17
Minimizing residues and strain in 2D materials transferred from PDMS Achint Jain 1 , Palash Bharadwaj 2 , Sebastian Heeg 1 , Markus Parzefall 1 , Takashi Taniguchi 3 , Kenji Watanabe 3 and Lukas Novotny 1 1 Photonics Laboratory, ETH Zürich, 8093 Zürich, Switzerland 2 Department of Electrical and Computer Engineering, Rice University, Houston, TX 77005, United States of America 3 National Institute for Material Science, 1-1 Namiki, Tsukuba 305-0044, Japan E-mail: [email protected] Received 9 February 2018, revised 29 March 2018 Accepted for publication 12 April 2018 Published 3 May 2018 Abstract Integrating layered two-dimensional (2D) materials into 3D heterostructures offers opportunities for novel material functionalities and applications in electronics and photonics. In order to build the highest quality heterostructures, it is crucial to preserve the cleanliness and morphology of 2D material surfaces that come in contact with polymers such as PDMS during transfer. Here we report that substantial residues and up to 0.22% compressive strain can be present in monolayer MoS 2 transferred using PDMS. We show that a UV-ozone pre-cleaning of the PDMS surface before exfoliation signicantly reduces organic residues on transferred MoS 2 akes. An additional 200 C vacuum anneal after transfer efciently removes interfacial bubbles and wrinkles as well as accumulated strain, thereby restoring the surface morphology of transferred akes to their native state. Our recipe is important for building clean heterostructures of 2D materials and increasing the reproducibility and reliability of devices based on them. Supplementary material for this article is available online Keywords: TMDCs, van der Waals heterostructures, transfer techniques, material synthesis strain (Some gures may appear in colour only in the online journal) 1. Introduction Two-dimensional (2D) materials like graphene, hexagonal boron nitride (hBN), transition metal dichalcogenides (TMDCs), etc., have gained immense attention in recent years due to the wealth of novel fundamental properties and fascinating physical phe- nomena exhibited by them [1, 2]. A unique possibility existing with these materials is that of assembling them into 3D hetero- structures to create articial materials which do not exist naturally and thus give rise to new functionalities which seem promising for many future applications [3, 4]. These heterostructures can be built by a variety of methods [5] like chemical vapor deposition (CVD) [6], epitaxial growth [7, 8], inkjet printing [9] or more generally by mechanical stacking of individual layers [10, 11]. Among the many available techniques for stacking 2D materials [1116], one approach that has become common recently is to mechanically exfoliate bulk 2D crystals onto a stamp made of a viscoelastic material, such as poly-dimethylsiloxane (PDMS), bring them in contact with a desired substrate and then slowly detach the stamp to leave the 2D akes behind [17, 18]. Although this procedure is very versatile, deterministic and fairly simple to perform, not much attention has been paid to the sur- face cleanliness of the akes transferred this way. PDMS is a widely used polymer for contact printing [19], microuidics [20] as well as stretchable electronics [21], and Nanotechnology Nanotechnology 29 (2018) 265203 (9pp) https://doi.org/10.1088/1361-6528/aabd90 Original content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI. 0957-4484/18/265203+09$33.00 © 2018 IOP Publishing Ltd Printed in the UK 1

29 // Minimizing residues and strain in 2D materials ... · its chemistry has been studied extensively [22]. PDMS is composed of a network of cross-linked dimethylsiloxane oligomers

Embed Size (px)

Citation preview

Minimizing residues and strain in 2Dmaterials transferred from PDMS

Achint Jain1 , Palash Bharadwaj2 , Sebastian Heeg1 ,Markus Parzefall1 , Takashi Taniguchi3, Kenji Watanabe3

and Lukas Novotny1

1 Photonics Laboratory, ETH Zürich, 8093 Zürich, Switzerland2Department of Electrical and Computer Engineering, Rice University, Houston, TX 77005, United Statesof America3National Institute for Material Science, 1-1 Namiki, Tsukuba 305-0044, Japan

E-mail: [email protected]

Received 9 February 2018, revised 29 March 2018Accepted for publication 12 April 2018Published 3 May 2018

AbstractIntegrating layered two-dimensional (2D) materials into 3D heterostructures offers opportunitiesfor novel material functionalities and applications in electronics and photonics. In order to buildthe highest quality heterostructures, it is crucial to preserve the cleanliness and morphology of2D material surfaces that come in contact with polymers such as PDMS during transfer. Here wereport that substantial residues and up to ∼0.22% compressive strain can be present inmonolayer MoS2 transferred using PDMS. We show that a UV-ozone pre-cleaning of the PDMSsurface before exfoliation significantly reduces organic residues on transferred MoS2 flakes. Anadditional 200 ◦C vacuum anneal after transfer efficiently removes interfacial bubbles andwrinkles as well as accumulated strain, thereby restoring the surface morphology of transferredflakes to their native state. Our recipe is important for building clean heterostructures of 2Dmaterials and increasing the reproducibility and reliability of devices based on them.

Supplementary material for this article is available online

Keywords: TMDCs, van der Waals heterostructures, transfer techniques, material synthesisstrain

(Some figures may appear in colour only in the online journal)

1. Introduction

Two-dimensional (2D) materials like graphene, hexagonal boronnitride (hBN), transition metal dichalcogenides (TMDCs), etc.,have gained immense attention in recent years due to the wealthof novel fundamental properties and fascinating physical phe-nomena exhibited by them [1, 2]. A unique possibility existingwith these materials is that of assembling them into 3D hetero-structures to create artificial materials which do not exist naturallyand thus give rise to new functionalities which seem promising

for many future applications [3, 4]. These heterostructures can bebuilt by a variety of methods [5] like chemical vapor deposition(CVD) [6], epitaxial growth [7, 8], inkjet printing [9] or moregenerally by mechanical stacking of individual layers [10, 11].Among the many available techniques for stacking 2D materials[11–16], one approach that has become common recently is tomechanically exfoliate bulk 2D crystals onto a stamp made of aviscoelastic material, such as poly-dimethylsiloxane (PDMS),bring them in contact with a desired substrate and then slowlydetach the stamp to leave the 2D flakes behind [17, 18].Although this procedure is very versatile, deterministic and fairlysimple to perform, not much attention has been paid to the sur-face cleanliness of the flakes transferred this way.

PDMS is a widely used polymer for contact printing [19],microfluidics [20] as well as stretchable electronics [21], and

Nanotechnology

Nanotechnology 29 (2018) 265203 (9pp) https://doi.org/10.1088/1361-6528/aabd90

Original content from this work may be used under the termsof the Creative Commons Attribution 3.0 licence. Any

further distribution of this work must maintain attribution to the author(s) andthe title of the work, journal citation and DOI.

0957-4484/18/265203+09$33.00 © 2018 IOP Publishing Ltd Printed in the UK1

its chemistry has been studied extensively [22]. PDMS iscomposed of a network of cross-linked dimethylsiloxaneoligomers which do not get fully cross-linked even afterextensive curing [23] and depending on the curing time andtemperature, up to 5% of oligomers can remain uncrosslinkedwithin the PDMS bulk [23, 24]. It is well-known that theseuncrosslinked species are even present on the surface ofPDMS stamps and get transferred to the target substrateduring contact printing [25, 26], thereafter acting as a surfacecontamination layer. PDMS oligomer residues have beencharacterized previously by many techniques such as non-linear spectroscopy [27], x-ray photoelectron spectroscopy[25, 28], atomic force microscopy (AFM) and ToF-SIMS[28], confirming that PDMS can indeed degrade the surfacecleanliness of transferred materials. Residues were also foundto occur on 2D materials, e.g.graphene [29] and TMDCs[30], transferred using PDMS. Owing to their 2D structure,the properties of 2D materials are quite sensitive to surfacecontaminants and residues trapped at the interfaces within 3Dheterostructures [31]. Hence, it is essential to stack thesematerials with minimum residues.

In this work, the cleanliness of 2D materials transferredusing PDMS was investigated with the help of AFM andphotoluminescence (PL) measurements. Using monolayer(1L) MoS2 as a test layer, we show evidence for substantialresidues being present on MoS2 transferred onto hBN fromPDMS. For transferring MoS2 in a cleaner way, we developedan ultraviolet-ozone (UV-O3) treatment recipe to clean thePDMS surface before exfoliating MoS2 on it and found asignificant reduction in residues compared to transfer fromuntreated PDMS. Using PL and Raman spectroscopy, wefurther reveal that a small compressive strain can be present inMoS2 after transfer from PDMS. We found that subsequentvacuum annealing leads to an almost pristine MoS2 surface onhBN, mostly free from interfacial bubbles, wrinkles andstrain. Although here we chose 1L-MoS2 for demonstrationsince its sensitive PL and Raman signals allow for systematicoptical characterization, our cleaning recipe is general and canbe used with other 2D materials as well.

2. Results and discussion

Naturally occurring MoS2 crystals purchased from SPI Sup-plies were exfoliated on commercial PDMS films (Gel-Film®

PF-40-X4 sold by Gel-Pak) using a blue tape (BT-150E-KL).Figure 1(a) is an optical microscope image of a monolayerMoS2 flake on PDMS. Bulk hBN crystals were separatelyexfoliated on O2 plasma cleaned p+Si/SiO2 (285 nm) sub-strates. The PDMS stamp with MoS2 was placed on a trans-parent quartz plate and aligned on top of a suitable hBN flakeon SiO2 using a mask aligner as depicted schematically infigure 1(b). Upon slowly bringing MoS2 in contact with hBNat room temperature, the entire stack was heated to ∼65 ◦Cfor two minutes using a Peltier module kept underneath theSi/SiO2 substrate. After allowing the stack to cool down, thePDMS stamp was slowly detached as described in [18],leaving behind the MoS2 flake on hBN (figure 1(c)).

To characterize the surface of the flakes after transfer, weperformed topography mapping using an AFM operating intapping mode. Figure 1(d) shows an AFM topography map ofthe region outlined in black in figure 1(c). The MoS2 flakeappears mostly ‘clean’ on this large scale apart from the usualwrinkles and bubbles which are frequently seen in PDMStransferred flakes [18]. However, if we examine the regionoutlined in red more closely in the higher resolution map infigure 1(e), a dense network of residue islands is clearlyvisible on the entire MoS2 surface, making it difficult to evenobtain an accurate estimate of the 1L-MoS2 thickness from ahorizontal cross-section (figure 1(h)). Note that these ±0.6 nm(rms) variations in topography cannot arise from substrateroughness alone as the hBN layer below is atomically smooth.Moreover, on the narrower 1L-MoS2 flake on the right infigure 1(d), a thick residue layer can be clearly identifiedwhich is unmistakably distinct from the MoS2 flake itself.Better visible in the higher resolution AFM map in figure 1(f),this additional layer left-over by PDMS can be as thick as∼2.5 nm in some areas (magenta profile in figure 1(i)).Together with the topography map in figure 1(e), we alsorecorded the corresponding AFM phase map as shown infigure 1(g) and observed a poor phase contrast between theMoS2 and hBN surfaces which also indicates the presence ofPDMS residues everywhere.

We tested several flakes and similar residues were foundon all of them (see supplementary figure S1 available onlineat stacks.iop.org/NANO/29/265203/mmedia for residuespresent on another flake). The amount of residues transferredwas also affected by the temperature and pressure appliedduring the transfer. In case of transfers done without applyingany heat, residues were visible even in an optical microscopedue a change in the color of the SiO2 (285 nm) layer whicharises from interference and is sensitive to a few nanometersthick transparent organic layer on top (see supplementaryfigure S2). These results unambiguously demonstrate, inagreement with previous reports [25–30], that PDMS canindeed leave a significant amount of residues behind and anefficient method for eliminating them is genuinely needed.

Although a high temperature (400 ◦C) vacuum annealafter transfer has been reported to reduce PDMS residues ongraphene [29], such high temperatures are known to introduceadditional defects in TMDCs [32, 33] which are more fragilecompared to graphene. Moreover, we found that annealing byitself is not sufficient to fully restore the transferred flakes totheir pristine state. The MoS2 flake shown in figure 1(c) wasannealed at 200 ◦C in vacuum for 3 h. In the AFM maps infigures 2(a), (b) it can be noticed that although bubbles andwrinkles are mostly gone, the MoS2 surface is still smearedwith residues. This is also evident from the AFM cross-section in figure 2(b) revealing the total thickness of the flaketo be 1.6 nm, which is significantly higher than the expectedvalue of 0.7 nm for monolayer MoS2. An additional vacuumanneal using the same parameters did somewhat reduce thethickness to ∼0.9 nm as depicted in figure 2(c). However, itcan be easily seen that the surface topography is quite inho-mogeneous and does not approach the cleanliness level of afreshly exfoliated MoS2 flake.

2

Nanotechnology 29 (2018) 265203 A Jain et al

Alternatively, dissolving PDMS residues in organic sol-vents such as dichloromethane and toluene [24] is also notvery effective as the solvent molecules themselves tend to getadsorbed at exposed 2D material surfaces/edges and can evenchemically dope TMDCs [34]. This brings us to the questionof how to eliminate PDMS residues in a way that does notcompromise the 2D material being transferred in any way. Inthis regard, it appears more reasonable to clean the PDMSsurface itself before it comes in contact with the 2D material.

To achieve this, we developed a UV-O3 treatment recipe tomodify the PDMS surface prior to MoS2 exfoliation. Thisrecipe was found to significantly reduce transfer residueswithout having any negative effect on the MoS2. UV-O3

cleaning of PDMS has been proposed in the past [25] and themechanism behind it can be understood as follows. Oxygen free

radicals and ozone (O3) created from atmospheric oxygen in thepresence of UV-radiation, break down organic species on thePDMS surface into CO2, H2O and simpler volatile organicproducts. At the same time, the silicon present in PDMS getsoxidized and forms a thin (20–30 nm) layer of silicon oxide(SiOx) on PDMS [35]. This SiOx surface layer besides having alow carbon content, also acts as a diffusion barrier for uncros-slinked oligomers still present within the PDMS bulk.

To optimize the treatment time, we exposed severalPDMS stamps to UV-O3 in a Bioforce Nanosciences UVOzone ProCleaner for varying times, and found a duration of30–40 min to be the optimum. Shorter times did not fullyclean the PDMS and longer exposure led to a poorer coverageof flakes on the PDMS upon exfoliation. We also observedthat if the exfoliation was done immediately after UV-O3

Figure 1. Characterization of PDMS residues. (a) Optical microscope image of a 1L-MoS2 flake exfoliated on PDMS. (b) Schematicillustration of an inverted PDMS stamp with MoS2 being transferred onto hBN. (c) The same MoS2 flake as in (a) after transfer to hBN. (d)Large area AFM topography map of the region outlined in (c) displaying an ‘apparently clean’ surface together with some bubbles andwrinkles. (e) Higher resolution AFM map of the red-outlined region in (d) exhibiting a dense distribution of residues over the entire MoS2flake. (f) Higher resolution AFM map of the green-outlined region in (d) with a thick layer of residues covering a majority of the area. (g)AFM phase map recorded together with the topography in (e) depicting a poor phase contrast between MoS2 and hBN due to PDMS residueson both surfaces. (h) Height profile across the MoS2 edge along the dashed line marked in (e). An accurate estimation of the thickness ofmonolayer MoS2 is hindered by topography variations due to surface residues. (i) Height profile across the PDMS residue layer and PDMS +MoS2 along the dashed lines marked in (f). The thickness of the residue layer is ∼2.5 nm.

3

Nanotechnology 29 (2018) 265203 A Jain et al

treatment, bonding often occurred between the PDMS and theblue tape used for exfoliation or in rare cases even betweenthe PDMS and O2 plasma treated SiO2 surface during thetransfer process. However, this bonding could be avoided byleaving the PDMS exposed to ambient air for a few hours tolet the surface undergo a partial hydrophobic recovery[36, 37]. This led us to the following optimized process flow,the results of which are presented below.

The PDMS stamp was exposed to UV-O3 for 30 min andthen after a wait interval of 2 h in air, MoS2 was exfoliated onPDMS as usual. Upon optical identification of suitable flakes,transfer was carried out the same way as described earlier.Figure 3(a) shows an optical microscope image of a largeMoS2 flake exfoliated on 30 min UV-O3 treated PDMS whichwas subsequently transferred to hBN (figure 3(b)). In theAFM topography map of the 1L region shown in figure 3(c),one can notice the absence of dense islands or thick layers ofPDMS residues unlike in figures 1(e) or (f). Although smallamounts of residues can still be detected, this MoS2 flake issignificantly cleaner than the one in figure 1 which wastransferred from untreated PDMS (also see supplementaryfigure S3 for better resolved AFM images of another cleanMoS2 flake).

The large number of bright spots in figure 3(c) arewrinkles and bubbles filled mainly with air molecules andorganic adsorbates that become trapped between MoS2 andhBN during transfer and get squeezed into small pockets via aself-cleaning effect [13, 38]. The amount of trapped bubblescan be reduced by performing the exfoliation and transferinside an Ar filled glove box to exclude molecular adsorbates[39] or even fully eliminated by stacking in vacuum as shownrecently [40]. Alternatively, we found that bubbles andwrinkles can be very efficiently removed by vacuumannealing at 200 ◦C, similar to another recent report [41]. Atthis temperature and under low pressure, the trapped speciesinside the bubbles become mobile and coalesce into biggerbubbles to minimize the total surface energy [16]. They alsotend to migrate towards the edge from where they eventuallyescape the interface, thereby lowering the overall density of

bubbles. Figure 3(d) is an AFM map of the same region asfigure 3(c) after 3 h vacuum annealing at 200 ◦C exhibiting acomplete removal of wrinkles and a remarkable reduction inbubbles. The higher resolution AFM map in figure 3(e) of thered-outlined region features a nearly pristine surface with aclean step of 7.2Å (figure 3(f)) and a surface roughness of±1.4Å (rms) over the entire flake. This is in striking contrastto the MoS2 flake transferred from untreated PDMS where thesurface quality remained compromised by PDMS residueseven after prolonged annealing as shown in figures 2(b), (c).The corresponding AFM phase map in figure 3(f) (inset)shows a clear phase contrast between MoS2 and hBN (unlikefigure 1(g)) providing further evidence for the decrease inresidues by UV-O3 treatment of PDMS. Detailed AFM ana-lysis of two additional 1L-MoS2 flakes transferred fromUV-O3 treated PDMS can be found in supplementary figuresS3–S4 showing similarly clean surfaces.

Thus our recipe provides a new method to obtain veryclean 2D material flakes on hBN. We would like to stress thata combination of UV-O3 pre-cleaning followed by vacuumannealing is crucial, and vacuum annealing by itself does notresult in a pristine surface as evident from figure 2. Note thathere we transferred MoS2 onto atomically smooth hBN flakesto decouple the SiO2 substrate roughness from the surfacetopography of MoS2 which allowed us to better resolve sur-face residues using AFM. Although the surface cleanlinessdoes not depend on the substrate used, we have observed thatthe removal of bubbles and wrinkles during vacuum anneal-ing is not very effective for MoS2 transferred directly on SiO2

(see discussion in supplementary section S4).

3. Optical characterization

To further characterize the effect of PDMS transfer on theoptical properties of MoS2, we performed PL and Ramanspectroscopy. Experimental details can be found in section 5(methods). Spatial maps of the spectrally integrated PL countswere recorded from the MoS2 flake in figure 1(d) transferred

Figure 2. Limited effect of annealing on PDMS residues. (a) AFM topography map of the dirty 1L-MoS2 flake shown in figure 1(d) after 3 hannealing in vacuum. (b) Higher resolution AFM map of the white-outlined region in (a). Even though the surface appears mostly free frombubbles and more homogeneous than before (see figure 1(e)), a thin layer of PDMS residues is still present over the entire MoS2 flake. Inset:height profile along the red-dashed line revealing a total thickness of 1.6 nm for the monolayer flake. (d) AFM map of the same region afterannealing again for 3 h. Although the total amount of residues does decrease after prolonged annealing, the surface looks far from pristine.Inset: height profile along the green dashed line.

4

Nanotechnology 29 (2018) 265203 A Jain et al

using untreated PDMS (labelled as ‘untreated’ in figure 4) aswell as from the clean MoS2 flake transferred using UV-O3

treated PDMS and subsequently vacuum annealed(figure 3(d)). PL maps of the two flakes depicted infigures 4(a), (b) show no apparent differences and the countrates were comparable for similar excitation powers. Thisindicates that PDMS residues (and/or vacuum annealing) donot seem to affect the overall PL quantum yield which couldbe one reason why residues have been overlooked in the past.The PL spectra of the two flakes, however, did show someinteresting differences.

Besides the well-known A, B-exciton and trion (X−)peaks, a new broad feature around 1.68 eV can be noticed infigure 4(c) in the PL spectra of untreated MoS2. This low-energy peak can be better visualized at the locations markedby green and yellow dots in figure 4(a) where thick PDMSlayers were present on MoS2 (as shown previously in theAFM map in figure 1(f)). At these spots, the main excitonicpeaks were weaker which made it easier to resolve the newpeak (see figure 4(c) inset). Surprisingly, this peak could notbe detected in MoS2 flakes transferred from UV-O3 treatedPDMS (light blue and orange curves in figure 4(d) inset) aswell as in the fluorescence spectra of an untreated PDMSstamp without MoS2 (data not shown here). It appeared in thePL spectra of MoS2 only when PDMS residues were present

on MoS2 but its true origin is unclear at present. Such a broadlow-energy emission could possibly be attributed to impuritybound excitons at room temperature [33, 42] whose creation,however, is still not well understood in literature and a moredetailed investigation is needed to elucidate the exact mech-anism that gives rise to this additional peak.

Apart from this, it can be clearly observed in figure 4(d)that the A-exciton peaks of as-transferred MoS2 (light anddark blue curves) are slightly blue-shifted with respect to thatof vacuum annealed MoS2 (orange curve) and lie at1.887± 0.002 eV which agrees very well with the value of1.89 eV measured previously on MoS2 transferred fromPDMS onto various substrates [43]. On the contrary, forvacuum annealed MoS2 the A-exciton peak lies at1.876±0.002 eV, very close to that of MoS2 exfoliateddirectly on SiO2 [44, 45]. By comparing the PL spectra atvarious locations on the two flakes in figures 4(a), (b) andperforming multi-Lorentzian fitting to decouple the trion peakfrom the exciton peak, we estimated a blue-shift of11±3 meV for the A-exciton peak of as-transferred MoS2.Moreover, one can also notice a blue-shift in the B-excitonpeak as highlighted by the Lorentzian fits in figure 4(e). Thecomplete set of fits for the entire spectra can be found insupplementary figure S6.

Figure 3. Clean transfer from UV-ozone cleaned PDMS. (a) Optical microscope image of 1L-MoS2 exfoliated on 30 min UV-O3 cleanedPDMS. (b) The same MoS2 flake after transfer to hBN with the monolayer segment demarcated. (c) AFM topography map of the regionoutlined in (b). Areas densely covered with residues like in figures 1(e) or (f) could no longer be detected although a considerable amount ofbubbles do occur which could also be efficiently removed. (d) AFM map of the same region showing a drastic reduction in bubbles andwrinkles from the entire flake after 200 ◦C vacuum annealing. (e) Higher resolution AFM map of the red-outlined region in (d) displaying avery clean and smooth MoS2 surface. The faint streaks correspond to migration paths of residual bubbles which could not reach the MoS2edge and get released. (f) Height profile along the red-dashed line in (e) exhibiting a clear monolayer MoS2 step of 7.2 Å unlike infigures 1(h) or 2(b), (c). Inset: AFM phase map recorded together with the topography in (e) revealing a clear phase contrast between MoS2and hBN which further indicates the absence of residues (compare with figure 1(g)).

5

Nanotechnology 29 (2018) 265203 A Jain et al

In order to understand the origin of the A, B-excitonblue-shift upon transfer, we performed Raman spectroscopyto gain further insight from the E2g

1 mode which is sensitive tostrain in MoS2. In exfoliated, unstrained 1L-MoS2 at roomtemperature, the E2g

1 peak should lie at 385 cm−1 [46] whereasfor our as-transferred MoS2 on hBN it lies at ∼386 cm−1

(blue curve in figure 4(f)). An E2g1 peak up-shift signifies an

increase in built-in compressive strain, thus implying thatPDMS not only leaves residues behind, but can also compressthe MoS2 during transfer. An up-shift of ∼1 cm−1 corre-sponds to an accumulation of ∼0.22% compressive strain inMoS2 after transfer [47], similar to estimate made by Bus-cema et al [43]. This induced strain causes an increase in thedirect bandgap at the K-point which leads to blue-shifted A,

B-exciton emission [48, 49]. According to previously reportedDFT calculations [50] as well as experimental results [47], a0.22% strain would induce an A-exciton shift of 10± 1.5 meVwhich is in good agreement with the shift of 11± 3meVestimated from our PL measurements. For the sake of com-pleteness, one can also compare the A1g peaks in the twoRaman spectra to characterize the effect of residues on doping.The strong electron–phonon coupling of the out-of-plane A1g

mode in MoS2 causes it to down-shift with increasing doping[51]. In figure 4f, the two A1g peaks lying at ∼406.5 cm−1

imply a low n-doping in both MoS2 flakes on hBN (comparedto 403–404 cm−1 for 1L-MoS2 on SiO2) [43, 46].

The origin of strain can possibly be attributed to theinherent lack of stiffness of PDMS. As discussed previously

Figure 4. Optical characterization. (a, b) Spatial maps of the spectrally integrated PL counts from the 1L-MoS2 flakes in figures 1(d), 3(d)displaying similar emission intensities. The dark spots and lines in (a) are interfacial bubbles and wrinkles respectively (see figure 1(d)) wherethe photon counts were somewhat decreased. (c) PL spectra of MoS2 transferred from untreated PDMS recorded at points marked bymatching colored dots in (a). Besides the indicated excitonic peaks, all spectra also show a low-energy feature (marked by arrow) around1.68 eV. (d) PL spectra of MoS2 transferred from UV-O3 treated PDMS showing the absence of any low-energy feature in comparison withMoS2 from untreated PDMS. This indicates that the new feature arises only in the presence of PDMS residues on MoS2. Apart from this, thePL spectra in light and dark blue show blue-shifted A, B excitonic peaks which reveals the presence of compressive strain in as-transferredMoS2 (irrespective of PDMS pre-treatment). This strain got released from the flake in (b) upon annealing. The PL spectra in light blue wasrecorded from the flake in figure S3a before annealing while the annealed MoS2 spectra (orange) was taken from the flake in (b). Insets:magnified plots of the spectra around 1.68 eV. (e) A magnified view of the B-exciton emission in (d). The filled area plots are Lorentzian fitsto the B-exciton peaks and display a clear shift in the center positions. (f) Raman spectra of the two flakes (vertically offset for clarity). Theempty circles are measured data points while the smooth lines are Lorentzian fits to estimate the center position of each peak. The E2g

1 peak(in blue) is up-shifted by ∼1 cm−1 which also indicates a build-up of compressive strain in as-transferred MoS2, in agreement with the PLspectra. All spectra have been normalized to the Raman peak of silicon for a better comparison.

6

Nanotechnology 29 (2018) 265203 A Jain et al

by Gomez et al [18], PDMS being soft can get slightlydeformed during transfer by the pressure exerted on PDMSupon coming in contact with the target substrate. It is quitelikely that this deformation of PDMS induces a compressivestrain in the flake being transferred as measured for MoS2 inour case. The induced strain eventually gets released uponvacuum annealing, down-shifting the E2g

1 peak to 385 cm−1

and at the same time red-shifting the A, B-exciton peaks tounstrained values close to that of directly exfoliated MoS2.The release of transfer induced compression in MoS2 can alsobe evidenced in the AFM scans in figures 3(c), (d) before andafter annealing where an increase in the total surface areaupon annealing can be clearly noticed.

4. Conclusions

To summarize, we have demonstrated that PDMS residues aswell as compressive strain can be present in 2D materialsexfoliated onto and transferred from PDMS stamps. Using1L-MoS2 as an example, we have observed evidence of sur-face contamination in both AFM maps and PL spectra.UV-O3 treatment of PDMS prior to exfoliation significantlyreduces the amount of unwanted surface oligomers whichresults in a cleaner transfer of MoS2 in comparison withuntreated PDMS. We showed that a 1L-MoS2 surface of avery high-quality can be obtained by a combination of UV-O3

pre-cleaning followed by vacuum annealing after transfer.AFM topography of annealed MoS2 flakes on hBN displayeda homogeneously smooth surface with a substantial reductionin interfacial bubbles and wrinkles. PL spectroscopy of as-transferred MoS2 revealed blue-shifted A, B-exciton peaksdue to an accumulation of compressive strain during thetransfer. This induced compression could be released by apost-transfer vacuum anneal.

It is advantageous to exfoliate MoS2 on PDMS forobtaining large area (>1000 μm2) monolayer MoS2 flakeswith a high yield unlike direct exfoliation on SiO2 whichresults in relatively smaller flakes. The recipe we providednow makes it possible to integrate these large area flakesexfoliated on PDMS into clean heterostructures for highperformance electronic and photonic devices. Our results arevaluable for future experimental studies and practical appli-cations utilizing clean 2D material heterostructures.

One must keep in mind that even though PDMS residuesdo not significantly influence the optical properties of MoS2,surface residues could still affect electrical transport byscattering charge carriers and thereby reduce carrier mobility.Moreover, trapped residues within TMDC heterostructurescould also weaken interlayer coupling [31] and adverselyaffect physical phenomena such as interlayer charge recom-bination or separation, interlayer electron–phonon coupling,polariton formation, out-of-plane tunneling, etc., which relyon high-quality interfaces. Hence, the importance of elim-inating residues while stacking 2D materials using PDMS, orany other technique in general, cannot be overstated. In thisdirection, our results highlight the significance of a careful

characterization and optimization of any transfer procedureand subsequent processing steps in order to preserve interfacequality and obtain unperturbed crystal structures with well-defined physical properties.

5. Methods

All transfers were done with a SÜSS MicroTec MJB4 maskaligner in air. Annealing was performed in a quartz tubefurnace (Carbolite Gero) at 200 ◦C for 3 h in a low vacuum of5×10−3 mbar (limited by our rotary pump). Before heating,the quartz tube was flushed several times with argon gas toremove any residual water or oxygen molecules. After 3 h, thefurnace was left for a few hours to cool down naturally toroom temperature before taking out the samples.

All measurements were carried out at room temperatureunder ambient conditions. For PL measurements, sampleswere mounted on a piezoelectric stage and excited with a532 nm Nd:YAG laser (attenuated to 15 μW power) using a50x air objective (0.8 NA) in a scanning confocal microscopysetup. The MoS2 flakes were raster scanned across the laserfocus and photon counts at each xy-position were recordedwith a single photon counting module (Perkin Elmer SPCM-AQRH-14) after passing through a 532 nm RazorEdge long-pass filter. PL spectra were measured with a PrincetonInstruments Acton SP300i spectrometer at 100 μW excitationpower (30 s integration). For Raman spectroscopy, excitationwas done with the 530.8 nm line of an Ar-Kr laser (Coherent)at 100 μW power and the back-scattered light was dispersedonto a grating with 1200 grooves mm–1. The grating wascalibrated with a Neon lamp before recording the Ramanspectra. Lorentzian fits were obtained using the freely avail-able peak-o-mat software [52]. All spectra were normalized tothe Raman peak of p+Si lying at 521.7 cm−1 for a bettercomparison.

Acknowledgments

This research was supported by the Swiss National ScienceFoundation (grant no. 200021_165841) and ETH Zürich(ETH-32 15-1). TT and KW acknowledge support from theElemental Strategy Initiative conducted by the MEXT, Japanand JSPS KAKENHI (grant no. JP15K21722).

Author contributions

The overall project was conceived and supervised by PB andLN. AJ developed the cleaning procedure, fabricated thesamples, performed the measurements and wrote the manu-script. PB built the confocal optical microscopy setup used forPL spectroscopy of MoS2. Raman characterization was car-ried out together with SH. The vacuum tube furnace wassetup by MP who also wrote the LabVIEW script for per-forming vacuum annealing. TT and KW synthesized the hBNcrystals used in this study. PB, SH, MP and LN also

7

Nanotechnology 29 (2018) 265203 A Jain et al

participated in discussion of results and contributed to writingthe manuscript.

ORCID iDs

Achint Jain https://orcid.org/0000-0002-3495-5935Palash Bharadwaj https://orcid.org/0000-0003-0722-6697Sebastian Heeg https://orcid.org/0000-0002-6485-3083Markus Parzefall https://orcid.org/0000-0003-2930-6698Lukas Novotny https://orcid.org/0000-0002-9970-8345

References

[1] Xu M, Liang T, Shi M and Chen H 2013 Graphene-like two-dimensional materials Chem. Rev. 113 3766–98

[2] Bhimanapati G R et al 2015 Recent advances in two-dimensionalmaterials beyond graphene ACS Nano 9 11509–39

[3] Geim A K and Grigorieva I V 2013 Van der Waalsheterostructures Nature 499 419–25

[4] Novoselov K, Mishchenko A, Carvalho A and Neto A C 20162D materials and van der Waals heterostructures Science353 461

[5] Lim H, Yoon S I, Kim G, Jang A-R and Shin H S 2014Stacking of two-dimensional materials in lateral and verticaldirections Chem. Mater. 26 4891–903

[6] Gong Y et al 2014 Vertical and in-plane heterostructures fromWS2/MoS2 monolayers Nat. Mater. 13 1135–42

[7] Koma A, Sunouchi K and Miyajima T 1985 Fabrication ofultrathin heterostructures with van der Waals epitaxy J. VacSci. Technol. B 3 724

[8] Lin Y-C et al 2015 Atomically thin resonant tunnel diodes builtfrom synthetic van der Waals heterostructures Nat.Commun. 6 7311

[9] McManus D et al 2017 Water-based and biocompatible 2Dcrystal inks for all-inkjet-printed heterostructures Nat.Nanotechnol. 12 343–50

[10] Dean C R et al 2010 Boron nitride substrates for high-qualitygraphene electronics Nat. Nanotechnol. 5 722–6

[11] Wang L et al 2013 One-dimensional electrical contact to a two-dimensional material Science 342 614–7

[12] Zomer P J, Guimarães M H D, Brant J C, Tombros N andvan Wees B J 2014 Fast pick up technique for high qualityheterostructures of bilayer graphene and hexagonal boronnitride Appl. Phys. Lett. 105 013101

[13] Kretinin A V et al 2014 Electronic properties of grapheneencapsulated with different two-dimensional atomic crystalsNano Lett. 14 3270–6

[14] Parzefall M et al 2015 Antenna-coupled photon emission fromhexagonal boron nitride tunnel junctions Nat. Nanotechnol.10 1058–63

[15] Pizzocchero F et al 2016 The hot pick-up technique for batchassembly of van der Waals heterostructures Nat. Commun. 711894

[16] Frisenda R et al 2018 Recent progress in the assembly ofnanodevices and van der Waals heterostructures bydeterministic placement of 2D materials Chem. Soc. Rev. 4753–68

[17] Meitl M A et al 2006 Transfer printing by kinetic control ofadhesion to an elastomeric stamp Nat. Mater. 5 33–8

[18] Castellanos-Gomez A et al 2014 Deterministic transfer of two-dimensional materials by all-dry viscoelastic stamping 2DMater. 1 011002

[19] Kumar A and Whitesides G M 1993 Features of gold havingmicrometer to centimeter dimensions can be formed througha combination of stamping with an elastomeric stamp and analkanethiol ink followed by chemical etching Appl. Phys.Lett. 63 2002–4

[20] Duffy D C, McDonald J C, Schueller O J A andWhitesides G M 1998 Rapid prototyping of microfluidicsystems in poly(dimethylsiloxane) Anal. Chem. 704974–84

[21] Rogers J A, Someya T and Huang Y 2010 Materials andmechanics for stretchable electronics Science 327 1603–7

[22] Xia Y and Whitesides G M 1998 Soft lithography Angew.Chem., Int. Ed. Engl. 37 550–75

[23] Regehr K J et al 2009 Biological implications ofpolydimethylsiloxane-based microfluidic cell culture LabChip 9 2132–9

[24] Lee J N, Park C and Whitesides G M 2003 Solventcompatibility of poly(dimethylsiloxane)-based microfluidicdevices Anal. Chem. 75 6544–54

[25] Glasmästar K, Gold J, Andersson A-S, Sutherland D S andKasemo B 2003 Silicone transfer during microcontactprinting Langmuir 19 5475–83

[26] Briseno A L et al 2006 Patterning organic semiconductorsusing dry poly(dimethylsiloxane) elastomeric stamps for thinfilm transistors J. Am. Chem. Soc. 128 3880–1

[27] Böhm I, Lampert A, Buck M, Eisert F and Grunze M 1999 Aspectroscopic study of thiol layers prepared by contactprinting Appl. Surf. Sci. 141 237–43

[28] Yunus S, de Looringhe C d C, Poleunis C and Delcorte A 2007Diffusion of oligomers from polydimethylsiloxane stamps inmicrocontact printing: surface analysis and possibleapplication Surf. Int. Anal. 39 922–5

[29] Allen M J et al 2009 Soft transfer printing of chemicallyconverted graphene Adv. Mat. 21 2098–102

[30] Liu K et al 2014 Elastic properties of chemical-vapor-deposited monolayer MoS2, WS2, and their bilayerheterostructures Nano Lett. 14 5097–103

[31] Tongay S et al 2014 Tuning interlayer coupling in large-areaheterostructures with CVD-grown MoS2 and WS2monolayers Nano Lett. 14 3185–90

[32] Tongay S et al 2013 Broad-range modulation of light emissionin two-dimensional semiconductors by molecularphysisorption gating Nano Lett. 13 2831–6

[33] Tongay S et al 2013 Defects activated photoluminescence intwo-dimensional semiconductors: interplay between bound,charged, and free excitons Sci. Rep. 3 2657

[34] Yang L et al 2014 Chloride molecular doping technique on 2Dmaterials: WS2 and MoS2 Nano Lett. 14 6275–80

[35] Ouyang M, Yuan C, Muisener R J, Boulares A andKoberstein J T 2000 Conversion of some siloxane polymersto silicon oxide by UV/ozone photochemical processesChem. Mater. 12 1591–6

[36] Bodas D and Khan-Malek C 2007 Hydrophilization andhydrophobic recovery of PDMS by oxygen plasma andchemical treatment - an SEM investigation SensorsActuators B 123 368–73

[37] Morent R et al 2007 Adhesion enhancement by a dielectricbarrier discharge of PDMS used for flexible and stretchableelectronics J. Phys. D: Appl. Phys. 40 7392–401

[38] Haigh S J et al 2012 Cross-sectional imaging of individuallayers and buried interfaces of graphene-basedheterostructures and superlattices Nat. Mater. 11 764–7

[39] Cao Y et al 2015 Quality heterostructures from two-dimensional crystals unstable in air by their assembly ininert atmosphere Nano Lett. 15 4914–21

[40] Kang K et al 2017 Layer-by-layer assembly of two-dimensional materials into wafer-scale heterostructuresNature 550 229–33

8

Nanotechnology 29 (2018) 265203 A Jain et al

[41] Wierzbowski J et al 2017 Direct exciton emission fromatomically thin transition metal dichalcogenideheterostructures near the lifetime limit Sci. Rep. 7 12383

[42] Chow P K et al 2015 Defect-induced photoluminescence inmonolayer semiconducting transition metal dichalcogenidesACS Nano 9 1520–7

[43] Buscema M, Steele G A, van der Zant H S J andCastellanos-Gomez A 2014 The effect of the substrate on theRaman and photoluminescence emission of single-layerMoS2 Nano Res. 7 561–71

[44] Schuller J A et al 2013 Orientation of luminescent excitons inlayered nanomaterials Nat. Nanotechnol. 8 271–6

[45] Amani M et al 2015 Near-unity photoluminescence quantumyield in MoS2 Science 350 1065–8

[46] Li H et al 2012 From bulk to monolayer MoS2: evolution oframan scattering Adv. Funct. Mater. 22 1385–90

[47] Conley H J et al 2013 Bandgap engineering of strainedmonolayer and bilayer MoS2 Nano Lett. 13 3626–30

[48] Nayak A P et al 2015 Pressure-dependent optical andvibrational properties of monolayer molybdenum disulfideNano Lett. 15 346–53

[49] Dou X, Ding K, Jiang D, Fan X and Sun B 2016 Probing spin–orbit coupling and interlayer coupling in atomically thinmolybdenum disulfide using hydrostatic pressure ACS Nano10 1619–24

[50] Liu Z et al 2014 Strain and structure heterogeneity in MoS2atomic layers grown by chemical vapour deposition Nat.Commun. 5 5246

[51] Chakraborty B et al 2012 Symmetry-dependent phononrenormalization in monolayer MoS2 transistor Phys. Rev. B85 161403

[52] peak-o-mat http://lorentz.sourceforge.net/

9

Nanotechnology 29 (2018) 265203 A Jain et al

Supplementary Information

Minimizing Residues and Strain in 2D Materials Transferred from PDMS

Achint Jain,1 Palash Bharadwaj,2 Sebastian Heeg,1 Markus Parzefall,1

Takashi Taniguchi,3 Kenji Watanabe,3 and Lukas Novotny1

1Photonics Laboratory, ETH Zurich, 8093 Zurich, Switzerland2Department of Electrical and Computer Engineering,

Rice University, Houston, TX 77005, USA3National Institute for Material Science, 1-1 Namiki, Tsukuba 305-0044, Japan

List of Figures

S1 Another MoS2 flake with PDMS residues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii

S2 Visualizing PDMS residues in an optical microscope . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii

S3 An example of clean MoS2 before annealing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

S4 Additional example of clean MoS2 transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

S5 Bubbles at the MoS2 - SiO2 interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi

S6 Lorentzian fits to the PL spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

S1. Ineffectiveness of annealing in removing PDMS residues

Residues are always present to a varying degree on flakes transferred from PDMS. Here we show

an exceptional case of residues on a 1L-MoS2 flake which was exfoliated on untreated PDMS (Fig.

S1a inset) and transferred to hBN (Fig. S1a). The AFM topography map of the as transferred

flake in Fig. S1b clearly reveals that it is covered in a thick coat of PDMS residues in most places.

As indicated by the red cross-section in Fig. S1c, the thickness of this residue layer is >5 nm near

the MoS2 edge. It is a common practice to anneal 2D materials to clean surface residues after

transfer and enhance interlayer coupling in heterostructures. However, as shown in the main text,

PDMS residues are difficult to completely remove simply by annealing. Figure S2d is an AFM

map of the same flake after 3 h annealing at 200 ◦C in vacuum. Bubbles and wrinkles were elimi-

nated as expected and the total thickness came down to ∼1.6 nm (green profile in Fig. S2c) upon

annealing. But this is still not comparable to the thickness of ∼0.7 nm exhibited by monolayer

MoS2 transferred from UV-O3 treated PDMS after annealing. These results demonstrate that

substantial residues are frequently present on MoS2 flakes transferred from untreated PDMS and

can’t be removed solely by annealing which highlights the need for a PDMS pre-cleaning technique

in order to build cleaner heterostructures.

ii

a b

c d

Bottom h-BN

6 nm

1L- MoS2

1.6 nm

b

Annealed

As transferredPDMS

10 μm

Figure S1. Another MoS2 flake with PDMS residues. Optical image of another MoS2 flake transferredonto hBN from untreated PDMS. The monolayer segment has been outlined. Inset: The same MoS2 flake onPDMS before transfer. (b) AFM topography map of the region outlined in a displaying thick, inhomogeneouslayers of residues covering most of the MoS2 flake. (c) Height profile along the red and green dashed lines inb, d indicating the total thickness before and after annealing respectively. (d) AFM map of the same regionafter 3 h vacuum annealing. PDMS residues are still present as evident from the green profile although theflake is nearly free from bubbles/wrinkles.

Besides UV-O3 treatment, we also tried other alternate approaches to clean the PDMS surface

before exfoliation. One possible way to clean PDMS is via extraction of uncrosslinked oligomers

from bulk PDMS by soaking it in organic solvents for several hours1. But we found that it is

difficult to completely extract all oligomers this way2 and being a lengthy procedure makes it

inconvenient for practical use. Another alternative to UV-O3 for breaking down organic species is

O2 plasma cleaning. However, we noticed that O2 plasma can be too harsh for PDMS and could

lead to the formation of fine cracks on the PDMS surface after few minutes exposure at 100 W

power, similar to observations made by Bodas et al.3. Moreover, O2 plasma treated PDMS surface

gave us a very poor yield of flakes upon exfoliation as very few of them stuck to it compared to

UV-O3 treated PDMS. The O2 plasma treatment process could possibly be optimized but we did

not investigate it further. It is also important to mention that the choice of PDMS didn’t make a

big difference. We also tried PDMS films synthesized using Sylgard 184 (Dow Corning) but found

similar residues as with Gel-Pak films.

iii

S2. Visualizing PDMS residues in an optical microscope

Under certain circumstances, PDMS residues on a Si/SiO2 (285 nm) substrate can be even

identified simply in an optical microscope. The apparent color of the Si/SiO2 substrate as seen

in reflection arises from interference and is quite sensitive to the thickness and refractive index of

an additional dielectric layer on top. In case of MoS2 transferred from PDMS to SiO2 without

the application of heat during the transfer (i.e. at room temperature), the residual PDMS layer

can sometimes be thick enough to cause a detectable change in the color of the Si/SiO2 substrate.

Figure S3a is an optical microscope image of MoS2 flakes exfoliated on untreated PDMS. The

PDMS was brought in contact with SiO2 and then slowly detached, without any intermediate

heating step while in contact. Among the flakes visible in Fig. S3a, the large thick MoS2 flake in

the lower part did not get transferred to SiO2. Interestingly, a clear outline of the missing flake

on the SiO2 substrate can be easily seen in Fig. S3b after transfer. In this region where the thick

MoS2 flake prevented the PDMS from coming in contact with the SiO2, the substrate retained

its original color whereas in areas where PDMS came in direct contact with SiO2, transferred

residues led to a change in color. This gave rise to a visible contrast between clean and PDMS

contaminated SiO2 surface which is easily noticeable in an optical microscope. For transfers done

at 65 ◦C, PDMS residues are harder to see directly but they can still be visualized by differential

interference contrast (DIC) microscopy (images not included here).

a bSiO2 withPDMS residues

Untreated PDMS

MoS2 flakemissing

10 μm 10 μm

dirty

Change in color

clean

Figure S2. Visualizing PDMS residues in an optical microscope. (a) Optical image of MoS2 flakesexfoliated on untreated PDMS. The image has been mirrored for an easier comparison with the right-handside image. (b) Optical image of the same flakes after transfer to SiO2. The large MoS2 flake did notget transferred but its outline is still clearly visible. This is due to a change in color of the surroundingareas which came in direct contact with PDMS and got contaminated. The clean SiO2 area has been partlydemarcated to serve as a guide to the eye.

iv

S3. Additional data on clean transfer from UV-O3 treated PDMS

In this section we show two more examples of clean MoS2 flakes transferred using our new

recipe. Figure S3a is an optical image of a 1L-MoS2 flake exfoliated on UV-O3 treated PDMS

(inset) and transferred to hBN. The AFM topography map of the as-transferred flake in Fig. S3b

shows a mostly clean surface (though slight traces of residues can be identified in the lower right

part). Remarkably, even a monolayer hBN terrace can be resolved in this image which highlights

the superior cleanliness of MoS2 transferred this way compared to that from untreated PDMS in

Fig. 1 of the main text. The higher resolution AFM map in Fig. S3c reveals a pristine surface

(except for bubbles) with a clean step of 6.9�A (Fig. S3d) and the corresponding AFM phase map

in Fig. S3e displays strong phase contrast between MoS2 and hBN. These images are quite unlike

Figs. 1e-i where the step height and phase contrast were both obscured by residues. It is important

to mention that these maps were recorded on as-transferred MoS2 before annealing but are still

remarkably clean. Figure S3f is an AFM topography map of the same flake after vacuum annealing

showing a relaxed surface without even a slight trace of residues. Moreover, the MoS2 in the lower

a b c

d e f

Bottom h-BN

nm

10 μm

1L-MoS2

PDMS(UV-O3 treated)

c

b

nm

As transferred

Step in

As transferred

1L-hBNterrace

hBN

6.9 Å

AFMPhase

nm

Vacuum annealed

Figure S3. An example of clean MoS2 before annealing. (a) Optical microscope image of another1L-MoS2 flake transferred from UV-O3 treated PDMS. (b) AFM topography map of the as-transferredflake taken from the region outlined in a. Even a 1L-hBN terrace is resolvable in this image which is a clearevidence of a significantly cleaner surface compared to Fig. 1 of the main text. (c) Higher resolution AFMmap of the region outlined in b showing the residue-free surface of the as-transferred MoS2 without anyannealing. (d) Height profile along the red dashed line in c displaying a clean step of 6.9�A correspondingto pristine monolayer MoS2. (e) AFM phase map corresponding to the topography in c exhibiting a strongphase contrast between MoS2 and hBN which again points towards the absence of residues. (f) AFM mapof the same region as in b after vacuum annealing.

v

right part makes much better contact with the hBN below after spreading out during annealing

leading to an apparent reduction in thickness. Hence, this leads us to conclude that a combination

of UV-O3 pre-treatment followed by vacuum annealing is the optimum for obtaining a pristine

MoS2 surface with an unperturbed lattice.

In Fig. 2 of the main text, we only presented AFM results obtained from the left (triangular)

1L-MoS2 flake in Fig. 2b after transfer from UV-O3 treated PDMS. Here we have included addi-

tional data from the right (trapezoidal) 1L-MoS2 attached to the same flake. AFM maps before

and after annealing are depicted in Fig. S4a, b respectively and exhibit a smooth, homogeneous,

residue-free surface after annealing in contrast to Figs. 2a and S2d. In the high-resolution map in

Fig. S4c, a pristine MoS2 surface with a thickness of 8.5�A (Fig. S4d) can be clearly seen. These

results in line with those in Fig. 3 and highlight the effectiveness of our UV-O3 pre-cleaning process

in reducing residues.

a b

c d

8.5 Å

AnnealedAs transferred

c

nm nm

nm

Annealed

Figure S4. Additional example of clean MoS2 transfer. (a) AFM map of the right 1L-MoS2 flakein Fig. 2b of the main text. (b) AFM map of the same area after vacuum annealing. (c) Higher resolutionmap of the region outlined in b exhibiting a pristine surface. The small undulations in topography are notvariations in the MoS2 thickness but arise from the substrate below. (d) Height profile along the red dashedline in c displaying a thickness (8.5�A) very close to that of monolayer MoS2.

vi

S4. Effect of annealing on MoS2 lying on SiO2

So far we have presented several examples demonstrating efficient removal of bubbles from the

MoS2-hBN interface via vacuum annealing. However, annealing doesn’t seem to be very effective

in getting rid of bubbles accumulated at the MoS2-SiO2 interface. Figure S5a is an optical image

of a long MoS2 flake transferred from UV-O3 treated PDMS onto a Si-SiO2 substrate (pre-cleaned

with O2 plasma). AFM map of the as transferred 1L-MoS2 region in Fig. S5b reveals a high density

of tiny bubbles trapped between MoS2 and SiO2. Unfortunately, after 3 h vacuum annealing at

200 ◦C, the distribution of bubbles in Fig. S5c looks quite similar to that before annealing, unlike

in case of MoS2 on hBN. To understand these contrasting behaviors, we must take into account

the strong adhesion between MoS2 and SiO2 due to Coulombic attraction arising from dangling

bonds and surface charges on SiO2. Moreover, it can be noticed in the high-resolution map in Fig.

S5d that the bubbles on SiO2 are much smaller in size than those seen previously in MoS2 on hBN

(cf. Fig. 1e) which also points towards a higher adhesion energy for MoS2 on SiO24,5. This strong

adhesion together with the increased sliding friction6 caused by the higher surface roughness of

SiO2 keeps MoS2 anchored in place during annealing. On the other hand, only a weak van der

Waals attraction exists between MoS2 and hBN which allows a greater freedom of movement for

a b

c d

SiO2

1L- MoS2

b

Annealed

As transferred

nm

d

1.0 nm

nm

nm

10 μm

Annealed

Figure S5. Bubbles at the MoS2-SiO2 interface. (a) Optical microscope image of a 1L-MoS2 flaketransferred to SiO2 (285 nm) from UV-O3 treated PDMS. (b) AFM topography map of the black outlinedregion in a displaying a high density of tiny bubbles. (c) AFM map of the same region after 3 h vacuumannealing with the distribution of bubbles largely unaffected by annealing. (d) Higher resolution scan ofthe white outlined region in c. Inset: height profile along the red dashed line

vii

MoS2. Moreover, the reduced sliding friction6 on the atomically smooth hBN surface makes it

easier for ripples in the MoS2 layer and mobile species trapped inside bubbles to slide out during

annealing which would otherwise have been immobilized on SiO2. This behavior illustrates that

the choice of substrate plays a crucial role in governing the surface morphology of transferred flakes

and deserves greater attention in future studies.

S5. Fitting the photoluminescence spectra

Here we show the complete set of fits for the photoluminescence (PL) spectra plotted in Fig.

3d of the main text. In Figs. S6a and b, the dark blue and orange data points were fitted with

a model comprising of a sum of 4 and 3 Lorentzian functions respectively, with each Lorentzian

representing one spectral feature. From the fits, a shift of 10 meV in the A-exciton peak position

was deduced for this pair of spectra. The trion peak showed a much smaller shift of ∼3 meV.

a

bUV-O3 + anneal

Untreated

Figure S6. Lorentzian fits to the PL spectra. (a) Sum of four Lorentzians (red) fitted to the PLspectra of MoS2 transferred from untreated PDMS. Along with the fitted sum, the individual fit componentsare also plotted separately. (b) Sum of three Lorentzians (red) fitted to the PL spectra of MoS2 transferredfrom UV-O3 treated PDMS and annealed. The empty circles are experimentally measured data points(reproduced from Fig. 3d of the main text) while the smooth curves are fits to the measured data.

viii

1 Lee, J. N., Park, C. & Whitesides, G. M. Solvent compatibility of poly(dimethylsiloxane)-based microflu-idic devices. Anal. Chem. 75, 6544–6554 (2003).

2 Regehr, K. J. et al. Biological implications of polydimethylsiloxane-based microfluidic cell culture. LabChip 9, 2132–2139 (2009).

3 Bodas, D. & Khan-Malek, C. Hydrophilization and hydrophobic recovery of PDMS by oxygen plasma andchemical treatment - An SEM investigation. Sensors and Actuators B: Chemical 123, 368 – 373 (2007).

4 Khestanova, E., Guinea, F., Fumagalli, L., Geim, A. K. & Grigorieva, I. V. Universal shape and pressureinside bubbles appearing in van der Waals heterostructures. Nature Commun. 7, 12587 (2016).

5 Lloyd, D. et al. Adhesion, stiffness, and instability in atomically thin MoS2 bubbles. Nano Lett. 17,5329–5334 (2017).

6 Quereda, J., Castellanos-Gomez, A., Agraıt, N. & Rubio-Bollinger, G. Single-layer MoS2 roughness andsliding friction quenching by interaction with atomically flat substrates. Appl. Phys. Lett. 105, 053111(2014).