anomalous metallic phase in tunable destructive

10
Anomalous Metallic Phase in Tunable Destructive Superconductors S. Vaitiek˙ enas, 1 P. Krogstrup, 1 and C. M. Marcus 1 1 Center for Quantum Devices and Microsoft Quantum Lab–Copenhagen, Niels Bohr Institute, University of Copenhagen, 2100 Copenhagen, Denmark (Dated: March 3, 2020) Multiply connected superconductors smaller than the coherence length show destructive super- conductivity, characterized by reentrant quantum phase transitions driven by magnetic flux. We investigate the dependence of destructive superconductivity on flux, transverse magnetic field, tem- perature, and current in InAs nanowires with a surrounding epitaxial Al shell, finding excellent agreement with mean-field theory across multiple reentrant transitions. Near the crossover between destructive and nondestructive regimes, an anomalous metal phase is observed with temperature- independent resistance, controlled over two orders of magnitude by a millitesla-scale transverse magnetic field. Quantum phase transitions (QPTs) [1, 2] in con- ventional superconductors serve as prototypes for re- lated effects in more complex, strongly-correlated sys- tems [3], including heavy-fermion materials [4] and high-temperature superconductors [5]. While low- temperature superconductors are well understood in bulk, new phenomena can arise in mesoscopic sam- ples and reduced dimensionality [3, 7]. For instance, in two-dimensional films, electrons theoretically con- dense into either a superconductor or insulator in the low-temperature limit [8]. Yet, in many instances, an anomalous metallic state with temperature-independent resistance is found at low temperatures [9]. In one- dimensional wires, incoherent phase slips can destroy su- perconductivity [10] or give rise to an anomalous metallic state [11], while coherent quantum phase slips can lead to superposition of quantum states enclosing different num- bers of flux quanta [12], potentially useful as a qubit [13]. Multiply connected superconductors provide an even richer platform for investigating phase transitions. Flux- oid quantization in units of Φ 0 = h/2e [14, 15], reveals not only electron pairing but a complex macroscopic or- der parameter, Δe [3, 16]. The same physical mech- anism underlies the Little-Parks effect, a periodic mod- ulation of the transition temperature, T C , of a super- conducting cylinder with magnetic flux period Φ 0 [17]. For hollow superconducting cylinders with diameter, d, smaller than the coherence length, the modulation am- plitude can exceed zero-field transition temperature, T C0 , leading to a reentrant destruction of superconductivity near odd half-integer multiples of Φ 0 [18–20]. Early experimental investigation of the destructive Little-Parks effect reported reentrant superconductivity interrupted by an anomalous-resistance phase around ap- plied flux Φ 0 /2 [21]. Subsequent experiments showed a low-temperature phase with normal-state resistance, R N , around Φ 0 /2, but did not display fully recovered super- conductivity at higher flux [22]. Several theoretical mod- els were proposed to interpret these different scenarios [23–25], but no consensus emerged. Here, we report a study of the Little-Parks effect across the transition from destructive to nondestructive regimes, in InAs nanowires with a thin epitaxial cylin- drical Al shell. Remarkable agreement with Ginzburg- Landau mean field theory is observed across multiple reentrant lobes as a function of flux, temperature, and current bias, using independently measured material and device parameters. We then investigate a field-tunable crossover from non-destructive to destructive regime. At the boundary, an anomalous metal phase is identified, characterized by a temperature-independent resistance that can be tuned over two orders of magnitude using small changes in perpendicular magnetic field, B . We interpret these results in terms of tunneling between ad- jacent fluxoid states with different phase winding num- bers giving rise to an anomalous metallic phase. As noted previously [24], the appearance of a field-tunable temperature-independent resistance does not emerge nat- urally from simple models. The basic mechanism leading to a field-tunable saturating resistance remains mysteri- ous. The devices we investigated were made using InAs nanowire grown by the vapor-liquid-solid (VLS) method using molecular beam epitaxy (MBE). Following wire growth, an epitaxial Al layer was grown within the MBE I S V S 1 μm dC dF tS (a) (b) V V G Ti/Au InAs Al B B FIG. 1. (a) Colorized material-sensitive scanning electron mi- crograph of InAs-Al hybrid nanowire cross-section. The full wire diameter dF, core diameter dC and shell thickness tS are indicated by dashed arrows. (b) Representative color- enhanced micrograph of a device (wire B) consisting of an InAs core (green) with Al shell (grey), contacted with Ti/Au leads (yellow). The device can be operated in voltage (V ) and current (IS) bias measurement set-ups. arXiv:1909.10654v2 [cond-mat.mes-hall] 29 Feb 2020

Upload: others

Post on 06-Feb-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Anomalous Metallic Phase in Tunable Destructive

Anomalous Metallic Phase in Tunable Destructive Superconductors

S. Vaitiekenas,1 P. Krogstrup,1 and C. M. Marcus1

1Center for Quantum Devices and Microsoft Quantum Lab–Copenhagen,Niels Bohr Institute, University of Copenhagen, 2100 Copenhagen, Denmark

(Dated: March 3, 2020)

Multiply connected superconductors smaller than the coherence length show destructive super-conductivity, characterized by reentrant quantum phase transitions driven by magnetic flux. Weinvestigate the dependence of destructive superconductivity on flux, transverse magnetic field, tem-perature, and current in InAs nanowires with a surrounding epitaxial Al shell, finding excellentagreement with mean-field theory across multiple reentrant transitions. Near the crossover betweendestructive and nondestructive regimes, an anomalous metal phase is observed with temperature-independent resistance, controlled over two orders of magnitude by a millitesla-scale transversemagnetic field.

Quantum phase transitions (QPTs) [1, 2] in con-ventional superconductors serve as prototypes for re-lated effects in more complex, strongly-correlated sys-tems [3], including heavy-fermion materials [4] andhigh-temperature superconductors [5]. While low-temperature superconductors are well understood inbulk, new phenomena can arise in mesoscopic sam-ples and reduced dimensionality [3, 7]. For instance,in two-dimensional films, electrons theoretically con-dense into either a superconductor or insulator in thelow-temperature limit [8]. Yet, in many instances, ananomalous metallic state with temperature-independentresistance is found at low temperatures [9]. In one-dimensional wires, incoherent phase slips can destroy su-perconductivity [10] or give rise to an anomalous metallicstate [11], while coherent quantum phase slips can lead tosuperposition of quantum states enclosing different num-bers of flux quanta [12], potentially useful as a qubit [13].

Multiply connected superconductors provide an evenricher platform for investigating phase transitions. Flux-oid quantization in units of Φ0 = h/2e [14, 15], revealsnot only electron pairing but a complex macroscopic or-der parameter, ∆eiϕ [3, 16]. The same physical mech-anism underlies the Little-Parks effect, a periodic mod-ulation of the transition temperature, TC, of a super-conducting cylinder with magnetic flux period Φ0 [17].For hollow superconducting cylinders with diameter, d,smaller than the coherence length, the modulation am-plitude can exceed zero-field transition temperature, TC0,leading to a reentrant destruction of superconductivitynear odd half-integer multiples of Φ0 [18–20].

Early experimental investigation of the destructiveLittle-Parks effect reported reentrant superconductivityinterrupted by an anomalous-resistance phase around ap-plied flux Φ0/2 [21]. Subsequent experiments showed alow-temperature phase with normal-state resistance, RN,around Φ0/2, but did not display fully recovered super-conductivity at higher flux [22]. Several theoretical mod-els were proposed to interpret these different scenarios[23–25], but no consensus emerged.

Here, we report a study of the Little-Parks effect

across the transition from destructive to nondestructiveregimes, in InAs nanowires with a thin epitaxial cylin-drical Al shell. Remarkable agreement with Ginzburg-Landau mean field theory is observed across multiplereentrant lobes as a function of flux, temperature, andcurrent bias, using independently measured material anddevice parameters. We then investigate a field-tunablecrossover from non-destructive to destructive regime. Atthe boundary, an anomalous metal phase is identified,characterized by a temperature-independent resistancethat can be tuned over two orders of magnitude usingsmall changes in perpendicular magnetic field, B⊥. Weinterpret these results in terms of tunneling between ad-jacent fluxoid states with different phase winding num-bers giving rise to an anomalous metallic phase. Asnoted previously [24], the appearance of a field-tunabletemperature-independent resistance does not emerge nat-urally from simple models. The basic mechanism leadingto a field-tunable saturating resistance remains mysteri-ous.

The devices we investigated were made using InAsnanowire grown by the vapor-liquid-solid (VLS) methodusing molecular beam epitaxy (MBE). Following wiregrowth, an epitaxial Al layer was grown within the MBE

IS

VS

1 μmdC

dF

tS

(a) (b)

V

VG

Ti/Au InAsAl BBⅡ

FIG. 1. (a) Colorized material-sensitive scanning electron mi-crograph of InAs-Al hybrid nanowire cross-section. The fullwire diameter dF, core diameter dC and shell thickness tSare indicated by dashed arrows. (b) Representative color-enhanced micrograph of a device (wire B) consisting of anInAs core (green) with Al shell (grey), contacted with Ti/Auleads (yellow). The device can be operated in voltage (V ) andcurrent (IS) bias measurement set-ups.

arX

iv:1

909.

1065

4v2

[co

nd-m

at.m

es-h

all]

29

Feb

2020

Page 2: Anomalous Metallic Phase in Tunable Destructive

2

0-0.4 -0.3 -0.2 -0.1 0.1 0.2 0.3 0.4

T (K

) 1.0

0

0.5

1.5

BⅡ (T)

(a) -2Φ0 -Φ0 Φ0 2Φ00

0

30

RS (Ω

)

0-0.3 -0.2 -0.1 0.1 0.2 0.3BⅡ (T)

T (K

)

1.0

0

0.5

(b) -2Φ0 -Φ0 0 Φ0 2Φ0

0

1.5

RS (Ω

)

wire A

wire B

theory

theory

FIG. 2. (a) Shell resistance, RS, measured for wire A withshell thickness tS = 7 nm as a function of axial magneticfield, B‖, and temperature, T . The superconducting transi-tion temperature of the shell is periodically modulated by B‖.The sample is superconducting for temperatures below 1 Kthroughout the whole measured B‖ range. The dashed theorycurve is Eq. 1 evaluated with α‖ from Eq. 2 and the corre-sponding fit parameters measured for the wire A. (b) Sameas (a), but measured for wire B with shell thickness aroundtS = 25 nm, showing the destructive regimes around ±Φ0/2and ±3Φ0/2 of the applied flux quantum.

chamber while rotating the sample stage, resulting in afull cylindrical Al shell coating the wire [1], as shown inFig. 1(a). Subsequent fabrication used standard electron-beam lithography, deposition, etching, and liftoff, as de-scribed elsewhere [27]. Devices were operated in two con-figurations [Fig. 1(b)]: In the first configuration, four Aucontacts were made to the Al shell allowing four-wire re-sistance measurements; In the second, an additional tun-neling contact to the InAs core at the end of the wire wasused as a tunnel probe, giving local density of states, asdiscussed in Ref. [27]. We investigated wires from threegrowth batches, denoted A, B, and C, with different corediameters, dC, and shell thicknesses, tS (see Supplemen-tal Material [28]). Transport measurements were car-ried out in a dilution refrigerator with a three-axis vectormagnet and base temperature of 20 mK.

Carrier density in the InAs core is predominantly atthe Al interface due to band bending [30, 31]. Moreover,the density of carriers in Al is orders of magnitude higherthan in InAs. We may therefore consider current to becarried by a hollow cylinder which is threaded by flux inan axial applied magnetic field. Induced circumferentialsupercurrents from the applied flux lead to Cooper pairbreaking, characterized by the parameter α, which con-trols the transition temperature TC(α), as described byAbrikosov-Gorkov expression,

0-0.3 -0.2 -0.1 0.1 0.2 0.3BⅡ (T)

I S (μ

A)

80

0

-80

(b) -2Φ0 -Φ0 0 Φ0 2Φ0

0

1.5

RS (Ω

)

0-2.0-2.5 -1.5 -1.0 -0.5 0.5 1.0 1.5 2.0 2.5

I S (μ

A)

0

-30

30

BⅡ (T)

(a) -10Φ0-15Φ0 -5Φ0 5Φ0 10Φ0 15Φ00

0

30R

S (Ω)

wire A theory

wire B theory

FIG. 3. (a) Shell resistance, RS, measured for wire A withshell thickness tS = 7 nm as a function of axial magnetic field,B‖, and current bias, IS. Both switching and re-trapping cur-rents are periodically modulated by B‖ up to B‖,C = 2.3 T,whereafter the supercurrent is suppressed. The dashed the-ory curve is Eq. 3 evaluated with α‖ from Eq. 2 and thecorresponding fit parameters measured for wire A. (b) Sameas (a), but measured for wire B with shell thickness aroundtS = 25 nm.

ln

(TC(α)

TC0

)= Ψ

(1

2

)−Ψ

(1

2+

α

2πkBTC(α)

), (1)

where Ψ is the digamma function [32]. FollowingRefs. [20, 22, 23], the pair-breaking parameter for a hol-low cylinder with wall thickness tS in a parallel magneticfield B‖ is given by

α‖ =4 ξS

2TC0

AF

[(n− Φ

Φ0

)2

+tS

2

dF2

(Φ2

Φ02

+n2

3

)], (2)

where ξS is the zero-field superconducting coherencelength, AF is the area of the cylinder cross section,the integer n is the fluxoid quantum number, Φ isthe applied flux, and dF is the diameter of the cylin-der [Fig. 1(a)]. Taking the dirty-limit expression forξS =

√π~vFle/24kBTC0 with the Fermi velocity vF

and mean free path le, we note that all parameters caneither be measured directly from the micrograph of thedevice or from independent transport measurements (seeSupplemental Material [28]).

Differential shell resistances, RS = dVS/dIS, for wiresA and B are shown in Fig. 2 as a function of B‖ andtemperature, T . Wires A and B have similar core di-ameters, dC ∼ 135 nm, but different shell thicknesses.For wire A, with tS = 7 nm, TC is finite throughout themeasured range, and varies periodically with applied ax-ial flux with amplitude ∼ 0.4 K with no clear envelope

Page 3: Anomalous Metallic Phase in Tunable Destructive

3

reduction up to B‖ = 0.4 T. Normal-state resistance ofthe wire yields a coherence length ξS = 70 nm, smallerthan dC (see Supplemental Material [28]). In contrast,wire B, with tS = 25 nm, has ξS = 180 nm > dC, andshows destructive regimes around ±Φ0/2 and ±3Φ0/2.Resistances in these destructive regimes remain equal tothe normal state resistance, RS = RN, to the lowest mea-sured temperatures.

The absence (presence) of the destructive regime inwire A (B) is consistent with the criterion of the su-perconducting coherence length being smaller (larger)than the wire diameter [18]. To be more quantita-tive, we plot in Fig. 2 theoretical curves marking thesuperconductor-metal transition based on Eqs. 1 and 2with independently measured wire parameters, using ei-ther the measured zero-field critical temperature, TC0

or, equivalently, the spectroscopically measured zero-fieldsuperconducting gap, ∆0, [Fig. S1 in Supplemental Ma-terial [28]], which was consistent with the BCS relation∆0 = 1.76 kBTC0 [3]. Figure 2 demonstrates the remark-ably good agreement found between experiment and the-ory. The observed increase of TC with decreasing tS isconsistent with enhanced energy gaps for thin Al films[33].

Similar to the effects of flux-induced circumferentialsupercurrent, a dc current, IS, applied along the wire canalso drive the shell normal. The field-dependent criticalcurrent IC(α) can be related to the corresponding criticaltemperature, TC(α),

IC(α) = IC0

(TC(α)

TC0

)3/2

, (3)

where IC0 is the zero-field critical current [34].

Base-temperature IS–B‖ phase diagrams for wires Aand B are shown in Fig. 3. The data are taken sweepingfrom negative to positive bias, so show re-trapping cur-rents for IS < 0 and switching current for IS > 0, both ofwhich are proportional to the critical current, IC [3]. Sim-ilar to the transition temperature, IC was observed to beΦ0-periodic in flux for both wires as expected from Eq. 3.For wire A, a bigger range of B‖ [Fig. 3(a)] shows thatthe thin shell remains non-destructive up to ∼ 2 T, cor-responding to ∼ 13Φ0, then enters the destructive regimetwice around 14Φ0 and finally turns fully normal aroundB‖,C = 2.3 T.

Figure 3 shows theoretical curves based on Eqs. 1–3superimposed on experimental data for both wire types.The zero-field switching and re-trapping currents weretaken as input parameters, with other parameters mea-sured independently. Again, excellent agreement be-tween experiment and theory for both thin (wire A) andthick (wire B) shells was found.

We next consider the effects of an applied transversemagnetic field, B⊥, which can be used to control a

-Φ0 Φ00

I S (μ

A)

0

-80

80 B = 0

wire C theory

I S (μ

A)

0

-40

40 B = 12 mT

wire C theory

0 50-50BⅡ (mT)

I S (μ

A)

0

-40

40 B = 13 mT

wire C theory

(a)

(b)

(c)

(d)BⅡ = 0

wire C theory-80

0

80

I S (μ

A)

B (mT)0 20-20

0.5

0

RS (Ω

)

FIG. 4. (a) Base-temperature shell resistance, RS, mea-sured for wire C as a function of current bias, IS, and paral-lel magnetic field, B‖, at zero perpendicular magnetic field,B⊥ = 0. Approximately equal switching and re-trapping cur-rents, which are proportional to the critical current, indicatenearly dissipationless supercurrent injection. The wire is non-destructive throughout the whole measured B‖ range. (b)Same as (a), but at B⊥ = 12 mT. Around half-flux quantumand zero-current bias, an anomalous phase develops with afinite, but smaller than normal state resistance. (c) Same as(a), but measured at B⊥ = 13 mT. Around the half-flux quan-tum RS remains at normal state value even at IS = 0. Thetheory curves in (a)-(c) are Eq. 3 evaluated with α = α‖+α⊥.(d) Critical current evolution as a function of B⊥ measuredat B‖ = 0. The theory curve in (d) is Eq. 3 computed withα⊥.

crossover between conventional and destructive Little-Parks regimes. We investigate the combined effects of B‖and B⊥ in wire C, with dC = 240 nm and tS = 40 nm.The larger diameter reduces the field value B‖ = Φ0/AF

and the thicker shell ensures a long ξS, such that initiallythe wire is nearly destructive. The transition of the wireC from being non-destructive at B⊥ = 0 to destructiveat B⊥ = 13 mT is depicted by IS–B‖ phase diagrams inFig. 4(a)-(c).

Theoretically, the effect of B⊥ on the superconductingtransition can be accounted for by introducing an addi-

Page 4: Anomalous Metallic Phase in Tunable Destructive

4

B = 4 mT

theory0

0.4

0.8

1.2

T (K

)

0.3

0

RS (Ω

)

theory0

0.4

0.8

T (K

)

B = 11 mT

B = 13 mT

0

0.4

0.8

T (K

)

0 40-40BⅡ (mT)

theory

RS (Ω

)

0.01

0.001

0.1

1

(b)

(c)

(d)(a) -Φ0 Φ00

wire C

wire C

wire C BⅡ = -19 mT

RS (Ω

)

0.01

0.001

0.1

1

10 205 15B (mT)

0.05 0.1 0.5 10.02 0.2T (K)

(e)

BⅡ (mT)

-16-19

-14-11.5-80

1

10010

RS (mΩ)-Φ0 Φ00

0

20

B (m

T)

40-40 BⅡ (mT)theory

T = 0.02 K

B (mT)

12.513

1211.51110.510

75.5

9

4.50

FIG. 5. (a) Differential shell resistance, RS, measured for wire C at B⊥ = 4 mT as a function of temperature, T , andparallel magnetic field, B‖. For small B⊥ the sample displays a non-destructive T–B‖ phase diagram. (b) Same as (a), butmeasured at B⊥ = 11 mT. Around ±Φ0/2 an anomalous-resistance phase develops at low T . (c) Same as (a), but measuredat B⊥ = 13 mT. The shell resistance increases to the normal state value as the applied flux passes ±Φ0/2, even at the basetemperature. Note that RS is finite for all temperatures above the mean-field theory predicted TC. The theory curves in (a)-(c)are Eq. 1 numerically solved for TC(α‖ + α⊥). (d) Half-flux quantum RS as a function of T measured at different B⊥ values.Close to B⊥ = 0, as the temperature is lowered, the sample displays a conventional normal-superconducting phase transition.Around B⊥ = 5 mT the shell resistance at low T saturates to a finite, B⊥-dependent value. Above B⊥ = 13 mT the shellresistance does not decrease below the normal state resistance. (e) Base-temperature RS as a function of B⊥ measured atdifferent B‖ or Φ values. The resistance increases with B⊥ in a step-like manner with the step feature mostly pronounced ataround −Φ0/2 of the applied flux. Inset: RS as a function of B⊥ and B‖. The theory curve was computed using Eq. 1, wherea critical B⊥ was found for each B‖, above which TC vanishes.

tional pair-breaking term [3],

α⊥ =4 ξS

2TC0

AF

Φ⊥2

Φ02, (4)

where Φ⊥ = B⊥AF. Figure 4 shows the theoretical tran-sitions based on Eqs. 1–4 using α = α‖ + α⊥ [35] super-imposed on experimental data.

Near the conventional-destructive crossover [Fig. 4(b)],a resistive state with RS smaller than RN was observedaround ±Φ0/2 and IS = 0. Figure 5 examines this re-sistive state close to the crossover, around B⊥ ∼ 12 mT,along with superimposed theory curves based on Eqs. 1–4. Note that unlike the situation far from the crossover[Fig. 5(a)], where theory and experiment agree well, inthe vicinity of the crossover [Fig. 5(b,c)] mean-field the-ory predicted TC deviates from the temperatures wherethe shell displays RN.

Temperature dependence of RS around −Φ0/2 forseveral values B⊥ near the conventional-destructive

crossover are shown in Fig. 5(d). Throughout this regime,RS was found to saturate to a temperature independentvalue, which can be tuned over two orders of magnitudewith small changes in B⊥. In contrast, a RS-T tracetaken close to the second destructive regime, not neara crossover (B⊥ = 12 mT and B‖ = 52 mT) remainstemperature dependence down to the base temperature[Fig. S2 in Supplemental Material [28]]. Qualitativelysimilar anomalous RS saturation was also observed fordifferent B‖ values at a fixed B⊥, see Fig. S3 in Supple-mental Material [28]. At base temperature the evolutionof RS as a function of B⊥ shows a step-like increase, thatis mostly pronounced around ±Φ0/2, see Fig. 5(e).

A possible explanation for the saturation of RS interms of disorder-induced variations of ∆, separating theshell into normal and superconducting segments [23] wastested by examining saturation effects in three segmentsof the same wire [Fig. S4 in Supplemental Material [28]].It was found that all segments behaved the same, argu-

Page 5: Anomalous Metallic Phase in Tunable Destructive

5

ing against long-range variation in ∆ on the scale of theseparation of contacts. We also note that the anomalousresistance develops predominantly above the theoreticalTC, where the sample is expected to be in the normalstate [Fig. S5 in Supplemental Material [28]].

The step-like increase of RS with B⊥ shown in Fig. 5(e)is reminiscent of phase slips, similar to the ones reportedin Refs. [10, 35], except here they are activated by per-pendicular field rather than temperature. This suggests apicture in which anomalous saturating resistance resultsfrom quantum fluctuations not captured by mean-fieldtheory. In general, the probability of a transverse phaseslip across a weak link is proportional to exp (−RQ/RN),with the resistance quantum RQ, and therefore is ex-ponentially small for wire C [36]. However, near onehalf flux quantum, states with consecutive phase wind-ings around the shell are degenerate, allowing quantumfluctuations to play a role. We note that both deep in thenondestructive regime [Fig. 2(a)] and deep in the destruc-tive regime [Fig. 2(b)], no anomalous phase is observed.

Previous theoretical work [16, 23] argued that the ratioof dFtS/2 to λ2 controls the order of the superconductor-metal transition. The present experiments span therange, with wires A and B having dFtS/2 < λ2, whereaswire C has dFtS/2 & λ2. We have not observed system-atic qualitative difference across this ratio. A detailedinvestigation of the order of the transition, and its affecton the anomalous metallic phase, would make an inter-esting future study.

In summary, we have investigated destructive andnondestructive Little-Parks effect in InAs nanowiresfully covered with epitaxial Al. Excellent agreementwith Ginzburg-Landau mean-field theory was obtainedacross multiple reentrant quantum phase transitions us-ing independently measured device and material pa-rameters. Millitesla-scale perpendicular magnetic fieldwas used to tune the crossover between destructive andnon-destructive regimes, yielding an anomalous metallicphase around one-half flux quantum with a temperatureindependent resistance ranging over two orders of magni-tude controlled by small changes in perpendicular field.This field-controllable anomalous phase is not explainedby existing theory, but presumably involves quantumfluctuations between winding numbers of superconduct-ing phase around the cylindrical superconducting shell.

We thank Mingtang Deng, Claus Sørensen, andShiv Upadhyay for materials and experimental contribu-tions, and Mikhail Feigelman, Karsten Flensberg, SteveKivelson, Yuval Oreg, Gil Refael, and Boris Spivak forvaluable discussions. Research was supported by Mi-crosoft, the Danish National Research Foundation, andthe European Commission.

[1] S. L. Sondhi, S. M. Girvin, J. P. Carini, and D. Shahar,Rev. Mod. Phys. 69, 315 (1997).

[2] T. Vojta, Ann. Phys. 9, 403 (2000).[3] N. Shah and A. Lopatin, Microscopic, Phys. Rev. B 76,

094511 (2007).[4] Q. Si and F. Steglich, Science 329, 1161 (2010).[5] M. R. Norman, Science 332, 196 (2011).[6] M. Tinkham, Introduction to Superconductivity (McGraw

Hill, New York, ed. 2, 1996).[7] A. Del Maestro, B. Rosenow and S. Sachdev, Ann. Phys.

324, 523 (2009).[8] A. M. Goldman, Int. J. Mod. Phys. B 24, 4081 (2010).[9] A. Kapitulnik, S. A. Kivelson, and B. Spivak,

Rev. Mod. Phys. 91, 011002 (2019).[10] A. Bezryadin, C. N. Lau, and M. Tinkham, Nature 404,

971 (2000).[11] A. D. Zaikin, D. S. Golubev, A. van Otterlo, and

G. T. Zimanyi, Phys. Rev. Lett. 78, 1552 (1997).[12] O. V. Astafiev, L. B. Ioffe, S. Kafanov, Yu. A. Pashkin,

K. Yu. Arutyunov, D. Shahar, O. Cohen, and J. S. Tsai,Nature 484, 355 (2012).

[13] J. E. Mooij and C. J. P. M. Harmans, N. J. Phys. 7, 219(2005).

[14] B. S. Deaver Jr. and W. M. Fairbanks, Phys. Rev. Lett.7, 43 (1961).

[15] H. Doll and M. Nabauer, Phys. Rev. Lett. 7, 51 (1961).[16] D. H. Douglass Jr. Phys. Rev. 132, 513 (1963).[17] W. A. Little and R. D. Parks, Phys. Rev. Lett. 9, 9

(1962).[18] P.-G. de Gennes, C. R. Acad. Sci. 292, 279 (1981).[19] R. M. Arutyunyan and G. F. Zharkov, Zh. Eksp. Teor.

Fiz. 79, 245 (1980).[20] G. Schwiete and Y. Oreg, Phys. Rev. Lett. 103, 037001

(2009).[21] Y. Liu, Yu. Zadorozhny, M. M. Rosario, B. Y. Rock,

P. T. Carrigan, and H. Wang, Sience 294, 2332 (2001).[22] I. Sternfeld, E. Levy, M. Eshkol, A. Tsukernik, M. Kar-

povski, H. Shtrikman, A. Kretinin, and A. Palevski,Phys. Rev. Lett. 107, 037001 (2011).

[23] V. H. Dao and L. F. Chibotaru, Phys. Rev. B 79, 134524(2009).

[24] O. Vafek, M. R. Beasley, and S. A. Kivelson,https://arxiv.org/abs/cond-mat/0505688 (2005).

[25] A. V. Lopatin, N. Shah, and V. M. Vinokur,Phys. Rev. Lett. 94, 037003 (2005).

[26] P. Krogstrup, N. L. B. Ziino, W. Chang, S. M. Albrecht,M. H. Madsen, E. Johnson, J. Nygard, C. M. Marcus, andT. S. Jespersen, Nat. Mater. 14, 400 (2015).

[27] S. Vaitiekenas, M.-T. Deng, P. Krogstrup, andC. M. Marcus, https://arxiv.org/abs/1809.05513 (2018).

[28] See Supplemental Material for more detailed device de-scription and additional measurements.

[29] C. Kittel, Introduction to Solid State Physics (Wiley, ed.8, 2005).

[30] E. G. Mikkelsen, P. Kotetes, P. Krogstrup, and K. Flens-berg, Phys. Rev. X 8, 031040 (2018).

[31] A. E. Antipov, A. Bargerbos, G. W. Winkler, B. Bauer,E. Rossi, and R. M. Lutchyn, Phys. Rev. X 8, 031041(2018).

[32] A. A. Abrikosov and L. P. Gorkov, Sov. Phys. JETP 12,1243 (1961).

Page 6: Anomalous Metallic Phase in Tunable Destructive

6

[33] N. A. Court, A. J. Ferguson and R. G. Clark, Super-cond. Sci. Technol. 21, 015013 (2008).

[34] J. Bardeen, Rev. Mod. Phys. 34, 667 (1962).[35] A. Rogachev, A. T. Bollinger, and A. Bezryadin,

Phys. Rev. Lett. 94, 017004 (2005).[36] M. Vanevic and Y. V. Nazarov, Phys. Rev. Lett. 108,

187002 (2012).

Page 7: Anomalous Metallic Phase in Tunable Destructive

7

Supplemental Information

METHODS

NANOWIRE GROWTH

The wires studied in this work were grown using molec-ular beam epitaxy on InAs(111)B substrate at 420 C, viastandard Au-catalized vapor-liquid-solid method. First,InAs wires were grown along the [0001] direction withwurtzite crystal structure. Subsequently to the semicon-ductor growth, a full Al shell was grown at −30 C on allsix facets by rotating the growth substrate with respectto the metal source, resulting in an epitaxial interface be-tween the Al and InAs [1]. The core diameter was tunedby changing the Au seed particle size. The shell thicknesswas controlled by the Al growth time.

DEVICE FABRICATION

For the device fabrication individual wires were trans-ferred onto a degenerately n-doped Si substrate cappedwith a 200 nm thermal oxide using a micro-manipulatorstation. Standard electron beam lithography techniqueswere used to pattern etching windows and contacts. Athin layer of AR 300-80 (new) adhesion promoter anddouble layer of EL6 copolymer resists was used to de-fine the etching windows. The Al was then selectivelywet-etch for ∼ 60 s in MF-321 photoresist developer. Tocontact the Al shell, a stack of A4 and A6 PMMA resistwas used. The Al oxide from under the contacts was re-moved by Ar-ion milling (RF ion source, 25 W, 18 mTorr,9 min) followed by normal Ti/Al (5/210 nm for wires Aand B, and 5/350 nm for wire C) ohmic contact metal-lization. To contact the InAs core a single layer of A6PMMA resist was used. A gentler Ar-ion milling (RF ionsource, 15 W, 18 mTorr, 6.5 min) was used to remove thenative oxide layer off the InAs core, followed by deposi-tion of the normal Ti/Al (5/180 nm for wires A and B,and 5/350 nm for wire C) ohmic contacts.

Wire dF (nm) dM (nm) dC (nm) tS (nm) L (nm)

A 157±5 146±4 137±5 7±3 945±5

B 195±5 163±4 135±5 24±3 945±5

C 340±5 288±4 240±5 41±3 920±5

Table. S 1. Wire dimensions measured from micrographs.Full-wire diameter dF, mean diameter dM, core diameter dC,shell thickness tS, and distance between the voltage probes L.

WIRE PARAMETERS

The main wire parameters, including the ones use tocompute the theory curves in the main text, are sum-marized in Tables S1, S2 and S3. The full-wire diam-eter, dF, and the core diameter, dC, [Fig. 1(a) in themain-text] as well as the distance between the voltageprobes, L, for each wire were measured from individ-ual micrographs. For all the wires the Al oxide wasassumed to be tOx = 2 nm. Using simple trigono-metrical considerations one can deduce the full cross-sectional area AF = 3

√3(dF − 2tOx)2/8, the shell thick-

ness tS =√

3 (dF − dC) /4 − tOx and the mean wire di-ameter dM = (dF − 2tOx + dC)/2. The normal stateresistance RN and the zero-field transition temperatureTC0 were measured while cooling down the sample. Zero-field switching IS0 and re-trapping IR0 currents weremeasured at the base temperature. The period of theLittle-Park oscillations in magnetic field can be calcu-lated using ∆B = Φ0/AM = 8 Φ0/3

√3 d2M. The shell

resistivity is given by ρ = RN(AF − AC)/L, whereAC = 3

√3d2C/2 is the core cross-sectional area. The

Drude mean free path for electrons in the shell is de-termined using le = mevF/e

2nρ, with electron massme, electron Fermi velocity in Al vF = 2.03 × 106 m/s[2], electron charge e and charge carrier density n =k3F/3π

2, where kF is the Fermi wave vector. The dirty-limit superconducting coherence length is given by [3]ξS =

√π~vFle/24kBTC0, where ~ is the reduced Planck

constant and kB is the Boltzmann constant. For a dirtysuperconductor, the Ginzburg-Landau penetration depthis [3] λ(T ) = λL(T )

√ξ0/1.33le, with the London penetra-

tion depth λL(T ) = λL(0)/√

2(1− T/TC0), and the co-

Wire RN (Ω) TC0 (K) ∆0 (µeV) IS0 (µA) IR0 (µA)

A 34.3±0.1 1.45±0.1 220±7 24±1 14±1

B 1.6±0.1 1.22±0.1 183±3 62±2 46±2

C 0.35±0.01 1.17±0.1 177±3 61±2 60±2

Table. S 2. Wire characteristics extracted from transportmeasurements. Normal-state resistance RN, zero-field criticaltemperature TC0, superconducting gap ∆0, zero-field switch-ing current IS0 and re-trapping current IR0.

Wire ∆B (mT) ρ (Ω nm) le (nm) ξS (nm) λ (nm)

A 150±7 110±40 4±1 71±8 200±60

B 120±5 20±3 20 ±3 180±10 100±20

C 38.4±0.9 14±1 29±2 224±8 89±7

Table. S 3. Calculated wire quantities. Flux period in mag-netic field ∆B = Φ0/AM, resistivity ρ, mean free path le,zero-field superconducting coherence length ξS and Ginzburg-Landau penetration depth λ.

Page 8: Anomalous Metallic Phase in Tunable Destructive

8

(a) -Δ0 Δ0

0-0.4 -0.2 0.2 0.4

dI/d

V (e

2 /h)

0.1

0.2

0.3

0

V (mV)(b)

0-0.4 -0.2 0.2 0.4

dI/d

V (e

2 /h)

0.2

0.4

0

V (mV)

-Δ0 Δ0

-Δ0 Δ0

wire CΔ0 = 0.18 meV

wire BΔ0 = 0.18 meV

wire AΔ0 = 0.22 meV

(c)

0-0.4 -0.2 0.2 0.4

dI/d

V (e

2 /h)

0.2

0.1

0.3

0

V (mV)

FIG. S 1. Differential conductance, dI/dV , as a function ofsource-drain voltage bias, V , measured for (a) wire A, (b)wire B, and (c) wire C.

herence length is ξS(T ) = 0.855√ξ0le/(1− T/TC0). This

gives λ = λLξS/1.39le, with the zero-temperature Lon-don penetration depth λL = 16 nm [2].

DENSITY OF STATES

Each of the measured wire is equipped with a tunnel-ing probe at its end, see the main-text Fig. 1(b). Apply-ing voltage to the back-gate, VBG, creates a tunnel bar-rier in the bare-semiconducting (InAs) segment, seper-ating the normal-metal (Ti/Au) contact and the wire(Al/InAs). In the tunneling regime, the change in thecurrent through the junction with the applied voltagebias corresponds to the local density of states. Differ-ential tunnelling conductance, dI/dV , measured for allthree wires as a function of source-drain voltage, V , isshown in Fig. S1. For wire A, with the thinnest shell, themeasured superconducting gap is ∆0 = 220 µeV, whereasboth wires B and C show a gap of around ∆0 = 180 µeV.All three gaps agree (within the experimental error) withthe BCS theory predicted value ∆0 = 1.764kBTC0. WiresA and B display additional peaks in density of states atthe energies below the main superconducting gap. Weidentify these with the proximity induced gaps inside thesemiconducting cores.

0

0.4

0.8

T (K

)

40 6020BⅡ (mT)

(a) Φ0/2 3Φ0/2Φ0

B = 12 mTwire C

B = 12 mTwire C

theory

0

0.3

RS (Ω

)

RS (Ω

)

0.40.3

0.2

0.1

(b)

0.10.050.02 0.2T (K)

BⅡ (mT)

5221

FIG. S 2. (a) Differential shell resistance, RS, as a function ofparallel magnetic field, B‖, and temperature, T , measured forwire C at perpendicular magnetic field B⊥ = 12 mT. The the-ory curve is the main-text Eq. 3 computed with α = α‖ +α⊥.(b) RS-T traces measured at B⊥ = 12 mT. At B‖ = 21 mT,close to Φ0/2 applied flux, RS saturates at low temperatures.At B⊥ = 51 mT, before the wire enter the destructive regimearound 3Φ0/2, RS shows temperature dependence even at thebase temperature.

RS (Ω

)

0.01

0.001

0.1

1

0.05 0.1 0.5 10.02 0.2

BⅡ (mT)

-18-19

-17-16-15-14-11

0

wire CB = 12 mT

T (K)

FIG. S 3. Differential shell resistance, RS as a function of Tmeasured at fixed B⊥ = 12 mT for wire C at different B‖values. Around B‖ = 0, as T is lowered, the sample displaysa conventional normal-superconducting phase transition. Asthe field is tuned to B‖ = −14 mT the shell resistance startsto saturate at low T to a finite, B‖-dependent value.

NON-SATURATING RESISTANCE

The observed low-temperature saturation of the half-flux quantum RS might rise a question whether it isnot an artifact of a deficient cooling. In other words,if the electron temperature upon cooling would saturateat some elivated temperature, so would the shell resis-tance. To rule out such an explanation we record twoRS-T traces for wire C at B⊥ = 12 mT, see Fig. S2:

Page 9: Anomalous Metallic Phase in Tunable Destructive

9

wire C

IS

V12 V34

1 μm

(a)

Ti/AuAl

BBⅡ

0.01

0.001

0.1

1

0.05 0.1 0.5 10.02 0.2

BⅡ = -19 mT(b)

B (mT)

12.513

1211.51110.510

85

9

40

0.01

0.001

0.1

1

0.05 0.1 0.5 10.02 0.2

BⅡ = -19 mT(c)

B (mT)

12.513

1211.51110.510

85

9

40

T (K)

T (K)

R12

-R12

(Ω)

0R

34-R

34 (Ω

)0

FIG. S 4. (a) Micrograph of wire C with the highlightedthree-terminal setups for the outer segments shell resistancemeasurements. (b) Differential shell resistance in the left-most wire C segment, R12 = dV12/dIS (with the subtractedcontact resistance R0

12) measured as a function of tempera-ture T at B‖ = −19 mT and different B⊥ values. (c) Similarto (b), but measured for the right-most segment. The con-tact resistances R0

12 and R034 were measured around the base

temperature at B⊥ = 0 and B‖ = −19 mT.

One at B‖ = 21 mT, close to Φ0/2 applied flux quan-tum, displaying the anomalous RS saturation; Anotherat B‖ = 52 mT, before the destructive regime around3Φ0/2, with a T -dependent RS down to the base temper-ature. Furthermore, the data in the main-text Fig. 5(d)shows that the RS starts to saturate at different temper-atures for different B⊥. Finally, it is unlikely for a poorelectron cooling to cause the observed broadening of theanomalous phase in flux, see the main-text Fig. 5(b) and(c), as well as Fig. S2(a).

0

0.4

0.8

1.2

0

0.4

0.8

0

0.4

0.8

(b)

(c)

(a)B = 4 mT

theory

T (K

)

100

10

1

RS (m

Ω)

theory

T (K

)

B = 11 mT

B = 13 mT

T (K

)

0 40-40BⅡ (mT)

theory

-Φ0 Φ00

wire C

wire C

wire C

FIG. S 5. Same data as in the main-text Fig. 5(a), (b) and (c),but in logarithmic color scale highlighting the low-resistancefeatures. At low T , the anomalous phase is present predomi-nantly above the TC predicted by the mean-field theory. Thefinite resistance at higher temperatures, below the arcs of thetheory curves, arise presumably due to thermal fluctuations.

FLUX-DEPENDENT RESISTANCESATURATION

The data shown in the main-text Fig. 5 demonstratethat at a fixed B‖ = −19 mT (around −Φ0/2 of theapplied flux) RS at low T saturates to a B⊥-dependentvalue. We observe a qualitatively similar B‖-dependentanomalous saturation of RS at a fixed B⊥ = 12 mT, seeFig. S3.

OUTER SEGMENTS

To demonstrate that the anomalous resistance satura-tion shown in the main-text Fig. 5(d) is not due to a localdisorder in the middle-wire segment, we investigate theouter two wire segments using three-terminal setup, seeFig. S4(a). Differential shell resistances R12 = dV12/dISand R34 = dV34/dIS with the subtracted correspond-ing contact resistances measured as a function of T atB‖ = −19 mT and different B⊥ values are shown inFig. S4(b) and (c). The contacts resistances R0

12 and R034

were measured around the base temperature at B⊥ = 0and B‖ = −19 mT. The observed B⊥-dependent, low-temperature anomalous shell resistances are qualitativelysimilar to the RS of the middle segment. The small quan-

Page 10: Anomalous Metallic Phase in Tunable Destructive

10

titative discrepancies between the segments might arisedue to the uncertainty in the applied B⊥ or a small wiretapering.

ANOMALOUS PHASE VS. MEAN-FIELDTHEORY

Figure S5 shows the same data as in the main-textFig. 5(a)-(c), but plotted in a logarithmic color scale tohighlight the low-resistance features. It appears that theanomalous resistance phase at low T develops predom-inantly above the mean-field theory predicted TC, closeto the ±Φ0/2 of the applied flux. At elevated T , the wire

shows finite, but reduced RS around 0 and ±Φ0 of theapplied flux, presumably arising due to thermal fluctua-tions.

[1] P. Krogstrup, N. L. B. Ziino, W. Chang, S. M. Albrecht,M. H. Madsen, E. Johnson, J. Nygard, C. M. Marcus, andT. S. Jespersen, Nat. Mater. 14, 400 (2015).

[2] C. Kittel, Introduction to Solid State Physics (Wiley, ed.8, 2005).

[3] M. Tinkham, Introduction to Superconductivity (McGrawHill, New York, ed. 2, 1996).