• No results found

A real-time Raman spectroscopy study of the dynamics of laser-thinning of MoS2 flakes to monolayers

N/A
N/A
Protected

Academic year: 2022

Share "A real-time Raman spectroscopy study of the dynamics of laser-thinning of MoS2 flakes to monolayers"

Copied!
10
0
0

Loading.... (view fulltext now)

Full text

(1)

A real-time Raman spectroscopy study of the dynamics of laser-thinning of MoS2 flakes to monolayers

Enyao Gu, Qiyuan Wang, Youwei Zhang, Chunxiao Cong, Laigui Hu, Pengfei Tian, Ran Liu, Shi-Li Zhang, and Zhi-Jun Qiu

Citation: AIP Advances 7, 125329 (2017);

View online: https://doi.org/10.1063/1.5008441

View Table of Contents: http://aip.scitation.org/toc/adv/7/12 Published by the American Institute of Physics

Articles you may be interested in

Magnetron sputtering cathode for low power density operation AIP Advances 7, 125225 (2017); 10.1063/1.5013341

Implementation of one-dimensional domain wall dynamics simulator AIP Advances 7, 125231 (2017); 10.1063/1.4996029

Edge-spin-derived magnetism in few-layer MoS2 nanomeshes AIP Advances 7, 125019 (2017); 10.1063/1.4989477

Prolonged silicon carbide integrated circuit operation in Venus surface atmospheric conditions AIP Advances 6, 125119 (2016); 10.1063/1.4973429

Research on c-HfO2 (0 0 1)/ -Al2O3 (1 -1 0 2) interface in CTM devices based on first principle theory AIP Advances 7, 125001 (2017); 10.1063/1.5001904

Modulating PL and electronic structures of MoS2/graphene heterostructures via interlayer twisting angle Applied Physics Letters 111, 263106 (2017); 10.1063/1.5011120

(2)

A real-time Raman spectroscopy study of the dynamics of laser-thinning of MoS

2

flakes to monolayers

Enyao Gu,1Qiyuan Wang,1Youwei Zhang,1,2Chunxiao Cong,1,aLaigui Hu,1 Pengfei Tian,1Ran Liu,1Shi-Li Zhang,2,aand Zhi-Jun Qiu1,a

1State Key Laboratory of ASIC & System, School of Information Science and Technology, Fudan University, Shanghai 200433, China

2Solid-State Electronics, The Ångstr¨om Laboratory, Uppsala University, Uppsala SE-751 21, Sweden

(Received 6 October 2017; accepted 21 December 2017; published online 29 December 2017)

Transition metal dichalcogenides (TMDCs) in monolayer form have attracted a great deal of attention for electronic and optical applications. Compared to mechanical exfoliation and chemical synthesis, laser thinning is a novel and unique “on-demand”

approach to fabricate monolayers or pattern desired shapes with high controllabil- ity and reproducibility. Its successful demonstration motivates a further exploration of the dynamic behaviour of this local thinning process. Here, we present an in-situ study of void formation by laser irradiation with the assistance of temporal Raman evolution. In the analysis of time-dependent Raman intensity, an empirical formula relating void size to laser power and exposure time is established. Void in thinner MoS2flakes grows faster than in thicker ones as a result of reduced sublimation tem- perature in the two-dimensional (2D) materials. Our study provides useful insights into the laser-thinning dynamics of 2D TMDCs and guidelines for an effective con- trol over the void formation. © 2017 Author(s). All article content, except where otherwise noted, is licensed under a Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).https://doi.org/10.1063/1.5008441

INTRODUCTION

Layered two-dimensional (2D) materials have gained interest for their unique properties and atomic-scale thin film structure. This family of materials has lately seen a rapid growth in number of family members, now extending far beyond graphene to also include black phosphorus, metal chalco- genides, metal carbides and transition metal oxides.1,2Transition metal dichalcogenides (TMDCs) alone can occur in more than 40 different categories and span the entire range of electronic struc- tures.1,3 Those that are semiconducting characterised by excellent electrical properties especially when they are thinned to monolayer thickness, are considered to be the most promising channel mate- rial for thin film transistors.1,3,4 As an example, bulk MoS2has an indirect band gap of 1.2∼1.3 eV.

It transforms to a semiconductor of direct band gap measuring 1.9 eV when thinned to a monolayer.4–7 Layered MoS2has a broad range of potential applications in gas sensing,8,9energy conversion (solar cell)10and energy storage (ion battery).11,12However, most of the aforementioned unique properties of 2D materials only persist when the number of repeating layers is reduced to one or close to one.

Effective fabrication routes for production of monolayers of 2D materials have motivated great efforts in the 2D-material community.

Several methods have been explored for producing ultrathin 2D materials with a suitable number of layers, such as mechanical exfoliation,13,14 liquid-phase exfoliation,15,16chemical vapour depo- sition (CVD),17 molecular beam epitaxy,18 electrochemical thinning19,20 and laser thinning.21–26

aAuthors to whom correspondence should be addressed. Electronic mail: cxcong@fudan.edu.cn; shili.zhang@

angstrom.uu.se;zjqiu@fudan.edu.cn.

2158-3226/2017/7(12)/125329/9 7, 125329-1 © Author(s) 2017

(3)

125329-2 Gu et al. AIP Advances 7, 125329 (2017)

Mechanical exfoliation, which can produce layers of the highest quality, is a random process with low yield. Liquid-phase exfoliation, though possible to economically produce large volumes of dis- persed layers, can unavoidably introduce impurities and residual solvent and yields a low fraction of monolayers. CVD is suitable for growth of large-area thin layers, but it is a high temperature process thereby limiting the choice of growth substrates. Laser thinning is a relatively new technique for

“on-demand” fabrication of monolayers and patterning of desired shapes with high controllability and reproducibility.3,22,25 This technique involves temporary illumination of a focused laser beam onto thin flakes without needing complex environment control. It presents thus an efficient and easy top-down method to thin and fabricate certain patterns via a convenient design and at a controllable laser power level. The accumulation of heat induced by laser irradiation can easily lead to sublima- tion of the upper few layers of the flake while the bottom-most monolayer can remain intact due to efficient heat dissipation to the substrate when the laser power, P0, is controlled below a certain level, e.g. 10 - 20 mW for MoS2thinning.22Additionally, the use of laser irradiation to thin or pattern some features in 2D flakes can avoid involving photoresist or mask, thus eliminating a source of organic contamination or change of electronic structure. The monolayer MoS2 produced by laser thinning has shown a comparable mobility to that of pristine one.9,22,26

Although the laser thinning technique has shown several advantages including rapid material processing, large scan area, high spatial resolution and single-step capability, its dynamic process is less understood. In the present work, confocal Raman spectroscopy is used both as the thinning laser source and as the light source for in-situ real-time monitoring of the laser thinning process.

This arrangement facilitates rapid characterisation of the same area before, during and after thinning, which allows for a systematic investigation of the dynamics of the laser-induced void formation as a function of P0and exposure time, t. An empirical formula to describe the relationship between void size and these two laser parameters is obtained.

METHODS

MoS2 flakes from natural crystalline MoS2 (SPI supplies) were deposited onto a SiO2/Si substrate by mechanical exfoliation. Prior to exfoliation, the Si substrate was pre-cleaned using sonication in acetone, isopropyl alcohol and deionized water. The number of MoS2 layers on the substrate was identified using an optical microscope (Keyence digital microscope VHX-600), atomic force microscopy (AFM, Dimension 3100, Veeco) and Raman spectroscopy (Renishaw inVia). Raman spectra of the MoS2 flakes were acquired in a backscattering configuration with a 514 nm laser. After careful location and focusing of the laser spot on each specimen, the Raman scattered signal was collected with a 100× objective (NA = 0.85) and dispersed by a grat- ing of 1800 lines/mm. The spectral resolution was ∼ 1.0 cm-1 and peak position accuracy was 0.1 cm-1. To identify the number of layers, Raman spectra were obtained with P0 below 1 mW and an exposure time of 10 s. However, for laser thinning process, P0 was increased up to 20 mW and the exposure time varied from 1 to 100 s. P0was directly measured using a laser power meter (Sanwa Laser Power Meter LP1 Mobiken Series) and carefully adjusted prior to each laser irradiation.

RESULTS AND DISCUSSION

The samples used in our laser thinning experiments were few-layer MoS2 flakes on SiO2/Si substrate exfoliated from a bulk MoS2crystal using a mechanical exfoliation technique. The location of the flakes was identified using optical microscope. The optical image of several MoS2 flakes is shown in Fig.1(a). The optical contrast of MoS2is found to increase with the number of layers. The layer number before thinning was determined by the frequency shift of specific vibrational modes using Raman spectroscopy with a low-power laser (<1 mW) in order to avoid heating effect. Fig.1(b) presents two prominent Raman peaks E2g1(in-plane vibration mode) and A1g(out-of-plane vibration mode) of the MoS2 flakes. It is known that the Raman shift difference between these two peaks increases with the number of layers,27as demonstrated in Fig.1(c).

(4)

FIG. 1. (a) Optical image of exfoliated MoS2flakes on SiO2/Si substrate. (b) Raman spectra of pristine MoS2flakes of 1 L to 7 L in thickness. (c) Variation of the frequency difference between the A1gand E2g1phonon modes with the number of layers. (d) Raman spectra of MoS2monolayers laser-thinned from pristine multilayers.

This same Raman setup was also used for the MoS2 thinning experiments, but with a higher incident P0ranging from 10 mW to 20 mW. The Si substrate is an efficient heat sink. Consequently, only the bottom monolayer directly in contact with the substrate can be preserved without being ablated by localised laser heating, while the upper layers are sublimed due to poor interlayer thermal coupling. The nominal temperature for sublimation of a natural MoS2 crystal is 450oC.28 Under laser irradiation at 10 mW or above, the lattice temperature of the irradiated part can exceed the sublimation temperature within a few seconds.29 This transient laser heating is sufficient to break the weak interlayer van der Waals bonds and then sublimes the upper layers readily. At an appro- priate P0, the bottom monolayer is preserved regardless of the initial number of layers. Fig.1(d) shows the Raman spectra of thinned few-layer MoS2 flakes after laser irradiation. The frequency difference, also depicted in Fig.1(c), is consistently reduced to ∼ 21 cm-1 from those of pristine MoS2flakes of different number of layers. This frequency difference is slightly higher than that of a pristine MoS2monolayer and can arise from shifts of Raman modes due to the presence of traces of unremoved MoS2.

Laser thinning of the exfoliated MoS2flakes to a single monolayer has been successfully demon- strated, but its dynamics is so far unexplored. The Raman microscopy performed in situ presents an efficient method for real-time monitoring of the laser thinning process. The temporal behaviour of Raman intensity is achieved when Raman spectra are collected continuously during laser thinning.

When a high-power laser is focused on the few-layer MoS2flakes, a small void is formed quickly with the bottom monolayer remaining intact. Fig.2(a)is an AFM image of such a void formed in a 7-layer (7 L) MoS2flake. About 3.6-nm-thick MoS2, corresponding to 6 L, is removed altogether while the bottom monolayer stays. The particles around the void rim are likely to originate from the atmo- sphere due to laser-induced heating effect.29The void is then found to laterally grow in size with t.

As a result of the upper layers being etched laterally, the integrated E2g1 peak intensity decreases with t. Fig.2(b) exhibits an initial rapid decay of the peak intensity followed by a slow decline.

However, it is important to note that the E2g1peak did not disappear completely over the duration of the experiment. Rather, it decreased to and then levelled off at a small value, indicative of thinning of the MoS2flake to the monolayer thickness that is resistive to further decomposition.

(5)

125329-4 Gu et al. AIP Advances 7, 125329 (2017)

FIG. 2. (a) AFM topographic image of a void formed in a 7 L MoS2flake by laser irradiation at 15 mW for 50-s. (b) Temporal evolution of the integrated E2g1peak intensity upon laser irradiation.

To gain additional insights into the decay dynamics, theoretical modelling was pursued by starting from this time-dependent Raman intensity. Here, the intensity distribution of the focused laser, p(r), is assumed to be Gaussian:

p(r)= P0 2πσ02exp *

,

r202+

-

(1)

where r the radial distance from the laser spot centre and r0 the laser spot size taken to be the full width at half maximum intensity given by r0= σ0(2ln2)1/2. To estimate the laser spot size, we used the acquired Raman intensity obtained by scanning a low-power laser spot from the SiO2

surface over to the edge of MoS2 flake. This intensity can be regarded as the convolution of the laser intensity profile and the topography of the MoS2 flake. In Fig.3(a,b), the optical image of a MoS2 flake with a sharp edge and the relative Raman intensity along a linecut are, respectively, depicted. By applying a numerical integration based on Eq. (1) with the best fit to the Raman data, we obtained a laser spot size of ∼ 0.2 µm, which is in good agreement with a diffraction limited spot radius (∼ 0.3 µm).

During the laser-thinning process, a void is expected to grow in size with t once it is formed.

Since the bottom-most monolayer in the void is persistently well preserved under laser irradia- tion at the laser-power range used here, the Raman intensity recorded during the in-situ analysis should include contributions from the bottom-most monolayer as well as the unremoved few-layers, as illustrated in Fig.4(a). According to Eq. (1), the time-dependent Raman intensity, I(t), can be

FIG. 3. (a) Optical image of the edge of a MoS2flake. The focused laser is scanned over a white linecut while the scattered Raman signal is collected. (b) Normalized Raman intensity and best fit using Eq. (1) as a function of the distance from the laser spot center to the MoS2edge.

(6)

FIG. 4. (a) Illustration of the distribution of P0density and the thinned monolayer embedded in a few-layer MoS2flake.

Raman signals obtained in-situ include contributions from both MoS2monolayer and multilayer. (b) Integrated E2g1intensity as a function of P0for pristine MoS2with different numbers of layers. Lines are linear fits to experimental data points.

(c) Dependence of K value on number of layers. Red circles represent K values of monolayer laser-thinned from pristine multilayers.

written as:

I(t)=K0P0 2πσ20

 rH(t)

0

exp * ,

r220+

-

2πrdr + K1P0 2πσ20

 rH(t)

exp * ,

r220+

- 2πrdr

= (K1K0)P0exp

rH2(t)02

+ K0P0 (2)

where rHis the radius of a laser-etched void and K proportionality coefficients reflecting the Raman intensity excited by a unit P0. The subscripts 0 and 1 refer to monolayer and few-layer MoS2, respectively. In order to obtain the proportionality coefficient, exfoliated MoS2 flakes of different numbers of layers were laser illuminated at P0below 1 mW. Fig.4(b)shows the variation of integrated E2g1Raman intensity with P0. A linear fit to the data points yields K for the various few layer films, as shown in Fig.4(c). The parabolic-like behaviour of K with number of layers is a result of multi- reflection and interference of the light-beams reflected from the MoS2 and the Si/SiO2 interface,27 which is also observed in graphene on SiO2/Si substrate.30Fig.4(c)also shows the K value of those thinned MoS2under a high P0of 15 mW. It is found that the K value after laser thinning is close to that of pristine monolayers, which confirms that few-layer MoS2are successfully reduced to monolayers by laser irradiation at the P0range used here.

Starting from Eq. (1) and (2), the power density at the void edge, p(rH), can be correlated to I(t), as the following:

p(rH)= 1 2πσ20

I(t) − K0P0

K1K0 (3)

In Fig.5, p(rH) is plotted as a function of t. In the logarithmic scale, p(rH) is found to first decrease linearly with t and then flattens out when t further increases and the void radius approaches its max- imum value. This behaviour is independent of the P0 used and it holds true for all the few-layer cases studied here. Hence, p(rH) is only determined by t but not by P0, which further confirms the validity of Eq. (1). According to Eq. (3), higher P0 induces larger void, keeping the resultant p(rH) constant. Taking 4 L MoS2 as an example, Fig. 6(a)illustrates the calculated p(rH) using

(7)

125329-6 Gu et al. AIP Advances 7, 125329 (2017)

FIG. 5. (a-f) Dependence of p(rH) on t for 2 L – 7 L MoS2flakes at different P0. Solid lines are linear fits in the logarithmic scale.

Eq. (1) as a function of void radius at 15-mW laser irradiation. In combination with Fig. 5(c), each p(rH) corresponds to a definite t and the thus-obtained void radius. For example, an irradia- tion for 5 s, 20 s and 50 s leads to a void radius of 0.12 µm, 0.17 µm and 0.20 µm, respectively.

Fig.6(b)shows the variation of p(rH) with void radius at different P0. Under the same t of 20 s,

FIG. 6. (a, b) Dependence of p(rH) on rHobtained from Eq. (1). In conjunction with Fig.5, p(rH), t and rHare all correlated.

Taking 4 L as an example, i.e. Fig.5(c), irradiations for 5 s, 20 s and 50 s lead to rHequal to 0.12 µm, 0.17 µm and 0.20 µm, respectively, as highlighted with illustrations and red arrows shown in (a). For t=20 s, P0of 13 mW, 15 mW and 17 mW induces rHof 0.13 µm, 0.17 µm and 0.20 µm, respectively, as highlighted with similar illustrations and red arrows shown in (b).

(8)

FIG. 7. Variation of parameters C and D in Eq. (4) with number of layers. In contrast to C, D is remarkably layer dependent.

P0 of 13 mW, 15 mW and 17 mW induces a void radius of 0.13 µm, 0.17 µm and 0.20 µm, respectively.

From Fig.5, an empirical formula correlating p(rH) to t can be obtained from the linear fitting in the logarithmic scale, as shown in the following:

t= Cp(rH)−D

(4) Here, C and D are two empirical parameters and their variations with number of layers are given in Fig.7. It can be seen that D is smallest for bilayer. It then increases with the number of layers.

The variation of C is, however, much less pronounced. From the empirical formula in Eq. (4), it can be deduced that for the same p(rH), parameter D is a major factor that determines t that is also the void growth time. Substituting Eq. (4) into Eq. (1), we can obtain a relationship between rH, P0and t as follows:

rH=√ 2σ0

vt ln *

, CP0 2πσ02+

- +lnt

D (5)

Hence, the void radius can be readily obtained with known P0 and t. Herein, the smallest D value in a bilayer MoS2 is equivalent to the quickest void expansion or easiest decomposition of

FIG. 8. (a-d) rHas a function of t for 2 L to 5 L MoS2flakes. Symbols are experimental data extracted from SEM measurements.

(9)

125329-8 Gu et al. AIP Advances 7, 125329 (2017)

FIG. 9. (a) AFM image of an etched void covered with particles. (b) SEM image of the same void in (a), where the etched profile is clearly visible.

the flake. Thinner layers are generally characterised by higher surface area to volume ratios that are essential for their chemical and thermal stability. Such thickness-dependent stability has been observed in graphene, MoS2 and black phosphorus.31–33 Thinner layers also have higher oxidiza- tion rates and lower sublimation temperatures, which explains why bilayer MoS2decomposes more quickly than its thicker counterparts. However, the strong thermal-coupling interaction of mono- layer MoS2 with the underlying substrate results in a relatively low temperature rise during laser irradiation. It is, therefore, immune to decomposition. The calculated rHfor the different MoS2few- layers is plotted in Fig. 8as a function of t. As expected, rH initially increases rapidly for about 10 s. It then gradually grows up to the approximate laser spot size. In some cases, the presence of adsorbed particles around the void makes it difficult to ascertain the void edge or topography by AFM measurement, as shown in Fig. 9(a). In comparison, scanning electron microscopy (SEM) can clearly reveal the void profile and, hence, its radius can be reliably extracted in Fig. 9(b).

It can be seen that the experimentally determined rHis consistent with the theoretical expectation in Fig.8.

CONCLUSIONS

In summary, a confocal Raman spectroscopy has been employed to induce void formation in MoS2 flakes of different numbers of layers, to in-situ characterise the void formation and to study the dynamics of the laser-thinning process as a function of laser power and exposure time. Raman analysis reveals that the void size is affected by these two laser parameters and eventually approaches the laser spot size at long exposure duration. In the analysis of time-dependent Raman intensity, an empirical formula correlating the void radius to the laser power and exposure time has been devel- oped. The calculation results are found to agree well with experimental data obtained by means of SEM analysis. Due to the higher surface area to volume ratio, thinner MoS2 flakes especially in bilayer form, are less stable and easier to decompose under a high-power laser irradiation, lead- ing to faster void expansion. Our finding provides an in-depth understanding of the laser-thinning process, which is essential to exploit for future fabrication of monolayer or nanostructure in other 2D TMDCs.

ACKNOWLEDGMENTS

This work was supported by National Natural Science Foundation of China (No. 61774042, 61774040), Shanghai Municipal Natural Science Foundation (No. 16ZR1402500, 17ZR1446500, 17ZR1446600), National Young 1000 Talent Plan of China, Shanghai Pujiang Program (No.

16PJ1401000), "First-Class Construction" project of Fudan University and the Swedish Research Council (VR, No. 621-2014-5591).

(10)

1M. Xu, T. Liang, M. Shi, and H. Chen,Chem. Rev. 113, 3766 (2013); S. Leb`egue, T. Bj¨orkman, M. Klintenberg, R. M. Nieminen, and O. Eriksson,Phys. Rev. X3, 031002 (2013).

2J. S. Ponraj, Z. Q. Xu, S. C. Dhanabalan, H. Mu, Y. Wang, J. Yuan, P. Li, S. Thakur, M. Ashrafi, K. McCoubrey, Y. Zhang, S. Li, H. Zhang, and Q. Bao,Nanotechnology27, 462001 (2016).

3S. Balendhran, S. Walia, H. Nili, J. Z. Ou, S. Zhuiykov, R. B. Kaner, S. Sriram, M. Bhaskaran, and K. Kalantar-zadeh, Adv. Funct. Mater.23, 3952 (2013).

4Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, and M. S. Strano,Nat. Nanotechnol.7, 699 (2012).

5T. Boker, R. Severin, A. Muller, C. Janowitz, R. Manzke, D. Voss, P. Kruger, A. Mazur, and J. Pollmann,Phys. Rev. B64, 235305 (2001).

6Y. Ding, Y. Wang, J. Ni, L. Shi, S. Shi, and W. Tang,Phys. B406, 2254 (2011).

7C. P. Lu, G. Li, J. Mao, L. M. Wang, and E. Y. Andrei,Nano Lett14, 4628 (2014).

8Q. He, Z. Zeng, Z. Yin, H. Li, S. Wu, X. Huang, and H. Zhang,Small8, 2994 (2012).

9H. Li, Z. Yin, Q. He, H. Li, X. Huang, G. Lu, D. W. Fam, A. I. Tok, Q. Zhang, and H. Zhang,Small8, 63 (2012).

10M. Bernardi, M. Palummo, and J. C. Grossman,Nano Lett13, 3664 (2013).

11X. Zhou, L.-J. Wan, and Y.-G. Guo,Chem. Commun.49, 1838 (2013).

12Y. Liang, R. Feng, S. Yang, H. Ma, J. Liang, and J. Chen,Adv. Mater.23, 640 (2011).

13H. Li, J. Wu, Z. Yin, and H. Zhang,Acc. Chem. Res.47, 1067 (2014).

14K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva, and A. A. Firsov,Science 306, 666 (2004).

15K. Zhou, S. Jiang, C. Bao, L. Song, B. Wang, G. Tang, Y. Hu, and Z. Gui,RSC Adv.2, 11695 (2012).

16Z. Zeng, Z. Yin, X. Huang, H. Li, Q. He, G. Lu, F. Boey, and H. Zhang,Angew. Chem. Int Ed.50, 11093 (2011).

17Y.-F. Dong, D.-W. He, Y.-S. Wang, H.-T. Xu, and Z. Gong, Acta Phys. Sinica 65, 128101 (2016).

18N. Sharma, D. Oh, H. Abernathy, M. Liu, P. N. First, and T. M. Orlando,Surf. Sci.604, 84 (2010).

19S. Wu, Z. Zeng, Q. He, Z. Wang, S. J. Wang, Y. Du, Z. Yin, X. Sun, W. Chen, and H. Zhang,Small8, 2264 (2012).

20T. Lin, B. Kang, M. Jeon, C. Huffman, J. Jeon, S. Lee, W. Han, J. Lee, S. Lee, G. Yeom, and K. Kim,ACS Appl. Mater.

Interfaces7, 15892 (2015).

21S. W. Kim, J. H. Na, W. L. Choi, H.-J. Chung, and S. H. Jhang,J. Korean Phys. Soc.69, 1497 (2016).

22A. Castellanos-Gomez, M. Barkelid, A. M. Goossens, V. E. Calado, H. S. van der Zant, and G. A. Steele,Nano Lett12, 3187 (2012).

23N.-Q. Deng, H. Tian, Z.-Y. Ju, H.-M. Zhao, C. Li, M. A. Mohammad, L.-Q. Tao, Y. Pang, X.-F. Wang, T.-Y. Zhang, Y. Yang, and T.-L. Ren,Carbon109, 173 (2016).

24G. H. Han, S. J. Chae, E. S. Kim, F. Guenes, I. H. Lee, S. W. Lee, S. Y. Lee, S. C. Lim, H. K. Jeong, M. S. Jeong, and Y. H. Lee,ACS Nano5, 263 (2011).

25Z. Lin, X. Ye, J. Han, Q. Chen, P. Fan, H. Zhang, D. Xie, H. Zhu, and M. Zhong,Sci. Rep.5, 11662 (2015).

26Z. Yin, H. Li, H. Li, L. Jiang, Y. Shi, Y. Sun, G. Lu, Q. Zhang, X. Chen, and H. Zhang,ACS Nano6, 74 (2012).

27C. Lee, H. Yan, L. E. Brus, T. F. Heinz, J. Hone, and S. Ryu,ACS Nano4, 2695 (2010).

28Y. Pan, M. Yang, Y. Li, Z. Wang, C. Zhang, Y. Zhao, J. Yao, Q. Wu, and J. Xu,Sci. Rep.6, 19571 (2016).

29B. C. Tran Khac, K. J. Jeon, S. T. Choi, Y. S. Kim, F. W. DelRio, and K. H. Chung,ACS Appl. Mater. Interfaces8, 2974 (2016).

30Y. Y. Wang, Z. H. Ni, Z. X. Shen, H. M. Wang, and Y. H. Wu,Appl. Phys. Lett.92, 043121 (2008).

31L. Liu, S. Ryu, M. R. Tomasik, E. Stolyarova, N. Jung, M. S. Hybertsen, M. L. Steigerwald, L. E. Brus, and G. W. Flynn, Nano Lett8, 1965 (2008).

32H. Zhou, F. Yu, Y. Liu, X. Zou, C. Cong, C. Qiu, T. Yu, Z. Yan, X. Shen, L. Sun, B. I. Yakobson, and J. M. Tour,Nano Res.

6, 703 (2013).

33X. Liu, J. D. Wood, K.-S. Chen, E. Cho, and M. C. Hersam,J. Phys. Chem. Lett.6, 773 (2015).

References

Related documents

The EU exports of waste abroad have negative environmental and public health consequences in the countries of destination, while resources for the circular economy.. domestically

46 Konkreta exempel skulle kunna vara främjandeinsatser för affärsänglar/affärsängelnätverk, skapa arenor där aktörer från utbuds- och efterfrågesidan kan mötas eller

Exakt hur dessa verksamheter har uppstått studeras inte i detalj, men nyetableringar kan exempelvis vara ett resultat av avknoppningar från större företag inklusive

För att uppskatta den totala effekten av reformerna måste dock hänsyn tas till såväl samt- liga priseffekter som sammansättningseffekter, till följd av ökad försäljningsandel

The increasing availability of data and attention to services has increased the understanding of the contribution of services to innovation and productivity in

Av tabellen framgår att det behövs utförlig information om de projekt som genomförs vid instituten. Då Tillväxtanalys ska föreslå en metod som kan visa hur institutens verksamhet

Generella styrmedel kan ha varit mindre verksamma än man har trott De generella styrmedlen, till skillnad från de specifika styrmedlen, har kommit att användas i större

Närmare 90 procent av de statliga medlen (intäkter och utgifter) för näringslivets klimatomställning går till generella styrmedel, det vill säga styrmedel som påverkar