Next Article in Journal
Study of the Hibridation of Ablation Casting and Laser Wire Metal Deposition for Aluminum Alloy 5356
Previous Article in Journal
Performance of Pico-Second Laser-Designed Silicon/Gold Composite Nanoparticles Affected by Precision of Focus Position
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Defect Pinning and Critical Current of Magnetic Vortex Cluster in Mesoscopic Type-1.5 Superconductors

1
School of Nuclear Science and Engineering, North China Electric Power University, Beijing 102206, China
2
Department of Mathematics and Physics, North China Electric Power University, Beijing 102206, China
*
Author to whom correspondence should be addressed.
Crystals 2025, 15(2), 133; https://doi.org/10.3390/cryst15020133
Submission received: 30 December 2024 / Revised: 20 January 2025 / Accepted: 22 January 2025 / Published: 25 January 2025
(This article belongs to the Section Inorganic Crystalline Materials)

Abstract

:
Based on two-band time-dependent Ginzburg–Landau theory, we study the electromagnetic properties of mesoscopic type-1.5 superconductors with different defect configurations. We perform numerical simulations with the finite element method, and give direct evidence for the existence of a vortex cluster phase in the presence of nonmagnetic impurity. In addition, we also investigate the depinning critical current of the magnetic vortex cluster induced by the isotropic or anisotropic defect structure under the external current. Our theoretical results thus indicate that the diversity of impurity deposition has a significant influence on the semi-Meissner state in type-1.5 superconductors.

1. Introduction

Over the past two decades, two-band superconductivity has become an important research subject in condensed matter physics. This field started from the discovery of superconductivity in MgB 2 [1], where the existence of two distinct superconducting gaps revealed the complexity of Fermi surface topology in the relevant system. Since then, extensive theoretical and experimental studies have been performed to provide novel insights into unconventional superconducting pairing mechanisms and physical properties in these materials. For example, multi-gap superconductivity signals a new pathway to achieve more superconducting pairing modes, which can induce phase competition or coexistence between multiple bands by adjusting the external magnetic field or impurity distribution. Furthermore, magnetic vortex behavior can be optimized through rational design of multi-band structures and its interaction with impurities can improve the overall performance of superconducting devices [2,3].
As we know, each condensate in two-band superconductors is predicted to support vortex excitation with fractional quantum flux [4,5]. Due to interband Josephson coupling, the vortices from different condensates are bound together with the string interaction and their normal cores will be locked to form a composite vortex with the standard integer quantum flux. Therefore, the vortex physics in two-band systems is influenced by the coherence lengths ξ 1 and ξ 2 as well as the magnetic field penetration depth λ . When the particular condition ξ 1 < 2 λ < ξ 2 is satisfied, a new superconducting state may be exhibited that combines characteristics of both type-1 and type-2 superconductors. This so-called semi-Meissner phase or vortex cluster phase is formed due to the interaction of long-range attraction and short-range repulsion between composite vortex excitations [6,7]. The existence of this novel vortex pattern was first visualized by Bitter decorations on high-quality MgB 2 single crystals in 2009 [8]. Thereafter, zero-field muon spin experiments have also revealed the presence of this type-1.5 superconducting state in unconventional superconductors Sr 2 RuO 4 [9,10] and LaPt 3 Si [11,12].
Recent seminal works of the Berger–Sardella group and Maksimova–Kashurnikov group on vortex dynamics and the critical current of mesoscopic superconductors have provided important inspiration for our theoretical research [13,14,15,16,17,18,19]. In the present paper, we study the electromagnetic effect of mesoscopic type-1.5 superconductors with different impurity distributions based on time-dependent Ginzburg–Landau (TDGL) theory. With the COMSOL Multiphysics software and the finite element method, our results directly show the crossover of this mesoscopic system from the diamagnetic Meissner state to the vortex cluster phase, and ultimately to the Abrikosov lattice phase. We also investigate the effects of isotropic and anisotropic defect structures on the pattern of magnetic vortex distribution. Furthermore, we discuss the depinning critical current of the magnetic vortex cluster induced by different pinning configurations under the external current. All of our theoretical results indicate that the diversity of impurity deposition has a significant influence on the collective behaviors of magnetic vortices in type-1.5 superconducting systems.
The rest of this article is organized as follows. In Section 2, we introduce the two-band TDGL theory and apply this formalism to type-1.5 superconductors. In Section 3, we describe the procedure of numerical simulations based on the finite element method. Then, in Section 4, we discuss the impurity effect and critical current of vortex clusters in the relevant mesoscopic systems. Finally, Section 5 offers a conclusion of the paper.

2. Theoretical Formalism

The simplest GL free energy functional of two-gap superconductors can be written as follows [20,21,22,23,24]:
F = i 1 2 m i Π Ψ i 2 α i | Ψ i | 2 + β i 2 | Ψ i | 4 + γ 1 | Ψ 1 | 2 | Ψ 2 | 2 + γ 2 2 ( Ψ 1 * Ψ 2 ) 2 + c . c . + 1 2 γ 3 ( Π Ψ 1 ) * · Π Ψ 2 + c . c . + B 2 8 π .
Here, Ψ i ( i = 1 , 2 ) represents the superconducting order parameter and m i is the effective mass for each band. γ 1 and γ 2 represent the interband quartic interactions, and γ 3 describes the gradient interaction between these two bands. The coefficient α i is a function of temperature, while β i is independent of temperature. In the presence of impurity, the parameters α 1 and α 2 can be approximately expressed as α i = α i 0 f ( r ) . Here, we introduce a function f ( r ) between + 1 and 1 to model the defect sites, which will deplete the superconducting state at specific positions [25,26]. We also define the covariant derivative operator Π = i 2 e A / c with the vector potential A and the magnetic field B = × A .
If the superconductor is driven out of equilibrium, the order parameter should relax back to its equilibrium value. It is well known that this deviation of superconducting materials can be conveniently described by TDGL theories. The single-band TDGL equations were first proposed by Schmid [27] and are derived from the microscopic BCS theory by Gor’kov and Éliashberg [28]. The extension of these TDGL equations to multi-component superconducting systems can be written as follows [29,30]:
Γ i Ψ i t = δ F δ Ψ i * and σ n A t = δ F δ A
where Γ i is the relaxation time of order parameters and σ n represents the electrical conductivity of the normal sample in the two-band case. Therefore, minimization of the free energy F with respect to Ψ i and A leads to the following dimensionless TDGL equations in the zero-electrostatic potential gauge:
Γ 1 Ψ 1 t = Π 2 Ψ 1 + m 1 γ 3 Π 2 Ψ 2 f ( r ) | Ψ 1 | 2 γ 1 β 1 | Ψ 2 | 2 Ψ 1 + γ 2 β 1 Ψ 2 2 Ψ 1 * ,
Γ 2 Ψ 2 t = m 1 m 2 Π 2 Ψ 2 + m 1 γ 3 Π 2 Ψ 1 α 20 α 10 f ( r ) β 2 β 1 | Ψ 2 | 2 γ 1 β 1 | Ψ 1 | 2 Ψ 2 + γ 2 β 1 Ψ 1 2 Ψ 2 *
and
A t = κ 1 2 × × A j s .
with the supercurrent
j s = Re ( Ψ 1 * Π Ψ 1 ) + m 1 m 2 Re ( Ψ 2 * Π Ψ 2 ) + m 1 γ 3 Re ( Ψ 1 * Π Ψ 2 + Ψ 2 * Π Ψ 1 ) .
Here, in the clean limit with the nonmagnetic impurity function f = 1 , we at first introduce the coherence length ξ i 2 = 2 / 2 m i α i 0 , the London penetration depth λ 2 = λ 1 2 + λ 2 2 with λ i 2 = 4 π e 2 Ψ i 0 2 / ( m i c 2 ) and Ψ i 0 = α i 0 / β i , and the GL parameter κ 1 = λ 1 / ξ 1 . We then take the coordinate r in units of ξ 1 , the time t in units of t 0 = m 1 σ n / 4 e 2 Ψ 10 2 , Γ i in units of α 10 t 0 , and the order parameter Ψ i in units of Ψ 10 . We also set the magnetic field B in units of H 0 = Φ 0 / 2 π ξ 1 2 with the flux quantum Φ 0 = π c / e and the vector potential A in units of A 0 = H 0 ξ 1 .
Following ref. [6], multi-component systems allow a type of superconductivity that is distinct from type-1 or type-2 superconductors. With the condition ξ 1 < 2 λ < ξ 2 , the type-1.5 superconducting state will originate from a peculiar vortex interaction that exhibits short-range repulsion and long-range attraction characteristics. The short-range repulsion prevents adjacent vortices from overlapping, while the long-range attraction facilitates the clustering of composite vortices. Consequently, this state is different from type-1 superconductors, which completely repel magnetic flux, and type-2 superconductors, which allow considerable magnetic flux penetration and the formation of a vortex lattice. In the ideal sample, the constraint mentioned above can be specifically expressed as
1 2 1 + m 1 m 2 α 20 α 10 β 1 β 2 < κ 1 < 1 2 m 1 m 2 α 10 α 20 + m 1 m 2 2 β 1 β 2 .
In this circumstance, the magnetic composite vortices will form vortex clusters and coexist with domains of the two-component Meissner state in the framework of the GL theory. Furthermore, note that for the energy to be positively defined, the kinetic terms should give the relation m 1 m 2 γ 3 2 < 0 . Also, for the free energy functional to be bounded from below, the fourth-order terms in the condensates satisfy the constraint β 1 β 2 ( γ 1 + γ 2 ) 2 > 0 [31].
In order to numerically solve Equations (3)–(5), we need to specify appropriate boundary conditions of the superconducting sample. We use the following superconductor–insulator (or vacuum) boundary conditions [32,33,34]:
Ψ i · n = 0 , A · n = 0 and × A = H e
where n is the outward unit vector normal to the boundary and the external applied magnetic field is set as H e = H e z ^ . The first two conditions simply indicate that any current passing through the interface between a superconducting domain and vacuum/insulator would be nonphysical. The third equation represents the continuity of magnetic field across the boundary. The partial differential Equations (3)–(5) will be solved numerically for the mesoscopic geometry in the two-dimensional space. The initial conditions at t = 0 are taken as | Ψ i | = 1 and A = ( 0 , 0 ) on the x y -plane, corresponding to the Meissner state and zero magnetic field inside the superconductor.
In this investigation, we also study the vortex dynamics and critical current in the presence of impurity by applying an external current density j n flowing through the sample in the x ^ -direction. Following ref. [35], the current will be modeled by introducing an additional magnetic field Δ B in the numerical calculations. In the chosen 15 ξ 1 × 15 ξ 1 geometry, this extra field is directed along the z ^ -axis and is constructed so that the total magnitude of the applied field is H e Δ H at y = 0 and H e + Δ H at y = 15 ξ 1 . From Equation (5), we have j n = A / t = κ 1 2 × ( Δ B ) , which can be explicitly written as j n = j n x ^ with j n = 2 κ 1 2 Δ H / ( 15 ξ 1 ) . Meanwhile, the electrical field is given by E = ( 1 / c ) A / t in the Weyl gauge. Then, the voltage V in the x ^ -direction is V = 0 15 ξ 1 E x d x = ( 1 / c ) 0 15 ξ 1 ( A x / t ) d x for the system. In the simulations, we set the normal current j n in units of j 0 = H 0 / ξ 1 and the sample voltage V in units of V 0 = H 0 ξ 1 2 / ( c t 0 ) .
At this point, we would like to briefly discuss the main limitation of the TDGL equations in multi-band superconducting systems. This phenomenological theory assumes that thermal excitations in the superconductor are slow-moving and in equilibrium with the local values of energy gaps and external fields. This occurs only in superconductors in which the characteristic interaction time between phonons and thermal excitations is faster than the characteristic time for normal to superfluid conversion [36].

3. Finite Element Method and Numerical Computations

Based on the COMSOL Multiphysics software platform [37], we will describe the procedure of the numerical simulations on the TDGL equations in this section. We first split the order parameters into the real and imaginary parts, i.e., Ψ 1 = u 1 + i u 2 and Ψ 2 = u 3 + i u 4 . The magnetic potential is also written in component form as A = ( u 5 , u 6 ) . In order to implement the boundary conditions, we will introduce an auxiliary variable u 7 ( x , y , t ) for reasons explained below. In the procedure of simulations, we set Γ 1 = Γ 2 = 5 , γ 1 = 0 and m 1 = 2 m 2 = γ 3 . To stabilize the semi-Meissner state, we also take α 10 = α 20 and β 1 = β 2 = 2 γ 2 in the calculations.
In this way, we can transform the TDGL equations into the general form of partial differential equations in this software package as follows:
k μ j k u k t + l l ν j l = η j .
Here, we have j , k = 1 , 2 , , 7 , l = 1 , 2 and ( 1 , 2 ) = ( x , y ) . The 7 × 7 matrix μ j k and the 7 × 2 column vector ν j l take the form
μ j k = 5 0 0 0 0 0 0 0 5 0 0 0 0 0 0 0 5 0 0 0 0 0 0 0 5 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0
and
ν j l = u 1 x u 3 x u 1 y u 3 y u 2 x u 4 x u 2 y u 4 y u 1 x 2 u 3 x u 1 y 2 u 3 y u 2 x 2 u 4 x u 2 y 2 u 4 y 0 κ 1 2 u 6 x u 5 y H e κ 1 2 u 5 y u 6 x + H e 0 u 5 u 6 .
Here, we note that the subscript x or y denotes the partial derivative with respect to the corresponding variable. Meanwhile, the driving force η j contains all other terms in the TDGL equations except the left-hand side of Equation (9), and detailed calculations will give all the components explicitly as
η 1 = f ( r ) u 1 1 2 2 u 1 2 + 2 u 2 2 + u 3 2 u 4 2 + 2 u 5 2 + 2 u 6 2 u 1 + u 5 x + u 6 y u 2 u 5 2 + u 6 2 u 3 + u 5 x + u 6 y u 4 + 2 u 2 x + u 4 x u 5 + 2 u 2 y + u 4 y u 6 u 2 u 3 u 4 ,
η 2 = f ( r ) u 2 1 2 2 u 1 2 + 2 u 2 2 u 3 2 + u 4 2 + 2 u 5 2 + 2 u 6 2 u 2 u 5 x + u 6 y u 1 u 5 x + u 6 y u 3 u 5 2 + u 6 2 u 4 2 u 1 x + u 3 x u 5 2 u 1 y + u 3 y u 6 u 1 u 3 u 4 ,
η 3 = f ( r ) u 3 1 2 u 1 2 u 2 2 + 2 u 3 2 + 2 u 4 2 + 4 u 5 2 + 4 u 6 2 u 3 u 5 2 + u 6 2 u 1 + u 5 x + u 6 y u 2 + 2 u 5 x + u 6 y u 4 + 2 u 2 x + 2 u 4 x u 5 + 2 u 2 y + 2 u 4 y u 6 u 1 u 2 u 4 ,
η 4 = f ( r ) u 4 1 2 u 2 2 u 1 2 + 2 u 3 2 + 2 u 4 2 + 4 u 5 2 + 4 u 6 2 u 4 u 5 x + u 6 y u 1 u 5 2 + u 6 2 u 2 2 u 5 x + u 6 y u 3 2 u 1 x + 2 u 3 x u 5 2 u 1 y + 2 u 3 y u 6 u 1 u 2 u 3 ,
η 5 = u 2 x + u 4 x u 1 u 1 x + u 3 x u 2 + u 2 x + 2 u 4 x u 3 u 1 x + 2 u 3 x u 4 u 1 2 + u 2 2 + 2 u 3 2 + 2 u 4 2 + 2 u 1 u 3 + 2 u 2 u 4 u 5 ,
η 6 = u 2 y + u 4 y u 1 u 1 y + u 3 y u 2 + u 2 y + 2 u 4 y u 3 u 1 y + 2 u 3 y u 4 u 1 2 + u 2 2 + 2 u 3 2 + 2 u 4 2 + 2 u 1 u 3 + 2 u 2 u 4 u 6 ,
η 7 = u 5 x + u 6 y + u 7 .
Now, we can examine the boundary conditions under this formalism. With the normal vector n = ( n 1 , n 2 ) and the column vector ν j l , the boundary conditions in Equation (8) can be simply cast into the compact form
l n l ν j l = 0
which is best suited to the COMSOL Multiphysics simulations. We also note that from the last equation ( j = 7 ) in (9), our manipulations will give a trivial solution u 7 = 0 for this auxiliary variable, ensuring the self-consistency of our problem.
COMSOL Multiphysics is a versatile and advanced simulation platform that is designed to tackle complex engineering and scientific problems. Its core principle is to numerically solve partial differential equations based on the finite element method [38,39,40]. The process begins with discretizing the computational domain and subdividing the lattice cell into small subregions called elements. Triangular elements are preferred due to their flexibility in handling complex and irregular shapes. The process will transform the continuous domain into a finite element mesh and enable precise numerical computations. Following this step, a function space typically composed of piecewise continuous polynomials is constructed to ensure smoothness across element boundaries. Subsequently, Lagrangian shape functions are selected as basis functions for their ability to achieve high computational accuracy and numerical stability [41]. Finally, the software employs an implicit solver that typically incorporates consistent initialization of the backward Euler method to ensure significant stability for time-dependent simulations. In our numerical computations, we take the time step Δ t = 0.5 t 0 and the relative tolerance 10 8 to control the convergence of the transient calculations for our system.

4. Results and Discussions

In this section, we will discuss the effect of impurity on the patterns of magnetic vortex distribution and the critical current for vortex cluster depinning in a 15 ξ × 15 ξ type-1.5 superconductor. Following Refs. [25,26], we have chosen the impurity function f to take the phenomenological form
f ( r ) = 0.5 , if | r r 0 | < R cos p θ + π 4 1 , otherwise .
It is easy to see that the impurity is centered at r 0 = ( x 0 , y 0 ) , and its shape depends on the angle θ and different integer values of p. This means that the defect sites can be isotropic with radius R when p = 0 and anisotropic with 2p-fold symmetry at p 0 . We take R = 0.5 ξ 1 for each pinning state in the simulations.
Meanwhile, based on Inequality (7), we expect to discover the semi-Meissner state within 1.22 < κ 1 < 1.73 in the clean limit. According to Ref. [42], the introduction of impurity into the material will enhance the effective magnetic penetration depth, thus leading to a larger effective GL parameter. Therefore, this allows the magnetic vortex phases to be observed at slightly smaller κ 1 values in the presence of impurity in type-1.5 superconductors. We set the value κ 1 = 1.35 for the following numerical calculations.
There are three main methods of trapping the magnetization in common use: zero field cooling (ZFC), field cooling (FC), and pulsed field magnetization (PFM). In our investigation, a specific numerical technique is applied to model the complete process of FC magnetization. This scheme is based on the two-dimensional H -formulation and the commercial software package COMSOL Multiphysics 4.3a [37]. The ac/dc module of COMSOL is employed for the electromagnetic analysis and the heat transfer module is used for the thermal analysis, which are coupled together to carry out the calculations of magnetic flux density in the FC regime. In the FC process, the magnetic field is applied to the superconducting sample at T > T c , which is then cooled below T c . We thus assume a smooth transition from the normal state to the superconducting state to avoid non-convergence around the transition temperature.
To verify the availability of the method, we first take the impurity function with N = 1 and p = 0 and insert this pinning site at the center of the superconducting square. We then plot the magnetic flux density B z = u 6 x u 5 y in units of H 0 (a–c) and the order parameter of the first condensate Ψ 1 = u 1 2 + u 2 2 in units of Ψ 10 (d–f) at t = 10 4 t 0 in Figure 1. With the external magnetic field H e taken as 0.25 H 0 , 0.55 H 0 , and 0.85 H 0 sequentially, we can clearly observe the crossover of this type-1.5 system from the perfect diamagnetism state to the vortex cluster phase, and ultimately to the Abrikosov vortex lattice. Our numerical simulations also show that the cluster phase presents the vortex pattern with octagonal symmetry and appears in the region of 0.45 H 0 < H e < 0.72 H 0 . Moreover, it is easy to see that the isotropic defect induces localized distortions of the flux lattice without breaking the C 4 rotational symmetry of the superconducting square.
Furthermore, we perform the simulations on this mesoscopic type-1.5 system with an anisotropic defect. For N = 1 and p = 2 , we still take the impurity site at the center of the superconducting square and plot the B z and Ψ 1 at t = 10 4 t 0 in Figure 2. By setting H e as 0.55 H 0 and 0.85 H 0 , we can observe the novel vortex cluster with C 4 (not C 8 in isotropic impurity case) symmetry shown in Figure 2b,e and the distorted flux lattice in Figure 2c,f, respectively. Our numerical data also indicate that the vortex cluster phase exists in the regime 0.35 H 0 < H e < 0.60 H 0 for this mesoscopic superconductor. As we see, the anisotropic defect occupies the smaller normal area compared with its isotropic counterpart, leading to a smaller H e for the emergence of the magnetic vortex phases.
At this stage, we also perform numerical computations on the depinning critical current of the magnetic vortex cluster in the presence of an isotropic or anisotropic defect structure. For H e = 0.55 H 0 and κ 1 = 1.35 , we plot the sample voltage V as a function of external current j n for different impurity configurations in Figure 3. From Figure 3, we can see that the system remains in the stable vortex cluster phase with zero voltage across the boundary until the threshold current is reached at j c 1 = 0.3 j 0 (the isotropic impurity case) and j c 2 = 0.15 j 0 (the anisotropic impurity case), respectively. With the further increase in current, the voltage increases from zero due to the energy dissipation of vortex motions in the depinning phase. In addition, the numerical result j c 1 > j c 2 is in agreement with our theoretical expectation that the isotropic defect possesses a larger effective normal area and will exhibit a stronger pinning force on vortices than the anisotropic counterpart.

5. Conclusions

Based on two-band TDGL theory, we investigated the impurity effect on vortex collective behaviors in mesoscopic type-1.5 superconductors. With the finite element method, our numerical results show a clear signature for the existence of a vortex cluster phase in the presence of nonmagnetic impurity in the studied system. We also numerically calculated the depinning critical current of the magnetic vortex cluster induced by an isotropic or anisotropic defect structure under the external current. We hope that our theoretical results will inspire further research on better understanding novel vortex patterns and phase transitions in two-band superconductors.

Author Contributions

Conceptualization, G.W., T.H., J.L., J.Z. and H.H.; methodology, G.W., T.H., J.L., J.Z. and H.H.; software, G.W.; validation, G.W., T.H., J.L., J.Z. and H.H.; formal analysis, G.W., T.H., J.L., J.Z. and H.H.; investigation, G.W., T.H., J.L., J.Z. and H.H.; resources, H.H.; data curation, G.W., T.H., J.L., J.Z. and H.H.; writing—original draft preparation, G.W.; writing—review and editing, G.W., T.H., J.L., J.Z. and H.H.; visualization, G.W.; supervision, T.H., J.L., J.Z. and H.H.; project administration, H.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Nagamatsu, J.; Nakagawa, N.; Muranaka, T.; Zenitani, Y.; Akimitsu, J. Superconductivity at 39 K in magnesium diboride. Nature 2001, 410, 63. [Google Scholar] [CrossRef] [PubMed]
  2. Zehetmayer, M. A review of two-band superconductivity: Materials and effects on the thermodynamic and reversible mixed-state properties. Supercond. Sci. Technol. 2013, 26, 043001. [Google Scholar] [CrossRef]
  3. Salamone, T.; Hugdal, H.G.; Jacobsen, S.H.; Amundsen, M. High magnetic field superconductivity in a two-band superconductor. Phys. Rev. B 2023, 107, 174516. [Google Scholar] [CrossRef]
  4. Tanaka, Y. Soliton in two-band superconductor. Phys. Rev. Lett. 2002, 88, 017002. [Google Scholar] [CrossRef]
  5. Babaev, E. Vortices with fractional flux in two-gap superconductors and in extended Faddeev model. Phys. Rev. Lett. 2002, 89, 067001. [Google Scholar] [CrossRef]
  6. Babaev, E.; Speight, M. Semi-Meissner state and neither type-I nor type-II superconductivity in multicomponent superconductors. Phys. Rev. B 2005, 72, 180502. [Google Scholar] [CrossRef]
  7. Carlström, J.; Garaud, J.; Babaev, E. Length scales, collective modes, and type-1.5 regimes in three-band superconductors. Phys. Rev. B 2011, 84, 134518. [Google Scholar] [CrossRef]
  8. Moshchalkov, V.; Menghini, M.; Nishio, T.; Chen, Q.H.; Silhanek, A.V.; Dao, V.H.; Chibotaru, L.F.; Zhigadlo, N.D.; Karpinski, J. Type-1.5 superconductivity. Phys. Rev. Lett. 2009, 102, 117001. [Google Scholar] [CrossRef]
  9. Hicks, C.W.; Kirtley, J.R.; Lippman, T.M.; Koshnick, N.C.; Huber, M.E.; Maeno, Y.; Yuhasz, W.M.; Maple, M.B.; Moler, K.A. Limits on superconductivity-related magnetization in Sr2RuO4 and PrOs4Sb12 from scanning SQUID microscopy. Phys. Rev. B 2010, 81, 214501. [Google Scholar] [CrossRef]
  10. Ray, S.J.; Gibbs, A.S.; Bending, S.J.; Curran, P.J.; Babaev, E.; Baines, C.; Mackenzie, A.P.; Lee, S.L. Muon-spin rotation measurements of the vortex state in Sr2RuO4: Type-1.5 superconductivity, vortex clustering, and a crossover from a triangular to a square vortex lattice. Phys. Rev. B 2014, 89, 094504. [Google Scholar] [CrossRef]
  11. Kawasaki, I.; Watanabe, I.; Amitsuka, H.; Kunimori, K.; Tanida, H.; Onuki, Y. Superconducting properties of noncentrosymmetric superconductor LaPt3Si studied by muon spin spectroscopy. J. Phys. Soc. Jpn. 2013, 82, 084713. [Google Scholar] [CrossRef]
  12. Fujisawa, T.; Yamaguchi, A.; Motoyama, G.; Kawakatsu, D.; Sumiyama, A.; Takeuchi, T.; Settai, R.; Onuki, Y. Magnetization measurements of non-centrosymmetric superconductor LaPt3Si: Construction of low temperature magnetometers with the SQUID and Hall sensor. Jpn. J. Appl. Phys. 2015, 54, 048001. [Google Scholar] [CrossRef]
  13. Cadorim, L.R.; Toledo, L.V.; Ortiz, W.A.; Berger, J.; Sardella, E. Closed vortex state in three-dimensional mesoscopic superconducting films under an applied transport current. Phys. Rev. B 2023, 107, 094515. [Google Scholar] [CrossRef]
  14. Cadorim, L.R.; Sardella, E.; Domínguez, D.; Berger, J. Stability limits of flux states in two-band superconductor rings. Phys. Rev. B 2024, 110, 144513. [Google Scholar] [CrossRef]
  15. Benites, T.N.S.; Presotto, A.; Ortega, J.B.; Sardella, E.; Zadorosny, R. Influence of surface defects on the vortex penetration and arrangement at mesoscopic superconducting samples. J. Phys. Conf. Ser. 2024, 2726, 012004. [Google Scholar] [CrossRef]
  16. Gokhfeld, D.M.; Maksimova, A.N.; Kashurnikov, V.A.; Moroz, A.N. Optimizing trapped field in superconductors with perforations. Phys. C 2022, 600, 1354106. [Google Scholar] [CrossRef]
  17. Moroz, A.; Rudnev, I.; Stepanenko, A.; Maksimova, A.; Kashurnikov, V. Features of vortex dynamics in a HTS with disordered pinning lattice. J. Supercond. Nov. Magn. 2024, 37, 339. [Google Scholar] [CrossRef]
  18. Maksimova, A.N.; Gokhfeld, D.M.; Moroz, A.N.; Kashurnikov, V.A. Relaxation of the trapped magnetic flux in a mesoscopic HTS with artificial pinning in the form of submicron holes. Chin. J. Phys. 2024, 88, 493. [Google Scholar] [CrossRef]
  19. Maksimova, A.N.; Moroz, A.N.; Rudnev, I.A.; Pokrovskii, S.V.; Kashurnikov, V.A. Critical current of a layered high-temperature superconductor with tilted irradiation defects. Phys. Scr. 2024, 99, 105938. [Google Scholar] [CrossRef]
  20. Yerin, Y.S.; Omelyanchouk, A.N. Coherent current states in a two-band superconductor. Low Temp. Phys. 2007, 33, 401. [Google Scholar] [CrossRef]
  21. Silva, R.M.; Milošević, M.V.; Domínguez, D.; Peeters, F.M.; Aguiar, J.A. Distinct magnetic signatures of fractional vortex configurations in multiband superconductors. Appl. Phys. Lett. 2014, 105, 232601. [Google Scholar] [CrossRef]
  22. Maiti, S.; Sigrist, M.; Chubukov, A. Spontaneous currents in a superconductor with s + is symmetry. Phys. Rev. B 2015, 91, 161102. [Google Scholar] [CrossRef]
  23. Garaud, J.; Silaev, M.; Babaev, E. Thermoelectric signatures of time-reversal symmetry breaking states in multiband superconductors. Phys. Rev. Lett. 2016, 116, 097002. [Google Scholar] [CrossRef] [PubMed]
  24. Vadimov, V.L.; Silaev, M.A. Polarization of the spontaneous magnetic field and magnetic fluctuations in s + is anisotropic multiband superconductors. Phys. Rev. B 2018, 98, 104504. [Google Scholar] [CrossRef]
  25. Lin, S.Z.; Maiti, S.; Chubukov, A. Distinguishing between s + id and s + is pairing symmetries in multiband superconductors through spontaneous magnetization pattern induced by a defect. Phys. Rev. B 2016, 94, 064519. [Google Scholar] [CrossRef]
  26. Sørensen, M.P.; Pedersenand, N.F.; Ögren, M. The dynamics of magnetic vortices in type II superconductors with pinning sites studied by the time dependent Ginzburg-Landau model. Phys. C 2017, 533, 40. [Google Scholar]
  27. Schmid, A. A time dependent Ginzburg-Landau equation and its application to the problem of resistivity in the mixed state. Phys. Kondens. Mater. 1966, 5, 302. [Google Scholar] [CrossRef]
  28. Gor’kov, L.P.; Éliashberg, G.M. Generlization of the Ginzburg-Landau equations for non-stationaryproblems in the case of alloys with paramagnetic impurities. Zh. Eksp. Teor. Fiz. 1968, 54, 612. [Google Scholar]
  29. Aguirre, C.A.; Joya, M.R.; Barba-Ortega, J. On the vortex matter in a two-band superconducting meso-prism. Phys. C 2021, 585, 1353867. [Google Scholar] [CrossRef]
  30. Du, S.Z.; Zhong, Y.N.; Yao, S.W.; Peng, L.; Shi, T.T.; Sang, L.N.; Liu, X.L.; Lin, J. The dynamics of current-driven vortex in two-band superconductor with s + d wave pairing. Phys. Lett. A 2022, 443, 128206. [Google Scholar] [CrossRef]
  31. Garaud, J.; Silaev, M.; Babaev, E. Microscopically derived multi-component Ginzburg-Landau theories for s + is superconducting state. Phys. C 2017, 533, 63. [Google Scholar] [CrossRef]
  32. Yao, S.W.; Peng, L.; Lin, J.; Chen, J.; Cai, C.B.; Zhou, Y. Properties of vortex configurations in two-band mesoscopic superconductors with Josphson coupling: The Ginzburg-Landau theory. J. Low. Temp. Phys. 2021, 202, 329. [Google Scholar] [CrossRef]
  33. Ryu, Y.G.; Mun, G.I.; Kwon, Y.N.; Kim, S.H.; Hong, S. Motion of magnetic vortices in type-II superconductor with randomly distributed pinning centers. Phys. C 2022, 602, 1354125. [Google Scholar] [CrossRef]
  34. Ryu, Y.G.; Om, J.H.; Kim, J.H.; Ro, G.I.; Mun, G.I.; Hong, S. The influence of surface defects on motion of magnetic vortices in mesoscopic type-II superconductor with randomly distributed pinning centers. J. Supercond. Nov. Magn. 2024, 37, 527. [Google Scholar] [CrossRef]
  35. Polo, A.S.M.; da Silva, R.M.; Vagov, A.; Shanenko, A.A.; Toro, C.E.D.; Aguiar, J.A. Nonequilibrium interband phase textures induced by vortex splitting in two-band superconductors. Phys. Rev. B 2017, 96, 054517. [Google Scholar] [CrossRef]
  36. Schuller, I.K.; Gray, K.E. Time-dependent Ginzburg-Landau: From single particle to collective behavior. J. Supercond. Nov. Magn. 2006, 19, 3. [Google Scholar] [CrossRef]
  37. COMSOL, Comsol Multiphysics Modeling Guide. 2012. Available online: https://www.comsol.com (accessed on 15 October 2024).
  38. Du, Q.; Gunzburger, M.D.; Peterson, J.S. Solving the Ginzburg-Landau equations by finite-element methods. Phys. Rev. B 1992, 46, 9027. [Google Scholar] [CrossRef]
  39. Alstrøm, T.S.; Sørensen, M.P.; Pedersen, N.F.; Madsen, S. Magnetic flux lines in complex geometry type-II superconductors studied by the time dependent Ginzburg-Landau equation. Acta. Appl. Math. 2011, 115, 63. [Google Scholar]
  40. Oripov, B.; Anlage, S.M. Time-dependent Ginzburg-Landau treatment of rf magnetic vortices in superconductors: Vortex semiloops in a spatially nonuniform magnetic field. Phys. Rev. E 2020, 101, 033306. [Google Scholar] [CrossRef] [PubMed]
  41. Li, J.C.; Huang, Y.Q. Time-Domain Finite Element Methods for Maxwell’s Equations in Metamaterials; Springer: Heidelberg, NY, USA, 2013. [Google Scholar]
  42. Tinkham, M. Introduction to Superconductivity; McGraw-Hill Inc: New York, NY, USA, 1996. [Google Scholar]
Figure 1. Evolution of the magnetic flux density B z (ac) and the order parameter of the first condensate Ψ 1 (df) in the presence of an isotropic defect in a 15 ξ 1 × 15 ξ 1 type-1.5 superconductor. The snapshots show the Meissner phase (a,d), vortex cluster phase (b,e), and vortex lattice phase (c,f) in the external magnetic fields H e = 0.25 H 0 , 0.55 H 0 , and 0.85 H 0 , respectively. The magnetization only has the component perpendicular to the superconducting plane.
Figure 1. Evolution of the magnetic flux density B z (ac) and the order parameter of the first condensate Ψ 1 (df) in the presence of an isotropic defect in a 15 ξ 1 × 15 ξ 1 type-1.5 superconductor. The snapshots show the Meissner phase (a,d), vortex cluster phase (b,e), and vortex lattice phase (c,f) in the external magnetic fields H e = 0.25 H 0 , 0.55 H 0 , and 0.85 H 0 , respectively. The magnetization only has the component perpendicular to the superconducting plane.
Crystals 15 00133 g001
Figure 2. Evolution of the magnetic flux density B z (ac) and the order parameter of the first condensate Ψ 1 (df) in the presence of an anisotropic defect in a 15 ξ 1 × 15 ξ 1 type-1.5 superconductor. The snapshots show the Meissner phase (a,d), vortex cluster phase (b,e), and vortex lattice phase (c,f) in the external magnetic field H e = 0.25 H 0 , 0.55 H 0 , and 0.85 H 0 , respectively. The magnetization only has the component perpendicular to the superconducting plane.
Figure 2. Evolution of the magnetic flux density B z (ac) and the order parameter of the first condensate Ψ 1 (df) in the presence of an anisotropic defect in a 15 ξ 1 × 15 ξ 1 type-1.5 superconductor. The snapshots show the Meissner phase (a,d), vortex cluster phase (b,e), and vortex lattice phase (c,f) in the external magnetic field H e = 0.25 H 0 , 0.55 H 0 , and 0.85 H 0 , respectively. The magnetization only has the component perpendicular to the superconducting plane.
Crystals 15 00133 g002
Figure 3. Variation in sample voltage V against external current j n with isotropic or anisotropic impurity in 15 ξ 1 × 15 ξ 1 mesoscopic superconducting system.
Figure 3. Variation in sample voltage V against external current j n with isotropic or anisotropic impurity in 15 ξ 1 × 15 ξ 1 mesoscopic superconducting system.
Crystals 15 00133 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, G.; Han, T.; Li, J.; Zhang, J.; Huang, H. Defect Pinning and Critical Current of Magnetic Vortex Cluster in Mesoscopic Type-1.5 Superconductors. Crystals 2025, 15, 133. https://doi.org/10.3390/cryst15020133

AMA Style

Wang G, Han T, Li J, Zhang J, Huang H. Defect Pinning and Critical Current of Magnetic Vortex Cluster in Mesoscopic Type-1.5 Superconductors. Crystals. 2025; 15(2):133. https://doi.org/10.3390/cryst15020133

Chicago/Turabian Style

Wang, Guo, Tianyi Han, Jie Li, Jiangning Zhang, and Hai Huang. 2025. "Defect Pinning and Critical Current of Magnetic Vortex Cluster in Mesoscopic Type-1.5 Superconductors" Crystals 15, no. 2: 133. https://doi.org/10.3390/cryst15020133

APA Style

Wang, G., Han, T., Li, J., Zhang, J., & Huang, H. (2025). Defect Pinning and Critical Current of Magnetic Vortex Cluster in Mesoscopic Type-1.5 Superconductors. Crystals, 15(2), 133. https://doi.org/10.3390/cryst15020133

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop