Huan
Lin†
ab,
Ridong
Wang†
cb,
Hamidreza
Zobeiri
b,
Tianyu
Wang
d,
Shen
Xu
*eb and
Xinwei
Wang
*b
aSchool of Environmental and Municipal Engineering, Qingdao University of Technology, Qingdao, Shandong 266033, P. R. China
bDepartment of Mechanical Engineering, Iowa State University, Ames, IA 50011, USA. E-mail: shxu16@sues.edu.cn; xwang3@iastate.edu
cState Key Laboratory of Precision Measuring Technology and Instruments, Tianjin University, Tianjin, 300072, P. R. China
dInstitute of Chemistry, Chinese Academy of Sciences, Beijing, 100190, P. R. China
eSchool of Mechanical and Automotive Engineering, Shanghai University of Engineering Science, Shanghai, 201620, P. R. China
First published on 22nd March 2021
Although 2D materials have been widely studied for more than a decade, very few studies have been reported on the in-plane structure domain (STD) size even though such a physical property is critical in determining the charge carrier and energy carrier transport. Grazing incidence X-ray diffraction (XRD) can be used for studying the in-plane structure of very thin samples, but it becomes more challenging to study few-layer 2D materials. In this work the nanosecond energy transport state-resolved Raman (nET-Raman) technique is applied to resolve this key problem by directly measuring the thermal reffusivity of 2D materials and determining the residual value at the 0 K-limit. Such a residual value is determined by low-momentum phonon scattering and can be directly used to characterize the in-plane STD size of 2D materials. Three suspended MoSe2 (15, 50 and 62 nm thick) samples are measured using nET-Raman from room temperature down to 77 K. Based on low-momentum phonon scattering, the STD size is determined to be 58.7 nm and 84.5 nm for 50 nm and 62 nm thick samples, respectively. For comparison, the in-plane structure of bulk MoSe2 that is used to prepare the measured nm-thick samples is characterized using XRD. It uncovers crystallite sizes of 64.8 nm in the (100) direction and 121 nm in the (010) direction. The STD size determined by our low momentum phonon scattering is close to the crystallite size determined by XRD, but still shows differences. The STD size by low-momentum phonon scattering is more affected by the crystallite sizes in all in-plane directions rather than that by XRD that is for a specific crystallographic orientation. Their close values demonstrate that during nanosheet preparation (peeling and transfer), the in-plane structure experiences very little damage.
Although 2D materials have been widely studied for decades, very few studies have been reported on the structure domain size (STD size) in the in-plane direction even though such physical information is critical in determining the charge carrier and energy carrier transport. It is generally known that XRD spectroscopy can provide detailed information about the crystallite size of materials in a particular direction. The average crystallite thickness along a specific lattice plane direction can be obtained from each XRD peak. However, for extremely thin 2D materials (e.g. mono-layer samples), it is very difficult to obtain the STD size in the in-plane direction from XRD spectroscopy as the cross-section is extremely small and X-ray scattering is too weak to be detected. It should be pointed out that phonon scattering in thin 2D materials is strongly hindered by the STD size and could be used to characterize it. In addition, it is clear that the thermal transport of most kinds of 2D materials is dominated by the phonon–phonon scattering (Umklapp scattering) at near RT. Therefore, the Umklapp scattering is very strong and the structure domain boundary scattering could be relatively weak if the domain size is large. The grain boundary will have different scattering effects on phonons of different wavelengths and it is difficult to characterize the boundary scattering effect at moderate or high temperatures. Instead of XRD spectroscopy, the thermal reffusivity theory can be used to characterize the STD size by considering the phonon scattering in different lattice directions. At low temperatures, the mean free path of phonon–phonon scattering becomes very long due to the decrease of phonon density. Because when the temperature approaches absolute zero, all phonons are frozen, the phonon–phonon scattering is weakened. The only scattering source is structural scattering by grain boundaries, and surface and point defects and this structural scattering is independent of temperature. Therefore, the mean free path at extremely low temperatures can reflect the structural information about the MoSe2 film. Wang's group first used the thermal reffusivity variation against temperature to determine the residual thermal reffusivity value at the 0 K limit and used this parameter to characterize the STD size of different materials.9–15 Xu et al. for the first time used the thermal reffusivity theory to determine the phonon thermal resistivity of DNA.9 Cheng et al. found that the STD size of an individual polycrystalline silver nanowire was 38.5 nm, which was larger than the value obtained from the XRD result [21 nm in the (311) direction].10 In Xie et al.'s work, the STD size of the 3D graphene foam material was determined to be 166 nm, smaller than the crystallite size determined from the XRD result [201.8 nm corresponding to the (002) plane].11 Xie et al. reviewed that the STD size of human hair (1.6 nm) agrees well with the crystallite size obtained from the XRD result (1.8 nm).12 The thermal reffusivity theory has also been successfully used in investigating the STD size of other materials including ultrahigh molecular weight polyethylene (UHMWPE) fibers,13 SiC microwires14 and graphene paper (GP).15 Liu et al. found that the STD size for the two UHMWPE fiber samples is 8.06 and 9.42 nm. They are smaller than the crystallite size determined by XRD [19.7 nm in the (002) direction].13 Zhu et al. determined the STD size of three kinds of SiC microwires as 9.35, 1.42 and 1.03 nm, respectively, which are proportional to the corresponding crystalline size determined by XRD: 67–113, 14.6–18.4, and 5.85–7.84 nm.14 Han et al. predicted a STD size of 375 nm for widely studied normal graphite materials, close to the 404 nm grain size uncovered by TEM. They also evaluated c-MFP induced by defect in the GP at 234 nm based on the thermal reffusivity theory.15 Based on the above materials, the thermal reffusivity theory has been fully verified.
In this work, we measure the thermal diffusivity of thin MoSe2 films (thickness from 15 nm to 62 nm) from RT down to 77 K using a novel nanosecond energy transport state-solved Raman (nET-Raman) technique. From thermal transport characterization we also obtain the thermal conductivity and optical properties of the MoSe2 samples. Then the thermal reffusivity model is used to study the structural defect levels. The STD size from the thermal reffusivity model is very close to the result uncovered by XRD of bulk MoSe2.
In the steady state, a CW laser (Excelsior, Spectra-Physics) is employed as the heating and probing source as shown in Fig. 1(b) and (d). During laser heating, Raman scattering induced by the heating laser could be collected and used to analyze the sample's thermal response. In the MoSe2 sample, the heat generated from photon absorption will then propagate along its in-plane and out-of-plane directions. As the thickness of the sample is only a few to tens of nm, which is much smaller than its lateral size, it is reasonable to neglect the energy transfer and temperature distribution in the thickness direction. In this state, the temperature rise only depends on the in-plane thermal conductivity. By using different laser powers (P), we can use the Raman shift power coefficient (RSC), ψCW = ∂ω/∂P = a·∂ω/∂T·f1(k), to describe the temperature rise against the incident power. ψCW is dependent on the laser absorption coefficient (a), the temperature coefficient of Raman shift (∂ω/∂T), and the in-plane thermal conductivity (k).
In the other energy transport state–transient energy transport state, a nanosecond (ns) laser (DCL AIO Laser, Photonics Industries, International, Inc.) is focused on the center of the same suspended sample. Fig. 1(c) illustrates that a ns laser is focused on the center of the suspended sample, which realizes local heating and transient Raman probing. Therefore, as shown in Fig. 1(e), in this transient state, the RSC has a different expression as ψns = ∂ω/∂P = a·∂ω/∂T·f2(k, ρcp). ψns is dependent on the volumetric heat capacity (ρcp) of the MoSe2 sample besides the abovementioned properties including the laser absorption coefficient, the temperature coefficient of the Raman shift, and the in-plane thermal conductivity.
With the values of ψCW and ψns, a normalized RSC can be determined as Θ = ψns/ψCW = f3(k, ρcp). In this way, the absorption coefficient and Raman temperature coefficient are eliminated. Assuming that the volumetric heat capacity of MoSe2 has the same value as the bulk counterpart, the value of Θ only depends on the in-plane thermal conductivity of MoSe2 films. Or if the specific heat is unknown, Θ will be determined by the sample's in-plane thermal diffusivity (α = k/ρcp). Note that since the sample is very thin, the laser absorption depth (τL) value has negligible effect on the finally determined thermophysical properties.16
For the steady-state heating, the energy transport governing equation is:17,18
k∇2TCW + ![]() | (1) |
![]() | (2) |
For transient-state heating, the pulse width of the nanosecond laser is 76 ns (full width at high maximum), and the interval between the two pulses is 10 μs. During the pulse heating, the diffusion length of MoSe2 in the thickness direction is about 3 μm (ref. 20) which is much longer than the sample thickness (about 62 nm or less). Therefore, it is reasonable to assume that the sample has a uniform temperature distribution in the thickness direction. On the other hand, the time interval between two pulses (10 μs) is long enough for the MoSe2 sample to cool down to the ambient temperature after a ns pulse heating. It is proved that there is no interference between ns pulses and no steady-state accumulated heat in the ns laser heating case. Therefore, the Fourier governing equation of nanosecond laser heating pulse can be written as:21
![]() | (3) |
![]() | (4) |
The theoretical ratio of the temperature rise of the sample in the two states could be obtained by solving eqn (1) and (3), while the measured ratio is equal to the normalized RSC from the acquired Raman spectra. Through comparing the experimental normalized RSC with the theoretical ratio, we could determine the in-plane thermal conductivity of the sample based on the different contributions of heat conduction in these two energy transport states.
In order to determine the in-plane thermal conductivity of the MoSe2 sample, a 3D heat conduction model is used to calculate the temperature rise in two energy transfer states. Through calculation, the theoretical relationship between the temperature rise ratio in the two cases and the in-plane thermal conductivity could be built. Finally, by varying the temperature of the environmental cell, the in-plane thermal conductivity of the samples against the temperature could be determined according to the measured Θ. More details on the simulation process are given in our previous work.20 It needs to point out that during experiments, the measured Raman shift change in fact reflects the temperature rise that is a Raman intensity-weighted average over space for the CW case and over time and space for the ns laser case. Such physics is taken into full consideration in our computer modeling.22
Fig. 2(a)–(c) show the atomic force microscopy (AFM) images of three suspended MoSe2 samples. In order to avoid sample damage of the suspended areas during AFM imaging, the thickness of the supported area near the suspended area is measured and used as the thickness of the sample. In Fig. 2(a1)–(c1), the red dashed line represents the thickness profile shown in Fig. 2(a2)–(c2). The thickness of MoSe2 samples is 15, 50 and 62 nm, respectively. The surface roughness is then expressed by the maximum thickness variation (Δlmax) along a line on the sample surface. As shown in Fig. 2(a3)–(c3), the values of Δlmaxof these samples are relatively small compared with the thickness of the samples. When the thickness increases, Δlmaxalso increases a little bit.
Sample thickness (nm) | Temperature (K) | CW laser power range (mW) | ns laser power range (mW) | CW laser spot diameter (μm) | ns laser spot diameter (μm) |
---|---|---|---|---|---|
15 | 296 | 0.15–0.62 | 0.01–0.04 | 2.967 | 2.412 |
50 | 77 | 0.75–2.97 | 0.19–0.48 | 3.030 | 2.175 |
62 | 77 | 0.82–3.25 | 0.37–0.95 | 3.268 | 2.071 |
Fig. 3(a) and (b) show the contour plots of the Raman peak at ∼240 cm−1 under varied powers of the two different lasers, indicating that the peak is redshifted with the increased laser power. This concludes that the local temperature of the sample increases with the increased laser power. It can be seen in Fig. 3(c) and (d) that the measured Raman shift of the A1g peak (∼240 cm−1) is almost linearly related to the laser power. Accordingly, the A1g peak of MoSe2 is used to determine the RSC value in this work. The linearly fitted slope, RSC of the A1g mode, is −0.345 ± 0.002 cm−1 mW−1 under a CW laser, and −2.318 ± 0.103 cm−1 mW−1 under a ns laser. The corresponding normalized RSC (Θ) is 6.72 ± 0.31. Fig. 3(e) shows the theoretical Θ curve of the 50 nm thick sample at 77 K, 225 K and 296 K. The value obtained in the experiments is used to interpolate the theoretical result to determine the thermal diffusivity. Taking the 50 nm thick MoSe2 at 77 K for example, the thermal diffusivity is determined as (1.52 ± 0.14) × 10−5 m2 s−1 as shown in the figure.
![]() | ||
Fig. 3 The 50 nm-thick sample at 77 K is used to illustrate the results of the nET-Raman experiment. (a) and (b) 2D contour plots to demonstrate the variation of the Raman spectrum with laser power of (a) CW laser and (b) ns laser. (c) and (d) The Raman wavenumber shift with laser power for CW and ns lasers. The solid lines in the figures are the linear fitting result for Raman power coefficient. The insets are the spots of two lasers. (e) Data fittings of experimental Θ against that obtained by 3D numerical modeling to determine the thermal diffusivity. These are for the 50 nm thick MoSe2 sample at 77 K, 225 K and 296 K. (f) The measured thermal diffusivity, specific heat of MoSe2 from the literature29 and calculated thermal conductivity of 50 nm-thick MoSe2. |
The measured thermal diffusivity for the 50 nm thick MoSe2 layer from RT to 77 K is presented in Fig. 3(f). As the temperature goes down, α increases slowly. The variation trend of thermal diffusivity with temperature will be explained later according to the concept of thermal reffusivity. The experimental result of the specific heat29 of MoSe2 is plotted in Fig. 3(f). Its density is 6900 kg m−3.30 Using the specific heat and density of MoSe2, the thermal conductivity knET (determined using this nET-Raman) of MoSe2 is calculated and the results are shown in Fig. 3(f). knET increases from 9.19 to 15.07 W m−1 K−1 from RT to 100 K, and drops to 12.24 W m−1 K−1 at 77 K. It should be pointed out that the temperature rise of the 50 nm thick sample under laser irradiation is less than 30 K. It can be seen from Fig. 3(c) that the change of Raman shift is about 0.7 cm−1 in the CW case. The average temperature coefficient of the Raman shift (∂ω/∂T) is around 0.008 cm−1 K−1 as shown in Fig. 4(b). Therefore, we can calculate the temperature rise of the sample under a CW laser spot as ΔT = Δω/(∂ω/∂T) = 88 K. The average temperature rise of the entire sample in all domains is about 29 K.22 The average temperature rise of the entire sample in all domains under a ns laser is less than 29 K as the laser heating time is very short and the sample experiences a transient process. Our measurement data agree well with those published using other Raman techniques. Differences are largely due to the sample-to-sample structure variation and experimental control deviations. For instance, Wang et al. measured the thermal conductivity of suspended MoSe2 increasing from 11.1 ± 0.4 to 20.3 ± 0.9 W m−1 K−1 with the increased thickness (45 nm to 140 nm) at RT.20 Zobeiri et al. measured the in-plane thermal conductivity of these films increased from 6.2 ± 0.9 to 25.7 ± 7.7 W m−1 K−1 when the sample thickness varied from 5 to 80 nm using frequency-domain energy transport state-resolved Raman (FET-Raman) technique.22 The reduction in the thermal conductivity is due to the enhancement of the phonon scattering effect on the surface of thinner samples. Using the optothermal Raman technique, Zhang et al. found that the room-temperature thermal conductivities are 59 ± 18 W m−1 K−1 and 43 ± 12 W m−1 K−1 for single-layer and bi-layer MoSe2, respectively.31 Due to the phonon–phonon scattering, the thermal conductivities decrease when supported on a substrate and decrease with the increased temperature. Gu et al. estimated that the phononic thermal conductivity of single-layer MoSe2 is 54 W m−1 K−1 using the first-principles-based PBTE approach.32
Here we also take the 50 nm thick sample for example to illustrate the results. Fig. 4(b) and (d) show the temperature coefficient of the Raman shift (∂ω/∂T) and the RSC (∂ω/∂P) against temperature. As we know, (∂ω/∂P)/(∂ω/∂T) = ∂T/∂P = ΔT1, which is the temperature rise in the sample under unit laser energy irradiation. From simulation, a temperature rise ΔT0 in each case can also be obtained considering the laser spot size and thermal conductivity knET. In this way, the laser absorption coefficient a can be determined as a = ΔT1[1 − exp(−Δz/τL)]/ΔT0, in which τL = λ/(4πkL)19 is the laser absorption depth of MoSe2, λ is the laser wavelength of 532 nm, and kL is the extinction coefficient. The term [1 − exp(−Δz/τL)] is the laser absorption used in the modeling. When λ is 532 nm, the refractive index and extinction coefficient of MoSe2 are 4.8 and 2.08,33 respectively, and we have τL = 20.4 nm. In this way, the calculated values of a are all around 0.25 as shown in Fig. 4(d).
The multiple reflection of the incident laser beam in a suspended film will enhance the absorption of the incident laser by the film, which has been studied in the previous work.34,35 Inspired by these references, we calculate the theoretical a to be 0.52 with rigorous consideration of the multi-reflection and interference within the 50 nm thick sample. The theoretical laser absorption coefficient is about twice that obtained from the experiment. We speculate that there are several factors accounting for this difference. First, in the studies reported by other researchers, the laser beam absorption was evaluated based on the refractive index of n = 4.8 (ref. 33) for MoSe2. However, this optical property varies in a large range due to the differences among the measured samples. For instance, from ref. 36 and 37, they found that the refractive index and extinction coefficient of MoSe2 are 2.25, 0.75 and 4.22, 1.61 at 532 nm. Second, our samples are housed in a cryogenic environment cell with a layer of glass on top of it. The laser irritates the sample after going through the glass as shown in Fig. 4(c), and the calculated reflectivity of glass is 6% when the laser is perpendicular to the surface of the sample. Additionally, the laser beam is focused with a limited numerical aperture. The reflectivity of glass for this converged laser beam will be higher than 6% which could bring in large errors in laser absorption calculation. Last but not least, the spacing between MoSe2 layers within the sample might be increased (due to delamination) during the sample preparation. The presence of spacing, although tiny ones, can greatly alter the absorption behavior. Our finding also raises a critical point: it is extremely challenging and difficult to determine the precise laser absorption in 2D materials’ thermal characterization. Instead, the nET-Raman technique completely eliminates this problem and determines the thermophysical properties of 2D materials with unprecedented accuracy.
Θt = Θ0 + C × e−θ/2T, | (5) |
It is clear that the thermal reffusivity drops when the temperature decreases and finally reaches a constant value (Θ0) at the 0 K limit. C is a constant, and Θ0 is termed the residual thermal reffusivity and describes the influence of defect scattering only. θ is a constant proportional to Debye temperature. By fitting the thermal reffusivity against temperature data using the thermal reffusivity model, the Θ0 value and structure domain size can be extracted. Taking the 62 nm thick sample for example, the results are shown in Fig. 5(a). The fitting curves by the thermal reffusivity model are also given in the figure, which provides a good fitting of the experimental data. From the fitting result, we have Θt = 5.10 × 104 + 4.46 × 105 × e−234/T, and the Θ0 value is 5.10 × 104 s m−2. To avoid damaging the 15 nm thick sample, we only conduct the nET-Raman experiment at RT since its Raman signal is not very good at low temperatures under ns laser heating. At other temperatures, CW Raman is used to obtain kCW from the relation of kCW = ∂ω/∂T·∂ω/∂P·f. As shown in Fig. 5(a), the thermal reffusivity of the 15 nm thick sample has large variations, mostly due to the large uncertainties in its weak Raman signal. However, the residual thermal reffusivity can still be determined based on the data trend, which is quite close to that of the 62 nm-thick sample.
![]() | ||
Fig. 5 (a) Variation of the thermal reffusivity with temperature for the three samples. The residual thermal reffusivities for 62 nm-thick and 50 nm-thick samples are determined to be 5.10 × 104 and 7.34 × 104 s m−2, respectively. This is caused by the defects in the films. (b) XRD pattern of bulk MoSe2. According to the XRD results, the crystallite sizes determined by the peaks (002), (004) (006), (008), (100), (110) and (010) are 94.2 nm, 67.4 nm, 64.8 nm, 52.6 nm, 31.3 nm, 112 nm and 121 nm, respectively. (c) The phonon dispersion relationship and irreducible representations of phonon modes at A, G and M points are shown for MoSe2. Note that the path Γ–A refers to the [001] direction, and Γ–M is parallel to the [010] direction38 (reproduced from ref. 38 with permission from The Royal Society of Chemistry). (d) Phonon scattering in the in-plane direction. The arrows indicate the direction of phonon scattering and propagation. |
There are three kinds of acoustic phonon modes: longitudinal acoustic (LA), flexural acoustic (ZA), and transverse acoustic (TA) mode. In the LA mode, the atomic displacements are along the wave propagation direction. In the TA mode, the in-plane displacements are perpendicular to the propagation direction, and the ZA mode corresponds to out-of-plane atomic displacements. The phonon dispersion gives the relationship between the phonon wave vector q and the phonon energy E or frequency ω (E = hω where h is the reduced Planck constant). At the 0 K-limit, the phonon momentum goes to zero, which corresponds to the low phonon wave vector q near the center of the Brillouin zone. At this location, the frequency of the LA and TA modes has approximately linear dispersions41,42 as ωLA ≈ vLAq and ωTA ≈ vTAq, respectively. The phonon group velocity is v = dω/dq. Therefore, the velocities vLA, vTA and vZA are the slopes of each dispersion curve at point Γ(0,0,0) in Fig. 5(c), which are 799, 313, and 494 m s−1, respectively. From the equation v−1 = 1/3(vLA−1 + vTA−1 + vZA−1), the average phonon velocity v can be calculated as 464 m s−1. For the in-plane thermal reffusivity, we have Θ0 = 2/(vl0). Θ0 is found to be 7.34 × 104 s m−2 for the 50 nm-thick sample, so the STD size l0 is determined to be 58.7 nm. Using the same procedure, the STD size for the 62 nm thick sample is calculated to be 84.5 nm.
To further understand the in-plane STD size, XRD is used to characterize the structure of MoSe2 bulk and the XRD pattern is shown in Fig. 5(b). The sharp peaks reveal the polycrystallinity of the MoSe2 bulk that is used to prepare the nanosheets. This is analyzed to determine the crystal structure of MoSe2. The observed interplane distance d values are used to determine the crystal structure. The MoSe2 bulk shows prominent (002), (004), (006), (008), (100), (110) and (010) peaks. The crystallite size in each direction is calculated using the Scherrer formula and the results are shown in Fig. 5(b). Specifically, for the in-plane direction, the crystallite size is determined to be 64.8 nm for (100), 31.3 nm for (110), and 121 nm for (010) direction. In the out-of-plane directions, the crystallite size is 94.2 nm for (002), 67.4 nm for (004), 52.6 nm for (006), and 112 nm for the (008) direction. These similar crystallite sizes determined from different peaks indicate that the crystallite in MoSe2 is sphere-like. Compared with the value of 64.8 nm [(100) plane] characterized by XRD, the STD size determined by low-momentum phonon scattering is very close, but still shows some differences. XRD can be used for providing detailed information about the crystallite size and structural order of materials in a specified direction. The average crystallite thickness along a specific lattice plane is obtained from each XRD peak. However, the thermal reffusivity model characterizes the STD size by considering the phonon scattering from all the lattice directions. Therefore, the STD size given by the thermal reffusivity model is actually the effective domain size with a combined microcrystalline effect from every in-plane direction as shown in Fig. 5(d). So, deviation could arise when compared with those by XRD. Theoretically, the grain sizes in different crystallographical directions can be obtained from the TEM results. However, TEM is rarely used in 2D material crystallite size characterization. The main reason is that the preparation of the sample is very difficult. Additionally, the sample will suffer from additional damage in the process of transfer and the anticipated information cannot be obtained. However, the STD size determined by low-momentum phonon scattering is very close to the in-plane crystallite size of bulk MoSe2. This indicates that during our nm-thick MoSe2 preparation, the sample experiences very little in-plane structure damage.
Footnote |
† Equal contribution authors. |
This journal is © The Royal Society of Chemistry 2021 |