Single crystals of Ca1−xLaxFe2As2 with x ranging from 0 to 0.25, have been grown and characterized by structural, transport, and magnetic measurements. Coexistence of two superconducting phases is observed, in which the phase with the lower superconducting transition temperature (Tc) has Tc ∼ 20 K and the higher Tc phase has Tc higher than 40 K. These data also delineate an x-T phase diagram in which the single magnetic/structural phase transition in undoped CaFe2As2 appears to split into two distinct phase transitions, both of which are suppressed with increasing La substitution. Superconductivity emerges when x is about 0.06 and coexists with the structural/magnetic transition until x is ∼ 0.13. With increasing concentration of La, the structural/magnetic transition is totally suppressed, and Tc reaches its maximum value of about 45 K for 0.15 ⩽ x ⩽ 0.19. A domelike superconducting region is not observed in the phase diagram, however, because no obvious over-doping region can be found. Two superconducting phases coexist in the x-T phase diagram of Ca1−xLaxFe2As2. The formation of the two separate phases and the origin of the high Tc in Ca1−xLaxFe2As2 have been studied and discussed in detail.
I. INTRODUCTION
The recent discovery of superconductivity at 26 K in the iron oxypnictide LaFeAs(O,F)1 has stimulated great interest in the condensed-matter physics community. A tremendous amount of work has been carried out, leading to the emergence of novel iron-based superconductor families with different crystal structures: 1111 (ReFeAsO, where Re = rare earth),1 122 (AeFe2As2, Ae = alkaline earth),2 111 (AFeAs, A = alkali metal)3 and 11 (Fe(Se,Te)).4 These compounds adopt a layered structure, based on FeAs(Se) layers, from which the superconducting carriers mainly flow. As in the cuprate superconductors,5 superconductivity in iron oxypnictide appears to be accompanied by the suppression of an antiferromagnetic (AFM) state upon doping or applying pressure, leading to very similar phase diagram. In cuprate and oxypnictide superconductors, electron and hole doping can both successfully induce superconductivity, although the symmetry that appears between electron- and hole-doping in cuprate superconductors6 is still not clear in oxypnictide superconductors. In the parent LaFeAsO, superconductivity can be induced by both F− and Sr2+ doping in O and La-sites, the results of which have proved to be electron- and hole-type carriers, respectively. The highest superconducting transition temperature, Tc, in the two compounds is about 25 K, and the phase diagrams are similar.7 On the other hand, in the 122 system, hole-doped Ba1−xKxFe2As22 was reported to attain superconductivity at 38 K, while electron-doping can only induce superconductivity at a lower temperature, around 22 K.8 One possible reason is that the electron-doping in the 122 system is usually realized by substituting other transition metals for the Fe ions in the FeAs layer, which is believed to be the main carrier-conducting layer. Thus, an electron-doped 122 superconductor with perfect FeAs layers will be an ideal candidate to study the symmetry between electron-doping and hole-doping. Recently, superconductivity above 40 K was discovered in rare-earth doped CaFe2As2,9–14 which has already proved to be dominated by electron-like charge carriers. Although these previous reports have revealed some of the structural, magnetic, and transport properties of this compound, the evolution of the structural/magnetic transition and the superconductivity with rare earth doping is still unknown. Furthermore, in order to understand the conditions for superconductivity and probe the symmetry between electron- and hole-doping in the 122 system, temperature versus doping phase diagrams must first be constructed, which are unfortunately still not clear in Ca1−xRExFe2As2, where RE is a rare earth element. In this article, we report a systematic investigation of the phase diagram of Ca1−xLaxFe2As2 single crystals for x ranging from 0 to 0.25 by using structural, transport and magnetic measurements. The origin and characteristics of the high transition temperature of Ca1−xRExFe2As2 have also been studied and are discussed in detail.
II. EXPERIMENTAL METHODS
Single crystals of Ca1−xLaxFe2As2, with x ranging from 0 to 0.5, were grown using the FeAs self-flux method. The FeAs precursor was first synthesized by reacting stoichiometric amounts of Fe and As inside a vacuum quartz tube at 750 °C for 24 h. High purity Ca grains, La bulks, and FeAs powders, mixed together in the ratio 1−x: x: 4, were put into alumina crucibles and sealed in a quartz tube under a 30% partial pressure atmosphere of Ar gas. The sealed quartz tube was quickly heated to 1180 °C, kept at this temperature for 2 h, and then slowly cooled down to 970 °C at a rate of 2 °C/hour. After that, the temperature was cooled down to room temperature by shutting down the furnace. Single crystals with a typical size of 5 × 5 × 0.2 mm3 were easily obtained by mechanically cleaving them from the flux. Single crystals were characterized by X-ray diffraction (XRD), with Cu Kα radiation from 10º to 70º. The actual La concentration was determined by a scanning electron microscope (SEM, Quanta 200) equipped with an energy dispersive X-ray spectroscope (EDX). Longitudinal and transverse (Hall) resistivities was measured by using a physical properties measurement system (PPMS, Quantum Design), and the magnetic susceptibility was measured by a commercial superconducting quantum interference device (SQUID).
III. RESULTS AND DISCUSSION
Figure 1 presents the actual La doping level x measured by EDX versus the nominal concentration. The actual values of x are almost the same as the nominal compositions at lower doping level, and they linearly increase until x ∼ 0.2. Then, the actual concentration x becomes smaller than the nominal one and saturates at the actual value of 0.25. This behavior is similar to that obtained by wavelength-dispersive spectroscopy (WDS),9 as well as matching EDX results13 from other reports. Figure 2 shows the single-crystal XRD patterns for Ca1−xLaxFe2As2. Only the (00l) peaks were observed, suggesting that the crystallographic c-axis is perfectly perpendicular to the plane of the single crystal. The lattice constant c has been calculated as a function of the La concentration x and is plotted in the inset of Figure 2. There is no obvious change in the lattice constant, regardless of the La doping level, which may be due to the close match between the ionic radii of La (130 pm) and Ca (126 pm).9,13
EDX results for the actual La concentration vs. nominal La concentration for Ca1−xLaxFe2As2.
EDX results for the actual La concentration vs. nominal La concentration for Ca1−xLaxFe2As2.
Single crystal X-ray diffraction pattern of Ca1−xLaxFe2As2 for different values of x. The inset shows the lattice constant c vs. the concentration of La.
Single crystal X-ray diffraction pattern of Ca1−xLaxFe2As2 for different values of x. The inset shows the lattice constant c vs. the concentration of La.
The temperature dependence of the in-plane resistivity ρ (T) of the Ca1−xLaxFe2As2 single crystals is shown in Figure 3(a). The data for each sample are normalized by the room-temperature value ρ (300K) and are subsequently shifted by 0.3 on the y-axis for clarity. The resistivity of the undoped parent compound CaFe2As2 exhibits metallic behavior over the entire temperature range with a sharp step-like increase at the temperature of about 160 K. The anomaly in the resistivity is associated with the structural and magnetic phase transitions at temperatures Ts and Tm, respectively.15 With La doping, the anomaly gradually broadens and shifts to lower temperature, and it disappears when the proportion of La is over 13%. The suppression of the resistivity anomaly can also be seen clearly in the temperature dependence of the d(ρ(T)/ρ(300K))/dT curve for x = 0.06, as shown in Figure 3(b). No distinct difference between the structural and magnetic transitions is manifested for the x = 0 and 0.04 samples (for clarity, just the data for x = 0 are shown), although, with increasing La doping, the combined structural/magnetic transition splits into two anomalies. Although no detailed study on the thermodynamic or transport properties has been conducted to distinguish these two phase transitions in Ca1−xRExFe2As2, based on the results on the analogous Ba-122 compounds16,17 and Co-doped Ca-122,18 it is natural to attribute the higher temperature to the structural phase transition and the lower temperature to the magnetic phase transition. The criteria of Ts and Tm are the same as those that are generally used for similar compounds.18–21 Further experiments on neutron scattering are hopefully to identify the structural and magnetic transitions, and verify their separation.
(a) Temperature dependence of the in-plane resistivity for Ca1−xLaxFe2As2, normalized to the room temperature value. Each subsequent curve for the next x value is shifted downward by 0.3 for clarity. (b) d(ρ(T)/ρ(300K))/dT for x = 0 and 0.06. The arrows indicate the magnetic (Tm) and the structural transitions (Ts). Temperature dependence of (c) ρ/ρ50K for 0.13 ⩽ x ⩽ 0.19, where the two steps in the superconducting transition can be clearly distinguished, and (d) d(ρ/ρ50K)/dT for x = 0.1, 0.15, 0.17, and 0.21. The arrows show the values of Tc that have been determined for the two superconducting phases.
(a) Temperature dependence of the in-plane resistivity for Ca1−xLaxFe2As2, normalized to the room temperature value. Each subsequent curve for the next x value is shifted downward by 0.3 for clarity. (b) d(ρ(T)/ρ(300K))/dT for x = 0 and 0.06. The arrows indicate the magnetic (Tm) and the structural transitions (Ts). Temperature dependence of (c) ρ/ρ50K for 0.13 ⩽ x ⩽ 0.19, where the two steps in the superconducting transition can be clearly distinguished, and (d) d(ρ/ρ50K)/dT for x = 0.1, 0.15, 0.17, and 0.21. The arrows show the values of Tc that have been determined for the two superconducting phases.
Along with the suppression of the anomaly, superconductivity emerges when x is about 0.06, and it coexists with the structural/magnetic transition. The resistivity behavior for La doping in the low doping region is very close to that of Ca(Fe1−xCox)2As2.18 With more La doping, however, Tc is not totally suppressed after the doping level reaches its maximum value, which is quite different from the case of Co-doped CaFe2As2, as well as other electron- or hole-doped iron-based 122 samples, for which a whole superconductivity dome was detected.18,20–22
Another distinct feature of the temperature dependent resistivity of Ca1−xLaxFe2As2 is the two superconducting transition steps, which are clearly shown in Figure 3(c). The feature of two superconducting phases is also observed in Pr-doped CaFe2As2.11 Here, we found that the two-phase feature exists at the medium doping level from x = 0.13 to 0.19, as shown in Figure 3(c). To further separate the two phases and accurately obtain the Tc of each transition, the derivative of the temperature dependent resistivity, d(ρ/ρ50K)/dT is plotted versus T in Figure 3(d). For clarity, just the data for x = 0.15 and 0.17 are shown, together with selected data for comparison from the two x values on either side, where the two-phase feature was not observed. The transition temperatures were obtained as the beginnings of the peaks in the d(ρ/ρ50K)/dT curves. TcH and TcL are defined as the higher and lower transition temperature, respectively. Thus, Ca1−xLaxFe2As2 can be divided into three doping regions: the low doping region with only one Tc, which is lower than 20 K; the medium doping region with two separate Tcs, where the lower one, TcL, is about 20 K, and the higher one, TcH, is over 40 K; and the high doping region with only one Tc at about 40 K.
To further confirm the transition temperature of the samples, the temperature dependence of the magnetization, M-T, was measured on several selected samples from the three different doping regions mentioned above, and the results are shown in Figure 4, together with the temperature dependence of the resistivity for comparison. In the low doping region, the transition temperature obtained from the magnetic susceptibility measurement, TcM, is almost the same as that from the transport results. In the medium doping region, TcM is close to the transport value for the lower Tc phase, TcL. In the high doping region, however, TcM is only around 20 K, much lower than the value of about 40 K obtained from the resistivity measurement and close to the value for the lower Tc phase in the medium doping region. Unlike the transport measurements, which only probe the superconducting percolative paths, the magnetic susceptibility results show the bulk properties of the sample. Thus, the lack of detection of the high Tc phase in M-T demonstrates that the superconducting phase with Tc higher than 40 K is not a bulk property. The much lower Tc observed from the magnetic susceptibility than from the transport measurements has also been reported previously.9,10,12 Although a Tc higher than 40 K from M-T was reported,11 it is easily suppressed by a very small field of about 100 Oe. The lower Tc phase was further detected by magnetic susceptibility measurements at different fields on the Ca0.83La0.17Fe2As2 sample, as shown in the inset of Figure 4(b), which demonstrates that the Tc remains constant with increasing field. Thus, the low Tc phase seems to be a robust global superconducting (SC) phase. The value of Tc ∼ 20 K is also observed in La doped SrFe2As2,23 and it is close to the values found in other electron doped 122 samples, such as Ca(Fe1−xCox)2As2,18 and Ba(Fe1−xMx)2As2.20,21 On the other hand, the high Tc above 40 K in Ca1−xRExFe2As2 may be coming from filamentary or interface superconductivity, as will be discussed later.
(a) Temperature dependence of ρ/ρ50K and (b) M-T results for the Ca1−xLaxFe2As2 samples. The inset of (b) shows M-T measurements on the Ca0.83La0.17Fe2As2 sample at different fields.
(a) Temperature dependence of ρ/ρ50K and (b) M-T results for the Ca1−xLaxFe2As2 samples. The inset of (b) shows M-T measurements on the Ca0.83La0.17Fe2As2 sample at different fields.
Now, we must reconsider the three different doping regions identified above in the discussion of resistivity and provide a short summary. In the low doping region, just one low Tc phase is observed. Then, with increasing La doping, the high Tc phase emerges. In this medium doping region, the volume of high Tc phase is still very low, and it is in disjoint patches separated by the low Tc phase, so both phases can be detected in the transport measurements. Finally, in the high doping region, the volume of high Tc phase is sufficient to form a continuous percolative path, so that the current will no longer pass through the low Tc part of the sample, and it cannot be detected by transport measurements. Even though the high Tc phase can form a continuous percolative path, its volume with respect to the whole sample is still too small to be observed in the magnetic susceptibility measurements.
Based on the resistivity and magnetic measurements described above, we can establish a doping-temperature (x-T) phase diagram for Ca1−xLaxFe2As2, as shown in Figure 5. On the under-doped side of the phase diagram (x⩽0.13), the structural/magnetic phase transition is monotonically suppressed with increasing La substitution, while at the same time, superconductivity emerges from x = 0.06, and then coexists with the orthorhombic/antiferromagnetic phase until x = 0.13, when the structural/antiferromagnetic phase transition is totally suppressed. The evolution of structural, magnetic, and SC phases with La doping in the under-doped region is similar to what occurs in other electron-doped iron-based 122 superconductors. The suppression rate of Ts/Tm, however, is roughly 6 K per atomic percent La substitution, much smaller than the value of 15 K per atomic percent Co doping in Ba(Fe1−xCox)2As2 or the value of 10 K for Ca(Fe1−xCox)2As2.18,21 The structural/magnetic phase transition can also persist to 13% La substitution, which is higher than 5.8% for Ba(Fe1−xCox)2As2 and 7.5% for Ca(Fe1−xCox)2As2.18,21
T-x phase diagram of Ca1−xLaxFe2As2 obtained from magnetic and transport results.
T-x phase diagram of Ca1−xLaxFe2As2 obtained from magnetic and transport results.
We now focus on the SC region of the phase diagram, which contains two separate Tc phases. As discussed above, the lower value of TcL obtained from the R-T measurements together with the value of TcM from the M-T measurements, corresponds to the low Tc phase. With further La substitution, the low Tc phase maintains its Tc value of roughly 20 K. Meanwhile, the high Tc phase emerges, and the value of TcH, which has been increasing with La doping, reaches its maximum value of about 45 K for 0.15 ⩽ x ⩽ 0.19 and then decreases to a lower value around 30 K. The high Tc phase shows an incomplete domelike appearance on the phase diagram, which does not reach the over-doping region, as Tc cannot be totally suppressed. Here, we must emphasize that because of its non-bulk nature, the absolute Tc value of the high Tc phase may exhibit some fluctuations between different samples. The two-phase property is independent of the crystal, however, and it will not affect the discussion on the origins of the high Tc phase below.
Before probing the origins of the two separate SC phases, the doping type of Ca1−xRExFe2As2 should be first considered. Figure 6 shows the temperature dependence of the Hall coefficient RH of Ca0.95La0.05Fe2As2 and Ca0.79La0.21Fe2As2. For Ca0.95La0.05Fe2As2, the sharp increase in the absolute value of RH below TM indicates a sudden drop in the carrier density with the magnetic transition, which is a common feature in iron-based superconductors.24 In the case of Ca0.79La0.21Fe2As2, the magnetic transition has been totally suppressed, and the strong temperature dependence of RH is often attributed to multiband effects or to the pseudogap.25,26 The negative values of RH for both samples indicate that the dominant carrier for Ca1−xRExFe2As2 is the electron. By using RH = 1/ne, the charge carrier density, n, was roughly estimated and is shown in the inset of Figure 6. Compared to Ca0.95La0.05Fe2As2, the charge carrier density in Ca0.79La0.21Fe2As2 is obviously enhanced, since more La doping induces more electrons into the system.
Temperature dependence of the Hall coefficient RH for x = 0.05 and 0.21. Inset shows the temperature dependence of the charge-carrier density n for these two samples.
Temperature dependence of the Hall coefficient RH for x = 0.05 and 0.21. Inset shows the temperature dependence of the charge-carrier density n for these two samples.
The origin of the observed high Tc phase in Ca1−xLaxFe2As2 may be explained by four possible scenarios: (i) Josephson junction coupling across the grains; (ii) filamentary superconductivity caused by local pinning strength or defects; (iii) minor foreign phase; and (iv) interface superconductivity. Scenario (i) usually accounts for the observation of two superconducting transitions in granular polycrystalline superconductors, which is in contradiction with the good crystallinity of the single crystal observed from the XRD results. In favor of scenario (ii), filamentary superconductivity is a common problem in 122 iron-based superconductors and has already been reported in BaFe2As2, SrFe2As2, and CaFe2As2.27–29 In this case, some local or mesoscopic structural defect or surface strain similar to that in some thin films will cause some pinning strength, which will induce superconductivity in a very small fraction of the sample. The filamentary superconductivity caused by local pinning strength or defects is usually sensitive to, or even easily removed by heat-treatment, pressure, or magnetic field.27–31
Annealing at high temperature is an easy and direct way, proved effective in SrFe2As229 and CaFe2As2,28 to eliminate filamentary superconductivity. Thus, we vacuum annealed the Ca1−xLaxFe2As2 single crystal at temperatures ranging from 300 °C to 800 °C for periods ranging from several hours to as long as two weeks. Figure 7 shows typical resistivity results for Ca0.79La0.21Fe2As2 annealed at 800 °C for two weeks. Although the residual resistivity ratio (RRR) increased a little due to the removal of some defects, Tc didn't change after annealing, as can be seen more clearly from the enlarged transition part in the inset of Figure 7. Recently Swee K. Goh32 reported pressure dependent resistivity measurements on Ca1−xLaxFe2As2, which showed that Tc is not suppressed, even under pressure greater than 40 kbar. Actually, an obvious two-step resistivity transition was also witnessed in their sample.
Temperature dependence of normalized resistivity of Ca0.79La0.21Fe2As2 before and after vacuum annealing at 800 °C for two weeks. Inset is an enlargement of the area containing the Tc.
Temperature dependence of normalized resistivity of Ca0.79La0.21Fe2As2 before and after vacuum annealing at 800 °C for two weeks. Inset is an enlargement of the area containing the Tc.
To check the influence of magnetic field on the high Tc phase, the temperature dependence of the resistivity with applied fields up to 9 T was measured on Ca0.9La0.1Fe2As2, Ca0.83La0.17Fe2As2, and Ca0.79La0.21Fe2As2 (selected from the three different doping regions of the phase diagram) and is plotted in Figure 8(a)–8(c). Magnetic fields were applied along the c-axis, which has already been proved to be more sensitive to the field than the ab-plane.10,12 Tc is gradually suppressed to lower temperature, and the transition is broadened with increasing magnetic field, which is similar to what has been reported for 1111 phase.33,34 The upper critical fields, Hc2, defined by 90 % of normal state resistivity is plotted in Figure 8(d) for three different x values. For Ca0.79La0.21Fe2As2, with only the high Tc phase, the slope of Hc2 is evaluated as −13.8 kOe/K from the linear fit to the Hc2 – T curve above 2 T. Using the Werthamer-Helfand-Hohenberg formula,35 |$H_{{\rm c2}} (0) = - 0.693T_{\rm c} {\rm d}H_{{\rm c2}} /{\rm d}T\left| {_{T = T_{\rm c} } } \right.$|, |$H_{c2}^{0.21} (0)$| is simply estimated as 38.5 T. In the case of Ca0.83La0.17Fe2As2, two upper critical fields were obtained, belonging to the high Tc phase and the low Tc phase respectively, which again proves the existence of two phases in the medium doping region of Ca1−xRExFe2As2. With the slopes of Hc2 determined to be −7.7 kOe/K and −11.0 kOe/K, approximately, the upper critical fields at 0 K for the high Tc and low Tc phases are 22.4 T and 13.7 T, respectively. Hc2 of 6.8 T at 0 K for Ca0.9La0.1Fe2As2 can be obtained by the same method, although the Hc2 for this sample shows a positive curvature different from the negative ones for Ca0.83La0.17Fe2As2 and Ca0.79La0.21Fe2As2, which may be due to inhomogeneity in the sample with the low doping level of La. A detailed study of the upper critical fields, as well as the resistivity tail, which is a common feature of the Ca1−xRExFe2As2,9–12 is reported elsewhere.36 Here, the results on the field dependence of the resistivity demonstrate that the two phases are not fragile with respect to the magnetic field. Thus, the heat-treatment, field, and pressure dependence of the superconducting transition temperature results indicate that the high Tc phase is not a filamentary-type superconductivity caused by local pinning strength or defects.
Temperature dependence of resistivity with applied fields up to 9 T measured on samples with (a) x = 0.1, (b) x = 0.17, and (c) x = 0.21 (selected from the three different doping regions of the phase diagram). (d) Upper critical fields of samples with x = 0.1, 0.17, and 0.21.
Temperature dependence of resistivity with applied fields up to 9 T measured on samples with (a) x = 0.1, (b) x = 0.17, and (c) x = 0.21 (selected from the three different doping regions of the phase diagram). (d) Upper critical fields of samples with x = 0.1, 0.17, and 0.21.
For scenario (iii), although the powder XRD patterns (data not shown) demonstrate that the diffraction peaks can be well indexed by the structure of Ca1−xLaxFe2As2, and only impurity peaks from FeAs flux can be observed, we cannot simply exclude the possibility of the existence of trace amounts of foreign phase which cannot be distinguished by XRD. The minor foreign phase that might possibly exist is not a polycrystal-like impurity phase, though, because the upper critical fields12,14 show anisotropy of about 3. It may be caused by chemical phase separation. With increasing La doping, small amounts of foreign phase with high Tc will emerge and coexist with the low Tc phase. In this case, two superconducting phases would be observed. Then, as La doping is further increased, the high Tc phase can form a continuous percolative path in the ab-plane, so that just one phase can be observed in the high doping region. With regards to scenario (iv), interface superconductivity can have an enhanced Tc.37 In the case of Ca1−xLaxFe2As2, the interfaces may be provided by alternate stacking of perfect and defective FeAs layers. Further work, especially on the microstructure/composition, will hopefully allow us to distinguish the origin of the high Tc phase in Ca1−xRExFe2As2 between the minor foreign phase and the interface superconductivity
IV. CONCLUSION
In summary, we have successfully grown single crystals of Ca1−xLaxFe2As2 (0 ⩽ x ⩽ 0.25), and determined the phase diagram based on transport and magnetic measurements. We find that the single magnetic/structural phase transition for CaFe2As2 splits with La substitution. The superconductivity emerges when x is about 0.06, and it coexists with the structural/magnetic transition until x ∼ 0.13, when the Tc reaches its maximum value of about 45 K. We did not observe a domelike SC region, however, because La substitution seems to saturate at x ∼ 0.25, and cannot reach the over-doping region. Another distinct feature of the phase diagram is the two separate SC phases, in which the low Tc phase exhibits robust global superconductivity and is close to the SC phases of other electron-doped iron-based 122 samples. On the other hand, the non-bulk high Tc phase with Tc higher than 40 K may be caused by a minor foreign phase or interface superconductivity.
ACKNOWLEDGMENTS
This work was supported by the Natural Science Foundation of China, the Ministry of Science and Technology of China (973 Project: No. 2011CBA00105), the Scientific Research Foundation of the Graduate School (Grant No. YBJJ1104) of Southeast University, the Scientific Innovation Research Foundation of College Graduates in Jiangsu Province (CXZZ_0135), and the Jiangsu Science and Technology Support Project (Grant No. BE2011027).