arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 20142, a represen-tative member of the TMD family that...

13
arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 2014 Exciton Binding Energy and Nonhydrogenic Rydberg Series in Monolayer WS 2 Alexey Chernikov, 1, Timothy C. Berkelbach, 2 Heather M. Hill, 1 Albert Rigosi, 1 Yilei Li, 1 Özgur B. Aslan, 1 David R. Reichman, 2 Mark S. Hybertsen, 3 and Tony F. Heinz 1, 1 Departments of Physics and Electrical Engineering, Columbia University, 538 West 120th Street, New York, NY 10027, USA 2 Department of Chemistry, Columbia University, 3000 Broadway, New York, NY 10027, USA 3 Center for Functional Nanomaterials, Brookhaven National Laboratory, Upton, NY 11973-5000, USA We have experimentally determined the energies of the ground and first four excited excitonic states of the fundamental optical transition in monolayer WS2, a model system for the growing class of atomically thin two-dimensional semiconductor crystals. From the spectra, we establish a large exciton binding energy of 0.32 eV and a pronounced deviation from the usual hydrogenic Rydberg series of energy levels of the excitonic states. We explain both of these results using a microscopic theory in which the non-local nature of the effective dielectric screening modifies the functional form of the Coulomb interaction. These strong but unconventional electron-hole interactions are expected to be ubiquitous in atomically thin materials. Atomically thin materials such as graphene and tran- sition metal dichalcogenides (TMDs) exhibit remarkable physical properties resulting from their reduced dimen- sionality [1]. The family of TMDs is an especially promis- ing platform for fundamental studies of two-dimensional (2D) systems, with potential applications in optoelec- tronics and valleytronics due to their direct gap, semi- conducting nature in the monolayer limit [2–7]. The recent advances in this emerging field include strongly enhanced photoluminescence [2, 4], efficient spin-valley coupling [8–11], pronounced many-body effects [6, 12], and high-performance in field-effect transistors [13]. The 2D character of monolayer TMDs suggests a strong enhancement of the Coulomb interaction. The resulting formation of bound electron-hole pairs, or ex- citons, can dominate the optical and charge-transport properties [14]. A microscopic understanding of how ex- citons are formed from otherwise free carriers is critical both for the elucidation of the underlying many-body physics in such materials and for their use in electronic and photonic devices, since the response of charged free carriers and neutral excitons to applied fields differs dra- matically. While theoretical and computational stud- ies have predicted exciton binding energies as high as 1 eV [15–19], a direct measurement of the exciton binding energy in atomically thin TMDs is still lacking (however, see Note added). In this work we experimentally and theoretically in- vestigate the properties of excitons in mono- and few- layer TMDs, identifying and characterizing not only the ground-state exciton, but the full sequence of excited (Rydberg) exciton states. Analyzing our sensitive mea- surements of the optical reflection spectra of these mate- rials, both empirically and using a physically motivated model for the non-local screening in TMDs, results in an estimate of 0.32(±0.04) eV for the 1s exciton bind- ing energy and 2.41(±0.04) eV for the quasiparticle gap of monolayer WS 2 . Remarkably, we also find significant deviations from the conventional hydrogenic model typi- cally employed for the description of Wannier excitons in inorganic semiconductors [20], and explain our findings in terms of a microscopic theory that highlights the pe- culiar form taken by the electron-hole interaction in this class of novel materials [21–23]. The specific material studied here is WS 2 , a represen- tative member of the TMD family that includes MoS 2 , MoSe 2 , and WSe 2 , all of which share similar properties with respect to atomic and electronic structure. The ad- vantage of WS 2 for this study is the large spin-orbit split- ting between the A and B excitons of about 0.4 eV [7], al- lowing for a study of the low-energy excitons unobscured by features from higher-lying transitions. In addition, the electronic transitions in the WS 2 samples exhibit narrow spectral features, permitting identification and analysis of many excited excitonic states and detailed quantitative comparison with theoretical predictions. Sample prepa- ration and characterization details can be found in the Supplemental Material [24]. Experimental and theoretical studies to date have clearly demonstrated that the basic excitonic properties of a three-dimensional bulk semiconductor differ funda- mentally from those of a 2D monolayer of the same ma- terial. The real-space origin of this behavior in TMDs is illustrated schematically in Fig. 1(a). In contrast to bulk, the electron and hole forming an exciton in monolayer TMDs are strongly confined to the plane of the monolayer and additionally experience reduced screening due to the change in the dielectric environment. These effects have two major implications for the electronic and excitonic properties of the material, shown by a schematic repre- sentation of the optical absorption in Fig. 1(b). First, the quasiparticle band gap is expected to increase for the monolayer. Second, the enhanced electron-hole interac- tion is expected to increase the exciton binding energy. In the absence of dielectric effects this yields an exci- ton binding energy that is a factor of four larger in 2D than in 3D. In the limit of atomically thin materials, however, the dielectric screening is also reduced because

Transcript of arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 20142, a represen-tative member of the TMD family that...

Page 1: arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 20142, a represen-tative member of the TMD family that includes MoS 2, MoSe 2, and WSe 2, all of which share similar properties with respect

arX

iv:1

403.

4270

v3 [

cond

-mat

.mtr

l-sc

i] 2

Sep

201

4

Exciton Binding Energy and Nonhydrogenic Rydberg Series in Monolayer WS2

Alexey Chernikov,1, ∗ Timothy C. Berkelbach,2 Heather M. Hill,1 Albert Rigosi,1 Yilei Li,1

Özgur B. Aslan,1 David R. Reichman,2 Mark S. Hybertsen,3 and Tony F. Heinz1, †

1Departments of Physics and Electrical Engineering, Columbia University,538 West 120th Street, New York, NY 10027, USA

2Department of Chemistry, Columbia University, 3000 Broadway, New York, NY 10027, USA3Center for Functional Nanomaterials, Brookhaven National Laboratory, Upton, NY 11973-5000, USA

We have experimentally determined the energies of the ground and first four excited excitonicstates of the fundamental optical transition in monolayer WS2, a model system for the growing classof atomically thin two-dimensional semiconductor crystals. From the spectra, we establish a largeexciton binding energy of 0.32 eV and a pronounced deviation from the usual hydrogenic Rydbergseries of energy levels of the excitonic states. We explain both of these results using a microscopictheory in which the non-local nature of the effective dielectric screening modifies the functional formof the Coulomb interaction. These strong but unconventional electron-hole interactions are expectedto be ubiquitous in atomically thin materials.

Atomically thin materials such as graphene and tran-sition metal dichalcogenides (TMDs) exhibit remarkablephysical properties resulting from their reduced dimen-sionality [1]. The family of TMDs is an especially promis-ing platform for fundamental studies of two-dimensional(2D) systems, with potential applications in optoelec-tronics and valleytronics due to their direct gap, semi-conducting nature in the monolayer limit [2–7]. Therecent advances in this emerging field include stronglyenhanced photoluminescence [2, 4], efficient spin-valleycoupling [8–11], pronounced many-body effects [6, 12],and high-performance in field-effect transistors [13].

The 2D character of monolayer TMDs suggests astrong enhancement of the Coulomb interaction. Theresulting formation of bound electron-hole pairs, or ex-citons, can dominate the optical and charge-transportproperties [14]. A microscopic understanding of how ex-citons are formed from otherwise free carriers is criticalboth for the elucidation of the underlying many-bodyphysics in such materials and for their use in electronicand photonic devices, since the response of charged freecarriers and neutral excitons to applied fields differs dra-matically. While theoretical and computational stud-ies have predicted exciton binding energies as high as1 eV [15–19], a direct measurement of the exciton bindingenergy in atomically thin TMDs is still lacking (however,see Note added).

In this work we experimentally and theoretically in-vestigate the properties of excitons in mono- and few-layer TMDs, identifying and characterizing not only theground-state exciton, but the full sequence of excited(Rydberg) exciton states. Analyzing our sensitive mea-surements of the optical reflection spectra of these mate-rials, both empirically and using a physically motivatedmodel for the non-local screening in TMDs, results inan estimate of 0.32(±0.04) eV for the 1s exciton bind-ing energy and 2.41(±0.04) eV for the quasiparticle gapof monolayer WS2. Remarkably, we also find significantdeviations from the conventional hydrogenic model typi-

cally employed for the description of Wannier excitons ininorganic semiconductors [20], and explain our findingsin terms of a microscopic theory that highlights the pe-culiar form taken by the electron-hole interaction in thisclass of novel materials [21–23].

The specific material studied here is WS2, a represen-tative member of the TMD family that includes MoS2,MoSe2, and WSe2, all of which share similar propertieswith respect to atomic and electronic structure. The ad-vantage of WS2 for this study is the large spin-orbit split-ting between the A and B excitons of about 0.4 eV [7], al-lowing for a study of the low-energy excitons unobscuredby features from higher-lying transitions. In addition, theelectronic transitions in the WS2 samples exhibit narrowspectral features, permitting identification and analysisof many excited excitonic states and detailed quantitativecomparison with theoretical predictions. Sample prepa-ration and characterization details can be found in theSupplemental Material [24].

Experimental and theoretical studies to date haveclearly demonstrated that the basic excitonic propertiesof a three-dimensional bulk semiconductor differ funda-mentally from those of a 2D monolayer of the same ma-terial. The real-space origin of this behavior in TMDs isillustrated schematically in Fig. 1(a). In contrast to bulk,the electron and hole forming an exciton in monolayerTMDs are strongly confined to the plane of the monolayerand additionally experience reduced screening due to thechange in the dielectric environment. These effects havetwo major implications for the electronic and excitonicproperties of the material, shown by a schematic repre-sentation of the optical absorption in Fig. 1(b). First,the quasiparticle band gap is expected to increase for themonolayer. Second, the enhanced electron-hole interac-tion is expected to increase the exciton binding energy.In the absence of dielectric effects this yields an exci-ton binding energy that is a factor of four larger in 2Dthan in 3D. In the limit of atomically thin materials,however, the dielectric screening is also reduced because

Page 2: arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 20142, a represen-tative member of the TMD family that includes MoS 2, MoSe 2, and WSe 2, all of which share similar properties with respect

2

FIG. 1. (a) Real-space representation of electrons andholes bound into excitons for the three-dimensional bulk anda quasi-two-dimensional monolayer. The changes in the di-electric environment are indicated schematically by differentdielectric constants ε3D, ε2D, and the vacuum permittivityε0. (b) Impact of the dimensionality on the electronic andexcitonic properties, schematically represented by optical ab-sorption. The transition from 3D to 2D is expected to lead toan increase of both the band gap and the exciton binding en-ergy (indicated by the red dashed line). The excited excitonicstates and Coulomb correction for the continuum absorptionhave been omitted for clarity.

the electric field lines joining the electron and hole beginto extend outside of the sample as shown in Fig. 1(a),potentially yielding an even greater enhancement factor.This so-called “dielectric confinement” or “image chargeeffect” [21, 25] was observed in nano-structured materialssuch as single-walled carbon nanotubes [26] and layeredorganic-inorganic perovskites [27]. The effectiveness ofthe dielectric screening thus depends on the separationbetween the electron and hole, giving rise to a non-localdielectric screening. This modifies the form of the in-teraction potential [21–23, 25] and causes a significantchange of the disposition of the energies of the excitonicstates, as discussed in more detail below.

To access these exciton properties experimentally westudy the so-called excitonic Rydberg series, i.e., theexcited states of the bound electron-hole pairs, labeledin analogy to the hydrogen series as 2s, 3s, and soon. In contrast to p- or d-like states with nonzero or-bital angular momentum, these transitions are dipole-allowed and are thus found in the linear optical spec-tra of many semiconductors with peak positions locatedbetween the quasiparticle band gap and the exciton 1sground state [14, 20]. The energy separation of theseresonances corresponds to a hydrogenic progression forWannier-like excitons. In addition, the coupling of the

excited states to light is reduced compared to the maintransition so that their spectral weight decreases withincreasing quantum number.

In our experiments we measure the reflectance con-trast ∆R/R = (Rsample − Rsubstrate)/Rsubstrate of theWS2 monolayer sample at a temperature of 5 K. The ex-perimental details are given in the Supplemental Mate-rial [24]. The spectrum, plotted in the inset of Fig. 2, ex-hibits several pronounced peaks on a broad background,the latter arising from interference effects induced by the300-nm thick SiO2 layer between the sample and the Sisubstrate [6]. The main transitions correspond to theso-called A, B, and C excitons in WS2 [7]. A small ad-ditional feature on the low-energy side of the A peak isidentified as a charged exciton (or trion), with a bind-ing energy on the order of 20–30 meV. Such a featurehas been observed in monolayers of other TMDs at lowtemperatures [6, 12] and indicates the presence of someunintentional residual doping in the WS2 sample. Here,we focus on the properties of the A exciton, related tothe fundamental band gap of the material. In order tohighlight the otherwise weak signatures of the higher-lying excitonic transitions, we plot in Fig. 2 the deriva-tive of the reflectance contrast d

dE (∆R/R) in the energyrange of interest. On the high-energy side of the exciton1s ground state, we observe multiple additional peaks,which we identify as the 2s, 3s, 4s, and 5s states of theA exciton, since the decrease of both the peak intensityand the energy spacing for increasing energy are charac-teristic features of an excitonic Rydberg series [14, 20].The peak positions extracted by taking the respective

FIG. 2. The derivative of the reflectance contrast spec-trum d

dE(∆R/R) of the WS2 monolayer. The exciton ground

state and the higher excited states are labeled by their re-spective quantum numbers (schematically shown at bottom-right). The spectral region around the 1s transition (AX) andthe trion peak (AXT ) of the A exciton is scaled by a factor of0.03 for clarity. Inset shows the as-measured reflectance con-trast ∆R/R for comparison, allowing for the identification ofthe A, B, and C transitions.

Page 3: arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 20142, a represen-tative member of the TMD family that includes MoS 2, MoSe 2, and WSe 2, all of which share similar properties with respect

3

points of inflection, corresponding to the zero-crossingsof the second derivative, are plotted in Fig. 3(a). Therespective energies are further confirmed by simulatingthe material response with a multiple-Lorentzian fit (seeSupplemental Material [24]).

To calculate the exciton binding energy, we mustfirst determine the quasiparticle band gap correspond-ing to the energy of a separated electron-hole pair.This is typically accomplished in semiconductors by fit-ting the excitonic peaks to a hydrogenic Rydberg se-ries [20]. In 2D, this hydrogen model employs an effec-tive mass Hamiltonian, H = −~

2∇

2r/2µ+ Veh(r), where

µ = 1/(m−1e + m−1

h ) is the exciton reduced mass andVeh(r) = −e2/εr is a locally-screened attractive electron-hole interaction. This model predicts exciton transitionenergies of Eg − E

(n)b , where Eg is the quasiparticle gap

and

E(n)b =

µe4

2~2ε2 (n− 1/2)2(1)

is the binding energy of the nth excitonic state. In con-trast, the exciton energies seen in our experiments exhibita much weaker scaling with the quantum number n, pre-cluding a simple fit to the data based on this model. How-ever, we observe that the n = 3− 5 peaks are reasonablyhydrogenic and by fitting to these data points only, weextract a quasiparticle band gap of Eg = 2.41(±0.04) eV,where the error bars originate from the fitting proce-dure. Subtracting the 1s transition energy of 2.09 eVfrom this band gap, we find an exciton binding energy ofEb = 0.32(±0.04) eV.

To provide insight into the non-hydrogenic physics ofthe n = 1, 2 excitons and the justification of a hydrogenicfit to the n = 3− 5 excitons, we first consider the use ofan effective dielectric constant in the hydrogenic Hamil-tonian. Using an exciton reduced mass of µ = 0.16m0 (asdetermined by density functional theory at the K or K ′

point [23, 28], see Supplemental Material [24]) and thequasiparticle band gap of Eg = 2.41 eV, we determinethe n-dependent dielectric constant εn required to repro-duce the experimental binding energy of the nth excitonE

(n)b,exp

, i.e., εn = [2~2E(n)b,exp

(n − 1/2)2/µe4]−1/2. Theresults plotted in the inset of Fig. 3(a) show a strong de-crease in this effective dielectric constant with increasingquantum number n. Because the exciton radius increaseswith n, we conclude that the physically correct electron-hole interaction is more strongly screened at short range,but only weakly screened at long range. In particular, theeffective dielectric is nearly constant for n = 3−5 (justify-ing our empirical use of the 2D hydrogen model for thesedata points), but shows significant deviations for n = 1, 2.This can be understood qualitatively in terms of a non-uniform dielectric environment schematically illustratedin Fig. 3(c). The electric field between an electron and ahole forming an exciton permeates both the thin layer of

FIG. 3. (a) Experimentally and theoretically obtained tran-sition energies for the exciton states as a function of the quan-tum number n. The fit of the n = 3, 4, 5 data to the 2Dhydrogen model for Wannier excitons is shown for compari-son. Grey bands represent uncertainty in the quasi-particleband gap from the fitting procedure. Corresponding effec-tive dielectric constants are shown in the inset. (b) Screened2D interaction Eq. (2) used in the model Hamiltonian (black)compared to the 2D hydrogen interaction 1/r (red); a semilog-arithmic plot is given in the inset. Also shown are the corre-sponding energy levels and radial wavefunctions up to n = 3.(c) Schematic representation of electron-hole pairs forming 1sand 2s excitonic states in a non-uniform dielectric environ-ment.

material with comparably strong screening and the sur-rounding medium with much weaker screening. As thespatial separation between the charges increases, a largerportion of the electric field is located in the surroundinglow-dielectric medium and the effective screening is re-duced. This phenomenon of “anti-screening,” giving riseto non-hydrogenic exciton behavior, has previously beenpredicted in carbon nanotubes, a quasi one-dimensionalsemiconductor [26].

To understand this behavior quantitatively, we ap-ply our recently developed theory of excitons in tran-sition metal dichalcogenides [23]. The treatment is againbased on a 2D effective mass Hamiltonian, but witha nonlocally-screened electron-hole interaction describedby the potential

Veh(r) = −πe2

2r0

[H0

(r

r0

)− Y0

(r

r0

)], (2)

where H0 and Y0 are Struve and Bessel functions.This interaction form describes the electrostatic inter-action of two charges within a thin 2D dielectric contin-

Page 4: arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 20142, a represen-tative member of the TMD family that includes MoS 2, MoSe 2, and WSe 2, all of which share similar properties with respect

4

uum [21, 22, 25]. The screening length r0, which canbe related to the 2D polarizability of the monolayer ma-terial [22], gives a crossover length scale between a 1/rCoulomb interaction at large separation and a weakerlog(r) interaction at small separation. This modifiedfunctional form of the interaction, which is a manifesta-tion of the strong dielectric contrast between the mono-layer WS2 and its surroundings, is responsible for thealtered disposition of the low-lying excitonic states ob-served experimentally.

Using the above functional form for the screened in-teraction we have numerically calculated the radiallysymmetric, s-type eigenstates of the excitonic Hamilto-nian, again using the calculated exciton reduced massµ = 0.16m0. Taking only the screening length r0 and theband gap Eg as free parameters, we find that we can veryaccurately fit the entire n = 1− 5 series of exciton levelswith the values r0 = 75 Å and Eg = 2.41 eV. These arethe parameters which minimize the root-mean-squareddeviation between theory and experiment. For this choiceof parameters, the 1s exciton binding energy is found tobe 0.32 eV, and so both the band gap and the bindingenergy are found to agree with the values determinedabove. We emphasize that the adopted screening lengthshould be understood as one that partially accounts foradditional screening due to the substrate, such that theintrinsic binding energy of WS2 is expected to be largerthan the value found here, in qualitative agreement withab initio calculations [15–17] (see Supplemental Materialfor a discussion of the microscopic origin of the precisevalue of r0 [24]). Fig. 3(b) depicts the noticeably weak-ened interaction at small electron-hole separations, alongwith the first three calculated radial wavefunctions. Theexciton radius is calculated to be approximately 30 Åfor the 1s exciton and even larger for the higher-lyingexcitons, which supports a strictly 2D treatment whencompared to the monolayer width of about 6 Å. Simi-larly, this relatively large in-plane spatial extent impliesa narrow reciprocal space distribution, justifying an ef-fective mass approximation centered around the K andK ′ points of the Brillouin zone. The above success offitting to a hydrogenic model is also explained by thepresent microscopic approach because the n = 3 − 5 ex-citon wavefunctions are large enough in spatial extentto predominantly probe the asymptotic 1/r form of thepotential given in Eq. (2).

Finally, to study the influence of the material thicknesswe monitor the spectral position of the 2s resonance forvarying thickness of the WS2 sample. Individual deriva-tives of the reflection contrast are plotted in Fig. 4(a)for the monolayer (1L), bilayer (2L), tetralayer (4L), andbulk. The corresponding energies of the 1s and 2s tran-sitions are shown in Fig. 4(b), with higher exited statesmasked by additional spectral broadening. Unlike forthe case of the monolayer, the bulk excitons are accu-rately treated with an anisotropic 3D hydrogenic Hamil-

FIG. 4. (a) The derivative of the reflectance contrast spec-tra for the WS2 monolayer (1L), bilayer (2L), tetralayer (4L)and bulk. The positions of the 2s exciton resonance are indi-cated by dotted arrows. (b) Energies of the 1s and 2s statesfor various layer thicknesses. Band gaps of the bulk and themonolayer are represented by the dashed lines.

tonian [29] that accounts for anisotropy in both the elec-tron and hole band masses and in the dielectric tensor.Using ab initio calculated values, we obtain a bulk ex-citon binding energy of 0.05 eV (see Supplemental Ma-terial [24]), implying a band gap of Eg = 2.10 eV,both of which are in agreement with literature resultsfor bulk WS2 [30]. As the layer thickness decreases, the2s resonance shifts to higher energies, while the 1s res-onance remains relatively unchanged, implying a strongincrease in both the exciton binding energy and the quasi-particle band gap. Both shifts are found to be large inabsolute energies, but opposite in sign. This explainsthe small change in the transition energy of the excitonground state, similar to findings reported for quasi-one-dimensional systems [26].

The large binding energy of 0.32 eV and the non-hydrogenic behavior of the intra-excitonic states in mono-layer WS2 are expected to be features common to otherTMD materials, based on the strong similarity in theirelectronic structure. Even larger values of the bindingenergy may be expected for suspended and undoped lay-ers, although the studied system represents the typicalscenario for experimental investigations. These energieslead to a high thermal stability of excitons and combinedwith their high oscillator strength should facilitate theapplication of these materials in photonic and excitonicdevices. The non-hydrogenic series of the excited statesresulting in large energy spacings between the individ-ual states is expected to allow for efficient exploitationof the intra-excitonic processes with potential applica-tions in the far-IR and THz spectral range. High-densityeffects scaling with the strength of the Coulomb scat-tering such as screening and band-gap renormalization,generally important for devices operating in this regime,e.g., lasers and concentrator solar-cells, can be potentially

Page 5: arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 20142, a represen-tative member of the TMD family that includes MoS 2, MoSe 2, and WSe 2, all of which share similar properties with respect

5

harvested for optical modulators, saturable absorbers,and tunable emitters operating in the limit of atomicthickness. Finally, the observed properties of the WS2

excitons render the material system highly suitable asa link between inorganic semiconductors with spatiallyextended, weakly-bound Wannier excitons and organicsemiconductors with spatially localized, strongly-boundFrenkel excitons. The presence of the strong Coulombinteraction opens up possibilities for both fundamentalstudies of the many-body physics in 2D materials as wellas for distinctive applications in optoelectronic devices.

This work was supported in part by the Center for Re-Defining Photovoltaic Efficiency through Molecule ScaleControl, an Energy Frontier Research Center funded bythe U.S. Department of Energy, Office of Science, Officeof Basic Energy Sciences under Award No. de-sc0001085.Support for the reflection contrast measurements mea-surements was also provided by the National ScienceFoundation through grant DMR-1123894. A.C. grate-fully acknowledges funding from the Alexander von Hum-boldt Foundation within the Feodor-Lynen Fellowshipprogram. T.C.B. was supported in part by the U.S. De-partment of Energy, Office of Science under ContractNo. DE-AC05-06OR23100. This work was carried outin part at the Center for Functional Nanomaterials,Brookhaven National Laboratory, which is supportedby the U.S. Department of Energy, Office of Basic En-ergy Sciences under Contract No. DE-AC02-98CH10886(M.S.H).

Note added. — After submission of our paper, sev-eral additional experimental and theoretical reports onthe excitonic properties of monolayer TMDs have ap-peared [31].

[email protected][email protected]

[1] K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth,V. V. Khotkevich, S. V. Morozov, and A. K. Geim,Proc. Natl. Acad. Sci. U.S.A., 102, 10451 (2005).

[2] K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz,Phys. Rev. Lett., 105, 136805 (2010).

[3] T. Li and G. Galli, J. Phys. Chem. C, 111, 16192 (2007).[4] A. Splendiani, L. Sun, Y. Zhang, T. Li,

J. Kim, C.-Y. Chim, G. Galli, and F. Wang,Nano Lett., 10, 1271 (2010).

[5] L. Britnell, R. M. Ribeiro, A. Eckmann, R. Jalil, B. D.Belle, A. Mishchenko, Y.-J. Kim, R. V. Gorbachev,T. Georgiou, S. V. Morozov, A. N. Grigorenko, A. K.Geim, C. Casiraghi, A. H. Castro Neto, and K. S.Novoselov, Science, 340, 1311 (2013).

[6] J. S. Ross, S. Wu, H. Yu, N. J. Ghimire, A. M. Jones,G. Aivazian, J. Yan, D. G. Mandrus, D. Xiao, W. Yao,and X. Xu, Nature Comm., 4, 1474 (2013).

[7] W. Zhao, Z. Ghorannevis, L. Chu, M. Toh, C. Kloc, P.-H.Tan, and G. Eda, ACS Nano, 7, 791 (2013).

[8] K. F. Mak, K. He, J. Shan, and T. F. Heinz,

Nature Nanotech., 7, 494 (2012).[9] H. Zeng, J. Dai, W. Yao, D. Xiao, and X. Cui,

Nature Nanotech., 7, 490 (2012).[10] T. Cao, G. Wang, W. Han, H. Ye, C. Zhu, J. Shi,

Q. Niu, P. Tan, E. Wang, B. Liu, and J. Feng,Nature Comm., 3, 887 (2012).

[11] A. M. Jones, H. Yu, N. J. Ghimire, S. Wu, G. Aivazian,J. S. Ross, B. Zhao, J. Yan, D. G. Mandrus, D. Xiao,W. Yao, and X. Xu, Nature Nanotech., 8, 634 (2013).

[12] K. F. Mak, K. He, C. Lee, G. H. Lee, J. Hone, T. F.Heinz, and J. Shan, Nature Mater., 11, 1 (2012).

[13] D. Lembke and A. Kis, ACS Nano, 6, 10070 (2012).[14] H. Haug and S. W. Koch, Quantum theory of the opti-

cal and electronic properties of semiconductors, 5th ed.(World Scientific, Singapore, 2009).

[15] A. Ramasubramaniam, Phys. Rev. B, 86, 115409 (2012).[16] H.-P. Komsa and A. V. Krasheninnikov,

Phys. Rev. B, 86, 241201(R) (2012).[17] H. Shi, H. Pan, Y.-W. Zhang, and B. I. Yakobson,

Phys. Rev. B, 87, 155304 (2013).[18] D. Y. Qiu, F. H. da Jornada, and S. G. Louie,

Phys. Rev. Lett., 111, 216805 (2013).[19] G. Berghäuser and R. Malic, Phys. Rev. B, 89, 125309

(2014).[20] C. Klingshirn, Semiconductor Optics, 2nd ed. (Springer,

Berlin Heidelberg New York, 2007).[21] L. V. Keldysh, JETP Lett., 29, 658 (1979).[22] P. Cudazzo, I. V. Tokatly, and A. Rubio,

Phys. Rev. B, 84, 085406 (2011).[23] T. C. Berkelbach, M. S. Hybertsen, and D. R. Reichman,

Phys. Rev. B, 88, 045318 (2013).[24] See Supplemental Material at [url], which includes

Refs. [32]–[44], for additional details on sample character-ization, optical microscopy, spectroscopy, and theoreticaland computational methodology.

[25] E. Hanamura, N. Nagaosa, M. Kmagai, and T. Takaga-hara, Mat. Sci. Eng. B, 1, 255 (1988).

[26] J. Deslippe, M. Dipoppa, D. Prendergast, M. V. O.Moutinho, R. B. Capaz, and S. G. Louie, Nano Lett., 9,1330 (2009).

[27] K. Tanaka, T. Takahashi, T. Kondo, T. Umebayashi,K. Asai, and K. Ema, Phys. Rev. B, 71, 045312 (2005).

[28] D. Xiao, G.-B. Liu, W. Feng, X. Xu, and W. Yao, Phys.Rev. Lett., 108, 196802 (2012).

[29] A. Baldereschi and M. G. Diaz, Nuovo Cimento, 68B,217 (1970).

[30] A. R. Beal and W. Y. Liang, J. Phys. C, 9, 2459 (1976).[31] B. Zhu, X. Chen, and X. Cui, (2014), arXiv:1403.5108;

Z. Ye, T. Cao, K. O’Brien, H. Zhu, X. Yin, Y. Wang,S. G. Louie, and X. Zhang, (2014), arXiv:1403.5568;G. Wang, X. Marie, I. Gerber, T. Amand, D. Lagarde,L. Bouet, M. Vidal, A. Balocchi, and B. Urbaszek,(2014), arXiv:1404.0056; M. M. Ugeda, A. J. Bradley,

S.-F. Shi, F. H. da Jornada, Y. Zhang, D. Y. Qui, S.-K.Mo, Z. Hussain, Z.-X. Shen, F. Wang, S. G. Louie, andM. F. Crommie, (2014), arXiv:1404.2331; T. Strouckenand S. W. Koch, (2014), arXiv:1404.4238; K. He, N. Ku-mar, L. Zhao, Z. Wang, K. F. Mak, H. Zhao, and J. Shan,Phys. Rev. Lett., 113, 026803 (2014).

[32] P. Tonndorf, R. Schmidt, P. Böttger, X. Zhang,J. Börner, A. Liebig, M. Albrecht, C. Kloc, O. Gor-dan, D. R. T. Zahn, S. Michaelis de Vasconcellos, andR. Bratschitsch, Opt. Express, 21, 4908 (2013).

[33] J. Wilson and A. Yoffe, Adv. Phys., 18, 193 (1969).

Page 6: arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 20142, a represen-tative member of the TMD family that includes MoS 2, MoSe 2, and WSe 2, all of which share similar properties with respect

6

[34] H. R. Gutiérrez, N. Perea-López, A. L. Elías,A. Berkdemir, B. Wang, R. Lv, F. López-Urías,V. H. Crespi, H. Terrones, and M. Terrones,Nano Lett., 13, 3447 (2012).

[35] W. Zhao, Z. Ghorannevis, A. K. Kumar, J. R. Pang,M. Toh, X. Zhang, C. Kloc, P. H. Tan, and G. Eda,(2013), arXiv:1304.0911.

[36] C. Lee, H. Yan, L. E. Brus, T. F. Heinz, J. Hone, andS. Ryu, ACS Nano, 4, 2695 (2010).

[37] E. Hecht, Optics, 4th ed. (Addison-Wesley, 2001).[38] S. Byrnes, Multilayer thin film optics calculator,

http://sjbyrnes.com/ (2012).[39] I. H. Malitson, J. Opt. Soc. Am., 55, 1205 (1965).[40] M. Bass, C. DeCusatis, J. Enoch, V. Lakshminarayanan,

G. Li, C. MacDonald, V. Mahajan, and E. V. Stryland,eds., Handbook of Optics, (vol. IV), 3rd ed. (McGraw-Hill, 2009).

[41] C. Zhang, H. Wang, W. Chan, C. Manolatou, andF. Rana, Phys. Rev. B, 89, 205436 (2014).

[42] P. Giannozzi, et al., J. Phys. Condens. Matter., 21,395502 (2009).

[43] J. Deslippe, G. Samzonidze, D. A. Strubbe, M. Jain,M. L. Cohen, and S. G. Louie, Comput. Phys. Com-mun., 183, 1269 (2012).

[44] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev.Lett., 77, 3865 (1996).

Page 7: arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 20142, a represen-tative member of the TMD family that includes MoS 2, MoSe 2, and WSe 2, all of which share similar properties with respect

7

SUPPLEMENTAL MATERIAL

Experimental methods

Spatial images of the samples were obtained using an optical microscope (Nikon Eclipse ME600) with a 100xobjective and a CCD camera for detection. Photoluminescence (PL) and Raman measurements were performed in ina commercial Raman microscope (Renishaw inVia) under ambient conditions at room temperature with the respectivespectral resolutions of 1 meV and 1 cm−1. The spectra were calibrated using the Si Raman line at 520 cm−1. Acontinuous-wave solid-state laser with a central wavelength of 532 nm was used for excitation. The power was set to50 µW and the laser was focused to a 1 µm diameter spot.

For the reflectance contrast measurements, broadband radiation from a tungsten quartz halogen source was focusedon the sample by a 40× objective, yielding a spot of about 2 µm in diameter. The reflected signal was collected andanalyzed with a grating spectrometer and a liquid nitrogen cooled CCD. The measurements were performed bothat room temperature and at 20 K. The data were oversampled with 10 pixels corresponding to the experimentalresolution of the setup of about 10 meV. Normalizing the reflectance from the sample to that of the bare substrate, weobtain the reflection contrast with a signal-to-noise ratio (SNR) of about 103. We took advantage of the oversamplingto average the derivative spectra over an interval of ∆ = 10 pixels. This further improved the SNR without anysignificant impact on the spectral resolution. We explicitly confirmed that an increase or decrease of ∆ by a factorof two did not affect the measured peak positions. The latter were obtained by taking the respective points ofinflection, corresponding to the zero-crossings of the second derivative. We also carried out simulations of the fulloptical response of the sample and substrate, including the influence of multiple reflections from the substrate andoxide layer, as well as the WS2 response. The exciton transition energies inferred from this procedure agreed withinexperimental error with those obtained from the simple point of inflection analysis (see section “Simulation of thelow-temperature reflection contrast” for details).

In general, not only the high SNR of the experiment but also the quality of the samples is crucial for the observationof higher excited states reported in the manuscript. Only the samples with sufficiently narrow line-widths of theexcitonic transitions allow for the detection of weaker resonances above the ground state, which can be furtherfacilitated by performing the measurements at low temperatures. The line-width of the higher transitions is found toscale with the line-width of the ground state, yet it is generally significantly broader than the main resonance due tothe presence of additional relaxation processes. Moreover, the residual doping of the samples significantly broadenshigher lying exciton states, precluding their observation for samples with high intrinsic doping. Nevertheless, higherexciton states were observed in four different WS2 monolayers, including room-temperature measurements.

Optical microscopy images

Figures S1(a), (b), and (c) present optical microscopy images of the WS2 monolayer (1L), bilayer (2L), andtetralayer (4L) samples, respectively. The average area of the studied flakes is in the range of 20 µm2, suitablefor optical studies using spots with diameters up to several µm. Representative grey scale contrast profiles are takenalong the dashed lines and plotted on the right-hand side of the images, normalized to the contrast of the SiO2/Sisubstrate. The discrete steps of about 0.1 observed in the grey-scale contrast correspond to the number of WS2 layers.

Room-temperature photoluminescence and reflection contrast measurements

Relying on previous studies of optical properties of the TMDs and the WS2 system in particular [1–4], we are ableto identify unambiguously the layer thickness of our samples using a combination of PL and reflection contrast data.PL spectra of the studied WS2 samples are shown in Fig. S2(a), normalized and offset for clarity. The correspondingtotal luminescence yield is plotted in the inset. The main K-point resonance and the indirect band gap emission areidentified by the labels A, and I, respectively. The emission from the monolayer is characterized by an increase of thePL intensity by about two orders of magnitude compared to the bilayer and by the absence of an indirect transition inthe luminescence spectra. The measured positions of the indirect gap at 1.73 eV and 1.49 eV for the bi-and tetralayer,respectively, agree well with the published energies of 1.72 eV and 1.48 eV from Ref. 4. For comparison, the energiesfor the tri- and pentalayer were previously found to be 1.53 eV and 1.42 eV [4]. Also, the red shift of the indirecttransition is accompanied by a further decrease of the emission intensity.

Page 8: arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 20142, a represen-tative member of the TMD family that includes MoS 2, MoSe 2, and WSe 2, all of which share similar properties with respect

8

FIG. S1. Optical micrographs of the studied WS2 monolayer (1L), bilayer (2L), and tetralayer (4L) samples. Exemplarygrey-scale contrast profiles are taken along the dashed lines and shown on the right-hand side of the respective images. Thedata is normalized to the contrast of the SiO2/Si substrate.

FIG. S2. (a) Room-temperature photoluminescence spectra of the WS2 mono- and few-layer samples. The data are normalizedand offset for clarity. The main K-point transition and the indirect band gap are indicated by A and I, respectively. Theoverall luminescence yield is plotted in the inset as function of the layer thickness. (b) Corresponding spectra of the reflectancecontrast with the spectral positions of the direct transitions marked by A, B, and C. The inset shows the smoothed derivativesof the reflectance contrast, normalized and offset for clarity, in the range of the C resonance.

Figure S2(b) shows the measured reflectance contrast spectra for the three samples. The overall sinusoidal shapeof the spectra with values above and below zero is due to the change in the interference pattern typical for layerssupported by SiO2/Si substrates (see e.g. Refs. 5 and 6). The spectral positions of the direct transitions are marked byA, B, and C according to the usual labeling convention [4, 7]. The inset shows smoothed derivatives of the reflectancecontrast in the range of the C resonance, normalized and offset for clarity. The energy positions of the C peaktransition of 2.86 eV, 2.78 eV, and 2.72 eV for the mono-, bi-, and tetralayer, respectively, agree with the previouslyreported values of 2.83 eV, 2.75 eV, and 2.70 eV [4] within the experimental uncertainty. The blue-shift of the A andB resonances, as well as the overall decrease of the reflection contrast with decreasing thickness, further corroboratethe assignment of the layer number.

Page 9: arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 20142, a represen-tative member of the TMD family that includes MoS 2, MoSe 2, and WSe 2, all of which share similar properties with respect

9

Raman spectroscopy

FIG. S3. Room temperature Raman spectra of the WS2 mono- and few-layer samples. The spectra are normalized and offsetfor clarity. The energy difference between the A1g and E1

2g modes is plotted in the inset as function of the layer thickness.

The Raman spectra for the three samples are plotted in Fig. S3, normalized and offset for clarity. The observedpeaks are identified as follows: the A1g(Γ) optical mode at 417 cm−1, E1

2g(Γ) optical mode at 354.4 cm−1 merged withthe second order acoustical mode 2xLA(M) at 350.6 cm−1, and a coupled acoustical-optical mode 2xLA(M)-3xE2

2g(M)at 295.4 cm−1 [8, 9]. The precise origin of the resonances at 322 cm−1 and 312.5 cm−1 is still under discussion [9]. Asthe layer number increases, the A1g(Γ) mode blue shifts by a few wavenumbers and the E1

2g(Γ) and 2xLA(M) modesredshift very gradually, as previously reported [8]. The energy difference between the A1g and E1

2g modes, shown inthe inset, thus increases with the layer number, as is typically observed for the TMDs [8, 10].

Simulation of the low-temperature reflection contrast

To verify the energy positions of the exciton peaks in the presence of the interference from the underlying SiO2/Sisubstrate we simulate the optical response of the monolayer WS2 at 20 K using the standard transfer-matrix methodfor evaluating light propagation in planar thin films (see e.g., Ref. 11). The calculations are performed using anopen-source software [12]. The transfer-matrix technique simulates the linear optical response of a given multi-layeredthin-film system taking into account multiple internal reflections and interference for layers of arbitrary thicknessesand complex refractive indices. The wavelength-dependent refractive indices of the SiO2 and Si are taken from Refs. 13and 14, respectively. The A, B, and C exciton resonances of the WS2 monolayer as well as the excited states of theA-exciton (2s, 3s, 4s, and 5s) are parametrized with Lorentzian peak functions; a constant background permittivityis included to take into account higher-lying transitions. The peak energies, widths, and spectral weights of theresonances are adjusted to obtain an optimal overlap of the simulated and measured reflectivity. Real and imaginaryparts of the dielectric function produced by the simulation to provide the best fit of the experimental data are shownin Fig. S4(a). The corresponding derivative of the simulated reflection contrast is plotted in Fig. S4(b), together withthe measured spectrum. Fig. S4(c) presents the same data magnified in the spectral region of the A-exciton Rydbergseries. The simulated data have been additionally shifted vertically by a constant value for better comparison of thelineshapes. The A-exciton peak energies produced by the simulation are plotted in Fig. S4(d) and compared with thepeak positions obtained by taking the points of inflection (POI) in the measured reflection contrast derivative. Theapproaches are in agreement within the experimental error, as indicated by the error bars for the POI values. Thesefindings corroborate the initial assumption that the interference from the multi-layered substrate does not significantlyaffect the measured energies of the spectrally narrow exciton resonances.

Page 10: arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 20142, a represen-tative member of the TMD family that includes MoS 2, MoSe 2, and WSe 2, all of which share similar properties with respect

10

FIG. S4. Simulations of the optical response based on the transfer-matrix method to account for multiple reflections. (a) Thereal and imaginary parts of the dielectric function used in the simulation to fit the measured reflectivity of the WS2 monolayeron the SiO2/Si substrate at temperature of 20 K. The permittivity at the excitonic resonances is parametrized with multipleLorentzians. (b) Measured and simulated derivatives of the reflection contrast. (c) Magnified derivatives of the reflectioncontrast in the spectral region of the A-exciton fine-structure. The simulated data have been also shifted by a constant valuefor better comparison with the measured lineshape. (d) Exciton peak energies extracted by taking the points of inflection arecompared with the values from the simulation.

Theoretical modeling and exciton properties

As discussed in the text, we use a microscopic exciton Hamiltonian (r = (x, y) is the 2D in-plane coordinate)

H = −~2∇

2r

2µ+ Veh(r) (S3)

where µxy = 1/(m−1e + m−1

h ) is the exciton reduced mass in the 2D xy plane, which for WS2 has been previouslycalculated to be in the range of 0.15–0.22 m0 [15–17], depending on the level of theory used (m0 is the rest mass ofthe electron). Here we use µ = 0.16 m0, but similar results could be obtained with any value in this range. Thescreened electron-hole interaction is given in reciprocal-space by

Veh(q) = −2πe2

qε(q)= −

2πe2

q(1 + r0q)(S4)

or in real-space by

Veh(r) = −πe2

2r0

[H0

(r

r0

)− Y0

(r

r0

)](S5)

where H0 and Y0 are Struve and Bessel functions. The eigenvalues and eigenfunctions of the Hamiltonian are calculatednumerically by diagonalization on a one-dimensional real-space grid.

The above interaction has been derived by Keldysh for quasi-2D semiconductors [18] and by Cudazzo et al. for strictly2D semiconductors [19]. In these derivations, the screening lengthscale r0 is given by r0 = dε/2 and r0 = 2πχ2D,respectively, where d is the layer thickness, ε is an isotropic macroscopic dielectric constant, and χ2D is the 2Dpolarizability. We have recently applied such a formalism to the family of TMDs to investigate the binding energy ofboth neutral excitons and charged trions [20], and found an approximate equivalence between these two definitions.Such an interaction can be understood as the large wavelength (small q) approximation to the electrostatic potentialof a charge inside a layer of thickness d and dielectric constant ε. Retaining the full q-dependence of the dielectricfunction ε(q) via an image-charge solution – as has been done in other works on quantum wells [21], inorganic-organicperovskites [22], and TMDs [23] – was not found to modify the results presented here. Using ab initio calculations

Page 11: arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 20142, a represen-tative member of the TMD family that includes MoS 2, MoSe 2, and WSe 2, all of which share similar properties with respect

11

1 2 3 4 5Quantum number n

1.9

2.0

2.1

2.2

2.3

2.4

2.5

Ene

rgy

(eV

)

εs = 2.1, Eg = 2.39

1 2 3 4 5Quantum number n

1.9

2.0

2.1

2.2

2.3

2.4

2.5

ExperimentTheory2D Hydrogen

εs = 3.9, Eg = 2.36

(b)(a)

FIG. S5. Predicted transition energies of monolayer WS2, explicitly accounting for screening due to the SiO2 substrate viaEq. S6. We have considered use of the relative permittivity evaluated at the optical frequency, εs = 2.1 (a) and evaluated atzero frequency, εs = 3.9 (b). In each case the band-gap of the theoretical result has been adjusted to give the best agreementwith experiment. The 2D hydrogen model result is unchanged from the result in the main text.

carried out based on density functional theory and the random phase approximation (DFT+RPA), as described indetail in Ref. 20, we would predict a screening length of r0 = 38 Å for intrinsic, suspended WS2. This value of r0is roughly half as large as that used in the present paper, and accordingly yields a larger exciton binding energy of0.50 eV, which is relevant for comparisons with fully ab initio calculations or with future measurements on suspendedsamples. A variational calculation gives a predicted lower bound on the WS2 trion binding energy of 26 meV (15meV) for the smaller (larger) value of r0, to be compared with our own preliminary experimental estimate of about30 meV.

We attribute the discrepancy in the screening length r0 to a combination of substrate and local-field effects, thelatter of which cannot be captured by the long wavelength theory outlined above. Specifically, we envisage an effectivehigher material polarizability at short- to mid-range length scales. Preliminary calculations based on a more detailedclassical electrostatic model and based on ab initio screening calculations of the full dielectric matrix support thishypothesis. Substrate screening effects can be approximately included in our theory with the modified interaction [18]

Veh(r) = −πe2

2r0

[H0

((1 + εs)r

2r0

)− Y0

((1 + εs)r

2r0

)], (S6)

where r0 is the screening length in the absence of a substrate and εs is the dielectric constant of the substrate. Thissubstrate screening has subtle, nontrivial effects on the effective electron-hole interaction: in addition to reducingthe overall strength of the interaction, the renormalization of the screening length reduces the range over whichthe potential exhibits a logarithmic form. We have calculated the binding energy using the above interaction, withtwo different possible dielectric constants for the SiO2 substrate: the relative permittivity at the optical frequency,εs = 2.1, and at zero frequency, εs = 3.9. The results are shown in Fig. S5, panels (a) and (b) respectively. Thescreening length r0 has been fixed to the ab initio calculated value of r0 = 38 Å, such that the band gap Eg is

the only free parameter. For these two cases, we find 1s exciton binding energies of 0.40 and 0.28 eV, which are inreasonable agreement with the value of 0.32 eV obtained in the main text. Both models underestimate the energy ofthe measured n = 5 transition, however this experimental data point carries the largest degree of uncertainty. Notably,each theoretically-predicted excitonic Rydberg series shown in Fig. S5 is still significantly non-hydrogenic. Togetherthese findings provide a strong indication that substrate effects can be partially accounted for by a phenomenologicallyincreased screening length as is done in the main text.

Compared to other theoretical treatments of excitons in TMDs, we point out that two other model-based studiesby Zhang et al. [23] and by Berghäuser and Malic [24] have employed our treatment of 2D dielectric behavior [20],which follows Keldysh [18] and Cudazzo et al. [19]. These two recent works are in good qualitative agreement withour own: the former focused on the ground-state exciton and found a binding energy of 0.28–0.33 eV, while the latteridentified the existence of an s-type excitonic Rydberg series, but did not comment on its non-hydrogenic behavior.Fully ab initio studies based on the GW+BSE approach [16, 17, 25, 26] do not require a model-based form for thedielectric function; they have found similar characteristics for the spectrum of excited states, although with higher

Page 12: arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 20142, a represen-tative member of the TMD family that includes MoS 2, MoSe 2, and WSe 2, all of which share similar properties with respect

12

overall binding energies.For the binding energy of bulk WS2, we employ the anisotropic exciton Hamiltonian

Hbulk =p2x + p2y2µxy

+p2z2µz

−e2

[εzεxy(x2 + y2) + ε2xyz

2]1/2 (S7)

which can be transformed via the change of variable z → (µz/µxy)1/2z into

H̃bulk =p2x + p2y + p2z

2µxy−

e2

[εzεxy(x2 + y2 + γz2)]1/2

(S8)

where the anisotropy parameter is defined as γ = εxyµxy/(εzµz). Using the values calculated by DFT+RPA (µxy =0.16 m0, µz = 1.2 m0, εxy = 13, εz = 6.3), we find γ ≈ 0.3. Although not analytically tractable, the eigenvaluesof this anisotropic Hamiltonian have been numerically calculated previously [27]. Using the results of Ref. 27 withγ = 0.3, we find a binding energy of 45 meV for bulk WS2.

Computational details

All single-particle electronic structure calculations (for the extraction of effective masses and RPA screening pa-rameters) were carried out for the experimental crystal structure of WS2 using the Quantum Espresso [28] andBerkeleyGW [29] software packages. DFT calculations were performed with 12 × 12 × 3 and 12 × 12 × 1 k-pointgrids for the bulk and monolayer, respectively, using the PBE exchange-correlation functional [30], norm-conservingpseudopotentials, and a plane-wave cutoff of 40 Ry (∼ 550 eV). Subsequent RPA calculations were utilized to extractthe dielectric properties, summing over approximately 50 unoccupied bands per layer.

[email protected][email protected]

[1] K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, Phys. Rev. Lett., 105, 136805 (2010).[2] T. Cao, G. Wang, W. Han, H. Ye, C. Zhu, J. Shi, Q. Niu, P. Tan, E. Wang, B. Liu, and J. Feng,

Nature Comm., 3, 887 (2012).[3] P. Tonndorf, R. Schmidt, P. Böttger, X. Zhang, J. Börner, A. Liebig, M. Albrecht, C. Kloc, O. Gordan, D. R. T. Zahn,

S. Michaelis de Vasconcellos, and R. Bratschitsch, Opt. Express, 21, 4908 (2013).[4] W. Zhao, Z. Ghorannevis, L. Chu, M. Toh, C. Kloc, P.-H. Tan, and G. Eda, ACS Nano, 7, 791 (2013).[5] K. F. Mak, K. He, C. Lee, G. H. Lee, J. Hone, T. F. Heinz, and J. Shan, Nature Mater., 11, 1 (2012).[6] J. S. Ross, S. Wu, H. Yu, N. J. Ghimire, A. M. Jones, G. Aivazian, J. Yan, D. G. Mandrus, D. Xiao, W. Yao, and X. Xu,

Nature Comm., 4, 1474 (2013).[7] J. Wilson and A. Yoffe, Adv. Phys., 18, 193 (1969).[8] H. R. Gutiérrez, N. Perea-López, A. L. Elías, A. Berkdemir, B. Wang, R. Lv, F. López-Urías, V. H. Crespi, H. Terrones,

and M. Terrones, Nano Lett., 13, 3447 (2012).[9] W. Zhao, Z. Ghorannevis, A. K. Kumar, J. R. Pang, M. Toh, X. Zhang, C. Kloc, P. H. Tan, and G. Eda, (2013),

arXiv:1304.0911.[10] C. Lee, H. Yan, L. E. Brus, T. F. Heinz, J. Hone, and S. Ryu, ACS Nano, 4, 2695 (2010).[11] E. Hecht, Optics, 4th ed. (Addison-Wesley, 2001).[12] S. Byrnes, Multilayer thin film optics calculator, http://sjbyrnes.com/ (2012).[13] I. H. Malitson, J. Opt. Soc. Am., 55, 1205 (1965).[14] M. Bass, C. DeCusatis, J. Enoch, V. Lakshminarayanan, G. Li, C. MacDonald, V. Mahajan, and E. V. Stryland, eds.,

Handbook of Optics, (vol. IV), 3rd ed. (McGraw-Hill, 2009).[15] D. Xiao, G.-B. Liu, W. Feng, X. Xu, and W. Yao, Phys. Rev. Lett., 108, 196802 (2012).[16] A. Ramasubramaniam, Phys. Rev. B, 86, 115409 (2012).[17] H. Shi, H. Pan, Y.-W. Zhang, and B. I. Yakobson, Phys. Rev. B, 87, 155304 (2013).[18] L. V. Keldysh, JETP Lett., 29, 658 (1979).[19] P. Cudazzo, I. V. Tokatly, and A. Rubio, Phys. Rev. B, 84, 085406 (2011).[20] T. C. Berkelbach, M. S. Hybertsen, and D. R. Reichman, Phys. Rev. B, 88, 045318 (2013).[21] E. Hanamura, N. Nagaosa, M. Kmagai, and T. Takagahara, Mat. Sci. Eng. B, 1, 255 (1988).[22] K. Tanaka, T. Takahashi, T. Kondo, T. Umebayashi, K. Asai, and K. Ema, Phys. Rev. B, 71, 045312 (2005).[23] C. Zhang, H. Wang, W. Chan, C. Manolatou, and F. Rana, Phys. Rev. B, 89, 205436 (2014).

Page 13: arXiv:1403.4270v3 [cond-mat.mtrl-sci] 2 Sep 20142, a represen-tative member of the TMD family that includes MoS 2, MoSe 2, and WSe 2, all of which share similar properties with respect

13

[24] G. Berghäuser and R. Malic, Phys. Rev. B, 89, 125309 (2014).[25] H.-P. Komsa and A. V. Krasheninnikov, Phys. Rev. B, 86, 241201(R) (2012).[26] D. Y. Qiu, F. H. da Jornada, and S. G. Louie, Physical Review Letters, 111, 216805 (2013).[27] A. Baldereschi and M. G. Diaz, Nuovo Cimento, 68B, 217 (1970).[28] P. Giannozzi, et al., J. Phys. Condens. Matter., 21, 395502 (2009).[29] J. Deslippe, G. Samzonidze, D. A. Strubbe, M. Jain, M. L. Cohen, and S. G. Louie, Comput. Phys. Commun., 183, 1269

(2012).[30] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett., 77, 3865 (1996).