1. Introduction
The superconducting state is characterized by two hallmarks: the vanishing electrical resistance below the superconducting transition temperature,
, and the Meissner-Ochsenfeld effect [
1], describing the expulsion of magnetic flux from the superconducting sample when cooling it in an applied magnetic field (field cooling, FC), creating a diamagnetic state. The implications of the Meissner-Ochsenfeld effect led directly to the development of the basic theories of superconductivity (London, Ginzburg-Landau and BCS), and are intensively described in all textbooks on superconductivity (see, e.g., [
2,
3,
4,
5,
6,
7,
8,
9,
10]). Thus, the first observations of superconducting transitions of Bi-based, high-
superconductors (HTSc) to a paramagnetic state instead of a diamgnetic one were more treated as experimental mishaps and went mostly unnoticed by the community [
11,
12]. The situation changed with more detailed measurements on granular Bi
Sr
CaCu
O
(Bi-2212) samples by Braunisch
et al. [
13,
14], linking the oservation of a superconducting transition towards the paramagnetic state with unique features of the HTSc, i.e., the so-called d-wave superconductivity and effects of granularity (
-junctions between the grains). This was soon followed by others, applying also different measurement techniques to exclude possible experimental artifacts and providing some theories to explain these observations [
15,
16,
17,
18,
19,
20,
21,
22,
23,
24]. Following these works, several theoretical approaches were published concerning this effect, now named paramagnetic Meissner effect (PME) or Wohlleben effect [
18].
Thus, it came as a big surprise as Thompson
et al. presented an observation of PME on bulk niobium disks, a classical s-wave superconductor [
25,
26], or often called conventional or low-
superconductor (LTSc). This work was soon followed by Kostić
et al. presenting a thorough investigation of the PME in Nb materials [
27]. This work resulted in a comment [
28] and a reply [
29], where the inherent differences between the PME in metallic, s-wave superconductors and the HTSc were clarified. Furthermore, several reports presented details of the superconducting transitions on different Nb samples with the magnetic field applied in parallel and perpendicular directions [
26,
30], the vanishing of the PME by surface treatments [
27,
31] and the enhancement of the PME by ion implantantion [
32]. In this way, a new research direction was born.
The isotropic LTSc used for these studies were compact, bulk and homogeneous materials in stark contrast to HTSc that are typically granular materials with their inherent complicated crystal structures, mostly tetragonal ones. Therefore, the LTSc may serve – owing to their relative simplicity – as a superconducting model system to perform detailed studies in order to clarify the physical origin of the paramagnetic moment appearing when crossing through from higher temperatures.
Since then, the PME was observed in a variety of metallic superconductors [
33,
34,
35,
36,
37,
38,
39,
40,
41,
42,
43,
44,
45,
46,
47,
48], and in different forms like thin films [
49,
50,
51,
52], Pb nanowire arrays [
53], multilayer systems [
54,
55,
56,
57,
58,
59,
60], and very importantly, in nanocomposites [
61,
62], and mesoscopic structured samples [
63,
64,
65,
66,
67]. Of course, since the discovery of MgB
– the metallic superconductor with the highest transition temperature [
68,
69] – it was only a matter of time that reports of PME in this system appeared in the literature as well [
70,
71,
72,
73,
74,
75]. Very recently, PME was also observed in superconducting boron-doped diamond thin films [
76]. All these observations comprise a large variety of superconducting materials and in various shapes. Thus, it is the aim of the present review to summarize all these experimental efforts, which will contribute to a better comprehension of the observations of the PME in metallic superconductors.
Several theoretical approaches [
77,
78,
79,
80,
81,
82,
83,
84,
85,
86,
87,
88,
89,
90,
91,
92,
93,
94,
95,
96,
97,
98,
99] to explain the PME were already reviewed by Li [
100], but only a short summary of the experimental data were given in this article. Therefore, the present review focuses on the experimental observations of the PME in metallic superconductors in bulk or mesoscopic forms. We further work out the distinct differences between the PME of HTSc and the metallic ones, and discuss the experimental difficulties for the observation of the PME.
This manuscript is organized as follows: In
Section 2, we present the various experimental observations found in the literature.
Section 2.1 gives an introduction to the PME found first in polycrystalline Bi-2212 HTSc samples and compares these reults to the observations of the PME on bulk Nb disks. A third type of PME was observed in YBCO thin films, patterned YBCO structures and YBCO nanowires, which has only some features in common with the other two cases. In
Section 2.2, the observations of paramagnetic signals upon field-cooling of various metallic superconductors are reviewed. Then, in
Section 2.3, we discuss the various methods applied to observe PME, which is very important to understand the data published in the literature.
Section 3 focusses on details of the specific experiments performed in Detroit, Tokyo and Nancy on bulk Nb disks to elucidate the nature of the PME. Firstly, the parameters of the investigated samples are presented in
Section 3.1. The measurements described in detail include the
-behavior with PME (
Section 3.2), the magnetic hysteresis loops at temperatures close to
(
Section 3.3) and discusses the experiments performed in the literature to enhance or reduce the PME (
Section 3.4). Furthermore, the following sections
Section 3.5 present investigations of PME as function of time, and
Section 3.6 presents AC susceptibility measurements on the Nb disks. In
Section 4, the attempts to imagine the giant vortex state and flux structures close to the superconducting transition are discussed. In
Section 5, the most important models to describe the PME in the metallic superconductors are outlined. Finally,
Section 6 presents the conclusions and an outlook to further investigations of the PME.
4. Magnetic Imaging
Several different imaging techniques were already applied to superconductors exhibiting PME, among them are magneto-optic imaging [
8,
142,
143], the scanning SQUID technique [
144,
145], the imaging using color centers in diamond [
146,
147] and low-energy muon spin spectroscopy (LE-
SR) [
56].
We will start the discussion with the scanning SQUID experiments performed by Kirtley et al. [
104] on HTSc Bi-2212 samples.
Figure 18a–f present the spatial distribution of magnetic flux in and around the HTSc sample. The grey contrast scale is chosen so that white corresponds to the largest and black to the smallest (often negative) flux value. In all cases the flux is plotted relative to the flux introduced by the external field, which sets the grey level outside the sample. One overall feature, which is observable by the human eye, is the difference between the paramagnetic magnetization (i.e., the sample is brighter than the background) at weak applied magnetic fields, and a diamagnetic signal (i.e., the sample is darker than the background), in the pickup loop at strong applied magnetic fields. For weak external magnetic fields the inhomogeneity of the magnetic flux is clearly visible and gives rise to a broad distribution of the local fluxes as the sample is a polycrystalline bulk. Here, it must be noted that all images were taken at
4.2 K (allowing the performance of the SQUID loop in the same condition as the sample). The sample was cooled through
with the external magnetic field applied, and the field was still on when reaching 4.2 K. Thus, in this form of the scanning SQUID technique, measurements in the very interesting region around
cannot be performed. Nevertheless, the images obtained nicely reproduce the situation shown in
Figure 1b, where the fields smaller than 30
T (0.3 G) produce a positive
, and fields of 100
T (1 G) and 300
T (3 G) yield a diamagnetic signal. The authors of [
104] claim that the images provide direct evidence for spontaneous orbital currents, but it is not possible to determine the origin of such orbital currents from the images. An attempt to repeat the same scanning SQUID experiment on the Nb disks from Detroit is represented in
Figure 18g,h.
Figure 18g gives the result obtained at
4.2 K when field-cooling the sample (abraded Nb disk without PME) in a field of ∼1.6
T (1.6 mG). The image shows many vortices (white and black colors denote vortices of opposite direction), which are quite well separated from each other (annihilation effects). In contrast, the same image of an Nb disk with PME (h) taken at ∼2.8
T (2.8 mG) reveals a much larger number of vortices being present within the image frame, and also the number of black vortices is much higher than the number of white vortices. Furthermore, there is a field gradient from left to right, but the two types of vortices are not so well separated as in (g). Hence, the scanning SQUID technique can produce very interesting images of the vortex distribution, but the temperature limitation to 4.2 K does not allow an observation of the giant vortex state around
, i.e. close to
.
Figure 18.
Scanning SQUID microscope images. (
a–
f) present images of a granular Bi-2212 sample, cooled and imaged in various applied magnetic fields. Each image comprises 512 × 512 pixels, and each pixel represents a distance of 6
m. The individual images are labelled by the cooling field, and by the maximal range of variation of the flux (in units of the flux quantum,
) in each case. The small rectangles indicate where profiles were taken in Ref. [
104], which are not shown here. All images were taken at
4.2 K. (
a). Cooling field 1
T (0.01 G), 0.21
. (
b). Cooling field 3
T (0.03 G), 0.21
. (
c). Cooling field 10
T (0.1 G), 0.21
. (
d). Cooling field 30
T (0.3 G), 0.39
. (
e). Cooling field 100
T (1 G), 1.65
. (
f). Cooling field 300
T (3 G), 7.6
. (
g)–(
h) give images obtained on a Nb disk from Detroit at
4.2 K. (
g). Sample with no PME, applied field of ∼ 1.6
T (1.6 mG). (
h) sample with strong PME, applied field ∼ 2.8
T (2.8 mG). Images (a)–(f) reproduced with permission from Ref. [
104], images (g) and (h): Courtesy of J.R. Kirtley.
Figure 18.
Scanning SQUID microscope images. (
a–
f) present images of a granular Bi-2212 sample, cooled and imaged in various applied magnetic fields. Each image comprises 512 × 512 pixels, and each pixel represents a distance of 6
m. The individual images are labelled by the cooling field, and by the maximal range of variation of the flux (in units of the flux quantum,
) in each case. The small rectangles indicate where profiles were taken in Ref. [
104], which are not shown here. All images were taken at
4.2 K. (
a). Cooling field 1
T (0.01 G), 0.21
. (
b). Cooling field 3
T (0.03 G), 0.21
. (
c). Cooling field 10
T (0.1 G), 0.21
. (
d). Cooling field 30
T (0.3 G), 0.39
. (
e). Cooling field 100
T (1 G), 1.65
. (
f). Cooling field 300
T (3 G), 7.6
. (
g)–(
h) give images obtained on a Nb disk from Detroit at
4.2 K. (
g). Sample with no PME, applied field of ∼ 1.6
T (1.6 mG). (
h) sample with strong PME, applied field ∼ 2.8
T (2.8 mG). Images (a)–(f) reproduced with permission from Ref. [
104], images (g) and (h): Courtesy of J.R. Kirtley.
The flux compression effects and the giant vortex state would be interesting objects to be directly observed by magnetic imaging methods. However, there are two main problems: The required high field sensitivity and the low temperatures of the superconducting transition of Pb and Nb. For magneto-optic imaging [
8,
142,
143], the spatial sensitivity depends on the distance of the indicator layer to the sample surface, and as the sample surface cannot be treated by polishing, the resulting distance is quite large, thus leading to reduced spatial resolution. This also affects the field sensitvity, which must be quite high to resolve single flux quanta. The temperature problem, which is often seen in magneto-optic cryostat systems operating with a Helium gas stream is nowadays nicely overcome by optical cryostats being cryocooled which enable even much lower temperatures to be reached while the sample being illuminated [
148]. After ensuring the presence of the PME on the Nb disks in Nancy, we also performed magneto-optic imaging in Liège (group of Prof. Silhanek, [
149]).
Figure 19a,b present the MO imaging results on a Nb square cut from the original Nb sheet. The applied magnetc field is 5 mT, which was kept on during the temperature sweep from 10 K to about 8 K.
Figure 19a gives the MO image, where the outer edge indicates the edges of the MO indicator film, and the inner L-shape indicates the edges of the Nb sample.
Figure 19b gives the recorded flux in the two yellow boxes shown in (a). However, this experiment does not reveal a clear difference between the outside and the inside, which means that the available field sensitivity is not high enough to reveal a signature of a giant vortex state. Thus, further new experiments are required to achieve more information on the giant vortex state.
Another interesting experiment was carried out by Nusran et al. [
150] (
Figure 20) employing the non-invasive magnetic field sensing using optically-detected magnetic resonance of nitrogen-vacancy centers in diamond, short NV magnetometry. The experimental apparatus incorporates a confocal microscope optimized for NV fluorescence detection. The fluorescence is stimulated by the green off-resonant 532 nm laser excitation and low-energy levels are populated by the microwave radiation applied using a single silver wire loop antenna coupled to a MW frequency generator. A thin diamond plate with an ensemble of NV centers embedded near the surface (∼20 nm depth) is used as the magneto-optical sensor. The spatial resolution of the sensor is determined by the effective size of the probe, which is essentially a convolution of the focal volume with the NV distribution in the diamond plate. This leads to a disk-like probing volume of thickness ≈20 nm and diameter of ≈500 nm. A detailed review of the NV-centers and NV magnetometry can be found in Refs. [
146,
147].
Figure 19.
MO images obtained at University of Liège of an Nb disk from Detroit. (a). The MO image at 5 mT applied field, and the temperature is swept from 10 K to about 8 K. The top L-shape indicates the MO indicator film, the inner L-shape represents the sample edges. The two yellow boxes give the regions (, ), where the profiles are taken for (b). (b). Profiles taken in the two yellow boxes marked in (a). The black curves are for , and the red curves for .
Figure 19.
MO images obtained at University of Liège of an Nb disk from Detroit. (a). The MO image at 5 mT applied field, and the temperature is swept from 10 K to about 8 K. The top L-shape indicates the MO indicator film, the inner L-shape represents the sample edges. The two yellow boxes give the regions (, ), where the profiles are taken for (b). (b). Profiles taken in the two yellow boxes marked in (a). The black curves are for , and the red curves for .
Figure 20.
Diamond center imaging by AMES on an Nb sample with PME. (
a). Temperature-dependent total magnetic moment measured using Quantum Design MPMS. Shown are ZFC-W and FC-C curves measured in Nb. (
b). FC profiles of the magnetic induction for applied magnetic fields of 1.0, 1.2, and 5 mT (10, 12, and 50 Oe) recorded for an Nb disk-shaped crystal. Measurement performed at
4.2 K. The signatures of the paramagnetic Meissner effect (PME) are observed in various random regions for low magnetic fields. Cooling in 1 mT (10 Oe, red datapoints) and 1.2 mT (12 Oe, green datapoints) applied magnetic fields results in very similar profiles. The inset to (b) presents a temperature-resolved measurement in a 1 mT (10 Oe) magnetic field upon FC-C. The peak (‘
P’) and valley positions (‘
V’) of the inset are indicated by arrows in the main panel. Image reproduced with permission from Ref. [
150].
Figure 20.
Diamond center imaging by AMES on an Nb sample with PME. (
a). Temperature-dependent total magnetic moment measured using Quantum Design MPMS. Shown are ZFC-W and FC-C curves measured in Nb. (
b). FC profiles of the magnetic induction for applied magnetic fields of 1.0, 1.2, and 5 mT (10, 12, and 50 Oe) recorded for an Nb disk-shaped crystal. Measurement performed at
4.2 K. The signatures of the paramagnetic Meissner effect (PME) are observed in various random regions for low magnetic fields. Cooling in 1 mT (10 Oe, red datapoints) and 1.2 mT (12 Oe, green datapoints) applied magnetic fields results in very similar profiles. The inset to (b) presents a temperature-resolved measurement in a 1 mT (10 Oe) magnetic field upon FC-C. The peak (‘
P’) and valley positions (‘
V’) of the inset are indicated by arrows in the main panel. Image reproduced with permission from Ref. [
150].
Figure 20a presents
measured on a Nb disk-shaped crystal using a Quantum Design MPMS SQUID system. Shown are the ZFC-W and FC-C curves measured for fields of 1 and 5 mT (10 and 50 Oe). Both curves reveal the PME, but no details of the superconducting transition as seen in
Figure 5.
Figure 20b presents recorded FC profiles of the magnetic induction for applied magnetic fields of 1.0, 1.2, and 5 mT (10, 12, and 50 Oe). Measurement performed at
4.2 K. The signatures of the paramagnetic Meissner effect (PME) are observed in various random regions for low magnetic fields. Cooling in 1 mT (10 Oe, red datapoints) and 1.2 mT (12 Oe, green datapoints) applied magnetic fields results in very similar profiles. The inset to (b) presents a temperature-resolved measurement in a 1 mT (10 Oe) magnetic field upon FC-C. The peak (‘
P’) and valley positions (‘
V’) of the inset are indicated by arrows in the main panel.
To investigate the Meissner effect in a superconductor/magnet system, the depth profile of the local magnetic susceptibility of a Au(27.5 nm)/Ho(4.5 nm)/Nb(150 nm) trilayer was measured by low-energy muon spin spectroscopy (LE-
SR) [
56]. The antiferromagnetic rare-earth metal Ho breaks time-reversal symmetry of the pair correlations in Au and has a thickness that is comparable to the known coherence length for singlet pairs in Ho to ensure pair transmission into Au. The Au layer is necessary since a Meissner state cannot be probed by muons directly in a magnetic material due to their rapid depolarization in a strong magnetic field. (LE-
SR) offers extreme sensitivity to magnetic fluctuations and spontaneous fields of less than 10
T (0.1 G) with a depth resolved sensitivity of a few nanometers [
151,
152]. To probe the depth dependence of the Meissner response in Au/Ho/Nb by LE-
SR, an external magnetic field (
) is applied parallel to the sample plane and perpendicular to the muon initial spin polarization (oriented in the
plane). To investigate the paramagnetic Meissner effect, implantation energies in the 3–6 keV range are used to determine the
profile in the Au layer.
Figure 21a–c present the results from such an experiment. The measurement at 3 K shows the most significant PME response (labelled here Inverse Meissner) in the Au layer. A conventional Meissner effect (i.e., negative
) is measured in Nb up to the interface with Ho, where the contribution of spin-singlet Cooper pairs to the screening supercurrent
is larger than that due to the long-ranged spin-triplet pairs. The advantage of this technique is the high spatial resolution
inside the trilayer structure, but requires the antiferromagnetic Ho layer to break time-reversal symmetry of the pair correlations in Au.
Figure 21.
Average local magnetic field in Au/Ho/Nb trilayers as a function of the muon implantation energy and mean stopping distance. (
a).
values from single-energy asymmetry fits versus implantation energy (bottom
x-axis) and mean stopping distance (top
x-axis) in the normal state (
•,
10 K) and superconducting state (
•,
5 K). (
b). Data showing the inverse Meissner state in Au. The continuous lines are a guide to the eye. (
c). Magnetization and screening current profile from the Ho/Nb interface in Au/Ho/Nb. Local magnetic field
determined as global-energy fit of the LE-
SR measurement data (
, left
y-axis) and theoretical model (
, left
y-axis); calculated dimensionless screening current density,
flowing parallel to the
x-axis in (a) inside the plane of the thin film heterostructure (gray curve, right side
y-axis). Dashed lines show that the position of the maximum in
coincides with that of the null in
. Image reproduced with permission from Ref. [
56].
Figure 21.
Average local magnetic field in Au/Ho/Nb trilayers as a function of the muon implantation energy and mean stopping distance. (
a).
values from single-energy asymmetry fits versus implantation energy (bottom
x-axis) and mean stopping distance (top
x-axis) in the normal state (
•,
10 K) and superconducting state (
•,
5 K). (
b). Data showing the inverse Meissner state in Au. The continuous lines are a guide to the eye. (
c). Magnetization and screening current profile from the Ho/Nb interface in Au/Ho/Nb. Local magnetic field
determined as global-energy fit of the LE-
SR measurement data (
, left
y-axis) and theoretical model (
, left
y-axis); calculated dimensionless screening current density,
flowing parallel to the
x-axis in (a) inside the plane of the thin film heterostructure (gray curve, right side
y-axis). Dashed lines show that the position of the maximum in
coincides with that of the null in
. Image reproduced with permission from Ref. [
56].
5. Discussion
After the presentation of the various PME results obtained on a large variety of metallic superconductors, it is necessary to discuss the various models applied to explain the PME in these samples. The most important models being applied to the PME of metallic superconductors are the flux compression [
85] and the giant vortex state [
88,
89]. As shown by Koshelev [
85], a kind of inhomogeneous cooling of a superconducting sample, i.e., when field-cooling a sample, vortices may be expulsed from the sample along the edges, and other flux is then trapped in the sample center. Continuous cooling then leads to a broadening of the flux-free regions and compresses the flux in the center even further by the vortex Nernst effect as described by Huebener [
5]. The same picture would apply if the sample surfaces have a higher
than the remaining bulk of the sample. Such situations may be realized by specific sample surfaces, e.g., due to oxidation effects [
140].
Figure 22.
Magnetic susceptibility,
, of mesoscopic Al disks, see also
Figure 2f. (
a). Field dependence of the Meissner response for a Al disk with 1.0
m diameter (□) and for one with 2.5
m diameter (•). The thickness is for both disks 0.1
m. The strongly oscillating behaviour clearly seen (the dashed line is a guide to the eye) for the larger sample is due to size quantization. Each jump corresponds to a change in the number of vortices inside the disk, which can either form an array of single quantum vortices or assemble into a single giant vortex. The inset to (a) compares the field-cooling (FC) and zero-field-cooling (ZFC) magnetization for the 2.5-mm disk at the field where the paramagnetic response is close to its maximum value. The ZFC response is always diamagnetic, and the jumps in the ZFC curve correspond to entry of individual vortices into the disk interior. (
b). Comparison of the magnetization states reached by cooling in a field and by sweeping the field at a constant temperature (arrows) at
0.4 K. The field-cooling (FC-C) data shown by □ are for the 2.5-
m disk shown in (a). The filled boxes (
▪) indicate the low-temperature states as shown in the inset to (a), ZFC and FC. The inset to (b) illustrates the compression of a giant vortex (
T close but below
) into a smaller volume (
T further away from
,) which enables extra flux to enter the sample at the surface. Image reproduced with permission from Ref. [
63].
Figure 22.
Magnetic susceptibility,
, of mesoscopic Al disks, see also
Figure 2f. (
a). Field dependence of the Meissner response for a Al disk with 1.0
m diameter (□) and for one with 2.5
m diameter (•). The thickness is for both disks 0.1
m. The strongly oscillating behaviour clearly seen (the dashed line is a guide to the eye) for the larger sample is due to size quantization. Each jump corresponds to a change in the number of vortices inside the disk, which can either form an array of single quantum vortices or assemble into a single giant vortex. The inset to (a) compares the field-cooling (FC) and zero-field-cooling (ZFC) magnetization for the 2.5-mm disk at the field where the paramagnetic response is close to its maximum value. The ZFC response is always diamagnetic, and the jumps in the ZFC curve correspond to entry of individual vortices into the disk interior. (
b). Comparison of the magnetization states reached by cooling in a field and by sweeping the field at a constant temperature (arrows) at
0.4 K. The field-cooling (FC-C) data shown by □ are for the 2.5-
m disk shown in (a). The filled boxes (
▪) indicate the low-temperature states as shown in the inset to (a), ZFC and FC. The inset to (b) illustrates the compression of a giant vortex (
T close but below
) into a smaller volume (
T further away from
,) which enables extra flux to enter the sample at the surface. Image reproduced with permission from Ref. [
63].
Based on a numerical, self-consistent solution of the Ginzburg–Landau equations for a fixed orbital quantum number
L, Moshchalkov et al. [
88,
89] proposed that the PME arises from the flux compression with integral number of quantum flux
, where
denotes the flux quantum, trapped in the sample interior. This model refines the approach of Koshelev and adds the surface superconductivity, described by the third upper critical field,
, into the model. When field-cooling a sample, one crosses the phase diagram (see e.g.,
Figure 9 and
Figure 23 below) on a horizontal line at
. Thus, we come first through an area of surface superconductivity as
, before entering the region with Abrikosov vortices below
. These vortices may form a liquid as the flux pinning sets in only below
, the irreversibility field. This region may be small for the metallic superconductors, but it does exist especially at low
T. For lower temperatures, the Shubnikov phase prevails as commonly described in the textbooks.
By studying the PME in mesoscopic superconducting Al and Nb disks, Geim et al. [
63] gave a clear indication that the quantized flux trapped at the third critical field,
, is responsible for the PME, which further supports the theoretical predictions of Refs. [
88,
89]. Thus, the flux compression mechanism seems to be more universal to explain the PME observed in both HTSc (here, the superconducting grains of polycrystalline sampes also form mesoscopic objects) and conventional superconductors (LTSc).
To explain the ideas in more detail, we assume now that inside a superconductor there are no pinning centers with the size comparable to the giant vortex core. In this case a giant vortex state is stabilized only by the sample surface and this state is reversible as long as the orbital quantum number L is kept constant. But as the temperature goes down, the multiquanta vortex state may decay rather quickly into Abrikosov vortices with , once the conservation of L is violated. As soon as the Abrikosov vortex lattice is formed at , flux pinning centers, which are relatively small in comparison to the giant core, can be very efficient to pin the vortices, thus leading to the onset of irreversibility. The irreversibility should then be considered as the consequence of the onset of the variation of L, initiating the crossover between the giant vortex state ( const.) and the Abrikosov vortex state (1), which should occur around the line. Spoken differently, in superconducting samples, where the surface pinning plays the dominant role in stabilizing the giant vortex state, the irreversibility line seems to lie in fact, quite close to the upper critical field .
By cooling down a superconductor in a fixed applied field (field-cooling mode), the
boundary in the phase diagram (see
Figure 9 and
Figure 23) is crossed at a particular point which corresponds to a certain orbital quantum number,
L. Note here that the
-line is not a homogeneous line, but cusplike according to
L[
88]. In the calculations performed in Refs. [
88,
89] was assumed that the orbital quantum number
L, found according to the location of the crossing point between
and
const., is kept constant also below the
line. The conservation of the orbital quantum number
L in the superconducting state can result from pinning of the giant vortex state, corresponding to a ringlike superconducting order parameter nucleated at the sample boundary at
. In this case, the sample boundary is the source for pinning the giant vortex state. The conserved value
const. is determined by the applied magnetic field. If sufficiently small fields are applied, a state with
1 (i.e.,
) can be realized. It also must be mentioned here that for
5 the vortex core and the area, where additional field
is generated due to the flux compression, are considerably larger than for
1.
As the temperature is further going down, the order parameter grows and pushes the magnetic field into the core. It can be clearly seen from the calculations of Refs. [
88,
89] that for the trapped
1 vortex the field
is localized in the area where the superconducting order parameter is strongly reduced. This reflects a very general flux expulsion property of a superconductor which causes either normal diamagnetic Meissner effect with complete flux expulsion for the state
0 without a core or flux compression (i.e., PME) in the vortex core for
. Topologically,
0 and
0 states are qualitatively different, since for the latter flux is expelled both inwards and outwards. When the former dominates, PME can appear.
Figure 23.
(
a). Temperature dependence of the magnetization in the FFCW mode at 0.5 mT (5 Oe). (
b). Phase diagramfor the PME in FFCW mode. The third critical field
is defined from the onset of DPE in the in-phase ac susceptibility measurements;
is defined from the intersection of two linear fits of the
curves above and below the onset;
is derived as the onset of a diamagnetic signal on the in-phase ac susceptibility curve.The solid and dashed lines are plots of the empirical formula
The crosses show the field locations where the magnetic relaxation curves are measured. Image reproduced with permission from Ref. [
37].
Figure 23.
(
a). Temperature dependence of the magnetization in the FFCW mode at 0.5 mT (5 Oe). (
b). Phase diagramfor the PME in FFCW mode. The third critical field
is defined from the onset of DPE in the in-phase ac susceptibility measurements;
is defined from the intersection of two linear fits of the
curves above and below the onset;
is derived as the onset of a diamagnetic signal on the in-phase ac susceptibility curve.The solid and dashed lines are plots of the empirical formula
The crosses show the field locations where the magnetic relaxation curves are measured. Image reproduced with permission from Ref. [
37].
Figure 22a plots the magnetic flux captured inside or expelled out of a Hall magnetometer of 2.5
m width at 0.4 K, due to the presence of superconducting disks (
, Ref. [
153]). The two disks were fabricated simultaneously by thermal evaporation and differ only in their diameters. The strongly oscillating behaviour clearly seen for the larger sample is due to size quantization. Each jump corresponds to a change in the number of vortices inside the disk, which can either form an array of single quantum vortices or assemble into a single giant vortex. The latter configuration is generally expected at applied magnetic fields between the second and third critical fields,
, that is, it corresponds to the surface superconductivity in a confined geometry. The smaller sample (□) does not exhibit the rapidly oscillating field dependence, and its Meissner response remains negative over the entire field interval. Such qualitatively different behavior can be related to the fact that, in the smaller sample, the superconductivity is suppressed by only ∼ 3 flux quanta,
, entering the disk area while∼ 20
are necessary to destroy superconductivity of the larger disk.
The observations of
Figure 22a,b may seem to be in contrast to the various studies of the PME on macroscopic samples, where the PME was normally found in very low fields and gradually disappeared with increasing field. However, one should take into account that even the lowest fields in these experiments allowed many thousands of flux quanta inside the sample interior, which is in stark contrast to the mesoscopic samples. One can observe that, with decreasing the thickness of the mesoscopic disks, the sign reversal of the Meissner effect tends to occur at lower fields and the magnitude of the PME gets larger. In contrast, no qualitative difference in behavior is observed between disks of circular and square shapes.
The origin of the PME became evident [
63] when comparing the field dependence of the Meissner effect discussed above with the magnetization response measured by sweeping the magnetic field at a constant temperature (abbreviated: C-T regime). Instead of a single magnetization curve characteristic of macroscopic superconductors, the spatial confinement of a mesoscopic sample gives rise to an entire family of magnetization curves, which are corresponding to different vortex states. Several superconducting states can be realized at the same applied magnetic field (up to five such states as can be seen in
Figure 22b), but only the state with the most negative C-T magnetization is the thermodynamically stable one [
154,
155]. All other states are metastable and become observable due to the presence of the surface barrier of the Bean-Livingston-type [
156]. The recent theory is in good agreement with similar C-T curves as reported previously.
Figure 22b demonstrates that the paramagnetic states reached via field-cooling are all metastable. Indeed, the FC data predictably fall on the C-T curves because only these distributions of the order parameter are allowed by quantization. However, among all possible states at a given magnetic field, the system unexpectedly `chooses’ the metastable state with the most positive possible magnetization. Only if we remove the proper screening in the experimental setup, a metastable high-magnetization state can eventually relax to the corresponding stable state on the lowest curve. The same result was obtained when the experiment was carried out in a more controllable manner by applying an oscillating magnetic field at a constant field,
H. One can verify that, according to
Figure 22b, an oscillating (fluctuating) field moves the system down the ladder of curves towards the equilibrium state. The following consideration may answer the question how, when cooling down, the system can end up in the most thermodynamically unfavourable state. Superconducting states in a confined geometry can be characterized by a quantum number
L, which corresponds to the number of nodes in the distribution of the complex order parameter
along the sample circumference. For the case of a giant vortex and an array of single-quantum vortices,
L has a simpler meaning:
L then represents the angular momentum and the number of vortex cores, respectively. Transitions between the various states with different
L are of first order and lead to (little) jumps in the magnetization data.
Close to the third critical field (i.e., the critical field of surface superconductivity),
, the magnetic field is distributed homogeneously and it requires the `high-temperature’ magnetic flux
[
157] to initiate a giant vortex state with the momentum
L inside a superconducting disk of radius
r. As the temperature decreases below the surface superconducting transition, the superconducting sheath at the disk perimeter rapidly expands inside, compressing the giant vortex into a small volume (see the inset to
Figure 22b; here, the case
is considered). The compressed flux inside a giant vortex is equal to
, that is,
is practically conserved for
. When at
the giant vortex splits up into
L single-quantum vortices, the captured flux changes little. At this point, it must be taken into account that the magnetic field also penetrates at the disk boundary, giving rise to an additional flux through the disk of the order of
, where
is the field strength in the
-layer at the surface. The magnetization response of the sample is paramagnetic as long as the low-temperature value of the total flux,
, is larger than
. For a superconducting cylinder,
and the PME appears at relatively large
and its amplitude is rather small (
). The plate geometry significantly enhances the PME because the field
is increased by demagnetization effects [
85]. In case the central region of the sample is occupied by a vortex or vortices is small as compared to the disk area, one can approximate
, where
t denotes the sample thickness. This results in a paramagnetic response
, which is considerable even for macroscopic thin disks. This result for
can be directly compared with the result of
Figure 8a, where
1.33 × 10
Am
(disk with
0.127 mm) was obtained for the additional paramagnetic moment appearing. Furthermore, the plate geometry also leads to an earlier start of PME [
63].
From these observations, it can be concluded that this persistence of L down to low temperatures is responsible for the PME. Thus, especially the various observations of the PME made on the Nb disks (i.e., the measurements of and the special shape of the magnetic hysteresis loops, ), can be well explained using the giant vortex model and flux compression.
Another interesting experiment in the literature considered the vortex state in the presence of the PME, which is still unclear. In the literature is reported that not all samples show PME, even with similar nominal composition. Furthermore, it was demonstrated that the PME disappears after abrading/polishing the surface of the sample or even can be created by irradiation, indicating that surface configurations, such as defects and pinning centers, do play an important role for the PME. Recently, a broad region of non-monotonic vortex interactions was discussed for multi-band and type-II/1 superconductors [
158,
159,
160,
161]. A giant PME may appear due to such non-monotonic vortex interactions, which may facilitate the trapping of magnetic flux [
162]. As experimental evidence of PME in type-II/1 superconductors was lacking in the literature, Ge et al. [
37] have studied the PME in a ZrB
single crystal. In this work, the authors have introduced the concept of fast and slow cooling the samples, i.e., (i) Fast (Slow)-field cool-warming (F(S)FC-W): the sample was cooled with a large (small) cooling rate 5 K min
(0.03 K min
) to the required temperature under a magnetic field of
H, then the magnetization was measured with increasing temperature. (ii) Slow-field cool-cooling (SFC-C): the magnetization was measured with decreasing temperature at a rate of 0.03 K min
to the desired temperature under various magnetic fields, see also
Figure 2e.
The fast cooling enabled the PME to be observed only when cooling down the sample at a sufficiently high cooling rate (5 K min). At the small cooling rate of 0.03 K min), the extra flux trapped through surface superconductivity has enough time to escape from the sample interior due to flux diffusion, resulting in a stable and more ordered vortex state at low temperatures. This is another important aspect for many other observations of the PME, which must be considered in the planning of the experiments.
From their data obtained on the ZrB
crystal, the authors have constructed a phase diagram, where they have interpreted the
as
, and
as the onset of a diamagnetic response in the measurement of
. The resulting phase diagram (
Figure 23) shows all characteristic fields/temperatures for a sample exhibiting PME, and can directly be compared to
Figure 9, showing the results of the bulk Nb disk. Thus it is obvious that Nb and ZrB
share the same origin of the PME, but the vortex interactions of both materials are different.
All these observations of PME in Nb, ZrB and the mesoscopic Al samples directly imply that the PME is not an `uncommon’ feature, but fits fully into the picture of a superconducting material when the contribution of surface superconductivity as described by is of significance. Furthermore, effects of demagnetization may play an important role for the observation of PME.