Quantized dynamics in a square-wave-driven coherent single-electron source

Jingjing Cheng , Jiayan Zhang , Fuming Xu , Yanxia Xing

Front. Phys. ›› 2026, Vol. 21 ›› Issue (3) : 035203

PDF (3395KB)
Front. Phys. ›› 2026, Vol. 21 ›› Issue (3) : 035203 DOI: 10.15302/frontphys.2026.035203
RESEARCH ARTICLE

Quantized dynamics in a square-wave-driven coherent single-electron source

Author information +
History +
PDF (3395KB)

Abstract

Through analytic derivation within the nonequilibrium Green’s function (NEGF) formalism, we present a comprehensive study of a quantum-coherent single-electron emitter. This emitter is based on a quantum RC circuit driven by periodic square-wave potentials with temporal period T and angular frequency ω=2π/T. Our theoretical results reveal three key characteristics of the single-electron emitter: (i) the characteristic time α of the exponentially decaying current J(t)(et/α) matches the intrinsic coherence time τ of the quantum dot; (ii) the Fourier spectrum of the current J(mω) carries information of both external driving amplitude U and internal energy levels ϵd of the quantum dot; (iii) the quantization of fundamental Fourier component |J1|=2ef (f=1/T) accompanies the half-quantized relaxation resistance Rq=h/(2e2), featuring quantized dynamics in AC transport through the single-electron emitter. These findings offers new perspectives into the dynamic properties of quantum-coherent AC transport.

Graphical abstract

Keywords

nonequilibrium Green’s function

Cite this article

Download citation ▾
Jingjing Cheng, Jiayan Zhang, Fuming Xu, Yanxia Xing. Quantized dynamics in a square-wave-driven coherent single-electron source. Front. Phys., 2026, 21(3): 035203 DOI:10.15302/frontphys.2026.035203

登录浏览全文

4963

注册一个新账户 忘记密码

1 Introduction

The quantum hierarchy in electrical transport manifests through three cardinal quantization paradigms: (i) Conductance quantization, emerging from discrete electronic states in ballistic channels [1, 2], forming the basis of Landauer−Büttiker formalism. (ii) Flux quantization [3], such as fluxoid Φ=h/e in Aharonov−Bohm effect or superconducting fluxoid Φ=h/(2e). (iii) Quantum Hall hierarchy, exhibiting ν-quantized plateaus through Landau levels filling in two-dimensional electron gases (2DEGs) [4, 5], establishing the von Klitzing constant RK=h/e2 as resistance metrology’s cornerstone [68]. While traditional DC transport regimes have unveiled foundational quantum phenomena, the exploration of time-resolved quantum transport reveals novel phenomena absent in DC regimes as well. A striking example is the half-quantized relaxation resistance Rq=h/(2e2) which is originated from the frequency-dependent admittance [9] and confirmed through single-electron capacitance spectroscopy [10]. It establishes fundamental constraints on frequency-dependent impedance parameters (capacitance C, resistance R and inductance L) [11]. On the other hand, if frequency and amplitude are sufficiently high, coherent control of single electrons [12] can also be achieved. This can be realized through single-electron pumps [1317] or single-electron sources [1820]. The single-electron pump have significantly advanced the development of quantum current standards [2126], achieving uncertainties of less than 1ppm.

Single-electron source enables on-demand electron emission with temporal precision comparable to photon manipulation in optical systems. This capability forms the foundation for emerging quantum technologies including quantum sensing [27, 28], electron motion capture [29], solid-state flying qubit control [30], electron control in quantum optics [3133]. Four primary methodologies have been developed for single-electron source: (i) Tunable-barriers pumps through utilizing time-dependent gate voltages [16, 3436]; (ii) Single-electron transfer through Coulomb interaction [37, 38]; (iii) Itinerant quantum dots obtained using surface acoustic waves caused by piezoelectric effect [3942]; (iv) Leviton generation, i.e., the creation of minimal-excitation wave packets through Lorentzian voltage pulses [4346]. These sources have been demonstrated across diverse material systems, including two-dimensional graphene [47], molecular junctions [48], and silicon-based nanostructures [14, 29]. For quantum information applications, given the critical importance of temporal precision in electron emission, the paradigmatic single-electron source employs alternating current [10, 18, 49], where radio-frequency voltage drives induce single-electron emission/absorption during alternating half-cycles of the capacitive charging/discharging sequence. Several theoretical approaches have been employed to solve the dynamic response of quantum system. For harmonic driving sources, the Floquet scattering matrix formalism [45, 5052] provides a non-perturbed description of time-periodic transport phenomena in low frequency. The quantum master equation approach [16] is effective for modeling dissipative dynamics in weakly coupled systems. The nonequilibrium Green’s function (NEGF) techniques [5362] and the scattering matrix [6365] also provide various perturbative treatment of dynamic response in energy representation.

In this work, we present a theoretical investigation for a quantum-coherent single-electron emitter using the NEGF formalism. Based on a quantum RC circuit setup [66] featuring a single-electron reservoir [67] driven by periodic square-wave potentials, we reproduce the existing experimental findings [18], i.e., the coherent emission-absorption cycles and the quantized alternating current. Remarkably, our results achieve superior current quantization precision beyond existing benchmarks [20, 68], exhibiting enhanced operational robustness in open quantum environments. It demonstrates that square-wave signals can also effectively drive single-electron emission, providing an alternative to established approaches based on smooth AC driving potentials [20, 59] or transient square waves [19, 69]. Employing NEGF techniques, we systematically derive the time-resolved spectral function A(t,ϵ), thereby enabling rigorous characterization of the non-stationary AC current J(t). Frequency-domain analysis through Fourier decomposition yields essential harmonic coefficients J(mω), providing information into the dynamic response of the quantum system. These findings include: (i) the characteristic time α of the exponentially decaying current J(t)(et/α) matches the intrinsic coherence time τ of the quantum dot; (ii) the decaying current oscillates with a frequency proportional to the external driving amplitude U, which is also supported by the Fourier component of the current J(Ω) occurring precisely at Ω=U; (iii) quantized fundamental Fourier component |J1|=2ef is directly related to the half-quantized relaxation resistance Rq=h/(2e2), revealing the quantum dynamics of AC transport through this single-electron emitter. These theoretical results derived from the quantum RC circuit provide new insights for understanding AC transport phenomena.

The rest of this paper is organized as follows: Section 2 introduces the system Hamiltonian and establishes the theoretical framework for analyzing alternating current under periodic square wave voltage signals. In Section 3, we perform numerical calculations and discuss the physical interpretations of the obtained results. Section 4 extends our analysis beyond the conventional wideband regime, exploring more generalized operational conditions. Finally, a comprehensive summary is shown in Section 5.

2 Theoretical formalism

Our analysis begins with the quantum RC circuit configuration illustrated in Fig.1(a). The system comprises a semi-infinite lead and a quantum dot with gate-tunable energy levels ϵd. The quantum dot is electrically coupled to a semi-infinite lead through quantum point contact technology [10, 18]. This setup employs a single-lead design as documented in Refs. [19, 55, 67], where the lead is subjected to a time-periodic external potential U(t)=U(t+T). The quantum dot’s energy spectrum can be precisely controlled via gate voltage Vg modulation, following established methodologies in Refs. [7072]. The periodic square wave

U(t)={U,0<tT2U,T2<tT

is selected as the driving potential. The configuration of U(t) is shown in Fig.1(b). The abrupt transition edges provide optimal temporal resolution for investigating AC response dynamics. Within each oscillation cycle, two distinct operational phases emerge. During the first half-period, electron injection occurs from the external terminal into the quantum dot, while retrieval processes dominate the latter half-period [Fig.1(c) and (d)]. This bidirectional charge transfer mechanism manifests as alternating-polarity current pulses in the temporal domain, thereby giving rise to temporally resolved current signatures.

The total Hamiltonian is constituted from three components: the Hamiltonian of the isolated quantum dot Hd, the Hamiltonian of the isolated semi-infinite lead Hl, and their coupling term Ht. It is expressed as H=Hl+Hd+Ht where

Hd=nϵd,ndndn,Hl=k[ϵk+U(t)]ckck,Ht=k,ntkckdn+h.c..

Here, dn (dn) represents particle annihilation (creation) operator for n-th quantum level of quantum dot. The intrinsic quantum energy levels of the central quantum dot, denoted by ϵd,n, can be adjusted through gate voltage Vg. ck (ck) is the annihilation (creation) operator for electrons with momentum state k in the semi-infinite lead. The quantum dot is coupled to an external Fermi reservoir through semi-infinite lead. The Fermi level of Fermi reservior is modulated by the external excitation U(t). Performing a unitary transformation [73]

P(t)=exp[ki0tdτU(τ)ckck],

Hl becomes time independent, which is convenient for current calculation. Finally, the transformed Hamiltonian becomes

H~=k,nϵkckck+ϵndndn+[tk(t)ckdn+h.c.],

where the time dependence of Hl is effectively transferred to the coupling terms Ht with tk(t)=tkeie0tdτU(τ).

The system’s configuration employs a single-electron reservoir setup inherently eliminating conventional conduction current pathways. Within this framework, our analysis focuses exclusively on the displacement current component. Adopting the rigorous NEGF formalism, the displacement current flowing through the semi-infinite lead is formulated as

J(t)=2eRenGnrΣ<+Gn<Σa.

Here, we have introduced the integral abbreviation

Gnr/<Σ</a=tdt1[Gnr/<(t,t1)Σ</a(t1,t)],

where Gnr/< is retarded or less Green’s function originated from n-th energy level of central quantum dot, and Σ</a is self energy from the semi-infinite lead. In the presence of time-dependent external excitation U(t), self energy is

Σγ(t1,t)=eitt1dτU(τ)Σγ,0(t1t),

where γ=r,a,< and Σαγ,0 is the self energy when U(t)=0, which are expressed as

Σr/a,0(t1t)=i2dϵ2πeiϵ(t1t)Γ(ϵ),Σ<,0(t1t)=idϵ2πeiϵ(t1t)f(ϵ)Γ(ϵ),

where we have introduced the linewidth function

Γ(ϵ)=2πkδ(ϵϵk)|tk|2=2πρ(ϵ)|t(ϵ)|2

and the Fermi distribution f(ϵ). In the wide-band limit (WBL), ρ(ϵ) and t(ϵ) are energy independent, and Γ=2πρ|t|2, which is also energy independent. Then Σr/a,0(t1t)=i2δ(t1t)Γ. For notational simplicity, we henceforth set the elementary charge e and the reduced Planck constant to unity, i.e., e==1. Subsequently, by introducing the spectral function [74]

An(t,ϵ)=tdt1Gnr(t,t1)eiε(tt1)eitt1U(τ)dτ

and considering the WBL, Eq. (1) becomes

J(t)=Γn[Imdϵπf(ϵ)An(t,ϵ)+iGn<(t,t)]

with

Gn<(t,t)=tdt1dt2Gnr(t,t1)Σ<(t1,t2)Gna(t2,t)=iΓdϵ2πfα(ϵ)|An(t,ϵ)|2.

The first and second term in Eq. (5) are corresponding to the current flowing into and out of quantum dot, respectively. Once the spectral function A(t,ϵ) is ready, the calculation of the current J(t) is straightforward.

Now, we proceed to solve the spectral function A(t,ϵ). Given that Σr/a,0(t1t)=i2δ(t1t)Γ, from Eq. (2) we have Σr/a=Σr/a,0. This implies that the external excitation U(t) exerts no effect on the Green’s function Gnr. Consequently, we have

Gnr(tt1)=dE2πeiE(tt1)Gnr(E)

with Gnr(E)=1/(Eϵ~n). Here, the complex-valued energy ϵ~n=ϵd,niΓ/2 represents the broadened energy level of the quantum dot system. The parameter Γ quantitatively describes the spectral width of this broadened state. Substituting Eq. (6) into Eq. (4), we have

An(t,ϵ)=tdt1dE2πGnr(E)ei(ϵE)(tt1)eitt1U(τ)dτ.

Integrating over time t1, we have

An(t,ϵ)=idE2πGnr(E)[1B±U±2Uei(B±U)t¯(BU)(B+U)(1ei(BU)T/2)]

with B=ϵE+i0+. In the WBL, the parameter Γ is energy independent. So the retarded Green’s function Gnr(E) possesses a single pole at ϵ~n, with an associated residue of unity 1. This analytical structure enables explicit evaluation of the integral in Eq. (7) through systematic application of the Residue Theorem. Through the semi-infinite lead, electrons can escape from quantum dot to the Fermi sea, and the life time of the energy level of quantum dot is τ1/Γ. Under radio-frequency driving conditions, where the period T is much greater than the relaxation time τ (Tτ). It means ΓT1, so the exponential phase factor ei(ϵϵd,n+iΓ/2U)T/2 exhibits asymptotic decay to zero. This temporal hierarchy ensures the vanishing of non-stationary contributions in the long-time limit. Consequently, the spectral function is simplified to

An(t,ϵ)=1ϵϵ~n±U±2Uei(ϵϵ~n±U)t¯(ϵϵ~nU)(ϵϵ~n+U),

where “±” denote the first and second half cycle and t¯ is the time duration within each half-period. While Eq. (8) is formulated within WBL and does not fully account for realistic spectral constraints, it captures essential physical mechanisms governing the system’s nonequilibrium dynamics.

3 Results and discussion

In our calculations, all quantities are expressed in dimensionless arbitrary units. The Fermi level of the free electron reservoir serves as our reference zero energy, with all other energy values measured relative to this baseline. The system is driven by a periodic square wave potential U(t) with radio-frequency ω applied to the single lead. Unlike conventional harmonic excitations, this abruptly switching square wave not only probes the intrinsic response speed of nanoscale devices, but also enables precise characterization of transient current decay dynamics. Within the single-terminal RC circuit configuration under square wave excitation, coherent electrons shuttle between the lead and quantum dot. This charge transfer manifests as a distinct positive current pulse during the rising edge of the square wave, followed by a negative pulse during the falling edge. Fig.2 specifically illustrates the transient current J(t) profile during the first half-period excitation. While the external drive U(t) theoretically induces an instantaneous Fermi level shift, practical device limitations prevent abrupt Fermi surface modifications. Consequently, the current cannot instantaneously achieve the maximum due to finite temporal response. The switching dynamics exhibit a strong dependence on the U/Γ ratio. Increasing this parameter enhances switching speed through enhanced field-effect dominance. Fig.2(a) reveals three distinct temporal regimes: (i) rapid current ascent from baseline to maximum amplitude, (ii) peak current, and (iii) exponential relaxation to equilibrium. For sufficiently large U values [Fig.2(b)−(d)], the charging process becomes nearly instantaneous, and region (i) is invisible. After reaching maximum quickly, the current begins to decay slowly. The decaying dynamics are governed by quantum dot energy level broadening effects, enabling electrons/holes escape into electron reservoir from the quantum dot. It results in the decay of the J(t) with characteristic time constant α.

The transient current J(t) can be effectively modeled through analogy with classical RC circuit behavior, where the exponential relaxation process follows

JRC(t)=±1αet¯α

for respective half-periods. The characteristic decay time α corresponds to the 1/e reduction timescale of the maximum current. Fig.2 juxtaposes our theoretical predictions for J(t) (black solid curves) with numerically fitted RC-model simulations JRC(t) (red dotted curves). The normalization condition |JRC(t)|dt=1 per half-period ensures consistent comparison across pulse events. Our results reveal complete current relaxation J(t)0 within t30, demonstrating αT where T is the driving period. This temporal scaling justifies our truncated visualization (t40) in Fig.2 and subsequent analyses. Systematic fitting procedures yield the U-dependence of α, revealing a saturation phenomenon: For U>3, α converges to 5.2. This converged value 5.2 reveals the timescale of the current decay for quantum RC circuit. It is shown the current decaying time is governed by the quantum dot’s intrinsic state lifetime τ that is derived from energy level broadening Γ through τ=/Γ. With Γ=0.2 in our calculations, we obtain τ=5, matching the saturated α value within numerical precision. The equivalence ατ demonstrates that the quantum RC circuit’s relaxation timescale is fundamentally dictated by the lifetime of the particles in quantum dot. It implies that the current response J(t) encapsulates crucial quantum information of quantum RC circuit.

Upon comparing the RC current profile shown in Fig.2, we observe the following fundamental departure from the basic exponential decay predicted by Eq. (9): The current displays persistent decaying oscillations following its maximum peak, with the oscillation frequency showing direct proportionality to the external excitation amplitude U. To systematically investigate these coherent quantum phenomena, we perform spectral decomposition of the alternating current component through Fourier analysis with the following spectral components

J(Ω)=0TdteiΩtJ(t).

The quasi-continuous frequency spectrum Ω=mω with finite small radio-frequency (ωΓ). The distinct traces correspond to different combinations of excitation strength U and quantum dot energy level ϵd. Our spectral analysis reveals two distinct regimes. (i) Low-frequency regime (Ω<U): the Fourier components exhibit universal behavior independent of both intrinsic (ϵd) and extrinsic (U) parameters, with all curves collapsing into a single rapidly decaying profile. (ii) High-frequency regime (Ω>U): The spectral response shows Ω1 scaling with superimposed parameter-dependent features. In this regime, both quantum dot energy ϵd and driving amplitude U significantly influence the spectral characteristics. Detailed examination of the high-frequency regime (Fig.3) reveals following striking signatures. (i) Resonant case (ϵd=0, left panels): Spectral discontinuous higher-order derivatives of J(Ω) occurs precisely at the fundamental quantum resonance condition Ω=U, manifested as a predictable dip in |ΩJ(Ω)|. (ii) Detuned case (ϵd0, right panels): The single derivative discontinuity splits symmetrically into dual resonances at Ω=U±ϵd, reflecting modified quantum interference conditions under energy detuning. These discontinuous higher-order derivatives represent fundamental quantum phenomena with no classical counterparts in RC circuit analogs, serving as unambiguous fingerprints of coherent charge transfer dynamics. This dual-regime behavior offers powerful diagnostic tools for characterizing both intrinsic quantum dot properties (ϵd) and external driving parameters (U) through spectral measurements, establishing Fourier analysis as an essential technique for probing quantum coherence in masoscopic charge transport.

After analyzing the Fourier spectrum in the high-frequency range, let us now turn our attention to the low-frequency range. Given the discrete nature of the frequency spectrum Ω=mω, we adopt the notation J(Ω)Jm for simplicity. Numerical simulations reveal that the current magnitudes during successive half-cycles are nearly identical in amplitude but opposite in polarity. This symmetry suppresses even-order Fourier components, rendering them negligible. Consequently, our analysis focuses exclusively on the odd-order Fourier components (m=1,3,5,), which dominates the spectral response. Fig.5 highlights the maxima of Jm for the first four odd Fourier components (J1,3,5,7). Fig.5(a) plots modulus |Jm| against the real components Re(Jm), while Fig.5(b) presents a Nyquist diagram, where Im(Jm) is plotted against Re(Jm). In contrast to DC transport, the AC impedance Z=R+iX inherently possesses a complex character. Here, the real part R quantifies energy dissipation, whereas the imaginary part X reflects energy storage during periodic charging and discharging processes. The Fourier components Jm display a clear correlation with the AC impedance properties: The real part Re(Jm) corresponds to the dissipative resistance, while the imaginary part Im(Jm) reflects the non-dissipative storage. Fig.5 reveals a significant trend: Higher-order Fourier components show a pronounced dominance of imaginary components, with reactance increasing monotonically as m increases. This hierarchy implies the dissipative effects predominantly reside in low-order harmonics, while reactive behavior, governed by periodic energy redistribution, emerges strongly in higher-order modes. Such presentation of the Fourier spectrum demonstrates the dissipative-storage nature of AC quantum transport, inaccessible to conventional DC measurements.

Fig.5 systematically maps the data of Fourier harmonics Jm when quantum dot energy levels ϵd sweeps from 2 to 2. Distinct symbols mark three excitation strengths: U=0.5 (black squares), 1.0 (red circles), and 3.0 (blue rhombus). There are two distinct spectral hierarchy: (i) Strong driving regime (U=3.0): When U>(|ϵd|+Γ), the broadened quantum dot level (|ϵd|±Γ) lies entirely within the reservoir’s Fermi window [U,U]. This full spectral overlap enables maximum electron tunneling efficiency, causing all Jm values to converge to a universal amplitude independent of ϵd (overlapping blue symbols). Notably, |Jm| decays linearly with harmonic order m [Fig.5(a), blue]. (ii) Moderate driving regime (U=0.5,1.0): For |ϵd|>U, the quantum dot’s spectral broadening (|ϵd|±Γ) exceeds the Fermi window, resulting in partial tunneling and sub-maximal currents. The |Jm| becomes ϵd-dependent, approaching its peak only when ϵd0 (black/red curves). In this regime, the relative contribution of the resistive (Re(Jm)) and reactive (Im(Jm)) also exhibit two distinct trends: (i) for fundamental mode (m=1), resistive and reactive components vary synchronously with ϵd, producing a linear correlation between |J1| and Re(J1) [Fig.5(a), black/red]; (ii) for higher harmonics (m>1), reactance progressively dominates, culminating in Re(J1)0 with near-pure reactive response at m=7 [Fig.5(b)]. The distinct Fourier spectrum Jm reveals the dual role of AC impedance dynamics: Lower harmonics mediate irreversible dissipation while higher harmonics govern reversible quantum capacitance effects, a character absent in DC regimes.

We now analyze the dominant fundamental harmonic component J1, where resistive behavior prevails. Fig.6 contrasts two configurations: single-level quantum dot (ϵd=0) in panels (a) and (b) and multi-level quantum dot (ϵd=±1) in panels (c) and (d). For both cases, Im(J1) and |(J1)| are plotted against Re(J1) across excitation strengths U=0.5,1.0,3.0. Data points trace trajectories parameterized by coupling strength Γ[0.05,1], with the starting point of Γ=0.05 indicated by cyan triangle. For single-level dynamics [Fig.6(a) and (b)], with the increasing of Γ, resistance is continuously suppressed. Specifically, Im(J1) decreases monotonically, reflecting diminished quantum phase storage. While Re(J1) initially rises with Γ and peaks near Γ=0.7. Beyond this critical coupling, Re(J1) declines, signaling a crossover from quantum-coherent transport to classical resistive dominance. When ΓU, strong hybridization suppresses quantum interference effects, leaving only Ohmic dissipation. In the multi-level system with ϵd=1 and 1, three distinct curves emerge for different U. For U=0.5 (black symbols), the energy levels ϵd=±1 lie outside the effective bias window [U,U]. As a result, the cyclic energy storage and release processes is inactive. Then, the imaginary component Im(J1), representing reactive current storage, approaches zero, while the resistive component Re(J1) exhibits a linear dependence on tunneling strength Γ. When U exceeds 1.0, the ϵd=±1 states begin entering the bias window, establishing a complete AC response mechanism. Both reactive [Im(J1)] and resistive [Re(J1)] components demonstrate Γ-dependence in this regime. At U>2.0, the ϵd=±1 levels fully reside within the bias window U(t), enabling periodic storage and release of two quantum charges. This configuration produces nearly doubled current magnitudes compared to lower U values, as evidenced by the U=3 case in Fig.6(d) where |J1| reaches maximum amplitudes of 4. This contrasts markedly with single-level systems [Fig.6(b)], where maximum current amplitudes remain below 2 due to single-charge cycling limitations. The enhanced current capacity in the two-level system directly reflects the doubled quantum charge participation in the storage-release cycle.

Under the condition αT, electrons achieve complete occupation/depletion in the quantum dot during each half-period, manifesting quantized current transport. As derived from Eq. (9), the m-th Fourier component is given by

JRC(Ω)=0dteiΩtJRC(t)=21iαΩ,

where Ω=mω. For fundamental frequency, Ω=ω, and (αωτω)1, the quantized current approaches |J1|2(ef), reflecting two mutually inverse processes across half-periods. Furthermore, for the single-lead RC circuit, the electrochemical capacitance is [11]

Cμ(ω)=Cμ1iωCμRq,

where Cμ=(1C0+πΓe2)1 is static electrochemical capacitance [75]. If the geometric capacitance C0 is not taken into account, Cμ=e2/(πΓ), then with the aid of half-quantized relaxation resistance Rq=h/(2e2), we obtain

Cμ(ω)=e2πΓ11iωΓ.

Comparing with fundamental Fourier dynamic current J(ω)=2ef/(1iαω) where α=/Γ, we can find Cμ(ω) and J(ω) share the same frequency dependence. It means the even-quantized current |J1|2ef in low frequency is precisely originated from the half-quantized relaxation resistance Rq=h/(2e2) [10], they are both the universal features of quantum AC circuits, and together form the two sides of the dual nature of the AC transport. The quantum RC circuit driven by a periodic square wave perfectly illustrates this correspondence between half-integer quantum relaxation resistance and even-integer quantum current, providing new insights and perspectives for understanding AC transport.

Numerical verification of this quantization is performed using Eqs. (5) and (10). Fig.7 displays the fundamental Fourier component |J1| versus excitation strength U for linearly spaced quantum energy levels ϵd,n=Vg+[3.0,1.0,1.0,3.0] [Fig.7(a)] and non-linearly spaced levels ϵd,n=Vg+[3.0,1.5,1.2,3.0] [Fig.7(b)] with parameters ω=0.01 and Γ=0.1 which satisfies τT. Energy level positions are globally tuned via gate voltage Vg. The following conclusions hold universally, regardless of the distribution of the energy levels in quantum dots. (i) Plateau formation: The current |J1| displays distinct step-like plateaus at values of 2, 4, and 6 as U varies. The step-like progression reflects sequential activation of quantum dot levels, where n active levels produce |J1|2nef. This contrasts with single-level systems limited to |J1|2, demonstrating how multi-level configurations amplify quantized current through additive state participation. (ii) Multi-level resolution: When level broadening Γ remains smaller than inter level spacing Δϵd, discrete quantization emerges. Each energy level entering the excitation window U(t) contributes a 2-unit current increment. (iii) Transition points: Multi-levels result in step-like quantized current plateaus. At transition points, the quantized current plateau transitions from n1 to n+1. The invariance of quantization at these points represents their robustness to gate-induced energy shifts.

By comparing Fig.7(a) and (b), it is also evident that the evolution of quantized plateaus depend critically on energy level alignment. As shown in Fig.7(a), the linearly spaced ϵd,n enter the excitation window U(t)[U,U] sequentially with equal intervals. This results in linearly spaced transition points (U=1,2,). If ϵd,n is randomly arranged, with the increasing of U, ϵd,n enter the excitation window U(t) randomly. It means the arrangement of the quantized current directly reflects the distribution of the quantum energy levels, as shown in Fig.7(b). For the convenience of physical analysis, we have inset the schematic diagram illustrating the positional relationship between the excitation window U(t) (depicted as gray region with U=1,2,3,4 in insets) and the quantum dot energy level ϵd,n (color lines in insets). Different colors and line types in insets correspond to the parameters of the curves with matching colors and line types.

For linearly spaced ϵd,n [Fig.7(a)], Vg=1 (red solid lines) corresponds to ϵd=2,0,2,. In this case, the excitation window U(t) fully encompasses the ϵd=0 level as long as U>Γ, producing a distinct 2-unit plateau at interval of U[Γ,2Γ]. As U increases, higher-energy states, i.e., states with energies ±2,±4,, are sequentially activated. This process leads to the formation of plateaus of 6,10, at U[2+Γ,4Γ], [4+Γ,6Γ], , respectively. On the other hand, for Vg=0 (black lines), with ϵd,n=[3,1,1,3], pairs of symmetric levels (±1,±3,) enter the U(t) window as soon as U exceeds Γ+1,3,, yielding quantized plateaus of 4,8, surrounding U=2,4,. Intermediate gate voltages (e.g.,Vg=0.5, black red lines) shift the energy levels as ϵd,n=[2.5,0.5,1.5,], causing sequential stepwise plateaus of 2,4,6, at U=1,2,3,. At the transition points, i.e., U=1,2,3,, quantization is maintained regardless of any energy shift of ϵd,n, representing the robustness against the gate-induced energy shifts. These transition points satisfy the condition U=nΔ/2, where Δ=|ϵd,i+1ϵd,i|=2 defines the fixed level spacing.

As a contrast, the randomly spaced ϵd,n lead random arrangement of the quantized current. Taking Vg=1.0 as an example [the first inset of Fig.7(b)], ϵd,n=[2,0.5,2.2,]. When U<2, the excitation window U(t) can only encompasses the single level ϵd=0.5. As U passes the value of 2, the two quantum dot levels ϵd,n=2 and 2.2 successively enter the excitation window U(t). Consequently, the current J1 briefly stabilizes at the plateau of |J1|=4, then rapidly transitions to the plateau of |J1|=6. For Vg=0 and 0.3 [the third and second inset of Fig.7(b)], the energy levels are ϵd,n=[3.0,1.5,1.2,3.0] and ϵd,n=[2.7,1.2,1.5,3.3], respectively. In this case, the corresponding two current curves merge before the third energy level enters the excitation window, i.e., when U<2.5. This occurs because the first lowest energy level |ϵd,1|=1.2 and the second lowest energy level |ϵd,2|=1.5 enter excitation window U(t) simultaneously. Similar to the case of linearly spaced quantum dot energy levels, discrete multi-level causes step-like quantized current plateaus J1=2,4,6,. Moreover, the transition points of the quantum plateaus remain at U=Δ/2. Here Δ=|ϵiϵ0| defines the distance from any energy level ϵi within [U,U] to energy level ϵ0 that is closest to the Fermi level. For the first and second transition points, Δ=|1.51.2| and |3.01.2|, corresponding to U=1.35 and 2.1, respectively. Therefore, the distribution of the quantum dot energy levels can also be determined by examining the distribution of the step-like currents as a function of U.

In Fig.8, we present the fundamental Fourier component |J1| as a function of gate voltage Vg for varying excitation strengths U. For the convenience of physical analysis, we only consider the linear spaced energy levels. From Fig.8, it is evident that the remarkable stability of quantized J1 persists across wide Vg ranges at the transition points U=1,2,3 (solid curves). This highlights the robustness of quantum current against Vg-induced shifts in energy levels. The weaker the external field U is, the fewer the quantum plateaus are, but the stronger their robustness is. At transition points U=nΔ/2 (solid lines with n=1,2,3), n energy levels remain continuously within the bias window regardless of Vg, contributing stable quantized plateaus with maximum quantized value 2n. The quantized value is determined by the number of broadened discrete levels fully contained within the excitation window [U,U]. As Vg sweeps, quantum dot levels sequentially pass through the transport window. The number of broadened energy levels within [U,U] increases successively from 1 to 3, and then decreases back to 1. As a result, quantized value of J1 exhibits a step-like pattern in the shape of 2, 4, 6, 4, 2 for U=3 (thick red solid curve). Deviating from transition point, i.e., UnΔ/2 (dashed and dotted lines), the broadened quantum energy levels cannot be fully included in the excitation window, and the current cannot be quantized. This behavior highlights the delicate balance required between excitation strength U and particle level spacing Δ for achieving robust quantized current.

Fig.9 presents the modulus of the m-th Fourier component, |Jm|, plotted as a function of gate voltage Vg for odd indices m=1,3,5,,119, in a multi-level quantum dot system. Each panel corresponds to a distinct excitation strength U, with all Fourier components J1 to J119 systematically ordered by descending amplitude. Discrete current peaks and plateaus emerge as Vg varies, signifying the sequential entry of quantum energy levels into the excitation window [U,U]. For U values ranging from 0.1 to 1.0 [Fig.9(a)−(d)], the initial sharp peaks progressively broaden into quantized plateaus. These plateaus exhibit continuous expansion until they coalesce into a unified platform at U=1.0. This behavior confirms U=Δ/2 as a stability threshold, where current quantization persists robustly against Vg-induced shifts in energy levels. When U exceeds 1.0 [Fig.9(e)−(h)], a second quantum level enters the [U,U] energy window. Consequently, Jm transitions from discontinuous spikes to discrete plateaus, ultimately forming a dominant quantized plateau with magnitude 4 at the stable point U=2.0. Prior to the inclusion of the second energy level, a distinct first plateau with magnitude 2 is observed, consistent with the analysis in Fig.8. The red dashed lines in Fig.9(e)−(h) demarcate the position of this initial plateau with magnitude 2.

4 Beyond the wide-band limit

The preceding analysis within WBL provides an intuitive elucidation of quantum current generation mechanisms under periodic pulse excitation. Building upon this foundation, we now proceed to systematically investigate the generalized self-energy effects that transcend WBL. The linewidth function, defined as Γ(ϵ)2πρ(ϵ)|t(ϵ)|2, adopts a Lorentzian parametrization [76]

Γ(ϵ)=W2ϵ2+W2Γ,

where Γ represents the linewidth magnitude and W characterizes the bandwidth. In the asymptotic limit W, the energy dependence vanishes and Γ(ϵ)Γ, recovering the conventional WBL. The analytic structure of this function reveals two first-order poles in the complex plane at ϵ=±iW, with corresponding residues

Res±[Γ(E)]=±Γ2iW.

The spectral representation of the self-energies Σr/a(E) is fundamentally embodied by the linewidth function. This relationship enables the derivation of the retarded/advanced self-energy components through the integral representation

Σr/a(E)=dϵ2πΓ(ϵ)Eϵ±i0+=Γ2WE±iW.

The energy-dependent linewidth function Γ(ϵ) exhibits smooth spectral variations, ensuring that Σr/a,0(t1t) maintains dominant contributions from the time-localized region t1t. This temporal localization justifies the approximate relationship: Gn<Σai2Γ(0)Gn<(t,t). Considering the energy dependence of line width function Γ(ϵ), the current pulse formulation becomes

J(t)=ndϵ2πΓ(ϵ)[iGn<(t,t)2f(ϵ)ImAn(t,ϵ)],

where

Gn<(t,t)=dϵ2πfα(ϵ)iΓ(ϵ)|An(t,ϵ)|2.

Next, we solve the spectral function An(tϵ) through Eq. (7). The energy-dependent nature of the self-energy Σr(E) manifests in modifications to the retarded Green’s function Gnr(E) under external perturbations. For the single-terminal RC circuit configuration, the application of driving potential U at the lead electrode induces an equivalent energy shift U at the quantum dot through electrostatic coupling. This leads to time-periodic shift of the quantum level ϵ¯n=ϵd,nU, where the upper (lower) sign corresponds to the first (second) half-period. Consequently, the retarded Green’s function acquires the modified form

Gnr(E)=1Eϵ¯nΣr(E)=E+iW(Eϵ~d+)(Eϵ~d)

with two poles

ϵ~d±=ϵ¯niW±(ϵ¯n+iW)2+2ΓW2.

The corresponding residues of Gnr(E) are given by

Res±[Gnr(E)]=±ϵ~d±+iWϵ~d+ϵ~d.

The contour integration in Eq. (7) can be analytically evaluated through residue theorem by enclosing the two poles residing in the lower half complex plane. Importantly, while the Fermi−Dirac distribution in Eq. (11) corresponds to equilibrium conditions (U=0), the spectral function requires energy-axis adjustment to account for nonequilibrium driving. This is implemented through the variable substitution EEU applied respectively to the first and second half-period components. Consequently, the generalized spectral function beyond WBL takes the form

An(t,ϵ)=±Res±[Gnr(E)]Bn±(t,ϵ)

with

Bn±(t,ϵ)=1ϵϵ~d±+2Uei(ϵϵ~d±)t¯(ϵϵ~d±)(ϵϵ~d±2U)

for the first half cycle and

Bn±(t,ϵ)=1ϵϵ~d±2Uei(ϵϵ~d±)t¯(ϵϵ~d±)(ϵϵ~d±+2U)

for the second half cycle.

By implementing the modified spectral function from Eq. (12) within the current formulation of Eq. (5), we numerically evaluate the pulse current characteristics beyond the WBL. Fig.10(a) and (b) present the time-resolved current profiles J(t) during the first and second half-periods, respectively. Various curves correspond to distinct bandwidth parameters W. The cyan dashed curves demonstrate perfect current antisymmetry between half-periods under WBL (W). However, finite bandwidth effects induce significant current magnitude disparities between the first and the second half-periods, particularly pronounced when W is small. This departure from perfect antisymmetry manifests as deviations in quantization precision of |J1|, as quantified in the inset displaying |J1(U)| dependencies for different W. The parametric study reveals a systematic improvement in quantization fidelity with the increasing of bandwidth, culminating in complete WBL at W=5.0 (red-solid vs cyan-dashed curve coincidence). Notably, despite inducing subtle quantitative deviations, finite-bandwidth corrections maintain the integrity of core qualitative features: decaying time ατ, oscillating frequency U, and quantized current |J1|2ef.

5 Conclusions

In summary, based on a quantum RC circuit and employing the NEGF formalism, we theoretically investigate the dynamic response of single-electron emitters under square-wave driving potentials. The analytic solution reproduces the coherent emission-absorption cycles with J(t)±et¯/α. Remarkably, our theoretical results reveal the dynamic properties of the single-electron emitter, which include: (i) the characteristic time α of the decaying current J(t)(et/α) matches the quantum dot’s intrinsic coherence time τ; (ii) the decaying current oscillates with a frequency proportional to the external driving amplitude U; (iii) the quantized Fourier component |J1|=2ef is directly related to the half-quantized relaxation resistance Rq=h/(2e2), where both of them label the quantized dynamics of the AC transport through single-electron emitters. These findings provide new insights for understanding quantized AC transport phenomena.

References

[1]

B. J. van Wees , H. van Houten , C. W. J. Beenakker , J. G. Williamson , L. P. Kouwenhoven , D. van der Marel , and C. T. Foxon , Quantized conductance of point contacts in a two-dimensional electron gas, Phys. Rev. Lett. 60(9), 848 (1988)

[2]

D. A. Wharam , T. J. Thornton , R. Newbury , M. Pepper , H. Ahmed , J. E. F. Frost , D. G. Hasko , D. C. Peacock , D. A. Ritchie , and G. A. C. Jones , One-dimensional transport and the quantisation of the ballistic resistance, J. Phys. C 21(8), L209 (1988)

[3]

R. A. Webb , S. Washburn , C. P. Umbach , and R. B. Laibowitz , Observation of Aharonov−Bohm oscillations in normal-metal rings, Phys. Rev. Lett. 54(25), 2696 (1985)

[4]

K. Klitzing , G. Dorda , and M. Pepper , New method for high-accuracy determination of the fine-structure constant based on quantized Hall resistance, Phys. Rev. Lett. 45(6), 494 (1980)

[5]

D. C. Tsui , H. L. Stormer , and A. C. Gossard , Two-dimensional magnetotransport in the extreme quantum limit, Phys. Rev. Lett. 48(22), 1559 (1982)

[6]

W. Poirier and F. Schopfer , Resistance metrology based on the quantum Hall effect, Eur. Phys. J. Spec. Top. 172(1), 207 (2009)

[7]

W. Poirier , F. Schopfer , J. Guignard , O. Thévenot , and P. Gournay , Application of the quantum Hall effect to resistance metrology, C. R. Phys. 12(4), 347 (2011)

[8]

S. Djordjevic , R. Behr , and W. Poirier , A primary quantum current standard based on the Josephson and the quantum Hall effects, Nat. Commun. 16(1), 1447 (2025)

[9]

M. Büttiker , H. Thomas , and A. Prêtre , Mesoscopic capacitors, Phys. Lett. A 180(4−5), 364 (1993)

[10]

J. Gabelli , G. Fève , J. M. Berroir , B. Plaçais , A. Cavanna , B. Etienne , Y. Jin , and D. C. Glattli , Violation of Kirchhoff’s laws for a coherent RC circuit, Science 313(5786), 499 (2006)

[11]

J. Wang , B. Wang , and H. Guo , Quantum inductance and negative electrochemical capacitance at finite frequency in a two-plate quantum capacitor, Phys. Rev. B 75(15), 155336 (2007)

[12]

C. Bäuerle , D. Christian Glattli , T. Meunier , F. Portier , P. Roche , P. Roulleau , S. Takada , and X. Waintal , Coherent control of single electrons: A review of current progress, Rep. Prog. Phys. 81(5), 056503 (2018)

[13]

S. P. Giblin , M. Kataoka , J. D. Fletcher , P. See , T. J. B. M. Janssen , J. P. Griffiths , G. A. C. Jones , I. Farrer , and D. A. Ritchie , Towards a quantum representation of the ampere using single electron pumps, Nat. Commun. 3(1), 930 (2012)

[14]

G. Yamahata , K. Nishiguchi , and A. Fujiwara , Gigahertz single-trap electron pumps in silicon, Nat. Commun. 5(1), 5038 (2014)

[15]

H. Howe , M. Blumenthal , H. E. Beere , T. Mitchell , D. A. Ritchie , and M. Pepper , Single-electron pump with highly controllable plateaus, Appl. Phys. Lett. 119(15), 153102 (2021)

[16]

G. Yamahata , N. Johnson , and A. Fujiwara , Understanding the mechanism of tunable-barrier single-electron pumping: Mechanism crossover and optimal accuracy, Phys. Rev. B 103(24), 245306 (2021)

[17]

K. T. Wang , F. Xu , B. Wang , Y. Yu , and Y. Wei , Transport features of topological corner states in honeycomb lattice with multihollow structure, Front. Phys. (Beijing) 17(4), 43501 (2022)

[18]

G. Fève , A. Mahé , J. M. Berroir , T. Kontos , B. Plaçais , D. C. Glattli , A. Cavanna , B. Etienne , and Y. Jin , An on-demand coherent single-electron source, Science 316(5828), 1169 (2007)

[19]

M. Albert , C. Flindt , and M. Büttiker , Accuracy of the quantum capacitor as a single-electron source, Phys. Rev. B 82(4), 041407 (2010)

[20]

Y. Xing , Z. Yang , Q. F. Sun , and J. Wang , Coherent single-spin source based on topological insulators, Phys. Rev. B 90(7), 075435 (2014)

[21]

J. P. Pekola , O. P. Saira , V. F. Maisi , A. Kemppinen , M. Möttönen , Y. A. Pashkin , and D. V. Averin , Single-electron current sources: Toward a refined definition of the ampere, Rev. Mod. Phys. 85(4), 1421 (2013)

[22]

B. Kaestner and V. Kashcheyevs , Non-adiabatic quantized charge pumping with tunable-barrier quantum dots: A review of current progress, Rep. Prog. Phys. 78(10), 103901 (2015)

[23]

F. Stein , H. Scherer , T. Gerster , R. Behr , M. Götz , E. Pesel , C. Leicht , N. Ubbelohde , T. Weimann , K. Pierz , H. W. Schumacher , and F. Hohls , Robustness of singleelectron pumps at sub-ppm current accuracy level, Metrologia 54(1), S1 (2017)

[24]

R. Zhao , A. Rossi , S. P. Giblin , J. D. Fletcher , F. E. Hudson , M. Möttönen , M. Kataoka , and A. S. Dzurak , Thermal-error regime in high-accuracy gigahertz single-electron pumping, Phys. Rev. Appl. 8(4), 044021 (2017)

[25]

S. P. Giblin , A. Fujiwara , G. Yamahata , M. H. Bae , N. Kim , A. Rossi , M. Möttönen , and M. Kataoka , Evidence for universality of tunable-barrier electron pumps, Metrologia 56(4), 044004 (2019)

[26]

S. P Giblin , E. Mykkänen , A. Kemppinen , P. Immonen , A. Manninen , M. Jenei , M. Möttönen , G. Yamahata , A. Fujiwara , and M. Kataoka , Realisation of a quantum current standard at liquid helium temperature with sub-ppm reproducibility, Metrologia 57(2), 025013 (2020)

[27]

C. L. Degen , F. Reinhard , and P. Cappellaro , Quantum sensing, Rev. Mod. Phys. 89(3), 035002 (2017)

[28]

N. Johnson , J. D. Fletcher , D. A. Humphreys , P. See , J. P. Griffiths , G. A. C. Jones , I. Farrer , D. A. Ritchie , M. Pepper , T. J. B. M. Janssen , and M. Kataoka , Ultrafast voltage sampling using single-electron wavepackets, Appl. Phys. Lett. 110(10), 102105 (2017)

[29]

G. Yamahata , S. Ryu , N. Johnson , H. S. Sim , A. Fujiwara , and M. Kataoka , Picosecond coherent electron motion in a silicon single-electron source, Nat. Nanotechnol. 14(11), 1019 (2019)

[30]

M. Yamamoto , S. Takada , C. Bäuerle , K. Watanabe , A. D. Wieck , and S. Tarucha , Electrical control of a solid state flying qubit, Nat. Nanotechnol. 7(4), 247 (2012)

[31]

N. Ubbelohde , F. Hohls , V. Kashcheyevs , T. Wagner , L. Fricke , B. Kästner , K. Pierz , H. W. Schumacher , and R. J. Haug , Partitioning of on-demand electron pairs, Nat. Nanotechnol. 10(1), 46 (2015)

[32]

N. Johnson , C. Emary , S. Ryu , H. S. Sim , P. See , J. D. Fletcher , J. P. Griffiths , G. A. C. Jones , I. Farrer , D. A. Ritchie , M. Pepper , T. J. B. M. Janssen , and M. Kataoka , LO-phonon emission rate of hot electrons from an ondemand single-electron source in a GaAs/AlGaAs heterostructure, Phys. Rev. Lett. 121(13), 137703 (2018)

[33]

J. D. Fletcher , N. Johnson , E. Locane , P. See , J. P. Griffiths , I. Farrer , D. A. Ritchie , P. W. Brouwer , V. Kashcheyevs , and M. Kataoka , Continuous-variable tomography of solitary electrons, Nat. Commun. 10(1), 5298 (2019)

[34]

L. P. Kouwenhoven , A. T. Johnson , N. C. van der Vaart , C. J. P. M. Harmans , and C. T. Foxon , Quantized current in a quantum-dot turnstile using oscillating tunnel barriers, Phys. Rev. Lett. 67(12), 1626 (1991)

[35]

A. Fujiwara , N. M. Zimmerman , Y. Ono , and Y. Takahashi , Current quantization due to single-electron transfer in si-wire charge-coupled devices, Appl. Phys. Lett. 84(8), 1323 (2004)

[36]

M. D. Blumenthal , B. Kaestner , L. Li , S. Giblin , T. J. B. M. Janssen , M. Pepper , D. Anderson , G. Jones , and D. A. Ritchie , Gigahertz quantized charge pumping, Nat. Phys. 3(5), 343 (2007)

[37]

L. D. Contreras-Pulido , J. Splettstoesser , M. Governale , J. König , and M. Büttiker , Time scales in the dynamics of an interacting quantum dot, Phys. Rev. B 85(7), 075301 (2012)

[38]

V. I. Kleshch , V. Porshyn , A. S. Orekhov , A. S. Orekhov , D. Lützenkirchen-Hecht , and A. N. Obraztsov , Carbon single-electron point source controlled by Coulomb blockade, Carbon 171, 154 (2021)

[39]

J. M. Shilton , V. I. Talyanskii , M. Pepper , D. A. Ritchie , J. E. F. Frost , C. J. B. Ford , C. G. Smith , and G. A. C. Jones , High-frequency single-electron transport in a quasi-one-dimensional GaAs channel induced by surface acoustic waves, J. Phys. : Condens. Matter 8(38), L531 (1996)

[40]

V. I. Talyanskii , J. M. Shilton , M. Pepper , C. G. Smith , C. J. B. Ford , E. H. Linfield , D. A. Ritchie , and G. A. C. Jones , Single-electron transport in a one-dimensional channel by high-frequency surface acoustic waves, Phys. Rev. B 56(23), 15180 (1997)

[41]

S. Hermelin , S. Takada , M. Yamamoto , S. Tarucha , A. D. Wieck , L. Saminadayar , C. Bäuerle , and T. Meunier , Electrons surfing on a sound wave as a platform for quantum optics with flying electrons, Nature 477(7365), 435 (2011)

[42]

C. J. B. Ford , Transporting and manipulating single electrons in surface-acoustic-wave minima, Phys. Status Solidi B 254(3), 1600658 (2017)

[43]

L. S. Levitov , H. Lee , and G. B. Lesovik , Electron counting statistics and coherent states of electric current, J. Math. Phys. 37(10), 4845 (1996)

[44]

J. Keeling , I. Klich , and L. S. Levitov , Minimal excitation states of electrons in one-dimensional wires, Phys. Rev. Lett. 97(11), 116403 (2006)

[45]

J. Dubois , T. Jullien , C. Grenier , P. Degiovanni , P. Roulleau , and D. C. Glattli , Integer and fractional charge Lorentzian voltage pulses analyzed in the framework of photon-assisted shot noise, Phys. Rev. B 88(8), 085301 (2013)

[46]

B. Bertin-Johannet , J. Rech , T. Jonckheere , B. Grémaud , L. Raymond , and T. Martin , Microscopic theory of photo-assisted electronic transport in normal-metal/BCS-superconductor junctions, Phys. Rev. B 105(11), 115112 (2022)

[47]

M. R. Connolly , K. L. Chiu , S. P. Giblin , M. Kataoka , J. D. Fletcher , C. Chua , J. P. Griffiths , G. A. C. Jones , V. I. Fal’ko , C. G. Smith , and T. J. B. M. Janssen , Gigahertz quantized charge pumping in graphene quantum dots, Nat. Nanotechnol. 8(6), 417 (2013)

[48]

V. Siegle , C. W. Liang , B. Kaestner , H. W. Schumacher , F. Jessen , D. Koelle , R. Kleiner , and S. Roth , A molecular quantized charge pump, Nano Lett. 10(10), 3841 (2010)

[49]

C. Mora and K. Le Hur , Universal resistances of the quantum resistance–capacitance circuit, Nat. Phys. 6(9), 697 (2010)

[50]

M. Moskalets and M. Büttiker , Floquet scattering theory of quantum pumps, Phys. Rev. B 66(20), 205320 (2002)

[51]

M. Moskalets , P. Samuelsson , and M. Büttiker , Quantized dynamics of a coherent capacitor, Phys. Rev. Lett. 100(8), 086601 (2008)

[52]

P. Portugal , F. Brange , and C. Flindt , Heat pulses in electron quantum optics, Phys. Rev. Lett. 132(25), 256301 (2024)

[53]

S. Datta, Electronic Transport in Mesoscopic Systems, Cambridge University Press, Cambridge, 1995

[54]

H. Haug,A. P. Jauho, Quantum Kinetics in Transport and Optics of Semiconductors, Springer Series in Solid-State Sciences No. 123, Springer Berlin, Heidelberg, 2008

[55]

M. I. Alomar , J. S. Lim , and D. Sánchez , Coulomb-blockade effect in nonlinear mesoscopic capacitors, Phys. Rev. B 94(16), 165425 (2016)

[56]

C. Zhang , F. Xu , and J. Wang , Full counting statistics of phonon transport in disordered systems, Front. Phys. (Beijing) 16(3), 33502 (2021)

[57]

Z. Z. Yu , G. H. Xiong , and L. F. Zhang , A brief review of thermal transport in mesoscopic systems from nonequilibrium Green’s function approach, Front. Phys. (Beijing) 16(4), 43201 (2021)

[58]

J. Zeng , R. Xue , T. Hou , Y. Han , and Z. Qiao , Formation of topological domain walls and quantum transport properties of zero-line modes in commensurate bilayer graphene systems, Front. Phys. (Beijing) 17(6), 63503 (2022)

[59]

K. Fukuzawa , T. Kato , T. Jonckheere , J. Rech , and T. Martin , Minimal alternating current injection into carbon nanotubes, Phys. Rev. B 108(12), 125307 (2023)

[60]

L. Zhang , F. Xu , J. Chen , Y. Xing , and J. Wang , Quantum nonlinear ac transport theory at low frequency, New J. Phys. 25(11), 113006 (2023)

[61]

G. Li , F. Xu , and J. Wang , Universal behaviors of magnon-mediated spin transport in disordered nonmagnetic metal-ferromagnetic insulator heterostructures, Front. Phys. (Beijing) 18(3), 33310 (2023)

[62]

M. Wei , B. Wang , Y. Yu , F. Xu , and J. Wang , Nonlinear hall effect induced by internal coulomb interaction and phase relaxation process in a four-terminal system with time-reversal symmetry, Phys. Rev. B 105(11), 115411 (2022)

[63]

P. W. Brouwer , Scattering approach to parametric pumping, Phys. Rev. B 58(16), R10135 (1998)

[64]

Y. Blanter and M. Büttiker , Shot noise in mesoscopic conductors, Phys. Rep. 336(1−2), 1 (2000)

[65]

J. Wang and B. Wang , Quantization of adiabatic pumped charge in the presence of superconducting lead, Phys. Rev. B 65(15), 153311 (2002)

[66]

J. Gabelli , G. Fève , J. M. Berroir , and B. Plaçais , A coherent RC circuit, Rep. Prog. Phys. 75(12), 126504 (2012)

[67]

S. U. Cho , W. Park , B. K. Kim , M. Seo , D. T. Park , H. Choi , N. Kim , H. S. Sim , and M. H. Bae , One-lead single-electron source with charging energy, Nano Lett. 22(23), 9313 (2022)

[68]

Y. Xing , B. Wang , and J. Wang , First-principles investigation of dynamical properties of molecular devices under a steplike pulse, Phys. Rev. B 82(20), 205112 (2010)

[69]

T. Jonckheere , T. Stoll , J. Rech , and T. Martin , Realtime simulation of finite-frequency noise from a single-electron emitter, Phys. Rev. B 85(4), 045321 (2012)

[70]

S. E. Nigg , R. López , and M. Büttiker , Mesoscopic charge relaxation, Phys. Rev. Lett. 97(20), 206804 (2006)

[71]

Y. Liu , Y. Sun , X. Yan , B. Li , L. Wang , J. Li , J. Sun , Y. Guo , W. Liu , B. Hu , Q. Lin , F. Fan , and H. Shen , Realizing low voltage-driven bright and stable quantum dot light-emitting diodes through energy landscape flattening, Light Sci. Appl. 14(1), 50 (2025)

[72]

B. Harpt , J. Corrigan , N. Holman , P. Marciniec , D. Rosenberg , D. Yost , R. Das , R. Ruskov , C. Tahan , W. D. Oliver , R. McDermott , M. Friesen , and M. A. Eriksson , Ultra-dispersive resonator readout of a quantum-dot qubit using longitudinal coupling, npj Quantum Inf. 11, 5 (2025)

[73]

Q. F. Sun , J. Wang , and T. H. Lin , Photon-assisted Andreev tunneling through a mesoscopic hybrid system, Phys. Rev. B 59(20), 13126 (1999)

[74]

N. S. Wingreen , A. P. Jauho , and Y. Meir , Time-dependent transport through a mesoscopic structure, Phys. Rev. B 48(11), 8487 (1993)

[75]

F. Xu and J. Wang , Statistical properties of electrochemical capacitance in disordered mesoscopic capacitors, Phys. Rev. B 89(24), 245430 (2014)

[76]

J. Maciejko , J. Wang , and H. Guo , Time-dependent quantum transport far from equilibrium: An exact non-linear response theory, Phys. Rev. B 74(8), 085324 (2006)

RIGHTS & PERMISSIONS

Higher Education Press

AI Summary AI Mindmap
PDF (3395KB)

462

Accesses

0

Citation

Detail

Sections
Recommended

AI思维导图

/