoluting the polarization function with the bare Coulomb
vibration computations were calculated using the extended potential. The dynamic screening is then computed by
PHonon package within the same DFT code. For both the convoluting the inverse of the microscopic function again
structures, a regular dense phonon q-grid 18 × 18 × 1 was with the bare Coulomb potential. A final convolution between
chosen at a rigid self-consistent energy threshold 10−16 Ry and this Coulombic screening and the noninteracting Green’s
single iteration mixing factor of 0.7 Ry. In order to compute function (G0) leads to the full frequency G0W0 self-energy. We
the electron−phonon couplings, the entire irreducible BZ was note here that most open-source code splits this self-energy
first randomly sampled to a fine 200 phonon momenta q-grids. into a pure exchange term and a correlational term for
14935 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 2. Electronic valence band dispersion of NAs (a, b) and NSb (c, d) monolayers showing the effect of spin−orbit splitting. In case of NAs,
the energy differences in the absence of SoC results in Γ3v − Γ1v = 61 meV. In the presence of SoC, the spin−orbit splitting is Δ0 = Γ5v − Γ4v = 44
meV. The strength of SoC shifts Γ1v to Γcf4v resulting in Γ1v − Γcf4v = −55 meV, Γ4v − Γcf4v = 70 meV. The spin-splitting along the M−Γ direction in
the top valence band is Δ5,6 = 33 meV. In case of NSb the respective energy differences are Γ3v − Γ1v = 318 meV, Δ0 = 57 meV, Γ1v − Γcf4v = −104
meV, Γ4v − Γcf4v = 298 meV, and Δ5,6 = 74 meV.

computational convenience. The former is static, while the be more meaningful to capture optical transitions properly
latter is dynamic in frequency and is thus a complex quantity. along the high symmetry routes. For this purpose, we fine
The inverse microscopic dielectric function inside the dynamic sample the entire BZ into 48 × 48 × 1 on a shifted grid. First,
screening term demands some comments. This inverse we obtain the kernels for the independent-particle cases on
function is plagued with numerous poles near the real axis, these grids. Next, we obtain the interacting BSE kernels on the
making the screening convolution integral very expensive to unshifted 12 × 12 × 1-course grids. The BSE kernels on the
compute. To implement this practically, a plasmon-pole model fine grids are now interpolated using a Wannier interpolator
is often used, which mimics the inverse dielectric function by which maps the fine and the course grids. This methodology39
approximating it at the most prominent pole, which is the makes the expensive fine-grid BSE computation relatively
plasmonic frequency. The plasmon-pole model by Godby and quick without losing accuracy. A manual broadening of 0.1 eV
Needs38 is used in the package, which uses one pole at zero is applied to construct a Lorentzian shape spectra together with
frequency while the other at the plasmon frequency. The latter an in-plane perturbing electric field. The quasi-particle energy
is to be chosen such that the dielectric function converges. corrections and the static screening are then added from the
This correlational self-energy is then updated for each energy previous G0W0 calculation. In the presence of lattice vibrations,
starting from the noninteracting Kohn−Sham eigenvalue and instead, the corrections corresponding to the electron−phonon
put back again into the nonlinear quasi-particle equation. interactions on the energy bands are implemented. The G0W0
However, such an iteration would create redundancy in gaps are now opened up using a scissor operator that provides
computing the self-energy. Thus, the nonlinear equation is a rigid shift to the bands. No external broadening is required as
linearized using Taylor’s series up to the first order under the the exciton line widths are now computed using the electron−
assumption that the renormalized energies are not far from the phonon matrix elements. We go beyond the standard Tamm−
Kohn−Sham mean-fields. If the noninteracting Green’s Dancoff approximation40 by including both the resonant and
function and the dynamic screening are also updated at each antiresonant electron−hole matrix elements in the BSE
iteration process, then the computation scheme becomes GW. Hamiltonian. The BSE matrix is then diagonalized and solved
If the dynamic screening is updated only once, then it is a to obtain the poles corresponding to the transition energies.
G0W0 process or a single-shot GW scheme. Coulombic Nonlinear SHG Calculations. The presence of a finite
divergences occurring in all self-energies are fixed by external electric field breaks down the crystalline symmetry.
integrating them over the irreducible BZ space by assuming We chose an in-plane electric field and removed the symmetry
first that the density matrices are the smooth function of of the crystal (the initial 12 × 12 × 1 course-grid). We then
momenta. The diverging quantities are then numerically map this to a dense symmetry removed 48 × 48 × 1 shifted
integrated using a Monte Carlo technique that creates grid in the presence of the same electric field. The time-
numerous randomly small BZs about each momenta vector. dependent BSE was then solved at these fine grids. The linear
We, therefore, used 107 random points distributed all over the coefficient was extracted by applying a delta-like pulse. A
irreducible BZ with a cutoff of about 3 Ry to converge the monochromatic sinusoidal wave was used to obtain the higher-
integrals. In addition to these, a Coulomb truncation (similar order nonlinear coefficients. The time-dependent polarization
to the DFT case) is applied to cut down the periodic was obtained from the geometric Berry’s curvature. This TD-
interactions between the repeated monolayer images. BSE was finally solved using the Crank−Nicholson algorithm41
The absorption spectra are obtained by solving the time- at a time step of 0.01 fs. The sudden switching-on of the
independent BSE. The same cut-offs are used to build up the electric field results in spurious noise in the polarization signal
exchange electron−hole attractive and repulsive kernels in the during the initial time. We, therefore, add a dephasing time of
BSE matrix. An important note is that electronic transitions are about 8 fs, which is approximate to a spectral broadening of
very sensitive to BZ samplings. A dense sampling would then about 0.17 eV. We then run this simulation for up to about 7-
14936 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 3. Bare electronic dispersion of NAs monolayer showing the partial density of states projected on the former. Plots (a) nitrogen 3PJ=(1/2), Ms
= −(1/2) and (b) arsenic 4PJ=(3/2), Ms = (1/2) show the spin−orbital occupancy in valence and conduction bands, respectively. The arrows in (b)
demonstrate how the electron that takes part in optical transition gets scattered due to the thermal broadening of the energy bands.

Figure 4. Bare electronic dispersion of NSb monolayer showing the partial density of states projected on the former. Plots (a) nitrogen 3PJ=(1/2), Ms
= −(1/2) and (b) antimony 4PJ=(3/2), Ms = (1/2) show the spin−orbital occupancy in valence and conduction bands, respectively. The arrows in
(b) demonstrate how the electron that takes part in optical transition gets scattered due to the thermal broadening of the energy bands. Note that
the optical transitions occurs between Γ and M instead at K. This happens because of the modification of the gaps when the GW corrections are
applied.

time constants ∼55−60 fs until a clean signal was obtained. energy bands. The splittings observed in these two materials
Such a dephasing also mimics the inadvertent addition of the are nonmagnetic Zeeman-type (shift in the energy) across the
experimental broadening due to thermal effect, defects, etc. BZ, except in the region around the high symmetry M point,
Once this clean polarization signal is obtained, the nonlinear where it can be observed as crossing of the bands along the
coefficients can be extracted by truncating the Fourier series. momentum k-space, resulting in a Rashba-type splitting.42 The

■ RESULTS AND DISCUSSION


Ground-State DFT Dispersion. Figures 2−4 demonstrate
valence and conduction splittings are even small in NAs
compared to NSb, signifying a smaller SoC. In general, in the
presence of a crystal field only, the C3v point group symmetry
the ground-state electronic dispersion of β-NAs and β-NSb. As splits up in energy at the Γ point of the BZ into a 2-fold high in
both of these materials contain heavier atoms, relativistic spin− energy degenerate level (Γ3v) and a lower in energy
orbit calculations were performed, resulting in splitting of the nondegenerate level (Γ1v). When the SoC is switched on,
14937 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 5. Lattice vibration spectra, phonon density of states, and electron−phonon Eliashberg function of (a)−(c) NAs and (d)−(f) NSb
monolayers, respectively. The individual Eliashberg functions at the ground-state conduction maxima and valence minima for both the NAs and
NSb are calculated, justifying why the indirect gap should decrease with increasing temperature.

this double group level Γ3v further splits into three energy the presence of SoC) is obtained as 104 meV. In the case of
levels Γ4v, Γ5v, and Γ6v. Group theoretic calculations tell us that NAs, the corresponding crystal-field and Δ0 splittings are 61
Γ5v and Γ6v are degenerate and higher in energy levels, while and 44 meV, respectively. The shift of Γ1v to Γcf4v is an energy
Γ4v is nondegenerate and lower in energy level. Consequently, difference of 55 meV. Therefore, we see that because of small
Γ5v and Γ6v are called heavy-hole and act as a Kramer’s doublet. spin−orbit energy splittings in NAs the use of spin−orbit
The lower state Γ4v is a light-hole. The additional presence of coupling is not essential.
SoC shifts the state Γ1v to a lower level Γcf4v and is called the Figures 3 and 4 demonstrate the full dispersion with orbital
Crystal-field split-off hole. occupancy. Interestingly, both of them exhibit valley
This is indicated in Figure 2, which demonstrates the degenerated indirect band-gaps. The location of the valence
valence dispersion in both the monolayers, both in the absence band maxima (VBM) for these materials are along the path
and presence of an SoC. We find that all the aforementioned K Γ of the BZ, while the valley degenerated VBM are along
splittings are actually seen in the case of NSb. The crystal-field Γ M. Similarly, the conduction band minima (CBM) are at
splitting (in the absence of SoC) and the spin−orbit energy the M point of the BZ, whereas the valley degenerated CBMs
splittings (Δ0) are found to be 318 and 57 meV, respectively. are at the K point. The location of direct gap for NAs is along
The shift of energy from Γ1v (in the absence of SoC) to Γcf4v (in M Γ, while in NSb it is at the K point. The local density-of-
14938 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 6. Ground-state (blue) and G0W0 (green) electronic band structure in (a) NAs and (b) NSb monolayers. The horizontal dashed lines are
with respect to the DFT top and G0W0 bottom of the valence and conduction bands, respectively. All the vertical arrows represent direct quasi-
particle gaps with most of the electronic transition weights.

states (DOS) projected on the respective electronic dispersion that these energies for β-NAs and β-NSb could be as large as
are also demonstrated in the same figures. We observed that −10.7 and −9.4 eV/atom, respectively, when compared to the
the valence bands in both the materials are mainly occupied by isolated atoms. One crucial observation is that there are small
the P orbital of nitrogen (3PJ=(1/2), Ms=−(1/2)), whereas the out-of-plane acoustic (ZA) soft-modes located along M Γ and
conduction bands are filled with the P orbital of the second K Γ directions. In fact, the presence of these soft modes
sublattice atom (4PJ=(3/2), Ms=(1/2)), i.e., As and Sb atoms, simply indicates that the exfoliation mechanism may not lead
respectively. To get more insight, we have shown the to a standalone monolayer system. Other preparations such as
contribution of all the local DOS for all atoms of both the direct-grow on a suitable substrate may stabilize these
materials in the Supporting Information in Figures S2 and S3. monolayers (see also stabilization of β-NAs and β-NSb by
Comparing Figures S2 and S3 with Figure 2, we see that in the strain in the appropriate section below). A similar situation
both materials there are the partial occupancy at points Γcf4v, is also observed during the stabilization of monolayer β-NP,
Γ4v, Γ5v, and Γ6v coming from 2PJ=(3/2), Ms=(1/2), and which falls under the same family.27 Also, buckled hexagonal
3PJ=(1/2), Ms = −(1/2) of the nitrogen atom. structures of silicene and germanene, which show negative
Electronic and Thermal Stability. We now discuss the phonon frequencies,45 were experimentally prepared on
electrothermal stability of these isolated monolayers. The substrate of Ag46 and Al.47 We finally note that several
lattice dispersions for both materials are shown in Figure 5a,d. buckled V−V binary materials were experimentally prepared
We note here that the C3v 3m group materials are polar. In recently.48 Nevertheless, in Tables 1 and 2, we have
their bulk structures, the coupling between the optical summarized the electronic and thermal gaps for both of
longitudinal (LO) modes and the polarization charge density these materials. The discussions about the phonon DOS and
leads to a finite splitting between them at the Γ point; the electron−phonon Eliashberg function are outlined in the
however, in the case of 2D systems, this splitting disappears, zero-point energy gap section below.
and the two modes become degenerated. Sohier et al.44 Excited-State G0W0+BSE Spectra. To gain more accurate
demonstrated this disappearance by appropriately calculating values of the electronic band gap of a material, it is crucial to
the screened Coulomb interactions, which eliminated the use many-body methods, such as GW. Corrected energies from
sudden discontinuities in the slope of the LO mode at the BZ the dynamically screened electron−electron correlation fix
center. We find that the LO and the optical transverse (TO) these gaps, which is beyond the scope of the standard DFT. In
modes are, in fact, quite stable and also degenerate at Γ. All the Figure 6 we have shown the comparison between the bare and
optical modes at Γ are also both infrared and Raman active. the G0W0 band structures. We observe that the inclusion of the
However, along the off-Γ direction (near M), particularly the exited-state corrections leads to a significant change in the
in-plane longitudinal acoustic and optical branch frequencies eigen-energies near the Γ point of BZ. This process results in
are infected with complicated torsional motions (see changing the position of NSb indirect and direct band-gaps.
phonons.mp4 in the Supporting Information for a schematic The NSb indirect band gap VBM is now shifted from K Γ to
visualization of the vibrational modes in these materials). We Γ, whereas the direct band gap is shifted from K to Γ.
have also computed the binding energies of these two isolated However, in the case of NAs, the position of band-gaps does
monolayers in their free-standing form. Our calculations show not change. Interestingly, both these materials now show a
14939 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 7. Absorption spectra and joint density of states in (a) NAs and (b) NSb monolayers. Solid line corresponds to the G0W0+BSE calculations
in the absence of lattice vibration. The effect of lattice vibrations on the spectra changes the widths and positions and is shown for temperatures 0 K
(dashed curve) up to 450 K at a step of 50 K. The vertical dashed lines are the direct quasi-particle gaps. The dots are the guide to eye showing the
shifts of the peak positions.

Figure 8. |∇k(EC − EV) |and EC − EV in (a) NAs and (b) NSb monolayer along the BZ route. EC and EV represent the lowest conduction and
highest valence band, respectively.

three times valley degeneracy in their respective valence band. of only one dominant bound excitonic peak (B1) at about 3.15
We report here a large indirect as well as direct G0W0 gaps for eV. The exciton binding energies for these two materials are
both the materials. The important optical transitions for NAs, obtained as 1.62 and 1.30 eV, respectively. We note here that
as we shall see below, is mainly located along the route M Γ the effect of spin-splittings in both valence and conduction
and with a gap of about 5.00 eV. For NSb, the transitions are bands (relatively small in both cases) does not show any
distributed at Γ and M Γ (labeled by I) with a gap of about additional splittings in the absorption spectra. This could be
4.45 and 4.77 eV, respectively. In summary, we have presented the reason since an additional manual Lorentzian spectral
all of the exited-state band-gaps in Table 1. The additional broadening of about 0.1 eV (which is reasonable when actual
energy convergences are demonstrated in Figures S4 and S5 of experiments are performed) is applied, which might have
the Supporting Information. overshadowed the spectra splittings. In the same figure, we
The two-body Bethe−Salpeter equation of motion com- have plotted the joint density of states (JDOS) to understand
pletely describes the excitonic character. In fact, it is the pole of the presence of dark excitons. We carefully observe that below
the two-body Green’s function that gives direct access to the the bound excitons in both materials, the JDOS changes the
absorption spectra. In Figure 7, we present the imaginary part slope. This indicates that there are possibilities of finding dark
of the dielectric function (which is proportional to the excitons. We analyzed the character of these dark excitons and
absorption spectra) for both the materials. Our calculations noted that in the case of NAs, there are degenerate dark
show that the NAs spectrum is dominated by two prominent excitons located at 3.27 (double) and 3.25 (triple) eV. There is
peaks, one at 3.38 eV and the other at 5.37 eV. The former is also a singlet positioned at 3.34 and 3.24 eV. All these are
the bound exciton B1, housed below the onset of the located below the main peak at 3.38 eV. In the case of NSb, we
continuum, while the latter is the resonant exciton R1 and is found that there are two double degenerate dark excitons
located above the continuum. Instead, the NSb spectra consist located at 2.92, 2.96, and 3.07 eV, while the singlets are at 3.12,
14940 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

2.94, and 2.89 eV. In addition, we also found one triply in angular momentum ΔL = 0; however, the valence and
degenerate exciton (two bright and one dark) at 3.05 eV and conduction bands in both materials have an opposite quantum
one double degenerate bright exciton at 3.08 eV. These bright number of spin Ms (with the difference ΔMs = 1) and
excitons at 3.05 and 3.08 eV, however, did not reflect on the implicates the nonforbidden transitions in the intermediate
corresponding absorption spectra. It was, however, suspected coupling. We can now access the information about the inset
that these are due to the spin-splittings, but our careful pictures of the respective excitonic weights in the reciprocal
analyses did not reveal such possibilities. Additionally, we have space. Comparing Figure 9 with Figures 3 and 4 we see that in
also shown how the spectra would be underestimated in case both the materials the optical transition comes from the
the excitonic theory is absent. The equation of motion when valence occupancy with spin Ms = −(1/2) to the conduction
contains only the Hartree kernel, the spectra are governed band with Ms = (1/2). This means that the development of the
through the one-body equation of motion and the peak first bright exciton in these materials is because of the coupling
relocates itself above the continuum. It is also important to between antiparallel spins within the intravalley. This is in stark
note that Carvalho et al.31 presented a lucid picture in order to contrast with the contemporary monolayer TMDCs, where an
understand why an atomically thin layer should absorb light opposite situation prevails.8,30 In Figure 10 we map these first
strongly. Their theory, using a variety of crystal structures,
demonstrated that the optical transitions actually come from
those areas of the BZ where the conduction and valence bands
run parallel to each other. At these critical points, the JDOS
suffers discontinuities and hence contributes to a very large
optical conductivity. This is known as band-nesting and
provides a quick visual inspection about the origin of the
spectrum peaks. For various monolayer TMDCs, they show
that this presence of the band-nesting might be the reason for
obtaining large conductivities. In order to see whether the
band-nesting is present inβ-NAs and β-NSb, we have plotted in
Figure 8 the gradients of the differences between the top
valence and bottom conduction band along the BZ. We
observe that between Γ and M, there are no critical points Figure 10. Isosurface of the first bound exciton with fixed hole (the
where the band runs parallel to each other. This implies that boundary value of the surface is a tenth of the highest value) over the
real space lattice of both NAs ((a) and (c) at 3.38 eV) and NSb ((b)
the JDOS do not suffer discontinuity along this route, and and (d) at 3.15 eV). Subplots (a) and (b) are seen from the top,
therefore, the strong absorption observed in Figure 7 are whereas (c) and (d) are from the side view.
mainly due to the excitonic formation. This particular
information is displayed in Figure 9 where we have
bright bound excitonic isosurfaces onto the real-space lattice of
the respective materials. In order to evaluate this texture, we
fixed the position of the hole on the top of the nitrogen atom
(which contributes mainly to the top of the valence bands, see
the Supporting Information in Figures S2 and S3) separated by
about 1 Å within the unit cell. The extent to which isosurface
spreads gives valuable information on whether the exciton is a
Mott−Wannier (more delocalized) or Frenkel (more local-
ized) type. We observe that the first bright exciton is in the
NAs case more localized than in the NSb case (Figure 10),
which agrees with stronger exciton binding energy in the first
NAs exciton (1.62 eV) than in the first NSb exciton (1.30 eV).
A careful analysis further reveals that these doubly degenerate
bright excitons in fact preserve the crystalline 3-fold symmetry.
De-excitation can be overwhelmed by means of the scattering
of electrons, accounted for, in the structural model, by the
Figure 9. Fundamental exciton formation in (a) NAs and (b) NSb thermal broadening of the bands. In panels (b) of Figure 3 and
monolayers. The electronic transitions representing exciton weights 4 we show possible patterns for the scattering of the electrons,
are projected onto the ground-state electronic dispersion and in both where the scattering event arrows (represented by II and II′)
cases are shown by the yellow shade. The inset exhibits the
corresponding excitonic weights in the BZ of the two cases.
were resolved by studying the life times of all states and so the
de-excitation process can be expected from the bottom of the
conduction band. We discuss this in the next section. The
superimposed the bound excitonic weights onto the bare energy convergences for the BSE spectra are demonstrated in
dispersion diagram to understand better. The yellow shades Figures S6−S8 in the Supporting Information.
demonstrate the vertical electronic transition from the direct Zero-Point Energy-Gaps. All the aforementioned dis-
gaps that contribute to the exciton formation and, cussions are presented under the condition when the influence
consequently, the absorption spectra. In NAs, this is along of lattice vibrations is absent. However, in reality, experiments
M Γ route, whereas in the case of NSb, it is contributed are performed at finite temperatures, and therefore, one should
mainly due to transitions at Γ and along M Γ route. The look onto the temperature-dependent optical properties. In
states of the valence and conduction bands poses no difference 1950, Fan49 first proposed the coupled electronic and lattice
14941 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 11. Spectral functions dependencies on temperature in NAs and NSb monolayer. (a), (b) Spectral function NAs monolayer at the direct
valence and conduction state (see Figure 3b). (c), (d) Same at the indirect valence and conduction state, respectively. The spectra in (e)−(l)
depict the cases for the NSb monolayer (see Figure 4b at high-symmetry points). The vertical dashed lines represents the band edges in the absence
of electron−phonon couplings.

vibration self-energy using the first-order perturbation theory. otherwise distinct electronic state. Below we address how the
Later, based on the crystalline X-ray diffraction on nonpolar bare states |nk⟩ and energies ϵnk of β-NAs and β-NSb
diamond and ZnS type crystals, Antončı ́k50 modified the monolayers are affected in the presence of this electron−
Debye−Waller (DW) second-order perturbation theory by phonon coupling.
adding empirical factors to it. In a chronological series of Following eqs S.4 and S.6 in the Supporting Information of
papers, Keffer,51 Walter,52 and Kasowski53 used the DW Kolos et al.,27 we first compute the Fan (∑Fan
nk (ω,T)) and DW
factors to estimate temperature-dependent band-gaps of PbTe, (∑DW
nk (T)) self-energies. The former is the dynamic one and is
GaAs, and NiS, respectively. However, it was not until 1976 a complex quantity, while the later is a static and purely real.
when Allen and Heine54 made the DW term a transitional The temperature dependencies on these two self-energies
invariant and thus were able to unify the Fan theory with the comes from their respective Bose occupation numbers. Once
DW. This approach allows them to expand the DW self-energy these self-energies are computed, the interacting electronic
as a product of the first-order derivatives of the self-consistent Green’s function (Gnk(ω,T)) can then be constructed from
potential. Following, Cardona and Allen,55−57 in a series of 1
papers, developed the electron−phonon renormalization Gnk (ω , T ) = Fan DW
theory. Together with Heine, this is now known as the ω − ϵn0k − ∑nk (ω , T ) − ∑nk (T ) (1)
Heine, Allen, and Cardona (HAC) theory. They argued for the in which ϵ0n,k is the bare electronic energy. The dressed

ÄÅ ÉÑ
fundamental fact that even at zero temperature, due to the

ÅÅ ÑÑ
electronic energies are the outcome of real part of the pole

ÅÅ Ñ
inherent quantum mechanical atomic spatial uncertainties, the

ÅÅω − ϵnk − ℜ ∑ (ω , T ) − ∑ (T )ÑÑÑ − iℑ ∑ (ω , T )


ÅÅ ÑÑ
couplings between electron and atomic vibrations can modify Fan DW Fan

ÅÇ ÑÖ
0
the eigenenergies of the former. This electronic energy
correction at zero Kelvin is the zero-point renormalization nk nk nk
(ZPR) or motion (ZPM). However, because of the presence of =0 (2)
an immense number of computational complexities of these
self-energies, the scientific community abandoned this topic in whereas the imaginary portion gives the electronic lifetime in
Fan
the early days in an anticipation that these corrections might that state. If the pole of ℜ ∑nk (ω , T ) and ϵ0n,k are not far
be extremely small when compared to the dynamic electron− from each other, then the real portion of eq 2 can be linearized
electron couplings. Eventually, the importance of ZPR by using the Taylor’s expansion up to first order followed and
resurfaced when it was observed that these corrections led to assigning a weight factor Znk. The dressed energy uncertainties

ÄÅ DW ÉÑ
a giant ZPR in diamond58,59 (in range of 615−622 meV) and
ÅÅ ÑÑ
(now called as quasi-particle energy spreading) now become

ÅÅ Ñ
LiF36 (718 meV). Marini21 and co-workers28,60,61 demon-
Δϵnk = ϵnk + ZnkÅÅ∑ (T ) + ℜ ∑ (ϵnk , T )ÑÑÑ
ÅÅ ÑÑ
Fan
strated the computational strategies of the ZPR in a purely ab
ÅÇ nk ÑÖ
0 0
initio way within the framework of DFPT. The inclusion of
these self-energies adds a finite energy spreading of the nk (3)

14942 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

ij yz
where the dimensionless weight factor 1

j z
constructed from A nk (ω , T ) = π |Gnk (ω , T )| leading to a
Znk jj = 1 − ∂ω ∑nk (ω , T ) z is a value between 0
ω =ϵnk z
−1

k {
∂ Fan

and 1. The spectral function at each state and band can also be Lorentzian-like distribution function

Fan
1 ℑ ∑nk (ϵn0k , T )
A n k (ω , T ) =
π [ω − ϵ0 − ℜ ∑Fan (ϵ0 , T ) − ∑DW (T )]2 + [ℑ ∑Fan (ϵ0 , T )]2
nk nk nk nk nk nk (4)

whose peak is centered at Δϵnk. A Zn,k → 0 implies a very indirect gaps and the ZPR respectively. We have also listed the
strong quasi-particle (electron−phonon) interaction leading to room temperature optical gaps in Table 4. The necessary
a flattened (distorted Lorentzian) Ank(ω,T) spectra. The well- spectral function convergence with respect to the number of
defined electronic state now spreads more in energy and, electronic bands is shown in Figure S9 of the Supporting
therefore, as per Heisenberg’s uncertainty rule, acquires a Information. In Figure S10, we have additionally shown the full
shorter lifetime. This is indicated through a quick scattering of electronic dispersion at 0 K.
the electron to a new accessible state. Instead, a Zn,k → 1 Another interesting observation from Figure 11a−d can be
means that the interaction is weak and is characterized by a seen for NAs valence and conduction energy line widths at the
sharp Lorentzian peak. The energy spreading then tends to a same temperatures. We highlight those cases of electronic
well-defined state, allowing the electron there to live longer. transitions which contribute to the formation of excitons. One
Figure 11 shows exactly this scenario. We have shown the can see that at a particular temperature the line width of the
spectral function for three different temperatures, 0, 300, and conduction band at the direct point is more than that
450 K. These are computed at the direct and indirect (valence compared to the corresponding indirect point. This means
and conduction) points of the BZ (see Figures 3 and 4). The that, after a transition (vertical) to the conduction band, an
direct points are those which contribute to the optical electron quickly gets scattered to the corresponding indirect
absorption, whereas the indirect points are included to state. This mechanism path (I−II) is shown in Figure 4b by an
broaden the view of the temperature-dependent electronic orange color arrow. This is not the case with NSb. There are
gaps. We observe that the spectral function in all the cases two major pathways one at Γ (labeled in magenta as I′−II′)
diminishes and become more asymmetric with an increase in and the other between Γ and M (labeled in orange as I−II).
temperature, signifying a strong coupling. For example, analysis The optical gap at Γ is more than the direct gap between Γ and
on NAs indirect valence band shows that the ℜZnk drops to M. However, since the line width of the conduction band at
almost 70% at 450 K. Effectively, Znk physically represents the the direct point (between Γ and M) is comparatively large
quasi-particle charge distribution and is the area under the than that at Γ (see Figure 11f,j) it is more likely that the
spectral function Ank(ω,T) . The energy distance between the electron from pathway I will scatter along the pathway II at M.
valence and conduction spectral function peaks is the This leaves the excited electron at Γ to scatter into state K
renormalized band gap. There are some other interesting (pathway II′). This is also plausible since both the scattered
activities at the direct gap point of NAs. We see that the electrons cannot be in the same state with same quantum
indirect gap actually reduces as the temperature increases and numbers. We now justify the temperature dependencies of the
is characterized by a right (left) movement of the valence gaps within the purturbative theory. To understand this, we

ÄÅ ÉÑ
(conduction) spectral function peak. The deviation of the compute the Eliashberg function63,104

ÅÅ |g qν |2 Ñ
ÅÅ 1 Λ nn ′ k ÑÑÑ
peaks at 0 K from their respective frozen atom (DFT) band

g Fnk(ω) = ∑ ÅÅÅ ÑÑδ


n′qν Å 2 ϵnk − ϵn ′ k ÑÑÑ
edges (see the vertical lines in Figure 11) gives the ZPR. This qν

ÅÇÅ nk ÑÖ
2 nn ′ k
trend is observed generally in most semiconductors. Instead, −
we found that the direct gap between between M and Γ ϵ − ϵn ′ k − q
increases. We show these results in the same figure for NSb
(ω − ωqν) (6)
also, in which the indirect gap (from M to Γ through K and
direct gap (labeled as I in Figures 4 and 6b) decreases, but the
direct gap at Γ increases. To account for the decreasing band within the adiabatic approximation, i.e., |ϵnk − ϵn′k−q| ≫ ωqν, in
gap trends, we used the Varshni’s empirical model71 which the “primes” denotes the scattered states. The
numerator of the first term inside the parentheses stems
αT 2 from the first-order electron−phonon matrix elements of the
Eg (T ) = Eg (0K) − Fan self-energy, while the numerator in the second term is due
(T + β ) (5)
to the second-order DW self-energy. The sum is over the
in which α and β are the numerical fits and Eg(0K) is the band empty electronic states and for each phonon branch and
gap at 0 K. We see that in both materials, the decrease of the momenta. The delta function simply states that g2Fnk(ω) is
indirect gap with temperature is rather slow. However, at the proportional to the phonon DOS. We evaluate g2Fnk(ω) at the
direct gap of NAs, the rate is quite large. The net deviations of VBM and CBM and obtained their difference g2Fc(ω) −
the VBM and CBM at 0 K with respect to the frozen atom g2Fv(ω) for both NAs and NSb. This is shown in Figure 5c,f,
edges (i.e., the ZPR) for NAs and NSb are found to be 105 and respectively. We observe that the difference is negative. Such
80 meV, respectively. The reason for a low ZPR of the latter is behavior are commonly found in most semiconductors where
because of the heavier atomic mass of Sb, which indicates that the g2Fc(ω) is negative while g2Fv(ω) is positive. This
heavy atoms vibrate less. In Tables 2 and 3, we have eventually cancels out any increment in the gap with
summarized the values of the temperature rate of change of temperature, and therefore, the indirect gaps reduces.
14943 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 12. (a) NAs and (b) NSb excitonic oscillator strengths as a function of temperature for all exciton with strengths more than 20%.

Figure 13. Linear optical properties of NAs (left column) and NSb (right column) in ultraviolet and visible regions for frozen atom approach and
300 K: the refractive index n, the extinction coefficient k, and the absorbance A as a function of wavelength.

However, this does not seems to be true at the direct gap tonian. The sharp distinction seems to appear in the 0 K
points in NAs (between M and Γ) and NSb (at Γ), which spectrum when compared with the corresponding frozen-atom
shows an increasing trend. We discuss this situation below case. We see that the effect of the ZPM drops the intensity
during the exciton−phonon interaction. significantly, revealing the true spectral broadening. In the case
Temperature-Dependent Absorption Spectra and of NAs, we see that the bound B1 and the resonant R1 excitonic
Excitonic Line Widths. Marini21 demonstrated how the peaks blue shift in energy, indicating that the direct gap
quasi-particle energy spreading (Δϵnk) due to the electron− increases with temperature. At 0 K, the shift of B1 is only about
phonon couplings can be introduced in a fully ab initio way in 20 meV. This is not the case with NSb. Instead, we see that the
the development of excitonic formalism through the BSE. This effect of ZPM has clearly red-shifted the 0 K B1 peak by a giant
careful study has indicated that the energy corrections from the 430 meV compared to its frozen-atom condition. We also
dynamical HAC theory in the development of excitons cannot observe that up to 350 K, there is not much change in the
be avoided if one aims to justify the optical properties properly. energy position of this peak, suggesting that the fundamental
Figure 7 shows the optical absorption spectra in the excitonic energy in NSb is a very slow function of temperature
temperature regime 0−450 K at a step of 50 K. All spectra compared to NAs. The two NSb peaks above 4 eV broaden to
are computed using the BS kernel. As polaronic excitation a single peak with a large width as the temperature increases
energies are involved now, we go beyond the Tamm-Dancoff further. Figure 12 shows interesting characters about the
approximation105 in order to correctly account for the excitonic oscillator strengths in the range 0−450 K. For better
electron−hole pairs and antipairs in Fock-space BS Hamil- clarity, we have plotted only those excitonic energies which
14944 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

have more than 20% oscillator strengths. Excitons with >20%


are shown in this figure with filled circles of varying colors. We
now focus on the first pair of excitons of NAs at 0 K, located
between 3 and 3.5 eV. We see that the first exciton of this pair
is dark. However, as the temperature rises, the oscillator
strength increases and makes it bright. Interestingly, the other
exciton adjacent on it, it is initially bright and starts losing its
strength as the temperature increases. As these two excitons
are very close to each other, we understand this effect through
an exchange of oscillator strength. In the case of NSb, we see
that the first exciton (between 2 and 3 eV) is always dark,
whereas the second one always remains bright. However, the
exchange of oscillator strengths starts between the third and
fourth excitons when they come close to each other above 200
K. Globally, we observe that most excitons in the two
monolayers are dark (except a few resonant excitons between
5.5 and 6.5 eV in NAs and a few bound excitons between 4 and
4.5 eV in NSb). We also observe the same blue and redshifts of Figure 15. (a) Exciton−phonon coupling (black solid line) as a
the fundamental bright excitonic energies in NAs and NSb, function of phonon frequency, showing the most prominent mode
that builds the NSb fundamental bright exciton. The shaded areas
respectively. In Figure 13, we also provide information on the represents the coupling function of valence and conduction bands at
influence of temperature on the refractive index n, the the direct-point Γ (see Figure 4b). (b) Inhomogeneous line width of
extinction coefficient k, and the absorbance spectra of both this fundamental exciton. The symbols are the ab initio data while the
materials. We observe that the refractive index converges to dashed line is the phenomenological excitonic width equation.
1.63 and 1.81 for NAs and NSb, respectively. This figure also
shows a strong absorbance peak in the visible light region for
NSb, on the other hand, the absorbance peaks for NAs case are Comparing with the phonon dispersion and DOS in Figure
mostly in the UV region. Temperature effects on these 5a,b, we see that in NAs, the most prominent exciton−phonon
parameters are mostly on the broadening of the peaks. interactions are located around the optical frequencies near
Particularly in the influence of temperature on the absorbance 580 and 680 cm−1 respectively. These reflect an in-plane E and
spectra is a visible lowering of the maximal values from 37% out-of-plane A1 infrared modes at the phonon zone center. In
(NAs) and 26% (NSb) to values between 15 and 20%. the acoustic region, the interactions are mainly from the mode
In order to understand which phonon mode couples to the near 270 cm−1 and are infected mostly with an in-plane
exciton during its development, we have computed the torsional motion (see Supporting Information for a schematic
exciton−phonon coupling function, which is similar to the visualization of the vibrational modes in these materials). We
previously described generalized Eliashberg function.21 These also note that near lower frequency, the coupling function
are shown in subplots (a) of Figures 14 and 15. We evaluated crosses the zero line and becomes positive. This is an
this coupling function for both the valence and conduction anomalous behavior that justifies the increase of the direct
state at the direct gap (between M and Γ, see Figure 3b). gap with temperature (see Figure 16a). Similarly, in NSb, we
see that the main exciton−phonon interaction is located
around 560 and 620 cm−1 of the optical frequency,
respectively. These correspond mostly to the in-plane and
out-of-plane vibrations. In the lower acoustic region at about
50 cm−1, we note a prominent interaction from the in-plane
mode. There is, however, a small encroachment of the coupling
function (around 220 cm−1) in the upper positive region. This
again leads to an anomalous behavior and justifies the increase
of the direct gap with temperature at Γ (see Figures 4b and
16b). The temperature-dependent excitonic inhomogeneous
line widths are shown in Figures 14b and 15b. We write here
the phenomenological nonradiative line width equation

Γ(T ) = Γ0(0K) + γacT + ÅÄ ÑÉÑ


ÅÅ
γop

ÅÅexp op − 1ÑÑÑ
ÅÅÇ ÑÑÖ
ℏω
( )
kBT (7)
in which Γ0(0K) is the 0K broadening. The constants γac and
γop are the exciton-acoustic and exciton-optical interaction
strengths, respectively, ωop is the optical frequency and kB is
Figure 14. (a) Exciton−phonon coupling (black solid line) as a
function of phonon frequency, showing the most prominent mode
the Boltzmann’s constant. We fit eq 7 to our result shown in
that builds the NAs fundamental bright exciton. The shaded areas Figures 14b and 15b to report that the interaction strengths γac
represent the coupling function of valence and conduction bands at and γop are 0.4 μeV K−1 and 36 meV for NAs at ωop = 580
the direct point (see Figure 3b). (b) The inhomogeneous line width cm−1 and 0.18 μeV K−1 and 52 meV for NSb at ωop = 620
of this fundamental exciton. The symbols are the ab initio data, while cm−1, respectively. We obtained that Γ0(0K) in NAs and NSb
the dashed line is the phenomenological excitonic width equation. are 105 and 129 meV. At room temperature, the line widths
14945 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 16. Temperature dependence of the DFT gaps in the presence of electron−phonon couplings in (a) NAs and (b) NSb monolayers. The
direct gap in (a) corresponds to transition I (in M Γ, see Figure 3(b)), while the indirect gap is between K Γ and M points of the BZ. The direct
gaps in (b) correspond to I and I′ transitions (at M Γ and Γ, respectively, see Figure 4(b)), while the indirect gap is between K Γ and M points
of the BZ.

are in the range of 250−280 meV. The residual width at 0 K calculated by equating this power to either one of the following
can be obtained by taking the slope of Γ(T) at the bending two empirical models namely, “bulk” and “sheet” models.85
knee down to 0 K. This corresponds to almost 94 and 110 Clark et al.84 compared these two models and showed them to
meV for NAs and NSb, respectively. Effectively, these are large (2) n2D(ω) n2D(2ω) (2)
widths when compared to only 1−10 meV found in be related as χbulk = 32πF χsheet in which n2D is
(ns(ω) + 1)3
conventional TMDCs monolayers on various substrates (see the refractive index of the monolayer at frequencies ω and 2ω
Table 5) and therefore suggest that the excitonic motions in and ns is the refractive index of the underlying substrate. The
NAs and NSb are diffusive rather than ballistic. factor F (which appears in the bulk model) is the so-called
Nonlinear SHG Spectra. We now present the optical “scaling factor” in the reflectance and transmittance signals and
nonlinear second harmonic coefficients generated from these stems due to the large mismatch of the refractive index
monolayers. In the presence of an intense light, the induced between the sample-substrate and sample-air. This often gives
macroscopic polarization in the frequency domain is related to rise to a severe overestimation of the SHG coefficient. For
the external time-dependent electric field ,(t ) as15 example, when the MoS2 experimental intensity counts are
used in the bulk model, the SHG is estimated to be on the
7i(ω) = ε0{χij(1) (ω),j(ω) + χijk(2) (−ω; ω1, ω2),j(ω1), k(ω2) + ...} order of 103 (CVD)−105 (exfoliated) pm/V.83 The sheet
(8) model instead predicts an order of 102 pm/V82,108 only.
where the index i denotes the polarization direction and j, k, Similarly, in the case of monolayer TMMCs, the sheet model
etc. are the directions of the electric field. The first term in the results an SHG of the order 102 pm/V109 which is
parentheses holds true for linear dielectrics, whereas the next overestimated by 103 when the bulk model is used.101 In
term onward represents nonlinearity. The linear response fact, the prediction of SHG coefficient using purely first-
principle calculations, however, indicates that the sheet model
function ℑχij(1) (ω) is proportional to ℑεM at low intensity. At is more accurate.75,109,110
large intensities and under the condition ω1 = ω2 = ω, Nevertheless, following Attacallite et al.,20 the nonlinear
χijk
(2)
(−ω;ω1,ω2) denotes the second harmonic coefficient. optical responses can be predicted by integrating the time
Furthermore, these tensorial components under the C3v dependent Schrödinger’s equation of motion (EOM)
ı
(3m) crystalline symmetry converges as χ(2) bbb = −χbba = −χaab
(2) (2)

= −χaba in which a and b are the in-plane crystal axes.


(2) d
iℏ |υnk ⟩ = [/k + i ,(t ) ·∂ k]|υnk ⟩
It should be noted that the optical nonlinear responses are dt (9)
not measured directly. Instead, light from a pump with a
tunable power is incident on the sample, and the for the time-dependent occupied periodic Bloch electronic
corresponding reflected average power is then measured. If states |υnk⟩. The second term in the parentheses represents the
the sample is monolayer, the response χ(2) ijk (−2ω;ω,ω) is then coupling between the time-dependent external electric field
14946 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 17. (a) Convergence demonstration of the absorption spectra using the approach of nonlinear real-time BSE and static BSE for NAs
monolayer. The shaded region (yellow) demonstrates that in the presence of a delta-like field with low intensity (500 KWcm−2 in this case), the
linear response (black) can be obtained from the solution of the time-dependent BSE. A scissor is added in the exciton Hamiltonian to mimic the
G0W0 gap in both cases. The time-dependent induced polarization along the crystalline a-axis is also shown due to (b) delta-like field and (c) a
quasi-monochromatic field along the b-axis, respectively.

Figure 18. (a) Convergence demonstration of the absorption spectra using the approach of nonlinear real-time BSE and static BSE for NSb
monolayer. The shaded region (yellow) demonstrates that in the presence of a delta-like field with low intensity (500 KW cm−2 in this case), the
linear response (black) can be obtained from the solution of the time-dependent BSE. A scissor is added in the exciton Hamiltonian to mimic the
G0W0 gap in both the cases. The time-dependent induced polarization along the crystalline a-axis are also shown due to (b) delta-like field and (c)
a quasi-monochromatic field along the b-axis respectively.

and the Bloch electrons. Here the coupling ∂̃k is the k- electron−hole correlation is used. The self-consistent Hartree
derivative operator due to the Born-von Kármán periodic potential VH (as functional of time-dependent electronic
boundary conditions. The states |υnk⟩ under such conditions density, ρ(r, t)) and static screen exchange (SEX) can then
remain invariant under unitary rotation.67,111 /k is the model be convoluted with density fluctuation as20
Hamiltonian where a number of hierarchy can be used to
approximate the electron−hole correlation. In case of a single- /k = / KS
k + ∑ Δn|υnk0 ⟩⟨υnk0 | + VH[Δρ(r , t )] + ∑ [Δγ ]
particle Kohn−Sham states (i.e., no correlation between the nk SEX
(10)
electron and hole), /k is simply / KS
k with bare energy eigen-
value ϵ0nk. This is an independent-particle picture. If the in which Δρ(r,t) = ρ(r,t) − ρ(r,0) is the incremental density
dynamic electron−electron correlation is added (i.e., the G0W0 and Δγ(r,r′,t) = γ(r,r′,t) − γ(r,r′,0) is the density fluctuation
correction), the picture stands then as an independent quasi- matrix. The second term in the above equation represents a
particle approximation. Excitonic picture is captured when the scissor operator that mimics the electron−electron self-energy
14947 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 19. Phase-delay between the induced polarization and the applied field in monolayers of (a) NAs and (b) NSb respectively. The real and
imaginary parts of the linear spectra is obtained from the real-time approach. All the analyses are obtained within the TD-BSE level of theory. Note
the convergence of the delay to π at low frequencies.

correction through Δn = ϵGn 0W0 − ϵ0n. Equation 10 when used as macroscopic polarization over time. An important information
a kernel in eq 9 is the TD-BSE. In-fact it was also shown that in can be obtained if one plots the phasor diagram of the
the presence of an impulse-like field with weak intensity, eq 9 response χ(1)
ij (ω). The complex phase delay ϕ between 7 and
converges to the solution of the time independent (static) , can be evaluated from χ(1) aa (ω) = |χaa (ω)|exp(iϕ). This is
(1)

BSE.20 Since the position operator in the reciprocal space of shown in Figure 19. The boundary ϕ = π/2 corresponds to the
periodic structures appearing in the coupling EOM above is in-phase oscillation of the in-plane macroscopic polarization
not defined properly, a Berry’s geometric phase is therefore ∂P
current − ∂t with , .20 We see that ϕ closely follows ℜχaa(1) (ω).
used to construct an in-plane polarization
ij 2π yzzz
At low frequencies where the absorption is minimum, ϕ → π,
j
detSjjjk, k + z
Nk zzz
Nk − 1 signifying plasmonic oscillations. The delay of the first bound
jj
eg |a|

k {
7=− s ∑ ℑlog ∏ B1 absorption peak appears to be greater than π/2 and thus
2π Ω Nk⊥ signatures clearly a localized excitation. Instead, at the peak
k⊥ k (11)
near 5.4 eV (R1 in case of NAs), the delay decreases below π/2
once |υnk⟩ are known. This was prescribed by Resta and 112
indicating a delocalized exciton.
Vanderbilt19 in their modern theory of polarization, and later The subplot (c) in Figures 17 and 18 shows the in-plane

i 2π y
spin degeneracy, Sjjjk, k + N zzz is the overlap matrix between
used by Attaccalite and co-workers.18,20,109,110,113 Here gs is the macroscopic polarization over time when the in-plane electric
field is harmonic. Careful observation reveals that the sudden
k {
k switching on of the electric field results in spurious initial
the states |υnk⟩ and |υmk + 2π / Nk ⟩, and Nk∥ and Nk⊥ are the spikes in the polarization spectra. To filter out these spikes and
get a clean polarization signal, a dephasing time of about 8 fs,
respective in and out-of plane k-points to the polarization equivalent to a damping of 0.2 eV, is applied. This
direction and Ω is the cell volume. Equation 9 is then solved phenomenological damping also imitates the presence of any
numerically using the Crank−Nicolson algorithm41 experimental ambient conditions like damping due to thermal
broadening, defects, etc. We observed a clean signal after about
I − i(△t /2)/k(t )
|υnk (t + Δt )⟩ = |υnk (t )⟩ 30 fs (about 4 time constants). The simulation was allowed to
I + i(Δt /2)/k(t ) (12) run for a total of 55 fs. Within this time window, the extraction
of χ(2)
aab(−2ω;ω,ω) was done at equally spaced 100 frequencies
at a time step of Δt where I is the identity element. between 0.5 and 4 eV. To get some meaningful insights from
First, in order to probe the response at all frequencies, we the SHG spectra, we first note that the nonlinear responses are
choose an in-plane δ-like impulse electric field in our model computed using various choices of gauge. These responses, for
Hamiltonian. We then show that at weak field intensities, the example, χaab (2)
(−2ω;ω,ω), are plagued with unphysical
linear response function ℑχij(1) (ω) obtained from the solution divergences of the form ω−1 and ω−2 as their leading orders.
of eq 9 converges to the ℑϵM spectra (i.e., the static BSE). For Sipe and Aversa,114 however, demonstrated that these
this, we used a low (500 KW cm−2) intensity. To compute the divergences could be avoided if one uses a length gauge.
self-consistent VH and the SEX convolution integrals, the same Instead, if the gauge choice is velocity, then at low frequencies
cutoff of 10 Ry was used (similar to BSE). The EOM was then (ω → 0), such divergences can be removed either by
integrated at a time step of 10 as Figures 17 and 18 separating the intra- and interband transition processes or by
demonstrates this scenario for NAs and NSb monolayers, applying a sum over the transition band rule. The length gauge,
respectively. We see from the subplots (a) in both cases that although rules out this problem but could only be applied to an
isolated system where the boundaries are aperiodic, like in
the linear coefficient ℑχij(1) (ω) is in excellent agreement to the clusters and molecules. Sipe114−118 along with Ghahramani119
respective absorption spectra ℑϵM . The subplots (b) show the showed that these divergences can be overcome in a periodic
typical corresponding behavior of the induced in-plane system and can also be independent of any choice of gauge.
14948 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 20. (a) Absolute SHG coefficient in NAs monolayer within the time-dependent (TD) BSE (TD-BSE) level of theory exhibiting the peaks
A1, and A′1 due to the bound and resonant excitons B1 and R′1 respectively. The x axis is the laser frequency. (b) The absorption spectra at ω and ω/
2 under the same TD-BSE level of theory. The vertical dashed line exhibits the ω/2 gap. One can see that the ℑχaab (2)
goes to zero below half of the
gap. (c), (d) All respective cases in the NSb monolayer.

The sum-over transition bands demonstrated that the again a 2ω resonance and is due to the resonant exciton R1′.
aab(−2ω;ω,ω) resonance occurs at ω and 2ω. However, this
χ(2) The corresponding SHGs at 810 and 1560 nm are close to
is only possible when the electron−hole interaction is at their each other at 27 and 31 pm/V, respectively. The SHG
lowest hierarchy, i.e., at the independent particle approx- coefficients of these lossless windows come out to be at par
imation. In such a case, it becomes easy to separate the intra with those obtained from monolayer TMDCs, TMMCs, and h-
and interband transition processes. However, in the presence BN (extracted after using the sheet model). The maximum
of local fields and excitons, such separation cannot be done SHG in each case occurs at 3.2 eV or 387.5 nm (636 pm/V)
very easily and therefore to understand the ω and 2ω for NAs and 2.6 eV or 476.9 nm (270 pm/V) for NSb. These
resonances, its stands reasonable to compare the are near the elemental nitrogen (337.1 nm) and Argon (476.5
χ(2)
aab(−2ω;ω,ω) spectra with ℑε(ω) and ℑε(ω /2). Such a nm) gas lasers. In addition, we have also shown the real and
comparison, although a qualitative approach, provides great imaginary part of the SHG spectra for a complete under-
details about the nonlinear responses.18,26,27,67,109,120,121 Figure standing of how these are coupled together through the
20 represents such a comparison between |χ(2) aab(ω) |and the Kramers and Kronig relation.15
absorption spectra ℑε(ω) and ℑε(ω/2) of NAs and NSb. We Optical Coefficients in the Presence of an in-Plane
use the real-time BSE kernel in the Hamiltonian to evaluate the Biaxial Strain. We now focus our attention on the
SHG spectra. In the case of NAs, the first peak A1 at 1.68 eV is improvement of these optical coefficients by applying a
in excellent agreement with the peak B1 of the ℑε(ω/2) suitable in-plane strain. Recently, there have been some
spectrum. By comparing ω/2 spectra, we observe that A1 benchmark experiments which show that it is possible to strain
actually stems due to a 2ω resonance. Similarly, the shoulder the traditional monolayers122 both uniaxially32,123,124 and
peak A1′ at 2.81 eV stems out due to the resonant exciton A1′. biaxially125,126 and are supported by various ab initio
This is also a 2ω resonance. Comparing with the 1ω resonance studies.127−131 In principle, any application of strain on a
spectra, we see that its contribution starts beyond 3.3 eV. In two-dimensional system tends to alter the lattice constant and
addition, we see that the SHG spectra are dominated by a primarily changes the electronic overlap wave functions,
largest peak at 3.20 eV, which is due to the various resonant thereby tuning the electron and phonon eigenenergies.
excitonic process. As ω → 0, χ(2) aab(0) → 16 pm/V. At this Heuristically, a uniaxial strain can be applied mechanically by
point, we note here that the losses inside a silica fiber optic are depositing the monolayer onto a flexible substrate and bending
minimum at a wavelength of 810 and 1560 nm. These two the latter. This procedure may produce a homogeneous strain
lossless windows are therefore favored much to send optical but breaks down the initial crystalline symmetry. A biaxial
signals. We find that at 810 and 1560 nm, χ(2)aab(ω) → 56 and 17 strain instead can be applied by depositing the monolayer onto
pm/V, respectively. In the case of NSb, we observe that the a substrate with a smaller thermal expansion coefficient. Upon
first 2ω SHG peaks at 1.32 eV and are due to the B1 exciton. heating or cooling, a homogeneous biaxial tensile or
This is followed by a broad spectrum in the region 2−3 eV, compressive strain can respectively be achieved, which also
where we see that A1′ is spread over large energies. This is preserves the initial crystalline symmetry. Since the thermal
14949 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

expansion or cooling is a time-bounded interplay, thus the that a biaxial tensile strain might prove to thermally stabilize
biaxial strain becomes a time-dependent effect. This is contrary these monolayers.
to the uniaxial case, where the time-dependency could be made Figure 22 shows the G0W0 electronic gap variations at
very slow depending on the substrate stiffness toward bending. different strain levels. The strain impact on the excited-state
As a result, the biaxial strain may find usefulness in dedicated energies are clear from subplots (a) and (b) exhibiting a
optical devices where real-time dynamics are primarily decrease in the direct band gap as the monolayer stretches.
important, like optical sensors, etc. However, the gap tends to increase with compression also until
In the following, we apply an in-plane biaxial strain σij by it starts to decrease beyond −8% for NAs and −4% for NSb. A
altering the equilibrium in-plane lattice constant. Here i and j careful analysis indicates that in the case of NAs (NSb), the
denote the usual in-plane directions. While carrying out an ab direct gap between M and Γ shifts to K below −8% (−4%)
initio computation, this in-plane strain is evaluated from the strain. The indirect gap, however, varies slowly at compressive
strains for NAs, but significantly reduces upon stretching. A
real-space translation lattice vector ⎯→
⎯a ′ = (1 + σ ) ⎯→
ij

ij a ij . This similar trend is seen in the case of NSb, which indicates that
also preserves the initial crystalline symmetry, and thus the during compressive strain, the dynamic correlation is less
optical nonlinear coefficients are expected to remain as sensed toward the energy differences between indirect gap
nonvanishing quantities. We note here that since the points.
traditional monolayers are reported to withstand up to a strain The influence of strain on the absorption spectra and JDOS
as large as 11% while in their suspended conditions;132 are shown in Figures 23 and 24. In the subplots, we have also
therefore, up to a ± 10% strain is used here. All convergence included the RPA spectra for each strain level for comparison.
energy thresholds used in the strainless conditions are found to We see that in the case of NAs, the tensile strain tends to
be sufficient in the above-mentioned strain range. Table 6 redshift the first bound exciton. This exciton in the presence of
shows the dynamic stability of the optical phonons at Γ of the strain (B*) is now also reduced in intensity compared to its
BZ. We observe that the influence of a tensile strain up to strain-less counterpart B1. The excitonic peak, however,
+10% actually reduces the optical mode frequencies in both continues to red-shift until +8%. At 10%, we see some
the monolayers without splitting ωLO − ωTO at Γ. However, emergence of shoulder peaks on the left of B* indicating
the difference between the optical phonon-gap frequency, i.e., electronic transitions from different parts of the BZ (see Figure
between ωLO − ωTO and ωZO, increases with strain. In the case 26). Additionally, we see the effect of the decrease in the
of a compressive strain, these scenarios behave in just opposite corresponding G0W0 gaps, which anticipate the decreasing
trends in the binding energies (shown in Figure 25).
ways. The mode frequency increases together with the
Surprisingly, the resonant peak near 5.37 eV gets more
reduction of the optical phonon gap. The acoustic modes,
intense, indicating it to become a more bright exciton. In
however, show an interesting behavior. It appears that the
effect, we see that the tensile strain (between 0% to 10%) tends
effect of tensile strain removes the imaginary frequencies in the to reduce the optical transparency (on-set of 10% of the
phonon dispersion of both the monolayers. A tensile strain fundamental excitonic peak position) region by about 0.9 eV.
beyond +4% starts to recover the well-known quadratic Instead, the compressive strain shows a completely opposite
behavior of the out-of-plane acoustic mode, indicating better behavior. The fundamental exciton B* now blue shifts and
dynamic stability. For example, see the phonon dispersion at tends to become more bright with increasing strain, whereas
+8% tensile strain demonstrated in Figure 21. We found that the resonant peak gradually dies off. Between −8% to −10%
the compressive strain tends to put more imaginary frequencies strain, we rather see a dramatic movement. The peak B* at
in the out-of-plane acoustic mode, thereby destabilizing both −8% has now emerged as two prominent peaks (see Figure
the monolayers (see Figures S11 and S12 in the Supporting 23j) specifying the on-set of optical transitions from other BZ
Information for a complete phonon dispersion in the presence points (see Figure 27 below). We, therefore, see that the
of tensile and compressive strains). Such a tendency indicates compressive strain (between 0% and −10%) tends to increase
the optical transparency by about 0.4 eV in NAs. In the case of
NSb, we observe similar trends except for the emergence of the
secondary peak C* at higher tensile strain. At compressive
strains, we see that the peak B* gradually diminishes, and while
many new secondary peaks A*, C*, and D* start to emanate.
These are possible due to various optical transitions occurring
at different BZ points. In effect, the optical transparencies
reduce by about 1 eV in the case of tensile, while increase with
minor modulations at higher negative strains (see Figure 24i,j).
The fundamental bound excitonic energies are shown in Figure
25a, where we observe the movements of the peaks with strain
in both monolayers. The binding energies in Figure 25b rather
show a slow variation in NAs (between +2 and +8% strain).
This also indicates that these energies in NAs remain almost
immune to tensile strain. A possible justification for this
behavior is because of the simultaneous change in the G0W0
gap and the excitonic energies that try to keep the difference
constant. This does not seem to be true for NSb at larger
Figure 21. Phonon dispersion in monolayers NAs (left) and NSb tensile strain where the binding energies reduce sharply below
(right) for biaxial strain of 8%. 50% of its value at the 0 strain. It is worth mentioning here that
14950 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 22. Quasi-particle G0W0 gaps (in eV) in biaxially strained (a) NAs and (b) NSb monolayers.

Figure 23. Absorption spectra and joint density of states of monolayer NAs under biaxial (a)−(e) tensile and (f)−(j) compressive strains,
respectively. The vertical dotted lines are the most important electronic transitions with maximum relative excitonic intensity, while the dashed
lines are the direct quasi-particle gaps.

the trend of red and blue shifts of the fundamental peaks with in band structures induced by strain. To explain this, one may,
tensile and compressive biaxial strains, respectively, was also however, look at the GW band structures, which are shown
found similar to both theoretically and experimentally for below in Figures 28 and 29. The most significant influence of
monolayer MoS2.126 The influence of the biaxial strain on the tensile biaxial strain on the band structures of both materials is
position of the first exciton weights in the BZ is shown in in the region of Γ point of the BZ, where the gap is
Figures 26 and 27. In both materials, the effect of tensile strain significantly lowering, resulting in nearly another valley
is the gradual shifting of the first bound exciton weight from degeneracy of conduction band at Γ under +10% biaxial
the region between M Γ to the Γ point of BZ. The strain. The influence of the tensile biaxial strain on the valence
compressive strain is resulting in a shift from M Γ to the K bands and the compressive strain on both bands is less
point in NSb. This shifting behavior is mainly due to changes significant. However, this interplay of changes in valence and
14951 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 24. Absorption spectra and joint density of states of monolayer NSb under biaxial (a)-(e) tensile and (f)-(j) compressive strains,
respectively. The vertical dotted lines are the most important electronic transitions with maximum relative excitonic intensity, while the dashed
lines are the direct quasi-particle gaps.

Figure 25. (a) First exciton energies and (b) excitonic binding energies in biaxially strained NAs and NSb monolayers.

conduction bands is the leading cause of shifting the excitonic (∼40 pm/V) compared to the zero-strain conditions,
weights of the first exciton. respectively. In addition, we also see some adversity in which
Variations in nonlinear SHG coefficients in the presence of the tensile strain degrades the brightness of the most
the aforementioned strains are shown in Figures 30 and 31 for prominent peak, which slows down from 3.2 eV (zero strain)
NAs and NSb, respectively. We see that the influence of a to about 2.7 eV (+10% strain). In the case of compressive
tensile strain is to again redshift the SHG spectra in both cases, strain, as shown in Figure 30b, we find that the most
while the compressive strain tends to blueshift. This behavior is prominent peak in the region 3.0−4.0 eV reduces from −8 to
similar to the corresponding linear spectra shown in Figures 23 −10% which could possibly be because of the change in
and 24. In the case of NAs, we find from Figure 30a that at position of the optical transitions in the BZ (see Figure 26)
these 810 and 1560 nm loss-less windows, a strain to +10% above). The spectra of NSb show more dramatic results. We
improves |χ(2)aab(ω)| to about 88% (∼105 pm/V) and 135% see from Figure 31a that the strain to +10% improves |χ(2)
aab(ω)|

14952 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 26. Excitonic weights showing the most important electronic transitions along the high symmetry points of the BZ of monolayer NAs under
biaxial (a)−(e) tensile and (f), (g) compressive strains respectively. The horizontal lines indicate zero strain band edges.

Figure 27. Excitonic weights showing the most important electronic transitions along the high symmetry points of the BZ of monolayer NSb under
biaxial (a)−(e) tensile and (f), (g) compressive strains respectively. The horizontal lines indicate zero strain band edges.

to about 193 pm/V (∼615%) and 79 pm/V (∼155%) at 810 TMDCs.131 The influence of the strain on the absorption
and 1560 nm, respectively. The results shown here, particularly spectra, the extinction coefficient k, and the refractive index n is
in the lossless windows, seem to be better than most traditional shown in Figures S13 and S14.
14953 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 28. Ground-state (gray) and G0W0 (red) band structure of monolayer NAs under various biaxial (a)−(e) tensile and (f), (g) compressive
strains respectively.

Figure 29. Ground-state (gray) and G0W0 (red) band structure of monolayer NSb under various biaxial (a)−(e) tensile and (f), (g) compressive
strains respectively.

We finally note that all the important resonances discussed avoided the effects from lattice thermal expansion (LTE). It is
here are 2ω resonances based on the qualitative arguments. A well-known that the band-gaps also depend on the LTE,133
more general theory on Sipe’s117 sum-over states might shine however, for most semiconductors, the effects from LTE are
important guidelines on excitonic transitions. Moreover, the minor toward the electronic eigenvalues,63 and therefore, they
discussion of the temperature-dependent spectra in this work have been excluded.
14954 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 30. Improvements in second harmonic coefficients in monolayer NAs under biaxial (a) tensile and (b) compressive strains respectively. The
dotted and dashed lines are the two important wavelengths 1560 and 810 nm, respectively.

Figure 31. Improvements in second harmonic coefficients in monolayer NSb under biaxial (a) tensile and (b) compressive strains, respectively.
The dotted and dashed lines are the two important wavelengths 1560 and 810 nm, respectively.

■ CONCLUSIONS
We present a detailed study of the fundamental properties of
resilient to the temperature variations between 0 and 450 K
(with the dominant bound excitonic peak around 2.8 eV only
two new monolayer materials, nitrogen arsenide NAs, and slightly red-shifting). In NAs, however, we showed stronger
nitrogen antimony NSb in their β-phases. We used a blue-shifting and broadening of spectra region around
combination of advanced many-body techniques to study dominant bound excitonic 3.4 eV peak. The other important
electron−electron, electron−hole, electron−phonon, and even NAs spectra region around dominant 5.4 eV resonant exciton
exciton−phonon interactions, leading to unique descriptions
was almost flattened by higher temperatures. The maximum
and deciphering of extraordinary properties of materials under
investigation. We provided deep insights into the effects of nonlinear second harmonic optical coefficient is large: 636
temperature and biaxial strain on the electronic structure, pm/V at 3.2 eV (387.5 nm) for NAs and 270 pm/V at 2.6 eV
accurate optical absorption spectra, and nonlinear optics. (476.9 nm) for NSb. We further proved that even a small
Both wide-gap semiconductors, NAs and NSb (indirect tensile biaxial strain stabilizes considered materials with no loss
quasiparticle gap ∼4.5 and ∼3.7 eV, respectively), embody in absorbance spectra; therefore, the practical preparation of
strong exciton binding energies, ∼1.6 and ∼1.3 eV,
these materials on support or included in heterostructures is
respectively. While we observed giant zero-point exciton
binding energy renormalization of 430 meV in NSb (red- possible. Generally, NAs and NSb seem as promising
shift), we detected almost no renormalization in NAs. In semiconductors with potential for use in optoelectronic
opposite, finite-temperature optical absorption in NSb is quite applications.
14955 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C


pubs.acs.org/JPCC Article

ASSOCIATED CONTENT (4) Karlický, F.; Turoň, J. Fluorographane C2FH: Stable and wide
* Supporting Information
sı band gap insulator with huge excitonic effect. Carbon 2018, 135,
134−144.
The Supporting Information is available free of charge at (5) Arnaud, B.; Lebègue, S.; Rabiller, P.; Alouani, M. Huge Excitonic
https://pubs.acs.org/doi/10.1021/acs.jpcc.2c03708. Effects in Layered Hexagonal Boron Nitride. Phys. Rev. Lett. 2006, 96,
Kinetic energy cutoff convergences, bare electronic 026402.
orbital dispersion, single-shot GW band gap conver- (6) Ramasubramaniam, A. Large excitonic effects in monolayers of
gence plots, absorption spectra convergence plots, molybdenum and tungsten dichalcogenides. Phys. Rev. B 2012, 86,
spectral function plots, zero Kelvin electronic dispersion, 115409.
strain effect on phonons, and strain effects on refractive (7) Wang, X.; Jones, A. M.; Seyler, K. L.; Tran, V.; Jia, Y.; Zhao, H.;
index, extinction coefficient and absorbance (PDF) Wang, H.; Yang, L.; Xu, X.; Xia, F. Highly anisotropic and robust
visualization of phonon modes of both materials (ZIP) excitons in monolayer black phosphorus. Nat. Nanotechnol. 2015, 10,


517.
(8) Palummo, M.; Bernardi, M.; Grossman, J. C. Exciton Radiative
AUTHOR INFORMATION Lifetimes in Two-Dimensional Transition Metal Dichalcogenides.
Corresponding Authors Nano Lett. 2015, 15, 2794.
František Karlický − Department of Physics, Faculty of (9) Echeverry, J. P.; Urbaszek, B.; Amand, T.; Marie, X.; Gerber, I.
Science, University of Ostrava, 701 03 Ostrava, Czech C. Splitting between bright and dark excitons in transition metal
Republic; orcid.org/0000-0002-2623-3991; dichalcogenide monolayers. Phys. Rev. B 2016, 93, 121107R.
Email: frantisek.karlicky@osu.cz (10) Wang, G.; Chernikov, A.; Glazov, M. M.; Heinz, T. F.; Marie,
Sitangshu Bhattacharya − Electronic Structure Theory Group, X.; Amand, T.; Urbaszek, B. Colloquium: Excitons in atomically thin
transition metal dichalcogenides. Phys. Rev. B 2018, 90, 021001.
Department of Electronics and Communication Engineering,
(11) Biswas, T.; Singh, A. K. Excitonic effects in absorption spectra
Indian Institute of Information Technology, Allahabad, Uttar of carbon dioxide reduction photocatalysts. Npj Comput. Mater. 2021,
Pradesh 211015, India; orcid.org/0000-0002-4442- 7, 189.
5225; Email: sitangshu@iiita.ac.in (12) Rasmussen, A.; Deilmann, T.; Thygesen, K. S. Towards fully
Authors automatized GW band structure calculations: What we can learn from
60.000 self-energy evaluations. Npj Comput. Mater. 2021, 7, 22.
Miroslav Kolos − Department of Physics, Faculty of Science, (13) Dubecký, M.; Karlický, F.; Minárik, S.; Mitas, L. Fundamental
University of Ostrava, 701 03 Ostrava, Czech Republic; gap of fluorographene by many-body GW and fixed-node diffusion
orcid.org/0000-0003-3821-7814 Monte Carlo methods. J. Chem. Phys. 2020, 153, 184706.
Rekha Verma − Nanoscale Electro-thermal Laboratory, Indian (14) Shen, Y. R. Surface properties probed by second-harmonic and
Institute of Information Technology, Allahabad, Uttar sum-frequency generation. Nature 1989, 337, 519.
Pradesh 211015, India; orcid.org/0000-0003-1636-2092 (15) Boyd, R. W. Nonlinear Opt., 3rd ed.; Academic Press: New
Complete contact information is available at: York, 2008.
https://pubs.acs.org/10.1021/acs.jpcc.2c03708 (16) Runge, E.; Gross, E. K. U. Density-Functional Theory for
Time-Dependent Systems. Phys. Rev. Lett. 1984, 52, 997.
(17) Ketolainen, T.; Macháčová, N.; Karlický, F. Optical Gaps and
Author Contributions
Excitonic Properties of 2D Materials by Hybrid Time-Dependent
The manuscript was written through the contributions of all
Density Functional Theory: Evidences for Monolayers and Prospects
authors. All authors have given their approval to the final for van der Waals Heterostructures. J. Chem. Theory Comput. 2020,
version of the manuscript. 16, 5876−5883.
Notes (18) Attaccalite, C.; Grüning, M.; Marini, A. Real-time approach to
The authors declare no competing financial interest. the optical properties of solids and nanostructures: Time-dependent

■ ACKNOWLEDGMENTS
The present work is carried out under the India-Czech
Bethe-Salpeter equation. Phys. Rev. B 2011, 84, 245110.
(19) King-Smith, R. D.; Vanderbilt, D. Theory of polarization of
crystalline solids. Phys. Rev. B 1993, 47, 1651.
(20) Attaccalite, C.; Grüning, M. Nonlinear optics from an ab-initio
Bilateral Scientific and Technological Cooperation (Indian No.
approach by means of the dynamical Berry phase: Application to
DST/INT/Czech/P-14/2019, Czech InterAction program No.
second- and third-harmonic generation in semiconductors. Phys. Rev.
LTAIN19138). Part of the work is also supported under B 2013, 88, 235113.
Science and Engineering Research Board (SERB)-MATRICS (21) Marini, A. Ab initio finite-temperature excitons. Phys. Rev. Lett.
MTR/2021/000017, India. Portion of the calculations were 2008, 101, 106405.
performed at IT4 Innovations National Supercomputing (22) Molina-Sànchez, A.; Palummo, M.; Marini, A.; Wirtz, L.
Center through the e-INFRA CZ (ID:90140). The authors Temperature-dependent excitonic effects in the optical properties of
acknowledge the Central Computation Facility at Indian single-layer MoS2. Phys. Rev. B 2016, 93, 155435.
Institute of Information Technology-Allahabad for all the (23) Mishra, H.; Bose, A.; Dhar, A.; Bhattacharya, S. Exciton-
major computations. phonon coupling and band-gap renormalization in monolayer WSe2.

■ REFERENCES
(1) Aryasetiawan, F.; Gunnarsson, O. The GW method. Rep. Prog.
Phys. Rev. B 2018, 98, 045143.
(24) Cannuccia, E.; Monserrat, B.; Attaccalite, C. Theory of
phonon-assisted luminescence in solids: application to hexagonal
Phys. 1998, 61, 237. boron nitride. Phys. Rev. B 2019, 99, 081109.
(2) Rohlfing, M.; Louie, S. G. Electron-hole excitations and optical (25) Shen, T.; Zhang, X.-W.; Shang, H.; Zhang, M.-Y.; Wang, X.;
spectra from first principles. Phys. Rev. B 2000, 62, 4927. Wang, E.-G.; Jiang, H.; Li, X.-Z. Influence of high-energy local orbitals
(3) Onida, G.; Reining, L.; Rubio, A. Electronic excitations: density- and electron-phonon interactions on the band gaps and optical
functional versus many-body Green’s-function approaches. Rev. Mod. absorption spectra of hexagonal boron nitride. Phys. Rev. B 2020, 102,
Phys. 2002, 74, 601. 045117.

14956 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

(26) Mishra, H.; Bhattacharya, S. Giant exciton-phonon coupling (47) Derivaz, M.; Dentel, D.; Stephan, R.; Hanf, M.-C.; Mehdaoui,
and zero-point renormalization in hexagonal monolayer boron nitride. A.; Sonnet, P.; Pirri, C. Continuous Germanene Layer on Al(111).
Phys. Rev. B 2019, 99, 165201. Nano Lett. 2015, 15, 2510−2516.
(27) Kolos, M.; Cigarini, L.; Verma, R.; Karlický, F.; Bhattacharya, S. (48) Guo, S.; Zhang, Y.; Ge, Y.; Zhang, S.; Zeng, H.; Zhang, H. 2D
Giant Linear and Nonlinear Excitonic Responses in an Atomically V-V Binary Materials: Status and Challenges. Adv. Mater. 2019, 31,
Thin Indirect Semiconductor Nitrogen Phosphide. J. Phys. Chem. C 1902352.
2021, 125, 12738−12757. (49) Fan, H. Y. Temperature Dependence of the Energy Gap in
(28) Kawai, H.; Yamashita, K.; Cannuccia, E.; Marini, A. Electron- Monatomic Semiconductors. Phys. Rev. 1950, 78, 808.
electron and electron-phonon correlation effects on the finite- (50) Antončík, E. On the theory of temperature shift of the
temperature electronic and optical properties of zinc-blende GaN. absorption curve in non-polar crystals. Czech. J. Phys. 1955, 5, 449−
Phys. Rev. B 2014, 89, 085202. 461.
(29) Cannuccia, E.; Marini, A. Ab-initio study of the effects induced (51) Keffer, C.; Hayes, T. M.; Bienenstock, A. PbTe Debye-Waller
by the electron-phonon scattering in carbon based nanostructures. Factors and Band-Gap Temperature Dependence. Phys. Rev. Lett.
arXiv 2013, V. 1968, 21, 1676.
(30) Mueller, T.; Malic, E. Exciton physics and device application of (52) Walter, J. P.; Zucca, R. R. L.; Cohen, M. L.; Shen, Y. R.
two-dimensional transition metal dichalcogenide semiconductors. npj Temperature Dependence of the Wavelength-Modulation Spectra of
2D Mater. Appl. 2018, 2, 29. GaAs. Phys. Rev. Lett. 1970, 24, 102.
(31) Carvalho, A.; Ribeiro, R. M.; Castro Neto, A. H. Band nesting (53) Kasowski, R. V. Band Structure of NiS as Calculated Using a
and the optical response of two-dimensional semiconducting Simplified Linear-Combination-of-Muffin-Tin-Orbitals Method. Phys.
transition metal dichalcogenides. Phys. Rev. B 2013, 88, 115205. Rev. B 1973, 8, 1378.
(32) Mennel, L.; Smejkal, V.; Linhart, L.; Burgdörfer, J.; Libisch, F.; (54) Allen, P. B.; Heine, V. Theory of the temperature dependence
Mueller, T. Band nesting in two-dimensional crystals: an exceptionally of electronic band structures. J. Phys. C 1976, 9, 2305.
sensitive probe of strain. Nano Lett. 2020, 20, 4242. (55) Allen, P. B.; Cardona, M. Theory of the temperature
(33) Taheri, A.; Da Silva, C.; Amon, C. H. Phonon thermal transport dependence of the direct gap of germanium. Phys. Rev. B 1981, 23,
in β−NX (X = P, As, Sb) monolayers: A first-principles study of the 1495.
interplay between harmonic and anharmonic phonon properties. Phys. (56) Allen, P. B.; Cardona, M. Temperature dependence of the
Rev. B 2019, 99, 235425. direct gap of Si and Ge. Phys. Rev. B 1983, 27, 4760.
(34) Giannozzi, P.; Andreussi, O.; Brumme, T.; Bunau, O.; Nardelli, (57) Lautenschlager, P.; Allen, P. B.; Cardona, M. Temperature
M. B.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Cococcioni, dependence of band gaps in Si and Ge. Phys. Rev. B 1985, 31, 2163.
(58) Giustino, F.; Louie, S. G.; Cohen, M. L. Electron-phonon
M.; et al. Advanced capabilities for materials modelling with Quantum
renormalization of the direct band gap of diamond. Phys. Rev. Lett.
ESPRESSO. J. Phys.: Condens. Matter 2017, 29, 465901.
2010, 105, 265501.
(35) Hamann, D. Optimized norm-conserving Vanderbilt pseudo-
(59) Antonius, G.; Poncé, S.; Boulanger, P.; Côté, M.; Gonze, X.
potentials. Phys. Rev. B 2013, 88, 085117.
Many-body effects on the zero-point renormalization of the band
(36) Antonius, G.; Poncé, S.; Lantagne-Hurtubise, E.; Auclair, G.;
structure. Phys. Rev. Lett. 2014, 112, 215501.
Gonze, X.; Côté, M. Dynamical and anharmonic effects on the
(60) Cannuccia, E.; Marini, A. Zero point motion effect on the
electron-phonon coupling and the zero-point renormalization of the
electronic properties of diamond, trans-polyacetylene and poly-
electronic structure. Phys. Rev. B 2015, 92, 085137. ethylene. Eur. Phys.J. B 2012, 85, 320.
(37) Sangalli, D.; Ferretti, A.; Miranda, H.; Attaccalite, C.; Marri, I.; (61) Cannuccia, E.; Marini, A. Effect of the quantum zero-point
Cannuccia, E.; Melo, P.; Marsili, M.; Paleari, F.; Marrazzo, A.; et al. atomic motion on the optical and electronic properties of diamond
Many-body perturbation theory calculations using the yambo code. J. and trans-polyacetylene. Phys. Rev. lett. 2011, 107, 255501.
Phys.: Condens. Matter 2019, 31, 325902. (62) Qiao, J.; Kong, X.; Hu, Z.-X.; Yang, F.; Ji, W. High-mobility
(38) Godby, R. W.; Needs, R. J. Metal-insulator transition in Kohn- transport anisotropy and linear dichroism in few-layer black
Sham theory and quasiparticle theory. Phys. Rev. Lett. 1989, 62, 1169. phosphorus. Nat. Commun. 2014, 5, 4475.
(39) Kammerlander, D.; Botti, S.; Marques, M. A. L.; Marini, A.; (63) Villegas, C. E.; Rocha, A. R.; Marini, A. Anomalous temperature
Attaccalite, C. Speeding up the solution of the Bethe-Salpeter dependence of the band-gap in black phosphorus. Nano Lett. 2016,
equation by a double-grid method and Wannier interpolation. Phys. 16, 5095.
Rev. B 2012, 86, 125203. (64) Molina-Sánchez, A.; Sangalli, D.; Hummer, K.; Marini, A.;
(40) Cannuccia, E. Giant polaronic effects in polymers: breakdown Wirtz, L. Effect of spin-orbit interaction on the optical spectra of
of the quasiparticle picture. Ph.D. thesis, Rome Tor Vergata single-layer, double-layer, and bulk MoS2. Phys. Rev. B 2013, 88,
University, 2011. 045412.
(41) Crank, J.; Nicolson, P. A practical method for numerical (65) Rohlfing, M.; Krüger, P.; Pollmann, J. Efficient scheme for GW
evaluation of solutions of partial differential equations of the heat- quasiparticle band-structure calculations with applications to bulk Si
conduction type. Math. Proc. Camb. Philos. Soc. 1947, 43, 50. and to the Si(001)-(2 × 1) surface. Phys. Rev. B 1995, 52, 1905.
(42) Mera Acosta, C.; Fazzio, A.; Dalpian, G. M. Zeeman-type spin (66) Monserrat, B.; Needs, R. J. Comparing electron-phonon
splitting in nonmagnetic three-dimensional compounds. npj Quantum coupling strength in diamond, silicon, and silicon carbide: First-
Materials 2019, 4, 41. principles study. Phys. Rev. B 2014, 89, 214304.
(43) Yu, W.; Niu, C.-Y.; Zhu, Z.; Wang, X.; Zhang, W.-B. Atomically (67) Attaccalite, C.; Nguer, A.; Cannuccia, E.; Grüning, M. Strong
thin binary V−V compound semiconductor: a first-principles study. J. second harmonic generation in SiC, ZnO, GaN two-dimensional
Mater. Chem. C 2016, 4, 6581. hexagonal crystals from first-principles many-body calculations. Phys.
(44) Sohier, T. D. P.; Gibertini, M.; Calandra, M.; Mauri, F.; Chem. Chem. Phys. 2015, 17, 9533.
Marzari, N. Breakdown of optical phonons’ splitting in two- (68) Chrenko, R. M. Ultraviolet and infrared spectra of cubic boron
dimensional materials. Nano Lett. 2017, 17, 3758. nitride. Solid State Commun. 1974, 14, 511.
(45) Roome, N. J.; Carey, J. D. Beyond Graphene: Stable Elemental (69) Hybertsen, M. S.; Louie, S. G. Electron correlation in
Monolayers of Silicene and Germanene. ACS Appl. Mater. Interfaces semiconductors and insulators: Band gaps and quasiparticle energies.
2014, 6, 7743−7750. Phys. Rev. B 1986, 34, 5390.
(46) Chen, M. X.; Weinert, M. Revealing the Substrate Origin of the (70) Chaney, R. C.; Lafon, E. E.; Lin, C. C. Energy band structure of
Linear Dispersion of Silicene/Ag(111). Nano Lett. 2014, 14, 5189− lithium fluoride crystals by the method of tight binding. Phys. Rev. B
5193. 1971, 4, 2734.

14957 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

(71) Varshni, Y. Temperature dependence of the energy gap in (90) Selig, M.; Berghäuser, G.; Raja, A.; Nagler, P.; Schüller, C.;
semiconductors. Physica 1967, 34, 149−154. Heinz, T. F.; Korn, T.; Chernikov, A.; Malic, E.; Knorr, A. Excitonic
(72) Zhang, G.; Chaves, A.; Huang, S.; Wang, F.; Xing, Q.; Low, T.; linewidth and coherence lifetime in monolayer transition metal
Yan, H. Determination of layer-dependent exciton binding energies in dichalcogenides. Nat. Commun. 2016, 7, 13279.
few-layer black phosphorus. Sci. Adv. 2018, 4, 9977. (91) Zhang, Y.; Chang, T.-R.; Zhou, B.; Cui, Y.-T.; Yan, H.; Liu, Z.;
(73) He, K.; Kumar, N.; Zhao, L.; Wang, Z.; Mak, K. F.; Zhao, H.; Schmitt, F.; Lee, J.; Moore, R.; Chen, Y.; et al. Direct observation of
Shan, J. Tightly Bound excitons in monolayer WSe2. Phys. Rev. Lett. the transition from indirect to direct bandgap in atomically thin
2014, 113, 026803. epitaxial MoSe2. Nat. Nanotechnol. 2014, 9, 111.
(74) Arora, A.; Koperski, M.; Nogajewski, K.; Marcus, J.; Faugerasa, (92) Pacuski, W.; Grzeszczyk, M.; Nogajewski, K.; Bogucki, A.;
C.; Potemski, M. Excitonic resonances in thin films of WSe2: from Oreszczuk, K.; Kucharek, J.; Połczyńska, K. E.; Seredyński, B.; Rodek,
monolayer to bulk material. Nanoscale 2015, 7, 10421. A.; Bożek, R.; et al. Narrow excitonic lines and large-scale
(75) Autere, A.; Jussila, H.; Marini, A.; Saavedra, J. R. M.; Dai, Y.; homogeneity of transition-metal dichalcogenide monolayers grown
Säynätjoki, A.; Karvonen, L.; Yang, H.; Amirsolaimani, B.; Norwood, by molecular beam epitaxy on hexagonal boron nitride. Nano Lett.
R. A.; et al. Optical harmonic generation in monolayer group-VI 2020, 20, 3058.
transition metal dichalcogenides. Phys. Rev. B 2018, 98, 115426. (93) Le, C. T.; Clark, D. J.; Ullah, F.; Senthilkumar, V.; Jang, J. I.;
(76) Park, S.; Mutz, N.; Schultz, T.; Blumstengel, S.; Han, A.; Aljarb, Sim, Y.; Seong, M.-J.; Chung, K.-H.; Park, H.; Kim, Y. S. Nonlinear
A.; Li, L.-J.; List-Kratochvil, E. J. W.; Amsalem, P.; et al. Direct optical characteristics of monolayer MoSe2. Ann. Phys. 2016, 528, 551.
determination of monolayer MoS2 and WSe2 exciton binding energies (94) Robert, C.; Picard, R.; Lagarde, D.; Wang, G.; Echeverry, J. P.;
on insulating and metallic substrate. 2D Mater. 2018, 5, 025003. Cadiz, F.; Renucci, P.; Högele, A.; Amand, T.; Marie, X.; et al.
(77) Hanbicki, A.T.; Currie, M.; Kioseoglou, G.; Friedman, A.L.; Excitonic properties of semiconducting monolayer and bilayer
Jonker, B.T. Measurement of high exciton binding energy in the MoTe2. Phys. Rev. B 2016, 94, 155425.
monolayer transition-metal dichalcogenides WS2and WSe2. Solid State (95) Janisch, C.; Wang, Y.; Ma, D.; Mehta, N.; Elías, A. L.; Perea-
Commun. 2015, 203, 16. López, N.; Terrones, M.; Crespi, V.; Liu, Z. Extraordinary second
(78) Cadiz, F.; Courtade, E.; Robert, C.; Wang, G.; Shen, Y.; Cai, harmonic generation in tungsten disulfide monolayers. Sci. Rep. 2015,
H.; Taniguchi, T.; Watanabe, K.; Carrere, H.; Lagarde, D.; et al. 4, 5530.
Excitonic linewidth approaching the homogeneous limit in MoS2- (96) Stier, A. V.; M, K.; McCreary; Jonker, B. T.; Kono, J.; Crooker,
based van der Waals heterostructures. Phys. Rev. X 2017, 7, 021026. S. A. Exciton diamagnetic shifts and valley Zeeman effects in
(79) Miyauchi, Y.; Konabe, S.; Wang, F.; Zhang, W.; Hwang, A.; monolayer WS2 and MoS2 to 65 T. Nat. Commun. 2016, 7, 10643.
Hasegawa, Y.; Zhou, L.; Mouri, S.; Toh, M.; Eda, G.; Matsuda, K. (97) Zhu, B.; Chen, X.; Cui, X. Exciton binding energy of monolayer
Evidence for line width and carrier screening effects on excitonic WS2. Sci. Rep. 2015, 5, 9281.
valley relaxation in 2D semiconductors. Nat. Commun. 2018, 9, 2598. (98) Guo, L.; Chen, C.-A.; Zhang, Z.; Monahan, D. M.; Lee, Y.-H.;
(80) Ribeiro-Soares, J.; Janisch, C.; Liu, Z.; Elías, A. L.; Dresselhaus, Fleming, G. R. Lineshape characterization of excitons in monolayer
M. S.; Terrones, M.; Cançado, L. G.; Jorio, A. Second harmonic WS2 by two-dimensional electronic spectroscopy. Nanoscale Adv.
generation in WSe2. 2D Mater. 2015, 2, 045015. 2020, 2, 2333.
(81) Rigosi, A. F.; Hill, H. M.; Rim, K. T.; Flynn, G. W.; Heinz, T. F. (99) Hamer, M. J.; Zultak, J.; Tyurnina, A. V.; Zólyomi, V.; Terry,
Electronic band gaps and exciton binding energies in monolayer D.; Barinov, A.; Garner, A.; Donoghue, J.; Rooney, A. P.; Kandyba, V.;
MoxW1‑xS2 transition metal dichalcogenide alloys probed by scanning et al. Indirect to direct gap crossover in two-dimensional InSe
tunneling and optical spectroscopy. Phys. Rev. B 2016, 94, 075440. revealed by angle-resolved photoemission spectroscopy. ACS Nano
(82) Li, Y.; Rao, Y.; Mak, K. F.; You, Y.; Wang, S.; Dean, C. R.; 2019, 13, 2136.
Heinz, T. F. Probing symmetry properties of few layer MoS2 and h- (100) Leisgang, N.; Roch, J. G.; Froehlicher, G.; Hamer, M.; Terry,
BN by optical second-harmonic generation. Nano Lett. 2013, 13, D.; Gorbachev, R.; Warburton, R. J. Optical second harmonic
3329. generation in encapsulated single-layer InSe editors-pick. AIP Adv.
(83) Kumar, N.; Najmaei, S.; Cui, Q.; Ceballos, F.; Ajayan, P. M.; 2018, 8, 105120.
Lou, J.; Zhao, H. Second harmonic microscopy of monolayer MoS2. (101) Zhou, X.; Cheng, J.; Zhou, Y.; Cao, T.; Hong, H.; Liao, Z.;
Phys. Rev. B 2013, 87, 161403R. Wu, S.; Peng, H.; Liu, K.; Yu, D. Strong second-harmonic generation
(84) Clark, D. J.; Senthilkumar, V.; Le, C. T.; Weerawarne, D. L.; in atomic layered GaSe. J. Am. Chem. Soc. 2015, 137, 7994.
Shim, B.; Jang, J. I.; Shim, J. H.; Cho, J.; Sim, Y.; Seong, M.-J.; et al. (102) Elias, C.; Valvin, P.; Pelini, T.; Summerfield, A.; Mellor, C.;
Strong optical nonlinearity of CVD-grown MoS2 monolayer as probed Cheng, T.; Eaves, L.; Foxon, C.; Beton, P.; Novikov, S.; et al. Direct
by wavelength-dependent second-harmonic generation. Phys. Rev.B band-gap crossover in epitaxial monolayer boron nitride. Nat.
2014, 90, 121409R. Commun. 2019, 10, 2639.
(85) Malard, L. M.; Alencar, T. V.; Barboza, A. P. M.; Mak, K. F.; de (103) Kim, S.; Fröch, J. E.; Gardner, A.; Li, C.; Aharonovich, I.;
Paula, A. M. Observation of intense second harmonic generation from Solntsev, A. S. Second-harmonic generation in multilayer hexagonal
MoS2 atomic crystals. Phys. Rev. B 2013, 87, 201401R. boron nitride flakes. Opt. Lett. 2019, 44, 5792.
(86) Hill, H. M.; Rigosi, A. F.; Roquelet, C.; Chernikov, A.; (104) Mahan, G. D. Many-Particle Physics, 3rd ed.; Springer
Berkelbach, T. C.; Reichman, D. R.; Hybertsen, M. S.; Brus, L. E.; International ed.: New York, 2014.
Heinz, T. F. Observation of excitonic Rydberg states in monolayer (105) Grüning, M.; Marini, A.; Gonze, X. Exciton-plasmon States in
MoS2and WS2 by photoluminescence excitation spectroscopy. Nano nanoscale materials: breakdown of the Tamm-Dancoff approximation.
Lett. 2015, 15, 2992. Nano Lett. 2009, 9, 2820−2824.
(87) Zhang, C.; Johnson, A.; Hsu, C.-L.; Li, L.-J.; Shih, C.-K. Direct (106) Moody, G.; Dass, C. K.; Hao, K.; Chen, C.-H.; Li, L.-J.; Singh,
imaging of band profile in single layer MoS2 on Graphite: A.; Tran, K.; Clark, G.; Xu, X.; Berghaüser, G.; et al. Intrinsic
quasiparticle energy gap, metallic edge states, and edge band bending. homogeneous linewidth and broadening mechanisms of excitons in
Nano Lett. 2014, 14, 2443. monolayer transition metal dichalcogenides. Nat. Commun. 2015, 6,
(88) Chiu, M.-H.; Zhang, C.; Shiu, H.-W.; Chuu, C.-P.; Chen, C.- 8315.
H.; Chang, C.-Y. S.; Chen, C.-H.; Chou, M.-Y.; Shih, C.-K.; Li, L.-J. (107) Dey, P.; Paul, J.; Wang, Z.; Stevens, C. E.; Liu, C.; Romero, A.
Determination of band alignment in the single-layer MoS2/WSe2 H.; Shan, J.; Hilton, D. J.; Karaiskaj, D. Optical coherence in atomic-
heterojunction. Nat. Commun. 2015, 6, 7666. monolayer transition-metal dichalcogenides limited by electron-
(89) Arora, A.; Nogajewski, K.; Molas, M.; Koperskiab, M.; phonon interactions. Phys. Rev. Lett. 2016, 116, 127402.
Potemski, M. Exciton band structure in layered MoSe2: from a (108) Hsu, W.-T.; Zhao, Z.-A.; Li, L.-J.; Chen, C.-H.; Chiu, M.-H.;
monolayer to the bulk limit. Nanoscale 2015, 7, 20769. Chang, P.-S.; Chou, Y.-C.; Chang, W.-H. Second Harmonic

14958 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Generation from Artificially Stacked Transition Metal Dichalcogenide (130) Khatibi, Z.; Feierabend, M.; Selig, M.; Brem, S.; Linderälv, C.;
Twisted Bilayers. ACS Nano 2014, 8, 2951−2958. Erhart, P.; Malic, E. Impact of strain on the excitonic linewidth in
(109) Attaccalite, C.; Palummo, M.; Cannuccia, E.; Grüning, M. transition metal dichalcogenides. 2D Mater. 2019, 6, 015015.
Second-harmonic generation in single-layer monochalcogenides: A (131) Beach, K.; Lucking, M. C.; Terrones, H. Strain dependence of
response from first-principles real-time simulations. Phys. Rev. Mat second harmonic generation in transition metal dichalcogenide
2019, 3, 074003. monolayers and the fine structure of the C exciton. Phys. Rev. B
(110) Grüning, M.; Sangalli, D.; Attaccalite, C. Dielectrics in a time- 2020, 101, 155431.
dependent electric field: A real-time approach based on density- (132) Bertolazzi, S.; Brivio, J.; Kis, A. Stretching and breaking of
polarization functional theory. Phys. Rev. B 2016, 94, 035149. ultrathin MoS2. ACS Nano 2011, 5, 9703.
(111) Souza, I.; Iniguez, J. I.; Vanderbilt, D. Dynamics of Berry- (133) Mounet, N.; Marzari, N. First-principles determination of the
phase polarization in time-dependent electric fields. Phys. Rev. B 2004, structural, vibrational and thermodynamic properties of diamond,
graphite, and derivatives. Phys. Rev. B 2005, 71, 205214.
69, 085106.
(112) Resta, R. Macroscopic polarization in crystalline dielectrics:
the geometric phase approach. Rev. Mod. Phys. 1994, 66, 899.
(113) Attaccalite, C.; Cannuccia, E.; Grüning, M. Excitonic effects in
third-harmonic generation: The case of carbon nanotubes and
nanoribbons. Phys. Rev. B 2017, 95, 125403.
(114) Aversa, C.; Sipe, J. E. Nonlinear optical susceptibilities of
semiconductors: Results with a length-gauge analysis. Phys. Rev. B
1995, 52, 14636.
(115) Moss, D. J.; Ghahramani, E.; Sipe, J. E.; van Driel, H. M.
Band-structure calculation of dispersion and anisotropy in χ →(3) for
third-harmonic generation in Si, Ge, and GaAs. Phys. Rev. B 1990, 41,
1542.
(116) Hughes, J. L. P.; Sipe, J. E. Calculation of second-order optical
response in semiconductors. Phys. Rev. B 1996, 53, 10751.
(117) Sipe, J. E.; Shkrebtii, A. I. Second-order optical response in
semiconductors. Phys. Rev. B 2000, 61, 5337.
(118) Virk, K. S.; Sipe, J. E. Semiconductor optics in length gauge: A
general numerical approach. Phys. Rev. B 2007, 76, 035213.
(119) Sipe, J. E.; Ghahramani, E. Nonlinear optical response of
semiconductors in the independent-particle approximation. Phys. Rev.
B 1993, 48, 11705.
(120) Takimoto, Y.; Vila, F. D.; Rehri, J. J. Real-time time-
dependent density functional theory approach for frequency-depend-
ent nonlinear optical response in photonic molecules. J. Chem. Phys. Recommended by ACS
2007, 127, 154114.
(121) Narsimha Rao, E.; Vaitheeswaran, G.; Reshak, A. H.; Auluck, Lone Pair Electrons with Weak Nuclear Binding Inducing
S. Role of spin-orbit interaction on the nonlinear optical response of Sensitive Nonlinear Optical Responses in Phosphorus
CsPbCO3F using DFT. Phys. hem. Chem. Phys. 2017, 19, 31255. Clusters
(122) Peng, Z.; Chen, X.; Fan, Y.; Srolovitz, D. J.; Le, D. Strain
Quanjie Zhong.
engineering of 2D semiconductors and graphene: from strain fields to
JULY 07, 2023
band-structure tuning and photonic applications. Nat. Light Sci. Appl. THE JOURNAL OF PHYSICAL CHEMISTRY LETTERS READ
2020, 9, 1.
(123) Aslan, O. B.; Datye, I. M.; Mleczko, M. J.; Cheung, K. S.;
Full Temporal Characterization of Ultrabroadband Few-
Krylyuk, S.; Bruma, A.; Kalish, I.; Davydov, A. V.; Pop, E.; Heinz, T.
Cycle Laser Pulses Using Atomically Thin WS2
F. Probing the optical properties and strain-tuning of ultrathin
Mo1−xWxTe2. Nano Lett. 2018, 18, 2485. Óscar Pérez-Benito, Rosa Weigand, et al.
(124) Aslan, O. B.; Deng, M.; Heinz, T. F. Strain tuning of excitons JUNE 15, 2023
ACS PHOTONICS READ
in monolayer WSe2. Phys. Rev. B 2018, 98, 115308.
(125) Ahn, G. H.; Amani, M.; Rasool, H.; Lien, D.-H.; Mastandrea,
J. P.; J, W. A., III; Dubey, M.; Chrzan, D. C.; Minor, A. M.; Javey, A. Strain Engineered Electrically Pumped SiGeSn Microring
Strain-engineered growth of two-dimensional materials. Nat. Lasers on Si
Commun. 2017, 8, 608. Bahareh Marzban, Dan Buca, et al.
(126) Frisenda, R.; Drüppel, M.; Schmidt, R.; de Vasconcellos, S. DECEMBER 16, 2022
M.; de Lara, D. P.; Bratschitsch, R.; Rohlfing, M.; Castellanos-Gomez, ACS PHOTONICS READ
A. Biaxial strain tuning of the optical properties of single-layer
transition metal dichalcogenides. npj 2D Mater. Appl. 2017, 10, 1. Ultrafast Thermionic Electron Injection Effects on Exciton
(127) Villegas, C. E. P.; Rocha, A. R. Elucidating the optical Formation Dynamics at a van der Waals
properties of novel heterolayered materials based on MoTe2 InN for Semiconductor/Metal Interface
photovoltaic applications. J. Phys. Chem. C 2015, 119, 11886. Kilian R. Keller, Nicolò Maccaferri, et al.
(128) Rhim, S. H.; Kim, Y. S.; Freeman, A. J. Strain-induced giant JULY 20, 2022
second-harmonic generation in monolayered 2H-MoX2 (X = S, Se, ACS PHOTONICS READ
Te). Appl. Phys. Lett. 2015, 107, 241908.
(129) Zollner, K.; Junior, P. E. F.; Fabian, J. Strain-tunable orbital,
Get More Suggestions >
spin-orbit, and optical properties of monolayer transition-metal
dichalcogenides. Phys. Rev. B 2019, 100, 195126.

14959 https://doi.org/10.1021/acs.jpcc.2c03708
J. Phys. Chem. C 2022, 126, 14931−14959

You might also like