<<

CEJP 3(1) 2005 77–103

Superconducting states of the cylinder with a single vortex in magnetic field according to the Ginzburg–Landau theory

Gely F. Zharkov∗ P.N. Lebedev Physical Institute, Russian Academy of Sciences, Leninsky pr., 53, 119991, Moscow, Russia

Received 31 May 2004; accepted 8 October 2004

Abstract: Self-consistent of the nonlinear Ginzburg–Landau (GL) equations are investigated numerically for a superconducting (SC) cylinder, placed in an axial magnetic field, with a single vortex on the axis (m = 1). Two modes, which show the original state of the cylinder, SC or normal (s0 and n0), are studied. The field increase (FI) and the field decrease (FD) regimes are studied. The critical fields destroying the SC state with m = 1 are found in both regimes. It is shown that in a cylinder of radius R and GL- parameter κ, there exist a number of solutions depending only on the radial co-ordinate r corresponding to different states such as M, e, d, p,i, n, n, n∗, and the state diagram on the plane of the variables (κ, R) is described. The critical fields corresponding to intra- state transitions and the onset of hysteresis are obtained. The critical field H0(R) dividing the paramagnetic and diamagnetic states of the cylinder with m = 1 is determined. The limiting fields of supercooling or of the normal state at which the restoration of the SC state occurs are established. It is shown, that (in both cases m = 1, 0) there exist two critical parameters, κ0 = 1/√2 = 0.707 and κc = 0.93, which divide bulk SC into three groups (with κ < κ , κ κ κ and κ > κ ), in accordance with the behavior 0 0 ≤ ≤ c c in a magnetic field. The parameters κ0 and κc mark the boundary for the existence of a supercooled normal n-state in FD-regime and a superheated SC M-state in FI-regime respectively. It is shown, that the value κ∗ = 0.417, which was claimed in a number of papers as related to type-I superconductors, is illusory. c Central European Science Journals. All rights reserved.

Keywords: GL-equations, critical fields, hysteresis PACS (2000): 74.25.-q, 74.25.Dw

∗ We regret to inform that Professor Gely Zharkov passed away on 9th July 2004. 78 G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103

1 Introduction

The macroscopic GL-theory is widely used for studying the behavior of superconduc- tors in a magnetic field. This theory leads to two coupled nonlinear three-dimensional equations for a complex order parameter Ψ(r) = ψ(r)eiΘ(r) and the vector potential of the magnetic field A(r) (r is the three-dimensional space coordinate, ψ is the modulus and Θ is the of the order parameteter). The analysis of these equations in case of infinite superconductors [1] leads to the conclusion that there exist two groups (or types) of superconductors, which differ by the value of GL-parameter κ. For κ < κ0 = 1/√2 the surface free energy of the interface between two co-existing n and s phases is positive,

σns > 0 (type-I superconductors), and for κ > κ0 the surface free energy is negative,

σns < 0 (type-II superconductors). [The value σns > 0 points to the instability of the magnetically supercooled n-phase relative the s-phase .] It has been shown that in a bulk type-II superconductor (at κ κ with the field increasing) the vortices ≫ 0 appear, which form the regular lattice (see [2]), and other singularities of the behavior of bulk superconductors were studied [2–5]. However, in case of a finite extent system it is generally impossible to find the exact of the nonlinear GL-equations without us- ing numerical methods. To understand the predictions, which follow from the GL-theory, the one-dimensional solutions (depending only on one space co-ordinate x) are frequently studied. Such solutions are of interest also in view of the attempts to manufacture small size measuring devices on the base of mesoscopic superconductors. The one-dimensional solutions of GL-equations for a finite extent system (the plate of width 2D and the cylinder of radius R in a parallel magnetic field H) were studied recently in [6,7]. It was found that even in the simplest case, when there are no vortices inside the superconductor (m = 0), there exist a number of vortex-free states (such as M-, e, d, p, i, n, n), which correspond to two possible regimes of the magnetic field action on the superconductor: the field increase (FI) and the field decrease (FD) regimes. It was shown, that a bulk superconductor in each regime is characterized by its own critical parameter κ, so that in FD-regime the value κ0 = 1/√2 = 0.707 corresponds to the stability threshold of a supercooled n-state (with ψ 0), and in FI-regime the value ≡ κ = 0.93 corresponds to the stability threshold of a superconducting state (with ψ 1). c ≈ In the original GL-paper [1], the distinction between κ0 and κc was not mentioned, since they considered in fact the intermediate state of infinite superconductor: the difference between κ0 and κc manifests itself only in finite dimension samples (see [6,7] for details). In the present paper the one-dimensional solutions of GL-equations are studied when a single vortex (m = 1) is situated on a cylinder axis. Some properties of such axially symmetric solutions are analogous to those described in [6,7] (see also [8–12]), however, to assist the reader, the essential points will be explained anew, so the preliminary ac- quaintance of the reader with [6–12] is not assumed. The main attention is devoted below to the questions not discussed previously. In contrast to the case m = 0 [6,7], we will differentiate two modes and two regimes, in which the external field acts on the super-

conductor. In s0-mode (the index 0 denotes the initial moment of time) the specimen G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 79

is initially in a superconducting s0-state in some field H, which then either increases

(FI-regime) or decreases (FD-regime). In n0-mode a superconductor (at a

T < Tc) is initially in normal n-state under a strong field (of arbitrary sign), which then either increases (FI-regime) or decreases (FD-regime). The sequence of the states, emerging in each of two modes and each of two regimes, is different, which points to the possibility of a hysteresis in the system. The solutions of GL-equations are studied below in detail for arbitrary values of the external field H, the cylinder radii R and the parameter κ. The emerging picture in the case m = 1 turns out to be more complicated than in the case m = 0. It will be shown, that on the plane of parameters (κ,R) there are 6 critical lines, which divide several regions of a characteristic behavior of the sample magnetization, 4πM(H). For − a bulk cylinder with a vortex on the axis (as in the case m = 0) there exist two critical parameters: κ0 = 1/√2 = 0.707 (in n0-mode) and κc = 0.93 (in s0-mode), which mark the instability boundaries of the system in magnetic field. The hysteresis states are studied and the hysteresis boundaries are found, as well as other peculiarities of the behavior of a cylinder of finite radius in an external field H. The paper is organized as follows. In Sec. 2 the basic equations and the boundary conditions are presented. Examples of self-consistent solutions of GL-equations (m = 1), together with an explanation of the terminology used are given in Sec. 3. In Sec. 4 the critical fields are determined for a cylinder of arbitrary radius R and various κ. In Sec. 5 the critical lines (µ, π, <ζ, ζ>, i and ν), which exist on the plane of parameters (κ,R) at m = 1 are obtained, and their meaning is clarified. In Sec. 6 the connection of the exact results with the linearized theory is discussed. Here it is demonstrated also, that the value κ∗ = 0.417 (which is discussed in a number of papers as having to do with type-I superconductors) is illusory. In Sec. 7 the solutions, which lie on the line π, and their connection with the Bogomolnyi equations are discussed. Here the solutions belonging to the line µ and their connection with the critical parameter κc are also discussed. In Sec. 8 the jumps in the total flux and in the free energy of the system (which accompany the transitions between different states of the cylinder in magnetic field) are discussed. The main results are briefly summarized in Sec. 9.

2 Basic equations and boundary conditions

The GL-equations for the order parameter ψ and the ϕ-component of the vector potential A can be written in the cylindrical co-ordinate system (r, ϕ, z) in the form

d2a 1 da ψ2 a = 0, (1) dρ2 − ρ dρ − κ2

d2ψ 1 dψ a2 + +(ψ ψ3) ψ = 0. (2) dρ2 ρ dρ − − ρ2 Here ψ(ρ) is a real function (0 ψ 1), ρ = r/ξ is a dimensionless co-ordinate, a(ρ) is ≤ ≤ the dimensionless magnetic field potential, while 80 G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103

a + m B 1 da φ0 A = ξHξ , bξ = = , Hξ = 2 , ρ Hξ ρ dρ 2πξ ξ is the coherence length, λ = κξ is the London penetration depth, κ is the GL-parameter,

Hξ is the unit for measuring the field, φ0 = hc/2e is the flux quantum. The boundary conditions associated with Eqs. (1), (2) are

1 da a ρ=0 = m, = hξ, (3) | − ρ dρ ρ=Rξ

dψ ψ ρ=0 = 0, = 0, (4) | dρ ρ=Rξ

where Rξ = R/ξ, hξ = H/Hξ and m = 1. The magnetic moment (or the cylin- der magnetization) Mξ = M/Hξ is found from the formula bξ = hξ + 4πMξ, where 2 bξ = B/Hξ = 2a(Rξ)/Rξ , the overline indicates the averaging over the specimen volume. (Notice, that the potential a(x), as well as the number m, do not depend on the choice of the field scale, i.e., they are scale invariants). In the case ψ 1 and B = H, the equations (1) and (2) reduce [13,14] to a single ≪ κ-independent linear equation

d2ψ 1 dψ a2(ρ) + + 1 0 ψ =0 (5) dρ2 ρ dρ − ρ2 !

(where a (ρ) = m + h ρ2/2), whose solutions shall be denoted by ψ(ρ) = K(m)(ρ). 0 − ξ The functions K(m)(ρ) have the form K(m)(ρ) = ym/2e−y/2 , where y = ρh /2 and F ξ F is a combination of two linearly independent hypergeometric functions (the Kummer functions, see [16]). The difference ∆g of the Gibbs free energies for the cylinder in superconducting and 2 normal states, [related to the unity volume and normalized by Hcb/8π, where Hcb = φ0/(2π√2λξ)], can be expressed in the form [17] (Mλ = M/Hλ, bλ = B/Hλ):

2 Hcb 8πMλ 4m bλ(0) hλ ∆g =∆G V = g0 hλ + − , (6) 8π ! − κ2 κ2 R2 . λ 2 2 Rλ 1 dψ g = ρdρ ψ4 2ψ2 + ,R = κR , h = κ2h . 0 2  κ2  λ ξ λ ξ Rλ Z0 − dρ !   Notice, that in (1)–(5) ξ is chosen as the unit of length, however, one can use 2 2 λ = κξ instead, and the field Hλ = φ0/(2πλ ) = Hξ/κ as the unit (or, for instance,

Hκ = φ0/(2πλξ) = Hξ/κ, or thermodynamic field Hcb = Hξ/√2κ). In presenting the results of the calculations we shall use different variants of normalization. The self-

consistent solutions were found by the iteration method [18], where in s0-mode the trial

order parameter value was taken in a form ψ(ρ)= ψ0th ρ with ψ0 = 1, but in n0-mode it was ψ0 = 0.001. The results of calculations do not depend on the choice of the concrete numerical algorithm, or on the form of the trial functions ψ(ρ). G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 81

3 Examples of self-consistent solutions of GL-equations

As was mentiond above, the full picture of the superconducting solutions which arise is rather complicated and in this section we provide some examples of such states and elucidate the terminology used below. In Figs. 1(a,b) the self-consistent solutions are displayed for the order parameter

ψ(x) and the field bλ(x) (x = r/R) inside the cylinder with Rλ = 6 and κ = 0.5 for various external fields H (in the normalization h h = H/H , b = B/H ). The curve ≡ λ λ λ λ h = 0 in Fig. 1(a) corresponds to the vortex state (m = 1, s0-mode) in absence of the field. With the field increasing (FI-regime) the order parameter ψ(x) diminishes and at h = 0.5110 the solution takes the form presented by the line M. Such M-solutions will be mentioned as the Meissner-type, because the external field in this state is screened out from the superconductor interior. This is evident from Fig. 1(b), where the solid line

M shows the magnetic field configuration in the sample, bλ(x), for h = 0.5110. When the field h increases by 1 10−4 the M-state gets unstable and transforms by a jump into a · stable normal n-state (with ψ 0). The critical field of the first-order jump from s- to ≡ n-state will be denoted as h1. In the n0-mode with the field decreasing (FD-regime) the initial n-state passes into a supercooled (metastable) n-state, which persist down to the

field hr = 0.3165, where n-state becomes absolutely unstable and the M-states restored by a jump. The dotted lines n and M in Figs. 1(a,b) correspond to the restoration field hr in FD-regime (n0-mode).

In Figs. 1(c,d) the solutions ψ(x) and bλ(x) are depicted at the point Rλ = 6,

κ = 0.8. Here in s0-mode (FI-regime) the Meissner M-solution at h1 = 0.8043 takes the form, depicted by the solid line M, after which the jump from M- to n-state occurs.

In FD-regime (n0-mode) the supercooled n-state persists down to the field hp = 0.6420, after which the precursor p-solution appears, having originally a very small amplitude (ψ 1), which grows with the field decreasing, and at the point h = 0.6286 the max ≪ r solution takes the form represented by the dotted line p. The point hr (FD-regime, n0- mode) is critical, at h

The states, which exist at the point Rλ = 6, κ = 1.5, are represented in Fig. 1(e,f).

In s0-mode (FI-regime) the M-state becomes unstable at h1 = 1.3704, and the jump to e the edge-suppressed e-state (with finite amplitude ψmax = 0.88) occurs. [It should be stressed, that the transition M e illustrates the presence of special (“edge”) mechanism → for the field penetration inside the finite extent superconductor, which is different from the usual (“vortex”) mechanism, when the field penetrates inside the specimen in the form of vortices (see also [6,7])]. If the field h is further increased, the amplitude of e-state 2 diminishes and at the point h2 = 2.2514 (hξ = h2/κ = 1.0006) it vanishes by a second order (ψ 0). In n -mode (FD-regime) the e-state nucleates at the max → 0 field h2, and at smaller field h1 it passes smoothly into the “depressed” d-state. Such a d-state exists down to the point hr = 1.2300, where the M-state is restored by a jump 82 G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 d M (the dotted lines). [Evidently, in transitions d M (or p M as in Fig. 1(d)) → → → the additional (non-vortex) mechanism for the field expulsion from the superconductor interior is realized.] What was described above, is further illustrated in Fig. 2, where the dependencies of the order parameter ψmax(h)(h = H/Hλ) and magnetization Mλ(h) are given at the same points (κ,Rλ), as in Fig. 1. One can see, that these dependences are asymmetric relative the axis h = 0, thus it is necessary to distinguish between the right and left branches of > the curves, and below we shall denote the critical fields h1, h2 and hr by the index (for < the right branch, FI-regime, s0-mode) and by the index (for the left branch, FD-regime, s0-mode). In addition, one should distinguish the metastable supercooled n-state (on the ∗ right branch, n0-mode, FD-regime) from the superheated n -state (on the left branch, n -mode, FI-regime), and also the absolutely unstablen ˙ -state with ψ 0 (see Sec. 6). 0 ≡ By the label h0 on the curves in Fig. 2 the value of the field is marked, at which the magnetization vanishes (4πM = B H = 0). The points, where M(h) < 0 for h > 0, − correspond to the diamagnetic state (B 0 for h> 0, correspond to the paramagnetic state of the cylinder (B>H, the magnetization is directed along the field). The points, where M(h) > 0 for h < 0, correspond to the diamagnetic state (magnetization is directed opposite the field). [In the case m = 0 the paramagnetic region is absent and both branches of the curves M(h) for h> 0 and h< 0 correspond to diamagnetism.]

The analogous dependencies are shown in Fig. 3 for smaller radius Rλ = 3. One can see (Figs. 3(a,b)), that for κ = 0.5 the superconducting M-state can exist only in finite fields h> 0 (the field stimulation effect, due to the retention of by applying the external field). Here the M-state vanishes by a second order phase transition > < in both regimes (h2 = 0.5841 in FI-regime, and h2 = 0.1633 in FD-regime). < For κ = 0.8 (Figs. 3(c,d)) there are the first order jumps at the field h1 = 0.2907 > − (FD-regime, s0-mode), and also at the field h1 = 1.0238 (FI-regime, s0-mode). In n0- < mode the jumps take place at the field h˙ r = 0.2054 (FI-regime) and at the field ˙ > − hr = 1.0022 (FD-regime). For κ = 1.5 (Figs. 3(e,f)) there is no jump on the right branches of the curves, but i the inflexion point i remains (at the field hi = 1.88, ψmax = 0.72), so one can distiguish between the M-state (at hhi), which ends at > h2 = 2.2700 (hξ = 1.0009) by a second order phase transition to the n-state. To conclude this section, notice that in addition to the stable and metastable su- perconducting states there exist also the absolutely unstable (which cannot be realized in practice) u-states with ψ = 0 (they are drawn schematically in Fig. 2 by dotted max 6 u-lines). The iteration algorithm [18] enables us to find only stable (relative to small perturbations) solutions, but to obtain unstable u-states one should use other methods. However, the reason for the principal difference of the stable and unstable branches of solutions is clear. Indeed, to find the stable solution we have used the physical setting of the problem: the field at the boundary is specified by the experimentalist and one looks for the corresponding solution for the order parameter ψ. The convergence of our G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 83 iteration procedure has to do with the stability of the solution relative to small pertur- bations. To find the unstable branch of solutions one should use the unphysical setting of the problem: the order parameter at the boundary is kept fixed (what is impossible to do in experiment) and one looks for the corresponding value of the external field. The solution, obtained in this way exists, it is, however, unphysical, since it grows monotoni- cally with the field increasing (the order parameter in Fig. 2 grows monotonically on the right branches of the u-curves, i.e. dψmax/dh > 0; analogously for the left branches). In addition, one can verify that the unstable solution has the positive increment, i.e. small deviations from u-solution grow in time. Thus, u-solutions exist mathematically, but they cannot be realized in practice (the analogy is evident here with the unstable state of the ball placed on top of the convex surface, and with the turnover of the wave front in hydrodynamics [19]). The absolutely unstable u-solutions will not be considered below. < (In Fig. 3 u-solutions are not shown. It is evident from Figs. 2(a,b,c), that the fields h1 > and h1 correspond to the bifurcation points, where M- and u-states coincide).

4 Critical fields for a cylinder with m = 1 (the phase diagrams)

What was said above is further clarified in Fig. 4, where the critical fields (or phase diagrams) for a cylinder of arbitrary radius R and for several values of κ are shown. A κ < > closer inspection of Fig. 4(a) ( = 0.5)reveals that the solid lines h1 and h1 bound the region, where the Meissner M-states exist in s0-mode. With the field increasing, the jump > from M- to n-state occurs at the field h1 ; with the field decreasing, the jump from M- to < n- occurs at the field h1. The dotted lines correspond to the n0-mode, they mark the < ˙ ˙ > critical fields hr and hr , at which the restoration of the superconducting M-state from the superheated n∗- or supercooled n-state take place. (The additional dot ˙ above some > < ∗ of the fields hr and hr marks the place, where the metastable n - or n-states adjoin with < > the absolutely unstablen ˙ -state; see also Sec. 6.) In the region between hr and hr the same M-state is realized in both modes. The curve h0 divides the paramagnetic region (where the magnetization M > 0) from the diamagnetic region (where M < 0), on this curve M(h0) = 0. With the radius R diminishing, the region of the supercooled n-state diminishes, and at point ζ> the hysteresis vanishes. On the left branch of the curves the hysteresis vanishes at point <ζ, where the region of existence of the superheated n∗-state ends. For even smaller radii the transition from M- to n-state (and vice versa) proceeds < > reversibly by a second order phase transition in the fields h2 and h2 . The point ν corresponds to the minimal cylinder radius (Rξ = 1.32) at which the superconducting state with a single vortex (m = 1) becomes impossible. Fig.4(b) (κ = 0.8) is analogous to the preceeding one. Here one can see additionally the region of existence of the hysteresis p-states, which nucleate at the field hp by a second > order phase transition (in n0-mode) and terminate at the field hr with a jump from p- to M-state. The point π marks the end of the region of p-states. The curve h0, which divides para- and diamagnetic regions, is present also in Fig. 4(b) (as in all other figures), but is not displayed in order to simplify the presentation. 84 G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103

In Fig. 4(c) (κ = 1) one can furthermore see the region of existence of e-states, > which appear in s0-mode at the field h1 (in a jump M e), with the tail of e-states →> vanishing by a second order phase transition at the field h2 (FI-regime). In n0-mode the > > e-state nucleates again at the field h2 , at the field h1 the e-state transforms smoothly > into a hysteresis d-state, which exists down to the line hr , where the jump from d- to M- state occurs (FD-regime). Above the point µ1 there are no supercooled n-states, but the hysteresis is ensured by the existence of d-states. Below the point µ1 the n-state emerges, from which p-state nucleates (at the field hp) and then p-state transforms in a jump (at > the field hr ) into M-state. At the point µ1 the states p- and d- are not distinguishable. Point π marks the end of p-states region, below this point the transition from n- to M- ˙ > takes place in a jump at the field hr , without formation of p-state preliminary. Point ζ> marks the end of the region of the hysteresis n-states, below point ζ> the transition > from n- to M- proceeds reversibly at the field h2 . Point ν (Rξ = 1.32) has the same significance in all figures. The left branch of the curves in Fig. 4(a) are analogous to that depicted in Fig. 4(a,b)and are not explicitly shown. > Above the point ζ in Fig. 4(d) the transition from M- to e-state in s0-mode is accompanied by jumps (the region j). Below point ζ> the jumps and hysteresis vanish, but on the magnetization curve (in the region i) the inflexion point remains (see Figs. 3(e,f)). At point i (marked by a circle) the dependence M(h) becomes monotonic, without the inflexion points (see Fig.3(f)). The remaining notations were explained earlier. The complicated picture of various states and transitions between them reflects the complexity and diversity of the nonlinear problem under investigation. Further explana- tions will be given below. Here we shall only point out that for large R the equality holds h> =

Notice also, that the curve h0, which divides the para- and diamagnetic regions, is

present in all Figs. 4(a–d). One can verify, that on the plane of variables (Rξ,hξ) this curve for κ 1 does not depend on κ and is described by the universal formula: γ ≥ Rξ = α(β/hξ) , where α = 1.32, β = 1.92, γ = 0.47265. [Thus, to the point ν with the minimally admissible radius Rξ = 1.32 corresponds the value hξ = 1.92]. For κ < 1 the deviations from this formula appear and the curve h0 approaches the axis hξ = 0, so at κ 0 the paramagnetic region vanishes. →

5 Critical lines on the plane (κ,Rλ) (the state diagram)

The solutions of the nonlinear system of equations (1)–(4) (with m = 1) depend on the co- ordinate ρ and three parameters: κ, Rξ and hξ. The results of the numerical calculations are summarized in Fig. 5, where the state diagram of the cylinder is shown on the plane of variables (κ,Rλ = Rξ/κ). To each point of this plane corresponds some solution ψ(ρ; hξ) and a(ρ; hξ), which depends on the field. It is convenient to imagine that in any point of G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 85 this plane one can make a peephole to see the field dependencies of the solutions, which are hidden beneath the surface (κ,Rλ). Studying such dependencies, one can learn, in particular, the peculiarities in the behavior of ψmax(hξ) and M(hξ) in every point of the plane (κ,Rλ). Proceeding in this manner, one can find 6 critical lines (denoted in Fig. 5 by the letters π, µ, <ζ, ζ> and ν) and 6 characteristic regions (denoted by the letters A, B, C, D, E and n), the meaning of which is elucidated below. ∗ The region A (for small κ and large Rλ) is characterized by the presence of states n , M-, n (see Fig. 4(a) at κ = 0.5), and by a first order jumps in transitions at the fields < > > h1 and h1 . If Rλ diminishes, the hysteresis n-state in Fig. 4(a) ends at point ζ . The curve ζ> in Fig. 5 consists of critical points ζ>, found for other κ < 1. The underlying curve <ζ consists of critical points <ζ, and the curve ν in Fig. 5 consists of the critical points ν, shown in Fig. 4(a).

The region B (for κ > 0.707 and large Rλ) is characterized by the presence of p-states

(see Fig. 4(b) at κ = 0.8). If Rλ diminishes, the hysteresis p-state ends at point π. The curve π in Fig. 5 consists of critical points π, it bounds the region of existence of hysteresis p-states. Beneath the line π the already explained curves <ζ, ζ> and ν exist.

The region C (for κ > 0.93 and large Rλ) is characterized by states M-, d, e (in

FI-regime, s0-mode; see Fig. 4(d) at κ = 1.5). The hysteresis here is due to d-states, > > > which end at point ζ with Rλ diminishing. The curve ζ in Fig. 5 consists of points ζ , shown in Fig. 4(d). Above this line there are the magnetization jumps at the transitions > from M- to e-, what indicate to the instability of M-state in the field h1 . Beneath the line ζ> the magnetization jumps disappear and are replaced by the inflexion points of the function M(h) (the region D in Fig. 5). Above the line <ζ (in regions A, B, C, D) there exist magnetization jumps at the left branches of the curves, as shown in Figs. 3 and 4. Below the line <ζ (in region E) the jumps disappear and are replaced by the inflexion points. [With Rλ diminishing, the points of inflexion vanish and the magnetization curve becomes monotonic (as in Fig. 4(d) at point i). Such i-points constitute the i-curve, which is depicted in Fig. 5 by a thin dotted line.] The region E is situated below the line <ζ, where s-state is destroyed by a second order phase transition in both regimes FI- and FD- (see Figs. 3(a,b)). The limiting curve ν κ ν in Fig. 5 (Rλ = 1.32/ ) bounds the region of existence of superconducting states with the vortex (m = 1) at the cylinder axis. Below this line only the normal states n (with ψ 0) are possible. ≡ Thus, each region in Fig. 5 is characterized by a specific behavior of the order param- eter ψ(h) and magnetization M(h), by the presence of characteristic jumps and hysteresis loops. This figure, in fact, is the main result of the laborious and time consuming cal- culations. Notice, that each curve, represented in Fig. 5, depends on κ and therefore it cannot be found from the κ-independent linear equation, Eq. (5); for that it would be necessary to solve the full system of nonlinear equations, Eqs. (1)–(4). The question of applicability of the linearized theory [14,15] is discussed in detail in the next Section. 86 G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103

6 Connection with the linearized theory

As was stated above, the linearized theory [14,15] corresponds to the approximation ψ 1, at that the nucleation of superconductiviy proceeds as a second order phase ≪ transition and is described by κ-independent Eq. (5), solutions of which are expressed via the Kummer functions [16]. The critical fields, determined with the help of Eq. (5), will be marked below by the letter K. To trace the connection of the exact results with the linear theory, let us return again to Fig. 4. We clarify once more the meaning of the additional point over some of the critical > < fields hr and hr in Fig. 4. The index r denotes the first order jump at full restoration of superconducting M-state in n0-mode. If the jump takes place from the intermediate p- > state with finite amplitude, the corresponding field (hr ) cannot be found from the linear Eq. (5). However, if the jump into M-state happens directly from a supercooled n-state (ψ 0 in the instability point), then such jump (which is accompanied by the time ≡ transformation of the solution) begins by the appearance of small superconducting state (ψ 1), so the field h˙ > (which marks the boundary between a supercooled metastable ≪ r n-state and dynamically unstablen ˙ -state) can be described by the linear theory. The analogous situation exists on the left branches of the critical fields. The field of the first < ∗ order jump h˙ r (which marks the bound between a superheated metastable n -state and dynamically unstablen ˙ -state) also can be described by the linear Eq. (5). Thus, the κ-independent linear equation (5) describes those segments of the curves in Fig. 4, which correspond to the solutions with ψ 0. max → κ ˙ > > For instance, in Fig. 4(a) ( = 0.5) K-curve consists of segment hr (up to point ζ ), > < < < ˙ next the K-segments h2 and h2, and after ζ-point, the is K-segment hr. In Fig. 4(b) (κ = 0.8) K-curve consists of segment hp (which ends at point π), then ˙ > > > < < ˙ the segment hr (which lies between points π and ζ ), and segments h2 , h2 and hr. κ > In Fig. 4(c) ( = 1) K-curve consists of segment h2 (down to point µ1, where a ˙ > supercooled n-state appears and the precursor p-state nucleates, segment hr (between points π and ζ>, where there are no p-states any more and the transitions from n- to M- > < < ˙ < ˙ happen by a first order jump), segments h2 , h2 and hr (segment hr is not shown). [Compare with Fig. 5, where the line κ = 1 crosses the critical curves in points µ1, π, µ2 ( ζ>), <ζ and ν, marked by the open circles.] ≡ κ > < In Fig. 4(d) ( = 1.5) K-curve consists of segments h2 and h2 (superconductivity < nucleates here by a second order phase transition from n- to e-state) and of segment h˙ r (here in FI-regime a superheated n∗-state becomes dynamically unstable). Thus, although K-curve (as a whole) does not depend on κ and can be found from the linear κ-independent relation, Eq. (5), but indeed consists of separate κ-dependent segments, the meaning and extent of which can be established only by solving the full system of nonlinear equations (1)–(4). [Notice, that κ-independent curve h0, mentioned in the end of Sec. 4, cannot be found from linear Eq. (5), because the states, belonging to this curve, possess values ψ(x) > 0. However, at point ν, lying on the boundary with n-region, one has ψ 0, therefore the minimal value R = 1.32 can be found from Eq. → ξ G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 87

(5).] The analogous situation exists in the vortex-free case m = 0 [6,7]. In Figs. 6(a,b) the critical fields are shown, found for m = 0 and m = 1 (the upper index (m) indicates (1) (1) the corresponding field). Thin solid lines h1 and hr correspond to first order phase transitions at m = 1. Thick solid lines K(1) correspond to a second order phase transitions

(ψmax 0), which are described by the Kummer functions with m = 1. Thin dotted (0)→ (0) lines h1 and hr correspond to a first order phase transitions at m = 0. Thick dotted lines K(0) correspond to second order phase transitions at m = 0. Only the branches of the curves with h> 0 are shown (compare Fig. 4). (Notice, by the way, that Saint-James and deGennes [15,16] also considered only positive critical fields, h> 0. The presence of solutions for h< 0 were mentioned first by Fink and Presson [20].) We stress the qualitative difference of the meaning of the κ-independent curves K(m) in Fig. 6(a,b) for small and large κ (m = 0; 1). One can see, that for small κ (κ = 0.5) (m) (m) the curves K correspond to the minimal field hr , at which (in FD-regime) from a supercooled normal n-state a normal superconducting state restores (in a jump at the instabiliy point ofn ˙ -state), but for larger κ (κ = 1.2) the curves K(m) correspond to the (m) maximal field h2 , at which (in FI-regime) a superconducting state is finally destroyed by a second order phase transition e M. (1) → (1) The open circle and letter h3 in Fig. 6 mark the maximal value of the curves K (h). (m) One can verify, that with m increasing, point h3 in Fig. 6 moves upward and to the left (see, for instance, Fig. 1 in [17]) and approaches (in bulk samples with fixed, but large

R) the value hc3 = 1.69, which corresponds to the maximal field of the existence of the surface superconductivity [14,15]

H = 1.69H , H H = φ /(2πξ2). (7) c3 c2 c2 ≡ ξ 0 [One can say, that the surface superconductivity is due to the fact, that at m 1 ≫ the extent of the vortex core (where ψ(x) 0) grows with m as xm, so the remaining ≈ superconducting state (with ψ(x) 1) is pushed toward the specimen surface, where it ≪ is held by the external field. Notice, that the relation (7) does not depend on κ, because it was obtained in [14,15] in the assumption ψ(x) 1 using Eq. (5).] ≪ It is appropriate to make here the following important remark. The formula (7) can be rewritten in the equivalent form, using a different normalization:

H = 1.69 √2κH = 2.4κH , (8) c3 · cb cb

where Hcb = φ0/(2π√2λξ)= Hξ/(√2κ) is a thermodynamic critical field for bulk super- conductor [1]. Considering formula (8) as an equation for κ, in a number of papers (for instance, [21–25], see also [2–5]) the conclusion was made, that there exist (allegedly) −1 some “special” value κ∗ = (2.4) = 0.417, which (allegedly) divides type-I supercon-

ductors into two groups, those with Hc3 < Hcb (at κ < κ∗), and superconductors with

Hc3 > Hcb (at κ > κ∗). However, such a conclusion is based on a misunderstanding.

Indeed, the field Hcb in (8) depends itself on κ, so Eq. (8), in fact, does not depend on κ 88 G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 and therefore it is impossible to extract from it any additional information, besides those already contained in (7).

According to (7), in bulk superconductors there exist two critical fields, Hc2 and Hc3. The field H [2] is the maximal field for infinite type-II superconductor (κ 1), when c2 ≫ the number of vortices inside is so great, that the vortex cores overlap and only the residual (volume) superconductivity remains with ψ 1. The field H [14,15] describes ≪ c3 the residual (surface) superconductivity with ψ(x) 1, which remains only near the ≪ specimen surface. As a result, the relation (7) between Hc2 and Hc3 depends neither on κ, nor on the choice of the field scale. The relation (8) in fact agrees with the evident meaning of (7), which claims that the surface superconducting state always appears at the

field Hc3, and its shape is the same for any κ. Thus, a “special” value κ∗ = 0.417, in our opinion, is illusory, having neither a physical, nor a mathematical significance. The same is seen from Fig. 5, where the value κ∗ = 0.417 is just a regular point of GL-equations.

7 π- and µ-states; GL-parameters κ0 and κc

As is seen from Figs. 4 and 5, the region B, inside which the metastable p-states with m = 1 exist, is bounded by the critical lines π and µ. On µ-line p-states originate (with the amplitude ψ 1), and on π-line p-states end, reaching the maximal amplitude max ≪ ψ 1. (For R to π-line corresponds the value κ = 1/√2 = 0.707, and max ≈ → ∞ 0 to µ-line corresponds the value κc = 0.93.) It was shown by Bogomolnyi [26], that at R 1, κ = 1/√2, h = 1 the GL-equations degenerate and reduce to the system of ξ ≫ 0 ξ two nonlinear differential equations of first order, which have an analytic solution (see also [27,28].) We shall now consider these π-states in more detail. The self-consistent p-solutions (with m = 1), found in the neighborhood of π-line

(Rξ = 10, κ = 0.708) are shown in Fig. 7(a). It is evident that the curve ψ(x) (x = r/R) consists of two branches 1 and 2, divided by the maximum point of the function ψ(x). One can ascertain, that with larger R the flat upper section of the curve ψ(x) widens. As a result, for R the monotonically growing branch 1 will correspond to the isolated → ∞ vortex (situated at the cylinder axis, far from its boundary), which can be described by

the Bogomolnyi equations with m = 1, κ = 1/√2, hξ = 0. The monotonically abating branch 2 will describe the region near the cylinder boundary (where the influence of the remote vortex is negligible) and obey the Bogomolnyi equations with m = 0, κ = 1/√2, hξ = 1. The curve bξ(x) in Fig. 7(a) can be divided analogously into two branches 1 and 2. What is described above is further illustrated in Fig. 7(b), where the self-consistent

p-solutions ψ(x) and bξ(x)(m = 0) are shown at the point Rλ = 12, κ = 0.7075 (which is near π-line with m = 0). The comparison of these graphs shows, that the branch 2 in Fig. 7(a) is really described by the Bogomolnyi equations with m = 0. Thus the functions 2 2 ψ(x) and bξ(x) are bound by the relation ψ (x)+ bξ(x) = 1, and to the inflexion point i of the function ψ(x) in Fig. 7(b) corresponds the value ψi = 0.451 [6] (in accordance with the theory [27,28]). (One can assertain, that π-state in Fig. 7(b) at R coincides → ∞ G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 89 with the (s, n)-interface, considered by Ginzburg and Landau [1].) Thus, in the infinite system the Bogomolnyi equations (with fixed m = 1) describe the isolated vortex with its field bξ(x) falling to zero at “infinity” (i.e. far away from the vortex center). However, in a large, but finite system the question arises as to what is going on “beyond the infinity”, i.e. near the remote boundary. Because in a finite system at r = R there is always the external field H = 0 (recall, that p-states arise in n -mode 6 0 from a supercooled n-state while the large external field diminishes), then it is clear, that the Bogomolnyi equations (with fixed value m = 1) cannot be obtained from considering the finite cylinder, even in the limit R . Therefore, in a bulk (but finite) system the → ∞ self-consistent GL-solutions with fixed m = 1 (and κ = κ0) are represented in fact by a superposition of independent Bogomolnyi solutions with m = 1 and m = 0 (analogously for m> 1).

The analogous situation is near µ-line (Fig. 5) for µ-states (in s0-mode, FI-regime).

In Fig. 8(a) the self-consistent GL-solutions ψ(x) and bξ(x) are shown (for m = 1,

κ = 0.938, Rλ = 12). As in Fig. 7(a), here also exist two “infinite” branches 1 and 2,

divided by the maximum point of the function ψ(x) (or the minimum point of bξ(x)).

The branch 1 describes the isolated vortex (m = 1), with a finite field b0 at the cylinder axis, and zero field far from the vortex core (i.e. at “infinity”). The branch 2 describes

the vortex-free µ-state (m = 0), where the field at the cylinder boundary is bξ(R) = 1, and far away (i.e. near the cylinder center) the field vanishes (see Fig. 8(b)). Notice, that the vortex states 1 in Figs. 7(a) and 8(a) are different. So, in π-state 1 in Fig. 7(a) (κ κ ) we have b (0) = 1, but in µ-state 1 in Fig. 8(a) (κ κ ) we have b (0) = 0.6724. ≈ 0 ξ ≈ c ξ The branches 2 in Figs. 7(a) and 8(a) also differ. So, in π-state 2 in Fig. 7(a) we have ψ(R) 0, but in µ-state 2 in Fig. 8(a) we have ψ(R) = 0.5928. The question whether ≈ the limiting µ-solutions at κ = κc are subject to some degenerate equations (as is the case

for π-solutions at κ = κ0) remains open.

To conclude this Section, we stress once more that the critical values κ0 and κc in Fig. 5 (for R 1, m = 1) divide all superconductors into two groups or types in each of λ ≫ two field regimes. In superconductors with κ > κc (group-II, FI-regime) the destruction of a superconducting state happens with a jump to e-state, followed by a second order phase transition e n. In superconductors with κ < κ (group-I, FI-regime) the transition M → c n is a first order phase transition. In superconductors with κ > κ (type-II, FD- → 0 regime) the nucleation of superconductivity begins by a second order phase transition n e, followed by the restoration of M-state. In superconductors with κ < κ (type-I, → 0 FD-regime) the complete restoration of M-state happens immediately by a first order phase transition. As a result, the division of superconductors into groups and types has some sense only in the parameter interval κ0 < κ < κc. Outside this interval the terms “type” or “group” are equivalent. In other words, in GL-equations three intervals of parameters κ should be distinguished: κ < κ , κ κ κ and κ > κ . In each of these 0 0 ≤ ≤ c c intervals the solutions of one-dimensional GL-equations behave in the external magnetic

field differently. [One can prove that the critical parameters κ0 and κc are the singular points of the GL-equations, when different (in the general case) values of the bifurcation 90 G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103

> κ κ > κ κ fields coincide (h1 = hp = h2 at = 0 and h1 = h2 at = c; see Figs. 2(a,c,e).]

8 The jumps of total flux and free energy

Apart from the magnetization jumps, described above, the jumps of the total flux and free energy of the system are also of considerable interest. Using the notations of Sec. 2, the total flux in the system is

Φ= Adl = Bds = φ0(aR + m), (9) IR ZS where S = πR2 and φ = 2 10−7G cm2 is the flux quantum. So, the value of the potential 0 · · function a a(R) = Φ/φ m equals the total flux inside the cylinder minus m. R ≡ 0 − The behavior of the total flux (9) and the Gibbs free energy (5) are shown in Fig. 9 as functions of the external field hλ = H/Hλ for a cylinder with Rλ =6 at κ = 0.5 and

κ = 1.5 (m = 1). The critical fields (h1, hr) and different states of the system (M, e, d, n, n, n∗,n ˙ ) are marked. The hysteresis loops are indicated by vertical arrows. As can be seen from Fig. 9(a) (κ = 0.5), superconducting cylinder in FI-regime passes to normal n-state in a jump at the field h> 0.51. Consequently, the total flux 1 ≈ contained inside the cylinder, increases in a jump by a quantity ∆Φ 4φ0. In FD-regime ≈ > the jump from a supercooled n-state to M-state take place at the field hr = 0.3 and thus the flux diminishes by a quantity ∆Φ 2φ0. The analogous jumps happen on the left ≈ < < branches of the curves (in negative fields h1 and h˙ r). Here, in FD-regime the jump M n happens at the field 0, imply the instability (more exactly, metasta- > < bility) of a superheated (at the field h1 ) or supercooled (at the field h1) superconducting state relative the transition to normal (n or n∗) state. It is seen, that with κ increasing, such metastable states become more stable. [The presence of the metastable states with ∆g > 0 was noticed for the first time by Fink and Presson [20]. One can verify, that the states with ∆g > 0 are due to the large values of the diamagnetic moment of the currents, which flow near the cylinder boundary and prevent the external field penetra- tion inside the superconductor (in formula (6) the term with Mλ introduces large positive contribution)].

The behavior aR(h) and ∆g(h) for smaller value Rλ = 3 is shown in Fig. 10. One can see that with Rλ diminishing, the jump amplitudes diminish also (they vanish finally at < h1 = 1.22). In Figs. 10(a,b) the points h0 are marked, where the magnetization is zero, G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 91

M(h0) = 0; the paramagnetic and diamagnetic regions are also shown (compare with Fig. 3). Figs. 10(a,c) correspond to a field-stimulated superconducting M-state, which exists only at finite values h> 0.

9 Conclusions

In conclusion we summarize the main results of the present investigation. In this work we have studied the predictions in detail, which follow from the GL-theory for superconducting cylinders in a magnetic field, when a single vortex is situated at the cylinder axis. We found, that at the same external field H two different superconducting states can exist in a cylinder of radius R, which have different space configurations of the order parameter ψ(x) and of the field inside the cylinder B(x). These states relate to the experimentally well known hysteresis phenomenon. With the field increasing (or decreasing) the transitions between these states become possible. Such transitions can proceed either by the way of the irreversible first order phase transition (in a jump at the instability point, as the transitions M e), or reversibly by second order phase → transition (as the transitions e n, or n p). As was said above, the presence of → → transitions M e confirms further the existence of the additional “edge” mechanism for → the field penetration into a finite extent superconductor, which is different from the usual “vortex” mechanism. The experimental determination of such “edge” mechanism may be important in view of the attempts to manufacture small measuring devices on the basis of mesoscopic superconductors. The critical fields of the cylinder are determined (as functions of its radius R and parameter κ, Fig. 4), which correspond to the destruction of superconductivity in the increasing field (FI) regime, or to restoration of superconductivity in the decreasing field (FD) regime. The interval of fields is found where it is possible to observe the paramag- netic effect, i.e. the positive magnetization M(H) > 0, due to the presence of the vortex (m = +1) at the cylinder axis. The critical lines on the plane of parameters (κ,R) (or, the state diagram, Fig. 5) are found, which mark the various peculiarities in the behavior M(H), including the hysteresis boundaries and the boundaries of the first and second order phase transitions. It is shown, that at R λ and m = 1 (as in the case m = 0 [6,7]) in the GL-theory there ≫ exist two critical parameters (κ0 = 1/√2 and κc = 0.93, Fig. 5), which correspond to two possible regimes (FD and FI) of the field action on a superconductor and to different types of hysteresis. We should stress, however, that these (and other) results were obtained on the base of one-dimensional solutions of GL-equations, without the detailed analysis of the longitu- dinal (along z-axis) or transverse (along the radius) stability of the solutions. It is rather evident, that with R increasing, the edge e-layer would become unstable and break into separate regions, containing the vortices, which gradually would fill up the cylinder inte- rior and ensure the system transition to the ordinary vortex state [2–5]. So, the e-states, apparently, can be realized only in the cylinders of sufficiently small radius. However, 92 G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 the detailed analysis of the stability of the solutions in finite extent systems (accounting for the arbitrary positions of the vortex axes and for possible hysteresis states) demands using the partial differential equations,which is beyond the scope of the present investi- gation. (Some examples of such calculations for superconducting disks of small thickness can be found in [29] and other literature cited therein). Thus, it is clear, that further theoretical and experimental study of the questions touched upon in the present paper are indicated as necessary.

Acknowledgment

I am grateful to V.L. Ginzburg for his interest in this work and valuable comments, and also to V.G. Zharkov, A.Yu. Tsvetkov, A.L. Dyshko and N.B. Konukhova for useful discussions. The work was supported through the grant RFFI 02-02-16285.

References

[1] V.L. Ginzburg and L.D. Landau: “To the theory of superconductivity”, Zh.Exp.Teor.Fyz., Vol. 20, (1950), pp. 1064–1082. [2] A.A. Abrikosov: Fundamentals of the Theory of Metals, North-Holland, Amsterdam, 1988. [3] M. Tinkham: Introduction to Superconductivity, McGraw Hill, New York, 1975. [4] P.G. de Gennes: Superconductivity of Metals and Alloys, Addison-Wesley, New York, 1989. [5] D. Saint-James, G. Sarma and E.J. Thomas: Type II superconductivity, Pergamon, Oxford, 1969. [6] G.F. Zharkov: “First and second order phase transitions and magnetic hysteresis in a superconducting plate”, J.Low Temp.Phys, Vol. 130(1/2), (2003), pp.45–67. [7] G.F. Zharkov: “Critical fields of a superconducting cylinder”, Centr. Europ. Journ. Phys., Vol. 2(1), (2004), pp. 220–240 (Erratum: ibid., Vol. 2(2), (2004), p. 419). [8] A.Yu. Tsvetkov, G.F. Zharkov and V.G. Zharkov: “Superconducting plate in a magnetic field”, Krat. Soob. Fys. FIAN, Vol. 1/2, (2003), pp. 42–50. [9] G.F. Zharkov, V.G. Zharkov and A.Yu. Tsvetkov: “GL-calculations for superconducting cylinder in a magnetic field”, Phys.Rev.B, Vol. 61, (2000), pp. 12293– 12312. [10] G.F. Zharkov: “The emergence of superconductivity and hystersis in a type-I superconducting cylinder”, Zh.Exp.Teor.Fys., Vol. 122, (2002), pp. 600–609. [11] G.F. Zharkov, V.G. Zharkov and A.Yu. Tsvetkov: “Self-consistent solutions of GL- equations and superconducting edge-states in a magnetic field”, Krat. Soob. Fys. FIAN, Vol. 11, (2001), pp. 35–48; “One-dimensional solutions of GL-equations for superconducting cylinder in magnetic field”, ibid, Vol. 12, (2001), pp. 31–38. [12] G.F. Zharkov: “Transitions of I- and II-order in magnetic field for superconducting cylinder obtained from self-consistent solution of GL-equations”, Phys.Rev.B, Vol. 63, (2001), pp. 224513–224519. G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 93

[13] G.F. Zharkov: “Onset of superconductivity in magnetic field for a long cylinder as obtained from a self-consistent solution of GL-equations”, J.Low Temp.Phys., Vol. 128(3/4), (2002), pp. 87–113. [14] D. Saint-James and P. de Gennes: “Onset of superconductivity in decreasing field”, Phys.Lett., Vol. 7, (1963), pp. 306–308.

[15] D. Saint-James: “Etude du champ critique Hc3 dans une geometrie cylindrique”, Phys.Lett. Vol. 15, (1965), pp. 13–15. [16] M.A. Abramovitz and I.A. Stegun: Handbook of Mathematical Functions, Dover, New York, 1970. [17] G.F. Zharkov: “Paramagnetic Meissner effect in superconductors from self-consistent solution of GL-equations”, Phys.Rev.B, Vol. 63, (2001), pp. 214502–214509. [18] G.F. Zharkov and V.G. Zharkov: “Magnetic vortex in superconducting wire”, Physica Scripta, Vol. 57, (1998), pp. 664–667. [19] L.D. Landau and E.M. Lifshits: Fluid Mechanics, Oxford, Pergamon Press, 1982. [20] H.J. Fink and A.G. Presson: “Superheating of the Meissner state and the giant vortex state of a cylinder of finite extent”, Phys.Rev., Vol. 168, (1968), pp. 399–402. [21] J.G. Park: “Critical currents and surface superconductivity”, Phys.Rev.Lett., Vol. 16, (1966), pp. 1196–1200. [22] J. Feder: “Comments on the supercooling field for superconductors with κ values near 0.4”, Sol.State.Comm., Vol. 5, (1967), pp. 299–301. [23] J.G. Park: “Metastable states of the superconducting surface sheath in decreasing fields”, Sol.State.Comm., Vol. 5, (1967), pp. 645–648. [24] P.V. Cristiansen and H. Smith: “ Ginzburg–Landau theory of surface superconductivity and magnetic hysteresis”, Phys. Rev., Vol. 171, (1968), pp. 445– 458. [25] J.P. McEnvoy, D.P. Jones and J.G. Park: “Nucleation of superconductivity in tantalum in a decreasing magnetic field”, Phys.Rev.Lett., Vol. 22, (1969), pp. 229–231. [26] E.B. Bogomolnyi: “Stability of classical solutions”, Yad.Phys., Vol. 24, (1976), pp. 861–870; Sov.J.Nucl.Phys., Vol. 24(449), (1976). [27] A.T. Dorsey: “Dynamics of interfaces in superconductors”, Annals of Physics, Vol. 233, (1994), pp. 248–269. [28] I. Luk’yanchuk: “Theory of superconductors with κ close to 1/√2 ”, Phys.Rev.B, Vol. 63, (2001), pp. 174504-174517. [29] S.V. Jampolsky and F.M. Peeters: “Vortex structure of thin mesoscopic disks with enhanced surface superconductivity”, Phys.Rev.B, Vol. 62, (2000), pp. 9663–9674. 94 G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103

Fig. 1 The co-ordinate dependencies of solutions ψ(x) and bλ(x) (x = r/R) for a cylinder with Rλ = 6 and different κ. Solid curves correspond to s0-mode (FI-regime). Dotted curves correspond to n0-mode (FD-regime). (a,b) – κ = 0.5; (c,d) – κ = 0.8; (e,f) – κ = 1.5. Shown are the states M, p, d, e, n, n, and also M-solution in the abscence of the field (h = H/Hλ = 0). The nucleation field of p-state is hp = 0.6420. G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 95

Fig. 2 The field dependencies ψmax(h) and 4πMλ(h) (h = hλ) for Rλ = 6 and different κ: (a,b) > – κ = 0.5; (c,d) – κ = 0.8; (e,f) – κ = 1.5. The critical fields h1 (s0-mode, FI-regime, the < right branch of the curves with h > 0) and h1 (s0-mode, FD-regime, left branch of the curves > < with h < 0) are different. The critical fields hr (n0-mode, FD-regime, right branch) and hr (n0-mode, FI-regime, left branch) also differ. Shown are the regions of metastable supercooled (n) and superheated (n∗) states, of the absolutely stable n-state, and also the states p, d and e. In addition, the absolutely unstablen ˙ - and u-states are shown. The field h0 divides the paramagnetic and diamagnetic regions, at this field M(h0) = 0. In the absence of the field (h = 0) the vortex state is always paramagnetic, Mλ > 0. For h < 0 the vortex state (m = 1) is always diamagnetic (Mλ < 0, magnetization is directed along the field). 96 G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103

Fig. 3 The field dependencies ψmax(h) and 4πMλ(h) for Rλ = 3: (a,b) – κ = 0.5; (c,d) – κ = 0.8; (e,f) – κ = 1.5. The labels at the curves are analogous to that in Fig. 2. The absolutely unstable u-solutions are not shown. At κ = 0.5 superconductivity is destroyed by a second order > < phase transition in both regimes FD (h2 ) and FD ( h2). At κ = 1.5 there is a second order > < phase transition in FI-regime (h2 ) and a first order phase transition in FD-regime ( h1). G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 97

Fig. 4 The critical fields (phase diagrams) of the cylinder with the vortex at the cylinder axis as functions of the radius R for different κ: (a) – κ = 0.5; (b) – κ = 0.8; (c) – κ = 1; (d) – κ = 1.5. The notations are explained in the text. 98 G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103

11 ν π 10 µ R 9 m=1 λ A B C 8 7 æ =0.707 6 0 µ 1 5 o æ =0.93 4 ζ> c ζ> o µ 3 2 G o D o < 2 n ζ ν E 1 i 0 0,0 0,2 0,4 0,6 0,8 1,0æ 1,2 1,4 1,6 1,8 2,0

Fig. 5 The critical lines (state diagram) on the plane (κ,Rλ). The notations are explained in the text. G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 99

Fig. 6 The comparison of the critical fields for m = 0 (dotted lines) and m = 1 (solid lines). (a) – κ = 0.5; (b) – κ = 1.2. The notations are explained in the text. In the case m = 1 beneath the line Rξ = 1.32 only the normal state exists. 100 G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103

Fig. 7 Self-consistent π-solutions ψ(x) and bξ(x) (x = r/R). (a) – Supercooled vortex π-state (m = 1, κ = 0.708, Rλ = 14.1); the label divides the curves into two branches 1 and 2, which correspond to m = 0 and m = 1. (b) –∞ Superheated vortex-free π-state (m = 0, κ = 0.7075, Rλ = 12). G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 101

Fig. 8 Self-consistent µ-solutions ψ(x) and bξ(x) (x = r/R). (a) – Superheated vortex µ-state (m = 1, κ = 0.938, Rλ = 14.1). (b) – Supercooled vortex-free µ-state (m = 0, κ = 0.94, Rλ = 12). The label divides the curves into two branches 1 and 2, which correspond to m = 0 and m = 1. ∞ 102 G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103

Fig. 9 (a,b) – The jumps of the potential function aR = a(R) (or, of the total flux Φ/φ0 = aR + m), and (c,d) – of the free energy ∆g, for Rλ = 6, κ = 0.5 and κ = 1.5 (compare with Fig. 2). The dotted lines in Figs. 9(a,b) (which correspond to normal state) are described by 2 the function aR = φ m with φ Φ/φ0 = hλR /2. In the field h0 the magnetization vanish, − ≡ λ M(h0) = 0. G.F. Zharkov / Central European Journal of Physics 3(1) 2005 77–103 103

Fig. 10 (a,b) – The dependencies of the potential function aR(hλ), and (c,d) – of the free energy ∆g(hλ) for Rλ = 3, κ = 0.5 and κ = 1.5 (compare with Fig. 3). The dotted lines in Fig. 10(a,b) 2 (which correspond to normal state) are described by the function aR = φ m with φ = hλRλ/2. The regions of para- and diamagnetism at m = 1 are marked. −