Skip to main content

BRIEF RESEARCH REPORT article

Front. Phys., 20 April 2022
Sec. Condensed Matter Physics
Volume 10 - 2022 | https://doi.org/10.3389/fphy.2022.853717

Elastoresistivity of Heavily Hole-Doped 122 Iron Pnictide Superconductors

www.frontiersin.orgXiaochen Hong1,2* www.frontiersin.orgSteffen Sykora2,3* www.frontiersin.orgFederico Caglieris4,2,5* www.frontiersin.orgMahdi Behnami2 www.frontiersin.orgIgor Morozov2,6 www.frontiersin.orgSaicharan Aswartham2 www.frontiersin.orgVadim Grinenko2,7,8 www.frontiersin.orgKunihiro Kihou9 www.frontiersin.orgChul-Ho Lee9 www.frontiersin.orgBernd Büchner2,8 www.frontiersin.orgChristian Hess1,2*
  • 1Fakultät für Mathematik und Naturwissenschaften, Bergische Universität Wuppertal, Wuppertal, Germany
  • 2Leibniz-Institute for Solid State and Materials Research (IFW-Dresden), Dresden, Germany
  • 3Institute for Theoretical Physics and Würzburg-Dresden Cluster of Excellence ct.qmat, Technische Universität Dresden, Dresden, Germany
  • 4Dipartimento di Fisica, University of Genoa, Genoa, Italy
  • 5Consiglio Nazionale Delle Ricerche (CNR)-SPIN, Genova, Italy
  • 6Department of Chemistry, Lomonosov Moscow State University, Moscow, Russia
  • 7Tsung-Dao Lee Institute, Shanghai Jiao Tong University, Shanghai, China
  • 8Institute of Solid State and Materials Physics and Würzburg-Dresden Cluster of Excellence ct.qmat, Technische Universität Dresden, Dresden, Germany
  • 9National Institute of Advanced Industrial Science and Technology (AIST), Tsukuba, Japan

Nematicity in heavily hole-doped iron pnictide superconductors remains controversial. Sizeable nematic fluctuations and even nematic orders far from magnetic instability were declared in RbFe2As2 and its sister compounds. Here, we report a systematic elastoresistance study of a series of isovalent- and electron-doped KFe2As2 crystals. We found divergent elastoresistance on cooling for all the crystals along their [110] direction. The amplitude of elastoresistivity diverges if K is substituted with larger ions or if the system is driven toward a Lifshitz transition. However, we conclude that none of them necessarily indicates an independent nematic critical point. Instead, the increased nematicity can be associated with another electronic criticality. In particular, we propose a mechanism for how elastoresistivity is enhanced at a Lifshitz transition.

1 Introduction

The “122” family, an abbreviation coined for BaFe2As2 and its substituted sister compounds, played a central role in the study of iron-based superconductors [1]. Those tetragonal ThCr2Si2-type structured compounds have the advantage that sizeable single crystals with continuous tunable doping can be prepared in a wide range, which is a crucial merit for the systematic investigation of various ordered states. Within the extended phase diagram of 122 compounds, the heavily hole-doped region, including the end-members K/Rb/CsFe2As2,, is of particular interest. The superconducting transition temperature Tc of Ba1−xKxFe2As2 peaks at optimal doping x = 0.4 and continuously decreases toward the overdoped (larger x) region. Tc remains finite in the end-member x = 1, while a change of the Fermi surface topology (Lifshitz transition) exists around x = 0.6 ∼ 0.8 [2]. Although the Tc vs. x trend seems to be smooth across the Lifshitz transition, there are quite a lot of things happening here. Vanishing electron pockets for x > 0.8 destroy the basis of the interpocket scattering induced-S± pairing symmetry which is generally believed as a feature of most iron-based superconductors. As a result, a change in the superconducting gap structure across the Lifshitz transition was observed experimentally [24]. Comparable pairing strength at the transition can foster a complex pairing state that breaks the time-reversal symmetry. Such an exotic state was also demonstrated to exist around the Lifshitz transition [5, 6]. Very recently, a so-called “Z2 metal state” above Tc at the Lifshitz transition has been unveiled, with an astonishing feature of spontaneous Nernst effect [7].

Electronic nematicity, a strongly correlated electronic state of electrons breaking the underlying rotational symmetry of their lattice but preserving translation symmetry, has been a wave of research in unconventional superconductors, particularly in iron-based superconductors [8, 9]. Consistent experimental efforts have identified nematicity in all the different iron-based superconductor families [1015], accompanied by theoretical proposals of the intimate relationship between nematicity and superconducting pairing [1619]. However, according to the previous background, we should not simply extend what is known in the under- and optimal-doped 122s to the over-doped region. Whether nematicity exists and how it develops in this region needs independent censoring.

Indeed, nematicity in the heavily hole-doped 122 turns out to be more elusive. Heavily hole-doped 122s stand out as a featured series because of their peculiar Fermi surface topology, isostructural phase transition, and possible novel pairing symmetries [2023]. Nematically ordered states were suggested by nuclear magnetic resonance spectroscopy and scanning tunneling microscopy on CsFe2As2 and RbFe2As2, and they were found to develop in different wave vectors other than the underdoped 122s [24, 25]. Such a nematic state far away from magnetic ordering challenges the prevailing idea that nematicity is some kind of vestigial order of magnetism [26]. An elastoresistance study further claims that a tantalizing isotropic (or XY-) nematicity is realized in the crossover region from dominating [100] nematicity in RbFe2As2 to [110] nematicity at the optimal doping [27]. However, many works pointed out that elastoresistance in K/Rb/CsFe2As2 is actually contributed by the symmetric A1g channel, having little to do with the B1g or B2g channels which are related to nematicity [28, 29]. Overall, the debate is still on for this topic.

In this brief report, we will not touch upon the nature of the possible nematicity of K/Rb/CsFe2As2. Instead, we confirm phenomenologically the existence of elastoresistance (χer) in K/Rb/CsFe2As2 and find that its amplitude diverges exponentially with growing substituted ion size. Moreover, we present χer data on a series of Ba1−xKxFe2As2 crystals crossing the Lifshitz transition. We observe, unexpectedly, a clear enhancement of χer from both sides of the Lifshitz point. Although a presumptive nematic quantum critical point (QCP) might be of relevance, here we propose a rather more conventional explanation based on a small Fermi pocket effect. Our results add a novel phenomenon to the Lifshitz transition of the Ba1−xKxFe2As2 system and highlight another contributing factor of elastoresistance which has been almost ignored so far.

2 Experimental Details

Single crystals of heavily hole-doped Ba1−xKxFe2As2 were grown by the self-flux method [3032]. The actual doping level x was determined by considering their structural parameters and Tc values. Elastoresistance measurements were performed as described in Ref. s [1012]. Thin stripe-shaped samples were glued on the surface of piezo actuators. The strain gauge were glued on the other side of the piezo actuators to monitor the real strain generated. In most cases, the samples were mounted to let the electric current flow along the polar direction of the piezo actuators (Rxx), along which direction the strain was measured by the gauge. For one sample (x = 0.68), an additional crystal was mounted at 90° rotated according to the polar direction (Ryy). More details are described in Section 3.3. The sample resistance was collected with a combination of a high-precision current source and a nanovoltage meter. Because of the very large RRR (R300K/R0) values of the samples, special care was taken to avoid a temperature drift effect, and the electric current was set in an alternating positive/negative manner to avoid artifact.

We point out that noisy and irreproducible elastoresistance results can be acquired if DuPont 4922N silver paint is used for making contacts to the samples. On the other hand, samples contacted with EPO-TEK H20E epoxy or directly tin-soldering gave perfectly overlapping results. Given that DuPont 4922N silver paint is widely used for transport measurements and is indeed suitable for elastoresistance experiments of other materials (for example) the LaFe1−xCoxAsO series [12], we have no idea why it does not work for heavily hole-doped Ba1−xKxFe2As2 crystals. In this work, the presented data were collected by using the H20E epoxy. To avoid sample degradation, the epoxy was cured inside an Ar-glove box. A similar silver paint contact problem of K/Rb/CsFe2As2 crystals was also noticed by another group [29].

3 Results and Discussions

3.1 Elastoresistance Measurement

The elastoresistance measured along the [110] direction of the KFe2As2 single crystal is shown in Figure 1. The sample resistance closely followed the strain change of the piezo actuator when the voltage across the piezo actuator is tuned. As presented in Figure 1B, the relationship between resistance change (ΔR/R) and strain (ΔL/L) is linear. This fact ensures that our experiments were performed in the small strain limit. In such a case, the elastoresistance χer, defined as the ratio between ΔR/R and the strain, acts as a measurement of the nematic susceptibility [10]. It is worthwhile to note that χer in KFe2As2 is positive (sample under tension yields higher resistance), consistent with the previous reports [27, 28] and opposite to that of BaFe2As2 [10]. It is to be noted that sign reversal of the elastoresistance was reported to occur in the underdoped region [33].

FIGURE 1
www.frontiersin.org

FIGURE 1. Representative example of elastoresistance under strain for KFe2As2. (A) Resistance and strain change according to the voltage applied across the piezo actuator at a fixed temperature T = 50 K. The strain was applied along the [110] direction. (B) Change of resistance ΔR/R as a function of strain ΔL/L at several temperatures.

3.2 Elastoresistance of K/Rb/CsFe2As2

We start by showing our χer(T) data measured along the [110] direction (χ[110]er) for a set of (K/Rb/Cs)Fe2As2 crystals. As shown in Figure 2, all the χ[110]er(T) curves follow a divergent behavior over the whole temperature range. A Curie–Weiss (CW) fit

χer=χ0+λ/a0TTnem(1)

can record the data. A slight deviation can be discriminated at low temperature, which is typical for elastoresistance data and is understood as a disorder effect [13]. It is to be noted that the amplitude of the elastoresistance grows substantially from KFe2As2 to CsFe2As2, nearly 5-fold at 30 K. The extracted parameters from the CW fit are shown in Figures 2E,F. While the amplitude term shows a diverging trend, the Tnem of all four samples is of a very small negative value, which practically remains unchanged if experimental and fit uncertainties are taken into account, which is at odds with a possible nematic criticality in this isovalent-doping direction. The enhanced χ[110]er might be a result of a presumptive QCP of an unknown kind or a coherence–incoherence crossover [3436]. These cannot be discriminated by our technique and thus are beyond the scope of this report.

FIGURE 2
www.frontiersin.org

FIGURE 2. Temperature dependence of elastoresistance χ[110]er for (A) KFe2As2, (B) RbFe2As2, (C) K0.046Cs0.095Fe2As2, and (D) CsFe2As2. Solid lines are CW-fit of the data (see text). The fit parameters Tnem and λ/a0 are summarized in (E,F), respectively. The dotted line in (F) is a guide to the eye.

3.3 Elastoresistance of Overdoped Ba1−xKxFe2As2

Next, we present a set of χer(T) data of five overdoped Ba1−xKxFe2As2 (0.55 ≤ x ≤ 1) across the Lifshitz point. The elastoresistance, measured only for the Rxx direction, as has been performed regularly in many reports [10, 12, 14], has been argued to be inconclusive for the end members (K/Rb/Cs)Fe2As2, as a result of the dominating A1g contribution, instead of a B2g (or B1g) component which is related to nematicity [28, 29]. However, such complexities are ruled out by taking Ryy into account for calculating χer(T) for one representative example x = 0.68 (Figure 3B). The χer(T) curves calculated by the two different methods match well.

FIGURE 3
www.frontiersin.org

FIGURE 3. Doping evolution of the elastoresistance in overdoped Ba1−xKxFe2As2. χer measured along the [110] direction is presented in the upper panels for Ba1−xKxFe2As2 single crystals with (A) x = 0.55, (B) x = 0.68, (C) x = 0.78, (D) x = 0.86, and (E) x = 0.97. The red dashed lines are CW-fit to the data. χer was also measured along the [100] direction for three of the samples. The data are presented in the lower panels of (A) and (C,D). In panel (B), χer extracted by using both (ΔR/R)xx and (ΔR/R)yy (filled light blue circles) and (ΔR/R)xx (open blue squares) shows indistinguishable results. A doping dependence of the fit parameters is displayed in (F) λ/a0 and (G) Tnem of the [110] χ[110]er data. The thick lines are a guide to the eye. A Lifshitz transition around x = 0.8 is highlighted by the green bar. Hole pockets around the M point of the Brillouin zone transform into electron lobes across the Lifshitz point [2]. Tnem at x = 0.4 is extracted from Ref. [13].

After checking the potential A1g contribution to χer for a doping level close to the Lifshitz transition, we turn now to the data. As shown in Figure 3, the χ[110]er(T) curves of Ba1−xKxFe2As2 also follow a CW-like feature. One can see a clear dip at around 50 K in Figure 3A for the x = 0.55 sample. In some reports [27], such a feature was taken as a signal for a nematic order. Since no other ordering transition (structural, magnetic, and so on) has been ever reported in this doping range, we refrain from claiming an incipient nematic order solely based on such a feature. This might be equally well-explained by different origins. However, we also cannot exclude its possibility.

On the other hand, we measured χer along the [100] direction (χ[100]er) for several samples. They are shown in the shaded panels of Figure 3. All of them are negative and small in amplitude, and none of them shows a CW-like feature. Such an observation is incompatible with the existence of B1g nematicity in the heavily doped Ba1−xKxFe2As2 series, which is in sharp contrast to what is reported for the closely related Ba1−xRbxFe2As2 series [27, 37, 38]. As a result, the so-called XY-nematicity is clearly ruled out in the Ba1−xKxFe2As2 series.

One remarkable feature, however, that can be safely concluded is that the amplitude of χ[110]er has a clear tendency to peak around x = 0.8, close to the Lifshitz transition. This becomes clearer in Figure 3F, where the CW fit parameters are plotted against the doping level. The question is why χ[110]er is increased at the Lifshitz transition? A nematic QCP is a potential explanation. However, as Figure 3G shows, Tnem drops from 45 K of the x = 0.4 (optimal-doped) to 0 K at the Lifshitz transition. Further doping does not drive Tnem to the more negative side within our experimental resolution. This is not typical QCP behavior. Moreover, since Tc across the Lifshitz transition is quite smooth, it seems not to be boosted by pertinent potential nematic fluctuations. Furthermore, three-point bending experiments did not show any anomaly in this doping range [39]. All these facts seem to be incompatible with the more understood nematic QCP in the electron-doped side [40]. Hence, if it is a nematic QCP, novel mechanisms need to be invoked. This motivated us to seek for alternative explanations for enhanced χer in heavily doped Ba1−xKxFe2As2. In the following section, we propose a conventional argument based on the small pocket effect, exempting from invoking a QCP to exist at the Lifshitz transition.

3.4 Theory for Enhanced Elastoresistance at the Lifshitz Transition

To study the effect of a Lifshitz transition to elastoresistance, we have calculated this quantity based on a minimal model of iron-based superconductors [41] with a very small Fermi surface. The corresponding dispersion which was used is shown in Figure 4A for the normal state along a cut (π, ky). We considered the two orbital models in Ref. [41] with the same hopping matrix elements but having set the nematic interaction equal to a very small value. Thus, the nematic interaction accounts here only for the temperature dependence of the susceptibility according to the Curie–Weiss law. Moreover, we introduced a very small lattice distortion in the x direction which is coupled with the electron system. Using the first-order perturbation theory with respect to this coupling (linear response), we then calculated the elastoresistivity. We have considered two different cases of the coupling between distortion and electrons (strength g): (i) The conventional coupling with the local electron density (electron–phonon coupling), where we denote the corresponding response with χph. (ii) A direct coupling of the distortion with the hopping matrix element tx in the x direction. The corresponding response is denoted by χt.

χtglimΔtx0σxx1tx+Δtxσyy1tx+ΔtxΔtxχph1Nkgω+εkεk+q2fkfk+qεk+qεk.(2)

Here, tx is the hopping matrix element in the x direction and σxx, σyy are the conductivities in the x and y directions [41]. The phonon energy ω in χpher is in general renormalized by the coupling to the electrons and becomes soft for a particular mode if the system is near a structural phase transition[42, 43]. The electron dispersion ɛk considered here is shown in Figure 4A. The function fk is the Fermi distribution with respect to ɛk. Thus, for most systems investigated, this term dominates χer. It is to be noted how magnetic fluctuations impact the phonons and the nematicity has been investigated [9, 16].

FIGURE 4
www.frontiersin.org

FIGURE 4. Theoretical consideration of the elastoresisitivity. (A) Dispersion of the hole-like band leading to a very small hole pocket around the point (π, π) when the chemical potential μsmall (upper dotted line) is placed near the Lifshitz point. The range of momentum vectors contributing to the elastoresisitivity arising from the Fermi function is schematically shown by the dashed lines. A much larger hole pocket is indicated by a lower value of the chemical potential μlarge (lower dotted line). (B) Calculated parts of the elastoresistivity according to Eq. 3. Around the Lifshitz point μsmall the first-order part χter dominates the second-order part which is, however, the most important contribution for μ away from the Lifshitz point.

Figure 4B shows the two parts χter and χpher which were calculated separately as a function of the chemical potential μ to simulate different doping values. To compare with Figure 3F, we extracted the temperature behavior according to Eq. 2 and plotted the calculated value λ/a0 in energy units of ty. It is seen that the first-order part χter dominates the second-order part only in the narrow range of μ, where the Fermi surface around (π, π) becomes very small. Thus, only when the system has very small Fermi surfaces, as in the case of Ba1−xKxFe2As2 at the Lifshitz transition, the term χter becomes important. However, it is also seen that if the chemical potential is chosen away from the Lifshitz point corresponding to a proper doping, the second-order part χpher is mostly important as expected.

The enhancement of χter in the presence of a very small Fermi surface can be explained by the existence of low-energy excitations in a relatively wide range of momentum vectors. Since the conductivities are proportional to Fermi distribution functions fk as follows

σiikεkki2fk1fk,(3)

we find that at low temperature, if the Fermi surface is small, the momentum range k, where fk(1 − fk) is non-zero, is much larger because of the tendency of the band to rapidly change the Fermi surface topology near the Lifshitz transition (compare the red dashed lines in Figure 4A) than for a usual Fermi surface.

4 Conclusion

To summarize, we reported that a CW-like χer(T) is observed for all kinds of heavily hole-doped 122s. There is an unexpected enhancement of the elastoresistance around the Lifshitz transition. We explained it as a small Fermi pocket effect on the nematicity. We expect that our explanation of an alternative contribution to the enhanced elastoresistance other than a nematic QCP will be considered in other systems, in particular for those with small Fermi pockets.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors, without undue reservation.

Author Contributions

IM, SA, VG, KK, and C-HL prepared the samples. XH, FC, and MB performed the experiments. SS proposed the theoretical model. CH and BB supervised the study. XH, SS, FC, and CH analyzed the data and wrote the manuscript with input from all authors.

Funding

This work has been supported by the Deutsche Forschungsgemeinschaft (DFG) through SFB 1143 (Project No. 247310070), through the Research Projects CA 1931/1-1 (FC) and SA 523/4-1 (SA). SS acknowledges funding by the Deutsche Forschungs gemeinschaft via the Emmy Noether Program ME4844/1-1 (project id 327807255). This project has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 Research and Innovation Program (grant agreement No. 647276-MARS-ERC-2014-CoG).

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Acknowledgments

We would like to thank Anna Böhmer, Ian Fisher, Suguru Hosoi, Rüdiger Klingeler, Christoph Meingast, Jörg Schmalian, Christoph Wuttke, Paul Wiecki, and Liran Wang for helpful discussions. We would like to thank Christian Blum and Silvia Seiro for their technical support.

References

1. Chen X, Dai P, Feng D, Xiang T, Zhang FC. Iron-based High Transition Temperature Superconductors. Natl.Sci.Rev. (2014) 1:371–95. doi:10.1093/nsr/nwu007

CrossRef Full Text | Google Scholar

2. Xu N, Richard P, Shi X, van Roekeghem A, Qian T, Razzoli E. Possible Nodal Superconducting gap and Lifshitz Transition in Heavily Hole-Doped Ba0. 1K0.9Fe2As2 Phys.Rev.B (2013) 88:220508. doi:10.1103/PhysRevB.88.220508

CrossRef Full Text | Google Scholar

3. Malaeb W, Shimojima T, Ishida Y, Okazaki K, Ota Y, Ohgushi K, et al. Abrupt Change in the Energy gap of Superconducting Ba1−xKxFe2As2 Single Crystals with Hole Doping. Phys.Rev.B (2012) 86:165117. doi:10.1103/PhysRevB.86.165117

CrossRef Full Text | Google Scholar

4. Hong XC, Wang AF, Zhang Z, Pan J, He LP, Luo XG, et al. Doping Evolution of the Superconducting Gap Structure in Heavily Hole-Doped Ba1−xKxFe2As2: a Heat Transport Study. Chin.Phys.Lett. (2015) 32:127403. doi:10.1088/0256-307X/32/12/127403

CrossRef Full Text | Google Scholar

5. Grinenko V, Materne P, Sarkar R, Luetkens H, Kihou K, Lee CH, et al. Superconductivity with Broken Time-Reversal Symmetry in Ion-Irradiated Ba0.27K0.73Fe2As2 Single Crystals. Phys.Rev.B (2017) 95:214511. doi:10.1103/PhysRevB.95.214511

CrossRef Full Text | Google Scholar

6. Grinenko V, Sarkar R, Kihou K, Lee CH, Morozov I, Aswartham S, et al. Superconductivity with Broken Time-Reversal Symmetry inside a Superconducting S-Wave State. Nat.Phys. (2020) 16:789. doi:10.1038/s41567-020-0886-9

CrossRef Full Text | Google Scholar

7. Grinenko V, Weston D, Caglieris F, Wuttke C, Hess C, Gottschall T, et al. State with Spontaneously Broken Time-Reversal Symmetry above the Superconducting Phase Transition. Nat.Phys. (2021) 17:1254. doi:10.1038/s41567-021-01350-9

CrossRef Full Text | Google Scholar

8. Fisher IR, Degiorgi L, Shen ZX. In-plane Electronic Anisotropy of Underdoped ’122’ Fe-Arsenide Superconductors Revealed by Measurements of Detwinned Single Crystals. Rep.Prog.Phys. (2011) 74:124506. doi:10.1088/0034-4885/74/12/124506

CrossRef Full Text | Google Scholar

9. Fernandes RM, Chubukov AV, Schmalian J. What Drives Nematic Order in Iron-Based Superconductors? Nat.Phys. (2014) 10:97. doi:10.1038/nphys2877

CrossRef Full Text | Google Scholar

10. Chu JH, Kuo HH, Aanlytis JG, Fisher IR. Divergent Nematic Susceptibility in an Iron Arsenide Superconductor. Science (2012) 337:710. doi:10.1126/science.1221713

PubMed Abstract | CrossRef Full Text | Google Scholar

11. Hosoi S, Matsuura K, Ishida K, Wang H, Mizukami Y, Watashige T, et al. Nematic Quantum Critical point without Magnetism in FeSe1−xSx Superconductors. Proc.Natl.Acad.Sci.U.S.A. (2016) 113:8139. doi:10.1073/pnas.1605806113

PubMed Abstract | CrossRef Full Text | Google Scholar

12. Hong XC, Caglieris F, Kappenberger R, Wurmehl S, Aswartham S, Scaravaggi F, et al. Evolution of the Nematic Susceptibility in LaFe1−xCoxAsO. Phys.Rev.Lett. (2020) 125:067001. doi:10.1103/PhysRevLett.125.067001

PubMed Abstract | CrossRef Full Text | Google Scholar

13. Kuo HH, Chu JH, Palmstrom JC, Kivelson SA, Fisher IR. Ubiquitous Signatures of Nematic Quantum Criticality in Optimally Doped Fe-Based Superconductors. Science (2016) 52:958. doi:10.1126/science.aab0103

PubMed Abstract | CrossRef Full Text | Google Scholar

14. Gu YH, Liu ZY, Xie T, Zhang WL, Gong DL, Hu D, et al. Unified Phase Diagram for Iron-Based Superconductors. Phys.Rev.Lett. (2017) 119:157001. doi:10.1103/PhysRevLett.119.157001

PubMed Abstract | CrossRef Full Text | Google Scholar

15. Terashima T, Matsushita Y, Yamase H, Kikugawa N, Abe H, Imai M, et al. Elastoresistance Measurements on CaKFe4As4 and KCa2Fe4As4F2 with the Fe Site of C2v Symmetry. Phys.RevB (2020) 102:054511. doi:10.1103/PhysRevB.102.054511

CrossRef Full Text | Google Scholar

16. Fernandes RM, Schmalian J. Manifestations of Nematic Degrees of freedom in the Magnetic, Elastic, and Superconducting Properties of the Iron Pnictides. Supercond.Sci.Technol. (2012) 25:084005. doi:10.1088/0953-2048/25/8/084005

CrossRef Full Text | Google Scholar

17. Lederer S, Schattner Y, Berg E, Kivelson SA. Enhancement of Superconductivity Near a Nematic Quantum Critical Point. Phys.Rev.Lett. (2015) 114:097001. doi:10.1103/PhysRevLett.114.097001

PubMed Abstract | CrossRef Full Text | Google Scholar

18. Labat D, Paul I. Pairing Instability Near a Lattice-Influenced Nematic Quantum Critical point. Phys.Rev.B (2017) 96:195146. doi:10.1103/PhysRevB.96.195146

CrossRef Full Text | Google Scholar

19. Maslov DL, Chubukov AV. Fermi Liquid Near Pomeranchuk Quantum Criticality. Phys.Rev.B (2010) 81:045110. doi:10.1103/PhysRevB.81.045110

CrossRef Full Text | Google Scholar

20. Tafti FF, Juneau-Fecteau A, Delage M-E, René de Cotret S, Reid J-P, Wang AF, et al. Sudden Reversal in the Pressure Dependence of Tc in the Iron-Based Superconductor KFe2As2. Nat.Phys. (2013) 9:349. doi:10.1038/nphys2617

CrossRef Full Text | Google Scholar

21. Wang YQ, Lu PC, Wu JJ, Liu J, Wang XC, Zhao JY, et al. Phonon Density of States of Single-crystal SrFe2As2 across the Collapsed Phase Transition at High Pressure. Phys.Rev.B (2016) 94:014516. doi:10.1103/PhysRevB.94.014516

CrossRef Full Text | Google Scholar

22. Ptok A, Sternik M, Kapcia KJ, Piekarz P. Structural, Electronic, and Dynamical Properties of the Tetragonal and Collapsed Tetragonal Phases of KFe2As2. Phys.Rev.B (2019) 99:134103. doi:10.1103/PhysRevB.99.134103

CrossRef Full Text | Google Scholar

23. Ptok A, Kapcia KJ, Cichy A, Oleś AM, Piekarz P. Magnetic Lifshitz Transition and its Consequences in Multi-Band Iron-Based Superconductors. Sci.Rep. (2017) 7:41979. doi:10.1038/srep41979

PubMed Abstract | CrossRef Full Text | Google Scholar

24. Li J, Zhao D, Wu YP, Li SJ, Song DW, Zheng LX, et al. Reemergeing Electronic Nematicity in Heavily Hole-Doped Fe-Based Superconductors (2016). arXiv:1611.04694.

Google Scholar

25. Liu X, Tao R, Ren MQ, Chen W, Yao Q, Wolf T, et al. Evidence of Nematic Order and Nodal Superconducting gap along [110] Direction in RbFe2As2. Nat.Commun. (2019).

Google Scholar

26. Fernandes RM, Orth PP, Schmalian J. Intertwined Vestigial Order in Quantum Materials: Nematicity and beyond. Annu.Rev.Condens.MatterPhys. (2019) 10:133. doi:10.1146/annurev-conmatphys-031218-013200

CrossRef Full Text | Google Scholar

27. Ishida K, Tsujii M, Hosoi S, Mizukami Y, Ishida S, Iyo A, et al. Novel Electronic Nematicity in Heavily Hole-Doped Iron Pnictide Superconductors. Proc.Natl.Acad.Sci.U.S.A. (2020) 117:6424. doi:10.1073/pnas.1909172117

PubMed Abstract | CrossRef Full Text | Google Scholar

28. Wiecki P, Haghighirad AA, Weber F, Merz M, Heid R, Böhmer AE. Dominant In-Plane Symmetric Elastoresistance in CsFe2As2. Phys.Rev.Lett. (2020) 125:187001. doi:10.1103/PhysRevLett.125.187001

PubMed Abstract | CrossRef Full Text | Google Scholar

29. Wiecki P, Frachet M, Haghighirad AA, Wolf T, Meingast C, Heid R, et al. Emerging Symmetric Strain Response and Weakening Nematic Fluctuations in Strongly Hole-Doped Iron-Based Superconductors. Nat.Commun. (2021) 12:4824. doi:10.1038/s41467-021-25121-5

PubMed Abstract | CrossRef Full Text | Google Scholar

30. Aswartham S, Abdel-Hafiez M, Bombor D, Kumar M, Wolter AUB, Hess C, et al. Hole Doping in BaFe2As2: The Case of Ba1−xNaxFe2As2 Single Crystals. Phys.Rev.B (2012) 85:224520. doi:10.1103/PhysRevB.85.224520

CrossRef Full Text | Google Scholar

31. Abdel-Hafiez M, Aswartham S, Wurmehl S, Grinenko V, Hess C, Drechsler S-L, et al. Specific Heat and Upper Critical fields in KFe2As2 Single Crystals. Phys.Rev.B (2012) 85:134533. doi:10.1103/PhysRevB.85.134533

CrossRef Full Text | Google Scholar

32. Moroni M, Prando G, Aswartham S, Morozov I, Bukowski Z, Büchner B, et al. Charge and Nematic Orders in AFe2As2 (A = Rb, Cs) Superconductors. Phys.Rev.B (2019) 99:235147. doi:10.1103/PhysRevB.99.235147

CrossRef Full Text | Google Scholar

33. Blomberg EC, Tanatar MA, Fernandes RM, Mazin II, Shen B, Wen HH, et al. Sign-reversal of the In-Plane Resistivity Anisotropy in Hole-Doped Iron Pnictides. Nat.Commun. (2013) 4:1914. doi:10.1038/ncomms2933

PubMed Abstract | CrossRef Full Text | Google Scholar

34. Eilers F, Grube K, Zocco DA, Wolf T, Merz M, Schweiss P, et al. Strain-Driven Approach to Quantum Criticality in AFe2As2 with A=K, Rb, and Cs. Phys.Rev.Lett. (2016) 116:237003. doi:10.1103/PhysRevLett.116.237003

PubMed Abstract | CrossRef Full Text | Google Scholar

35. Zhang ZT, Dmytriieva D, Molatta S, Wosnitza J, Khim S, Gass S, et al. Increasing Stripe-type Fluctuations in AFe2As2 (A=K, Rb, Cs) Superconductors Probed by 75As NMR Spectroscopy. Phys.Rev.B (2018) 97:115110. doi:10.1103/PhysRevB.97.115110

CrossRef Full Text | Google Scholar

36. Hardy F, Böhmer AE, Aoki D, Burger P, Wolf T, Schweiss P, et al. Evidence of Strong Correlations and Coherence-Incoherence Crossover in the Iron Pnictide Superconductor KFe2As2. Phys.Rev.Lett. (2013) 111:027002. doi:10.1103/PhysRevLett.111.027002

PubMed Abstract | CrossRef Full Text | Google Scholar

37. Onari S, Kontani H. Origin of Diverse Nematic Orders in Fe-Based Superconductors: 45° Rotated Nematicity in AFe2As2 (A=Cs,Rb). Phys.Rev.B (2019) 100:020507. doi:10.1103/PhysRevB.100.020507

CrossRef Full Text | Google Scholar

38. Borisov V, Fernandes RM, Valentí R. Evolution from B2g Nematics to B1g Nematics in Heavily Hole-Doped Iron-Based Superconductors. Phys.Rev.Lett. (2019) 123:146402. doi:10.1103/PhysRevLett.123.146402

PubMed Abstract | CrossRef Full Text | Google Scholar

39. Böhmer AE, Burger P, Hardy F, Wolf T, Schweiss P, Fromknecht R, et al. Nematic Susceptibility of Hole-Doped and Electron-Doped BaFe2As2 Iron-Based Superconductors from Shear Modulus Measurements. Phys.Rev.Lett. (2014) 112:047001. doi:10.1103/PhysRevLett.112.047001

PubMed Abstract | CrossRef Full Text | Google Scholar

40. Fernandes RM, VanBebber LH, Bhattacharya S, Chandra P, Keppens V, Mandrus D, et al. Effects of Nematic Fluctuations on the Elastic Properties of Iron Arsenide Superconductors. Phys.Rev.Lett. (2010) 105:157003. doi:10.1103/PhysRevLett.105.157003

PubMed Abstract | CrossRef Full Text | Google Scholar

41. Wuttke C, Caglieris F, Sykora S, Steckel F, Hong X, Ran S, et al. Ubiquitous Enhancement of Nematic Fluctuations across the Phase Diagram of Iron Based Superconductors Probed by the Nernst Effect. arXiv:2202.00485.

42. Sykora S, Hübsch A, Becker KW. Dominant Particle-Hole Contributions to the Phonon Dynamics in the Spinless One-Dimensional Holstein Model. Europhys.Lett. (2006) 76:644. doi:10.1209/epl/i2006-10327-x

CrossRef Full Text | Google Scholar

43. Sykora S, Hübsch A, Becker KW. Generalized Diagonalization Scheme for many-particle Systems. Phys.Rev.B (2020) 102:165122. doi:10.1103/PhysRevB.102.165122

CrossRef Full Text | Google Scholar

Keywords: elastoresistance, nematicity, Lifshitz transition, iron-based superconductors, quantum criticality

Citation: Hong X, Sykora S, Caglieris F, Behnami M, Morozov I, Aswartham S, Grinenko V, Kihou K, Lee C-, Büchner B and Hess C (2022) Elastoresistivity of Heavily Hole-Doped 122 Iron Pnictide Superconductors. Front. Phys. 10:853717. doi: 10.3389/fphy.2022.853717

Received: 12 January 2022; Accepted: 21 March 2022;
Published: 20 April 2022.

Edited by:

Anna Böhmer, Ruhr-University Bochum, Germany

Reviewed by:

Konrad Jerzy Kapcia, Adam Mickiewicz University, Poland
Marcin Matusiak, Institute of Physics (PAN), Poland

Copyright © 2022 Hong, Sykora, Caglieris, Behnami, Morozov, Aswartham, Grinenko, Kihou, Lee, Büchner and Hess. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Xiaochen Hong, xhong@uni-wuppertal.de; Steffen Sykora, steffen.sykora@tu-dresden.de; Federico Caglieris, federico.caglieris@spin.cnr.it; Christian Hess, c.hess@uni-wuppertal.de

Download