You are on page 1of 27

View Article Online

View Journal

Journal of
Materials Chemistry A
Materials for energy and sustainability

Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: M. Kang, S. Kang,
W. Lee, N. Park, K. C. Kwon, S. Choi, G. Kim, J. Nam, K. S. Kim, E. Saitoh, H. W. Jang and S. Lee, J. Mater.
Chem. A, 2020, DOI: 10.1039/D0TA02629H.
Volume 6
Number 12
28 March 2018
This is an Accepted Manuscript, which has been through the
Pages 4883-5230
Royal Society of Chemistry peer review process and has been
Journal of accepted for publication.
Materials Chemistry A
Materials for energy and sustainability
rsc.li/materials-a
Accepted Manuscripts are published online shortly after acceptance,
before technical editing, formatting and proof reading. Using this free
service, authors can make their results available to the community, in
citable form, before we publish the edited article. We will replace this
Accepted Manuscript with the edited and formatted Advance Article as
soon as it is available.

You can find more information about Accepted Manuscripts in the


Information for Authors.

Please note that technical editing may introduce minor changes to the
text and/or graphics, which may alter content. The journal’s standard
ISSN 2050-7488 Terms & Conditions and the Ethical guidelines still apply. In no event
COMMUNICATION
Zhenhai Wen et al.
An electrochemically neutralized energy-assisted low-cost
shall the Royal Society of Chemistry be held responsible for any errors
acid-alkaline electrolyzer for energy-saving electrolysis
hydrogen generation
or omissions in this Accepted Manuscript or any consequences arising
from the use of any information it contains.

rsc.li/materials-a
Page 1 of 26 Journal of Materials Chemistry A
View Article Online
DOI: 10.1039/D0TA02629H

Large-scale MoS2 thin films with chemically formed


holey structure for enhanced Seebeck thermopower
and their anisotropic properties

Journal of Materials Chemistry A Accepted Manuscript


Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

Min-Sung Kang a,*, Soo-Young Kang a,*, ,Won-Yong Leea,*, No-Won Park a, Ki Chang Kownb,
Seokhoon Choib, Gil-Sung Kima, Jungtae Namc,d, Keun Soo Kimc, Eiji Saitohe,f,g, Ho Won Jangb,
and Sang-Kwon Leea,e**
a Department of Physics, Chung-Ang University, Seoul 06974, Republic of Korea
b Department of Materials Science and Engineering, Seoul National University, Seoul 08826, Republic of Korea
c Department of Physics and Astronomy, Sejong University, Seoul 05006, Republic of Korea
d Institute of Advanced Composite Materials, Korea Institute of Science and Technology, Wanju 55324, Republic of Korea
e Institute for Materials Research, Tohoku University, Sendai 980-8577, Japan
f WPI Advanced Institute for Materials Research, Tohoku University, Sendai 980-8577, Japan
g Department of Applied Physics, The University of Tokyo, Tokyo 113-8656, Japan

* These authors contributed equally to this work


**Address correspondence to sangkwonlee@cau.ac.kr
Journal of Materials Chemistry A Page 2 of 26
View Article Online
DOI: 10.1039/D0TA02629H

Abstract
The thermoelectric (TE) effect in nanoscale materials is of great interest as an ideal platform for
small-scale energy harvesting and micro-cooling technologies. In this regard, two-dimensional (2D)

Journal of Materials Chemistry A Accepted Manuscript


metal dichalcogenide (TMDC) can be considered as a promising material for TE applications owing
to its superior electronic and phonon transport properties provided by a favorable large energy
bandgap and an atomically thin layer, which serves as an ideal quantum well structure. We
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

investigate and conduct direct comparison of the cross-plane Seebeck coefficients of various TMDC
films prepared by using the chemical vapor deposition method. In particular, these coefficients for
MoS2, WSe2, and WS2 thin films are determined to be approximately 115, 129, and 211 µV/K at 300
K, respectively. The chemically formed holey structure of MoS2 thin films with a thickness of ~7 nm
exhibits superior in-plane Seebeck coefficients of ~742 µV/K, showing strong anisotropic behavior
with a ratio of ~6.5 along in- and cross-plane directions at 300 K. Such behavior can be explained
by the energy filtering effect in the holey MoS2 film in the in-plane direction. Moreover, an extremely
high anisotropic ratio (~2.4  108) of the power factor was observed owing to the large in-plane
Seebeck coefficients and the electrical conductivity of the MoS2 films across the samples. This study
is the first to assess the cross-plane Seebeck coefficients of large-scale MoS2, WS2, and WSe2 thin
films at 300 K which analyzing the anisotropic behavior of the MoS2 film. The results reported here
confirm the importance of providing reliable Seebeck coefficient information on films formed by
TMDC materials for further applications on TE devices.

KEYWORDS: In- and cross-plane Seebeck coefficients; Anisotropy; Two-dimensional layered


materials; Quantum confinement; Carrier filtering; Density of states

2
Page 3 of 26 Journal of Materials Chemistry A
View Article Online
DOI: 10.1039/D0TA02629H

1. Introduction

Thermoelectric (TE) energy conversion is an important technology for energy-harvesting device

Journal of Materials Chemistry A Accepted Manuscript


applications, which convert waste heat into electrical power and vice versa.1-6 The efficiency of TE

materials is generally characterized by a dimensionless figure of merit quantity, 𝑍𝑇 = 𝑆2𝜎𝑇 𝜅, where


Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

σ is the electrical conductivity, S is the Seebeck coefficient, T is the absolute temperature, and κ is

the thermal conductivity.4, 7-11 In particular, the Seebeck coefficient is a significant factor in the design

of high ZT TE materials. Low-dimensional materials and structures, such as two-dimensional (2D)12-


14 or one-dimensional/zero-dimensional (1D/0D) structures4,15-17 have promising advantages for

increasing the ZT value because S is strongly related to the profile of the density of states (DOS) at

the Fermi energy. For instance, HgTe quantum dots,16 SiGe nanowires,17 Te nanosheet,13 and

Bi2Te3/Bi0.5Sb1.5Te3 superlattice thin film14 were recently fabricated to enhance their TE properties.

Transition metal dichalcogenide (TMDC) materials have an MX2 stoichiometry where M and X

are a transition metal and chalcogen, respectively. In the past decade, low-dimensional TMDC

materials have attracted the attention of researchers due to their interesting optical and electronic

properties.18-21 For example, porous MoS2 and MoSe2 nanosheets19-21 were used as catalysts for

hydrogen evolution reaction, and NbSe2 nanosheets, nanorods, and nanoparticles were used as

counter electrodes in dye-sensitized solar cells.18 In the case of the TE energy conversion, recently

emerging 2D TMDC materials such as MoS2, WS2, WSe2 are ideal platforms for TE energy

generation applications because they naturally form 2D-layered materials with an atomic thickness

of <1 nm. These materials also provide superior electronic and phonon transport properties, which

significantly leads to high TE ZT values. Several research groups have recently reported remarkably

high TE Seebeck coefficients (i.e., thermopower) of single layers of MoS2 and WSe2 produced by

mechanical exfoliation22, 23 and chemical vapor deposition (CVD).24, 25 For example, Wu et al.

investigated electrically modulated TE Seebeck coefficients of CVD-grown single layer MoS2 and

3
Journal of Materials Chemistry A Page 4 of 26
View Article Online
DOI: 10.1039/D0TA02629H

obtained superior Seebeck coefficients of up to ~3 × 104 µV/K at room temperature.25 Similarly,

Buscema et al. reported remarkable Seebeck coefficients of ~1 × 104 µV/K for mechanically

Journal of Materials Chemistry A Accepted Manuscript


exfoliated single-layer MoS2 flakes produced by laser heating.23

Previous relevant studies have focused Seebeck coefficients of mechanically exfoliated and
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

CVD-grown 2D TMDC layers only in the in-plane direction (i.e., parallel to the sample surface). To

the best of our knowledge, no experimental results have been reported regarding the TE properties of

films formed by 2D TMDC materials in the cross-plane direction (i.e., perpendicular to the sample

surface). Moreover, although recent works investigated the in-plane electrical and Seebeck TE power

of 2D MoS2 layers, an extensive and systematic study of the anisotropic properties of films formed

by large-area MoS2 films has not been conducted.

Here, we investigate the cross-plane TE properties of CVD-prepared large-area MoS2, WSe2,

and WS2 films by measuring the Seebeck coefficients of the samples. In addition, the Seebeck

coefficients of the MoS2 films with chemically formed holey structures are measured along the two

directions, i.e., perpendicular (cross-plane) and parallel (in-plane) to the sample surfaces, to evaluate

the anisotropic properties in a vacuum chamber at temperature differences of 1 to 10 K. Our study

incorporates several distinct features compared with previous studies of the TE transport of TMDC

bulk and single-crystal materials. First, this study is the first to assess the cross-plane Seebeck

coefficients of CVD-prepared large-scale MoS2, WS2, and WSe2 thin films at 300 K. Second, our

approach for evaluating the Seebeck properties of films formed by TMDC materials yields reliable

measurements and is well suited for future device applications because the typical size of single

crystal samples is less than 10 µm2. Third, the anisotropic Seebeck coefficients and power factors of

the MoS2 films is an entirely first study. Fourth, the chemically formed holey MoS2 thin films show

extremely high in-plane Seebeck coefficients at 300 K. Fifth, the proposed methodology for

evaluating the TE properties of TMDC films will be extremely useful for understanding additional

4
Page 5 of 26 Journal of Materials Chemistry A
View Article Online
DOI: 10.1039/D0TA02629H

electronic and phononic properties in future studies.

Journal of Materials Chemistry A Accepted Manuscript


2. Experimental details

2.1 Sample growth of large-area MoS2, WS2, and WSe2 films


Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

Thin films of MoO3 and WO3 with 99.9% purity (iTASCO) at various thicknesses of 5, 10, and 15

nm were deposited by using an e-beam evaporator (Korea Vacuum Tech Co, Ltd). For synthesis of

the MoS2, WS2, and WSe2 thin films, we followed a simple sulfurization/selenization procedure with

H2 and N2 gases used for the sulfurization/selenization process in a low-pressure CVD system. First,

the CVD furnace temperature was raised to 900 °C (sulfurization)/800°C (selenization) and was

maintained for 40 min under the flow of H2 and N2 at 0.8 Torr. The flow rates of the H2 and N2 were

fixed to 100 and 500 cm3/min, respectively. Then, sublimation of S (99.998%, Sigma-Aldrich)/Se

power (99.5%, Sigma-Aldrich) was initiated in the other heating zone upstream at a temperature of

200 °C/300 °C. This process was repeated for 30 min to obtain successful substitution of O2 to S/Se

in the precursor thin films. For the surface morphology and roughness of the samples, atomic force

microscopy (AFM) and scanning electron microscopy (SEM) measurement were performed.

Furthermore, the topological characteristics, crystal structures, and elemental distributions in holey

MoS2 films were investigated by Raman spectroscopy and transmission electron microscopy (TEM)

combined with scanning TEM (STEM) analysis.

2.2 In- and cross-plane Seebeck coefficient measurements of 2D TMDC films

Fig. 1a shows a schematic of the Seebeck coefficient measurement in 2D-layered MoS2 films in the

perpendicular direction to the sample surface. As shown in the figure, the CVD-grown MoS2 films

were first wet-transferred onto the Cu/SiO2/Si substrate by sputtering, and then the top ~200 nm of

5
Journal of Materials Chemistry A Page 6 of 26
View Article Online
DOI: 10.1039/D0TA02629H

the Cu metals were sputtered onto the MoS2/Cu/SiO2/Si at 300 K. (Further details of the wet-transfer

process are given in the Supporting Information). The sandwiched structures of Cu/MoS2/Cu and

Journal of Materials Chemistry A Accepted Manuscript


SiO2/Si were finally created to evaluate the cross-plane Seebeck coefficient; details of the

experimental setup can be found in previous research.26 Briefly, as shown in Fig. 1a, the samples
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

were held between an AlN layer, which was used as a heat bath, and a micro-Peltier module, which

was used as a heating source. The upper plate was attached to the heat bath, and the lower plate was

placed on the top of the micro-Peltier module, which was thermally connected to the heat bath through

a Mo screw (Fig. 1a). By applying an electric current to the micro-Peltier module, a temperature

gradient (T) was generated across the Cu/MoS2/Cu samples. This temperature difference (T) was

measured by using two T-type thermocouples in which the cross-plane TE Seebeck voltages (V)

were measured by two shielded W needles between the upper and lower Cu plates (Fig. 1a).

Accordingly, the cross-plane Seebeck coefficients of 2D MoS2 films at a given temperature were

calculated by linearly fitting V versus T using 𝑆 = ― ∆𝑉 ∆𝑇. All Seebeck coefficients of only

MoS2 films (SMoS2) were evaluated by subtracting the Seebeck coefficient of the Cu/Cu metal layer26

(SCu ~+1.28 µV/K) from the measured Seebeck coefficient of the Cu sandwiched samples

(Cu/MoS2/Cu, SCu/MoS2/Cu).

Fig. 1b shows a schematic image of the in-plane Seebeck coefficient measurement of the 2D

MoS2 films, in which a temperature gradient was applied from left to right. As shown in the figure,

the CVD-grown 2D MoS2 films were transferred onto the insulating SiO2/Si substrate, and the top

electrodes were then formed at the ends of the samples by using a metal shadow mask. In the in-plane

Seebeck coefficient measurement (Fig. 1b), a strain gauge of 120  as heat and two T-type

thermocouples for controlling the temperature were attached to the top of the Cu plate and to the

dummy substrate, which was bridged between the Cu plate and block.27 All measurements were

performed in a microprobe system, as reported in the literature, for measuring the longitudinal and
6
Page 7 of 26 Journal of Materials Chemistry A
View Article Online
DOI: 10.1039/D0TA02629H

transverse spin Seebeck effect.28-31 The T and measured V values of the samples were monitored

and controlled using CompactDAQ with LabVIEW software (National Instruments, USA). In this

Journal of Materials Chemistry A Accepted Manuscript


system, the temperature difference varied from 1 to 10 K with excellent thermal stability using the

micro-Peltier modules and the stain gauge in the microprobe system.26 For other CVD-grown 2D-
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

layer TMDC films, specifically WSe2 and WS2 films, we used the same measurement procedure for

in- and cross-plane Seebeck coefficients as that used for the MoS2 films. Fig. 1a and 1b show the heat

flow distribution and electrical potential (i.e., TE Seebeck voltage) in the cross- and in-plane Seebeck

coefficient measurement system T = 10 K at the Cu/MoS2/Cu layers and the Cu/MoS2/SiO2/Si

substrate, respectively, using a multiphasic finite element method (FEM) simulator (COMSOL

Multiphysics, Inc., USA) with a heat transfer module for thermal simulation.

3. Results and discussion


3.1 Material characteristics of holey MoS2 thin films

The transfer of CVD-grown TMDCs from SiO2/Si onto target substrates has been successfully

achieved in previous research by using sacrificial poly(methyl methacrylate) (PMMA) film to support

the TMDCs and subsequent etching of SiO2 in alkali- or HF-based solutions.32-34 In particular, Kwon

et al. observed the number of the pinholes on the MoS2 thin films after the transfer process in HF-

only etchant.32 Based on this observation, we intentionally immersed MoS2-grown SiO2/Si substrates

into dilute HF etchant (5 wt%) for 60 s to create a holey structure and subsequently transferred them

onto the target substrates. Fig. 2a shows a field emission scanning electron microscopy (FESEM)

image of a chemically formed holey MoS2 thin film, indicating the presence of a large number of

holes with various diameters of ~1 to 5 μm. The formation of holey structure was assumed to be

related to fast SiO2 etch rate in HF-only etchant (SiO2 etch rate ~30 nm/s). Generally, the HF-only

etchant has higher wettability for buried SiO2 layer between PMMA polymer and Si substrate, leading

7
Journal of Materials Chemistry A Page 8 of 26
View Article Online
DOI: 10.1039/D0TA02629H

to accelerating chemical etching on the defected Mo sites caused by rapid detachment of

PMMA/MoS2 layer and consequently forming holey structure. To further investigate the surface

Journal of Materials Chemistry A Accepted Manuscript


topographic feature, we employed AFM in the non-contact mode. Fig. 2b shows the AFM mapping

image of the MoS2 thin film with holes that was transferred to the SiO2 (300 nm)/p-Si substrate. The
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

color AFM image shows uniform color distribution on the surface of the MoS2 layer, indicating that

the chemically formed holey MoS2 thin film had a smooth and clean surface. In addition, the average

thickness of the MoS2 thin film was about ~7 nm, corresponding to the AFM height profile (blue-

dotted line, A–B) as shown in Fig. 2c. The topological characteristics of the holey MoS2 thin film

were also identified by Raman mapping of the intensity of E2g1 and A1g, with wavenumbers of ~383.3

cm-1 and ~407.4 cm-1, respectively, as shown in Fig. 2d and 2e. These results are highly consistent

with the AFM topographic image (Fig. 2b). Fig. 2f shows the Raman spectra from two different

positions labeled C and D representing the surface and hole regions, respectively. The two peaks

appeared at ~383.3 and ~ 407.4 cm-1 on the surface of MoS2 layer, which were assigned to the E2g1

and A1g mode frequencies as excited by a 514.5 nm layer source.35 On contrary, the absence of both

modes at the hole region was generally considered to represent the disappearance of the MoS2 layer

after the transfer process in the dilute HF etchant. In order to gain an in-depth understanding of the

configuration of the MoS2 thin film, a cross-sectional TEM specimen was prepared by a focused-ion

beam (FIB) ion milling. Fig. 2g and 2h show cross-sectional HR-TEM images for the MoS2 thin film,

indicating the formation of ~6 nm-thick MoS2 consisting of 11 multilayers. According to the intensity

profile in Fig. 2i, the average interplanar d-spacing of the (002) planes was determined to be 0.626

nm, indicating that our MoS2 thin film is a 2D-layered material with layers bounded to each other by

van der Waals force (d-spacing = ~ 0.62).32, 36, 37 Furthermore, cross-sectional STEM imaging with

EDS mapping was performed to confirm the existence of Mo and S elements constituting the 2D-

layered MoS2 thin film. Fig. 2j shows the elemental mapping images for Si (red), Mo (green), and S

8
Page 9 of 26 Journal of Materials Chemistry A
View Article Online
DOI: 10.1039/D0TA02629H

(pink), corresponding to the STEM image. Thus, the Mo and S elemental distributions were mostly

consistent with the thickness of the 2D-layered MoS2 thin film; Si elements from the insulating

Journal of Materials Chemistry A Accepted Manuscript


SiO2/Si substrate were also observed. These results confirm the successful fabrication of chemically

formed holey MoS2 thin film comprising 11 2D-stacked multilayers using dilute HF solution, which
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

can be helpful for understanding the in- and cross-plane thermoelectric properties depending on

morphological characteristics such as holey structures.

3.2 Cross-plane Seebeck coefficients of MoS2, WSe2, and WS2 films

Fig. 3a shows the TE Seebeck potential difference in the measured voltage in the cross-plane direction

between the top and bottom Cu metal (inset image in Fig. 3a) as a function of the time and temperature

difference from 1 to 10 K. The results showed excellent thermal stability at each temperature

difference, which confirm that the temperature difference in the cross-plane Seebeck thermopower

measurement system was effectively controlled with low variation, <0.4%. In addition, such a

measurement system is amenable for reliable in- and cross-plane Seebeck thermopower

measurements of very thin films formed by TMDC material. Fig. 3b shows the photo-switching

behavior of films formed by the MoS2 layer in the in-plane direction based on the application of

electric light illumination at a fixed temperature difference of T = 1 K. The results showed a change

of approximately 3 µV in the in-plane voltage measurement owing to the photo-current response to

the light illumination. These results further confirm that CVD-grown MoS2 film can be a promising

photo-detector because it optimizes the device structure and thickness of the samples. This is because

the absorption spectrum of MoS2 is well matched to the solar spectrum with proper absorption

coefficients of 104–106 cm-1. Previous research on the single-layer MoS2 photo-detector demonstrated

high photo-responsivity with a fast response time.38, 39 Further research on new photo-controlled

Seebeck coefficients or TE devices with a more specific light source will be conducted on the basis

9
Journal of Materials Chemistry A Page 10 of 26
View Article Online
DOI: 10.1039/D0TA02629H

of the results obtained from this experiment.

Fig. 3c shows the potential difference of large-area MoS2 films across the in- and cross-plane

Journal of Materials Chemistry A Accepted Manuscript


directions as a function of the temperature difference of 1 to 10 K, showing good linearity with R2 =

0.9999 and R2 = 0.99 for the in- and cross-plane measurements, respectively. As a result, the in- and
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

the cross-plane Seebeck coefficients can be evaluated by linearly fitting V versus T, as shown in

Fig. 3c. In Fig. 3d, the measured Seebeck coefficients across the in- and cross-plane directions of the

CVD-grown MoS2 were determined to be 742 and 115 µV/K, respectively at 300 K, achieving high

Seebeck coefficients with a high anisotropic ratio of ~6.5 in these films. This positive value of the

Seebeck coefficient demonstrates that CVD-grown MoS2 has p-type conduction. This result can be

explained the interface interaction between MoS2 layer and CuO substrate. During the wet transfer

process, O dangling bonds are generated at the CuO surface by oxygen plasma treatment, which leads

to p-type electrical properties in MoS2 without any doping process. This phenomenon is well

consistent with previous literatures.40-44 In Section 3.3, we discuss the cross-plane Seebeck coefficient

(S) and in-plane Seebeck coefficient (S‖) of large-area MoS2 films. Fig. 3d shows the values of the

Seebeck coefficient (~115 µV/K) of films formed by the MoS2 layer for the first time using a reliable

measurement system. As discussed in Section 1, no experimental information is available for S of

films formed by MoS2 layer. As reported previously, the measured S and S‖ of CVD-grown MoS2

films were approximately 2.3 and 44 times higher than those for pristine MoS2 bulk at 300 K,

respectively.45 This high S can be explained by their low dimensionality with low electrical

conductivity (, Fig. 3e) and naturally formed atomic-scale 2D MoS2 films, whereas the DOS was

strongly increased owing to the quantum confinement in the MoS2 films. For low-dimensional

materials, the Seebeck coefficient can be expressed by the Bethe–Sommerfeld expansion of the Mott

relation:46, 47

10
Page 11 of 26 Journal of Materials Chemistry A
View Article Online
DOI: 10.1039/D0TA02629H

𝜋2𝑘2𝐵𝑇 𝐷𝑂𝑆(𝐸)
𝑆= 3𝑒 [ 𝑛(𝐸) + 𝜇(𝐸) ]
1 𝑑𝜇(𝐸)
𝑑𝐸 𝐸 = 𝐸 ,
𝐹
(1)

where e is the electron charge, DOS (E) is the energy-dependent electronic DOS, n(E) is the energy-

Journal of Materials Chemistry A Accepted Manuscript


dependent number of states, kB is the Boltzmann constant, and µ(E) is the energy-dependent electron

mobility. In the layered structure of MoS2 films, the quantum confinement effect in the MoS2 films
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

leads to an increase in the DOS density at the Fermi level, resulting in increased Seebeck

thermopower in the layered MoS2 films. Similar behavior has been observed previously in

superlattice films.2, 4, 8 In Fig. 3e, the in- (‖) and cross-plane () electrical conductivity of the MoS2

film at 300 K were determined to be ~786 and ~1 × 10-4 S/m, respectively, resulting in enhancement

of the S‖ value of the films by suppression of the σ‖ values. These results agree well with Equation

(1), which describes the relationship between the ‖ and S‖ of the sample. Furthermore, we compared

the measured S of our MoS2 thin film and those of other TMDC materials, specifically WSe2 and

WS2, through curve fitting using a linear polynomial, as shown in Fig. 3f. The average thicknesses of

the WS2 and WSe2 thin films were ~8 and ~14 nm, respectively (Fig. S1†), and the evaluated S at

300 K was determined to be 129 and 211 µV/K, respectively, as shown in Fig. 3g. As with the

measured S of MoS2 thin film (115 μV/K), the measured S of WSe2 and WS2 was also larger than

that of each pristine bulk, which corresponds to the thickness effect on the DOS for the TMDCs

sample.48-50

In Fig. 3d, the anisotropy in Seebeck thermopower was determined to be 6.5 at room temperature.

As discussed above, we measured the S of the MoS2 thin film using a reliable measurement system

for the first time; thus no experimental data of MoS2 thin film are available to compare the anisotropy

in the Seebeck coefficient. Instead, we compared the anisotropy using the measurement data for WS2

film, which is a well-known TMDC material.24 Because TMDC materials are formed by stacking of

monolayers and retention by van der Waals force,51, 52 the trend of the anisotropy value in the Seebeck
11
Journal of Materials Chemistry A Page 12 of 26
View Article Online
DOI: 10.1039/D0TA02629H

coefficient is similar to that of other TMDC films. The measured S‖ and S were to 620 and 112 µV/K

at 300 K, resulting in anisotropy of ~5.5 for the WS2 film. When comparing the anisotropy value, our

Journal of Materials Chemistry A Accepted Manuscript


result was slightly larger than that of the WS2 film.24 The thickness effect on the DOS has little

influence on anisotropy in Seebeck thermopower because it effects both S‖ and S, respectively. Thus,
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

this result is mainly attributed to the increase in S‖ in the MoS2 thin film caused by the holey structure,

as shown in Fig. 2a. This result is discussed further in Section 3.3.

On the basis of the measured electrical conductivity (Fig. 3e) and the Seebeck coefficient (Fig.

3d) in both the in- and cross-plane directions, the power factor (𝑃𝐹 = 𝑆2𝜎) of CVD-grown TMDC

films was calculated. As shown in Fig. S2a† in the ESI, the calculated in- (PF‖ ) and cross-plane (PF)

power factors of the MoS2 films were ~432.6 and ~1.8 × 10-6 µW/m·K2, respectively at 300 K,

achieving an extremely high anisotropic ratio ( ) of ~2.4 × 10 . The calculated PF


𝑃𝐹 ∥
𝑃𝐹 ⊥
8
‖ of the large-

area MoS2 films was approximately twice that of a similar CVD-grown MoS2 monolayer24 with a

carrier density of 1013 cm-2, at ~200 µW/m·K2. This superior result corresponds to the previously

reported value for O2-doped MoS2 bulk materials, at ~300  400 µW/m·K2 in both the in- and cross-
𝑃𝐹 ∥
plane directions.45 Accordingly, this high PF‖ and 𝑃𝐹 ⊥ clearly indicate that large-area MoS2 films

can be ideal TE devices in the in-plane direction when optimizing the structure and carrier

concentration. In addition, the calculated PF of WSe2 and WS2 films were ~2.8 × 10-7 and ~1.9 ×

10-7 µW/ m·K2 at 300 K (Fig. S2b in Supporting Information), respectively, based on the measured

cross-plane Seebeck coefficient (Fig. 3g) and electrical conductivity (Fig. 3h) of the films.

3.3 In-plane Seebeck coefficients of CVD-grown MoS2 films

As previously mentioned, the in-plane Seebeck thermopower of our MoS2 thin film was larger than

12
Page 13 of 26 Journal of Materials Chemistry A
View Article Online
DOI: 10.1039/D0TA02629H

that of other reported MoS2 films including MoS2 bulk,24, 45, 50, 53 which was measured to be 742 μV/K

as shown in Fig. 3d. This phenomenon can be explained by morphological characteristics such as the

Journal of Materials Chemistry A Accepted Manuscript


holey structure in MoS2 thin film. As shown in Fig. 2a and 4a, we confirmed through SEM

observation that our MoS2 thin film contains holes in various diameters of ~1 to ~5 μm. Two notable
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

effects of these holes on the in-plane Seebeck thermopower of the MoS2 thin film were noted. First,

an increase in the in-plane Seebeck thermopower observed for the MoS2 thin film with holes over

that without holes can be explained by an increase in the porosity of the film. When employing the

effective medium theories for the MoS2 thin film with holes, the Seebeck thermopower can be

expressed as 54

|ln (1 ― 𝜙eff
𝑝 )|
𝑘B
[
𝑆 = 𝑞 (𝜂 ― 𝜇 ∗ ) 1 + 𝜂 ― 𝜇∗ ], (2)

where 𝑞 is the carrier charge; 𝜇 ∗ is the reduced chemical potential (𝜇 ∗ = 𝜇/𝑘B𝑇); 𝜂 is the scattering

parameter depending on the dominant scattering mechanism, e.g., 𝜂 = 2 for carrier scattering on

acoustic phonons and 𝜂 = 4 for scattering on ionized impurities; and 𝜙eff


𝑝 is the effective porosity (0

< 𝜙eff
𝑝 < 1). According to Equation (2), the S‖ of the MoS2 thin film with holes was always lager

than that of the MoS2 thin film owing to its porosity, which is consistent with that reported in the

literature.54-56 Specifically, Lee et al. showed thorough Boltzmann transport calculation that Seebeck

thermopower increases not only when the pore size is on the nanometer-scale but also on the

micrometer-scale.56 Second, the increase in the Seebeck thermopower in MoS2 films with holey

structures is attributed to the energy filtering effect in the films, which increases the S‖ of the MoS2

thin film. In Fig. 4b and c, the chemically formed holes in the MoS2 thin film are regarded as

scattering sites due to a potential difference, which leads to an energy filtering effect near the hole

sites. Thus, the Seebeck thermopower of the MoS2 thin film with holes can be expressed from

Boltzmann transport equation as 57-63

13
Journal of Materials Chemistry A Page 14 of 26
View Article Online
DOI: 10.1039/D0TA02629H

[ ( ∂𝑓(𝐸)
)𝑑𝐸
]

∫𝐸 𝜏(𝐸)𝐷𝑂𝑆(𝐸)𝐸2 ― ∂𝐸
1 b
𝑆= 𝑒𝑇 ― 𝜇∗ , (3)

(
∫𝐸 𝜏(𝐸)𝑔(𝐸)𝐸 ―
b ∂𝐸 )𝑑𝐸
∂𝑓(𝐸)

Journal of Materials Chemistry A Accepted Manuscript


where 𝑇 is the absolute temperature, 𝐸 is the energy of the carriers, 𝜏(𝐸) is the relaxation time, 𝑓(𝐸)

is the Fermi distribution function, and 𝐸b is the potential barrier height obtained from the holes. As
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

shown in Fig. 4c, low-energy carriers are filtered out at the boundary of the charge trap at the interface

of the holes, whereas the carriers with high energy can easily pass through them without disruption

of the hole. As a result, the increase in average energy of the carriers (Fig. 4c), enhances the in-plane

Seebeck thermopower of the MoS2 thin films with holes. The in-plane electrical conductivity of the

this MoS2 thin film was determined to be ~785 S/m at room temperature (Fig. 3e), which is more than

100 times lower than that of MoS2 thin film without holes.64, 65 This result can be fully explained by

the energy filtering effect on the Seebeck thermopower of the films.

4. Conclusions

We investigated the cross-plane Seebeck thermopower properties of CVD-prepared 2D large-area

MoS2, WSe2, and WS2 TMDC thin films at 300 K. In particular, the Seebeck coefficients of the large-

area MoS2 films with chemically formed holey structures were measured along the in- and cross-

plane directions for evaluating anisotropic properties at temperature differences of 1 to 10 K. The

chemically formed holey structure of the MoS2 thin films exhibited superior in-plane Seebeck

coefficients of ~742 µV/K and strong anisotropic behavior with a ratio of ~6.5. This can be explained

by the porosity of the holey structure and the energy filtering effect in the film. Therefore, the large-

area MoS2 films can be ideal TE devices in the in-plane direction is their structure and carrier

concentration are optimized. Furthermore, the methodology proposed here for evaluating the TE

properties of TMDC films will be extremely useful for understanding additional electronic and

phononic properties in future studies.

14
Page 15 of 26 Journal of Materials Chemistry A
View Article Online
DOI: 10.1039/D0TA02629H

Conflict of interest

Journal of Materials Chemistry A Accepted Manuscript


There are no conflicts to declare.
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

Acknowledgements

This study was supported by the Priority Research Centers Program and by the Basic Science

Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry

of Education, Science and Technology (NRF-2020R1A2C1004979 and 2017R1D1A1B03031010).

Appendix A. Supplementary information

† Electronic supplementary information (ESI) is available: Wet transfer method for TMDC films and

supplementary Fig. S1-2 See DOI: 10.1039/

15
Journal of Materials Chemistry A Page 16 of 26
View Article Online
DOI: 10.1039/D0TA02629H

References

1. F. J. DiSalvo, Science, 1999, 285, 703-706.

Journal of Materials Chemistry A Accepted Manuscript


2. R. Venkatasubramanian, E. Siivola, T. Colpitts and B. O'Quinn, Nature, 2001, 413, 597-602.
3. L. D. Zhao, S. H. Lo, Y. S. Zhang, H. Sun, G. J. Tan, C. Uher, C. Wolverton, V. P. Dravid and
M. G. Kanatzidis, Nature, 2014, 508, 373-377.
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

4. G. Bulman, P. Barletta, J. Lewis, N. Baldasaro, M. Manno, A. Bar-Cohen and B. Yang, Nat.


Commun., 2016, 7, 10302.
5. A. T. Duong, V. Q. Nguyen, G. Duvjir, V. T. Duong, S. Kwon, J. Y. Song, J. K. Lee, J. E. Lee,
S. Park, T. Min, J. Lee, J. Kim and S. Cho, Nat. Commun., 2016, 7, 13713.
6. M. J. Lee, J. H. Ahn, J. H. Sung, H. Heo, S. G. Jeon, W. Lee, J. Y. Song, K. H. Hong, B.
Choi, S. H. Lee and M. H. Jo, Nat. Commun., 2016, 7, 12011.
7. O. Bubnova, Z. U. Khan, A. Malti, S. Braun, M. Fahlman, M. Berggren and X. Crispin, Nat.
Mater., 2011, 10, 429-433.
8. I. Chowdhury, R. Prasher, K. Lofgreen, G. Chrysler, S. Narasimhan, R. Mahajan, D. Koester,
R. Alley and R. Venkatasubramanian, Nat. Nanotechnol., 2009, 4, 235-238.
9. A. I. Boukai, Y. Bunimovich, J. Tahir-Kheli, J. K. Yu, W. A. Goddard and J. R. Heath, Nature,
2008, 451, 168-171.
10. J. R. Heath, Nature, 2007, 445, 492-493.
11. J. K. Yang, Y. Yang, S. W. Waltermire, X. X. Wu, H. T. Zhang, T. Gutu, Y. F. Jiang, Y. F.
Chen, A. A. Zinn, R. Prasher, T. T. Xu and D. Y. Li, Nat. Nanotechnol., 2012, 7, 91-95.
12. S. Shimizu, J. Shiogai, N. Takemori, S. Sakai, H. Ikeda, R. Arita, T. Nojima, A. Tsukazaki
and Y. Iwasa, Nat. Commun., 2019, 10.
13. G. Qiu, S. Huang, M. Segovia, P. K. Venuthurumilli, Y. Wang, W. Wu, X. Xu and P. D. Ye,
Nano Lett., 2019, 19, 1955-1962.
14. N.-W. Park, W.-Y. Lee, Y.-S. Yoon, G.-S. Kim, Y.-G. Yoon and S.-K. Lee, ACS Appl. Mater.
Interfaces, 2019, 11, 38247-38254.
15. T. C. Harman, P. J. Taylor, M. P. Walsh and B. E. LaForge, Science, 2002, 297, 2229-2232.
16. X. Lan, M. Chen, M. H. Hudson, V. Kamysbayev, Y. Wang, P. Guyot-Sionnest and D. V.
Talapin, Nat. Mater., 2020, 19, 323-329.
17. I. Donmez Noyan, G. Gadea, M. Salleras, M. Pacios, C. Calaza, A. Stranz, M. Dolcet, A.
Morata, A. Tarancon and L. Fonseca, Nano Energy, 2019, 57, 492-499.
18. M. A. Ibrahem, W.-C. Huang, T.-w. Lan, K. M. Boopathi, Y.-C. Hsiao, C.-H. Chen, W.
Budiawan, Y.-Y. Chen, C.-S. Chang, L.-J. Li, C.-H. Tsai and C. W. Chu, J. Mater. Chem. A,
2014, 2, 11382-11390.
16
Page 17 of 26 Journal of Materials Chemistry A
View Article Online
DOI: 10.1039/D0TA02629H

19. Z. Lei, S. Xu and P. Wu, Physical Chemistry Chemical Physics, 2016, 18, 70-74.
20. X. T. Tran, S. Poorahong and M. Siaj, RSC Advances, 2017, 7, 52345-52351.
21. K. C. Kwon, S. Choi, K. Hong, C. W. Moon, Y.-S. Shim, D. H. Kim, T. Kim, W. Sohn, J.-M.

Journal of Materials Chemistry A Accepted Manuscript


Jeon, C.-H. Lee, K. T. Nam, S. Han, S. Y. Kim and H. W. Jang, Energ Environ Sci, 2016, 9,
2240-2248.
22. M. Yoshida, T. Iizuka, Y. Saito, M. Onga, R. Suzuki, Y. J. Zhang, Y. Iwasa and S. Shimizu,
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

Nano Lett., 2016, 16, 2061-2065.


23. M. Buscema, M. Barkelid, V. Zwiller, H. S. J. van der Zant, G. A. Steele and A. Castellanos-
Gomez, Nano Lett., 2013, 13, 358-363.
24. J. Pu, K. Kanahashi, N. T. Cuong, C. H. Chen, L. J. Li, S. Okada, H. Ohta and T. Takenobu,
Phys. Rev. B, 2016, 94, 014312.
25. J. Wu, H. Schmidt, K. K. Amara, X. F. Xu, G. Eda and B. Ozyilmaz, Nano Lett, 2014, 14,
2730-2734.
26. N. W. Park, D. Y. Kang, W. Y. Lee, Y. S. Yoon, G. S. Kim, E. Saitoh, T. G. Kim and S. K.
Lee, ACS Appl. Mater. Interfaces, 2019, 11, 23303-23312.
27. K. Uchida, T. Ota, H. Adachi, J. Xiao, T. Nonaka, Y. Kajiwara, G. E. W. Bauer, S. Maekawa
and E. Saitoh, J. Appl. Phys., 2012, 111, 103903.
28. T. Kikkawa, K. Uchida, Y. Shiomi, Z. Qiu, D. Hou, D. Tian, H. Nakayama, X. F. Jin and E.
Saitoh, Phys. Rev. Lett., 2013, 110, 067207.
29. K. Uchida, S. Takahashi, K. Harii, J. Ieda, W. Koshibae, K. Ando, S. Maekawa and E. Saitoh,
Nature, 2008, 455, 778-781.
30. K. Uchida, J. Xiao, H. Adachi, J. Ohe, S. Takahashi, J. Ieda, T. Ota, Y. Kajiwara, H. Umezawa,
H. Kawai, G. E. W. Bauer, S. Maekawa and E. Saitoh, Nat. Mater., 2010, 9, 894-897.
31. J. Xiao, G. E. W. Bauer, K. Uchida, E. Saitoh and S. Maekawa, Phys. Rev. B, 2010, 81,
214418.
32. K. C. Kwon, S. Choi, K. Hong, C. W. Moon, Y. S. Shim, D. H. Kim, T. Kim, W. Sohn, J. M.
Jeon, C. H. Lee, K. T. Nam, S. Han, S. Y. Kim and H. W. Jang, Energ Environ Sci, 2016, 9,
2240-2248.
33. S. Najmaei, Z. Liu, W. Zhou, X. L. Zou, G. Shi, S. D. Lei, B. I. Yakobson, J. C. Idrobo, P. M.
Ajayan and J. Lou, Nat. Mater., 2013, 12, 754-759.
34. Y. C. Lin, W. J. Zhang, J. K. Huang, K. K. Liu, Y. H. Lee, C. T. Liang, C. W. Chu and L. J. Li,
Nanoscale, 2012, 4, 6637-6641.
35. M. Ye, D. Winslow, D. Zhang, R. Pandey and Y. Yap, Photonics, 2015, 2, 288-307.
36. K. C. Kwon, S. Choi, J. Lee, K. Hong, W. Sohn, D. M. Andoshe, K. S. Choi, Y. Kim, S. Han,

17
Journal of Materials Chemistry A Page 18 of 26
View Article Online
DOI: 10.1039/D0TA02629H

S. Y. Kim and H. W. Jang, J. Mater. Chem. A, 2017, 5, 15534-15542.


37. L. K. Tan, B. Liu, J. H. Teng, S. F. Guo, H. Y. Low and K. P. Loh, Nanoscale, 2014, 6, 10584-
10588.

Journal of Materials Chemistry A Accepted Manuscript


38. Z. Y. Yin, H. Li, H. Li, L. Jiang, Y. M. Shi, Y. H. Sun, G. Lu, Q. Zhang, X. D. Chen and H.
Zhang, ACS Nano, 2012, 6, 74-80.
39. O. Lopez-Sanchez, D. Lembke, M. Kayci, A. Radenovic and A. Kis, Nat. Nanotechnol., 2013,
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

8, 497-501.
40. K. Dolui, I. Rungger and S. Sanvito, Phys. Rev. B, 2013, 87, 165402.
41. G. Gregori, R. Merkle and J. Maier, Prog. Mater. Sci., 2017, 89, 252-305.
42. Y.-J. Lin and T.-H. Su, Appl. Surf. Sci., 2016, 387, 661-665.
43. C. N. R. Rao, K. Gopalakrishnan and U. Maitra, ACS Appl. Mater. Interfaces, 2015, 7, 7809-
7832.
44. Y. Zhan, Z. Liu, S. Najmaei, P. M. Ajayan and J. Lou, Small, 2012, 8, 966-971.
45. S. Kong, T. M. Wu, M. Yuan, Z. W. Huang, Q. L. Meng, Q. K. Jiang, W. Zhuang, P. Jiang
and X. H. Bao, J. Mater. Chem. A, 2017, 5, 2004-2011.
46. V. Y. Irkhi and Y. P. Irkhin, Electronic Structure, Correlated Effects and Physical Properties
of d- and f-Metals and Their Compounds, Cambridge International Science Pub, 2007.
47. J. He and T. M. Tritt, Science, 2017, 357.
48. A. Pisoni, J. Jacimovic, R. Gaal, B. Nafradi, H. Berger, Z. Revay and L. Forro, Scr. Mater.,
2016, 114, 48-50.
49. D. Wickramaratne, F. Zahid and R. K. Lake, J. Chem. Phys., 2014, 140, 124710.
50. K. Hippalgaonkar, Y. Wang, Y. Ye, D. Y. Qiu, H. Zhu, Y. Wang, J. Moore, S. G. Louie and X.
Zhang, Phys. Rev. B, 2017, 95, 115407.
51. A. Arab and Q. L. Li, Sci. Rep., 2015, 5, 13706.
52. H. H. Guo, T. Yang, P. Tao and Z. D. Zhang, Chin. Phys. B, 2014, 23, 017201.
53. M. Kayyalha, J. Maassen, M. Lundstrom, L. Shi and Y. P. Chen, J. Appl. Phys., 2016, 120,
134305.
54. R. H. Tarkhanyan and D. G. Niarchos, J. Mater. Res., 2013, 28, 2316-2324.
55. K. Valalaki, P. Benech and A. G. Nassiopoulou, Nanoscale Res. Lett., 2016, 11, 201.
56. H. Lee, D. Vashaee, D. Z. Wang, M. S. Dresselhaus, Z. F. Ren and G. Chen, J. Appl. Phys.,
2010, 107, 094308.
57. W. Y. Lee, N. W. Park, S. Y. Kang, G. S. Kim, J. H. Koh, E. Saitoh and S. K. Lee, J. Phys.
Chem. C, 2019, 123, 14187-14194.
58. N. W. Park, W. Y. Lee, Y. S. Yoon, G. S. Kim, Y. G. Yoon and S. K. Lee, ACS Appl. Mater.

18
Page 19 of 26 Journal of Materials Chemistry A
View Article Online
DOI: 10.1039/D0TA02629H

Interfaces, 2019, 11, 38247-38254.


59. J. H. Bahk, Z. X. Bian and A. Shakouri, Phys. Rev. B, 2013, 87, 075204.
60. D. Vashaee and A. Shakouri, Phys. Rev. Lett., 2004, 92.

Journal of Materials Chemistry A Accepted Manuscript


61. K. Kishimoto, M. Tsukamoto and T. Koyanagi, J. Appl. Phys., 2002, 92, 5331-5339.
62. M. Hong, Z. G. Chen and J. Ziou, Chin. Phys. B, 2018, 27, 048403.
63. G. D. Mahan and J. O. Sofo, Proc. Natl. Acad. Sci. U. S. A., 1996, 93, 7436-7439.
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

64. B. W. H. Baugher, H. O. H. Churchill, Y. F. Yang and P. Jarillo-Herrero, Nano Lett., 2013, 13,
4212-4216.
65. Y. J. Zhan, Z. Liu, S. Najmaei, P. M. Ajayan and J. Lou, Small, 2012, 8, 966-971.

19
Journal of Materials Chemistry A Page 20 of 26
View Article Online
DOI: 10.1039/D0TA02629H

Figure captions
Fig. 1. Seebeck thermopower measurement system of the MoS2 thin films in each direction.
(a) Schematics of measuring the cross-plane Seebeck thermopower of the MoS2 thin film. The

Journal of Materials Chemistry A Accepted Manuscript


electrical potential difference (V) between the bottom and top of the sample under the temperature
gradient was measured to obtain the cross-plane Seebeck thermopower of the samples (top image).
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

The schematic (bottom-left image) shows the measurement system used for both the Seebeck
coefficient and the electrical conductivity of the sample along the cross-plane direction (i.e., z-axis
direction). The micro-Peltier was used as a heater. The temperature difference (T) was generated
between the ends of the films along the z-direction (i.e., bottom to top). The temperature distribution
in this system was calculated by using COMSOL Multiphysics software at T = 6 K (bottom-right
image). (b) Schematic images for measuring the in-plane Seebeck thermopower of the MoS2 thin
film. The V between the ends of the samples under the temperature gradient was measured to
obtain the in-plane Seebeck coefficient of the samples (top image). The schematic (bottom-left
image) shows the measurement system used for both the Seebeck coefficient and the electrical
conductivity of the sample along the in-plane direction (i.e., x-axis direction). The strain gauge was
used as a heater. The T was generated between the ends of the films along the x-direction (i.e.,
left to right). The temperature distribution in the MoS2 film on the SiO2/Si substrate during the in-
plane Seebeck coefficient measurement at T = 10 was calculated by using COMSOL Multiphysics
software (bottom-left image).

Fig. 2. Material characteristics of holey MoS2 thin films. (a) FESEM image of chemically formed
holey MoS2 thin film. (b) and (c) AFM mapping image of the MoS2 thin film with holes transferred to
the SiO2 (300 nm)/p-Si substrate and its height profile (blue dotted line, A–B) in (b). (d) and (e)
Raman mapping images of the intensity of E2g1 and A1g, with wavenumbers of ~383.3 cm-1 and ~
407.4 cm-1, respectively, for the MoS2 thin film with holey structure. (f) Raman spectra from two
different positions labeled as C (surface region) and D (hole region). (g)–(i) Cross-sectional HR-TEM
images of ~ 6-nm-thick MoS2 thin film and its intensity profile (red dotted line) in (h). (j) Cross-
sectional STEM image and EDS mapping images for Si (red), Mo (green), and S (pink).

Fig. 3. TE and characterization of films formed by TMDC materials. (a) TE voltage measurement
data for MoS2 thin film in the cross-plane direction (inset image) as a function of the time and
temperature difference from 1 to 10 K. (b) Photo-switching behavior of the MoS2 thin film along the
in-plane direction when applying electric light illumination with a fixed temperature difference of T
= 1 K. (c) TE Seebeck potential difference (V) measured along in- and cross-plane directions for

20
Page 21 of 26 Journal of Materials Chemistry A
View Article Online
DOI: 10.1039/D0TA02629H

MoS2 thin film as a function of temperature difference (T) from 1 to 12 K for in-plane and 1 to 6 K
for cross-plane directions. Insets show photographs of the each Seebeck coefficient measurement
setup. (d) Measured Seebeck thermopower across the in- and cross-plane directions for MoS2 thin

Journal of Materials Chemistry A Accepted Manuscript


film at 300 K with its anisotropy. (e) Electrical conductivities of the MoS2 film in each direction. (f)
Cross-plane V for WS2, WSe2, and MoS2 thin films as a function of T from 1 to 6 K. Insets show
photographs of the measurement module and schematic. The measured (g) cross-plane Seebeck
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

thermopower and (h) the values for MoS2, WSe2, and WS2, are also shown.

Fig. 4. In-plane Seebeck coefficient increase in MoS2 thin films with holey structure. (a) Low-
magnification FESEM image of the MoS2 thin film with holes (top-view image). (b) Schematic image
measurement of the in-plane Seebeck coefficient of the MoS2 thin film with holey structures. (c)
Schematic showing how the trapped charge area and the energy barrier can lead to increases in the
in-plane Seebeck coefficient of the MoS2 thin film with holes. The position of the MoS2 layer is
represented by the area marked A in (b).

21
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

22
Fig. 1
Journal of Materials Chemistry A
DOI: 10.1039/D0TA02629H
View Article Online

Journal of Materials Chemistry A Accepted Manuscript


Page 22 of 26
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.
Page 23 of 26

23
Fig. 2
Journal of Materials Chemistry A
DOI: 10.1039/D0TA02629H
View Article Online

Journal of Materials Chemistry A Accepted Manuscript


Journal of Materials Chemistry A Page 24 of 26
View Article Online
DOI: 10.1039/D0TA02629H

Fig. 3

a b

Journal of Materials Chemistry A Accepted Manuscript


10 682
T=1K
TC
8 T=10 K V⊥ 680
9K
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.

V (mV)

6 8K Light-on

V‖ (V)
7K TH 678
6K
4 5K
4K 676
3K
2 2K
Light-off V||
1K 674
0 Large-scale MoS2 thin film TH TC

672
0 100 200 300 400 0 50 100 150 200
Counts Counts

c d e
Large-scale MoS2
1.0
1.0 10
10
8 MoS2 films

Electrical conductivity,  (S/m)


800
0.8
0.8 ~742 88
6 6.5
V(mV)

5.9 x 106 times


700
0.6 66
S (mV/K)

0.6
S / S ‖
Anisotropy in Seebeck

4
0.4 S‖ 44 600
2
0.0002
0.2 22
~115
0 0.0001
S ‖ 
0.0 00 0.0000
0 2 4 6 8 10
T (K) MoS2 films In-plane Cross-plane

f g h
1.6
1.4 250
211 0.2
1.2
200
(×10-3 S/m)

1.0
V(mV)

S (V/K)

0.8 150 129


115
0.6 0.1
100
0.4 TC
V⊥
MoS2
WSe2
MoS2

0.2
WS2

MoS2
50 WSe2
TH
0.0 WS2
0 0.0
-1 0 1 2 3 4 5 6 7 8 9 TMDC
T (K)
TMDC

24
Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.
Page 25 of 26

25
Fig. 4
Journal of Materials Chemistry A
DOI: 10.1039/D0TA02629H
View Article Online

Journal of Materials Chemistry A Accepted Manuscript


Published on 06 April 2020. Downloaded by University of Birmingham on 4/11/2020 5:33:44 PM.
Journal of Materials Chemistry A
DOI: 10.1039/D0TA02629H
View Article Online

Journal of Materials Chemistry A Accepted Manuscript


Page 26 of 26

You might also like