Skip to main content

ORIGINAL RESEARCH article

Front. Electron. Mater, 07 September 2023
Sec. Superconducting Materials
This article is part of the Research Topic Progress on Superconducting Materials for SRF Applications View all 8 articles

Superheating field in superconductors with nanostructured surfaces

W. P. M. R. Pathirana
W. P. M. R. Pathirana1*A. GurevichA. Gurevich2
  • 1Department of Physics and Astronomy, Virginia Military Institute, Lexington, VA, United States
  • 2Department of Physics and Center for Accelerator Science, Old Dominion University, Norfolk, VA, United States

We report calculations of a DC superheating field Hsh in superconductors with nanostructured surfaces. Numerical simulations of the Ginzburg–Landau (GL) equations were performed for a superconductor with an inhomogeneous impurity concentration, a thin superconducting layer on top of another superconductor, and superconductor–insulator–superconductor (S-I-S) multilayers. The superheating field was calculated taking into account the instability of the Meissner state with a non-zero wavelength along the surface, which is essential for the realistic values of the GL parameter κ. Simulations were performed for the material parameters of Nb and Nb3Sn at different values of κ and the mean free paths. We show that the impurity concentration profile at the surface and thicknesses of S-I-S multilayers can be optimized to enhance Hsh above the bulk superheating fields of both Nb and Nb3Sn. For example, an S-I-S structure with a 90-nm-thick Nb3Sn layer on Nb can boost the superheating field up to ≈500 mT, while protecting the superconducting radio-frequency (SRF) cavity from dendritic thermomagnetic avalanches caused by local penetration of vortices.

1 Introduction

The superconducting radio-frequency (SRF) resonant cavities in particle accelerators enable high accelerating gradients with low power consumption. The best Nb cavities can have high quality factors Q ∼ 1010–1011 and sustain accelerating fields up to 50 MV/m at T = 1.5–2 K and 0.6–2 GHz (Padamsee et al., 2018; Gurevich, 2023). The peak RF fields B0 ≃ 200 mT at the equatorial surface of Nb cavities can approach the thermodynamic critical field Bc ≈ 200 mT at which the screening current density flowing at the inner cavity surface is close to the depairing current density JcBc/μ0λ—the maximum DC current density a superconductor can carry in the Meissner state (Tinkham, 2004), where λ is the penetration depth of the magnetic field. Thus, the breakdown fields of the best Nb cavities have nearly reached the DC superheating field BshBc (Galaiko, 1966; Matricon and Saint-James, 1967; Christiansen, 1969; Chapman, 1995; Catelani and Sethna, 2008; Transtrum et al., 2011; Lin and Gurevich, 2012). The Q factors can be increased by material treatments such as high-temperature annealing followed by low-temperature baking which not only increase Q (B0) and the breakdown field but also reduce the deterioration of Q at high fields (Ciovati et al., 2010; Posen et al., 2020). High-temperature treatments combined with the infusion of nitrogen, titanium, or oxygen can produce an anomalous increase of Q (B0) with RF field amplitude B0 = μ0H0 (Ciovati et al., 2016; Grassellino et al., 2017; Dhakal, 2020; Lechner et al., 2021). These advances raise the question about the fundamental limit of the breakdown fields of SRF cavities and the extent to which it can be pushed by surface nanostructuring and impurity management (Gurevich and Kubo, 2017; Gurevich, 2023).

Several ways of increasing the SRF breakdown fields by surface nanostructuring have been proposed. They include depositing high-Tc superconducting multilayers with thin dielectric interlayers (Gurevich, 2006; Gurevich, 2015; Kubo et al., 2014; Liarte et al., 2017; Kubo, 2021) or a dirty overlayer with a higher impurity concentration at the surface (Ngampruetikorn and Sauls, 2019). The DC superheating field of such structures has been evaluated using the London, Ginzburg–Landau (GL), and quasiclassical Usadel and Eilenberger equations in the limit of an infinite GL parameter κ = λ/ξ in which the breakdown of the Meissner state at H0 = Hsh occurs once the current density at the surface reaches the depairing limit (Gurevich, 2006; Gurevich, 2015; Kubo et al., 2014; Liarte et al., 2017; Kubo, 2021; Ngampruetikorn and Sauls, 2019). Yet, it has been well-established that in a more realistic case of a finite κ, the breakdown of the Meissner state at H = Hsh occurs due to the exponential growth of periodic perturbations of the order parameter and the magnetic field with a wavelength λc(ξ3λ)1/4 along the surface, where ξ is the coherence length (Christiansen, 1969; Chapman, 1995; Transtrum et al., 2011). The effect of such periodic Turing instability (Cross and Hohenberg, 1993) on Hsh can be particularly important for Nb cavities with κ ∼ 1. Addressing the effect of finite κ (which in turn depends on the mean free path) on Hsh in superconductors with nanostructured surfaces is the goal of this work.

We present the results of numerical calculations of a DC superheating field for different superconducting geometries in materials with finite κ and determine optimal surface nanostructures that can withstand the maximum magnetic field in the vortex-free Meissner state. In particular, we consider a bulk superconductor with a thin impurity diffusion layer, a clean superconducting overlayer separated by an insulating layer from the bulk (e.g., Nb3Sn-I-Nb3Sn), a thin dirty superconducting layer on the top of the same superconductor (e.g., dirty Nb3Sn-I-clean Nb3Sn), and a thin high-Tc superconducting layer on the top of a low-Tc superconductor (e.g., Nb3Sn-I-Nb). We calculate Hsh and determine an optimal layer thickness for each geometry by numerically solving the GL equations, taking into account both the non-linear screening of the applied magnetic field and the periodic instability of the Meissner state in a nanostructured superconductor.

The paper is organized as follows. The GL equations and methods of numerical detection of Hsh and the wavelength λc of a critical perturbation causing the instability of the Meissner state are presented in Section 2. The results of numerical calculations of Hsh for impurity diffusion layers and various S-I-S structures are given in Section 3 and Section 4, respectively. Section 5 contains discussion of the results, and Section 6 provides the conclusion with a summary. Computational details are given in Section Method and other technical details are given in Supplementary Appendices A, B.

2 GL equations and numerical detection of Hsh and kc

We first consider a semi-infinite uniform superconductor in a magnetic field H0 applied along the z-axis, parallel to the planar surface. In this case, the induced supercurrents flow in the xy plane and GL equations for the complex order parameter ψ = Δe, and two components of the vector potential Ax and Ay can be reduced to two coupled partial differential equations for the amplitude Δ(x, y, t) and the z-component of the magnetic field H (x, y, t). As shown in Supplementary Appendix A, these equations can be written in the following dimensionless form:

ḟf+f32f+κ2f3xh2+yh2=0,(1)
hf2=hκ2.(2)

Here, f (x, y) = Δ(x, y)/Δ0, Δ0(T) is the equilibrium order parameter in the bulk, h(x,y)=H(x,y)/2Hc, and all lengths are in units of the coherence length ξ and κ = λ/ξ. Despite the presence of the time derivative ḟ in Eqs (1, 2), they are essentially the quasi-static GL equations, but not the true time-dependent Ginzburg–Landau (TDGL) equations (Watts-Tobin et al., 1981; Sheikhzada and Gurevich, 2020) which describe a non-equilibrium superconductor at TTc. Here, ḟ is added just to detect the instability of the Meissner state in numerical simulations upon slow ramping the applied magnetic field. This procedure allows us to find the field H0 = Hsh above which the GL equations no longer have stationary solutions. Another way of numerical calculation of Hsh is based on finding the applied field at which the linearized Eqs (1, 2) have zero eigenmode with ḟ=0 (Christiansen, 1969; Chapman, 1995; Transtrum et al., 2011; Liarte et al., 2017), as summarized in Supplementary Appendix B. It turns out that the direct solving Eqs (1, 2) with the ad hoc term ḟ is much faster than solving the eigenmode problem. For slow magnetic ramp rates Ḣ0 used in our simulations, the resulting Hsh calculated by these two methods only differ by ≃ 1%, as it is shown in the next Sections. Equations (1, 2) were solved with the following boundary conditions:

h0,y=h0t,hx,0=hx,Ly,fx,0=fx,Ly,fLx,y=1,hLx,y=0,(3)

where h0=H0/2Hc and the lengths Lx and Ly of the simulation box Lx × Ly were chosen to be ≃ (50–150)ξ depending on κ. The details of the numerical calculations are given in the Supplementary Method.

Shown in Figures 1A,B are f (x, y) calculated at κ = 10 and the applied fields H0 slightly below and above Hsh. At H0 < Hsh, the order parameter f(x) is reduced at the surface by the flowing screening currents. At H0 > Hsh, the stationary f(x) becomes unstable with respect to spontaneously growing periodic perturbations δf (x, y, t) along the surface, as shown in Figure 1B. This Turing instability (Cross and Hohenberg, 1993) occurs with respect to a small disturbance δf (x, y, t) ∝ δf(x)eikyt, where the increment Γ(H0, k) depends on the wave vector k of spatial oscillations of f (x, y) along the surface as shown in Figure 2A. Below the superheating field, Γ(k) is negative so perturbations with all k decay exponentially and the Meissner state is stable. At the superheating field, Γ(k) first vanishes at a critical wave number k = kc at which Γ(k) is maximum. At H0 = Hsh + 0, the increment Γ(k) becomes positive at k = kc, making the Meissner state unstable with respect to a growing critical perturbation δf(x,y,t)δf(x)eikcy+Γ(kc)t with the wavelength λc = 2π/kc, while all other perturbations with kkc decay exponentially. We calculated Hsh by slowly ramping the applied field and detecting the onset of the exponential growth of f (x, y, t) with time as described in Method. The critical wavelength λc was evaluated from the maximum peak in the spatial Fourier transform of δf (0, y), as shown in Figure 2B. This instability is a precursor of the penetration of the vortex structure with the initial period λc(ξ3λ)1/4 smaller than the stationary vortex spacing λξ at H0Hsh and κ > 1. The aforementioned direct method for the calculation of Hsh is based on the numerical detection of the field threshold above which the stationary Meissner state does not exist. Here, the time scales of the transition to the vortex state at H0 > Hsh are irrelevant, provided that Hsh is calculated at low enough magnetic ramp rates at which Hsh is independent of ḣ0. We calculated Hsh at ḣ0105 and verified that Hsh is indeed practically independent of ḣ0, which is also consistent with the calculations of Hsh using both the TDGL and full non-equilibrium equations for dirty superconductors (Sheikhzada and Gurevich, 2020).

FIGURE 1
www.frontiersin.org

FIGURE 1. Spatial distribution of the order parameter calculated at κ =10, H0= Hsh −0 (A), and H0= Hsh +0 (B).

FIGURE 2
www.frontiersin.org

FIGURE 2. (A) Qualitative dependence of the instability increment Γ(H0, k) on the wave vector of perturbation k at different applied fields H0. (B) Snapshot of δf(y) at x =0, and H0= Hsh +0 calculated at κ =10.

We then compare some of our numerical results with the known analytical approximations for Hsh and kc at κ ≫ 1, given as follows (Christiansen, 1969; Chapman, 1995; Transtrum et al., 2011; Liarte et al., 2017).

HshHc53+0.545κ,(4)
λkc0.956κ3/4.(5)

At κ = 10–20, Eqs (4, 5) give the instability wavelength λc=6.57(ξ3λ)1/4(1.170.21)λ and Hsh approximately (23–16)% higher than Hsh = 0.745Hc in the limit of κ in which λc → 0 and the breakdown of the Meissner state at H = Hsh occurs once the current density at the surface reaches the depairing limit (Christiansen, 1969). Thus, even at κ = 10–20, characteristic of a dirty Nb or a clean stoichiometric Nb3Sn (Orlando et al., 1979; Posen and Hall, 2017), the periodic instability along the surface occurs on the scale of the order of the field penetration depth, so the self-consistent GL calculation of Hsh is required.

3 Superconductor with an impurity diffusion layer

We consider a dirty layer at the surface with a higher impurity concentration, as shown in Figure 3A. In our simulations, such a layer was modeled by a spatially varying coherence length and penetration depths ξ2(x)/ξ2=λ2/λ2(x)=1αexp(x/ld), as shown in Figure 3B. Here, ξ and λ are the corresponding bulk values far away from the surface, ld is the thickness of the diffusion layer, and the parameter α < 1 quantifies the reduction of ξ(0) = (1 − α)1/2ξ and the enhancement of λ(0) = (1 − α)−1/2λ at the surface. The ratio ξ2(x)/ξ2=λ2/λ2(x) is controlled by the impurity function χ(ℏvF/2πTcl(x)) (Werthamer, 1969) with an inhomogeneous mean free path l(x), which is defined in Supplementary Appendix A. The resulting GL equations take the following form:

ḟ=ff3+Sγfκ2Sγf3xh2+yh2,(6)
hSγf2hκ2=0,(7)

where κ = λ/ξ, Sγ=ξ2(x)/ξ2=1αexp(x/ld), and the lengths are in units of ξ. The boundary conditions are the same as in Eq. (3). Different impurity profiles were investigated by changing α and ld at κ = 2 and κ = 10, respectively, representing a cleaner and dirtier Nb.

FIGURE 3
www.frontiersin.org

FIGURE 3. (A) Impurity diffusion layer at the surface shown by the dark gray contrast. (B) Variations of normalized coherence length and penetration depth across a dirty layer with α =0.5.

The calculated dependencies of Hsh(ld) on the diffusion layer thickness at different α for κ = 2 and κ = 10 are shown in Figures 4A, B, respectively. One can see that Hsh(ld) first increases with ld, reaches a maximum, and then decreases with ld approaching a lower value of Hsh at ldξ. At κ = 2, Hsh(ld) is maximum at ld/ξ = 0.8, 0.9, 1.5 for α = 0.2, 0.5, 0.8. Similarly, at κ = 10, Hsh(ld) is maximum at ld/ξ = 4, 5, 10. Here, the diffusion layer can increase Hsh by 9% at κ = 2 and by 14% at κ = 10 as compared to a superconductor with an ideal surface. A qualitatively similar non-monotonic dependence of Hsh on ld was also obtained by solving the quasiclassical Eilenberger equations in the entire temperature range 0 < T < Tc (Ngampruetikorn and Sauls, 2019). The maximum in Hsh(d) results from a current counterflow induced in the dirty surface layer by a cleaner substrate with a smaller λ < λ(0) (Kubo et al., 2014; Gurevich, 2015), the magnitude of the peak increases as the diffusion layer gets dirtier. The curves Hsh(ld) cross over at larger ld for which Hsh() is determined by the surface GL parameter κ(0) = κ/(1 − α). As a result, Hsh() decreases as the material gets dirtier, in agreement with Eq. (4).

FIGURE 4
www.frontiersin.org

FIGURE 4. Superheating field Hsh(ld) as a function of the dirty layer thickness calculated at (A) κ =2 (B) κ =10 for different α. The dashed line shows Hsh(ld) calculated from the condition δ2F =0 and kδ2F =0 at (A) κ =2 and (B) κ =10 at α =0.5.

The direct numerical calculation of Hsh involves detecting the instability of the Meissner state with respect to an infinitesimal perturbation δf(x,y)=δf(x)eikcy+Γt with a finite wave number kc and an increment Γ(H0) changing the sign from negative at H0 < Hsh to positive at H0 > Hsh. Figure 5 shows the fast Fourier transform of a snapshot of δf (0, y, t) along y calculated at H0 = Hsh + 0 at α = 0.5, κ = 10, and different ratios ld/ξ. One can see that δf (0, y) has several harmonics even at the slow field ramp rate ḣ0=5105 used in our simulations. Such multi-mode temporal oscillations of δf (0, y) can result from a nonlinear mode coupling above the Turing instability threshold (Cross and Hohenberg, 1993), as well as a finite size of the computational box. In this case, the critical wave number kc would correspond to the highest peak in the Fourier spectrum of δfk (0, t). Yet, Figure 5 reveals two uneven peaks whose heights change differently as the ratio ld/ξ is varied. For instance, at ld/ξ = 5, the critical wave number kc is determined by the higher left peak observed in Figure 5, but as ld/ξ is increased to 6, the right peak becomes higher than the left one, so kc changes jumpwise at ld/ξ ≈ 5.5. The peak shifts toward higher λk values, providing a constant kc in this range of ld/ξ. The switching of kc between two values as ld/ξ is increased can be seen in “Fourier Transform. mp4” in Supplementary Video S1. To see the extent to which this ambiguity in kc may affect Hsh, we have also calculated kc and Hsh from the sign change of the second variation of the free energy δ2F caused by small perturbations of δf (x, y) and δh (x, y). In this method (Christiansen, 1969; Chapman, 1995; Transtrum et al., 2011), Hsh is determined by the conditions δ2F (kc, Hsh) = 0 and ∂δ2F/∂kc = 0. Figures 6A, B show λkc as a function of ld/ξ at κ = 2 and κ = 10 and different α computed from the second variation δ2F, as described in Supplementary Appendix B. One can see that the peaks in kc shown in Figure 5 are in the range of λkc ≈ 5–7 qualitatively matching λkc (ld) ≈ 5 at α = 0.5, as shown in Figure 6. Yet, the Hsh(ld/ξ) curves calculated by these two methods at α = 0.5 turned out to be very similar (the difference is approximately 1%), as shown by the dashed lines in Figure 4.

FIGURE 5
www.frontiersin.org

FIGURE 5. Discrete fast Fourier transform of δf (0, y) at H0= Hsh +0, α =0.5, and κ =10; (A) ld/ξ=5, (B) ld/ξ=5.5, and (C) ld/ξ=6. Evolution of these peaks with ld/ξ is shown in “Fourier Transform.gif” in Supplementary Video S1.

FIGURE 6
www.frontiersin.org

FIGURE 6. Dependencies of λkc on ld calculated from the conditions δ2F =0 and kδ2F =0 at (A) κ =2 and (B) κ =10 for different α.

4 S-I-S structures

Using the direct simulation method outlined in Section 2, we have calculated Hsh(d) and the critical wave number kc(d) for various S-I-S structures: the S layer of thickness d is separated by an insulator from the S substrate of the same material, a dirty S layer on the top of a cleaner superconductor (e.g., dirty Nb-I-clean Nb), and a thin high Tc overlayer on the top of a low Tc superconductor (e.g., Nb3Sn-I-Nb). Here, the I layer is assumed to be thick enough to fully suppress the Josephson coupling between the S overlayer and the bulk, but thinner than the S overlayer. The screening of the applied field in an S-I-S multilayer is shown in Figure 7.

FIGURE 7
www.frontiersin.org

FIGURE 7. Superconductor–insulator–superconductor structure. The vertical black line represents the insulating layer, and the red line shows the screened magnetic field profile H(x).

4.1 S overlayer on the top of the S-substrate

The GL Eqs (1, 2) for the S overlayer separated by the I layer from the substrate made of the same superconductor were solved in both S-domains with the boundary conditions given by Eq. (3) supplemented by the conditions of continuity of h (d + 0, y) = h (d − 0, y), parallel electric field Ey (d + 0, y) = Ey (d − 0, y), and zero current yh (d + 0) = yh (d − 0) = 0 through the I layer. Figure 8 shows that Hsh is a function of the thickness of the S overlayer d calculated at κ = 17 representing Nb3Sn. Here, a very thin S overlayer reduces Hsh(d), which then gradually increases with d, reaching a higher bulk value of Hsh at d > 9ξ2, where ξ2 is the coherence length in the S-substrate. The reduction of Hsh(d) at small d results from the I layer blocking the perpendicular currents produced by the critical perturbation and reducing its decay length in the bulk from λξ to d. In turn, the critical wave number kc(d) along the surface shown in Figure 9 increases jumpwise from kc ≈ 1.8/λ2 at d < 9ξ2 in a thin overlayer to kc ≈ 7.2/λ2 at d > 9ξ2, corresponding to the instability of the Meissner state in a semi-infinite superconductor. The calculated kc at d > 9ξ2 is approximately 10% smaller than kc ≈ 8/λ2 given by the asymptotic Eq. (5) at κ2 = 17.

FIGURE 8
www.frontiersin.org

FIGURE 8. Superheating field Hsh(d) calculated for the Nb3Sn-I-Nb3Sn structure.

FIGURE 9
www.frontiersin.org

FIGURE 9. Critical wave number kc(d) calculated for the Nb3Sn-I-Nb3Sn structure by solving the quasistatic GL equations directly as described in Section 2.

4.2 Dirty S overlayer on a cleaner S-substrate

A dirty S overlayer with a higher concentration of non-magnetic impurities on a cleaner S substrate of the same material is considered, assuming that both have the same Tc unaffected by non-magnetic impurity scattering (Tinkham, 2004). Superconductivity in the bulk is described by the following GL equations

ḟ2=2f2+f2f23κ22f23xh22+yh22,(8)
h2f22=h2κ22,(9)

where index 2 corresponds to the substrate parameters in which the lengths and f2 are in units of their respective bulk values of ξ2 and Δ2. In turn, the GL equations in the overlayer are as follows:

ḟ1=f1f13+ξ12ξ222f1ξ12ξ22κ12f13xh12+yh12,(10)
h1f12=h1ξ22ξ12κ12.(11)

Equations (811) were solved for a dirty Nb3Sn overlayer on a cleaner Nb3Sn with a mean free path l = 2 nm, λ1λ2(ξ2/l)1/2135 nm, ξ1(lξ2)1/23 nm, κ1 = 45 in the overlayer, and κ2 = 17 in the bulk. Figure 10 shows the calculated dependence of Hsh on the overlayer thickness which has a maximum at the optimum thickness dm ≈ 9ξ2. Such an optimal dirty overlayer can increase Hsh by approximately 10% as compared to the bulk Hsh. The behavior of Hsh(d) at a finite κ turns out to be similar to that calculated using the London and GL theories in the limit of κ in which the enhancement of Hsh at ddm results from the current counterflow induced by the substrate with a shorter λ2 in the overlayer with a larger λ1 (Kubo et al., 2014; Gurevich, 2015). Here, the cusp-like dependence of Hsh(d) is controlled by the instability of the Meissner state in the substrate at d < dm and by the instability of the Meissner state in the overlayer at d > dm, the overlayer partly screening the substrate and allowing it to withstand external fields higher than the bulk Hsh. The corresponding critical wave number kc(d) is shown in Figure 11. The jumpwise change of kc(d) reflects the switch from the instability of the Meissner state at the inner surface of the substrate at d < dm to the instability at the outer surface in the overlayer at d > dm. Such a jump in kc also follows Eq. (5) which gives kcλ2=0.956κ23/28 at d < dm and kcλ2=0.956κ11/4κ21/210.2.

FIGURE 10
www.frontiersin.org

FIGURE 10. Superheating field Hsh(d) calculated for the Nb3Sn(dirty)-I-Nb3Sn structure. The red dashed line shows Hsh(d) calculated from the London Eqs (1517) with Hsh1=0.855Hc and Hsh2=0.91Hc taken from the asymptotic limits of Hsh(d) at dλ1 and d =0, respectively.

FIGURE 11
www.frontiersin.org

FIGURE 11. The critical wave number kc(d) calculated for the Nb3Sn(dirty)-I-Nb3Sn structure by solving the quasi-static GL equations directly as described in Section 2.

4.3 High-Tc superconducting overlayer

Finally, we consider an S-I-S structure comprising a high-Tc layer on the top of a lower-Tc substrate. The order parameter f2 and the field h2 in the substrate are described in Eqs (8, 9), and the GL equations for f1 and h1 in the overlayer are given by

ḟ1=ζf1f13+s2f1κ̃2f13xh12+yh12,(12)
h1f12=λ22h1λ12ζκ22,(13)
ζ=1T/Tc11T/Tc2,s=ξ12ξ22ζ,κ̃2=ξ12λ14ξ22λ24κ22ζ3,(14)

where Tc1 and Tc2 are the critical temperatures of the overlayer and the substrate, respectively, and the order parameter and lengths are normalized to the respective parameters of the substrate. Equations (1214) are supplemented by the boundary conditions given by Eq. (3) and the conditions of field continuity and zero current through the I layer.

We solved the GL equations for a clean Nb3Sn overlayer on a bulk Nb using κ2 = 50/22 and κ1 = 17 (Orlando et al., 1979; Posen and Hall, 2017). The calculated superheating field Hsh(d) shown in Figure 12 has a maximum at dm ≈ 4ξ2. This behavior of Hsh(d) is similar to that of Hsh(d) considered in the previous section: Hsh(d) at d < dm is limited by the instability of the Meissner state in the Nb substrate partly screened by the Nb3Sn overlayer, while Hsh at d > dm is determined by the superheating field of Nb3Sn enhanced at ddm by the current counterflow caused by the Nb substrate. The corresponding dependence of kc(d) on the overlayer thickness is shown in Figure 13. The jumpwise change of kc(d) reflects the switch from the instability of the Meissner state at the inner surface of the low-Tc substrate at d < dm to the instability at the outer surface in the high-Tc overlayer at d > dm, which is similar to that shown in Figure 11. For the parameters used in the simulations, such Nb3Sn-I-Nb structures with ddm can boost the superheating field up to 2.2 times higher than the bulk Hsh2 of Nb (Gurevich, 2006; Gurevich, 2015) and approximately 5.3% higher than the bulk Hsh1 of Nb3Sn.

FIGURE 12
www.frontiersin.org

FIGURE 12. Superheating field Hsh(d) calculated for the Nb3Sn-I-Nb structure. The red dashed line shows Hsh(d) calculated from the London Eqs (1517) with Hsh1=2.28Hc and Hsh2=1.08Hc taken from the asymptotic limits of Hsh(d) at dλ1 and d =0, respectively.

FIGURE 13
www.frontiersin.org

FIGURE 13. The critical wave number kc(d) calculated for the Nb3Sn-I-Nb structure by solving the quasistatic GL equations directly as described in Section 2.

5 Discussion

The GL calculations of the DC superheating field at TTc self-consistently take into account the essential non-linear field screening and the periodic instability of the Meissner state in the entire range of the GL parameters which can be tuned by the impurities. This approach shows that the thicknesses of the impurity diffusion layer or S-I-S layers can be optimized to increase Hsh above the superheating fields of individual components. For instance, optimizing the diffusion length can enhance Hsh by ≃ 5–20% at κ = 10 and by ≃ 2–9% at κ = 2. An optimized dirty Nb3Sn overlayer deposited onto the Nb3Sn field by ≃ 10% as compared to Hsh of a clean Nb3Sn. This effect manifests itself in a non-monotonic dependence of Hsh on the dirty layer thickness due to the current counterflow induced at the surface by a superconducting substrate with a shorter field penetration depth. Such behavior of Hsh(d) is consistent with the previous calculations of Hsh based on the London (Gurevich, 2006; Gurevich, 2015; Kubo et al., 2014) or Usadel (Kubo, 2021) and Eilenberger (Ngampruetikorn and Sauls, 2019) theories at κ. To see the extent to which the London model is consistent with the GL results, we consider Hsh(d) calculated for an S-I-S multilayer in the London limit (Gurevich, 2015).

Hshd=coshd/λ1+λ2/λ1sinhd/λ1Hsh2,d<dm(15)
Hshd=λ1+λ2tanhd/λ1λ1tanhd/λ1+λ2Hsh1,d>dm(16)
dm=λ1lnλ1λ1+λ2Hsh1Hsh2+Hsh12Hsh22+1λ22λ12,(17)

where Hsh1 and Hsh2 are the bulk superheating fields of the overlayer and the substrate, respectively. Equation (15) describes Hsh(d) of S-I-S structures with thin overlayers (d < dm), where the Meissner state first breaks down at the surface of the substrate. Here, the high-Hc overlayer partly screens the substrate, allowing it to stay in the Meissner state at a higher applied field H0 = Hsh(d) than the bare substrate. If d > dm, the Meissner state first breaks down at the outer surface of the overlayer so that Hsh(d) → Hsh1 at dλ1. The maximum Hsh(dm) is given by:

Hshdm=Hsh12+1λ22λ12Hsh221/2.(18)

The maximum Hsh(dm)=(Hsh12+Hsh22)1/2 in an S-I-S multilayer occurs if (λ2/λ1)21.

Equations (1517) do not take into account the reduction of the superfluid density by current, non-linear field screening and the periodic instability of the Meissner state at a finite κ. The London model does not account for the size effect of reducing Hsh(d) if the overlayer thickness is smaller than the decay length λ1ξ1 of the critical perturbation, as was discussed in Section 4.1. Indeed, for identical materials of the substrate and overlayer (λ1 = λ2, Hsh1 = Hsh2), Eqs (1517) give dm = 0 and Hsh(d) = Hsh1 independent of d, which is inconsistent with the reduction of Hsh(d) at d ≲ 10ξ2, as shown in Figure 8. Yet, the London model captures the non-monotonic thickness dependence Hsh(d) calculated from the GL theory if the overlayer has different properties than those of the substrate, and the input parameters Hsh1 and Hsh2 in Eqs (1517) are exact bulk superheating fields for given values of κ and Hc, respectively (Gurevich, 2015). For instance, Figure 10 compares the GL and the London Hsh(d) calculated for an Nb3Sn(dirty)-I-Nb3Sn structure with a dirty overlayer for which the London model works reasonably well. For the Nb3Sn-I-Nb multilayers considered in Section 4.3, we observed a rather good agreement between Hsh(d) calculated from the GL theory and Eqs (1517), as shown in Figure 12. Such surprising accuracy of Eqs (1517) was also observed by Kubo (2021) in the Usadel simulations of dirty S-I-S multilayers in the entire temperature range of 0 < T < Tc at κ.

In the GL region TTc2, Eqs (1517) predict a significant change in the temperature dependence of Hsh(T) of the S-I-S multilayer with a higher-Tc overlayer as compared to Hsh2(T) of the bare substrate. If TTc2, the penetration depth λ2(T)(Tc2T)1/2 diverges and Hsh2(T) ∝ Tc2T vanishes, while λ1 and Hsh1 remain nearly independent of T. This case is characteristic of Nb3Sn-I-Nb for which Tc1 ≃ 2Tc2, the crossover thickness dm(T) increases with T and diverges logarithmically at TTc2. In turn, Hsh(d, T) obtained by Eq. (15) is limited by the small superheating field of the substrate partially screened by the high-Tc overlayer:

HshTλ2/λ1sinhd/λ1Hsh2TTc2T,(19)

Hence, Hsh(T) can be significantly higher than Hsh2(T) ∝ Tc2T at TTc2, particularly if d > λ1. As an illustration, Figure 14 shows Hsh(T) calculated from Eqs (1517) for different ratios d/λ1 and the parameters of Nb3Sn-I-Nb specified in Section 4.3. One can see both the square root temperature dependence given by Eq. (19) at TTc2 and a sharp change in Hsh(T) upon decreasing T as dm(T) becomes shorter than d and Hsh(T) crosses over to a nearly constant Hsh1(T) of the overlayer. For d/λ1 < 2, such a transition in Hsh(T) happens at lower T outside the GL temperature range shown in the figure.

FIGURE 14
www.frontiersin.org

FIGURE 14. Temperature dependencies of Hsh(T) calculated from Eqs (1517) for different ratios d/λ1 and the superconducting parameters of Nb3Sn-I-Nb specified in the text. The sharp change in the behavior of Hsh(T) upon decreasing T at d/λ1=2 occurs as dm(T) becomes shorter than d and Hsh(T) crosses over to a nearly constant Hsh1(T) of the overlayer. For smaller d/λ1, such a transition in Hsh(T) takes place at lower T outside the GL temperature range shown in the figure. Here, the blue line with d/λ1=0 represents Hsh2(T) of the bare substrate.

The relation between the static Hsh calculated here and the dynamic superheating field Hsd (T, ω) representing the fundamental field limit of superconductivity breakdown in SRF cavities depends on the rf frequency ω, temperature, and the material purity (Gurevich, 2023). The calculation of Hsd for S-I-S structures generally requires solving complex equations of non-equilibrium superconductivity, which in some cases can be reduced to TDGL equations at TTc (Watts-Tobin et al., 1981). The dynamic superheating field of an alloyed superconductor with an ideal surface at TTc was calculated from the microscopic theory (Sheikhzada and Gurevich, 2020), where it was shown that Hsd(T) approaches the static Hsh(T) at low frequencies ωωc but can be by a factor 2 larger than Hsh at ωωc. Here, the crossover frequency ωcmin(τϵ1,τΔ1) is set by the inelastic electron–phonon scattering time τϵ(T) ∝ T−3 and the TDGL relaxation time of the order parameter τΔ = πℏ/8kB(TcT) (Gurevich, 2023). For Nb and Nb3Sn at TTcNb=9.2 K, both τϵ(9K) ∼ 10–11 s and τΔ ∼ 10–11 s at TcT = 0.2 K are much shorter than the rf period at 1 GHz. In this case, the superconducting and quasiparticle screening currents follow practically instantaneously the driving rf field, and the quasistatic Hsh(T) considered here is applicable. The dynamic superheating field at lower temperatures T ≃ 2 K at GHz frequencies has not yet been calculated from a microscopic theory.

In this work, Hsh was calculated for S-I-S structures with ideal surfaces and interfaces without topographical and material defects or weakly coupled grain boundaries in the overlayer and the substrate. Topographical and other surface defects can locally reduce the field onset of the dissipative penetration of vortices and reduce the global Hsh, as was shown by TDGL simulations (Vodolazov, 2000; Pack et al., 2020; Wang et al., 2022). Likewise, Hsh(T) can be reduced by weakly coupled grain boundaries causing premature proliferation of mixed Abrikosov–Josephson vortices or phase slips (Sheikhzada and Gurevich, 2017). The I interlayer in S-I-S coating can mitigate these detrimental effects by the following: 1. increasing the cavity breakdown field by thin high-Hc overlayers and 2. confining vortices penetrating at surface defects in a thin overlayer and blocking flux penetration in the cavity wall, where it can trigger thermo-magnetic avalanches, causing global superconductivity breakdown (Gurevich, 2015; Gurevich, 2023). The S-I-S coating can provide these two goals if the overlayer thickness does not exceed λ1 (Gurevich, 2006). In this work, we calculated the upper limit of Hsh and showed how the S-I-S geometry can be optimized to increase Hsh at dλ1.

6 Conclusion

Our numerical GL calculations of the DC superheating field in superconductors with nanostructured surfaces cover the entire range of 1 < κ < and account for both the non-linear Meissner screening and the instability with a finite wave number kc at H0 = Hsh. We showed that there are optimum thicknesses of the impurity diffusion layer and the superconducting overlayer which maximize Hsh. These results suggest the possible ways of increasing the breakdown fields by surface nanostructuring and can help understand the ways of optimizing SRF cavities to achieve higher accelerating gradients.

Data availability statement

The original contributions presented in the study are included in the article/Supplementary Material; further inquiries can be directed to the corresponding author.

Author contributions

WP performed all numerical simulations, analyzed the results, and wrote the first draft of the manuscript. AG initiated and supervised the project and wrote the revised manuscript. All authors contributed to the article and approved the submitted version.

Funding

This work was supported by DOE under grant DE-SC 100387–020 (ODU) and by the Virginia Military Institute (VMI).

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Supplementary material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/femat.2023.1246016/full#supplementary-material

References

Catelani, G., and Sethna, J. P. (2008). Temperature dependence of the superheating field for superconductors in the high-κ London limit. Phys. Rev. B 78, 224509. doi:10.1103/PhysRevB.78.224509

CrossRef Full Text | Google Scholar

Chapman, S. J. (1995). Superheating field of type II superconductors. SIAM J. Appl. Math. 55 (5), 1233–1258. doi:10.1137/S0036139993254760

CrossRef Full Text | Google Scholar

Christiansen, P. V. (1969). Magnetic superheating field in high-κ superconductors. Solid State Commun. 7 (10), 727–729. doi:10.1016/0038-1098(69)90652-8

CrossRef Full Text | Google Scholar

Ciovati, G., Myneni, G., Stevie, F., Maheshwari, P., and Griffis, D. (2010). High field Q slope and the baking effect: review of recent experimental results and new data on Nb heat treatments. Phys. Rev. St. Accel. Beams 13, 022002. doi:10.1103/PhysRevSTAB.13.022002

CrossRef Full Text | Google Scholar

Ciovati, G., Dhakal, P., and Myneni, G. (2016). Superconducting radio-frequency cavities made from medium and low-purity niobium ingots. Supercond. Sci. Technol. 29, 064002. doi:10.1088/0953-2048/29/6/064002

CrossRef Full Text | Google Scholar

COMSOL COMSOL Multiphysics modeling software. Available at: https://www.comsol.com (Accessed March 15, 2017).

Google Scholar

Cross, M. C., and Hohenberg, P. C. (1993). Pattern formation outside of equilibrium. Rev. Mod. Phys. 65, 851–1112. doi:10.1103/RevModPhys.65.851

CrossRef Full Text | Google Scholar

Dhakal, P. (2020). Nitrogen doping and infusion in SRF cavities: A review. Open Phys. J. 5, 100034. doi:10.1016/j.physo.2020.100034

CrossRef Full Text | Google Scholar

Galaiko, V. P. (1966). Stability limits of the superconducting state in a magnetic field for superconductors of the second kind. Zh. Exp. Teor. Fiz. 50, 717–723.

Google Scholar

Grassellino, A., Romanenko, A., Trenikhina, Y., Checchin, M., Martinello, M., Melnychuk, P., et al. (2017). Unprecedented quality factors at accelerating gradients up to 45 MVm-1 in niobium superconducting resonators via low temperature nitrogen infusion. Supercond. Sci. Technol. 30, 094004. doi:10.1088/1361-6668/aa7afe

CrossRef Full Text | Google Scholar

Gurevich, A. (2006). Enhancement of rf breakdown field of superconductors by multilayer coating. Appl. Phys. Lett. 88 (1), 012511. doi:10.1063/1.2162264

CrossRef Full Text | Google Scholar

Gurevich, A., and Kubo, T. (2017). Surface impedance and optimum surface resistance of a superconductor with an imperfect surface. Phys. Rev. B 96, 184515. doi:10.1103/PhysRevB.96.184515

CrossRef Full Text | Google Scholar

Gurevich, A. (2015). Maximum screening fields of superconducting multilayer structures. AIP Adv. 5 (1), 017112. doi:10.1063/1.4905711

CrossRef Full Text | Google Scholar

Gurevich, A. (2023). Tuning microwave losses in superconducting resonators. Supercond. Sci. Technol. 36, 063002. doi:10.1088/1361-6668/acc214

CrossRef Full Text | Google Scholar

Kubo, T., Iwashita, Y., and Saeki, T. (2014). Radio-frequency electromagnetic field and vortex penetration in multilayered superconductors. Appl. Phys. Lett. 104, 032603. doi:10.1063/1.4862892

CrossRef Full Text | Google Scholar

Kubo, T. (2021). Superheating fields of semi-infinite superconductors and layered superconductors in the diffusive limit: structural optimization based on the microscopic theory. Supercond. Sci. Technol. 34, 045006. doi:10.1088/1361-6668/abdedd

CrossRef Full Text | Google Scholar

Lechner, E. M., Angle, J. W., Stevie, F. A., Kelley, M. J., Reece, C. E., and Palczewski, A. D. (2021). RF surface resistance tuning of superconducting niobium via thermal diffusion of native oxide. Appl. Phys. Lett. 119, 082601. doi:10.1063/5.0059464

CrossRef Full Text | Google Scholar

Liarte, D. B., Posen, S., Transtrum, M. K., Catelani, G., Liepe, M., and Sethna, J. P. (2017). Theoretical estimates of maximum fields in superconducting resonant radio frequency cavities: stability theory, disorder, and laminates. Supercond. Sci. Technol. 30, 033002. doi:10.1088/1361-6668/30/3/033002

CrossRef Full Text | Google Scholar

Lin, F. P. J., and Gurevich, A. (2012). Effect of impurities on the superheating field of type-II superconductors. Phys. Rev. B 85 (5), 054513. doi:10.1103/PhysRevB.85.054513

CrossRef Full Text | Google Scholar

Matricon, J., and Saint-James, D. (1967). Superheating fields in superconductors. Phys. Lett. A 24 (5), 241–242. doi:10.1016/0375-9601(67)90412-4

CrossRef Full Text | Google Scholar

MATLAB Matlab. Available at: https://www.mathworks.com/products/matlab.html (Accessed March 15, 2017).

Google Scholar

Ngampruetikorn, V., and Sauls, J. A. (2019). Effect of inhomogeneous surface disorder on the superheating field of superconducting RF cavities. Phys. Rev. Res. 1, 012015. doi:10.1103/PhysRevResearch.1.012015

CrossRef Full Text | Google Scholar

Orlando, T. P., McNiff, E. J., Foner, S., and Beasley, M. R. (1979). Critical fields, Pauli paramagnetic limiting, and material parameters of Nb3Sn and V3Si. Phys. Rev. B 19, 4545–4561. doi:10.1103/physrevb.19.4545

CrossRef Full Text | Google Scholar

Pack, A. R., Carlson, J., Wadsworth, S., and Transtrum, M. K. (2020). Vortex nucleation in superconductors within time-dependent Ginzburg-Landau theory in two and three dimensions: role of surface defects and material inhomogeneities. Phys. Rev. B 101, 144504. doi:10.1103/PhysRevB.101.144504

CrossRef Full Text | Google Scholar

Padamsee, H., Knobloch, J., and Hays, T. (2018). RF superconductivity for accelerators. 2nd ed. Germany: Wiley-VCH Verlag GmbH and Co. KGaA.

Google Scholar

Posen, S., and Hall, D. L. (2017). Nb3Sn superconducting radiofrequency cavities: fabrication, results, properties, and prospects. Supercond. Sci. Technol. 30, 033004. doi:10.1088/1361-6668/30/3/033004

CrossRef Full Text | Google Scholar

Posen, S., Romanenko, A., Grassellino, A., Melnychuk, O. S., and Sergatskov, D. A. (2020). Ultralow surface resistance via vacuum heat treatment of superconducting radio-frequency cavities. Phys. Rev. Appl. 13, 014024. doi:10.1103/PhysRevApplied.13.014024

CrossRef Full Text | Google Scholar

Sheikhzada, A., and Gurevich, A. (2020). Dynamic pair-breaking current, critical superfluid velocity, and nonlinear electromagnetic response of nonequilibrium superconductors. Phys. Rev. B 102, 104507. doi:10.1103/PhysRevB.102.104507

CrossRef Full Text | Google Scholar

Sheikhzada, A., and Gurevich, A. (2017). Dynamic transition of vortices into phase slips and generation of vortex-antivortex pairs in thin film Josephson junctions under dc and ac currents. Phys. Rev. B 95, 214507. doi:10.1103/PhysRevB.95.214507

CrossRef Full Text | Google Scholar

Tinkham, M. (2004). Introduction to superconductivity. New York: Dover Publication.

Google Scholar

Transtrum, M. K., Catelani, G., and Sethna, J. P. (2011). Superheating field of superconductors within Ginzburg-Landau theory. Phys. Rev. B 83 (9), 094505. doi:10.1103/PhysRevB.83.094505

CrossRef Full Text | Google Scholar

Vodolazov, D. Yu. (2000). Effect of surface defects on the first field for vortex entry in type-II superconductors. Phys. Rev. B 62, 8691–8694. doi:10.1103/PhysRevB.62.8691

CrossRef Full Text | Google Scholar

Wang, Q-Y., Xue, C., Dong, C. C., and Zhou, Y-H. (2022). Effects of defects and surface roughness on the vortex penetration and vortex dynamics in superconductor–insulator–superconductor multilayer structures exposed to RF magnetic fields: numerical simulations within TDGL theory. Supercond. Sci. Technol. 35, 045004. doi:10.1088/1361-6668/ac4ad1

CrossRef Full Text | Google Scholar

Watts-Tobin, R. J., Krähenbühl, Y., and Kramer, L. (1981). Nonequilibrium theory of dirty, current-carrying superconductors: phase-slip oscillators in narrow filaments near Tc. J. Low. Temp. Phys. 42, 459–501. doi:10.1007/BF00117427

CrossRef Full Text | Google Scholar

Werthamer, N. R. (1969). in Superconductivity. Editor R. D. Parks (New York: Marcel Dekker Inc.), 321–370.

Google Scholar

Keywords: superheating field, superconductors, multilayered superconductors, vortices, Ginzburg–Landau theory

Citation: Pathirana WPMR and Gurevich A (2023) Superheating field in superconductors with nanostructured surfaces. Front. Electron. Mater. 3:1246016. doi: 10.3389/femat.2023.1246016

Received: 24 June 2023; Accepted: 15 August 2023;
Published: 07 September 2023.

Edited by:

Pashupati Dhakal, Jefferson Lab (DOE), United States

Reviewed by:

Akira Miyazaki, UMR9012 Laboratoire de Physique des 2 infinis Irène Joliot-Curie (IJCLab), France
Gianluigi Catelani, Forschungszentrum Jülich, Germany

Copyright © 2023 Pathirana and Gurevich. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: W. P. M. R. Pathirana, walivepathiranagemr@vmi.edu

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.