Linear ac response of diffusive SNS junctions

6
Linear AC Response of Diffusive SNS Junctions Pauli Virtanen, 1, 2 F. Sebasti´ an Bergeret, 3, 4 Juan Carlos Cuevas, 5 and Tero T. Heikkil¨ a 2 1 Institute for Theoretical Physics and Astrophysics, University of W¨ urzburg, D-97074 W¨ urzburg, Germany 2 Low Temperature Laboratory, Aalto University, P.O. Box 15100, FI-00076 AALTO, Finland 3 Centro de F´ ısica de Materiales (CFM), Centro Mixto CSIC-UPV/EHU, Edificio Korta, Avenida de Tolosa 72, E-20018 San Sebasti´ an, Spain 4 Donostia International Physics Center (DIPC), Manuel de Lardizbal 4, E-20018 San Sebasti´an, Spain 5 Departamento de F´ ısica Te´ orica de la Materia Condensada, Universidad Aut´ onoma de Madrid, E-28049 Madrid, Spain (Dated: December 21, 2010) We explore the behavior of the ac admittance of superconductor-normal metal-superconductor (SNS) junctions as the phase difference of the order parameters between the superconductors is varied. We find three characteristic regimes, defined by comparing the driving frequency ω to the inelastic scattering rate Γ and the Thouless energy ET of the junction (typically ~Γ ET ). Only in the first regime ω Γ the usual picture of the kinetic inductance holds. We show that the ac admittance can be used to directly access some of the characteristic quantities of the SNS junctions, in particular the phase dependent energy minigap and the typically phase dependent inelastic scattering rate. Our results partially explain the recent measurements of the linear response properties of SNS Superconducting Quantum Interference Devices (SQUIDs) and predict a number of new effects. The frequency dependent susceptibility typically re- veals information about the internal dynamics of the studied systems. In the electronic case, this susceptibility is more often measured as admittance, whose frequency dependence in semiclassical models for bulk metals is due to scattering and appears for frequencies exceeding some tens of THz [1], and in wires it is dictated by stray ca- pacitances and geometric inductance. Understanding the frequency-dependent response is moreover of importance for high-frequency devices. In superconductors or superconducting tunnel junc- tions (SIS), the admittance at low frequencies is domi- nated by the superconducting kinetic inductance [2]. For SIS junctions, this Josephson inductance is related to the supercurrent I S (ϕ) through the junction, which depends on the superconducting phase difference ϕ: L J (ϕ) -1 = 2e ~ ϕ I S (ϕ) . (1) This relation allows characterizing the current-phase re- lation of Josephson junctions via measurements of their ac admittance [3]. The remaining dissipative part of the admittance is due to quasiparticles, is proportional to exp[-|Δ - ~ω|/(k B T )] [4], and is important only for the high frequencies of the order of the superconducting gap Δ or temperatures T close to the critical temperature. For other types of Josephson junctions than SIS the admittance, however, can deviate from the above simple picture. In this Letter we show how combining normal metal wires (N) and superconductors (S) into SNS weak links results into an admittance that entails characteris- tics of the inelastic scattering rates and the inverse dif- fusion times through the structure. At frequencies of the order or larger than the inelastic scattering rate Γ, the simple kinetic inductance picture has to be revised to in- clude non-adiabatic effects associated with the dynamics of the electron distribution. The dissipative response, de- scribing microwave absorption, is moreover finite for tem- peratures or frequencies exceeding the phase-dependent minigap in the spectrum of excitations inside the junc- tion. It probes the density of states in the junction, and is related to the physics of stimulation and suppression of the supercurrent [5, 6]. Here, we study diffusive SNS junctions, whose length L is longer than the superconducting coherence length ξ 0 = p ~D/Δ, where D is the diffusion constant. The proximity effect induces a gap in the density of states inside the normal metal, E g (ϕ =0)=3.12E T Δ, where E T = ~D/L 2 is the inverse diffusion time. The minigap depends on the phase difference approximately as cos(ϕ/2) and vanishes for ϕ = π. The coupling to the electric field is modeled by assuming an ac bias voltage, V (t)= δV sin(ωt), which induces an oscillating phase difference ϕ(t)= ϕ + δφ cos(ωt) across the junction. To find quantitative results, we describe the SNS junc- tion dynamics with the Keldysh-Usadel equation [7], used also in Ref. [6]. In this approach, physical quantities are obtained from the Retarded, Advanced and Keldysh Green’s functions ˆ g R/A/K (E,E 0 ), which depend on two energy arguments. These functions are matrices in the Nambu (electron-hole) space, and the Keldysh (K) part can be parameterized in terms of an electron distribu- tion function matrix ˆ h(E,E 0 )= h L (E,E 0 )+h T (E,E 0 τ 3 : ˆ g K g Rˆ h- ˆ hˆ g A , where matrix products involve also con- volutions, ( ˆ B ˆ C)(E,E 0 )= R -∞ dE1 2π ˆ B(E,E 1 ) ˆ C(E 1 ,E 0 ). In the presence of the harmonic drive, these functions arXiv:1012.4294v1 [cond-mat.supr-con] 20 Dec 2010

Transcript of Linear ac response of diffusive SNS junctions

Linear AC Response of Diffusive SNS Junctions

Pauli Virtanen,1, 2 F. Sebastian Bergeret,3, 4 Juan Carlos Cuevas,5 and Tero T. Heikkila2

1Institute for Theoretical Physics and Astrophysics,University of Wurzburg, D-97074 Wurzburg, Germany

2Low Temperature Laboratory, Aalto University, P.O. Box 15100, FI-00076 AALTO, Finland3Centro de Fısica de Materiales (CFM), Centro Mixto CSIC-UPV/EHU,

Edificio Korta, Avenida de Tolosa 72, E-20018 San Sebastian, Spain4Donostia International Physics Center (DIPC),

Manuel de Lardizbal 4, E-20018 San Sebastian, Spain5Departamento de Fısica Teorica de la Materia Condensada,Universidad Autonoma de Madrid, E-28049 Madrid, Spain

(Dated: December 21, 2010)

We explore the behavior of the ac admittance of superconductor-normal metal-superconductor(SNS) junctions as the phase difference of the order parameters between the superconductors isvaried. We find three characteristic regimes, defined by comparing the driving frequency ω tothe inelastic scattering rate Γ and the Thouless energy ET of the junction (typically ~Γ � ET ).Only in the first regime ω � Γ the usual picture of the kinetic inductance holds. We show thatthe ac admittance can be used to directly access some of the characteristic quantities of the SNSjunctions, in particular the phase dependent energy minigap and the typically phase dependentinelastic scattering rate. Our results partially explain the recent measurements of the linear responseproperties of SNS Superconducting Quantum Interference Devices (SQUIDs) and predict a numberof new effects.

The frequency dependent susceptibility typically re-veals information about the internal dynamics of thestudied systems. In the electronic case, this susceptibilityis more often measured as admittance, whose frequencydependence in semiclassical models for bulk metals is dueto scattering and appears for frequencies exceeding sometens of THz [1], and in wires it is dictated by stray ca-pacitances and geometric inductance. Understanding thefrequency-dependent response is moreover of importancefor high-frequency devices.

In superconductors or superconducting tunnel junc-tions (SIS), the admittance at low frequencies is domi-nated by the superconducting kinetic inductance [2]. ForSIS junctions, this Josephson inductance is related to thesupercurrent IS(ϕ) through the junction, which dependson the superconducting phase difference ϕ:

LJ(ϕ)−1 =2e

~∂ϕIS(ϕ) . (1)

This relation allows characterizing the current-phase re-lation of Josephson junctions via measurements of theirac admittance [3]. The remaining dissipative part of theadmittance is due to quasiparticles, is proportional toexp[−|∆− ~ω|/(kBT )] [4], and is important only for thehigh frequencies of the order of the superconducting gap∆ or temperatures T close to the critical temperature.

For other types of Josephson junctions than SIS theadmittance, however, can deviate from the above simplepicture. In this Letter we show how combining normalmetal wires (N) and superconductors (S) into SNS weaklinks results into an admittance that entails characteris-tics of the inelastic scattering rates and the inverse dif-fusion times through the structure. At frequencies of the

order or larger than the inelastic scattering rate Γ, thesimple kinetic inductance picture has to be revised to in-clude non-adiabatic effects associated with the dynamicsof the electron distribution. The dissipative response, de-scribing microwave absorption, is moreover finite for tem-peratures or frequencies exceeding the phase-dependentminigap in the spectrum of excitations inside the junc-tion. It probes the density of states in the junction, andis related to the physics of stimulation and suppressionof the supercurrent [5, 6].

Here, we study diffusive SNS junctions, whose lengthL is longer than the superconducting coherence lengthξ0 =

√~D/∆, where D is the diffusion constant. The

proximity effect induces a gap in the density of statesinside the normal metal, Eg(ϕ = 0) = 3.12ET � ∆,where ET = ~D/L2 is the inverse diffusion time. Theminigap depends on the phase difference approximatelyas cos(ϕ/2) and vanishes for ϕ = π. The coupling to theelectric field is modeled by assuming an ac bias voltage,V (t) = δV sin(ωt), which induces an oscillating phasedifference ϕ(t) = ϕ+ δφ cos(ωt) across the junction.

To find quantitative results, we describe the SNS junc-tion dynamics with the Keldysh-Usadel equation [7], usedalso in Ref. [6]. In this approach, physical quantitiesare obtained from the Retarded, Advanced and KeldyshGreen’s functions gR/A/K(E,E′), which depend on twoenergy arguments. These functions are matrices in theNambu (electron-hole) space, and the Keldysh (K) partcan be parameterized in terms of an electron distribu-tion function matrix h(E,E′) = hL(E,E′)+hT (E,E′)τ3:

gK = gRh−hgA, where matrix products involve also con-volutions, (BC)(E,E′) =

∫∞−∞

dE1

2π B(E,E1)C(E1, E′).

In the presence of the harmonic drive, these functions

arX

iv:1

012.

4294

v1 [

cond

-mat

.sup

r-co

n] 2

0 D

ec 2

010

2

can be written in a matrix representation gn,m(E)=g(E+n~ω,E + m~ω) [8] that reduces convolutions to matrixproducts.

The ac admittance is Y (ω) = 2iωδφI(ω), for a linear-

response drive δφ � 2π. The ac current harmonic isI(ω) = σNS

4

∫∞−∞ dE Trτ3j

K01(E), where σN and S are the

normal-state conductivity and the cross section of thejunction, and the Keldysh current is jK = gR∂xg

K +gK ∂xg

A. Here x is the position in the junction, and∂xB = ∂xB− i[Aτ3, B] the gauge-covariant gradient con-taining the vector potential A(E,E′) = A(E)δ(E − E′).

In the above approach, the admittance splits naturallyinto three gauge-invariant parts, Y = Ysc + Ydy + Yqp,where (hereafter, ~ = e = kB = 1)

Ysc =σNS

2iωδφ

∫ ∞−∞

dE Tr{τ3(jR01h11 − h00jA01)} , (2a)

Ydy =σNS

2iωδφ

∫ ∞−∞

dE Tr{τ3(jR00h01 − h01jA11)} , (2b)

Yqp =σNS

2iωδφ

∫ ∞−∞

dE Tr{

(1− τ3gA11τ3gR00) (2c)

× [τ3∂xh01 −iA0

2(h11 − h00)]

}.

Here jR/A = gR/A∂xgR/A describe spectral (su-

per)currents, h00(E) = h11(E − ω) = tanh(E/2T )

the equilibrium electron distribution, and h01 its time-dependent part. The contributions describe (a) the acsupercurrent, (b) effect of the dynamic variation of thepopulations of the Andreev levels, and (c) the quasiparti-cle current driven directly by the field. Below, we mostlywork in a gauge in which the electric field is contained inthe vector potential, A(t) = A0 cos(ωt), A0 = −δφ/(2L).Requiring charge neutrality leads to a finite position de-pendent scalar potential, but our numerics indicate thatthis can be disregarded for ω . 10ET .

In the following, our aim is to relate the contributions(2) to quantities that can be calculated in the absenceof the ac drive [9]. The regimes where the different con-tributions are relevant depend on the particular valuesof the phase difference, frequency and temperature. Themain results are summarized in Fig. 1, which shows thephase dependence of the admittance in different regimesof frequencies and temperatures.

Low frequency: For frequencies satisfying Γ � ω �Eg(ϕ), the superconducting correlations follow the time-dependent phase difference, but the electron distribu-tion is driven out of equilibrium. In this limit, thetime-dependent supercurrent (2a) yields Eq. (1), i.e.,Ysc ' −2i∂ϕIS(ϕ)/ω. Since the supercurrent IS decaysexponentially as the temperature increases, this contri-bution to reactance becomes unimportant at high tem-peratures (T & 10ET ), unless the frequency is very low.

The second major contribution to reactance comesfrom a dynamic variation of the population, as givenby Ydy. For this, the time dependent component h01 ≡

00

204060

600040002000

02000

0048

12

6020

2060

02468

1006020

2060

0

1

2

0.50.51.52.5

0

1

2

0.50.00.51.0

0.00.51.01.5

0.0

0.1

0.2

low 10ET highT

adiab

atic

low

E Thi

gh

FIG. 1. (Color online): Phase dependence of the admittanceY in units of the normal state conductance GN (dissipativepart: solid, reactive part: dashed), in different regimes ofinterest. The “low” temperature results have been calculatedfor T = ET , the “high” temperature results for T = 16ET .The adiabatic frequency is ω = ET /200 � Γ = ET /20, thelow ω = ET /4 � Γ, and the high frequency is ω = 10ET .Dotted lines show the contribution from dominant parts, Ysc

[Eqs. (1), (8)], Ydy [Eq. (5)], Yqp [Eqs. (7), (9)], indicated inthe lower left corners. The adiabatic frequency results areobtained from analytical approximations, the others from thefull numerics.

h′L+ τ3h′T of the distribution function needs to be solved

from a kinetic equation. Assuming again simple time de-pendence for the superconducting correlations, the firstharmonic of the Usadel kinetic equation reads (cf. [6])

D∂x · (DL∂xh′L − T ∂xh′T + jSh′T ) = (3a)

iA0

2(jS + ∂xT )(h11 − h00) + i(ω − 2iΓ)Nh′L ,

D∂x · (DT∂xh′T + T ∂xh′L + jSh′L) = 0. (3b)

Here DL/T are the spectral heat/charge diffusion coef-ficients, T is an anomalous kinetic coefficient and Nis the local density of states. These quantities arerelated to the equilibrium Retarded Green’s function,gR(E) = g(E)τ3 + f(E)τ↑ − f(E)τ↓, e.g., jS = Im jE ,

jE = 12i Tr τ3j

R00, N = Re g, and DT = 1

2 [1+|g|2+|f |2/2+

|f |2/2].At low frequencies, Eqs. (3) should be solved with

Andreev reflection boundary conditions amounting toh′T = 0 and ∂x ·h′Ln = 0 at the two NS interfaces. The re-sulting function h′T is in the vector potential gauge finitebut small, and as a first approximation we can disregardit. Moreover, gradients of h′L are small due to Andreev

3

002468

101214

Q(

)

03210123456

P()

(a) (b)

FIG. 2. (Color online): (a) Function Q(ϕ, T ) (see Eq. (5))describing the phase-dependent dynamic contribution to thereactive response and (b) function P (ϕ, T ) (Eq. (7)) de-scribing the phase dependence of the dissipative part ofthe admittance at low frequencies. From top to bottomin (a) and marked with symbols in (b): T/ET = 16(M), 8(O), 4(�), 2(◦), 1(+), 0.5(×). The dotted lines representthe analytic high-temperature approximations to whichQ andP tend for T � ET . The inset of (b) shows the temperaturedependence of G0(T ).

reflection, and we get a fairly good estimate for the av-erage h′L by averaging Eqs. (3) over the normal-metal

junction, defining 〈·〉 =∫ L0dx · /L. As a result, we get

h01 ≈ 〈h′L〉 ≈A0

2

jS(h11 − h00)

(ω − 2iΓ)〈N〉. (4)

Substituting this to Eq. (2b) and assuming ω . ET ,

Ydy ≡−iGNω − 2iΓ

ETTQ(ϕ, T ) (5)

≈ −iGNω − 2iΓ

∫ ∞−∞

dEj2S

4T 〈N〉 cosh2[E/(2T )].

This is similar to a correction to the dc conductancedescribed by Lempitskii [10]. Its origin can be under-stood as follows [11]: the current is carried by a densespectrum {εn(ϕ)} of discrete bound states with popula-tions fn, j(ϕ) =

∑n(∂ϕεn)fn(εn). With ac bias one finds

δj/δV ∝∑n ∂

2ϕεnfn/iω+

∑n(∂ϕεn)2∂εfn/iω, where the

first term is equivalent to Ysc and the second one to Ydy,as jS ∼

∑n δ(E − εn)∂ϕεn ∼ N∂ϕε.

The dynamic contribution is purely dissipative andconstant for ω � Γ, contains both reactive and dissipa-tive components for ω ≈ Γ, and becomes purely reactiveand decays for ω � Γ, as visible in Fig. 1. In general,Γ also depends on the phase difference [12], so that thephase-dependent response at frequencies of the order ofthe inelastic scattering rate may be quite complicated.On the other hand, Eq. (5) offers a way to probe suchphase-dependent scattering rates via an admittance mea-surement.

The function Q(ϕ, T ) is shown in Fig. 2(a) fordifferent temperatures. At T � ET the re-sponse can be fitted with the function Q(ϕ) ≈

8.9(| saw(ϕ)| − 0.32 sin2(ϕ/2)

)| sin(ϕ)| where saw(ϕ) =

(ϕ + π mod 2π) − π. At low temperatures, Q(ϕ, T ) issuppressed for phases at which T < Eg(ϕ). Note thatQ is positive definite, has almost a double periodicitycompared to the kinetic inductance term, and has a min-imum around ϕ = 0, where the kinetic inductance is atmaximum.

The dissipative part of the impedance originates fromtwo additional sources: the quasiparticle part Yqp, and,importantly at low temperatures, a part of the AC super-current oscillating in phase with the voltage. The quasi-particle contribution is easiest to derive in a gauge wherethe vector potential vanishes. Then, h′T in Eqs. (3) has

the boundary conditions h′T (x = ±L/2) = ± iδφ4 (h11 −h00), and solving Eq. (3) assuming h′L ≈ 0 yields

Yqp + ReYsc ' GN∫ ∞−∞

dEK0(E)

4T cosh2[E/(2T )], (6)

≡ G0(T ) +GNETT

P (ϕ, T ), (7)

where we assumed ω . ET , and defined G0(T ) as theϕ = 0 value [P (0, T ) = 0]. The kernel is K0 ≈〈D−1T 〉−1 − Re ∂ϕjE ∼ 〈N2 + 1

4 |f + f∗|2〉 [13]. It de-scribes the spectrum of excitations in the junction avail-able for receiving energy — this has a minigap Eg(ϕ), sothat for ω, T < Eg(ϕ) all dissipation vanishes. Note thatthe appearance of the minigap is related to the presenceof the AC supercurrent contribution: Yqp ≥ GN and issimilar to the usual proximity-enhanced conductance.

The functions P (ϕ, T ) and G0(T ) are shown inFig. 2(b). At low temperatures, the temperature andphase dependence shows a clear signature of the pres-ence of a minigap in the density of states; note that thisalso applies to the Ydy contribution. The dissipation isconcentrated at phase differences close to π where theminigap is small. In the high-temperature limit, the dissi-pative term consists of a phase independent contributionG0(T ) and the phase-dependent part has a simple phaseand temperature dependence, −0.23 saw(ϕ)2GNET /T .

High frequency: When the frequency becomes of the or-der of the Thouless energy ET , the above semi-adiabaticexpressions break down as the spectral quantities becomefrequency dependent. We can however construct approx-imations also in this limit (for full expressions, see Ap-pendix). The supercurrent contribution is fairly approx-imated by

Ysc ≈σNS

∫ ∞−∞

dE

{∂ϕ[jE(E) + jE(E + ω)]∗h00(E)

− ∂ϕ[jE(E) + jE(E + ω)]h11(E)

},

(8)

4

100 101

/ET

10 1

100

101Y/

GN

100 101

/ET

10 1

100

101

102

FIG. 3. (Color online): Amplitude δY =maxϕ[Re / Im]Y (ϕ) − minϕ[Re / Im]Y (ϕ) of the phase-dependence in the nonadiabatic dissipative (Re) and reactive(Im) admittance of a long SNS junction, obtained by solvingthe time-dependent Usadel equations numerically. The linescorrespond to temperatures T/ET = 0.5, 1, 2, 4, 8, 16 (top tobottom) and were calculated with Γ = 0.05ET .

and the quasiparticle part of the impedance by

Yqp ' GN∫ ∞−∞

dEK(E,ω)h11(E)− h00(E)

2ω, (9)

K(E,ω) =1

2

⟨[1 + g(E)g(E + ω)∗ (10)

+1

2f(E)f(E + ω)∗ +

1

2f(E)f(E + ω)∗

]−1⟩−1.

The dynamic contribution Ydy can be neglected for ω �ET . The above is compared to full numerical solutionsin the top panels in Fig. 1.

At high temperatures, Ysc is exponentially suppressedeven for high frequencies, similarly as the equilibrium su-percurrent. Consequently, Yqp dominates for ω, T � ET(see Fig. 1). At low temperatures, both the supercurrentand quasiparticle contributions are important.

The amplitude of the phase dependence in the ad-mittance is illustrated in Fig. 3, up to high frequencies.As the frequency increases, the proximity-induced phasedependence in the reactive and the dissipative compo-nents decays, as the admittance approaches the constantnormal-state value, Y (ω, ϕ)→ GN . Similarly, the phasedependence vanishes as the temperature increases. Atlow frequencies, on the other hand, the reactive compo-nent diverges due to the Josephson inductance, and thedissipative component is dominated by Ydy.

In a recent experiment probing directly the ac admit-tance of SNS junctions [11], it was found that at frequen-cies Γ� ω � ET (semi-adiabatic limit), the reactive re-sponse follows closely our prediction consisting of the sumof the Josephson inductance and the dynamical correc-tion (5). However, both the dissipative contribution andthe dependence at high frequencies ω & ET are different:in [11], the dissipative contribution is directly related tothe reactive contribution, and moreover the amplitude ofphase oscillations in susceptibility δχ = iωδY decays asfrequency increases. The characteristic frequency scale

for this was found to be temperature independent and ofthe order of the Thouless energy. In contrast, in this fre-quency range our Usadel model predicts δχ ∼ const. —however, it is for example possible that the simple relax-ation time approximation does not include all interactionmechanisms playing a role in the experiment.

Finally, we remark that our work amounts essentiallyto deriving the parameters for the Resistively ShuntedJunction (RSJ) model of SNS junctions [2]: there, theJosephson inductance resulting from the supercurrentterm should be modified to include the nonadiabatic cor-rection, Eq. (5), and the shunt resistor describing thedissipation in the junction should be replaced by the dis-sipative terms presented in Eqs. (5,7).

In conclusion, we have described the frequency-dependent admittance of diffusive superconductor-normal metal-superconductor junctions and shown howthe simple adiabatic Josephson inductance picture ismodified once the frequency is increased. Besides study-ing the dynamics of the system, the detailed frequencydependence can be used to study directly the inelasticscattering rates. Our results are also relevant for devicesutilizing high-frequency properties of SNS junctions, suchas those used in metrology and radiation detection.

We thank H. Bouchiat, S. Gueron, K. Tikhonov, andM. Feigelman for discussions that in part motivated thiswork, and CSC (Espoo) for computer resources. Thiswork was supported by the Academy of Finland, the ERC(Grant No. 240362-Heattronics), the Spanish MICINN(Contract No. FIS2008-04209), and the Emmy-Noetherprogram of the DFG.

[1] N. Ashcroft and N. Mermin, Solid State Physics (Saun-ders College, Philadelphia, 1976).

[2] M. Tinkham, Introduction to Superconductivity, 2nd ed.(McGraw-Hill, New York, 1996).

[3] A. A. Golubov, M. Y. Kupriyanov, and E. Il’ichev, Rev.Mod. Phys. 76, 411 (2004).

[4] J. Tucker and M. Feldman, Rev. Mod. Phys. 57, 1055(1985).

[5] J. M. Warlaumont, J. C. Brown, T. Foxe, and R. A.Buhrman, Phys. Rev. Lett. 43, 169 (1979); M. Fuechsle,J. Bentner, D. A. Ryndyk, M. Reinwald, W. Wegschei-der, and C. Strunk, Phys. Rev. Lett. 102, 127001 (2009);L. G. Aslamazov and S. Lempitskii, Sov. Phys. JETP 55,967 (1982).

[6] P. Virtanen, T. T. Heikkila, F. S. Bergeret, and J. C.Cuevas, Phys. Rev. Lett. 104, 247003 (2010).

[7] K. D. Usadel, Phys. Rev. Lett. 25, 507 (1970); W. Belzig,F. K. Wilhelm, C. Bruder, G. Schon, and A. D. Zaikin,Superlatt. Microstruct. 25, 1251 (1999).

[8] J. C. Cuevas, J. Hammer, J. Kopu, J. K. Viljas, andM. Eschrig, Phys. Rev. B 73, 184505 (2006).

[9] These quantities may be calculated for exam-ple with the publicly available Usadel solver, seehttp://ltl.tkk.fi/˜theory/usadel1/.

5

[10] S. Lempitskii, Sov. Phys. JETP 58, 624 (1983).[11] F. Chiodi, M. Ferrier, K.Tikhonov, P. Virtanen,

T. Heikkila, M. Feigelman, S. Gueron, and H.Bouchiat,(2010), arXiv:1005.0406.

[12] T. T. Heikkila and F. Giazotto, Phys. Rev. B 79, 094514(2009).

[13] The latter approximation is the one used in [6] and cap-tures the minigap correctly, whereas the former gives abetter approximation to the integral, but is inaccurate atenergies E . Eg(ϕ).

On approximations

Although one cannot solve the time-dependent linear-response Usadel equations analytically in closed form, the

admittance can be obtained with quantitative accuracy ifthe solution at equilibrium is known. The approximationprocedure is outlined in the main text. Below, we showintermediate results before taking the ω → 0 limit, anddemonstrate that the results agree with the full numericalapproach which solves the ac Usadel equation exactly (seethe Supplementary information of Ref. 6 for details of thenumerics).

Following the procedure outlined in the main text, thefollowing approximations can be derived:

Y = σS

∫ ∞−∞

dE Y (E) = σS

∫ ∞−∞

dE [Ysc(E) + Ydy(E) + Yqp(E)] , (11a)

Ysc(E) ' −1

(∂φ[jE(E) + jE(E + ω)]

2h(E + ω)− ∂φ[jE(E) + jE(E + ω)]∗

2h(E)

), (11b)

Ydy(E) ' 1

i

ω − 2iΓ

(jE(E)− jE(E + ω)∗)2

12 〈g(E) + g(E + ω)∗〉

(h(E + ω)− h(E)) , (11c)

Yqp(E) ' 1

⟨[1 + g(E)g(E + ω)∗ +

1

2f(E)f(E + ω)∗ +

1

2f(E)f(E + ω)∗]−1

⟩−1(h(E + ω)− h(E)) . (11d)

Above, h(E) = tanh[E/(2T )], and the energy-dependentquantities jE(E) = −i

2 Tr[τ3gR∇gR] and gR(E) =

g(E)τ3 + f(E) τ1+iτ22 − f(E) τ1−iτ22 are obtained from theequilibrium Usadel equations [7], which are much fasterto solve than the full ac equations.

Figures 4, 5, and 6 illustrate that the above approxima-tions reproduce all qualitative features visible in the fullynumerical solution, and are quantitatively accurate in alarge part of the parameter regime we are interested in.In particular, the deviations between the two approaches

are miniscule for frequencies lower than ET . At large fre-quencies there are clear quantitative differences, but thequalitative phase and frequency dependence is the same,and the resulting admittances are of the same order ofmagnitude. This demonstrates that to a fair accuracy theac admittance can be well described by using Eqs. (11)and standard solvers for the equilibrium Usadel equation.Besides that, Eq. (11) provides a possibility for makinganalytical estimates for the admittance contributions.

6

00

204060

600040002000

02000

0048

12

6020

2060

02468

1006020

2060

0

1

2

0.50.51.52.5

0

1

2

0.50.00.51.0

0.00.51.01.5

0.0

0.1

0.2

low 10ET highT

adiab

atic

low

E Thi

gh

00

204060

600040002000

02000

0048

12

6020

2060

02468

1006020

2060

0

1

2

0.50.51.52.5

0

1

2

0.50.00.51.0

0.00.51.01.5

0.0

0.1

0.2

low 10ET highT

adiab

atic

low

E Thi

ghFIG. 4. Left: Fig. 1 in the main text. Right: Fig. 1 with all data computed from Eqs. (11).

100 101

/ET

10 1

100

101

Y/G

N

100 101

/ET

10 1

100

101

102

100 101

/ET

10 1

100

101

Y/G

N

100 101

/ET

10 1

100

101

102

FIG. 5. Left: Fig. 3 in the main text. Right: Fig. 3 with all data computed from Eqs. (11).

0 2 4 6 8 10/ET

0

5

10

15

20

k BT

/ET

0 2 4 6 8 10/ET

2.01.61.20.80.4

0.00.40.81.21.62.0

log 1

0|

Y/G

N|

0 2 4 6 8 10/ET

0

5

10

15

20

k BT

/ET

0 2 4 6 8 10/ET

2.01.61.20.80.4

0.00.40.81.21.62.0

log 1

0|

Y/G

N|

FIG. 6. As Fig. 3 in the main text, but expressed as a color plot. Left: numerically computed admittance oscillations. Right:admittance oscillations from Eqs. (11).