This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Scaling theory of charge transport and thermoelectric response in disordered 2D electron systems: From weak to strong localization

Takahiro Yamamoto1,2    Hiroki Kaya1    Manaho Matsubara1,2    Hidetoshi Fukuyama2 1Department of Physics, Tokyo University of Science, Kagurazaka 1-3 Shinjuku, Tokyo 162-8601, Japan 2RIST, Tokyo University of Science, Kagurazaka 1-3 Shinjuku, Tokyo 162-8601, Japan
Abstract

We develop a new theoretical scheme for charge transport and thermoelectric response in two-dimensional disordered systems exhibiting crossover from weak localization (WL) to strong localization (SL). The scheme is based on the scaling theory for Anderson localization combined with the Kubo–Luttinger theory. Key aspects of the scheme include introducing a unified β\beta function that seamlessly connects the WL and SL regimes, as well as describing the temperature (TT) dependence of the conductance from high to low TT regions on the basis of the dephasing length. We found that the Seebeck coefficient, SS, behaves as STS\propto T in the WL limit and as ST1pS\propto T^{1-p} (p<1p<1) in the SL limit, both with possible logarithmic corrections. The scheme is applied to analyze experimental data for thin films of the p-type organic semiconductor poly[2,5-bis(3-alkylthiophen-2-yl)thieno(3,2-b)thiophene] (PBTTT).

pacs:
73.50.Lw, 73.20.Fz, 68.35.bm

I Introduction

The development of high-performance thermoelectric (TE) materials is a focus of extensive research, particularly in relation to sustainable energy production. Since Hicks and Dresselhaus suggested in 1993 that a substantial increase in TE performance could be achieved by using low-dimensional semiconductors [1], numerous low-dimensional TE materials have been discovered. Recently, poly[2,5-bis(3-alkylthiophen-2-yl)thieno(3,2-b)thiophene] (PBTTT) thin films, composed of π\pi-stacked PBTTT polymers, have garnered significant attention as a new class of two-dimensional (2D) p-type organic semiconductors, both experimentally [2, 3, 4, 10, 11, 12, 9, 14, 15, 6, 5, 8, 7, 13] and theoretically [16, 17], due to their high TE performance. Because of the inherent structural disorder of polymers, understanding and controlling their charge transport and TE response pose important challenges in condensed matter physics and materials engineering.

Interesting experimental results on charge transport in p-type PBTTT thin film were reported by several groups [10, 12, 7, 13, 15]. Recently, Ito et al. measured the temperature (TT) dependence of electrical conductivity, σ(T)\sigma(T), for various hole densities controlled by an electrochemical transistor (ECT) [15]. They found that σ(T)\sigma(T) exhibits a logarithmic TT dependence, characteristic of weak localization (WL), in the range 200–100K. As TT decreases from 100 to 30K, σ(T)\sigma(T) deviates from this logarithmic behavior, which was attributed to a precursor to the crossover from WL to strong localization (SL), characterized by an exponential TT dependence of the conductivity. Similarly, experimental data on σ(T)\sigma(T) for similar hole density achieved by chemical doping have been reported by Watanabe et al. [9], who analyzed the data under the assumption of variable range hopping (VRH)-type exponential TT dependence.

In addition to electrical conductivity, the Seebeck coefficient, S(T)S(T), has also been measured in a wide TT range by Ito et al. [15] and Watanabe et al. [9]. Despite the differences in carrier-doping methods, both groups conclude that S(T)S(T) is proportional to TT at high carrier densities, apparently consistent with the Mott formula for metals [18], whereas S(T)S(T) deviates from TT-linear behavior at low carrier densities. A coherent understanding of this behavior has not yet been achieved.

The purpose of the present study is to theoretically elucidate, on the basis of the scaling theory of Anderson localization, the TT dependence of σ(T)\sigma(T) and S(T)S(T) in the WL-SL crossover regime in 2D systems. This approach aims to provide a unified understanding of various experimental data obtained through different carrier-doping methods [15, 9], for which a consensus on the physical interpretation has not been attained.

II Theoretical scheme

II.1 Scaling theory at finite temperature

The scaling theory for the Anderson localization [19], i.e., Abraham–Anderson–Licciardello–Ramakrishnan (AALR) theory [20], indicates that the conductance in a disordered system at T=0T=0 follows a differential equation:

dlngdlnL=β(g),\displaystyle\dfrac{d\ln g}{d\ln L}=\beta(g), (1)

where gg is the dimensionless conductance scaled by the universal conductance e2/he^{2}/h (ee: elementary charge, hh: Planck’s constant), LL is the length of the system, and β(g)\beta(g) is assumed to be a smoothly and monotonically increasing function with respect to gg.

In the large-gg limit, gg obeys the classical Ohm’s law as g=σLd2g=\sigma L^{d-2}, where σ\sigma is conductivity and dd is the spatial dimension of the system. From Eq. (1), we see that the β\beta function in this limit is given by β(g)=d2\beta(g)=d-2; the asymptotic behavior of β(g)\beta(g) is therefore expressed as β(g)d2α/g\beta(g)\approx d-2-\alpha/g, where the expansion coefficient α\alpha is a positive constant on the order of unity. On the other hand, in the small-gg limit, the electronic states are localized because of disorder and gg decreases exponentially with LL as geL/ξg\sim e^{-L/\xi}, where ξ\xi is the localization length, resulting in β(g)lng\beta(g)\approx\ln g in this limit. Thus, in the two limits, the β\beta function is given by

β(g){lng(g0)d2αg(g)\displaystyle\beta(g)\approx\left\{\begin{array}[]{ll}\ln g&(g\to 0)\\ d-2-\dfrac{\alpha}{g}&(g\to\infty)\\ \end{array}\right. (4)

for dd-dimensional systems. In three-dimensional (3D) systems (d=3d=3), where conducting and localized states are sharply separated across the mobility edge, different theoretical schemes are needed for each state. By contrast, the 2D systems (d=2d=2) of present interest exhibit g(L)0g(L)\to 0 in the limit of LL; that is, the electronic states are always localized. If LL is finite and varied, g(L)g(L) is expected to crossover from metallic to insulating behaviors as LL is increased. Actually, this expectation has been confirmed through studies on the frequency (ω\omega) dependences of conductance for L=L=\infty at T=0T=0 by Vollhardt and Wolfle [21] and by Kawabata [22] using the self-consistent (SC) theory. Notably, the SC theory is governed by a single parameter, λ=εFτ\lambda=\frac{\hbar}{\varepsilon_{\rm F}\tau} (where εF\varepsilon_{\rm F} is the Fermi energy and τ\tau is the elastic scattering time), which characterizes the scattering strength in quantum transport, and turns out to be capable of describing not only metallic but also insulating states. Regarding the TT dependence of conductance, however, we need to develop a different theoretical scheme. Once at finite temperatures, g(T)g(T) can be non zero even for L=L=\infty and the TT dependence of g(T)g(T) exhibits a crossover from essentially metallic WL with lnT\ln T to SL with exponential TT dependences. The characteristic temperature of this crossover depends on the carrier density.

To address the WL–SL crossover in 2D systems, we first consider T=0T=0 and introduce the following β\beta function as a smoothly and monotonically increasing function that asymptotically behaves as in Eq. (4) in both limits of g0g\to 0 and gg\to\infty:

β(g)=ln(αg+1).\displaystyle\beta(g)=-\ln\left(\frac{\alpha}{g}+1\right). (5)

To solve the differential equation in Eq. (1) with β(g)\beta(g) in Eq. (5), we impose g(L0)=g0g(L_{0})=g_{0} as the boundary condition, where L0L_{0} is the microscopic characteristic length that is much larger than the mean free path and g0g_{0} is the conductance at length scale L0L_{0}. Under this condition, Eq. (1) can be rewritten as

g0g(L)dlngβ(g)=L0LdlnL.\displaystyle\int_{g_{0}}^{g(L)}\dfrac{d\ln g}{\beta(g)}=\int_{L_{0}}^{L}d\ln L. (6)

By solving Eq. (6) with Eq. (5), we obtain the LL dependence of gg at T=0T=0.

To extend this theory to finite TT, Anderson, Abrahams, and Ramakrishnan (AAR) proposed replacing the system length LL in Eq. (6) with a characteristic length LϕL_{\phi} (hereafter referred to as the dephasing length) over which an electron remains in an eigenstate [23] . In the diffusive transport in the WL regime, LϕL_{\phi} is given by Lϕ=DτϕL_{\phi}=\sqrt{D\tau_{\phi}} with diffusion coefficient DD and dephasing time τϕ\tau_{\phi}. Because the dephasing time generally depends on TT as τϕT2p\tau_{\phi}\propto T^{-2p} with a positive constant pp, the TT-dependence of LϕL_{\phi} is given by LϕTpL_{\phi}\propto T^{-p}. In the present work, we adopt LϕL_{\phi} as a single parameter to describe the TT dependences, similar to λ=εFτ\lambda=\frac{\hbar}{\varepsilon_{\rm F}\tau} in L=L=\infty at T=0T=0 in the SC theory, and express it as

Lϕ(ε,T)=Ld(TTd)p\displaystyle L_{\phi}(\varepsilon,T)=L_{\rm d}\left(\frac{T}{T_{\rm d}}\right)^{-p} (7)

in terms of three different parameters: the temperature TdT_{\rm d} characterizing the WL-SL crossover and the dephasing length LdL_{\rm d} at T=TdT=T_{\rm d} and the exponent pp of the temperature dependence. This expression is possible because the TT dependence of conductance in a 2D system, unlike that in a 3D system with the mobility edge, exhibits a gradual change from logarithmic to exponential behavior in the ω\omega dependences of the conductance as TT decreases.

In the WL limit, where g0g_{0} is large and g0αln(Lϕ/L0)=αpln(TWL/T)g_{0}\gg\alpha\ln(L_{\phi}/L_{0})=\alpha p\ln(T_{\rm WL}/T) is satisfied, the differential equation in Eq. (6), with Lϕ(ε,T)L_{\phi}(\varepsilon,T) substituted for LL, can be solved analytically. The spectral conductance g(ε,T)g(\varepsilon,T) is given by

g(ε,T)\displaystyle g(\varepsilon,T) =\displaystyle= g0(ε)αln(Lϕ(ε,T)L0)\displaystyle g_{0}(\varepsilon)-\alpha\ln\left(\frac{L_{\phi}(\varepsilon,T)}{L_{0}}\right) (8)
\displaystyle\equiv g0(ε)αp(ε)ln(TWL(ε)T)\displaystyle g_{0}(\varepsilon)-\alpha p(\varepsilon)\ln\left(\frac{T_{\rm WL}(\varepsilon)}{T}\right)

with

TWL(ε)Td(Ld(ε)L0)1/p(ε),\displaystyle T_{\rm WL}(\varepsilon)\equiv T_{\rm d}\left(\frac{L_{\rm d}(\varepsilon)}{L_{0}}\right)^{1/p(\varepsilon)}, (9)

resulting in the well-known logarithmic correction to conductance as long as T<TWL(ε)T<T_{\rm WL}(\varepsilon). On the other hand, in the SL limit where g0g_{0} is small (α/g01\alpha/g_{0}\gg 1), the differential equation can also be analytically solved, enabling g(ε,T)g(\varepsilon,T) to be expressed as

g(ε,T)=α(g0α)LϕL0αexp[(TSL(ε)T)p(ε)]\displaystyle g(\varepsilon,T)=\alpha\left(\frac{g_{0}}{\alpha}\right)^{\frac{L_{\phi}}{L_{0}}}\equiv\alpha\exp\left[-\left(\frac{T_{\rm SL}(\varepsilon)}{T}\right)^{p(\varepsilon)}\right] (10)

for TTSL(ε)T\ll T_{\rm SL}(\varepsilon), where TSL(ε)T_{\rm SL}(\varepsilon) is defined as

TSL(ε)Td(Ld(ε)L0)1/p(ε)(lnαg0(ε))1/p(ε).\displaystyle T_{\rm SL}(\varepsilon)\equiv T_{\rm d}\left(\frac{L_{\rm d}(\varepsilon)}{L_{0}}\right)^{1/p(\varepsilon)}\left(\ln\frac{\alpha}{g_{0}(\varepsilon)}\right)^{1/p(\varepsilon)}. (11)

II.2 Thermoelectric response in WL-SL crossover

The electrical conductivity and the Seebeck coefficient can be expressed in terms of g(ε,T)g(\varepsilon,T) as follows. In linear response theory, the electrical current density 𝒋{\bm{j}} under the electric field 𝑬{\bm{E}} and the temperature gradient T\nabla T can be described by 𝒋=L11𝑬L12(T/T)\displaystyle{\bm{j}}=L_{11}{\bm{E}}-L_{12}({\nabla T}/{T}). Here, the electrical conductivity L11(=σ)L_{11}(=\sigma) and the thermoelectric conductivity L12L_{12} can be expressed by the following Sommerfeld-Bethe (SB) relation based on the Kubo–Luttinger theory [24, 25] under the assumption that the heat current is carried only by electrons.

L11(T)\displaystyle L_{11}(T) =\displaystyle= 2e2hLAc𝑑ε(f0ε)g(ε,T),\displaystyle\frac{2e^{2}}{h}\frac{L}{A_{\rm c}}\int_{-\infty}^{\infty}\!\!\!d\varepsilon\left(-\frac{\partial f_{0}}{\partial\varepsilon}\right)g(\varepsilon,T), (12)
L12(T)\displaystyle L_{12}(T) =\displaystyle= 2ehLAc𝑑ε(f0ε)(εμ)g(ε,T),\displaystyle-\frac{2e}{h}\frac{L}{A_{\rm c}}\int_{-\infty}^{\infty}\!\!\!d\varepsilon\left(-\frac{\partial f_{0}}{\partial\varepsilon}\right)(\varepsilon-\mu)g(\varepsilon,T), (13)

where the factor of 2 accounts for the spin degree of freedom, LL and AcA_{\rm c} represent the length and cross-sectional area of the system, respectively, f0f_{0} denotes the Fermi–Dirac distribution function, and μ\mu is the chemical potential. Notably, the SB relation was originally derived from the Boltzmann transport theory (BTT); however, it is valid even for strongly disordered systems that cannot be treated by the BTT [26]. Using Eq. (12) and Eq. (13), we can describe the Seebeck coefficient by

S=1TL12(T)L11(T)=1eT𝑑ε(f0ε)(εμ)g(ε,T)𝑑ε(f0ε)g(ε,T).\displaystyle S=\dfrac{1}{T}\dfrac{L_{12}(T)}{L_{11}(T)}=\dfrac{-1}{eT}\dfrac{\int_{-\infty}^{\infty}\!d\varepsilon\left(-\frac{\partial f_{0}}{\partial\varepsilon}\right)(\varepsilon-\mu)g(\varepsilon,T)}{\int_{-\infty}^{\infty}\!d\varepsilon\left(-\frac{\partial f_{0}}{\partial\varepsilon}\right)g(\varepsilon,T)}. (14)

Here, we note that the ε\varepsilon dependences of g(ε,T)g(\varepsilon,T) results from those of g0g_{0}, pp, TWLT_{\rm WL} and TSLT_{\rm SL} in the present scheme. When the ε\varepsilon dependence of g(ε,T)g(\varepsilon,T) within kBTk_{\rm B}T is weak and can be approximated as linear with respect to ε\varepsilon, i.e., g(ε,T)g(μ,T)+g(μ,T)(ϵμ)g(\varepsilon,T)\approx g(\mu,T)+g^{\prime}(\mu,T)(\epsilon-\mu), the Seebeck coefficients in the two limits are respectively given as

SWLS0WL(μ)TTWL(μ)(1+A(μ)lnTWL(μ)T)\displaystyle S_{\rm WL}\approx S_{0}^{\rm WL}(\mu)\frac{T}{T_{\rm WL}(\mu)}\left(1+A(\mu)\ln\frac{T_{\rm WL}(\mu)}{T}\right) (15)

and

SSL\displaystyle S_{\rm SL} \displaystyle\approx S0SL(μ)(TTSL(μ))1p(μ)\displaystyle S_{0}^{\rm SL}(\mu)\left(\frac{T}{T_{\rm SL}(\mu)}\right)^{1-p(\mu)} (16)
×(1+B(μ)lnTSL(μ)T).\displaystyle\times\left(1+B(\mu)\ln\frac{T_{\rm SL}(\mu)}{T}\right).

Here, S0WL(μ)S_{0}^{\rm WL}(\mu), A(μ)A(\mu) and S0SL(μ)S_{0}^{\rm SL}(\mu), B(μ)B(\mu) are respectively given by

S0WL(μ)\displaystyle S_{0}^{\rm WL}(\mu) =\displaystyle= π2kB2TWL(μ)3eg0(μ)g0(μ)(1Δ(μ)),\displaystyle-\frac{\pi^{2}k_{\rm B}^{2}T_{\rm WL}(\mu)}{3e}\frac{g^{\prime}_{0}(\mu)}{g_{0}(\mu)}\left(1-\Delta(\mu)\right), (17)
A(μ)\displaystyle A(\mu) =\displaystyle= αp(μ)g0(μ)(1Δ(μ))p(μ)g0(μ)1Δ(μ)\displaystyle\alpha\frac{\frac{p(\mu)}{g_{0}(\mu)}(1-\Delta(\mu))-\frac{p^{\prime}(\mu)}{g^{\prime}_{0}(\mu)}}{1-\Delta(\mu)} (18)

with Δ(μ)=αp(μ)g0(μ)TWL(μ)TWL(μ)\Delta(\mu)=\alpha\frac{p(\mu)}{g^{\prime}_{0}(\mu)}\frac{T^{\prime}_{\rm WL}(\mu)}{T_{\rm WL}(\mu)}, and

S0SL(μ)\displaystyle S_{0}^{\rm SL}(\mu) =\displaystyle= π2kB23ep(μ)TSL(μ),\displaystyle\frac{\pi^{2}k_{\rm B}^{2}}{3e}p(\mu)T_{\rm SL}^{\prime}(\mu), (19)
B(μ)\displaystyle B(\mu) =\displaystyle= p(μ)p(μ)TSL(μ)TSL(μ).\displaystyle\frac{p^{\prime}(\mu)}{p(\mu)}\frac{T_{\rm SL}(\mu)}{T_{\rm SL}^{\prime}(\mu)}. (20)

Thus, SWL(T)S_{\rm WL}(T) is proportional to TT, whereas SSL(T)S_{\rm SL}(T) is proportional to T1pT^{1-p} and both of which have logarithmic corrections.

Refer to caption
Figure 1: (Color online) (a) TT dependence of the electrical conductivity of a PBTTT thin film under various gate voltages. The marks are experimental data acquired by Ito et al. using the ECT [15]. The solid curves are the present theoretical results. The left and right vertical axes correspond to the experimental and theoretical results, σexp\sigma_{\rm exp} and σth\sigma_{\rm th}, respectively. (b) TT dependence of the electrical conductivity of chemically carrier-doped PBTTT thin films. The marks are experimental data reported by Watanabe et al. [9].

III Comparison with experiments

III.1 TT dependence of electrical conductivity

We here compare the present theory to the experimental data for PBTTT thin films obtained by two different carrier-doping methods: the electrochemical doping method [15] and the chemical doping method [9]. For the ECT fabricated by Ito et al. [15], the channel length between the source and drain electrodes is L=140L=140 μ\mum, the channel width is W=2W=2 mm, and the film thickness is t=17t=17 nm. On the other hand, for the PBTTT device fabricated by Watanabe et al. [9], L=300L=300 μ\mum, W=80W=80 μ\mum, and t=40t=40 nm. Note that the lamellar spacing of PBTTT (a distance between two layers in the PBTTT with a multi layered 2D structure) is dl=2d_{l}=2 nm; thus, the cross-sectional area of a single layer in Eqs. (12) and (13) is given by AcWdlA_{\rm c}\equiv Wd_{l}.

Figure 1(a) presents the experimental electrical conductivity, σexp(T)\sigma_{\rm exp}(T), under various gate voltages, VgV_{g}, as measured using the ECT setup [15]. σexp(T)\sigma_{\rm exp}(T) increases monotonically with increasing |Vg||V_{g}| at a fixed TT, which is reasonable because the hole density increases. We also observe that σexp(T)\sigma_{\rm exp}(T) is proportional to lnT\ln T in the higher-TT region and deviates from this logarithmic behavior as TT decreases, which we ascribe to a precursor to the WL–SL crossover. The solid curves in Fig. 1(a) represent the TT dependences of theoretical results σth(T)\sigma_{\rm th}(T) based on the assumption of L0=4L_{0}=4nm, α=1.0\alpha=1.0, and Td=20T_{\rm d}=20K, where TdT_{d} is chosen as a typical temperature at which σexp(T)\sigma_{\rm exp}(T) in Fig. 1 deviates from the logarithmic behavior. The three qualitatively different parameters (g0g_{0}, LdL_{\rm d}, and pp) obtained by fitting to experimental data are listed in TABLE 1, showing convincing agreement, except for the difference between the absolute values of σexp\sigma_{\rm exp} and σth\sigma_{\rm th}. We consider that the difference is due to uncontrolled sample conditions, such as layer numbers and types of spatial disorder.

Figure 1(b) shows the electrical conductivity of PBTTT thin films chemically doped at two different doping levels. The marks represent experimental data [9] and the solid curves are the present theoretical results based on the same assumption of L0=4L_{0}=4nm, α=1.0\alpha=1.0, Td=20T_{\rm d}=20K. The values of the three parameters (g0g_{0}, LdL_{\rm d} and pp) are listed in TABLE 2. The theoretical curves are in excellent agreement with the experimental data, except for the difference between their absolute values, similar to Fig. 1(a).

Table 1: Parameters g0g_{0}, LdL_{d} and pp as a function of VgV_{g} for the PBTTT-based ECT in Ref. [15]. Here, L0=4L_{0}=4nm, α=1.0\alpha=1.0 and Td=20T_{\rm d}=20K are assumed.
Vg(V)V_{g}~{}({\rm V}) g0g_{0} Ld(nm)L_{\rm d}~{}({\rm nm}) pp
1.0-1.0 0.55 49.5 0.449
1.1-1.1 0.88 54.0 0.437
1.2-1.2 1.24 59.3 0.422
1.3-1.3 1.51 64.4 0.410
1.5-1.5 2.02 73.9 0.393
1.7-1.7 2.31 77.3 0.392
2.0-2.0 2.55 80.5 0.391
Table 2: Parameters g0g_{0}, LdL_{d}, and pp for the chemically carrier-doped PBTTT thin films in Ref. [9]. Here, L0=4L_{0}=4 nm, α=1.0\alpha=1.0, and Td=20T_{\rm d}=20 K are assumed.
Doping level g0g_{0} Ld(nm)L_{d}~{}({\rm nm}) pp
low 0.92 42.7 0.441
high 2.08 66.7 0.383

As VgV_{g} changes from Vg=1.0V_{g}=-1.0 V to 2.0-2.0 V, g0g_{0} increases monotonically from 0.550.55 to 2.552.55, which reflects the carrier density and the strength of disorder scattering, as characterized by parameter λ\lambda under the SC theory. This result suggests that, under these experimental conditions, the electrons in the PBTTT thin film are situated more or less in the WL regime rather than in the SL regime. The data in TABLE 1 also show that the dephasing length LdL_{\rm d} increases and the exponent pp decreases because of the delocalization of electrons with increasing |Vg||V_{g}|. Similar tendencies in pp and LdL_{d} are observed in the data in TABLE 2. Thus, the experimental data obtained by different carrier-doping methods can be understood in a unified manner using the three parameters g0g_{0}, LdL_{\rm d} and pp.

Refer to caption
Figure 2: (Color online) (a) TT dependence of Seebeck coefficient S(T)S(T) of a PBTTT thin film under various gate voltages. The marks are experimental data measured by Ito et al. using the ECT [15], and the solid lines denote the TT-linear behavior in the WL regime in the high-TT region. (b) TT-dependence of S(T)S(T) of chemically carrier-doped PBTTT thin films. The marks are experimental data reported by Watanabe et al. [9]. The lines in the high-TT region represent the TT-linear behavior in the WL regime, and the line in the low-TT region denotes the power-law behavior as S(T)T1pS(T)\propto T^{1-p^{*}} with p0.3p^{*}\sim 0.3.

As shown in TABLE 1 and TABLE 2, the exponent is p0.4p\sim{0.4} which implies that τϕ1\tau^{-1}_{\phi} is almost proportional to TT in the crossover regime where diffusive motions are assumed in the microscopic region. This TT dependence might indicate that Coulomb interaction is the dominant scattering processes in this regime [28, 29, 30]. It is of interest to see that the present result of p1/2p\sim 1/2 based on quantum transport is close to that argued by Efros and Shklovskii [31], whose approach was based on the idea of VRH caused by electron–electron scattering. Such an interesting possibility of interaction effects in the crossover regime between conducting and localized regimes is a very particular property of 2D systems, in contrast to a sharp, though continuous, critical transition between them in 3D systems as analyzed recently for the Seebeck coefficient taking into account the energy dependences of the localization length near the mobility edge [33]. It is to be noted that p=1/3p=1/3 in a 2D Mott VRH [32], where the electron hopping between localized states is caused by electron–phonon scattering.

III.2 TT dependence of Seebeck coefficient

We here discuss the TT dependence of S(T)S(T) data for PBTTT thin films obtained by the electrochemical doping method [15] and the chemical doping method [9]. Figure 2(a) shows the experimental S(T)S(T) of the PBTTT-based ECT under various VgV_{g} [15]. The S(T)S(T) for all VgV_{g} are proportional to TT in the high-TT region, where the electrical conductivity varies as σ(T)lnT\sigma(T)\propto\ln T, reflecting the WL. By fitting the experimental data using S(T)=S0WLTTWLS(T)=S_{0}^{\rm WL}\frac{T}{T_{\rm WL}} in Eq. (15), we determined S0WLS_{0}^{\rm WL} and TWLT_{\rm WL}, as listed in TABLE 3, where TWLT_{\rm WL} can be obtained from Eq. (9). Both S0WLS_{0}^{\rm WL} and TWLT_{\rm WL} increase as VgV_{g} (or g0g_{0}) increases, as expected from Eqs. (9) and (17). In the low-TT region for low-doped cases of Vg=1.0V_{g}=1.0V and 1.11.1V, S(T)S(T) is expected to deviate from TT-linear behavior; however, corresponding experimental data are not available.

Figure 2(b) shows the S(T)S(T) for chemically carrier-doped PBTTT films [9]. In the case of high doping, S(T)S(T) is proportional to TT within a wide TT region from 30 to 200 K, reflecting the WL. By contrast, in the case of low doping, S(T)S(T) is also linear with respect to TT in the high-TT region but deviates from the TT-linear behavior in the low-TT region (the fitting parameters S0WLS_{0}^{\rm WL} and TWLT_{\rm WL} for the high- and low-doped cases are summarized in TABLE 4). In the low-TT region, S(T)S(T) appears to vary as S(T)T1pS(T)\propto T^{1-p^{*}} with p0.3p^{*}\sim 0.3. Notably, the TT-dependence of S(T)S(T) follows a power-law behavior, which is a characteristic feature of SL as shown in Eq. (16), even though the electrons in this low-doped PBTTT thin film are not in the fully SL regime but are somewhat more localized than in the WL regime.

Table 3: S0WLS_{0}^{\rm WL} and TWLT_{\rm WL} for the PBTTT-based ECT used in Ref. [15]. Here, L0=4L_{0}=4 nm, α=1.0\alpha=1.0 and Td=20T_{\rm d}=20 K are assumed.
Vg(V)V_{\rm g}~{}({\rm V}) S0WLS_{0}^{\rm WL} (mV/K) TWLT_{\rm WL} (K) S0WL/TWLS_{0}^{\rm WL}/T_{\rm WL} (nV/K2)
1.0-1.0 1.171.17 5.42×1035.42\times 10^{3} 215.8
1.1-1.1 1.251.25 7.72×1037.72\times 10^{3} 161.9
1.2-1.2 1.441.44 11.9×10311.9\times 10^{3} 121.0
1.3-1.3 1.801.80 17.6×10317.6\times 10^{3} 102.3
1.5-1.5 2.622.62 33.5×10333.5\times 10^{3} 78.3
1.7-1.7 2.502.50 38.1×10338.1\times 10^{3} 65.6
2.0-2.0 2.362.36 43.3×10343.3\times 10^{3} 54.5
Table 4: S0WLS_{0}^{\rm WL} and TWLT_{\rm WL} for the chemically carrier-doped PBTTT thin films investigated in Ref. [9]. Here, L0=4L_{0}=4 nm, α=1.0\alpha=1.0, and Td=20T_{\rm d}=20 K are assumed.
doping level S0WLS_{0}^{\rm WL} (mV/K) TWLT_{\rm WL} (K) S0WL/TWLS_{0}^{\rm WL}/T_{\rm WL} (nV/K2)
low 0.700.70 4.29×1034.29\times 10^{3} 163.2
high 2.612.61 31.0×10331.0\times 10^{3} 84.1

IV Summary and outlook

We developed a new theoretical scheme for charge transport and TE response in disordered 2D electron systems. The proposed scheme can describe the WL–SL crossover on the basis of the AALR+AAR theory [23, 20] combined with the Kubo–Luttinger theory [24, 25]. The two key aspects of the scheme are (i) the introduction of the unified β\beta function, which seamlessly connects the WL and SL regimes, and (ii) the description of the TT dependence of the conductance from high and low TT regions in terms of Lϕ(ε,T)L_{\phi}(\varepsilon,T) in Eq. (7). Using this scheme, we predicted S(T)TS(T)\propto T in WL and S(T)T1pS(T)\propto T^{1-p} in SL, both with logarithmic corrections.

In addition, we applied this scheme to interpret recent experimental σ(T)\sigma(T) and S(T)S(T) data for PBTTT thin films, which are realized by two different carrier doping methods [15, 9]. As a result, we could provide a unified theoretical interpretation for both experimental data based on the new scheme. We found that the electrons in the PBTTT thin films used in both experiments [15, 9] are located more or less in the WL regime, not in the SL regime. We also found that S(T)S(T) in PBTTT thin films exhibits TT-linear behavior in the WL regime with a high carrier density but power-low behavior when the system begins to enter the SL regime from the WL regime as the carrier density decreases. Finally, we expect that the current theoretical scheme is applicable to other 2D materials (e.g., atomic layered materials) apart from present organic ones.

We thank Taishi Takenobu and Shun-ichiro Ito for fruitful discussions and for providing experimental data acquired using an ECT, We also thank Junichi Takeya and Shun Watanabe for valuable discussions regarding the experimental data obtained through the chemical doping method. This work was partly supported by JSPS KAKENHI (Grant Nos. 22K18954 and 23H00259).

References

  • [1] L. D. Hicks and M. S. Dresselhaus, Phys. Rev. B 47, 12727 (1993).
  • [2] S. Wang, M. Ha, M. Manno, C. D. Frisbie, and C. Leighton, Nat. Commun. 3, 1210 (2012).
  • [3] D. Venkateshvaran, M. Nikolka, A. Sadhanala, et al., Nature 515, 384–388 (2014).
  • [4] A. M. Glaudell, J. E. Cochran, S. N. Patel and M. L. Chabinyc, Adv. Energy Mater. 5, 1401072 (2015).
  • [5] S. Zanettini, J. F. Dayen, C. Etrillard, N. Leclerc, and M. V. Kamalakar, Appl. Phys. Lett. 106, 063303 (2015).
  • [6] K. Kang, S. Watanabe, K. Broch, A. Sepe, A. Brown, I. Nasrallah, M. Nikolka, Z Fei, M. Heeney, D. Matsumoto, K. Marumoto, H. Tanaka, S. Kuroda, and H. Sirringhaus, Nat. Mater, 15, 896-902 (2016).
  • [7] S. D. Kang and G. J. Snyder, Nat. Mater. 16, 252 (2017).
  • [8] Y. Yamashita, J. Tsurumi, M. Ohno, R, Fujimoto, S. Kumagai, T. Kurosawa, T. Okamoto, J. Takeya, and S. Watanabe, Nature. 572, 634–638 (2019).
  • [9] S. Watanabe, M. Ohno, Y. Yamashita, T. Terashige, H. Okamoto, and J. Takeya, Phys. Rev. B 100, 241201(R) (2019).
  • [10] H. Tanaka, K. Kanahashi, N.Takekoshi, H. Mada, H. Ito, Y. Shimoi, H. Ohta, and T. Takenobu, Sci. Adv. 6, eaay8065 (2020).
  • [11] S. Fratini, M. Nikolka, A. Salleo, G. Schweicher, and H. Sirringhaus, Nat. Mater. 19, 491–502 (2020).
  • [12] H. Ito, H. Mada, K. Watanabe, H. Tanaka, and T. Takenobu, Commun. Phys. 4, Article number: 8 (2021).
  • [13] Y. Huang, D. H. L. The, I. E. Jacobs, X. Jiao, Q. He, M. Statz, X. Ren, X. Huang, I.McCulloch, M. Heeney, C. McNeill, and H. Sirringhaus, Appl. Phys. Lett. 119, 111903 (2021).
  • [14] C. Chen, I. E. Jacobs, K. Kang, Y. Lin, C. Jellett, B. Kang, S. B. Lee, Y. Huang, M. B. Qarai, R. Ghosh, M. Statz, W. Wood, X. Ren, D. Tjhe, Y. Sun, X. She, Y. Hu, L. Jiang, F. C. Spano, I. McCulloch, and H. Sirringhaus, Adv. Energy Mater. 13, 2202797 (2023).
  • [15] S-i. Ito and T. Takenobu (in preparation).
  • [16] J.E. Northrup, Phys. Rev. B 76, 245202 (2007).
  • [17] L.-H. Li, O. Y. Kontsevoi, S. H. Rhim, and A. J. Freeman, J. Chem. Phys. 138, 164503 (2013).
  • [18] N.F. Mott and E.A. Davis,“Electronic Processes in Non-Crystalline Materials”, (Clarendon, Oxford, 1971), p. 47.
  • [19] Anderson Localization, ed. Y. Nagaoka: Prog. Theor. Phys. Supplement 84 (1985).
  • [20] E. Abrahams, P.W. Anderson, D.C. Licciardello and T.V. Ramakrishnam, Phys. Rev. Lett., 42, 673 (1979).
  • [21] D. Vollhardt and P. Wölfle, Phys. Rev. B, 22, 4666 (1980).
  • [22] A. Kawabata, Solid State Commun., 38, 823 (1981).
  • [23] E. Abrahams, P.W. Anderson, and T.V. Ramakrishnam, Phys. Rev. Lett., 43, 718 (1979).
  • [24] R. Kubo, J. Phys. Soc. Jpn. 12, 570 (1957).
  • [25] J. M. Luttinger, Phys. Rev. 135, A1505 (1964).
  • [26] M. Ogata and H. Fukuyama, J. Phys. Soc. Jpn. 88, 074703 (2019).
  • [27] A.G. Zabrodskii and K. N. Zeninova, Eksp. Teor. Fiz. 86, 727 (1984).
  • [28] E. Abrahams, P.W. Anderson, D.C. Licciardello and T.V. Ramakrishnam, Phys. Rev., 24, 6783 (1981).
  • [29] H. Fukuyama and E. Abrahams, Phys. Rev. B 27, 5976 (1983).
  • [30] H. Fukuyama, in A.L. Efros and M. Pollak (Eds.), “Electron-electron interactions in disordered systems” (North Holland, 1985), Chap. 5, p. 155.
  • [31] A. L. Efros and B. I. Shklovskii, J. Phys. C8, L49 (1975).
  • [32] N.F. Mott, J. Non-Cryst. Solids 1, 1 (1968).
  • [33] T. Yamamoto, M. Ogata and H. Fukuyama: J. Phys. Soc. Jpn. 91, 044704 (2022).