Next Article in Journal
Obtaining Boron Carbide and Nitride Matrix Nanocomposites for Neutron-Shielding and Therapy Applications
Next Article in Special Issue
Vortex Dynamics and Pinning in CaKFe4As4 Single Crystals from DC Magnetization Relaxation and AC Susceptibility
Previous Article in Journal
Effects of the Exciton Fine Structure Splitting on the Entanglement-Based Quantum Key Distribution
Previous Article in Special Issue
Superconducting and Mechanical Properties of the Tl0.8Hg0.2Ba2Ca2Cu3O9−δ Superconductor Phase Substituted by Lanthanum and Samarium Fluorides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

The Depairing Current Density of a Fe(Se,Te) Crystal Evaluated in Presence of Demagnetizing Factors

1
Department of Physics “E.R. Caianiello”, University of Salerno, Via Giovanni Paolo II, 132, I-84084 Fisciano, Salerno, Italy
2
CNR-SPIN Salerno, Via Giovanni Paolo II, 132, I-84084 Fisciano, Salerno, Italy
3
Institute of Solid State Physics, Bulgarian Academy of Sciences, 72 Tzarigradsko Chaussee, 1784 Sofia, Bulgaria
4
National Institute of Materials Physics, 405A Atomistilor Str., 077125 Magurele, Romania
*
Authors to whom correspondence should be addressed.
Condens. Matter 2023, 8(4), 91; https://doi.org/10.3390/condmat8040091
Submission received: 27 September 2023 / Revised: 20 October 2023 / Accepted: 20 October 2023 / Published: 23 October 2023

Abstract

:
The effect of the demagnetizing factor, regarding the determination of the de-pairing current density J d e p , has been studied in the case of a Fe(Se,Te) crystal, using DC magnetic measurements as a function of a magnetic field (H) at different temperatures (T). First, the lower critical field H c 1 (T) values were obtained, and the demagnetization effects acting on them were investigated after calculating the demagnetizing factor. The temperature behaviors of both the original H c 1 values and the ones obtained after considering the demagnetization effects ( H c 1 d e m a g ) were analyzed, and the temperature dependence of the London penetration depth λ L (T) was obtained in both cases. In particular, the λ L (T) curves were fitted with a power law dependence, indicating the presence of low-energy quasiparticle excitations. Furthermore, by plotting λ L 2 as a function of T, we found that our sample behaves as a multigap superconductor, which is similar to other Fe-11 family iron-based compounds. After that, the coherence length ξ values were extracted, starting with the H c 2 (T) curve. The knowledge of λ L and ξ allowed us to determine the J d e p values and to observe how they are influenced by the demagnetizing factor.

1. Introduction

The initial discovery of superconductivity in fluorine-doped LaFeAsO [1] attracted a lot of interest in the scientific community. Nevertheless, it was an iron-based superconductor, but with oxygen in its stoichiometric formula. This prompted us to correlate this compound with previously discovered cuprates superconductors [2]. Later, with the important discovery of the Ba1−xKxFe2As2 compound, it became clear that the importance of this new class of materials (which can present not-oxide superconductors) breaks the link with cuprates. This highlights the fact that, for fabricating high-temperature superconductors, oxygen does not necessary play a crucial role. The 11 family in the class of iron-based superconductors has been intensively investigated in recent years so that we may understand the basic mechanism governing its superconductivity [3]. These materials have a layered structure, like cuprates, and they show interesting peculiarities such as high values in the upper critical field, critical current density, and irreversibility field [4,5,6,7]. Moreover, they have lower anisotropy values than cuprates, with associated high pinning energy values [8,9,10,11]. Among these properties, there are also non-monotonic responses to magnetization with applied magnetic fields; this has led to particular phenomena that have been deeply studied in recent years [12,13,14,15,16]. These phenomena are strictly correlated with vortex dynamics [17,18,19,20,21]. On the other hand, the presence of vortices inside type-II superconductors can also be investigated by analyzing the lower critical field, H c 1 , together with the London penetration depth, λ L . H c 1 and λ L are useful parameters for gathering information regarding the bulk thermodynamic properties of a sample, and they have been investigated for iron-based compounds in the past [22,23,24]. In particular, compared with other physical quantities, penetration depth is a useful parameter to study the superconductivity of a compound intrinsically; this is because it is not sensitive to the aspects related to surface conditions. In particular, the study of the behavior of λ L 2 , as a function of temperature, can provide information on the superconductivity typology characterizing the sample (e.g., single gap BCS theory, the two gap model, etc.) [25,26,27]. Moreover, the λ L values indicate the strength of the interactions between vortices, thus allowing us to deduce the magnitude of the effective pinning energy of the sample. λ L , together with the coherence length, ξ , represent the fingerprints of a superconductor, and this study becomes crucial when first characterizing a new superconductor. Moreover, by combining them, it is possible to obtain the de-pairing current density, J d e p , which fixes the upper limit of the presence of superconductivity inside a superconducting material. The de-pairing current density is of significant importance for understanding the existing limits for increasing J c [28,29], and since it directly provides data on the critical velocity of the superfluid, it is essential for the investigation of the superconducting mechanism and the symmetry of the superconducting gap [30]. Using this framework, by following the Ginzburg–Landau theory, the de-pairing current density, J d e p , depends on the characteristic critical parameters H c 1 and H c 2 , and more specifically, the London penetration depth, λ L , and the coherence length, ξ [30,31]. In this work, the influence of the demagnetization effects on the de-pairing current density, J d e p , has been analyzed by studying a Fe(Se,Te) iron-based superconductor. We started by measuring the first magnetization curve at different temperatures in order to obtain the lower critical field H c 1 values. We have noted that the demagnetization effects acting on the sample were significant, and they resulted in an underestimation of the real H c 1 values. From the H c 1 values, the London penetration depth, λ L , as a function of the temperature, was obtained, and it was noted that it is not possible to fit the penetration depth with the typical exponential behavior that characterizes the s-wave superconductor. In this context, the plot of λ L 2 , as a function of T, confirmed that our sample shows peculiarities which can be ascribed to a multigap superconducting behavior. Finally, after determining the coherence length, ξ , from the upper critical field, H c 2 , the J d e p values were calculated as a function of the temperature, by considering the demagnetization effects and not considering them; this provided very high values in the framework of the iron-based superconductors.

2. Results and Discussion

In order to study the lower critical field, H c 1 behavior, as a function of temperature T, the virgin magnetic moment vs. magnetic field was measured at different temperatures. In Figure 1, the first branch of the superconducting hysteresis loop was reported as reaching up to 0.3 T in the temperature range of 2.5 K to 10 K. The initial linear decrease of the magnetic moment m was visible due to the Meissner state, and the reduction of the superconducting signal was visible due to the increase in temperature. To determine H c 1 at a fixed temperature, the first value that deviated from the linear trend of the Meissner state (black dashed line) meaning that the vortices have penetrated the sample building up a critical state [32,33,34]. Considering all the temperatures, the H c 1 (T) behavior was obtained and then fitted using the following equation [35]:
H c 1 T = H c 1 0 1 T T c n
where H c 1 (0) is the value of H c 1 at T = 0 K, T c = 14.5 K [4], and n is the exponent. As per the fitting procedure, H c 1 (0) = 143 Oe and n = 1.54 were obtained. The H c 1 (T) curve, together with its fit, is presented in Figure 2a (black squares and black solid line). It is worth noting that the H c 1 (T) curve shows an upward trend with a negative curvature; this was not predicted in the single-band gap description of the mean-field theory, therefore, it is evident that two energy gaps exist [36]. Generally, a superconductor immersed in a magnetic field is subject to demagnetization effects at low fields due to its finite dimensions [37,38]. Considering that the demagnetization effects are stronger when the sample is in a perpendicular field configuration, these effects cannot be overlooked in our case. In particular, the H c 1 values are underestimated because, due to the demagnetization effects, the sample experiences a magnetic field higher than the applied one. Therefore, in order to take into account the demagnetization effects acting on the sample, the H c 1 values must be properly scaled using the so-called demagnetizing factor. More specifically, the demagnetized H c 1 values H c 1 d e m a g , as a function of temperature, were calculated using the formula reported by Yeshurun et al. [39].
H c 1 d e m a g = H c 1 1 N
where N is the demagnetizing factor which can assume values ranging between 0 and 1. The N value can be determined using the following formula [37,38]:
N = H 4 π M + 1  
where M is the magnetization in the Meissner state and H is the applied field. In our case, N has been already estimated to be approximately 0.76 in Ref. [4], which is quite a high value; this aligns with the perpendicular field configuration. Therefore, using Equation (2) with N = 0.76, the demagnetized values for H c 1 were calculated and fitted with the following equation:
H c 1 d e m a g T = H c 1 d e m a g 0 1 T T c n
where H c 1 d e m a g 0 is the value of H c 1 d e m a g at T = 0 K, T c = 14.5 K [4], and n is the exponent. From the fitting procedure, H c 1 d e m a g (0) = 597 Oe and n = 1.54 were obtained. The H c 1 d e m a g T curve, together with its fit, is presented in Figure 2a (red circles and red dashed line).
As previously mentioned, the H c 1 d e m a g values are higher than the H c 1 ones, as reported in Figure 2a. In Figure 2b, the ratio of the H c 1 d e m a g ( T ) and H c 1 ( T ) values was reported to show a value equal to approximately four. On the other hand, the H c 1 d e m a g ( 0 ) / H c 1 (0) value was also reported to be a red star in Figure 2b, and it is still equal to approximately four, thus confirming the viability of the fitting procedures that were previously performed. From the determination of the lower critical field values, the temperature dependence of the London penetration depth, λ L , can be obtained using the following formula [25]:
H c 1 = ϕ 0 4 π λ L 2 ln k
where ϕ 0 = 2.07 × 10 7 Oe cm2 is the magnetic flux quantum and k is the Ginzburg–Landau parameter. Using the Ginzburg–Landau theory, and by following the approach reported in Ref. [40], k can be calculated using the relation k = H c 2 (0)/(21/2 H c (0)), where H c (0) is the thermodynamic critical field. H c (0) is calculated using H c 1 , H c 1 d e m a g , and H c 2 values at a temperature of zero (i.e., H c 0 = H c 1 ( 0 ) × H c 2 ( 0 ) 1 / 2 ≈ 8 kOe and H c d e m a g 0 = H c 1 d e m a g ( 0 ) × H c 2 ( 0 ) 1 / 2 ≈ 16.5 kOe); therefore, obtaining k ≈ 40 and kdemag ≈ 20 aligns with other Fe-chalcogenide superconductors [40,41,42,43]. For the calculation of λ L , Equation (5), both H c 1 and H c 1 d e m a g , together with the k and kdemag values, were used, and the results are reported in Figure 3. Both the λ L T curves were fitted using the following equation:
λ L ( T ) = λ L 0 1 T T c n
where λ L 0 is the London penetration depth at T = 0 K, T c = 14.5 K [4], and n is the exponent. From the fitting procedure, λ L (0) = 208 nm, n = −0.72, and λ L d e m a g (0) = 92 nm, and n = −0.72 were obtained, having considered the H c 1 , k, and H c 1 d e m a g , kdemag values for the λ L calculation, respectively. At low temperatures, λ L does not show the typical exponential behavior expected for a fully gapped clean s-wave superconductor [44]. In general, a power law temperature dependence of λ L implies the presence of low-energy quasiparticle excitations [45].
In this framework, when plotting λ L 2 as a function of T (see Figure 4), the curve behavior was noted as being completely different from the behavior predicted using the single-gap BCS theory, which should have an opposite concavity; this indicates a more probable multigap superconductivity in our sample [25,27,46,47]. This aligns with the results reported in the literature for Fe-based pnictides [48,49,50,51]. On the other hand, it should be noted that the λ L 2 (T) curvature may be also a sign that the sample has an anisotropic single gap nature [52]. Finally, it is worth noting that the λ L d e m a g values obtained from H c 1 d e m a g appear low in respect to other Fe(Se,Te) samples reported in the literature [45,46,53,54], indicating a low vortex–vortex interaction that, in the framework of vortex dynamics, usually characterizes a single vortex state. This vortex lattice configuration, due to the strength of the effective pinning energy, typically allows the sample to carry high transport currents in high fields that are suitable for power applications of this class of superconducting materials. After the calculation of the λ L values, the coherence length ξ values were extracted, starting with the H c 2 (T) curve reported in Figure 12 of Ref. [4], using the following equation [55,56]:
H c 2 T = ϕ 0 2 π ξ 2 T  
where ϕ 0 = 2.068 × 10−15 Tm2 is the magnetic flux quantum.
The ξ values, as a function of temperature, are shown in Figure 5 together with the fit with the equation [57]:
ξ T = ξ 0 1 T T c n
where ξ 0 is the coherence length at 0 K, T c = 14.5 K [4], and n is the exponent. From the fitting procedure, we obtained ξ 0 ≈ 3 nm, which coheres with the value, ξ 0 ≈ 2.7 nm, which was calculated using H c 2 (0) = 46.5 T from Ref. [4], and n 0.64 , from Ref. [57]. The determination of the penetration depth, and the coherence length values, allows us to estimate the de-pairing current density, J d e p . The de-pairing current density is the current value above which the superconductivity of the superconductor is completely broken; this is because it is the value wherein the kinetic energy of the superconducting carriers equals the binding energy of the Cooper pair. The de-pairing current density can be expressed, thanks to the Ginzburg–Landau theory, as follows [30,31]:
J d e p T = ϕ 0 3 3 / 2 π μ 0 λ L 2 ξ  
where μ 0 is the vacuum permeability. It is worth underlining that this equation is usually only considered valid for temperatures approaching T c . Nevertheless, by taking into account the calculations performed by Bardeen [58], the de-pairing current density deviates from the GL theory by a maximum factor of 1.5 at T = 0 K; this indicates that it is possible to study its temperature behavior with reasonable error. Moreover, Equation (9) is usually valid for the single band superconductor, whereas Fe(Se,Te) shows peculiarities which can be ascribed to a multigap superconducting behavior, as previously mentioned. In this context, another important parameter to consider is γ H = H c 2 , a b / H c 2 , c , and in particular, its temperature dependence. In fact, in a single gap superconductor, the γ H parameter is temperature independent, while in a multiband superconductor, the contributions of different bands lead to a non-constant γ H , with each band contributing differently as T is changed. For Fe(Se,Te), it is evident in several works [59,60,61] that γ H is slightly dependent on temperature, therefore, making reasonable the use of the GL formula, although it gives overestimated Jdep values, especially at low temperatures. The J d e p values. obtained using Equation (9), are reported in Figure 6. It is worth noting the presence of two J d e p (T) curves. In particular, the black curve was obtained using λ L without the demagnetization correction, whereas the red curve was obtained by considering λ L d e m a g in conjunction with the demagnetization correction. Both sets of J d e p values align with the highest values reported for the different iron-based families; this demonstrates the very good quality of this crystal [62,63,64,65,66]. It is evident that the J d e p d e m a g values that take the demagnetizing factor into account are five times higher than the J d e p values obtained without considering the demagnetizing factor. This helps us understand how important the demagnetizing factor is when estimating different important superconducting parameters such as H c 1 , λ L , and J d e p . It is important to note that the role of λ L is crucial to Equation (9) since it is squared, and therefore, a small λ L change generates a large J d e p variation. In conclusion, in light of the fact that H c 1 and λ L can strongly depend on stoichiometry, the possibility of tuning it by modifying the fabrication process and parameters could be exploited to enhance the de-pairing current density.

3. Materials and Methods

We analyzed a FeSe0.5Te0.5 (nominal composition) crystal with the following dimensions: 3 × 3 × 0.2 mm3. The crystal was created using the Bridgman technique, and T c = 14.5 K. Details concerning the creation of the crystal are reported elsewhere [4]. A SEM-EDX analysis was performed on the sample, which showed the presence of twin boundaries and a slight deviation from the nominal composition in terms of stoichiometry (Fe0.96Te0.59Se0.45) [67]. This is probably due to micro inhomogeneity and the phase separation of magnetic premises, which is typical for crystal growth and synthesis in FeSeTe [68,69,70] and its basic compound FeSe [71,72,73]. The sample was characterized in a dc magnetic field that was applied perpendicularly to its largest face (H||c). In particular, the dc magnetic moment, as a function of the field, m(H), was measured using a Quantum Design PPMS-9T equipped with a VSM option. To avoid the effect on the sample response caused by the residual trapped field inside the PPMS dc magnet [74], this field was reduced below 1 × 10−4 T [75]. Regarding the m(H) measurements, the sample was first cooled down to the measurement temperature in the zero field and thermally stabilized for at least 20 min. Then, the field was ramped with the fixed sweep rate value to +9 T, then it was reduced to −9 T, and finally, to +9 T again in order to acquire the complete hysteresis loop.

4. Conclusions

By studying the DC magnetic moment as a function of the magnetic field (H) at different temperatures (T) on a Fe(Se,Te) crystal created using the Bridgman technique, the effect of the demagnetizing factor on the de-pairing current density J d e p values was estimated. In order to achieve this, the London penetration depth and the coherence length were evaluated. Initially, the lower critical field, H c 1 , was obtained as a function of temperature that focused on the impact that demagnetization effects have on its values. In particular, it was found that H c 1 values were underestimated by a factor equal to four, even at T = 0 K. Starting with these results, the London penetration depth, λ L , was calculated as a function of temperature using both the original H c 1 values and the values obtained after considering the demagnetization effects ( H c 1 d e m a g ). The λ L ( T ) curves did not show the typical exponential behavior that was expected for a fully gapped clean s-wave superconductor, but rather, they exhibited a power law dependence that indicated the presence of low-energy quasiparticle excitations. In this framework, λ L 2 , as a function of T, was graphed, and multigap-like behavior was found in our sample, which aligns with the behavior of other iron-based samples reported in the literature. Additionally, we found that the λ L d e m a g values obtained from H c 1 d e m a g are lower than the values obtained from other Fe(Se,Te) samples reported in the literature, which suggests a possible single vortex state. In this vortex lattice configuration, the effective pinning energy is high, therefore, reducing the dissipations inside the material, and improving the current transport properties is suitable for power applications of this class of superconducting materials. After that, the coherence length ξ values were extracted, starting with the H c 2 (T) curve, as reported in our previous work. Combining the λ L and ξ values, the J d e p values that took the demagnetization effects into account, as a function of temperature, were five times higher than the values that did not take the demagnetization effects into account. Tuning the stoichiometry of the compounds by modifying the fabrication procedures, and thus, changing the H c 1 and λ L values by considering the demagnetizing factor, can push the limits of the de-pairing current density values of the materials.

Author Contributions

Conceptualization, A.G., K.B., V.T. and E.N.; methodology, A.G. and K.B.; software, A.G., V.T. and A.L.; validation, K.B., E.N., G.G. and A.C.; formal analysis, A.G. and K.B.; investigation, A.G. and K.B.; resources, A.C. and M.P.; data curation, A.G., K.B. and V.T.; writing—original draft preparation, A.G.; writing—review and editing, K.B., E.N., G.G., A.C. and M.P.; visualization, K.B., A.L. and G.G.; supervision, A.C. and M.P.; project administration, A.C. and M.P; funding acquisition, A.C. and M.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the EU COST Actions CA19108 Hi-SCALE, CA20116 OPERA and CA21144 SUPERQUMAP. A.C. gratefully acknowledges the funding from the Core Program of the National Institute of Materials Physics, granted by the Romanian Ministry of Research, Innovation, and Digitalization under the Project PC2-PN23080202.

Data Availability Statement

All the data reported are available from the authors upon reasonable request.

Acknowledgments

This work has been carried out within the framework of the inter-academic Italian-Bulgarian research project 2019–2021 (Department of Physics ‘E R Caianiello’ CNR-SPIN unit, University of Salerno and the Institute of Solid State Physics ‘Georgi Nadjakov’, Bulgarian Academy of Sciences). The authors would like to thank M. Gospodinov from ISSP-BAS for the valuable discussions and the support for the sample preparation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kamihara, Y.; Watanabe, T.; Hirano, M.; Hosono, H. Iron-based layered superconductor La[O1−xFx]FeAs (x = 0.05–0.12) with Tc = 26 K. J. Am. Chem. Soc. 2008, 130, 3296–3297. [Google Scholar] [CrossRef]
  2. Bednorz, J.G.; Müller, K.A. Possible high Tc superconductivity in the Ba-La-Cu-O system. Z. Phys. B Condens. Matter 1986, 64, 189–193. [Google Scholar] [CrossRef]
  3. Hsu, F.-C.; Luo, J.-Y.; Yeh, K.-W.; Chen, T.-K.; Huang, T.-W.; Wu, P.M.; Lee, Y.-C.; Huang, Y.-L.; Chu, Y.-Y.; Yan, D.-C.; et al. Superconductivity in the PbO-type structure α-FeSe. Proc. Natl. Acad. Sci. USA 2008, 105, 14262–14264. [Google Scholar] [CrossRef] [PubMed]
  4. Galluzzi, A.; Buchkov, K.; Tomov, V.; Nazarova, E.; Leo, A.; Grimaldi, G.; Nigro, A.; Pace, S.; Polichetti, M. Evidence of pinning crossover and the role of twin boundaries in the peak effect in FeSeTe iron based superconductor. Supercond. Sci. Technol. 2018, 31, 015014. [Google Scholar] [CrossRef]
  5. Galluzzi, A.; Buchkov, K.; Nazarova, E.; Tomov, V.; Grimaldi, G.; Leo, A.; Pace, S.; Polichetti, M. Transport properties and high upper critical field of a Fe(Se,Te) iron based superconductor. Eur. Phys. J. Spec. Top. 2019, 228, 725–731. [Google Scholar] [CrossRef]
  6. Galluzzi, A.; Buchkov, K.; Nazarova, E.; Tomov, V.; Leo, A.; Grimaldi, G.; Pace, S.; Polichetti, M. Magnetic field sweep rate influence on the critical current capabilities of a Fe(Se,Te) crystal. J. Appl. Phys. 2020, 128, 073902. [Google Scholar] [CrossRef]
  7. Hosono, H.; Yamamoto, A.; Hiramatsu, H.; Ma, Y. Recent advances in iron-based superconductors toward applications. Mater. Today 2018, 21, 278–302. [Google Scholar] [CrossRef]
  8. Yuan, H.Q.; Singleton, J.; Balakirev, F.F.; Baily, S.A.; Chen, G.F.; Luo, J.L.; Wang, N.L. Nearly isotropic superconductivity in (Ba,K)Fe2As2. Nature 2009, 457, 565–568. [Google Scholar] [CrossRef] [PubMed]
  9. Yamamoto, A.; Jaroszynski, J.; Tarantini, C.; Balicas, L.; Jiang, J.; Gurevich, A.; Larbalestier, D.C.; Jin, R.; Sefat, A.S.; McGuire, M.A.; et al. Small anisotropy, weak thermal fluctuations, and high field superconductivity in Co-doped iron pnictide Ba(Fe1−xCox)2As2. Appl. Phys. Lett. 2009, 94, 062511. [Google Scholar] [CrossRef]
  10. Grimaldi, G.; Leo, A.; Martucciello, N.; Braccini, V.; Bellingeri, E.; Ferdeghini, C.; Galluzzi, A.; Polichetti, M.; Nigro, A.; Villegier, J.-C.; et al. Weak or Strong Anisotropy in Fe(Se,Te) Superconducting Thin Films Made of Layered Iron-Based Material? IEEE Trans. Appl. Supercond. 2019, 29, 1–4. [Google Scholar] [CrossRef]
  11. Leo, A.; Braccini, V.; Bellingeri, E.; Ferdeghini, C.; Galluzzi, A.; Polichetti, M.; Nigro, A.; Pace, S.; Grimaldi, G. Anisotropy effects on the quenching current of Fe(Se,Te) Thin Films. IEEE Trans. Appl. Supercond. 2018, 28, 8234633. [Google Scholar] [CrossRef]
  12. Eley, S.; Willa, R.; Chan, M.K.; Bauer, E.D.; Civale, L. Vortex phases and glassy dynamics in the highly anisotropic superconductor HgBa2CuO4+δ. Sci. Rep. 2020, 10, 10239. [Google Scholar] [CrossRef]
  13. Polichetti, M.; Galluzzi, A.; Buchkov, K.; Tomov, V.; Nazarova, E.; Leo, A.; Grimaldi, G.; Pace, S. A precursor mechanism triggering the second magnetization peak phenomenon in superconducting materials. Sci. Rep. 2021, 11, 7247. [Google Scholar] [CrossRef] [PubMed]
  14. Llovo, I.F.; Sónora, D.; Mosqueira, J.; Salem-Sugui, S.; Sundar, S.; Alvarenga, A.D.; Xie, T.; Liu, C.; Li, S.L.; Luo, H.Q. Vortex dynamics and second magnetization peak in the iron-pnictide superconductor Ca0.82La0.18Fe0.96Ni0.04As2. Supercond. Sci. Technol. 2021, 34, 115010. [Google Scholar] [CrossRef]
  15. Yi, X.; Xing, X.; Meng, Y.; Zhou, N.; Wang, C.; Sun, Y.; Shi, Z. Anomalous Second Magnetization Peak in 12442-Type RbCa2Fe4As4F2 Superconductors. Chin. Phys. Lett. 2023, 40, 027401. [Google Scholar] [CrossRef]
  16. Lopes, P.V.; Sundar, S.; Salem-Sugui, S.; Hong, W.; Luo, H.; Ghivelder, L. Second magnetization peak, anomalous field penetration, and Josephson vortices in KCa2Fe4As4F2 bilayer pnictide superconductor. Sci. Rep. 2022, 12, 20359. [Google Scholar] [CrossRef]
  17. Prozorov, R.; Ni, N.; Tanatar, M.A.; Kogan, V.G.; Gordon, R.T.; Martin, C.; Blomberg, E.C.; Prommapan, P.; Yan, J.Q.; Bud’ko, S.L.; et al. Vortex phase diagram of Ba(Fe0.93Co0.07)2As2 single crystals. Phys. Rev. B 2008, 78, 224506. [Google Scholar] [CrossRef]
  18. Sun, Y.; Taen, T.; Tsuchiya, Y.; Pyon, S.; Shi, Z.; Tamegai, T. Magnetic relaxation and collective vortex creep in FeTe0.6Se0.4 single crystal. Europhys. Lett. 2013, 103, 57013. [Google Scholar] [CrossRef]
  19. Pramanik, A.K.; Harnagea, L.; Nacke, C.; Wolter, A.U.B.; Wurmehl, S.; Kataev, V.; Büchner, B. Fishtail effect and vortex dynamics in LiFeAs single crystals. Phys. Rev. B 2011, 83, 094502. [Google Scholar] [CrossRef]
  20. Sundar, S.; Salem-Sugui, S.; Amorim, H.S.; Wen, H.H.; Yates, K.A.; Cohen, L.F.; Ghivelder, L. Plastic pinning replaces collective pinning as the second magnetization peak disappears in the pnictide superconductor Ba0.75K0.25Fe2As2. Phys. Rev. B 2017, 95, 134509. [Google Scholar] [CrossRef]
  21. Taen, T.; Tsuchiya, Y.; Nakajima, Y.; Tamegai, T. Critical current densities and vortex dynamics in FeTexSe1−x single crystals. Phys. C Supercond. Appl. 2010, 470, 1106–1108. [Google Scholar] [CrossRef]
  22. Ren, C.; Wang, Z.S.; Luo, H.Q.; Yang, H.; Shan, L.; Wen, H.H. Evidence for two energy gaps in superconducting Ba0.6K0.4Fe2As2 single crystals and the breakdown of the uemura plot. Phys. Rev. Lett. 2008, 101, 257006. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, X.L.; Dou, S.X.; Ren, Z.A.; Yi, W.; Li, Z.C.; Zhao, Z.X.; Lee, S.I. Unconventional superconductivity of NdFeAsO0.82F0.18 indicated by the low temperature dependence of the lower critical field Hc1. J. Phys. Condens. Matter 2009, 21, 205701. [Google Scholar] [CrossRef]
  24. Martin, C.; Gordon, R.T.; Tanatar, M.A.; Kim, H.; Ni, N.; Bud’Ko, S.L.; Canfield, P.C.; Luo, H.; Wen, H.H.; Wang, Z.; et al. Nonexponential London penetration depth of external magnetic fields in superconducting Ba1−xKxFe2As2 single crystals. Phys. Rev. B 2009, 80, 020501. [Google Scholar] [CrossRef]
  25. Abdel-Hafiez, M.; Ge, J.; Vasiliev, A.N.; Chareev, D.A.; Van De Vondel, J.; Moshchalkov, V.V.; Silhanek, A.V. Temperature dependence of lower critical field Hc1(T) shows nodeless superconductivity in FeSe. Phys. Rev. B 2013, 88, 174512. [Google Scholar] [CrossRef]
  26. Song, Y.J.; Ghim, J.S.; Yoon, J.H.; Lee, K.J.; Jung, M.H.; Ji, H.S.; Shim, J.H.; Bang, Y.; Kwon, Y.S. Small anisotropy of the lower critical field and the s±- wave two-gap feature in single-crystal LiFeAs. Europhys. Lett. 2011, 94, 57008. [Google Scholar] [CrossRef]
  27. Prozorov, R.; Kogan, V.G. London penetration depth in iron-based superconductors. Rep. Prog. Phys. 2011, 74, 124505–124525. [Google Scholar] [CrossRef]
  28. Rosenstein, B.; Li, D. Ginzburg-Landau theory of type II superconductors in magnetic field. Rev. Mod. Phys. 2010, 82, 109–168. [Google Scholar] [CrossRef]
  29. Tahara, S.; Anlage, S.M.; Halbritter, J.; Eom, C.B.; Fork, D.K.; Geballe, T.H.; Beasley, M.R. Critical currents, pinning, and edge barriers in narrow YBa2Cu3O7−δ thin films. Phys. Rev. B 1990, 41, 11203–11208. [Google Scholar] [CrossRef] [PubMed]
  30. Tinkham, M. Introduction to Superconductivity; Dover Publications: Mineola, NY, USA, 2004; ISBN 0486134725. [Google Scholar]
  31. Arpaia, R.; Nawaz, S.; Lombardi, F.; Bauch, T. Improved nanopatterning for YBCO nanowires approaching the depairing current. IEEE Trans. Appl. Supercond. 2013, 23, 1101505. [Google Scholar] [CrossRef]
  32. Wang, T.; Ma, Y.; Li, W.; Chu, J.; Wang, L.; Feng, J.; Xiao, H.; Li, Z.; Hu, T.; Liu, X.; et al. Two-gap superconductivity in CaFe0.88Co0.12AsF revealed by temperature dependence of the lower critical field Hc1c(T). npj Quantum Mater. 2019, 4, 33. [Google Scholar] [CrossRef]
  33. Musolino, N.; Bals, S.; Van Tendeloo, G.; Clayton, N.; Walker, E.; Flükiger, R. Modulation-free phase in heavily Pb-doped (Bi,Pb)2212 crystals. Phys. C Supercond. Its Appl. 2003, 399, 1–7. [Google Scholar] [CrossRef]
  34. Galluzzi, A.; Leo, A.; Masi, A.; Varsano, F.; Nigro, A.; Grimaldi, G.; Polichetti, M. Magnetic Vortex Phase Diagram for a Non-Optimized CaKFe4As4 Superconductor Presenting a Wide Vortex Liquid Region and an Ultra-High Upper Critical Field. Appl. Sci. 2023, 13, 884. [Google Scholar] [CrossRef]
  35. Felner, I.; Kopelevich, Y. Magnetization measurement of a possible high-temperature superconducting state in amorphous carbon doped with sulfur. Phys. Rev. B 2009, 79, 233409. [Google Scholar] [CrossRef]
  36. Yadav, C.S.; Paulose, P.L. Upper critical field, lower critical field and critical current density of FeTe0.60Se0.40 single crystals. New J. Phys. 2009, 11, 103046. [Google Scholar] [CrossRef]
  37. Stoner, E.C. XCVII. The demagnetizing factors for ellipsoids. Lond. Edinb. Dublin Philos. Mag. J. Sci. 1945, 36, 803–821. [Google Scholar] [CrossRef]
  38. Prando, G.; Giraud, R.; Aswartham, S.; Vakaliuk, O.; Abdel-Hafiez, M.; Hess, C.; Wurmehl, S.; Wolter, A.U.B.; Büchner, B. Evidence for a vortex–glass transition in superconducting Ba(Fe0.9Co0.1)2As2. J. Phys. Condens. Matter 2013, 25, 505701. [Google Scholar] [CrossRef] [PubMed]
  39. Yeshurun, Y.; Malozemoff, A.P.; Shaulov, A. Magnetic relaxation in high-temperature superconductors. Rev. Mod. Phys. 1996, 68, 911–949. [Google Scholar] [CrossRef]
  40. Maheshwari, P.K.; Gahtori, B.; Gupta, A.; Awana, V.P.S. Impact of Fe site Co substitution on superconductivity of Fe1−xCoxSe0.5Te0.5 (x = 0.0 to 0.10): A flux free single crystal study. AIP Adv. 2017, 7, 15006. [Google Scholar] [CrossRef]
  41. Murugesan, K.; Lingannan, G.; Ishigaki, K.; Uwatoko, Y.; Sekine, C.; Kawamura, Y.; Junichi, H.; Joseph, B.; Vajeeston, P.; Maheswari, P.K.; et al. Pressure Dependence of Superconducting Properties, Pinning Mechanism, and Crystal Structure of the Fe0.99Mn0.01Se0.5Te0.5 Superconductor. ACS Omega 2021, 6, 30419–30431. [Google Scholar] [CrossRef] [PubMed]
  42. Lei, H.; Hu, R.; Petrovic, C. Critical fields, thermally activated transport, and critical current density of β-FeSe single crystals. Phys. Rev. B 2011, 84, 014520. [Google Scholar] [CrossRef]
  43. Dutta, P.; Pramanick, S.; Chatterjee, S. Effect of S-doping on the magnetic and electrical properties of FeSe superconductor. Phys. C Supercond. Appl. 2022, 602, 1354126. [Google Scholar] [CrossRef]
  44. Fletcher, J.D.; Serafin, A.; Malone, L.; Analytis, J.G.; Chu, J.H.; Erickson, A.S.; Fisher, I.R.; Carrington, A. Evidence for a nodal-line superconducting state in LaFePO. Phys. Rev. Lett. 2009, 102, 147001. [Google Scholar] [CrossRef] [PubMed]
  45. Takahashi, H.; Imai, Y.; Komiya, S.; Tsukada, I.; Maeda, A. Anomalous temperature dependence of the superfluid density caused by a dirty-to-clean crossover in superconducting FeSe0.4Te0.6 single crystals. Phys. Rev. B 2011, 84, 132503. [Google Scholar] [CrossRef]
  46. Bendele, M.; Weyeneth, S.; Puzniak, R.; Maisuradze, A.; Pomjakushina, E.; Conder, K.; Pomjakushin, V.; Luetkens, H.; Katrych, S.; Wisniewski, A.; et al. Anisotropic superconducting properties of single-crystalline FeSe0.5Te0.5. Phys. Rev. B 2010, 81, 224520. [Google Scholar] [CrossRef]
  47. Milošević, M.V.; Perali, A. Emergent phenomena in multicomponent superconductivity: An introduction to the focus issue. Supercond. Sci. Technol. 2015, 28, 060201. [Google Scholar] [CrossRef]
  48. Weyeneth, S.; Puzniak, R.; Mosele, U.; Zhigadlo, N.D.; Katrych, S.; Bukowski, Z.; Karpinski, J.; Kohout, S.; Roos, J.; Keller, H. Anisotropy of superconducting single crystal SmFeAsO0.8F0.2 studied by torque magnetometry. J. Supercond. Nov. Magn. 2009, 22, 325–329. [Google Scholar] [CrossRef]
  49. Gonnelli, R.S.; Daghero, D.; Tortello, M.; Ummarino, G.A.; Stepanov, V.A.; Kremer, R.K.; Kim, J.S.; Zhigadlo, N.D.; Karpinski, J. Point-contact Andreev-reflection spectroscopy in ReFeAsO1−xFx (Re = La, Sm): Possible evidence for two nodeless gaps. Phys. C Supercond. Appl. 2009, 469, 512–520. [Google Scholar] [CrossRef]
  50. Szabó, P.; Pribulová, Z.; Pristáš, G.; Bud’ko, S.L.; Canfield, P.C.; Samuely, P. Evidence for two-gap superconductivity in Ba0.55K 0.45Fe2As2 from directional point-contact Andreev-reflection spectroscopy. Phys. Rev. B 2009, 79, 012503. [Google Scholar] [CrossRef]
  51. Mu, G.; Luo, H.; Wang, Z.; Shan, L.; Ren, C.; Wen, H.H. Low temperature specific heat of the hole-doped Ba0.6K0.4Fe2As2 single crystals. Phys. Rev. B 2009, 79, 174501. [Google Scholar] [CrossRef]
  52. Bekaert, J.; Vercauteren, S.; Aperis, A.; Komendová, L.; Prozorov, R.; Partoens, B.; Milošević, M.V. Anisotropic type-I superconductivity and anomalous superfluid density in OsB2. Phys. Rev. B 2016, 94, 144506. [Google Scholar] [CrossRef]
  53. Klein, T.; Braithwaite, D.; Demuer, A.; Knafo, W.; Lapertot, G.; Marcenat, C.; Rodière, P.; Sheikin, I.; Strobel, P.; Sulpice, A.; et al. Thermodynamic phase diagram of Fe(Se0.5Te0.5) single crystals in fields up to 28 tesla. Phys. Rev. B 2010, 82, 184506. [Google Scholar] [CrossRef]
  54. Diaconu, A.; Martin, C.; Hu, J.; Liu, T.; Qian, B.; Mao, Z.; Spinu, L. Possible nodal superconducting gap in Fe1+y(Te1−xSex) single crystals from ultralow temperature penetration depth measurements. Phys. Rev. B 2013, 88, 104502. [Google Scholar] [CrossRef]
  55. Kumar, R.; Varma, G.D. Study of TAFF and vortex phase of FexTe0.60Se0.40 (0.970 ≤ x ≤ 1.030) single crystals. Phys. Scr. 2020, 95, 045814. [Google Scholar] [CrossRef]
  56. Poole, C.; Farach, H.; Creswick, R.; Prozorov, R. Superconductivity; Academic Press: Cambridge, MA, USA, 2007; ISBN 0080550487. [Google Scholar]
  57. Peri, A.; Mangel, I.; Keren, A. Superconducting Stiffness and Coherence Length of FeSe0.5Te0.5 Measured in a Zero-Applied Field. Condens. Matter 2023, 8, 39. [Google Scholar] [CrossRef]
  58. Bardeen, J. Critical fields and currents in superconductors. Rev. Mod. Phys. 1962, 34, 667–681. [Google Scholar] [CrossRef]
  59. Maiorov, B.; Mele, P.; Baily, S.A.; Weigand, M.; Lin, S.Z.; Balakirev, F.F.; Matsumoto, K.; Nagayoshi, H.; Fujita, S.; Yoshida, Y.; et al. Inversion of the upper critical field anisotropy in FeTeS films. Supercond. Sci. Technol. 2014, 27, 044005. [Google Scholar] [CrossRef]
  60. Her, J.L.; Kohama, Y.; Matsuda, Y.H.; Kindo, K.; Yang, W.H.; Chareev, D.A.; Mitrofanova, E.S.; Volkova, O.S.; Vasiliev, A.N.; Lin, J.Y. Anisotropy in the upper critical field of FeSe and FeSe0.33Te0.67 single crystals. Supercond. Sci. Technol. 2015, 28, 045013. [Google Scholar] [CrossRef]
  61. Sun, Y.; Pan, Y.; Zhou, N.; Xing, X.; Shi, Z.; Wang, J.; Zhu, Z.; Sugimoto, A.; Ekino, T.; Tamegai, T.; et al. Comparative study of superconducting and normal-state anisotropy in Fe1+yTe0.6Se0.4 superconductors with controlled amounts of interstitial excess Fe. Phys. Rev. B 2021, 103, 224506. [Google Scholar] [CrossRef]
  62. Mishev, V.; Nakajima, M.; Eisaki, H.; Eisterer, M. Effects of introducing isotropic artificial defects on the superconducting properties of differently doped Ba-122 based single crystals. Sci. Rep. 2016, 6, 27783. [Google Scholar] [CrossRef] [PubMed]
  63. Kondo, K.; Motoki, S.; Hatano, T.; Urata, T.; Iida, K.; Ikuta, H. NdFeAs(O,H) epitaxial thin films with high critical current density. Supercond. Sci. Technol. 2020, 33, 09LT01. [Google Scholar] [CrossRef]
  64. Li, J.; Yuan, J.; Yuan, Y.H.; Ge, J.Y.; Li, M.Y.; Feng, H.L.; Pereira, P.J.; Ishii, A.; Hatano, T.; Silhanek, A.V.; et al. Direct observation of the depairing current density in single-crystalline Ba0.5K0.5Fe2As2 microbridge with nanoscale thickness. Appl. Phys. Lett. 2013, 103, 62603. [Google Scholar] [CrossRef]
  65. Bristow, M.; Knafo, W.; Reiss, P.; Meier, W.; Canfield, P.C.; Blundell, S.J.; Coldea, A.I. Competing pairing interactions responsible for the large upper critical field in a stoichiometric iron-based superconductor CaKFe4As4. Phys. Rev. B 2020, 101, 134502. [Google Scholar] [CrossRef]
  66. Sun, Y.; Ohnuma, H.; Ayukawa, S.-Y.; Noji, T.; Koike, Y.; Tamegai, T.; Kitano, H. Achieving the depairing limit along the c axis in Fe1+yTe1−xSex single crystals. Phys. Rev. B 2020, 101, 134516. [Google Scholar] [CrossRef]
  67. Galluzzi, A.; Buchkov, K.; Tomov, V.; Nazarova, E.; Kovacheva, D.; Leo, A.; Grimaldi, G.; Pace, S.; Polichetti, M. Mixed state properties of iron based Fe(Se,Te) superconductor fabricated by Bridgman and by self-flux methods. J. Appl. Phys. 2018, 123, 233904. [Google Scholar] [CrossRef]
  68. Tsurkan, V.; Deisenhofer, J.; Günther, A.; Kant, C.; Klemm, M.; von Nidda, H.-A.; Schrettle, F.; Loidl, A. Physical properties of FeSe0.5Te0.5 single crystals grown under different conditions. Eur. Phys. J. B 2011, 79, 289–299. [Google Scholar] [CrossRef]
  69. Wittlin, A.; Aleshkevych, P.; Przybylińska, H.; Gawryluk, D.J.; Dłuzewski, P.; Berkowski, M.; Puźniak, R.; Gutowska, M.U.; Wiśniewski, A. Microstructural magnetic phases in superconducting FeTe0.65Se0.35. Supercond. Sci. Technol. 2012, 25, 065019. [Google Scholar] [CrossRef]
  70. Sivakov, A.G.; Bondarenko, S.I.; Prokhvatilov, A.I.; Timofeev, V.P.; Pokhila, A.S.; Koverya, V.P.; Dudar, I.S.; Link, S.I.; Legchenkova, I.V.; Bludov, A.N.; et al. Microstructural and transport properties of superconducting FeTe0.65Se0.35 crystals. Supercond. Sci. Technol. 2017, 30, 015018. [Google Scholar] [CrossRef]
  71. McQueen, T.M.; Huang, Q.; Ksenofontov, V.; Felser, C.; Xu, Q.; Zandbergen, H.; Hor, Y.S.; Allred, J.; Williams, A.J.; Qu, D.; et al. Extreme sensitivity of superconductivity to stoichiometry in Fe1+δSe. Phys. Rev. B 2009, 79, 014522. [Google Scholar] [CrossRef]
  72. Onar, K.; Yakinci, M.E. Solid state synthesis and characterization of bulk β-FeSe superconductors. J. Alloys Compd. 2015, 620, 210–216. [Google Scholar] [CrossRef]
  73. Fiamozzi Zignani, C.; De Marzi, G.; Grimaldi, G.; Leo, A.; Guarino, A.; Vannozzi, A.; della Corte, A.; Pace, S. Fabrication and Physical Properties of Polycrystalline Iron-Chalcogenides Superconductors. IEEE Trans. Appl. Supercond. 2017, 27, 1–5. [Google Scholar] [CrossRef]
  74. Galluzzi, A.; Nigro, A.; Fittipaldi, R.; Guarino, A.; Pace, S.; Polichetti, M. DC magnetic characterization and pinning analysis on Nd1.85Ce0.15CuO4 cuprate superconductor. J. Magn. Magn. Mater. 2019, 475, 125–129. [Google Scholar] [CrossRef]
  75. Galluzzi, A.; Mancusi, D.; Cirillo, C.; Attanasio, C.; Pace, S.; Polichetti, M. Determination of the Transition Temperature of a Weak Ferromagnetic Thin Film by Means of an Evolution of the Method Based on the Arrott Plots. J. Supercond. Nov. Magn. 2018, 31, 1127–1132. [Google Scholar] [CrossRef]
Figure 1. The field dependence of the initial magnetic moment curve is plotted for different temperatures. The black dashed line provides the linear fit for the low field m(H) curves.
Figure 1. The field dependence of the initial magnetic moment curve is plotted for different temperatures. The black dashed line provides the linear fit for the low field m(H) curves.
Condensedmatter 08 00091 g001
Figure 2. (a) Comparison between the temperature dependence of the lower critical field H c 1 (black squares) (where the black solid line is the fit of the data using the following equation, H c 1 T = H c 1 0 1 T T c n ) and the temperature dependence of the lower critical field after considering the demagnetization effects H c 1 d e m a g (red circles) (where the red dashed line is the fit of the data using the following equation H c 1 d e m a g T = H c 1 d e m a g 0 1 T T c n ). When performing the fit, H c 1 (0) = 143 Oe and n = 1.54 were obtained without considering the demagnetizing factor, and H c 1 d e m a g (0) = 597 Oe and n = 1.54 were obtained by considering the demagnetizing factor. (b) Ratio of the H c 1 d e m a g ( T ) and H c 1 ( T ) values. The red star indicates the ratio of H c 1 d e m a g ( 0 ) / H c 1 (0).
Figure 2. (a) Comparison between the temperature dependence of the lower critical field H c 1 (black squares) (where the black solid line is the fit of the data using the following equation, H c 1 T = H c 1 0 1 T T c n ) and the temperature dependence of the lower critical field after considering the demagnetization effects H c 1 d e m a g (red circles) (where the red dashed line is the fit of the data using the following equation H c 1 d e m a g T = H c 1 d e m a g 0 1 T T c n ). When performing the fit, H c 1 (0) = 143 Oe and n = 1.54 were obtained without considering the demagnetizing factor, and H c 1 d e m a g (0) = 597 Oe and n = 1.54 were obtained by considering the demagnetizing factor. (b) Ratio of the H c 1 d e m a g ( T ) and H c 1 ( T ) values. The red star indicates the ratio of H c 1 d e m a g ( 0 ) / H c 1 (0).
Condensedmatter 08 00091 g002
Figure 3. Temperature dependence of the London penetration depth λ L obtained from H c 1 (black squares) and H c 1 d e m a g (red circles). Both the curves were fitted with λ L T = λ L 0 1 T T c n . When performing the fit, λ L (0) = 208 nm and n = −0.72 were obtained without considering the demagnetizing factor, whereas λ L d e m a g (0) = 92 nm and n = −0.72 were obtained by considering the demagnetizing factor.
Figure 3. Temperature dependence of the London penetration depth λ L obtained from H c 1 (black squares) and H c 1 d e m a g (red circles). Both the curves were fitted with λ L T = λ L 0 1 T T c n . When performing the fit, λ L (0) = 208 nm and n = −0.72 were obtained without considering the demagnetizing factor, whereas λ L d e m a g (0) = 92 nm and n = −0.72 were obtained by considering the demagnetizing factor.
Condensedmatter 08 00091 g003
Figure 4. Temperature dependence of the London penetration depth λ L 2 , obtained from H c 1 (black squares) and H c 1 d e m a g (red circles). The solid lines are a guide for the eyes.
Figure 4. Temperature dependence of the London penetration depth λ L 2 , obtained from H c 1 (black squares) and H c 1 d e m a g (red circles). The solid lines are a guide for the eyes.
Condensedmatter 08 00091 g004
Figure 5. Temperature dependence of the coherence length, ξ (black squares), together with its fit and the following equation ξ T = ξ 0 1 T T c n . After performing the fit, ξ 0 ≈ 3 nm and n 0.64 were obtained.
Figure 5. Temperature dependence of the coherence length, ξ (black squares), together with its fit and the following equation ξ T = ξ 0 1 T T c n . After performing the fit, ξ 0 ≈ 3 nm and n 0.64 were obtained.
Condensedmatter 08 00091 g005
Figure 6. Temperature dependence of the de-pairing current density, J d e p with (red closed circles) and without (black closed squares) consideration of the demagnetizing factor. The solid lines are a guide for the eyes.
Figure 6. Temperature dependence of the de-pairing current density, J d e p with (red closed circles) and without (black closed squares) consideration of the demagnetizing factor. The solid lines are a guide for the eyes.
Condensedmatter 08 00091 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Galluzzi, A.; Buchkov, K.; Tomov, V.; Nazarova, E.; Leo, A.; Grimaldi, G.; Crisan, A.; Polichetti, M. The Depairing Current Density of a Fe(Se,Te) Crystal Evaluated in Presence of Demagnetizing Factors. Condens. Matter 2023, 8, 91. https://doi.org/10.3390/condmat8040091

AMA Style

Galluzzi A, Buchkov K, Tomov V, Nazarova E, Leo A, Grimaldi G, Crisan A, Polichetti M. The Depairing Current Density of a Fe(Se,Te) Crystal Evaluated in Presence of Demagnetizing Factors. Condensed Matter. 2023; 8(4):91. https://doi.org/10.3390/condmat8040091

Chicago/Turabian Style

Galluzzi, Armando, Krastyo Buchkov, Vihren Tomov, Elena Nazarova, Antonio Leo, Gaia Grimaldi, Adrian Crisan, and Massimiliano Polichetti. 2023. "The Depairing Current Density of a Fe(Se,Te) Crystal Evaluated in Presence of Demagnetizing Factors" Condensed Matter 8, no. 4: 91. https://doi.org/10.3390/condmat8040091

Article Metrics

Back to TopTop