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

Dynamic Response of Wigner Crystals

Lili Zhao International Center for Quantum Materials, Peking University, Haidian, Beijing 100871, China    Wenlu Lin International Center for Quantum Materials, Peking University, Haidian, Beijing 100871, China    Yoon Jang Chung Department of Electrical Engineering, Princeton University, Princeton, New Jersey 08544, USA    Adbhut Gupta Department of Electrical Engineering, Princeton University, Princeton, New Jersey 08544, USA    Kirk W. Baldwin Department of Electrical Engineering, Princeton University, Princeton, New Jersey 08544, USA    Loren N. Pfeiffer Department of Electrical Engineering, Princeton University, Princeton, New Jersey 08544, USA    Yang Liu [email protected] International Center for Quantum Materials, Peking University, Haidian, Beijing 100871, China
Abstract

The Wigner crystal, an ordered array of electrons, is one of the very first proposed many-body phases stabilized by the electron-electron interaction. This electron solid phase has been reported in ultra-clean two-dimensional electron systems at extremely low temperatures, where the Coulomb interaction dominants over the kinetic energy, disorder potential and thermal fluctuation. We closely examine this quantum phase with capacitance measurements where the device length-scale is comparable with the crystal’s correlation length. The extraordinarily high performance of our technique makes it possible to quantitatively study the dynamic response of the Wigner crystal within the single crystal regime. Our result will greatly boost the study of this inscrutable electron solid.

Interacting two-dimensional electron system (2DES) subjected to high perpendicular magnetic fields (BB) and cooled to low temperatures exhibits a plethora of exotic states Jain (2007). The Wigner crystal (WC) Wigner (1934) terminates the sequence of fractional quantum Hall states at very small landau level filling factor Jiang et al. (1990); Goldman et al. (1990); Li et al. (1991); Santos et al. (1992); Sajoto et al. (1993); Pan et al. (2002); Maryenko et al. (2018); Hossain et al. (2020); Chung et al. (2022); Lozovik and Yudson (1975); Lam and Girvin (1984); Levesque et al. (1984); Andrei et al. (1988); Williams et al. (1991); Li et al. (1997); Ye et al. (2002); Chen et al. (2004); Li et al. (1995); Deng et al. (2019); Chen et al. (2006); Drichko et al. (2016); Tiemann et al. (2014). This electron solid is pinned by the ubiquitous residual disorder, manifests as an insulating phase in DC transport Jiang et al. (1990); Goldman et al. (1990); Li et al. (1991); Santos et al. (1992); Sajoto et al. (1993); Pan et al. (2002); Maryenko et al. (2018); Hossain et al. (2020); Chung et al. (2022), and the electrons’ collective motion is evidenced by a resonance in AC transport Lozovik and Yudson (1975); Lam and Girvin (1984); Levesque et al. (1984); Andrei et al. (1988); Williams et al. (1991); Li et al. (1997); Ye et al. (2002); Chen et al. (2004). A series of experiments have been applied to investigate this correlated solid, such as the nonlinear IVI-V response Goldman et al. (1990); Williams et al. (1991), the noise spectrum Li et al. (1991), the huge dielectric constant Li et al. (1995), the weak screening efficiency Deng et al. (2019), the melting process Chen et al. (2006); Drichko et al. (2016); Deng et al. (2019), the nuclear magnetic resonance Tiemann et al. (2014) and the optics Zhou et al. (2021); Smoleński et al. (2021).

Capacitance measurements have revealed a series of quantum phenomena Mosser et al. (1986); Ashoori et al. (1992); Smith et al. (1986); Yang et al. (1997); Eisenstein et al. (1994); Zibrov et al. (2017); Irie et al. (2019); Eisenstein et al. (1992); Jo et al. (1993); Li et al. (2011); Zibrov et al. (2018); Tomarken et al. (2019); Deng et al. (2019). In this work, we examine the WC formed in an ultra-high mobility 2DES at ν\nu\lesssim 1/5 using high-precision capacitance measurement Zhao et al. (2022a, b). We find an exceedingly large capacitance at low measurement frequency ff while the conductance is almost zero. This phenomenon is inconsistent with transporting electrons, but rather an evidence that the synchronous vibration of electrons induces a polarization current. When we increase ff, our high-precision measurement captures the fine structure of the resonance response with a puzzling ”half-dome” structure. Our systematic, quantitative results provide an in-depth insight of this murky quantum phase.

Refer to caption
Figure 1: (color online) (a) CC and GG measured from the r1=r2=r_{1}=r_{2}= 60 μ\mum sample with 7 MHz excitation at 30 mK. The horizontal dashed lines represent the zeros of CC or GG. The blue shaded regions mark the presence of WC. Inset is the cartoon of our device. (b) The correlation between CC and GG in panel (a) data. Transporting current dominates at B<B< 8 T where CG3/2C\propto G^{3/2}, indicated the red solid line. When the WC polarization current dominates, C0.2C\simeq 0.2 pF and GG is about zero (the blue box). (c) The schematic model describing the collective motion of electrons in the pinned WC. hh is the depth of 2DES. The equally spaced (by the lattice constant a0a_{0}) vertical bars represent the equilibrium position of electrons. The gray-scaled solid circles represent the electron position at finite external electric field 𝐄\mathbf{E}. The darker gray corresponds to larger electron displacement 𝐱\mathbf{x}. The radius of individual electron is about the magnetic length lBl_{\text{B}}. The accumulated charge QQ is proportional to 𝐱\nabla\cdot\mathbf{x}, and decays exponentially as a function of the distance dd from the gate boundary. ζ\zeta is the decay length. CWCC_{\text{WC}} is the effective capacitance of WC in the un-gated region between the two gates. (d) CC v.s. ν\nu of the r2r_{2}=100 μ\mum sample. The black dashed line is the zero of CC. The red dashed line is the linear extension of data, showing that C=0C=0 at the extreme quantum limit ν=0\nu=0. (e) 1/CWC1/C_{\text{WC}} v.s. ln(r2/r1r_{2}/r_{1}) at two different magnetic field.

Our sample consists an ultra-clean low-density 2DES confined in a 70-nm-wide GaAs quantum well with electron density n4.4×1010n\simeq 4.4\times 10^{10} cm-2 and mobility μ\mu\simeq 17 ×106\times 10^{6} cm2/(V\cdots). Each device has a pair of front concentric gates G1 and G2, whose outer and inner radius are r1r_{1} and r2r_{2}, respectively; see the inset of Fig. 1(a) 111See Supplemental Material for detailed description of our sample information and measurement techniques.. We study four devices with r1=r_{1}=60 μ\mum and r2=r_{2}= 60, 80, 100 and 140 μ\mum, respectively. We measure the capacitance CC and conductance GG between the two gates using a cryogenic bridge and analyze its output with a custom-made radio-frequency lock-in amplifier Zhao et al. (2022a, b); Note (1).

Fig. 1(a) shows the CC and GG measured from the r1=r2=r_{1}=r_{2}= 60 μ\mum sample. Both CC and GG decrease as we increase the magnetic field BB, owing to the magnetic localization where the 2DES conductance σ(ne2τ)/m(1+ωc2τ2)\sigma\propto(ne^{2}\tau)/m^{\star}(1+\omega_{c}^{2}\tau^{2}), mm^{\star}, ωc\omega_{c} and τ\tau are the effective mass, cyclotron frequency and transport scattering time of the electrons, respectively Zhao et al. (2022b). The CC and GG are finite at ν=1/2\nu=1/2 and 1/4 where the 2DES forms compressible composite Fermion Fermi sea. When ν\nu is an integer or a certain fraction such as 1/3 and 1/5, the 2DES forms incompressible quantum Hall liquids so that both CC and GG vanish 222The zero of CC and GG can be defined either by extrapolating their field dependence to B=B=\infty, or by their values at strong quantum hall states such as ν=1\nu=1. These two approaches are consistent with each other and the dash lines in Fig. 1(a) represent the deduced zero..

In all the above cases, the current is carried by transporting electrons, so that CC has a positive dependence on GG, i.e. CG3/2C\propto G^{3/2}, as shown in Fig. 1(b) Zhao et al. (2022b). Such a correlation discontinues when the WC forms at very low filling factors ν1/5\nu\lesssim 1/5, see the blue shaded regions of Fig. 1(a). The vanishing conductance GG suggests that the electrons are immovable, however, the surprisingly large capacitance CC evidences that the WC hosts a current even surpassing the conducting Fermi sea at ν=1/2\nu=1/2 and 1/4 at much lower magnetic field! The phase transition between the WC and the liquid states are clearly evidenced by spikes in GG (marked by solid circles in Fig. 1(a)) and sharp raises in CC. A developing minimum is seen in GG at 1/5<ν<2/91/5<\nu<2/9 (marked by the up-arrow) when CC has a peak. This GG minimum develops towards zero and the CC peak saturates when the solid phase is stronger (see black traces in Fig. 3(a)). This is consistent with the reentrant insulating phase Jiang et al. (1990); Goldman et al. (1990); Williams et al. (1991); Li et al. (1991); Chen et al. (2004); Shayegan (2006, 1998).

It is important to mention that the 2DES in our devices is effectively “isolated” and we are merely transferring charges between different regions within one quantum phase. Similar to the dielectric materials which also have no transporting electrons, the collective motion of all electrons, i.e. the k0k\to 0 phonon mode of WC, can generate polarization charges and corresponding polarization current in response to the in-plane component of applied electric field. An infinitesimally small but ubiquitous disorder pins the WC so that electrons can only be driven out of their equilibrium lattice site by a small displacement 𝐱\mathbf{x}, as shown in Fig. 1(c). During the experiments, we use excitation VinV_{\text{in}}\simeq 0.1 mVrms{}_{\text{rms}} and the measured WC capacitance is \sim 0.15 pF at 13.5 T. The polarization charge accumulated under the inner gate is Q=CVinQ=CV_{\text{in}}\sim 100 ee. The corresponding electron displacement at the boundary of the inner gate, |𝐱(r1)|Q/(2πr1ne)0.6|\mathbf{x}(r_{1})|\simeq Q/(2\pi r_{1}ne)\sim 0.6 nm, is much smaller than the magnetic length lB=/eB8l_{B}=\sqrt{\hbar/eB}\sim 8 nm, substantiating our assumption that the electrons vibrate diminutively around their equilibrium lattice sites.

An ideal, disorder-free WC is effectively a perfect dielectric with infinite permittivity, so that the device capacitance should be close to its zero-field value C0C_{0}\sim 1 pF when 2DES is an excellent conductor. We note that C0C_{0} is consistent with the device geometry, ϵ0ϵGaAsπr12/h\epsilon_{0}\epsilon_{\text{GaAs}}\pi r_{1}^{2}/h\simeq 1.3 pF, where ϵGaAs=12.8\epsilon_{\text{GaAs}}=12.8 is the relative dielectric constant of GaAs and hh\simeq 960 nm is the depth of 2DES. However, the measured C0.15C\sim 0.15 pF in the WC regime is much smaller than C0C_{0}. This discrepancy is likely caused by the friction-like disorder which poses a pinning force β𝐱\simeq-\beta\mathbf{x} on the electrons. When the crystal’s inversion symmetry is broken, i.e. 𝐱\mathbf{x} is non-uniform and 𝒥(𝐱)\mathcal{J}(\mathbf{x}) is finite, the electron-electron interaction generates a restoring force a0μij𝒥(𝐱)\simeq-a_{0}\mu_{ij}\mathcal{J}(\mathbf{x}), where μij\mu_{ij}, a0a_{0} and 𝒥(𝐱)\mathcal{J}(\mathbf{x}) are the elastic tensor, WC lattice constant and the Jacobi matrix of 𝐱\mathbf{x}, respectively. At the low frequency limit, the WC is always at equilibrium and all forces are balanced, e𝐄a0μij𝒥(𝐱)β𝐱=0e\mathbf{E}-a_{0}\mu_{ij}\mathcal{J}(\mathbf{x})-\beta\mathbf{x}=0, 𝐄\mathbf{E} is the total parallel electric field on the WC.

𝐄\mathbf{E} is approximately zero under the metal gates, since the gate-to-2DES distance hh is small. Therefore, 𝐱\mathbf{x} decreases exponentially when the distance from the gate boundary dd increases, 𝐱exp(d/ζ)\mathbf{x}\propto\exp(-d/\zeta), where ζ=μa0/β\zeta=\mu a_{0}/\beta is the decay length. Deeply inside the gates, electrons feel neither parallel electric field nor net pressure from nearby electrons, so that their displacement 𝐱\mathbf{x} remains approximately zero. This region does not contribute to the capacitive response, and the effective gate area reduces to about 2πr1ζ2\pi r_{1}\zeta and 2πr2ζ2\pi r_{2}\zeta at the inner and outer gate, respectively. Because r1=r2=r_{1}=r_{2}= 60 μ\mum in Fig. 1(a), the experimentally measured Cϵ0ϵGaAs/h2πr1ζ/2C\approx\epsilon_{0}\epsilon_{\text{GaAs}}/h\cdot 2\pi r_{1}\zeta/2\simeq 0.15 pF at 13.5 T corresponds to a decay length ζ\zeta\simeq 6.7 μ\mum. Interestingly, our result shows a linear dependence C1/BC\propto 1/B in Fig. 1(d), suggesting that βlB2\beta\propto l_{B}^{-2} if we assume μij\mu_{ij} is independent on BB. Especially, the pinning becomes infinitely strong, i.e. β\beta\to\infty, at the extreme quantum limit lB0l_{B}\to 0.

The permittivity of a disorder-pinned WC is no longer infinitely large, since a non-zero electric field 𝐄\mathbf{E} is necessary to sustain a finite 𝐱\mathbf{x}. If we assume 𝐱\mathbf{x} is a constant in the ring area between the two gates, so that e𝐄=β𝐱e\mathbf{E}=\beta\mathbf{x}. The residual 𝐄\mathbf{E} can be modeled as a serial capacitance CWC2πne2/β[ln(r2/r1)]1C_{\text{WC}}\approx 2\pi ne^{2}/\beta\cdot[\ln(r_{2}/r_{1})]^{-1} in our device. We then measure different devices with r1r_{1}= 60 μ\mum and r2=60r_{2}=60, 80, 100 and 140 μ\mum, and calculate the corresponding CWCC_{\text{WC}} through CWC1=C1(r1+r2)/r2Cr1=r21C_{\text{WC}}^{-1}=C^{-1}-(r_{1}+r_{2})/r_{2}\cdot C_{r_{1}=r_{2}}^{-1}, see Fig. 1(e). By fitting the linear dependence CWC1ln(r2/r1)C_{\text{WC}}^{-1}\propto\ln(r_{2}/r_{1}), we estimate the pinning strength β\beta to be about 1.3 ×109\times 10^{-9} and 1.1 ×109\times 10^{-9} N/m at B=13.5B=13.5 and 12 T, respectively 333Alternatively, CWCC_{\text{WC}} can be modeled as a cylinder capacitor whose height equals the effective thickness of the 2DES, Z045Z_{0}\approx 45 nm. The WC dielectric constant is ϵWC=(2πϵ0Z0(CWC1)/ln(r2/r1))12×104\epsilon_{\text{WC}}=(2\pi\epsilon_{0}Z_{0}\partial(C_{\text{WC}}^{-1})/\partial\ln(r_{2}/r_{1}))^{-1}\approx 2\times 10^{4} at 13.5 T, consistent with previous reported value in ref. Li et al. (1995).. Finally, assuming μijμδij\mu_{ij}\approx\mu\cdot\delta_{ij}, we can estimate the WC elastic modulus μβζ/a0\mu\approx\beta\cdot\zeta/a_{0}. For example, μ\mu is about 1.6×1071.6\times 10^{-7} N/m at 13.5 T.

Refer to caption
Figure 2: (color online) (a) CC and GG vs. ν\nu measured at various temperatures from the r2=r_{2}= 80 μ\mum sample with 17 MHz excitation. (b) Summarized CC and GG vs. TT at ν=0.14\nu=0.14 and 0.18 from the panel (a) data. A critical temperature TcT_{c} at certain ν\nu is defined either as the temperature when GG has a peak at ν\nu in panel (a) or as the temperature when GG vs. TT trace reaches maximum in panel (b); marked by the black and red arrows. The panel (b) inset summarizes the TcT_{c} using the two equivalent definitions using black and red circles, respectively. The diagram can be separated into three different regions corresponding to the WC, the fractional quantum Hall (FQH) liquid and the compressible liquid.
Refer to caption
Figure 3: (color online) (a) CC and GG vs. ν\nu taken from the r2r_{2}=100 μ\mum sample using different excitation frequencies ff. We see a violent change of CC and GG at different ff in the blue region where the WC appears. (b) The CC and GG vs. ff extracted from the panel (a) trace at ν=0.14\nu=0.14 and 0.213. The resonance frequency frf_{r}, defined as the frequency when CC changes its sign, is about 26 MHz. (c) The CC and GG vs. ff at ν\nu = 0.14 and different temperatures, data taken from the r2r_{2}=80 μ\mum sample. The resonance disappears at TT\simeq 280 mK when CC and GG remain nearly zero.

Fig. 2 reveals an intriguing temperature-induced solid-liquid phase transition when the WC melts. Fig. 2(a) shows CC and GG taken from the r2=80r_{2}=80 μ\mum sample at various temperatures. At a certain temperature, e.g. at T110T\approx 110 mK, C0.2C\sim 0.2 pF when the 2DES forms WC at ν0.16\nu\lesssim 0.16 and vanishes when it is a liquid phase at ν0.18\nu\gtrsim 0.18. GG has a peak at ν0.175\nu\simeq 0.175 when CC vs. ν\nu has the maximal negative slope, and it is small when the 2DES is either a WC at ν<0.17\nu<0.17 or a liquid at ν>0.19\nu>0.19 444We observe developing minimum at ν=1/7,2/11\nu=1/7,2/11 during the solid-liquid phase transition, signaling that the fractional quantum Hall state emerges Pan et al. (2002); Chung et al. (2022).. At very high temperature TT\gtrsim 200 mK, both CC and GG are close to zero. In Fig. 2(b), we summarized CC and GG as a function of TT at two different filling factors to better illustrate this solid-liquid transition. At ν0.14\nu\simeq 0.14, for example, CC is large and GG is small at T100T\lesssim 100 mK when the WC is stable 555CC vs. TT has a slightly positive slope in the WC region, possibly due to the softening of disorder pinning., while both of them become small at T200T\gtrsim 200 mK when the 2DES is a liquid. The GG has a peak at a critical temperature TCT_{C}, marked by the red arrows, around which the precipitous decrease of CC happens. Alternatively, TCT_{C} at a certain filling factor ν\nu can be defined as the temperature when the GG has a peak (black arrow in Fig. 2(a)) at ν\nu. We summarize TCT_{C} obtained using these two equivalent procedures in the Fig. 2(b) inset with corresponding red and black symbols. TCT_{C} has a linear dependence on ν\nu whose two intercepts are TC340T_{C}\simeq 340 mK at the extreme quantum limit ν=0\nu=0, and ν\nu\simeq 1/4 at TC=0T_{C}=0 mK.

The Fig. 2(b) evolution can be qualitatively understood by the coexistence of transport and polarization currents at the solid-liquid transition. The large CC reduces to almost zero when the transport current dominates over the polarization current. GG is a measure of the 2DES’s capacity to absorb and dissipate power. It is negligible if either of these two currents dominates, since the polarization current is dissipation-less and the dissipating transport current is difficult to excite. GG becomes large when these two currents coexist nip and tuck at intermediate TT when the excited polarization charge can be just dissipated by the transport current.

The WC exhibits a resonance when we increase the excitation frequency. In Fig. 3(a), the CC and GG measured from the r2=100r_{2}=100 μ\mum sample using different excitation frequencies change enormously when the WC presents (blue shaded region). GG is almost zero and CC is large at f7f\simeq 7 MHz, and GG becomes finite and CC becomes even larger at f23f\simeq 23 MHz. At slightly higher frequency 27 MHz, GG reaches its maximum and CC drops to about zero. Further increasing ff, GG gradually declines while CC first becomes negative at 35 MHz and then gradually approaches zero. The summarized CC and GG vs. ff at two certain fillings in Fig. 3(b), resembles qualitatively a resonant behavior with resonance frequency fr26f_{r}\simeq 26 MHz (when C=0C=0). Fig. 3(c) studies this resonance at different temperatures. The data is taken from the r2r_{2}\simeq 80 μ\mum sample whose resonance frequency is about 35 MHz 666frf_{r} has no obvious dependence with sample geometry, which is about 35, 35, 26 and 29 MHz for samples with r2r_{2} = 60, 80, 100, 140 μ\mum, respectively.. The abrupt change of CC near frf_{r} becomes gradual and the GG peak flattens at higher temperatures. Both CC and GG become flat zero at T280T\gtrsim 280 mK. It is noteworthy that, as long as a resonance is seen, frf_{r} is nearly independent on the filling factor (Fig. 3(b)) and temperatures (Fig. 3(c)). This is consistent with another experimental study using surface acoustic wave Drichko et al. (2016).

The resonance of WC is usually explained by the pinning mode Fogler and Huse (2000); Ye et al. (2002). The resonance frequency is related to the mean free path LTL_{T} of the transverse phonon through LT=(2πμt,cl/neBfr)1/2L_{T}=(2\pi\mu_{t,cl}/neBf_{r})^{1/2}, where μt,cl=0.245e2n3/2/4πϵ0ϵGaAs\mu_{t,cl}=0.245e^{2}n^{3/2}/4\pi\epsilon_{0}\epsilon_{\text{GaAs}} is the classical shear modulus of WC. fr=26f_{r}=26 MHz corresponds to LTL_{T}\simeq 3.2 μ\mum, very similar to ζ6.7\zeta\simeq 6.7 μ\mum in our Fig. 1(c) discussion. This is justifiable because both LTL_{T} and ζ\zeta describe the length-scale within which the collective motion of WC is damped/scattered by the random pinning potential.

Before ending the discussion, we would like to highlight the puzzling ”half-dome” structure of the resonance. GG has a regular-shaped resonance peak, i.e. GG decreases gradually on both sides of frf_{r}, when either the WC is weak ( ν0.213\nu\simeq 0.213 in Fig. 3(b)) or the temperature is high (T140T\simeq 140 mK in Fig. 3(c)). Surprisingly, the resonance peak becomes quite peculiar when the WC is strong at ν0.14\nu\simeq 0.14 and T30T\simeq 30 mK. GG gradually decreases from its peak at frf_{r} on the high frequency side f>frf>f_{r}, while it vanishes instantly when the frequency is lower than frf_{r}, resulting in a ”half-dome” GG vs. ff trace. Meanwhile, the CC increases by 2\sim 2 times and then abruptly changes to negative at frf_{r}. This anomalous ”half-dome” feature is seen in all of our devices as long as the WC is strong and temperature is sufficiently low, suggesting a threshold frequency for the power dissipation.

In conclusion, using the extraordinarily high-precision capacitance measurement technique, we investigate the dynamic response of WC systematically. From the quantitative results and using a simple model, we can study several physical properties of the WC such as elastic modulus, dielectric constant, pinning strength, etc., and discover a puzzling ”half-dome” feature in the resonance peak. Our results certainly shine light on the study of WC and provides new insight on its dynamics.

Acknowledgements.
We acknowledge support by the National Nature Science Foundation of China (Grant No. 92065104 and 12074010) and the National Basic Research Program of China (Grant No. 2019YFA0308403) for sample fabrication and measurement. This research is funded in part by the Gordon and Betty Moore Foundation’s EPiQS Initiative, Grant GBMF9615 to L. N. Pfeiffer, and by the National Science Foundation MRSEC grant DMR 2011750 to Princeton University. We thank L. W. Engel, Bo Yang and Xin Lin for valuable discussion.

References

  • Jain (2007) J. K. Jain, Composite Fermions (Cambridge University Press, Cambridge, UK, 2007).
  • Wigner (1934) E. Wigner, Phys. Rev. 46, 1002 (1934).
  • Jiang et al. (1990) H. W. Jiang, R. L. Willett, H. L. Stormer, D. C. Tsui, L. N. Pfeiffer,  and K. W. West, Phys. Rev. Lett. 65, 633 (1990).
  • Goldman et al. (1990) V. J. Goldman, M. Santos, M. Shayegan,  and J. E. Cunningham, Phys. Rev. Lett. 65, 2189 (1990).
  • Li et al. (1991) Y. P. Li, T. Sajoto, L. W. Engel, D. C. Tsui,  and M. Shayegan, Phys. Rev. Lett. 67, 1630 (1991).
  • Santos et al. (1992) M. B. Santos, Y. W. Suen, M. Shayegan, Y. P. Li, L. W. Engel,  and D. C. Tsui, Phys. Rev. Lett. 68, 1188 (1992).
  • Sajoto et al. (1993) T. Sajoto, Y. P. Li, L. W. Engel, D. C. Tsui,  and M. Shayegan, Phys. Rev. Lett. 70, 2321 (1993).
  • Pan et al. (2002) W. Pan, H. L. Stormer, D. C. Tsui, L. N. Pfeiffer, K. W. Baldwin,  and K. W. West, Phys. Rev. Lett. 88, 176802 (2002).
  • Maryenko et al. (2018) D. Maryenko, A. McCollam, J. Falson, Y. Kozuka, J. Bruin, U. Zeitler,  and M. Kawasaki, Nature Communications 9, 4356 (2018).
  • Hossain et al. (2020) M. S. Hossain, M. K. Ma, K. A. V. Rosales, Y. J. Chung, L. N. Pfeiffer, K. W. West, K. W. Baldwin,  and M. Shayegan, Proceedings of the National Academy of Sciences 117, 32244 (2020).
  • Chung et al. (2022) Y. J. Chung, D. Graf, L. W. Engel, K. A. V. Rosales, P. T. Madathil, K. W. Baldwin, K. W. West, L. N. Pfeiffer,  and M. Shayegan, Phys. Rev. Lett. 128, 026802 (2022).
  • Lozovik and Yudson (1975) Y. Lozovik and V. Yudson, JETP Lett. 22, 11 (1975).
  • Lam and Girvin (1984) P. K. Lam and S. M. Girvin, Phys. Rev. B 30, 473 (1984).
  • Levesque et al. (1984) D. Levesque, J. J. Weis,  and A. H. MacDonald, Phys. Rev. B 30, 1056 (1984).
  • Andrei et al. (1988) E. Y. Andrei, G. Deville, D. C. Glattli, F. I. B. Williams, E. Paris,  and B. Etienne, Phys. Rev. Lett. 60, 2765 (1988).
  • Williams et al. (1991) F. I. B. Williams, P. A. Wright, R. G. Clark, E. Y. Andrei, G. Deville, D. C. Glattli, O. Probst, B. Etienne, C. Dorin, C. T. Foxon,  and J. J. Harris, Phys. Rev. Lett. 66, 3285 (1991).
  • Li et al. (1997) C.-C. Li, L. W. Engel, D. Shahar, D. C. Tsui,  and M. Shayegan, Phys. Rev. Lett. 79, 1353 (1997).
  • Ye et al. (2002) P. D. Ye, L. W. Engel, D. C. Tsui, R. M. Lewis, L. N. Pfeiffer,  and K. West, Phys. Rev. Lett. 89, 176802 (2002).
  • Chen et al. (2004) Y. P. Chen, R. M. Lewis, L. W. Engel, D. C. Tsui, P. D. Ye, Z. H. Wang, L. N. Pfeiffer,  and K. W. West, Phys. Rev. Lett. 93, 206805 (2004).
  • Li et al. (1995) Y. Li, D. Tsui, T. Sajoto, L. Engel, M. Santos,  and M. Shayegan, Solid State Communications 95, 619 (1995).
  • Deng et al. (2019) H. Deng, L. N. Pfeiffer, K. W. West, K. W. Baldwin, L. W. Engel,  and M. Shayegan, Phys. Rev. Lett. 122, 116601 (2019).
  • Chen et al. (2006) Y. P. Chen, G. Sambandamurthy, Z. H. Wang, R. M. Lewis, L. W. Engel, D. C. Tsui, P. D. Ye, L. N. Pfeiffer,  and K. W. West, Nature Physics 2, 452 (2006).
  • Drichko et al. (2016) I. L. Drichko, I. Y. Smirnov, A. V. Suslov, Y. M. Galperin, L. N. Pfeiffer,  and K. W. West, Phys. Rev. B 94, 075420 (2016).
  • Tiemann et al. (2014) L. Tiemann, T. D. Rhone, N. Shibata,  and K. Muraki, Nature Physics 10, 648 (2014).
  • Zhou et al. (2021) Y. Zhou, J. Sung, E. Brutschea, I. Esterlis, Y. Wang, G. Scuri, R. J. Gelly, H. Heo, T. Taniguchi, K. Watanabe, G. Zaránd, M. D. Lukin, P. Kim, E. Demler,  and H. Park, Nature 595, 48 (2021).
  • Smoleński et al. (2021) T. Smoleński, P. E. Dolgirev, C. Kuhlenkamp, A. Popert, Y. Shimazaki, P. Back, X. Lu, M. Kroner, K. Watanabe, T. Taniguchi, I. Esterlis, E. Demler,  and A. Imamoğlu, Nature 595, 53 (2021).
  • Mosser et al. (1986) V. Mosser, D. Weiss, K. Klitzing, K. Ploog,  and G. Weimann, Solid State Communications 58, 5 (1986).
  • Ashoori et al. (1992) R. C. Ashoori, H. L. Stormer, J. S. Weiner, L. N. Pfeiffer, S. J. Pearton, K. W. Baldwin,  and K. W. West, Phys. Rev. Lett. 68, 3088 (1992).
  • Smith et al. (1986) T. P. Smith, W. I. Wang,  and P. J. Stiles, Phys. Rev. B 34, 2995 (1986).
  • Yang et al. (1997) M. J. Yang, C. H. Yang, B. R. Bennett,  and B. V. Shanabrook, Phys. Rev. Lett. 78, 4613 (1997).
  • Eisenstein et al. (1994) J. P. Eisenstein, L. N. Pfeiffer,  and K. W. West, Phys. Rev. B 50, 1760 (1994).
  • Zibrov et al. (2017) A. A. Zibrov, C. Kometter, H. Zhou, E. M. Spanton, T. Taniguchi, K. Watanabe, M. P. Zaletel,  and A. F. Young, Nature 549, 360 (2017).
  • Irie et al. (2019) H. Irie, T. Akiho,  and K. Muraki, Applied Physics Express 12, 063004 (2019).
  • Eisenstein et al. (1992) J. P. Eisenstein, L. N. Pfeiffer,  and K. W. West, Phys. Rev. Lett. 68, 674 (1992).
  • Jo et al. (1993) J. Jo, E. A. Garcia, K. M. Abkemeier, M. B. Santos,  and M. Shayegan, Phys. Rev. B 47, 4056 (1993).
  • Li et al. (2011) L. Li, C. Richter, S. Paetel, T. Kopp, J. Mannhart,  and R. C. Ashoori, Science 332, 825 (2011).
  • Zibrov et al. (2018) A. A. Zibrov, P. Rao, C. Kometter, E. M. Spanton, J. I. A. Li, C. R. Dean, T. Taniguchi, K. Watanabe, M. Serbyn,  and A. F. Young, Phys. Rev. Lett. 121, 167601 (2018).
  • Tomarken et al. (2019) S. L. Tomarken, Y. Cao, A. Demir, K. Watanabe, T. Taniguchi, P. Jarillo-Herrero,  and R. C. Ashoori, Phys. Rev. Lett. 123, 046601 (2019).
  • Zhao et al. (2022a) L. Zhao, W. Lin, X. Fan, Y. Song, H. Lu,  and Y. Liu, Review of Scientific Instruments 93, 053910 (2022a).
  • Zhao et al. (2022b) L. Zhao, W. Lin, Y. J. Chung, K. W. Baldwin, L. N. Pfeiffer,  and Y. Liu, Chinese Physics Letters 39, 097301 (2022b).
  • Note (1) See Supplemental Material for detailed description of our sample information and measurement techniques.
  • Note (2) The zero of CC and GG can be defined either by extrapolating their field dependence to B=B=\infty, or by their values at strong quantum hall states such as ν=1\nu=1. These two approaches are consistent with each other and the dash lines in Fig. 1(a) represent the deduced zero.
  • Shayegan (2006) M. Shayegan, in High Magnetic Fields: Science and Technology, Vol. 3, edited by F. Herlach and N. Miura (World Scientific, Singapore, 2006) pp. 31–60.
  • Shayegan (1998) M. Shayegan, in Perspectives in Quantum Hall Effects, edited by S. D. Sarma and A. Pinczuk (Wiley, New York, 1998) pp. 343–383.
  • Note (3) Alternatively, C\REV@textWCC_{\REV@text{WC}} can be modeled as a cylinder capacitor whose height equals the effective thickness of the 2DES, Z045Z_{0}\approx 45 nm. The WC dielectric constant is ϵ\REV@textWC=(2πϵ0Z0(C\REV@textWC1)/\symoperatorsln(r2/r1))12×104\epsilon_{\REV@text{WC}}=(2\pi\epsilon_{0}Z_{0}\partial(C_{\REV@text{WC}}^{-1})/\partial\mathop{\symoperators ln}\nolimits(r_{2}/r_{1}))^{-1}\approx 2\times 10^{4} at 13.5 T, consistent with previous reported value in ref. Li et al. (1995).
  • Note (4) We observe developing minimum at ν=1/7,2/11\nu=1/7,2/11 during the solid-liquid phase transition, signaling that the fractional quantum Hall state emerges Pan et al. (2002); Chung et al. (2022).
  • Note (5) CC vs. TT has a slightly positive slope in the WC region, possibly due to the softening of disorder pinning.
  • Note (6) frf_{r} has no obvious dependence with sample geometry, which is about 35, 35, 26 and 29 MHz for samples with r2r_{2} = 60, 80, 100, 140 μ\mum, respectively.
  • Fogler and Huse (2000) M. M. Fogler and D. A. Huse, Phys. Rev. B 62, 7553 (2000).

I Supplementary Materials

I.1 Samples

The sample we studied is made from a GaAs/AlGaAs heterostructure wafer grown by molecular beam epitaxy. A 70 nm-wide GaAs quantum well is bound by AlGaAs spacer-layers and δ\delta-doped layers on each side, and locates hh\simeq 960 nm below the sample surface. The as-grown density of the 2DES is n4.4×1010n\simeq 4.4\times 10^{10} cm-2, and its mobility at 300 mK is μ\mu\simeq 17 ×106\times 10^{6} cm2/(V\cdots). Our sample is a 2 mm ×\times 2 mm square piece with four In/Sn contacts at each corner. The contacts are grounded through a resistor to avoid signal leaking. We evaporate concentric, Au/Ti front gate pair G1 and G2 using standard lift-off process, whose outer and inner radius is r1r_{1} and r2r_{2}, respectively. We deposit a 20 nm thick Al2O3 layer between the two gates to prevent them from shorting with each other. The four outer-gates are merged into one piece so that the area of the outer gate G2 is much larger than the inner gate G1.

I.2 Capacitance Measurement Setup

Refer to caption
Figure S1: (color online) (a) Circuit diagram of measurement bridge with 50 Ω\Omega impedance match networks. (b) The VXV_{\text{X}} and VYV_{\text{Y}} from a typical “V-curve” procedure. CC is about 0.25 pF from the balance condition Eq. (5). (c) The calibration results, by measuring commercial capacitors with different frequencies.

The capacitance and conductance response is measured with a cryogenic bridge similar to refs. Zhao et al. (2022a, b).

The kernel of the bridge consists four devices, RhR_{\text{h}}, RrR_{\text{r}}, CrC_{\text{r}} and CC, as shown in Fig. S1(a). CC is the capacitance of sample. We change the value of RhR_{\text{h}} to reach the balance condition

CCr=RhRr.\frac{C}{C_{\text{r}}}=\frac{R_{\text{h}}}{R_{\text{r}}}. (1)

The bridge output VoutV_{\text{out}} is minimum at the balance condition, from which we calculate the CC. This is the so-call “V-curve” procedure, see refs. Zhao et al. (2022a, b) for more information.

In order to expand the allowed bandwidth of the excitation frequency, we add an impedance match network to the input of the bridge, shown as the Fig. S1(a). VextV_{\text{ext}} is the signal source with 50 Ω\Omega output impedance. VextV_{\text{ext}} drives a signal splitter box (the red dashed box) located at the top of the dilution refrigerator through a \sim2 m-long semi-rigid coaxial cable. The box input is a 1:5 transformer in series with a 50 Ω\Omega resistor. The transformer output drives two serial connected 50 Ω\Omega resistors differentially. The differential signals are transmitted to the cryogenic sample holder (the blue dotted box) by two rigid coaxial cables of \sim2 m length. Another pair of impedance matching 50 Ω\Omega resistors are added at the input of the cryogenic bridge, and the 360 Ω\Omega resistors are chosen by balancing the competition between the performance and heating. The characteristic impedance of all coaxial cables in the work is 50 Ω\Omega.

The low-frequency signals Vquasi-DC1V_{\text{quasi-DC1}} and Vquasi-DC2V_{\text{quasi-DC2}} used to measure the value of RhR_{\text{h}} and RrR_{\text{r}}, respectively. The 0.1 μ\muF capacitors are used to separate the high-frequency excitation signals and the quasi-DC signal.

The output VoutV_{\text{out}} is approximately

VoutS(Rh360+RhCCrRr360+Rr)Vext.V_{\text{out}}\propto S\cdot(\frac{R_{\text{h}}}{360+R_{\text{h}}}-\frac{C}{C_{\text{r}}}\cdot\frac{R_{\text{r}}}{360+R_{\text{r}}})\cdot V_{\text{ext}}. (2)

SS can be obtain from the “V-curve” procedure by linear fitting the VXV_{\text{X}} vs. Rh/(360+Rh)R_{\text{h}}/(360+R_{\text{h}}), as shown in Fig. S1(b). VXV_{\text{X}} and VYV_{\text{Y}} are the orthogonal component of VoutV_{\text{out}},

VX=|Vout|cos(θ),\displaystyle V_{\text{X}}=|V_{\text{out}}|\cdot\cos(\theta), (3)
VY=|Vout|sin(θ),\displaystyle V_{\text{Y}}=|V_{\text{out}}|\cdot\sin(\theta), (4)

where θ\theta is the phase of VoutV_{\text{out}}. We can derive the value of CC using Eq. (2) and (3). The new balance condition of the revised bridge is

CCr=RhRr360+Rr360+Rh,\frac{C}{C_{\text{r}}}=\frac{R_{\text{h}}}{R_{\text{r}}}\cdot\frac{360+R_{\text{r}}}{360+R_{\text{h}}}, (5)

where the VX=0V_{\text{X}}=0.

Note that the capacitance CC and the conductance GG of sample lead to the orthogonal component VXV_{\text{X}} and VYV_{\text{Y}}, respectively. Therefore, the GG can be obtained from Eq. (2) and (4) by replacing C/CrC/{C_{\text{r}}} with G/2πfCrG/2\pi f{C_{\text{r}}}, where ff is the excitation frequency.

Fig. S1(c) shows our calibration measurement using different excitation frequencies. The data is almost flat from 7 to \sim100 MHz. The measured capacitance begins to decline slowly above \sim100 MHz, possibly due to the parasitic inductance of bonding wires.