Introduction

The Coulomb interaction in one-dimensional (1D) conductors plays an essential role in non-equilibrium transport characteristics1,2,3,4,5,6. Two distinct transport regimes, Tomonaga–Luttinger (TL) liquid in the low-energy regime and single-particle hot-electron transport in the high-energy regime, appear in chiral 1D channels of an integer quantum Hall (QH) region at Landau-level filling factor ν = 2 in AlGaAs/GaAs heterostructures7,8,9. In the low-energy regime, the energy dispersion near the Fermi energy can be approximated to be linear and the system is well described with collective excitation modes in the TL liquid model10,11. Arbitrary electronic excitation can be represented by spin and charge collective modes. The collective excitations propagate at different velocities for each mode, a phenomenon known as spin-charge separation12,13,14,15,16. Owing to the integrable nature of the TL model, the electronic excitation will be equilibrated into an ensemble of spin and charge excitations but not fully relaxed into a thermal equilibrium state17,18,19,20,21. This non-thermal metastable state (prethermalized state) survives even after traveling over a distance (>20 μm) much longer than required for the spin-charge separation (<0.1 μm)22,23,24. In these experiments, the initial state is excited with low energies (<0.2 meV from the Fermi energy) where the TL model holds well. The validity of the TL model, or the system’s integrability, is no longer guaranteed in the high-energy regime, where due to the deviation from the linear energy dispersion the Coulomb interaction induces weak electron–electron (e–e) scattering between the hot and cold electrons25,26.

In the high-energy regime, the hot electron travels for a long distance (>10 μm) without losing energy when the Coulomb interaction is suppressed, for example, by screening the interaction with metal on the surface27,28,29,30,31. Such ballistic hot electron transport can be seen at high energy (>50 meV) from the Fermi energy. In the intermediate energy region, the e–e scattering becomes significant, where the crossover between the TL liquid and single-particle hot electron transport is expected. However, the crossover regime remained veiled as the cold- and hot-electron dynamics have been studied separately by using different experimental schemes.

Here, we investigate the crossover dynamics from high-energy hot electrons to a non-thermal TL liquid, where the hot electrons lose their energy by exciting the TL liquid under the Coulomb interaction. The resulting state is unusual as warm and cold electrons coexist in the same channel even in the quasi-steady state. The emergence of the non-thermal many-body state is striking as the hot-electron energy is well beyond the low-energy regime for the TL liquid. The obtained non-thermal state is stable for long transport even after full relaxation of hot electrons, which represents the robustness of the integrable nature. We find that the final state depends on the initial hot-electron energy even after the equilibration, while energy-dependent dissipation is observed at higher energy. The crossover includes intra-channel scattering generating a hot spot in the spin-up channel and subsequent inter-channel scattering generating another hot spot in the spin-down channel. The characteristics are experimentally obtained by using a quantum point contact (QPC) as a spin-dependent bolometer to extract the energy-space trajectory of hot electrons and a quantum dot (QD) as a narrow-band spectrometer to evaluate the non-thermal TL liquid. The QH channel is a promising platform to study the interplay between single-particle and many-body physics.

Results and discussion

Crossover dynamics

We investigate the dynamics of hot and cold electrons in the two chiral edge channels at ν = 2 in a high magnetic field B. Figure 1a illustrates our finding on how the energy distribution functions f and f of up and down spins, respectively, change with traveling along the x-axis. With the spin-up and down channels equilibrated with cold electrons prepared at the base temperature TB at x < 0, we inject hot electrons with energy eVinj selectively to the spin-up channel at x = 0 [panel (i)]. For moderate eVinj, the Coulomb interaction induces intra-channel e–e scattering between the hot and cold spin-up electrons (e–e), and electron-hole plasma (the light magenta regions) is generated in the spin-up channel [panel (ii)]26. We neglect three-body e–e scattering throughout this paper for simplicity32. The hot-electron relaxation can be characterized by the energy decay rate per unit length, γ = −dEh/dx, for hot-electron’s energy Eh and the maximum energy exchange in a single scattering event, ΔE33,34. γ is obtained by measuring the energy(Eh)-space(x) trajectory of the hot electrons, and ΔE is roughly estimated from the spread of the distribution function, Wd, near the chemical potential μ. As we show later, both γ and ΔE increase rapidly as the hot electrons lose energy. The resultant sharp drop in the hot electron’s energy creates a hot spot at x = xH, where the spin-up electrons are maximally excited [panel (iii)]. In addition, the inter-channel e–e scattering between up and down spins (e–e) excites plasma (the dark magenta regions) and creates another hot spot at x = xH in the spin-down channel [panel (iv)]. The e–e scattering equilibrates the two channels further during transport. Importantly, once the electron-hole plasma relaxes to fit in the energy range where the TL physics governs, the plasma cannot equilibrate further and remains in a non-thermal metastable state in which warm and cold electrons coexist in the same channel [panel (v)]. Such crossover from hot electrons into a non-thermal TL liquid is the subject of this paper. Due to the coupling to the environment, the system finally reaches thermal equilibrium at TB after long transport [panel (vi)].

Fig. 1: Crossover dynamics of single-particle scattering processes and non-thermal Tomonaga–Luttinger (TL) liquid.
figure 1

a Schematic spatial evolution of energy distribution functions f and f for up- and down-spins, respectively. Hot electrons are injected into the spin-up channel at energy eVinj (i) and relax with intra-channel electron–electron scattering (e–e) and inter-channel scattering (e–e) (i)–(iv). Electron-hole plasma is excited in the channels, and hot spots appear at xH and xH in the spin-up and -down channels, respectively. The system ends up in a non-thermal metastable state (v) before reaching thermal equilibrium (vi). b Energy profiles for the lowest Landau levels (LLLs) for up and down spins. Intra-channel electron–electron scattering (e–e) is shown for a hot electron at energy Eh. c Intra-channel scattering (e–e) for hot and cold electrons at energies Eh and Ec, respectively, and final states at Eh − ε and Ec + ε with energy exchange ε. d Inter-channel scattering (e–e) and suppressed intra-channel scattering (e–e) with destructive interference for Eh < ΔE. e Formation of a TL liquid within the energy range ΔETL.

The crossover dynamics can be understood by considering microscopic processes in the edge states, where the Landau levels are bent upwards by the edge potential, as schematically shown in Fig. 1b. This energy(E)–space(y) diagram can be understood with the energy(E)–wavenumber(kx) dispersion, as the relation kx = −eBy/ holds in the QH regime under the Landau gauge7. Due to the nonlinearity of the dispersion, e–e scattering is suppressed as it cannot conserve both energy and momentum, particularly when the energy Eh and the energy exchange ε are large (Fig. 1c). In reality, the scattering is allowed by the presence of disorder that breaks the translational invariance and lifts the momentum conservation to some extent. This should determine the cut-off ΔE in the e–e scattering. For the concave dispersion relation, ΔE should increase with decreasing Eh. The Coulomb interaction increases with decreasing the distance at lower Eh. Both effects induce the rapid reduction of Eh with large γ and ΔE near x = xH.

When Eh < ΔE is reached, the system enters a new regime, where the scattering involves two competing processes with the swapped final states (the dashed lines with an arrow labeled e–e in Fig. 1d). They destructively interfere with each other, and thus the intra-channel scattering of the same spins should be suppressed at low Eh (<ΔE)34. When the intra-channel scattering is inefficient, the inter-channel scattering between different spins should dominate the relaxation. This region can be referred to as the inter-channel scattering regime. The hot electrons in the spin-up channel lose their energies by exciting an electron-hole plasma in the spin-down channel, and no competing process is present for scattering between different spins. The inter-channel scattering subsequently excites the spin-up channel and thus equilibrates the two channels in the end. As the spin-down channel is located on the inner side, where the edge potential is more gentle, the maximum energy exchange ΔE and the energy loss rate γ for different spins should be smaller than ΔE and γ, respectively.

The above single-particle picture fails when the hot electron is further relaxed within the energy range ΔETL of the TL regime (Eh ΔETL). The dispersion can be approximated to be linear, as shown in Fig. 1e. All electrons within the energy range ΔETL near the Fermi level interact with each other. Such interacting electrons can be understood with non-interacting bosonic excitations in the TL model10,20. The inter-channel interaction induces spin-charge separation, and the intra-channel interaction enhances the charge velocity.16 Once all electron-hole plasma enters the TL regime, no further transformation is expected in the plasma. Therefore, a non-thermal metastable state eventually shows up after stochastic processes. In the following experiment, the coexistence of warm and cold electrons remains even after long-distance transport. While the TL model is based on low-energy physics, such non-thermal states can emerge from high-energy hot electrons. This demonstrates the robustness of the prethermalized state against the excitation energy.

Electronic distribution functions of non-thermal states have been studied previously. When an initial state with a double-step distribution function is prepared by a QPC with bias voltage Vb and low transmission probability, the non-thermal state shows an arctangent distribution function in the low-energy limit at \(\left\vert E-\mu \right\vert \ll e{V}_{{{{{{{{\rm{b}}}}}}}}}\)21,23. Phenomenologically, the non-thermal states in the high-energy region can be approximated by a binary distribution function:

$$f\left(E\right)=\left(1-p\right){f}_{{{{{{{{\rm{FD}}}}}}}}}\left(E;{k}_{{{{{{{{\rm{B}}}}}}}}}{T}_{0}\right)+p{f}_{{{{{{{{\rm{FD}}}}}}}}}\left(E;{k}_{{{{{{{{\rm{B}}}}}}}}}{T}_{1}\right)$$
(1)

consisting of high- (T1) and low-temperature (T0, close to the base temperature TB) components with Fermi distribution \({f}_{{{{{{{{\rm{FD}}}}}}}}}\left(E;{k}_{{{{{{{{\rm{B}}}}}}}}}T\right)=1/\left[1+\exp \left(E/{k}_{{{{{{{{\rm{B}}}}}}}}}T\right)\right]\), which is valid for small fraction p (1)22. The binary form captures the exponential current profiles \(\sim {e}^{-E/{k}_{{{{{{{{\rm{B}}}}}}}}}{T}_{1}}\) observed in the low-energy regime22,23,24. As we do not know the theoretical form for hot-electron injection, we use this binary form to analyze the non-thermal states in this paper.

Experimental setup

We investigated the crossover dynamics with the device fabricated in a standard AlGaAs/GaAs heterostructure with an electron density of 2.5 × 1011 cm−2 and low-temperature mobility of about 106 cm2/Vs, as shown in Fig. 2a. The electron system is divided into three conductive regions that serve as the source (S), base (B), and drain (D) by applying large negative gate voltages VGi on gate i = 1, 2, 3,  . A perpendicular magnetic field B = 5.0 T is applied to set the electron system in the ν = 2 QH regime and form two parallel edge channels (the blue lines) in each region.

Fig. 2: Schematic measurement setup.
figure 2

a Device structure. By applying gate voltages on the gates (yellow), an injector point contact (IPC) and a detector quantum dot (QD) are formed between the source (S), base (B), and drain (D) regions. The blue parallel lines show edge channels at ν = 2. b Device configuration for a quantum point contact (QPC) detector. c Scanning electron micrograph with the false color of a control device with several IPC gates. d Schematic energy diagram of the measurement. Electrons are filled with the Fermi energy εF for up- and down-spin channels. The energy-space trajectory of hot electrons is shown by the red arrows. Electron-hole plasma near the chemical potential in B is analyzed by measuring \({I}_{\det }\) through the QD or QPC.

Hot electrons with up spin and energy eVinj (=1 − 100 meV) are injected from S to B, by applying voltage −Vinj to the ohmic contact of S and tuning the transmission of the injector point contact (IPC). The electronic distribution function of the up-spin channel near the chemical potential is investigated with a QD, and electron-hole plasma in up- and down-spin channel are studied by activating a QPC as shown in Fig. 2b. The QD or QPC is located at a distance L from the IPC, where several injection gates at different distances L = 1, 2, 5, 10, 20, and 30 μm were selectively activated (see the scanning electron micrograph of Fig. 2c). As shown in the schematic energy profile of Fig. 2d, the hot spot (xH) can be changed by tuning eVinj. The non-thermal state in the TL regime can be investigated at x = L > xH, and e–e scattering processes can be studied at x = L < xH. The current \({I}_{\det }\) through the QD or QPC is measured with a bias voltage VD on the detector (D) ohmic contact. The base current IB is monitored to ensure \({I}_{{{{{{{{\rm{inj}}}}}}}}}\simeq {I}_{\det }+{I}_{{{{{{{{\rm{B}}}}}}}}}\) (see Supplementary Fig. 1).

Energy-space trajectory of hot electrons

First, we investigate the hot-spot positions xH and xH by using QPC detection26,35. As shown in Fig. 3a, the QPC conductance G obtained without injecting hot electrons (by opening the IPC with Vinj = 0 and VGinj = 0) shows standard quantized conductances as a function of the gate voltage VG4. The onsets for spin-up and -down transport are seen at VG4 ~ −0.4 V and −0.2 V, respectively. Additional peaks at VG4 ~ −0.5 V are associated with parasitic impurity states, which play minor roles in the following analysis. The following QPC detection was performed at VD = 0. When hot spin-up electrons at energy eVinj are injected from the IPC for L = 1 μm, distinct features appear in the detector current profile \({I}_{\det }\), as shown in Fig. 3b. For the bottom trace at small eVinj = 2 meV, a single current step with a height comparable to Iinj = 1 nA is seen at the opening of the spin-up transport (VG4 ~ −0.4 V). This ensures no tunneling between spin-up and -down channels. With increasing eVinj, current peaks (rather than steps) show up for both spin-up and -down transports (in the pink stripes). As shown in the insets (i) and (ii) to Fig. 3a, each peak can be understood by considering the bolometric detection for each channel, where higher-energy electrons are transmitted but lower-energy holes are reflected26. Therefore, the peak current is proportional to the density of the electron-hole plasma in the respective channel, while the sensitivity for up spins may be greater than that for down spins. Since the hot spot is the point where the detected current is maximal, the strategy to determine the hot spot position (for a given L) is to look for the voltage Vinj that maximizes the detected current. The peak currents for up spins (VG4 = −0.42 V) and down spins (VG4 = −0.18 V) are plotted as a function of eVinj in Fig. 3c. The spin-up current is maximized at eVinj = 22 meV, which indicates that the position of the hot spot in the spin-up channel coincides with that of the QPC detector (xH = 1 μm for eVinj = 22 meV). Similarly, the spin-down current is maximized at eVinj = 14 meV, where the hot spot in the spin-down channel is considered to be located at the position of the QPC detector (xH = 1 μm for eVinj = 14 meV).

Fig. 3: Quantum point contact (QPC) detection of electron-hole plasma.
figure 3

a Conductance \(G={I}_{\det }/{V}_{{{{{{{{\rm{D}}}}}}}}}\) of the detector QPC measured with VD = 0.1 mV. Transmission and reflection of hot electrons (the blue arrow) and cold electrons and holes (the magenta arrows) are illustrated in the insets (i), (ii), and (iii). b \({I}_{\det }\) profiles obtained at various eVinj. VG4 in the horizontal axis was simultaneously swept with VG5 (=VG4). c The peak current of \({I}_{\det }\) for spin-up (red) and -down (blue) are plotted as a function of eVinj. The maximum peak current appears when the hot spot is located close to the detector.

We repeated similar measurements for various L, and the obtained hot-spot positions are summarized in Fig. 4a. In addition to the data obtained with the QPC (the circles), data obtained with the QD only for spin-up (described below) are also plotted by the squares. Both data for up spins agree well with each other within the experimental reproducibility. The smooth connection of the data points can be used to extract the energy-space trajectory of hot electrons. γ can be obtained from \({\gamma }_{\uparrow \uparrow }=d(e{V}_{{{{{{{{\rm{inj}}}}}}}}})/dL\), as the hot spot xH (=L) increases with eVinj at rate \({\gamma }_{\uparrow \uparrow }^{-1}\). Empirically, the power-law dependence in a form of \({\gamma }_{\uparrow \uparrow }=a{E}_{{{{{{{{\rm{h}}}}}}}}}^{-\lambda }\) for hot-electron energy Eh explains the data with parameters a and λ26. This suggests an energy(Eh)-space(x) trajectory:

$${E}_{{{{{{{{\rm{h}}}}}}}}}(x)={[{(e{V}_{{{{{{{{\rm{inj}}}}}}}}})}^{\lambda +1}-a(\lambda +1)x]}^{1/(\lambda +1)},$$
(2)

and the hot-spot condition:

$$e{V}_{{{{{{{{\rm{inj}}}}}}}}}={[a\left(\lambda +1\right){x}_{{{{{{{{\rm{H}}}}}}}}\uparrow }]}^{1/\left(\lambda +1\right)}.$$
(3)

The latter reproduces our data very well with λ = 1.8, as shown by the red curve in Fig. 4a. Corresponding energy loss rate \({\gamma }_{\uparrow \uparrow }\left({E}_{{{{{{{{\rm{h}}}}}}}}}\right)\) increases with decreasing Eh, as shown in Fig. 4b. The energy-space trajectory \({E}_{{{{{{{{\rm{h}}}}}}}}}\left(x\right)\) is shown for several eVinj values in Fig. 4c. As the energies of incident electrons should be distributed at around eVinj mostly due to the energy-dependent tunneling probability of the IPC, possible dispersion of the trajectories for ±1 meV distribution is shown by pink stripes. Notice that hot-electron relaxation proceeds rapidly just before the hot spot xH, which can be tuned with Vinj. This is useful for investigating the decay length of non-thermal states, as one can change the distance D between the hot spot and the detector.

Fig. 4: Energy-space trajectory of hot electrons.
figure 4

a The hot-spot conditions for up (red) and down (blue) spins measured with quantum point contact (QPC) (circles) and quantum dot (QD) (squares) detectors. The error bar shows the variations due to the injector and detector conditions. The solid lines are fitted to the data with λ = 1.8 (the uncertainty between 1.6 and 2.2) and  = 0.9 μm. The inset shows the data in the double logarithmic plot. b The estimated energy loss rate γ by assuming the power dependence of hot-electron energy Eh. c The estimated energy(Eh)-space(x) trajectory of hot electrons for eVinj = 40, 60, and 80 meV. Each stripe represents the spread associated with variation of eVinj by ±1 meV.

As for the spin-down hot spot xH shown by the blue circles in Fig. 4a, the L dependence can be reproduced by the blue curve which is obtained just by shifting the red curve horizontally by  = 0.9 μm. This suggests that finite length , independent of eVinj, is required for exciting spin-down electrons through the inter-channel interaction. This is much longer than the length lSC required for spin-charge separation in the TL regime. Similar devices show lSC 0.18 μm at eVinj = 0.6 meV, and lSC decreases with increasing eVinj (lSC 1/Vinj)23. The difference ( > lSC) indicates the existence of the stochastic inter-channel e–e scattering regime between xH and xH.

The transition between intra- and inter-channel scattering regimes can be confirmed from the data in Fig. 3c. In the range of eVinj between 30 and 70 meV for L < xH, spin-up electrons are highly excited (\({I}_{\det } > {I}_{{{{{{{{\rm{inj}}}}}}}}}\)) but no spin-down electrons are excited (\({I}_{\det }\simeq {I}_{{{{{{{{\rm{inj}}}}}}}}}\)) because the intra-channel scattering is dominant. In contrast, both signals become large at eVinj < 25 meV and change in a similar manner at eVinj < 15 meV (L > xH), where the inter-channel interaction equilibrates the two channels. The sequence of the intra-channel scattering regime [from (i) to (iii) in Fig. 1a] followed by the inter-channel scattering regime [(iii)–(iv)] is justified with the result.

QD spectroscopy

We investigate electronic excitation in the spin-up channel by using the QD detector23,36. Coulomb diamond characteristics of the QD are shown in Supplementary Fig. 2. Figure 5a shows the Coulomb blockade (CB) oscillations obtained with VD = 0.2 mV under hot-electron injection at L = 10 μm. The horizontal axis ΔVG5 is the relative voltage of VG5 from the left onset of the right CB peak. The width of the CB peaks in the bottom trace measures the transport window of eVD. When hot electrons are injected at various eVinj and fixed Iinj = 1.8 pA, excess current appears in the entire CB region, as highlighted by the pink color. This current measures the electrons excited above the QD level. For instance, the current at ΔVG5 = −19 mV (the dashed line) becomes maximum at eVinj = 62 meV, as shown in Fig. 5b. This measures the spin-up hot-spot condition, where the hot electrons injected at eVinj = 62 meV just relaxed to the Fermi energy at L = 10 μm. This coincides with the QPC data, as shown by the squares in Fig. 4a. In the following analysis, we use the hot-electron trajectory given by Eq. (2) at λ = 1.8, while the parameter a for each geometry (L) is determined from eVinj that maximizes \({I}_{\det }\) by using Eq. (3).

Fig. 5: Quantum dot (QD) detection of electron-hole plasma.
figure 5

a Coulomb oscillations of the QD observed in detector current \({I}_{\det }\) as a function of relative gate voltage ΔVG5 under hot-electron injection at various injection voltage Vinj. The excess current is highlighted by the pink color. The inset shows the energy diagram of the QD. b \({I}_{\det }\) measured in the deep Coulomb blockade (CB) region at ΔVG5 = −19 mV (the dashed line in a). The left inset shows the energy diagram for eVinj 60 meV, where the hot electrons relax before reaching the QD. The right inset shows the diagram for eVinj 60 meV, where the hot electrons pass over the QD.

The broad current profile in Fig. 5a suggests that the electrons are excited well above the charging energy of the QD (U 1.5 meV) as well as the single-particle energy spacing of the QD (\(\bar{\Delta }\simeq\) 0.2 meV on average). The current decreases with decreasing ΔVG5 reflecting the electron energy distribution function \(f\left(E\right)\) of the channel. To evaluate \(f\left(E\right)\), the current profiles are plotted in the logarithmic scale in Fig. 6a for eVinj ≤ 60 meV, where the hot electrons are fully relaxed before reaching the QD detector (xH < L). The current profile in the CB region shows exponential dependence \({I}_{\det }\propto \exp \left(\alpha \Delta {V}_{{{{{{{{\rm{G5}}}}}}}}}/{W}_{{{{{{{{\rm{c}}}}}}}}}\right)\) as shown by the dashed lines. Here, α 0.044e is the lever arm factor to convert ΔVG5 into the electrochemical potential μN = − αΔVG5 of the N-electron QD, and Wc is the characteristic energy of the current profile. For eVinj > 62 meV, the hot electrons pass over the QD potential (xH > L), as shown in the inset to Fig. 6b. Noting that all the hot electrons are eventually absorbed by the ammeter, we plot in Fig. 6b \({I}_{\det }-{I}_{{{{{{{{\rm{inj}}}}}}}}}\) as the net current in the logarithmic scale. The exponential profile, \(\propto \exp \left(\alpha \Delta {V}_{{{{{{{{\rm{G5}}}}}}}}}/{W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\right)\), in the CB region is characterized by \({W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\). We confirmed that Wc and \({W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\) do not change significantly with Iinj, as shown in Fig. 6c, d, as well as N (partially seen at ΔVG5 < −30 mV in Fig. 6a, b). Some other data obtained at different L and Iinj are shown in Supplementary Fig. 3.

Fig. 6: Quantum dot (QD) spectroscopy for energy distribution functions.
figure 6

a Detector current \({I}_{\det }\) as a function of relative gate voltage ΔVG5 at various injection energy eVinj ≤ 60 meV for studying non-thermal metastable states after long transport. b Excess current \({I}_{\det }-{I}_{{{{{{{{\rm{inj}}}}}}}}}\) for eVinj ≥ 62 meV for studying non-equilibrium states in the course of electron–electron scattering. The dashed lines, \(\propto \exp \left(\alpha {V}_{{{{{{{{\rm{G5}}}}}}}}}/{W}_{{{{{{{{\rm{c}}}}}}}}}\right)\) in (a) and \(\propto \exp \left(\alpha {V}_{{{{{{{{\rm{G5}}}}}}}}}/{W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\right)\) in (b), are fitted to the data. The current tails in (a) and (b) are characterized by cut-off energies Wc and \({W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\), respectively. The insets show the energy diagrams around the QD. c \({I}_{\det }\) for several injection current Iinj at eVinj = 20 meV. No significant change in the slope (the dashed lines) is seen. d Injection current Iinj dependence of cut-off energies Wc (red) and \({W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\) (blue).

To analyze the exponential tail in the CB region, we rely on the orthodox CB theory with a continuous density of QD states37, as all features associated with the excited states of the QD are smeared out by the broad profile. We also assume the QD is relaxed to the ground state before accepting a single hot-electron. Under these assumptions, the detector current yields:

$${I}_{\det }=\int\nolimits_{{\mu }_{N}}^{\infty }\frac{{G}_{{{{{{{{\rm{QD}}}}}}}}}}{e}f\left(E\right)dE$$
(4)

with GQD the tunneling conductance (see “Methods”). If \({G}_{{{{{{{{\rm{QD}}}}}}}}}\propto {e}^{E/{W}_{{{{{{{{\rm{t}}}}}}}}}}\) increases exponentially with energy E at characteristic energy Wt of the tunneling, the model suggests exponential energy dependence of the distribution function \(f\left(E\right)\propto \exp \left(-E/{W}_{{{{{{{{\rm{d}}}}}}}}}\right)\) with \({W}_{{{{{{{{\rm{d}}}}}}}}}^{-1}={W}_{{{{{{{{\rm{t}}}}}}}}}^{-1}+{W}_{{{{{{{{\rm{c}}}}}}}}}^{-1}\). While we do not know Wt of our device, the spread of the current tail (Wc) should reflect the spread of the distribution function (Wd) when Wc is not too large.

Non-thermal metastable state

The state after equilibration at x > xH can be studied with the data in Fig. 6a. The coexistence of the small exponential current tail with large Wc and the large CB current peak with a sharp onset associated with cold electrons manifests the non-thermal metastable state of the TL liquid22,23,24. In the present case, the non-thermal states are analyzed with the binary form of Eq. (1). The low temperature (T0) is always close to the base temperature (TB), as the CB peak remains sharp at any Vinj in Fig. 6a. The high temperature (T1) is related to the spread of the current tail (kBT1 ~ Wc). Significantly different temperatures (T1T0) are seen at higher Vinj. These current profiles are qualitatively the same as those reported for small excitation energies (10–100 μeV) where the low-energy TL physics governs. Surprisingly, low-energy physics emerges from the high-energy source (1–60 meV). Note that the non-thermal metastable state is distinct from the ordinary non-equilibrium state often seen in higher dimensions38,39.

As several QD levels are involved in the transport for the highly non-equilibrium state, p in Eq. (1) cannot be estimated from the current profile. Instead, p can be obtained from the electrical power Pinj = IinjVinj injected from the IPC. It is known that a single edge channel carries heat current \(J=\frac{{\pi }^{2}}{6h}{\left({k}_{{{{{{{{\rm{B}}}}}}}}}T\right)}^{2}\) if the electrons are in thermal equilibrium at temperature T40. Therefore, if the spin-up and -down channels were thermally equilibrated with Pinj, the thermal energy would increase to \({k}_{{{{{{{{\rm{B}}}}}}}}}{T}_{{{{{{{{\rm{th}}}}}}}}}=\sqrt{\frac{3h}{{\pi }^{2}}{P}_{{{{{{{{\rm{inj}}}}}}}}}}\) by neglecting the thermal energy of the base temperature (TB). This yields kBTth = 10 μeV for Pinj = 13 fW (Iinj = 1.8 pA and Vinj = 7 mV), which is much smaller than Wc = 0.3 meV at Vinj = 7 mV for the third trace from the bottom in Fig. 6a. The heat conservation suggests that only a small fraction, about \(p={\left({k}_{{{{{{{{\rm{B}}}}}}}}}{T}_{{{{{{{{\rm{th}}}}}}}}}/{W}_{{{{{{{{\rm{c}}}}}}}}}\right)}^{2}\simeq\) 10−3 of electrons, is highly excited to the high-temperature T1. Such an unusual energy distribution is a consequence of the crossover from hot electrons.

The exponential current tail in Fig. 6a starts to develop from eVinj 2 meV, and the corresponding Wc value rapidly increases with eVinj. The Wc values obtained at various eVinj and L are summarized in Fig. 7. First, we focus on the low eVinj (8 meV) regime for the shortest L = 1 μm, where Wc increases linearly with eVinj (the dashed line). As the hot electron relaxes within a short propagation length (xH < 0.1 μm as shown in the auxiliary scale) in this case, the distribution function should measure the non-thermal metastable state after traveling of ~1 μm. Actually, the slope Wc/eVinj = 0.2 is comparable to the values obtained in the low-energy limit of the TL regime (Wc/eVinj = 0.17 for Wc = 34 μeV at eVinj = 200 μeV with QPC injection in ref. 23, Wc/eVinj = 0.22–0.25 for Wc = 14–17 μeV at eVinj = 63 μeV with QD injection in ref. 24, and Wc/eVinj = 0.17 for Wc = 20–35 μeV at eVinj = 100–200 μeV with indirect QPC injection in ref. 22). Our experiment confirms that the same ratio (Wc/eVinj) holds up to Wc 1 meV, which is more than ten times greater than the values in the previous works. The linearity implies that the injection energy fits in the energy range of the TL liquid without experiencing e–e scattering. All electrons within the range interact with each other to form the collective plasmon mode, as discussed in Fig. 1e. The memory of the initial state at eVinj remains in the broad electronic distribution function with Wd (Wc 0.2eVinj). This can be understood with the integrable nature of holding many conserved quantities.

Fig. 7: The cut-off energy Wc for the non-thermal states.
figure 7

The dashed line shows the linear relation, Wc 0.2eVinj, obtained in the low-energy limit of Tomonaga–Luttinger (TL) liquids. The inset shows the energy-space trajectory of hot electrons. The hot-spot position, xH, obtained with Eq. (3) at λ = 1.8 is shown in the upper scale.

However, Wc deviates from the linear dependence to lower values at eVinj 8 meV for L = 1 μm, as seen in Fig. 7. Wc obtained with longer L may follow the linear dependence at small eVinj 1 meV and becomes substantially smaller at larger eVinj. The reduction in Wc might originate from the dissipation41,42 or the departure from the TL regime. To characterize the decay of non-thermal states, the Wc data are plotted as a function of the distance D = L − xH from the hot spot to the detector, as shown in Fig. 8. The corresponding eVinj values are shown in the auxiliary scale for each L. Each trace features a gradual reduction of Wc (except for L = 1 μm) at eVinj > 50 meV followed by a sudden reduction at eVinj < 10 meV (DL). Both features can be explained consistently with the energy-dependent dissipation during traveling.

Fig. 8: Decay of non-thermal metastable states.
figure 8

The characteristic energy Wc is plotted as a function of the distance D from the hot spot to the quantum dot (QD). D is obtained from Eq. (3) at λ = 1.8. The inset shows the energy-space trajectory. The eVinj value for each L is shown in the respective auxiliary scale. The decay of Wc is shown by the short dashed lines connecting the data points obtained with the same eVinj (≤10 meV) and the long dashed line as a guide to the eye for the data at eVinj 50 meV.

For eVinj < 10 meV, the electron-hole excitation at the hot spot is developing with increasing eVinj, while D (L) is almost constant for each L. Therefore, the dissipation can be studied by varying L for a given eVinj, as shown by the two dotted lines connecting the data points for eVinj = 5 and 10 meV. The steeper slope at higher Wc suggests increased dissipation at higher excitation energy.

In contrast, the excitation at the hot spot does not increase further for eVinj > 30 meV. As seen in the energy-space trajectory of Fig. 4c, the excess energy beyond 20–30 meV is lost during the long-distance transport with the remaining 20–30 meV used to build the hot spot with the highest excitation energy. Therefore, the eVinj dependence at eVinj > 30 meV in Fig. 8 should be understood as representing the D dependence of Wc. The overall feature at eVinj > 50 meV follows a monotonic reduction of Wc with D, as shown by the long dashed line. This slope is comparable to the slope of the short-dashed lines if the data in the same range of Wc are compared. The data shows how the electronic distribution function changes with transport. The characteristic energy Wd (Wc) decays fast if Wd is large (the decay length d ~ 5 μm at Wc 1 meV), but the decay length becomes longer for smaller Wd (~30 μm at 0.2 meV). This is comparable to d ~ 20 μm obtained from the low-energy experiment at Wc = 34 μeV23. The energy dependent d may be explained by energy-dependent dissipation associated with the characteristics of the environment. If the reduction signifies the departure of the TL regime, the energy range for the TL physics could be ΔETL ~ 1 meV in our device.

e–e scattering

The electron-hole plasma generated in the course of the e–e scattering was analyzed with the current profile, \({I}_{\det }-{I}_{{{{{{{{\rm{inj}}}}}}}}}\), at xH > L, in Fig. 6b. The data indicates that \({W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\) (the inverse of the slope) decreases with increasing eVinj. Considering that the energy loss rate γ increases as Eh decreases during the transport from the IPC to the detector (see Fig. 4b), it is likely that the obtained \({W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\) value reflects most sensitively the e–e scattering near the detector QD. From this point of view, we plot in Fig. 9\({W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\) as a function of \({E}_{{{{{{{{\rm{h}}}}}}}}}\left(L\right)\), the hot-electron energy at the detector position. The two data sets for L = 1 and 10 μm show similar dependency, which justifies this approach. The data show that \({W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\) decreases with increasing Eh, as shown by the dashed line. Namely, the electron-hole plasma is highly excited in the vicinity of the hot spot, as illustrated in the inset.

Fig. 9: The characteristic energy \({W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\) for the electron–electron scattering.
figure 9

The hot-electron energy \({E}_{{{{{{{{\rm{h}}}}}}}}}\left(L\right)\) at the quantum dot (QD) detector is obtained from Eq. (2) at λ = 1.8. The eVinj value for each L is shown in the respective auxiliary scale. The dashed line is a guide to the eye by using a function proportional to \({E}_{{{{{{{{\rm{h}}}}}}}}}^{-1.5}\). The inset shows the energy-space trajectory and buildup of electron-hole plasma.

The physical meaning of \({W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\) can be understood by considering the electron-hole generation process. Suppose that an intra-channel e–e scattering between a hot electron with energy Eh and a cold electron with Ec (<0) generates an electron-hole pair with a hole at Ec and an electron near the Fermi energy [at Ec + ε (>0)] by the energy exchange ε, as shown in Fig. 1c. Scattering with various combinations of Ec and ε induces the plasma. Here, ε is bounded by the momentum conservation. For a given Eh, we assume that the scattering probability \(P\left({E}_{{{{{{{{\rm{c}}}}}}}}},\varepsilon ;{E}_{{{{{{{{\rm{h}}}}}}}}}\right)\propto {e}^{-\varepsilon /\Delta {E}_{\uparrow \uparrow }}\) for Ec and ε depends only on ε with the upper bound ΔE that depends on Eh. For the present case, we are interested in the electronic energy distribution function \(f\left(E\right)\) above the Fermi energy (E > 0). We further assume that \(f\left(E\right)\) is determined by the e–e scattering in the vicinity of the detector by neglecting the scattering that occurs in the upstream region. Then, \(f\left(E\right)\) is given by the instantaneous e–e scattering with Eh at x = L as:

$$f\left(E\right)\propto \int\nolimits_{-\infty }^{0}P\left({E}_{{{{{{{{\rm{c}}}}}}}}},\varepsilon ;{E}_{{{{{{{{\rm{h}}}}}}}}}\right)d{E}_{{{{{{{{\rm{c}}}}}}}}}\propto {e}^{-E/\Delta {E}_{\uparrow \uparrow }},$$
(5)

where E = Ec + ε was used in the integration. Therefore, provided that \({W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\) measures the spread Wd of the distribution function, the obtained \({W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\) can be understood as the energy cut-off ΔE of the intra-channel e–e scattering.

The monotonic reduction of \({W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\) with Eh(L) in Fig. 9 is consistent with the concave dispersion in Fig. 1b, as explained above. In this way, the energy loss of hot electrons (Eh > 20 meV) and excitation of cold electrons (Eh < 5 meV) are consistently explained with the intra-channel e–e scattering. It should be noted that the energy dependency (larger \({W}_{{{{{{{{\rm{c}}}}}}}}}^{{\prime} }\) for smaller Eh) of e–e scattering in Fig. 9 is opposite to that (larger Wc for larger eVinj) for TL states in Fig. 7.

Conclusions

We have investigated how high-energy hot electrons lose their energy by exciting cold electrons. A QPC was used as a broad-band spin-dependent bolometer to extract the energy-space trajectory of hot electrons. A QD was employed as a narrow-band spectrometer to detect non-equilibrium electronic excitation near the Fermi energy. The hot electrons lose their energy with the e–e scattering and eventually create spin-dependent hot spots. The cold electrons are maximally excited at the hot spot and equilibrate while traveling the channels. Interestingly, the system always ends up with a non-thermal metastable state without fully relaxing into thermal equilibrium, even when the initial hot-electron energy is well beyond the TL regime. This demonstrates the robustness of non-thermal states against the excitation energy. The non-thermal state depends on the excitation energy even after traveling 20 μm. This suggests that the final state remembers the initial state even after the equilibration, in consistency with the integrable nature with many conserved quantities. In addition, the non-equilibrium state generated with the e–e scattering is in-situ investigated to evaluate the maximum energy exchange in the scattering. The analysis implies the significance of inter-channel scattering with different spins as well as intra-channel scattering with the same spins. The experiment provides a novel crossover from single particles to a non-equilibrium many-body state. The scheme can be used for studying non-linear TL liquids in the presence of non-linear dispersion43. The robustness of non-thermal states is attractive for transporting information and energy for long distances44,45,46.

Methods

Hot-electron injection

We use custom-made voltage source and ammeter to inject hot electrons. A large input impedance of 100 kΩ in the ammeter prevents unwanted large current during adjustments. VGinj is precisely adjusted by a software to obtain the desired Iinj value (1 pA–1 nA) within a few % error before measuring \({I}_{\det }\) at each condition (Vinj, VG4, etc.). VGinj depends strongly on Vinj (typically, from VGinj −1.1 V at Vinj = 5 mV to VGinj −1.9 V at Vinj = 80 mV with an almost linear relation in between) for Iinj = 1 nA. Slightly more negative VGinj by  −0.01 V was needed for Iinj = 10 pA. While the location of the injector should be pushed toward the other gate G1 with more negative VGinj, we neglected possible change in transport length L. Iinj is kept sufficiently low, 1–10 pA for the QD detection to minimize the heating inside the QD and ~1 nA for the QPC detection for better visibility.

As the e–e scattering is sensitive to the distance from the channel to the gate metal, we fixed the gate voltages defining the channel (VG1 = VG3 = −0.6 V and VG2 = −0.45 V) throughout this paper. These gate voltages were chosen in such a way that the edge potential becomes gentle enough to suppress unwanted LO phonon emission while high enough to confine hot electrons26,35,47. While Vinj (1–100 meV) is much larger than the cyclotron energy of about 9 meV at 5 T, we did not observe any tunneling transition to higher Landau levels in the present conditions.

QPC detection

We focused on the current peaks appearing at the onsets of the conductance steps of the QPC, as described in the main text. Additional data that support our observation are shown in Supplementary Note 1. However, one can investigate the hot-electron energy from the current step appearing at VG4 < −0.8 V and eVinj > 30 meV in Fig. 3b. This current step appears when the potential barrier height coincides with the hot-electron energy, as illustrated in the inset (iii) to Fig. 3a, and disappears at eVinj 25 meV in consistency with the hot spot condition obtained above. In our previous report, the hot-electron energy is obtained from VG4 at the current step by using a conversion factor determined from the LO phonon replicas of the current step26,35. The obtained λ in the range between 1 and 2 is consistent with the present value (λ = 1.8), while λ may depend on the device and the side gate voltage. However, the hot-electron energy cannot be obtained for the present case, as we suppressed LO phonon emission. Instead, the whole hot-electron energy is used for generating the electron-hole plasma, which is used to estimate p in this study.

QD detection

The QD characteristics (U 1.5 meV, \(\bar{\Delta }\simeq\) 0.2 meV, and α 0.044e) were obtained from standard Coulomb diamond measurements at Vinj = 0 and VGinj = 0. The base electron temperature TB 110 mK was also obtained from the onset of the CB peak. The backaction from the change in VGinj and Vinj to the detector conditions is present. While this is negligible for the QPC detection, the backaction to the QD is significant, as the CB peaks were shifted by ~30 mV in VG5 for L = 1 μm and ~8 mV for L = 10 μm when Vinj was changed from 1 to 90 mV under the adjustment of VGinj for Iinj = 5 pA. The relative value ΔVG5 from the CB peak is used in the main text. Additional data that support our observation are shown in Supplementary Note 2.

Analysis of the current profiles

We use the orthodox CB theory to analyze the broad current profile in the CB region at μN > μ > μD (see inset to Fig. 5a). As the current is sufficiently low, the dot should remain close to the base temperature. In this case, the tunneling rate from the edge channel of interest to the dot is given by \({\Gamma }_{{{{{{{{\rm{dc}}}}}}}}}\simeq \int\nolimits_{{\mu }_{N}}^{\infty }\frac{1}{{e}^{2}}{G}_{{{{{{{{\rm{L}}}}}}}}} \, f\left(E\right)dE\) for the distribution function \(f\left(E\right)\), where the tunneling conductance GL may depends on energy. By assuming TB = 0 for simplicity, the tunneling rates from the dot to the channel and drain read \({\Gamma }_{{{{{{{{\rm{cd}}}}}}}}}\simeq \int\nolimits_{\mu }^{{\mu }_{N}}\frac{1}{{e}^{2}}{G}_{{{{{{{{\rm{L}}}}}}}}}dE\) and \({\Gamma }_{{{{{{{{\rm{Dd}}}}}}}}}\simeq \int\nolimits_{{\mu }_{{{{{{{{\rm{D}}}}}}}}}}^{{\mu }_{N}}\frac{1}{{e}^{2}}{G}_{{{{{{{{\rm{R}}}}}}}}}dE\), where GL and GR are tunneling conductances of the respective barriers. By solving the master equation, we obtain the steady current \(I=e{\Gamma }_{{{{{{{{\rm{dc}}}}}}}}}{\Gamma }_{{{{{{{{\rm{Dd}}}}}}}}}/\left({\Gamma }_{{{{{{{{\rm{dc}}}}}}}}}+{\Gamma }_{{{{{{{{\rm{Dd}}}}}}}}}+{\Gamma }_{{{{{{{{\rm{cd}}}}}}}}}\right)\). By assuming energy-dependent conductances \({G}_{{{{{{{{\rm{L}}}}}}}}}\propto {e}^{E/{W}_{{{{{{{{\rm{t}}}}}}}}}}\) and \({G}_{{{{{{{{\rm{R}}}}}}}}}\propto {e}^{E/{W}_{{{{{{{{\rm{t}}}}}}}}}}\) with the same characteristic energy Wt for the tunneling, we obtain \(I\simeq \int\nolimits_{{\mu }_{N}}^{\infty }\frac{1}{e}{G}_{{{{{{{{\rm{QD}}}}}}}}}f\left(E\right)dE\) with \({G}_{{{{{{{{\rm{QD}}}}}}}}}\propto {e}^{E/{W}_{{{{{{{{\rm{t}}}}}}}}}}\) by neglecting factors that weakly depend on energy. This form was used in the analysis of distribution functions.