This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Normal-state resistivity and the depairing current density of BaFe2(As,P)2 nanobridges along the cc axis

Yuki Mizukoshi1, Kotaro Jimbo1, Akiyoshi Park2, Yue Sun3, Tsuyoshi Tamegai2, Haruhisa Kitano1 1Department of Physics, Aoyama Gakuin University, Sagamihara 252-5258, Japan
2Department of Applied Physics, The University of Tokyo, Tokyo 113-8656, Japan
3Department of Physics, Southeast University, Nanjing 211189, China
Abstract

We report precise measurements to obtain the normal-state resistivity and the depairing current density of BaFe2(As1-xPx)2(x0.290.32x\sim 0.29-0.32) nanobridges along the cc axis, by using focused ion beam (FIB) techniques. We obtained both of the abab-plane and cc-axis resistivity (ρab\rho_{ab} and ρc\rho_{c}) in the same part of a specimen, by fabricating the cc-axis nanobridge in the middle of a narrow bridge extended in abab-plane, in spite of the slight deficiency of P dopant due to the additional FIB fabrication. The normal-state resistivity anisotropy agreed with the previous results for bulk samples, showing ρc/ρab<8\rho_{c}/\rho_{ab}<8 just above the superconducting transition temperature, TcT_{c}, in the slightly underdoped region and a slight decrease with increasing temperatures. The critical current density obtained in the cc-axis nanobridges near the optimal doping reaches \sim8 MA/cm2 at 0.15TcT_{c}, corresponding to about 87% of a depairing limit derived by the Eilenberger equations. An extrapolation to T=T=0 K using the Ginzburg-Landau model suggests that the anisotropy of the depairing current density roughly corresponds to that of the normal-state resistivity. At low temperatures, we also observed a step-like voltage jump before arriving at the depairing limit, suggesting the occurrence of phase-slip phenomena near the depairing processes.

I Introduction

The phase diagram of BaFe2As2-based superconductors has been extensively investigated to explore key fluctuations giving rise to high temperature superconductivity in iron-based superconductors (IBSs) [1]. There are three phase diagrams in K, Co, and P-doped BaFe2As2, which are regarded as the hole-doped, electron-doped, and isovalent substituted system, respectively. In these systems, the superconducting (SC) transition temperature, TcT_{c}, has a dome-shaped dependence on the concentration of K, Co and P and the SC phase is in proximity to the antiferromagnetic ordered phase, showing strong similarity to cuprate superconductors.

However, in contrast to cuprates, the relative low anisotropy of transport properties in the normal-state and superconducting state was reported by several studies on the electron-doped BaFe2As2 [2, 3, 4] and P-doped BaFe2As2 [5]. In addition, the softness of this material makes the determination of transport properties along the cc axis quite difficult, since partial cracks introduced by cutting and exfoliating single crystals influence the current distribution seriously [2]. Thus, the normalized data of the cc-axis resistivity by the measured value at 300 K were often plotted in the previous studies [4, 5].

Recently, we demonstrated that a narrow bridge structure fabricated by using focused ion beam (FIB) techniques was quite useful to determine the normal-state resistivity anisotropy of Fe1+yTe1-xSex single crystals [6]. In contrast to the conventional methods using two samples with different electrode for the resistivity measurements in the abab-plane and along the cc-axis (ρab\rho_{ab} and ρc\rho_{c}, respectively), the advantage of FIB method is that both of ρab\rho_{ab} and ρc\rho_{c} can be measured in the same region of a specimen, by additionally making a short nanobridge along the cc axis into the narrow bridge extended to the abab-plane. In the previous study on Fe1+yTe0.6Se0.4[6], we found that the normal-state anisotropy determined by the ratio of ρc/ρab\rho_{c}/\rho_{ab} was systematically decreased with the removal of interstitial excess iron.

By using the same narrow bridge structure, we also succeeded in measuring the critical current density, jcj_{c}, for the depairing of Cooper pairs moving along the cc axis of Fe1+yTe1-xSex single crystals [7]. The transport jcj_{c} for the cc-axis microbridge with a width smaller than the Pearl length reached 1.3 MA/cm2 at 0.3Tc\sim 0.3T_{c}. This value is at least one order of magnitude larger than the depinning jcj_{c}, obtained by magnetization hysteresis loop (MHL) measurements, while the result for transport jcj_{c} is comparable to a calculated value of the depairing jcj_{c} based on the Ginzburg-Landau (GL) theory. The temperature dependence of jcj_{c} showed a quantitative agreement with the GL theory down to 0.83TcT_{c} and a qualitative agreement with Kupriyanov-Lukichev (KL) theory [8] down to 0.3Tc\sim 0.3T_{c}. In the KL theory, the contribution of a growth of the SC gap is considered through a numerical solution of Eilenberger equations.

In this study, we applied the FIB method to BaFe2(As1-xPx)2 (x=0.32x=0.32) single crystals, in order to measure both ρab\rho_{ab} and ρc\rho_{c} as well as the depairing jcj_{c} along the cc axis. We found that this method facilitated the detection and avoidance of partial cracks affecting the local current distribution, compared to the conventional method using bulk samples. The magnitude and temperature dependence of ρab\rho_{ab} show an excellent agreement with the previously-reported results using bulk crystals [9]. On the other hand, ρc(T)\rho_{c}(T) shows a slightly negative curvature in ρc(T)\rho_{c}(T) at around 250 K, indicating that the concentration of P after the additional FIB fabrication is slightly in the underdoped side of the SC dome [5]. We actually confirmed the slight decrease of xx (less than 0.01) using the energy-dispersive X-ray (EDX) analyses for the fabricated narrow bridges. The normal-state anisotropy ρc/ρab\rho_{c}/\rho_{ab} for the slightly underdoped side near the optimal doping (xx\sim0.32) is at most 8 just above TcT_{c} and shows a slight decrease with increasing temperature. The magnitude of ρc/ρab\rho_{c}/\rho_{ab} semiquantitatively agrees with the previously-reported results for bulk crystals [5].

We also found that the cc-axis transport jcj_{c} for the nanobridge showing the highest TcT_{c}(=28 K) reached 8 MA/cm2 at \sim0.15TcT_{c}. It is at least 9 times larger than the depinning jcj_{c} in the abab-plane determined by MHL for the optimally-doped bulk crystals with x=0.33x=0.33 [10, 11], and corresponds to \sim 87 % of the depairing jcj_{c} derived from the KL theory. At low temperatures, we also discovered a sharp jump in the current-voltage (II-VV) characteristics just below the depairing jcj_{c}, suggesting the contribution of phase slip phenomena near the depairing limit.

II Experiment

Single crystals of BaFe2(As1-xPx)2 were grown by using Ba2As3/Ba2P3 flux [11]. Several samples with x=0.32x=0.32 are carefully cleaved and cut into rectangular shape (typically, 700 μ\mum×\times150 μ\mum×\times10-20 μ\mum). The value of xx for the cleaved samples before the FIB microfabrication was determined by energy-dispersive X-ray (EDX) analysis and the cc-axis lattice parameter determined by the (00ll) peaks in the X-ray diffraction (XRD) pattern [11]. The cleaved sample is glued on a sapphire substrate using an epoxy (Stycast 1266). Four Au electrodes are prepared on each sample by using a sputtering technique.

A narrow bridge is fabricated between two voltage electrodes by using the FIB etchings (Hitachi High-Tech MI4050), as shown in Fig. 1(a). The existence of partial cracks is directly checked by observing scanning ion microscope (SIM) images of the bridge. Even if a tiny crack hides into the narrow bridge part, we can easily find its influence through the dc resistance measurements down to low temperatures at a rate of 15-20 K/h, since the narrow bridge including such tiny cracks cannot suffer stress due to thermal shrinkage. After measuring the dc resistance (R1R_{1}) for the microbridge along the abab-plane, we fabricate the cc-axis nanobridge by adding two slits, as shown in Fig. 1(b). Then, we measure the dc resistance (R2R_{2}) for the cc-axis nanobridge. By using R1R_{1} and R2R_{2}, both of ρab\rho_{ab} and ρc\rho_{c} in the same region of the microbridge can be obtained, as described in the previous study [6]. As described in Supplementary Information [12, 13, 14], we went through the limitation of this method for materials with small anisotropy. Finally, the transport jcj_{c} is measured for the cc-axis nanobridge by using a combination of pulse current source and nanovoltmeter (Delta mode of Keithley 6221/2182A) [7]. The pulse width and the repetition period are 100 μ\mus and 2\sim3 s, respectively.

In this paper, we focus on 5 nanobridges, as listed in Table I. Here, the value of xx for each bridge was determined by EDX analyses (Zeiss ULTRA55/Bruker QUANTAX) after measuring the transport jcj_{c}. The value of TcT_{c} was determined by a maximum temperature of the zero-resistance in R2R_{2} measurements (TcRT_{c}^{R}) or the onset temperature of the critical current density (JcJ_{c}) in the JcJ_{c}-VV characteristics (TcJT_{c}^{J}). We found that the FIB microfabrication removed P dopants slightly in the bridge part of BaFe2(As1-xPx)2 single crystals, leading to a distribution of TcT_{c} in the fabricated bridge samples, as shown in Table I [15].

Table 1: Sizes of the fabricated bridges and properties in the SC state. xx is the concentration of P, determined by EDX analyses for three points on the side wall of the narrow bridges. ww and hh are the lateral sizes of the nanobridge along the cc axis, and ll is a length of the nanobridge, as shown in Fig. 1. TcRT_{c}^{R} and TcJT_{c}^{J} are determined by a rising temperature of R2(T)R_{2}(T) and jc(T)j_{c}(T), respectively. jcGL(0)j_{c}^{\rm GL}(0) is an extrapolated value by the fitting of the measured data just below TcT_{c} to a behavior of (1T/Tc)1.5(1-T/T_{c})^{1.5} in the GL theory.
sample xx w(μw~{}(\mum) h(μh~{}(\mum) l(μl~{}(\mum) TcRT_{c}^{R} (K) TcJT_{c}^{J} (K) jcGL(0)j_{c}^{\rm GL}(0) (MA/cm2)
bp06 0.324 0.3 1.3 0.8 27 27 17.6
bp07 0.317 0.9 0.8 0.4 26 26 6.0
bp16 0.287 0.3 1.4 2.0 16 18 14.7
bp18 0.320 0.3 1.0 0.2 26 27.5 15.2
ij12 0.321±\pm0.001 0.6 0.8 0.2 28 28 24.3

III Results and Discussion

Figures 1(a) and 1(b) show the temperature dependence of R1R_{1} and R2R_{2} for a sample (bp07), respectively. As described in the previous study [6], the value of R1R_{1} is given by a sum of several resistances in the abab plane (RabR_{ab}), while there is a resistance along the cc axis (RcR_{c}) as well as the sum of RabR_{ab} in the measurements of R2R_{2}. Note that the contribution of RcR_{c} is enhanced by the small cross-sectional area of the cc-axis nanobridge, compared to the conventional Corbino electrode configuration to obtain the out-of-plane resistivity in a plate-like sample. Thus, ρab\rho_{ab} and ρc\rho_{c} are obtained with sufficient accuracy from R1R_{1} and R2R_{2}, as shown in Fig. 2(a). The magnitude and the temperature dependence of ρab\rho_{ab} show almost perfect agreement with the previous results for the optimally-doped sample (xx=0.33) [9]. On the other hand, the temperature dependence of ρc\rho_{c} seems to be slightly underdoped region in the SC dome, since a small negative curvature is observed in ρc(T)\rho_{c}(T) at around 250 K. In the previous study using bulk crystals [5], only a TT-linear behavior was observed in ρc(T)\rho_{c}(T) with xx=0.33, while ρc(T)\rho_{c}(T) in the underdoped samples showed a shallow maximum at higher temperatures. In addition, it is found that the superconducting transition observed in ρc\rho_{c} occurs at a slightly lower temperature than that observed in ρab\rho_{ab}. EDX analyses for this sample also indicate a slight decrease of xx (<1<1%). These results suggest an influence of the slight deficiency of P dopant due to the additional FIB fabrication.

The normal-state resistivity anisotropy is at most 8 just above TcT_{c}, as shown in Fig. 2(b). Note that this value is presumably overestimated, since the slight decrease of xx is expected to push up ρc(T)\rho_{c}(T). Nevertheless, the obtained anisotropy of ρc/ρab\rho_{c}/\rho_{ab} seems to agree with the previous results for a bulk crystal of xx=0.33 (ρc/ρab6±2\rho_{c}/\rho_{ab}\approx 6\pm 2)[5]. We also investigated the dependence of ρc\rho_{c} on the length of nanobridge along the cc axis [12], and found that the additional FIB fabrication to lengthen the nanobridge leads to an increase RabR_{ab} as well as RcR_{c}, since the elongation of a slit made in the side wall of narrow bridge gives rise to both an increase of the length of the cc-axis nanobridge and a decrease of the thickness of in-plane bridge. Thus, in the case of small anisotropy of ρc/ρab\rho_{c}/\rho_{ab}, the nanobridge with a shorter length is more useful to obtain the contribution of RcR_{c}.

Figure 3(a) shows the II-VV characteristics of another sample (ij12) measured at several temperatures below TcT_{c}, which shows the maximal values of TcT_{c} and jcj_{c} among the measured samples. R2(T)R_{2}(T) measurements for this sample showed almost TT linear behavior and no shallow maximum at higher temperatures, suggesting that this sample (ij12) was actually near the optimal doping. In contrast to other samples, we observed two-staged voltage jump at low temperatures below 14 K, as shown in Fig. 3(a). After the first II-VV measurements shown in Fig. 3(a), we performed the 2nd measurements below 14 K again, in order to check the reproducibility of the two-staged voltage jump. As shown in Fig. 3(b), we found the stochastic nature of the 1st voltage jump observed at a lower current, since the current position corresponding to the 1st jump was decreased for TT=10 K and 8 K, while it was increased for TT=6 K. On the other hand, we found that the variation of the current positions corresponding to the 2nd jump between two measurements was much smaller than those for the 1st jump. Thus, we determined the critical current, IcI_{c}, by using two criteria of 10 μ\muV and 50 μ\muV, as shown in Fig. 3(c). Here, we used the firstly-measured data shown in Fig. 3(a), since the SC properties in the 2nd II-VV measurements seemed to be slightly degraded. The magnitude of jcj_{c} was obtained by using a cross-sectional area of the cc-axis nanobridges, given by a product of w×hw\times h. Since there wa no report on λL\lambda_{L} along the cc axis for the optimally doped BaFe2(As1-xPx)2, we roughly estimated the Pearl length (Λ=2λL2/h\Lambda=2\lambda_{L}^{2}/h) by assuming that λL\lambda_{L}\simμ\mum. Note that this value is comparable to the reported value for λL\lambda_{L} along the cc axis in the optimally doped Ba(Fe1-xCox)2As2 [16]. By using such an estimation, we confirmed that the cross-sectional area (\sim0.48 μ\mum2 for sample ij12) of the cc-axis nanobridge was smaller than 2λL22\lambda_{L}^{2} along the cc axis. Thus, it is considered that the spatial distribution of the supercurrent flowing along the cc-axis in the fabricated nanobridges is almost homogeneous [7, 17], except for sample bp07 with a larger cross-sectional area than that of sample ij12.

As shown in Fig. 3(c), the temperature dependence of jcj_{c} determined by the lower criterion of 10 μ\muV shows a small dip between 10 Kand 15 K, which is probably due to the two-staged jumps occurring below 14 K. On the other hand, jc(T)j_{c}(T) determined by the higher criterion of 50 μ\muV shows a monotonic increase down to the lowest temperature and seems to be more smoothly connected to jc(T)j_{c}(T) above 15 K, which was obtained by the threshold of 10 μ\muV. This suggests that the second jumps observed below 14 K are closer to the depairing limit and that the first jumps are rather caused by another origin, as discussed below. By adopting the values of jcj_{c} at the second jumps below 12 K, we compared the temperature dependence of jcj_{c} normalized by jc(0)j_{c}(0) with the GL and KL theories, as shown in Fig. 3(d). Here, jcGL(0)j_{c}^{\rm GL}(0)(=24.3 MA/cm2) was determined by an extrapolation of the fitting data obtained just below TcT_{c} to a behavior of (1T/Tc)1.5(1-T/T_{c})^{1.5} predicted by the GL theory. We compared this value with a calculated value of jcGL(0)j_{c}^{\rm GL}(0)(=79 MA/cm2[7]) in the abab plane, based on the GL theory (jc=cϕ0/123π2ξλ2j_{c}=c\phi_{0}/12\sqrt{3}\pi^{2}\xi\lambda^{2}). The obtained anisotropy (jcab(0)/jcc(0)j_{c}^{ab}(0)/j_{c}^{c}(0)\sim3.3) in the SC state seems to be slightly larger than the normal-state anisotropy determined by ρc/ρab\sqrt{\rho_{c}/\rho_{ab}}. Note that the depairing current density is proportional to a critical velocity of Cooper pairs transporting the SC current, which is determined by a crossover of the kinetic energy of Cooper pairs with the SC condensation energy [18]. Thus, the anisotropy of jcab/jccj_{c}^{ab}/j_{c}^{c} is expected to be equal to a square root of the anisotropy of the effective mass, mc/mab\sqrt{m_{c}/m_{ab}}. However, since IBS is a multiband system with several Fermi surfaces, an issue of the anisotropy in the charge transport properties should be more carefully discussed.

Next, we discuss jc(T)/jcGL(0)j_{c}(T)/j_{c}^{\rm GL}(0) by comparing with the GL and KL theories. As shown in Fig. 3(d), jc(T)/jc(0)j_{c}(T)/j_{c}(0) quantitatively agrees with both theories down to \sim0.8 TcT_{c} and with the KL theory down to \sim0.55 TcT_{c}. The experimental result of jcj_{c} obtained at the lowest temperature (TT\sim0.15 TcT_{c}) is \sim8 MA/cm2, which is about 87% of a value obtained by the KL theory and about an order of magnitude larger than the depinning jcj_{c} determined by MHL [10, 11]. These results strongly suggest that the obtained jc(T)j_{c}(T) actually represents the depairing current density along the cc axis of BaFe2(As1-xPx)2 near the optimal doping and that the depairing current density for single-crystalline samples was experimentally explored down to \sim0.15 TcT_{c}. This temperature range is larger than that for the cc-axis depairng current density of Fe1+yTe0.6Se0.4 single crystals (down to \sim0.3 TcT_{c}) [7] and that for the abab-plane depairing current density of the optimally doped (Ba0.5K0.5)Fe2As2 single-crystalline microbridges (down to \sim0.87 TcT_{c}) [19]. In addition, the temperature dependence of jc(T)/jcGL(0)j_{c}(T)/j_{c}^{\rm GL}(0) showing a better agreement with the KL theory indicates that the first jump in the two-staged voltage jumps observed at low temperatures is rather caused by a different origin from the depairing events.

We consider that the first jump is attributed to phase-slip (PS) phenomenon, which has been often observed in narrow SC wires [20, 21], SC strips [22] and the two-dimensional superconductors with a few layers [23]. The characteristic step structure in the II-VV curves was observed in the SC filaments just below TcT_{c} and has been successfully explained by the formation of phase-slip centers (PSCs) and the interactions between PSCs [20, 21]. The two-dimensional analogue of PSC has been observed in wide strips and films [22, 23] and is called as phase-slip line (PSL), characterized by the formation of kinematic vortices or kinematic vortex-antivortex pairs [24, 25].

Unfortunately, the details of such resistive SC states existing between the SC state with zero voltage and the fully normal-state are still poorly understood. In numerical studies based on the time-dependent GL (TDGL) theory and its generalization [26, 27], oscillatory PS solutions intrinsically emerge both below and above the depairing current and the system exhibits a hysteretic behavior. This suggests an intrinsic instability of the SC filament transporting the SC current with a high speed. In this work, the two-staged steps were observed at low temperatures, in contrast to early studies on narrow SC wires [20, 21], where the consecutive steps appeared just below TcT_{c}. In addition, the presence or absence of the step structure was dependent on samples, implying the stochastic and unstable nature of the resistive SC state near the depairing limit. These results clearly show that the PS phenomena with stochastic and unstable nature occur even in the three-dimensional nanobridges made of IBS single crystals. We will discuss this issue in more depth in the near future.

We compared the temperature dependence of jcj_{c} between the fabricated samples except for the sample (bp07), which had the largest width (ww=0.9 μ\mum) and was used to obtain the normal-state anisotropy. The value of jcGL(0)j_{c}^{\rm GL}(0) for bp07 was much smaller than other samples, as shown in Table I. Although the magnitude of jc(<j_{c}(<2 MA/cm2) at the lowest temperature was still larger than the in-plane depinning jcj_{c} [11], a moderate increase of voltage was already observed before a sharp jump even at low temperatures, in contrast to other samples. Thus, we concluded that the critical current density obtained for bp07 was far below the depairing limit and that it was dominated by the vortex intrusion into the nanobridge or some PS phenomena.

Figure 4(a) shows the temperature dependence of jcj_{c} for other 4 samples. Note that the values of jcj_{c} for bp16 and bp18 at the lowest temperature, which are about a half of the value for ij12, are sufficiently larger than the in-plane depinning jcj_{c}. We found that the values of TcJT_{c}^{J} for both bp16 and bp18 were slightly higher than those of TcRT_{c}^{R}, as shown in Table I. This is explained by the fact that R2(T)R_{2}(T) in both samples showed the multi-stage SC transition, which is considered to be induced by the slight spatial distribution of P dopants in the FIB microfabrication. This leads to a depressed value of TcRT_{c}^{R}, since a rising temperature of R2(T)R_{2}(T) measurements is sensitive to a small part of the nanobridge with the lowest TcT_{c}. On the other hand, the value of TcJT_{c}^{J} is determined by jcj_{c} at a threshold voltage with a sharp transition in the II-VV characteristics. The sharp transition is rather affected by a large part of the nanobridge showing the largest decrease in the R2(T)R_{2}(T) measurements. Thus, we used the reduced temperature normalized by TcJT_{c}^{J} in Fig. 4(b).

In spite of the large difference of TcJT_{c}^{J} and jcGL(0)j_{c}^{\rm GL}(0) between bp06, bp16 and ij12, the plots of jc(T)/jc(0)GLj_{c}(T)/j_{c}(0)^{\rm GL} versus T/TcJT/T_{c}^{J} for these samples seem to fall into the almost identical curve in a broad temperature range from TcT_{c} to \sim0.15 TcT_{c}, as shown in Fig. 4(b). This strongly suggests that the depairing jcj_{c} along the cc axis can be directly determined down to 0.15 TcT_{c} for these samples. Meanwhile, the plots for bp18 show a slightly depressed growth below 0.7 TcT_{c}. As shown in Table I, the lateral sizes (ww and hh) and the length (ll) of the cc-axis nanobridge for bp18 is not so different from other samples, suggesting that the depressed jcj_{c} is not derived from the nonuniform distribution of the supercurrent in the cross-sectional area. Thus, jc(T)j_{c}(T) for bp18 may be dominated by the PS events below 0.7 TcT_{c}, similar to the first jump for ij12 at low temperatures. Although we could not observe any step-like structure in the II-VV measurements for this sample, we consider that the oscillatory PS solution giving the step structure is not always stable in the condition of our measurements.

In the previous studies on the measurements of the depairing current density [7, 17, 28], the smaller values of (jc/jcGL)2/3(j_{c}/j_{c}^{\rm GL})^{2/3} than the theoretical values (=0.53=0.53 at TT=0) predicted by the KL theory have often been discussed. Nawaz et al. [17] and Rusanov et al. [28] have focused on the current crowding effects around the corners of YBa2Cu3Oy microbridges and the sample heating effects via contacts on Nb microbridges, respectively. In the cc-axis nanobridges used in this study, there is a 90 turn near the current-gate, as shown in the lower inset of Fig. 1(b), which is expected to cause the current crowding [29]. However, such an effect of the current crowding seems to be nearly common to all the measured samples, since the shape and size of the slits giving the 90 turn are all similar. Furthermore, the contacts of current lead in this study are far enough from the cc-axis nanobridges, as shown in the upper inset of Fig. 1(b). Thus, both the current crowding and the sample heating via contacts never explain the depressed behavior of jcj_{c} for bp18.

On the other hand, Sun et al. have dealt with the multiband effects of Fe1+yTe1-xSex [7]. Such effects attract much attention as a possible explanation for a TT linear behavior observed in the temperature dependence of the upper critical field μ0Hc2\mu_{0}H_{c2} for magnetic field parallel to the cc axis [30]. Although BaFe2(As1-xPx)2 is also a multiband system, the obtained results of jcj_{c} along the cc axis showed a much closer behavior to the calculated curve by the KL theory than the previous results on Fe1+yTe1-xSex. Thus, the pair-breaking effects in nonequilibrium superconductivity including multiband effects should be microscopically understood to resolve the difference between BaFe2(As1-xPx)2 and Fe1+yTe1-xSex.

The recent theoretical study for a dirty superconductor shows that the broadening of quasiparticle density of states (DOS), given by Γ/Δ0\Gamma/\Delta_{0}, reduces the depairing current density [31]. Here, Γ\Gamma is a Dynes parameter representing the broadening in the DOS peak. The comparison between such a calculation and our results suggests that 0.05<Γ/Δ0<0.20.05<\Gamma/\Delta_{0}<0.2. It should be noted that the theoretical decrease of the depairing current density due to the finite value of Γ/Δ0\Gamma/\Delta_{0} seems to appear in a whole temperature region below TcT_{c}, while our results suggest that the deviation from the KL theory becomes prominent below 0.7 TcT_{c}, as shown in Fig. 4(b). Therefore, we conclude that the most plausible candidate for explaining the observed difference from the KL theory is the influence of the PS phenomena. This also implies that the suppression of the PS oscillation giving the multi-staged structure in the II-VV curves is quite important for the precise determination of the depairing current density. Even for ij12 with the largest jc(0)GLj_{c}(0)^{\rm GL}, it is possible that the second jump as well as the first jump are caused by the occurrence of PSCs. In this sense, it is preferable that the depairing current density is measured in the hysteretic region, since the zero-voltage state is expected to be broken at the depairing jcj_{c} when we start from the SC state with j<jcj<j_{c} in the hysteretic region [32].

Finally, we refer to an interesting issue relevant to this work. As well known, BaFe2(As1-xPx)2 is considered to be one of ideal systems to investigate the signature of a quantum critical point (QCP) in the phase diagram [33]. The isovalent substitution of P for As does not induce serious disorders in the charge transport, as proved by the observation of clear quantum oscillations in a wide range of xx [34]. To date, the experimental evidence of QCP located at xc0.3x_{c}\sim 0.3 is given by several results including the non-Fermi liquid behaviors of the resistivity [9], the NMR relaxation rate [35], and the anomalous enhancement of the effective mass, mm^{*} [34, 36, 37]. Particularly, the xx dependence of London penetration depth (λLm\lambda_{L}\propto\sqrt{m^{*}}) attracted much attention as the QCP beneath the superconducting (SC) dome [33, 37]. Recently, a similar peak was also observed in the electron-doped Ba(Fe1-xCox)2As2. This indicates that the signature of QCP is not limited in the clean sytem, since Co-doped BaFe2As2 is considered to be significantly disordered by electron doping, showing a sharp contrast to P-doped BaFe2As2 [38]. We note that the depairing current density given by the GL theory is dependent on mm^{*}, since jc(ξλL2)1/mj_{c}(\propto\xi\lambda_{L}^{2})\propto 1/\sqrt{m^{*}}, while a critical velocity depairing Cooper pairs is proportional to Δ/kF\Delta/k_{F} in the BCS theory, where Δ\Delta is the superconducting gap and kFk_{F} is a Fermi wavenumber, respectively. Thus, it is expected that the xx dependence of the depairing current density in both BaFe2(As1-xPx)2 and Ba(Fe1-xCox)2As2 will provide an important insight to settle this issue.

IV Conclusion

In conclusion, we investigated the cc-axis charge transport of BaFe2(As1-xPx)2 single crystals by using the cc-axis nanobridge structure, additionally fabricated into the narrow abab-plane bridge. This technique is excellent for obtaining the interlayer resistivity of soft materials such as BaFe2(As1-xPx)2, where small cracks are partially introduced by cutting and exfoliating crystals. Although the additional FIB etching brought the slight decrease of P concentration in the narrow bridges of BaFe2(As1-xPx)2, the normal-state resistivity anisotropy in the slightly underdoped side near the optimal doping was found to be at most 8 just above TcT_{c}, showing a good agreement with the previous results for bulk crystals.

We also confirmed that the magnitude of jcj_{c} along the cc axis determined by the II-VV characteristics was much larger than the abab-plane depinning jcj_{c} determined by the previous magnetization measurements. The temperature dependence of jcj_{c} near the optimal doping showed a good agreement with the KL theory, indicating the achievement of the depairing limit along the cc axis. These results ensure that we can explore the xx dependence of the depairing jcj_{c} of BaFe2(As1-xPx)2 and Ba(Fe1-xCox)2As2 by using the cc-axis nanobridge, in order to settle the unresolved issue of quantum critical behavior.

In addition, we observed the two-staged voltage jumps in the II-VV characteristics at low temperatures, strongly suggesting the contribution of phase-slip phenomena near the depairing limit. The slight deviation of the depairing jcj_{c} from the KL theory observed at low temperatures may also be attributed to the contribution of phase-slip phenomena. Although such phenomena are still far from the complete understanding, our results manifestly show that they occur not only in narrow wires and two-dimensional strips but also in three-dimensional bridges with small cross-sectional area. Detailed studies on the II-VV characteristics near the depairing limit will provide useful information to approach the complete comprehension.

Acknowledgements.
This work was partly supported by KAKENHI from JSPS (JP20H05164) and Aoyama Gakuin University Research Institute grant programs for research unit and creation of innovative research. FIB microfabrication in this work was supported by Center for Instrumental Analysis, College of Science and Engineering, Aoyama Gakuin University.

References

  • Fernandes et al. [2014] R. M. Fernandes, A. V. Chubukov, and J. Schmalian, What drives nematic order in iron-based superconductors?, Nature Physics 10, 97 (2014).
  • Tanatar et al. [2009a] M. A. Tanatar, N. Ni, C. Martin, R. T. Gordon, H. Kim, V. G. Kogan, G. D. Samolyuk, S. L. Bud’ko, P. C. Canfield, and R. Prozorov, Anisotropy of the iron pnictide superconductor Ba(Fe1-xCox)2As2 (xx=0.074, Tc{T}_{c} = 23 K), Physical Review B - Condensed Matter and Materials Physics 7910.1103/PhysRevB.79.094507 (2009a).
  • Tanatar et al. [2009b] M. A. Tanatar, N. Ni, G. D. Samolyuk, S. L. Bud’Ko, P. C. Canfield, and R. Prozorov, Resistivity anisotropy of AAFe2As2 (AA=Ca, Sr, Ba): Direct versus Montgomery technique measurements, Physical Review B - Condensed Matter and Materials Physics 7910.1103/PhysRevB.79.134528 (2009b).
  • Tanatar et al. [2011] M. A. Tanatar, N. Ni, A. Thaler, S. L. Bud’Ko, P. C. Canfield, and R. Prozorov, Systematics of the temperature-dependent interplane resistivity in Ba(FeMx1x{}_{1-x}{M}_{x})2As2 (M{M}=Co, Rh, Ni, and Pd), Physical Review B - Condensed Matter and Materials Physics 8410.1103/PhysRevB.84.014519 (2011).
  • Tanatar et al. [2013] M. A. Tanatar, K. Hashimoto, S. Kasahara, T. Shibauchi, Y. Matsuda, and R. Prozorov, Interplane resistivity of isovalent doped BaFe2(As1-xPx)2, Physical Review B - Condensed Matter and Materials Physics 8710.1103/PhysRevB.87.104506 (2013).
  • Sun et al. [2021] Y. Sun, Y. Pan, N. Zhou, X. Xing, Z. Shi, J. Wang, Z. Zhu, A. Sugimoto, T. Ekino, T. Tamegai, and H. Kitano, Comparative study of superconducting and normal-state anisotropy in Fe1+yTe0.6Se0.4 superconductors with controlled amounts of interstitial excess Fe, Physical Review B 10310.1103/PhysRevB.103.224506 (2021).
  • Sun et al. [2020] Y. Sun, H. Ohnuma, S. ya Ayukawa, T. Noji, Y. Koike, T. Tamegai, and H. Kitano, Achieving the depairing limit along the cc axis in Fe1+yTe1-xSex single crystals, Physical Review B 10110.1103/PhysRevB.101.134516 (2020).
  • Lukichev and Kupryanov [1980] M. Y. Lukichev and V. F. Kupryanov, Temperature dependence of pair-breaking current in superconductors, Fiz. Nizk. Temp 6, 445 (1980), [Sov. J. Low Temp. Phys. 6, 210 (1980)].
  • Kasahara et al. [2010] S. Kasahara, T. Shibauchi, K. Hashimoto, K. Ikada, S. Tonegawa, R. Okazaki, H. Shishido, H. Ikeda, H. Takeya, K. Hirata, T. Terashima, and Y. Matsuda, Evolution from non-Fermi- to Fermi-liquid transport via isovalent doping in BaFe2(As1-xPx)2 superconductors, Physical Review B - Condensed Matter and Materials Physics 8110.1103/PhysRevB.81.184519 (2010).
  • Ishida et al. [2017] S. Ishida, D. Song, H. Ogino, A. Iyo, H. Eisaki, M. Nakajima, J. I. Shimoyama, and M. Eisterer, Doping-dependent critical current properties in K, Co, and P-doped BaFe2As2 single crystals, Physical Review B 9510.1103/PhysRevB.95.014517 (2017).
  • Park et al. [2020] A. Park, I. Veshchunov, A. Mine, S. Pyon, T. Tamegai, and H. Kitamura, Effects of 6 Mev proton irradiation on the vortex ensemble in BaFe2(As0.67P0.33)2 revealed through magnetization measurements and real-space vortex imaging, Physical Review B 10110.1103/PhysRevB.101.224507 (2020).
  • [12] See Supplemental Material for the details of experiments and analyses.The Supplemental Material also contains Refs. [13, 14].
  • Ota et al. [2009] K. Ota, K. Hamada, R. Takemura, M. Ohmaki, T. MacHi, K. Tanabe, M. Suzuki, A. Maeda, and H. Kitano, Comparative study of macroscopic quantum tunneling in Bi2Sr2CaCu2Oy intrinsic Josephson junctions with different device structures, Physical Review B - Condensed Matter and Materials Physics 7910.1103/PhysRevB.79.134505 (2009).
  • Kakizaki et al. [2017] Y. Kakizaki, J. Koyama, A. Yamaguchi, S. Umegai, S. Y. Ayukawa, and H. Kitano, Transmission electron microscopy study of focused ion beam damage in small intrinsic Josephson junctions of single crystalline Bi2Sr2CaCu2Oy, Japanese Journal of Applied Physics 5610.7567/JJAP.56.043101 (2017).
  • [15] Note that P atoms have the smallest mass among the compositional atoms in BaFe2(As1-xPx)2. We consider that P atoms is more easily etched during the FIB microfabrications than other atoms included in BaFe2(As1-xPx)2.
  • Prozorov and Kogan [2011] R. Prozorov and V. G. Kogan, London penetration depth in iron-based superconductors, Reports on Progress in Physics 7410.1088/0034-4885/74/12/124505 (2011).
  • Nawaz et al. [2013] S. Nawaz, R. Arpaia, F. Lombardi, and T. Bauch, Microwave response of superconducting YBa2Cu3O7-δ nanowire bridges sustaining the critical depairing current: Evidence of josephson-like behavior, Physical Review Letters 11010.1103/PhysRevLett.110.167004 (2013).
  • Clem and Kogan [2012] J. R. Clem and V. G. Kogan, Kinetic impedance and depairing in thin and narrow superconducting films, Physical Review B - Condensed Matter and Materials Physics 8610.1103/PhysRevB.86.174521 (2012).
  • Li et al. [2013] J. Li, J. Yuan, Y. H. Yuan, J. Y. Ge, M. Y. Li, H. L. Feng, P. J. Pereira, A. Ishii, T. Hatano, A. V. Silhanek, L. F. Chibotaru, J. Vanacken, K. Yamaura, H. B. Wang, E. Takayama-Muromachi, and V. V. Moshchalkov, Direct observation of the depairing current density in single-crystalline Ba0.5K0.5Fe2As2 microbridge with nanoscale thickness, Applied Physics Letters 10310.1063/1.4818127 (2013).
  • Skocpol et al. [1974] W. J. Skocpol, M. R. Beasley, and M. Tinkham, Phase-Slip Centers and Nonequilibrium Processes in Superconducting Tin Microbridges, Journal of Low Temperature Physics 16 (1974).
  • [21] R. Tidecks, Current-Induced Nonequilibrium Phenomena in Quasi-One-Dimensional Superconductors, Springer Berlin Heidelberg (1990) 10.1007/BFb0048849.
  • Sivakov et al. [2003] A. G. Sivakov, A. M. Glukhov, A. N. Omelyanchouk, Y. Koval, P. Müller, and A. V. Ustinov, Josephson behavior of phase-slip lines in wide superconducting strips, Physical Review Letters 9110.1103/PhysRevLett.91.267001 (2003).
  • Paradiso et al. [2019] N. Paradiso, A. T. Nguyen, K. E. Kloss, and C. Strunk, Phase slip lines in superconducting few-layer NbSe2 crystals, 2D Materials 610.1088/2053-1583/ab0bcc (2019).
  • Andronov et al. [1993] A. Andronov, I. Gordion, V. Kurin, I. Nefedov, and I. Shereshevsky, Kinematic vortices and phase slip lines in the dynamics of the resistive state of narrow superconductive thin film channels, Physica C 213, 193 (1993).
  • Berdiyorov et al. [2009] G. R. Berdiyorov, M. V. Miloševic̀, and F. M. Peeters, Kinematic vortex-antivortex lines in strongly driven superconducting stripes, Physical Review B - Condensed Matter and Materials Physics 7910.1103/PhysRevB.79.184506 (2009).
  • Kramer and Baratoff [1977] L. Kramer and A. Baratoff, Lossless and Dissipative Current-Carrying States in Quasi-One-Dimensional Superconductors, Physical Review Letters 38, 518 (1977).
  • Kramer and Watts-Tobin [1978] L. Kramer and R. J. Watts-Tobin, Theory of Dissipative Current-Carrying States in Superconducting Filaments, Physical Review Letters 40, 1041 (1978).
  • Rusanov et al. [2004] A. Y. Rusanov, M. B. Hesselberth, and J. Aarts, Depairing currents in superconducting films of Nb and amorphous MoGe, Physical Review B - Condensed Matter and Materials Physics 7010.1103/PhysRevB.70.024510 (2004).
  • Clem and Berggren [2011] J. R. Clem and K. K. Berggren, Geometry-dependent critical currents in superconducting nanocircuits, Physical Review B - Condensed Matter and Materials Physics 8410.1103/PhysRevB.84.174510 (2011).
  • Pan et al. [2024] Y. Pan, Y. Sun, N. Zhou, X. Yi, Q. Hou, J. Wang, Z. Zhu, H. Mitamura, M. Tokunaga, and Z. Shi, Novel anisotropy of upper critical fields in Fe1+yTe0.6Se0.4, Journal of Alloys and Compounds 97610.1016/j.jallcom.2023.173262 (2024).
  • Kubo [2020] T. Kubo, Superfluid flow in disordered superconductors with Dynes pair-breaking scattering: Depairing current, kinetic inductance, and superheating field, Physical Review Research 210.1103/PhysRevResearch.2.033203 (2020).
  • Baranov et al. [2013] V. V. Baranov, A. G. Balanov, and V. V. Kabanov, Dynamics of resistive state in thin superconducting channels, Physical Review B - Condensed Matter and Materials Physics 8710.1103/PhysRevB.87.174516 (2013).
  • Shibauchi et al. [2014] T. Shibauchi, A. Carrington, and Y. Matsuda, A quantum critical point lying beneath the superconducting dome in iron pnictides, Annual Review of Condensed Matter Physics 5, 113 (2014).
  • Shishido et al. [2010] H. Shishido, A. F. Bangura, A. I. Coldea, S. Tonegawa, K. Hashimoto, S. Kasahara, P. M. Rourke, H. Ikeda, T. Terashima, R. Settai, Y. Onuki, D. Vignolles, C. Proust, B. Vignolle, A. McCollam, Y. Matsuda, T. Shibauchi, and A. Carrington, Evolution of the fermi surface of BaFe2(As1-xPx)2 on entering the superconducting dome, Physical Review Letters 10410.1103/PhysRevLett.104.057008 (2010).
  • Nakai et al. [2010] Y. Nakai, T. Iye, S. Kitagawa, K. Ishida, H. Ikeda, S. Kasahara, H. Shishido, T. Shibauchi, Y. Matsuda, and T. Terashima, Unconventional superconductivity and antiferromagnetic quantum critical behavior in the isovalent-doped BaFe2(As1-xPx)2, Physical Review Letters 10510.1103/PhysRevLett.105.107003 (2010).
  • Walmsley et al. [2013] P. Walmsley, C. Putzke, L. Malone, I. Guillamón, D. Vignolles, C. Proust, S. Badoux, A. I. Coldea, M. D. Watson, S. Kasahara, Y. Mizukami, T. Shibauchi, Y. Matsuda, and A. Carrington, Quasiparticle mass enhancement close to the quantum critical point in BaFe2(As1-xPx)2, Physical Review Letters 11010.1103/PhysRevLett.110.257002 (2013).
  • Hashimoto et al. [2012] K. Hashimoto, K. Cho, T. Shibauchi, S. Kasahara, Y. Mizukami, R. Katsumata, Y. Tsuruhara, T. Terashima, H. Ikeda, M. A. Tanatar, H. Kitano, N. Salovich, R. W. Giannetta, P. Walmsley, A. Carrington, R. Prozorov, and Y. Matsuda, A sharp peak of the zero-temperature penetration depth at optimal composition in BaFe2(As1-xPx)2Science 336, 1554 (2012).
  • Joshi et al. [2020] K. R. Joshi, N. M. Nusran, M. A. Tanatar, K. Cho, S. L. Bud’Ko, P. C. Canfield, R. M. Fernandes, A. Levchenko, and R. Prozorov, Quantum phase transition inside the superconducting dome of Ba(Fe1-xCox)2As2 from diamond-based optical magnetometry, New Journal of Physics 2210.1088/1367-2630/ab85a9 (2020).
Refer to caption
Figure 1: (a) Temperature dependence of the resistance (R1R_{1}) for the microbridge in the abab-plane. Upper and lower insets are a schematic viewgraph and a scanning ion microscope (SIM) image of the microbridge, respectively. Here, ww is a width of the microbridge. (b) Temperature dependence of the resistance (R2R_{2}) for the nanobridge with two additional slits along the cc axis. Upper and lower insets are a schematic viewgraph and a SIM image, respectively. Here, hh and ll are the distance between the two slits and the length of the cc-axis nanobridge, respectively.
Refer to caption
Figure 2: (a) Temperature dependence of the resistivity in the abab-plane (black circles) and along the cc axis (red circles), which were obtained from R1R_{1} and R2R_{2}. (b) Temperature dependence of the normal-state resistivity anisotropy.
Refer to caption
Figure 3: (a) The current-voltage (II-VV) characteristics for the cc-axis nanobridge (ij12), measured at several temperatures below TcT_{c}. (b) The II-VV curves below 14 K. Solid symbols are the same data as those in (a), while open symbols represent the 2nd II-VV measurements. (c) Temperature dependence of the cc-axis critical current density, which is determined from the 1st measurements by using a threshold voltage at 10 μ\muV (black circles) or 50 μ\muV (red circles). (d) Reduced temperature dependence of the cc-axis critical current density, normalized by a value of jcGL(0)j_{c}^{\rm GL}(0). Black and orange lines are the calculation of the GL and KL theories, respectively.
Refer to caption
Figure 4: (a) Temperature dependence of the cc-axis critical current density for all samples. (b) Reduced temperature dependence of the cc-axis critical current density, normalized by a value of jcGL(0)j_{c}^{\rm GL}(0). The reduced temperature is obtained by using TcJT_{c}^{J}. Black and orange lines are the calculation of the GL and KL theories, respectively.