<<

Journal of Materials Chemistry C

Dimensionality and Anisotropicity Dependence of Radiative Recombination in Nanostructured

Journal: Journal of Materials Chemistry C

Manuscript ID TC-ART-04-2019-002214.R1

Article Type: Paper

Date Submitted by the 28-May-2019 Author:

Complete List of Authors: Wu, Feng; University of California Santa Cruz Rocca, Dario; Universite de Lorraine Ecole Doctorale LTS Ping, Yuan; University of California Santa Cruz, Chemistry and Biochemistry

Page 1 of 8 Journal of Materials Chemistry C

Dimensionality and Anisotropicity Dependence of Ra- diative Recombination in Nanostructured Phospho- rene

Feng Wua, Dario Roccab and Yuan Ping∗c

The interplay between dimensionality and anisotropicity leads to intriguing optoelectronic prop- erties and exciton dynamics in low dimensional . In this study we use nanos- tructured phosphorene as a prototypical example to unfold such complex physics and develop a general first-principles framework to study exciton dynamics in low dimensional systems. Specif- ically we derived the radiative lifetime and light emission intensity from 2D to 0D systems based on many body perturbation theory, and investigated the dimensionality and anisotropicity effects on radiative recombination lifetime both at 0 K and finite temperature, as well as polarization and angle dependence of emitted light. We show that the radiative lifetime at 0 K increases by an order of 103 with the lowering of one dimension (i.e. from 2D to 1D nanoribbons or from 1D to 0D quantum dots). We also show that obtaining the radiative lifetime at finite temperature requires accurate exciton dispersion beyond the effective mass approximation. Finally, we demonstrate that monolayer phosphorene and its nanostructures always emit linearly polarized light consistent

with experimental observations, different from in-plane isotropic 2D materials like MoS2 and h-BN that can emit light with arbitrary polarization, which may have important implications for quantum information applications.

1 Introduction “quasi-1D" excitons 17. The complex nature of excited states and Two-dimensional (2D) materials are known for their unprece- the anisotropic behavior of excitonic wavefunctions may lead to dented potential in ultrathin electronics, photonics, intriguing effects on the exciton dynamics in MBP, which are dis- 1–11 and valleytronics applications . Strong light-matter interac- tinctive from the in-plane isotropic 2D systems as we will discuss tions and atomically-thin thickness lead to exotic physical prop- later. Further lowering dimensionality is expected to show an erties which do not exist in traditional 3D semiconductors. In interplay between quantum confinement and quasi-1D excitonic particular, the monolayer black phosphorus (MBP), commonly nature which affects its optoelectronic properties in a complex known as phosphorene, has recently attracted significant interest manner. 19 However, because lower dimensional MBP nanostruc- due to its emerging optoelectronic applications and the develop- tures (such as nanoribbons and quantum dots) are more difficult 12–14 ment of large scale fabrication methods . to stabilize than two-dimensional MBP 20,21, the optical measure- Unlike other common 2D materials such as , h-BN, ments and determination of excited state lifetime of these nanos- and transition metal dichalcogenides, MBP has a strong in-plane tructures are more challenging. Only the exciton recombination anisotropic behavior along two distinctive directions, denoted as lifetime of 2D MBP has been determined to be 211 ps through 12,15–18 “armchair”and “zigzag” . For example, the lowest exciton time-resolved photoluminescence measurements with light polar- transition is only bright when light is polarized along the arm- ized along the armchair direction at -10◦C 21. This calls for the- chair direction and the corresponding excitonic wavefunction is oretical studies to predict exciton dynamics of low dimensional much more extended along this direction, leading to so-called MBP nanostructures. Previously, the radiative lifetime of excitons of typical in-plane isotropic 2D systems have been computed by coupling Fermi’s a The Department of Chemistry and Biochemistry, University of California, Santa Cruz, golden rule with model Hamiltonians or the Bethe-Salpeter equa- 95064 CA, United States tion 22–25. However, the radiative lifetime and light emission in- b CNRS, LPCT, UMR 7019, 54506 Vandœuvre-lès-Nancy, France c The Department of Chemistry and Biochemistry, University of California, Santa Cruz, tensity of in-plane anisotropic systems (e.g. phosphorene), which 95064 CA, United States. E-mail: [email protected] have unique dependence on the polarization direction of light,

Journal Name, [year], [vol.], 1–8 | 1 Journal of Materials Chemistry C Page 2 of 8 have not been investigated in-depth. Additional outstanding 3 Results and Discussion questions involve the dimensionality (i.e. going from the 2D MBP to 1D and 0D nanostructures) and temperature dependence of 3.1 Radiative lifetime in nanostructred MBP excited state lifetime. In particular, temperature effects are deter- In order to study the dimensionality dependence of radiative life- mined by the exciton dispersion in momentum space, where the time in MBP nanostructures, we derived the general radiative de- typically used effective mass approximation may break down for cay rate expressions for anisotropic materials from 2D to 0D; de- low dimensional systems 26. By answering the above questions, tailed derivations can be found in Supporting Information (SI). In we will propose general principles and pathways of engineering general, the radiative decay rate can be written as radiative lifetime in anisotropic low-dimensional systems. γ(Qex) = γ0Y(Qex), (4) 2 Methods where γ0, which does not depend on the directions of photon or The rate of emission of a photon with specific wave-vector qL and exciton wavevectors, is the exciton decay rate at Qex = 0; the polarization direction λ from an exciton with wave-vector Qex is dependence on the wavevector direction is instead contained in Y(Qex), which satisfies Y(Qex = 0) = 1. To understand the effects of anisotropicity and dimensional- 2π D E 2 γ(Q ,q ,λ) = G,1 Hint S(Q ),0 ity of nanostructured MBP, we computed electronic structure, ab- ex L h¯ qLλ ex sorption spectra and radiative lifetimes of various MBP nanos-

× δ(E(Qex) − hcq¯ L), (1) tructures from 2D to 0D as shown in Figure 1 and Table 1. We used open source plane-wave code Quantum-Espresso 30,31 where 0 and 1 denote absence and presence of a photon, re- λqL with Perdue-Burke-Ernzehorf (PBE) exchange-correlation func- spectively; G is the ground-state; S(Qex) is the exciton state; tional 32, ONCV norm-conserving pseudopotentials 33,34. The E(Qex) is the exciton energy and V is the system volume. band structure is computed from GW approximation with the WEST-code 35,36. The absorption spectra and exciton properties Two important quantities can be derived from γ(Qex,qL,λ). One is the radiative decay rate γ(Qex) of the exciton that only are computed by solving the Bethe-Salpeter equation (BSE) in the 37 depends on exciton wavevector Qex; the other is the light emis- Yambo-code for MBP nanoribbons and monolayer systems. For MBP quantum dots, a BSE implementation without explicit empty sion intensity I(qL,λ), which only depends on photon wavevector states and inversion of dielectric matrix is applied instead 38–40 to qL and polarization λ for a given system. Both quantities can be directly measured from experiments. speed up the convergence. More computational details can be found in SI. The radiative decay rate of a specific exciton with wave-vector Optical absorption spectra of the nanostructures considered in Qex can be obtained from: this work are provided in SI; here we will summarize the main γ(Qex) = ∑ γ(Qex,qL,λ); (2) features that are relevant to the discussion of lifetimes. The first qLλ=1,2 exciton in phosphorene is bright with light polarized along the armchair direction, but dark along the zigzag direction. This be- the corresponding lifetime can be simply computed as the inverse havior is inherited by phosphorene nanostructures (nanoribbons of γ(Qex). and quantum dots). However, for these systems the light absorp- By using the dipole approximation and the relation p = tion and dipole matrix elements along a non-periodic direction i −m h¯ [r,H], the exciton transition matrix element in Eq. 1 can be are significantly weakened by a “depolarization effect", which written explicitly as comes from the microscopic electric fields induced by polariza- tion charge in an external field. This effect is included in our absorption spectra calculations by taking into account the local D int E G,1 H S(Q ),0 41,42 qLλ ex field effects in the Bethe-Salpeter Equation . s The radiative lifetimes for various MBP nanostructures at Qex = 2 e 2π 1 0 (obtained by 1/γ0) are summarized in Table 1. The lifetime of = εq λ · hG|r|Si. (3) m2cVh¯ q L the first exciton at 0 K changes dramatically from 0D (≈ 10-100 ns) to 1D (≈ 10-100 ps) then to 2D (≈ 100 fs), and generally The radiative decay rate γ(Qex) can be obtained by combin- decreases with increasing system sizes. Comparing the radiative ing Eqs. 1-3. The exciton energy E(Qex) and exciton dipole mo- decay rate expressions in SI we can see that with every decrease ments hG|r|Si necessary as inputs are computed by solving the of dimension (e.g. from 2D to 1D), an additional Ω0li/c multi- 27 Bethe-Salpeter equation , which accurately takes into account plicative factor appears in the decay rate expression, where li is the electron-hole interaction (this is mandatory for 2D semicon- the unit cell size and Ω0 is the exciton energy at Qex = 0. This ductors where large exciton binding energy > 0.5 eV has been is due to the momentum conservation between photon qi and ex- 28,29 observed ). citon Qi along the periodic direction i and energy conservation The details of BSE calculations can be found in the Supporting (hcq¯ L = h¯Ω0). Information (SI). Considering typical values in a solid such as

2 | Journal Name, [year], [vol.], 1–8 Page 3 of 8 Journal of Materials Chemistry C

Table 1 Radiative lifetime of the first exciton of MBP with 2D, 1D and 0D dimensions at 0 K and 263 K, with light polarized along the armchair direction.

Width Width h¯Ω µ2 Lifetime System Armchair Zigzag 0 A (eV) (a.u.) at 0 K (Å) (Å)

0D-2a2z 9 9 2.80 0.75 57 ns 0D-3a2z 13 9 2.39 3.49 20 ns 0D-4a2z 18 9 2.16 6.43 14 ns 1D-a-2z ∞ 9 1.73 7.42 21 ps 1D-a-3z ∞ 12 1.52 9.78 21 ps 1D-a-4z ∞ 15 1.51 11.89 17 ps 1D-z-2a 9 ∞ 2.72 0.04 1.0 ns 1D-z-3a 13 ∞ 2.40 0.45 130 ps 1D-z-4a 18 ∞ 2.17 0.64 110 ps 2D* ∞ ∞ 1.54 2.56 99 fs 2D(Exp)* 211 ps 21 2D(noEM)* 101 ps 2D(EM)* 1.8 ns * Exciton lifetime measured or computed at 263K. 2D(EM) is computed from effective mass approximation. 2D(noEM) is computed from Eq. 8.

h¯Ω0 ≈ 0.5 − 5 eV and l = 5 − 10 Bohr unit cell lattice parameter along one dimension, we can estimate that

−2 −3 Ω0l/c ≈ 10 − 10 . (5)

Accordingly, each reduction of dimension will give approximately 103 times longer lifetime at 0 K (this qualitative estimation does not take into account the change of dipole matrix elements and exciton energy). The above discussion explains the increase of the lifetime by several orders of magnitude when decreasing the di- mensionality. A similar trend can also be deduced from equations in Ref. 43. For systems of the same dimensionality, the lifetime also varies with the size and periodic direction of nanostructures. Specifi- Fig. 1 Structural models and exciton wavefunctions of MBP monolayer cally, the radiative lifetime is inversely proportional to the prod- (a), nanoribbons (b)(c) and quantum dots (d) with dangling bonds termi- 2 2 uct of squared dipole matrix element (µ A = |hG|r|Si| ) and the nated by hydrogen atoms. MBP monolayer has two different directions, n 2 exciton energy (Ω0), i.e. 1/γ0 ∝ 1/(Ω µ ), where n depends on armchair and zigzag. We constructed two kinds of nanoribbons: “1D- 0 A a-Nz" where the armchair direction is periodic, and N is the number of the dimensionality. We found that in most cases, by changing the 2 unit cells along the non-periodic zigzag direction. “1D-z-Na" where the system sizes, the change of µ A is much larger than that of Ω0 zigzag direction is periodic, and N is the number of unit cells along the and dominates the lifetime, as shown in Table 1. For example, non-periodic armchair direction. Quantum dots are denoted as “0D-Na- the fact that the radiative lifetime of nanoribbons periodic along Mz" where N and M give the number of unit cells along the specific direc- zigzag (“1D-z") is 101 − 102 times longer than their counterparts tions. The wavefunctions of the lowest exciton of different structures are 2 shown by a yellow isosurface with a value of 0.13 e/Å3. periodic along armchair (“1D-a") is directly determined by µ A. Because of the anisotropicity of MBP structure, the only non-zero component of hG|r|Si of the first e xciton i s a long t he armchair direction. In 1D-a nanoribbons the armchair direction is peri- 2 odic, so the quasi-1D exciton µ A is similar to 2D. On the other hand, in 1D-z nanoribbons the armchair direction is not periodic 2 and µ A is strongly reduced due to the depolarization effect; how- ever, the dipole moments with light polarized along the armchair 2 direction is still the main contribution to µ A. This is different from previous work on 1D nanostructures such as carbon nan-

Journal Name, [year], [vol.], 1–8 | 3 Journal of Materials Chemistry C Page 4 of 8

44 otubes where only the component µz in the periodic direction coupled: 2 dominates µA. Furthermore, the quasi-1D nature of the first ex- citon (extended along the armchair direction and confined along E(Qex) =E0 + (Ex(Qex,x) − E0) the zigzag in Figure 1) introduces a much larger change in life- + (E (Q ) − E ). (9) times as a function of the width for 1D-z nanoribbons compared y ex,y 0 to 1D-a nanoribbons. The radiative lifetimes of 0D systems are significantly longer The temperature-dependent radiative lifetimes of phosphorene with respect to the other nanostructures, ranging from 2.0×104 ps (specifically a t 2 63 K a nd 0 K) a re l isted i n T able 1 . I f t he tra- to 1.2 × 105 ps due to the dimensionality dependent factor for de- ditional approach based on the effective mass (“EM" value in Ta- cay rates in Eq. 5 (see Table 1). Similarly to the 1D-z nanoribbons, ble 1) approximation is used, a lifetime of 1.8 ns is obtained, the 0D-Na2z (N=2,3,4) series shows a clear trend of lifetime de- which overestimates the experimental result by an order of mag- creasing by increasing the quantum dot size along the armchair nitude. If the more complex model in Eq. 8 is used (“noEM" value direction. in Table 1), the computed lifetime of phosphorene at 263 K is 123 ps, which is in good agreement with the experimental result 211 ps 21. The remaining discrepancy (less than a factor of 2) 3.2 Radiative lifetime at finite temperature could be due to the use of a Si/SiO2 substrate in experimental Next we will discuss the exciton recombination lifetime at finite measurements, which introduces an additional dielectric screen- temperature for 2D phosphorene in order to compare with the ex- ing in the material compared to our free-standing system. Our perimental results. The radiative lifetime at finite temperature T findings show the importance of the accurate exciton dispersion is computed by assuming that the recombination process is slow E(Qex) beyond the effective mass approximation. enough to allow excitons at different Qex to reach thermal equi- librium 44. The temperature-dependent radiative lifetime is then Besides the temperature effect on exciton radiative recombina- written as a thermal statistical average: 22,44 tion lifetime at thermal equilibrium, other quasi-particles in the Z system may couple to excitons and lead to additional effects, for −1 γ(T) ≈ Z dQexγ(Qex) (6) example the band gap renormalization and nonradiative decay Qex

4 | Journal Name, [year], [vol.], 1–8 Page 5 of 8 Journal of Materials Chemistry C

We consider intensity along two polarization directions: through a fast decoherence process that leads to a significant loss of polarization 23. However, such decoherence mechanism does 2 I(θL,ϕL,IP) = n0Γ0 −Mx sinϕL + My cosϕL (12) not exist in MBP due to its anisotropicity which may be advan- tageous for quantum information applications that require long I(θ ,ϕ ,OOP) = n Γ (−M cosθ cosϕ L L 0 0 x L L coherence time of quantum states. The above conclusions agree 2 with the experimental photoluminescence measurements of the − My cosθL sinϕL + Mz sinθL) , (13) first exciton in MBP 16,58,59, which show a nearly perfect linear where IP (in-plane) and OOP (out-of-plane) denote two po- polarization. larization directions that are both perpendicular to the photon

wavevector qL (see Figure1 in SI), and Mi = µi/µA represent- ing the components of the normalized dipole moment along the directions i = x,y,z. The formulation in Eqs. 11-13 is general and only the angle independent term Γ0 describes the difference among 2D, 1D and 0D nanostructures (detailed derivations can be found in SI). The emitted photon polarization can be described by Stokes 54–57 parameters (S0,S1,S2,S3), which can be written explicitly as the following:

S0 = I(θL,ϕL,IP) + I(θL,ϕL,OOP),

S1 = 2I(θL,ϕL,IP) − S0,

S2 = 2I(θL,ϕL,45°) − S0,

S3 = 2I(θL,ϕL,L) − S0, (14)

where λ = 45° bisects the angle between IP and OOP, and λ = L denotes the left circularly polarized light. When |S3/S0| = 0 (1), the light is fully linearly polarized (circularly polarized). By eval- uating Eq. 14 with inputs from Eqs. 11-13, we can get the Stokes parameters as a function of normalized exciton dipole moments M. We found that because of the in-plane anisotropicity in MBP and its nanostructures, the emitted light is completely linearly po- larized. As we can consider Mz ≈ 0 (where z is along the perpen- dicular direction to the material plane), the circular polarization component S3 in Eq. 14 reduces to the following form (detailed derivations can be found in SI):

∗ ∗ S3 = n0Γ0icosθL(Mx My − MxMy ). (15)

When one of the following conditions is fulfilled: Mx = 0, My = 0, ∗ or MxMy is non-zero but real, S3 is zero and the light is fully lin- early polarized. Note that the specific choice of the x,y,z direc- tions does not change the conclusion obtained from Eq. 15, i.e. the polarization of the emitted light 54. In the case of anisotropic systems such as MBP nanostructures, by choosing the y axis along the dark transition direction and the x axis along the bright tran- sition direction (namely M = (1,0,0)), from Eq. 15 it can be im- mediately deduced that the emitted light is linearly polarized. In contrast, in-plane isotropic 2D systems with valley degener- acy like MoS2 may have possible normalized dipole moments M that give other types of polarization (circular or elliptical) due to the mixed exciton states between K and K0 valleys. For example, √ when the exciton has M = 1/ 2(1,i,0), we obtain S3/S0 = 1, and the light is fully circularly polarized. Mixed exciton states can go

Journal Name, [year], [vol.], 1–8 | 5 Journal of Materials Chemistry C Page 6 of 8

wavevector. This is the case regardless of the dimensionality and the polarization of the laser that excites the system. This behavior is different from in-plane isotropic 2D systems like BN and MoS2 that can have spherical symmetric (Figure 2(e) and 2(f) when M = √1 (1,i,0)) or cos2(ϕ ) angle dependence ((Figure 2(c) and 2 L 2(d) when M = √1 (1,1,0)) based on the mixture between K and 2 K0 valleys. 4 Conclusion

In conclusion, we presented a general framework to study exciton radiative lifetimes and light emission intensities in 2D, 1D, and 0D systems from many body perturbation theory. Based on it, impor- tant insights were provided on dimensionality and anisotropicity effects on exciton dynamics in phosporene nanostructures. For each dimensionality reduction the energy and momentum conser- vation leads to an additional factor of the order of 10−3 in the de- cay rates. This explained the general experimental observations that the exciton radiative lifetime is much longer in lower dimen- sional systems. Furthermore, in order to obtain the accurate ra- diative lifetime at finite temperature and quantitatively compare with experiments, accurate exciton energy dispersion in momen- tum space beyond the effective mass approximation must be con- sidered. Finally, we demonstrated that the exciton anisotropic- ity in MBP nanostructures always leads to linearly polarized light emission unlike in-plane isotropic 2D materials such as MoS2 and BN, which can emit light with arbitrary polarization. Acknowledgement

We thank Francesco Sottile and Matteo Gatti for helpful dis- cussions. This work is supported by NSF award DMR-1760260 and Hellman Fellowship. D.R. acknowledges financial support from Agence Nationale de la Recherche (France) under Grant No. ANR-15-CE29-0003-01. This research used resources of the Center for Functional Nanomaterials, which is a US DOE Office of Science Facility, and the Scientific Data and Comput- ing center, a component of the Computational Science Initia- tive, at Brookhaven National Laboratory under Contract No. DE- Fig. 2 The angle dependence of light emission intensity. The intensity SC0012704, the National Energy Research Scientific Computing along IP and OOP polarization directions at different ϕL (azimuthal angle Center (NERSC), a DOE Office of Science User Facility supported that defines photon wavevector) with given θL (polar angle) of different by the Office of Science of the US Department of Energy un- M ,M ,M combination computed from Eq. 12 and Eq. 13. x y z der Contract No. DEAC02-05CH11231, the Extreme Science and • (a)(b): M = (M ,0,0) (|M | = 1) x x Engineering Discovery Environment (XSEDE) through allocation • (c)(d): M = √1 (1,1,0) 2 TG-DMR160106, which is supported by National Science Foun- 60 • (e)(f): M = √1 (1,i,0) dation Grant No. ACI-1548562 . 2 • (a)(c)(e) θL = 0° References • (b)(d)(f) θL = 45° 1 X. Liu, Q. Guo and J. Qiu, Adv. Mater., 2017, 29, 1605886. 2 N. Youngblood, C. Chen, S. J. Koester and M. Li, Nat. Photon-

In addition, the angle dependence of I(θL,ϕL,IP/OOP) is ics, 2015, 9, 247. also different in in-plane anistropic and isotropic systems as 3 R.-J. Shiue, D. K. Efetov, G. Grosso, C. Peng, K. C. Fong and shown in Figure 2. For in-plane anisotropic systems like the D. Englund, Nanophotonics, 2017, 6, 1329–1342. MBP nanostructures we found that the light emission intensity 4 Q. Guo, A. Pospischil, M. Bhuiyan, H. Jiang, H. Tian, 2 (I/(n(Qex)Γ0)) always has a cos (ϕL) angle dependence as in Fig- D. Farmer, B. Deng, C. Li, S.-J. Han, H. Wang, Q. Xia, T.-P. ure 2(a) and 2(b), where ϕL is the azimuth angle of photon Ma, T. Mueller and F. Xia, Nano Lett., 2016, 16, 4648–4655.

6 | Journal Name, [year], [vol.], 1–8 Page 7 of 8 Journal of Materials Chemistry C

5 M. Bernardi, C. Ataca, M. Palummo and J. C. Grossman, 2437–2469. Nanophotonics, 2017, 6, 479–493. 28 F. Hüser, T. Olsen and K. S. Thygesen, Phys. Rev. B, 2013, 88, 6 G. Eda and S. A. Maier, ACS Nano, 2013, 7, 5660–5665. 245309. 7 L. Britnell, R. M. Ribeiro, A. Eckmann, R. Jalil, B. D. Belle, 29 D. Y. Qiu, F. H. da Jornada and S. G. Louie, Phys. Rev. B, 2016, A. Mishchenko, Y.-J. Kim, R. V. Gorbachev, T. Georgiou, S. V. 93, 235435. Morozov, A. N. Grigorenko, A. K. Geim, C. Casiraghi, A. H. C. 30 P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, Neto and K. S. Novoselov, Science, 2013, 340, 1311–1314. C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococ- 8 J. Wong, D. Jariwala, G. Tagliabue, K. Tat, A. R. Davoyan, cioni, I. Dabo, A. Dal Corso, S. de Gironcoli, S. Fab- M. C. Sherrott and H. A. Atwater, ACS Nano, 2017, 11, 7230– ris, G. Fratesi, R. Gebauer, U. Gerstmann, C. Gougoussis, 7240. A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri, 9 W. Han, APL Mater., 2016, 4, 032401. R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbrac- 10 Y. P. Feng, L. Shen, M. Yang, A. Wang, M. Zeng, Q. Wu, cia, S. Scandolo, G. Sclauzero, A. P. Seitsonen, A. Smogunov, S. Chintalapati and C.-R. Chang, Wiley Interdiscip. Rev.: Com- P. Umari and R. M. Wentzcovitch, J. Phys.: Condens. Matter, put. Mol. Sci., 2017, 7, e1313. 2009, 21, 395502. 11 A. Avsar, J. Y. Tan, M. Kurpas, M. Gmitra, K. Watanabe, 31 P. Giannozzi, O. Andreussi, T. Brumme, O. Bunau, M. B. T. Taniguchi, J. Fabian and B. Özyilmaz, Nat. Phys., 2017, Nardelli, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, 13, 888. M. Cococcioni, N. Colonna, I. Carnimeo, A. D. Corso, 12 A. Cupo and V. Meunier, J. Phys.: Condens. Matter, 2017, 29, S. de Gironcoli, P. Delugas, R. A. D. Jr, A. Ferretti, A. Floris, 283001. G. Fratesi, G. Fugallo, R. Gebauer, U. Gerstmann, F. Giustino, 13 M. Akhtar, G. Anderson, R. Zhao, A. Alruqi, J. E. T. Gorni, J. Jia, M. Kawamura, H.-Y. Ko, A. Kokalj, E. Küçük- Mroczkowska, G. Sumanasekera and J. B. Jasinski, npj 2D benli, M. Lazzeri, M. Marsili, N. Marzari, F. Mauri, N. L. Mater. Appl., 2017, 1, 5. Nguyen, H.-V. Nguyen, A. O. de-la Roza, L. Paulatto, S. Poncé, D. Rocca, R. Sabatini, B. Santra, M. Schlipf, A. P. Seitsonen, 14 A. Carvalho, M. Wang, X. Zhu, A. S. Rodin, H. Su and A. H. A. Smogunov, I. Timrov, T. Thonhauser, P. Umari, N. Vast, Castro Neto, Nat. Rev. Mater., 2016, 1, 16061 EP. X. Wu and S. Baroni, J. Phys.: Condens. Matter, 2017, 29, 15 J. Pang, A. Bachmatiuk, Y. Yin, B. Trzebicka, L. Zhao, L. Fu, 465901. R. G. Mendes, T. Gemming, Z. Liu and M. H. Rummeli, Adv. 32 J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, Energy Mater., 2017, 8, 1702093. 77, 3865–3868. 16 L. Li, J. Kim, C. Jin, G. J. Ye, D. Y. Qiu, H. Felipe, Z. Shi, 33 D. R. Hamann, Phys. Rev. B, 2013, 88, 085117. L. Chen, Z. Zhang, F. Yang, K. Watanabe, T. Taniguchi, W. Ren, S. G. Louie, X. H. Chen, Y. Zhang and F. Wang, Nat. Nanotech- 34 M. Schlipf and F. Gygi, Comput. Phys. Commun., 2015, 196, nol., 2017, 12, 21. 36–44. 17 V. Tran, R. Soklaski, Y. Liang and L. Yang, Phys. Rev. B, 2014, 35 M. Govoni and G. Galli, J. Chem. Theory Comput., 2015, 11, 89, 235319. 2680–2696. 18 D. Y. Qiu, F. H. da Jornada and S. G. Louie, Nano Lett., 2017, 36 T. A. Pham, H.-V. Nguyen, D. Rocca and G. Galli, Physical Re- 17, 4706–4712. view B, 2013, 87, 155148. 19 Z. Nourbakhsh and R. Asgari, Phys. Rev. B, 2016, 94, 035437. 37 A. Marini, C. Hogan, M. Grüning and D. Varsano, Comput. Phys. Commun., 2009, 180, 1392 – 1403. 20 A. Castellanos-Gomez, L. Vicarelli, E. Prada, J. O. Island, K. L. Narasimha-Acharya, S. I. Blanter, D. J. Groenendijk, 38 D. Rocca, D. Lu and G. Galli, J. Chem. Phys., 2010, 133, M. Buscema, G. A. Steele, J. V. Alvarez, H. W. Zandbergen, 164109. J. J. Palacios and H. S. J. van der Zant, 2D Mater., 2014, 1, 39 Y. Ping, D. Rocca, D. Lu and G. Galli, Phys. Rev. B, 2012, 85, 025001. 035316. 21 J. Yang, R. Xu, J. Pei, Y. W. Myint, F. Wang, Z. Wang, S. Zhang, 40 D. Rocca, Y. Ping, R. Gebauer and G. Galli, Phys. Rev. B, 2012, Z. Yu and Y. Lu, Light: Sci. Appl., 2015, 4, e312 EP. 85, 045116. 22 M. Palummo, M. Bernardi and J. C. Grossman, Nano Lett., 41 A. G. Marinopoulos, L. Reining, A. Rubio and N. Vast, Phys. 2015, 15, 2794–2800. Rev. Lett., 2003, 91, 046402. 23 H.-Y. Chen, M. Palummo, D. Sangalli and M. Bernardi, Nano 42 M. Bruno, M. Palummo, A. Marini, R. Del Sole, V. Olevano, Lett., 2018, 18, 3839–3843. A. N. Kholod and S. Ossicini, Phys. Rev. B, 2005, 72, 153310. 24 P. San-Jose, V. Parente, F. Guinea, R. Roldán and E. Prada, 43 H.-Y. Chen, V. A. Jhalani, M. Palummo and M. Bernardi, arXiv Phys. Rev. X, 2016, 6, 031046. preprint arXiv:1901.08747, 2019. 25 S. Ayari, A. Smiri, A. Hichri, S. Jaziri and T. Amand, Phys. Rev. 44 C. D. Spataru, S. Ismail-Beigi, R. B. Capaz and S. G. Louie, B, 2018, 98, 205430. Phys. Rev. Lett., 2005, 95, 247402. 26 P. Cudazzo, L. Sponza, C. Giorgetti, L. Reining, F. Sottile and 45 P. Cudazzo, F. Sottile, A. Rubio and M. Gatti, J. Phys.: Con- M. Gatti, Phys. Rev. Lett., 2016, 116, 066803. dens. Matter, 2015, 27, 113204. 27 Y. Ping, D. Rocca and G. Galli, Chem. Soc. Rev., 2013, 42, 46 J. Koskelo, G. Fugallo, M. Hakala, M. Gatti, F. Sottile and

Journal Name, [year], [vol.], 1–8 | 7 Journal of Materials Chemistry C Page 8 of 8

P. Cudazzo, Phys. Rev. B, 2017, 95, 035125. T. Heine, D. K. Maude, A. Kis and P. Plochocka, 2D Mater., 47 J. Noffsinger, E. Kioupakis, C. G. Van de Walle, S. G. Louie and 2018, 6, 015012. M. L. Cohen, Phys. Rev. Lett., 2012, 108, 167402. 54 E. Hecht, Optics,4/e, Pearson Education, 2002. 48 D. Y. Qiu, F. H. da Jornada and S. G. Louie, Phys. Rev. Lett., 55 W. H. McMaster, Am. J. Phys., 1954, 22, 351–362. 2013, 111, 216805. 56 W. H. McMaster, Rev. Mod. Phys., 1961, 33, 8–28. 49 V. Perebeinos and P. Avouris, Phys. Rev. Lett., 2008, 101, 57 G. Peach, N. Feautrier and H. E. Saraph, J. Phys. A: Math. 057401. Theor., 2009, 42, 445501. 50 H. Long, X. Peng, K. Lin, L. Xie, J. Lu, B. Zhang, L. Ying and 58 J. Qiao, X. Kong, Z.-X. Hu, F. Yang and W. Ji, Nat. Commun., Z. Wei, Appl. Phys. Express, 2019, 12, 052003. 2014, 5, 4475. 51 H. Mishra, A. Bose, A. Dhar and S. Bhattacharya, Phys. Rev. B, 59 X. Wang, A. M. Jones, K. L. Seyler, V. Tran, Y. Jia, H. Zhao, 2018, 98, 045143. H. Wang, L. Yang, X. Xu and F. Xia, Nat. Nanotechnol., 2015, 52 S. Shree, M. Semina, C. Robert, B. Han, T. Amand, A. Baloc- 10, 517. chi, M. Manca, E. Courtade, X. Marie, T. Taniguchi, K. Watan- 60 J. Towns, T. Cockerill, M. Dahan, I. Foster, K. Gaither, abe, M. M. Glazov and B. Urbaszek, Phys. Rev. B, 2018, 98, A. Grimshaw, V. Hazlewood, S. Lathrop, D. Lifka, G. D. Pe- 035302. terson, R. Roskies, J. R. Scott and N. Wilkins-Diehr, Comput. 53 J. M. Urban, M. Baranowski, A. Kuc, Ł. Kłopotowski, A. Sur- Sci. Eng., 2014, 16, 62–74. rente, Y. Ma, D. Włodarczyk, A. Suchocki, D. Ovchinnikov,

8 | Journal Name, [year], [vol.], 1–8