License: CC BY-NC-ND 4.0
arXiv:2404.05890v1 [cond-mat.supr-con] 08 Apr 2024
\affiliation

Department of Physics, Northwestern University, Evanston, Illinois. 60208, USA

Current dependence of the low bias resistance of small capacitance Josephson junctions

Venkat Chandrasekhar
(April 8, 2024)
Abstract

The dc current-voltage characteristics of small Josephson junctions reveal features that are not observed in larger junctions, in particular, a switch to the finite voltage state at current values much less than the expected critical current of the junction and a finite resistance in the nominally superconducting regime. Both phenomena are due to the increased sensitivity to noise associated with the small capacitance of the Josephson junction and have been extensively studied a few decades ago. Here I focus on the current bias dependence of the differential resistance of the junction at low current bias in the nominally superconducting regime, using a quantum Langevin equation approach that enables a physically transparent incorporation of the noise environment of the junction. A similar approach might be useful in modeling the sensitivity of superconducting qubits to noise in the microwave regime.

Josephson junctions (JJs) form the heart of superconducting qubits, arguably one of the leading platforms for quantum computing and quantum sensing applications. Indeed, JJs and devices incorporating JJs were one of the earliest systems used to study macroscopic quantum tunneling and macroscopic quantum coherence voss_macroscopic_1981 ; devoret_measurements_1985 ; clarke_quantum_1988 ; ambegaokar_quantum_1982 ; fisher_quantum_1985 ; clarke_macroscopic_1986 ; blackburn_survey_2016 . In particular, the role of noise in the environment and its effect on quantum tunneling were extensively studied almost four decades ago, where it was found that the dissipation introduced by the noise reduced the probability of tunneling ambegaokar_quantum_1982 ; caldeira_quantum_1983 . This analysis was applied to studies of the switching current of JJs as a function of temperature voss_macroscopic_1981 ; devoret_measurements_1985 ; clarke_quantum_1988 , and later to understand the appearance of a finite resistance in the nominally zero-voltage regime in very small junctions iansiti_charging_1987 ; iansiti_crossover_1988 ; kautz_noise-affected_1990 ; martinis_classical_1989 ; joyez_josephson_1999 ; ivanchenko_josephson_1969 ; biswas_effect_1970 ; hu_low-voltage_1992 , The latter phenomena has been modeled using Monte Carlo methods primarily in the regime where thermal activation is dominant kautz_noise-affected_1990 ; buttiker_thermal_1983 ; joyez_josephson_1999 . Here I focus on the low temperature regime where quantum tunneling is dominant, using the Wentzel-Kramer-Brillouin (WKB) approximation merzbacher to calculate the tunneling rate. This approach gives better physical insight into the different factors affecting the zero-bias resistance, although the results still need to be calculated numerically.

Our starting point is the so-called resistively and capacitively shunted junction (RCSJ) model tinkham2004introduction , where the dynamics of the phase difference ϕitalic-ϕ\phiitalic_ϕ across the junction is determined by the equation

(2e)2Cd2ϕdt2+(2e)21Rdϕdt+2eIcsinϕ=2eI,superscriptPlanck-constant-over-2-pi2𝑒2𝐶superscript𝑑2italic-ϕ𝑑superscript𝑡2superscriptPlanck-constant-over-2-pi2𝑒21𝑅𝑑italic-ϕ𝑑𝑡Planck-constant-over-2-pi2𝑒subscript𝐼𝑐italic-ϕPlanck-constant-over-2-pi2𝑒𝐼\left(\frac{\hbar}{2e}\right)^{2}C\frac{d^{2}\phi}{dt^{2}}+\left(\frac{\hbar}{% 2e}\right)^{2}\frac{1}{R}\frac{d\phi}{dt}+\frac{\hbar}{2e}I_{c}\sin\phi=\frac{% \hbar}{2e}I,( divide start_ARG roman_ℏ end_ARG start_ARG 2 italic_e end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_C divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ϕ end_ARG start_ARG italic_d italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + ( divide start_ARG roman_ℏ end_ARG start_ARG 2 italic_e end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG italic_R end_ARG divide start_ARG italic_d italic_ϕ end_ARG start_ARG italic_d italic_t end_ARG + divide start_ARG roman_ℏ end_ARG start_ARG 2 italic_e end_ARG italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT roman_sin italic_ϕ = divide start_ARG roman_ℏ end_ARG start_ARG 2 italic_e end_ARG italic_I , (1)

where Planck-constant-over-2-pi\hbarroman_ℏ is the reduced Planck’s constant, e𝑒eitalic_e the electron charge, t𝑡titalic_t the time, C𝐶Citalic_C the intrinsic capacitance of the junction, R𝑅Ritalic_R the shunt resistance in parallel with C𝐶Citalic_C, Icsubscript𝐼𝑐I_{c}italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is the critical current of the junction and I𝐼Iitalic_I the applied current bias (see Fig. 1). R𝑅Ritalic_R here includes the quasiparticle resistance orr_phase_1986 as well as any contributions arising from the measuring setup, such as the source resistance of the driving electronics kautz_noise-affected_1990 ; joyez_josephson_1999 ; hu_low-voltage_1992

Refer to caption
Figure 1: Equivalent circuit of a Josephson junction, corresponding to the RCSJ model. The element with the cross represents the Josephson element with current related to the phase drop ϕitalic-ϕ\phiitalic_ϕ between the superconductors by I=Icsinϕ𝐼subscript𝐼𝑐italic-ϕI=I_{c}\sin\phiitalic_I = italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT roman_sin italic_ϕ. In the transmon configuration, the junction is shunted by a capacitor Cs>>Cmuch-greater-thansubscript𝐶𝑠𝐶C_{s}>>Citalic_C start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT > > italic_C.

As has been noted, this equation is identical to that for a particle of mass m=(/2e)2C𝑚superscriptPlanck-constant-over-2-pi2𝑒2𝐶m=(\hbar/2e)^{2}Citalic_m = ( roman_ℏ / 2 italic_e ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_C in a tilted washboard potential U=EJcosϕ(/2e)Iϕ𝑈subscript𝐸𝐽italic-ϕPlanck-constant-over-2-pi2𝑒𝐼italic-ϕU=-E_{J}\cos\phi-(\hbar/2e)I\phiitalic_U = - italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT roman_cos italic_ϕ - ( roman_ℏ / 2 italic_e ) italic_I italic_ϕ, EJ=(/2e)Icsubscript𝐸𝐽Planck-constant-over-2-pi2𝑒subscript𝐼𝑐E_{J}=(\hbar/2e)I_{c}italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = ( roman_ℏ / 2 italic_e ) italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT being the Josephson energy, in the presence of dissipation that couples to the velocity of the particle through a parameter γ=(1/RC)𝛾1𝑅𝐶\gamma=(1/RC)italic_γ = ( 1 / italic_R italic_C )

mx¨+mγx˙+U(x)=0𝑚¨𝑥𝑚𝛾˙𝑥superscript𝑈𝑥0m\ddot{x}+m\gamma\dot{x}+U^{\prime}(x)=0italic_m over¨ start_ARG italic_x end_ARG + italic_m italic_γ over˙ start_ARG italic_x end_ARG + italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_x ) = 0 (2)

In analogy with the Brownian motion of a particle, the dissipation can be thought of as arising from the influence of forces random in time balakrishnan_elements_2021 which need to be included in Eqn. (2). A more general form of this equation including these random forces can then be written as a quantum Langevin equation ford_dissipative_1988

mx¨+0μ(tt)x˙(t)𝑑t+U(x)=F(t)𝑚¨𝑥superscriptsubscript0𝜇𝑡superscript𝑡˙𝑥superscript𝑡differential-dsuperscript𝑡superscript𝑈𝑥𝐹𝑡m\ddot{x}+\int_{0}^{\infty}\mu(t-t^{\prime})\dot{x}(t^{\prime})dt^{\prime}+U^{% \prime}(x)=F(t)italic_m over¨ start_ARG italic_x end_ARG + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_μ ( italic_t - italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) over˙ start_ARG italic_x end_ARG ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_d italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT + italic_U start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_x ) = italic_F ( italic_t ) (3)

where the autocorrelation of the random force F(t)𝐹𝑡F(t)italic_F ( italic_t ) is related to the real part of the Fourier transform of the so-called memory function μ𝜇\muitalic_μ by ford_dissipative_1988

1212\displaystyle\frac{1}{2}divide start_ARG 1 end_ARG start_ARG 2 end_ARG <F(t)F(t)+F(t)F(t)>expectation𝐹𝑡𝐹superscript𝑡𝐹superscript𝑡𝐹𝑡\displaystyle<F(t)F(t^{\prime})+F(t^{\prime})F(t)>< italic_F ( italic_t ) italic_F ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) + italic_F ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_F ( italic_t ) >
=1π0[μ~(ω)]ωcoth(ω/2kBT)cosω(tt).absent1𝜋superscriptsubscript0~𝜇𝜔Planck-constant-over-2-pi𝜔hyperbolic-cotangentPlanck-constant-over-2-pi𝜔2subscript𝑘𝐵𝑇𝜔𝑡superscript𝑡\displaystyle=\frac{1}{\pi}\int_{0}^{\infty}\Re[\tilde{\mu}(\omega)]\;\hbar% \omega\coth(\hbar\omega/2k_{B}T)\cos\omega(t-t^{\prime}).= divide start_ARG 1 end_ARG start_ARG italic_π end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_ℜ [ over~ start_ARG italic_μ end_ARG ( italic_ω ) ] roman_ℏ italic_ω roman_coth ( roman_ℏ italic_ω / 2 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T ) roman_cos italic_ω ( italic_t - italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) .

In the context of the RCSJ model, F(t)𝐹𝑡F(t)italic_F ( italic_t ) would correspond to the current noise fluctuations δI(t)𝛿𝐼𝑡\delta I(t)italic_δ italic_I ( italic_t ) introduced by the shunt resistor R𝑅Ritalic_R kautz_noise-affected_1990 ; joyez_josephson_1999 , but the more general formulation here could also be applied to other cases with a more complex environmental impedance, or where the response of the system depends in a more complicated way on past history hu_low-voltage_1992 , hence the moniker ‘memory function’ for μ𝜇\muitalic_μ.

The solution of Eqn. 3 is particularly illuminating in frequency space. To simplify the analysis, we consider the case of a phase particle in the tilted washboard potential. For small values of the current I𝐼Iitalic_I, there will be local minima of the potential at ϕ=arcsinI/Icitalic-ϕ𝐼subscript𝐼𝑐\phi=\arcsin{I/I_{c}}italic_ϕ = roman_arcsin italic_I / italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT mod 2π2𝜋2\pi2 italic_π, around which the potential is locally quadratic. The potential about this minimum can be modeled as a harmonic oscillator potential U(x)=U0+(1/2)mωI02x2𝑈𝑥subscript𝑈012𝑚superscriptsubscript𝜔𝐼02superscript𝑥2U(x)=U_{0}+(1/2)m\omega_{I0}^{2}x^{2}italic_U ( italic_x ) = italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + ( 1 / 2 ) italic_m italic_ω start_POSTSUBSCRIPT italic_I 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, where the current-dependent harmonic oscillator frequency is given by ford_dissipative_1988

ωI02=ω021(I/Ic)2.superscriptsubscript𝜔𝐼02superscriptsubscript𝜔021superscript𝐼subscript𝐼𝑐2\omega_{I0}^{2}=\omega_{0}^{2}\sqrt{1-(I/I_{c})^{2}}.italic_ω start_POSTSUBSCRIPT italic_I 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT square-root start_ARG 1 - ( italic_I / italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (4)

Here ω0subscript𝜔0\omega_{0}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the frequency for I=0𝐼0I=0italic_I = 0, and is given by ω0=4e2EJ/C=8EcEJPlanck-constant-over-2-pisubscript𝜔04superscript𝑒2subscript𝐸𝐽𝐶8subscript𝐸𝑐subscript𝐸𝐽\hbar\omega_{0}=\sqrt{4e^{2}E_{J}/C}=\sqrt{8E_{c}E_{J}}roman_ℏ italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = square-root start_ARG 4 italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT / italic_C end_ARG = square-root start_ARG 8 italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT end_ARG, where Ec=e2/2Csubscript𝐸𝑐superscript𝑒22𝐶E_{c}=e^{2}/2Citalic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 italic_C is the single electron charging energy.

Fourier transforming Eqn. 3, the solution can be represented as x~(ω)=α(ω)F~(ω)~𝑥𝜔𝛼𝜔~𝐹𝜔\tilde{x}(\omega)=\alpha(\omega)\tilde{F}(\omega)over~ start_ARG italic_x end_ARG ( italic_ω ) = italic_α ( italic_ω ) over~ start_ARG italic_F end_ARG ( italic_ω ), where x~(ω)~𝑥𝜔\tilde{x}(\omega)over~ start_ARG italic_x end_ARG ( italic_ω ) and F~(ω)~𝐹𝜔\tilde{F}(\omega)over~ start_ARG italic_F end_ARG ( italic_ω ) are the Fourier transforms of x(t)𝑥𝑡x(t)italic_x ( italic_t ) and F(t)𝐹𝑡F(t)italic_F ( italic_t ) respectively, and the response function α(ω)𝛼𝜔\alpha(\omega)italic_α ( italic_ω ) is given by ford_dissipative_1988

α(ω)=1mω2+iωμ~(ω)+mωI02.𝛼𝜔1𝑚superscript𝜔2𝑖𝜔~𝜇𝜔𝑚superscriptsubscript𝜔𝐼02\alpha(\omega)=\frac{1}{-m\omega^{2}+i\omega\tilde{\mu}(\omega)+m\omega_{I0}^{% 2}}.italic_α ( italic_ω ) = divide start_ARG 1 end_ARG start_ARG - italic_m italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_i italic_ω over~ start_ARG italic_μ end_ARG ( italic_ω ) + italic_m italic_ω start_POSTSUBSCRIPT italic_I 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (5)

As usual, the poles of α(ω)𝛼𝜔\alpha(\omega)italic_α ( italic_ω ) give the excitation energies of the system. For our problem, μ~(ω)~𝜇𝜔\tilde{\mu}(\omega)over~ start_ARG italic_μ end_ARG ( italic_ω ) can be related to the environmental impedance coupled to the JJ, μ~(ω)=(/2e)2Y(ω)~𝜇𝜔superscriptPlanck-constant-over-2-pi2𝑒2𝑌𝜔\tilde{\mu}(\omega)=(\hbar/2e)^{2}Y(\omega)over~ start_ARG italic_μ end_ARG ( italic_ω ) = ( roman_ℏ / 2 italic_e ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_Y ( italic_ω ), where Y(ω)𝑌𝜔Y(\omega)italic_Y ( italic_ω ) is the effective admittance of the environment connected in parallel to the JJ hu_low-voltage_1992 ; joyez_josephson_1999 .

As an example, consider the case for I=0𝐼0I=0italic_I = 0 when there is only a shunt capacitor Cssubscript𝐶𝑠C_{s}italic_C start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT connected across the JJ, as is done for a qubit in the transmon configuration (see Fig. 1). In this case, Y(ω)=iωCs𝑌𝜔𝑖𝜔subscript𝐶𝑠Y(\omega)=i\omega C_{s}italic_Y ( italic_ω ) = italic_i italic_ω italic_C start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. It is easy to show that the excitation energy is then modified to ω0=4e2EJ/C4e2EJ/Ctsubscript𝜔04superscript𝑒2subscript𝐸𝐽𝐶4superscript𝑒2subscript𝐸𝐽subscript𝐶𝑡\omega_{0}=\sqrt{4e^{2}E_{J}/C}\rightarrow\sqrt{4e^{2}E_{J}/C_{t}}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = square-root start_ARG 4 italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT / italic_C end_ARG → square-root start_ARG 4 italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT / italic_C start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_ARG, where Ct=Cs+Csubscript𝐶𝑡subscript𝐶𝑠𝐶C_{t}=C_{s}+Citalic_C start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = italic_C start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + italic_C is the parallel combination of the intrinsic capacitance C𝐶Citalic_C and the shunt capacitance Cssubscript𝐶𝑠C_{s}italic_C start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, as expected. If we now include also ohmic dissipation in the form of a resistor R𝑅Ritalic_R connected in parallel with the junction as shown in Fig. 1, Y(ω)𝑌𝜔Y(\omega)italic_Y ( italic_ω ) is now modified to Y(ω)=(1/R)+iωCs𝑌𝜔1𝑅𝑖𝜔subscript𝐶𝑠Y(\omega)=(1/R)+i\omega C_{s}italic_Y ( italic_ω ) = ( 1 / italic_R ) + italic_i italic_ω italic_C start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and the oscillation frequency is modified to ωI02γ2/4superscriptsubscript𝜔𝐼02superscript𝛾24\sqrt{\omega_{I0}^{2}-\gamma^{2}/4}square-root start_ARG italic_ω start_POSTSUBSCRIPT italic_I 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 4 end_ARG, where γ=1/RCt𝛾1𝑅subscript𝐶𝑡\gamma=1/RC_{t}italic_γ = 1 / italic_R italic_C start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT. The ohmic dissipation also introduces a finite lifetime of τ=2/γ𝜏2𝛾\tau=2/\gammaitalic_τ = 2 / italic_γ. Classically, this would be the time scale over which the particle stops oscillating and comes to rest at the local minimum of the tilted washboard potential. One should note that is the behaviour of the average position or velocity: the instantaneous position will continue to exhibit random fluctuations centered about the minimum in potential driven by noise, with corresponding fluctuations in the instantaneous velocity about zero.

Quantum mechanically, one expects to have a series of quantized energy levels, with the ground state energy ωI0/2Planck-constant-over-2-pisubscript𝜔𝐼02\hbar\omega_{I0}/2roman_ℏ italic_ω start_POSTSUBSCRIPT italic_I 0 end_POSTSUBSCRIPT / 2. One can think of the dissipation due to noise as leading to a finite lifetime for the state arising from transitions to other states. For a particle in the ground state in a tilted washboard potential, this would occur through transitions out of the local minimum to states in neighboring local minima, which might occur through thermal activation or quantum tunneling caldeira_quantum_1983 ; ambegaokar_quantum_1982 . This of course is the model that has been used to describe the stochastic switching of a current-biased JJ out of the zero-voltage state at currents close to the critical current Icsubscript𝐼𝑐I_{c}italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. In this case, once the particle tunnels out of the local minimum, it then continues down the tilted washboard potential in the so-called free-running state. Consequently, the phase is continually evolving, and a finite voltage results.

Here we are concerned with the low bias regime, where the current I𝐼Iitalic_I is close to zero. There may still be a finite probability of transition to neighboring minima, but the presence of dissipation means that the particle loses its small amount of excess energy after passage to the nearest minimum, or perhaps after traversing a few such minima, and so never goes into the free-running state. However, this process of phase diffusion will still result in a finite voltage across the junction. For simplicity, let us consider the case when it stops at the neighboring minimum. Each such transition would result in a phase change of 2π2𝜋2\pi2 italic_π; if the rate of such transitions is ΓΓ\Gammaroman_Γ, then the average dϕ/dt=2πΓ𝑑italic-ϕ𝑑𝑡2𝜋Γd\phi/dt=2\pi\Gammaitalic_d italic_ϕ / italic_d italic_t = 2 italic_π roman_Γ would result in a finite voltage (/2e)dϕ/dt=(h/2e)ΓPlanck-constant-over-2-pi2𝑒𝑑italic-ϕ𝑑𝑡2𝑒Γ(\hbar/2e)d\phi/dt=(h/2e)\Gamma( roman_ℏ / 2 italic_e ) italic_d italic_ϕ / italic_d italic_t = ( italic_h / 2 italic_e ) roman_Γ.

Both thermal activation and quantum tunneling can lead to transitions. To calculate the tunneling rate, I use the WKB approximation to calculate the tunneling probability P𝑃Pitalic_P merzbacher

P=Ae2abκ(x)𝑑x,𝑃𝐴superscript𝑒2superscriptsubscript𝑎𝑏𝜅𝑥differential-d𝑥P=Ae^{-2\int_{a}^{b}\kappa(x)dx},italic_P = italic_A italic_e start_POSTSUPERSCRIPT - 2 ∫ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_b end_POSTSUPERSCRIPT italic_κ ( italic_x ) italic_d italic_x end_POSTSUPERSCRIPT , (6)

where

κ(x)=2m(U(x)E)2.𝜅𝑥2𝑚𝑈𝑥𝐸superscriptPlanck-constant-over-2-pi2\kappa(x)=\sqrt{\frac{2m(U(x)-E)}{\hbar^{2}}}.italic_κ ( italic_x ) = square-root start_ARG divide start_ARG 2 italic_m ( italic_U ( italic_x ) - italic_E ) end_ARG start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG . (7)

Here A𝐴Aitalic_A is a constant that is determined by the normalization of the wavefunction in the WKB approximation and is essentially unity, E=ωI0/2𝐸Planck-constant-over-2-pisubscript𝜔𝐼02E=\hbar\omega_{I0}/2italic_E = roman_ℏ italic_ω start_POSTSUBSCRIPT italic_I 0 end_POSTSUBSCRIPT / 2, and a𝑎aitalic_a and b𝑏bitalic_b are the positions in the potential where E=U(x)𝐸𝑈𝑥E=U(x)italic_E = italic_U ( italic_x ). The probability of thermal activation over the barrier for a state at energy E𝐸Eitalic_E for I0similar-to𝐼0I\sim 0italic_I ∼ 0 should go as e(2EJE)/kBTsimilar-toabsentsuperscript𝑒2subscript𝐸𝐽𝐸subscript𝑘𝐵𝑇\sim e^{-(2E_{J}-E)/k_{B}T}∼ italic_e start_POSTSUPERSCRIPT - ( 2 italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT - italic_E ) / italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_POSTSUPERSCRIPT at a temperature T𝑇Titalic_T: here we consider temperatures where this term is much smaller than the tunneling term Eqn. 6. The transition rate is then given by the attempt frequency multiplied by P𝑃Pitalic_P. Following Gamow Gamow1928ZurQD , we take the attempt frequency to be ωI0/2subscript𝜔𝐼02\omega_{I0}/2italic_ω start_POSTSUBSCRIPT italic_I 0 end_POSTSUBSCRIPT / 2, so that Γ=ωI0P/2Γsubscript𝜔𝐼0𝑃2\Gamma=\omega_{I0}P/2roman_Γ = italic_ω start_POSTSUBSCRIPT italic_I 0 end_POSTSUBSCRIPT italic_P / 2. For small currents, the particle can tunnel to the left and right with rates ΓlsubscriptΓ𝑙\Gamma_{l}roman_Γ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT and ΓrsubscriptΓ𝑟\Gamma_{r}roman_Γ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT. The resulting voltage is then V=(h/2e)(ΓrΓl)𝑉2𝑒subscriptΓ𝑟subscriptΓ𝑙V=(h/2e)(\Gamma_{r}-\Gamma_{l})italic_V = ( italic_h / 2 italic_e ) ( roman_Γ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT - roman_Γ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT ).

There are two relevant energy scales in the problem, the charging energy Ecsubscript𝐸𝑐E_{c}italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and the Josephson energy EJsubscript𝐸𝐽E_{J}italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT with the results depending on the ratio Ec/EJsubscript𝐸𝑐subscript𝐸𝐽E_{c}/E_{J}italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT. For a fixed EJsubscript𝐸𝐽E_{J}italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT, a larger Ecsubscript𝐸𝑐E_{c}italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT corresponds classically to a particle with a smaller mass and thus one more susceptible to noise; quantum mechanically, since the energy of the state increases with respect to the barrier, the particle is more likely to tunnel. We should therefore expect the voltage to increase with increasing Ec/EJsubscript𝐸𝑐subscript𝐸𝐽E_{c}/E_{J}italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT, as has been verified in recent experiments lu_phase_2023 . Dissipation reduces the energy of the state, hence we expect increased dissipation should reduce the probability of tunneling, in line with previous analyses caldeira_quantum_1983 , and therefore reduce the voltage.

Refer to caption
Figure 2: The calculated zero bias resistance due to quantum tunneling as a function of Ec/EJsubscript𝐸𝑐subscript𝐸𝐽E_{c}/E_{J}italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT for three different values of the resistance R𝑅Ritalic_R. Icsubscript𝐼𝑐I_{c}italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT here is assumed to be 30 nA, which sets the value of EJsubscript𝐸𝐽E_{J}italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT

At I=0𝐼0I=0italic_I = 0, V=0𝑉0V=0italic_V = 0, but the zero bias resistance R0=dV/dIsubscript𝑅0𝑑𝑉𝑑𝐼R_{0}=dV/dIitalic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_d italic_V / italic_d italic_I can be finite and increases rapidly with Ec/EJsubscript𝐸𝑐subscript𝐸𝐽E_{c}/E_{J}italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT. This is shown in Fig. 2 for three different values of the shunt resistance R𝑅Ritalic_R, assuming a value of Ic=30subscript𝐼𝑐30I_{c}=30italic_I start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 30 nA, consistent with design parameters on recent transmon qubits wisne_jj_2024 . For Ec/EJ<0.5subscript𝐸𝑐subscript𝐸𝐽0.5E_{c}/E_{J}<0.5italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT < 0.5, the dependence is superexponential, with R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT drop** many orders of magnitude over a narrow range of Ec/EJsubscript𝐸𝑐subscript𝐸𝐽E_{c}/E_{J}italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT. For Ec/EJ>1subscript𝐸𝑐subscript𝐸𝐽1E_{c}/E_{J}>1italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT > 1 the dependence is weaker but still exponential. Changing the shunt resistance R𝑅Ritalic_R seems to have little effect until R𝑅Ritalic_R is of the order of a few kΩΩ\Omegaroman_Ω or less, and even then the major effect is at larger values of Ec/EJsubscript𝐸𝑐subscript𝐸𝐽E_{c}/E_{J}italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT. A similar calculation using thermal activation instead of tunneling to generate the phase jumps gives R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT values many orders of magnitude smaller than those shown in Fig, 2, assuming a temperature of 25 mK, showing that phase jumps due to thermal activation can be neglected at these temperatures. These results are similar to those obtained earlier by Hu and O’Connell hu_low-voltage_1992 .

Refer to caption
Figure 3: Differential resistance dV/dI𝑑𝑉𝑑𝐼dV/dIitalic_d italic_V / italic_d italic_I of a Josephson junction as a function of bias current I𝐼Iitalic_I for three different values of the shunt resistance R𝑅Ritalic_R. Ec/EJ=0.9subscript𝐸𝑐subscript𝐸𝐽0.9E_{c}/E_{J}=0.9italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = 0.9.

Let us now consider the dependence of the differential resistance dV/dI𝑑𝑉𝑑𝐼dV/dIitalic_d italic_V / italic_d italic_I as a function of I𝐼Iitalic_I near the zero bias regime. (At larger values of I𝐼Iitalic_I, the JJ will transition to the free-running state, which we do not concern ourselves with here.) Figure 3 shows the differential resistance as a function of bias current for three different values of the shunt resistor R𝑅Ritalic_R, for a specific value of Ec/EJ=0.9subscript𝐸𝑐subscript𝐸𝐽0.9E_{c}/E_{J}=0.9italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = 0.9. There are two competing effects that modify the tunneling as the bias current is increased. The height of the barrier decreases as I𝐼Iitalic_I is increased, which might be expected to give a rapid increase in the differential resistance as it is easier for the phase particle to transition to a neighboring minimum. However, the energy of the ground state also decreases, which counterbalances somewhat the decrease in barrier height, leading to only a relatively gradual increase in resistance with bias. As seen earlier above, for a fixed value of Ec/EJsubscript𝐸𝑐subscript𝐸𝐽E_{c}/E_{J}italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT in this range, the overall value of the differential resistance decreases when the shunt resistance R𝑅Ritalic_R drops below a few kΩΩ\Omegaroman_Ω. One should also note that the overall curvature of dV/dI𝑑𝑉𝑑𝐼dV/dIitalic_d italic_V / italic_d italic_I vs. I𝐼Iitalic_I also decreases as R𝑅Ritalic_R falls below a few kΩΩ\Omegaroman_Ω. These curves are similar to those obtained on small Josephson junctions wisne_jj_2024 .

In principle, one should be able to determine the effective capacitance and shunt resistance in the measurement by fitting measurements of dV/dI𝑑𝑉𝑑𝐼dV/dIitalic_d italic_V / italic_d italic_I vs I𝐼Iitalic_I to this model. In reality, however, the detailed impedance environment of the Josephson junction will likely be different from the simple model we have assumed. The power of using the quantum Langevin approach is that a variety of enviromental impedances can be modeled readily by appropriately specifying the admittance Y(ω)𝑌𝜔Y(\omega)italic_Y ( italic_ω ), as has been discussed earlier.

While we have focused here on dc electrical transport Josephson junctions, the parameters that determine their sensitivity to noise should also be applicable in the microwave regime, the usual frequency regime for superconducting qubits. It should be possible to include various noise sources by modeling them as components of a frequency dependence admittance Y(ω)𝑌𝜔Y(\omega)italic_Y ( italic_ω ). In particular, it might be possible to model the influence of two-level-systems (TLSs), which have been identified as a major source of decoherence in superconducting qubits Kjaergaard_2020 .

This work is supported by the U.S. Department of Energy, Office of Science, National Quantum Information Science Research Centers, Superconducting Quantum Materials and Systems Center (SQMS) under Contract No. DEAC02-07CH11359.

References

  • (1) R. F. Voss, R. A. Webb, Macroscopic Quantum Tunneling in 1-μ𝜇\muitalic_μm Nb Josephson Junctions, Phys. Rev. Lett. 47 (4) (1981) 265–268. doi:10.1103/PhysRevLett.47.265.
  • (2) M. H. Devoret, J. M. Martinis, J. Clarke, Measurements of Macroscopic Quantum Tunneling out of the Zero-Voltage State of a Current-Biased Josephson Junction, Phys. Rev. Lett. 55 (18) (1985) 1908–1911. doi:10.1103/PhysRevLett.55.1908.
  • (3) J. Clarke, A. N. Cleland, M. H. Devoret, D. Esteve, J. M. Martinis, Quantum Mechanics of a Macroscopic Variable: The Phase Difference of a Josephson Junction, Science 239 (4843) (1988) 992–997. doi:10.1126/science.239.4843.992.
  • (4) V. Ambegaokar, U. Eckern, G. Schön, Quantum Dynamics of Tunneling between Superconductors, Phys. Rev. Lett. 48 (25) (1982) 1745–1748. doi:10.1103/PhysRevLett.48.1745.
  • (5) M. P. A. Fisher, W. Zwerger, Quantum Brownian motion in a periodic potential, Phys. Rev. B 32 (10) (1985) 6190–6206. doi:10.1103/PhysRevB.32.6190.
  • (6) J. Clarke, G. Schön, Macroscopic Quantum Phenomena in Josephson Elements, Europhys. News 17 (7-8) (1986) 94–96. doi:10.1051/epn/19861707094.
  • (7) J. A. Blackburn, M. Cirillo, N. Grønbech-Jensen, A survey of classical and quantum interpretations of experiments on Josephson junctions at very low temperatures, Physics Reports 611 (2016) 1–33. doi:10.1016/j.physrep.2015.10.010.
  • (8) A. O. Caldeira, A. J. Leggett, Quantum tunnelling in a dissipative system, Annals of Physics 149 (2) (1983) 374–456. doi:10.1016/0003-4916(83)90202-6.
  • (9) M. Iansiti, A. T. Johnson, W. F. Smith, H. Rogalla, C. J. Lobb, M. Tinkham, Charging energy and phase delocalization in single very small Josephson tunnel junctions, Phys. Rev. Lett. 59 (4) (1987) 489–492. doi:10.1103/PhysRevLett.59.489.
  • (10) M. Iansiti, A. T. Johnson, C. J. Lobb, M. Tinkham, Crossover from Josephson Tunneling to the Coulomb Blockade in Small Tunnel Junctions, Phys. Rev. Lett. 60 (23) (1988) 2414–2417. doi:10.1103/PhysRevLett.60.2414.
  • (11) R. L. Kautz, J. M. Martinis, Noise-affected I - V curves in small hysteretic Josephson junctions, Phys. Rev. B 42 (16) (1990) 9903–9937. doi:10.1103/PhysRevB.42.9903.
  • (12) J. M. Martinis, R. L. Kautz, Classical phase diffusion in small hysteretic Josephson junctions, Phys. Rev. Lett. 63 (14) (1989) 1507–1510. doi:10.1103/PhysRevLett.63.1507.
  • (13) P. Joyez, D. Vion, M. Götz, M. H. Devoret, D. Esteve, The Josephson Effect in Nanoscale Tunnel Junctions, Journal of Superconductivity 12 (6) (1999) 757–766. doi:10.1023/A:1007733009637.
  • (14) Y. M. Ivanchenko, A. Zil’Berman, THE JOSEPHSON EFFECT IN SMALL TUNNEL CONTACTS, Sov. Phys. JETP. (1969) 1272.
  • (15) A. C. Biswas, S. S. Jha, Effect of Thermal Noise on the dc Josephson Effect, Phys. Rev. B 2 (7) (1970) 2543–2547. doi:10.1103/PhysRevB.2.2543.
  • (16) G. Y. Hu, R. F. O’Connell, Low-voltage resistance in small Josephson junctions, J. Phys.: Condens. Matter 4 (48) (1992) 9635–9642. doi:10.1088/0953-8984/4/48/017.
  • (17) M. Büttiker, E. P. Harris, R. Landauer, Thermal activation in extremely underdamped Josephson-junction circuits, Phys. Rev. B 28 (3) (1983) 1268–1275. doi:10.1103/PhysRevB.28.1268.
  • (18) E. Merzbacher, Quantum Mechanics, 2nd Edition, John Wiley & Sons, 1961.
  • (19) M. Tinkham, Introduction to Superconductivity, 2nd Edition, Dover Publications, 2004.
  • (20) B. G. Orr, J. R. Clem, H. M. Jaeger, A. M. Goldman, Phase fluctuations in Josephson junctions, Phys. Rev. B 34 (5) (1986) 3491–3494. doi:10.1103/PhysRevB.34.3491.
  • (21) Elements of Nonequilibrium Statistical Mechanics, Cham.
  • (22) G. Ford, J. Lewis, R. O’Connell, Dissipative quantum tunneling: quantum Langevin equation approach, Physics Letters A 128 (1-2) (1988) 29–34. doi:10.1016/0375-9601(88)91037-7.
  • (23) G. Gamow, Zur quantentheorie des atomkernes, Zeitschrift für Physik 51 (1928) 204–212.
  • (24) W.-S. Lu, K. Kalashnikov, P. Kamenov, T. J. DiNapoli, M. E. Gershenson, Phase Diffusion in Low-EJ Josephson Junctions at Milli-Kelvin Temperatures, Electronics 12 (2) (2023) 416. doi:10.3390/electronics12020416.
  • (25) M. Wisne, in preparation (2024).
  • (26) M. Kjaergaard, M. E. Schwartz, J. Braumüller, P. Krantz, J. I.-J. Wang, S. Gustavsson, W. D. Oliver, Superconducting qubits: Current state of play, Annual Review of Condensed Matter Physics 11 (1) (2020) 369–395. arXiv:https://doi.org/10.1146/annurev-conmatphys-031119-050605, doi:10.1146/annurev-conmatphys-031119-050605.