First Principle Investigation on Thermoelectric ... - ACS Publications

May 25, 2017 - First Principle Investigation on Thermoelectric Properties of Transition Metal Dichalcogenides: Beyond the Rigid Band Model. Ch. Adessi...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/JPCC

First Principle Investigation on Thermoelectric Properties of Transition Metal Dichalcogenides: Beyond the Rigid Band Model Ch. Adessi,* S. Thebaud, R. Bouzerar, and G. Bouzerar

Downloaded via KAOHSIUNG MEDICAL UNIV on August 2, 2018 at 20:21:27 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Univ Lyon, Université Claude Bernard Lyon 1, CNRS, Institut Lumière Matière, F-69622, Villeurbanne, France ABSTRACT: The thermoelectric properties (electrical conductivity, Seebeck coefficient, and power factor) of single layer transition metal dichalcogenides (MoS2, MoSe2, WS2, and WSe2) are investigated theoretically on the basis of ab initio quantum transport using the Landauer Büttiker formalism. The often used rigid band model is compared to realistic doping, namely substitution and adsorption, it is found that several important physical insights governing the transport are missing in this approximation. The rigid band model appears to clearly overestimate the thermoelectic efficiency, hampering its relevance for thermoelectric studies. Substitution doping by chloride or phosphorus leads to poor power factor due to drastic changes of the pristine band structure. In contrast, adsorption doping by alkalies (Li, Na, K, and Rb) favors larger power factor. Realistic treatment of the disorder induced by the dopants is also investigated and reveals that Cl doping leads to very short localization length of 3.5 nm while K comes with micrometer length scale. The Anderson localization phenomenon in thermoelectric properties of single layer transition metal dichalcogenides definitely comes out as a main issue.



INTRODUCTION Two-dimensional (2D) direct band gap1,2 transition-metal dichalcogenides (TMDC) with chemical composition MX2 (M = Mo or W and X = S or Se) have led to a growing interest in the past decade. For instance, in the field of nanoelectronic, significant accomplishments have been done in the development of MoS2 based transistors.3−5 More recently people have paid attention to this material in the prospect of thermoelectricity on both theoretical6−11 and experimental side.12−14 Thermoelectric generators which convert heat into electrical energy have gained a tremendous attention due to advances in nanoscale engineering. The efficiency of a thermoelectric material can be described by the dimensionless quantity called the figure of merit ZT = S2σT/κ, where S is the Seebeck coefficient, σ the electrical conductivity, T the temperature. and κ the thermal conductivity. Two ways to improve this figure of merit is to decrease κ or to increase the power factor (PF) i.e., S2σ. Low dimension materials like 2D compounds are expected to have higher thermoelectric figure of merit compared to bulk materials due to poor thermal conductivity.15,16 However, this is not so obvious for TMDC considering that the observed thermal conductivity of MoS2 is of the order of 30−50 W·m−1· K−1.17−19 More recently a lower value of 13.3 W·m−1·K−1 has been obtained for suspended monolayer.20 Nevertheless, even if the thermal conductivity of TMDC is not as low as some other promising thermoelectric materials,21 this issue is actually not critical for 2D materials due to recent progress in phonon engineering.22 In order to use semiconducting materials as thermoelectric devices, it is mandatory to perform doping. However, doping issues in TMDC are actually very intriguing. When dealing with © 2017 American Chemical Society

2D material or layered material, there are three ways to dope the material, substitution doping is the most conventional one and it has been successfully used experimentally. For example, p-type doping has been achieved on MoS2 by substituting Mo atoms by Nb23 or S by P.24 In the case of metal MoS2 contact a significant reduction of the Schottky barrier has been reached by chloride substitution of a sulfur corresponding to an electron doping.25 Another doping mechanism is vacancy induced doping which is suspected to have a major role in contact resistance lowering between metal and MoS2.26 Finally, a very efficient way to obtain a significant doping with 2D material is based on chemical adsorption. For instance, Kong et al.27 have recently shown that oxygen adsorption strategy leads to an enhancement of the thermoelectric properties of MoS2. Furthermore, adsorbed potassium has been successfully used as electron donor in the topic of field effect transistor.28 A comprehensive review of doping technics of MoS2 can be found in ref29. In this paper we investigate theoretically the influence of substitution and adsorption doping on the properties of the Seebeck coefficient, the electric conductance and the power factor within the Landauer Büttiker formalism (it is portrayed in Figure 1 for a TMDC single layer) at room temperature (T = 300 K). Within this framework, we are going beyond the usual semiclassical Boltzmann formalism in order to analyze accurately the influence of the disorder. The guiding line of the paper is to detail which doping procedure can lead to the largest power factor. Received: March 18, 2017 Revised: May 3, 2017 Published: May 25, 2017 12577

DOI: 10.1021/acs.jpcc.7b02570 J. Phys. Chem. C 2017, 121, 12577−12584

Article

The Journal of Physical Chemistry C

TRANSIESTA.38 For all the transport calculations, we have used 80 k-points in the transverse direction. On the basis of the Landauer-Büttiker formalism, the transmission reads: T (E) = tr[ΓLGr ΓRGa]

where Gr(a) is the retarded (advanced) Green’s function and ΓL(R) the coupling between the left (right) lead and the device (see Figure 1). The main assets of this formalism are to allow to deal with disordered systems and to fully include the quantum backscattering processes. Rigid Band Model. In order to have a starting point to evaluate the doping efficiency of the various technics and dopants, we first present some results for the PF obtained within the rigid band model. In this model, the Seebeck coefficient and the electrical conductivity are obtained from the Landauer formalism for a pristine layer by shifting rigidly the Fermi level of the system. When the Fermi level is shifted toward the valence bands, it mimics hole doping, and when shifted toward the conduction bands, it mimics electron doping. This approach allows one to easily get a first insight on how the material is behaving under doping providing that the band structure (or at least the bands of interest) of the host material is not significantly perturbed. This kind of modeling has been used to study the thermoelectric properties of TMDC in ref 39 with a various number of layers but without realistic doping. In Figure 2 is depicted the PF (the Seebeck coefficient and the electrical conductivity are inset) for the various TMDC. We have considered two major transport directions, the armchair and the zigzag ones. These directions are depicted on top of Figure 2 accordingly to the TMDC structures. These plots are given as a function of the Fermi level position but they can be linked with the carrier charge density (see Table 1 for specific values) by integration of the density of states ρ(E). The negative Fermi level is equivalent to hole doping and the positive Fermi level to electron doping. Basically, the maximum observed for the PF either for hole or electron doping (respectively the red and the blue curves) arises from the two opposite behavior of the Seebeck coefficient and the conductivity. As the carrier charge is increased, the Seebeck coefficient tends to zero (negatively or positively depending of the doping) and conversely, the electrical conductivity tends to increase with the carrier density. The maximum of the PF results from the combination of these two opposite trends. At first one can notice that the PF is almost similar for both transport directions, the anisotropy of these materials is actually quite small. Second, there is a strong asymmetry between hole and electron doping. Indeed, hole doping is always leading to a lower PF than electron doping. In Table 1 is given the Seebeck coefficient and the conductivity obtained for the maximum of the PF in the case of the zigzag transport direction (these values are pointed out in Figure 2 by a dashed line). By inspecting these values, one can notice that the difference between p and n doping is mainly due to the conductivity, it is always lower for p doping compared with n doping. This can be directly linked with the density of states which is smaller at the top of the valence band compared with that of the bottom of conduction band. Therefore, the thermoelectric performance of a full p−n thermoelectric device based on these materials would necessarily suffer from the limitations due to the p side. Anyhow, focusing at the n side, the maximum of the PF can reach a value almost as large as 2600 μW/K2m for WSe2 in the armchair direction. This value is in good accordance with

Figure 1. Schematic representation of the decomposition of the system used within the Landauer formalism and corresponding to a generic MX2 layer with adatoms.



RESULTS AND DISCUSSION Theoretical Background. All the calculations are performed using the SIESTA ab initio package30 under the generalized gradient approximation of Perdew, Burke, and Ernzerhof31 and using Troullier−Martin norm-conserving pseudopotentials.32 The basis used corresponds to a doubleζ-polarized (dζp) basis optimized using the simplex tool of the SIESTA package. All the atomic structures were optimized up to forces less than 0.01 eV/Å. A Monkhorst-pack of 10 × 10 × 10 k-points have been used for the calculations using a mesh cutoff of 400 Ry. The spin−orbit coupling is not taken into account in these calculations. One of the effect of this coupling is a band splitting33,34 leading to a degeneracy removal at some high symmetry points of the band structure. The expected influence of this band splitting on the transport properties is to lower a bit the Seebeck coefficient and thus the PF. However, this effect is not supposed to affect significantly our results. On the basis of ab initio calculations, the thermoelectric properties i.e. the Seebeck coefficient and the electrical conductance are calculated using the Landauer formalism35,36 for which the key ingredient is the transmission through the system. Within this formalism, the electrical conductance and the Seebeck coefficient can be expressed respectively as 2e 2 L0(EF ) ℏ

(1a)

1 L1(EF )/L0(EF ) kBT

(1b)

G(EF ) = − S(EF ) = −

with +∞

Ln(EF ) =

∫−∞

⎛ ∂f ⎞ (E − EF )n T (E)⎜ ⎟ dE ⎝ ∂E ⎠

(3)

(2)

where T(E) is the transmission, f(E) the Fermi distribution, and EF the Fermi energy. The temperature dependence comes from the Fermi distribution, and in the following, we have focused on room temperature (T = 300 K) calculations. G The conductivity σ is defined for these 2D systems as σ = L , ⊥

where L⊥ is the dimension of the layer perpendicular to the transport direction. Under this definition of σ for 2D systems, the physical dimension of the conductivity is equivalent to the conventional one. This definition is underpinned by the fact that under ballistic regime, the conductance does not depend on the length Lz of the device. The transmission is computed on the basis of Green’s function formalism using the Hamiltonian computed by SIESTA. The present ab initio quantum transport theory, fully described in ref 37, is similar to 12578

DOI: 10.1021/acs.jpcc.7b02570 J. Phys. Chem. C 2017, 121, 12577−12584

Article

The Journal of Physical Chemistry C

voltage. The results obtained with this technic are directly comparable to the rigid band model as the carrier concentration is solely tuned by a field effect. They have also measured the PF in Ta-doped WSe2, that was found 1 order of magnitude lower than that obtained by electrostatic doping. Kong et al.41 have also reported values of the same order of magnitude, 300 μW/ K2/m, in MoS2 doped by MoO2 nanoinclusions. However, this results must be qualified by the fact that, if the maximum is occurring for large value of the Fermi level, it will correspond to a larger dopant concentration (see Table 1 for the carrier charge per unit cell at the maximum of the PF). A material with a large PF at unrealistic dopant concentration is not useful. For instance, both MoSe2 and WS2 require a large carrier charge to achieve the maximum of PF (this carrier charge can be directly linked with the dopant concentration in the case a substitution doping). Thus, in the search of a material which allows to achieve a large PF, we must also take into account the energy difference between the maximum of the PF and the bottom of the conduction band. As a matter of fact, the smaller this difference the lower the needed concentration of dopant. Within this consideration, it appears that the best candidates for electron doping would be MoS2 and WSe2 for which the PF maxima are obtained for Fermi levels closer to the bottom of the conduction band corresponding to a carrier charge per unit cell of 0.01 and 0.015 respectively. For comparison, Fang et al.28 have measured for potassium doped MoS2 an electron concentration of 1.0 × 1013 cm−2 which corresponds to 0.009 e−/UC (using a cell parameter of 3.16 Å for MoS2). In order to discuss the validity of the rigid band scenario, in the following sections we now consider the effects of the different kinds of doping in a more realistic manner. Substitution Doping. There are several ways to dope TMDC by substitution, for instance it is possible to substitute a X atom (S or Se) by an atom in the neighboring columns of the periodic table. Indeed, the substitution by a phosphorus atom is leading to a hole doping and the substitution by a chloride atom is leading to an electron doping. The same is also valid for the transition metal atom. For example, the substitution by a rhenium atom is leading experimentally to a n-type semiconductor.42 In Figure 3 is given the density of states (DOS) obtained for two substitutional doping with either phosphorus or chloride in MoS2, for each case we have varied the dopant concentration. For x = 17% it corresponds to a 2 × 3 rectangular cell and for x = 4% to a 4 × 6 cell. At high concentration, a significant DOS is observed at the Fermi level either at the top of the valence band (P doping) or at the bottom of the conduction band (Cl doping). However, when the concentration is decreased, it can be noticed in both cases that the DOS tends to display sharp peaks around the Fermi level. These sharp peaks are associated with almost flat bands which are an indication on the localized nature of these states. Actually, when we inspect the projected density of states (PDOS) on the significant atomic orbitals, it should be stressed that the DOS at the Fermi level is induced by the dopant states and is absent in the intrinsic DOS of pure MoS2. The dopant hybridizes locally with Mo and S and is leading to specific bands with almost no dispersion for low dopant concentration, the only way to obtain a significant dispersion at the Fermi level is to use unrealistic doping concentration of the order of 17%. This is even more obvious when we compare the DOS obtained with the phosphorus dopant and the DOS of pristine MoS2. We can clearly see that

Figure 2. Power factors for (from top to bottom) MoS2, MoSe2, WS2 and WSe2. In inset is given the Seebeck coefficient and the electrical conductivity. Red and blue curves correspond respectively to hole and electron doping. On top is represented the transport direction in the TMDC layer. The black and yellow atoms are respectively the M and X atoms of MX2.

Table 1. Seebeck Coefficient, Conductivity, and PF for the Zigzag Transport Direction Obtained at the Maximum of the PF (Depicted by Dashed Lines in Figure 2) for Electron and Hole Dopinga MoS2 MoSe2 WS2 WSe2

S (μV/K)

σ (μS/Å)

PF (μW/K2/m)

carrier charge per UC

−156.6 +163.1 −86.9 +182.8 −85.8 +167.6 −172.7 +179.0

7.7 2.7 19.6 1.9 16.2 1.9 6.6 1.8

1888 718 1480 635 1193 534 1968 577

+0.010 −0.012 +0.090 −0.005 +0.065 −0.007 +0.015 −0.007

a

The last column is the carrier charge per unit cell. Negative and positive charge corresponds respectively to p- and n-type.

experimental one, for instance Yoshida et al.40 have obtained a PF of the order of 3200 μW/K2/m and a Seebeck coefficient of the order of −150 μV/K for an ultrathin WSe2 crystal by electrostatic doping (or field effect carrier doping) using a gate 12579

DOI: 10.1021/acs.jpcc.7b02570 J. Phys. Chem. C 2017, 121, 12577−12584

Article

The Journal of Physical Chemistry C

site is actually the most energetically favorable. However, when we consider an adsorption site at the center of an hexagon (almost as favorable as at the top of Mo), it is found that it makes no difference on the thermoelectric properties. In Figure 4 is presented the DOS obtained in the case of MoS2 doped with either potassium or lithium. In both cases, an

Figure 3. Electronic density of states (DOS) for P (top) and Cl (bottom) doped MoS2. For comparison purpose, the pristine DOS is also shown (it is shifted accordingly to the 4.2% case). The projected DOS (PDOS) is given for the 4.2% Cl doped case. The green-shaded backgrounds depict the location of the conduction band in the pristine case.

Figure 4. Electronic density of states for MoS2 doped by adsorption with K (top) and Li (bottom). The pristine case corresponds to the red curve.

the DOS close to the Fermi level does not correspond to that of pristine MoS2. Both kind of substitutional dopant clearly modify the band structure around the Fermi level. This already suggests that the rigid band model fails for those cases. This local states are actually not specific to chloride and phosphorus and are also observed with Br, F, N, and As. Adsorption Doping. One other promising strategy to dope 2D materials is chemical adsorption; this technic has been investigated theoretically for MoS243−45 on the basis of ab initio calculations. However, these calculations were not focusing specifically on the link between electronic and thermoelectric properties. A variety of adatoms have been examined, and only few atoms lead to a significant doping. Alkalies seem very promising because they are leading to a significant electron doping with no local states at the Fermi level. On the basis of this statement, we have thus focused on alkali doping for this study on transport properties. In this work, the adsorption site used for the alkalies is at the top of a molybdenum atom; this

electron doping results due to the donor nature of alkalies with respect to the MoS2 layer. The influence of these adatoms concentration is mainly to shift the Fermi level with no significant modification of the pristine DOS. This can be opposed to the case of substitution doping with phosphorus and chloride for which the DOS was significantly affected by the dopant. No indication of local states is observed, and the DOS at the Fermi level corresponds to bands already existing in the pristine material. Indeed, the electronic states of the alkalies are located at much higher energy and do not have any contribution at the bottom of the conduction band. This can be clearly seen by comparison with the DOS of the pristine MoS2 layer. We have observed this behavior with all alkali atoms until rubidium. The only influence of the dopant on the DOS shape is a light spreading of the bottom of the conduction band toward lower energy. This spreading of the DOS at the bottom 12580

DOI: 10.1021/acs.jpcc.7b02570 J. Phys. Chem. C 2017, 121, 12577−12584

Article

The Journal of Physical Chemistry C

and potassium and lithium dopants to adsorption doping. For comparison purpose, we have also plotted the result within the rigid band approximation. The PF is represented as a function of the carrier charge per unit cell. It actually corresponds for alkalies to the charge transferred by the adatom to the MoS2 layer, this charge is estimated by means of Mulliken analysis. We have also checked with Voronoi ̈ and Hirshfeld analysis the consistency of the estimated charge and the resulting relative uncertainty is about 10% for the largest concentration and about 5% for the smallest one. One should notice that under substitution doping by chloride the dopant concentration (referenced as x on Figure 3) is equivalent to the carrier charge per unit cell. However, this is not true for adsorption doping for which the carrier charge (or the transferred charge) is linked with the affinity between the adatom and the host layer. Note that this charge is always lower than the carrier charge obtained with an equivalent substitution doping concentration. Concerning the substitution with chloride, one can notice that the PF is much lower than the hypothetical one obtained from the rigid band model. On one side, starting from high concentration and decreasing it, the enhancement of the Seebeck coefficient leads first to an increase of the PF. On the other side, due to the local states, the conductivity then decreases dramatically and leads to a very rapid drop of the PF. Conversely, doping with alkali adatoms leads to larger values of the PF and are actually closer to the rigid band model. As argued previously, in the case of alkali doping, the bands involved in the transport are almost unaffected by the dopants and so we obtain a PF which is behaving more or less as the rigid band model but with smaller values. Indeed, due to the degeneracy removal induced by the dopants, the Seebeck coefficient obtained with alkalies is smaller than that obtained with the rigid band model (the conductivity is almost not affected). This is the reason why the PF is always smaller than that obtained within the rigid band model. When we compare the different alkalies, a significant difference can be noticed. Li is leading to poorer PF compared with K. In Table 2 is given the PF for various alkali adatoms

of the conduction band observed with alkali dopants has its origin in the degeneracy removal of some conduction bands. It is depicted in Figure 5 which corresponds to the band structure

Figure 5. Band structure of a 2 × 3 rectangular MoS2 lattice for pristine and Li doped (17%). For Li doping, the energy is shifted for comparison with the pristine case.

for a 2 × 3 rectangular cell of a 17% Li-doped MoS2 layer and a pristine one. The bottom of the conduction band of the pristine MoS2 single layer corresponds to four bands which are degenerated at the Γ and U points. This degeneracy is removed when the dopant is adsorbed on the layer. This behavior is definitely expected considering the symmetry breaking introduced by the dopant. However, it can be noticed that two bands are shifted toward lower energy at the U point leading to the spreading of DOS. The splitting at this point is 0.26 eV and it has, as it is described hereafter, a significant influence on the Seebeck coefficient. Power Factor for Realistic Doping. In this section we provide an overview of the thermoelectric properties of MoS2 realistically doped either with substitution or adsorbed atoms introduced in the previous sections. Transport properties of these doped materials are still performed within the Landauer formalism. On Figure 6 is represented the PF for three different dopants: Chloride dopant corresponds to a substitution doping

Table 2. Conductivity, Seebeck Coefficient, and Power Factor for Various Alkali Adatoms with a 4.2% Concentration MoS2 MoS2 MoS2 MoS2 MoSe2 WS2 WSe2

Li Na K Rb K K K

σ (μS/Å)

S (μV/K)

PF (μ W/K2/m)

3.6 1.8 5.2 5.7 4.6 5.9 4.1

−104.0 −105.5 −116.9 −113.0 −79.1 −85.3 −150.5

389.4 200.3 710.6 727.8 287.8 429.3 928.7

(concentration of 4.2%) and for the four different TMDC. One can notice that small alkalies like Li or Na lead to smaller PF than large adatoms such as K or Rb. The difference is actually coming from the conductivity and linked with the charge transfer between the alkali and the MoS2 layer which is smaller for Na and Li than K and Rb. This is leading, for the same dopant concentration, to a lower DOS at the Fermi level and thus to a lower conductance. The comparison between the four TMDC clearly points out that MoS2 and WSe2 are the best candidate for thermoelectric applications considering their high power factors at rather low dopant concentration.

Figure 6. Power factor for MoS2 with three different dopants: K, Li, and Cl plotted as a function of the carrier charge per unit cell. The rigid band calculation is added for comparison purpose. 12581

DOI: 10.1021/acs.jpcc.7b02570 J. Phys. Chem. C 2017, 121, 12577−12584

Article

The Journal of Physical Chemistry C Influence of the Disorder. One issue which have not yet been argued is the influence of the disorder. All the results presented previously are made artificially periodic by supercell construction and do not fully take into account the natural disordered location of the dopants. In order to have an insight on the influence of disorder, we have now considered systems where the dopants are randomly distributed. Indeed, it is possible within the Green’s function formalism to define an aperiodic device (see Figure 1). Moreover, using the tridiagonal shape of the Hamiltonian, we have dealt with system with length up to 50 nm, corresponding to 100 unit cells or more than 2500 atoms. We have thus defined systems on which the dopants are randomly distributed among the adsorption/ substitution sites and by varying the length of the system, it is then possible to evaluate the effects of the disorder on the thermoelectric properties. One should notice that we have voluntarily focused on scattering by static disorders and other scattering mechanisms as phonon scattering are not considered. We justify this approach by the fact that it has previously been shown46 that phonon scattering is not likely to be the limiting mechanism in MoS2 for the dopant concentration used here. The results are presented in Figure 7 for chloride substitution and potassium adsorption. In these plots, each point

influence of the disorder is very small and the PF is only reduced by a factor of 3. One key quantity to estimate quantitatively the influence of the disorder is the localization length.35,47 For disordered systems, the conductivity can be expressed as a decaying law with the length for which, the localization length (Lc) comes as a reference length of the system. It can be expressed as σ σ ( L ) = 2L / L 0 (4) e c−1 We have represented the corresponding law for the two dopants in Figure 7 as dotted lines (for K, it is almost not distinguishable from the data). This law is rather well verified, in the chloride case it is leading to a localization length of 3.5 nm and for potassium the derived Lc is about 1.5 μm. However, this last value is extracted from a very short length range compared with Lc and it must thus be taken as a lower bound of the possible Lc value for potassium. Anyhow, the 3 orders of magnitude between these two values is clearly illustrating the localization phenomenon occurring with chloride dopant.



CONCLUSIONS In summary, we have investigated the thermoelectric properties of single layer transition metal dichalcogenide on the basis of ab initio quantum transport within the Landauer Büttiker formalism. We conclude, on the basis of rigid band model, that MoS2 and WSe2 are leading, for equivalent doping concentration, to the largest PF under electron doping compared to MoSe2 and WS2. Moreover, by performing realistic doping, we observe that the substitution doping mechanism by chloride is inducing local states at the Fermi level, leading at low dopant concentration to a poor PF. In contrast, alkalies are acting as almost perfect electron donors, the band structure at the Fermi level is then very close to the band structure of the pristine TMDC layer. However, due to a degeneracy removal phenomenon, the PF obtained with alkali doping is not as large as the one obtained within the rigid band model. This point is tempering a bit the attractivity of TMDC for thermoelectric applications. Finally, we have investigated the influence of the disorder induced by the dopant. In case of adsorption doping by potassium, the PF is barely affected by the disorder. Conversely, the disorder induced by substitution doping by chloride leads to an exponential drop of the PF within a typical length scale of 30 nm. This last point highlights the utmost importance of the Anderson localization phenomenon for TMDC materials when evaluating the transport properties.

Figure 7. Power factor for Cl- and K-doped MoS2 as a function of the device length (Lz). The dopant concentration is set to 12.5%. Inset: Seebeck coefficient and the conductivity (the dotted lines corresponds to the fits). Each point corresponds to an average over 10 different configurations (random locations of the dopants).



corresponds to an average over 10 different systems with same length but with different random locations of the dopants. The uncertainty on these values is of the order of the marker size for K. For Li the uncertainty is increasing with length, below 10 nm it is of the order of the marker size and for 32 nm the PF value is (9 ± 6) × 10−4 μW/K2/m. For these calculations, the length perpendicular to the transport direction (L⊥) is kept constant and only the length Lz in the transport direction is varied. As expected, chloride doping is leading to very poor power factor when the length of the disordered system is increased. With raising the length of the system, an Anderson localization phenomenon is occurring, leading to an exponential decrease of the conductivity which is not compensated by the Seebeck increase. In these conditions, the PF is dropping very rapidly by 5 orders of magnitude with the length of the system. Conversely, for potassium the

AUTHOR INFORMATION

Corresponding Author

*(C.A.) E-mail: [email protected]. ORCID

Ch. Adessi: 0000-0001-5325-1039 Notes

The authors declare no competing financial interest.



REFERENCES

(1) Mak, K. F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T. F. Atomically Thin MoS2: A New Direct-Gap Semiconductor. Phys. Rev. Lett. 2010, 105, 136805. (2) Kadantsev, E. S.; Hawrylak, P. Electronic structure of a single MoS2 monolayer. Solid State Commun. 2012, 152, 909−913. (3) Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147−150. 12582

DOI: 10.1021/acs.jpcc.7b02570 J. Phys. Chem. C 2017, 121, 12577−12584

Article

The Journal of Physical Chemistry C

Low Energy Phosphorus Implantation. ACS Nano 2016, 10, 2128− 2137. (25) Yang, L.; Majumdar, K.; Liu, H.; Du, Y.; Wu, H.; Hatzistergos, M.; Tieckelmann, P. Y. H.; Tsai, W.; Hobbs, C.; Ye, P. D.; et al. Chloride Molecular Doping Technique on 2D Materials: WS2 and MoS2. Nano Lett. 2014, 14, 6275−6280. (26) McDonnell, S.; Addou, R.; Buie, C.; Wallace, R. M.; Hinkle, C. L. Defect-dominated doping and contact resistance in MoS2. ACS Nano 2014, 8, 2880−2888. (27) Kong, S.; Wu, T.; Yuan, M.; Huang, Z.; Meng, Q.-L.; Jiang, Q.; Zhuang, W.; Jiang, P.; Bao, X. Dramatically enhanced thermoelectric performance of MoS2 by introducing MoO2 nanoinclusions. J. Mater. Chem. A 2017, 5, 2004−2011. (28) Fang, H.; Tosun, M.; Seol, G.; Chang, T. C.; Takei, K.; Guo, J.; Javey, A. Degenerate n-Doping of Few-Layer Transition Metal Dichalcogenides by Potassium. Nano Lett. 2013, 13, 1991−1995. (29) Pham, V. P.; Yeom, G. Y. Recent Advances in Doping of Molybdenum Disulfide: Industrial Applications and Future Prospects. Adv. Mater. 2016, 28, 9024−9059. (30) Soler, J. M.; Artacho, E.; Gale, J. D.; García, A.; Junquera, J.; Ordejón, P.; Sánchez-Portal, D. The Siesta method for ab initio orderN materials simulation. J. Phys.: Condens. Matter 2002, 14, 2745−2779. (31) Perdew, J.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865−3868. (32) Troullier, N.; Martins, J. L. Efficient pseudopotentials for planewave calculations. Phys. Rev. B: Condens. Matter Mater. Phys. 1991, 43, 1993−2006. (33) Zhu, Z.; Cheng, Y.; Schwingenschlögl, U. Ultralow thermal conductivity and high thermoelectric figure of merit in SnSe crystals. Phys. Rev. B: Condens. Matter Mater. Phys. 2011, 84, 153402. (34) Kadantsev, E. S.; Hawrylak, P. Electronic structure of a single MoS2 monolayer. Solid State Commun. 2012, 152, 909−913. (35) Datta, S. Electronic transport in mesoscopic systems; Cambridge University Press: Cambridge, U.K., 1995. (36) Datta, S. Quantum transport: Atom to transistor; Cambridge University Press: Cambridge, U.K., 2005. (37) Adessi, C.; Roche, S.; Blase, X. Reduced backscattering in potassium-doped nanotubes: Ab initio and semiempirical simulations. Phys. Rev. B: Condens. Matter Mater. Phys. 2006, 73, 125414. (38) Brandbyge, M.; Mozos, J.-L.; Ordejón, P.; Taylor, J.; Stokbro, K. Density-functional method for nonequilibrium electron transport. Phys. Rev. B: Condens. Matter Mater. Phys. 2002, 65, 165401. (39) Wickramaratne, D.; Zahid, F.; Lake, R. K. Electronic and thermoelectric properties of few-layer transition metal dichalcogenides. J. Chem. Phys. 2014, 140, 124710. (40) Yoshida, M.; Iizuka, T.; Saito, Y.; Onga, M.; Suzuki, R.; Zhang, Y.; Iwasa, Y.; Shimizu, S. Gate-Optimized Thermoelectric Power Factor in Ultrathin WSe2 Single Crystals. Nano Lett. 2016, 16, 2061− 2065. (41) Kong, S.; Wu, T.; Yuan, M.; Huang, Z.; Meng, Q.-L.; Jiang, Q.; Zhuang, W.; Jiang, P.; Bao, X. Dramatically enhanced thermoelectric performance of MoS2 by introducing MoO2 nanoinclusions. J. Mater. Chem. A 2017, 5, 2004−2011. (42) Tiong, K.; Liao, P.; Ho, C.; Huang, Y. Growth and characterization of rhenium-doped MoS2 single crystals. J. Cryst. Growth 1999, 205, 543−547. (43) Li, X. D.; Fang, Y. M.; Wu, S. Q.; Zhu, Z. Z. Adsorption of alkali, alkaline-earth, simple and 3d transition metal, and nonmetal atoms on monolayer MoS2. AIP Adv. 2015, 5, 057143. (44) Rastogi, P.; Kumar, S.; Bhowmick, S.; Agarwal, A.; Chauhan, Y. S. Doping Strategies for Monolayer MoS2 via Surface Adsorption: A Systematic Study. J. Phys. Chem. C 2014, 118, 30309−30314. (45) Chang, J.; Larentis, S.; Tutuc, E.; Register, L. F.; Banerjee, S. K. Atomistic simulation of the electronic states of adatoms in monolayer MoS2. Appl. Phys. Lett. 2014, 104, 141603−141603−4. (46) Kaasbjerg, K.; Thygesen, K. S.; Jacobsen, K. W. Phonon-limited mobility in n-type single-layer MoS2 from first principles. Phys. Rev. B: Condens. Matter Mater. Phys. 2012, 85, 115317.

(4) Lembke, D.; Kis, A. Breakdown of High-Performance Monolayer MoS2 Transistors. ACS Nano 2012, 6, 10070−10075. (5) Das, S.; Chen, H.-Y.; Penumatcha, A. V.; Appenzeller, J. High Performance Multilayer MoS2 Transistors with Scandium Contacts. Nano Lett. 2013, 13, 100−105. (6) Jin, Z.; Liao, Q.; Fang, H.; Liu, Z.; Liu, W.; Ding, Z.; Luo, T.; Yang, N. A Revisit to High Thermoelectric Performance of Single-layer MoS2. Sci. Rep. 2016, 5, 18342. (7) Huang, W.; Da, H.; Liang, G. Thermoelectric performance of MX2 (M = Mo,W; X = S,Se) monolayers. J. Appl. Phys. 2013, 113, 104304. (8) Fan, D. D.; Liu, H. J.; Cheng, L.; Jiang, P. H.; Shi, J.; Tang, X. F. MoS2 nanoribbons as promising thermoelectric materials. Appl. Phys. Lett. 2014, 105, 133113. (9) Huang, W.; Luo, X.; Gan, C. K.; Quek, S. Y.; Liang, G. Theoretical study of thermoelectric properties of few-layer MoS2 and WSe2. Phys. Chem. Chem. Phys. 2014, 16, 10866−10874. (10) Babaei, H.; Khodadadi, J. M.; Sinha, S. Large theoretical thermoelectric power factor of suspended single-layer MoS2. Appl. Phys. Lett. 2014, 105, 193901. (11) Wang, R.-N.; Dong, G.-Y.; Wang, S.-F.; Fu, G.-S.; Wang, J.-L. Variations of thermoelectric performance by electric fields in bilayer MX2 (M = W, Mo; X = S, Se). Phys. Chem. Chem. Phys. 2017, 19, 5797−5805. (12) Wu, J.; Schmidt, H.; Amara, K. K.; Xu, X.; Eda, G.; Ö zyilmaz, B. Large Thermoelectricity via Variable Range Hopping in Chemical Vapor Deposition Grown Single-Layer MoS2. Nano Lett. 2014, 14, 2730−2734. (13) Yoshida, M.; Iizuka, T.; Saito, Y.; Onga, M.; Suzuki, R.; Zhang, Y.; Iwasa, Y.; Shimizu, S. Gate-Optimized Thermoelectric Power Factor in Ultrathin WSe2 Single Crystals. Nano Lett. 2016, 16, 2061− 2065. (14) Wang, T.; Liu, C.; Xu, J.; Zhu, Z.; Liu, E.; Hu, Y.; Li, C.; Jiang, F. Thermoelectric performance of restacked MoS2 nanosheets thin-film. Nanotechnology 2016, 27, 285703. (15) Hicks, L.; Dresselhaus, M. Effect of quantum-well structures on the Thermoelectric figure of merit. Phys. Rev. B: Condens. Matter Mater. Phys. 1993, 47, 12727. (16) Hicks, L.; Dresselhaus, M. Thermoelectric figure of merit of a one-dimensional conductor. Phys. Rev. B: Condens. Matter Mater. Phys. 1993, 47, 16631. (17) Yan, R.; Simpson, J. R.; Bertolazzi, S.; Brivio, J.; Watson, M.; Wu, X.; Kis, A.; Luo, T.; Hight Walker, A. R.; Xing, H. G. Thermal Conductivity of Monolayer Molybdenum Disulfide Obtained from Temperature-Dependent Raman Spectroscopy. ACS Nano 2014, 8, 986−993. (18) Jo, I.; Pettes, M. T.; Ou, E.; Wu, W.; Shi, L. Basal-plane thermal conductivity of few-layer molybdenum disulfide. Appl. Phys. Lett. 2014, 104, 201902. (19) Sahoo, S.; Gaur, A. P. S.; Ahmadi, M.; Guinel, M. J.-F.; Katiyar, R. S. Temperature-Dependent Raman Studies and Thermal Conductivity of Few-Layer MoS2. J. Phys. Chem. C 2013, 117, 9042−9047. (20) Bae, J. J.; Jeong, H. Y.; Han, K. H.; Kim, J.; Kim, H.; Kim, M. S.; Moon, B. H.; Lim, S. C.; Lee, Y. H. Thickness-dependent in-plane thermal conductivity of suspended MoS2 grown by chemical vapor deposition. Nanoscale 2017, 9, 2541−2547. (21) Zhao, L.-D.; Lo, S.-H.; Zhang, Y.; Sun, H.; Tan, G.; Uher, C.; Wolverton, C.; Dravid, V. P.; Kanatzidis, M. G. Ultralow thermal conductivity and high thermoelectric figure of merit in SnSe crystals. Nature 2014, 508, 373−377. (22) Balandin, A. A.; Nika, D. L. Phononics in low-dimensional materials. Mater. Today 2012, 15, 266−275. (23) Suh, J.; Park, T.-E.; Lin, D.-Y.; Fu, D.; Park, J.; Jung, H. J.; Chen, Y.; Ko, C.; Jang, C.; Sun, Y.; et al. Doping against the Native Propensity of MoS2: Degenerate Hole Doping by Cation Substitution. Nano Lett. 2014, 14, 6976−6982. (24) Nipane, A.; Karmakar, D.; Kaushik, N.; Karande, S.; Lodha, S. Few-Layer MoS2 p-Type Devices Enabled by Selective Doping Using 12583

DOI: 10.1021/acs.jpcc.7b02570 J. Phys. Chem. C 2017, 121, 12577−12584

Article

The Journal of Physical Chemistry C (47) Landauer, R. Electrical resistance of disordered one-dimensional lattices. Philos. Mag. 1970, 21, 863−867.

12584

DOI: 10.1021/acs.jpcc.7b02570 J. Phys. Chem. C 2017, 121, 12577−12584