This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

How High a Field Can Be and Has Been Achieved in Superconducting Bulk Niobium Cavities Across Different RRR Values?

Takayuki Kubo [email protected] High Energy Accelerator Research Organization (KEK), Tsukuba, Ibaraki 305-0801, Japan The Graduate University for Advanced Studies (Sokendai), Hayama, Kanagawa 240-0193, Japan
Abstract

This Brief Note explores the relationship between residual resistivity ratio (RRR) and the maximum surface magnetic field in superconducting bulk niobium (Nb) cavities. Data from the 1980s to 2020s, covering RRR values from 30 to 500, are compared with theoretical performance limits, including the lower critical field (Bc1B_{c1}), superheating field (BshB_{\rm sh}), and thermal runaway field (BrunB_{\rm run}). The results show that modern Nb cavities are approaching BrunB_{\rm run} and the metastability region above Bc1B_{c1} across the entire RRR range but remain below the fundamental limit at BshB_{\rm sh}. Achieving BshB_{\rm sh} requires not only advanced high-gradient surface processing but also improved thermal stability with low surface resistance, ultra-pure Nb, and optimized Kapitza conductance to ensure Brun>BshB_{\rm run}>B_{\rm sh}.

Superconducting radio-frequency (SRF) cavities accelerate particles using an electric field along the axis [1]. The accelerating gradient, EaccE_{\rm acc}, is a key performance measure, as higher gradients shorten the accelerator length required for a given energy. However, the maximum EaccE_{\rm acc} is constrained by the material properties of the cavity.

The first limiting factor is the superconducting properties of the material, particularly the lower critical field (Bc1B_{c1}) and the superheating field (BshB_{\rm sh}[2, 3, 4, 5]. As EaccE_{\rm acc} increases, the peak surface magnetic field B0B_{0} rises, where B0=gEaccB_{0}=gE_{\rm acc}, with gg set by the cavity design. Initially, the cavity remains in the Meissner state, but as the field increases, vortices penetrate, leading to RF losses and quenching. The Meissner state becomes metastable at Bc1B_{c1}, with an upper limit at BshB_{\rm sh}. Thus, the maximum achievable field, B0(max)B_{0}^{\rm(max)}, is constrained within the metastable band between Bc1B_{c1} and BshB_{\rm sh}. Both fields depend on the electron mean free path, tied to the residual resistivity ratio (RRR).

Another limitation comes from the material’s thermal stability. Even without surface defects, like normal conducting residues, topographical irregularities, or weak superconducting precipitates, the exponential temperature dependence of the surface resistance RsR_{s} creates a positive feedback loop [6, 7]. This feedback between the absorbed power, (1/2)RsH02(1/2)R_{s}H_{0}^{2}, and the resulting temperature rise leads to defect-independent thermal runaway above a threshold field, BrunB_{\rm run}. The threshold BrunB_{\rm run} depends on factors such as surface resistance, cavity wall thickness, thermal conductivity, and Kapitza conductance.

These fundamental limits, Bc1B_{c1}, BshB_{\rm sh}, and BrunB_{\rm run}, can be enhanced by using high-purity niobium with a high RRR. Although the link between higher RRR and increased theoretical field is well known, a comprehensive summary of decades of cavity tests with varying RRR values remains unavailable.

This Brief Note compiles data [8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21] from the 1980s to 2020s, with RRR values ranging from 30 to 500, to investigate the relationship between RRR and the maximum achievable fields in Nb cavities. The results are compared with rough theoretical estimates of the fields as a function of RRR. This work establishes the foundation for discussing field limits in bulk Nb technologies, including both simple bulk Nb and multilayer approaches with thin films deposited on Nb [22, 23, 24, 25, 26, 27].

Refer to caption
Figure 1: Maximum surface magnetic field B0maxB_{0}^{\rm max} (mT) for 1.31.3 and 1.5GHz1.5\,{\rm GHz} cavities as a function of RRR. The cavities were processed using chemical polishing and/or electropolishing, with or without low-temperature baking. The 147 data points are sourced from Refs. [8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21], where RRR values were explicitly reported. Some high-performing cavities, such as those in Ref. [28], which achieved 190mT190\,{\rm mT}, are not included due to the absence of RRR data.

Figure 1 compiles cavity test results of B0(max)B_{0}^{\rm(max)} for various RRR values collected over several decades [8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21]. The data show a broad range of B0(max)B_{0}^{\rm(max)}, with some cavities achieving notable results (180mT\gtrsim 180\,{\rm mT}). Cavities with RRR200{\rm RRR}\simeq 200 and even RRR100{\rm RRR}\simeq 100 reached B0150mTB_{0}\simeq 150\,{\rm mT}, corresponding to Eacc=35MV/mE_{\rm acc}=35\,{\rm MV/m} in Tesla-shaped cavities. However, concluding that RRR100{\rm RRR}\sim 100 is enough to consistently reach such gradients would be incorrect. The data do not reflect the success rate of cavities reaching these fields. For instance, the highest result for RRR130{\rm RRR}\sim 130, achieving B0max=175mTB_{0}^{\rm max}=175\,{\rm mT}, required 12 tests with repeated surface treatments [12]. It’s important to note this figure shows the fields achieved in one or more tests without indicating the yield of high-performing cavities.

This dataset is compared with rough theoretical estimates of the maximum achievable fields (BshB_{\rm sh}, Bc1B_{c1}, BrunB_{\rm run}) as functions of RRR, derived in the following.

BshB_{\rm sh} represents the ultimate stability limit for the Meissner state, beyond which it becomes fully unstable. For dirty niobium, Ginzburg-Landau (GL) theory gives Bsh|TTc0.745Bc|TTcB_{\rm sh}|_{T\simeq T_{c}}\simeq 0.745B_{c}|_{T\simeq T_{c}} (see, e.g., Ref. [25] and references therein). As temperature decreases, this coefficient rises, with microscopic theory predicting Bsh0.8BcB_{\rm sh}\sim 0.8B_{c} at TTcT\ll T_{c} [2]. In the dirty limit, Bsh=0.795Bc160mTB_{\rm sh}=0.795B_{c}\simeq 160\,{\rm mT} at T=0T=0 [3, 4, 26]. On the other hand, for clean niobium (κ1\kappa\sim 1), GL theory predicts Bsh|TTc(1.21.3)Bc|TTcB_{\rm sh}|_{T\simeq T_{c}}\simeq(1.2-1.3)B_{c}|_{T\simeq T_{c}} [29], and extrapolating to TTcT\ll T_{c} yields Bsh(1.21.3)Bc240260mTB_{\rm sh}\simeq(1.2-1.3)B_{c}\simeq 240-260\,{\rm mT}. Since BshB_{\rm sh} at low temperatures remains uncertain in most cases except for dirty niobium [2, 3, 26], we adopt the standard practice of extrapolating the GL-based formula. We use Christiansen’s formula [30]:

Bsh(κ)=5Bc3(1+0.731κ),\displaystyle B_{\rm sh}(\kappa)=\frac{\sqrt{5}B_{c}}{3}\biggl{(}1+\frac{0.731}{\sqrt{\kappa}}\biggr{)}, (1)

originally derived for κ>κc\kappa>\kappa_{c} (where κc1.1\kappa_{c}\sim 1.1 separates one- and two-dimensional critical perturbations). Numerical calculations [5] show this formula remains accurate for BshB_{\rm sh} even at κ0.6\kappa\gtrsim 0.6. The Pade approximation [29], derived for κ<κc\kappa<\kappa_{c}, gives similar results to Eq. (1) in the range 0.6κ10.6\lesssim\kappa\lesssim 1.

Bc1B_{c1} represents the field at which the Meissner state becomes metastable. To estimate Bc1B_{c1}, we use the approximate GL expression Bc1=f(κ)BcB_{c1}=f(\kappa)B_{c} [31], where f(κ)={lnκ+C(κ)}/2κf(\kappa)=\{\ln\kappa+C(\kappa)\}/\sqrt{2}\kappa, and C(κ)=0.5+(1+ln2)/(2κ2+2)C(\kappa)=0.5+(1+\ln 2)/(2\kappa-\sqrt{2}+2). Here, BcB_{c} denotes the thermodynamic critical field. This formula is chosen for its balance between computational simplicity and accuracy. Though derived from GL theory and technically valid near TcT_{c}, the formula can be rewritten to provide a good approximation even at low temperatures. Since BcB_{c} is independent of nonmagnetic impurity concentration (Anderson’s theorem), we get Bc1(κ)/Bc1clean=f(κ)/f(κclean)B_{c1}(\kappa)/B_{c1}^{\rm clean}=f(\kappa)/f(\kappa_{\rm clean}) or

Bc1(κ)=f(κ)f(κclean)Bc1clean.\displaystyle B_{c1}(\kappa)=\frac{f(\kappa)}{f(\kappa_{\rm clean})}B_{c1}^{\rm clean}. (2)

Here, Bc1cleanB_{c1}^{\rm clean} and κclean\kappa_{\rm clean} refer to pure niobium. Using Bc1clean=185mTB_{c1}^{\rm clean}=185\,{\rm mT} at T=0T=0 and κclean=0.8\kappa_{\rm clean}=0.8 for Nb [32], this approximation matches experimental data for Bc1B_{c1} at T=0T=0 across κ<1.5\kappa<1.5 [32], fitting our range of interest (as discussed, κ<1.5\kappa<1.5 corresponds to RRR20{\rm RRR}\gtrsim 20). Hence, extrapolating this formula to lower temperatures appears effective.

The GL parameter κ\kappa is related to the mean free path \ell via microscopic theory [33]:

κ()=1χ(aimp)κclean,\displaystyle\kappa(\ell)=\frac{1}{\chi(a_{\rm imp})}\kappa_{\rm clean}, (3)

where χ\chi is the Gor’kov function,

χ(aimp)=87ζ(3)n=01(2n+1)2(2n+1+aimp),\displaystyle\chi(a_{\rm imp})=\frac{8}{7\zeta(3)}\sum_{n=0}^{\infty}\frac{1}{(2n+1)^{2}(2n+1+a_{\rm imp})}, (4)
aimp=π2eγEξ00.882ξ0,\displaystyle a_{\rm imp}=\frac{\pi}{2e^{\gamma_{E}}}\frac{\xi_{0}}{\ell}\simeq 0.882\frac{\xi_{0}}{\ell}, (5)
κclean=2eγEπ67ζ(3)λ0ξ00.957λ0ξ0.\displaystyle\kappa_{\rm clean}=\frac{2e^{\gamma_{E}}}{\pi}\sqrt{\frac{6}{7\zeta(3)}}\frac{\lambda_{0}}{\xi_{0}}\simeq 0.957\frac{\lambda_{0}}{\xi_{0}}. (6)

Here, ζ\zeta is the Riemann zeta function, γE=0.577\gamma_{E}=0.577 is Euler’s constant, ξ0=vf/πΔ0\xi_{0}=\hbar v_{f}/\pi\Delta_{0} is the BCS coherence length, and λ02=(2/3)μ0N0e2vf2\lambda_{0}^{-2}=(2/3)\mu_{0}N_{0}e^{2}v_{f}^{2} is the BCS penetration depth in the clean limit at T=0T=0, with vfv_{f} as the Fermi velocity and N0N_{0} the normal electron density of states at the Fermi energy. The parameter aimpa_{\rm imp} characterizes the material’s dirtiness. The mean free path \ell is expressed in terms of RRR as =(3.7×1016Ωm2)/ρn\ell=(3.7\times 10^{-16}\,{\rm\Omega\cdot m^{2}})/\rho_{n}, where ρnρn(295K)/RRR\rho_{n}\simeq\rho_{n}(295\,{\rm K})/{\rm RRR} and ρn(295K)=1.45×107Ωm\rho_{n}(295\,{\rm K})=1.45\times 10^{-7}\,{\rm\Omega\cdot m}. This leads to:

=RRR×2.55nm.\displaystyle\ell={\rm RRR}\times 2.55\,{\rm nm}. (7)

These formulas [Eqs. (1)-(7)], along with empirical temperature dependencies of (Bc1clean,Bc)1(T/Tc)2(B_{c1}^{\rm clean},B_{c})\propto 1-(T/T_{c})^{2} with Tc=9.2KT_{c}=9.2\,{\rm K}, form the poor man’s formula for calculating BshB_{sh} and Bc1B_{c1} as functions of RRR (or \ell) at any temperature.

Fig. 2 (a) shows BshB_{\rm sh} (purple) and Bc1B_{c1} (blue) as functions of RRR at 2K2\,{\rm K}. The input parameters are Bc1clean|T=0=185mTB_{c1}^{\rm clean}|_{T=0}=185\,{\rm mT}, κclean=0.8\kappa_{\rm clean}=0.8, and ξ0=45nm\xi_{0}=45\,{\rm nm} [32]. The blue area between the curves represents the metastability region, beyond which the Meissner state becomes completely unstable. Although many cavities surpass Bc1B_{c1} and enter this region, none have yet reached BshB_{\rm sh}.

Refer to caption
Refer to caption
Figure 2: (a) BshB_{\rm sh} (purple) and Bc1B_{c1} (blue) at 2K2\,{\rm K} as functions of RRR, estimated using Eqs. (1)-(7). (b) BrunB_{\rm run} (red) at 2K2\,{\rm K} as a function of RRR, calculated for RBCS(2K)10nΩR_{\rm BCS}(2\,{\rm K})\leq 10\,{\rm n\Omega}. The 147 data points from Fig. 1 are overlaid (gray points) in both figures for comparison.

BrunB_{\rm run}, the threshold field for defect-independent thermal runaway, is another critical limit for superconducting cavities. This breakdown field [6, 7] is analyzed using the heat balance equation for the inner surface temperature. To focus on the theoretical field limit, we assume the residual surface resistance, which lowers BrunB_{\rm run}, is negligible compared to the thermally activated quasiparticle contribution (i.e., RresRBCSR_{\rm res}\ll R_{\rm BCS}), a condition approximately met in well-controlled 2K2\,{\rm K} experiments, especially when the magnetic environment is optimized to minimize trapped flux, a key source of residual resistance (see, e.g., Ref. [34] and references therein). Under these conditions, BrunB_{\rm run} is expressed as [6, 7]:

Brun=μ02T/TcαeTrRBCS,\displaystyle B_{\rm run}=\mu_{0}\sqrt{\frac{2T/T_{c}}{\alpha e}\frac{T}{rR_{\rm BCS}}}, (8)

where e=2.718e=2.718 is Napier’s constant, α=Δ/kTc1.9\alpha=\Delta/kT_{c}\simeq 1.9 for Nb, r(T,RRR)=hK1+d/Kr(T,{\rm RRR})=h_{\rm K}^{-1}+d/K is the thermal resistance, KK is the thermal conductivity, and hKh_{\rm K} is the Kapitza conductance. The BCS surface resistance is given by RBCS=(A/T)eΔ/kTR_{\rm BCS}=(A/T)e^{-\Delta/kT}, where AA is a constant. Although RBCSR_{\rm BCS} depends on the field [35, 36], we do not account for this in the current analysis. Instead, we vary AA to estimate BrunB_{\rm run} with associated uncertainties.

The thermal conductivity KK depends on RRR. For our calculations, we set d=2.8mmd=2.8\,{\rm mm} and use the formula from Ref. [37] to determine K(RRR)K({\rm RRR}), applying the room temperature resistivity ρn(295K)=1.45×107Ωm\rho_{n}(295\,{\rm K})=1.45\times 10^{-7}\,{\rm\Omega\cdot m} and the phonon mean free path ph=75μm\ell_{\rm ph}=75\,{\rm\mu m}. For the Kapitza conductance hKh_{\rm K}, we adopt hK=5000Wm2K1h_{\rm K}=5000\,{\rm W\,m^{-2}\,K^{-1}} from Ref. [38]. With these values, the thermal resistance (r=hK1+d/Kr=h_{\rm K}^{-1}+d/K) equals approximately 7Km2W17\,{\rm K\,m^{2}\,W^{-1}} at 2K2\,{\rm K} for RRR=300{\rm RRR}=300. This estimate aligns well with recent measurements of thermal resistance in the range of 4-6Km2W14\text{-}6\,{\rm K\,m^{2}\,W^{-1}} at 2K2\,{\rm K} for RRR>250{\rm RRR}>250 [39].

Fig. 2 (b) shows BrunB_{\rm run} at 2K2\,{\rm K} as a function of RRR. Although the absolute value of BrunB_{\rm run} contains some uncertainty due to variations in the thermal resistance rr, and RBCSR_{\rm BCS} generally exhibits intrinsic field dependence, ranging from 1nΩ1\,{\rm n\Omega} to 10nΩ10\,{\rm n\Omega} [35, 36], which introduces some imprecision in BrunB_{\rm run} within the red region of the figure, the overall trend suggests that BrunB_{\rm run} remains on the same order of magnitude as Bc1B_{c1} and BshB_{\rm sh}.

To achieve BshB_{\rm sh}, it is essential to ensure that BrunB_{\rm run} exceeds BshB_{\rm sh}. If BrunB_{\rm run} remains below BshB_{\rm sh}, thermal runaway will occur before reaching the superheating field, posing an insurmountable barrier. Achieving this will require not only advanced high-gradient surface processing but also enhanced thermal stability, characterized by low surface resistance, adequate thermal conductivity (i.e., bulk RRR), and optimized Kapitza conductance to guarantee Brun>BshB_{\rm run}>B_{\rm sh}.

Before concluding, it’s important to note the role of defects. Cavity performance is impacted by resistance to local heating at defect sites, governed by thermal conductivity KK, which depends on RRR. Defects acting as heat sources cause temperature spikes, but higher RRR materials mitigate these, enabling higher fields. This drove efforts in the 1980s and 1990s to raise niobium’s RRR from reactor-grade (RRR20{\rm RRR}\simeq 20) to over 300. This improvement in RRR and its impact on cavity performance is well documented during this period.

Acknowledgements.
I am deeply grateful to those who supported my three-year paternity leave [40], during which I revisited an unfinished project initiated in 2018 for the TESLA collaboration meeting held at RIKEN, leading to the development of this Brief Note. This work was supported by JSPS KAKENHI Grants No. JP17KK0100 and Toray Science Foundation Grants No. 19-6004.

References

  • [1] H. Padamsee, J. Knobloch, and T. Hays, RF Superconductivity for Accelerators (Wiley-VCH, Weinheim, 2008).
  • [2] F. Pei-Jen and A. Gurevich, Effect of impurities on the superheating field of type-II superconductors, Phys. Rev. B 85, 054513 (2012).
  • [3] T. Kubo, Superfluid flow in disordered superconductors with Dynes pair-breaking scattering: Depairing current, kinetic inductance, and superheating field, Phys. Rev. Res. 2, 033203 (2020).
  • [4] T. Kubo, Erratum: Superfluid flow in disordered superconductors with Dynes pair-breaking scattering: Depairing current, kinetic inductance, and superheating field [Phys. Rev. Research 2, 033203 (2020)], Phys. Rev. Res. 6, 039002 (2024).
  • [5] M. K. Transtrum, G. Catelani, and J. P. Sethna, Superheating fields of superconductors within Ginzburg-Landau theory, Phys. Rev. B 83, 094505 (2011).
  • [6] A. Gurevich, Thermal RF breakdown of superconducting cavities, in Proceedings of the Workshop on Pushing the Limits of RF superconductivity, Argonne National Laboratory, IL, USA, 2004 (Argonne National Laboratory, IL, USA, 2005), p.17.
  • [7] A. Gurevich and G. Ciovati, Effect of vortex hotspots on the radio-frequency surface resistance of superconductors, Phys. Rev. B 87, 054502 (2013).
  • [8] H. Padamsee, The Technology of Nb Production and Purification, in Proceedings of SRF1984, Geneva, Switzerland (JACoW, CERN, Geneva, 1984), p. 339.
  • [9] H. Padamsee, M. Hakimi, J. Kirchgessner, P. Kneisel, D. Moffat, G. Mueller, F. Palmer, L. Phillips, C. Reece, D. Rubin, R. Sundelin, and Q. S. Shu, Superconducting RF Activities at Cornell University, in Proceedings of SRF1987, Argonne National Laboratory, Illinois, USA (JACoW, CERN, Geneva, 1987), p. 13.
  • [10] E. Kako, S. Noguchi, M. Ono, K. Saito, T. Shishido, T. Fujino, Y. Funahashi, H. Inoue, M. Matsuoka, T. Higuchi, T. Suzuki and H. Umezawa, Characteristics of the Results of Measurement on 1.3 GHz High Gradient Superconducting Cavities, in Proceedings of SRF1995, Gif-sur-Yvette, France (JACoW, CERN, Geneva, 1995), p. 425.
  • [11] E. Kako, M. Bolore, Y. Boudigou, J. P. Charrier, B. Coadou, E. Jacques, M. Juillard, J. P. Poupeau, H. Safa, S. Noguchi, M. Ono, K. Saito and T. Shishido, Cavity Performances in the 1.3 GHz Saclay / KEK Nb Cavities, in Proceedings of SRF1997, Abano Terme Padova, Italy (JACoW, CERN, Geneva, 1997), p. 491.
  • [12] T. Shishido, T. Fujino, H. Inoue, E. Kako, S. Noguchi, M. Ono, K. Saito, and T. Higuchi, Test Results of the L-Band Superconducting Cavity made from Twice Melted Niobium, in Proceedings of SRF1999, Santa Fe, New Mexico, USA (JACoW, CERN, Geneva, 1999), p. 246.
  • [13] P. Kneisel, G. R. Myneni, G. Ciovati, J. Sekutowicz, and T. Carneiro, Performance of Large Grain and Single Crystal Niobium Cavities, in Proceedings of SRF2005, Cornell University, Ithaca, New York, USA (JACoW, CERN, Geneva, 2005), p. 134.
  • [14] A. Brinkmann, J. Iversen, D. Reschke, W. Singer, X. Singer, K. Twarowski, and J. Ziegler, Progress of the Test Cavity Program for the European XFEL, in Proceedings of SRF2007, Peking University, Beijing, China (JACoW, CERN, Geneva, 2007), p. 327.
  • [15] R. L. Geng, G. V. Eremeev, H. Padamsee, V. D. Shemelin, High Gradient Studies for ILC with Single Cell Re-entrant Shape and Elliptical Shape Cavities made of Fine-grain and Large-grain Niobium, in Proceedings of PAC07, Albuquerque, New Mexico, USA (JACoW, CERN, Geneva, 2007), p. 2337.
  • [16] H. Padamsee, RF Superconductivity: Science, Technology, and Applications (Wiley-VCH, Weinheim, 2009).
  • [17] W. Singer, S. Aderhold, J. Iversen, G. Kreps, A. Matheisen, X. Singer, K. Twarowski, H. Weise, M. Pekeler, F. Scholz, B. Spaniol, and E. Stiedl, Advances in Large Grain Resonators for the European XFEL, AIP Conference Proceedings 1352, 13 (2011).
  • [18] T. Kubo, Y. Ajima, H. Inoue, K. Umemori, Y. Watanabe, and M. Yamanaka, In-house Production of a Large-Grain Single-Cell Cavity at Cavity Fabrication Facility and Results of Performance Tests, in Proceedings of IPAC2014, Dresden, Germany (JACoW, CERN, Geneva, 2014), p. 2519.
  • [19] G. Ciovati, P. Dhakal, and G. R. Myneni, Superconducting radio-frequency cavities made from medium and low-purity niobium ingots, Supercond. Sci. Technol. 29, 064002 (2016).
  • [20] H. Shimizu, T. Dohmae, M. Egi, K. Enami, H. Inoue, E. Kako, G. Park, H. Sakai, K. Umemori, Y. Watanabe, S. Yamaguchi, and M. Yamanaka, Fabrication and Evaluation of Superconducting Single-Cell Cavities Manufactured Using Various Materials and Methods, IEEE Transactions on Applied Superconductivity 27, 3500714 (2017).
  • [21] K. Howard, Y.-K. Kim, D. Bafia, A. Grassellino, The Collaborative Effects of Intrinsic and Extrinsic Impurities in Low RRR SRF Cavities, in Proceedings of SRF2023, Grand Rapids, MI, USA (JACoW, CERN, Geneva, 2023), p. 162.
  • [22] A. Gurevich, Enhancement of rf breakdown field of superconductors by multilayer coating, Appl. Phys. Lett. 88, 012511 (2006).
  • [23] T. Kubo, Y. Iwashita, and T. Saeki, Radio-frequency electromagnetic field and vortex penetration in multilayered superconductors, Appl. Phys. Lett. 104, 032603 (2014).
  • [24] A. Gurevich, Maximum screening fields of superconducting multilayer structures, AIP Advance 5, 017112 (2015).
  • [25] T. Kubo, Multilayer coating for higher accelerating fields in superconducting radio-frequency cavities: a review of theoretical aspects, Supercond. Sci. Technol. 30, 023001 (2017).
  • [26] T. Kubo, 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 (2021).
  • [27] C. Z. Antoine, M. Aburas, A. Four, F. Weiss, Y. Iwashita, H. Hayano, S. Kato, T. Kubo, and T. Saeki, Optimization of tailored multilayer superconductors for RF application and protection against premature vortex penetration, Supercond. Sci. Technol. 32, 085005 (2019).
  • [28] A. Grassellino, A. Romanenko, Y. Trenikhina, M. Checchin, M. Martinello, O. S. Melnychuk, S. Chandrasekaran, D. A. Sergatskov, S. Posen, A. C. Crawford, S. Aderhold, and D. Bice, 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 (2017).
  • [29] A. J. Dolgert, S. J. Di. Bartolo, and A. T. Dorsey, Erratum: Superheating fields of superconductors: Asymptotic analysis and numerical results [Phys. Rev. B 53, 5650 (1996)], Phys. Rev. B 56, 2883 (1997).
  • [30] P. V. Christiansen, Magnetic superheating of high-κ\kappa superconductors, Solid State Commun. 7, 727 (1969).
  • [31] E. H. Brandt, Properties of the ideal Ginzburg-Landau vortex lattice, Phys. Rev. B 68, 054506 (2003).
  • [32] A. Ikushima and T. Mizusaki, Superconductivity in niobium and niobium-tantalum alloys, J. Phys. Chem. Solids 30, 873 (1969).
  • [33] L. P. Gor’kov, Microscopic Derivation of the Ginzburg-Landau Equations in the theory of superconductivity, Sov. Phys. JETP 9, 1364 (1959).
  • [34] S. Huang, T. Kubo, R. L. Geng, Dependence of trapped-flux-induced surface resistance of a large-grain Nb superconducting radio-frequency cavity on spatial temperature gradient during cooldown through T c Physical Review Accelerators and Beams 19, 082001 (2016).
  • [35] A. Gurevich, Reduction of Dissipative Nonlinear Conductivity of Superconductors by Static and Microwave Magnetic Fields, Phys. Rev. Lett. 113, 087001 (2014).
  • [36] T. Kubo and A. Gurevich, Field-dependent nonlinear surface resistance and its optimization by surface nanostructuring in superconductors, Phys. Rev. B 100, 064522 (2019).
  • [37] F. Koechlin and B. Bonin, Parametrization of the niobium thermal conductivity in the superconducting state, Supercond. Sci. Technol. 9, 453 (1996).
  • [38] A. Boucheffa and M. X. Francois, Kapitza conductance of niobium for superconducting cavities in the temperature range 1.6 K, 2.1 K, in Proceedings of SRF1995, Gif-sur-Yvette, France (JACoW, CERN, Geneva, 1995), p. 659.
  • [39] P. Dhakal, G. Ciovati, and G. R. Myneni Role of thermal resistance on the performance of superconducting radio frequency cavities Phys. Rev. Accel. Beams 20, 032003 (2017).
  • [40] T. Kubo, An Encouraging of Paternity Leave: A Physicist Who Has Become a Stay-at-Home Dad in New York, KASOKUKI, 20, 50 (2023) [Journal of the Particle Accelerator Society of Japan 20, 50 (2023)].