Propriedades Excitônicas de Camadas de Fosforeno Empilhadas

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

pubs.acs.

org/JPCC Article

Theoretical Exploration of Structural and Excitonic Properties in


Black Phosphorus: From First-Principles to a Semi-Empirical
Approach
Diego Guedes-Sobrinho,* Celso R. Caldeira Rêgo,* Gabriel Reynald Da Silva, Henrique R. Da Silva,
W. Wenzel, Maurício J. Piotrowski,* and Alexandre Cavalheiro Dias*
Cite This: https://doi.org/10.1021/acs.jpcc.3c08414 Read Online
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
Downloaded via UNIV FED DO PARANA on April 24, 2024 at 14:24:41 (UTC).

ABSTRACT: Black phosphorus serves as an exemplary stacked bidimensional semiconductor,


exhibiting anisotropic features in electronic and optical properties that demand special attention
in theoretical investigations. Herein, we employed a series of computational protocols, starting
with first-principles approaches (particularly density functional theory�DFT), combined with
the solution of the Bethe−Salpeter equation within the tight-binding method to explore the
structural stability and optoelectronic properties (bandgap, exciton binding energies, and optical
absorption) of black phosphorus (Pn) across layers ranging from n = 1 to 6 and bulk. In our DFT
investigations, we observed that empirical and semiempirical van der Waals models, contributing
a dispersion energy component, revealed a myriad of differences and similarities in properties,
such as interlayer nonbonded interactions. Notably, the many-body dispersion correction
exhibited superior performance in connecting layered systems with the bulk. The magnitude of
dispersion energies correlated with the stability during the aggregation process P(n−1) + P1 → Pn.
Additionally, the bandgap, properly corrected through relativistic quasi-particle calculations,
narrowed due to enhanced interlayer wave function overlap, a result of the dispersion energies promoting the shortening of interlayer
distances. Subsequently, we utilized the band structure relativistically corrected as a starting point to obtain the Hamiltonian,
achieved through the generation of maximally localized Wannier functions. This facilitated a screening of the electron−hole (e−h)
pairwise interaction Coulomb potential, specifically the exciton binding energy. We identified an indirect impact of the dispersion
energies on excitonic properties, which were effectively described by the Rytova−Keldysh model for the e−h Coulomb potential,
aligning well with photoluminescence experiments.

1. INTRODUCTION approximately 2.0 eV for monolayers to 0.30−0.40 eV for the


Since 2014, when it was first proposed as a channel material for bulk.14−17 Simultaneously, the carrier mobility undergoes
field-effect transistor (FET) applications by Li et al.,1 black significant changes, varying within the interval of 1000−
phosphorus (BP) has captured the attention of the two- 10,000 cm2/V·s from monolayer to bulk.7,8,18 This underscores
dimensional (2D) semiconductor materials community.2−7 the challenge of spatial confinement effects on optoelectronic
Arranged in a puckered honeycomb lattice (orthorhombic, features, particularly the exciton energy spectrum, where
space group Cmca), phosphorus atoms introduce in-plane information regarding electron−hole (e−h) interactions is
anisotropy across electrical, optical, thermal, and mechanical captured.19−21 In this context, computational design based on
characteristics. This distinctive profile positions BP not only as ab initio calculations, particularly through density functional
a promising material for FET applications8,9 but also for a theory (DFT), has been instrumental in providing crucial
myriad of electronic and photonic applications.10−12 However, insights into the electronic−atomistic interface.22−26 However,
due to its nature as multilayers weakly interacting through van standard DFT studies employed to investigate BP systems face
der Waals (vdW) forces�similar to graphene and achievable certain limitations of the typical exchange−correlation (xc)
through mechanical exfoliation of bulk material13 �the functionals used in crystal solids. These limitations include: (i)
computational design of bandgap and excitonic properties,
achieved by controlling the layer number, necessitates a precise
description of interlayer interactions. Additionally, the impact Received: December 26, 2023
of the chosen theoretical approach on predicting BP properties Revised: March 17, 2024
should be carefully considered. Accepted: April 1, 2024
Experimental findings have consistently reported a sizable
dependence of the BP bandgap on layer thickness. This
dependence exhibits an exponential decrease, ranging from

© XXXX American Chemical Society https://doi.org/10.1021/acs.jpcc.3c08414


A J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

the underestimation of bandgap energies27,28 and (ii) the 2. THEORETICAL APPROACH AND COMPUTATIONAL
inadequate description of vdW interactions among BP DETAILS
interlayers, currently inserted via the EKS+vdW = EKS + Edisp Our sequential calculation protocol, depicted in Figure 1,
relationship,29,30 wherein EKS is the standard Kohn−Sham follows a systematic strategy:
(KS) energy and Edisp is the vdW dispersive energy obtained a
posteriori through vdW correction models.
The Edisp is accepted as an improvement of the standard
DFT approach when just the EKS is considered. At this context,
several vdW models have been proposed to Edisp calculation,
such as the most known empirical class proposed by
Grimme,31−33 which is based on the addition of the
coordination number dependence (named D3)34 as well as
its version by replacing the damping function by the Becke−
Johnson (BJ) model (D3BJ).35 A representative semiempirical
models are based on the set of Tkatchenko−Scheffler (TS)
methods, composed by the simplest TS model as dependent
on the charge density (but considering some parametrized
terms),36 TS by including self-consistent screening (named by
TSSCS),37,38 TS model based on iterative Hirshfeld partition-
ing (TSHP),39,40 and the many body dispersion (MBD).37
Another semiempirical method is the Steinmann approach41,42
also density-dependent (named by dDsC) from the general-
ized gradient approximation through dipole moment calcu-
lation.
While the BP excitonic properties in mono- and few-layer
systems hold crucial significance for optoelectronic device
applications, understanding the influence of a substrate or the
effects of layer stacking on exciton binding energy calculations
remains inconclusive. For example, quasi-particle (QP)
calculations within the GW approach have unveiled the
suppression of BP binding energy (Eb) from n = 1 → 4 layers
Figure 1. Sequential protocol for calculating excitonic properties in
due to the environmental screening effects of hexagonal boron BP (Pn) systems (n = 1, 2, 3, 4, 5, 6, and ∞). DFT calculations were
nitride (hBN) and sapphire.43 Conversely, even with the state- carried out using the VASP code, followed by the utilization of
of-the-art GW method, known for capturing many-body Wannier90 to derive the TB Hamiltonian. This Hamiltonian was then
effects, its prohibitive computational cost limits systematic incorporated into the WanTiBEXOS code for the computation of
studies involving large systems. For BP monolayers as excitonic properties.
freestanding systems, the calculated Eb = 0.48 eV was found
to be underestimated compared to the experimentally reported • Ab initio geometry relaxation: perform ab initio
value of 0.9 ± 0.1 eV obtained via polarization-resolved calculations for geometry relaxations of multilayer BP
photoluminescence (PR−PL).16 Consequently, exploring systems, denoted as Pn, where n = 1, 2, 3, 4, 5, 6, and ∞
computationally more economical alternatives within DFT (bulk). Utilize several empirical and semiempirical vdW
approaches becomes crucial to understanding how several vdW corrections (Etot = EKS + Edisp), such as D3, D3BJ, dDsC,
methods operate in BP and their impact on structural, TS, TSSCS, TSHP, and MBD. The detailed descriptions
electronic, and optical properties. can be found in the Supporting Information.
In this study, we systematically investigate the impact of • Relativistic QP correction (DFT-1/2): optimize the cut
several empirical and semiempirical vdW methods on the function within the DFT-1/2 method to prevent
optoelectronic properties of BP (Pn), ranging from n = 1, 2, 3, interpenetration between neighboring atoms. Determine
4, 5, 6 up to bulk. The structures are optimized using vdW- the CUT parameter variationally for bandgap calcu-
based methods, and we employ the computationally efficient lations in mono- and bilayers, maximizing the gap energy
relativistic QP correction known as DFT-1/2,44 to calculate without empirical parameters.
bandgap energies within a single-particle theory. Utilizing the • DFT-1/2 bandgap calculations: perform DFT-1/2
DFT-1/2 bandgap as a starting point, we proceed to calculate calculations for bandgaps without excitonic effects for
excitonic properties through the solution of the Bethe− all Pn systems obtained from the optimized vdW-based
Salpeter equation (BSE) within a tight-binding (TB) approach. methods.
Our methodology involves employing the maximally localized • MLWF-TB: employ the MLWF-TB approach to
Wannier functions TB (MLWF-TB) approach to derive the calculate the orthogonal TB Hamiltonian, denoted as
orthogonal TB Hamiltonian. This Hamiltonian is then used for H(k). This Hamiltonian is obtained through the
a comprehensive screening of a set of 2D and 3D Coulomb Wannier90 package45 and serves as input for the
potentials, providing a detailed description of the exciton e−h WanTiBEXOS (W) code,46 which handles QP effects
interaction. and optical properties.

B https://doi.org/10.1021/acs.jpcc.3c08414
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 2. (a) P−P distances marked along both zigzag (d1) and armchair (d2) directions. Corresponding quantifications (e) are depicted for all
optimizations using vdW corrections. (b) Dispersion energies with respect to the number of layers (n), given by eV/n, for all vdW correction
models. (c) Interlayer distances represented for both top-bottom (L1) and top−top (L2), and (f) the corresponding average values quantified for n
= 1 up to bulk for each vdW model. (d) Lattice parameters for layers and bulk, denoted as a and b vectors, where the numbers listed correspond to
the c vector for the bulk (∞). Red diamonds highlight the experimental values for the lattice parameters59−61 and L2 interlayer distance,8 and vdW-
color association for the lines is defined as no vdW-black, D3-pink, D3BJ-red, dDsC-orange, TS-brown, TSSCS-violet, TSHP-green, and MBD-
blue.

• Scanning Coulomb potential models: use H(k) to scan The Hellmann−Feynman forces were relaxed to values of
2D and 3D Coulomb potential models, denoted as less than 1 meV/A on every atom. The optimization of forces
V(Q), to describe exciton e−h interactions within the for phosphorene multilayers included an additional dispersion
WanTiBEXOS code (TB-BSE). The models include energy term in the KS total energy. Several Edisp treatments
bare (V3D) and screened (V3DL) for 3D Coulomb were employed, incorporating empirical D3 corrections and its
potentials, as well as 2D models like those proposed by BJ damping function version (D3BJ) proposed by
Keldysh (V2DK),47 Rytova−Keldysh (V2DRK),48,49 Grimme,32,34 as well as semiempirical approaches based on
truncated (V2DT), 50 and its modified version dDsC41,42 and Tkatchenko−Scheffler methods, such as TS,
(V2DT2).51 Detailed descriptions of all models and TSHP, TSSCS, and MBD.36−38
parameters used, particularly in the 2D V(Q) models, 2.2. Gap Calculation via DFT-1/2. For bandgap energies,
can be found in the Supporting Information. At this
excluding excitonic effects, we employed the DFT-1/2 QP
stage, we utilized the Pn systems optimized without vdW
corrections. correction method,44 which maintains computational costs at
the standard DFT level. DFT-1/2 utilizes the Slater transition
• (W)TB-BSE excitonic properties calculation: employ the
technique extended to crystalline systems simulated via
defined V(Q) within (W)TB-BSE for all Pn systems
obtained from vdW corrections to calculate excitonic DFT,44,58 where a half-occupation scheme is employed,
properties. considering the valence band maximum hybridization for 3sp
states in the P atoms (weighted as 0.1 from s and 0.4 from p).
2.1. Total Energy, Geometry Optimization, and vdW The half-occupation is determined through the modified KS
Corrections. For the full relaxation of multilayers, we
potential, defined as Vmod ,KS = VKS( r ) VS( r ), where VKS( r )
employed the DFT52,53 semilocal formulation proposed by
Perdew, Burke, and Ernzerhof (PBE)54 for the xc energy represents the standard KS potential and VS( r ) is the self-
functional, as implemented in the Vienna Ab initio Simulation energy potential given by
Package (VASP).55,56 In the VASP implementation, we utilized VS( r ) = ( r , CUT)[Vae( r ) Vae 1/2( r )]. The ( r , CUT)
the projector augmented-wave method57 to solve the KS parameter is a cutoff function determined variationally to
equations, incorporating a fully relativistic approach to describe prevent an overlap of the Coulomb tails into neighboring atom
the core states. The phosphorus valence electrons were
sites, maximizing the gap energy without considering empirical
modeled with electron configurations of 3s2 and 3p3. In the
VASP calculations, Brillouin zone integration was performed parameters. Figure S1 in Supporting Information illustrates the
with a k-point mesh of 15 × 11 × Z, where Z = 3 for n = 1, Z = good structural transferability (n = 1 → 2 → bulk) of the
2 for n = 2 and 3, and Z = 1 for n = 4, 5, and 6. For bulk, we set variationally optimized CUT parameter, which remains at a
15 × 11 × 4. These grids were automatically generated through value of 3.20a0, wherein a0 is the Bohr radius.
the Monkhorst−Pack protocol. The plane waves were 2.3. Excitonic Properties Determined Using the
expanded up to a cutoff energy of 500 eV for all 2D systems. Bethe−Salpeter Formalism within TB Calculations. The
The total energy convergence criterion of 10−6 eV was orthogonal TB Hamiltonian, obtained through the Wannier90
employed throughout the calculations. framework, is expressed as follows
C https://doi.org/10.1021/acs.jpcc.3c08414
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

N same order as the Edisp magnitude for the c vector: D3BJ <
H(k) = H0 + eik·R iHR i dDsC < MBD < D3 < TSSCS, TSHP, and TS. By comparing
i=1 (1) with experimental values for the lattice parameters,59−61 we
where H0 represents the matrix elements of the Hamiltonian found that a vector for zigzag direction fits very well with
for the unit cell, incorporating on-site energies and hopping experiments (a = 3.31 Å) independently of the vdW model
parameters and HRi is the hopping matrices describing the employed. However, only D3, dDsC, and MBD models
d

resulted in b for armchair direction and c vectors very close to


interaction between the unit cell and its neighboring cells. The the experiments (i.e., b = 4.37 Å and c = 10.47 Å).
matrices H0 and HRi, obtained using the Wannier90 package,45
d

For stacked layers, our findings suggest that all vdW models
were utilized as input for the WanTiBEXOS code46 to calculate have negligible effects on the zigzag direction of Pn multilayers,
optical properties, incorporating excitonic effects. while they significantly impact the armchair direction,
The solution of the TB-BSE equation involved solving the including the layer normal c vector for the bulk. We observed
exciton Hamiltonian (Hexc), defined as a correlation between the Edisp magnitude and the b vector
Hexc = He + Hh + V (Q) concerning the number of layers, stabilizing at n = 4 when no
(2)
vdW approach is taken as a reference. The highest contractions
where He represents the single-particle electron Hamiltonian in b occur in the sequence D3BJ (0.31 Å), D3, MBD, and
component, Hh is the hole component, and V(Q) is the dDsC (0.22 Å) for n = 1 → bulk, while for the others,
Coulomb interaction potential that characterizes the binding of contractions fall within the interval of 0.06−0.15 Å. The P−P
e−h pairs. A comprehensive explanation of the TB-BSE distances in the zigzag plane (d1) and armchair (d2), quantified
formalism, as implemented in WanTiBEXOS, is provided in in Figure 2e, follow a similar pattern in response to the impact
the Supporting Information, including details about the of the vdW models on the a and b lattice vectors. Changes in
Coulomb potential models and the strategies employed for d1 are minimal compared to those in d2 for armchair bonds.
calculating the dielectric constant when necessary. Conversely, in alignment with the observation on the c vector
for the bulk, interlayer distances, such as top-bottom (L1) and
3. RESULTS top−top (L2) distances, are highly sensitive to the effects of
3.1. Structural Analysis. We conducted an analysis of Edisp magnitude contribution, given the expected vdW
several structural parameters for the Pn systems (n = 1, 2, 3, 4, homostructures morphology for n > 1 layers. Thus, the
5, and 6 as the number of layers) as a slab model with 17 A of sequence of shortening for L1 and L2 correlates with the Edisp
vacuum, including ab-plane lattice parameters, average bond magnitude, specifically, L1(L2): D3BJ < dDsC < MBD < D3 <
lengths (d1 and d2), and interlayer distances (for n > 1) for top- TSSCS, TSHP, and TS < no vdW. Furthermore, even though
bottom (L1) and top−top (L2), as depicted in Figure 2a,c. all the L2 values from the vdW model are into 5.1−5.7 Å
Additionally, Figure 2b quantifies the contribution of energy interval, D3, dDsC, and MBD models (in agreement with the
dispersion (Edisp in eV/n) from the employed vdW models, lattice parameter result) are the closest to the 5.3 Å value as
both in and intra layers. It is observed that the Edisp values the L2 experimental interlayer distance.8 This set of structural
exhibit a trend similar to that of n for Pn systems. Specifically, consequences, ordered by the Edisp intensity, impacts the
from n = 4 onward, the Edisp values remain relatively constant, thermodynamic stability and optoelectronic properties of Pn
adding only a small energy portion in the module for the bulk multilayer systems, as discussed in the following sections.
system. Regarding the magnitude of Edisp, which plays a crucial 3.2. Stability Analysis. To investigate the thermodynamic
role in the structural profile and indirectly affects the calculated stability of the Pn multilayers, we calculated the enthalpy of
optoelectronic properties of Pn, we observe three distinct aggregation (ΔHag), representing the heat released during the
regions of values among the vdW models: (i) D3BJ imparts the formation of multilayers (nP1 → Pn), for all employed vdW
highest dispersion energy contribution, with a difference of models and as a function of the number of layers. Figure 3a
approximately 0.61 eV/n for n = 1 → bulk. (ii) dDsC and illustrates the ΔHag values for each interface Pn−P1 formed
MBD exhibit similar behavior, and the difference in energy (i.e., n − 1 as the number of interfaces). Throughout the
contribution for n = 1 → bulk decreases to around 0.31 eV/n. aggregation process from n = 2 to 6, the calculated ΔHag values
(iii) D3, TS, TSSCS, and TSHP form a set with the lowest for all vdW models indicate an exothermic process. This
dispersion energy contribution portion. The difference for n = suggests that the heat released during the inclusion of one
1 → bulk remains around 0.31 eV/n within this group. monolayer is consistent and independent of the number of
In accordance with the Edisp contributions, Figure 2d layers in the substrate, reflecting the layer−layer interaction in
illustrates the lattice parameters for the n = 1, 2, 3, 4, 5, and the P(n−1) + P1 → Pn aggregation process. Additionally, the
6 layers (a and b vectors) and for the bulk (a, b, and c vectors) magnitude of the released heat roughly follows the sequence of
for all employed vdW models. Thanks to the attractive Edisp magnitude observed in the structural discussion. D3BJ
performance of the vdW interaction, all models induce a exhibits the highest exothermic aggregation process, while the
contraction of the b vector (i.e., armchair direction contraction no vdW calculations show the lowest, with a difference of
as layers are added) compared to systems without vdW around 450 meV/(n − 1). The ΔHag values for other vdW
correction, while the a vector (zigzag direction) remains models lie between −300 and −400 meV/(n − 1).
slightly constant for both the vdW model and the number of As structural distortion is expected with the addition of a
layers. This behavior is in agreement with results obtained new monolayer in the Pn system, we calculated the enthalpy
through optB88-vdW functional, in which a = 3.32−3.34 Å and portion related to distortion (ΔHdis) to investigate the
b = 4.58−4.47 Å from monolayer (throughout n = 2, 3, 4, and associated heat contribution. Figure 3b presents the average
5) up to bulk.18 For the bulk structure, all vdW models ΔHdis for the multilayer systems, along with their standard
promote a contraction of the c vector [as numerically depicted deviation indicated by the vertical bars, in relation to the P1
in panel (d)]. Notably, the sequence of shortening follows the pristine monolayer, as described by the equation. It is evident
D https://doi.org/10.1021/acs.jpcc.3c08414
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 4. (a) Gap energy (Eg) calculated without excitonic effects via
Figure 3. (a) Enthalpy of aggregation (ΔHag in meV) with respect to DFT-1/2 for Pn and bulk optimized using all vdW correction models.
the number of interfaces (n − 1), as indicated by the equation shown Solid lines represent the decreasing trend from n = 1 → 6, based on
in the figure in which Hn is the enthalpy of the layered systems and the indicated model and optimized a, b, and c parameters. (b) Eg for
Hi(n) is the enthalpy of the individual layers as part of a given n. P80 (reaching the bulk gap value)17,62 calculated through the fitting
Decrease in these values toward negative numbers signifies the model derived from all vdW correction models. Red shaded region
exothermic energy released during the aggregation process of n layers. represents the experimental gap interval of Eg = 0.3−0.4 eV reported
(b) Average enthalpy of distortion (ΔHdis in meV) for each layer through PL63 and scanning tunneling64 spectroscopy experiments.
along with the standard deviation (represented by vertical bars).
These values are associated with the heat involved in the distortion
contribution of individual layers with respect to the P1 pristine
monolayer, i.e., H(1). the lowest set of deviations (|ΔEg| < 5%) compared to the 1.45
eV obtained via PL,14−16 as illustrated in Figure 4b. In this
that the aggregation of monolayers (nP1 → Pn) induces a context, our DFT-1/2 protocol (based on PBE) at the
distortion process in each monolayer, contributing heat in an standard DFT level demonstrates good agreement with
endothermic manner. In line with the Edisp contributions computationally expensive HSE06 + vdW (1.51 eV) and
influencing structural parameters, such as contractions of the b mBJ + vdW (1.41 eV) xc functionals.18
vector, L1(L2) interlayer distances, and a higher dispersion of For n = 2, 3, 4, 5, 6, and bulk, our findings demonstrate that
the d2 bond length values, D3BJ and dDsC resulted in the the DFT-1/2 bandgap magnitude strongly depends on the
highest ΔHdis values, as expected. employed vdW approximation. However, the trend maintains a
3.3. Gap Energy via DFT-1/2+vdW. To employ an large gap for the monolayer, gradually decreasing to narrower
economical and effective method for bandgap correction, bandgaps for the bulk. Specifically, D3, D3BJ, dDsC, TSSCS,
DFT-1/2 was utilized with the optimization of the CUT and MBD yielded Eg values lower than those obtained via PL
parameter for both P1 and P2 systems (Figure S1). The in the n = 2−5 range, with |ΔEg| < 30% (with TS presenting
consistent CUT = 3.20a0 for both systems indicates the the lowest deviation below 20%). On the contrary, no vdW
method’s good transferability. Based on this, Figure 4a calculations yielded higher Eg values (except for n =
illustrates the bandgap values (Eg) for Pn multilayers obtained 2).15,63,66,67 Our results unveil a strong correlation between
through DFT-1/2 (without excitonic effects) for all vdW structural parameters (such as stretching b and L1,2 interlayer
methods using the optimized CUT parameter. The DFT-1/ distance contraction) and Eg. The wave function overlap18
2+vdW, based on the PBE functional, significantly improves between different layers is indirectly influenced by the Edisp
the Eg values compared to plain PBE + vdW, while keeping the magnitude; for instance, D3BJ maximizes Edisp, resulting in the
computational cost at the level of standard DFT based on the narrowest Eg values.
PBE xc functional (as depicted in Table S1 in Supporting For the bulk explicitly calculated, the inclusion of vdW
Information). Thus, our proposed bandgap calculation corrections leads to a striking scenario, with Eg decreasing to
protocol without excitonic effects serves as a preliminary step values below 0.11 eV (observed in no vdW and no QP
for obtaining excitonic properties through the wannierization correction, see Table S1). This trend is expected as a
process, subsequently employing the TB-BSE protocol. consequence of the Eg underestimation inherent to semilocal
First, a strong dependence of Eg on the number of layers xc functionals, which falls within 0.04−0.08 eV for plain DFT
from n = 1 up to 6 and the bulk is observed, reflecting an + vdW calculations. In contrast, DFT-1/2+vdW yields Eg
exponential decay relation indicative of thickness dependence values in the range of 0.08−0.90 eV. The narrowest Eg value
and quantum confinement effects. This trend aligns with is observed for the D3BJ method, aligning with the highest
findings from previous experimental and theoretical stud- wave function overlap interpretation. This method provides
ies.43,63,65−67 For n = 1, DFT-1/2 provides Eg values in the the strongest dispersion energy, resulting in the shortest b
1.37−1.47 eV range for all vdW models. These values exhibit parameter and L1,2 interlayer distances, thereby maximizing the
E https://doi.org/10.1021/acs.jpcc.3c08414
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 5. Comparison between DFT-1/2 (blue curves) and wannierization (red dots) band structures (A−Y−Γ−X−A−Γ) for systems with n = 1,
2, 3, 4, 5, and 6 layers and bulk system geometrically optimized without vdW correction (i.e., plain DFT). Valence band maximum (VBM) is set at
the Fermi level (0 eV), and the corresponding Eg values obtained through DFT-1/2 are indicated.

wave function overlap. However, this leads to a considerable suggests that charge carriers are confined in the effective 1D
deviation from experimental results based on PL.63 environment along the armchair direction (Γ → Y).70
Additionally, based on several fitting functions reported on For the (W)TB + BSE approach, we conducted convergence
theoretical and experimental studies on the literature,14,19,68,69 studies and screened the ground-state exciton binding energy
we employed a fitting model based on Eg(n) ∝ 1/na (a is an (Eb). This parameter is defined as the difference between the
optimized parameter) with our set of DFT-1/2+vdW gap direct bandgap, EDFT‑1/2
g , and the excitonic ground state, ETB opt.
energy results for Pn systems throughout the n = 1 → 6 We explored various Coulomb potentials for the e−h pairwise
interval, since those systems were calculated with V(Q) = interactions [V(Q) = V3D, V3DL, V2DK, V2DRK, V2DT, and
V2DRK; all the optimized parameters for the fitting model are V2DT2] for the Pn systems optimized without the vdW
depicted in Table S2 (Supporting Information). Since Eg method. It is noteworthy that EDFT‑1/2
g and ETB g (as the DFT
reaches the bulk value for P80,17,62 Figure 4b shows the and TB electronic gap energies) are coincident, justifying the
Eg(n) estimated from n = 80 through our fitting model, and the procedure of calculating Eb = EDFT‑1/2
g − ETB TB
opt, where Eopt is the
numerical data are available in Table S2. D3 and MBD models optical band gap obtained via (W)TB-BSE. Detailed optical
presented concurrent agreement with DFT-1/2+vdW calcu- (and QP) bandgaps under different k-mesh densities and Eb
lations and experiments (red shaded region) based on PL63 values for all V(Q) models can be found in Tables S3 and S4 in
and scanning tunneling64 spectroscopy. Supporting Information. We compared our results with PR−
The DFT-1/2+vdW results reveal a high sensitivity of the PL16 and QP-GW calculations43 from the literature. Among
gap energy in Pn layers and the bulk concerning the employed the various Coulomb potentials, we found that V(Q) =
vdW method. While all protocols summarize the gap energy V2DRK yielded the best comparative values of Eb, with 0.93
dependence on thickness, one must be mindful of the indirect eV for the monolayer (n = 1). This result aligns well with the
effects of empirical and semiempirical vdW methods on gap 0.9 ± 0.1 eV obtained via PR−PL (see Table S4).
energies due to their impact on the structure, even when Consequently, we employed the V2DRK Coulomb potential
working at the same level of theory. Conversely, our protocol, for all Pn systems optimized in our set of vdW methods, given
combining DFT + vdW with QP correction, stands as a the good transferability of these potentials. For the bulk, a
compelling option for initiating investigations into excitonic comparison between V3D and V3DL potentials revealed that
effects, offering more accurate gap energy values within a cost- the first is the suitable one since its topology requires
effective computational framework. periodicity throughout all dimensions.
3.4. Wannierization and Exciton Calculations. Our Figure 6 illustrates the ground-state Eb calculated by using
investigation of excitonic effects in Pn is based on constructing the V2DRK Coulomb potential for e−h pairwise interactions
the MLWF Hamiltonian. We utilized the MLWF-TB protocol within the (W)TB-BSE protocol for all Pn systems, considering
to incorporate the Hamiltonian into the (W)TB + BSE various vdW methods. For n = 1, our results are consistent
formalism, allowing us to obtain optical properties with with experimental data from PR−PL,16 in contrast to the
excitonic effects. Beginning with no vdW calculations, we significantly underestimated and computationally intensive
derived the scissors operator (SCS) from our TB Hamiltonian. QP-GW calculations.43 As for n > 1, Eb values were fitted for
This was constructed through the MLWF-TB parametrization each vdW model using the same shape of equation employed
(s and p orbitals of P atoms) over the DFT-1/2 band for EDFT‑1/2
g in Figure 5; the b1, b2, and b3 fitting parameters are
structures, as illustrated in Figure 5. In this figure, all DFT-1/2 depicted in Table S5 in Supporting Information, from which all
band gap values are indicated, showcasing a perfect fit of the functions yield R2 = 0.99. It is evident that Eb exponentially
electronic bands around and away from the Fermi level (EF) decreases in accordance with the reduction in quantum
along the A−Y−Γ−X−A−Γ path for Pn systems with n = 1, 2, confinement resulting from the additional stacking layers,
3, 4, 5, 6 layers, as well as bulk; Figure S2 provides a full k-path providing evidence of wave function overlap in the Pn+1
for the bulk by involving the kz direction. The bandgap is systems. The Eb values for the layers align with the magnitude
approximately located at the Γ-point, indicating a very flat of the Edisp energies, consistent with their impact on EDFT‑1/2
g
band dispersion along the Γ → X (zigzag) direction. This values. For instance, D3BJ yields the lowest set of Eb values for
F https://doi.org/10.1021/acs.jpcc.3c08414
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

D3BJ presented a spurious result with a negative Eb = −0.04


eV, which is a consequence of the great underestimation of the
band gap. Consequently, considering all the protocols
employed�initiating with the relativistic treatment of the
bandgap at the DFT level and progressing to the TB-BSE
approach for excitonic calculations�we conclude that the D3
and MBD vdW corrections exhibit notable performance in
describing the indirect effects of the structural profile on the
optoelectronic properties of multilayer BP systems.
Given that the MBD provides one of the best performances
for the set of electronic and structural properties in agreement
with the experiment, as observed in Figure 2d,f, we advance to
Figure 6. Exciton binding energy (Eb) calculated using the V(Q) = the optoelectronic properties of Pn systems by aiming for a
V2DRK Coulomb potential for electron−hole pairs within the better understanding of the BP stacking and the impact of
(W)TB-BSE approach for all Pn layered systems, whereas for the MBD vdW correction, which is depicted in Figure 7 for the
bulk, the V(Q) = V3D was employed, taking each structure optimized
using various vdW methods. PR−LP experiments16 reporting a
exciton absorption spectrum (BSE) through the light polar-
monolayer Eb = 0.9 ± 0.1 eV and the underestimation of theoretical izations along the x direction (αx) and y direction (αy). A
calculations based on the QP-GW approach43 are indicated. While n = comparison is made for all Pn systems from n = 2 to bulk with
∞ correspond to the Eb explicitly calculated for the bulk, the particular emphasis on the monolayer. The vertical lines in the
highlighted numbers show the Eb values for n = 80 based on the fitting spectra represent the single-particle gap energies (Eg) obtained
model by employing the range n = 1 → 6. through DFT-1/2. In our results for αx,y, the previously
observed orthogonality between heat and electron transport in
the exciton ground state, while no vdW protocol provides the monolayer BP is extended to systems with a few layers up to
highest set of values. the bulk. This orthogonality effect, reported in previous
By utilizing the fitting function applied to the Eb values studies,71 continues to influence the optical properties in
across the range n = 1 → 6, we extrapolated the multilayer systems. As a consequence of the wave function
correspondence between the bulk and the Eb values attained overlap in Pn+1 systems, a significant in-plane anisotropy is
at n = 80 for all vdW models. This procedure is similar to that observed when comparing αx and αy. For n > 1, the αx spectra
one employed for the band gap, which permits us to verify the are notably suppressed, indicating the presence of collective
matching between Eg(n = 80) and Eb(n = 80) through the electronic excitations below Eg values, especially relative to the
fitting models. The highlighted numbers in Figure 6 reveal that armchair (y) direction. This result reinforces the impact of
D3, TS, TSSCS, TSHP, and MBD are the vdW methods for linear polarization direction on the incident light, demonstrat-
which the bulk Eb value aligns with the fitting, consistent with ing that even with just a few stacked layers, BP’s optoelectronic
their effective calculations, such as Eb for n = ∞. Conversely, properties can be finely tuned.

Figure 7. Absorption coefficients (α) calculated for light polarizations along the x direction (αx) and y direction (αy). For Pn layered systems with
various numbers of layers (n = 1, 2, 3, 4, 5, and 6) and bulk, all geometrically optimized with MBD vdW correction, so that V(Q) = V2DRK
Coulomb potential was used for the layers, and for the bulk, the V(Q) = V3D was employed. Panels (a−f) show αx, while panels (g−l) depict αy.
Plot organization facilitates the comparison between systems with more than one layer (n > 1) and the monolayer (n = 1), emphasizing their
respective single-particle (Eg) values obtained through DFT-1/2.

G https://doi.org/10.1021/acs.jpcc.3c08414
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

4. SUMMARY *
sı Supporting Information
The Supporting Information is available free of charge at
We conducted a comprehensive and systematic investigation at
https://pubs.acs.org/doi/10.1021/acs.jpcc.3c08414.
various theoretical levels for BP systems spanning from
monolayer (n = 1) to bulk, including n = 2, 3, 4, 5, and 6 Empirical and semi-empirical vdW methods and
layers, with structural optimizations performed using different Coulomb potentials for e−h interaction (PDF)
empirical (D3 and D3BJ) and semiempirical (dDsC, TS,
TSHP, TSSCS, and MBD) vdW methods to enhance
interlayer nonbonded interactions. The characterization of
■ AUTHOR INFORMATION
Corresponding Authors
optoelectronic properties involved a combination of protocols, Diego Guedes-Sobrinho − Chemistry Department, Federal
starting with bandgap corrections using relativistic QP University of Paraná, Curitiba 81531-980, Brazil;
correction (DFT-1/2). The resulting corrected band structures orcid.org/0000-0002-3313-2822;
were then used to construct a TB Hamiltonian through a Email: [email protected]
MLWF-TB approach, followed by the solution of the BSE in Celso R. Caldeira Rêgo − Institute of Nanotechnology,
the TB framework. Our DFT-1/2+vdW approach revealed a Karlsruhe Institute of Technology (KIT), Eggenstein-
high sensitivity of the gap energy in Pn layers and bulk to the Leopoldshafen 76344, Germany; orcid.org/0000-0003-
choice of vdW methods. This emphasizes the need for careful 1861-2438; Email: [email protected]
consideration when employing empirical and semiempirical Maurício J. Piotrowski − Department of Physics, Federal
vdW methods to minimize significant underestimations of gap University of Pelotas, Pelotas, Rio Grande do Sul 96010-900,
energies by enhancing interlayer wave function overlap. Brazil; orcid.org/0000-0003-3477-4437;
Additionally, our semiempirical approach for calculating Email: [email protected]
excitonic properties, particularly the screening of e−h Alexandre Cavalheiro Dias − Institute of Physics and
interactions based on TB 2D and 3D Coulomb potential International Center of Physics, University of Brasília,
models, requires attention. The Rytova−Keldysh potential Brasília 70910900, Brazil; orcid.org/0000-0001-5934-
proved to be effective in describing the transition from the 8528; Email: [email protected]
monolayer to bulk, providing exciton binding energies that
were in excellent agreement with PL results from the literature, Authors
unlike the more computationally expensive QP-GW alter- Gabriel Reynald Da Silva − Chemistry Department, Federal
natives. Furthermore, our fitting model, applied to both University of Paraná, Curitiba 81531-980, Brazil
bandgap and exciton binding energy results, demonstrates that Henrique R. Da Silva − Chemistry Department, Federal
in particular for phosphorene, the performance of the D3 and University of Paraná, Curitiba 81531-980, Brazil
MBD method was coherent with experiments based on W. Wenzel − Institute of Nanotechnology, Karlsruhe Institute
monolayer and bulk, in order that its application for a 2D of Technology (KIT), Eggenstein-Leopoldshafen 76344,
system with different chemical composition and topology Germany; orcid.org/0000-0001-9487-4689
needs some care. On the other hand, our result provides a Complete contact information is available at:
particular evidence based on the achieving equivalence https://pubs.acs.org/10.1021/acs.jpcc.3c08414
between bulk calculations and the values obtained for n = 80
layers, indicating the reliability of D3 and MBD throughout the Notes
stacking of phosphorene layered. The authors declare no competing financial interest.

■ ASSOCIATED CONTENT ■ ACKNOWLEDGMENTS


The authors are thankful for the financial support provided by
Data Availability Statement
the National Council for Scientific and Technological
All the DFT calculations were carried out using VASP 5.4.4 Development (CNPq, grant nos. 307345/2021-1, 408144/
(https://www.vasp.at/). The creation of conformers for the 2022-0, and 305174/2023-1), the Federal District Research
scan calculations was generated using the open-source PyMOL Support Foundation (FAPDF, grant nos. 00193-00001817/
Application Programming Interface (API) (https://github. 2023-43 and 00193-00001817/2023-43), the Coordination for
com/schrodinger/pymol-opensource). The bandgap correc- Improvement of Higher Level Education (CAPES), the Rio
tion performed to all the systems based in all vdW correction Grande do Sul Research Foundation (FAPERGS), and the
was based on QP relativistic correction DFT-1/2 (https:// German Federal Ministry of Education and Research (BMBF)
github.com/Celso0408/DFT-1-2). The maximally localized for funding the Innovation-Platform MaterialDigital (www.
Wannier functions to orthogonal TB Hamiltonian�H(k)� materialdigital.de) project under project no. FKZ 13XP5094A.
were obtained through Wannier90 (https://wannier.org/), in The authors acknowledge the use of the HoreKa super-
which a proper set of parameters was used (https://github. computer, supported by the Ministry of Science, Research and
com/dguedessob/phosphorene_wannier90.git). Thus, H(k) the Arts Baden-Württemberg and the Federal Ministry of
was used as inputs into WanTiBEXOS to resolutions of the Education and Research. Additionally, the authors extend their
BSE (https://github.com/dguedessob/wantibexos_ thanks to the “Centro Nacional de Processamento de Alto
phosphorene.git), which were performed through the Desempenho em São Paulo” (CENAPAD-SP, UNICAMP/
WantiBEXOS code (https://wantibexos.readthedocs.io/en/ FINEP�MCTI project) for providing resources for projects
latest/ and https://github.com/ac-dias/wantibexos). Visual- 897 and 570, to Lobo Carneiro HPC (NACAD) at the Federal
ization and pictures were made using Visualization for University of Rio de Janeiro (UFRJ) for resources for project
Electronic and Structural Analysis (VESTA, https://jp- 133, to CIMATEC SENAI at Salvador�BA, Brazil, for their
minerals.org/vesta/en/). partnership and support through the Ogun Supercomputer,
H https://doi.org/10.1021/acs.jpcc.3c08414
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

and to the “Laboratório Central de Processamento de Alto (19) Zhang, G.; Chaves, A.; Huang, S.; Wang, F.; Xing, Q.; Low, T.;
Desempenho” (LCPAD) financed by FINEP through CT- Yan, H. Determination of Layer-Dependent Exciton Binding Energies
INFRA/UFPR projects. in Few-Layer Black Phosphorus. Sci. Adv. 2018, 4, No. eaap9977.
(20) Wei, Y.; Jiang, Y.; Pang, K.; Xu, X.; Kaner, N. T.; Li, X.; Yang,

■ REFERENCES
(1) Li, L.; Yu, Y.; Ye, G. J.; Ge, Q.; Ou, X.; Wu, H.; Feng, D.; Chen,
J.; Zhang, G.; Li, W.; Jiang, Y.; et al. Long Radiation Lifetime and
Quasi-Isotropic Excitons in Antioxidant V−V Binary Phosphorene
Allotropes with Intrinsic Dipole. J. Phys. Chem. C 2020, 124, 14787−
X. H.; Zhang, Y. Black Phosphorus Field-Effect Transistors. Nat. 14796.
Nanotechnol. 2014, 9, 372−377. (21) Pham, P. V.; Bodepudi, S. C.; Shehzad, K.; Liu, Y.; Xu, Y.; Yu,
(2) He, G.; Dong, T.; Yang, Z.; Ohlckers, P. Tuning 2D Black B.; Duan, X. 2D Heterostructures for Ubiquitous Electronics and
Phosphorus: Defect Tailoring and Surface Functionalization. Chem. Optoelectronics: Principles, Opportunities, and Challenges. Chem.
Mater. 2019, 31, 9917−9938. Rev. 2022, 122, 6514−6613.
(3) Yan, W.; Shresha, V. R.; Jeangros, Q.; Azar, N. S.; Balendhran, S.; (22) Eslamibidgoli, M. J.; Eikerling, M. H. Mechanical and Chemical
Ballif, C.; Crozier, K.; Bullock, J. Spectrally Selective Mid-Wave Stability of Monolayer Black Phosphorous Studied by Density
Infrared Detection Using Fabry-Pérot Cavity Enhanced Black Functional Theory Simulations. J. Phys. Chem. C 2018, 122,
Phosphorus 2D Photodiodes. ACS Nano 2020, 14, 13645−13651. 22366−22373.
(4) Seiler, H.; Zahn, D.; Zacharias, M.; Hildebrandt, P.-N.; (23) Peng, Y.; Lu, B.; Wang, N.; Lu, J. E.; Li, C.; Ping, Y.; Chen, S.
Vasileiadis, T.; Windsor, Y. W.; Qi, Y.; Carbogno, C.; Draxl, C.; Oxygen Reduction Reaction Catalyzed by Black-Phosphorus-
Ernstorfer, R.; et al. Accessing the Anisotropic Nonthermal Phonon Supported Metal Nanoparticles: Impacts of Interfacial Charge
Populations in Black Phosphorus. Nano Lett. 2021, 21, 6171−6178. Transfer. ACS Appl. Mater. Interfaces 2019, 11, 24707−24714.
(5) Han, D.; Han, X.; Liu, L.; Li, D.; Liu, Y.; Liu, Z.; Liu, D.; Chen, (24) Barik, G.; Pal, S. Energy Gap-Modulated Blue Phosphorene as
Y.; Zhuo, K.; Sang, S. Sub-ppb-Level Detection of Nitrogen Dioxide Flexible Anodes for Lithium- and Sodium-Ion Batteries. J. Phys. Chem.
Based on High-Quality Black Phosphorus. ACS Appl. Mater. Interfaces C 2019, 123, 2808−2819.
2022, 14, 13942−13951. (25) Wang, Z.; Lv, P.; Becton, M.; Hong, J.; Zhang, L.; Chen, X.
(6) Du, K.; Lv, Q.; Liang, Z.; Liu, G.; Hussain, S.; Liu, J.; Qiao, G. Mechanically Tunable Near-Field Radiative Heat Transfer between
Trends in the Preparation and Passivation Techniques of Black Monolayer Black Phosphorus Sheets. Langmuir 2020, 36, 12038−
Phosphorus Nanostructures for Optoelectronics Applications: A 12044.
Review. ACS Appl. Nano Mater. 2023, 6, 3159−3183. (26) Gómez-Pérez, J. F.; Correa, J. D.; Pravda, C. B.; Kónya, Z.;
(7) Buscema, M.; Groenendijk, D. J.; Blanter, S. I.; Steele, G. A.; van Kukovecz, A. Dangling-to-Interstitial Oxygen Transition and Its
der Zant, H. S. J.; Castellanos-Gomez, A. Fast and Broadband Modifications of the Electronic Structure in Few-Layer Phosphorene.
Photoresponse of Few-Layer Black Phosphorus Field-Effect Tran- J. Phys. Chem. C 2020, 124, 24066−24072.
sistors. Nano Lett. 2014, 14, 3347−3352. (27) Cohen, A. J.; Mori-Sánchez, P.; Yang, W. Fractional Charge
(8) Xia, F.; Wang, H.; Jia, Y. Rediscovering Black Phosphorus as an Perspective on the Band Gap in Density-Functional Theory. Phys. Rev.
Anisotropic Layered Material for Optoelectronics and Electronics. B 2008, 77, 115123.
Nat. Commun. 2014, 5, 4458. (28) Crowley, J. M.; Tahir-Kheli, J.; Goddard, W. A. Resolution of
(9) Wei, Q.; Peng, X. Superior mechanical flexibility of phosphorene the Band Gap Prediction Problem for Materials Design. J. Phys. Chem.
and few-layer black phosphorus. Appl. Phys. Lett. 2014, 104, 251915. Lett. 2016, 7, 1198−1203.
(10) Castellanos-Gomez, A. Black Phosphorus: Narrow Gap, Wide (29) Dion, M.; Rydberg, H.; Schröder, E.; Langreth, D. C.;
Applications. J. Phys. Chem. Lett. 2015, 6, 4280−4291. Lundqvist, B. I. Van Der Waals Density Functional for General
(11) Su, B.-W.; Li, X.-K.; Jiang, X.-Q.; Xin, W.; Huang, K.-X.; Li, D.- Geometries. Phys. Rev. Lett. 2004, 92, 246401.
K.; Guo, H.-W.; Liu, Z.-B.; Tian, J.-G. Carrier Engineering in (30) Klimeš, J.; Michaelides, A. Perspective: Advances and
Polarization-Sensitive Black Phosphorus van der Waals Junctions. Challenges in Treating van der Waals Dispersion Forces in Density
ACS Appl. Mater. Interfaces 2018, 10, 35615−35622. Functional Theory. J. Chem. Phys. 2012, 137, 120901.
(12) Xia, F.; Wang, H.; Hwang, J. C. M.; Neto, A. H. C.; Yang, L. (31) Grimme, S. Accurate Description of van der Waals Complexes
Black Phosphorus and its Isoelectronic Materials. Nat. Rev. Phys. by Density Functional Theory Including Empirical Corrections. J.
2019, 1, 306−317. Comput. Chem. 2004, 25, 1463−1473.
(13) Dong, B.; Huang, L.; Lee, C.; Ang, K.-W. Black Phosphorus (32) Grimme, S. Semiempirical GGA-Type Density Functional
Based Photodetectors. In Fundamentals and Applications of Phosphorus Constructed with a Long-range Dispersion Correction. J. Comput.
Nanomaterials; American Chemical Society, 2019; pp 135−153. Chem. 2006, 27, 1787−1799.
Chapter 7. (33) Grimme, S.; Hansen, A.; Brandenburg, J. G.; Bannwarth, C.
(14) Yang, J.; Xu, R.; Pei, J.; Myint, Y. W.; Wang, F.; Wang, Z.; Dispersion-Corrected Mean-Field Electronic Structure Methods.
Zhang, S.; Yu, Z.; Lu, Y. Optical Tuning of Exciton and Trion Chem. Rev. 2016, 116, 5105−5154.
Emissions in Monolayer Phosphorene. Light Sci. Appl. 2015, 4, (34) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and
No. e312. Accurate Ab Initio Parametrization of Density Functional Dispersion
(15) Liu, H.; Neal, A. T.; Zhu, Z.; Luo, Z.; Xu, X.; Tománek, D.; Ye, Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010,
P. D. Phosphorene: An Unexplored 2D Semiconductor with a High 132, 154104.
Hole Mobility. ACS Nano 2014, 8, 4033−4041. (35) Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the Damping
(16) Wang, X.; Jones, A. M.; Seyler, K. L.; Tran, V.; Jia, Y.; Zhao, H.; Function in Dispersion Corrected Density Functional Theory. J.
Wang, H.; Yang, L.; Xu, X.; Xia, F. Highly Anisotropic and Robust Comput. Chem. 2011, 32, 1456−1465.
Excitons in Monolayer Black Phosphorus. Nat. Nanotechnol. 2015, 10, (36) Tkatchenko, A.; Scheffler, M. Accurate Molecular Van Der
517−521. Waals Interactions from Ground-State Electron Density and Free-
(17) Chen, C.; Chen, F.; Chen, X.; Deng, B.; Eng, B.; Jung, D.; Guo, Atom Reference Data. Phys. Rev. Lett. 2009, 102, 073005.
Q.; Yuan, S.; Watanabe, K.; Taniguchi, T.; et al. Bright Mid-Infrared (37) Tkatchenko, A.; DiStasio, R. A.; Car, R.; Scheffler, M. Accurate
Photoluminescence from Thin-Film Black Phosphorus. Nano Lett. and Efficient Method for Many-Body van der Waals Interactions.
2019, 19, 1488−1493. Phys. Rev. Lett. 2012, 108, 236402.
(18) Qiao, J.; Kong, X.; Hu, Z.-X.; Yang, F.; Ji, W. High-Mobility (38) Ambrosetti, A.; Reilly, A. M.; DiStasio, R. A.; Tkatchenko, A.
Transport Anisotropy and Linear Dichroism in Few-Layer Black Long-Range Correlation Energy Calculated from Coupled Atomic
Phosphorus. Nat. Commun. 2014, 5, 4475. Response Functions. J. Chem. Phys. 2014, 140, 18A508.

I https://doi.org/10.1021/acs.jpcc.3c08414
J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

(39) Bučko, T.; Lebègue, S.; Hafner, J.; Á ngyán, J. G. Improved (63) Zhang, G.; Huang, S.; Wang, F.; Yan, H. Layer-Dependent
Density Dependent Correction for the Description of London Electronic and Optical Properties of 2D Black Phosphorus:
Dispersion Forces. J. Chem. Theory Comput. 2013, 9, 4293−4299. Fundamentals and Engineering. Laser Photonics Rev. 2021, 15,
(40) Bucko, T.; Lebègue, S.; Á ngyán, J. G.; Hafner, J. Extending the 2000399.
applicability of the Tkatchenko-Scheffler dispersion correction via (64) Zhang, C. D.; Lian, J. C.; Yi, W.; Jiang, Y. H.; Liu, L. W.; Hu,
iterative Hirshfeld partitioning. J. Chem. Phys. 2014, 141, 034114. H.; Xiao, W. D.; Du, S. X.; Sun, L. L.; Gao, H. J. Surface Structures of
(41) Steinmann, S. N.; Corminboeuf, C. Comprehensive Bench- Black Phosphorus Investigated with Scanning Tunneling Microscopy.
marking of a Density-Dependent Dispersion Correction. J. Chem. J. Phys. Chem. C 2009, 113, 18823−18826.
Theory Comput. 2011, 7, 3567−3577. (65) Li, L.; Kim, J.; Jin, C.; Ye, G. J.; Qiu, D. Y.; da Jornada, F. H.;
(42) Steinmann, S. N.; Corminboeuf, C. A Generalized-Gradient Shi, Z.; Chen, L.; Zhang, Z.; Yang, F.; et al. Direct observation of the
Approximation Exchange Hole Model for Dispersion Coefficients. J. layer-dependent electronic structure in phosphorene. Nat. Nano-
Chem. Phys. 2011, 134, 044117. technol. 2017, 12, 21−25.
(43) Qiu, D. Y.; da Jornada, F. H.; Louie, S. G. Environmental (66) Zhou, Q.; Zhou, H.; Tao, W.; Zheng, Y.; Chen, Y.; Zhu, H.
Screening Effects in 2D Materials: Renormalization of the Bandgap, Highly Efficient Multiple Exciton Generation and Harvesting in Few-
Electronic Structure, and Optical Spectra of Few-Layer Black Layer Black Phosphorus and Heterostructure. Nano Lett. 2020, 20,
Phosphorus. Nano Lett. 2017, 17, 4706−4712. 8212−8219.
(44) Ferreira, L. G.; Marques, M.; Teles, L. K. Approximation to (67) Joshi, P. P.; Li, R.; Spellberg, J. L.; Liang, L.; King, S. B.
Density Functional Theory for the Calculation of Band Gaps of Nanoimaging of the Edge-Dependent Optical Polarization Anisotropy
Semiconductors. Phys. Rev. B 2008, 78, 125116. of Black Phosphorus. Nano Lett. 2022, 22, 3180−3186.
(45) Mostofi, A. A.; Yates, J. R.; Lee, Y.-S.; Souza, I.; Vanderbilt, D.; (68) Zhang, G.; Huang, S.; Chaves, A.; Song, C.; Ö zçelik, V. O.;
Marzari, N. wannier90: A Tool for Obtaining Maximally-Localised Low, T.; Yan, H. Infrared Fingerprints of Few-Layer Black
Wannier Functions. Comput. Phys. Commun. 2008, 178, 685−699. Phosphorus. Nat. Commun. 2017, 8, 14071.
(46) Dias, A. C.; Silveira, J. F. R. V.; Qu, F. WanTiBEXOS: A (69) Zhou, J.; Cai, T.-Y.; Ju, S. Anisotropic Exciton Excitations and
Wannier Based Tight Binding Code for Electronic Band Structure, Optical Properties of Hittorf’s Phosphorene. Phys. Rev. Res. 2020, 2,
Excitonic and Optoelectronic Properties of Solids. Comput. Phys. 033288.
Commun. 2023, 285, 108636. (70) Margot, F.; Lisi, S.; Cucchi, I.; Cappelli, E.; Hunter, A.;
(47) Dias, A. C.; Bragança, H.; Zeng, H.; Fonseca, A. L. A.; Liu, D.- Gutiérrez-Lezama, I.; Ma, K.; von Rohr, F.; Berthod, C.; Petocchi, F.;
S.; Qu, F. Large Room-Temperature Valley Polarization by Valley- et al. Electronic Structure of Few-Layer Black Phosphorus from μ-
Selective Switching of Exciton Ground State. Phys. Rev. B 2020, 101, ARPES. Nano Lett. 2023, 23, 6433−6439.
085406. (71) Low, T.; Roldán, R.; Wang, H.; Xia, F.; Avouris, P.; Moreno, L.
(48) Chaves, A.; Sousa, G. O.; Khaliji, K.; da Costa, D. R.; Farias, G. M.; Guinea, F. Plasmons and Screening in Monolayer and Multilayer
A.; Low, T. Signatures of Subband Excitons in Few-Layer Black Black Phosphorus. Phys. Rev. Lett. 2014, 113, 106802.
Phosphorus. Phys. Rev. B 2021, 103, 165428.
(49) Van Tuan, D.; Yang, M.; Dery, H. Coulomb Interaction in
Monolayer Transition-Metal Dichalcogenides. Phys. Rev. B 2018, 98,
125308.
(50) Rozzi, C. A.; Varsano, D.; Marini, A.; Gross, E. K. U.; Rubio, A.
Exact Coulomb Cutoff Technique for Supercell Calculations. Phys.
Rev. B 2006, 73, 205119.
(51) Ismail-Beigi, S. Truncation of Periodic Image Interactions for
Confined Systems. Phys. Rev. B 2006, 73, 233103.
(52) Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys.
Rev. 1964, 136, B864−B871.
(53) Kohn, W.; Sham, L. J. Self-Consistent Equations Including
Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133−
A1138.
(54) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient
Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865−3868.
(55) Kresse, G.; Hafner, J. Ab Initio Molecular Dynamics for Open-
Shell Transition Metals. Phys. Rev. B 1993, 48, 13115−13118.
(56) Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab
Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys.
Rev. B 1996, 54, 11169−11186.
(57) Blöchl, P. E. Projector Augmented-Wave Method. Phys. Rev. B
1994, 50, 17953−17979.
(58) Slater, J. C.; Johnson, K. H. Self-Consistent-Field X−α Cluster
Method for Polyatomic Molecules and Solids. Phys. Rev. B 1972, 5,
844−853.
(59) Brown, A.; Rundqvist, S. Refinement of the Crystal Structure of
Black Phosphorus. Acta Crystallogr. 1965, 19, 684−685.
(60) Cartz, L.; Srinivasa, S. R.; Riedner, R. J.; Jorgensen, J. D.;
Worlton, T. G. Effect of Pressure on Bonding in Black Phosphorus. J.
Chem. Phys. 1979, 71, 1718−1721.
(61) Akai, T.; Endo, S.; Akahama, Y.; Koto, K.; Marljyama, Y. The
Crystal Structure and Oriented Transformation of Black Phosphorus
under High Pressure. High Pressure Res. 1989, 1, 115−130.
(62) Ling, X.; Wang, H.; Huang, S.; Xia, F.; Dresselhaus, M. S. The
renaissance of black phosphorus. Proc. Natl. Acad. Sci. U.S.A. 2015,
112, 4523−4530.

J https://doi.org/10.1021/acs.jpcc.3c08414
J. Phys. Chem. C XXXX, XXX, XXX−XXX

You might also like