Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Excitation mechanism of A1g mode and origin of nonlinear temperature dependence of Raman shift of CVD-grown mono- and few-layer MoS2 films

Open Access Open Access

Abstract

MoS2 films are grown on SiO2/Si substrates by chemical vapor deposition. The vibrational properties of optical phonons of mono-, bi- and multilayer MoS2 are studied by Raman scattering spectroscopy over temperature range from 90 to 540 K with 514.5 nm and 785 nm lasers. The Raman peaks of E2g1 and A1g modes are observed simultaneously for mono-, bi- and multilayer MoS2 with 514.5 nm laser, but only the Raman peak of E2g1 mode is seen for monolayer MoS2 as 785 nm laser is used, revealing electron-phonon exchange excitation mechanism of A1g mode for the first time. The Raman shifts of E2g1 and A1g modes present obvious nonlinear temperature dependence. A semi-quantitative model is used to fit the nonlinear temperature dependence of Raman shifts which matches well to experimental data. Meanwhile, the fitting results reveal that the nonlinear temperature dependence of Raman shifts of E2g1 mode mainly originates from three-phonon anharmonic effect, while one of A1g mode is contributed by stronger three- and weaker four-phonon anharmonic effects cooperatively but two contributions are comparable in intensity.

© 2016 Optical Society of America

1. Introduction

Two-dimensional transition-metal dichalcogenide (TMDC) materials, such as molybdenum disulfide (MoS2), are promising for the next generation of electronic and optoelectronic devices [1] because of their unique electrical, optical, and mechanical properties [2–9]. In contrast to graphene, MoS2 atomic layers have a band gap which transits from indirect 1.2 eV to direct 1.9 eV as the thickness of MoS2 reduces from bulk to monolayer [3,5]. Those unique properties of MoS2 make it very suitable for fabrication of electronic and optoelectronic devices. Single-layer MoS2 field effect transistors [10–12], heterojunction [13] and ultrasensitive photodetectors [14], have been reported. Field emission and photoresponse of MoS2 thin films are also detailed investigated [15].Even the integrated circuits based on bilayer MoS2 transistors have also been demonstrated [16]. It may be predicted that the degree of integration of MoS2 transistor-based integrated circuits will be much higher than one of current Si-based microelectronic integrated circuits because MoS2 transistors are in nanometer scale. Heat dissipation or thermal conductivity has been one of the most significant constraints on design and fabrication of current Si-based integrated electronic circuits. Consequently, it may be predicted that heat dissipation or thermal conductivity will be critical in near future in design and fabrication of single layer MoS2 integrated circuits. For thermal manipulation, it is very necessary to study electron-phonon interactions, dynamics and thermodynamics of phonons in MoS2 thin films.

Raman spectroscopy is a powerful tool to access thermal properties of materials, especially nanomaterials, and has been extensively used to investigate thermal properties and microstructures of TMDC films and other materials [17–23], revaling their doping effect [24], electronic and structural properties [25], and the vibration property of nanoribbons and nanosheets [26–28]. Several comprehensive studies on the vibration frequency and peak width evolution of phonon modes of MoS2 mono- and few-layers with temperature have also been reported so that thermal conductivity and phonon dynamics of MoS2 could be revealed [29–35]. However, some controversial experimental results were reported [29–35]. Both Taube et al. [29] and Yan et al. [30] investigated the temperature dependence of Raman shifts of E2g1 and A1g modes of monolayer MoS2 within a low temperature range of 80 – 350 K, respectively on SiO2/Si and sapphire (suspended) substrates. Although the MoS2 monolayer samples studied by the two groups were all fabricated from mechanical exfoliation of MoS2 single crystals, Taube et al. [29] found nonlinear temperature dependence of Raman shifts of both E2g1 and A1g modes, while Yan et al. [30] observed linear temperature dependence for either a sapphire substrate or the suspended. Inversely, Lanzillo et al. [31] and Najmaei et al. [32] studied the temperature dependence of Raman shifts of E2g1 and A1g modes of monolayer MoS2 on SiO2/Si substrates within a high temperature range of 300 – 500 K, respectively prepared by chemical vapor deposition (CVD) and mechanical exfoliation. Both groups observed linear temperature dependence of Raman shifts of both E2g1 and A1g modes. Lanzillo et al. [31] obtained the first-order temperature coefficient of E2g1 mode as −0.013 cm−1/K, but Najmaei et al. [32] gave out one as −0.0179 cm−1/K. An obvious difference on the first-order temperature coefficient of E2g1 mode existed between the two reports. A more disputed conclusion given by Najmaei et al. [32] was that linear temperature dependence of Raman shifts of E2g1 mode was dominated by anharmonic four-phonon processes rather than three-phonon processes. It was unexpected physically because four-phonon processes were higher-order than three-phonon processes. Sahoo et al. [33] and Thripuranthaka et al. [34] studied the temperature dependence of Raman shifts of E2g1 and A1g modes of CVD-grown few-layer MoS2, respectively on SiO2/Si substrates and without substrates (suspended) in a widely variable temperature range of 80 – 600 K. Both of them claimed Raman shifts of E2g1 and A1g modes varied linearly with temperature. Two slightly different first-order temperature coefficients of E2g1 mode, −0.0132 cm−1/K [33] and −0.016 cm−1/K [34], were given out. Su et al. [35] studied temperature-dependent Raman shifts of mono- and bilayer MoS2 on sapphire and Si substrates in the range of above room temperature, and found that the temperature dependence of Raman shifts of E2g1 and A1g modes was complex, including linear and nonlinear, depending on not only the number of layer of MoS2 films, but also substrate type. Thus more experimental studies are necessary very much to extract the accurate temperature coefficient of Raman shift of Raman active modes because the temperature coefficient directly reflects the strength change of Raman vibration bond with varying temperature. It is also an important parameter to differentiate layer number of layered films.

To distinguish the disputes mentioned above, it is very necessary to further study the temperature dependence of Raman shifts of mono- and bilayer MoS2, especially the dependence of monolayer MoS2 directly CVD-grown on a Si substrate over a wide temperature range over 80 – 600 K, because CVD growth can fabricate large-area monolayer MoS2 and Si substrates have excellent compatibility with Si-based microelectronics. Consequently, monolayer MoS2 directly CVD-grown on a Si substrate has widely potential applications. In addition, it was already reported Raman shift strongly depended on substrate type and bonding strength between films and substrates [35,36]. To our knowledge, no such data are available yet for mono- and bilayer MoS2 directly grown on SiO2/Si substrate over a wide temperature range of the below and above room temperature. In addition, physical mechanism of nonlinear temperature dependence of Raman shifts was absent and excitation mechanism of E2g1 and A1g modes of monolayer MoS2 was also unknown.

In this article, we present the temperature dependence of Raman shifts of E2g1 and A1g modes of mono-, bi- and multilayer MoS2 directly CVD-grown on SiO2/Si substrates over a wide temperature range of 90-540 K. We observe nonlinear temperature dependence of Raman shifts for both E2g1 and A1g modes. It is different from the reported temperature dependence of Raman shifts of CVD-grown MoS2 transferred onto SiO2/Si substrate [35]. The nonlinear temperature dependence of Raman shifts is analyzed with a semi-quantitative physical model [32,35,37], which includes thermal expansion and pure temperature effects. The analyzed results show that the nonlinear temperature dependence of Raman shifts mainly originates from anharmonic effects of three- and four-phonon processes. For E2g1 mode, the contribution of three-phonon process is much stronger than four-phonon process. For A1g mode, the contributions of three- and four-phonon processes are comparable but the former is still stronger than the latter. Furthermore, we perform a temperature-dependent non-resonant Raman scattering experiment on monolayer MoS2 with 785 nm laser whose photon energy is lower than the band gap (1.8 eV) of monolayer MoS2, and only observe the Raman peak of E2g1 mode. Thus we deduce A1g mode is active only via electron-phonon interaction. These findings cannot only enrich the knowledge on vibrational properties of optical phonons of MoS2 films but also may be useful in thermal manipulation of MoS2 integrated circuits in the future.

2. Experiment results with 514.5nm laser

The optical micrograph of CVD-grown MoS2 films on SiO2/Si substrate is shown in Fig. 1(a).

 figure: Fig. 1

Fig. 1 (a) Optical micrograph of CVD-grown MoS2 on SiO2/Si substrate. (b) Raman spectra of mono-, bi- and multilayer MoS2. (c) Temperature-dependent Raman spectra of monolayer MoS2 under the excitation of 1.6mW laser at 514.5 nm wavelength.

Download Full Size | PDF

Monolayer MoS2 is crystalline well and in triangular shape. Single crystal domain size is large up to 50 μm. The temperature-dependent Raman spectra are collected using a Renishaw InVia micro-Raman spectrometer with laser line of 514.5 nm and 785nm. It is a confocal micro-Raman spectroscope system with a 2400lp/mm grating. Spectral resolution of the system is up to 0.6 cm−1. The maximum power of 514.5 nm and 785nm lasers used in our experiments are 1.6 mW and 7.5mW respectively. Laser spot is focused down to 1 μm which is much smaller than MoS2 sample size.The number of MoS2 layers can be roughly estimated by observing the color and shape of homogenous domain. Mono-, bi- and multilayer MoS2 are indicated in Fig. 1(a). Raman spectrum is used quantitatively to differentiate the number of MoS2 film layers. Figure 1(b) shows the Raman spectra of three regions in Fig. 1(a) labeled by monolayer, bilayer and multilayer. Raman spectra are collected first using 514.5 nm laser. The wave-number difference of 20 cm−1 between Raman peaks of E2g1 and A1g modes, calculated from the bottom Raman spectrum in Fig. 1(b), is obtained, agrees well with the value of 20.3 cm−1 measured by Zhan et al. on CVD-grown monolayer MoS2 [38]. Similarly, a 22 cm−1 difference calculated from the middle Raman spectrum in Fig. 1(b) is acquired, which is also in accordance with 22.3 cm−1 difference of CVD-grown bilayer MoS2 reported by Zhan et al. [38].

The MoS2 sample is mounted in a cryostat cooled by liquid nitrogen. Temperature-dependent Raman spectra are taken over a wide temperature range of 90 – 540 K for mono-, bi- and multilayer MoS2. Typical Raman spectra of MoS2 monolayer are plotted in Fig. 1(c) for different lattice temperature over 90 - 540 K.

3. Analyses of experimental results

3.1 First- and second-order temperature coefficients extracted by linear and quadratic polynomial fittings

The positions of Raman peaks of E2g1 and A1g modes are extracted by Lorentzian function fitting to all Raman spectra recorded at different temperatures, as shown in Fig. 1(c). The peak positions of E2g1 and A1g modes versus temperature are plotted in Fig. 2(a) by triangle, circle and square points, respectively for mono-, bi- and multilayer MoS2 thin films. Error bars are not plotted out in Fig. 2(a) because they are smaller than 0.1 cm−1 and too small to be displayed. It is very apparent that the Raman shifts of E2g1 and A1g modes decrease nonlinearly with increasing temperature for all of mono-, bi- and multilayer MoS2. However, they appear linear if one only looks the graph in the range of above room temperature. Therefore, the linear dependence reported in [31,35] is only an approximate of our results here in the range of above room temperature.

 figure: Fig. 2

Fig. 2 (a) The temperature dependence of Raman shifts of E2g1 and A1g modes of mono-, bi- and multilayer MoS2 under the excitation of 1.6mW laser at 514.5 nm wavelength. (b) The temperature dependence of Raman shifts of E2g1 and A1g modes of multilayer MoS2 for three different laser powers at 514.5 nm wavelength.

Download Full Size | PDF

The following quadratic polynomial

Ω(T)=ω0+χ1T+χ2T2
is used to best fit the nonlinear temperature-dependent Raman shift data, where ω0 is Raman shift at zero Kelvin . The best fittings are also plotted by solid lines in Fig. 2(a). Obviously, the solid lines fit the experimental points very well. The best fittings give out the first-order and second-order temperature coefficients, χ1 (cm−1/K) and χ2 (cm−1/K2), which are listed in Table 1 for the E2g1 and A1g modes of mono-, bi- and multilayer MoS2.

Tables Icon

Table 1. First- and second-order temperature coefficients extracted by linear and quadratic polynomial fittings.

However, if we ignore the nonlinearity of temperature-dependent Raman shifts, as done in some literatures [30,31,33,34], the following linear formula,

Ω(T)=ω0+χ1T
can be used to best fit the temperature-dependent Raman shifts in Fig. 2(a), which gives out linear temperature coefficients χ1 (cm−1/K) of E2g1 and A1g modes for mono-, bi- and multilayer MoS2. All values of χ1 are also listed in Table 1. One can find the first-order temperature coefficients (χ1) given by the linear and quadratic polynomial fittings are very different for either monolayer and bilayer or multilayer MoS2. The values of χ1 obtained by linear fitting agree well with ones reported in the literature [30–34], but are significantly larger than ones given by quadratic fitting. The fact that the first-order temperature coefficient χ1 given by linear fiting agrees well with the reported values shows the reliability of our Raman experimental results, whereas obvious difference in χ1 given by linear and nonlinear fittings in turn implies the true existence of significant nonlinearity in the temperature dependence. Therefore, nonlinear fitting results are more justified. The first- and second-order temperature coefficients listed in Table 1 are first experimental results given by quadratic fitting to wide temperature range Raman shifts. Su et al. [35] reported the experimental results of χ1 and χ2 given by cubic fitting to temperature-dependent Raman shifts in above room temperature. For E2g1 mode of monolayer MoS2, our results (χ1 = −0.00369, χ2 = −1.32 × 10−5) are comparable to the reported results (χ2 = −0.0143, χ2 = −1.44 × 10−5, χ3 = 7.71 × 10−9 which may be transformed to χ1 = −0.00366, χ2 = −2.129 × 10−5 by shifting origin of temperature coordinate from 298 K to 0 K due to different temperature origin used between this work and [35]) of monolayer MoS2 CVD-grown on a sapphire substrate, but not to the reported results of CVD-grown monolayer MoS2 transferred onto a SiO2/Si substrate. Maybe the bonding between MoS2 and SiO2/Si substrates weakened in transferred films. The obviously strange temperature dependence of A1g mode reported in [35] for all transferred monolayer MoS2 films may be the evidence because out of plane mode is more sensitive to coupling change between films and substrates [36]. As shown in Table 1, in our results the extracted all temperature coefficients of E2g1 and A1g modes are consistent well. Therefore, we believe our results are more reliable because one reason is good self-consistency and the other one is our results are extracted from a wider range temperature-dependent Raman shifts over below to above room temperature.

Because the band gap and absorption coefficient of MoS2 are also temperature-dependent, laser heating effect may be different at different lattice temperature. In other words, the temperature dependence of the Raman shifts observed may contain the contribution from different heating effects of the same excitation laser power. In order to test the effect of possibly different laser heating on Raman shifts, we have performed the measurement of temperature-dependent Raman spectra under another two lower excitation powers of 0.16 and 0.6 mW. The temperature dependence of Raman shifts of E2g1 and A1g modes of multilayer MoS2 is plotted in Fig. 2(b) simultaneously for three excitation powers of 1.6, 0.6 and 0.16 mW. One can see almost the same Raman shifts at each lattice temperature with three different laser powers. That implies the temperature-dependent laser heating effect may be negligible in the range of our low laser power.

3.2 Physical origin of nonlinear temperature dependence

In order to understand the physical origin of the nonlinear temperature dependence of Raman shifts, a physical model including thermal volume expansion and pure temperature effects is used to analyze temperature-dependent Raman shift data [32,35,37]. Under the condition of constant atmospheric pressure, the model can be written as

ω(T)=ω0'+ΔωE(T)+ΔωA(T),
where ΔωE and ΔωA are Raman shift change induced by lattice thermal expansion and pure temperature effects, respectively.

Volume expansion-induced contribution to the change of Raman shift can be described by Gruneisen constant model [32,35]

ΔωE(T)=ω0exp(nγ0TαdT)ω0,
where γis the Gruneisen parameter, ω0 is the extrapolated Raman shift at 0 K, and αis the thermal expansion coefficient of the material. n is the degeneracy, 1 for A1g mode and 2 for E2g1 mode.

Gruneisen parameter γ slightly depend on the layer number of material. However, here we take γ of bulk MoS2 as an approximation of γ of mono- and few-layer MoS2, as done in [32], because γ itself is weakly layer number-dependent and its accurate thickness-dependent relation is unavailable so far. The reported in-plane γ(E2g1) = 0.21 and out-of-plane γ(A1g) = 0.42 in bulk MoS2 are used later in our fitting calculation [39]. The thermal expansion coefficient α of E2g1 and A1g modes were given out by El-Mahalawy and Evans as [40]

αa=(0.6007×105+0.6958×107Ta)(1c°),
and
αc=(0.1064×103+1.5475×107Tc)(1c°),
respectively, where T is temperature in °C, a and c are lattice constants of MoS2 in angstroms.

Contribution of pure temperature effects to the change of Raman shift is mainly from anharmonic effects of three- and four-phonon processes, and was simply modeled by Kelmens [41], written as [32,37,41]

ΔωA(T)=A[1+2ex1]+B[1+3ey1+3(ey1)2]=ΔωA3p(T)+ΔωA4p(T),
wherex=ω/2kT,y=ω/3kT. The first term describes the contribution (ΔωA-3p) of three-phonon process, while the second term denotes one (ΔωA-4p) of four-phonon process. A and B are strength coefficients to describe the contributions of three- and four-phonon processes, respectively, and can be obtained by fitting experimental data of temperature-dependent Raman shifts with Eq. (3).

The best fittings to nonlinear temperature dependence of Raman shifts of E2g1 and A1g modes of monolayer MoS2 with Eq. (3) are plotted in Fig. 3(a) by black solid lines. One can see the model of Eq. (3) fits temperature-dependent nonlinear Raman shifts (open squares) quite well. Meanwhile, component contributions of thermal expansion, three- and four-phonon processes are also potted in Fig. 3(a) by color dashed lines. For E2g1 mode, one can discern nonlinear temperature dependence of Raman shift mainly originates from three-phonon process (red dashed line), while thermal expansion (green dashed line) and four-phonon process (blue dashed line) contribute very weakly. However, for A1g mode, one can find that the contributions from three- and four-phonon processes are comparable though three-phonon contribution (ΔωA-3p) is slightly stronger than four-phonon one (ΔωA-4p). For either E2g1 or A1g mode, thermal expansion mainly contributes a nearly linear weak temperature-dependent component. As a result, nonlinear temperature dependence of Raman shifts is mainly from anharmonic effect of three-phonon process plus a weak four-phonon process. A quantitative comparison of three- and four-phonon contributions to Raman shifts is plotted in Fig. 3(b) by the ratio ((ΔωA-3pA)/ (ΔωA-4p-B)) of three to four phonon contributions. It is obvious that three-phonon contribution is at least thirty five times stronger than the contribution of four phonon process for E2g1 mode, whereas the contribution of three phonon process is only a few times stronger than one of four phonon process for A1g mode. It is worth emphasizing that such ratios of three- to four-phonon contributions are reasonable physically because four-phonon process is a higher-order process than three-phonon process. Contrarily, the conclusion obtained in [32], the four-phonon process was dominant in temperature dependence of Raman shift, was disputed and inappropriate physically. One can also notice from Fig. 3(b) that the ratio of three- to four-phonon contributions first increases with lattice temperature below ~140 K, peaks at ~140 K, and then decreases beyond ~140 K, which imply that three-phonon contribution enhances faster than four-phonon contribution below 140 K, while above 140 K the enhancement trend of three- and four-phonon processes reverse.

 figure: Fig. 3

Fig. 3 (a) The best fitting to experimental Raman shift data (open squares) of monolayer MoS2 with Eq. (3) described in text. Component contributions from thermal expansion, three- and four-phonon processes are plotted by color dashed lines. (b) The ratio of three- to four-phonon process contributions of E2g1 and A1g modes.

Download Full Size | PDF

In similar way, we also fit the nonlinear temperature dependence of Raman shifts of E2g1 and A1g modes of bi- and multilayer MoS2, as shown in Fig. 2(a). The strength coefficients, A and B, of three- and four-phonon processes are shown in Fig. 4. For E2g1 mode, one can still find three-phonon contribution (A) is much stronger than four-phonon one (B) for both bi- and multilayer MoS2, but for A1g mode A and B are comparable although B is smaller than A. Therefore, three-phonon process is the main contribution to nonlinear temperature dependence of Raman shifts of E2g1 and A1g modes for either mono- and bilayer or multilayer MoS2.

 figure: Fig. 4

Fig. 4 Strength coefficients of three- and four-phonon processes versus layer number of MoS2 films in temperature-dependent Raman shifts of E2g1 and A1g modes.

Download Full Size | PDF

4. Experiment results with 785nm laser and analysis

To understand excitation mechanism of E2g1 and A1g modes, we re-measure temperature- dependent Raman spectra of monolayer MoS2 with 785 nm laser instead of 514.5nm laser. In this experiment, hot phonon excitation via electron-phonon exchange interaction is avoided because the photon energy (1.58 eV) of 785 nm laser is lower than the band gap (~1.82 eV at room temperature) of monolayer MoS2. The measured temperature-dependent Raman spectra are plotted in Fig. 5(a). One can find that only the Raman peak of E2g1 mode occurs, whereas A1g mode disappears. It seems to imply that A1g phonon mode can only be excited by electron-phonon exchange interplay. To further test the electron-phonon exchange excitation of A1g mode, we measure Raman spectrum of multilayer MoS2 with 785 nm laser, and again observe the Raman peaks of E2g1and A1g modes simultaneously. This provides direct evidence to support the electron-phonon exchange excitation mechanism of A1g mode because the band gap of multilayer MoS2 is about 1.2 eV which is smaller than the photon energy (1.58 eV) of 785 nm laser and hence electron excitation takes place via interband transition.

 figure: Fig. 5

Fig. 5 (a)Temperature-dependent Raman spectra of monolayer MoS2 measured with 785nm laser. (b) Nonlinear temperature dependence of Raman shifts of E2g1 mode measured with 514.5 nm (filled circles) and 785 nm (open squares) lasers. Black lines are the best fittings with the model of Eq. (3) in text.

Download Full Size | PDF

On the other hand, one can find in Fig. 5(b) that the Raman shifts measured with 785 nm laser are larger than those measured with 514.5 nm laser for E2g1 mode although the shift data measured with 785 nm laser are noisy a little because nonresonant Raman scattering by 785 nm laser is much weaker than resonant Raman scattering with 514.5 nm laser. The difference in Raman shift decreases with increasing lattice temperature of MoS2, and approaches to zero above 500 K. This phenomenon can be explained by laser heating to MoS2 via electron-phonon excitation. When Raman spectra are measured with 514.5 nm laser, a great amount of hot optical phonons are excited continuously via electron-phonon exchange due to interband transition of electrons so that the equilibrium temperature of optical phonon system is actually higher than the lattice temperature. Generally speaking, the lifetime of hot optical phonons (decayed into acoustic phonons) decreases with the increase of lattice temperature [33], and hence heat dissipation of optical phonon system becomes fast with the rise of lattice temperature so that the temperature difference between optical phonon system and lattices becomes smaller at higher lattice temperature, just as shown in Fig. 5(b). Actually, such double-wavelength Raman spectra can be used to detect the temperature of optical phonon system or the temperature difference between optical phonon system and lattices. For example, the Raman shifts at O and P points are measured at the same lattice temperature of 200 K but different laser wavelength, respectively at 785 nm and 514.5 nm. The Raman shift at P is smaller than that at O, which means the temperature of optical phonon system at P is actually higher than the lattice temperature of 200 K. The real temperature of optical phonon system can be read out from point N as ~311 K, which is the equilibrium temperature of optical phonon system that is 110K higher than lattice temperature due to heating of 514.5 nm laser via electron-phonon exchange.

Finally, we analyze the nonlinear temperature dependence of Raman shifts of E2g1 mode measured with 785 nm laser using Eq. (3). One can see the temperature-dependent nonlinear Raman shifts mainly originates from three-phonon process. Four-phonon process is very weak and negligible. This result is similar to one measured with 514.5 nm laser in Fig. 3.

5. Conclusion

In this work, we have studied the temperature dependence of vibrational properties of optical phonon modes of CVD-grown MoS2 mono- and few-layer thin films over a wide temperature range of 90 - 540 K using backscattering micro-Raman spectroscopy under the excitation of 514.5 nm and 785 nm lasers. We have observed nonlinear temperature dependence of Raman shifts of E2g1 and A1g modes for mono-, bi- and multilayer MoS2. A semi-quantitative model including thermal expansion and anharmonic effects is adopted to analyze the nonlinear temperature dependence of Raman shifts. We have found the nonlinear temperature dependence of Raman shifts of E2g1 and A1g modes can be fit well by the model for either mono-, bi- or multilayer MoS2. The fitting results reveal that thermal expansion contribution is very weak and almost linear. The nonlinear temperature dependence of Raman shifts mainly originate from anharmonic contributions of three- and four-phonon processes. The nonlinear temperature dependence of Raman shift of E2g1 mode is dominated by three-phonon process, and the contribution of four-phonon process is very weak and negligible in comparison to one of three-phonon process. This result is much different from one reported in the literature where four-phonon process was considered as main origin of nonlinear temperature dependence of Raman shifts. In contrast, we have found that the nonlinear temperature dependence of Raman shift of A1g mode originates from cooperative three- and four-phonon processes, but three-phonon process contributes still stronger than four-phonon process. For the excitation of 785 nm laser, we have observed only Raman peak of E2g1 mode but not A1g mode over the whole range of 90~540 K. The fact that the Raman peak of A1g mode is absent with 785 nm laser reveals electron-phonon exchange excitation mechanism of A1g optical phonon mode for the first time. We have also found the temperature of optical phonon system is actually higher than lattice temperature. The temperature difference between them can be measured by double-wavelength Raman spectra, where photon energy of one wavelength laser is higher than band gap of MoS2, and photon energy of the other wavelength laser is lower than band gap.

Acknowledgments

This work was supported by National Basic Research Program of China under grant no. 2013CB922403, National Natural Science Foundation of China under grant nos. 11274399 and 61475195 as well as Guangdong Natural Science Foundation, China under grant no. 2014A030311029.

References and links

1. Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, and M. S. Strano, “Electronics and Optoelectronics of Two-Dimensional Transition Metal Dichalcogenides,” Nat. Nanotechnol. 7(11), 699–712 (2012). [CrossRef]   [PubMed]  

2. C. Lee, H. Yan, L. E. Brus, T. F. Heinz, J. Hone, and S. Ryu, “Anomalous Lattice Vibrations of Single- and Few-layer MoS2.,” ACS Nano 4(5), 2695–2700 (2010). [CrossRef]   [PubMed]  

3. K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, “Atomically thin MoS₂: a new direct-gap semiconductor,” Phys. Rev. Lett. 105(13), 474–479 (2010). [CrossRef]   [PubMed]  

4. G. Kioseoglou, A. T. Hanbicki, M. Currie, A. L. Friedman, D. Gunlycke, and B. T. Jonker, “Valley Polarization and Intervalley Scattering in Monolayer MoS2,” Appl. Phys. Lett. 101(22), 221907 (2012). [CrossRef]  

5. T. Cheiwchanchamnangij and W. R. L. Lambrecht, “Quasiparticle Band Structure Calculation of Monolayer, Bilayer, and Bulk MoS2,” Phys. Rev. B 85(20), 205302 (2012). [CrossRef]  

6. Z. Huang, X. Peng, H. Yang, C. He, L. Xue, G. Hao, C. Zhang, W. Liu, X. Qi, and J. Zhong, “The Structural, Electronic and Magnetic Properties of Bi-layered MoS2 with Transition-Metals Doped in The Interlayer,” RSC Advances 3(31), 12939–12944 (2013). [CrossRef]  

7. T. Cao, G. Wang, W. Han, H. Ye, C. Zhu, J. Shi, Q. Niu, P. Tan, E. Wang, B. Liu, and J. Feng, “Valley-Selective Circular Dichroism of Monolayer Molybdenum Disulphide,” Nat. Commun. 3, 177–180 (2012). [CrossRef]   [PubMed]  

8. Y. H. Ho, Y. H. Wang, and H. Y. Chen, “Magnetoelectronic and Optical Properties of A MoS2 Monolayer,” Phys. Rev. B 89(15), 1361–1377 (2014). [CrossRef]  

9. N. Scheuschner, O. Ochedowski, A. M. Kaulitz, R. Gillen, M. Schleberger, and J. Maultzsch, “Photoluminescence of Freestanding Single- and Few-Layer MoS2,” Phys. Rev. B 89(12), 125406 (2014). [CrossRef]  

10. B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, and A. Kis, “Single-Layer MoS2 Transistors,” Nat. Nanotechnol. 6(3), 147–150 (2011). [CrossRef]   [PubMed]  

11. D. J. Late, B. Liu, H. S. Matte, V. P. Dravid, and C. N. R. Rao, “Hysteresis in Single-Layer MoS2 Field Effect Transistors,” ACS Nano 6(6), 5635–5641 (2012). [CrossRef]   [PubMed]  

12. D. J. Late, Y. K. Huang, B. Liu, J. Acharya, S. N. Shirodkar, J. Luo, A. Yan, D. Charles, U. V. Waghmare, V. P. Dravid, and C. N. R. Rao, “Sensing behavior of atomically thin-layered MoS2 transistors,” ACS Nano 7(6), 4879–4891 (2013). [CrossRef]   [PubMed]  

13. H. Li, J. Shao, D. Yao, and G. Yang, “Gate-voltage-controlled spin and valley polarization transport in a normal/ferromagnetic/normal MoS₂ junction,” ACS Appl. Mater. Interfaces 6(3), 1759–1764 (2014). [CrossRef]   [PubMed]  

14. O. Lopez-Sanchez, D. Lembke, M. Kayci, A. Radenovic, and A. Kis, “Ultrasensitive Photodetectors Based on Monolayer MoS2.,” Nat. Nanotechnol. 8(7), 497–501 (2013). [CrossRef]   [PubMed]  

15. D. J. Late, P. A. Shaikh, R. Khare, R. V. Kashid, M. Chaudhary, M. A. More, and S. B. Ogale, “Pulsed laser-deposited MoS₂ thin films on W and Si: field emission and photoresponse studies,” ACS Appl. Mater. Interfaces 6(18), 15881–15888 (2014). [CrossRef]   [PubMed]  

16. H. Wang, L. Yu, Y. H. Lee, Y. Shi, A. Hsu, M. L. Chin, L. J. Li, M. Dubey, J. Kong, and T. Palacios, “Integrated circuits based on bilayer MoS₂ transistors,” Nano Lett. 12(9), 4674–4680 (2012). [CrossRef]   [PubMed]  

17. B. C. Windom, W. G. Sawyer, and D. W. Hahn, “A Raman Spectroscopic Study of MoS2 and MoO3: Applications to Tribological Systems,” Tribol. Lett. 42(3), 301–310 (2011). [CrossRef]  

18. M. Virsek, M. Krause, A. Kolitsch, and M. Remškar, “Raman Characterization of MoS2 Microtube,” Phys. Status Solidi 246(11–12), 2782–2785 (2009). [CrossRef]  

19. Y. Wang, C. Cong, C. Qiu, and T. Yu, “Raman Spectroscopy Study of Lattice Vibration and Crystallographic Orientation of Monolayer MoS2 under Uniaxial Strain,” Small 9(17), 2857–2861 (2013). [CrossRef]   [PubMed]  

20. X. Zhang, W. P. Han, J. B. Wu, S. Milana, Y. Lu, Q. Q. Li, A. C. Ferrari, and P. H. Tan, “Raman Spectroscopy of Shear and Layer Breathing Modes in Multilayer MoS2,” Phys. Rev. B 87(11), 115413 (2013). [CrossRef]  

21. S. V. Bhatt, M. P. Deshpande, V. Sathe, R. Rao, and S. H. Chaki, “Raman Spectroscopic Investigations on Transition-Metal Dichalcogenides MX2 (M= Mo, W; X = S, Se) at High Pressures and Low Temperature,” J. Raman Spectrosc. 45(10), 971–979 (2014). [CrossRef]  

22. M. Praveena, C. D. Bain, V. Jayaram, and S. K. Biswas, “Total Internal Reflection (tir) Raman Tribometer: A New Tool for In Situ Study of Friction-Induced Material Transfer,” RSC Advances 3(16), 5401–5411 (2013). [CrossRef]  

23. K. Gołasa, M. Grzeszczyk, R. Bożek, P. Leszczyński, A. Wysmołek, M. Potemski, and A. Babiński, “Resonant Raman scattering in MoS2-from bulk to monolayer,” Solid State Commun. 197, 53–56 (2014). [CrossRef]  

24. D. J. Late, U. Maitra, L. S. Panchakarla, U. V. Waghmare, and C. N. R. Rao, “Temperature effects on the Raman spectra of graphenes: dependence on the number of layers and doping,” J. Phys. Condens. Matter 23(5), 055303 (2011). [CrossRef]   [PubMed]  

25. D. J. Late, S. N. Shirodkar, U. V. Waghmare, V. P. Dravid, and C. N. R. Rao, “Thermal expansion, anharmonicity and temperature-dependent Raman spectra of single- and few-layer MoSe₂ and WSe₂,” ChemPhysChem 15(8), 1592–1598 (2014). [CrossRef]   [PubMed]  

26. A. S. Pawbake, J. O. Island, E. Flores, J. R. Ares, C. Sanchez, I. J. Ferrer, S. R. Jadkar, H. S. J. van der Zant, A. Castellanos-Gomez, and D. J. Late, “Temperature-dependent Raman spectroscopy of titanium trisulfide (TiS3) nanoribbons and nanosheets,” ACS Appl. Mater. Interfaces 7(43), 24185–24190 (2015). [CrossRef]   [PubMed]  

27. T. M and D. J. Late, “Temperature dependent phonon shifts in single-layer WS2,” ACS Appl. Mater. Interfaces 6(2), 1158–1163 (2014). [CrossRef]   [PubMed]  

28. D. J. Late, “Temperature dependent phonon shifts in few-layer black phosphorus,” ACS Appl. Mater. Interfaces 7(10), 5857–5862 (2015). [CrossRef]   [PubMed]  

29. A. Taube, J. Judek, C. Jastrzębski, A. Duzynska, K. Świtkowski, and M. Zdrojek, “Temperature-dependent nonlinear phonon shifts in a supported MoS2 monolayer,” ACS Appl. Mater. Interfaces 6(12), 8959–8963 (2014). [CrossRef]   [PubMed]  

30. R. Yan, J. R. Simpson, S. Bertolazzi, J. Brivio, M. Watson, X. Wu, A. Kis, T. Luo, A. R. Hight Walker, and H. G. Xing, “Thermal conductivity of monolayer molybdenum disulfide obtained from temperature-dependent raman spectroscopy,” ACS Nano 8(1), 986–993 (2014). [CrossRef]   [PubMed]  

31. N. A. Lanzillo, A. Glen Birdwell, M. Amani, F. J. Crowne, P. B. Shah, S. Najmaei, Z. Liu, P. M. Ajayan, J. Lou, M. Dubey, S. K. Nayak, and T. P. O’Regan, “Temperature-dependent phonon shifts in monolayer MoS2,” Appl. Phys. Lett. 103(9), 093102 (2013).

32. S. Najmaei, P. M. Ajayan, and J. Lou, “Quantitative analysis of the temperature dependency in Raman active vibrational modes of molybdenum disulfide atomic layers,” Nanoscale 5(20), 9758–9763 (2013). [CrossRef]   [PubMed]  

33. S. Sahoo, A. P. S. Gaur, M. Ahmadi, M. J.-F. Guinel, and R. S. Katiyar, “Temperature-dependent Raman studies and thermal conductivity of few-layer MoS2,” J. Phys. Chem. C 117(17), 9042–9047 (2013). [CrossRef]  

34. M. Thripuranthaka, R. V. Kashid, C. Sekhar Rout, and D. J. Late, “Temperature dependent Raman spectroscopy of chemically derived few layer MoS2 and WS2 nanosheets,” Appl. Phys. Lett. 104(8), 081911 (2014).

35. L. Su, Y. Zhang, Y. Yu, and L. Cao, “Dependence of coupling of quasi 2-D MoS2 with substrates on substrate types, probed by temperature dependent Raman scattering,” Nanoscale 6(9), 4920–4927 (2014). [CrossRef]   [PubMed]  

36. L. Q. Su, Y. F. Yu, L. Y. Cao, and Y. Zhang, “Effects of substrate type and material-substrate bonding on high-temperature behavior of monolayer WS2,” Nano Res. 8(8), 2686–2697 (2015). [CrossRef]  

37. M. Balkanski, R. F. Wallis, and E. Haro, “Anharmonic effects in light scattering due to optical phonons in silicon,” Phys. Rev. B 28(4), 1928–1934 (1983). [CrossRef]  

38. Y. Zhan, Z. Liu, S. Najmaei, P. M. Ajayan, and J. Lou, “Large-area vapor-phase growth and characterization of MoS2 atomic layers on a SiO2 substrate,” Small 8(7), 966–971 (2012). [CrossRef]   [PubMed]  

39. S. Sugai and T. Ueda, “High-pressure Raman spectroscopy in the layered materials 2H-MoS2, 2H -MoSe2, and 2H -MoTe2,” Phys. Rev. B 26(12), 6554–6558 (1982). [CrossRef]  

40. S. H. El-Mahalawy and B. L. Evans, “The thermal expansion of 2H-MoS2,2H-MoSe2 and 2H -WSe2 Between 20 and 800°C,” J. Appl. Cryst. 9(5), 403–406 (1976). [CrossRef]  

41. P. G. Klemens, “Anharmonic decay of optical phonons,” Phys. Rev. 148(2), 845–848 (1966). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 (a) Optical micrograph of CVD-grown MoS2 on SiO2/Si substrate. (b) Raman spectra of mono-, bi- and multilayer MoS2. (c) Temperature-dependent Raman spectra of monolayer MoS2 under the excitation of 1.6mW laser at 514.5 nm wavelength.
Fig. 2
Fig. 2 (a) The temperature dependence of Raman shifts of E 2 g 1 and A1g modes of mono-, bi- and multilayer MoS2 under the excitation of 1.6mW laser at 514.5 nm wavelength. (b) The temperature dependence of Raman shifts of E 2 g 1 and A1g modes of multilayer MoS2 for three different laser powers at 514.5 nm wavelength.
Fig. 3
Fig. 3 (a) The best fitting to experimental Raman shift data (open squares) of monolayer MoS2 with Eq. (3) described in text. Component contributions from thermal expansion, three- and four-phonon processes are plotted by color dashed lines. (b) The ratio of three- to four-phonon process contributions of E 2 g 1 and A1g modes.
Fig. 4
Fig. 4 Strength coefficients of three- and four-phonon processes versus layer number of MoS2 films in temperature-dependent Raman shifts of E 2 g 1 and A1g modes.
Fig. 5
Fig. 5 (a)Temperature-dependent Raman spectra of monolayer MoS2 measured with 785nm laser. (b) Nonlinear temperature dependence of Raman shifts of E 2 g 1 mode measured with 514.5 nm (filled circles) and 785 nm (open squares) lasers. Black lines are the best fittings with the model of Eq. (3) in text.

Tables (1)

Tables Icon

Table 1 First- and second-order temperature coefficients extracted by linear and quadratic polynomial fittings.

Equations (7)

Equations on this page are rendered with MathJax. Learn more.

Ω ( T ) = ω 0 + χ 1 T + χ 2 T 2
Ω ( T ) = ω 0 + χ 1 T
ω ( T ) = ω 0 ' + Δ ω E ( T ) + Δ ω A ( T ) ,
Δ ω E ( T ) = ω 0 exp ( n γ 0 T α d T ) ω 0 ,
α a = ( 0.6007 × 10 5 + 0.6958 × 10 7 T a ) ( 1 c ° ) ,
α c = ( 0.1064 × 10 3 + 1.5475 × 10 7 T c ) ( 1 c ° ) ,
Δ ω A ( T ) = A [ 1 + 2 e x 1 ] + B [ 1 + 3 e y 1 + 3 ( e y 1 ) 2 ] = Δ ω A 3 p ( T ) + Δ ω A 4 p ( T ) ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.