Shao-Pin
Chiu‡
a,
Vivek
Mishra‡
b,
Yu
Li
b,
Fu-Chun
Zhang
c,
Stefan
Kirchner
*d and
Juhn-Jong
Lin
*e
aDepartment of Electrophysics & Center for Emergent Functional Matter Science, National Yang Ming Chiao Tung University, Hsinchu 30010, Taiwan
bKavli Institute for Theoretical Sciences, University of Chinese Academy of Sciences, Beijing 100190, China
cKavli Institute for Theoretical Sciences & CAS Center for Excellence in Topological Quantum Computation & HKU-UCAS Joint Institute of Theoretical and Computational Physics at Beijing, University of Chinese Academy of Sciences, Beijing 100190, China
dDepartment of Electrophysics & Center for Emergent Functional Matter Science & Center for Theoretical and Computational Physics, National Yang Ming Chiao Tung University, Hsinchu 30010, Taiwan. E-mail: stefan.kirchner@correlated-matter.com
eDepartment of Electrophysics & Center for Emergent Functional Matter Science, National Yang Ming Chiao Tung University, Hsinchu 30010, Taiwan. E-mail: jjlin@nycu.edu.tw
First published on 5th May 2023
We report the observation of enhanced interfacial two-component superconductivity possessing a dominant triplet component in nonmagnetic CoSi2/TiSi2 superconductor/normal-metal planar heterojunctions. This is accomplished through the detection of odd-frequency spin-triplet even-parity Cooper pairs in the diffusive normal-metal component of T-shaped proximity junctions. We show that by modifying the diffusivity of the normal-metal part, the transition temperature enhancement can be tuned by a factor of up to 2.3 while the upper critical field increases by up to a factor of 20. Our data suggest that the C49 phase of TiSi2, which is stabilized in confined geometries, underlies this enhancement. These findings are addressed via a Ginzburg–Landau model and the quasi-classical theory. We also relate our findings to the enigmatic 3-K phase reported in Sr2 RuO4.
A three-terminal T-shaped proximity structure comprised of a superconductor (S) in contact with a diffusive normal metal (N), schematically shown in Fig. 1(a), had been proposed to phase-sensitively distinguish triplet from singlet superconductors.10 This T-shaped proximity structure detects triplet pairing through a zero-bias conductance peak (ZBCP) that results from the generation of odd-frequency spin-triplet even-parity pairs in the normal diffusive side of the S/N interface as illustrated in Fig. 1(b).11–14 This is referred to as the anomalous proximity effect (APE) to distinguish it from the case of a singlet superconductor where a conductance dip ensues.
![]() | ||
Fig. 1 (a) Illustration of a T-shaped proximity structure proposed in ref. 10 to distinguish a triplet from a singlet superconductor (S) and where S forms the arm of the letter ‘T’. N is the diffusive normal metal component of the structure. This sketch also illustrates the S/N interface (in red) where the interfacial superconductivity with its enhanced onset temperature forms. RB denotes the barrier resistance of this interface. RN is the resistance of the N component. (b) The predicted zero-bias conductance peak (dip) for a spin-triplet (spin-singlet) S. (c) SEM image of device B1 together with the schematic 4-probe measurement configuration. The magnetic field direction is indicated. (d) Schematics of our CoSi2/TiSi2 T-shaped proximity devices and their various voltage-electrode (VE) pairs. 2L1 is the voltage-electrode (VE) separation. L2 is the length of the normal metal (N) segment connecting to superconductor (S). |
Yet, the fabrication of such T-shaped structures has proven difficult for many materials including superconductors.15,16 The successful fabrication of high-quality CoSi2/TiSi2 T-shaped proximity structures on silicon was only recently achieved.17,18 The availability of these structures established the existence of triplet pairing in CoSi2 on Si(100) via the observation of the APE.17 CoSi2 is a superconducting material that is widely used in the semiconductor industry with a superconducting transition temperature (Tbulkc) of 1.3 K. Bulk specific heat measurements as well as theoretical estimates suggest that CoSi2 is a phonon-mediated spin-singlet superconductor.19–21
Generalizing the proposal of Asano et al.10 to include the effect of spin–orbit coupling (SOC), the findings reported in ref. 17 have been interpreted in terms of two-component superconductivity with a dominant triplet component.22 This interpretation is in line with the finding that the accompanying interface between CoSi2 and the Si(100) substrate23 gives rise to a SOC which exceeds the bulk CoSi2 superconducting energy gap Δ0 by a factor of ∼30, see ref. 17. We note in passing that yet another route to triplet pairing is based on proximity structures involving ferromagnets.24 These systems too offer a way of realizing odd-frequency pairing states near their superconductor/ferromagnetic metal interfaces.25,26
In this manuscript, we demonstrate that an enhanced two-component superconductivity emerges in the vicinity of the S/N interface formed by CoSi2 (S) and TiSi2 (N) which is reminiscent of the enigmatic ‘3-K phase’ reported in Sr2RuO4.27 We establish that the transition temperature and upper critical field can be tuned by the diffusive properties of the normal metal component. This is accomplished via the APE that has been reported in these structures.17 An SEM image of a typical device is shown in Fig. 1(c) while a sketch of it is provided in Fig. 1(d). Clearly discernible are the various voltage-electrode (VE) pairs which allow us to relate the ZBCP to the proximity of superconducting CoSi2/Si. The device parameters of all devices used in the present study are compiled in Table 1. Fig. 1(a) also highlights the CoSi2/Ti2Si2 interface (in red) where the interfacial superconductivity forms as we discuss below.
Device | T a (°C) | L 1 (μm) | L 2 (μm) | R N (Ω) | ρ N (μΩ cm) | ΔGn (at 0.37 K) | FWHM (meV) | E Th (meV) | D (cm2 s−1) | T c onset (K) | G J,□ (4 K) (Ω−1 μm−2) |
---|---|---|---|---|---|---|---|---|---|---|---|
a T onsetc was extracted from Gn(0, T, 0) of the S/N junction geometry to minimize any possible experimental uncertainty. b Device B4 became unstable. Conductance spectra could not be completed. | |||||||||||
B1 | 750 | 0.42 | 0.39 | 57.5 | 197 | 112% | 0.18 | 0.0012 | 3.1 | 2.94 | 1.46 |
B1m | 750 | 0.42 | 0.39 | 55.8 | 191 | 31% | 0.22 | 0.0012 | 3.2 | 2.84 | 1.47 |
B2 | 750 | 0.20 | 0.66 | 12.0 | 79.8 | 9.3% | 0.04 | 0.013 | 7.7 | 2.33 | 0.34 |
B2m | 750 | 0.20 | 0.66 | 12.7 | 84.3 | 1.0% | 0.056 | 0.012 | 7.3 | 2.03a | 0.32 |
B3 | 780 | 0.23 | 0.10 | 0.346 | 2.14 | 5.9% | 0.026 | 0.13 | 106 | 2.14 | 1.03 |
B4b | 750 | 0.34 | 0.18 | 73.3 | 195 | 1.8% | 0.24 | 0.0019 | 3.2 | 1.70 | — |
B5 | 800 | 0.45 | 0.32 | 1.31 | 3.34 | 0.13% | 0.03 | 0.023 | 67.9 | 1.50 | 5.18 |
B5m3 | 800 | 0.45 | 0.32 | 19.1 | 48.6 | 66% | 0.20 | 0.004 | 12.7 | 2.10 | 0.93 |
![]() | (1) |
We here report our observation of interfacial superconductivity in CoSi2/TiSi2 heterostructures with an onset temperature (Tonsetc) that exceeds Tbulkc by up to a factor 2.3 and that also possesses a dominant triplet component. We present a Ginzburg–Landau (GL) analysis in terms of this two-component superconductivity order parameter that encompasses the Tonsetc enhancement and explores its particular features. Implications and possible applications are also mentioned.
This analysis also sheds light on the different roles played by the CoSi2/Si(100) and the CoSi2/TiSi2 interfaces forming CoSi2/TiSi2 heterostructures. The thickness of the CoSi2/Si films in this work is ≃105 nm while that of the TiSi2 part is ≃125 nm thick and ≃(0.2–0.4) μm wide. TiSi2 is a diffusive normal metal and remains metallic down to at least a temperature (T) of 50 mK.17 The primary quantity of interest of the CoSi2/TiSi2 T-shaped proximity junction is the normalized conductance Gn(V, T, B) = G(V, T, B)/G0 where G denotes the differential conductance, G(V, T, B) = dI(V, T, B)/dV with I the current, V the bias voltage, B the magnetic field, and G0 is the residual conductance of the device in the absence of superconductivity. Practically, we take G0 = G(0, 4 K, 0).
Fig. 2(a) illustrates the basic feature of the APE in the representative device B5m3. It depicts Gn(V, 0.365 K, 0) measured by using different VE pairs which define different segments of the TiSi2 component [cf.Fig. 1(c) and (d)]. In all cases, I was applied through the outermost electrode pair (1,6). Fig. 2(a) reveals large amplitudes of the ZBCPs measured with VE pairs (3,4), (2,5) and (1,6). The inset shows a zoom-in of the ZBCPs measured with the VE pairs (1,2) and (2,3), which are located away from the S/N interface. We see that in this case the amplitudes of the ZBCPs are small. In fact, the Gn curve measured with the VE pair (1,2) is flat. It then follows that the ZBCP must arise from odd-frequency spin-triplet Cooper pairs through the APE10,17,22 and cannot be of any other origin (cf. additional discussion in ESI†). Owed to the properties of the T-shaped proximity structure, the ZBCP thus serves as a diagnostics for the presence of triplet pairing in CoSi2/Si and/or CoSi2/TiSi2 interface.
![]() | ||
Fig. 2 (a) Normalized conductance spectra Gn(V, 0.365 K, 0) of device B5m3 measured with various voltage-electrode (VE) pairs indicated by the electrode numbers in Fig. 1(d). I was always applied using the electrode pair (1,6). Inset: a zoom-in of Gn(V, 0.365 K, 0) measured with VE pairs (1,2) and (2,3). (b) Gn(V, T, 0) of device B1 recorded at several T values. Inset: a zoom-in of the 2.5 K curve. (c) Gn(V, 0.37 K, B) of device B1 in several B values. Inset: a zoom-in of the 2.0 T curve. |
Fig. 2(b) shows Gn(V, T, 0) vs. V for device B1 at several T values. ZBCPs are clearly discernible, which are gradually suppressed with increasing T, thus establishing the presence of dominant triplet pairing in the junction. The inset reveals that the ZBCP persists up to at least 2.5 K. Fig. 2(c) displays Gn(V, 0.37 K, B) of the same device in several B fields and at T = 0.37 K. While the ZBCP is gradually suppressed with increasing B, it persists up to at least 2 T§ (inset).
We find that in a number of devices the highest T at which the ZBCP is discernable is significantly enhanced over Tbulkc, i.e., Tonsetc > Tbulkc in these devices. The enhanced Tonsetc value is not only enhanced over those in the T-shaped structures studied in ref. 17 (Tonsetc ≲ 1.4 K), but also with respect to the Tc value (≈1.5 K) of epitaxial CoSi2/Si films23 while the amplitude of the ZBCP is strongly enhanced up to ≈210% with respect to G0. This is illustrated in Fig. 3(a). Devices B1, B2, B3 and B5 were measured as grown (cf. ESI†).
![]() | ||
Fig. 3 (a) Gn(0, T, 0) in six devices vs. temperature T. (b) Gn(0, T, B) of device B1 in several B fields. |
The upper critical field is significantly enhanced as well. Fig. 3(b) depicts Gn(0, T, B) of device B1 for several B fields. Gn(0, T, B) is gradually suppressed with increasing B. Yet, a small proximity effect is still visible in B = 2.0 T which is much higher than the in-plane upper critical field (≤0.12 T) of CoSi2/Si films.17 Additional conductance spectra are provided in the ESI.†
In order to address the origin of this phenomenon the junctions underwent thermal cyclings up to room temperature followed by the cooling down for further measurements, see the ESI† for additional information. (Following thermal cycling, the devices were renamed such that device Bi (i = 1, … 5) became Bim after the first thermal cycling and Bimj (j > 1) after the jth subsequent thermal cycling.) The Gn (0, T, 0) of device B1 (B1m) increases with decreasing T below Tonsetc = 2.94 K (2.84 K), reaching 212% (131%) at 0.365 K. There are visible conductance fluctuations which are probably induced by dynamical structural defects in this particular device,30 see Fig. 3.
The quasi-classical theory discussed below associates the height of the ZBCP with the ratio of the normal-metal and interface conductances which leads us to study the relation between Tonsetc and the diffusivity of the device. Fig. 4 brings out this relation and reveals that the measured Tonsetc value is strongly correlated with the high residual resistivity ρN, i.e., low electron diffusion constant (D) of the N part. By and large, a higher Tonsetc is found in devices with higher ρN or in other words systems with lower D. The reason for the variation in ρN of TiSi2 is connected to its occurrence in mainly two phases, a base-centered phase (C49) and a face-centered phase (C54).31,32 C49 is metastable and contains large amounts of stacking faults33 which lead to an order of magnitude larger resistivity [ρ(300 K) ≃ (100–200) μΩ cm] than the C54 phase [ρ(300 K) ≃ (17–25) μΩ cm].
![]() | ||
Fig. 4 Onset temperature Tonsetcvs. inverse residual resistivity 1/ρN of TiSi2, i.e., the normal-metal component. Devices of the present study are marked by red bullets (![]() |
Although C49 is metastable in the thermodynamic limit it can be stabilized in spatial confinement and systems at the microscale due to a lack of nucleation sites for C54. The annealing temperature, Ta, at which the TiSi2 in the junction forms, see Table 1, affects its composition with respect to C54/C49 such that Ta ≳ 800° C favors formation of C54.33,34 The C49 structure possesses a slightly larger volume than the C54 structure so that phase changes from C49 to C54 and vice versa are accompanied by strain.35 This gives a qualitative understanding of the effects of Ta and thermal cycling on Tonsetc. Devices primarily composed of C54 such as B5 are transformed into a C49 rich composition and a concomitantly large Tonsetc as in device B1. The changes induced by thermal cycling in devices primarily composed of C49 are less pronounced (see ESI† for details). This leads us to suspect that C49 grains are more stable near the CoSi2/TiSi2 interface and that its microscopic properties drive the formation of triplet pairing at the interface at the elevated Tonsetc compared with Tbulkc (CoSi2). The data point of Fig. 4 with the largest diffusivity seemingly defies the overall trend between Tonsetc and ρN. This outlier may represent an interface whose properties are not well characterized by the diffusivity of the wire or this data point indicates that the onset temperature goes through a minimum as a function of the diffusivity reminiscent of what has been observed in Sr2RuO4 under strain.36
We address the experimental findings in terms of a quasi-classical theory for three-terminal proximity structures as formed by the CoSi2/TiSi2 heterojunctions. The diffusive character of the N part of our junctions requires a treatment of the transport equations in the Usadel limit37 where the results can differ from those obtained in the weakly disordered limit.38 For the superconducting order parameter, we consider a SOC induced s + p pairing state with a dominant triplet component.22 Within the formulation of the circuit theory,39–42 an insulating barrier is expressed as a delta function [ZEFkF−1δ(x)] at the S/N interface. The weight of the barrier can be expressed in terms of a dimensionless parameter Z, Fermi energy EF, and the Fermi wavenumber kF. A higher ratio of RN/RB, i.e., resistance of the N component (RN) over barrier resistance (RB), results in an APE over a broader energy range and thus in an increase in the full-width-half-minimum (FWHM) of the ZBCP.10 The zero-bias value itself, however, is (roughly) independent of RN/RB if electron dephasing is ignored in the diffusive N component of the junction.
In Fig. 5(a), we keep RB, Δt, Δs and Z fixed, and vary L1 according to the experiments as shown in Fig. 2(a). Importantly, when VE configuration is changed, the S/N interface properties remain fixed but as L1 changes so will RN. We find that the ZBCP becomes broader, in qualitative agreement with Fig. 2(a). While this variation of FWHM with L1 agrees with the experiment, we find that the inclusion of a small amount of electron dephasing is required to reproduce the experimental behavior of Gn (V = 0), i.e., a suppression of Gn (V = 0) with increasing L1. In diffusive metals, existence of a finite electron dephasing rate in low-T regimes has been reported for long, but its microscopic origin(s) are yet to be fully identified.43 Magnetic impurities as a source of dephasing however seems unlikely as their presence in our heterojunctions has been ruled out.17 Thus, the quasi-classical approach supports the interpretation that for the C49 phase, RN is higher than in the C54 phase, giving rise to a larger ZBCP provided other parameters are kept constant. The higher ρN value of the devices indicating the presence of the C49 phase also results in an enhanced Tonsetc as shown in Fig. 4 and a strong proximity-induced odd-frequency pairing in the N part of the junctions. The robustness of ZBCPs in these devices supports the interpretation in terms of the SOC generated two-component superconductivity with a dominant triplet component, which is essential for the ZBCP.
Here we resort to a GL approach to understand this phenomena qualitatively. This Rashba SOC causes a band splitting that gives rise to two bands with opposite helicities and mixing of singlet and triplet pairing channels in the superconducting state. The free energy for such a system in a one-dimensional representation reads,46, here
is the usual GL free energy for each individual component, the subscript ν = ± represents the two components of the order parameter, and M[Φ] contains all higher order and gradient terms (technical details are relegated to the ESI†). The coefficient av is αv(T − Tcv). We take the superconductor to be in the clean limit.
is the coupling between the two components, a negative value of the coefficient c ensures that both order parameter have identical phases and
is the free energy density of the non-superconducting state in the absence of a magnetic field (H).
is the contribution from the magnetic field (
denotes the total magnetic field while
refers to the external field as is common in the GL literature. H has to be identified with B of Fig. 2 and 3).
is a Lifshitz invariant term that leads to s magneto-electric coupling and
determines the effect of superconducting order on spin-polarization (cf. ESI†).
We take the S/N interface at x = 0 and assume the system to be homogeneous along the other directions. We restrict ourselves to the experimental field configuration, i.e., along the interface in the plane (ŷ). The coefficients of the quadratic terms are taken to be spatially varying to model the enhanced superconductivity near the interface. The information about the interface is embedded in the spatial dependence of these quadratic coefficients.47,48a±(x) is α±[T − Tc±(x)], where and
. Here η±/c are the dimensionless parameters determining the amount of Tc enhancement, and d is the width of the effective interface. Minimizing the free energy leads to a set of differential equations (see ESI† for details). At the S/N interface we apply the De Gennes's boundary conditions,49,50
, where the extrapolation length
is a characteristic length scale associated with the induced superconducting correlations. Note, in contrast to conventional superconductors, here
cannot be identified with the dirty limit of the superconducting correlation length in the N segment, because it does not account for the physics of odd-frequency pairs, which are vital in the present case. We therefore treat
as a phenomenological parameter.
Fig. 5(b) shows the temperature and magnetic field phase diagram for appearance of onset order at the interface for various d values. The onset temperature and magnetic field are obtained by minimizing the free energy with De Gennes's boundary conditions with extrapolation , where ξ0 is the coherence length at T = 0 for the bulk superconductor. Within the GL formalism, Tonsetc exceeds Tbulkc in the low-field limit, and the Tonsetc decreases with increasing field. In our calculations, we find that the onset magnetic field in the low-T limit is comparable to
bulkc2 ≡ Φ0/(2πξ02), which is a magnetic field scale of the order of orbital upper critical fields. However, quantitatively the onset magnetic field is much smaller compared to the experiment, despite a reasonable Tonsetc obtained from theoretical calculations.
The experiment finds that in the low T limit, the onset magnetic field exceeds the upper critical field of the bulk superconductor by a factor ∼100, see ref. 19 and that of CoSi2/Si films by ∼20, see ref. 17. The experimentally observed critical magnetic field is higher than 2 T and this value is comparable to the Pauli-limited field for the bulk superconductor (∼2.4 T). The interface induced order survives up to (10–12) ξ0 from the surface (cf. ESI†). As shown in Fig. 5(b), Tonsetc for the appearance of such order above Tbulkc is very sensitive to the width of the interface. As the interface region becomes thinner, Tonsetc drops rapidly. In the devices with C49 phase, the interface region is expected to be relatively more disordered and its effective width is expected to be high compared to the low-ρN devices, due to smaller grain sizes or possible incomplete C49–C54 transformation (see ESI† for additional information).
Thus, the GL functional provides an effective model for the CoSi2/TiSi2 heterojunctions, which can guide future investigations. is the magneto-electric coupling term. The presence of this term suggests that the (CoSi2/Si)/TiSi2 system is an ideal system to explore, e.g., the superconducting diode effect and charge transport effects.51,52 Consequences of this term with respect to the magnetic field direction are currently explored. Our results also demonstrate the stability of triplet dominant pairing, and thus the prevalence of odd-frequency pairs in heterostructures.13,22,53,54 There is however a noticeable difference between its earlier realizations and the present case. The formation of odd-frequency pairs in (CoSi2/Si)/TiSi2 junctions does not require the proximity of magnetic order.
The enhanced superconductivity in CoSi2/TiSi2 is reminiscent of the elusive ‘3-K phase’ in the enigmatic transition metal compound Sr2RuO4.27,55–57 In fact, a similar model to our GL functional has originally been used to address the 3-K phase.47,48 There, it has been speculated that superconducting Ru islands are responsible for the 3-K phase. In contrast, here the highly diffusive C49 phase appears to drive the interface superconductivity while the C49 and C54 TiSi2 phases remain metallic down to lowest temperatures (see Fig. S2 of the ESI†). Moreover, here, the interface superconductivity possesses a dominant triplet component which is robust with regard to external magnetic fields.
Unconventional and functional properties of superconductivity near interfaces and in confined systems are of considerable interest. In part this is fueled by potential future applications. The silicon-based heterojunctions with dominant triplet pairing reported here are amenable to existing microfabrication techniques and are promising building blocks for the fabrication of superconducting devices and thus for quantum technology applications.
Footnotes |
† Electronic supplementary information (ESI) available: Additional data, technical details, auxiliary discussions supporting the main conclusions. See DOI: https://doi.org/10.1039/d2nr05864b |
‡ These authors contributed equally to this work. |
§ In this work, the B field was applied in the CoSi2/Si plane and parallel to the S/N interface. |
This journal is © The Royal Society of Chemistry 2023 |