Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2020 Jun 10;117(25):13929–13936. doi: 10.1073/pnas.2007495117

Large enhancement of thermoelectric performance in MoS2/h-BN heterostructure due to vacancy-induced band hybridization

Jing Wu a,b,1, Yanpeng Liu b,c,d,1, Yi Liu b,e, Yongqing Cai f, Yunshan Zhao b,e,g, Hong Kuan Ng a,h, Kenji Watanabe i, Takashi Taniguchi i, Gang Zhang j, Cheng-Wei Qiu b,e,k, Dongzhi Chi a, A H Castro Neto b,h, John T L Thong b,e,2, Kian Ping Loh b,c,k,2, Kedar Hippalgaonkar a,b,l,2
PMCID: PMC7321979  PMID: 32522877

Significance

The study of correlated phenomena in 2D semiconductors opens up new pathways toward understanding and engineering material functionalities (such as thermoelectrics) in easily accessible van der Waals solids. Local structural defects such as vacancies inevitably exist in natural as well as synthetic TMD crystals and have been predicted to serve as magnetic impurities capable of enhancing the strongly correlated effect. Herein we discover unusual thermoelectric behavior in sulfur vacancy-enriched MoS2 by rationally selecting h-BN as the substrate. We demonstrate that the thermoelectric transport properties can be strongly manipulated by vacancy-induced Kondo hybridization. A significant enhancement of thermoelectric power factor by two orders of magnitude is achieved in the MoS2/h-BN device.

Keywords: thermoelectric, Kondo, 2D materials, Seebeck, phonon drag

Abstract

Local impurity states arising from atomic vacancies in two-dimensional (2D) nanosheets are predicted to have a profound effect on charge transport due to resonant scattering and can be used to manipulate thermoelectric properties. However, the effects of these impurities are often masked by external fluctuations and turbostratic interfaces; therefore, it is challenging to probe the correlation between vacancy impurities and thermoelectric parameters experimentally. In this work, we demonstrate that n-type molybdenum disulfide (MoS2) supported on hexagonal boron nitride (h-BN) substrate reveals a large anomalous positive Seebeck coefficient with strong band hybridization. The presence of vacancies on MoS2 with a large conduction subband splitting of 50.0 ± 5.0 meV may contribute to Kondo insulator-like properties. Furthermore, by tuning the chemical potential, the thermoelectric power factor can be enhanced by up to two orders of magnitude to 50 mW m−1 K−2. Our work shows that defect engineering in 2D materials provides an effective strategy for controlling band structure and tuning thermoelectric transport.


Thermoelectrics are solid state energy converters that can be used to harvest electrical energy from waste heat; thus, they are attractive as a sustainable energy resource. The performance of thermoelectric materials is characterized by the thermoelectric figure of merit (ZT, defined as S2σT/κ, where T is the absolute temperature): the key bottleneck is the interdependency between the Seebeck coefficient (S), electrical (σ), and thermal (κ) conductivities. An ideal thermoelectric material should exhibit simultaneously large S and σ (1). It has been proposed that van der Waals materials, in their two-dimensional (2D) form with discretized density of states (DOS) (quantum confinement), may show better performance in thermoelectrics compared to bulk materials (2). For instance, bulk layered SnSe has been found to have a figure of merit (ZT) of 2.8 ± 0.5. The remarkably high ZT is not due to quantum confinement but instead due to the anharmonicity of its chemical bonds, giving it an ultralow thermal conductivity at high temperature (700 to 900 K) (3, 4). However, the power factor (PF = S2σ) is still relatively modest (1 mW m−1 K−2) (3). New transport mechanisms are needed to push the PF of thermoelectric materials beyond the well-known S-σ anticorrelation limit (1). For example, the violation of the Mott relation results in enhanced PF in the hydrodynamic charge transport regime in graphene due to strong inelastic scattering among electrons (5). It is well known that impurity states, especially magnetic impurities, not only strongly couple with itinerant charge carriers but can also significantly affect the band structure and even reverse charge transport behavior (68). Few-atoms-thick layered materials are highly sensitive to such extrinsic effects, and therefore, they serve as a good platform for exploiting thermoelectric properties by introducing such impurities. Among these materials, MoS2 has attracted special attention because of its well-defined spin-splitting under light illumination and/or applied magnetic fields (911). Extrinsic magnetic impurities in MoS2 can be introduced by vacancies, dislocations, edges, strain, or doping by magnetic ions (1216). However, due to its extreme sensitivity to external fluctuations and turbostratic interfaces, there has not yet been a clear correlation between the thermoelectric properties of layered 2D materials and impurities.

In this work, we report that when few-layer MoS2 flakes are interfaced to hexagonal boron nitride (h-BN) substrate, a correlation between charge carriers and vacancy impurities in MoS2 is clearly observed. This is in part due to reduced scattering by charged surface states and surface roughness on the atomically smooth and chemically inert h-BN substrate. Using low-temperature scanning tunneling microscopy (LT-STM), single-atom sulfur vacancies (up to ∼1.81 ± 0.4 × 1012 cm−2) were observed on MoS2. The conduction subband near the vicinity of such a sulfur vacancy shows a colossal band splitting of 50 ± 5 meV. Density functional theory (DFT) simulations reveal that this band hybridization may originate from sulfur vacancy-induced magnetic states. When fabricated into a MoS2/h-BN field effect transistor (FET) device, we observed an anomalous sign change of the Seebeck coefficient and an extremely large positive Seebeck value (∼2 mV/K) even in the metallic regime at 60 K. Furthermore, an electrostatically tunable PF of MoS2/h-BN device larger by two orders of magnitude is achieved due to strong interactions between conducting charges and the observed vacancy states, attaining a value as high as 50 mW m−1 K−2 in the temperature range from 30 to 50 K. Our results demonstrate that engineering vacancies in 2D layered crystals provide a material platform for next generation thermoelectric and energy applications.

Results

Vacancy-Induced Band Hybridization in MoS2/h-BN Heterostructure.

Inversion symmetry breaking at vacancy or edge sites of MoS2 has been theoretically predicted to enable local magnetic moments and induces band splitting of both valence and conduction bands (17). To probe band splitting induced by vacancies in MoS2, an atomically flat h-BN substrate is employed to avoid spurious effects from impurities at the interfaces. The MoS2/h-BN heterostructure device was fabricated by a well-established transfer technique and then loaded into the LT-STM chamber and annealed in ultrahigh vacuum (<1×1010torr) to ensure better electrical and thermal contact and a cleaner surface (more details in SI Appendix, Notes S1 and S2).

Fig. 1A shows a typical STM topographic image of MoS2 flake on h-BN for a scan area of 75 nm × 75 nm at liquid helium temperatures. The thickness of MoS2 flake was confirmed to be six layers using atomic force microscopy and verified subsequently by scanning tunneling spectroscopy (STS) (Fig. 1C). The dark topographic contrast in the STM image (Fig. 1A) and lattice discontinuity are hallmarks of single sulfur vacancies (18), with a concentration of ∼1.81 ± 0.4 × 1012 cm−2. In the high-resolution image inserted in Fig. 1C, Center, a hexagonal lattice with a periodicity of 3.15 ± 0.5 Å is seen (SI Appendix, Fig. S3), corresponding to the atomic lattice of the top sulfur layer in the MoS2 crystal (19). The defects are experimentally observed at the center of sulfur lattice sites, excluding other defect types such as interstitials, Mo vacancy, or antisite defects (a Mo atom substituting a S2 column or vice versa) (20). STS, which probes the local density states, was applied to study the effect of such sulfur vacancy defects on the electronic properties. To better resolve the effect of sulfur vacancies, a lower set point (VS = −0.7 V, I = 1.3 nA) was used to collect the dI/dV spectrum from the vicinity of sulfur vacancy as well as far away from it (Fig. 1C). At regions where the MoS2 is pristine (Fig. 1C, Center), a series of oscillation peak/dip features were observed at the conduction band region spaced ∼98.0, ∼134.3, and ∼169.0 meV apart, in good agreement with the subband (C1, C2, and C3; Fig. 1C, Right) structure of six-layer MoS2 predicted by DFT calculations. Intrinsically, such subbands are associated with quantum confinement of electronic states. When the STM tip is located on top of a vacancy, however, a splitting (energy of 50.0 ± 5.0 meV) of these subbands (into C1a and C1b, C2a and C2b, and C3a and C3b) was clearly captured, which is similar to observations of band hybridization in Kondo insulators (21). A survey scan using X-ray photoelectron spectroscopy over the whole MoS2 flake shows no evidence of magnetic elemental impurities (SI Appendix, Fig. S4).

Fig. 1.

Fig. 1.

Electronic properties and band structure of six-layer MoS2 with sulfur vacancy. (A) Large-area STM image of the bare MoS2 on h-BN substrate (VS = 1.0 V, I = 0.7 nA). The dark topographic contrast, which shows the single sulfur vacancy features, can be found over the whole scanned area. (B) DFT calculation of six-layer MoS2. The zoomed-in image shows the conduction subbands (Cn) near the CBM at point of six-layer MoS2. (C) dI/dV spectrum (VS = −0.7 V, I = 1.3 nA) at the vicinity of a sulfur vacancy (Left side) (scale bar, 0.6 nm) and pristine region (Right) (scale bar, 0.8 nm) of MoS2 near CBM. Conduction subbands C1, C2, and C3 can be revealed from the resonance peaks in STS curve of pristine MoS2. However, each of those subbands split into two bands (Cna and Cnb) due to band hybridization in the vicinity of a sulfur vacancy. (D) Top and side views of spin density associated with single sulfur vacancy in monolayer MoS2. (Purple larger spheres represent molybdenum atoms, and smaller yellow spheres represent sulfur atoms.) The spin density associated with VSul−1, with a significant component localized at the three exposed Mo atoms. (E) Electronic DOS of the sulfur vacancy in neutral (VSul0) and −1 charged state (VSul−1) (for intuitive understanding, the VBMs for the two cases were aligned).The charging of the vacancy leads to the lifting of the degeneracy of defective level indicated by the splitting of curves of spin-up (dashed line) and spin-down (solid line) states for VSul−1.

In order to investigate the local effect of vacancies on electronic structure in MoS2, first-principle calculations were conducted on monolayer MoS2 (more calculation details in SI Appendix, Note S4). The removal of one sulfur atom in the MoS2 sheet creates a single sulfur vacancy (VSul), accompanied by dangling states and exposure of Mo atoms in the vacancy core. Single neutral VSul has two states: a fully occupied singlet A state and an empty doubly degenerate E state (22). When carriers are injected into the MoS2, the VSul state becomes negatively charged, and the VSul−1 is spin-polarized, accompanied by a spin moment of μ = 1/2. The spin density associated with VSul−1, with a significant component localized at the three exposed Mo atoms, is shown in Fig. 1 D and E. Charging via gating alters the local magnetic moment and state-splitting of the VSul, which in turn, tunes the scattering of charge carriers, especially at low temperatures (23).

Electrical Performance Due to Band Hybridization.

To investigate the influence of vacancy-induced band hybridization, temperature-dependent transport measurements were carried out. Fig. 2A shows the top and section view of the MoS2/h-BN heterostructure, and a representative optical image is shown in Fig. 2B. To be consistent, a MoS2 flake with thickness 4.2 ± 0.3 nm was selected (six layers; Fig. 1C) (more transport results of various thicknesses are provided in SI Appendix, Note S15). Four-probe measurements were performed to exclude the effects of electrical contact resistance, and the nanofabricated heater allows thermoelectric measurements. The linear IsdVsd curves collected at room temperature (Fig. 2C) indicate ohmic contacts between metal electrode and MoS2 flake (SI Appendix, Fig. S6). The carrier concentration (n) in MoS2 is modulated by applying a back-gate voltage (Vg). Fig. 2D shows an increasing conductance (G) with Vg for the MoS2/h-BN sample within the measured temperature (T) range, indicating typical n-type FET behavior. From 300 to 100 K, a clear crossing point (at Vg20V) appeared, indicating that MoS2 undergoes a routine metal-to-insulator transition (MIT), where the conductance changes from decreasing with temperature to increasing with temperature (2426). At T < 100 K, the conductance drops anomalously as T decreases, in contrast with the trend exhibited by MoS2 on Si/SiO2 devices. This can be explained by the strong hybridization between localized impurities states and conduction electrons, which will be further discussed later.

Fig. 2.

Fig. 2.

Structural and electronic properties of MoS2/h-BN heterostructure. (A) Schematic diagram of the device. Electrode 1 acts as an electrical heater. Electrodes 2 and 5 act as a current source for four-probe electrical measurements, while electrodes 3 and 4 act as thermometers. (Bottom) The section view of the heterostructure. (B) Optical image of a complete device. The dotted red dashed box outlines the MoS2 flake. (Scale bar, 10 μm.) (Bottom) AFM scan gives the thickness of MoS2 flakes as 4.2 ± 0.3 nm. (C) Four-probe Isd-Vsd curves at different Vg values at 300 K. (D) Four-probe FET characterization at different temperatures. Clear crossover can be found at Vg20V for T > 100 K, which is defined as the MIT. For Vg>VMIT, MoS2 shows metallic behavior, and conductance decreases with increasing T. For Vg<VMIT, conductance increases with increasing T, which is a typical insulating behavior. When T < 100 K, the conductance (in the metallic region) drops anomalously as T decreases.

In order to better understand the electronic transport in MoS2/h-BN, temperature-dependent four-probe sheet conductivity (σs = G·L/W, where L and W are the length and width of sample channel, respectively) was plotted as a function of Vg for both MoS2/h-BN (Fig. 3A) and MoS2/SiO2 (Fig. 3B) devices for comparison. In the high temperature range (T > 100 K), and close to σse2/h, the metal-insulator transition driven by electron correlations can be observed clearly in both devices. For MoS2/SiO2 device, such behavior exists over the temperature range from 300 to 20 K. However, for the MoS2/h-BN device, instead of saturating at a residual value as T→0, a conductance peak (temperature at this critical point is defined as Tmax; red dashed line in Fig. 3A) in the metallic region is observed. This observed anomalous resistance minimum at low temperature is another signature of a Kondo insulator behavior.

Fig. 3.

Fig. 3.

Carrier density and temperature-dependent properties of defective MoS2 on h-BN substrate. (A) Four-probe sheet conductivity of MoS2/h-BN devices as a function of T and Vg. MIT can be observed when σse2/h. The anomalous peaks Tmax at low temperatures are marked out by the red dashed line. (B) Four-probe sheet conductivity of MoS2/SiO2 devices as a function of T and Vg. (C) Temperature-dependent field effect mobility. The solid line shows the phonon-limited power law μphTγ. γ=1.7 and μ200cm2V1s1 for MOS2/SiO2 sample. For MoS2 on h-BN, γ=2.1 and μ reaches 405cm2V1s1. For T < 100 K, dμFE/dT>0 is observed for MoS2/h-BN sample indicating an anomalous scattering mechanism. (D) Temperature dependence of four-probe resistance at Vg = 70, 50, and 30 V for MoS2/h-BN device, with the resistance minima at 70, 89, and 135 K, respectively. (E) Gate tuning of the Kondo temperature Tk. (F) Normalized Kondo resistance versus normalized temperature for the data from Vg = 10 V to Vg = 70 V. The red dashed line describes the universal Kondo behavior from numerical renormalization group calculations (35).

This effect is suppressed in the MoS2/SiO2 device due to the imperfect interface between MoS2 and SiO2 but clearly detectable for the six-layer MoS2/h-BN devices (more data on six-layer MoS2/SiO2 and MoS2/h-BN as well as monolayer and bilayer MoS2/h-BN are in SI Appendix, Note S15). At low temperatures, the hybridization of the conduction band with the localized impurity state manifests in a large band splitting. Kondo resonant scattering can dominate charge transport in this temperature range and give rise to the anomalous component of the resistance. Such a picture is also complemented by the change in four-probe mobility (μ) with temperature (Fig. 3C). For MoS2 on SiO2, μ is mainly limited by optical phonon scattering when T > 100 K, and its temperature dependence follows a power law of μTγ (27). Once T < 100 K, most optical phonons are deactivated, and saturation is observed, resulting in dμ/dT0 as T→ 0 (27). On the contrary, MoS2/h-BN device deviates from this trend where its mobility decreases as T < 100 K and shows dμ/dT>0.

Fig. 3D shows the four-probe resistance plots of MoS2/h-BN device at three different back gates. For each curve, a clear resistance minimum can be observed at 70, 89, and 135 K. For a better understanding of this data, apart from 2D Block–Gruneisen resistance (T4) (28, 29) and electron–phonon high temperature resistance (T), we employ a Kondo resistance model [RK(TK,T), full expression in SI Appendix, Note S11]:

R=AT4+BT+RK(TK,T)+R0, [1]

where R0 is the temperature-independent term arising from a residual zero temperature resistance (Fig. 3D). In the Kondo picture, close to a characteristic temperature (Kondo temperature, TK), magnetic impurities quantum-mechanically exchange spin with conduction electrons of many wave vectors (momenta) (30), thereby creating a resonant scattering state at the Fermi level with a width kBTK [kBTKΔexp(1D(EF)J0), where D(EF) is the electron DOS at the Fermi level, Δ is the bandwidth, and J0 is a constant representing a Kondo scattering exchange energy] (31, 32). As a result, a dip in the resistance vs. temperature curve appears. Fig. 3E summarizes the TK variation as a function of Vg for MoS2/h-BN sample. At Vg > VMIT, TK remains nearly unchanged because of the constant 2D DOS. When Vg < VMIT, the Fermi level shifts toward the gap region; therefore, TK starts to increase with lower doping, similar to the behavior observed in heavily doped fermion bulk systems such as CePd3 (33). We also scale the R(T) curves at each Vg and observe a universal Kondo behavior (6, 34), in which the normalized Kondo resistance Rk/Rk0 vs. T/Tk at all gate voltages collapses onto a single universal Kondo behavior curve (Fig. 3F). This is in accordance with numerical renormalization group theory (35), which attests to the Kondo-driven transport in MoS2/h-BN devices.

Thermoelectric Performance Driven by Band Hybridization.

To investigate further if the thermoelectric values show a similar effect, a DC current is introduced via a nanofabricated heater to create a temperature gradient along the devices (Fig. 2A). Fig. 4A shows the Seebeck coefficient (S) as a function of temperature at Vg = 70, 50, and 30 V for both MoS2/h-BN and SiO2 devices. The absolute |S| decreases with increasing carrier concentration in the MoS2 channel (details in SI Appendix, Note S13 and Fig. S19), comparable to the reported trends in earlier work (36, 37). From the temperature-dependent behavior of S for both samples, a clear dip located around 100 to 150 K can be observed which is very close to the thermal conductivity peak of MoS2 (38, 39), indicating the contribution of acoustic phonons in MoS2 to the Seebeck coefficient via the phonon drag effect, Sph (40). Both MoS2/SiO2 and MoS2/h-BN samples show similar values of negative S and behavior for T > 120 K. However, when T < 120 K, an anomalous sign change (from negative to positive) in S is observed in MoS2/h-BN sample even though electrons remain as the dominant charge carriers. The positive S reaches a magnitude as high as 2 mV/K for Vg = 30 V at 60 K, which is almost one order of magnitude enhancement compared to MoS2/SiO2 sample (−270 μV/K) under the same experimental conditions. This is a remarkably large value for a material in the metallic state. To better illustrate the sign change of S, a 2D map of S as a function of both T and Vg is plotted in Fig. 4B. Clearly, S changes from negative to positive values at all gate voltages. It is also found that the critical temperature (defined as the temperature at which S = 0) and the maximum positive S (highest intensity; red in Fig. 4B) increases as Vg decreases, suggesting that the Seebeck coefficient can be tuned by electrostatic gating. Importantly, this critical temperature across all temperatures for S = 0 is identical to the conduction maximum line Tmax in Fig. 3A, further evidencing that this anomalous thermoelectric transport has the same origins as the electronic transport.

Fig. 4.

Fig. 4.

Thermoelectric transport and Seebeck coefficient measurement of defective MoS2 on h-BN substrate. (A) Comparison of S vs. T of MoS2/SiO2 and MoS2/h-BN device at Vg = 70 V (circle), 50 V(square), and 30 V (diamond) (from left to right; black arrow). The solid lines guide the eye. MoS2/SiO2 sample shows negative values due to the diffusive type n-type charge carriers. For MoS2/h-BN sample, at low temperatures, an anomalous sign change in S occurs where the majority carriers are electrons. (B) Color contour plot of S values versus Vg and T for MoS2/h-BN device. The black dotted line indicates the point with S ∼ 0 mV/K. (C) Total S values of MoS2/SiO2 sample at n = 2 × 1012 cm−2 and its respective contributions from the energy-dependent diffusion SN and phonon-drag Sph.

Since the transfer curve shows unambiguous n-type behavior and the quasiparticle bandgap of six-layer MoS2 is relatively large (41), the valence band states of MoS2 are inaccessible and hence cannot result in a positive S, especially at low temperatures. When electrons are the majority carriers, the single-particle Boltzmann transport equation (BTE) can be employed to describe thermoelectric transport and is defined as SBTE. In the degenerate state (highly conducting on-state), the BTE reduces to the Mott formula described below (42):

SBTE=π23kB2T|e|[lnτE+(lngE+lnv2E)]|E=EF=Sτ+SN, [2]

with Sτ=π23kB2T|e|lnτE|E=EF and SN=π23kB2T|e|(lngE+lnv2E)|E=EF.

Here kB, τ, g, and v indicate the Boltzmann constant, relaxation time, DOS, and group velocity, respectively. Sτ and SN are the contributions arising from the energy-dependent scattering and diffusion. The sign of SN only depends on the charge carrier type; thus, an n-type MoS2 will only exhibit a negative SN. As such, the Seebeck coefficient (S) comprises the sum of SBTE and Sph. For the MoS2/SiO2 sample, the low-temperature thermoelectric transport is dominated by acoustic phonon scattering where the acoustic phonon scattering relaxation time (τ) is energy independent (43) in the 2D limit. Hence, the SBTE of our MoS2/SiO2 is dominated by negative SN and phonon drag Sph. Here SN can be calculated by solving the linearized BTE under a relaxation time approximation (detailed calculations are provided in SI Appendix, Note S12) (36, 44). Consequently, the overall phonon drag Seebeck coefficient (Sph=StotalSN) can be estimated. Fig. 4C shows the measured Seebeck coefficient (S) as a function of temperature of MoS2/SiO2 device at a carrier concentration of n = 2 × 1012 cm−2. The phonon drag contribution Sph shows clear peaks around 100 K as expected where the largest concentration of heat-carrying phonons (acoustic phonons) can interact with carriers (45). Hence, the peak position is not changing across different back gate voltages (SI Appendix, Fig. S18). At temperatures below 100 K, Sph diverges from unity. A first-order quantitative estimation of phonon drag S is given by Sph=βvsls/μT (4648), where vs is the velocity of sound and ls is the phonon mean free path. β(0,1) indicates the relative contribution of electron–phonon interaction to electron mobility. The phonon mean free path can be evaluated using the Debye equation κ=13cvsls, yielding ls5.5to14nm which is consistent with the calculated values from literature (49). We remark that this estimated ls is an effective lower bound involving all phonons. The resulting estimate of Sph102μV/K is also comparable to our results (details in SI Appendix, Note S12).

However, an anomalous sign change in S is observed for the MoS2/h-BN sample accompanied by large positive values in the low-temperature region. Typically, a sign change in S occurs when the type of charge carrier changes, like in ambipolar materials [graphene (50) and black phosphorus (51)] or in complicated Fermi surface nesting systems such as charge density wave phase transition materials [NbSe3 (52) and TaSe3 (53)]. In addition, it has been observed that longer phonon mean free paths such as those in h-BN may contribute to an anomalously large thermoelectric behavior through the phonon drag effect (45). In such cases, the phonon drag Sph peak should occur at higher temperatures due to the higher Debye temperature of h-BN. On the contrary, the observed peak in Stotal is located at lower temperatures in our experiments (Fig. 4A). Moreover, our experiment further shows that the sign change in Stotal is strongly dependent on the modulation of electron carrier concentration (gate effect) and vacancy concentration in MoS2 (controlled via annealing). Hence, the anomalous thermoelectric transport can be mainly attributed to the strong correlation transport induced by the hybridized band structure, as shown in our STS results, in addition to phonon drag in MoS2.

In fact, the vacancy-induced band hybridization and Kondo effect have been studied in 2D materials like graphene (6, 54, 55). In such a case, the scattering is highly energy-dependent, and the mobility is known to be proportional to 1/τ (56), which is opposite to the conventional one-band system where μτ. Approximating 1/E as 1/(kBT) (57) in Sτ, lnτTdμdT could be derived to describe the relationship between Seebeck coefficient and mobility. From Fig. 5A, we can see that the measured S is positive in the region where dμdT<0, in good agreement with the Kondo physics model. In this model, the interaction between electrons at Fermi energy and hybridized band strongly influence the nonequilibrium energy spectrum, thereby generating a Kondo resonance that reverses the sign of Seebeck coefficient. This is further corroborated by the peak position of the S values occurring near the Kondo temperature.

Fig. 5.

Fig. 5.

Thermolectric performance of MoS2/h-BN heterostructure. (A) Seebeck coefficient and mobility as function of temperature at Vg = 70 V for MoS2/h-BN device. (B) Total S values of MoS2/h-BN sample at n = 2 × 1012 cm−2 are contributions from the energy-dependent diffusive part SN, phonon-drag part Sph, and the Kondo scattering part Sτ. At a fixed n, the total S first exhibits a diffusive negative value at high temperatures from conducting electrons described by SN+Sph. As the temperature decreases, the conventional diffusive contribution is weakened, and the Kondo scattering term Sτ starts to dominate and shows large positive values. As temperature is decreased further, all of the physical interactions start to freeze, and the total S goes back to zero as expected (64). (C) Comparison of PF of MoS2/h-BN and MoS2/SiO2 sample as a function of temperature at different gate voltages. Additional band hybridization-induced peaks as high as 50 mW m−1 K−2 can be observed at 30 K ∼ 50 K for MoS2/h-BN sample. (D) The PF value of our MoS2/h-BN heterostructure shows superior thermoelectric performance compared to even the highest reported values in other 2D materials at room temperature. [G/SiO2 and h-BN (50), MoS2 (36, 37), WSe2 (60), TiS2 (61, 62), and BP (51).]

To understand how the S is related to resonant scattering (Sτ), the two-band approach originating from the Hirst model is adopted (58):

Sτ=ATε0ε02+[TΓexp(TΓ/T)]2. [3]

Here the conduction electrons are scattered by an extraband induced by the hybridization. This band is described using a Lorentzian form with a width Γ0 located at ε0 with respect to the Fermi level. Γ0 varies with temperature as Γ0=TΓexp(TΓ/T). TΓ is a temperature-dependent parameter and represents the quasi-elastic linewidth of the Kondo resonance. ε0 can be positive or negative depending on whether the position of the extraband is above or below the Fermi level, respectively. Consequently, the overall Seebeck coefficient (Stotal=SN+Sτ+Sph) can be estimated. Fig. 5B shows that the measured Seebeck coefficient as a function of temperature of MoS2/h-BN device for n = 2 × 1012 cm−2 is well captured by this scattering model (more details in SI Appendix, Fig. S18). At higher gate voltages, the Fermi level of the MoS2/h-BN sample is raised toward the conduction band, causing resonant scattering to be present only at lower temperatures (Fig. 4B). Due to the strong interaction of electrons with the hybridized band through such resonant scattering, the larger S values in the on state of MoS2 lead to an unusually strong enhancement in the thermoelectric PF (S2σ, where σ is calculated from the four-probe sheet conductivity σs by considering the thickness of MoS2) at an equivalent high bulk carrier concentration of ∼4.7 × 1018 cm−3. This is distinct from the S enhanced only due to the phonon drag effect, which occurs at a much lower doping concentration (∼1015 cm−3) without a sign change (59), thereby translating to a much lower PF. Fig. 5C shows the PF comparison between MoS2/h-BN and MoS2/SiO2 samples as a function of temperature. When T > 120 K, the PF of the MoS2/h-BN sample is comparable with that of the MoS2/SiO2 sample. However, with decreasing temperature, the PF of the MoS2/h-BN sample drops significantly due to the positive Sτ contribution which cancels the contribution from traditional negative SN. When temperature further decreases, Kondo scattering starts to dominate the thermoelectric transport. A large enhancement in PF up to two orders of magnitude is observed. Such a high PF value of 50 mW m−1 K−2 (T = 30 to 50 K), originating from band hybridization-enhanced S, is two to four orders of magnitude larger than the PF of other 2D materials (36, 37, 50, 51, 6062) in the same temperature regime and exhibits the highest PF in 2D materials at all reported temperatures (Fig. 5D).

Conclusion

In summary, we discover that band hybridization due to sulfur vacancies exerts a strong influence on the thermoelectric properties of few-layer MoS2 supported on h-BN substrate, whereas these effects are suppressed on SiO2 substrate. The electric and thermoelectric transport in few-layer MoS2 on h-BN match well with a strongly correlated Kondo-like behavior, leading to a large anomalous positive Seebeck coefficient of 2 mV/K in n-type MoS2 in the on state. Importantly, our work demonstrates that such an effect can be electrostatically tuned to manipulate the Seebeck coefficient, leading to two orders of magnitude enhancement in thermoelectric PF. The ability to exhibit both negative Seebeck (diffusive) and positive Seebeck (band hybridized) coefficients in n-type MoS2 suggests that a singly doped material could be used to fabricate thermoelectric devices, thus bypassing the need for multiple materials in thermoelectric devices. This work also points to the possibility of dopant engineering to make Kondo lattices (63) for nanothermoelectric devices.

Methods

MoS2 and h-BN flakes were separately mechanically exfoliated onto SiO2/Si substrates. The flakes are then examined by observing the contrast an optical microscope (in both bright-field and dark-field mode) and atomic force microscopy (AFM) to ensure that the surface, geometry, and thickness are suitable for subsequent measurements. MoS2 flake was transferred onto h-BN flakes using standard wet transfer techniques by poly(methyl methacrylate) (PMMA). After that, electrodes were defined by standard electron beam lithography (EBL), followed by deposition of titanium/gold (Ti/Au) with thickness of 3 nm for Ti and 70 nm for Au. Ti was chosen owing to its low work function, comparable to the electron affinity of MoS2 to form ohmic contacts. The STM measurements were performed using Omicron ultrahigh-vacuum LT-STM (<10−10 torr). Before the measurement, the tungsten tip was calibrated by Au (111) crystal. During STM measurements, the sample is grounded through the deposited Ti/Au electrode on top of the MoS2 flake. The gate electrode is lead out by an aluminum wire. After wire bonding, the MoS2/h-BN sample was annealed at 220 °C in the preparation chamber (<10−10 torr) to remove absorbents and form a better contact. STM and STS data were collected at liquid helium temperatures. Constant-current mode was employed to capture STM images, and external lock-in technique was employed to record STS data. Four-probe measurements were employed to eliminate contact resistance and extract a precise electrical conductivity of MoS2 flakes for different gate voltages. The temperature-dependent mobility is extracted from transfer characteristics, via μFE=1CoxdσdVg, where Cox=12×109F/cm2 is the capacitance for 285-nm-thick SiO2. The sheet conductivity (σs=LWG) is converted from G by considering the length (L) and width (W) of the MoS2 flake, and electrical conductivity is calculated from σ=σsd by considering its thickness (d). To measure the Seebeck coefficient S = VTEP/ΔT, a temperature gradient is realized by applying a DC current bias (I) through a metal heater using a source meter (Keithley 6221); therefore, Joule heat and a temperature gradient along the sample are generated. The temperature gradient generates the thermoelectric voltage VTEP, which is measured by a voltmeter (Keithley 6430). As a series of DC current bias with increasing values are applied to the heater, the corresponding VTEP were recorded. The temperature gradient is measured through the four-probe resistance of thermometers.

Data and Materials Availability.

All data supporting the findings of this study are available within the article or SI Appendix.

Supplementary Material

Supplementary File

Acknowledgments

This research was supported by Agency for Science, Technology and Research (A*STAR) Pharos Funding from the Science and Engineering Research Council (SERC) (Grant 152 70 00015). K.P.L. acknowledges Ministry of Education (MOE) Tier 3 Grant “Two-dimensional crystal quantum exciton photonic,” Project MOE2018-T3-1-005. C.-W.Q. acknowledges the A*STAR SERC Pharos Grant 152 70 00014 with Project R-263-000-B91-305. A.H.C.N. acknowledges the National Research Foundation (NRF) Competitive Research Programme Award R-144-000-295-281. We also thank Professor Zheng Yi for the helpful discussion and comments.

Footnotes

The authors declare no competing interest.

This article is a PNAS Direct Submission.

This article contains supporting information online at https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.2007495117/-/DCSupplemental.

References

  • 1.Snyder G. J., Toberer E. S., Complex thermoelectric materials. Nat. Mater. 7, 105–114 (2008). [DOI] [PubMed] [Google Scholar]
  • 2.Wu J., Chen Y., Wu J., Hippalgaonkar K., Perspectives on thermoelectricity in layered and 2D materials. Adv. Electron. Mater. 14, 1800248 (2018). [Google Scholar]
  • 3.Chang C., et al. , 3D charge and 2D phonon transports leading to high out-of-plane ZT in n-type SnSe crystals. Science 360, 778–783 (2018). [DOI] [PubMed] [Google Scholar]
  • 4.Zhao L. D., et al. , Ultralow thermal conductivity and high thermoelectric figure of merit in SnSe crystals. Nature 508, 373–377 (2014). [DOI] [PubMed] [Google Scholar]
  • 5.Ghahari F., et al. , Enhanced thermoelectric power in graphene: Violation of the Mott relation by inelastic scattering. Phys. Rev. Lett. 116, 136802 (2016). [DOI] [PubMed] [Google Scholar]
  • 6.Chen J. H., Li L., Cullen W. G., Williams E. D., Fuhrer M. S., Tunable Kondo effect in graphene with defects. Nat. Phys. 7, 535–538 (2011). [Google Scholar]
  • 7.Demion A., Verga A. D., Nonlinear electric transport in graphene with magnetic disorder. Phys. Rev. B Condens. Matter Mater. Phys. 90, 1–9 (2014). [Google Scholar]
  • 8.Xiao D., Bin Liu G., Feng W., Xu X., Yao W., Coupled spin and valley physics in monolayers of MoS 2 and other group-VI dichalcogenides. Phys. Rev. Lett. 108, 1–5 (2012). [DOI] [PubMed] [Google Scholar]
  • 9.Wang Q. H., Kalantar-Zadeh K., Kis A., Coleman J. N., Strano M. S., Electronics and optoelectronics of two-dimensional transition metal dichalcogenides. Nat. Nanotechnol. 7, 699–712 (2012). [DOI] [PubMed] [Google Scholar]
  • 10.Zhu Z. Y., Cheng Y. C., Schwingenschlögl U., Giant spin-orbit-induced spin splitting in two-dimensional transition-metal dichalcogenide semiconductors. Phys. Rev. B Condens. Matter Mater. Phys. 84, 1–5 (2011). [Google Scholar]
  • 11.Komider K., González J. W., Fernández-Rossier J., Large spin splitting in the conduction band of transition metal dichalcogenide monolayers. Phys. Rev. B Condens. Matter Mater. Phys. 88, 1–7 (2013). [Google Scholar]
  • 12.Cai L., et al. , Vacancy-induced ferromagnetism of MoS2 nanosheets. J. Am. Chem. Soc. 137, 2622–2627 (2015). [DOI] [PubMed] [Google Scholar]
  • 13.Rodin A. S., Neto A. H. C., Localized magnetic states in two-dimensional semiconductors. Phys. Rev. B 97, 235428 (2018). [Google Scholar]
  • 14.Zhang J., et al. , Magnetic molybdenum disulfide nanosheet films. Nano Lett. 7, 2370–2376 (2007). [DOI] [PubMed] [Google Scholar]
  • 15.Pan H., Zhang Y. W., Tuning the electronic and magnetic properties of MoS 2 nanoribbons by strain engineering. J. Phys. Chem. C 116, 11752–11757 (2012). [Google Scholar]
  • 16.Han S. W., et al. , Electron beam-formed ferromagnetic defects on MoS2 surface along 1 T phase transition. Sci. Rep. 6, 38730 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Mak K. F., He K., Shan J., Heinz T. F., Control of valley polarization in monolayer MoS2 by optical helicity. Nat. Nanotechnol. 7, 494–498 (2012). [DOI] [PubMed] [Google Scholar]
  • 18.Pető J., et al. , Spontaneous doping of the basal plane of MoS2 single layers through oxygen substitution under ambient conditions. Nat. Chem. 10, 1246–1251 (2018). [DOI] [PubMed] [Google Scholar]
  • 19.Huang Y. L., et al. , Bandgap tunability at single-layer molybdenum disulphide grain boundaries. Nat. Commun. 6, 6298 (2015). [DOI] [PubMed] [Google Scholar]
  • 20.Hong J., et al. , Exploring atomic defects in molybdenum disulphide monolayers. Nat. Commun. 6, 6293 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Neupane M., et al. , Surface electronic structure of the topological Kondo-insulator candidate correlated electron system SmB6. Nat. Commun. 4, 2991 (2013). [DOI] [PubMed] [Google Scholar]
  • 22.Cai Y., Zhou H., Zhang G., Zhang Y., Modulating carrier density and transport properties of MoS2 by organic molecular doping and defect engineering. Chem. Mater. 28, 8611–8621 (2016). [Google Scholar]
  • 23.Nair R. R., et al. , Spin-half paramagnetism in graphene induced by point defects. Nat. Phys. 8, 199–202 (2012). [Google Scholar]
  • 24.Baugher B. W. H., Churchill H. O. H., Yang Y., Jarillo-Herrero P., Intrinsic electronic transport properties of high-quality monolayer and bilayer MoS2. Nano Lett. 13, 4212–4216 (2013). [DOI] [PubMed] [Google Scholar]
  • 25.Radisavljevic B., Kis A., Mobility engineering and a metal-insulator transition in monolayer MoS2. Nat. Mater. 12, 815–820 (2013). [DOI] [PubMed] [Google Scholar]
  • 26.Schmidt H., et al. , Transport properties of monolayer MoS2 grown by chemical vapor deposition. Nano Lett. 14, 1909–1913 (2014). [DOI] [PubMed] [Google Scholar]
  • 27.Cui X., et al. , Multi-terminal transport measurements of MoS2 using a van der Waals heterostructure device platform. Nat. Nanotechnol. 10, 534–540 (2015). [DOI] [PubMed] [Google Scholar]
  • 28.Peng B., et al. , Thermal conductivity of monolayer MoS2, MoSe2, and WS2: Interplay of mass effect, interatomic bonding and anharmonicity. RSC Advances 6, 5767–5773 (2016). [Google Scholar]
  • 29.Efetov D. K., Kim P., Controlling electron-phonon interactions in graphene at ultrahigh carrier densities. Phys. Rev. Lett. 105, 256805 (2010). [DOI] [PubMed] [Google Scholar]
  • 30.Kouwenhoven L., Glazman L., Revival of the Kondo effect. Phys. World 14, 33–38 (2001). [Google Scholar]
  • 31.Grewe N., Steglich F., “Heavy fermions” in Handbook on the Physics and Chemistry of Rare Earths, Gschneidner K. A. Jr., Eyring L., Eds. (Elsevier Science Publishers, Amsterdam, 1991), pp. 343–474. [Google Scholar]
  • 32.Phillips P., Advanced solid State Physics (Westview Press, 2003). [Google Scholar]
  • 33.Gumeniuk R., Schnelle W., Kvashnina K. O., Leithe-Jasper A., Kondo effect and thermoelectric transport in CePd3Be x. J. Phys. Condens. Matter 28, 165603 (2016). [DOI] [PubMed] [Google Scholar]
  • 34.Goldhaber-Gordon D., et al. , From the Kondo regime to the mixed-valence regime in a single-electron transistor. Phys. Rev. Lett. 81, 5225–5228 (1998). [Google Scholar]
  • 35.Costi T. A., Hewson A. C., Zlatic V., Transport coefficients of the Anderson model via the numerical renormalization group. J. Phys. Condens. Matter 6, 2519 (1994). [Google Scholar]
  • 36.Hippalgaonkar K., et al. , High thermoelectric power factor in two-dimensional crystals of MoS2. Phys. Rev. B 95, 115407 (2017). [Google Scholar]
  • 37.Kayyalha M., Maassen J., Lundstrom M., Shi L., Chen Y. P., Gate-tunable and thickness-dependent electronic and thermoelectric transport in few-layer MoS2. J. Appl. Phys. 120, 134305 (2016). [Google Scholar]
  • 38.Yarali M., et al. , Effects of defects on the temperature-dependent thermal conductivity of suspended monolayer molybdenum disulfide grown by chemical vapor deposition. Adv. Funct. Mater. 27, 1704357 (2017). [Google Scholar]
  • 39.Aiyiti A., et al. , Thermal conductivity of suspended few-layer MoS2. Nanoscale 10, 2727–2734 (2018). [DOI] [PubMed] [Google Scholar]
  • 40.Takahashi H., et al. , Colossal Seebeck effect enhanced by quasi-ballistic phonons dragging massive electrons in FeSb2. Nat. Commun. 7, 12732 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Cheiwchanchamnangij T., Lambrecht W. R. L., Quasiparticle band structure calculation of monolayer, bilayer, and bulk MoS2. Phys. Rev. B Condens. Matter Mater. Phys. 85, 205302 (2012). [Google Scholar]
  • 42.Cutler M., Mott N. F., Observation of Anderson localization in an electron gas. Phys. Rev. 181, 1336–1340 (1969). [Google Scholar]
  • 43.Lundstrom M., Fundamentals of Carrier Transport (Cambridge University Press, ed. 2, 2000). [Google Scholar]
  • 44.Ng H. K., Chi D., Hippalgaonkar K., Effect of dimensionality on thermoelectric powerfactor of molybdenum disulfide. J. Appl. Phys. 121, 204303 (2017). [Google Scholar]
  • 45.Wang G., Endicott L., Chi H., Lošt’ák P., Uher C., Tuning the temperature domain of phonon drag in thin films by the choice of substrate. Phys. Rev. Lett. 111, 046803 (2013). [DOI] [PubMed] [Google Scholar]
  • 46.Takahashi H., Okazaki R., Yasui Y., Terasaki I., Low-temperature magnetotransport of the narrow-gap semiconductor FeSb2. Phys. Rev. B Condens. Matter Mater. Phys. 84, 205215 (2011). [Google Scholar]
  • 47.Sales B. C., et al. , Transport, thermal, and magnetic properties of the narrow-gap semiconductor CrSb2. Phys. Rev. B Condens. Matter Mater. Phys. 86, 235136 (2012). [Google Scholar]
  • 48.Herring C., Theory of the thermoelectric power of semiconductors. Phys. Rev. 96, 1163–1187 (1954). [Google Scholar]
  • 49.Cai Y., Lan J., Zhang G., Zhang Y., Lattice vibrational modes and phonon thermal conductivity of monolayer MoS2. Phys. Rev. B Condens. Matter Mater. Phys. 89, 035438 (2014). [Google Scholar]
  • 50.Duan J., et al. , High thermoelectricpower factor in graphene/hBN devices. Proc. Natl. Acad. Sci. U.S.A. 113, 14272–14276 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Saito Y., et al. , Gate-tuned thermoelectric power in black phosphorus. Nano Lett. 16, 4819–4824 (2016). [DOI] [PubMed] [Google Scholar]
  • 52.Yang L., et al. , Distinct signatures of electron-phonon coupling observed in the lattice thermal conductivity of NbSe3 nanowires. Nano Lett. 19, 415–421 (2019). [DOI] [PubMed] [Google Scholar]
  • 53.Dominko D., Optical phonon drag and variable range hopping mechanisms of thermoelectric power generation in charge density wave system o-TaS 3. arXiv:1408.6190 (26 August 2014).
  • 54.Mitchell A. K., Fritz L., Kondo effect with diverging hybridization: Possible realization in graphene with vacancies. Phys. Rev. B Condens. Matter Mater. Phys. 88, 075104 (2013). [Google Scholar]
  • 55.Jiang Y., et al. , Inducing Kondo screening of vacancy magnetic moments in graphene with gating and local curvature. Nat. Commun. 9, 2349 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Sun P., Steglich F., Nernst effect: Evidence of local Kondo scattering in heavy fermions. Phys. Rev. Lett. 110, 216408 (2013). [DOI] [PubMed] [Google Scholar]
  • 57.Sun P., et al. , Large Seebeck effect by charge-mobility engineering. Nat. Commun. 6, 7475 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Hirst L. L., Electronic properties of 4f substances at configuration crossover. Phys. Rev. B 15, 1–9 (1977). [Google Scholar]
  • 59.Zhou J., et al. , Ab initio optimization of phonon drag effect for lower-temperature thermoelectric energy conversion. Proc. Natl. Acad. Sci. U.S.A. 112, 14777–14782 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Yoshida M., et al. , Gate-optimized thermoelectric power factor in ultrathin WSe2 single crystals. Nano Lett. 16, 2061–2065 (2016). [DOI] [PubMed] [Google Scholar]
  • 61.Wan C., et al. , Flexible n-type thermoelectric materials by organic intercalation of layered transition metal dichalcogenide TiS2. Nat. Mater. 14, 622–627 (2015). [DOI] [PubMed] [Google Scholar]
  • 62.Wan C., et al. , Flexible thermoelectric foil for wearable energy harvesting. Nano Energy 30, 840–845 (2016). [Google Scholar]
  • 63.Jiang Y., Zhang Y. N., Cao J. X., Wu R. Q., Ho W., Real-space imaging of Kondo screening in a two-dimensional O2 lattice. Science 333, 324–328 (2011). [DOI] [PubMed] [Google Scholar]
  • 64.Gegenwart P., Si Q., Steglich F., Quantum criticality in heavy-fermion metals. Nat. Phys. 4, 186–197 (2008). [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary File

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES