Next Article in Journal
Experimental Study of Auxetic Structures Made of Re-Entrant (“Bow-Tie”) Cells
Previous Article in Journal
Development and Modification of New or Recycled Materials and Technological Processes towards Sustainable Development
Previous Article in Special Issue
Fe(Se,Te) Thin Films Deposited through Pulsed Laser Ablation from Spark Plasma Sintered Targets
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Properties and Applications of Iron–Chalcogenide Superconductors

by
Jianlong Zhao
1,2,
Junsong Liao
1,2,
Chiheng Dong
1,2,3,*,
Dongliang Wang
1,2,3,* and
Yanwei Ma
1,2,3
1
Key Laboratory of Applied Superconductivity, Institute of Electrical Engineering, Chinese Academy of Sciences, Beijing 100190, China
2
School of Electronic, Electrical and Communication Engineering, University of Chinese Academy of Sciences, Beijing 100049, China
3
Institute of Electrical Engineering and Advanced Electromagnetic Drive Technology, Qilu Zhongke, Jinan 250013, China
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(13), 3059; https://doi.org/10.3390/ma17133059
Submission received: 28 March 2024 / Revised: 10 May 2024 / Accepted: 10 May 2024 / Published: 21 June 2024

Abstract

:
Iron–chalcogenide superconductors continue to captivate researchers due to their diverse crystalline structures and intriguing superconducting properties, positioning them as both a valuable platform for theoretical investigations and promising candidates for practical applications. This review begins with a comprehensive overview of the fabrication techniques employed for various iron–chalcogenide superconductors, accompanied by a summary of their phase diagrams. Subsequently, it delves into the upper critical field, anisotropy, and critical current density. Furthermore, it discusses the successful fabrication of meters-long coated conductors and explores their applications in superconducting radio-frequency cavities and coils. Finally, several prospective avenues for future research are proposed.

1. Introduction

Iron-based superconductors [1], discovered in 2008, are considered to be a new platform to study high-temperature superconducting mechanisms and potential candidates for practical applications [2,3,4]. From the crystalline structure point of view, iron-based superconductors can be divided into two major categories, iron–pnictides with the [FeAs]−1 layer and iron–chalcogenides with the electroneutral FeSe layer. The [FeAs]−1 layers in iron–pnictides are responsible for charge carrier transfer, which is alternated by the charge reservoir layers, e.g., the [ReO]+1, Ae2+, and A+1 layers (Re = rare earth elements, Ae = alkaline earth elements, A = alkaline elements) [5]. Consequently, there are varieties of iron–pnictides with multiple structures and abbreviations, such as 111, 122, 1111, 12,442, et al., as shown in Figure 1a. On the contrary, the iron–chalcogenides’ 11 system contains non-toxic elements and possesses the simplest crystal structure among iron-based superconductors, making them more attractive for theoretical studies and practical applications. Here, in this review, we start with the fabrication methods for iron–chalcogenide and discuss the superconducting transition temperatures related to different crystal structures and charge carrier densities. Then, the important factors for determining the applications, the upper critical field and the critical current density, are summarized for single crystals and films. At last, recent applications of FeTe1−xSex-coated conductors are discussed.

2. Fabrications, Crystal Structures, and Superconducting Transition Temperature of Iron–Chalcogenide Superconductors

The substitution of Se by Te or S induces novel superconductivity with higher superconducting transition temperatures, Tc [14,15]. However, there is a common characteristic in the 11 system, which is the excess Fe atoms (Fe(2) in Figure 1a) existing between the FeSe layers, as shown by the bright spots in the Scanning Tunneling Microscope (STM) images in Figure 2b,i [16]. They cause carrier localization and provide magnetic order deteriorating superconductivity in the Fe(1) plane [17], which masks the intrinsic superconductivity. It was soon found that annealing in air, O2, S, Se, Te, P, As, Sb, and I atmospheres can effectively remove the excess iron [18], as shown in Figure 2c,d. Correspondingly, the semiconductor-like normal state, namely, the resistivity (ρ), increases with cooling and changes to metallic-like behavior [19] with the increased annealing temperature, as shown in Figure 2a. Moreover, the Tc is enhanced from 9 K to 14 K, and the superconducting transition becomes sharper. The superconducting volume fraction of FeTe0.6Se0.4 approaches 100% after optimizing the annealing conditions. Gu et al. reported that through Mn element doping, the suppressed superconductivity in the pristine FeTe0.55Se0.45 single crystals can largely recover. As shown in the inset of Figure 2e, 1% Mn doping into the Fe site increases the Tconset to 15 K. The quick saturation of susceptibility and the obvious jump in the temperature dependence of specific heat in Figure 2f,g suggest that bulk superconductivity emerges after the excess Fe atoms are removed, as evidenced by Figure 2j. They are summarized in a Mn-doping phase diagram, as shown in Figure 2h. Minor Mn doping of 0.5% has increased the Tc to a plateau value of 14.5 K, while further doping suppresses Tc and the jump of specific heat.
The interlayer coupling of FeSe is dominated by the Van der Waals Force. The intercalation with metal atoms can be fulfilled through the traditional high-temperature synthesis routes and low-temperature ammonothermal techniques. The (Tl,K,Rb)xFeySe2 single crystals are usually fabricated through slow cooling from the melt at temperatures exceeding 900 °C. By increasing the Fe content, the normal state of (Tl,K)FeySe2 was tuned from an insulator to a semiconductor, accompanied by a superconducting transition at Tc~31 K [9]. However, systematic investigation using a scanning electron microscope (SEM) and STM reveals a severe phase separation existing in this system. The filament KFe2Se2 superconducting phase (as shown by the rectangles in Figure 3b,c is sandwiched by the insulating K2Fe4Se5 phase with a Néel temperature of 560 K (the dark matrix) [20]. Li et al. found that the STM image in region I of Figure 3d exhibits a centered rectangular lattice structure with very few defects, as shown by Figure 3e. The Scanning Tunneling Spectroscopy (STS) reveals a double-gap structure, which is uniform along the white line in Figure 3e [21]. In region II, the STS indicates an energy gap of 0.43 eV across the Fermi level, suggesting an insulation phase. There are obvious Fe vacancies in the STM images, constructing a so-called 5 × 5 superlattice.
It is known that liquid ammonia dissolves alkali, alkaline earth, and some rare earth elements. Burrard-Lucas et al. first synthesized the Fe1+δSe polycrystals using the solid-state-reaction method. The Fe1+δSe and the Li metal were then placed in a Schlenk tube with a magnetic stirrer, which was attached to the ammonia cylinder. The tube was cooled to allow the ammonia to condense onto the reactants. The mixtures were then stirred at −78 °C for half an hour, after which the Schlenk tube was warmed up and the ammonia was evaporated. Finally, the Lix(NH2)y(NH3)1-yFe2Se2 (x~0.6; y~0.2) with Tc = 43 K was obtained [11]. Ying et al., instead, directly poured the liquid ammonia into an autoclave filled with β-FeSe powders and alkali, alkaline earth, Yb, and Eu at liquid nitrogen temperature. The whole device was then kept at room temperature for 2–17 days [22]. Varieties of FeSe-based superconductors intercalated with Li, Na, Ba, Sr, Ca, Yb, and Eu were obtained, as proved by the Rietveld refinement shown in Figure 4a. The highest Tc is 45 K for NaFe2Se2. As shown in Figure 4b, the sharp superconducting transition and the large magnetization hysteresis loop indicate bulk superconductivity. By using the high-purity FeSe powders, they continued to intercalate K in FeSe via the ammonothermal route [23]. At least two superconducting phases were discovered, of which the Tc was controlled by the nominal content of K rather than the NH3. Discrete superconducting regions were revealed with abrupt changes of Tc with doping, as shown in Figure 5a,b. When x in KxFe2Se2(NH3)y is below 0.3, Tconset is 44 K. When 0.3 < x < 0.6, a two-step transition was observed at Tc1~44 K and Tc2~30 K, indicating the phase separation in this doping region. There is only one superconducting phase with Tc~30 K when x > 0.6. Moreover, there is no sign of antiferromagnetic (AFM) transition in the temperature-dependence of magnetization, indicating that the Fe-vacancy phase, such as the 245 phase, disappears in KxFe2Se2(NH3)y. To further investigate the intrinsic superconducting properties, Sun et al. synthesize Lix(NH3)yFe2Se2 single crystals through the low-temperature ammonothermal technique. The FeSe single crystals fabricated through the eutectic AlCl3 and NaCl (=0.52:0.48) flux method were used as the template. The as-synthesized thin plate-like Lix(NH3)yFe2Se2 single crystals are shown in the inset of Figure 5d, with sharp (00l) diffraction peaks in the XRD patterns. The lattice parameters are a = 3.7704(6) Å and c = 16.973(7) Å. The Tconset is determined to be 44.3 K. Surprisingly, the crystals exhibit a huge anisotropy in resistivity. The ab-plane resistivity ρab exhibits a semiconductor behavior above 150 K, while the c-axis resistivity ρc presents a metallic behavior. Furthermore, the ratio ρc/ρab increases from 900 at 300 K to 8000 at 50 K, as shown in Figure 5e.
Guo et al. utilized a similar method to synthesize the NH3-free Na-intercalated FeSe1-zSz superconductors. The difference is that the Taiatsu Glass TVS-N2 high-pressure vessel continued to be evacuated to ~10−2 Pa using a molecular pump after the reaction was finished [24]. The X-ray diffraction pattern shown in Figure 4c,d can be fitted well by Na0.65Fe1.93Se2 (91.3%), with minor impurities of FeSe (3.5%) and Fe7Se8 (5.2%). As depicted in Figure 4e, the NH3-free phase has a Tc of 37 K, while the NH3-poor intercalates show a Tc of 45 K. Figure 4f summarizes the S-doping-dependence of Tc for the FeSe1-zSz and two intercalates. With S doping, the Tc decreases quicker in the NH3-free phase than in the NH3-poor phase, indicating that S doping effect is weaker in the NH3-poor phase.
Except for the low-temperature ammonothermal method, Hatakeda et al. dissolved lithium and sodium metal in 2-phenethylamine (2-PEA) and reacted them with FeSe powders at 45 °C for 7 and 28 days, respectively [26]. The c-axis length is 19.04(6) Å and 18.0(1) Å for the Li- and Na-cases, respectively, which is the largest among the iron–chalcogenide superconductors. Tc is 39 K for the Li-case, while a two-step transition is observed in the Na-case. The thermalgravimetric measurements show a 27% mass loss below 300 °C, which is ascribed to the deintercalation or desorption of 2-PEA, and a nearly 60% mass loss above 700 °C. This implies a high instability of the intercalated FeSe at high temperatures.
Shi et al. performed an electrochemical intercalation method to successfully synthesize a cetyltrimethyl ammonium (CTA+)-intercalated FeSe-based superconductor (CTA)0.3FeSe with Tc = 45 K [12]. The crystal structure is exhibited in Figure 1b. The FeSe single crystals were used as the starting materials and fixed on an In wire as the positive electrode. The electrolyte is N-methyl-2-pyrrolidone added by hexadecyl trimethyl ammonium bromide (CTAB). The Transmission Electron Microscope (TEM) indicated that the distance between the FeSe layers increased from 0.56 nm for the pristine FeSe to 1.45 nm for the CTA+ intercalated FeSe. The CTA+ organic layer can be clearly seen, as shown by the darker layer in Figure 6g. Figure 6a shows the temperature dependence of susceptibility measured under the fields of 5 Oe and 40 Oe. At 5 Oe, the Tc is determined to be 45 K, and the small distance between the FC and ZFC branches implies weak flux-pinning properties, confirming the high degree of homogeneity of the intercalated sample. The Tc also exhibits a negative pressure effect up to 1 GPa, which is consistent with other pressure experiments on the intercalated FeSe systems under low pressure. Meng et al. cleaved the FeSe single crystals obtained through the chemical vapor transport (CVT) method into thin flakes with a thickness of ~15 μm [27]. Several crystals were attached to the negative electrode using silver paint. The ionic liquid 1-ethyl-3-methylimidazolium tetrafluoroborate (EMIM-BF4) was used as the electrolyte, as schematized in Figure 6i. By using the optimized voltage 3 V and a temperature of 330 K provided by a thermostatic hot plate, the intercalated FeSe samples with different H+ content were obtained after different protonation times. As depicted in Figure 6j–l, the nematic transition at Ts~89 K marked by the black arrows in (j) was gradually suppressed after protonation and totally disappeared after 5 days. Meanwhile, the Tc is enhanced from 10 K for the pristine FeSe single crystals to Tc1 = 25 K at 5 days of protonation. In the intermediate protonation time, there is a coexistence of three superconducting transitions with Tc2 = 25 K, Tc3 = 34 K, and Tc4 = 44 K. After 15 days, only the Tc4 = 44 K phase survived. Moreover, the magnetization measurements prove the 100% volume fraction of superconductivity for all of the samples at 2 K. They also proved that this effect was reversible after deprotonation, as the Tc4 phase gradually transformed to the Tc1 phase, accompanied by the reappearance of the nematic order.
Lei et al. applied the electric-double-layer transistor (EDLT) using ionic liquids as the gate dielectric to the single crystalline FeSe flakes with typical thicknesses of ~10 nm [28], as shown in Figure 7a,b. The ionic liquid DEME-TFSI was used as the dielectric, and the cations and anions were DEME+ and TFSI-, respectively. After introducing carriers to the FeSe flakes at 220 K with the application of gate voltages, the sheet resistance was measured, as depicted in Figure 7e. After increasing the gate voltage Vg, the pristine Tc = 5.2 K was enhanced to 7.5 K at Vg = 4.0 V. Further increase of Vg induces a two-step transition. Above Vg > 5.0 V, there is only one superconducting phase with Tc above 40 K. The highest Tc~48.2 K is realized after applying Vg = 6 V. Similar results were observed in the Li/Na intercalated FeSe superconductors by using the ionic gating technique [29]. The lithium-based Li1+x+yAx(Ti,Ge)2−xSiyP3−yO12 and the sodium-based Na3.4Zr1.8Mg0.2Si2PO12 were adopted as the substrates, as shown in Figure 7c,d. The 5–38 nm thick FeSe flake was sandwiched between the solid ionic substrate and a 100 nm SiO2 layer. The SiO2 layer is crucial to insulate the electrodes from the substrate to avoid metal accumulation and accurate determination of the content of the intercalated metals. It can be observed that with intercalation of Li cations, the Tc1 = 8 K from the FeSe matrix gradually increases to Tc2 = 36 K and finally enters into the phase with Tc3 = 44 K, accompanied by decreased normal state resistivity. However, the normal state resistance gradually increases upon Li doping, and, finally, the Tc3 = 44 K phase disappears and the sample transforms into an insulator, as shown in Figure 7g. Similar intercalation behavior was found in NaxFeSe. The common discrete superconducting phase in these systems was found to be robust against S substitution but vulnerable to the Fe site substitution by Cu, implying the importance of the intact Fe plane to the high-temperature superconductivity.
(Li1−xFex)OHFeSe was first prepared through the hydrothermal reaction method. Selenourea, Fe powders, LiOH·H2O, and de-ionized water were put into an autoclave and heated at 160 °C for 3–10 days [30]. The crystal structure was determined through XRD and Rietveld refinements (Figure 8a). As depicted in Figure 1b, the anti-PbO-type FeSe layers are alternated with the anti-PbO-type LiFeO2. The crystal lattice parameters are determined to be a = 3.7926(1) Å, c = 9.2845(1) Å. The susceptibility shows a superconducting transition at 40 K. Sun et al. systematically studied the superconducting properties of (Li1−xFex)OHFe1-ySe with different iron vacancy concentrations [31]. It was found that the synthesized compounds through hydrothermal synthesis lead to x~0.2, but with a variable y. Interestingly, superconductivity emerges when Fe vacancy (y < 0.05) is decreased. The lithiation process reduces y to zero by replacing some Fe ions in the reservoir layer by Li, while the replaced Fe fills into the vacancy in the conducting layer. As a result, a large superconducting volume was reached with the highest Tc. Due to the powder form, the resistivity measurements on the pressed bulk show a low Tczero. Dong et al. successfully synthesized single crystalline (Li0.84Fe0.16)OHFe0.98Se via the hydrothermal ion-exchange technique based on the K0.8Fe1.6Se2 (245 insulating phase) single crystals [32]. During the hydrothermal reaction process, the K ions in the 245 single crystals were completely released into the solution, and (Li/Fe)OH layers constructed by ions from the solution were squashed into the matrix, which connects with the adjacent edge-sharing FeSe tetrahedra via weak hydrogen bonding, as schematized in Figure 8c. Despite the sharp superconducting transition at 42 K in the susceptibility data, clear resistivity evidence was also observed, as shown in Figure 8g.
The highest Tc in iron–chalcogenide superconductors was obtained in the single-layer FeSe film prepared through the Molecular Beam Epitaxy (MBE) method [33,34]. The FeSe single-layer film was grown on the TiO2-terminated and Nb-doped SrTiO3 (STO-001) substrate at 450 °C by co-evaporating Fe and Se from Knudsen cells with a flux ratio of 1:10. RHEED was used to monitor the growth process, as shown in Figure 9a. Figure 9b shows the temperature dependence of resistance derived from the I–V curves measured at fixed temperatures. The Tc is determined to be 109 K. Another R–T curve with Tc = 99 K was obtained through measurements conducted while sweeping the temperature, as shown in Figure 9c. As a result, it can be claimed that this is the second high-temperature superconductor with Tc surpassing the liquid nitrogen temperature, indicating great significance not only for the superconducting mechanism but also for potential applications in the liquid nitrogen temperature region. Xue’s group continues to investigate the mechanism [35] of high-temperature superconductivity [35]. They suggest that band bending of STO occurs due to the larger work function of FeSe compared to STO, which contributes to the charge transfer to the FeSe side, as shown in Figure 9d. Moreover, annealing induces stoichiometry of FeSe and shifts the Fermi level upward, thus increasing the electron density (Figure 9e,f). Consequently, the high-temperature superconductivity emerges in the superconducting/dielectric heterostructure, where the dielectric layer enhances the charge density of FeSe and provides the electron–phonon coupling with its intrinsic high-energy phonon mode.

3. Phase Diagram of Iron–Chalcogenide Superconductors

The phase diagrams of iron–chalcogenide superconductors are summarized in Figure 10. The left side of Figure 10a shows the FeSe1−xSx single crystals fabricated through the hydrothermal route (0.29 ≤ x ≤ 1) and the CVT method (0 ≤ x ≤ 0.29) [37]. In the nematic phase below Ts, the normal state resistivity presents a linear temperature dependence, namely, the non-Fermi liquid behavior [38]. With S doping, the nematic order is gradually suppressed, and the Fermi liquid behavior is recovered. A small superconducting dome appears within the nematic phase, with the highest Tc~9 K. It reaches a minimum value at x~0.45 and increases again to the FeS end. The right side of Figure 10a shows the Te doping dependence of Tc and TN. The post-annealing process applied to the high Te content of FeSe1−xTex single crystals prepared through the self-flux method proves the coexistence of long-range AFM order and superconductivity. However, high Se side crystals must be synthesized through the CVT method to avoid phase separation [39,40]. The Tc of FeSe1−xTex presents a minimum around x~0.2 due to the sample disorder. The nematic order is also gradually suppressed by Te doping and disappears at x~0.5, where the Tc reaches the maximum value of 15 K. Zhuang et al. also obtained a full doping phase diagram of FeSe1−xTex by synthesizing the film on the CaF2 (100) substrate using the Pulsed Laser Deposition (PLD) method [41]. As marked by the green pentagon in Figure 10a, the Tc shows a maximum value of 20 K within 0.6 ≤ x ≤ 0.8, where phase separation should be apparent. High transition temperatures with Tc~24 K were obtained in FeTe0.8Se0.2 film [42,43]. The difference between the film and the CVT single crystals may originate from the strain effect induced by the substrate. Figure 10b shows the Te doping phase diagram of FeS1−xTex. Similarly to the doped FeSe, S doping in FeTe gradually suppresses the AFM order, while superconductivity appears when x < 0.95 [44]. There is a region of coexistence of AFM and superconductivity. The maximum Tc is 9 K. However, the solubility limit of S in FeTe prevents further investigation into this system. Zhao et al. prepared the FeS1−xTex single crystals with 0 ≤ x ≤ 0.15 via the hydrothermal technique [45]. The Tc = 4.5 K in FeS [46] was quickly suppressed after Te doping due to the scattering from the impurities, and the superconductivity was totally suppressed when x > 0.1.
Ying et al. summarized a common electron doping phase diagram of the intercalated FeSe superconductors, as shown in Figure 10c. An obvious discreet superconducting phase diagram can be seen [27,29]. The corresponding Fermi surface topologies are also incorporated at the top. For the pristine FeSe system, the hole pockets are at the center of the Brillouin zone (Γ point), and the electron pockets are at the corner (M points). Upon electron doping to a critical level, the Lifshitz transition occurs and leads to an abrupt appearance of the Tc2 phase. Further increasing the doping level induces the enlarged electron pocket and the emergence of the Tc3 phase. Figure 10d depicts Tc as a function of the interlayer distance, d, for all of the FeSe-based superconductors published so far. It seems that the Tc is enhanced with the interlayer distance when d < 9 Å, but it saturates with the higher d. For the protonated FeSe sample, the Tc is enhanced by increasing the electron density, rather than the interlayer distance, as proved by the enhancement from Tc1 to Tc4 with the same interlayer distance. Rebec et al. explain the phase diagram from the viewpoint of the superconducting energy gap, Δ [47]. As schematized in Figure 10e, there is a linear Tc dependence on the gap 2Δ/kBTc = 5.87, which demarcates the phase diagram into three regions. For region I, the FeSe/Te bulks exhibit a Δ < 5 meV, and the Tc is less than 20 K. In the second region, a gap of Δ~10 meV is opened at a low temperature, and the Tc of the intercalated or electron-doped FeSe is enhanced to 31–48 K [48]. The 1-unic cell FeSe on the STO substrate belongs to the third group, where interfacial electron–phonon interaction plays an important part in high-temperature superconductivity.

4. Upper Critical Field, Anisotropy, and Critical Current Density of Practical Iron–Chalcogenide Superconductors

Two crucial parameters related to practical applications are the upper critical field, above which Cooper pairs are depaired at high magnetic fields, and the anisotropy parameter, γ = Bc2ab/Bc2c = (mc/mab)1/2 = λcab = ξcab, where m, λ, and ξ are the effective mass of electrons, the penetration depth, and the coherence length, respectively. Figure 11a summarizes the temperature dependence of the upper critical field, Bc2. One similarity is that the Bc2 with the field parallel to the ab-plane of the crystal structure is larger than the field parallel to the c-axis. This is caused by the effective mass anisotropy of electrons. For the FeSe (100 nm thick flake), the Bc2 is 16 T at 4 K when the field is parallel to the ab-plane. [49]. The Bc2 with the field parallel to the c-axis is smaller, achieving a Bc2 = 12 T at 2 K. Te doping in the Se site not only enhances the Tc but also increases the Bc2 to 50 T at 2 K. Moreover, the Bc2ab is closer to the Bc2c below Tc than FeSe, and there is a crossover between them at 3 K. For the intercalated FeSe superconductors, the enhanced Tc makes the application in liquid hydrogen possible. For the (Li,Fe)OHFeSe single crystals, the Bc2c is increased to 45 T and 65 T at the liquid hydrogen and liquid helium temperature regions, respectively [50]. The Bc2ab, on the other hand, is 68 T at 27 K, indicating good application potential. The data of Lix(NH3)yFe2Se2 are limited below 14 T, but there is an apparent large gap between the two directions [25].
The anisotropy parameters of the above-mentioned superconductors are depicted as a function of normalized temperature, t = T/Tc, in Figure 11b. For the FeTe0.6Se0.4 single crystals, the γ increases from values smaller than 1 (reversed anisotropy with Bc2ab < Bc2c) below 0.25 Tc to 1 < γ < 2, and finally above γ = 2 near Tc. The FeSe flakes with infinite conducting layers exhibit more two-dimensional characteristics and exhibit a γ > 2 in the whole studying range. The nearly isotropic behavior of Bc2 is beneficial to the magnet design. For the (Li,Fe)OHFeSe with larger interlayer space, the γ increases from 2.5 to 5.5 when t = T/Tc moves from 0.6 to 0.8. The largest γ is observed in Lix(NH3)yFe2Se2, ranging from 8 to 18 above t = 0.38 [25]. The extremely large electronic anisotropy in both (Li,Fe)OH and Li-NH3 layers of intercalated FeSe compounds is suggested to result from the structure’s characteristics.
In the next section, we will discuss the most important parameter, the critical current density of iron–chalcogenide superconductors, reported so far. Although the Jc has already been studied in the (Tl,K,Rb)xFeySe2 system [52,53,54], the discussion will not be included here because of the severe phase separation and the small portion of the superconducting phase, which leads to a small Jc~104 A/cm2 calculated through the Bean model. Instead, the (Li,Fe)OHFeSe system with an intact Fe plane and 100% volume fraction of superconductivity will be discussed.
Generally, the upper limit of critical current density is determined through the depairing critical current density, which is described by
J d 0   K = ϕ 0 3 3 μ 0 λ a b 2 0   K ξ a b ( 0   K )
where Φ0 is the flux quantum. The penetration depth and coherence length of three typical iron–chalcogenide superconductors are listed in Table 1. The ξ of FeSe is two to three times those of FeTe0.5Se0.5 and (Li,Fe)OHFeSe, leading to a small Jd of 11 MA/cm2. Comparatively, the Jd of the FeTe1−xSex and (Li,Fe)OHFeSe system achieves 36 and 57 MA/cm2 at 0 K.
However, the actual Jc a superconductor can carry is mainly controlled by the flux pinning efficiency of the pinning landscape, in which the dimensionality also determines the Jc anisotropy. As summarized in Table 1, the pristine FeSe single crystals only achieve a small Jc of 0.043 MA/cm2. After the irradiation of 2.4 GeV U ions, the Jc approaches the same level of the pristine FeTe0.61Se0.39 and (Li,Fe)OHFeSe single crystals [61]. Meanwhile, the irradiated FeTe0.61Se0.39 single crystals by 200 MeV Au achieve 0.5 MA/cm2 at 0 T and 2 K. The FeTe1−xSex film and coated conductor have a higher Jc above 1 MA/cm2 because of the more abundant crystal defects incorporated after the PLD fabrication process.
Figure 12 summarizes the critical current density of the state-of-the-art iron–chalcogenide superconductors. Generally speaking, the iron–chalcogenide single crystals have much fewer crystalline defects than the film, leading to a weak flux pinning force, strong field dependence, and low Jc at high fields. The FeSe single crystals show a Jc(2 K) = 4 × 104 A/cm2 at 0 T, which decreases quickly with the field. The FeTe1−xSex single crystals, on the contrary, exhibit a robust field dependence above 2 T. Irradiation is an effective way to introduce artificial pinning centers into the ‘11’ system, leading to a pronounced enhancement of its current carrying ability. The Jc of FeSe at high fields has been enhanced four times. For the irradiated FeTe1−xSex single crystals, the Au irradiation has increased the Jc below 5 T, above which the Jc before and after irradiation intersect with each other. The (Li,Fe)OHFeSe single crystals have a higher Jc than the FeTe1−xSex single crystals in the testing range due to the intrinsic pinning, which will be discussed below [63]. Comparatively, the iron–chalcogenides in the film form have higher Jc because of the strong flux pinning centers constructed during the film growth [42,64,66,67,68]. The FeTe1−xSex film achieves a high Jc(0 T) = 1.36 MA/cm2 and exhibits a rather robust field dependence [69]. Even at 9 T, the Jc remains 0.97 MA/cm2. The (Li,Fe)OHFeSe film grown through the hydrothermal method shows a Jc below 105 A/cm2 above 20 T [58]. However, the extrapolation to the low field region seems higher than the (Li,Fe)OHFeSe single crystals. Moreover, the Mn doping considerably enhances the high field Jc by nearly 10 times at 33 T, which is ascribed to the extra pinning centers introduced by Mn doping [58]. A similar Mn doping effect on Jc is observed in the FeTe1−xSex [15] and KxFe2−ySe2 [53] systems, but the former is due to the extinguished excess Fe irons. Interestingly, the single-layer FeSe film has a Jc above 1 MA/cm2 at self-field. Such a high Jc in the one-unit cell FeSe slowly declines with fields, as shown in the inset.
The flux pinning mechanism is important to unravel the factors governing the critical current density. Figure 13 presents the flux pinning properties of the FeTe1−xSex and (Li,Fe)OHFeSe single crystals and films. The flux pinning type of superconductors is usually studied through the Dew–Hughes model or the angle dependence of Jc. The flux pinning mechanism can be obtained according to the peak of the fp-h curve, where fp is the normalized flux pinning force fp = Fp/Fpmax and h is the reduced magnetic field h = H/Hirr. The Fe1.04Te0.6Se0.4 single crystals have a hpeak = 0.3, indicating mixed pinning from the point and surface defects [70]. After 3at. % Co doping into the Fe sites, the enhanced Jc is still dominated by the point defects [71]. The FeTe0.5Se0.5 film deposited on the CaF2 film also shows a peak at hpeak = 0.33, suggesting a point pinning mechanism, as shown in Figure 13d. The angular Jc at 15 K in Figure 13e has a peak at the fields parallel to the ab-plane [67] due to the intrinsic pinning. As evidenced in Figure 13f, the Jc(Θ, H) curves collapse into one curve when it is plotted against the effective field, Heff = Hε(Θ), where ε(Θ) = (cos2Θ + γ−2sin2Θ)1/2. The fitted anisotropy parameter is γ = 1.28. It is suggestive that the flux pinning of the film can be attributed to the mixed randomly distributed point defects and mass anisotropy. The fp-h curves of the (Li,Fe)OHFeSe film before and after Mn doping are both close to the surface pinning [58]. This intrinsic pinning is contributed by the insulating (Li,Fe)OH spacer layer.
The naturally formed pinning centers are quite rare in iron–chalcogenide superconductors, making it necessary to introduce artificial pinning centers. Ion irradiation is considered to be an effective method for introducing point and column defects depending on the type and energy of the irradiation particles [72]. So far, several groups have performed irradiation experiments on the Fe(Te,Se) single crystals and films. As summarized in Figure 14, the large-sized particles, such as U, Xe, and Au ions, created continuous columnar defects in the FeSe (Figure 14a) and FeTe0.61Se0.39 (Figure 14b,c) single crystals. As shown in Figure 14d, the U-irradiation suppressed the Tc of FeSe even at a small dose, which demonstrates a nearly linear dependence of Tc on the dose [61]. The quick suppression of Tc when compared with iron pnictides can be explained by the much larger damaged areas (~10 nm) shown in the inset of Figure 14a. In the Ba0.6K0.4Fe2As2 single crystals, the damaged area is only ~2–5 nm. Consequently, one can observe a peak of the Jc-BΦ curve at a small BΦ =2 T, above which the Jc is quickly lowered due to the damaged superconducting matrix.
Ozaki et al. studied the irradiation effect in increasing Jc. For the low-energy Au irradiation, it produces cluster-like defects with sizes of 10–15 nm (Figure 15c,d), much larger than the coherence length [73]. The strong isotropic pinning leads to a nearly 70% increase of Jc at 9 T and 10 K with H//c. The low-energy proton irradiation introduces splayed cascade defects, which are about 1–2 nm in diameter and ~10 nm wide, in the FeTe1−xSex films deposited on the CeO2 buffer layer, as shown in Figure 15g. The cascade defects cause strained areas [74]. The blue valleys are the highly compressed regions, whereas the yellow peak regions indicate tensile strain (Figure 15h). The nanoscale strain and proximity effect causes an enhanced Tc from 18 K to 18.5 K, as well as the increased Hirr, Hc2, and Jc (Figure 15e,f). A similar proton irradiation effect can be found in the FeSe single crystals. Figure 15a,b show that the hpeak moves from 0.14 to 0.28, indicating the introduction of point defects [75]. The self-field Jc of the irradiated FeSe single crystals was enhanced from 3 × 104 A/cm2 to 8 × 104 A/cm2.

5. Practical Applications of Iron–Chalcogenide Superconductors

5.1. Fabrications of FeSe Film for Superconducting Radio-Frequency Applications

Bulk Nb superconducting radio-frequency (SRF) cavities have been widely utilized in accelerators. However, the gradient of acceleration and overall performance are greatly limited by the overheating field (Bsh). Lin et al. fabricated the FeSe-coated Nb as an alternative material. The out-of-plane θ–2θ XRD shows only the (00l) peaks of the FeSe film, indicating excellent c-axis orientation and crystallinity. As shown in Figure 16a–d, SEM and AFM images demonstrate a homogeneous thickness distribution of Nb (117 nm) and FeSe (130 nm) layers throughout the film, with root mean square roughness of 0.773 nm for Nb/CaF2 and 1.89 nm for FeSe/Nb/CaF2. This is beneficial to the effectiveness of the Bean–Livingston barrier, which can prevent vortex penetration and increase the Bsh.
As shown in Figure 16k–m, the m(H) curves of the Nb/CaF2, FeSe/CaF2, and FeSe/Nb/CaF2 film deviate at different temperatures and fields. The Bc1(T) is obtained based on the deviation point from the linear background (Figure 16n,o). After the deposition of FeSe on Nb/CaF2, the Bc1(0) is over ten times higher than that of Nb/CaF2 and FeSe/CaF2. In addition, the FeSe layer exhibits excellent bending ability. Figure 16e–h indicate that the FeSe layer maintained a strong adhesion to the surface, even when subjected to varying degrees of inward or outward bending, demonstrating extraordinary resilience. Furthermore, FeSe film can be fabricated on a 10 × 10 mm2 curved Nb cut from a cavity, as shown in Figure 16i. The XRD pattern exhibits the (00l) crystal orientation on the polycrystalline Nb substrate (Figure 16j). It suggests that FeSe satisfies the basic requirements and can be an alternative material for SRF cavities.

5.2. Fabrications of Meters-Long FeTe1−xSex-Coated Conductors and the Applications in Coils

Like other siblings of iron-based superconductors, the 11 system exhibits a weak link effect. The critical current density across the junction of the FeSe0.5Te0.5 films on the [001]-tilt SrTiO3 (STO) bicrystals decreases exponentially when the misorientation angle is larger than a critical value of θc = 9° [77,78,79]. As a result, the epitaxial film is an accessible option to reach a large supercurrent. The initiative efforts were mainly devoted to growing films on single crystalline substrates. Bellingeri et al. tried growing FeTe1−xSex thin film on (001)-orientated MgO, STO, LaAlO3 (LAO), and yttria-stabilized zirconia (ZrO:Y) using the PLD method. They found that the Tc is independent on the film thickness but closely correlated to the compressive strain [80]. The highest Tc = 21 K was obtained in the film deposited on the LAO substrates. The CaF2 substrate was also used, and a 1 MA/cm2 critical current density was achieved at 4.2 K and 0 T [70]. The 5–20 nm lattice disorder contributes to the complete isotropy of the current carrying ability. Nowadays, significant progress has been made in research on Fe(Se,Te)-coated conductors (CC). Si et al. significantly enhanced the superconducting performance of Fe(Se,Te) CC by employing a CeO2 buffer layer [74]. The FeSe0.5Te0.5 on a rolling-assisted biaxially textured substrate (RABiTS) exhibited Tc > 20 K. The self-field Jc reached 1 MA/cm2 at 4.2 K, and Jc remained around 1 × 105 A/cm2 under a high magnetic field of 30 T. Sylva et al. found that the CeO2 layer plays an important role in preventing the element diffusion from the Ni-W substrate and inducing in-plane texture [81,82]. Song et al. adopted an interface engineering strategy to mitigate the thickness effect, successfully fabricating micron-thick FeSe0.5Te0.5 superconducting films on IBAD-LaMnO3-buffered metal tapes [65]. Unlike the continuous growth mode, the alternating growth of a 10 nm thick seed layer and a 400 nm thick FST superconducting layer ensures the crystalline quality of FeSe0.5Te0.5 films (see Figure 17a–c). Apart from the minor pits and droplets, the elemental distribution of the film is uniform, with a smooth surface morphology. Figure 17d depicts the normalized resistance curves of FST films under zero magnetic field, revealing superconductivity across all films. As the film thickness increases, the influence of the LMO buffer layer on Tc diminishes, and the introduction of the seed layer promotes the uniform growth of epitaxy film. Consequently, Tc gradually increases and approaches the bulk value of ~14–15 K. In Figure 17f, the highest Jc appears in the 400 nm thick film, with a self-field Jc of ∼1.3 MA/cm2 and Jc(9 T) of 0.71 and 0.5 MA/cm2 for H//ab and H//c, respectively. Upon further increasing the thickness, there is a rapid drop down of Jc, suggesting the presence of the thickness effect. But, with the help of the seed layer, the Jc revives to a second peak at 1 μm. A possible reason for the thickness effect is the sensitivity of pinning effects to temperature and thickness in type-II superconductors, where δl pinning contributes more to Jc than δTc pinning. When the film thickness is 200 nm, δl pinning dominates. With increasing thickness and temperature, δl pinning gradually transforms to δTc pinning and exhibits a mixed form.
Ye et al. investigated the influence of substrate temperature and thickness on the superconducting properties of FeSe0.5Te0.5 CC. Moderate substrate temperatures are conducive to achieving superconducting films with high performance due to the impact on the evaporation of Se and Te, leading to improved texture and stoichiometry. Based on this, increasing the thickness leads to a transition in the deposition mode from heteroepitaxial to homoepitaxial growth, and the prolonged deposition mitigates longitudinal elemental non-uniformity and the attraction of metal ions within the buffer layer, enabling the attainment of good texture even at low temperatures. However, as the thickness increases, the phenomenon of compositional inhomogeneity in the film becomes more severe. Additionally, the actual temperature at the deposition surface may be lower than the set temperature, leading to grain misorientation or the formation of non-superconducting phases, which is detrimental to the superconducting performance of Fe(Se,Te) CC. As shown in Figure 17g–i, the Tc first increases at 450 nm and then decreases for a little bit at 600 nm. The corresponding Jc peaks at 450 nm, but the Ic continues to grow larger with the increasing thickness. The analysis of the pinning mechanism revealed the dominance of point pinning, as proved by the hpeak~0.33 in Figure 17j and 3–10 nm defects in the TEM image in Figure 17k, which do not change with the temperature.
Liu et al. creatively utilized reel-to-reel PLD for the first time to fabricate 1 m long Fe(Se,Te) coated conductors with high superconducting performance [83], as shown in Figure 18a. Figure 18d demonstrates the smooth surface of the CC, except for some particles most likely formed by the condensation of Se vapor. The AFM image in Figure 18e also proves the smooth plane with a small square root mean square roughness of about 1 nm. Figure 18f,g illustrate the high-quality epitaxial growth of Fe(Se,Te) films on CeO2 buffer layers, where all (00l) planes are clearly visible with distinct interfaces between layers. Figure 18h,i demonstrate that the end-to-end Ic of the samples reached 108 A/cm at 4.2 K and self-field, with good performance observed even under high magnetic fields for short samples. Sharp superconducting transitions occurred in Fe(Se,Te) films under magnetic fields ranging from 0 to 9 T, with Tc reaching up to 17.5 K, and an estimated Hc2(0) = 53 T. The Jc(4.2 K) derived from the Bean model is over 2 MA/cm2 at self-field.
Liu et al. optimized the ability of the Fe(Se,Te)-coated conductors to withstand uniaxial tensile pressure by encapsulating them with a copper layer [84]. As illustrated in Figure 19a–c, regardless of considering the cooling process effects, the mechanical and electrical performance of the tape with copper encapsulation under tensile strain is consistently superior to that of the bare tape. As the strain gradually increased from 0% to 0.5%, the Ic of both the bare tape and the copper-coated tape decreased by approximately 58% and 43%, respectively. With the adoption of copper encapsulation, the irreversible tensile strain limit of the FeSe0.5Te0.5 tape reached 0.15%. Furthermore, considering the effects of the cooling process, the actual irreversible tensile strain limit was enhanced to 0.29%.
Wei et al. successfully fabricated the world’s first hybrid coil by using FeSe0.5Te0.5 CC. The schematic diagram of the FeSe0.5Te0.5 hybrid coil and the inserted magnet is shown in Figure 19d–g. This hybrid coil consists of a single-layer FeSe0.5Te0.5 coil with an inner diameter of 51 mm and eight double-layer YBCO coils with inner diameters of 45 mm. The YBCO coils are divided into two parts and connected to both sides of the FeSe0.5Te0.5 coil. Additionally, a copper layer is used to cover the FeSe0.5Te0.5 coil to withstand the high stress during cooling, thereby preventing substrate shrinkage from damaging the superconducting layer. Figure 19i shows that the maximum stress on the coil is less than 1 MPa up to 10 T, which is within the allowable pressure range of the tape. As shown in Figure 19h, based on the quench process of the FeSe0.5Te0.5 coil in the first cold test, the self-field Ic reached 108.1 A, and the Ic at a 10T field reached 17.4 A. The Jc was approximately 1.42 × 105 A/cm2, which is consistent with the results of the short sample test. Results of repeated cold tests demonstrate the stable performance of the FeSe0.5Te0.5 hybrid coil.

6. Conclusions and Prospects

Despite the electroneutral FeSe layer, iron–chalcogenide superconductors present varieties of crystal structures exhibiting different superconducting characteristics. The interesting properties of this system prompt scientists to invent different methodologies to modulate the crystalline and electrical structures through oxidation, intercalation, the electrochemical method, and interface engineering, which greatly enriche the fabrications and investigations of high-temperature superconductors. According to the summarized phase diagram, higher Tc could be achieved through charge transfer and electron–phonon coupling across heterostructures. More importantly, the FeSe1−xTex and the intercalated FeSe have a large upper critical field, which is the precondition for high-field applications. The successful fabrications of meters-long FeSe1−xTex CC facilitate the first fabrications of insert pancake coils. Compared with the poor texture in the iron-based superconducting wires fabricated through the PIT process, the in-plane texture can be easily obtained in the CC, leading to a higher Jc than the FeSe1−xTex wires at high fields [86,87,88]. Due to the larger critical misorientation angle compared to cuprates (θc = 5°), less textured substrates are required for the FeSe1−xTex CC. However, the vortex pinning is dominated by the scarce point defects, whether in single crystals or CC. Stronger vortex pinning centers, e.g., nanoparticles and correlated column defects introduced in the ReBa2Cu3O7−δ (Re = rare earth elements) CC by mature techniques, are highly desirable for further enhancement of the current carrying ability of the iron chalcogenide superconductors.

Author Contributions

Conceptualization, J.Z. and C.D.; methodology, J.Z.; formal analysis, J.Z.; investigation, J.Z. and C.D.; resources, Y.M.; writing—original draft preparation, J.Z.; writing—review and editing, J.Z., J.L. and C.D.; supervision, D.W. and Y.M.; project administration, D.W. and Y.M.; funding acquisition, D.W. and Y.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant Nos. 52172275, U1832213, 51861135311, 51721005, 52107031), Beijing Municipal Natural Science Foundation (Grant No. 3222061), Natural Science Foundation of Shandong Province (Grant No. ZR2021ME061).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

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] [PubMed]
  2. Mazin, I.I. Superconductivity gets an iron boost. Nature 2010, 464, 183–186. [Google Scholar] [CrossRef] [PubMed]
  3. Putti, M.; Pallecchi, I.; Bellingeri, E.; Cimberle, M.R.; Tropeano, M.; Ferdeghini, C.; Palenzona, A.; Tarantini, C.; Yamamoto, A.; Jiang, J.; et al. New Fe-based superconductors: Properties relevant for applications. Supercond. Sci. Technol. 2010, 23, 034003. [Google Scholar] [CrossRef]
  4. Yao, C.; Ma, Y. Recent breakthrough development in iron-based superconducting wires for practical applications. Supercond. Sci. Technol. 2019, 32, 023002. [Google Scholar] [CrossRef]
  5. Takahashi, H.; Igawa, K.; Arii, K.; Kamihara, Y.; Hirano, M.; Hosono, H. Superconductivity at 43 K in an iron-based layered compound LaO1−xFxFeAs. Nature 2008, 453, 376–378. [Google Scholar] [CrossRef] [PubMed]
  6. 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]
  7. Pelliciari, J.; Karakuzu, S.; Song, Q.; Arpaia, R.; Nag, A.; Rossi, M.; Li, J.; Yu, T.; Chen, X.; Peng, R.; et al. Evolution of spin excitations from bulk to monolayer FeSe. Nat. Commun. 2021, 12, 3122. [Google Scholar] [CrossRef] [PubMed]
  8. Bao, W.; Qiu, Y.; Huang, Q.; Green, M.A.; Zajdel, P.; Fitzsimmons, M.R.; Zhernenkov, M.; Chang, S.; Fang, M.; Qian, B.; et al. Tunable (δπ, δπ)-Type Antiferromagnetic Order in α-Fe(Te,Se) Superconductors. Phys. Rev. Lett. 2009, 102, 247001. [Google Scholar] [CrossRef] [PubMed]
  9. Fang, M.; Wang, H.; Dong, C.; Li, Z.; Feng, C.; Chen, J.; Yuan, H. Fe-based superconductivity with Tc = 31 K bordering an antiferromagnetic insulator in (Tl,K)FexSe2. Europhys. Lett. 2011, 94, 27009. [Google Scholar] [CrossRef]
  10. Bao, W.; Huang, Q.-Z.; Chen, G.-F.; Wang, D.-M.; He, J.-B.; Qiu, Y.-M. A Novel Large Moment Antiferromagnetic Order in K0.8Fe1.6Se2 Superconductor. Chin. Phys. Lett. 2011, 28, 086104. [Google Scholar] [CrossRef]
  11. Burrard-Lucas, M.; Free, D.G.; Sedlmaier, S.J.; Wright, J.D.; Cassidy, S.J.; Hara, Y.; Corkett, A.J.; Lancaster, T.; Baker, P.J.; Blundell, S.J.; et al. Enhancement of the superconducting transition temperature of FeSe by intercalation of a molecular spacer layer. Nat. Mater. 2013, 12, 15–19. [Google Scholar] [CrossRef]
  12. Shi, M.Z.; Wang, N.Z.; Lei, B.; Shang, C.; Meng, F.B.; Ma, L.K.; Zhang, F.X.; Kuang, D.Z.; Chen, X.H. Organic-ion-intercalated FeSe-based superconductors. Phys. Rev. Mater. 2018, 2, 074801. [Google Scholar] [CrossRef]
  13. Lu, X.F.; Wang, N.Z.; Wu, H.; Wu, Y.P.; Zhao, D.; Zeng, X.Z.; Luo, X.G.; Wu, T.; Bao, W.; Zhang, G.H.; et al. Coexistence of superconductivity and antiferromagnetism in (Li0.8Fe0.2)OHFeSe. Nat. Mater. 2014, 14, 325–329. [Google Scholar] [CrossRef] [PubMed]
  14. Fang, M.H.; Pham, H.M.; Qian, B.; Liu, T.J.; Vehstedt, E.K.; Liu, Y.; Spinu, L.; Mao, Z.Q. Superconductivity close to magnetic instability in Fe(Se1−xTex)0.82. Phys. Rev. B 2008, 78, 224503. [Google Scholar] [CrossRef]
  15. Dong, C.; Wang, H.; Yang, J.; Qian, B.; Chen, J.; Li, Z.; Yuan, H.; Fang, M. Effect of annealing on superconductivity in Fe1+y(Te1−xSx) system. Sci. China Phys. Mech. Astron. 2010, 53, 1216–1220. [Google Scholar] [CrossRef]
  16. Sun, Y.; Tsuchiya, Y.; Taen, T.; Yamada, T.; Pyon, S.; Sugimoto, A.; Ekino, T.; Shi, Z.; Tamegai, T. Dynamics and mechanism of oxygen annealing in Fe1+yTe0.6Se0.4 single crystal. Sci. Rep. 2014, 4, 4585. [Google Scholar] [CrossRef] [PubMed]
  17. Liu, T.J.; Ke, X.; Qian, B.; Hu, J.; Fobes, D.; Vehstedt, E.K.; Pham, H.; Yang, J.H.; Fang, M.H.; Spinu, L.; et al. Charge-carrier localization induced by excess Fe in the superconductor Fe1+yTe1−xSex. Phys. Rev. B 2009, 80, 174509. [Google Scholar] [CrossRef]
  18. Sun, Y.; Shi, Z.; Tamegai, T. Review of annealing effects and superconductivity in Fe1+yTe1−xSex superconductors. Supercond. Sci. Technol. 2019, 32, 103001. [Google Scholar] [CrossRef]
  19. Dong, C.; Wang, H.; Li, Z.; Chen, J.; Yuan, H.Q.; Fang, M. Revised phase diagram for the FeTe1−xSex system with fewer excess Fe atoms. Phys. Rev. B 2011, 84, 224506. [Google Scholar] [CrossRef]
  20. Ding, X.; Fang, D.; Wang, Z.; Yang, H.; Liu, J.; Deng, Q.; Ma, G.; Meng, C.; Hu, Y.; Wen, H.-H. Influence of microstructure on superconductivity in KxFe2−ySe2 and evidence for a new parent phase K2Fe7Se8. Nat. Commun. 2013, 4, 1897. [Google Scholar] [CrossRef]
  21. Li, W.; Ding, H.; Deng, P.; Chang, K.; Song, C.; He, K.; Wang, L.; Ma, X.; Hu, J.-P.; Chen, X.; et al. Phase separation and magnetic order in K-doped iron selenide superconductor. Nat. Phys 2012, 8, 126–130. [Google Scholar] [CrossRef]
  22. Ying, T.P.; Chen, X.L.; Wang, G.; Jin, S.F.; Zhou, T.T.; Lai, X.F.; Zhang, H.; Wang, W.Y. Observation of superconductivity at 30∼46K in AxFe2Se2 (A = Li, Na, Ba, Sr, Ca, Yb and Eu). Sci. Rep. 2012, 2, 426. [Google Scholar] [CrossRef]
  23. Ying, T.; Chen, X.; Wang, G.; Jin, S.; Lai, X.; Zhou, T.; Zhang, H.; Shen, S.; Wang, W. Superconducting Phases in Potassium-Intercalated Iron Selenides. J. Am. Chem. Soc. 2013, 135, 2951–2954. [Google Scholar] [CrossRef]
  24. Guo, J.; Lei, H.; Hayashi, F.; Hosono, H. Superconductivity and phase instability of NH3-free Na-intercalated FeSe1−zSz. Nat. Commun. 2014, 5, 4756. [Google Scholar] [CrossRef]
  25. Sun, S.; Wang, S.; Yu, R.; Lei, H. Extreme anisotropy and anomalous transport properties of heavily electron doped Lix(NH3)yFe2Se2 single crystals. Phys. Rev. B 2017, 96, 064512. [Google Scholar] [CrossRef]
  26. Hatakeda, T.; Noji, T.; Sato, K.; Kawamata, T.; Kato, M.; Koike, Y. New Alkali-Metal- and 2-Phenethylamine-Intercalated Superconductors Ax(C8H11N)yFe1−zSe(A = Li, Na) with the Largest Interlayer Spacings and Tc∼40 K. J. Phys. Soc. Jpn. 2016, 85, 103702. [Google Scholar] [CrossRef]
  27. Meng, Y.; Xing, X.; Yi, X.; Li, B.; Zhou, N.; Li, M.; Zhang, Y.; Wei, W.; Feng, J.; Terashima, K.; et al. Protonation-induced discrete superconducting phases in bulk FeSe single crystals. Phys. Rev. B 2022, 105, 134506. [Google Scholar] [CrossRef]
  28. Lei, B.; Cui, J.H.; Xiang, Z.J.; Shang, C.; Wang, N.Z.; Ye, G.J.; Luo, X.G.; Wu, T.; Sun, Z.; Chen, X.H. Evolution of High-Temperature Superconductivity from a Low-Tc Phase Tuned by Carrier Concentration in FeSe Thin Flakes. Phys. Rev. Lett. 2016, 116, 077002. [Google Scholar] [CrossRef]
  29. Ying, T.P.; Wang, M.X.; Wu, X.X.; Zhao, Z.Y.; Zhang, Z.Z.; Song, B.Q.; Li, Y.C.; Lei, B.; Li, Q.; Yu, Y.; et al. Discrete Superconducting Phases in FeSe-Derived Superconductors. Phys. Rev. Lett. 2018, 121, 207003. [Google Scholar] [CrossRef]
  30. Lu, X.F.; Wang, N.Z.; Zhang, G.H.; Luo, X.G.; Ma, Z.M.; Lei, B.; Huang, F.Q.; Chen, X.H. Superconductivity in LiFeO2Fe2Se2 with anti-PbO-type spacer layers. Phys. Rev. B 2014, 89, 020507. [Google Scholar] [CrossRef]
  31. Sun, H.; Woodruff, D.N.; Cassidy, S.J.; Allcroft, G.M.; Sedlmaier, S.J.; Thompson, A.L.; Bingham, P.A.; Forder, S.D.; Cartenet, S.; Mary, N.; et al. Soft Chemical Control of Superconductivity in Lithium Iron Selenide Hydroxides Li1–xFex(OH)Fe1–ySe. Inorg. Chem. 2015, 54, 1958–1964. [Google Scholar] [CrossRef]
  32. Dong, X.; Jin, K.; Yuan, D.; Zhou, H.; Yuan, J.; Huang, Y.; Hua, W.; Sun, J.; Zheng, P.; Hu, W.; et al. (Li0.84Fe0.16)OHFe0.98Se superconductor: Ion-exchange synthesis of large single-crystal and highly two-dimensional electron properties. Phys. Rev. B 2015, 92, 064515. [Google Scholar] [CrossRef]
  33. Wang, Q.-Y.; Li, Z.; Zhang, W.-H.; Zhang, Z.-C.; Zhang, J.-S.; Li, W.; Ding, H.; Ou, Y.-B.; Deng, P.; Chang, K.; et al. Interface-Induced High-Temperature Superconductivity in Single Unit-Cell FeSe Films on SrTiO3. Chin. Phys. Lett. 2012, 29, 037402. [Google Scholar] [CrossRef]
  34. Zhang, W.-H.; Sun, Y.; Zhang, J.-S.; Li, F.-S.; Guo, M.-H.; Zhao, Y.-F.; Zhang, H.-M.; Peng, J.-P.; Xing, Y.; Wang, H.-C.; et al. Direct Observation of High-Temperature Superconductivity in One-Unit-Cell FeSe Films. Chin. Phys. Lett. 2014, 31, 017401. [Google Scholar] [CrossRef]
  35. Zhang, H.; Zhang, D.; Lu, X.; Liu, C.; Zhou, G.; Ma, X.; Wang, L.; Jiang, P.; Xue, Q.-K.; Bao, X. Origin of charge transfer and enhanced electron–phonon coupling in single unit-cell FeSe films on SrTiO3. Nat. Commun. 2017, 8, 214. [Google Scholar] [CrossRef]
  36. Ge, J.-F.; Liu, Z.-L.; Liu, C.; Gao, C.-L.; Qian, D.; Xue, Q.-K.; Liu, Y.; Jia, J.-F. Superconductivity above 100 K in single-layer FeSe films on doped SrTiO3. Nat. Mater. 2015, 14, 285–289. [Google Scholar] [CrossRef]
  37. Hou, Q.; Sun, L.; Sun, Y.; Shi, Z. Review of Single Crystal Synthesis of 11 Iron-Based Superconductors. Materials 2023, 16, 4895. [Google Scholar] [CrossRef]
  38. Yi, X.; Xing, X.; Qin, L.; Feng, J.; Li, M.; Zhang, Y.; Meng, Y.; Zhou, N.; Sun, Y.; Shi, Z. Hydrothermal synthesis and complete phase diagram of FeSe1−xSx (0 ≤ x ≤ 1) single crystals. Phys. Rev. B 2021, 103, 144501. [Google Scholar] [CrossRef]
  39. Terao, K.; Kashiwagi, T.; Shizu, T.; Klemm, R.A.; Kadowaki, K. Superconducting and tetragonal-to-orthorhombic transitions in single crystals of FeSe1−xTex (0 ≤ x ≤ 0.61 ). Phys. Rev. B 2019, 100, 224516. [Google Scholar] [CrossRef]
  40. Sun, Q.; Xu, Z.; Dong, C.; Huang, H.; Wang, D.; Zhang, X.; Ma, Y. Macroscopically ordered phase separation: A new strategy for improving the superconducting performance in Fe(Se, Te). Mater. Today Adv. 2023, 17, 100354. [Google Scholar] [CrossRef]
  41. Zhuang, J.; Yeoh, W.K.; Cui, X.; Xu, X.; Du, Y.; Shi, Z.; Ringer, S.P.; Wang, X.; Dou, S.X. Unabridged phase diagram for single-phased FeSexTe1−x thin films. Sci. Rep. 2014, 4, 7273. [Google Scholar] [CrossRef] [PubMed]
  42. Fan, F.; Xu, Z.; Cheng, Z.; Huang, H.; Zhu, Y.; Liu, S.; Dong, C.; Zhang, X.; Ma, Y. Enhanced superconducting properties of FeSe0.8Te0.2 thin films grown by pulsed laser deposition. Phys. C Supercond. Its Appl. 2019, 564, 55–58. [Google Scholar] [CrossRef]
  43. Zhang, H.-F.; Chen, X.-H.; Xiao, Q.-L.; Chen, F.; Feng, Z.; Cao, S.; Zhang, J.; Qi, Y.; Shi, Z.; Ge, J.-Y. Evolution of Superconducting Properties in Fe1.1Se0.8Te0.2 Films Before and After Structure Avalanche. ACS Appl. Mater. Interfaces 2021, 13, 42138–42145. [Google Scholar] [CrossRef] [PubMed]
  44. Dong, C.; Wang, H.; Mao, Q.; Khan, R.; Zhou, X.; Li, C.; Yang, J.; Chen, B.; Fang, M. Phase diagram and annealing effect for Fe1+δTe1-xSx single crystals. J. Phys. Condens. Matter 2013, 25, 385701. [Google Scholar] [CrossRef] [PubMed]
  45. Zhao, C.; Yi, X.; Hou, Q.; Feng, J.; Zhang, Y.; Xu, M.; Shi, Z. Hydrothermal Synthesis and Transport Properties of FeS1−xTex (0 ≤ x ≤ 0.15) Single Crystals. J. Supercond. Nov. Magn. 2021, 34, 2565–2572. [Google Scholar] [CrossRef]
  46. Lai, X.; Zhang, H.; Wang, Y.; Wang, X.; Zhang, X.; Lin, J.; Huang, F. Observation of Superconductivity in Tetragonal FeS. J. Am. Chem. Soc. 2015, 137, 10148–10151. [Google Scholar] [CrossRef] [PubMed]
  47. Rebec, S.N.; Jia, T.; Zhang, C.; Hashimoto, M.; Lu, D.-H.; Moore, R.G.; Shen, Z.-X. Coexistence of Replica Bands and Superconductivity in FeSe Monolayer Films. Phys. Rev. Lett. 2017, 118, 067002. [Google Scholar] [CrossRef]
  48. Miyata, Y.; Nakayama, K.; Sugawara, K.; Sato, T.; Takahashi, T. High-temperature superconductivity in potassium-coated multilayer FeSe thin films. Nat. Mater. 2015, 14, 775–779. [Google Scholar] [CrossRef]
  49. Farrar, L.S.; Bristow, M.; Haghighirad, A.A.; McCollam, A.; Bending, S.J.; Coldea, A.I. Suppression of superconductivity and enhanced critical field anisotropy in thin flakes of FeSe. npj Quantum Mater. 2020, 5, 29. [Google Scholar] [CrossRef]
  50. Wang, Z.; Yuan, J.; Wosnitza, J.; Zhou, H.; Huang, Y.; Jin, K.; Zhou, F.; Dong, X.; Zhao, Z. The upper critical field and its anisotropy in (Li 1−xFex )OHFe1−ySe. J. Phys. Condens. Matter 2017, 29, 025701. [Google Scholar] [CrossRef]
  51. Pan, Y.; Sun, Y.; Zhou, N.; Yi, X.; Hou, Q.; Wang, J.; Zhu, Z.; Mitamura, H.; Tokunaga, M.; Shi, Z. Novel anisotropy of upper critical fields in Fe1+yTe0.6Se0.4. J. Alloys Compd. 2024, 976, 173262. [Google Scholar] [CrossRef]
  52. Lei, H.; Petrovic, C. Giant increase in critical current density of KxFe2−ySe2 single crystals. Phys. Rev. B 2011, 84, 212502. [Google Scholar] [CrossRef]
  53. Li, M.; Chen, L.; You, W.-L.; Ge, J.; Zhang, J. Giant increase of critical current density and vortex pinning in Mn doped KxFe2−ySe2 single crystals. Appl. Phys. Lett. 2014, 105, 192602. [Google Scholar] [CrossRef]
  54. Lei, H.; Petrovic, C. Critical current density and mechanism of vortex pinning in KxFe2−ySe2 doped with S. Phys. Rev. B 2011, 84, 052507. [Google Scholar] [CrossRef]
  55. 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]
  56. Lanoël, L.; Haberkorn, N.; Nieva, G. Flux creep regimes and vortex phase diagram in β-FeSe single crystals. Phys. C Supercond. Its Appl. 2024, 618, 1354466. [Google Scholar] [CrossRef]
  57. Khasanov, R.; Zhou, H.; Amato, A.; Guguchia, Z.; Morenzoni, E.; Dong, X.; Zhang, G.; Zhao, Z. Proximity-induced superconductivity within the insulating (Li0.84Fe0.16)OH layers in ( Li0.84Fe0.16)OHFe0.98Se. Phys. Rev. B 2016, 93, 224512. [Google Scholar] [CrossRef]
  58. Li, D.; Yuan, J.; Shen, P.; Xi, C.; Tian, J.; Ni, S.; Zhang, J.; Wei, Z.; Hu, W.; Li, Z.; et al. Giant enhancement of critical current density at high field in superconducting (Li,Fe)OHFeSe films by Mn doping. Supercond. Sci. Technol. 2019, 32, 12LT01. [Google Scholar] [CrossRef]
  59. Amigó, M.L.; Crivillero, V.A.; Franco, D.G.; Nieva, G. Multiband character of β-FeSe: Angular dependence of the magnetoresistance and upper critical field. J. Phys. Conf. Ser. 2014, 568, 022005. [Google Scholar] [CrossRef]
  60. 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]
  61. Sun, Y.; Park, A.; Pyon, S.; Tamegai, T.; Kambara, T.; Ichinose, A. Effects of heavy-ion irradiation on FeSe. Phys. Rev. B 2017, 95, 104514. [Google Scholar] [CrossRef]
  62. Tamegai, T.; Taen, T.; Yagyuda, H.; Tsuchiya, Y.; Mohan, S.; Taniguchi, T.; Nakajima, Y.; Okayasu, S.; Sasase, M.; Kitamura, H.; et al. Effects of particle irradiations on vortex states in iron-based superconductors. Supercond. Sci. Technol. 2012, 25, 084008. [Google Scholar] [CrossRef]
  63. Lin, H.; Xing, J.; Zhu, X.; Yang, H.; Wen, H.-H. Robust superconductivity and transport properties in (Li1−xFex)OHFeSe single crystals. Sci. China Phys. Mech. Astron. 2016, 59, 657404. [Google Scholar] [CrossRef]
  64. Xu, Z.; Yuan, P.; Ma, Y.; Cai, C. High-performance FeSe0.5Te0.5 thin films fabricated on less-well-textured flexible coated conductor templates. Supercond. Sci. Technol. 2017, 30, 035003. [Google Scholar] [CrossRef]
  65. Song, J.; Xu, Z.; Xiong, X.; Yuan, W.; Dong, C.; Sun, Q.; Tang, M.; Chen, W.; Tian, H.; Li, J.; et al. Critical Role Played by Interface Engineering in Weakening Thickness Dependence of Superconducting and Structural Properties of FeSe0.5Te0.5-Coated Conductors. ACS Appl. Mater. Interfaces 2023, 15, 26215–26224. [Google Scholar] [CrossRef] [PubMed]
  66. Si, W.; Han, S.J.; Shi, X.; Ehrlich, S.N.; Jaroszynski, J.; Goyal, A.; Li, Q. High current superconductivity in FeSe0.5Te0.5-coated conductors at 30 tesla. Nat. Commun. 2013, 4, 1347. [Google Scholar] [CrossRef] [PubMed]
  67. Yuan, P.; Xu, Z.; Ma, Y.; Sun, Y.; Tamegai, T. Angular-dependent vortex pinning mechanism and magneto-optical characterizations of FeSe0.5Te0.5 thin films grown on CaF2 substrates. Supercond. Sci. Technol. 2016, 29, 035013. [Google Scholar] [CrossRef]
  68. Braccini, V.; Kawale, S.; Reich, E.; Bellingeri, E.; Pellegrino, L.; Sala, A.; Putti, M.; Higashikawa, K.; Kiss, T.; Holzapfel, B.; et al. Highly effective and isotropic pinning in epitaxial Fe(Se,Te) thin films grown on CaF2 substrates. Appl. Phys. Lett. 2013, 103, 172601. [Google Scholar] [CrossRef]
  69. Yuan, P.; Xu, Z.; Zhang, H.; Wang, D.; Ma, Y.; Zhang, M.; Li, J. High performance FeSe0.5Te0.5 thin films grown at low temperature by pulsed laser deposition. Supercond. Sci. Technol. 2015, 28, 065009. [Google Scholar] [CrossRef]
  70. Shahbazi, M.; Wang, X.L.; Dou, S.X.; Fang, H.; Lin, C.T. The flux pinning mechanism, and electrical and magnetic anisotropy in Fe1.04Te0.6Se0.4 superconducting single crystal. J. Appl. Phys. 2013, 113, 17E115. [Google Scholar] [CrossRef]
  71. Sang, L.; Maheswari, P.; Yu, Z.; Yun, F.F.; Zhang, Y.; Dou, S.; Cai, C.; Awana, V.P.S.; Wang, X. Point defect induced giant enhancement of flux pinning in Co-doped FeSe0.5Te0.5 superconducting single crystals. AIP Adv. 2017, 7, 115016. [Google Scholar] [CrossRef]
  72. M Eisterer Radiation effects on iron-based superconductors. Supercond. Sci. Technol. 2018, 31, 013001. [CrossRef]
  73. Ozaki, T.; Wu, L.; Zhang, C.; Si, W.; Jie, Q.; Li, Q. Enhanced critical current in superconducting FeSe0.5Te0.5 films at all magnetic field orientations by scalable gold ion irradiation. Supercond. Sci. Technol. 2018, 31, 024002. [Google Scholar] [CrossRef]
  74. Ozaki, T.; Wu, L.; Zhang, C.; Jaroszynski, J.; Si, W.; Zhou, J.; Zhu, Y.; Li, Q. A route for a strong increase of critical current in nanostrained iron-based superconductors. Nat. Commun. 2016, 7, 13036. [Google Scholar] [CrossRef] [PubMed]
  75. Sun, Y.; Pyon, S.; Tamegai, T.; Kobayashi, R.; Watashige, T.; Shigeru Kasahara, Y.M.T.S. Enhancement of critical current density and mechanism of vortex pinning in H+-irradiated FeSe single crystal. Appl. Phys. Express 2015, 8, 113102. [Google Scholar] [CrossRef]
  76. Lin, Z.; Qin, M.; Li, D.; Shen, P.; Zhang, L.; Feng, Z.; Sha, P.; Miao, J.; Yuan, J.; Dong, X.; et al. Enhancement of the lower critical field in FeSe-coated Nb structures for superconducting radio-frequency applications. Supercond. Sci. Technol. 2021, 34, 015001. [Google Scholar] [CrossRef]
  77. Si, W.; Zhang, C.; Shi, X.; Ozaki, T.; Jaroszynski, J.; Li, Q. Grain boundary junctions of FeSe0.5Te0.5 thin films on SrTiO3 bi-crystal substrates. Appl. Phys. Lett. 2015, 106, 032602. [Google Scholar] [CrossRef]
  78. Sarnelli, E.; Nappi, C.; Camerlingo, C.; Enrico, E.; Bellingeri, E.; Kawale, S.; Braccini, V.; Leveratto, A.; Ferdeghini, C. Properties of Fe(Se, Te) Bicrystal Grain Boundary Junctions, SQUIDs, and Nanostrips. IEEE Trans. Appl. Supercond. 2017, 27, 1–4. [Google Scholar] [CrossRef]
  79. Iida, K.; Hänisch, J.; Yamamoto, A. Grain boundary characteristics of Fe-based superconductors. Supercond. Sci. Technol. 2020, 33, 043001. [Google Scholar] [CrossRef]
  80. Bellingeri, E.; Pallecchi, I.; Buzio, R.; Gerbi, A.; Marrè, D.; Cimberle, M.R.; Tropeano, M.; Putti, M.; Palenzona, A.; Ferdeghini, C. Tc = 21 K in epitaxial FeSe0.5Te0.5 thin films with biaxial compressive strain. Appl. Phys. Lett. 2010, 96, 102512. [Google Scholar] [CrossRef]
  81. Sylva, G.; Bellingeri, E.; Bernini, C.; Celentano, G.; Ferdeghini, C.; Leveratto, A.; Lisitskiy, M.; Malagoli, A.; Manca, N.; Mancini, A.; et al. The role of texturing and thickness of oxide buffer layers in the superconducting properties of Fe(Se,Te) Coated Conductors. Supercond. Sci. Technol. 2020, 33, 114002. [Google Scholar] [CrossRef]
  82. Sylva, G.; Augieri, A.; Mancini, A.; Rufoloni, A.; Vannozzi, A.; Celentano, G.; Bellingeri, E.; Ferdeghini, C.; Putti, M.; Braccini, V. Fe(Se,Te) coated conductors deposited on simple rolling-assisted biaxially textured substrate templates. Supercond. Sci. Technol. 2019, 32, 084006. [Google Scholar] [CrossRef]
  83. Liu, L.; Ye, J.; Mou, S.; Zhu, R.; Miao, C.; Li, Y.; Shi, Z.; Qi, Y.; Ge, J.; Liu, H.; et al. Fabrication of Meter-Long Class Fe(Se,Te)-Coated Conductors with High Superconducting Performance. Adv. Eng. Mater. 2023, 25, 2201536. [Google Scholar] [CrossRef]
  84. Liu, X.; Wei, S.; Shi, Y.; Liu, F.; Zhou, C.; Li, Q.; Li, Y.; Liu, L.; Shi, Z.; Ren, L.; et al. Reversible critical current performance of FeSe0.5Te0.5 coated conductor tapes under uniaxial tensile strain. Supercond. Sci. Technol. 2022, 35, 10LT01. [Google Scholar] [CrossRef]
  85. Wei, S.; Liu, X.; Zhang, Z.; Liu, F.; Qin, J.; Shi, Y.; Li, Y.; Liu, L.; Shi, Z.; Ren, L.; et al. First performance test of FeSe0.5Te0.5-coated conductor coil under high magnetic fields. Supercond. Sci. Technol. 2023, 36, 04LT01. [Google Scholar] [CrossRef]
  86. Zhang, S.; Feng, J.; Ma, X.; Liu, J.; Li, C.; Zhang, P. In-situ and ex-situ PIT fabrication of FeSe superconducting tapes. J. Mater. Sci. Mater. Electron. 2017, 28, 8366–8371. [Google Scholar] [CrossRef]
  87. Li, X.; Zhang, Y.; Yuan, F.; Zhuang, J.; Cao, Z.; Xing, X.; Zhou, W.; Shi, Z. Fabrication of Nb-sheathed FeSe0.5Te0.5 tape by an ex-situ powder-in-tube method. J. Alloys Compd. 2016, 664, 218–222. [Google Scholar] [CrossRef]
  88. Palombo, M.; Malagoli, A.; Pani, M.; Bernini, C.; Manfrinetti, P.; Palenzona, A.; Putti, M. Exploring the feasibility of Fe(Se,Te) conductors by ex-situ powder-in-tube method. J. Appl. Phys. 2015, 117, 213903. [Google Scholar] [CrossRef]
Figure 1. Crystal structures of iron-chalcogenide superconductors [2]. (a) FeSe-11 [6,7,8] and (b) intercalated FeSe superconductors [9,10,11,12,13].
Figure 1. Crystal structures of iron-chalcogenide superconductors [2]. (a) FeSe-11 [6,7,8] and (b) intercalated FeSe superconductors [9,10,11,12,13].
Materials 17 03059 g001
Figure 2. (a) R/R300K for the FeTe0.6Se0.4 crystals before and after annealing under different atmospheres [19]. The left inset shows the data near Tc, and the right inset plots the susceptibility χ(T). STM images for (b) as-grown and (c,d) O2-annealed Fe1+yTe0.6Se0.4 single crystal [16]. (e) Normalized resistance vs. temperature from 2 to 300 K for Fe1−xMnxTe0.55Se0.45 (x = 0, 0.005, 0.01, and 0.02). The inset shows the resistance near the transition [15]. (f) The temperature-dependent susceptibility under a 10 Oe field for the samples with different Mn contents. (g) Temperature dependence of specific heat plotted as C/T vs. T for the Mn-doped samples. (h) Doping dependence of Tczero of Fe1−xMnxTe0.55Se0.45 (x = 0–0.2) single crystals. (i,j) STM topographic images of the cleaved single crystals with x = 0 (left) and 0.01 (right) at 250 mK.
Figure 2. (a) R/R300K for the FeTe0.6Se0.4 crystals before and after annealing under different atmospheres [19]. The left inset shows the data near Tc, and the right inset plots the susceptibility χ(T). STM images for (b) as-grown and (c,d) O2-annealed Fe1+yTe0.6Se0.4 single crystal [16]. (e) Normalized resistance vs. temperature from 2 to 300 K for Fe1−xMnxTe0.55Se0.45 (x = 0, 0.005, 0.01, and 0.02). The inset shows the resistance near the transition [15]. (f) The temperature-dependent susceptibility under a 10 Oe field for the samples with different Mn contents. (g) Temperature dependence of specific heat plotted as C/T vs. T for the Mn-doped samples. (h) Doping dependence of Tczero of Fe1−xMnxTe0.55Se0.45 (x = 0–0.2) single crystals. (i,j) STM topographic images of the cleaved single crystals with x = 0 (left) and 0.01 (right) at 250 mK.
Materials 17 03059 g002
Figure 3. (a) The ab-plane resistivity ρ(T) as a function of temperature for (Tl,K)FexSe2 (x = 1.50, 1.64, 1.68, 1.69, 1.76, and 1.78) single crystals [9]. (b,c) Back-scattered electron images on the cleaved surface of the KxFe2-ySe2 single crystals [20]. (d) STM topographic image of the KxFe2-ySe2 film. Two distinct regions are marked by I and II [21]. (e) Atomic-resolution STM topography of region I. (f) Differential conductance spectrum along the white line in (e). (g) Atomic structure of the (110) plane. K and Se atoms are in the topmost layer. Fe atoms are in the second layer. (h) Differential conductance spectrum in region II. (i) Atomic-resolution STM topography of region II. (j) The structure of the 5 × 5 Fe vacancy pattern, as seen from (001) plane. (k) The positions of Fe vacancies are marked by crosses on the (110) plane.
Figure 3. (a) The ab-plane resistivity ρ(T) as a function of temperature for (Tl,K)FexSe2 (x = 1.50, 1.64, 1.68, 1.69, 1.76, and 1.78) single crystals [9]. (b,c) Back-scattered electron images on the cleaved surface of the KxFe2-ySe2 single crystals [20]. (d) STM topographic image of the KxFe2-ySe2 film. Two distinct regions are marked by I and II [21]. (e) Atomic-resolution STM topography of region I. (f) Differential conductance spectrum along the white line in (e). (g) Atomic structure of the (110) plane. K and Se atoms are in the topmost layer. Fe atoms are in the second layer. (h) Differential conductance spectrum in region II. (i) Atomic-resolution STM topography of region II. (j) The structure of the 5 × 5 Fe vacancy pattern, as seen from (001) plane. (k) The positions of Fe vacancies are marked by crosses on the (110) plane.
Materials 17 03059 g003
Figure 4. NaFe2Se2 system: (a) Powder XRD pattern and Rietveld refinement profile for nominal NaFe2Se2 at 297 K [23]. (b) Temperature dependence of susceptibility under ZFC and field-cooling (FC) procedures. The XRD patterns and Rietveld refinements for (c) the NH3-free Na0.65(1)Fe1.93(1)Se2 phase and (d) the NH3-poor Na0.80(4)(NH3)0.6Fe1.86(1)Se2 phase. (e) The magnetization curves of the samples with different NH3 contents. (f) Tc as a function of the S content in the FeSe1−zSz, Na0.65(1)Fe1.93(1)(Se1−zSz)2, and Na0.80(4)(NH4)0.6Fe1.86(1)(Se1−zSz)2 systems [24].
Figure 4. NaFe2Se2 system: (a) Powder XRD pattern and Rietveld refinement profile for nominal NaFe2Se2 at 297 K [23]. (b) Temperature dependence of susceptibility under ZFC and field-cooling (FC) procedures. The XRD patterns and Rietveld refinements for (c) the NH3-free Na0.65(1)Fe1.93(1)Se2 phase and (d) the NH3-poor Na0.80(4)(NH3)0.6Fe1.86(1)Se2 phase. (e) The magnetization curves of the samples with different NH3 contents. (f) Tc as a function of the S content in the FeSe1−zSz, Na0.65(1)Fe1.93(1)(Se1−zSz)2, and Na0.80(4)(NH4)0.6Fe1.86(1)(Se1−zSz)2 systems [24].
Materials 17 03059 g004
Figure 5. KxFe2Se2(NH3)y system. (a) M–T curves under a zero-field cooling (ZFC) procedure [24]. (b) Tc and (c) lattice constant c as a function of the nominal potassium content. Lix(NH3)yFe2Se2 single crystals. (d) XRD pattern; the insets show the single crystals and the crystal structure. (e) Temperature dependence of in-plane resistivity ρab(T) and c-axial resistivity ρc(T). (f) Temperature dependence of susceptibility under a 1 mT field for H//c and H//ab [25].
Figure 5. KxFe2Se2(NH3)y system. (a) M–T curves under a zero-field cooling (ZFC) procedure [24]. (b) Tc and (c) lattice constant c as a function of the nominal potassium content. Lix(NH3)yFe2Se2 single crystals. (d) XRD pattern; the insets show the single crystals and the crystal structure. (e) Temperature dependence of in-plane resistivity ρab(T) and c-axial resistivity ρc(T). (f) Temperature dependence of susceptibility under a 1 mT field for H//c and H//ab [25].
Materials 17 03059 g005
Figure 6. (a) Temperature dependence of susceptibility of (CTA)0.3FeSe under an external field of 5 and 40 Oe. (b) The normalized susceptibility of (CTA)0.3FeSe at different external pressures. (c,d) The high-resolution TEM (HRTEM) images and (e) the STEM-HAADF (high-angle annular dark field) image of the pristine FeSe crystals. (f,g) The HRTEM images and (h) the STEM-HAADF image of the (CTA)0.3FeSe crystals [12]. The brighter layer is the FeSe layer and the darker layer is the CTA+ organic layer. (i) Schematic of the experimental setup used for the protonation of FeSe. (jl) Temperature dependence of the normalized resistivity, ρ(T)/ρ200K, for Hx-FeSe single crystals during the protonation process [27].
Figure 6. (a) Temperature dependence of susceptibility of (CTA)0.3FeSe under an external field of 5 and 40 Oe. (b) The normalized susceptibility of (CTA)0.3FeSe at different external pressures. (c,d) The high-resolution TEM (HRTEM) images and (e) the STEM-HAADF (high-angle annular dark field) image of the pristine FeSe crystals. (f,g) The HRTEM images and (h) the STEM-HAADF image of the (CTA)0.3FeSe crystals [12]. The brighter layer is the FeSe layer and the darker layer is the CTA+ organic layer. (i) Schematic of the experimental setup used for the protonation of FeSe. (jl) Temperature dependence of the normalized resistivity, ρ(T)/ρ200K, for Hx-FeSe single crystals during the protonation process [27].
Materials 17 03059 g006
Figure 7. (a) A schematic illustration of the FeSe thin flake EDLT device in Ref. [28]. The ionic liquid DEME-TFSI serves as the dielectric, covering the sample and gate electrodes. (b) The optical image of the EDLT device. (c) Schematic structure and (d) optical image of the solid ionic gating device in Ref. [29]. From the bottom to the top: silver back gate layer, Li/Na-ion substrate, FeSe thin flake, 100 nm SiO2, and four electrodes. The gate voltage VG is applied at low temperatures (<155 K), where all of the ions in the substrate are frozen in place. (e) Temperature dependence of the resistance for a FeSe thin flake at different Vg. (f) Resistance of the Li-intercalated FeSe flake (17.5 nm) as a function of temperature. (g) Resistance of another 10 nm FeSe flake in the overdoped regime [29].
Figure 7. (a) A schematic illustration of the FeSe thin flake EDLT device in Ref. [28]. The ionic liquid DEME-TFSI serves as the dielectric, covering the sample and gate electrodes. (b) The optical image of the EDLT device. (c) Schematic structure and (d) optical image of the solid ionic gating device in Ref. [29]. From the bottom to the top: silver back gate layer, Li/Na-ion substrate, FeSe thin flake, 100 nm SiO2, and four electrodes. The gate voltage VG is applied at low temperatures (<155 K), where all of the ions in the substrate are frozen in place. (e) Temperature dependence of the resistance for a FeSe thin flake at different Vg. (f) Resistance of the Li-intercalated FeSe flake (17.5 nm) as a function of temperature. (g) Resistance of another 10 nm FeSe flake in the overdoped regime [29].
Materials 17 03059 g007
Figure 8. (a) XRD and Rietveld refinements of LiFeO2Fe2Se2 [30]. (b) Temperature dependence of susceptibility for the as-grown (Li0.8Fe0.2)OHFeSe [13]. (c) Hydrothermal ion-exchange process [32]. XRD of the (d) K2Fe4Se5 and (e) (Li0.84Fe0.16)OHFe0.98Se single crystals. (f) Susceptibility and (g) magnetoresistivity of the (Li0.84Fe0.16)OHFe0.98Se single crystals.
Figure 8. (a) XRD and Rietveld refinements of LiFeO2Fe2Se2 [30]. (b) Temperature dependence of susceptibility for the as-grown (Li0.8Fe0.2)OHFeSe [13]. (c) Hydrothermal ion-exchange process [32]. XRD of the (d) K2Fe4Se5 and (e) (Li0.84Fe0.16)OHFe0.98Se single crystals. (f) Susceptibility and (g) magnetoresistivity of the (Li0.84Fe0.16)OHFe0.98Se single crystals.
Materials 17 03059 g008
Figure 9. (a) Top left: RHEED pattern after the growth of single-layer FeSe. Bottom left: atomic-scale STM image of the film. Right: large-scale STM image. (b) R–T curves derived from the I–V curve measured at fixed temperatures [36]. (c) Similar R–T curve but measured by sweeping the temperature. (d) Energy bands of Nb-doped STO and 20 uc-FeSe separately. (e,f) Energy band profile across the FeSe/STO heterostructure at the non-superconducting and superconducting stages, respectively [35].
Figure 9. (a) Top left: RHEED pattern after the growth of single-layer FeSe. Bottom left: atomic-scale STM image of the film. Right: large-scale STM image. (b) R–T curves derived from the I–V curve measured at fixed temperatures [36]. (c) Similar R–T curve but measured by sweeping the temperature. (d) Energy bands of Nb-doped STO and 20 uc-FeSe separately. (e,f) Energy band profile across the FeSe/STO heterostructure at the non-superconducting and superconducting stages, respectively [35].
Materials 17 03059 g009
Figure 10. Phase diagrams of iron–chalcogenide superconductors. (a) The entire phase diagram of FeSe1−xTex and FeSe1−xSx single crystals synthesized through the optimal methods, hydrothermal for FeSe1−xSx (0.29 ≤ x ≤ 1), CVT for FeSe1−xSx (0 ≤ x ≤ 0.29) and FeSe1−xTex (0.29 ≤ x ≤ 1), and self-flux plus annealing for FeSe1−xTex (0.55 ≤ x ≤ 1) [37]. The data from the FeSe1−xTex film are also included, as marked by the green pentagon [41]. (b) The doping phase diagram of FeS1−xTex single crystals [37]. (c) Phase diagram of FeSe-derived superconductors [29]. The red lines denote the discrete SC phases of metal-intercalated FeSe superconductors, as represented by LixFeSe. The different Fermi surface topologies of the pristine and doped FeSe are superimposed. The yellow area on the right represents the insulating phase in the heavily overdoped region. (d) Tc vs. interlayer distance d in the intercalated FeSe superconductors [27]. (e) Tc as a function of the maximum superconducting gap of various FeSe-based superconductors [47].
Figure 10. Phase diagrams of iron–chalcogenide superconductors. (a) The entire phase diagram of FeSe1−xTex and FeSe1−xSx single crystals synthesized through the optimal methods, hydrothermal for FeSe1−xSx (0.29 ≤ x ≤ 1), CVT for FeSe1−xSx (0 ≤ x ≤ 0.29) and FeSe1−xTex (0.29 ≤ x ≤ 1), and self-flux plus annealing for FeSe1−xTex (0.55 ≤ x ≤ 1) [37]. The data from the FeSe1−xTex film are also included, as marked by the green pentagon [41]. (b) The doping phase diagram of FeS1−xTex single crystals [37]. (c) Phase diagram of FeSe-derived superconductors [29]. The red lines denote the discrete SC phases of metal-intercalated FeSe superconductors, as represented by LixFeSe. The different Fermi surface topologies of the pristine and doped FeSe are superimposed. The yellow area on the right represents the insulating phase in the heavily overdoped region. (d) Tc vs. interlayer distance d in the intercalated FeSe superconductors [27]. (e) Tc as a function of the maximum superconducting gap of various FeSe-based superconductors [47].
Materials 17 03059 g010
Figure 11. (a) Upper critical field, Bc2, of iron–chalcogenide superconductors [23,49,50,51]. The empty and solid symbols denote the field parallel to the ab-plane and the c-axis, respectively. (b) Anisotropy parameter, γ = Bc2ab/Bc2c, as a function of normalized temperature T/Tc.
Figure 11. (a) Upper critical field, Bc2, of iron–chalcogenide superconductors [23,49,50,51]. The empty and solid symbols denote the field parallel to the ab-plane and the c-axis, respectively. (b) Anisotropy parameter, γ = Bc2ab/Bc2c, as a function of normalized temperature T/Tc.
Materials 17 03059 g011
Figure 12. Field dependence of Jc for iron–chalcogenide superconductors [58,61,62,63,64,69]. The inset shows Jc of the single-layer FeSe [34].
Figure 12. Field dependence of Jc for iron–chalcogenide superconductors [58,61,62,63,64,69]. The inset shows Jc of the single-layer FeSe [34].
Materials 17 03059 g012
Figure 13. (ad) Normalized flux pinning force f = Fp/Fp,max as a function of reduced field h = H/Hmax for (a) the Fe1.04Te0.6Se0.4 single crystal [70], (b) the 3at. % Co-doped FeSe0.5Te0.5 single crystal [71], (c) the Mn-doped and Mn-free (Li,Fe)OHFeSe films [58], and (d) the FeSe0.5Te0.5 film. (e) Angle dependence of Jc and (f) Jc vs. effective field, Heff of the FeSe0.5Te0.5 film [67].
Figure 13. (ad) Normalized flux pinning force f = Fp/Fp,max as a function of reduced field h = H/Hmax for (a) the Fe1.04Te0.6Se0.4 single crystal [70], (b) the 3at. % Co-doped FeSe0.5Te0.5 single crystal [71], (c) the Mn-doped and Mn-free (Li,Fe)OHFeSe films [58], and (d) the FeSe0.5Te0.5 film. (e) Angle dependence of Jc and (f) Jc vs. effective field, Heff of the FeSe0.5Te0.5 film [67].
Materials 17 03059 g013
Figure 14. (a) Cross-sectional TEM micrograph of FeSe irradiated by uranium with Bφ = 2 T. The inset shows an enlarged view of the columnar defect [61]. (b,c) TEM images of the FeTe0.61Se0.39 single crystals irradiated by 800 MeV Xe at Bφ = 8 T. The inset in (c) is the Fourier transformation [62]. (d) Normalized Tc (Tc/Tc0, where Tc0 is for the pristine one) and the self-field Jc at 2 K as a function of the matching field Bφ (bottom axis) and damaged area (top axis) for the uranium-irradiated FeSe.
Figure 14. (a) Cross-sectional TEM micrograph of FeSe irradiated by uranium with Bφ = 2 T. The inset shows an enlarged view of the columnar defect [61]. (b,c) TEM images of the FeTe0.61Se0.39 single crystals irradiated by 800 MeV Xe at Bφ = 8 T. The inset in (c) is the Fourier transformation [62]. (d) Normalized Tc (Tc/Tc0, where Tc0 is for the pristine one) and the self-field Jc at 2 K as a function of the matching field Bφ (bottom axis) and damaged area (top axis) for the uranium-irradiated FeSe.
Materials 17 03059 g014
Figure 15. (a,b) show the normalized flux pinning force f = Fp/Fp,max as a function of reduced field h = H/Hmax of the FeSe single crystals before and after irradiation [75]. (c,d) High-resolution TEM image of Au-ion-irradiated FeSe0.5Te0.5 film [73]. (e,f) Temperature dependence of normalized resistivity of the FeSe0.5Te0.5 (FST) film before and after irradiation by the 190 keV protons [74]. (g) HRTEM image of FST film irradiated with protons. (h) In-plane strain map (εzz). The color bar in the middle indicates the strain from -0.1 (compressive) to 0.1 (tensile).
Figure 15. (a,b) show the normalized flux pinning force f = Fp/Fp,max as a function of reduced field h = H/Hmax of the FeSe single crystals before and after irradiation [75]. (c,d) High-resolution TEM image of Au-ion-irradiated FeSe0.5Te0.5 film [73]. (e,f) Temperature dependence of normalized resistivity of the FeSe0.5Te0.5 (FST) film before and after irradiation by the 190 keV protons [74]. (g) HRTEM image of FST film irradiated with protons. (h) In-plane strain map (εzz). The color bar in the middle indicates the strain from -0.1 (compressive) to 0.1 (tensile).
Materials 17 03059 g015
Figure 16. SEM cross-sectional images of (a) Nb film grown on CaF2 and (b) FeSe film on Nb-coated CaF2 films [76]. (c,d) Atomic force microscopy (AFM) images (5 × 5 µm2) of Nb/CaF2 and FeSe/Nb/CaF2 films. (eh) The FeSe-coated Nb foil bent to inward angles of 160°, 120°, and 60° and an outward angle of 90°. (i,j) Photo and θ–2θ scans of FeSe film deposited on a curved bulk Nb. (km) m(H) curves of Nb/CaF2, FeSe/CaF2, and FeSe/Nb/CaF2 at various temperatures for H parallel to the film surface. (n,o) T2 dependence of Bc1 for Nb/CaF2, FeSe/CaF2, and FeSe/Nb/CaF2.
Figure 16. SEM cross-sectional images of (a) Nb film grown on CaF2 and (b) FeSe film on Nb-coated CaF2 films [76]. (c,d) Atomic force microscopy (AFM) images (5 × 5 µm2) of Nb/CaF2 and FeSe/Nb/CaF2 films. (eh) The FeSe-coated Nb foil bent to inward angles of 160°, 120°, and 60° and an outward angle of 90°. (i,j) Photo and θ–2θ scans of FeSe film deposited on a curved bulk Nb. (km) m(H) curves of Nb/CaF2, FeSe/CaF2, and FeSe/Nb/CaF2 at various temperatures for H parallel to the film surface. (n,o) T2 dependence of Bc1 for Nb/CaF2, FeSe/CaF2, and FeSe/Nb/CaF2.
Materials 17 03059 g016
Figure 17. Fabrications of thick Fe(Se,Te) (FST) film. (a) The structure of a 2 μm thick FST film [65]. (b) Cross-sectional bright-field TEM image of 1 μm thick FST films. The arrow indicates the seed layer grown at high temperatures. (c) An enlarged image of the seed layer region in panel (b). (d) Normalized resistivity near Tc for the thick film with and without seed layer. (e) Scaling behavior of Jc(θ) at 4.2 K for the films with different thicknesses. (f) Jc(4.2 K) vs. film thickness under different magnetic fields with H//ab and H //c. (g) ρ-T curve, (h) self-field critical current, and (i) critical current density at 4.2 K of FST films with different thicknesses [74]. (j) Normalized pinning force densities of 450 nm thick FST films at different temperatures with a field parallel to the c-axis. (k) Cross-sectional TEM image of the 450 nm thick FST film.
Figure 17. Fabrications of thick Fe(Se,Te) (FST) film. (a) The structure of a 2 μm thick FST film [65]. (b) Cross-sectional bright-field TEM image of 1 μm thick FST films. The arrow indicates the seed layer grown at high temperatures. (c) An enlarged image of the seed layer region in panel (b). (d) Normalized resistivity near Tc for the thick film with and without seed layer. (e) Scaling behavior of Jc(θ) at 4.2 K for the films with different thicknesses. (f) Jc(4.2 K) vs. film thickness under different magnetic fields with H//ab and H //c. (g) ρ-T curve, (h) self-field critical current, and (i) critical current density at 4.2 K of FST films with different thicknesses [74]. (j) Normalized pinning force densities of 450 nm thick FST films at different temperatures with a field parallel to the c-axis. (k) Cross-sectional TEM image of the 450 nm thick FST film.
Materials 17 03059 g017
Figure 18. (a) Sketch of reel-to-reel PLD [83]. (b) Photo image of 1 m long Fe(Se,Te) CC. (c) Structure diagram, (d) SEM image, and (e) AFM image of the Fe(Se,Te) CC. (f) Low- and (g) high-magnification cross-sectional TEM images of the Fe(Se,Te) CC. (h) End-to-end Ic at 4.2 K and self-field of 1 m long Fe(Se,Te) CC. (i) Ic and Jc as a function of magnetic fields up to 10 T with H//c at 4.2 K for short sample.
Figure 18. (a) Sketch of reel-to-reel PLD [83]. (b) Photo image of 1 m long Fe(Se,Te) CC. (c) Structure diagram, (d) SEM image, and (e) AFM image of the Fe(Se,Te) CC. (f) Low- and (g) high-magnification cross-sectional TEM images of the Fe(Se,Te) CC. (h) End-to-end Ic at 4.2 K and self-field of 1 m long Fe(Se,Te) CC. (i) Ic and Jc as a function of magnetic fields up to 10 T with H//c at 4.2 K for short sample.
Materials 17 03059 g018
Figure 19. (a,b) The finite element method (FEM) calculations of the strain of the bare FeSe0.5Te0.5 tape and the copper-encapsulated FeSe0.5Te0.5 tape during the cooling process from 343 K to 4.2 K [84]. (c) The normalized critical current Ic/Ic0 of the bare and the copper-encapsulated FeSe0.5Te0.5 tapes after considering the influence of the cooling process. (d) The outer view of FeSe0.5Te0.5-encapsulated tape. (e) The FeSe0.5Te0.5 single pancake coils (SPC). (f) The insert hybrid magnet. (g) Part of the testing device [85]. (h) Magnetic field dependence of transport Ic for FeSe0.5Te0.5 insert coil in the first and second tests. (i) Maximum stress on the FeSe0.5Te0.5 coil and the YBCO coils at the Ic of FeSe0.5Te0.5 tape.
Figure 19. (a,b) The finite element method (FEM) calculations of the strain of the bare FeSe0.5Te0.5 tape and the copper-encapsulated FeSe0.5Te0.5 tape during the cooling process from 343 K to 4.2 K [84]. (c) The normalized critical current Ic/Ic0 of the bare and the copper-encapsulated FeSe0.5Te0.5 tapes after considering the influence of the cooling process. (d) The outer view of FeSe0.5Te0.5-encapsulated tape. (e) The FeSe0.5Te0.5 single pancake coils (SPC). (f) The insert hybrid magnet. (g) Part of the testing device [85]. (h) Magnetic field dependence of transport Ic for FeSe0.5Te0.5 insert coil in the first and second tests. (i) Maximum stress on the FeSe0.5Te0.5 coil and the YBCO coils at the Ic of FeSe0.5Te0.5 tape.
Materials 17 03059 g019
Table 1. Penetration depth, coherence length, and critical current density Jcab(0 T) of practical iron–chalcogenide superconductors. The anisotropy parameter is derived from Figure 11b. The pristine and irradiated samples are denoted as pri and irr, respectively. The flux pinning efficiency is calculated by η = Jc/Jd at 0 T.
Table 1. Penetration depth, coherence length, and critical current density Jcab(0 T) of practical iron–chalcogenide superconductors. The anisotropy parameter is derived from Figure 11b. The pristine and irradiated samples are denoted as pri and irr, respectively. The flux pinning efficiency is calculated by η = Jc/Jd at 0 T.
MaterialsFeSeFeTe1−xSex(Li,Fe)OHFeSe
λab(0 K), nm445 [55,56]430281 [57,58]
ξab(0 K), nm4.4 [56,59]1.5 [60]2.21 [50,58]
Jd(0 K), MA/cm211 [56]3657.7 [58]
γ = Bc2ab/Bc2c2–31–32–5
Jc-single crystal, MA/cm20.043
(2 K, pri)
0.2 (2 K, irr) [61]0.27 (2 K, pri)0.52 (2 K, irr) [62]0.24 [63]
Jc-film, MA/cm21.7 (2 K, Mono-layer) [34]1.36 (4.2 K)\
Jc coated conductor @ 4.2 K, MA/cm2\0.43 [64]1.3 [65]\
Maximum η15%3.7%0.4%
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

Zhao, J.; Liao, J.; Dong, C.; Wang, D.; Ma, Y. Properties and Applications of Iron–Chalcogenide Superconductors. Materials 2024, 17, 3059. https://doi.org/10.3390/ma17133059

AMA Style

Zhao J, Liao J, Dong C, Wang D, Ma Y. Properties and Applications of Iron–Chalcogenide Superconductors. Materials. 2024; 17(13):3059. https://doi.org/10.3390/ma17133059

Chicago/Turabian Style

Zhao, Jianlong, Junsong Liao, Chiheng Dong, Dongliang Wang, and Yanwei Ma. 2024. "Properties and Applications of Iron–Chalcogenide Superconductors" Materials 17, no. 13: 3059. https://doi.org/10.3390/ma17133059

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

Article Metrics

Back to TopTop