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

On the lineshapes of temperature-dependent transport measurements of superconductors under pressure

Alexander C. Mark Alexander C. Mark Department of Physics
University of Illinois Chicago
Chicago, Illinois 60607
Russell J. Hemley Departments of Physics, Chemistry, and Earth and Environmental Sciences
University of Illinois Chicago
Chicago, Illinois 60607
   Russell J. Hemley Alexander C. Mark Department of Physics
University of Illinois Chicago
Chicago, Illinois 60607
Russell J. Hemley Departments of Physics, Chemistry, and Earth and Environmental Sciences
University of Illinois Chicago
Chicago, Illinois 60607
(Received: date / Accepted: date)
Abstract

Recent reports of superconductivity in the vicinity of room temperature have been the subject of discussion by the community. Specifically, features in the resistance-temperature (RR-TT) relations have raised questions. We show that many of these features can arise from previously unaccounted-for dynamic effects associated with the AC transport techniques often used in high-pressure experiments. These dynamic AC effects can can cause the apparent resistance (R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡\mathit{R_{apparent}}) to diverge from the DC resistance (R𝐷𝐶\mathit{R_{DC}}), sharpen measured superconducting transitions, and produce other features in the measured RR-TT response. We also show that utilizing the full output of phase-sensitive transport measurements provides a valuable probe of superconducting samples in difficult to measure systems.

\ErrorsOff[latex]

{strip}

On the lineshapes of temperature-dependent transport
measurements of superconductors under pressure

Alexander C. Marka and Russell J. Hemleyb

a Department of Physics, University of Illinois Chicago, Chicago IL 60607
b
Departments of Physics, Chemistry, and Earth and Environmental Sciences,
University of Illinois Chicago, Chicago IL 60607

1 Introduction

The discovery of very high-temperature superconductivity in pressurized hydrides has been a major milestone in the route to ambient (i.e., room temperature and pressure) superconductivity. Discoveries include superconducting critical temperatures (TcT_{c}) above 200 K in H3S\mathrm{H_{3}S} drozdov_conventional_2015 (1), LaH10\mathrm{LaH_{10}} somayazulu_evidence_2019 (2, 3, 4), YH9\mathrm{YH_{9}} kong_superconductivity_2021 (5, 6), YH6\mathrm{YH_{6}} troyan_anomalous_2021 (7), and CaH6\mathrm{CaH_{6}} ma_high-temperature_2022 (8, 9) and a growing number of ternary hydrides such as C-S-H pasan_observation_nodate (10), La-Y-H semenok_superconductivity_2021 (11) and La-Al-H chen_high-temperature_2024 (12), many of which were theoretically predicted wang_superconductive_2012 (13, 14, 15, 16, 17). Understanding the origin and nature of superconductivity in these materials requires detailed analyses of electrical resistance, magnetic susceptibility, and other properties of these systems. Of particular interest have been features present in the RR-TT data near, above, and below TcT_{c} reported in different studies drozdov_conventional_2015 (1, 2, 3, 4, 5, 6, 8, 9, 10, 11, 12, 18, 19, 20).

Superconductors are characterized using both AC or DC transport methods tinkham_introduction_2004 (21). AC techniques, including phase-sensitive detection methods, are used to enhance the signal-to-noise ratio (SNR) when measurement signals are very low and often buried within noise or background signals. As a result, AC methods are often used in high-pressure experiments, especially at very high (e.g., megabar) pressures in diamond-anvil cells where sample sizes and therefore signal levels are necessarily small. While the superconducting RR drop at TcT_{c} may be intrinsically sharp (e.g., 2D superconductors guo_crossover_2020 (22)), the measured sharpness of the transition, as well as the temperature dependence in the normal state, may also be influenced by the parameters of the measurement,as pointed out in the original measurements for LaH10\mathrm{LaH_{10}} somayazulu_evidence_2019 (2). Being sensitive to both resistive and reactive effects, AC techniques can introduce frequency dependent features in the measured response that are not observed with DC probes stuller_introduction_2007 (23).

Figure 1 shows examples of apparent narrowing of the superconducting transition, peaks above Tc\mathit{T_{c}}, two-step resistance drops, and non-zero offsets below TcT_{c} in samples measured with AC transport techniques somayazulu_evidence_2019 (2, 3, 18, 24, 25, 19, 20, 26) . We show that these features, many of which have been subjects of discussion in the literature hirsch_nonstandard_2021 (27, 28, 29, 25, 30, 31, 24), can arise from the specific characteristics of the AC detection methods used in these high-pressure experiments. These features can be understood by considering simple electrodynamics, electrical circuit concepts, and signal processing effects that are not typically presented in published reports.

Refer to caption
Figure 1: Examples of R\mathit{R}-T\mathit{T} relations of hydride superconductors exhibiting features above and below Tc\mathit{T_{c}} that have been discussed in the literature. a) and b) LaH10\mathrm{LaH_{10}} somayazulu_evidence_2019 (2), c) LaH10\mathrm{LaH_{10}} drozdov_superconductivity_2019 (3), d) CeH9,10\mathrm{CeH_{9,10}} semenok_evidence_nodate (26), e) H3S\mathrm{H_{3}S} mozaffari_superconducting_2019 (18), f) C-S-H (from Ref. hirsch_enormous_2023 (24)), g) C-S-H (from Ref. hamlin_vector_nodate (25)), h) Lu-N-H dias_observation_2023 (19), i) Lu-N-H salke_evidence_nodate (20). Figures adapted from the references above.

2 Modeling AC Transport Measurements

A common AC measurement technique relies on phase-sensitive detection, i.e., lock-in amplification. Lock-in measurements can be modeled using standard in-phase (X\mathit{X}), quadrature (Y\mathit{Y}), amplitude (|V||\mathit{V}|), and phase angle (θ\theta) output signals from the amplifier. Mathematical definitions for each are provided in the Supplemental Information. The discussion will concern parasitic resistances, capacitances, and inductances that contribute to the signal. These parasitic effects are expected to be most prominent in small and/or inhomogeneous samples beltran_effect_2003 (32, 33).

2.1 Specific Experimental Setups

Transport measurements at high pressures are typically conducted using a four-probe geometry, sometimes referred to as a van der Pauw (VDP) configuration. It is important to note that while a VDP geometry is often stated for these experiments, this does not imply the VDP algorithm van_der_pauw_method_1958 (34) was used to determine the sheet resistivity of a sample unless specifically stated. The majority of the measurements exhibiting unusual features were not performed using the VDP method, but instead utilized techniques based on AC bridge circuitry snider_synthesis_2021 (6, 35) or a boxcar-averaged DC circuit drozdov_conventional_2015 (1, 3) (e.g., Quantum Design PPMS options noauthor_physical_nodate (36)). Other studies reported experiments using lock-in-based techniques with circuitry similar to that shown in Fig. 2 somayazulu_evidence_2019 (2, 37, 35). As all of the aforementioned techniques rely on periodic averaging, they are inherently susceptible to dynamic effects arising from sample inductance and capacitance. The relatively small size and granular nature of hydride superconductors often cause the typically negligible frequency-dependent contribution to the measured signals to approach the magnitude of the resistive contribution troyan_high-temperature_2022 (38, 39).

In a lock-in measurement, current-regulated sinusoidal signals of the form

I𝑖𝑛(t)=|I𝑚𝑎𝑥|sin(ωt)\mathit{I_{in}(t)}=|\mathit{I_{max}}|\sin(\mathit{\omega t}) (1)

is injected into the sample through two points (I and II) by a regulated current source (Fig. 2). A differential voltage of the form

V𝑜𝑢𝑡(t)=|V|sin(ωt+θ)\mathit{V_{out}(t)}=|\mathit{V}|\sin(\mathit{\omega t}+\theta) (2)

is then measured across the other two contacts (III and IV), buffered via an active pre-amplifier, and then measured by a lock-in amplifier. Before taking any measurements, it must be ensured that excitation frequencies are within the pass band of all circuitry in the setup.

Refer to caption
Figure 2: Example of the schematic block diagram of a setup used for small-signal transport measurements. Small parasitic resistances in the current and voltage lines are represented by R𝑐𝑢𝑟𝑟𝑒𝑛𝑡\mathit{R_{current}} and R𝑙𝑜𝑎𝑑\mathit{R_{load}}, respectively. All equipment in physical contact with the sample is electrically floating to ensure large values of R𝑙𝑜𝑎𝑑\mathit{R_{load}} (e.g. a failing contact) do not cause the pre-amplifier to output zero volts due to the sample being shunted to ground (see Ref. noauthor_model_2004 (40) and Supplemental Information).

Typically, effective in-phase and quadrature resistances are then defined in terms of X(t)\mathit{X(t)} and Y(t)\mathit{Y(t)} using Ohms law as

R𝑖𝑝(t,ω)=X(t,ω)I(t,ω)\mathit{R_{ip}(t,\omega)}=\frac{\mathit{X(t,\omega)}}{\mathit{I(t,\omega)}} (3)

and

R𝑞𝑑(t,ω)=Y(t,ω)I(t,ω),\mathit{R_{qd}(t,\omega)}=\frac{\mathit{Y(t,\omega)}}{\mathit{I(t,\omega)}}, (4)

respectively. A complex impedance,

𝐙(t,ω)=R𝑖𝑝(t,ω)+𝑖𝑅𝑞𝑑(t,ω)=|Z(t,ω)|eiθ(t,ω)\mathit{\mathbf{Z}(t,\omega)}=\mathit{R_{ip}(t,\omega)}+\mathit{iR_{qd}(t,\omega)}=|\mathit{Z(t,\omega)}|\ e^{\mathit{i}\theta(\mathit{t,\omega)}} (5)

is then used to estimate R𝐷𝐶\mathit{R_{DC}}. For ohmic materials R𝑖𝑝(t,ω)>>R𝑞𝑑(t,ω)\mathit{R_{ip}(t,\omega)}>>\mathit{R_{qd}(t,\omega)}, so the apparent resistance is estimated as R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡(t,ω)=|Z(t,ω)|R𝐷𝐶(t)\mathit{R_{apparent}(t,\omega)}=|\mathit{Z(t,\omega)}|\approx\mathit{R_{DC}(t)} for a wide range of conditions. While the voltage across ideal ohmic materials is expected to remain in phase with the excitation current, all real materials exhibit dynamic electromagnetic effects with at least a diamagnetic effect present jackson_john_2015 (41). For non-ohmic materials, such as superconductors, the actual II-VV relation may have a significant ω\mathit{\omega} dependence and this approximation is invalid kaiser_electromagnetic_2004 (42, 43).

To understand how these dynamic effects contribute to the measured X\mathit{X} and Y\mathit{Y}, the time-dependent II-VV relation may be written

|V(t,ω)|sin(ωt+θ)=Z(t,ω)|I|sin(ωt)|\mathit{V(t,\omega)}|\ \mathrm{sin}(\mathit{\omega t}+\theta)=\mathit{Z(t,\omega)}\ |\mathit{I}|\ \mathrm{sin}(\mathit{\omega t}) (6)

where Z(ω,t)\mathit{Z(\omega,t)} is the frequency-dependent impedance. This relation is then converted to phasor notation:

𝐕(t,ω)=𝐈(t,ω)(t,ω),\mathit{\mathbf{V}(t,\omega)}=\mathit{\mathbf{I}(t,\omega)}\mathcal{H}\mathit{(t,\omega)}, (7)

where 𝐕(t,ω)\mathit{\mathbf{V}(t,\omega)} is a complex sinusoidal voltage signal and (t,ω)\mathit{\mathcal{H}(t,\omega)} is a complex transfer function that is determined by the physical properties of the sample stuller_introduction_2007 (23, 44).

The presence of a finite offset from zero resistance in the R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡\mathit{R_{apparent}} of samples reported to be superconducting measured with AC techniques has been the subject of frequent discussion (e.g., Refs. hamlin_vector_nodate (25, 30)). Here it is important to note that the use of |V(t)||\mathit{V(t)}|, being a vector magnitude, to estimate R𝐷𝐶\mathit{R_{DC}} will result in a scalar voltage offset. While lock-in amplifiers can extract extremely small signals from background noise, there is still a finite noise floor in any real measurement. For extremely low voltage signals, e.g., zero voltage drop in a superconductor below Tc\mathit{T_{c}}, the signal is guaranteed to be below the lock-in noise floor of the amplifiers. It is naively expected that using Ohm’s law to calculate the temperature dependence of the DC resistance (R𝐷𝐶(T)\mathit{R_{DC}(T)}) based on |V(T,ω)||\mathit{V(T,\omega)}| will result in R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡(T<Tc)0\mathit{R_{apparent}(T<T_{c})}\approx 0 with environmental and instrumental noise causing individual readings to symmetrically fluctuate around 0. This discrepancy is addressed by recognizing that the lock-in amplifier only measures X(T)\mathit{X(T)} and Y(T)\mathit{Y(T)}, with both |V(T)||\mathit{V(T)}| and θ(T)\theta(\mathit{T}) being calculated in hardware. |V(T,ω)||\mathit{V(T,\omega)}|, being a positive definite phasor magnitude, is then guaranteed to produce a positive resistance with a scalar offset proportional to the total SNR. For well-isolated experiments the raw X(T)\mathit{X(T)} and Y(T)\mathit{Y(T)} signals are often observed to fluctuate around zero resistance even without accounting for any possible phase ambiguity (see Supplementary Information of Ref. salke_evidence_nodate (20)).

2.2 System Transfer Function

Modeling \mathit{\mathcal{H}} of a system undergoing a complex electronic phase change, such as a granular superconductor, is achieved using passive electrical components (Fig. 3).

Refer to caption
Figure 3: Idealized model of a sample. R𝐷𝐶\mathit{R_{DC}}, L𝑒𝑓𝑓\mathit{L_{eff}}, and Cs\mathit{C_{s}} are in general functions of T\mathit{T}, P\mathit{P}, and ω\mathit{\omega}.

To capture the dynamics inherent to the AC measurement we choose to model the system using effective resistances, capacitances, and inductances representative of the normal state bulk properties. Previous models disregarded these dynamic effects and considered only resistive networks raven_filamentary_1995 (45, 46). In this model the effective inductance takes the form L𝑒𝑓𝑓=L𝑠𝑒𝑙𝑓+L𝑡𝑟\mathit{L_{eff}=L_{self}+L_{tr}} with L𝑠𝑒𝑙𝑓\mathit{L_{self}} being the classical self-inductance and L𝑡𝑟\mathit{L_{tr}} being a transient inductance present at magnetic phase changes. L𝑠𝑒𝑙𝑓\mathit{L_{self}} is typically highly nonlinear with closed-form solutions only existing for relatively simple boundary conditions (e.g., a long thin conducting wire carrying a sinusoidal current rosa_self_1908 (47)). A finite self-capacitance (CsC_{s}) can arise from sample granularity, where adjacent grains are capacitively shunted to each other sang_interfacial_2004 (48, 49, 50, 51). In general, R𝐷𝐶\mathit{R_{DC}}, L𝑒𝑓𝑓\mathit{L_{eff}}, and Cs\mathit{C_{s}} are functions of extrinsic variables, e.g., T\mathit{T}, P\mathit{P}, ω\mathit{\omega}, sample size, and sample granularity.

The admittance of the circuit (Fig. 3) is then

1𝐙=1R𝐷𝐶+iωL𝑒𝑓𝑓+iωCs\frac{1}{\mathbf{Z}}=\frac{1}{\mathit{R_{DC}+\mathit{i}\omega L_{eff}}}+\mathit{i\omega C_{s}} (8)

which may be inverted to produce the complex transfer function (Eq. 9). {strip}   

(ω,L𝑒𝑓𝑓\displaystyle\mathcal{H}(\mathit{\omega,L_{eff}} ,Cs,R𝐷𝐶)=R𝐷𝐶3+ω2R𝐷𝐶L𝑒𝑓𝑓2R𝐷𝐶2+ω2[R𝐷𝐶4Cs2+L𝑒𝑓𝑓22R𝐷𝐶2L𝑒𝑓𝑓Cs]+ω4[2R𝐷𝐶2L𝑒𝑓𝑓2Cs2L𝑒𝑓𝑓3Cs]+ω6[L𝑒𝑓𝑓4Cs2]\displaystyle\mathit{,C_{s},R_{DC}})=\frac{\mathit{R_{DC}^{3}+\omega^{2}R_{DC}L_{eff}^{2}}}{\mathit{R_{DC}^{2}+\omega^{2}[R_{DC}^{4}C_{s}^{2}+L_{eff}^{2}-2R_{DC}^{2}L_{eff}C_{s}]+\omega^{4}[2R_{DC}^{2}L_{eff}^{2}C_{s}^{2}-L_{eff}^{3}C_{s}]+\omega^{6}[L_{eff}^{4}C_{s}^{2}]}}
iω[R𝐷𝐶4CsR𝐷𝐶2L𝑒𝑓𝑓]+ω2[2R𝐷𝐶2L𝑒𝑓𝑓2Cs]+ω4[L𝑒𝑓𝑓4Cs]R𝐷𝐶2+ω2[R𝐷𝐶4Cs2+L𝑒𝑓𝑓22R𝐷𝐶2L𝑒𝑓𝑓Cs]+ω4[2R𝐷𝐶2L𝑒𝑓𝑓2Cs2L𝑒𝑓𝑓3Cs]+ω6[L𝑒𝑓𝑓4Cs2]\displaystyle-\mathit{i\omega}\frac{\mathit{[R_{DC}^{4}C_{s}-R_{DC}^{2}L_{eff}]+\omega^{2}[2R_{DC}^{2}L_{eff}^{2}C_{s}]+\omega^{4}[L_{eff}^{4}C_{s}]}}{\mathit{R_{DC}^{2}+\omega^{2}[R_{DC}^{4}C_{s}^{2}+L_{eff}^{2}-2R_{DC}^{2}L_{eff}C_{s}]+\omega^{4}[2R_{DC}^{2}L_{eff}^{2}C_{s}^{2}-L_{eff}^{3}C_{s}]+\omega^{6}[L_{eff}^{4}C_{s}^{2}]}} (9)

  

{(t,ω)}\Re\{\mathcal{H}\mathit{(t,\omega)}\} is the system in-phase steady-state response, which is proportional to X(t,ω)\mathit{X(t,\omega)} up to an ambiguous phase offset. {(t,ω)}\Im\{\mathcal{H}\mathit{(t,\omega)}\} is the steady-state
quadrature response which is similarly proportional to Y(t,ω)\mathit{Y(t,\omega)} up to a phase steinmetz_reactance_1894 (52). The phase ambiguity may be accounted for by recognizing that lock-in amplifiers inherently take differential measurements on two orthogonal quantities, so Y(t=0)=0\mathit{Y}(\mathit{t}=0)=0 may be manually set through the use of a null detector to record only the phase change throughout the experiment. The signal magnitude and phase shift are then given by

|V(T)|={(T)}2+{(T)}2|\mathit{V(T)|}=\sqrt{\Re\{\mathcal{H}(T)\}^{2}+\Im\{\mathcal{H}(T)\}^{2}} (10)

and

θ(T)=atan({(T)}{(T)})\theta(T)=\mathrm{atan}\bigg{(}\frac{\Re\{\mathcal{H}(T)\}}{\Im\{\mathcal{H}(T)\}}\bigg{)} (11)

We first consider the case of R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡\mathit{R_{apparent}} deviating from R𝐷𝐶\mathit{R_{DC}} for different resistive loads in this model with constant L=1mH\mathit{L}=1\ \mathrm{mH} and C=1nF\mathit{C}=1\ \mathrm{nF}, values similar to those measured in DAC experiments han_electrostrictive_2021 (53) (Fig. 4a) where for R𝐷𝐶<0.5mΩ\mathit{R_{DC}}<0.5\ \mathrm{m\Omega} low excitation frequencies cause R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡\mathit{R_{apparent}} to accurately reproduce the true DC resistance. Although lower frequencies are less susceptible to resonant effects they require significantly larger integration times to obtain reasonable SNR so in practice higher frequencies are typically employed. As the load and/or frequency is increased electrically reactive effects become significant and R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡\mathit{R_{apparent}} begins to diverge from R𝐷𝐶\mathit{R_{DC}} due to a small resonance near ωR110\mathit{\omega_{R1}}\sim 10, and a larger resonance above ωR210000\mathit{\omega_{R2}}\sim 10000. Corresponding phase shifts also indicate large load-dependent effects (Fig. 4b). For ωR1<ω<ωR2\mathit{\omega_{R1}<\omega<\omega_{R2}}, R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡\mathit{R_{apparent}} will approximate the true R𝐷𝐶\mathit{R_{DC}}. Above ωR2\mathit{\omega_{R2}} the dynamic effects may cause the measured resistance to saturate the lock-in amplifier with the signal exceeding the dynamic range of the instrument, as discussed in Ref. somayazulu_evidence_2019 (2) (e.g., Fig. 4). For many experiments, ωR2\omega_{R2} is above the pass-band frequency of the lock-in and is not observed. We note that if a sample undergoes a significant change in electronic properties, such as a superconducting transition, as a function of temperature, the measurement will naturally sweep through one or more of these nonlinear resonances. During the transition, R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡\mathit{R_{apparent}} is expected to diverge from R𝐷𝐶\mathit{R_{DC}}, particularly near Tc\mathit{T_{c}}.

Refer to caption
Figure 4: (a) R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡\mathit{R_{apparent}} for the LRC network in Fig. 3 with constant LL=100 μH\mathrm{\mu H} and CC=50 nF using various resistive loads. ω\mathit{\omega} spans the pass-band of a typical lock-in amplifier. (b) Corresponding phase shifts for RapparentR_{apparent} values.

2.3 Modeling Superconductivity

We now model the resistance as a function of temperature and pressure with R𝐷𝐶(T,P)\mathit{R_{DC}(T,P)}, L𝑒𝑓𝑓(T,P)\mathit{L_{eff}(T,P)} and Cs\mathit{C_{s}} values assumed based on previously reported measurements of superconducting transitions under pressure salke_evidence_nodate (20) and ambient pressure measurements of Sn-whiskers miller_fluctuation_1973 (54), 2H\mathrm{2H}-NbSe2\mathrm{NbSe_{2}} perconte_low-frequency_2020 (55), and Bi2Sr2Ca2Cu3O8+δ\mathrm{Bi_{2}Sr_{2}Ca_{2}Cu_{3}O_{8+\delta}} (Bi-2223) (see Supplemental Information and Ref. mark_notitle_nodate (56)). A R𝐷𝐶\mathit{R_{DC}}-T\mathit{T} relation of the form

R(T)=R𝐷𝐶(T)exp(TTcΓ)1\mathit{R(T)}=\frac{\mathit{R_{DC}(T)}}{\exp\big{(}\frac{\mathit{T-\mathit{T_{c}}}}{\Gamma}\big{)}-1} (12)

is assumed, where Γ\Gamma is a constant broadening factor and R𝐷𝐶(T)\mathit{R_{DC}(T)} is the temperature dependence of the resistance above Tc\mathit{T_{c}}.

Inductive effects are magnetic and well described by Lenz’s law in which any change in magnetic flux (ΦB\Phi_{\mathit{B}}) generates a proportional back electromotive force (𝐵𝑎𝑐𝑘\mathcal{E}_{\mathit{Back}}) determined by the sample’s effective inductance (L𝑒𝑓𝑓(\mathit{L_{eff}}) graf_modern_1999 (57). For a sinusoidal signal, Lenz’s law implies

𝐵𝑎𝑐𝑘=dΦBdt=|I|ωL𝑒𝑓𝑓\mathcal{E}_{\mathit{Back}}=-\frac{d\Phi_{\mathit{B}}}{d\mathit{t}}=-|\mathit{I}|\mathit{\omega L_{eff}} (13)

with Back\mathcal{E}_{Back} generating a voltage 90o90^{o} out of phase from the driving current. L𝑡𝑟\mathit{L_{tr}} then arises from the sudden repulsion of ΦB\Phi_{\mathit{B}} within a superconductor at Tc\mathit{T_{c}} due to the Meissner effect. L𝑡𝑟\mathit{L_{tr}} is suppressed below Tc\mathit{T_{c}} due to the superconductor’s inability to sustain a voltage drop. This L𝑡𝑟\mathit{L_{tr}} is unrelated to the kinetic inductance that has been observed in high-frequency supercurrents meservey_measurements_2003 (58, 59) which are typically several orders of magnitude smaller at the frequencies used in these experiments and only exist well below Tc\mathit{T_{c}} vodolazov_nonlinear_2023 (60, 61, 62). Similar transient low-frequency impedances observed in 2H-NbSe2\mathrm{NbSe_{2}} perconte_low-frequency_2020 (55) were attributed to a coupling of the superconducting vortex state to the heat capacity. Whereas coupling to the heat capacity may cause a phase shift near Tc\mathit{T_{c}}, it does not explain the observed frequency-dependent deviation of X(T)\mathit{X(T)} from R𝐷𝐶(T)\mathit{R_{DC}(T)} or the high-frequency behavior where R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡\mathit{R_{apparent}}-T\mathit{T} can approach a step function in some measurements somayazulu_evidence_2019 (2). We cannot rule out variations (e.g., peaks) in RDC(T)R_{DC}(T) near and above TcT_{c} arising from sample or phase inhomogeneity (e.g., Ref. zhang_bosonic_2016 (63)), but effects of AC measurement techniques can introduce features in this regime as well.

In our model, the total effective inductance is described by a growing and decaying exponential surrounding Tc\mathit{T_{c}} modulated with a step function to account for the finite L𝑠𝑒𝑙𝑓\mathit{L_{self}} above Tc\mathit{T_{c}}. Model R(T)\mathit{R(T)} and L𝑒𝑓𝑓(T)\mathit{L_{eff}(T)} relations that exhibit this effect are presented in Fig. 5. In inhomogeneous materials, including powders and polycrystals, a universal dielectric response causes an effective capacitance (Cs\mathit{C_{s}}) between grain boundaries jonscher_universal_1977 (64, 65, 66, 67, 68). Eq. 9 implies that lock-in based techniques will produce a Y(T,ω)Y(T,\omega) that scales with CsC_{s}. Similar capacitive couplings between grains have been observed in ambient pressure Josephson junctions cheng_anomalous_2016 (69, 70) and in chains of capacitively coupled granular superconductors ilin_superconducting_2020 (71). In addition to the sample’s inherent Cs\mathit{C_{s}}, metal film contacts prepared similarly to those used in high-pressure experiments have been reported to exhibit very large contact capacitances that scale with pressure wang_stress-dependent_2021 (72), with some reported to exceed 1 F dervos_effect_1998 (73). In DAC experiments the collective capacitance of samples can be large (C>1μC>1\ \mathrm{\mu}F) and significantly impact low-frequency measurements when voltage signals are low he_situ_2007 (74, 75).

Refer to caption
Figure 5: Model R𝐷𝐶\mathit{R_{DC}}-T\mathit{T} (black) and L𝑒𝑓𝑓\mathit{L_{eff}}-T\mathit{T} (red) for a superconducting sample. Temperature dependence on DC resistance is assumed to be of the form R(T)=A+𝐵𝑇+𝐶𝑇3\mathit{R(T)}=\mathit{A}+\mathit{BT}+\mathit{CT^{3}} similar to the cubic relation employed in Ref. dusen_platinum-resistance_1925 (76). The temperature scale is normalized such that Tc,𝑚𝑖𝑑\mathit{T_{c,mid}}=1. The L𝑒𝑓𝑓\mathit{L_{eff}} lineshape is estimated based on observed X(T)\mathit{X(T)} and Y(T)\mathit{Y(T)} signals measured in Lu-N-H salke_evidence_nodate (20) and cuprate samples measured here. Inset: X(T)\mathit{X(T)} and Y(T)\mathit{Y(T)} signals from optimally doped Bi-2223 (Tc\mathit{T_{c}}=108 K) using the setup presented in Fig. 2. Bi-2223 data measured over a larger temperature range are presented in the Supplementary Information.

We now consider the frequency dependence of these effects using the range of frequencies employed in these measurements (17 Hz to 10 kHz salke_evidence_nodate (20, 2)). Significant effects have been reported in both low frequency perconte_low-frequency_2020 (55, 20) and high frequency somayazulu_evidence_2019 (2) experiments. These features can provide additional information about the material in both the superconducting and normal states. As Cs\mathit{C_{s}} is relatively insensitive to temperature or frequency, it is assumed to be constant for each pressure. Model signal curves for a superconducting transition using various frequencies are presented in Fig. 6. The calculated R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡(T)\mathit{R_{apparent}(T)} in the simulated curve does not approximate R𝐷𝐶\mathit{R_{DC}} for all frequencies, especially for small samples where voltages from dynamic effects approach the resistive voltage drops. As ωωR1\mathit{\omega}\rightarrow\mathit{\omega_{R1}} a peak emerges in R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡\mathit{R_{apparent}} that distorts the transition (Fig. 6). While the signal is influenced by ωR1\mathit{\omega_{R1}}, R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡\mathit{R_{apparent}} overestimates R𝐷𝐶\mathit{R_{DC}} above Tc\mathit{T_{c}} and is seen to sharpen the transition. As ω\mathit{\omega} is further increased, the influence of ωR2\mathit{\omega_{R2}} causes the voltage signal to grow exponentially, often saturating the measurement equipment resulting in the R\mathit{R}-T\mathit{T} relation approaching a step function, although this effect is not always observed due to ωR2\mathit{\omega_{R2}} often being outside the range of the lock-in amplifier. Similar peaks have been reported in the resistivity of thin disordered superconductors measured using an AC bridge (e.g., Sn/In ems_resistance_1971 (77) and TiN postolova_reentrant_2017 (78)).

Refer to caption
Figure 6: Model R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡\mathit{R_{apparent}}-T\mathit{T} relations for a sample with a constant Cs=1pF\mathit{C_{s}}=1\mathrm{\ pF} and the R𝐷𝐶(T)\mathit{R_{DC}(T)} and L𝑒𝑓𝑓(T)\mathit{L_{eff}(T)} relations from Fig. 5. Inset: Difference in R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡\mathit{R_{apparent}} and R𝐷𝐶\mathit{R_{DC}} for various frequencies.

Additional features in RR-TT curves are present if the reactive contribution from CsC_{s} is significant (e.g. Fig. 1b somayazulu_evidence_2019 (2)). Cs(P)C_{s}(P) is a function of pressure, with typical values ranging from 1 pF to 1 μ\mathrm{\mu}F wei_dielectric_2008 (49). This variability can cause lineshapes to be distorted under pressure (Fig. 7). For larger values of CsC_{s} and higher ω\mathit{\omega}, the system may sweep through ωR2\mathit{\omega_{R2}} during the superconducting transition, producing large X(T,ω)X(T,\omega) and Y(T,ω)Y(T,\omega) signals that saturate the lock-in. While significant enough to impact the measurement, the magnitude of the large resonant effects at frequencies near ωR2\mathit{\omega_{R2}} is non-physical due to significant non-linear effects not accounted for in this model.

Refer to caption
Figure 7: Model R\mathit{R}-T\mathit{T} relations for various high frequencies and a large Cs\mathit{C_{s}} = 1μ\mu F. Inset: lineshape presented on a logarithmic scale demonstrating the full extent of the resonance near Tc\mathit{T_{c}} (see text)

3 Discussion

We now discuss specific features in reported R\mathit{R}-T\mathit{T} measurements of hydride high Tc\mathit{T_{c}} superconductors that have given rise to criticisms and concerns (e.g., Refs. hirsch_nonstandard_2021 (27, 25, 28)) that can be understood in terms of the above considerations. First, we point out that many papers report utilizing a ‘van der Pauw geometry’, but do not use the VDP ‘technique’ itself (i.e., to measure DC sheet resistivity). Without performing a true VDP measurement resistivity cannot be determined using the VDP formula, and electrical transport properties must be determined by other means van_der_pauw_method_1958 (34); specifically, AC techniques imply R𝐷𝐶\mathit{R_{DC}} cannot be calculated using extensions of Ohm’s law. Instead the full AC transfer function such as Eq. (9) should be derived and analyzed to extract physical properties. We also point out that there is no evidence for significant broadening of the superconducting transitions shown in Fig. 1 due to pressure gradients in the sample (e.g., Ref. deemyad_dependence_2003 (79)). This observation is consistent with the expected relaxation of shear stresses during synthesis and further suggests that shear stresses in the hydride materials are small somayazulu_evidence_2019 (2).

We now discuss examples of various features observed in specific high Tc\mathit{T_{c}} hydride materials. As discussed above, in the original study that led to the discovery of superconductivity in LaH10\mathrm{LaH_{10}} somayazulu_evidence_2019 (2) the sharpness of the transition and slope of the measured RR-TT response depend on the properties of the measurement. Regular sample geometries produced R\mathit{R}-T\mathit{T} curves reminiscent of the expected R𝐷𝐶\mathit{R_{DC}} drop (Fig. 1a) somayazulu_evidence_2019 (2). The same study found that irregularly shaped inhomogeneous samples, determined using spatially resolved in-situ x-ray diffraction, exhibited different behavior. For these samples the R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡\mathit{R_{apparent}}-T\mathit{T} curves approached a step function (Fig. 1b). This behavior can readily be explained as arising from filtering effects due to a geometrically enhanced Cs\mathit{C_{s}} (Fig. 1b). The result of a subsequent study drozdov_superconductivity_2019 (3) show peaks within the resistance drop near Tc\mathit{T_{c}} (Fig. 1c), features that were attributed to competing superconducting phases and metal-insulator interface effects zhang_bosonic_2016 (63). Similar sharp step-like behavior in CeH9,10\mathrm{CeH_{9,10}} was observed in RapparentR_{apparent}-TT upon increasing ω\omega semenok_evidence_nodate (26) (Fig. 1d). We point out that inhomogeneity could increase sample CsC_{s} and enhances AC resonance effects at these temperatures.

Other interesting features are apparent in the R\mathit{R}-T\mathit{T} curves of compressed H3S\mathrm{H_{3}S} reportedly measured using an AC bridge circuit (f\mathit{f}=13 Hz) (Fig. 1e) mozaffari_superconducting_2019 (18). Peaks characteristic of the resonance at ωR1\mathit{\omega_{R1}} are apparent in the resistance drop of the 155 GPa runs but are absent in the 160 GPa data. It was reported that these samples were prepared using different methods; notably the 160 GPa sample being annealed to increase homogeneity whereas that studied at 155 GPa was not. It is expected that more inhomogeneous samples would have a larger CsC_{s} and therefore exhibit larger dynamic signals near Tc\mathit{T_{c}}.

Reported R\mathit{R}-T\mathit{T} curves for compressed C-S-H measured using a lock-in-based setup exhibit many of the above features associated with AC measurements (Fig. 1f) hirsch_enormous_2023 (24); see also Ref. pasan_observation_nodate (10). The curves show sharp transitions at Tc\mathit{T_{c}} representative of the large Y(T)\mathit{Y(T)} signal present at Tc\mathit{T_{c}}. Additionally, several experimental runs exhibit broad peaks or ‘ripples’ above Tc\mathit{T_{c}}, particularly the data for run 1 (blue boxes) at 174 GPa and 220 GPa. As these effects are highly dependent on sample geometry it is expected that repeated runs on the same samples would produce similar features.

Superconductivity in the Lu-N-H system near ambient conditions remains a subject of continued investigations. There is evidence of superconductivity in the vicinity of room temperature and pressure in selected Lu-N-H samples dias_observation_2023 (19, 20). The R\mathit{R}-T\mathit{T} relations measured at various pressures also exhibit narrow superconducting transitions (Fig. 1g). Measurements at 1.6 and 2.0 GPa in Ref. dias_observation_2023 (19) were performed on significantly larger samples than the 1.0 GPa measurements. As a result, the normal state resistance is over 100x larger for the higher pressure measurements. The smaller sample gave rise to a sharper transition, an effect that we ascribe to parasitic impedances.

4 Conclusions

We have shown that classical electrodynamics, circuit theory, and plausible values for the bulk electrical properties of the materials can explain features in reported RR-TT measurements of high TcT_{c} hydride superconductors under pressure. In particular, we demonstrate that lineshapes in reported AC-based transport measurements that have been pointed out to be anomalous may be explained by considering all signal-processing effects. Many of the features observed in these measurements are expected for a superconducting transition. Sample dependent geometric effects are expected to alter the measured signals,e.g., a larger CsC_{s} in an inhomogeneous material is expected to drastically enhance the measured Y(T,ω)Y(T,\omega) signal at all frequencies. While instrumental and dynamic electrical effects may result in R𝑎𝑝𝑝𝑎𝑟𝑒𝑛𝑡\mathit{R_{apparent}} diverging from R𝐷𝐶\mathit{R_{DC}}, particularly near and above Tc\mathit{T_{c}}, the physicality of the superconducting transition is unambiguous. While these results are fully consistent with superconductivity, additional measurements using different techniques are, of course, needed to confirm the materials are superconductors. To accurately interpret results and avoid confusion all aspects of the measurements, including the experimental setup and relevant experimental parameters, should be reported (e.g., X\mathit{X}, Y\mathit{Y}, |V||\mathit{V}| and θ\theta). Analysis of these data can provide additional information about the state of potential superconducting samples.

Acknowledgements.
We thank A. Denchfield, C. Mark, P. Melnikov, J.C. Campuzano, S. Deemyad, M. Somayazulu, D. Semenok, and G.W. Collins for helpful discussions. This work was supported by the U.S. National Science Foundation (DMR-2104881), DOE-NNSA (DE-NA0003975 Chicago/DOE Alliance Center), and DOE-SC (DE-SC0020340).
{Data availability}

Data Availability All data are available upon request.

References

  • (1) A.. Drozdov et al. “Conventional superconductivity at 203 kelvin at high pressures in the sulfur hydride system” Number: 7567 Publisher: Nature Publishing Group In Nature 525.7567, 2015, pp. 73–76 DOI: 10.1038/nature14964
  • (2) Maddury Somayazulu et al. “Evidence for superconductivity above 260 K in lanthanum superhydride at megabar pressures” Publisher: American Physical Society In Phys. Rev. Lett. 122.2, 2019, pp. 027001 DOI: 10.1103/PhysRevLett.122.027001
  • (3) A.. Drozdov et al. “Superconductivity at 250 K in lanthanum hydride under high pressures” Number: 7757 Publisher: Nature Publishing Group In Nature 569.7757, 2019, pp. 528–531 DOI: 10.1038/s41586-019-1201-8
  • (4) Fang Hong et al. “Superconductivity of lanthanum superhydride investigated using the standard four-probe configuration under high pressures” Publisher: Chinese Physical Society and IOP Publishing Ltd In Chinese Phys. Lett. 37.10, 2020, pp. 107401 DOI: 10.1088/0256-307X/37/10/107401
  • (5) Panpan Kong et al. “Superconductivity up to 243 K in the yttrium-hydrogen system under high pressure” Number: 1 Publisher: Nature Publishing Group In Nat. Commun. 12.1, 2021, pp. 5075 DOI: 10.1038/s41467-021-25372-2
  • (6) Elliot Snider et al. “Synthesis of yttrium superhydride auperconductor with a transition temperature up to 262 K by catalytic hydrogenation at high pressures” Publisher: American Physical Society In Phys. Rev. Lett. 126.11, 2021, pp. 117003 DOI: 10.1103/PhysRevLett.126.117003
  • (7) Ivan A. Troyan et al. “Anomalous high-temperature superconductivity in YH6{}_{\textrm{6}}” _eprint: https://onlinelibrary.wiley.com/doi/pdf/10.1002/adma.202006832 In Adv. Mater. 33.15, 2021, pp. 2006832 DOI: 10.1002/adma.202006832
  • (8) Liang Ma et al. “High-temperature superconducting phase in clathrate calcium hydride CaH6{}_{\textrm{6}} up to 215 K at a pressure of 172 GPa” Publisher: American Physical Society In Phys. Rev. Lett. 128.16, 2022, pp. 167001 DOI: 10.1103/PhysRevLett.128.167001
  • (9) Zhiwen Li et al. “Superconductivity above 200 K discovered in superhydrides of calcium” Number: 1 Publisher: Nature Publishing Group In Nat. Commun. 13.1, 2022, pp. 2863 DOI: 10.1038/s41467-022-30454-w
  • (10) Hiranya Pasan et al. “Observation of conventional near room temperature superconductivity in carbonaceous sulfur hydride” arXiv:2302.08622 [cond-mat] In arXiv.2302.08622 DOI: 10.48550/arXiv.2302.08622
  • (11) Dmitrii V. Semenok et al. “Superconductivity at 253 K in lanthanum–yttrium ternary hydrides” In Mater. Today 48, 2021, pp. 18–28 DOI: 10.1016/j.mattod.2021.03.025
  • (12) Su Chen et al. “High-temperature superconductivity up to 223 K in the Al stabilized metastable hexagonal lanthanum superhydride” In Nal. Sci. Rev, 11.1, 2024, pp. nwad107 DOI: 10.1093/nsr/nwad107
  • (13) Hui Wang et al. “Superconductive sodalite-like clathrate calcium hydride at high pressures” Publisher: Proceedings of the National Academy of Sciences In Proc. Nat. Acad. Sci. 109.17, 2012, pp. 6463–6466 DOI: 10.1073/pnas.1118168109
  • (14) Yinwei Li et al. “The metallization and superconductivity of dense hydrogen sulfide” In J. Chem. Phys. 140.17, 2014, pp. 174712 DOI: 10.1063/1.4874158
  • (15) Defang Duan et al. “Pressure-induced metallization of dense (H2{}_{\textrm{2}}S)2{}_{\textrm{2}}H2{}_{\textrm{2}} with high-Tc{}_{\textrm{c}} superconductivity” Publisher: Nature Publishing Group In Sci. Rep. 4.1, 2014, pp. 6968 DOI: 10.1038/srep06968
  • (16) Hanyu Liu et al. “Potential high-Tc{}_{\textrm{c}} superconducting lanthanum and yttrium hydrides at high pressure” Publisher: Proceedings of the National Academy of Sciences In Proc. Natl. Acad. Sci. 114.27, 2017, pp. 6990–6995 DOI: 10.1073/pnas.1704505114
  • (17) Feng Peng et al. “Hydrogen clathrate structures in rare earth hydrides at high pressures: Possible route to room-temperature superconductivity” Publisher: American Physical Society In Phys. Rev. Lett. 119.10, 2017, pp. 107001 DOI: 10.1103/PhysRevLett.119.107001
  • (18) Shirin Mozaffari et al. “Superconducting phase diagram of H3{}_{\textrm{3}}S under high magnetic fields” In Nat. Comm. 10, 2019, pp. 2522 DOI: 10.1038/s41467-019-10552-y
  • (19) Ranga P. Dias et al. “Observation of room temperature superconductivity in hydride at near ambient pressure” In Bull. Am. Phys. Soc. 68,.Abstract #K20-00002, 2023, pp. Abstract #K20–00002 URL: https://meetings.aps.org/Meeting/MAR23/Session/K20.2?s=09
  • (20) Nilesh P. Salke, Alexander C. Mark, Muhtar Ahart and Russell J. Hemley “Evidence for near ambient superconductivity in the Lu-N-H system” arXiv:2306.06301 [cond-mat] In arXiv.2306.06301 DOI: arXiv.2306.06301
  • (21) Michael Tinkham “Introduction to Superconductivity” Google-Books-ID: VpUk3NfwDIkC Dover, 2004
  • (22) Jing Guo et al. “Crossover from two-dimensional to three-dimensional superconducting states in bismuth-based cuprate superconductor” Number: 3 Publisher: Nature Publishing Group In Nat. Phys. 16.3, 2020, pp. 295–300 DOI: 10.1038/s41567-019-0740-0
  • (23) John A. Stuller “An Introduction to Signals and Systems” London: Thomson Learning, 2007
  • (24) J.. Hirsch “Enormous variation in homogeneity and other anomalous features of room temperature superconductor samples: a Comment on Nature 615, 244 (2023)” arXiv:2304.00190 [cond-mat] In J. Supercond. Nov. Magn. 36.6, 2023, pp. 1489–1494 DOI: 10.1007/s10948-023-06593-6
  • (25) James J. Hamlin “Vector graphics extraction and analysis of electrical resistance data in Nature volume 586, pages 373-377 (2020)” In arXiv.2210.10766 DOI: 10.48550/arXiv.2210.10766
  • (26) Dmitrii Semenok et al. “Evidence for pseudogap phase in cerium superhydrides: CeH10{}_{\textrm{10}} and CeH9{}_{\textrm{9}}” arXiv:2307.11742 [cond-mat] In arXiv.2307.11742 DOI: 10.48550/arXiv.2307.11742
  • (27) J.. Hirsch and F. Marsiglio “Nonstandard superconductivity or no superconductivity in hydrides under high pressure” Publisher: American Physical Society In Phys. Rev. B 103.13, 2021, pp. 134505 DOI: 10.1103/PhysRevB.103.134505
  • (28) Mehmet Dogan and Marvin L. Cohen “Anomalous behavior in high-pressure carbonaceous sulfur hydride” In Physica C 583, 2021, pp. 1353851 DOI: 10.1016/j.physc.2021.1353851
  • (29) Ranga P. Dias and Ashkan Salamat “Standard superconductivity in carbonaceous sulfur hydride” arXiv:2111.15017 [cond-mat] In arXiv.2111.15017 DOI: 10.48550/arXiv.2111.15017
  • (30) Dale R. Harshman and Anthony T. Fiory “Analysis of electrical resistance data from Snider et al., Nature 586, 373 (2020)” arXiv:2212.06237 [cond-mat] In arXiv.2212.06237 DOI: 10.48550/arXiv.2212.06237
  • (31) E.. Talantsev, V.. Minkov, F.. Balakirev and M.. Eremets “Broadening of in-field superconducting transitions in hydrides” arXiv:2311.07865 [cond-mat] In arXiv.2311.07865 DOI: 10.48550/arXiv.2311.07865
  • (32) Nicolás H. Beltrán, Ricardo A. Finger, Jorge Santiago-Aviles and Patricio Espinoza-Vallejos “Effect of parasitic capacitances on impedance measurements in microsensors structures: a numerical study” In Sens. Actuators B: Chem. 96.1, 2003, pp. 139–143 DOI: 10.1016/S0925-4005(03)00516-1
  • (33) Vahid Farmehini et al. “On-chip impedance for quantifying parasitic voltages during AC electrokinetic trapping” Conference Name: IEEE Transactions on Biomedical Engineering In IEEE Trans. Biomed. Eng. 67.6, 2020, pp. 1664–1671 DOI: 10.1109/TBME.2019.2942572
  • (34) L.J. Pauw “A method of measuring the resistivity and Hall coefficient on lamellae of arbitrary shape” In Phillips Tech. Rev. 20, 1958, pp. 220–224 URL: http://electron.mit.edu/~gsteele/vanderpauw/vanderpauw.pdf
  • (35) Sam Cross et al. “High-temperature superconductivity in La4{}_{\textrm{4}}H23{}_{\textrm{23}} below 100 GPa” Publisher: American Physical Society In Phys. Rev. B 109.2, 2024, pp. L020503 DOI: 10.1103/PhysRevB.109.L020503
  • (36) “Physical Property Measurement System PPMS” URL: https://www.qdusa.com/products/ppms.html
  • (37) A.. Grockowiak et al. “Hot hydride superconductivity above 550 K” In Front. Electron. Mat. 2, 2022 URL: https://www.frontiersin.org/articles/10.3389/femat.2022.837651
  • (38) I.. Troyan et al. “High-temperature superconductivity in hydrides” In Physics-Uspekhi 65.7, 2022, pp. 748–761 URL: https://link.springer.com/article/10.1007/s10948-022-06148-1
  • (39) M.. Eremets et al. “High-temperature superconductivity in hydrides: Experimental evidence and details” In J. Supercond. Nov. Magn. 35.4, 2022, pp. 965–977 DOI: 10.1007/s10948-022-06148-1
  • (40) “Model SR554 Transformer Preamplifier”, 2004, pp. 13 URL: https://www.thinksrs.com/products/sr554.html
  • (41) Roland Jackson “John Tyndall and the early history of diamagnetism” In Ann. Sci. 72.4, 2015, pp. 435–489 DOI: 10.1080/00033790.2014.929743
  • (42) Kenneth L. Kaiser “Electromagnetic Compatibility Handbook” Google-Books-ID: nZzOAsroBIEC Boca Raton: CRC Press, 2004
  • (43) Forbes T. Brown “Engineering System Dynamics: A Unified Graph-Centered Approach, Second Edition” Google-Books-ID: UzqX4j9VZWcC Boca Raton: CRC Press, 2006
  • (44) Earl D. Gates “Introduction to Electronics” Google-Books-ID: FLcJzgEACAAJ Delmar, 2001
  • (45) M.. Raven “A filamentary resistance model for mixed c-axis and a-axis orientated YBa2{}_{\textrm{2}}Cu3{}_{\textrm{3}}Ox{}_{\textrm{x}} thin films” _eprint: https://onlinelibrary.wiley.com/doi/pdf/10.1002/pssa.2211470116 In Phys. Stat. Solidi 147.1, 1995, pp. 155–164 DOI: 10.1002/pssa.2211470116
  • (46) J.. Phillips “Filamentary microstructure and linear temperature dependence of normal state transport in optimized high temperature superconductors” Publisher: Proceedings of the National Academy of Sciences In Proc. Nat. Acad. Sci. 94.24, 1997, pp. 12771–12775 DOI: 10.1073/pnas.94.24.12771
  • (47) Edward B. Rosa “The self and mutual inductances of linear conductors” In Bull. Bureau Stand. 4.2, 1908, pp. 301–344
  • (48) Zhi-Fang Sang and Zhen-Ya Li “Interfacial effect on effective dielectric response of spherical granular composites” In Phys. Lett. A 331.1, 2004, pp. 125–131 DOI: 10.1016/j.physleta.2004.08.045
  • (49) En-Bo Wei, G.. Gu and Y.. Poon “Dielectric responses of anisotropic graded granular composites having arbitrary inclusion shapes” Publisher: American Physical Society In Phys. Rev. B 77.10, 2008, pp. 104204 DOI: 10.1103/PhysRevB.77.104204
  • (50) Patrick Winkel et al. “Implementation of a transmon qubit using superconducting granular aluminum” Publisher: American Physical Society In Phys. Rev. X 10.3, 2020, pp. 031032 DOI: 10.1103/PhysRevX.10.031032
  • (51) Chongpu Zhai, Yixiang Gan, Dorian Hanaor and Gwénaëlle Proust “Stress-dependent electrical transport and its universal scaling in granular materials” In Extreme. Mech. Lett. 22, 2018, pp. 83–88 DOI: 10.1016/j.eml.2018.05.005
  • (52) Charles Proteus Steinmetz and Frederick Bedell “Reactance” Conference Name: Transactions of the American Institute of Electrical Engineers In Trans. AIEE XI, 1894, pp. 640–648 DOI: 10.1109/T-AIEE.1894.4763812
  • (53) Tao Han et al. “Electrostrictive effect of materials under high pressure revealed by electrochemical impedance spectroscopy” Publisher: American Chemical Society In J. Phys. Chem. C 125.16, 2021, pp. 8788–8793 DOI: 10.1021/acs.jpcc.1c01020
  • (54) John R. Miller and John M. Pierce “Fluctuation effects in the complex impedance of superconducting tin-whisker crystals near Tc{}_{\textrm{c}}” Publisher: American Physical Society In Phys. Rev. B 8.9, 1973, pp. 4164–4174 DOI: 10.1103/PhysRevB.8.4164
  • (55) David Perconte et al. “Low-frequency imaginary impedance at the superconducting transition of 2H-NbSe2{}_{\textrm{2}}” Publisher: American Physical Society In Phys. Rev. Appl. 13.5, 2020, pp. 054040 DOI: 10.1103/PhysRevApplied.13.054040
  • (56) Alexander C. Mark, A.. Manayil-Marathamkottil, Eduardo Henrique Toledo Poldi and Russell J. Hemley In in preparation
  • (57) Rudolf F. Graf “Modern Dictionary of Electronics” Google-Books-ID: o2I1JWPpdusC Boston: Newnes, 1999
  • (58) R. Meservey and P.. Tedrow “Measurements of the kinetic inductance of superconducting linear structures” In J. Appl. Phys 40.5, 2003, pp. 2028–2034 DOI: 10.1063/1.1657905
  • (59) Pratap Raychaudhuri and Surajit Dutta “Phase fluctuations in conventional superconductors” Publisher: IOP Publishing In J. Phys.: Condens. Matter 34.8, 2021, pp. 083001 DOI: 10.1088/1361-648X/ac360b
  • (60) D. Vodolazov “Nonlinear kinetic inductance sensor” arXiv:2312.08296 [cond-mat] In JTEP, 2023 DOI: 10.1134/S0021364023603251
  • (61) Wei Liu, Minsoo Kim, G. Sambandamurthy and N.. Armitage “Dynamical study of phase fluctuations and their critical slowing down in amorphous superconducting films” Publisher: American Physical Society In Phys. Rev. B 84.2, 2011, pp. 024511 DOI: 10.1103/PhysRevB.84.024511
  • (62) Mintu Mondal et al. “Enhancement of the finite-frequency superfluid response in the pseudogap regime of strongly disordered superconducting films” Number: 1 Publisher: Nature Publishing Group In Sci. Rep. 3.1, 2013, pp. 1357 DOI: 10.1038/srep01357
  • (63) Gufei Zhang et al. “Bosonic anomalies in boron-doped polycrystalline diamond” Publisher: American Physical Society In Phys. Rev. Appl. 6.6, 2016, pp. 064011 DOI: 10.1103/PhysRevApplied.6.064011
  • (64) A.. Jonscher “The ‘universal’ dielectric response” Number: 5613 Publisher: Nature Publishing Group In Nature 267.5613, 1977, pp. 673–679 DOI: 10.1038/267673a0
  • (65) D.. Almond and B. Vainas “The dielectric properties of random R - C networks as an explanation of the ‘universal’ power law dielectric response of solids” In J. Phys.: Condens. Matter 11.46, 1999, pp. 9081 DOI: 10.1088/0953-8984/11/46/310
  • (66) R. Bouamrane and D.. Almond “The ‘emergent scaling’ phenomenon and the dielectric properties of random resistor–capacitor networks” In J. Phys.: Condens. Matter 15.24, 2003, pp. 4089 DOI: 10.1088/0953-8984/15/24/302
  • (67) D.. Almond and C.. Bowen “Anomalous power law dispersions in AC conductivity and permittivity shown to be characteristics of microstructural electrical networks” Publisher: American Physical Society In Phys. Rev. Lett. 92.15, 2004, pp. 157601 DOI: 10.1103/PhysRevLett.92.157601
  • (68) K.. Murphy, G.. Hunt and D.. Almond “Evidence of emergent scaling in mechanical systems” Publisher: Taylor & Francis _eprint: https://doi.org/10.1080/14786430500197934 In Phil. Mag. 86.21-22, 2006, pp. 3325–3338 DOI: 10.1080/14786430500197934
  • (69) Bing Cheng et al. “Anomalous gap-edge dissipation in disordered superconductors on the brink of localization” Publisher: American Physical Society In Phys. Rev. B 93.18, 2016, pp. 180511 DOI: 10.1103/PhysRevB.93.180511
  • (70) S. Mukhopadhyay et al. “Superconductivity from a melted insulator in Josephson junction arrays” Number: 11 Publisher: Nature Publishing Group In Nat. Phys. 19.11, 2023, pp. 1630–1635 DOI: 10.1038/s41567-023-02161-w
  • (71) Eduard Ilin et al. “Superconducting phase transition in inhomogeneous chains of superconducting islands” Publisher: American Physical Society In Phys. Rev. B 102.13, 2020, pp. 134502 DOI: 10.1103/PhysRevB.102.134502
  • (72) Xu Wang, Chongpu Zhai and Yixiang Gan “Stress-dependent electrical impedance behaviours at fractal rough interfaces” Publisher: IOP Publishing In Surf. Topogr.: Metrol. Prop. 9.2, 2021, pp. 025014 DOI: 10.1088/2051-672X/abf84f
  • (73) Constantine T. Dervos and Joseph M. Michaelides “The effect of contact capacitance on current-voltage characteristics of stationary metal contacts” Conference Name: IEEE Transactions on Components, Packaging, and Manufacturing Technology: Part A In IEEE Trans. Compon. Packag. Manuf. Technol. 21.4, 1998, pp. 530–540 DOI: 10.1109/95.740045
  • (74) Chunyuan He et al. “In situ electrical impedance spectroscopy under high pressure on diamond anvil cell” In Appl. Phys. Lett. 91.9, 2007, pp. 092124 DOI: 10.1063/1.2778760
  • (75) Chunyuan He, Bingguo Liu, Ming Li and Chunxiao Gao “Alternating current impedance spectroscopy measurement under high pressure” In Rev. Sci. Instrum. 82.1, 2011, pp. 015104 DOI: 10.1063/1.3514988
  • (76) M.. Dusen “Platinum-resistance thermometry at low temperatures” Publisher: American Chemical Society In J. Am. Chem. Soc. 47.2, 1925, pp. 326–332 DOI: 10.1021/ja01679a007
  • (77) S.. Ems and J.. Swihart “Resistance peak at the superconducting transition of thin films of tin and indium” In Phys. Lett. A 37.3, 1971, pp. 255–256 DOI: 10.1016/0375-9601(71)90488-9
  • (78) Svetlana V. Postolova et al. “Reentrant resistive behavior and dimensional crossover in disordered superconducting TiN Films” Number: 1 Publisher: Nature Publishing Group In Sci. Rep. 7.1, 2017, pp. 1718 DOI: 10.1038/s41598-017-01753-w
  • (79) S. Deemyad et al. “Dependence of the superconducting transition temperature of single and polycrystalline MgB2{}_{\textrm{2}} on hydrostatic pressure” In Physica C 385.1, 2003, pp. 105–116 DOI: 10.1016/S0921-4534(02)02300-6

References

  • (1) Walter C. Michels and Norma L. Curtis “A pentode lock‐in amplifier of high frequency selectivity” In Rev. Sci. Instrum. 12.9, 1941, pp. 444–447 DOI: 10.1063/1.1769919
  • (2) C.. Stutt “Low-frequency spectrum of lock-in amplifiers” Accepted: 2004-03-03T21:52:25Z Publisher: Massachusetts Institute of Technology, Research Laboratory of Electronics In Technical Report (Massachusetts Institute of Technology. Research Laboratory of Electronics) 105, 1949 URL: https://dspace.mit.edu/handle/1721.1/4940

References

  • (1) Walter C. Michels and Norma L. Curtis “A pentode lock‐in amplifier of high frequency selectivity” In Rev. Sci. Instrum. 12.9, 1941, pp. 444–447 DOI: 10.1063/1.1769919
  • (2) C.. Stutt “Low-frequency spectrum of lock-in amplifiers” Accepted: 2004-03-03T21:52:25Z Publisher: Massachusetts Institute of Technology, Research Laboratory of Electronics In Technical Report (Massachusetts Institute of Technology. Research Laboratory of Electronics) 105, 1949 URL: https://dspace.mit.edu/handle/1721.1/4940

FigureFig. S​​

\newrefcontext

[labelprefix=S]

{strip}

Supplemental Information

Lock-in Amplification

In general, phase sensitive detection exploits the fact that many periodic waveforms are orthogonal up to a phase (see Refs. michels_pentode_1941 (82, 83)). In this work, we use sinusoidal waveforms which have the orthogonality relation,

02πsin(ωns+ϕn)\displaystyle\int_{0}^{2\pi}\mathrm{sin}(\mathit{\omega_{n}\,s}+\phi_{\mathit{n}}) sin(ωms+ϕm)ds\displaystyle\>\mathrm{sin}(\mathit{\omega_{m}\,s}+\phi_{\mathit{m}})\>d\mathit{s}
=\displaystyle= {cos(ϕnϕm):ωn=ωm0:ωnωm\displaystyle\begin{cases}\mathrm{cos}(\phi_{\mathit{n}}-\phi_{\mathit{m}}):\mathit{\omega_{n}=\omega_{m}}\\ 0:\mathit{\omega_{n}\neq\omega_{m}}\end{cases} (S1)

where ωn,m\mathit{\omega_{n,m}} are frequencies and ϕn,m\phi_{\mathit{n,m}} are phase shifts. This relation allows for the isolation and detection of small periodic signals, A(t,ω)\mathit{A(t,\omega)}. Even if the measured signal is noisy due to environmental or instrumental noise, by modulating it with a reference signal of known amplitude and ω\mathit{\omega} then integrating over at least one period, A(t,ω)\mathit{A(t,\omega)} can be recovered. In practice, the SNR can be increased further by integrating over many periods. The orthogonality of the sinusoidal carrier wave ‘averages out’ any component of A(t,ω)\mathit{A(t,\omega)} that is not oscillating at ω\mathit{\omega}. The amplifier then outputs the root mean square (RMS) of the signal that is in phase with the reference sinusoid as

X(t,ω)=2τtτtA(s,ϕA)sin(ωs+ϕ𝑟𝑒𝑓)𝑑s,\mathit{X(t,\omega)}=\frac{2}{\mathit{\tau}}\int_{\mathit{t}-\tau}^{\mathit{t}}\mathit{A}(\mathit{s},\phi_{\mathit{A}})\ \mathrm{sin}(\mathit{\omega s}+\phi_{\mathit{ref}})d\mathit{s}, (S2)

where τ\tau is an integration time, ϕ𝑟𝑒𝑓\phi_{\mathit{ref}} is the reference signal phase, and ϕA\phi_{\mathit{A}} is the signal phase. The phase difference is then θ(t)=ϕAϕ𝑟𝑒𝑓\theta(\mathit{t})=\phi_{\mathit{A}}-\phi_{\mathit{ref}}. Typically ϕ𝑟𝑒𝑓\phi_{\mathit{ref}} is adjusted manually through a unity gain null detector in the lock-in input to ensure Y(t=0,ω)=0\mathit{Y}(\mathit{t}=0,\mathit{\omega})=0 before measurements are taken.

As ϕ𝑖𝑝\phi_{\mathit{ip}} might drift throughout the experiment, whether due to sample properties or parasitic effects, a quadrature measurement (90o90^{o} out of phase from the reference signal) can also be measured and used to remove any phase ambiguity. This quadrature signal is given as

Y(t,ω)=2τtτtA(s,ϕA)cos(ωs+ϕ𝑟𝑒𝑓)𝑑s,\mathit{Y(t,\omega)}=\frac{2}{\mathit{\tau}}\int_{\mathit{t}-\tau}^{\mathit{t}}\mathit{A}(\mathit{s},\phi_{\mathit{A}})\ \mathrm{cos}(\mathit{\omega s}+\phi_{\mathit{ref}})d\mathit{s}, (S3)

For τ>>2πω\tau>>\frac{2\pi}{\mathit{\omega}} equations S2 and S3 reduce to

X(t,ω)|A𝑅𝑀𝑆(t,ω)|cos[θ(t,ω)]\mathit{X(t,\omega)}\approx|\mathit{A_{RMS}(t,\omega)}|\cos[\theta\mathit{(t,\omega)}] (S4)

and

Y(t,ω)|A𝑅𝑀𝑆(t,ω)|sin[θ(t,ω)],\mathit{Y(t,\omega)}\approx|\mathit{A_{RMS}(t,\omega)}|\ \sin[\theta\mathit{(t,\omega)}], (S5)

respectively. To relate the measurements to physical values, output signals can be converted to polar coordinates and root mean square values can be converted into amplitudes, e.g, multiplying by 2\sqrt{2} for sinusoidal carrier waves. Being mutually orthogonal, X(t,ω)\mathit{X(t,\omega)} and Y(t,ω)\mathit{Y(t,\omega)} can be defined on a complex plane to give

𝐀(t,ω)=X(t,ω)+𝑖𝑌(t,ω)=|A(t,ω)|eiθ(t,ω).\mathit{\mathbf{A}(t,\omega)}=\mathit{X(t,\omega)}+\mathit{iY(t,\omega)}=|\mathit{A(t,\omega)}|e^{\mathit{i}\theta(\mathit{t,\omega})}. (S6)

|A(t)||\mathit{A(t)}| is then the output signal magnitude and θ(t)\theta\mathit{(t)} is the phase offset from the lock in reference signal. In this work |A||\mathit{A}| is used to denote the output magnitude as opposed to the more common R\mathit{R} to avoid any ambiguity with resistance. Throughout this discussion, it is important to keep in mind that the lock-in amplifier only measures X(t,ω)\mathit{X(t,\omega)} and Y(tω)\mathit{Y(t\omega)} which are used to calculate

|A(t,ω)|=[X(t,ω)]2+[Y(t,ω)]2|\mathit{A(t,\omega)}|=\sqrt{[\mathit{X(t,\omega)}]^{2}+[\mathit{Y(t,\omega)}]^{2}} (S7)

and

θ(t,ω)=arctan(Y(t,ω)X(t,ω))\theta(\mathit{t,\omega})=\mathrm{arctan}\big{(}\frac{\mathit{Y(t,\omega)}}{\mathit{X(t,\omega)}}\big{)} (S8)

with all instrumental noise and uncertainty propagated accordingly.

Fault Detection

In order to prevent pre-amplifier loading effects, the sample was electrically floating (main text Fig. 2). In a grounded sample with inhomogeneities in the electrical resistance (Fig. S1) the current may bypass the buffered pre-amplifier entirely and short to ground, producing a zero-voltage signal for an ohmic material (Fig. S2). In order to ensure the signal is not shunted to ground the sample is electrically floating and only a differential voltage is measured (main text Fig. 2). Removing the path to ground prevents fictitious zero voltage measurements in the event of large in-line resistances or contact breakage. Increasing RloadR_{load} in the floating circuit produces increasing signal values and causes θ\theta to saturate near 80o (Fig. S3).

Refer to caption
Figure 1: Example of the schematic block diagram of a setup used for small-signal transport measurements with the sample grounded. Small parasitic resistances in the current and voltage lines are represented by R𝑐𝑢𝑟𝑟𝑒𝑛𝑡\mathit{R_{current}} and R𝑙𝑜𝑎𝑑\mathit{R_{load}}, respectively. Grounding the sample may result in fictitious zero resistance readings, and such a circuit topology should not be employed for low-resistance measurements.
Refer to caption
Figure 2: a) Voltage outputs for a grounded circuit (Fig. S1) measuring a sheet of copper. XX, YY and |V||V| as functions of RloadR_{load} (an in-line variable resistor) are plotted on a log scale. b) θ\theta response to loading effects.
Refer to caption
Figure 3: a) Voltage outputs for a floating circuit (main text Fig. 2) measuring a sheet of copper. XX, YY and |V||V| as functions of RloadR_{load} (an in-line variable resistor) are plotted on a log scale. b) θ\theta response to loading effects.

Bi-2223 Transport Data

Ambient pressure electrical transport data from sintered bars of Bi-2223 (Tc=108K\mathit{T_{c}}=108\ K) purchased from Quantum levitation. Measurements were taken using the measurement circuit presented in Fig. 2. Voltage and phase signals as functions of temperature are presented in Fig. S4.

Refer to caption
Figure 4: a) Bi-2223 lock-in signals vs temperature measured using the circuit presented in Fig. 2. b) Corresponding phase angle vs temperature from 100 K - 200 K. There is a shift in the phase angle at 185 K that is attributed to the pseudogap onset (T\mathit{T^{*}}). Below Tc\mathit{T_{c}} the signal is chaotic due to both X(T)\mathit{X(T)} and Y(T)\mathit{Y(T)} being noise resolved. Inset: Measurement over full temperature range.