17
PHYSICAL REVIEW B 93, 134519 (2016) Impact of uniaxial pressure on structural and magnetic phase transitions in electron-doped iron pnictides Xingye Lu, 1, 2 , * Kuo-Feng Tseng, 3, 4 T. Keller, 3, 4 Wenliang Zhang, 2 Ding Hu, 2 Yu Song, 1 Haoran Man, 1 J. T. Park, 5 Huiqian Luo, 2 Shiliang Li, 2, 6 Andriy H. Nevidomskyy, 1 and Pengcheng Dai 1 , 1 Department of Physics and Astronomy & Rice Center for Quantum Materials, Rice University, Houston, Texas 77005, USA 2 Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China 3 Max-Planck-Institut f¨ ur Festk¨ orperforschung, Heisenbergstrasse 1, D-70569 Stuttgart, Germany 4 Max Planck Society Outstation at the Forschungsneutronenquelle Heinz Maier-Leibnitz (MLZ), D-85747 Garching, Germany 5 Heinz Maier-Leibnitz Zentrum (MLZ), Technische Universit¨ at M ¨ unchen, 85748 Garching, Germany 6 Collaborative Innovation Center of Quantum Matter, Beijing, China (Received 15 July 2015; revised manuscript received 20 March 2016; published 28 April 2016) We use neutron resonance spin echo and Larmor diffraction to study the effect of uniaxial pressure on the tetragonal-to-orthorhombic structural (T s ) and antiferromagnetic (AF) phase transitions in iron pnictides BaFe 2x Ni x As 2 (x = 0,0.03,0.12), SrFe 1.97 Ni 0.03 As 2 , and BaFe 2 (As 0.7 P 0.3 ) 2 . In antiferromagnetically ordered BaFe 2x Ni x As 2 and SrFe 1.97 Ni 0.03 As 2 with T N and T s (T N T s ), a uniaxial pressure necessary to detwin the sample also increases T N , smears out the structural transition, and induces an orthorhombic lattice distortion at all temperatures. By comparing temperature and doping dependence of the pressure induced lattice parameter changes with the elastoresistance and nematic susceptibility obtained from transport and ultrasonic measurements, we conclude that the in-plane resistivity anisotropy found in the paramagnetic state of electron underdoped iron pnictides depends sensitively on the nature of the magnetic phase transition and a strong coupling between the uniaxial pressure induced lattice distortion and electronic nematic susceptibility. DOI: 10.1103/PhysRevB.93.134519 I. INTRODUCTION The parent compounds of iron pnictide superconductors such as BaFe 2 As 2 and SrFe 2 As 2 exhibit a tetragonal-to- orthorhombic structural transition at T s followed by devel- opment of collinear antiferromagnetic (AF) order along the a axis of the orthorhombic lattice below T N [left inset in Fig. 1(a) and T s T N ][16]. Upon electron doping via partially substituting Fe by Co or Ni to form BaFe 2x T x As 2 (T = Co, Ni), the nearly coupled structural and magnetic phase transitions in BaFe 2 As 2 become two separate second order phase transitions at T s and T N (T s >T N ) that decrease in temperature with increasing x [710]. On the other hand, the coupled first-order structural and magnetic phase transitions in SrFe 2 As 2 [4], while decreasing in temperature with increasing x in SrFe 2x T x As 2 , remain coupled first-order transitions leading up to superconductivity [11]. Because the structural and magnetic phase transitions in BaFe 2x T x As 2 and SrFe 2x T x As 2 occur below room temper- ature, iron pnictides in the orthorhombic AF ground state will form twin domains with AF Bragg peaks appearing at the in-plane (±1,0) and (0,±1) positions in reciprocal space [right inset in Fig. 1(a)][6]. To probe the intrinsic electronic properties of these materials, one can apply uniaxial pressure along one axis of the orthorhombic lattice to obtain single domain samples [1215]. Indeed, transport measurements on uniaxial pressure detwinned electron-doped BaFe 2x T x As 2 (T = Co, Ni) reveal in-plane resistivity anisotropy in the AF state that persists to temperatures above the zero-pressure * Present Address: Paul Scherrer Institut, Swiss Light Source, CH- 5232 Villigen PSI, Switzerland. [email protected] T N and T s [1215]. On the other hand, similar transport measurements on uniaxial pressure detwinned SrFe 2x T x As 2 (T = Co, Ni) indicate vanishingly small resistivity anisotropy at temperatures above the zero pressure coupled T N and T s [1618]. Figure 1(b) compares temperature dependence of the resistivity anisotropy [defined as ρ = (ρ b ρ a )/(ρ b + ρ a ), where ρ a and ρ b are resistivity along the a and b axis of the orthorhombic lattice, respectively] obtained under 20 MPa uniaxial pressure for BaFe 2 As 2 , BaFe 1.97 Ni 0.03 As 2 , SrFe 1.97 Ni 0.03 As 2 , and SrFe 2 As 2 . Consistent with earlier works [1218], we find that resistivity anisotropy is much larger in BaFe 2 As 2 and BaFe 1.97 Ni 0.03 As 2 at temperatures above T N . Although resistivity anisotropy in the paramagnetic state of the iron pnictides under applied uniaxial pressure suggests the presence of an electronic nematic phase that breaks the in-plane fourfold rotational symmetry (C 4 ) of the underly- ing tetragonal lattice [1924], much is unclear about the microscopic origin of the in-plane resistivity anisotropy and electronic nematic phase [2538]. Since neutron-scattering experiments reveal that uniaxial pressure necessary to detwin the sample also increases T N of the system, the observed in-plane resistivity anisotropy above the zero pressure T N and T s may arise from the increased T N and intrinsic anisotropic nature of the collinear AF phase [39,40]. Furthermore, while it is generally assumed that the uniaxial pressure for sample detwinning has negligible effect on the lattice pa- rameters of the iron pnictides [1215], the precise effect of uniaxial pressure on structural distortion of these materials is unknown. From neutron extinction effect measurements, a uniaxial pressure is suggested to push structural fluctuations related to the orthorhombic distortion to a temperature well above the zero-pressure value of T s [41], similar to the effect on the resistivity anisotropy [1215]. To understand 2469-9950/2016/93(13)/134519(17) 134519-1 ©2016 American Physical Society

Impact of uniaxial pressure on structural and magnetic phase ...sc8.iphy.ac.cn/papers/PRB_93_134519.pdfPHYSICAL REVIEW B 93, 134519 (2016) Impact of uniaxial pressure on structural

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

  • PHYSICAL REVIEW B 93, 134519 (2016)

    Impact of uniaxial pressure on structural and magnetic phasetransitions in electron-doped iron pnictides

    Xingye Lu,1,2,* Kuo-Feng Tseng,3,4 T. Keller,3,4 Wenliang Zhang,2 Ding Hu,2 Yu Song,1 Haoran Man,1 J. T. Park,5

    Huiqian Luo,2 Shiliang Li,2,6 Andriy H. Nevidomskyy,1 and Pengcheng Dai1,†1Department of Physics and Astronomy & Rice Center for Quantum Materials, Rice University, Houston, Texas 77005, USA

    2Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China3Max-Planck-Institut für Festkörperforschung, Heisenbergstrasse 1, D-70569 Stuttgart, Germany

    4Max Planck Society Outstation at the Forschungsneutronenquelle Heinz Maier-Leibnitz (MLZ), D-85747 Garching, Germany5Heinz Maier-Leibnitz Zentrum (MLZ), Technische Universität München, 85748 Garching, Germany

    6Collaborative Innovation Center of Quantum Matter, Beijing, China(Received 15 July 2015; revised manuscript received 20 March 2016; published 28 April 2016)

    We use neutron resonance spin echo and Larmor diffraction to study the effect of uniaxial pressure onthe tetragonal-to-orthorhombic structural (Ts) and antiferromagnetic (AF) phase transitions in iron pnictidesBaFe2−xNixAs2 (x = 0,0.03,0.12), SrFe1.97Ni0.03As2, and BaFe2(As0.7P0.3)2. In antiferromagnetically orderedBaFe2−xNixAs2 and SrFe1.97Ni0.03As2 with TN and Ts (TN � Ts), a uniaxial pressure necessary to detwin thesample also increases TN , smears out the structural transition, and induces an orthorhombic lattice distortion atall temperatures. By comparing temperature and doping dependence of the pressure induced lattice parameterchanges with the elastoresistance and nematic susceptibility obtained from transport and ultrasonic measurements,we conclude that the in-plane resistivity anisotropy found in the paramagnetic state of electron underdoped ironpnictides depends sensitively on the nature of the magnetic phase transition and a strong coupling between theuniaxial pressure induced lattice distortion and electronic nematic susceptibility.

    DOI: 10.1103/PhysRevB.93.134519

    I. INTRODUCTION

    The parent compounds of iron pnictide superconductorssuch as BaFe2As2 and SrFe2As2 exhibit a tetragonal-to-orthorhombic structural transition at Ts followed by devel-opment of collinear antiferromagnetic (AF) order along thea axis of the orthorhombic lattice below TN [left inset inFig. 1(a) and Ts ≈ TN ] [1–6]. Upon electron doping viapartially substituting Fe by Co or Ni to form BaFe2−xTxAs2(T = Co, Ni), the nearly coupled structural and magneticphase transitions in BaFe2As2 become two separate secondorder phase transitions at Ts and TN (Ts > TN ) that decreasein temperature with increasing x [7–10]. On the other hand, thecoupled first-order structural and magnetic phase transitions inSrFe2As2 [4], while decreasing in temperature with increasingx in SrFe2−xTxAs2, remain coupled first-order transitionsleading up to superconductivity [11].

    Because the structural and magnetic phase transitions inBaFe2−xTxAs2 and SrFe2−xTxAs2 occur below room temper-ature, iron pnictides in the orthorhombic AF ground statewill form twin domains with AF Bragg peaks appearing atthe in-plane (±1,0) and (0,±1) positions in reciprocal space[right inset in Fig. 1(a)] [6]. To probe the intrinsic electronicproperties of these materials, one can apply uniaxial pressurealong one axis of the orthorhombic lattice to obtain singledomain samples [12–15]. Indeed, transport measurements onuniaxial pressure detwinned electron-doped BaFe2−xTxAs2(T = Co, Ni) reveal in-plane resistivity anisotropy in the AFstate that persists to temperatures above the zero-pressure

    *Present Address: Paul Scherrer Institut, Swiss Light Source, CH-5232 Villigen PSI, Switzerland.

    [email protected]

    TN and Ts [12–15]. On the other hand, similar transportmeasurements on uniaxial pressure detwinned SrFe2−xTxAs2(T = Co, Ni) indicate vanishingly small resistivity anisotropyat temperatures above the zero pressure coupled TN and Ts[16–18]. Figure 1(b) compares temperature dependence of theresistivity anisotropy [defined as �ρ = (ρb − ρa)/(ρb + ρa),where ρa and ρb are resistivity along the a and b axisof the orthorhombic lattice, respectively] obtained under∼20 MPa uniaxial pressure for BaFe2As2, BaFe1.97Ni0.03As2,SrFe1.97Ni0.03As2, and SrFe2As2. Consistent with earlierworks [12–18], we find that resistivity anisotropy is muchlarger in BaFe2As2 and BaFe1.97Ni0.03As2 at temperaturesabove TN .

    Although resistivity anisotropy in the paramagnetic stateof the iron pnictides under applied uniaxial pressure suggeststhe presence of an electronic nematic phase that breaks thein-plane fourfold rotational symmetry (C4) of the underly-ing tetragonal lattice [19–24], much is unclear about themicroscopic origin of the in-plane resistivity anisotropy andelectronic nematic phase [25–38]. Since neutron-scatteringexperiments reveal that uniaxial pressure necessary to detwinthe sample also increases TN of the system, the observedin-plane resistivity anisotropy above the zero pressure TN andTs may arise from the increased TN and intrinsic anisotropicnature of the collinear AF phase [39,40]. Furthermore,while it is generally assumed that the uniaxial pressure forsample detwinning has negligible effect on the lattice pa-rameters of the iron pnictides [12–15], the precise effect ofuniaxial pressure on structural distortion of these materialsis unknown. From neutron extinction effect measurements, auniaxial pressure is suggested to push structural fluctuationsrelated to the orthorhombic distortion to a temperature wellabove the zero-pressure value of Ts [41], similar to theeffect on the resistivity anisotropy [12–15]. To understand

    2469-9950/2016/93(13)/134519(17) 134519-1 ©2016 American Physical Society

    http://dx.doi.org/10.1103/PhysRevB.93.134519

  • XINGYE LU et al. PHYSICAL REVIEW B 93, 134519 (2016)

    0

    30

    60

    90

    120

    150

    Tem

    pera

    ture

    (K)

    0.00 0.05 0.10 0.15 0.20 0.25

    AF Ort

    SCIC

    TN

    TcOrt PM Tet

    TSNematic

    a

    b

    L H

    K

    (400)

    (040)

    (101)

    (a)

    Detwinned

    x = 0.03

    x = 0.12 BaFe2-xNixAs2

    (c)

    (011)

    P

    Lb

    x = 0

    0

    1

    2

    3

    δ(P)

    - δ(

    P=0)

    (10-

    3 )

    x

    0 40 80 120 160 200 240Temperature (K)

    280

    (ρb-

    ρ a)/

    (ρb +

    ρ a)

    (b)

    BaFe1.97Ni0.03As2

    BaFe2As2

    SrFe1.97Ni0.03As2

    BaFe1.88Ni0.12As2

    P ~ 20 MPa

    P ~ 20 MPa BaFe1.97Ni0.03As2

    BaFe2As2

    SrFe2As2

    0.00

    0.05

    0.10

    0.15

    0.20

    SrFe1.97Ni0.03As2

    FIG. 1. (a) The schematic electronic phase diagram ofBaFe2−xNixAs2 with arrows marking x = 0, 0.03, and 0.12 samplesdescribed in the present study. The AF, PM, Ort, Tet, IC, SC areantiferromagnetic, paramagnetic, orthorhombic, tetragonal, incom-mensurate, and superconducting states, respectively [10]. The leftinset shows the direction of the applied uniaxial pressure (markedby the vertical arrows) and the spin arrangements of Fe in the AFordered iron pnictides, where a and b are the orthorhombic axes. Theright inset shows the corresponding reciprocal lattice. All the markedpositions have AF or nuclear Bragg peaks for a twinned sample, whilethe positions marked by open symbols have vanishing scatteringintensity for a detwinned sample. (b) Temperature dependence ofthe resistivity anisotropy for BaFe2−xNixAs2 and SrFe2−xNixAs2(x = 0,0.03) under P ≈ 20 MPa. (c) Summary of temperaturedependence of the uniaxial pressure induced lattice distortion atP ≈ 20 MPa [δ(P ≈ 20 MPa) − δ(P = 0 MPa)] for BaFe2−xNixAs2(x = 0,0.03,0.12) and SrFe1.97Ni0.03As2. The actual data for x =0.03,0.12 are normalized to 20 PMa assuming a linear relationshipbetween uniaxial pressure and δ. Uniaxial pressure induced latticedistortion vanishes rapidly below TN marked by the vertical dashedlines in (b) and (c).

    the microscopic origin of the in-plane resistivity anisotropyin the paramagnetic state [12–15], it is important to establishthe effect of a uniaxial pressure on the magnetic and structuralphase transitions of BaFe2−xTxAs2 and SrFe2−xTxAs2, anddetermine if the electronic anisotropy in the paramagnetic

    tetragonal phase of iron pnictides is intrinsic [42,43], orentirely due to the symmetry breaking uniaxial pressureapplied to the materials [44,45]. It is also important to deducewhat role the nature of the AF transition plays in the nematicsusceptibility [17,25,26,38] and how the latter depends on theuniaxial pressure.

    In this article, we use neutron resonance spin echo (NRSE)[46,47] and Larmor diffraction [48] to study the effectof uniaxial pressure on the structural and magnetic phasetransitions in electron-doped iron pnictides BaFe2−xNixAs2with x = 0,0.03,0.12 [9,10] and SrFe1.97Ni0.03As2 [18],and in the isovalently doped BaFe2(As0.7P0.3)2 [49]. Whilethe underdoped BaFe1.97Ni0.03As2 (TN = 109 K and Ts =114 K) exhibits a second-order AF transition below Ts ,SrFe1.97Ni0.03As2 has coupled first-order structural and mag-netic phase transitions at TN = Ts ≈ 175 K [50]. The electronoverdoped BaFe1.88Ni0.12As2 (Tc = 18.6 K) and isovalentlydoped BaFe2(As0.7P0.3)2 (Tc = 30 K) have a paramagnetictetragonal structure at all temperatures without static AForder. Figure 1(c) summarizes the key experimental resultof the present work, where the temperature dependenciesof the uniaxial pressure induced orthorhombic lattice distor-tions δ(P ≈ 20 MPa) − δ(P = 0 MPa) are determined usingneutron Larmor diffraction for BaFe2As2, BaFe1.97Ni0.03As2,SrFe1.97Ni0.03As2, and BaFe1.88Ni0.12As2 [we defined thelattice distortion δ = (a − b)/(a + b) with a and b being theorthorhombic lattice parameters]. Remarkably, the magnitudeof our determined structural nematic susceptibility dδ/dP ∝δ(P ) − δ(0) in Fig. 1(c) is comparable in all three materialsthat have a structural phase transition, unlike the very differentvalues of the resistivity anisotropy displayed in Fig. 1(b).Comparing these results with those of the elastoresistanceand nematic susceptibility obtained from transport [20,23,24]and from elastic shear modulus/ultrasound spectroscopymeasurements [38,51,52], we conclude that the resistivityanisotropy in the paramagnetic phase of the iron pnictidesdepends sensitively on whether the underlying magnetic phasetransition is first or second order. We also find a strong couplingbetween the uniaxial pressure induced lattice distortion andthe electronic nematic susceptibility, and have to be cautiousin directly relating resistivity anisotropy to the nematic orderparameter in the iron pnictides.

    II. RESULTS

    A. Experimental results

    Our experiments were carried out using conventionalthermal triple-axis spectrometer PUMA and three axes spin-echo spectrometer (TRISP) at the ForschungsneutronenquelleHeinz Maier-Leibnitz (MLZ), Garching, Germany. The princi-ples of NRSE and Larmor diffraction are described elsewhere[50]. Single crystals of BaFe2−xNixAs2, SrFe2−xNixAs2,and BaFe2(As0.7P0.3)2 were grown by self-flux method asdescribed before [49,53]. We define the momentum transferQ in the three-dimensional reciprocal space in Å−1 as Q =Ha∗ + Kb∗ + Lc∗, where H , K , and L are Miller indices anda∗ = â2π/a, b∗ = b̂2π/b, c∗ = ĉ2π/c with a ≈ b ≈ 5.6 Å,and c = 12.96 Å for BaFe2−xNixAs2. In this notation, theAF Bragg peaks should occur at (±1,0,L) (L = 1,3,5, . . . )

    134519-2

  • IMPACT OF UNIAXIAL PRESSURE ON STRUCTURAL AND . . . PHYSICAL REVIEW B 93, 134519 (2016)In

    tens

    ity (c

    ount

    s/0.

    01M

    )

    60 75 90 105 120 135 150

    0

    20

    40

    60 P ~15MPa@Puma

    0

    20

    40

    60(1,0,1)(0,1,1)

    P = 0 @Puma P ~15MPa@TRISP

    (1,0,1)

    (1,0,1)(0,1,1)

    TN = 109 K (P = 0)

    TN = 118 K (P ~15MPa)

    (a)

    -5

    0

    5

    10

    Γ ( μ

    eV)

    0 40 60 80 100 12020Temperature (K)

    Q = (1, 0, 1)(b)

    TN = 109 K (P = 0)TN = 118 K (P ~15MPa)

    P ~15MPa@TRISP

    FIG. 2. (a) Magnetic order parameters at Q = (1,0,1) for the zero(P = 0) and uniaxial pressured (P ∼ 15 MPa) BaFe1.97Ni0.03As2. TNis 109 K for an unpressured sample (blue diamonds). Upon applyinguniaxial pressure of P ≈ 15 MPa, the TN is enhanced to 118 K andthe sample becomes 100% detwinned as seen by PUMA and TRISPmeasurements. (b) The energy linewidth (half width at half maximum,�) of the magnetic Bragg peak Q = (1,0,1) measured by NRSE usingTRISP for BaFe1.97Ni0.03As2. The blue and red dashed lines indicateTN in P = 0 and 15 MPa unaixial pressure, respectively. The slightlarger errors of � near TN is due to low statistics data.

    positions in reciprocal space of a completely detwinned sample[right inset in Fig. 1(a)]. For neutron-scattering experiments,single crystals are aligned in either the [H,K,H + K] [41] or[H,K,0] zone.

    We first discuss the effect of uniaxial pressure on thecollinear AF order in BaFe2−xTxAs2. In previous neutron-scattering work on BaFe2−xCoxAs2, the Néel temperature (TN )was found to be pushed to higher temperature under uniaxialstrain field, forming a broader magnetic transition [39,40].Moreover, it seems that the increase in TN depends on theannealing condition [41,54]. Although the TN enhancementwas attributed to uniaxial strain aligned fluctuating magneticdomains, the effect of uniaxial pressure on the ordered momentremains elusive and the nature of the TN enhancement is stillunder debate [40,41,54].

    By aligning single crystals in the [1,0,1] × [0,1,1] scat-tering plane [41], we were able to determine TN , detwinningratio, as well as the ordered moment of the system under zeroand finite uniaxial pressures. Figure 2(a) shows temperaturedependence of the (1,0,1) and (0,1,1) magnetic scatteringintensity for BaFe1.97Ni0.03As2 obtained using PUMA [leftaxis in Fig. 2(a)] and TRISP (right axis). The two sets ofdata are in excellent quantitative agreement with each other.Under the applied uniaxial pressure of P ≈ 15 MPa, theNéel temperature increases from TN ≈ 109 K (at P = 0) toTN ≈ 118 K. The magnetic scattering intensity [Fig. 2(a)] inthe (1,0,1) peak becomes approximately twice as large asin the twinned sample, whereas the (0,1,1) peak vanishes,

    suggesting that the sample is completely detwinned and theapplied uniaxial pressure does not significantly affect theordered moment.

    To test whether the TN increase is an intrinsic featureof the system, we note that the magnetic order parameterunder uniaxial pressure has a round tail around TN [39,40],suggesting that the TN enhancement could arise from enhancedslow spin dynamics (critical scattering) under inhomogeneousuniaxial strain field and cannot be resolved by conventionaltriple-axis neutron diffraction due to its coarse energy resolu-tion (�E ≈ 0.3–1 meV). To clarify the nature of the increasein TN , we have measured the energy linewidth [� � 0; seeFig. 2(b)] of the quasielastic scattering for magnetic reflection(1,0,1) using high-energy resolution (�E ≈ 1 μeV) NRSEat TRISP [55]. As seen in Fig. 2(b), the � at all measuredtemperatures are resolution limited, indicating that the increasein magnetic scattering intensity below TN ≈ 118 K is elastic(� � 1 μeV), and an intrinsic nature of the system.

    To determine the effect of uniaxial pressure on thetetragonal-to-orthorhombic phase transition in iron pnictides,we carried out neutron Larmor diffraction experiments ca-pable of measuring minor change of lattice spacing d =2π/|Q(H,K,L)| and its spread �d with a resolution betterthan 10−5 in �d/d [inset in Fig. 3(a)] [48,50]. We focus on(4,0,0) and (0,4,0) nuclear Bragg reflections corresponding toa d spacing d ≈ a/4, which we measured in BaFe2−xNixAs2(x = 0,0.03,0.12), SrFe1.97Ni0.03As2, and BaFe2(As0.7P0.3)2both on freshly prepared samples (uniaxial pressure P = 0)and under uniaxial pressure (P ≈ 10,20 MPa). Figure 3 showsthe temperature and pressure dependence of the d spreadfor these samples. The d spread is characterized by theFWHM (full width at half maximum) of the lattice spacingdistribution f (�d/d), which is assumed to be a Gaussiandistribution [50]. The diamonds in Fig. 3(a) show temperaturedependence of the FWHM for BaFe2As2 at zero pressure.Similar to BaFe1.97Ni0.03As2 [45], temperature dependence ofFWHM follows a Curie-Wiess form and peaks around thezero-pressure value of TN ≈ Ts . Upon application of a uniaxialpressure P ≈ 20 MPa, the magnitude of FWHM increases atall temperatures and now peaks at an enhanced TN = 144 K[Fig. 3(a)].

    Figure 3(b) shows similar data for SrFe1.97Ni0.03As2,where there are coupled strong first-order structural andAF phase transitions at TN = Ts = 175 K [11]. Comparedwith BaFe2As2, where the AF phase transition is weaklyfirst order and structural transition is second order [5], theAF and structural transition induced changes in FWHM aremuch smaller and confined to temperatures close to TN ≈Ts in SrFe1.97Ni0.03As2 [Fig. 3(b)]. Under a uniaxial pres-sure P ≈ 20 MPa, however, both the FWHM and TN increasedramatically with solid lines showing Curie-Wiess fits to thedata. For BaFe1.97Ni0.03As2, application of a P ≈ 10 MPauniaxial pressure transforms temperature dependence of theFWHM, which forms a broad peak above the zero-pressurevalue of Ts . Upon releasing the uniaxial pressure [P released,filled green diamonds in Fig. 3(c)], the system goes back tothe original unpressured fresh state.

    Figure 3(d) compares temperature dependence of theFWHM for electron overdoped BaFe1.88Ni0.12As2 andBaFe2(As0.7P0.3)2, where both materials are in the param-

    134519-3

  • XINGYE LU et al. PHYSICAL REVIEW B 93, 134519 (2016)

    (4 0 0)(4 0 0)P released

    P = 0, fresh (4 0 0)(0 4 0)

    P ~10MPa

    100 120 140 160 180 200 220 240 260

    P ~ 20MPa

    P = 0

    Temperature (K)

    4

    6

    8

    10

    12

    14

    FWH

    M (Δ

    d/d

    10-4

    )

    16

    4

    6

    8

    10

    12

    14

    16

    FWH

    M (Δ

    d/d

    10-4

    )

    100 120 140 160 180 200 220 240 260

    4

    6

    8

    10

    12

    14

    FWH

    M (Δ

    d/d

    10-4

    )

    20 50 100 150 200 250 3004

    80 120 160 200 240

    P = 0

    P ~14MPa

    P = 0

    4004

    6

    8

    10

    12

    14

    16

    FWH

    M (Δ

    d/d

    10-4

    )

    BaFe2As2

    SrFe1.97Ni0.03As2

    BaFe1.97Ni0.03As2

    BaFe2(As0.7P0.3)2

    BaFe1.88Ni0.12As2

    P ~ 20MPa P = 0

    Q = (4 0 0)

    (a)

    (b)

    (c)

    (d)

    TN=TS=138(1)K

    T=144(2)K

    Q = (4 0 0)

    TN=TS=175(2)K

    T=180(2)K

    TN=109(2)K TS=114(2)K

    T=120(2)K

    1

    3

    5

    7

    9

    11

    13

    d/dave

    Δd/dΔd/d

    Q = (4 0 0)

    FIG. 3. Temperature dependence of full width at half maximum(FWHM) of �d/d in several iron pnictides under different uniaxialpressures obtained from neutron Larmor diffraction experiments [50].(a) Temperature dependence of �d/d in FWHM for the (4,0,0) Braggreflection of BaFe2As2 at P = 0 and 20 MPa. The solid line above TNis a fit using Curie-Wiess formalism [FWHM(T ) = A/(T − T1) + B,where A, B, and T1 are fitting parameters]. (b) Similar data forSrFe1.97Ni0.03As2. The vertical blue and red dashed lines in (a)and (b) mark the TN of the sample at zero and finite pressure,respectively. (c) Similar data for BaFe1.97Ni0.03As2, where the verticalgreen and blue dashed lines mark TN and Ts , respectively, at zeropressure. The open green and yellow squares mark measurementsof FWHM under zero pressure (fresh) at the (4,0,0) and (0,4,0)Bragg peaks, respectively. The pink solid circles are identicalmeasurements under P ≈ 10 MPa uniaxial pressure on (4,0,0). Thesolid green diamonds are data after uniaxial pressure is released. Thevertical red dashed line indicates the peak position of the FWHMunder P = 10 MPa. (d) Temperature dependence of FWHM in�d/d for BaFe2(As0.7P0.3)2 at P = 0 MPa (solid green circles),BaFe1.88Ni0.12As2 at P = 0 (solid green diamonds), and 14 MPa(solid red circles).

    agnetic tetragonal state without static AF order. The weaktemperature dependence of FHWM in these materials suggeststhat the large temperature dependence of FWHM in AFordered BaFe2−xNixAs2 (x = 0,0.03) and SrFe1.97Ni0.03As2 isdue to a strong magnetoelastic coupling. Although applicationof a P ≈ 14 MPa uniaxial pressure on BaFe1.88Ni0.12As2increases the absolute value of FWHM, it is still weaklytemperature dependent [Fig. 3(d)].

    To further demonstrate the impact of uniaxial pressureon the tetragonal-to-orthorhombic structural transition inBaFe2−xNixAs2 (x = 0,0.03,0.12) and SrFe1.97Ni0.03As2, wecompare in Fig. 4 temperature dependence of the latticeparameters along the orthorhombic a and b axis directionsunder zero and finite uniaxial pressure. We first discuss resultsfor BaFe2−xNixAs2 with x = 0 [Figs. 4(a) and 4(b)] and0.03 [Figs. 4(c) and 4(d)]. At P = 0, the lattice parametershave a = b at temperatures above Ts (tetragonal phase) anddecrease linearly with decreasing temperature [open diamondsand hexagons in Figs. 4(a) and 4(c)]. Upon application of auniaxial pressure, the system becomes orthorhombic at all tem-peratures and the orthorhombic structural transition becomesa crossover [filled diamonds and hexagons in Figs. 4(a) and4(c)]. Figures 4(b) and 4(d) show temperature dependence ofthe lattice orthorhombicity δ = (a − b)/(a + b) at differentuniaxial pressures for x = 0, and 0.03, respectively. Forunpressured fresh samples (P = 0), and after the pressure hasbeen released, the tetragonal structure becomes orthorhombicbelow Ts and the AF order below TN further enhancesthe lattice orthorhombicity [5]. Upon applying the uniaxialpressure P ≈ 10, 15, and 20 MPa, the temperature dependenceof the lattice orthorhombicity becomes remarkably similarto that of the B2g elastoresistance and nematic susceptibil-ity of BaFe2−xTxAs2 obtained from transport [23,24] andelastic shear modulus/ultrasound spectroscopy measurements[38,51,52], respectively.

    B. Theoretical Ginzburg-Landau analysis

    To understand the temperature dependence of the pressure-induced lattice orthorhombicity described in Figs. 4(b) and4(d), we consider the Ginzburg-Landau free-energy formalismused in previous works [20,38]:

    F [ϕ,δ] = F0 + a2

    (T − T0)ϕ2 + B̃4

    ϕ4 + C66,02

    δ2 − λδϕ−Pδ,(1)

    where the electronic nematic order parameter ϕ is coupledlinearly to the orthorhombic lattice distortion δ. It then followsthat (see the Appendix [50])

    δ = (λ〈ϕ〉 + P )/C66,0, (2)

    where C66,0 is the bare elastic constant that has no strong tem-perature dependence and P is the conjugate uniaxial pressure(stress) [38,41,50–52]. In the absence of the elasto-nematiccoupling (λ = 0), the nematic susceptibility χϕ = 1/[a(T −T0)] is characterized by the Curie-Weiss temperature T0. Uponconsidering the coupling between the nematic order parameterϕ and the structural lattice distortion δ (or equivalently, theelastic shear strain ε6), the elastic susceptibility takes on the

    134519-4

  • IMPACT OF UNIAXIAL PRESSURE ON STRUCTURAL AND . . . PHYSICAL REVIEW B 93, 134519 (2016)

    δ =

    (a-b

    )/(a

    +b) (

    10-3

    )

    (b)

    (a)Δa

    /a ,

    Δb/b

    (10-

    3 )

    P ~ 20 MPa, Δb/b

    P ~ 20 MPa, Δa/a

    BaFe2As2

    TS (P = 0)

    100 140 180 220 260

    0

    1

    2

    3

    4

    P ~ 20 MPa

    P = 0, fresh P released

    -4

    -2

    0

    2

    4

    -6

    -4

    -2

    0

    2

    4

    140 160 180 200 220 240 260-1

    0

    1

    2

    3

    4 P released P ~ 20 MPa

    P ~ 20 MPa, Δb/b

    P ~ 20 MPa, Δa/a

    SrFe1.97Ni0.03As2

    TS (P = 0)

    -1.6

    -1.2

    -0.8

    -0.4

    0

    0.2

    0 50 100 150 200 250Temperature (K)

    P ~ 14MPa

    0.1

    TC=18.6 K

    Δa/a

    Δb/b Aluminum

    Δa/a

    δ =

    (a-b

    )/(a

    +b) (

    10-3

    )

    (f )

    (e)

    Δa/a

    , Δb

    /b (1

    0-3 )

    0

    1

    2

    3

    -4

    -3

    -2

    -1

    0

    1

    2

    -5

    P = 0, fresh

    P ~ 15 MPa

    P released

    P ~ 10 MPa

    P ~ 15MPa, Δb/b

    P ~ 15MPa, Δa/a

    P = 0, fresh, Δa/a

    P = 0, fresh, Δb/b

    40 120 160 200 240

    Temperature (K)

    80

    BaFe1.88Ni0.12As2

    BaFe1.97Ni0.03As2

    -4

    -3

    -2

    -1

    0

    1

    2

    -5

    TN (P = 0)

    TS (P = 0)

    -1

    (h)

    (g)

    (d)

    (c)

    P released, Δa/aP released, Δa/a

    P released, Δa/aP released, Δa/a

    FIG. 4. Temperature dependence of the a and b lattice parameters and orthorhombicity δ under different uniaxial pressure conditions(P = 0 fresh, ∼10, ∼15, ∼20, and 0 released MPa) for BaFe2−xNixAs2 (x = 0,0.03,0.12) and SrFe1.97Ni0.03As2. (a) Temperature dependenceof the a and b under P = 0 and 20 MPa uniaxial pressure for BaFe2As2. (b) Temperature dependence of δ under different uniaxial pressure(P = 0 fresh, ∼20, and 0 released MPa). The vertical blue dashed line marks the TN/Ts . (c) Temperature dependence of the lattice parametersa and b at P = 0 and 15 MPa for BaFe1.97Ni0.03As2. (d) Temperature dependence of δ under different uniaxial pressure (P = 0 fresh, ∼10,∼15, and 0 released MPa). The open red diamonds and green hexagons are obtained by lattice thermal expansion measurements under uniaxialpressure. The pink squares are measurements of an unpressured fresh sample, while blue circles are obtained after releasing P ≈ 10 MPauniaxial pressure. The blue circles and two pink squares below Ts are from zero pressure Larmor diffraction measurements. The pink squaresabove Ts are obtained from thermal-expansion measurements. The vertical green and blue dashed lines in (c) and (d) mark the TN and Tsof the sample at zero pressure, respectively. (e) Temperature dependence of the a and b for BaFe1.88Ni0.12As2. The lattice thermal expansionof aluminum is plotted as a reference [56]. The vertical arrow marks Tc = 18.6 K and the solid lines are guides to the eye. (f) Temperaturedependence of the orthorhombic lattice distortions calculated from (c). (g) Temperature dependence of the a and b lattice parameters forSrFe1.97Ni0.03As2 under P = 0 fresh, 20, and 0 (released) MPa. (h) Temperature dependence of δ for the same pressure condition. The solidcurves in (b), (d), (f), and (h) are fits using a Curie-Weiss functional form [50].

    134519-5

  • XINGYE LU et al. PHYSICAL REVIEW B 93, 134519 (2016)

    form [20,38]

    dP= 1

    C66,0

    T − T0T − T CWs

    (3)

    with the renormalized nematic transition temperature T CWs =T0 + λ2/(aC66,0) that is increased compared to the bareCurie-Weiss temperature T0. The pressure-induced latticedistortions in Figs. 4(b) and 4(d) can be well described bythe Curie-Weiss functional form [50]. Therefore, uniaxial pres-sure induced orthorhombic lattice distortion and its tempera-ture dependence in undoped and underdoped BaFe2−xNixAs2are directly associated with the nematic susceptibility [24,38].Since the external uniaxial pressure explicitly breaks thetetragonal lattice symmetry, it turns the nematic transition atTs = T CWs into a crossover, as is clearly seen in Fig. 4.

    If the in-plane resistivity anisotropy in electron under-doped iron pnictides indeed arises from the coupling ofthe uniaxial-pressure induced lattice distortion δ with thenematic susceptibility, it would be interesting to determine theeffect of similar uniaxial pressure on the electron overdopedsample, where the resistivity anisotropy is known to be muchweaker [14]. Figures 4(e) and 4(f) summarize the outcomeof the neutron Larmor diffraction experiments on uniaxialpressured BaFe1.88Ni0.12As2, which is tetragonal (a = b) andnonmagnetic at all temperatures in zero pressure [10]. Fig-ure 4(e) shows temperature dependence of the lattice parameterchanges along the a axis (�a/a) and b axis (�b/b) under auniaxial pressure of P ≈ 14 MPa. For comparison, the thermalcontraction of aluminum is also shown [56]. Figure 4(f)shows the temperature dependence of the orthorhombic latticedistortion δ, which reveals a clear anomaly at Tc consistentwith ultrasonic spectroscopy measurements [51,52]. Whilethe applied uniaxial pressure induces orthorhombic latticedistortion at 230 K, the magnitude of the lattice distortion,δ ≈ 1.1 × 10−4, is about five times smaller than that ofBaFe2As2 and BaFe1.97Ni0.03As2 at 230 K. On cooling to20 K, δ in BaFe1.88Ni0.12As2 increases to ∼2 × 10−4, whileδ in BaFe2As2 and BaFe1.97Ni0.03As2 becomes ∼2.5 × 10−3near Ts [Figs. 4(b) and 4(d)], an order of magnitude larger thanthat of the electron overdoped compound.

    To understand how a uniaxial pressure affects the first-ordernature of the structural and magnetic phase transitions inSrFe1.97Ni0.03As2, we compare in Figs. 4(g) and 4(h) tempera-ture dependence of the lattice parameters and orthorhombicityunder the zero and finite uniaxial pressure. At zero pressure, thefirst-order nature of the structural transition is clearly seen inhysteresis of temperature dependence of the lattice parametersand distortion [Figs. 4(g) and 4(h)]. Upon application ofP ≈ 20 MPa uniaxial pressure, the lattice orthorhombicityno longer displays the first-order transition at Ts , but insteadbecomes a crossover, similar to that observed in the undopedand underdoped BaFe2−xNixAs2 [see Figs. 4(b) and 4(d)].

    Assuming that the application of the modest uniaxialpressure P ≈ 20 MPa can be considered in the linear-responseregime [45], we can estimate the elastic susceptibility from thefinite difference dδ/dP ∝ �(δ) = δ(P ≈ 20 MPa) − δ(P =0) and compare it among the different compounds in theiron pnictide family. Figure 1(c) compares temperature de-pendence of δ(P ≈ 20 MPa) − δ(P = 0) for BaFe2−xNixAs2

    (x = 0,0.03,0.12) and SrFe1.97Ni0.03As2 normalized for P ≈20 MPa. For AF ordered BaFe2−xNixAs2 (x = 0,0.03) andSrFe1.97Ni0.03As2, the magnitudes of the pressure-inducedlattice orthorhombicity are similar in the paramagnetic phaseand vanish rapidly upon entering into the AF ordered state.Furthermore, the δ(P ≈ 20 MPa) − δ(P = 0) decreases forthe iron pnictides with reduced TN , and are much smaller forBaFe1.88Ni0.12As2.

    III. DISCUSSION

    It is well known that the effect of increasing electron dopingin BaFe2−xTxAs2 is to suppress the static AF order and toeliminate the low-temperature lattice orthorhombicity [7–10].At zero pressure, BaFe2As2 first exhibits a second-orderstructural transition from the high-temperature paramagnetictetragonal phase to a paramagnetic orthorhombic phase at Ts ,followed by a discontinuous further orthorhombic structuraldistortion and weakly first-order AF phase transition at TN(TN < Ts) due to magnetoelastic coupling [5]. Upon Nidoping in BaFe2−xNixAs2, the structural and magnetic phasetransitions are gradually separated and suppressed [Fig. 1(a)],and become second order in nature [7–10]. Upon application ofa uniaxial pressure, the C4 rotational symmetry of the tetrag-onal lattice is broken. Since the tetragonal-to-orthorhombicsymmetry of the underlying lattice can only be broken once,Ts will become a crossover regardless of the magnitudeof the applied pressure, as our findings in Figs. 3 and 4corroborate. The same conclusion holds for SrFe1.97Ni0.03As2where the first-order structural transition becomes a crossover[see Figs. 4(g) and 4(h)]. Therefore, both BaFe2−xNixAs2 andSrFe1.97Ni0.03As2 under uniaxial pressure can only exhibitAF phase transition. We note that our measurements andtheoretical Landau-Ginzburg analysis do not rely on themicroscopic nature of the nematic order parameter ϕ. Inparticular, they apply equally well to the so-called Ising spinnematic scenario [25–29] or the orbital order interpretationof nematicity [30–36]. In fact, the ferro-orbital order ϕorb =〈nxz − nyz〉 is always linearly coupled [26,36] to the Isingspin nematic order parameter ϕspin = 〈Si · Si+x̂ − Si · Si+ŷ〉,so the orbital order is generically present whenever ϕspin �= 0,although there are theoretical indications that the converseis not always true. In other words, the orbital order canexist in the absence of static AF order [36], as is known tobe the case in FeSe [57–59]. In either case, the applicationof external uniaxial stress renders the nematic transition acrossover, so that the lattice distortion δ and consequentlyϕ are both finite above the zero-stress value of Ts . In thislight, the electronic anisotropy seen in the magnetic torque[21] and scanning tunneling microscopy [43] measurementsabove Ts without explicit external uniaxial pressure is likelydue to intrinsic local strain in these materials which breaks theC4 rotational symmetry of the paramagnetic tetragonal phase.Indeed, local strain-induced effect has recently been observedin free standing BaFe2As2 above TN and Ts [60].

    The key finding of the present work is that undopedBaFe2As2, as well as BaFe1.97Ni0.03As2 and SrFe1.97Ni0.03As2all exhibit similar magnitudes of the pressure-induced latticeorthorhombicity [Figs. 1(c) and 4] and FWHM of �d/d nearTN (Fig. 3). This indicates that these samples experience

    134519-6

  • IMPACT OF UNIAXIAL PRESSURE ON STRUCTURAL AND . . . PHYSICAL REVIEW B 93, 134519 (2016)

    similar strain field under nominally similar applied uniaxialpressure, thus suggesting that the doped Ni impurities do notplay an important role in determining the strain field insidethe sample. Theoretically, the electronic anisotropy of theiron pnictides is expected to couple linearly to the latticeorthorhombicity δ [26–29], as captured by the effective Landaufree energy in Eq. (1). The Curie-Weiss-like temperaturedependence of the uniaxial pressure induced lattice distortion[Fig. 1(c)] is consistent with the temperature dependence ofthe nematic susceptibility dδ/dP in Eq. (3) and agrees withthe results of Young’s modulus measurements [38]. This givesus confidence that in the effective Landau description [50], theuniaxial pressure-induced lattice distortion δ has a componentproportional to the electronic nematic order parameter ϕ viaEq. (2), where one expects δ ∝ ϕ in zero pressure (P = 0).Since δ has similar magnitude in BaFe2As2, BaFe1.97Ni0.03As2,and SrFe1.97Ni0.03As2 (see Fig. 4), one would also expectcomparable values of ϕ in all three compounds. So if oneuses the resistivity anisotropy �ρ = (ρb − ρa)/(ρb + ρa) asa proxy for the nematic order parameter, as has been widelyused in the literature [12,20,23,24], how does one then explainthe resistivity anisotropy differences in BaFe2−xTxAs2 [14]and a much smaller resistivity anisotropy above TN [Fig. 1(b)]in SrFe2−xTxAs2 family of materials [18]? The bare value ofthe elastic shear modulus C66,0 that enters Eq. (A16) has nostrong temperature dependence [38] and from the Curie-Weissfits of the nematic susceptibility to Eq. (A20), we find it to beroughly the same in all three compounds, C66,0 ≈ 50 GPa [50].The only remaining unknown variable is the elastonematiccoupling constant λ, which could be material dependent butnot temperature dependent [20,38]. It is thus very challengingto explain the qualitatively different temperature dependenceof the resistivity anisotropy in BaFe1.97Ni0.03As2 [monotonic,blue diamonds in Fig. 1(b)] from that in BaFe2As2 and inSrFe1.97Ni0.03As2 [both nonmonotonic, with a maximum at orjust below Ts]. One possible explanation for the nonmonotonictemperature dependence of the resistivity anisotropy, recentlyproposed in the context of FeSe [61], is to assume atemperature-dependent coefficient of proportionality between�ρ and ϕ:

    �ρ(T ) = ϒ(T )ϕ(T ), (4)

    such that ϒ(T ) tends to zero as T → 0, whereas ϕ(T ) isexpected to increase monotonically below Ts as the temper-ature is lowered (consider for instance the mean-field resultϕ(T ) ∝ √Ts − T for the second-order phase transition).

    Even with the introduction of ϒ(T ) in Eq. (4), which hasa meaning of the temperature-dependent scattering function,it is extremely difficult to explain the much lower valueof �ρ in SrFe1.97Ni0.03As2 compared to BaFe2As2 andBaFe1.97Ni0.03As2. In fact, from the Curie-Weiss fits of thesusceptibility data, we estimate the elastonematic couplingconstant λ to be a factor of ∼5 smaller in SrFe1.97Ni0.03As2compared to BaFe1.97Ni0.03As2 [50]. Given the comparablemagnitudes of δ between the two compounds [see Figs. 4(d)and 4(h)], one would then expect the nematic order parameterϕ to be a factor of ∼5 greater in SrFe1.97Ni0.03As2, toensure that the left-hand side of Eq. (2) remains of thesame magnitude. And yet the resistivity anisotropy �ρ ∝ ϕ

    paints a diametrically opposite picture, being much smaller inSrFe1.97Ni0.03As2.

    We propose that a likely resolution of this dilemma lies inthe nature of the magnetic phase transition which we have sofar neglected in our analysis. Indeed, it is well established thatstructural and magnetic phase transitions in SrFe2−xTxAs2 arecoupled first-order transitions that decrease with increasingx before vanishing near optimal superconductivity [11],while electron-doped BaFe2As2 has second-order magneticand structural phase transitions [7–10]. Although applicationof a uniaxial pressure renders the structural transition acrossover, the first-order nature of the magnetic transitionmeans a vanishing critical regime with suppressed low-energyspin fluctuations at temperatures near TN , compared withthose of BaFe2−xTxAs2 where the AF phase transition issecond order. One expects the scattering of electrons on themagnetic fluctuations, and hence the resistivity, to thereforebe smaller in the vicinity of the first-order magnetic transition,as is the case in SrFe2−xTxAs2. We thus conclude that thevanishing resistivity anisotropy above TN in the uniaxialpressure detwinned SrFe2−xTxAs2 (compared with those ofBaFe2−xTxAs2) is likely rooted in the first-order nature ofthe AF phase transition. This is also consistent with theincreased paramagnetic resistivity anisotropy on moving fromBaFe2As2 to BaFe1.96Co0.04As2 [15], where the magnetictransition changes from weakly first order to second order[5,7,8]. Similarly, the lack of large resistivity anisotropy in theparamagnetic state of uniaxial pressured Ba1−xKxFe2As2 [62],Ba1−xNaxFe2As2 [63], and Ca1−xLaxFe2As2 [64] is likely dueto the first-order nature of the paramagnetic to AF phasetransition in these materials. The phenomenological Landautheory can be extended to include the coupling betweennematicity ϕ and the magnetic order parameter [38,50], andour theoretical analysis shows [50] that the resulting uniaxialpressure-induced lattice distortion δ(P ) − δ(0) reproducessemiquantitatively the experimental findings in Fig. 1(c).

    We conclude that the in-plane resistivity anisotropy foundin the paramagnetic state of iron pnictides depends sensitivelyon the nature of the magnetic phase transition and a strongelastonematic coupling between the uniaxial pressure inducedlattice distortion and the electronic nematic susceptibility.We caution that while the resistivity anisotropy �ρ and itsdependence on the shear strain can be successfully used toextract the quantity proportional to the nematic susceptibility[20], care should be taken when equating �ρ with thenematic order parameter itself. In particular, the nonmonotonictemperature dependence of �ρ and its sensitivity to thenature of the magnetic phase transition remain relatively littleexplored and deserve further experimental and theoreticalstudies.

    ACKNOWLEDGMENTS

    We thank A. E. Böhmer, J. H. Chu, J. Hu, and Q. Sifor helpful discussions. The neutron-scattering work at Ricewas supported by the US NSF-DMR-1308603, NSF-DMR-1362219, and DMR-1436006 (P.D.). A.H.N. was supportedby US NSF CAREER Award No. DMR-1350237). This workis also supported by the Robert A. Welch Foundation GrantsNo. C-1839 (P.D.) and No. C-1818 (A.H.N.). The work at

    134519-7

  • XINGYE LU et al. PHYSICAL REVIEW B 93, 134519 (2016)

    0.8

    0.9

    1.0

    ZFC FC

    χ (a

    rb.u

    .)

    80 90 100 110 120 130 140

    0.0

    0.4

    0.8

    1.2

    1.6

    dχ/d

    T (a

    rb. u

    .)

    80 90 100 110 120 130 140

    0

    10

    20

    30

    (1 0 1)

    coun

    ts (a

    rb. u

    .)

    Temperature (K)

    600

    800

    1000

    1200

    1400 (2 -2 0)

    TN = 109(2)K

    TS = 114(2)K

    H = 10000 OeH // ab plane

    BaFe1.97

    Ni0.03

    As2

    TN = 109(2)K

    (a)

    (c)

    (b)

    TS = 114(2)K

    ZFC

    FIG. 5. (a) Temperature dependence of the magnetic susceptibil-ity under ZFC and FC cases for BaFe1.97Ni0.03As2. (b) Temperaturederivative of the magnetic susceptibility, showing clearly the struc-tural and magnetic phase transitions. (c) Temperature dependence ofthe Q = (1,0,1) magnetic (green diamonds) and (2,−2,0) nuclear(red squares) Bragg peak intensity [41]. The extinction release ofthe (2,−2,0) Bragg reflection is sensitive to the change of structuraldistortion and used to determine the Ts . Combining the results in(a)–(c), the TN and Ts are determined as shown in the green and redvertical dashed lines, respectively.

    the Institute of Physics, Chinese Academy of Sciences, wassupported by Ministry of Science and Technology of China(973 Projects No. 2012CB821400 and No. 2011CBA00110),National Natural Science Foundation of China Projects No.11374011 and No. 91221303, and The Strategic PriorityResearch Program (B) of the Chinese Academy of SciencesGrant No. XDB07020300. T.K. thanks for support by theDeutsche Forschungsgemeinschaft through SFB/TRR80.

    APPENDIX

    1. Sample Information

    The iron pnictide single crystals used in the presentstudy were prepared by the self-flux method [53]. Thesamples have been characterized by resistivity, magnetiza-tion, and neutron-scattering measurements. Figures 5 and 6show the basic characterizations of BaFe1.97Ni0.03As2 andSrFe1.97Ni0.03As2 samples, respectively. The basic character-istics of the BaFe2As2 and BaFe1.88Ni0.12As2 samples can befound elsewhere [41,53].

    TN

    = TS = 175(2)K

    SrFe1.97

    Ni0.03

    As2

    (a)

    (b)0.128

    0.132

    0.136

    0.140

    0.144

    0.148

    Resi

    stiv

    ity (a

    rb. u

    .)

    -6

    -4

    -2

    0

    2

    dR/d

    T (a

    rb. u

    .)

    155 160 165 170 175 180 185 190

    50

    60

    70

    80 Warming Cooling

    coun

    ts (a

    rb. u

    .)

    Temperature (K)

    (4 0 0)

    FIG. 6. (a) Temperature dependent resistivity and its temperaturederivative for SrFe1.97Ni0.03As2. (b) Temperature dependence of the(4,0,0) nuclear Bragg peak intensity measured for warming andcooling. The dramatic increase of the (4,0,0) Bragg reflection signalsthe change of structural distortion near Ts [41].

    Figure 5(a) shows temperature dependence of the zero-field-cooled (ZFC) and field-cooled (FC) magnetic sus-ceptibility χ . Figure 5(b) is temperature derivative of χ ,dχ/dT . Figure 5(c) shows temperature dependence of themagnetic (1,0,1) Bragg peak (green diamonds) and (2,−2,0)nuclear Bragg intensity (red squares). These results estab-lish TN (green dashed line) and Ts (red dashed line) ofBaFe1.97Ni0.03As2. The TN was determined as 109 ± 2 K frommagnetic order parameter of the (1,0,1) magnetic Bragg peak[Fig. 5(c)] and temperature-dependent magnetization mea-surements [Fig. 5(a)]. The structural transition temperatureTs is estimated from a feature shown in the magnetization,[Fig. 5(b)], and the neutron extinction release of the (2,−2,0)nuclear Bragg peak intensity [Fig. 5(c)] [41,65,66].

    Figure 6(a) shows temperature-dependent resistivity R andits derivative dR/dT for SrFe1.97Ni0.03As2. The dip of thedR/dT at T = 175 K indicates the concomitant structuraland magnetic transition, different from the two features forthe separated TN and Ts in BaFe1.97Ni0.03As2. Figure 6(b)is temperature dependence of the (4,0,0) nuclear Braggpeak intensity. The dramatic increase of the peak intensityalso signals the structural transition. The observed intensityhysteresis is consistent with first-order nature of the structuraland magnetic phase transition.

    Large single crystals were selected and cut into rectangularshapes along the orthorhombic [1,0,0] and [0,1,0] directionsby a high precession wire saw. The well-cut samples wereplaced inside a uniaxial pressure device with b axis being thedirection of the applied pressure [41]. The applied uniaxialpressures for the the samples range from P ≈ 10 MPa to P ≈20 MPa, as described before.

    In order to measure temperature and pressure dependenceof orthorhombic lattice distortion δ = (a − b)/(a + b), thesamples were mounted in the [H,K,0] scattering plane, whereorthorhombic (4,0,0) and (0,4,0) Bragg reflections can be

    134519-8

  • IMPACT OF UNIAXIAL PRESSURE ON STRUCTURAL AND . . . PHYSICAL REVIEW B 93, 134519 (2016)

    L1

    sample

    C1 C4

    L2

    Gin out AD

    θB

    α

    C2 C3LD

    (b)

    (a)

    L1 L2

    xy Bz

    v1

    BSample

    σ σy

    FIG. 7. (a) Basic principles for neutron spin echo with schematicsof the experimental setup. The magnetic guide field B directionsare clearly marked. (b) Schematic diagram for the neutron Larmordiffraction measurements. For NRSE, the neutron precession direc-tion in L1 is opposite to that of L2. In neutron Larmor diffraction, theneutron precession directions are same in L1 and L2.

    measured. The effects of uniaxial pressure on tetragonal-to-orthorhombic structural transition and uniaxial pressureinduced lattice orthorhombicity can be probed directly viameasuring temperature and pressure dependence of the (4,0,0)and (0,4,0) reflections by neutron Larmor diffraction. Formagnetic measurements, the BaFe1.97Ni0.03As2 sample wasmounted in the [1,1,2] × [1,−1,0] scattering plane, whereboth the (1,0,1) and (0,1,1) magnetic Bragg peaks can bereached [41].

    2. Neutron resonance spin-echo measurements

    Neutron spin-echo (NSE) technique has been demon-strated to be an effective method to measure the slowdynamics (quasielastic scattering) with an extremely high-energy resolution (∼1 μeV or even to ∼1 neV) [46]. Bycombining triple axis spectrometer and neutron resonancespin-echo (NRSE) techniques, the TRISP spectrometer at theForschungsneutronenquelle Heinz Maier-Leibnitz (MLZ) iscapable of measuring the lifetime of excitations with an energyresolution �E ∼ 1 μeV in the range of about 1–200 μeV [67].

    Compared with typical neutron-scattering experimentswhere S(Q,ω) (E = �ω) is usually measured, neutron spin-echo measures I (Q,τNSE) or P (Q,τNSE), where P is thepolarization of the scattered neutrons, which is the time Fouriertransform of the S(Q,ω) and thus provides direct informationof S(Q,ω) such as energy linewidth (lifetime) and intensity[48,68].

    The basic principle of NSE can be understood in a simplifiedpicture as shown in Fig. 7(a). We assume neutrons polarizedalong the y direction with a velocity v1 enter the first

    arm of NSE spectrometer with a constant magnetic field B[Fig. 7(a)]. The precession angle in the first arm (L1) is thenφ1 = ωLt = γ |B|L1/v1, where γ = 2μN/� = 2.916 kHz/Gis the gyromagnetic ratio of neutron, L1 is the length of thefirst neutron guide arm, and t is the time for neutron to travelthrough the first arm. After interactions with the sample, someneutrons are scattered into different energy with velocity v2.In the second arm (L2), the neutron spin will precess along theopposite direction, generating −φ2 = −ωLL2/v2. AssumingL1 = L2 = L and v2 = v1 + δv, δv � v1, the net phase afterpassing through both field regions will be φ = ωLL

    v21δv. Since

    neutron energy transfer is �ω = 12m(v22 − v21) ≈ mv1δv, thenet phase can be written as

    φ =(

    �ωLL

    mv31

    )ω ≡ ωτNSE, (A1)

    where τNSE is defined as

    τNSE =(

    �ωLL

    mv31

    )= 1.863 × 10−16B(G)L(cm)λ3(Å) (A2)

    Note τNSE is not a physical time but a quantity determined byspecific parameters of the spectrometer, with the dimension oftime.

    The polarization along y direction of the scattered neu-trons can be analyzed and detected [Fig. 7(a)]. The averagepolarization 〈σy〉 for neutrons with energy transfer �ω is

    〈σy〉 = 〈cosφ〉 =∫

    dωS(Q,ω)cosωτNSE. (A3)

    Thus 〈σy〉 is the cosine Fourier transform of S(Q,ω) for ω andhas been shown equal to the intermediate scattering functionI (Q,τ ). Therefore, the τNSE dependent polarization P (τ ), thatis, I (Q,τ ), provide direct information about S(Q,ω) [48].

    In the NRSE, the precession fields and spin flippers arereplaced by four short bootstrap rf spin flipper coils [C1–C4in Fig. 7(b)], which can improve the energy resolution by afactor of 4 compared with the NSE with the same B and L.The neutrons only precess in bootstrap while keep their spindirections in L1 and L2. L2 can be tuned by translating theflipper C4, by which the intensity with respect to the positionof C4, I (xc4), can be measured. For a fixed τ , the measuredintensity can be described as

    I (xc4) = I02

    [1 + P cos

    (2π

    �xc4(xc4 − xc4,0)

    )], (A4)

    where P is the polarization, I0 is the averaged intensity of thescattered beam, �xc4 is the period of the intensity modulation,and xc4,0 is the reference position of C4.

    The measurements of the P (τ ) for BaFe1.97Ni0.03As2 underP ≈ 15 MPa are summarized in Fig. 8. Figure 8(a) showsthe intensity modulations for τ = 16.29 and 39.25 ps ofQ = (1,0,1) at T = 102 K. The polarizations are obtainedthrough fitting the data by Eq. (A4). The fitted P (τ ) fordifferent temperatures are plotted in Fig. 8(b). Assuming thepossible broadening in energy of the magnetic reflections iscaused by some slow dynamics (quasielastic scattering), thecorresponding S(ω) can be described by a simple Lorentzian:

    S(ω) = 1π

    (ω − ω0)2 + �2 , (A5)

    134519-9

  • XINGYE LU et al. PHYSICAL REVIEW B 93, 134519 (2016)

    100

    150

    200

    250

    300

    350

    Inte

    nsity

    (cou

    nts/

    0.06

    M)

    (a)

    (b)

    -25 -24 -23 -22xc4 (cm)

    τ = 16.29 ps

    τ = 39.25 ps

    T = 101.8 K

    0.3

    0.5

    Pola

    rizat

    ion

    0.1

    0.6

    0.4

    0.2

    3.4 K101.8 K110.4 K115.2 K

    Γ = 5 μeV

    -10 100

    Γ = 1 μeV

    Γ = 5 μeV

    ΔE ∼1μeV

    E (μeV)

    τ (ps)0 10 20 30 40 50

    0

    -0.1

    BaFe1.97Ni0.03As2

    FIG. 8. (a) Measurements of the neutron polarization in oneperiod of φ for τ = 16.29 and 39.25 ps. The solid lines are cosinefittings of the data using Eq. (A4). (b) P (τ ) for various temperatures.The solid lines are fittings by Eq. (A6). The brown dashed line showsa drawing of Eq. (A6) with � = 5 μeV for comparison purpose.The corresponding Lorentzian S(ω) are shown as the solid green andbrown lines in the inset.

    where � (� � 0) (half width at half maximum) is the linewidthof the quasielastic scattering (ω0 = 0). Followed by Eq. (A5),the P (τ ) should be fitted by the Fourier transform of Eq. (A5):

    P (τ ) = P0exp(

    −�τ�

    ). (A6)

    All the P (τ ) in our measurements can be well describedby this exponential decay, as shown in Fig. 8(b). The fittedenergy linewidths � are less than 1 μeV, meaning the signalsare resolution limited at the measured temperatures. Thecomparison between S(ω) for resolution limited (� � 1 μeV)and � = 5 μeV is shown in the inset of Fig. 8(b), as a reference.The temperature dependence of � is shown in Fig. 2(b) of themain text, which must have values greater than zero. The largeerror bars for the values of � near TN are due to experimentaluncertainties of P (τ ) in Fig. 8(b).

    3. Larmor diffraction measurements

    We now turn to the neutron Larmor diffraction measure-ments. Larmor diffraction is a neutron Larmor precessiontechnique capable of measuring lattice spacing expansionand spread with a resolution better than 10−5 in terms of�d/d. It is sensitive to minor change of lattice spacing dbut insensitive to sample mosaicity and not much affectedby slight misalignment [69]. The total precession phase (φtot)

    -26 -25 -24 -23 -22 -21

    3000

    4000

    5000

    6000 T = 113 K T = 118.6 K

    Inte

    nsity

    (cou

    nts/

    0.06

    M)

    xc4 (mm)

    Q = (4, 0, 0)δLBaFe1.97Ni0.03As2

    FIG. 9. Measurement of the �φtot between T = 113 K and T =118.6 K, by which the relative change of d spacing can be tracked.

    dependent polarizations [P (φtot)] can be used to determinethe d spread and the splitting between multiple d spacingswith small differences, such as the peak splitting caused bythe tetragonal-to-orthorhombic structural transitions in ironpnictides [41].

    Figure 7(b) is a schematic of Larmor diffraction. The spinflippers are tuned to be parallel with the diffracting planesand the neutron precession directions in L1 and L2 are thesame. Assuming L1 = L2 = L, the total neutron precessionphase is φtot = 2ωLL/v. From the Bragg law |Q| = |G| =2ki sin θB,|G| = 2π/d, and the neutron velocity v = �ki/m,the total Larmor phase φtot can be written as

    φtot = 2ωLLm sin θBπ�

    d. (A7)

    Consequently, the variation of the Larmor phase is proportionalto the change of the d spacing (caused by external or thermaleffect), that is

    �φtot = φtot �dd

    . (A8)

    For d change induced thermal expansion, the evolution ofP (φtot) at different temperatures and the relative change of theφtot (�φtot) can be obtained by fitting the intensity modulationsI (φtot,0 + �φ) using Eq. (A4), where the modulations aremeasured by scanning xc4 near xc4,0 [Figs. 7(b) and 9]. The�φtot between different temperatures (or pressures/Qs) canbe used to determine the evolution of the lattice spacings.To facilitate data analysis, �φtot between two neighboringconditions should be kept within 2π . Figure 9 is an example oflattice thermal expansion at Q = (4,0,0) of BaFe1.97Ni0.03As2.The difference in xc4 (δL) between T = 113 and 118.6 K, δL,can be converted to lattice expansions according to Eqs. (A7)and (A8). In present measurements, 1 mm is equivalent to∼1 × 10−3 in �d/d, with resolution ∼1 × 10−5. Note latticeexpansion measurements are only valid for single d spacingat one Q. Systems showing coexisting multiple d spacings

    134519-10

  • IMPACT OF UNIAXIAL PRESSURE ON STRUCTURAL AND . . . PHYSICAL REVIEW B 93, 134519 (2016)

    (a)

    (b)

    Pola

    rizat

    ion

    0

    0.4

    0.8

    1.2

    (4 0 0)

    (0 4 0)

    0 1000 2000 3000 4000

    tot (rad.)

    T = 102.4 K

    0.0

    0.5

    1.0

    1.5

    2.0

    2.5

    3.0

    3.5288 K

    105.2 K

    107.5 K

    109.9 K

    Pola

    rizat

    ion

    141.3 K

    (c)

    0 1000 2000 3000 4000

    T = 4 K

    T = 121 K

    0 1000 2000 3000 4000

    Q = (4 0 0)

    tot (rad.)

    Pola

    rizat

    ion

    0

    0.4

    0.8

    1.2

    Q = (4 0 0)P ~ 10MPa

    P = 0

    P ~ 10MPa

    0Δd/d

    -0.001 0.001

    FIG. 10. Larmor diffraction measurements of P (τ ) forBaFe1.97Ni0.03As2. (a) Comparison of P (φtot) at Q = (4,0,0) betweenT = 4 K and 121 K with P ≈ 10 MPa. (b) P (φtot) for the (4,0,0)and (0,4,0) reflections measured at T = 102.4 K with P ≈ 10MPa. The inset shows the Gaussian distribution of the d spacing.(c) Temperature dependence of the P (φtot) for the (4,0,0) peak atP = 0. P (φtot) for the two d spacings shows clear modulation whichcan be fitted by Eq. (A10) (solid curves) due to the twinning causedby the tetragonal-to-orthorhombic structural transition.

    around the same Q with small differences between them cannotbe measured by this method.

    Besides the lattice expansion measurements, the P (φtot) =〈cos�φ(φtot)〉 in a wide range of φtot has been demonstratedto be the Fourier transform of the lattice spacing distribution[f (�d/d)] [47,69]. For a single-Gaussian distribution of dwith FWHM = εFW , the P (φtot) can be derived as

    P (φtot) = P0exp(

    − φ2totε

    2FW

    16ln2

    ), (A9)

    where the FWHM represents the magnitude of the d spread. Itis usually expressed in terms of �d/d. The data shown in Fig. 3of the main text are temperature dependence of the FWHM(lattice spacing spread). All of our P (φtot) for single d spacingare well described by this model, resulting in a Gaussiandistribution of the d values. Figure 10 shows the P (φtot) ofBaFe1.97Ni0.03As2 under P ≈ 10 MPa and their fittings byEqs. (A9) and (A10). A clear difference between T = 4 Kand T = 121 K in Fig. 10(a) indicates different FWHM ofthe d spread. Figure 10(b) compares P (φtot) for the (4,0,0)and (0,4,0) reflections at T = 102.4 K. The correspondingFWHMs of the d distributions are also shown as an inset. Theirdifferences suggest that the d spread along the pressure-appliedorientation is much larger.

    Figure 10(c) shows temperature dependence of P (φtot) foran unpressured BaFe1.97Ni0.03As2 sample. Its evolution at hightemperature (T > 109 K) indicates the broadening of the dspread. Below 109 K, clear modulations are seen in P (φtot).This is caused by the peak splitting of the orthorhombic (4,0,0)

    and (0,4,0) reflections in a twinned sample. For systemsshowing two or more coexisting d spacings around someQ, their d spacing distribution functions are superposition ofmultiple Gaussian distributions. In this case, the interferencebetween different d spacings will appear and can be usedto identify the specific values and spread of the involved dspacings.

    For the peak splitting (two d spacings) in BaFe1.97Ni0.03As2,interference between scattered neutrons from d1 and d2 givesrise to the modulating polarization

    P (φtot) = A√

    a2 + (1 − a)2 + 2a(1 − a)cos(φtot�ε), (A10)where

    A = P0exp(

    −φ2totε

    2FW

    16ln2

    ); (A11)

    here we assume both d spacings have the same εFW . a and(1 − a) denote the populations of the d1 and d2;

    �ε = d1 − d2(d1 + d2)/2 (A12)

    is the lattice distortion. The definition of �ε is similar tothe orthorhombic lattice distortion δ = (a − b)/(a + b) in ironpnictides, with �ε = 2δ [5]. The P (φtot) at T = 107.5 and105.2 K in Fig. 10(c) are well described by Eq. (A10). Thefitted lattice distortions and d spreads are shown in Figs. 3and 4 of the main text. The resolution in determining �ε heredepends on the range of φtot and the d spread of the samplesince the dips of the polarization are critical for fitting �ε. Theresolution of �ε for two d spacings is 7 × 10−4 in the presentwork. Thus the possible distortions at temperatures slightlylower than Ts = 114 K [such as the 109.5-K data shown inFig. 10(c)] in BaFe1.97Ni0.03As2 cannot be distinguished fromthe broadening of the d spread.

    Figure 11 is P (φtot) for temperatures across the structuraltransition of the BaFe2As2 sample. P (φtot) in Fig. 11(a)is a beating pattern caused by interference between two dspacings below Ts , similar with that shown in Fig. 10(c).The corresponding d spacing distributions are shown inFig. 11(e). The orthorhombic distortions can be determinedas δ = (d2 − d1)/(d2 + d1). Upon warming the sample toT = 137 K, a temperature slightly lower than Ts , a morecomplicated pattern [Fig. 11(b)] indicates the coexistence offour d spacings [Fig. 11(f)]. This is consistent with the coex-isting orthorhombic antiferromagnetic (δ1) and orthorhombicparamagnetic (δ2) phases revealed by high-resolution x-ray-diffraction measurements [5]. The four-d spacing model ofP (τ ) can be derived analytically (not shown here) and fits thedata very well. Figures 11(c) and 11(g) are results for T ∼ Ts ,where the δ2 is indistinguishable and only one broad d spreadcan be fitted. Here, the orthorhombic antiferromagnetic phase(δ1) is about to disappear, suggesting this temperature is nearTs . For temperature higher than Ts in Figs. 11(d) and 11(h),only one d spacing is observed, indicating the system entersinto the paramagnetic tetragonal phase. In Figs. 11(a) and11(b), the magnitude of the lattice distortions determines thebeating periods (overall line shape) and the relative populationsof different d spacings control whether the polarization canreach zero at dips. In the present study, the overall line shapesof all P (τ ) are well fitted by specific multiple (2–4) d spacing

    134519-11

  • XINGYE LU et al. PHYSICAL REVIEW B 93, 134519 (2016)

    0.2

    0.4

    0.6

    0.8

    Pola

    rizat

    ion

    (a)

    0 1 2 3 4 5

    0.0

    0.2

    0.4

    0.6

    0.8

    0.0

    0.2

    0.4

    0.6

    0.8

    0.0

    0.2

    0.4

    0.6

    0.8

    0.0

    0.2

    0.4

    0.6

    0.8

    0.0

    0.2

    0.4

    0.6

    0.8

    0.0

    T = 136 K

    T = 137 K

    T = 138 K~TS

    T = 139 K

    φtot (103 rad.)

    Pola

    rizat

    ion

    (b)

    Pola

    rizat

    ion

    (c)

    Pola

    rizat

    ion

    (d)

    0.2

    0.4

    0.6

    0.0

    0.996 1 1.004d/dave

    0.1

    0.2

    0.3

    0.0

    0.2

    0.4

    0.8

    0.0

    0.6

    1.0

    0.2

    0.4

    0.8

    0.0

    0.6

    1.0

    Inte

    nsity

    (arb

    . u.)

    Inte

    nsity

    (arb

    . u.)

    Inte

    nsity

    (arb

    . u.)

    Inte

    nsity

    (arb

    . u.)

    Δd

    2δ12δ2

    (e)

    (f)

    (g)

    (h)

    d2d1

    FIG. 11. (a)–(d) Precession phase (φtot) dependent polarizationsacross the structural transition for BaFe2As2. The solid green curvesare fits of the P (φtot) by one or multi-Gaussian d spacing distributionmodels. (e)–(h) d spacing distributions in d space. The orthorhombic,coexisting two different orthorhombic and tetragonal phases can bedetermined for T = 136–139 K, respectively.

    models [green curves in Figs. 11(a)–11(d)], indicating that thelattice distortions are well determined.

    The orthorhombic lattice distortions for BaFe2As2 andSrFe1.97Ni0.03As2 obtained from Larmor diffraction measure-ments of P (φtot) are shown in Fig. 12. These results areconsistent with previous results measured by x-ray diffraction[5]. The error bars in Figs. 3 and 4 of the main text are fittingerrors of the raw data at different temperatures according toformulas discussed above.

    4. d spread anisotropy between a and b

    Our another interesting discovery is the doping dependent dspread anisotropy under uniaxial pressure. The samples shownin Fig. 13 exhibit similar temperature dependence of the dspread between (4,0,0) and (0,4,0), suggesting that the dif-ference of d spread between a and b is trivial. However, we notethat the FWHM of (0,4,0), along the uniaxial pressure direc-tion, is much larger than a in underdoped samples [Figs. 13(a)–13(c)]. This may be attributed to an inhomogeneous distribu-tion of the pressure induced strain field. However, we find verysmall differences in d spread between (4,0,0) and (0,4,0) inthe overdoped BaFe1.88Ni0.12As2 [Fig. 13(d)], suggesting the

    130 135 140 145 150

    0

    1

    2

    3

    δ =

    (a-b

    )/(a

    +b) (

    10-3

    )Temperature (K)

    125

    (a)

    100 120 140 160 180 200 220 240 260

    0

    1

    2

    3

    δ =

    (a-b

    )/(a

    +b) (

    10-3

    )

    (b)

    4

    5

    BaFe2As2

    SrFe1.97Ni0.03As2

    TN = TS = 138(1)K

    TN = TS = 175(2)K

    FIG. 12. Temperature-dependent orthorhombic lattice distortionsfor BaFe2As2 and SrFe1.97Ni0.03As2. The open green diamond in (a)marks the temperature range showing four d spacings. The open reddiamonds in (b) show the persistence of the tetragonal phase into theorthorhombic phase, indicative of a first-order structural transition,consistent with previous reports. The vertical blue dashed lines markthe structural transitions.

    d spread anisotropy between a and b is nontrivial and maybe associated with antiferromagnetic/structural instability oreven nematic susceptibility in underdoped samples.

    5. Lattice distortions and Young’s modulus

    The Young’s modulus Y along the b axis (∼C66) can beestimated by Y = P/δ, where δ is pressure induced latticedistortion. At ∼250 K, the Y for BaFe2As2, BaFe1.97Ni0.03As2,and BaFe1.88Ni0.12As2 estimated from our neutron Larmordiffraction experiments are ∼50, ∼50, and ∼100 GPa, re-spectively. Compared with the shear modulus C66 obtainedby ultrasound spectroscopy [52], the estimated Y for x = 0and x = 0.03 are ∼30% larger. These differences are mainlycaused by the errors in our estimation of the applied pressure Pthrough measuring compressed spring distances and estimatedspring constant [41]. However, they will not affect temperaturedependence of the pressure-induced FHWM of �d/d and itscomparison with other iron pnictides, and thus will not alterthe conclusions of our experiments.

    6. Landau theory and effect of magnetismon nematicity and strain

    In order to understand the distinct behavior of the ob-served lattice distortion in SrFe1.97Ni0.03As2compared toBaFe1.97Ni0.03As2 [see Fig. 1(c) in the main text], we writedown the Landau free energy incorporating the electronicnematic order parameter ϕ, coupled magnetoelastically to the

    134519-12

  • IMPACT OF UNIAXIAL PRESSURE ON STRUCTURAL AND . . . PHYSICAL REVIEW B 93, 134519 (2016)

    (4 0 0), P ~15MPa

    100 150 200 250 300

    Temperature (K)

    FWH

    M (Δ

    d/d

    10-4

    )

    5

    15

    20

    25

    30

    FWH

    M (Δ

    d/d

    10-4

    )

    140 160 180 200 220 240 260

    4

    8

    12

    16

    20

    FWH

    M (Δ

    d/d

    10-4

    )

    40 80 120 160 200 240

    4007

    8

    10

    11

    FWH

    M (Δ

    d/d

    10-4

    )

    BaFe2As2

    SrFe1.97Ni0.03As2

    BaFe1.97Ni0.03As2

    BaFe1.88Ni0.12As2

    (4 0 0), P ~ 20MPa

    (a)

    (b)

    (c)

    (d)

    TN=TS=138(1)K

    T=144(2)K

    TN=TS=175(2)K

    T=180(2)K

    TN=109(2)KTS=114(2)K

    T=120(2)K

    5

    10

    15

    20

    25

    (0 4 0), P ~ 20MPa

    10

    (4 0 0), P ~ 20MPa(0 4 0), P ~ 20MPa

    (0 4 0), P ~15MPa

    9

    12080 160 200 240

    (4 0 0), P ~14MPa(0 4 0), P ~14MPa

    FIG. 13. Temperature and doping dependent d spread ofBaFe2−xNixAs2 and SrFe1.97Ni0.03As2 measured with finite uniaxialpressure. TN and Ts are marked as blue (and green) vertical dashedlines. The red dashed line marks the temperature with the maximumof the FWHM.

    lattice distortion δ ∝ ε6, as well as to the antiferromagneticorder parameter M:

    F = F0 + T0f [ϕ,δ] + T0f̃ [M,ϕ], (A13)where

    f [ϕ,δ] = a2

    T − T0T0

    ϕ2 + B4T0

    ϕ4 + C66,02T0

    δ2 − λT0

    δϕ − PδT0

    .

    (A14)Here we chose to normalize the free energy by the Curie-Weisstemperature T0 associated with the quadratic ϕ2 term (if T0is negative, it is replaced by |T0|). Treating the electronicnematic order parameter ϕ as a dimensionless variable, thishas an advantage that all the coefficients in the free energy are

    TABLE I. Curie-Weiss fitting parameters of the pressure-inducedlattice distortions in the tetragonal state; see Fig. 1(c) in the main text.

    Sample T CWs (K) T0(K) TCWs − T0(K)

    BaFe2As2 134.9 ± 0.3 51.3 ± 8.7 83.5 ± 8.7SrFe1.97Ni0.03As2 170.9 ± 1 135.7 ± 6 35.2 ± 6.1

    dimensionless (here we choose, without loss of generality, a=B = 1). The remaining Landau expansion parameters can befixed from the experiment. Indeed, it is convenient to expressthe external uniaxial stress P in terms of the dimensionlessstress variable σ = P/C66,0. Then, the last three terms inEq. (A14) can be written as follows:

    λ

    |T0|[(

    C66,0

    λ

    )(δ2

    2− δσ

    )− δϕ

    ]. (A15)

    Minimizing the free energy with respect to δ, we find

    δ = λC66,0

    ϕ + σ. (A16)

    From the minimization with respect to ϕ, it is easy to obtain

    dδ= λ

    a(T − T0) + 3Bφ2 (A17)

    and now the shear modulus C66 ≡ dP/dδ = d2F/dδ2 be-comes

    C66 = C66,0 − λdϕdδ

    = C66,0 − λ2

    a(T − T0) + 3Bφ2 ; (A18)

    in other words the elastic modulus gets renormalized fromits bare value C66,0 by virtue of the elastonematic couplingλ. Equivalently, it follows from the above equation that theinverse nematic susceptibility χ−1ϕ ≡ d2F/dϕ2|ϕ→0 also getsrenormalized from its bare value χ−1φ = a(T − T0):

    χ̃−1ϕ = χ−1φ −λ2

    C66,0= a(T − T CWs ), (A19)

    where T CWs = T0 + λ2

    aC66,0is the renormalized Curie-Weiss

    temperature. One can now cast Eq. (A18) above the transitiontemperature Ts = T CWs into the form

    dP≡ 1

    C66= 1

    C66,0

    T − T0T − T CWs

    ,(T > T CWs

    ), (A20)

    which is Eq. (3) in the main text. We now use this Eq. (A20) tofit the data for the pressure-induced distortion δ(P ) − δ(0) ≈P (dδ/dP ) [Fig. 1(c) in the main text], extracting the valuesof T0 and T CWs which we quote in Table I for BaFe2As2 andSrFe1.97Ni0.03As2.

    We now turn to the question of the strength of theelastonematic coupling constant λ. The unknown dimension-less parameter r = C66,0/λ in Eq. (A16) can be fixed fromthe ratio ϕ/δ in zero external stress (σ = 0). Substitutingthe typical value of δ ∼ 3 × 10−3 in SrFe1.97Ni0.03As2andBaFe1.97Ni0.03As2 and choosing the value of the nematicorder parameter ϕ = 1 deep inside the nematic phase forconvenience, we find r ∼ 300. As for the value of λ̃ = λ/|T0|itself, it can also be fixed from the experiment since λ enters

    134519-13

  • XINGYE LU et al. PHYSICAL REVIEW B 93, 134519 (2016)

    in Eq. (A19) to renormalize the Curie-Weiss temperature.

    Therefore, one obtains λ̃ = a (T CWs −T0)|T0| r . Plugging in the valuesof T CWs and T0 from our fittings of the lattice distortions(Table I), we obtain λ̃ ≈ 80 for SrFe1.97Ni0.03As2and λ̃ ≈ 490for BaFe2As2, in other words the effective electron-lattice cou-pling is about ∼6 times weaker in SrFe1.97Ni0.03As2comparedto the BaFe2−xTxAs2 compounds. For BaFe1.97Ni0.03As2, thequality of our data was insufficient to accurately determinethe bare Curie-Weiss temperature T0 (we were only able todetermine T CWs = 88.5 ± 1.0 K). However, from the esti-mated T CWs − T0 ≈ 50 K by the elastic measurements [38],we can deduce the approximate value of the coupling constantλ̃ ≈ 390, similar in magnitude to undoped BaFe2As2.

    We now turn our attention to the magnetonematic coupling.On symmetry grounds, nematic order parameter must coupleto M2 (since magnetization breaks time-reversal symmetry,and ϕ does not). This can be shown explicitly by consideringthe magnetization MA,B on the two sublattices composedof the next-nearest-neighbor sites of the square lattice, inwhich case the nematic order parameter couples linearlyto the product (MA · MB) [25,28,36,70]. Note that thisconclusion holds independently of whether the microscopicorigin of nematicity is purely magnetic [25,28,70] or orbital[31–33,36,71]. The magnetic phase transition itself may beintrinsically second order, as in BaFe2−xTxAs2 compounds, orfirst order, as in SrFe1.97Ni0.03As2, Ba1−x(K,Na)xFe2As2, andCa1−xLaxFe2As2. Below we consider both possibilities:

    f̃1[M,ϕ] = u2

    (T − TN,0)M2 − v4M4 + w

    6M6 − μϕM2,

    (A21)

    f̃2[M,ϕ] = u2

    (T − TN,0)M2 + v4M4 − μϕM2. (A22)

    Since we are after the qualitative consequences of themagnetonematic coupling, the precise values of the Landaucoefficients are not essential (we take u = v = w = 1 andμ = 0.1 for concreteness).

    Having introduced the Landau formalism above, we nowstudy the effect of the applied external stress P on thebehavior of the lattice distortion. The calculated temperaturedependence of δ(P ) − δ(0) is shown in Fig. 14 for the realisticstrain P = 20 MPa and is shown to depend crucially on thenature of the magnetic phase transition. Indeed, the onlydifference between the two curves is the sign in front ofthe quartic M4 terms in Eqs. (A21) and (A22), while all theother Landau expansion parameters are kept the same (the twocurves are offset horizontally for clarity). Note that for smallP , δ(P ) − δ(0) ≈ (dδ/dP )P is proportional to the nematicsusceptibility, which is expected to diverge at T CWs accordingto Eq. (A20). Both curves in Fig. 14 exhibit an enhancednematic susceptibility on approaching T CWs , as expected. Themain difference is the shape of the curve on approachingthe transition, which has a distinct asymmetric “λ” shapein the case of the second-order magnetic transition, andresembles closely the experimentally measured δ(P ) − δ(0)for BaFe1.97Ni0.03As2 in Fig. 1(c) (see main text). By contrast,the Néel transition is first order in SrFe1.97Ni0.03As2, and theexperimental behavior in Fig. 1(c) is close to the calculated

    0 0.5 1 1.5 20

    0.5

    1

    1.5

    2

    2.5 x 10−3

    Temperature T/ T0

    δ(P

    =20

    MP

    a) −

    δ(P

    =0)

    1st order2nd order

    FIG. 14. Change in the lattice distortion as a function of temper-ature, calculated from the Landau theory assuming either first-order[Eq. (A21)] or second-order [Eq. (A22)] magnetic phase transitioncoupled to the nematic order parameter ϕ. The blue dashed curve isshifted to the left by �T = 0.2T0 for clarity. These results should becompared to the neutron data in Fig. 1(c) in the main text.

    sharp increase seen in our model (solid line in Fig. 14). There-fore, the Landau free energy results corroborate our conclusionthat the nature of the magnetic transition is crucial to theobserved temperature dependence of the lattice distortion.

    We note in passing that for sufficiently strong couplingconstant μ, the magnetic transition becomes weakly first ordereven if the intrinsic free energy has a positive M4 term inEq. (A22). This is likely the explanation for the observedchange of the nature of the magnetic transition from weaklyfirst order in BaFe2As2 to second order upon Co doping [5,7,8].However for the values of the coupling constants in Fig. 14,this effect is imperceptible and the main difference betweenthe two curves is due to the different intrinsic nature of themagnetic phase transition depending on the sign of the quarticterm in Eqs. (A21) and (A22). We have verified that for thesignificantly larger values of the coupling constant [μ � 0.4in Eq. (A22)], it is indeed possible to obtain the shape similarto the dashed line in Fig. 14 because the magnetic transitionbecomes effectively first order. In either case, our conclusionsremain intact.

    7. Interpretation of the resistivity anisotropy

    The resistivity anisotropy �ρ = (ρa − ρb)/(ρa + ρb) hasbeen widely used as a proxy for the electronic nematic orderparameter in the iron pnictides [12,20]. However in somecompounds, in particular in SrFe1.97Ni0.03As2, the resistivityanisotropy is vanishingly small immediately above TN [seeFig. 1(b) in the main text], whereas it is much larger inBaFe2−xTxAs2. This is puzzling because the lattice distortionis comparable in both cases [Fig. 1(c)] and, according toEq. (A16), one expects the lattice distortion δ to be proportionalto the nematic order parameter.

    To shed more light on this apparent inconsistency, we haveplotted in Fig. 15 the temperature dependence of the nematicorder parameter ϕ under the uniaxial stress P = 20 MPa.The two curves correspond to the first- and second-order

    134519-14

  • IMPACT OF UNIAXIAL PRESSURE ON STRUCTURAL AND . . . PHYSICAL REVIEW B 93, 134519 (2016)

    0 0.5 1 1.5 20

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4

    Temperature T/ T0

    Nem

    atic

    O.P

    . φ a

    t P =

    20

    MP

    a

    1st order2nd order

    FIG. 15. Calculated temperature dependence of the nematic orderparameter ϕ coupled to either first- or second-order magneticphase transition. Note that in both cases, the bare ϕ4 term ispositive, however the nematic transition is rendered first order (solidcurve) when coupled to the first-order magnetic order parameter inEq. (A21).

    nature of the magnetic transition, respectively, and the Landauparameters were kept the same in both cases (except forthe sign of the quartic term in Eqs. (A21) and (A22).Above the transition temperature, T > Ts , the values of ϕare predictably small, but importantly, they are identical in thetwo cases. In fact, the main difference lies in the temperaturedependence immediately below Ts . From Fig. 15, it wouldappear that in this regime, the nematic order parameter shouldbe smaller for the second-order phase transition, howeverthis is diametrically opposite from the comparison betweenBaFe1.97Ni0.03As2 and SrFe2As2 in Fig. 1(b) (see the maintext), where the magnetic transition in BaFe1.97Ni0.03As2 issecond order, yet resistivity anisotropy is much larger. Thisqualitative observation can be made sharper by consideringEq. (A16), where the coupling constant λ is estimated fromexperiment to be a factor of ∼5 larger in BaFe1.97Ni0.03As2andBaFe2As2 compared to SrFe1.97Ni0.03As2, whereas the elastic

    modulus C66,0 ≈ 50 GPa is similar in all three materials. Then,BaFe2As2 is expected to have at least a factor of 5 largerlattice distortion compared to SrFe1.97Ni0.03As2, assumingthat ϕ is the same in both materials. If one now equatesthe resistivity anisotropy with the nematic order parameterϕ, as has commonly been done in the literature [12,20], thenone is forced to conclude that ϕ must be about four timeslarger in BaFe2As2 due to the larger resistivity anisotropy[see Fig. 1(b)]. Taken together, one would expect the latticedistortion δ to be a factor of ∼20 larger in BaFe2As2 and inBaFe1.97Ni0.03As2 compared to SrFe1.97Ni0.03As2 (a factor of4 due to larger resistivity anisotropy, times a factor of 5 dueto larger λ). And yet this clearly contradicts the experimentalevidence in Fig. 1(c), according to which the lattice distortionis almost the same in all three materials.

    One possible way out of this dilemma is that the Landautheory may not be applicable to describe the nematicity inthe pnictides. However, given the excellent semiquantitativeagreement that Landau theory provides for the lattice distortion(Fig. 14 above) and its well documented success describing theelastic shear modulus measurements [38], such a conclusionis perhaps not well justified. Rather, a much more plausibleconclusion is that resistivity anisotropy is a poor substitutefor the nematic order parameter. While it is plausible that thetwo quantities are proportional to each other, as follows fromthe nematic susceptibility measurements [20], the coefficientof proportionality need not be constant and can have a strongtemperature dependence (and likely material dependence), assuggested recently by Tanatar et al. in the recent study onFeSe [61]. This material displays a nonmonotonic temperaturedependence of the resistivity anisotropy with a peak belowTs , qualitatively similar to BaFe2As2 . Further theoretical andexperimental studies are necessary to elucidate the preciserelationship of the resistivity anisotropy and the nematicorder parameter in the iron pnictides and chalcogenides.Direct microscopic measurements of the nematic order pa-rameter, for instance using the angle-resolved photoemissionspectroscopy (ARPES) to probe the orbital splitting, com-bined with the uniaxial pressure measurements, would bedesirable.

    [1] Y. Kamihara, T. Watanabe, M. Hirano, and H. Hosono, J. Am.Chem. Soc. 130, 3296 (2008).

    [2] C. de la Cruz, Q. Huang, J. W. Lynn, J. Li, W. Ratcliff, II, J. L.Zarestky, H. A. Mook, G. F. Chen, J. L. Luo, N. L. Wang, andP. C. Dai, Nature (London) 453, 899 (2008).

    [3] Q. Huang, Y. Qiu, Wei Bao, M. A. Green, J. W. Lynn, Y. C.Gasparovic, T. Wu, G. Wu, and X. H. Chen, Phys. Rev. Lett.101, 257003 (2008).

    [4] J. Zhao, W. Ratcliff, J. W. Lynn, G. F. Chen, J. L. Luo, N. L.Wang, J. P. Hu, and P. C. Dai, Phys. Rev. B 78, 140504(R)(2008).

    [5] M. G. Kim, R. M. Fernandes, A. Kreyssig, J. W. Kim, A. Thaler,S. L. Bud’ko, P. C. Canfield, R. J. McQueeney, J. Schmalian,and A. I. Goldman, Phys. Rev. B 83, 134522 (2011).

    [6] P. C. Dai, Rev. Mod. Phys. 87, 855 (2015).[7] C. Lester, J.-H. Chu, J. G. Analytis, S. C. Capelli, A. S. Erickson,

    C. L. Condron, M. F. Toney, I. R. Fisher, and S. M. Hayden,Phys. Rev. B 79, 144523 (2009).

    [8] S. Nandi, M. G. Kim, A. Kreyssig, R. M. Fernandes, D. K. Pratt,A. Thaler, N. Ni, S. L. Bud’ko, P. C. Canfield, J. Schmalian, R.J. McQueeney, and A. I. Goldman, Phys. Rev. Lett. 104, 057006(2010).

    [9] H. Q. Luo, R. Zhang, M. Laver, Z. Yamani, M. Wang, X. Y. Lu,M. Y. Wang, Y. C. Chen, S. L. Li, S. Chang, J. W. Lynn, and P.C. Dai, Phys. Rev. Lett. 108, 247002 (2012).

    [10] X. Y. Lu, H. Gretarsson, R. Zhang, X. Liu, H. Q. Luo, W. Tian,M. Laver, Z. Yamani, Y.-J. Kim, A. H. Nevidomskyy, Q. Si, andP. C. Dai, Phys. Rev. Lett. 110, 257001 (2013).

    [11] R. W. Hu, S. L. Bud’ko, W. E. Straszheim, and P. C. Canfield,Phys. Rev. B 83, 094520 (2011).

    [12] J. H. Chu, J. G. Analytis, K. De Greve, P. L. McMahon, Z. Islam,Y. Yamamoto, and I. R. Fisher, Science 329, 824 (2010).

    [13] M. A. Tanatar, E. C. Blomberg, A. Kreyssig, M. G. Kim, N.Ni, A. Thaler, S. L. Bud’ko, P. C. Canfield, A. I. Goldman,I. I. Mazin, and R. Prozorov, Phys. Rev. B 81, 184508(2010).

    134519-15

    http://dx.doi.org/10.1021/ja800073mhttp://dx.doi.org/10.1021/ja800073mhttp://dx.doi.org/10.1021/ja800073mhttp://dx.doi.org/10.1021/ja800073mhttp://dx.doi.org/10.1038/nature07057http://dx.doi.org/10.1038/nature07057http://dx.doi.org/10.1038/nature07057http://dx.doi.org/10.1038/nature07057http://dx.doi.org/10.1103/PhysRevLett.101.257003http://dx.doi.org/10.1103/PhysRevLett.101.257003http://dx.doi.org/10.1103/PhysRevLett.101.257003http://dx.doi.org/10.1103/PhysRevLett.101.257003http://dx.doi.org/10.1103/PhysRevB.78.140504http://dx.doi.org/10.1103/PhysRevB.78.140504http://dx.doi.org/10.1103/PhysRevB.78.140504http://dx.doi.org/10.1103/PhysRevB.78.140504http://dx.doi.org/10.1103/PhysRevB.83.134522http://dx.doi.org/10.1103/PhysRevB.83.134522http://dx.doi.org/10.1103/PhysRevB.83.134522http://dx.doi.org/10.1103/PhysRevB.83.134522http://dx.doi.org/10.1103/RevModPhys.87.855http://dx.doi.org/10.1103/RevModPhys.87.855http://dx.doi.org/10.1103/RevModPhys.87.855http://dx.doi.org/10.1103/RevModPhys.87.855http://dx.doi.org/10.1103/PhysRevB.79.144523http://dx.doi.org/10.1103/PhysRevB.79.144523http://dx.doi.org/10.1103/PhysRevB.79.144523http://dx.doi.org/10.1103/PhysRevB.79.144523http://dx.doi.org/10.1103/PhysRevLett.104.057006http://dx.doi.org/10.1103/PhysRevLett.104.057006http://dx.doi.org/10.1103/PhysRevLett.104.057006http://dx.doi.org/10.1103/PhysRevLett.104.057006http://dx.doi.org/10.1103/PhysRevLett.108.247002http://dx.doi.org/10.1103/PhysRevLett.108.247002http://dx.doi.org/10.1103/PhysRevLett.108.247002http://dx.doi.org/10.1103/PhysRevLett.108.247002http://dx.doi.org/10.1103/PhysRevLett.110.257001http://dx.doi.org/10.1103/PhysRevLett.110.257001http://dx.doi.org/10.1103/PhysRevLett.110.257001http://dx.doi.org/10.1103/PhysRevLett.110.257001http://dx.doi.org/10.1103/PhysRevB.83.094520http://dx.doi.org/10.1103/PhysRevB.83.094520http://dx.doi.org/10.1103/PhysRevB.83.094520http://dx.doi.org/10.1103/PhysRevB.83.094520http://dx.doi.org/10.1126/science.1190482http://dx.doi.org/10.1126/science.1190482http://dx.doi.org/10.1126/science.1190482http://dx.doi.org/10.1126/science.1190482http://dx.doi.org/10.1103/PhysRevB.81.184508http://dx.doi.org/10.1103/PhysRevB.81.184508http://dx.doi.org/10.1103/PhysRevB.81.184508http://dx.doi.org/10.1103/PhysRevB.81.184508

  • XINGYE LU et al. PHYSICAL REVIEW B 93, 134519 (2016)

    [14] I. R. Fisher, L. Degiorgi, and Z. X. Shen, Rep. Prog. Phys. 74,124506 (2011).

    [15] S. Ishida, M. Nakajima, T. Liang, K. Kihou, C. H. Lee, A.Iyo, H. Eisaki, T. Kakeshita, Y. Tomioka, T. Ito, and S. Uchida,Phys. Rev. Lett. 110, 207001 (2013).

    [16] E. C. Blomberg, M. A. Tanatar, A. Kreyssig, N. Ni, A. Thaler,R. Hu, S. L. Bud’ko, P. C. Canfield, A. I. Goldman, and R.Prozorov, Phys. Rev. B 83, 134505 (2011).

    [17] E. C. Blomberg, A. Kreyssig, M. A. Tanatar, R. M. Fernandes,M. G. Kim, A. Thaler, J. Schmalian, S. L. Bud’ko, P. C. Canfield,A. I. Goldman, and R. Prozorov, Phys. Rev. B 85, 144509(2012).

    [18] S. D. Das, M. S. Laad, L. Craco, J. Gillett, V. Tripathi, and S. E.Sebastian, Phys. Rev. B 92, 155112 (2015).

    [19] E. Fradkin, S. A. Kivelson, M. J. Lawler, J. P. Eisenstein, and A.P. Mackenzie, Annu. Rev. Condens. Matter Phys. 1, 153 (2010).

    [20] J. H. Chu, H.-H. Kuo, J. G. Analytis, and I. R. Fisher, Science337, 710 (2012).

    [21] S. Kasahara, H. J. Shi, K. Hashimoto, S. Tonegawa, Y.Mizukami, T. Shibauchi, K. Sugimoto, T. Fukuda, T. Terashima,A. H. Nevidomskyy, and Y. Matsuda, Nature (London) 486, 382(2012).

    [22] Y. Gallais, R. M. Fernandes, I. Paul, L. Chauviere, Y.-X. Yang,M.-A. Measson, M. Cazayous, A. Sacuto, D. Colson, and A.Forget, Phys. Rev. Lett. 111, 267001 (2013).

    [23] H.-H. Kuo and I. R. Fisher, Phys. Rev. Lett. 112, 227001 (2014).[24] H.-H. Kuo, J.-H. Chu, S. A. Kivelson, and I. R. Fisher,

    arXiv:1503.00402.[25] R. M. Fernandes, E. Abrahams, and J. Schmalian, Phys. Rev.

    Lett. 107, 217002 (2011).[26] R. M. Fernandes, A. V. Chubukov, and J. Schmalian, Nat. Phys.

    10, 97 (2014).[27] P. Chandra, P. Coleman, and A. I. Larkin, Phys. Rev. Lett. 64,

    88 (1990).[28] C. Fang, H. Yao, W.-F. Tsai, J. P. Hu, and S. A. Kivelson,

    Phys. Rev. B 77, 224509 (2008).[29] J. Dai, Q. Si, J. X. Zhu, and E. Abrahams, Proc. Natl. Acad. Sci.

    USA 106, 4118 (2009).[30] M. Yi, D. H. Lu, J.-H. Chu, J. G. Analytis, A. P. Sorini, A.

    F. Kemper, B. Moritz, S.-K. Mo, R. G. Moore, M. Hashimoto,W.-S. Lee, Z. Hussain, T. P. Devereaux, I. R. Fisher, and Z.-X.Shen, Proc. Natl. Acad. Sci. USA 108, 6878 (2011).

    [31] C. C. Lee, W. G. Yin, and W. Ku, Phys. Rev. Lett. 103, 267001(2009).

    [32] F. Krüger, S. Kumar, J. Zaanen, and J. van den Brink, Phys. Rev.B 79, 054504 (2009).

    [33] W. C. Lv, J. S. Wu, and P. Phillips, Phys. Rev. B 80, 224506(2009).

    [34] C.-C. Chen, J. Maciejko, A. P. Sorini, B. Moritz, R. R. P. Singh,and T. P. Devereaux, Phys. Rev. B 82, 100504(R) (2010).

    [35] B. Valenzuela, E. Bascones, and M. J. Calderón, Phys. Rev. Lett.105, 207202 (2010).

    [36] Z. Wang and A. H. Nevidomskyy, J. Phys.: Condens. Matter 27,225602 (2015).

    [37] H. Kontani and Y. Yamakawa, Phys. Rev. Lett. 113, 047001(2014).

    [38] A. E. Böhmer and C. Meingast, C. R. Phys. 17, 90 (2016).[39] C. Dhital, Z. Yamani, W. Tian, J. Zeretsky, A. S. Sefat, Z. Q.

    Wang, R. J. Birgeneau, and S. D. Wilson, Phys. Rev. Lett. 108,087001 (2012).

    [40] C. Dhital, T. Hogan, Z. Yamani, R. J. Birgeneau, W. Tian, M.Matsuda, A. S. Sefat, Z. Wang, and S. D. Wilson, Phys. Rev. B89, 214404 (2014).

    [41] X. Y. Lu, J. T. Park, R. Zhang, H. Q. Luo, A. N. Nevidomskyy,Q. Si, and P. C. Dai, Science 345, 657 (2014).

    [42] M. P. Allan, T-M. Chuang, F. Massee, Yang Xie, N. Ni, S. L.Bud’ko, G. S. Boebinger, Q. Wang, D. S. Dessau, P. C. Canfield,M. S. Golden, and J. C. Davis, Nat. Phys. 9, 220 (2013).

    [43] E. P. Rosenthal, E. F. Andrade, C. J. Arguello, R. M. Fernandes,L. Y. Xing, X. C. Wang, C. Q. Jin, A. J.