8
Phonon induced luminescence decay in monolayer MoS 2 on SiO 2 /Si substrates Nihit Saigal and Sandip Ghosh a) Department of Condensed Matter Physics and Materials Science, Tata Institute of Fundamental Research, Mumbai 400005, India (Received 14 September 2015; accepted 7 December 2015; published online 16 December 2015) Exfoliated monolayer MoS 2 films on SiO 2 /Si substrates have been studied using photoluminescence (PL), Raman and reflectance contrast (RC) spectroscopies. With increase in temperature, the intensity of the two dominant PL spectral features A and D, attributed to A exciton/trion and to defects, seemingly decay in an activated fashion with an energy 50 meV, which is close to the energies of E 1 2g and A 1g phonons. Comparison of absorption spectrum derived from RC with circular polarization resolved PL spectrum suggests that both D and A emissions are associated with bound excitons, the A emission involving relatively weakly localized ones. The PL decay behaviour is explained using a phenomenological model where non-radiative loss of excitons is determined by the number of excited phonon modes. This corroborates the recent finding of strong A exciton and A 1g phonon coupling in monolayer MoS 2 . V C 2015 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4938141] Monolayer MoS 2 , unlike graphene, has a finite bandgap making it more suitable for electronic and opto-electronic device applications. 1 In monolayer MoS 2 the indirect-to- direct bandgap transition, 2 which enables efficient lumines- cence, and the coupling of valley and spin degree of free- dom 3,4 of carriers together are the most promising properties for spin based devices. An appropriate circularly polarized light can selectively populate either up or down spin states in monolayer MoS 2 , which in turn can be detected from the hel- icity of the luminescence when carriers from these states recombine radiatively. 5 Luminescence has also played a role in unraveling the presence of stable trions 6 in monolayer MoS 2 and is relevant for conventional light emitting device (LED) applications. 7,8 Understanding luminescence proper- ties of monolayer MoS 2 , including its decay with tempera- ture, is therefore important for both basic physics research and practical applications. The luminescence intensity normally decays from its value at low temperature when temperature is increased because the net non-radiative recombination lifetime of the excited carriers becomes smaller. 9 In group III-V semicon- ductors, which are used in all commercial LEDs and lasers, the luminescence at low temperature is dominated by radia- tive recombination of excitons which have relatively low binding energies (4 meV in GaAs and 26 meV in GaN), while at room temperature the excitons are ionized and free carrier recombination is more likely. In contrast, the exciton binding energy E ex b in monolayer MoS 2 is quite large, with experiments suggesting values ranging from 440 meV to 570 meV. 10,11 Thus luminescence from monolayer MoS 2 is likely to be dominated by exciton recombination even at room temperature where k BT ¼ 26 meV. For a monolayer in contact with a substrate, interaction with the substrate is important. This is known to influence properties like carrier mobility 12 and also luminescence throughput. 4,13 SiO 2 /Si is the most commonly used substrate for monolayer 2-dimensional materials since it aids visibility of the film and provides for back-gating through an insulator in many device configurations. Here we present results of a tempera- ture dependent photoluminescence (PL) decay study of monolayer MoS 2 on SiO 2 /Si. We observe some interesting similarities in the luminescence properties of the previously assumed free exciton/trion transitions and those related to defects. We provide a phenomenological model to describe the luminescence decay, which invokes the recently reported observation of strong exciton-phonon coupling 14 in mono- layer MoS 2 . Experiments on bilayer MoS 2 on SiO 2 /Si and sapphire were also performed to test our hypothesis. The monolayer and bilayer MoS 2 films were obtained by mechanical exfoliation from natural crystals and trans- ferred on to both SiO 2 (280 nm) covered Si(001) and sapphire(0001) substrates, using poly (dimethyl)-siloxane (PDMS) stamps. 15 Mono/bilayer verification was done using micro-Raman, circular polarization resolved micro-PL and also micro-reflectance contrast 16 (RC) spectroscopies. A spa- tial resolution of 1:5 lm was achieved by using a setup that has a 0.55 m focal length monochromator coupled to a microscope built using a long working distance (WD) objec- tive (magnification 100, numerical aperture 0.5, and WD ¼ 12 mm). PL was excited using either a 633 nm (1.96 eV) or a 532 nm (2.33 eV) laser with up to 2 mW power and detected using a thermo-electric cooled Si CCD. For cir- cularly polarized PL excitation, we used a Glan-Taylor (GT) polarizer with a Soleil-Babinet compensator and for detec- tion we used a GT polarizer and broadband liquid-crystal retarder combination. Raman measurements used an extra sharp edge cut-off filter at the detection end to block the elastic-scattered laser. For RC spectroscopy we used a 75 W Xe lamp, photo-multiplier tube detection, and a pinhole at the field-stop region of the microscope illumination path to expose only a small spot on the sample. The samples were cooled using a continuous flow liquid He cryostat. a) Electronic mail: [email protected] 0003-6951/2015/107(24)/242103/4/$30.00 V C 2015 AIP Publishing LLC 107, 242103-1 APPLIED PHYSICS LETTERS 107, 242103 (2015) This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 158.144.58.149 On: Mon, 21 Dec 2015 07:39:35

Phonon induced luminescence decay in monolayer MoS2 on ...sangho10/Publications/15APL-ns2.pdfPhonon induced luminescence decay in monolayer MoS 2 on SiO 2/Si substrates Nihit Saigal

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Phonon induced luminescence decay in monolayer MoS2 on ...sangho10/Publications/15APL-ns2.pdfPhonon induced luminescence decay in monolayer MoS 2 on SiO 2/Si substrates Nihit Saigal

Phonon induced luminescence decay in monolayer MoS2 on SiO2/Sisubstrates

Nihit Saigal and Sandip Ghosha)

Department of Condensed Matter Physics and Materials Science, Tata Institute of Fundamental Research,Mumbai 400005, India

(Received 14 September 2015; accepted 7 December 2015; published online 16 December 2015)

Exfoliated monolayer MoS2 films on SiO2/Si substrates have been studied using photoluminescence

(PL), Raman and reflectance contrast (RC) spectroscopies. With increase in temperature, the

intensity of the two dominant PL spectral features A and D, attributed to A exciton/trion and to

defects, seemingly decay in an activated fashion with an energy �50 meV, which is close to the

energies of E12g and A1g phonons. Comparison of absorption spectrum derived from RC with circular

polarization resolved PL spectrum suggests that both D and A emissions are associated with bound

excitons, the A emission involving relatively weakly localized ones. The PL decay behaviour is

explained using a phenomenological model where non-radiative loss of excitons is determined by

the number of excited phonon modes. This corroborates the recent finding of strong A exciton and

A1g phonon coupling in monolayer MoS2. VC 2015 AIP Publishing LLC.

[http://dx.doi.org/10.1063/1.4938141]

Monolayer MoS2, unlike graphene, has a finite bandgap

making it more suitable for electronic and opto-electronic

device applications.1 In monolayer MoS2 the indirect-to-

direct bandgap transition,2 which enables efficient lumines-

cence, and the coupling of valley and spin degree of free-

dom3,4 of carriers together are the most promising properties

for spin based devices. An appropriate circularly polarized

light can selectively populate either up or down spin states in

monolayer MoS2, which in turn can be detected from the hel-

icity of the luminescence when carriers from these states

recombine radiatively.5 Luminescence has also played a role

in unraveling the presence of stable trions6 in monolayer

MoS2 and is relevant for conventional light emitting device

(LED) applications.7,8 Understanding luminescence proper-

ties of monolayer MoS2, including its decay with tempera-

ture, is therefore important for both basic physics research

and practical applications.

The luminescence intensity normally decays from its

value at low temperature when temperature is increased

because the net non-radiative recombination lifetime of the

excited carriers becomes smaller.9 In group III-V semicon-

ductors, which are used in all commercial LEDs and lasers,

the luminescence at low temperature is dominated by radia-

tive recombination of excitons which have relatively low

binding energies (4 meV in GaAs and 26 meV in GaN),

while at room temperature the excitons are ionized and free

carrier recombination is more likely. In contrast, the exciton

binding energy Eexb in monolayer MoS2 is quite large, with

experiments suggesting values ranging from 440 meV to

570 meV.10,11 Thus luminescence from monolayer MoS2 is

likely to be dominated by exciton recombination even at

room temperature where kBT¼ 26 meV. For a monolayer in

contact with a substrate, interaction with the substrate is

important. This is known to influence properties like carrier

mobility12 and also luminescence throughput.4,13 SiO2/Si

is the most commonly used substrate for monolayer

2-dimensional materials since it aids visibility of the film

and provides for back-gating through an insulator in many

device configurations. Here we present results of a tempera-

ture dependent photoluminescence (PL) decay study of

monolayer MoS2 on SiO2/Si. We observe some interesting

similarities in the luminescence properties of the previously

assumed free exciton/trion transitions and those related to

defects. We provide a phenomenological model to describe

the luminescence decay, which invokes the recently reported

observation of strong exciton-phonon coupling14 in mono-

layer MoS2. Experiments on bilayer MoS2 on SiO2/Si and

sapphire were also performed to test our hypothesis.

The monolayer and bilayer MoS2 films were obtained

by mechanical exfoliation from natural crystals and trans-

ferred on to both SiO2 (�280 nm) covered Si(001) and

sapphire(0001) substrates, using poly (dimethyl)-siloxane

(PDMS) stamps.15 Mono/bilayer verification was done using

micro-Raman, circular polarization resolved micro-PL and

also micro-reflectance contrast16 (RC) spectroscopies. A spa-

tial resolution of �1:5 lm was achieved by using a setup that

has a 0.55 m focal length monochromator coupled to a

microscope built using a long working distance (WD) objec-

tive (magnification 100�, numerical aperture 0.5, and

WD¼ 12 mm). PL was excited using either a 633 nm

(1.96 eV) or a 532 nm (2.33 eV) laser with up to 2 mW power

and detected using a thermo-electric cooled Si CCD. For cir-

cularly polarized PL excitation, we used a Glan-Taylor (GT)

polarizer with a Soleil-Babinet compensator and for detec-

tion we used a GT polarizer and broadband liquid-crystal

retarder combination. Raman measurements used an extra

sharp edge cut-off filter at the detection end to block the

elastic-scattered laser. For RC spectroscopy we used a 75 W

Xe lamp, photo-multiplier tube detection, and a pinhole at

the field-stop region of the microscope illumination path to

expose only a small spot on the sample. The samples were

cooled using a continuous flow liquid He cryostat.a)Electronic mail: [email protected]

0003-6951/2015/107(24)/242103/4/$30.00 VC 2015 AIP Publishing LLC107, 242103-1

APPLIED PHYSICS LETTERS 107, 242103 (2015)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

158.144.58.149 On: Mon, 21 Dec 2015 07:39:35

admin
Highlight
admin
Sticky Note
k_{B}T
Page 2: Phonon induced luminescence decay in monolayer MoS2 on ...sangho10/Publications/15APL-ns2.pdfPhonon induced luminescence decay in monolayer MoS 2 on SiO 2/Si substrates Nihit Saigal

Figure 1 shows the Raman spectrum of a monolayer

MoS2 film on SiO2/Si with its optical image as an inset. The

two dominant phonon modes E12g and A1g have a separation

of 18 cm�1 indicating that it is a monolayer film.17 Figure 2

shows the circular polarization resolved PL spectrum of the

sample with 633 nm (1.96 eV) excitation. One sees two

prominent emission features peaked around 1.91 eV and

1.82 eV, labeled A and D, respectively. The feature A is

associated with unresolved A exciton and trion transitions at

the K-point of the Brillouin zone of monolayer MoS2.2,6 The

feature D is associated with defects whose origin has

been suggested to be either sulphur vacancies18 or surface

adsorbates.19 With increase in laser power, the D emission

intensity tends to saturate, indicating its association with

defects.20 For right circularly polarized (RCP) excitation

near resonant in energy with A, the emission is also predomi-

nantly RCP. The circular polarization anisotropy defined in

terms of intensity of right (IRCP) and left (ILCP) circularly

polarized emission as ½IRCP � ILCP�=½IRCP þ ILCP� reaches

>70% as shown in the inset of Fig. 2. This anisotropy arises

due to the coupling of spin and valley degrees of freedom in

monolayer MoS2 whereby RCP excitation can excite only

carriers with up spin in the K0 valley (and not down spin in

the K valley). These carriers, in the absence of momentum

and spin scattering, recombine to give out RCP emission.3,4

This effect is lost in non-monolayer MoS2 and therefore the

observation of this large polarization anisotropy is taken as

evidence for a monolayer film.

Figure 3 shows unpolarized PL spectra of (a) a monolayer

MoS2 film on SiO2/Si and (b) bilayer MoS2 film on sapphire

at temperatures between 30 K and 273 K for 2.33 eV excita-

tion. Here one also observes a weak feature at higher energies

due to the B exciton whose origin has been attributed to hot

luminescence.4 Figure 4 shows Arrhenius plots of PL intensity

decay with increase in temperature for the A, D, and B fea-

tures in Fig. 3(a) and A feature in Fig. 3(b). For this, the PL

spectra were fitted with a sum of Gaussians whose areas deter-

mine the PL intensity.20 The PL intensity (IPL) decay at high

temperatures looks like an Arrhenius type activated behaviour

given by IPL ¼ const:� eEa=kBT , where Ea is the activation

energy. The dashed line in Fig. 4 represents an activated decay

with slope corresponding to Ea¼ 50 meV. The data for A

(both mono/bilayer) and D run roughly parallel to it at high

temperatures, indicating Ea � 50 meV for both emissions.

The observation that both emissions have similar Ea is

FIG. 1. Micro-Raman spectrum of monolayer MoS2 on SiO2/Si substrate

whose optical image is shown as an inset. The two dominant monolayer

MoS2 phonon modes E12g and A1g are indicated along with the corresponding

atomic displacements.

FIG. 2. Circular polarization resolved micro-PL spectrum of monolayer

MoS2 on SiO2/Si, for right-circularly polarized 633 nm (1.96 eV) excitation.

The inset shows the circular polarization anisotropy of emission as a func-

tion of energy.

FIG. 3. Temperature dependence of the micro-PL spectrum of (a) monolayer

MoS2 on SiO2/Si and (b) bilayer (BL) MoS2 on sapphire for unpolarized

532 nm (2.33 eV) excitation. The inset in (b) is an optical image of the BL

sample.

FIG. 4. Dependence of PL intensity on temperature for a monolayer MoS2

on SiO2/Si [A, D, and B features in Fig. 3(a)] and a bilayer MoS2 on sap-

phire [A feature in Fig. 3(b)]. Square symbols represent bilayer data. The

dashed line indicates an Arrhenius decay with activation energy 50 meV.

The continuous lines are simulations using Eq. (1) with Ephn¼ 50 meV. The

inset shows Ephn estimated from different samples by fitting an Arrhenius

equation to only the high temperature tail part of the data.

242103-2 N. Saigal and S. Ghosh Appl. Phys. Lett. 107, 242103 (2015)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

158.144.58.149 On: Mon, 21 Dec 2015 07:39:35

Page 3: Phonon induced luminescence decay in monolayer MoS2 on ...sangho10/Publications/15APL-ns2.pdfPhonon induced luminescence decay in monolayer MoS 2 on SiO 2/Si substrates Nihit Saigal

surprising given their different origins and a �90 meV differ-

ence in their emission peak positions at low temperatures.

Since the A emission is known to be associated with

excitons, the similarity in the PL decay characteristics of D

and A emission suggests that the D emission too may be

associated with excitons but which are strongly bound to

some defect. This is likely also because the large Eexb in

monolayer MoS2 makes excitons relatively stable. It has

been suggested that depolarization in case of circular polar-

ization resolved PL studies can occur due to the Bir-Aranov-

Pikus electron-hole exchange mechanism with simultaneous

spin flip and momentum scattering between K and K0 val-

leys.4 Exciton localization, with consequent spread in crystal

momentum values, would make scattering between K and K0

valleys easier, thereby weakening the circular polarization

anisotropy of emission. Another approach suggests defect

induced disorder as giving rise to depolarization.21 The

observed absence of polarization anisotropy of the D emis-

sion is therefore likely to be another signature of emission

from strongly localized excitons in the presence of defects.

We next try to understand the similarity in the PL decay

behaviour of A and D emissions.

PL decay with increase in temperature can occur when

carriers which would have recombined radiatively are

instead excited to states from where recombination via non-

radiative pathways becomes more favourable. For instance

PL due to free excitons will decay if they can be separated

into electrons and holes which then can recombine non-

radiatively via traps, in which case Ea would be the same as

Eexb . However Eex

b in monolayer MoS2 mentioned earlier, is

much larger than 50 meV. In case of PL from carriers recom-

bining in confined systems, such as quantum wells22 (QW)

or from carriers bound to defects,23 if sufficient energy is

provided to make them free whereby they are delocalized in

the larger crystal, then they are more likely to encounter

traps/defects that facilitate non-radiative recombination. In

such cases, Ea corresponds to the energy required to make

the entity free. For example in QWs Ea is found to be the

energy required to excite the ground state electron-hole pair

in the well to the barrier region and so is equal to the differ-

ence between the QW PL emission peak and the barrier

bandgap energy.24 To check for the signature of such a

“free” state 50 meV above A and D emission peaks, we

measured the monolayer MoS2 sample’s absorption spec-

trum using RC spectroscopy. We used a Gaussian convolved

Lorentz oscillator model for exciton/trion dielectric function

in a transfer matrix calculation20 to simulate the RC spec-

trum and then varied its parameters to fit the measured RC

spectrum (inset of Fig. 5). The absorption spectrum thus

obtained is shown in Fig. 5 where features due to the ground

state exciton and trion transitions are now partially resolved.

Comparison with the PL spectrum plotted alongside shows

that the exciton (trion) absorption peak lies 60 meV

(30 meV) above the A emission peak and also there is no fea-

ture in absorption 50 meV above the D emission. This sug-

gests that Ea is perhaps not associated with thermal escape

mechanisms of the type discussed earlier, which has also

been suggested before.25 The other point to note is that the

lower energy of the A emission peak compared with exciton/

trion absorption peaks indicates that the A emission itself is

not from free exciton/trion but from bound ones. The larger

width of the emission feature compared with the absorption

feature also suggests this, since the different localizing envi-

ronments from where the exciton recombines adds to the dis-

tribution of emission energies. There is also evidence for

localization of carriers from electrical measurements.26

However, localization of the excitons contributing to A is

weak enough not to destroy completely the spin-valley cou-

pling and therefore preserves circular polarization anisot-

ropy, unlike the D emission.

Now note that Ea � 50 meV for the activation like decay

of both A and D emissions also happens to coincide with the

energies of the dominant phonon modes E12g (48 meV) and A1g

(50 meV) in monolayer MoS2, as seen in Fig. 1. Also, the

above analysis suggested that both emissions, although at dif-

ferent energies, involve excitons. Based on a resonant Raman

scattering experiment, it has been shown very recently14 that

there is a strong coupling of the A1g phonon with A excitons in

monolayer MoS2. This lattice vibration mode involves dis-

placement of the sulphur atoms perpendicular to the mono-

layer film plane, thereby increasing the scope for interaction

with the substrate. It was already mentioned that the choice of

substrates affects the PL throughput.4,13 These points together

lead us to suggest that perhaps PL decay here occurs due to

increased phonon induced interaction of the excitons with the

substrate, including impurities/traps therein, thereby providing

a path for them to break up and recombine non-radiatively.

Such a decay mechanism is also likely in a few-layer film and

on different substrates. The similar decay behaviour of PL

from the bilayer MoS2 film on sapphire substrate supports this

hypothesis. Our PL decay model implies that larger the num-

ber of phonons (nphn) excited, more would be the rate of non-

radiative recombination. We therefore express the lifetime for

non-radiative recombination as sNR ¼ sNRo=nphn, where sNRo

is a proportionality constant. The role of nphn in determining

optical signal magnitude is known for instance in case of tem-

perature dependence of Raman spectra.27 Here 1=sNRo can be

interpreted as the non-radiative decay rate induced per excited

phonon and includes exciton-phonon interaction strength.

Bose-Einstein statistics gives the average phonon occupation

number at temperature T as nphn ¼ 1=½eEphn=kBT � 1�, where

Ephn is the phonon mode energy. If DN number of excitons

are generated under steady state laser excitation, then

FIG. 5. Absorption spectrum of a monolayer MoS2 film on SiO2/Si showing

partially resolved exciton (Ex) and trion (Tr) related features. The unpolar-

ized PL spectrum of the sample is plotted alongside. The inset shows the

measured and simulated micro-reflectance contrast spectrum, which was

used for estimating the absorption spectrum.

242103-3 N. Saigal and S. Ghosh Appl. Phys. Lett. 107, 242103 (2015)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

158.144.58.149 On: Mon, 21 Dec 2015 07:39:35

Page 4: Phonon induced luminescence decay in monolayer MoS2 on ...sangho10/Publications/15APL-ns2.pdfPhonon induced luminescence decay in monolayer MoS 2 on SiO 2/Si substrates Nihit Saigal

the PL intensity IPL would be proportional to the ratio

½DN=sR�=½DN=sT �, where sR is the radiative lifetime and sT ¼½1=sNR þ 1=sR��1

is the net recombination lifetime.9 Putting

these together we get

IPL ¼ Io=½1þ c=ðeEphn=kBT � 1Þ�; (1)

where c ¼ sR=sNRo and Io is a constant representing the PL

intensity at the lowest temperature. The continuous lines in

Fig. 4 are simulated curves using Eq. (1) with Ephn¼ 50 meV

for both A and D emissions, with c values being 61 and 409,

respectively, for the monolayer sample. The simulations

match the experimental data fairly well. Note that Eq. (1)

is not identical to an activated behaviour of the type seen

in case of defect23 PL, but reduces to its form when

eEphn=kBT � 1.20 Depending on the c and Ephn values, over a

certain high temperature range (typically 125 K–275 K here)

Eq. (1) can be further approximated20 to an Arrhenius form

IPL ’ const:� eEphn=kBT . The Ephn values obtained by fitting

such an equation to the high temperature tail data for differ-

ent monolayer (A and D features) and bilayer (A feature)

samples on SiO2/Si and sapphire substrates are shown as an

inset in Fig. 4. Their average gives Ephn ’ 47:263 meV,

close to 50 meV used in the simulations. The higher value of

c for D compared with A, as mentioned earlier, implies that

the non-radiative decay rate for the relatively strongly local-

ized excitons is higher. This may be expected since increased

exciton-phonon coupling, which according to our model,

increases non-radiative decay rate, with increase in exciton

localization has been observed in other low dimensional sys-

tems such as carbon nanotubes.28 The decay of the mono-

layer B emission in Fig. 4 was also simulated using Eq. (1)

with Ephn¼ 50 meV; however, its decay range is too small to

test the correctness of Ephn value. Nevertheless, the estimated

small c ¼ 4:6 for the B emission suggests that the coupling

of A1g phonons with the B exciton in monolayer MoS2 is rel-

atively weaker. This is indeed observed in the resonant

Raman scattering study on monolayer MoS2.14

In conclusion, our results suggest that the two dominant

features in the PL spectrum of monolayer MoS2 on SiO2/Si

arise due to weakly localized exciton/trions and excitons

strongly localized at defects. At high temperatures, the decay

of both PL features resembles activation like behaviour with

similar activation energies. To explain this, we provided a

phenomenological model which invokes the role of certain

phonon modes of monolayer MoS2 in increasing the exciton-

substrate interaction and thereby leading to non-radiative

loss of excitons. Such a PL decay mechanism may be generic

to 2D monolayer materials given the inevitability of interac-

tion with the substrate in these systems.

The authors acknowledge the help of A. Mukherjee, V.

Sugunakar, and A. Raut in building the high spatial

resolution spectroscopy setup.

1Q. H. Wang, K. K. Zadeh, A. Kis, J. N. Coleman, and M. S. Strano, Nat.

Nanotechnol. 14, 699 (2012).2A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C. Y. Chim, G. Galli, and

F. Wang, Nano Lett. 10, 1271 (2010).3T. Cao, G. Wang, W. Han, H. Ye, C. Zhu, J. Shi, Q. Niu, P. Tan, E. Wang,

B. Liu, and J. Feng, Nat. Commun. 3, 887 (2012).4K. F. Mak, K. He, J. Shan, and T. F. Heinz, Nat. Nanotechnol. 7, 494

(2012).5X. Xu, W. Yao, D. Xiao, and T. F. Heinz, Nat. Phys. 10, 343 (2014).6K. F. Mak, K. He, C. Lee, G. H. Lee, J. Hone, T. F. Heinz, and J. Shan,

Nat. Mater. 12, 207 (2012).7R. S. Sundaram, M. Engel, A. Lombardo, R. Krupke, A. C. Ferrari, Ph.

Avouris, and M. Steiner, Nano Lett. 13, 1416 (2013).8F. Withers, O. D. Pozo-Zamudio, A. Mishchenko, A. P. Rooney, A.

Gholinia, K. Watanabe, T. Taniguchi, S. J. Haigh, A. K. Geim, A. I.

Tartakovskii, and K. S. Novoselov, Nat. Mater. 14, 301 (2015).9P. T. Landsberg, Phys. Status Solidi B 41, 457 (1970).

10H. M. Hill, A. F. Rigosi, C. Roquelet, A. Chernikov, T. C. Berkelbach, D.

R. Reichman, M. S. Hybertsen, L. E. Brus, and T. F. Heinz, Nano Lett. 15,

2992 (2015).11A. R. Klots, A. K. M. Newaz, B. Wang, D. Prasai, H. Krzyzanowska, J.

Lin, D. Caudel, N. J. Ghimire, J. Yan, B. L. Ivanov, K. A. Velizhanin, A.

Burger, D. G. Mandrus, N. H. Tolk, S. T. Pantelides, and K. I. Bolotin,

Sci. Rep. 4, 6608 (2014).12B. Radisavljevic and A. Kis, Nat. Mater. 12, 815 (2013).13M. Buscema, G. A. Steele, H. S. J. van der Zant, and A. Castellanos-

Gomez, Nano Res. 7, 561 (2014).14B. R. Carvalho, L. M. Malard, J. M. Alves, C. Fantini, and M. A. Pimenta,

Phys. Rev. Lett. 114, 136403 (2015).15A. Castellanos-Gomez, N. Agrait, and G. Rubio-Bollinger, Appl. Phys.

Lett. 96, 213116 (2010).16N. Saigal, A. Mukherjee, V. Sugunakar, and S. Ghosh, Rev. Sci. Instrum.

85, 073105 (2014).17C. Lee, H. Yan, L. E. Brus, T. F. Heinz, J. Hone, and S. Ryu, ACS Nano 4,

2695 (2010).18S. Tongay, J. Suh, C. Ataca, W. Fan, A. Luce, J. S. Kang, J. Liu, C. Ko, R.

Raghunathanan, J. Zhou, F. Ogletree, J. Li, J. C. Grossman, and J. Wu,

Sci. Rep. 3, 2657 (2013).19G. Plechinger, F. X. Schrettenbrunner, J. Eroms, D. Weiss, C. Sch€uller,

and T. Korn, Phys. Status Solidi RRL 6, 126 (2012).20See supplementary material at http://dx.doi.org/10.1063/1.4938141 for PL

lineshape, decay equation and reflectance contrast lineshape analysis.21Q. Wang, S. Ge, X. Li, J. Qiu, Y. Ji, J. Feng, and D. Sun, ACS Nano 7,

11087 (2013).22S. Weber, W. Limmer, K. Thonke, R. Sauer, K. Panzlaff, G. Bacher, H. P.

Meier, and P. Roentgen, Phys. Rev. B 52, 14739 (1995).23E. H. Bogardus and H. B. Bebb, Phys. Rev. 176, 993 (1968).24S. Ghosh, B. M. Arora, S. J. Kim, J. H. Noh, and H. Asahi, J. Appl. Phys.

85, 2687 (1999).25T. Korn, S. Heydrich, M. Hirmer, J. Schmutzler, and C. Sch€uller, Appl.

Phys. Lett. 99, 102109 (2011).26S. Ghatak, A. N. Pal, and A. Ghosh, ACS Nano 5, 7707 (2011).27C. Kittel, Introduction to Solid State Physics, 5th ed. (Wiley, New York,

1976).28V. Ardizzone, Y. Chassagneux, F. Vialla, G. Delport, C. Delcamp, N.

Belabas, E. Deleporte, Ph. Roussignol, I. Robert-Philip, C. Voisin, and J.

S. Lauret, Phys. Rev. B 91, 121410(R) (2015).

242103-4 N. Saigal and S. Ghosh Appl. Phys. Lett. 107, 242103 (2015)

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

158.144.58.149 On: Mon, 21 Dec 2015 07:39:35

Page 5: Phonon induced luminescence decay in monolayer MoS2 on ...sangho10/Publications/15APL-ns2.pdfPhonon induced luminescence decay in monolayer MoS 2 on SiO 2/Si substrates Nihit Saigal

1

Supplementary Information

Manuscript Title: Phonon induced luminescence decay in monolayer MoS2 on SiO2/Si substratesAuthors: Nihit Saigal and Sandip GhoshAffiliation: Department of Condensed Matter Physics and Materials Science, Tata Institute of Fundamental Re-search, Mumbai 400005, India

I: Photoluminescence (PL) lineshape fittingThe PL spectral lineshape IPL(E) is expected to be asymmetric in general.[1] However in the present case because ofthe large inhomogeneous broadening we considered a sum of normalized Gaussians as the lineshape, given by

IPL(E) =∑j

Aj2

∆Ej

√ln 2

πexp[−4 ln 2

(E − Eoj

∆Ej

)2

]. (S1)

Here E is the photon energy, Eoj and ∆Ej are the peak energy and full-width at half-maximum of the jth spectralfeature, respectively. The parameter Aj directly gives the area under the curve for the particular emission featureand is taken as the measure of its relative strength. The fairly good fits obtained with such a lineshape, as shown inFig. S1(a)-(f), indicates that this approximation is reasonable. Figure S1 also shows data for 632 nm excitation forwhich the temperature dependence of PL decay yields similar Ephn values as for 532 nm excitation. This similaritysuggests that temperature dependent changes in the joint density of states does not play a dominant role in thedecay of PL signal in the high temperature region which determines Ephn. A reliable estimate of D emission strengthbecomes difficult above 175 K since it becomes too weak and comparable in strength to the tail of A emission. Wetherefore did not consider D emission data above 175 K for the PL decay analysis. The measured spectra shown hereand others used in the analysis were all corrected for the spectral response of our measurement setup determinedseparately by measuring the spectrum of a black-body radiation source (a tiny 0.3 W tungsten lamp) kept at thesample’s position.

II: Pump power dependence of Photoluminescence intensityFigure S2(a) shows that with increase in pump laser (532 nm) power the D emission intensity tends to saturate whileA does not. This indicates D’s association with defect states whose density would be much lower than the densityof states of monolayer MoS2. Figure S2(b) shows only a negligible blue shift of the D emission peak with increasein laser power because it is the broad defect energy distribution which determines the spectral lineshape. The Aemission peak redshifts with increase in laser power and a similar behaviour reported earlier was attributed to local

PL si

gnal

cou

nt(x

103 )

0

5

10

15

0.0

1.0

2.0

0.0

0.5

Energy (eV)1.6 1.7 1.8 1.9 2.0 2.1

PL si

gnal

cou

nt(x

103 )

0

5

10

Energy (eV)1.6 1.7 1.8 1.9 2.0 2.1

0.0

1.0

2.0

Energy (eV)1.6 1.7 1.8 1.9 2.0 2.1

0.0

1.0

(a) 30K 633nm

(b) 164K 633nm

(c) 230K 633nm

(f) 230K 532nm

(e) 164K 532nm

(d) 30K 532nm

B

A

A

D

D

Figure S1 PL spectrum of monolayer MoS2 on SiO2/Si obtained using (a)-(c) 633 nm and (d)-(f) 532nm excitation, fittedwith a sum of normalized Gaussians. The grey dashed lines represent the individual fitted A, D and B components while theblack line is their sum.

admin
Sticky Note
x-scale was wrong for this plot in the file on the AIP website. It has been corrected here.
Page 6: Phonon induced luminescence decay in monolayer MoS2 on ...sangho10/Publications/15APL-ns2.pdfPhonon induced luminescence decay in monolayer MoS 2 on SiO 2/Si substrates Nihit Saigal

2

Sample MM4 MoS2 monolayer

Excit 532 nm (1.962 eV)

Laser power (mW)0.0 0.5 1.0 1.5 2.0

Inte

grat

ed P

L si

gnal

0

500

1000

1500

2000

A

D

Laser power (mW)0.0 0.5 1.0 1.5 2.0

Tran

sitio

n en

ergy

(eV

)

1.82

1.84

1.86

1.88

1.90

A

D

Figure S2 Variation of (a) intensity and (b) emission peak energy of A and D PL features in a monolayer MoS2 film onSiO2/Si at 4.5 K, with increase in pump laser (532 nm) power. The lines in (a) are fitted polynomials to guide the eye.

heating.[2] There are also reports that show transition from exciton to trion dominated emission in monolayer MoS2

with change in doping levels[3] which is also possible here locally with increase in laser power.

III: Photoluminescence decay aspectsIn the conventional model of activated PL decay[4, 5] with increase in temperature, carriers which would have recom-bined radiatively are instead thermally excited to some higher energy state from where their non-radiative decay, viadefects or trap levels, becomes more favourable. The states to which the electron/hole [or both] need to be excited tofor non-radiative decay to follow are at an energy [sum total energy] Ea higher than the PL emission peak energy. Thenon-radiative lifetime τNR is then expressed as τNR = τNR∞ eEa/kBT , which implies that as temperature increasesτNR decreases to the limiting value τNR∞. The PL intensity IPL is proportional to the ratio [∆N/τR]/[∆N/τT ],where ∆N in our case represents the number of exciton/trion generated, τR their radiative recombination lifetimeand τT = [1/τNR + 1/τR]−1 is the net recombination lifetime. This gives

IPL ∝ 1/[1 + τR/τNR] = Io/[1 + (τR/τNR∞) e−Ea/kBT ] (S2)

where Io is again a constant representing the PL intensity at the lowest temperature. The Eq.1 of the manuscriptunder the condition eEphn/kBT � 1 (which holds up till close to room temperature where kBT = 26 meV, forEphn = 50 meV) becomes

IPL = Io/[1 + γe−Ephn/kBT ]. (S3)

This has the same form as Eq. S2 above if we take τR/τNR∞ = γ′ as a constant. This explains why Eq.1 of themanuscript has a temperature dependence that looks similar to the conventional activated PL decay model with an

1/Temperature (K-1)0.01 0.02

Inte

grat

ed P

L si

gnal

10-2

10-1

100

γ=1000

γ=100

Ιο=1 Ephn=50meV

Temperature (K)50100300

γ=10

1/Temperature (K-1)0.002 0.004 0.006 0.008 0.010

Inte

grat

ed P

L si

gnal

101

102

103

A (BL) x5

Temperature (K)

D (ML)

100150200250300200

(a) (b)

A (ML)

Figure S3 (a) Simulated PL decay plots (continuous lines) for different values of γ in Eq. 1 of the manuscript with

Ephn = 50 meV. The dashed lines represent a simple Arrhenius decay of the form IPL = const.× eEphn/kBT , withEphn = 50 meV. (b) The simple Arrhenius decay equation fitted (lines) to the high temperature tail of several measured(symbol) monolayer (ML) and bilayer (BL) MoS2 PL decay curves.

Page 7: Phonon induced luminescence decay in monolayer MoS2 on ...sangho10/Publications/15APL-ns2.pdfPhonon induced luminescence decay in monolayer MoS 2 on SiO 2/Si substrates Nihit Saigal

3

activation energy the same as Ephn. The Ephn value essentially determines the slope of the decay curve at hightemperatures. Our simulation and measurement do not match well around 100 K in Fig. 4 of the manuscript. Toimprove the fitting of activated PL decay data at low temperatures, different modified versions of Eq. S2 have beenused.[6] These mostly involve multiplying a Tn term to account for temperature dependence of the carrier capturecross-section (related to γ′ in our case) or adding a second e−Ea2/kBT term to account for a second decay pathway, inthe denominator of Eq. S2 to better fit the data at low temperatures. However our concern here is mainly the hightemperature region around room temperature where PL decays rapidly.

Figure S3(a) shows simulated PL decay plots (continuous lines) using Eq. 1 of the manuscript with Ephn = 50 meVand different values of the parameter γ. The dashed lines indicate that on such a Log(IPL) Vs 1/T plot over a certainhigh temperature range depending on the γ value, the PL decay can be approximated by a line representing a simpleArrhenius form IPL = const. × eEphn/kBT . In our experimental data this linear region starts at temperatures above125 K. It is known that optical phonon scattering dominates the carrier scattering process in monolayer MoS2 attemperatures higher than 100 K while acoustic phonons are relevant only at low temperatures.[7, 8] This thereforealso ties in with our suggestion that optical phonons (A1g) play a primary role in the non-radiative loss of excitonsleading to the decay of A and D emissions at high temperature. We fitted the simple Arrhenius form to the hightemperature PL decay tail of 10 different A and D spectral features on 5 different sample combinations which includedmonolayers or bilayers on SiO2/Si or sapphire substrate. These fits are shown in Fig. S3(b) and the values of Ephn

obtained from this exercise are plotted in the inset of Fig. 4 of the manuscript.

IV: Reflectance Contrast spectral analysisTo extract the exciton/trion absorption spectrum from the reflectance contrast (RC) spectrum we need to simulate theRC spectrum by considering a model for the refractive index contribution of the exitons/trions in MoS2. Previouslythe Lorentz oscillator model[9] had been used to describe the complex dielectric function ε(E) of exitons/trions andrefractive indices obtained from it. However at 4.5 K (kBT ∼ 0.4 meV) the broadening of the exiton/trion related RCspectral features is> 20 meV indicating that inhomogeneous broadening is dominant. For ε(E) we therefore considereda Lorentz oscillator model convolved with a normalized Gaussian representing the inhomogeneous broadening as follows

ε(E) = εb +

2∑l=1

E0l+2∆El∑Ej=E0l−2∆El

al2

∆El

√ln 2

πexp[−4 ln 2

(E0l − Ej

∆El

)2

]δEj1

E2j − E2 − ıΓE

(S4)

where εb = (nb + ιkb)2 is the background dielectric constant of MoS2 (excluding trion/exciton contribution) with

background refractive indices taken as nb = 4.5 and kb ' 0 in the small spectral range of interest. E is the photonenergy and E0l, ∆El, al are mean transition energy, inhomogeneous broadening, amplitude parameters respectivelyof the lth transition where l = 1 represents the trion and l = 2 the exciton. δEj is the step size in the summationover Ej values. The Lorentzian broadening Γ representing lifetime broadening in this model will be small andwas taken to be 1 meV. The final complex refractive indices for MoS2 were obtained from ε(E) using standardrelations.[10] The absorption coefficient α (plotted in Fig.5 of the main manuscript) is given by α = 4πk/λ wherek is the imaginary part of the refractive index and λ the wavelength. For the other two layers in the problem SiO2

and Si, the wavelength (λ in µm) dependent refractive indices (valid in the range 0.4-0.8 µm) were approximatedas nSiO2

= [1.0 + 1.4278λ2/(λ2 − 7.2316 × 10−3)]0.5, kSiO2= 0, nSi = [4.9325 + 7.1714λ2/(λ2 − 0.11632)]0.5 and

Energy (eV)1.9 2.0

ε 1

15

20

25

Ex

Tr

4.5K

Energy (eV)1.9 2.0

ε 2

0

5

104.5K

Ex

Tr

Figure S4 The real ε1(E) and the imaginary ε2(E) parts of the dielectric function of monolayer MoS2 in the A exciton/triondominated region as determined from the analysis of the reflectance contrast spectrum.

Page 8: Phonon induced luminescence decay in monolayer MoS2 on ...sangho10/Publications/15APL-ns2.pdfPhonon induced luminescence decay in monolayer MoS 2 on SiO 2/Si substrates Nihit Saigal

4

kSi = exp[23.530−108.32λ+147.35λ2−70.817λ3].[11] Using these refractive index values we calculated the reflectanceof structures with (R′) and without (R) MoS2 on top of SiO2/Si using the Transfer Matrix method[10] for multilayeredstructures. Reflectance spectrum measured under a microscope can be distorted because for an objective lens with alarge numerical aperture (NA), a large range of angles of incidence is involved.[12] However in this measurement setupwe had an objective with relatively small NA=0.5 and therefore angle of incidence averaging was not considered inthe simulations. The MoS2, SiO2 and Si thicknesses used were 0.65 nm, 283 µm and semi-infinite respectively. Thesimulated RC spectrum is finally obtained as 100× (R−R′)/R.

The values of the parameters in Eq.S4 were varied to match the simulated RC spectrum with the measured one.This match is shown in the inset of Fig.5 of the main manuscript. The Lorentzian model does not account for theSommerfeld factor modified band-to-band transition at higher energies and this possibly causes the poorer match withthe measured RC spectrum towards the high energy end. The dielectric function parameter values thus obtained were,for the trion E01 = 1.94 eV, ∆E1 = 22 meV, a1 = 0.0935 eV2 and for the exciton E02 = 1.969 eV, ∆E2 = 25 meV,a2 = 0.2 eV2. The real and imaginary parts of ε(E) for MoS2 thus obtained are plotted in Fig. S4. Their magnitudescompare reasonably with the most recently reported estimation of ε(E) of MoS2 from combined transmission andreflectivity measurements.[13]

[1] E. W. Williams and H. B. Bebb, Semiconductors and Semimetals, Vol. 8, R. K. Willardson and A. C. Beer, Eds. p321(Academic Press, New York, 1972).

[2] P. J. Ko, A. Abderrahmane, T. V. Thu, D. Ortega, T. Takamura, and A. Sandhu, J. Nanosci. Nanotech. 15, 1 (2015).[3] S. Mouri, Y. Miyauchi, and K. Matsuda, Nano. Lett. 13, 5944 (2013).[4] P. T. Landsberg, Phys. Stat. Sol. 41, 457 (1970).[5] E. H. Bogardus and H. B. Bebb, Phys. Rev. 176, 993 (1968).[6] J. Krustok, H. Collan, and K. Hjelt, J. Appl. Phys. 81, 1442 (1997).[7] K. Kaasbjerg, K. S. Thygesen, and K. W. Jacobsen, Phys. Rev. B 85, 115317 (2012).[8] A. Thilagam, Physica B 464, 44 (2015).[9] A. Arora, M. Koperski, K. Nogajewski, J. Marcus, C. Faugerasa, and M. Potemski, Nanoscale 7, 10421 (2015).

[10] E. Hecht, Optics, 4th Edition (Addison-Wesley, New York, 2002).[11] E. D. Palik, Handbook of Optical Constants of Solids, Vol.1 (Academic Press, London, 1998).[12] N. Saigal, A. Mukherjee, V. Sugunakar, and S. Ghosh, Rev. Sci. Instrum. 85, 073105 (2014).[13] Y. V. Morozov and M. Kuno, Appl. Phys. Lett. 107, 083103 (2015).