Skip to main content
Log in

Magnetic-field frustration and chiral-glass to insulator transition in superconducting arrays with random \(\pi \) junctions

  • Regular Article - Solid State and Materials
  • Published:
The European Physical Journal B Aims and scope Submit manuscript

Abstract

We study the effects of random \(\pi \) Josephson junctions and applied magnetic field on the quantum phase transitions of superconducting arrays, described by a two-dimensional phase-glass model. The zero-temperature phase diagram and critical behavior are obtained from path-integral Monte Carlo simulations. A superconducting chiral-glass phase occurs for sufficiently large disorder resulting from frustration effects of the \(\pi \) junctions. Unlike the superconductor–insulator transition for weak disorder, the chiral-glass to insulator transition is insensitive to additional frustration introduced by an external magnetic field corresponding to half flux quantum per plaquette. The results are consistent with recent measurements on superconducting thin films with a periodic pattern of nanoholes in an applied magnetic field and doped with magnetic impurities. The magnetic-field frustration effects for the activation energy at low doping and its disappearance at higher doping, observed experimentally, is indicative of a transition to the chiral-glass regime.

Graphic abstract

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Data Availability Statement

This manuscript has no associated data or the data will not be deposited. [Authors’ comment: All our numerical data have already been presented in the figures.]

References

  1. M.P.A. Fisher, Phys. Rev. Lett. 62, 1415 (1989)

    Article  ADS  Google Scholar 

  2. D.A. Huse, H.S. Seung, Phys. Rev. B 42, 1059 (1990)

    Article  ADS  Google Scholar 

  3. L. Bulaevskii, V. Kuzii, A. Sobyanin, JETP Lett. 25, 290 (1977)

    ADS  Google Scholar 

  4. B.I. Spivak, S.A. Kivelson, Phys. Rev. B 43, 3740 (1991)

    Article  ADS  Google Scholar 

  5. M. Sigrist, T.M. Rice, Rev. Mod. Phys. 67, 503 (1995)

    Article  ADS  Google Scholar 

  6. H. Kawamura, J. Phys. Soc. Jpn. 64, 711 (1995)

    Article  ADS  Google Scholar 

  7. H. Kawamura, M. Li, Phys. Rev. Lett. 78, 1556 (1997)

    Article  ADS  Google Scholar 

  8. E. Granato, Phys. Rev. B 69, 012503 (2004)

    Article  ADS  Google Scholar 

  9. D. Dalidovich, P. Phillips, Phys. Rev. Lett. 89, 027001 (2002)

    Article  ADS  Google Scholar 

  10. P. Phillips, D. Dalidovich, Phys. Rev. B 68, 104427 (2003a)

    Article  ADS  Google Scholar 

  11. P. Phillips, D. Dalidovich, Science 302, 243 (2003b)

    Article  ADS  Google Scholar 

  12. P.W. Phillips, Nat. Phys. 12, 206 (2016)

    Article  Google Scholar 

  13. H. Jaeger, D. Haviland, B. Orr, A. Goldman, Phys. Rev. B 40, 182 (1989)

    Article  ADS  Google Scholar 

  14. D. Das, S. Doniach, Phys. Rev. B 60, 1261 (1999)

    Article  ADS  Google Scholar 

  15. C. Christiansen, L. Hernandez, A. Goldman, Phys. Rev. Lett. 88, 037004 (2002)

    Article  ADS  Google Scholar 

  16. A. Kapitulnik, S.A. Kivelson, B. Spivak, Rev. Mod. Phys. 91, 011002 (2019)

    Article  ADS  Google Scholar 

  17. C. Yang, Y. Liu, Y. Wang, L. Feng, Q. He, J. Sun, Y. Tang, C. Wu, J. Xiong, W. Zhang et al., Science 366, 1505 (2019)

    Article  ADS  Google Scholar 

  18. E. Granato, Phys. Rev. B 102, 184503 (2020)

    Article  ADS  Google Scholar 

  19. M.D. Stewart Jr., A. Yin, J.M. Xu, J.M. Valles Jr., Science 318, 1273 (2007)

    Article  ADS  Google Scholar 

  20. E. Granato, Eur. Phys. J. B 89, 68 (2016a)

    Article  MathSciNet  ADS  Google Scholar 

  21. E. Granato, Phys. Rev. B 94, 060504(R) (2016b)

    Article  ADS  Google Scholar 

  22. X. Zhang, J.C. Joy, C. Wu, J.H. Kim, J. Xu, J.M. Valles Jr., Phys. Rev. Lett. 122, 157002 (2019)

    Article  ADS  Google Scholar 

  23. V. Ryazanov, V. Oboznov, A.Y. Rusanov, A. Veretennikov, A.A. Golubov, J. Aarts, Phys. Rev. Lett. 86, 2427 (2001)

    Article  ADS  Google Scholar 

  24. R.M. Bradley, S. Doniach, Phys. Rev. B 30, 1138 (1984)

    Article  ADS  Google Scholar 

  25. R. Fazio, H. van der Zant, Phys. Rep. 355, 235 (2001)

    Article  ADS  Google Scholar 

  26. J. Villain, J. Phys. C. Solid State Phys. 10, 4793 (1977)

    Article  ADS  Google Scholar 

  27. H. Kawamura, M. Tanemura, J. Phys. Soc. Jpn. 60, 608 (1991)

    Article  ADS  Google Scholar 

  28. E. Granato, Phys. Rev. B 58, 11161 (1998)

    Article  ADS  Google Scholar 

  29. J.M. Kosterlitz, N. Akino, Phys. Rev. Lett. 82, 4094 (1999)

    Article  ADS  Google Scholar 

  30. E. Granato, Phys. Rev. B 96, 184510 (2017)

    Article  ADS  Google Scholar 

  31. E. Granato, J.M. Kosterlitz, Phys. Rev. Lett. 65, 1267 (1990)

    Article  ADS  Google Scholar 

  32. M.-C. Cha, S.M. Girvin, Phys. Rev. B 49, 9794 (1994)

    Article  ADS  Google Scholar 

  33. S. Sachdev, Quantum Phase Transitions (Cambridge University Press, Cambridge, 2000)

    Book  MATH  Google Scholar 

  34. M. Wallin, E.S. Sorensen, S.M. Girvin, A.P. Young, Phys. Rev. B 49, 12115 (1994)

    Article  ADS  Google Scholar 

  35. S.L. Sondhi, M. Girvin, J. Carini, D. Sahar, Rev. Mod. Phys. 69, 315 (1997)

    Article  ADS  Google Scholar 

  36. K. Hukushima, K. Nemoto, J. Phys. Soc. Jpn. 65, 1604 (1996)

    Article  ADS  Google Scholar 

  37. H. Rieger, A.P. Young, Phys. Rev. Lett. 72, 4141 (1994)

    Article  ADS  Google Scholar 

  38. L.W. Lee, A.P. Young, Phys. Rev. Lett. 90, 227203 (2003)

    Article  ADS  Google Scholar 

  39. H. Nguyen, S. Hollen, J. Shainline, J. Xu, J. Valles, Sci. Rep. 6, 38166 (2016)

    Article  ADS  Google Scholar 

  40. E. Granato, Physica B Condens. Matter 536, 442 (2018)

    Article  ADS  Google Scholar 

  41. H.G. Ballesteros, A. Cruz, L.A. Fenández, V. Martin-Mayor, J. Pech, J.J. Ruiz-Lorenzo, A. Tarancón, P. Téllez, C.L. Ullod, C. Ungil, Phys. Rev. B 62, 14237 (2000)

    Article  ADS  Google Scholar 

  42. R.N. Bhatt, A.P. Young, Phys. Rev. B 37, 5606 (1988)

    Article  ADS  Google Scholar 

  43. M.P.A. Fisher, G. Grinstein, S.M. Girvin, Phys. Rev. Lett. 64, 587 (1990)

    Article  ADS  Google Scholar 

  44. M.-C. Cha, M.P.A. Fisher, S.M. Girvin, M. Wallin, A.P. Young, Phys. Rev. B 44, 6883 (1991)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

The author thanks J. M. Valles, Jr. and X. Zhang for helpful discussions. This work was supported by Fundação de Amparo à Pesquisa do Estado de São Paulo-FAPESP (Grant no. 18/19586-9), CNPq (Conselho Nacional de Desenvolvimento Científico e Tecnológico) and computer facilities from CENAPAD-SP.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Enzo Granato.

Appendix

Appendix

Here we describe the results for critical behavior corresponding to the CG-I transition at \(x=0.2\) for \(f=1/2\), obtained from the scaling behavior of the correlation length and phase stiffness. We also obtain an estimate of the electrical conductivity at the transition.

The finite-size correlation length \(\xi \) can be defined as [41]

$$\begin{aligned} \xi (L,g) = \frac{1}{2\sin (k_0/2)}[S(0)/S(k_0) - 1]^{1/2}, \end{aligned}$$
(8)

where S(k) is the Fourier transform of the correlation function C(r) for the relevant order parameter, and \(k_0\) is the smallest nonzero wave vector. The correlation length \(\xi _\tau (L,g) \) in the imaginary-time direction is obtained by the analogous expression. For the phase-coherence transition, the phase correlation length \(\xi _\mathrm{G}\) is obtained from the correlation function defined in terms of the overlap order parameter [37, 42] of phase variables \(q_{\tau ,j}=\exp ( i(\theta ^1_{\tau , j} - \theta ^2_{\tau ,j}) )\), where 1 and 2 label two different copies of the system with the same coupling parameters. Likewise, the chiral correlation length \(\xi ^c_\mathrm{G}\) is obtained in terms of the overlap of the chiral variables defined in Eq. (3), as \(q^c_{\tau ,p}=\chi ^1_{\tau , p} \ \chi ^2_{\tau ,p} \).

For a continuous phase transition, \(\xi (L,g)\) should satisfy the scaling form

$$\begin{aligned} \xi /L = A(L^{1/\nu } \delta g ), \end{aligned}$$
(9)

where A(x) is a scaling function, \(\delta g = g-g_\mathrm{c}\) and \(\nu \) is the correlation-length critical exponent. It implies that data for \( \xi (L,g)/L \) as a function of g, for different system sizes L, should cross at the critical coupling \(g_\mathrm{c}\). In addition, a scaling plot in the form \(\xi /L \times L^{1/\nu } \delta g\) sufficiently close to \(g_\mathrm{c}\) should collapse the data on to the same curve.

In Fig. 7a, we show the behavior of the correlation length \(\xi _{\mathrm{G},\tau } \) for the phase-glass model with \(f=1/2\) at \(x=0.2\). The dynamic exponent is set to \(z=1.2\), the same value estimated for \(f=0\) since it gives consistent results for the scaling behavior [18]. The curves for \(\xi _{\mathrm{G},\tau }/L \) as a function of g for different system sizes cross approximately at a common point \(g_\mathrm{c}\), providing evidence of a continuous phase-coherence transition. Figure. 7b, shows the scaling plot according to Eq. (9), which gives estimates for \(g_\mathrm{c}\) and \(\nu \).

Fig. 7
figure 7

a Phase correlation length \(\xi _{\mathrm{G},\tau } \) near the CG-I transition at \(x=0.2\) and different system sizes L. b Corresponding scaling plot of the data near the transition with \(g_\mathrm{c}=1.50 \), \(\nu = 1.4 \)

The corresponding results for the chiral correlation length \(\xi ^c_{\mathrm{G},\tau } \) are shown in Fig. 8. Although the data are much noisier, the behavior is very similar to the phase variables with approximately the same \(\nu \) but a slightly smaller value for \(g_\mathrm{c}\).

Fig. 8
figure 8

a Chiral correlation length \(\xi ^c_{\mathrm{G},\tau } \) near the CG-I transition at \(x=0.2\) and different system sizes L . b Corresponding scaling plot of the data near the transition with \(g_\mathrm{c}=1.42 \), \(\nu = 1.1 \)

The scaling behavior of the phase correlation length \(\xi _{\mathrm{G},\tau }\) in Fig. 7 already indicates that the CG-I transition corresponds to a phase-coherence transition and therefore the chiral-glass phase is expected to be superconducting. To further investigate this property, we consider the scaling behavior of the phase stiffness \(\gamma \), a measure of the free energy cost to impose an infinitesimal phase twist along a given direction. In the imaginary time direction, \(\gamma _\tau \), corresponds to the compressibility of the bosonic system. It is given by [32, 34]

$$\begin{aligned} \gamma _\tau =\frac{1}{L^2 L_\tau g^2}[g<\epsilon _\tau> -< I_\tau ^2> + < I_\tau >^2]_d, \end{aligned}$$
(10)

where \(\epsilon _\tau = \sum _{\tau ,i} \cos (\theta _{\tau ,i} -\theta _{\tau +1,i})\) and \(I_\tau = \sum _{\tau ,i}\sin (\theta _{\tau ,i} -\theta _{\tau +1,i})\). In Eq. (10), \(< \ldots>\) represents a MC average for a fixed disorder configuration, and \([ \ldots ]_d\) represents an average over different disorder configurations. In the superconducting phase, \(\gamma \) should be finite, reflecting the existence of phase coherence, while in the insulating phase it should vanish in the thermodynamic limit. For a continuous phase transition, \(\gamma _\tau \) should satisfy the finite-size scaling form [34]

$$\begin{aligned} \gamma _\tau L^{2-z} = A(L^{1/\nu } \delta g). \end{aligned}$$
(11)

This scaling form implies that data for \( \gamma _\tau L^{2-z} \) as a function of g, for different system sizes L, should cross at the critical coupling \(g_\mathrm{c}\). In Fig. 9, we show the behavior of the phase stiffness in the time direction. Curves for different system sizes cross approximately at the same critical coupling \(g_\mathrm{c}\) estimated from the phase and chiral correlation lengths in Figs. 7 and 8. This behavior lends further support for a superconducting chiral-glass phase for \(f=1/2\), as also observed previously [18] for \(f=0\). From these different estimates, we obtain for the superconducting CG-I transition \(g_\mathrm{c}=1.49(6) \) and \(\nu =1.3(2) \).

Fig. 9
figure 9

a Phase stiffness \(\gamma _\tau \) near the CG-I transition at \(x=0.2\) and different system sizes L. b Corresponding scaling plot with \(g_\mathrm{c}= 1.535 \) and \(\nu =1.3 \)

Fig. 10
figure 10

Scaling plot of conductivity \(\sigma (iw_n)\) at the critical coupling \(g_\mathrm{c}\) for \(x=0.2\), with \(g_\mathrm{c}=1.535 \) and \(\alpha =0.06 \). The universal conductivity is given by the intercept with the \(x_o=0\) dashed line, giving \(\frac{\sigma ^*}{\sigma _Q}= 0.5 \)

Another critical quantity of interest is the value of the electrical conductivity \(\sigma ^*\) at the transition, in units of the quantum of conductance \(\sigma _Q=(2 e)^2/h\). It should be universal [43], but it could depend on the universality class of the transition. The universal conductivity can be determined from the Kubo formula

$$\begin{aligned} \sigma = 2 \pi \sigma _Q \lim _{w_n\rightarrow 0} \frac{\gamma (i w_n)}{w_n}, \end{aligned}$$
(12)

where \(\gamma (i w_n)\) is the frequency dependent phase stiffness in the spatial direction, evaluated at the finite frequency \(w_n=2 \pi n /L_\tau \), with n an integer. \(\gamma (i w_n)\) in the \({\hat{x}}\) direction is given by

$$\begin{aligned} \gamma (i\omega _n)= & {} \frac{1}{L^2 L_\tau g^2} [ g< \epsilon _x> -< |I(i w_n)|^2> \\&+ < |I(i w_n)| >^2]_d, \end{aligned}$$
(13)

where

$$\begin{aligned} \epsilon _x= & {} \sum _{\tau ,j} e_{j,j+{\hat{x}}}\cos (\Delta _x \theta _{\tau ,j}) ,\\I(i w_n)= & {} \sum _{\tau ,j} e_{j,j+{\hat{x}}} \sin (\Delta _x \theta _{\tau ,j}) e^{i w_n \tau }, \end{aligned}$$
(14)

and \(\Delta _x \theta _{\tau ,j}= \theta _{\tau ,j}-\theta _{\tau ,j+{\hat{x}}}\). Following the finite-size scaling method of Cha et al. [32, 44], at the transition \(\gamma (i w_n)\) vanishes linearly with frequency and \(\sigma \) assumes a universal value \(\sigma ^*\), which can be extracted from its frequency and finite-size dependence as [32]

$$\begin{aligned} \frac{\sigma (iw_n)}{\sigma _Q} = \frac{\sigma *}{\sigma _Q} - c \left( \frac{w_n}{2 \pi } - \alpha \frac{2 \pi }{w_n L_\tau }\right) \cdots \end{aligned}$$
(15)

The parameter \(\alpha \) is determined from the best data collapse of the frequency-dependent curves for different systems sizes in a plot of \(\frac{\sigma (iw_n)}{\sigma _Q}\) versus \(x_o=(\frac{w_n}{2\pi } - \alpha \frac{2\pi }{w_n L_\tau })\). The universal conductivity is obtained from the intercept of these curves with the line \(x_o=0\). From this scaling behavior, shown in Fig. 10, we obtain \(\sigma ^*/\sigma _Q = 0.53(4)\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Granato, E. Magnetic-field frustration and chiral-glass to insulator transition in superconducting arrays with random \(\pi \) junctions. Eur. Phys. J. B 94, 143 (2021). https://doi.org/10.1140/epjb/s10051-021-00162-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epjb/s10051-021-00162-3

Navigation