+ All Categories
Home > Documents > 5 arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013 · 2018. 11. 23. · Field-angle-resolved...

5 arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013 · 2018. 11. 23. · Field-angle-resolved...

Date post: 22-Jan-2021
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
13
Field-angle-resolved anisotropy in superconducting CeCoIn 5 using realistic Fermi surfaces Tanmoy Das, 1 A. B. Vorontsov, 2 I. Vekhter, 3 , and Matthias J. Graf 1 1 Los Alamos National Laboratory, Los Alamos, New Mexico 87545 2 Department of Physics, Montana State University, Bozeman, Montana 59717 3 Department of Physics and Astronomy, Louisiana State University, Baton Rouge, Louisiana 70803 (Dated: November 23, 2018) We compute the field-angle-resolved specific heat and thermal conductivity using realistic model band struc- ture for the heavy-fermion superconductor CeCoIn5 to identify the gap structure and location of nodes. We use a two-band tight-binding parametrization of the band dispersion as input for the self-consistent calculations in the quasiclassical formulation of the superconductivity. Systematic analysis shows that modest in-plane anisotropy in the density of states and Fermi velocity in tetragonal crystals significantly affects the fourfold oscillations in thermal quantities, when the magnetic field is rotated in the basal plane. The Fermi surface anisotropy sub- stantially shifts the location of the lines in the H-T plane, where the oscillations change sign compared to quasicylindrical model calculations. In particular, at high fields, the anisotropy and sign reversal are found even for isotropic gaps. Our findings imply that a simultaneous analysis of the specific heat and thermal conductivity, with an emphasis on the low energy sector, is needed to restrict potential pairing scenarios in multiband super- conductors. We discuss the impact of our results on recent measurements of the Ce-115 family, namely CeT In5 with T =Co,Rh,Ir. PACS numbers: 74.25.Uv,74.20.Rp,74.25.Bt,74.25.fc I. INTRODUCTION Many heavy-fermion and other novel superconductors are thought to possess nodes in the gap function on the Fermi sur- face. Since the gap shape is directly related to the symme- try of the pairing interaction, knowing the position of nodes can shed light on possible pairing mechanisms. Magnetic field-angle-resolved specific heat and thermal conductivity ex- periments are able to provide detailed information about the anisotropy of quasiparticle excitations near the Fermi surface, and hence help identify the nodal directions in the bulk. 1–4 To implement this procedure it is necessary to have high- precision probes that detect small variations under changes of the direction of the applied field. A series of remarkable experiments proved already the viability of this approach. 4–7 However it has proved non-trivial to interpret these exper- iments in general. The oscillations in physical quantities, as a function of the field direction, change sign depend- ing on the magnitude of the applied magnetic field and the temperature. 3,8,9 The location of these inversion lines de- pends sensitively on the topology of the Fermi surface and the material-specific details of the (multi-) band structure. 10,11 Obviously, this calls for the development of theoretical tools that take material-specific properties into account. Further- more, it suggests that a quantitative and unambiguous identi- fication of the structure of the superconducting (SC) gap re- quires the incorporation of realistic Fermi surface (FS) prop- erties. The unconventional heavy-fermion superconductor CeCoIn 5 is an ideal candidate for testing field-angle-resolved probes due to the existence of large high-quality crystals and accessible temperature and field ranges. Early field-angle- resolved thermal conductivity and specific heat measurements were controversial on whether CeCoIn 5 has a superconduct- ing gap with d x 2 -y 2 or d xy symmetry. 12,13 Recent specific heat measurements observed the predicted inversion of the oscillations at low temperature. 6 This seemed to have settled the dispute in favor of d x 2 -y 2 pairing symmetry. In this paper we incorporate first-principles electronic structure calculations to obtain the realistic tight-binding parametrization for Ce-115 (CeT In 5 with T =Co,Rh,Ir) ma- terials that reproduce the Fermi surface (FS) topology and yield the Fermi velocities, and the density of states (DOS) at the Fermi level. We use this FS parametrization as in- put for self-consistent calculations of thermal properties in the extended Brandt-Pesch-Tewordt (BPT) approximation of the quasiclassical Eilenberger equation. 8,9 Use of the tight- binding parametrization allows for a numerically efficient computation, while keeping the essential character of the low- energy band structure that reflects on the hybridization be- tween Ce 4f and In 5p states. Within this framework, we consider candidate s- and d-wave order parameters, perform a systematic study of the angle-resolved specific heat coeffi- cient, γ = C/T , and thermal conductivity, κ, in a magnetic field rotating in the Ce-In basal plane. Finally, we construct a field-temperature phase diagram of the fourfold oscillations. The main results of our calculations, which are applicable to a wide range of systems with tetragonal point group sym- metry, are: (1) For isotropic gap (s-wave) we find that mod- erate FS anisotropies are sufficient to introduce field-angle- dependent oscillations in the specific heat and thermal con- ductivity in the superconducting state over a significant range of temperatures and at intermediate to high magnetic fields. In addition, the inversion of the oscillation pattern as a func- tion of temperature shows that oscillations are not simply a di- rect consequence of the anisotropy of the upper critical field. Therefore not all such oscillations at intermediate fields can be taken as proof of strong anisotropy in the superconducting gap. This result agrees with our recent numerical study of the iron-based superconductor T Fe 2 Se 2 . 11 (2) The complex field- angle dependence of the specific heat and thermal conductiv- ity for systems with anisotropic Fermi surfaces suggests that arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013
Transcript
Page 1: 5 arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013 · 2018. 11. 23. · Field-angle-resolved anisotropy in superconducting CeCoIn 5 using realistic Fermi surfaces Tanmoy Das, 1A.

Field-angle-resolved anisotropy in superconducting CeCoIn5 using realistic Fermi surfaces

Tanmoy Das,1 A. B. Vorontsov,2 I. Vekhter,3, and Matthias J. Graf11Los Alamos National Laboratory, Los Alamos, New Mexico 87545

2Department of Physics, Montana State University, Bozeman, Montana 597173Department of Physics and Astronomy, Louisiana State University, Baton Rouge, Louisiana 70803

(Dated: November 23, 2018)

We compute the field-angle-resolved specific heat and thermal conductivity using realistic model band struc-ture for the heavy-fermion superconductor CeCoIn5 to identify the gap structure and location of nodes. We use atwo-band tight-binding parametrization of the band dispersion as input for the self-consistent calculations in thequasiclassical formulation of the superconductivity. Systematic analysis shows that modest in-plane anisotropyin the density of states and Fermi velocity in tetragonal crystals significantly affects the fourfold oscillationsin thermal quantities, when the magnetic field is rotated in the basal plane. The Fermi surface anisotropy sub-stantially shifts the location of the lines in the H-T plane, where the oscillations change sign compared toquasicylindrical model calculations. In particular, at high fields, the anisotropy and sign reversal are found evenfor isotropic gaps. Our findings imply that a simultaneous analysis of the specific heat and thermal conductivity,with an emphasis on the low energy sector, is needed to restrict potential pairing scenarios in multiband super-conductors. We discuss the impact of our results on recent measurements of the Ce-115 family, namely CeT In5

with T=Co,Rh,Ir.

PACS numbers: 74.25.Uv,74.20.Rp,74.25.Bt,74.25.fc

I. INTRODUCTION

Many heavy-fermion and other novel superconductors arethought to possess nodes in the gap function on the Fermi sur-face. Since the gap shape is directly related to the symme-try of the pairing interaction, knowing the position of nodescan shed light on possible pairing mechanisms. Magneticfield-angle-resolved specific heat and thermal conductivity ex-periments are able to provide detailed information about theanisotropy of quasiparticle excitations near the Fermi surface,and hence help identify the nodal directions in the bulk.1–4

To implement this procedure it is necessary to have high-precision probes that detect small variations under changesof the direction of the applied field. A series of remarkableexperiments proved already the viability of this approach.4–7

However it has proved non-trivial to interpret these exper-iments in general. The oscillations in physical quantities,as a function of the field direction, change sign depend-ing on the magnitude of the applied magnetic field and thetemperature.3,8,9 The location of these inversion lines de-pends sensitively on the topology of the Fermi surface andthe material-specific details of the (multi-) band structure.10,11

Obviously, this calls for the development of theoretical toolsthat take material-specific properties into account. Further-more, it suggests that a quantitative and unambiguous identi-fication of the structure of the superconducting (SC) gap re-quires the incorporation of realistic Fermi surface (FS) prop-erties.

The unconventional heavy-fermion superconductorCeCoIn5 is an ideal candidate for testing field-angle-resolvedprobes due to the existence of large high-quality crystals andaccessible temperature and field ranges. Early field-angle-resolved thermal conductivity and specific heat measurementswere controversial on whether CeCoIn5 has a superconduct-ing gap with dx2−y2 or dxy symmetry.12,13 Recent specificheat measurements observed the predicted inversion of the

oscillations at low temperature.6 This seemed to have settledthe dispute in favor of dx2−y2 pairing symmetry.

In this paper we incorporate first-principles electronicstructure calculations to obtain the realistic tight-bindingparametrization for Ce-115 (CeT In5 with T=Co,Rh,Ir) ma-terials that reproduce the Fermi surface (FS) topology andyield the Fermi velocities, and the density of states (DOS)at the Fermi level. We use this FS parametrization as in-put for self-consistent calculations of thermal properties inthe extended Brandt-Pesch-Tewordt (BPT) approximation ofthe quasiclassical Eilenberger equation.8,9 Use of the tight-binding parametrization allows for a numerically efficientcomputation, while keeping the essential character of the low-energy band structure that reflects on the hybridization be-tween Ce 4f and In 5p states. Within this framework, weconsider candidate s- and d-wave order parameters, performa systematic study of the angle-resolved specific heat coeffi-cient, γ = C/T , and thermal conductivity, κ, in a magneticfield rotating in the Ce-In basal plane. Finally, we construct afield-temperature phase diagram of the fourfold oscillations.

The main results of our calculations, which are applicableto a wide range of systems with tetragonal point group sym-metry, are: (1) For isotropic gap (s-wave) we find that mod-erate FS anisotropies are sufficient to introduce field-angle-dependent oscillations in the specific heat and thermal con-ductivity in the superconducting state over a significant rangeof temperatures and at intermediate to high magnetic fields.In addition, the inversion of the oscillation pattern as a func-tion of temperature shows that oscillations are not simply a di-rect consequence of the anisotropy of the upper critical field.Therefore not all such oscillations at intermediate fields canbe taken as proof of strong anisotropy in the superconductinggap. This result agrees with our recent numerical study of theiron-based superconductor TFe2Se2.11 (2) The complex field-angle dependence of the specific heat and thermal conductiv-ity for systems with anisotropic Fermi surfaces suggests that

arX

iv:1

303.

6258

v1 [

cond

-mat

.sup

r-co

n] 2

5 M

ar 2

013

Page 2: 5 arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013 · 2018. 11. 23. · Field-angle-resolved anisotropy in superconducting CeCoIn 5 using realistic Fermi surfaces Tanmoy Das, 1A.

2

comparison of both quantities with material-specific theoriesis required to identify the pairing symmetry and gap struc-ture. This is already important for materials, where the Fermisurface anisotropy is moderate, as is the case for the Ce-115family.

The rest of the paper is arranged as follows. In Sec. II,we present our tight-binding representation of the two FSs forCeCoIn5. Detailed analytical and computational formalism ofthe field-angle resolved specific heat and thermal conductivitycalculations is given in Sec. III. The results of the temperatureand magnetic field dependence of these quantities, and theirrelative sign reversal in the field-angle oscillation for s-andtwo d-wave pairing symmetries are given in Sec. IV. Somecomparison with the available data for CeCoIn5, CeRhIn5,and CeIrIn5 is also included. Finally, we conclude in Sec. V.

II. ELECTRONIC STRUCTURE

First-principles calculations of CeCoIn5 demonstrate thatthe bands crossing the Fermi level are dominated by stronglyhybridized 4f electrons of the Ce atom with weak overlapcoming from the 5p orbitals of the In atom.19,21 In this workour basic aim is to parameterize the true shape of the FSs only,while the overall dispersion feature at higher energy is irrele-vant for thermodynamic and transport properties. Therefore,we use an effective tight-binding model of the lowest energyof three 4f−orbitals in a tetragonal lattice. As we are onlyinterested in the eigenvalues and not the eigenvectors of eachband, we absorb the orbital symmetry of contributing orbitalsinto the tight-binding hopping parameters, which makes allbands decoupled from each other. With this motivation wewrite the tight-binding dispersion including up to third nearestneighbor hopping on the x−y plane and only nearest neighborhopping along c−axis to obtain

ξk = −2∑i

(tici + t2ic2i)− 4txycxcy − EF . (1)

Here cαi = cos (αki) with i = x, y, z. EF is the Fermi en-ergy. We obtain the values of the tight-binding parametersafter fitting to first-principles dispersions by Ref. 19 shownin Fig. 1(a): (tx=ty ,tz ,t2x=t2y ,txy ,EF )=(-0.12,-0.05,0,0.09,-

0.55), and (-0.17,0.06,0,0.15,-0.47) in eV for the α and βband, respectively. Note that t2z = 0. The other two bandscrossing the Fermi level have small areas and are not furtherconsidered in our two-band model description of CeCoIn5.

The α and β bands give two concentric electron pockets atthe zone corner (M -point), see Fig. 1(b). The kz dispersionof each band is more interesting and needs special attention.Along the kz direction, both α and β FSs are more like corru-gated cylinders: the α FS has a narrow waist at kz = 0, whilethe β FS has a belly. Note that only nearest neighbor hop-ping along the c axis is sufficient to obtain the qualitative kzdispersion of all bands in agreement with the ab-initio bandstructure22 and dHvA experiments20 [see Fig. 1(c)]. The op-posite sign of the tz parameter is responsible for the oppositeshape of the α and β FSs (narrow waist vs. belly).

III. THEORY AND COMPUTATIONAL METHOD

For magnetic field H applied at angle α with respect tothe (100) direction, we compute the field-angle induced su-perconducting DOS per spin, Nn(ω;H) (band index n =1, 2) by solving the Eilenberger equation3,8,9,23 within the ex-tended BPT quasiclassical approximation.24–26 The BPT ap-proximation implies a uniform field H over the unit cell ofthe Abrikosov vortex lattice (unit-cell averaged Green’s func-tion). This produces quantitatively correct results near the up-per critical field, and continues to yield semi-quantitativelycorrect description over the range 0.5Hc2(T ) . H ≤ Hc2(T )for isotropic gap,24,27,28 and to much lower fields for nodal andstrongly anisotropic gaps in single-band models.3,29,30

Here we summarize the key steps of the calculation, andhighlight the main technical differences between the single-and multi-band systems following Refs. 8, 9, and 23. Themain object of interest, the quasiclassical Green’s function, isassumed to be diagonal in the band space (n = 1, 2), sincebands are well separated in the Brillouin Zone, and have the4×4 Gor’kov-Nambu matrix structure corresponding to sin-glet pairing in each band,

G =

(g1 00 g2

), gn =

(gn iσ2fn

iσ2fn −gn

). (2)

The Green’s function in each band satisfies the Eilenbergerequation for given Matsubara frequency iων = iπT (2ν + 1),which has a simple commutator form:31

[(iων +

e

cvn(kf ) ·A(R)) τ3 − ∆n(R,kf )− σimpn (iων), gn(R,kf ; iων)

]+ ivn(kf ) ·∇R gn(R,kf ; iων) = 0 , (3)

where the Fermi velocity in band n is denoted by vn(kf ), with the wavevector kf on the respective FS. Since this is a homoge-neous equation, it has to be complemented by the normalization condition of the Green’s functions:

gn(R,kf ; iων)2 = −π2. (4)

Furthermore, the off-diagonal Green’s functions and self-energies are related by symmetry:31 fn(R,kf ; iων) =

fn(R,−kf ; iων)∗ = fn(R,kf ;−iων)∗; ∆impn (R,kf ; iων) = ∆imp

n (R,−kf ; iων)∗ = ∆impn (R,kf ;−iων)∗.

The equations for the Green’s functions in two bands are coupled indirectly through the self-energies entering the Eilenbergerequation. The scattering of quasiparticles off impurities with concentration nimp is taken into account via the self-energy in each

Page 3: 5 arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013 · 2018. 11. 23. · Field-angle-resolved anisotropy in superconducting CeCoIn 5 using realistic Fermi surfaces Tanmoy Das, 1A.

3

(a)

k z [π

/c]

α β

2

1

-1

0

Γ Γ M(π,π,0) X(π,0,0) Z(0,0,π) A(π,π,π) R(π,0,π) Z

Ener

gy (e

V)

(c)

α β α β

vn

A

0

1

(b)

-1

FIG. 1. (color online) Electronic structure of CeCoIn5: (a) Tight-binding fits of the two most relevant bands α (red) and β (green) to theelectronic dispersions of CeCoIn5 calculated in the local density approximation by Maehira et al.19 (b) α and β FSs at three representative kzvalues, colored by corresponding Fermi velocity from low (green) to high (yellow). (c) Three dimensional rendering of the computed FSs forα and β bands (bottom panel) compared with the dHvA experiments20 (top panel). The α FS has a narrow waist, while the β FS has a belly.The color map of the calculated FSs gives the anisotropy of the magnitude of the Fermi velocities ranging from low (blue) to high (red).

band, σimpn , which is evaluated in the T -matrix approximation for the two-band system,32,33

σimpn ≡(D + Σimpn iσ2∆imp

n

iσ2∆impn D − Σimpn

)= nimptnn , T = U + U〈Nf (kf )G(kf )〉FS T . (5)

The T matrix and the impurity scattering potential have the following structure in band space:

T =

(t11 t12

t21 t22

), U =

(u11 u12

u21 u22

). (6)

The angular brackets denote the integral over one or the other Fermi surface, as appropriate, e.g.:

〈 Nf (kf ) G(kf ) 〉FS = diagn=1,2

[∫FSn

d2kf Nf,n(kf ) gn(kf )

], (7)

and the corresponding normal-state DOS at the Fermi level is Nf,n(kf ) ∼ 1/|vn(kf )|. Sometimes we will omit the subscriptFS for brevity.

For each T and H the order parameters are calculated self-consistently from the coupled gap equations of the two-band model

∆n(R,kf ) = T∑ων

∑m

⟨Vnm(kf ,kf

′) Nf,m(kf′) fm(R,kf

′; iων)⟩FS. (8)

We use a factorized pairing potential at the Fermi surface as Vnm(kf ,kf′) = Vnm Yn(φ)Ym(φ′), with Yn(φ) the basis

Page 4: 5 arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013 · 2018. 11. 23. · Field-angle-resolved anisotropy in superconducting CeCoIn 5 using realistic Fermi surfaces Tanmoy Das, 1A.

4

function that depends only on the azimuthal angle, see Figs. 2and 3. This means that the order parameters are also factor-ized, ∆1,2(kf ) = ∆1,2Y1,2(φ). We couple the bands, forsimplicity, by purely interband pairing V12 = V21 = −V ,and assume same symmetries and angular variations on bothbands Y1(φ) = Y2(φ). This ensures that the order parametersin the two bands are strongly coupled, and the temperature andfield dependence of both gaps is similar, while keeping thenumber of parameters minimal. While there are indicationsthat in CeCoIn5 there is a small excitation gap that closes atvery low fields, of order 0.1% of the upper critical field,34 itseems likely that this gap is proximity induced on the parts ofthe Fermi surface with low f -electron content that we do notconsider here. Since all the experiments measuring the field-angle anisotropy are carried out at fields, which are at suffi-ciently high H/Hc2, we consider only Fermi surface sheetswith strong pairing and large gaps. It is worth mentioning thatthe interband pairing captures both nodeless s± and nodal dwave pairing scenarios.

Generally, for arbitrary interaction matrix Vnm, the cou-pled gap equations support two solutions for the amplitudes(∆1,∆2). The physical solution corresponds to the high-est transition temperature Tc0, that is, the greatest eigenvalueVmax of the interaction matrix,(

V11〈Nf,1Y21 〉 V12〈Nf,2Y2

2 〉V21〈Nf,1Y2

1 〉 V22〈Nf,2Y22 〉

)(e1

e2

)= Vmax

(e1

e2

).

(9)The effective interaction strength Vmax and the cutoff Ωc canbe eliminated using standard techniques in favor of the baretransition temperature, Tc0 = 1.13Ωc exp(−1/Vmax),31 andthe gap amplitudes in different bands are given by the eigen-vector of the interaction matrix,(

∆1

∆2

)=

(e1

e2

)∆ . (10)

Upon projecting out this vector from Eq. (8), the system of theself-consistency equations is reduced to a single equation forthe order parameter ∆ of the dominant instability.

Since the only coupling between bands is via the self-consistency equations of the order parameter and the self-energies, the solutions for the propagators in each bandcan be formally obtained from the transport equation (3)with given ∆n and σn in the same way as for single-band systems.8,9 We express the gradient term via the rais-ing and lowering operators (a†, a) for the vortex solu-tions corresponding to the superposition of different har-monic oscillator functions:26 vn(kf ) ·

(∇R − i 2e

c A(R))

=1√2Λ

(vn,−(kf )a† − vn,+(kf )a

). The (x, y) projections of

the Fermi velocity on the plane perpendicular to the directionof the field z have to be rescaled by the anisotropy factor Sf ,

vn,± = vn,x(kf )/√Sf ± ivn,y(kf )

√Sf . (11)

The relevant parameter that determines the excitations in theSC state at a particular point on the Fermi surface is the com-ponent of the (rescaled) Fermi velocity normal to the appliedfield,

v⊥n (kf ) =√vn,x(kf )2/Sf + vn,y(kf )2 Sf . (12)

The corresponding energy scale is

vf,n(φ,H) ≡ v⊥n (kf )

2Λ, (13)

where Λ = (~c/2|e|H)1/2 is the magnetic length, which is oforder of the intervortex distance, and φ is the FS angle withrespect to the kx axis. The anisotropy parameter Sf is chosento give the correct form of the vortex lattice in the linearizedGinzburg-Landau (GL) equations for ∆. This allows us toconsider only the lowest Landau level,8

∆(R) = ∆∑ky

C(n)ky

eiky√Sfy

4√SfΛ2

Φ0

(x− Λ2

√Sfky

Λ√Sf

).

(14)For tetragonal symmetry this parameter depends on the an-

gle that the applied field makes with the symmetry axis c (inthis paper θH = π/2),

Sf =

√cos2 θH +

K||

K⊥sin2 θH . (15)

Here K|| (along c-axis) and K⊥ (in-plane) are the coefficientsof the gradient terms in the GL expansion for the gradientsalong the c-axis and in the ab-plane respectively. For ourtwo-band system they depend on the degree of mixing of thebands in a particular superconducting state ∆. For the state(∆1,∆2) = (e1∆, e2∆) they are determined by the right,~e = (e1, e2)T , and left, ~e′ = (e′1, e

′2), eigenvectors of the in-

teraction matrix in Eq. (9), corresponding to eigenvalue Vmaxwith ~e′ · ~e = 1.

K|| =(e′1 , e

′2)

Vmax

(V11 V12

V21 V22

)(e1 〈v2

1cNf1Y21 〉

e2 〈v22cNf2Y2

2 〉

), (16)

K⊥ =(e′1 , e

′2)

Vmax

(V11 V12

V21 V22

)(e1 〈v2

1aNf1Y21 〉

e2 〈v22aNf2Y2

2 〉

).(17)

With these remarks in mind, we can directly use the single-band results for the unit-cell averaged Green’s functions in thesingle Landau level approximation [we follow the notation ofEqs. (46)-(48) in Ref. 8]:

gn(iων ,kf ;H) = −iπ√1−i√π 1

v2f,n

W ′(iων,nvf,n

)∆n∆n

, (18)

fn(iων ,kf ;H) = ign√π

vf,nW(iων,nvf,n

)∆n. (19)

Here iων,n = iων −Σimpn (iων ,kf ;H) and ∆n = ∆n(kf ) +∆impn (iων ,kf ;H) are the Matsubara frequency and the order

parameter renormalized by the impurity self-energies in eachband n, σimpn . W ′(z) is the first derivative of the complex-valued function W (z) = exp (−z2)erfc(−iz). One can fur-ther cast this in a form similar to that of a uniform supercon-ductor by introducing the new self-energy Σn according toi√π/v2

f,n W ′n(iων/vf,n) ≡ (iων − Σn)−2. The effectiveself-energy Σn now contains effects from both the impurityscattering and the effects of orbital magnetic field.

Page 5: 5 arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013 · 2018. 11. 23. · Field-angle-resolved anisotropy in superconducting CeCoIn 5 using realistic Fermi surfaces Tanmoy Das, 1A.

5

In contrast to the Doppler shift approximation, both the realand the imaginary parts of Σn contribute to the SC DOS, andtheir interplay as a function of energy, H and T , determinethe sign reversal in the fourfold oscillation of the SC DOS.These effects have been extensively studied earlier using asingle quasi-cylindrical FS and nodal gap, and a minimal 2Dmodel for two-band systems, see for example Refs. 3, 8, 9, and23.

The transport and thermodynamic coefficients are calcu-lated by using the retarded Green’s functions through ana-lytic continuation, iων → ω + i0, in the propagators foundabove. We begin with the total electronic specific heat fromboth bands, C = C1 + C2, which is given by the deriva-tive of the net entropy, C = T (∂S/∂T ). Because the low-

temperature approximation given by8

Cn(α) ≈ 2

∫ ∞−∞

dωω2〈Nn(ω,kf ;H)〉FS

4T 2cosh(ω/2T )2, (20)

remains valid almost up to the normal-state transition region,it can be employed to describe the behavior of the heat ca-pacity over most of the phase diagram. Detailed numericalcalculations show that the high-temperature sign reversal lineis robust, but will be shifted to slightly higher temperatures byapproximately 0.1Tc0 for the FS parametrization consideredhere compared to the calculations using the low-temperatureapproximation in Eq. (20).

Next, we consider the total electronic thermal conductivity,which is the sum of the contributions from both bands, κ =κ1 + κ2, with9,29

κxxn (α) = 2

∫ ∞−∞

dωω2

2T 2cosh(ω/2T )2(21)

×〈vxn(kf )2Nn(ω,kf ;H)τn(ω,kf ;H)〉FS .

Here the field-induced SC DOS per spin in each band, Nn(ω,kf )/Nf,n(kf ) = −Im gR(ω,kf ;H)n/π, the factor 2 accountsfor the spin degeneracy, and the transport lifetime is due to both impurity and vortex scattering9,23,29

1

2τn(ω,kf ;H)= −Im Σimpn (ω,kf ;H) +

√π

1

|vf,n(kf ;H)|Im [gRn (ω,kf ;H)W (ω/|vf,n(kf ;H)|)]

Im gRn (ω,kf ;H)|∆n(kf ;H)|2 .(22)

When T → 0 we recover the standard expressions for the Sommerfeld coefficient, γn ≡ Cn/T → 23π

2〈Nn(0,kf ;H)〉FS ,and for the linear coefficient of the thermal conductivity κxxn /T → 1

3π2〈vxn(kf )2Nn(0,kf ;H)τn(0,kf ;H)〉FS . Since the

Green’s function, given by Eq. (18), takes the standard BCS form at H = 0, we also recover the universal thermal conductivityfor gaps with nodes on the FS.35–39 At low fields the approximation breaks down, but for nodal superconductors it providesa good interpolation from low to high fields, and, in the regime 1 1/τimp∆n H/Hc2 reproduces the well-known

√H

field-dependence of the density of states in d-wave superconductors40,41 up to logarithmic corrections.8,9,29

Since the function x2/cosh(x/2)2 peaks at x ∼ 2.5T , the anisotropy of the heat capacity at low temperatures is qualitativelydetermined by the anisotropy in the total SC DOS, N(ω = 2.5T,kf ;H). Using the expansion of the error function, we obtaintwo limiting values for W ′(z): W ′(0) = 2i/

√π and W ′(z 1) ≈ −i/

√πz2. Thus the SC DOS for each band n becomes

Nn(ω;H) = 〈Nn(ω,kf ;H)〉FS ≈

⟨Nf,n(kf )

[1 + 2

(∆n(kf ;H)|vf,n(kf ;H)|

)2]−1/2

⟩FS

, ω vf,n,⟨Nf,n(kf )

[1−

(∆n(kf ;H)

ω

)2]−1/2

⟩FS

, ω vf,n.

(23)

The first line in Eq. (23) only makes physical sense when theBPT approximation is valid at low energies, i.e., for nodal andstrongly anisotropic gaps. In that case at low T (low energy)and low fields, where ∆n(kf ;H) only weakly depends onthe direction of the field, the SC DOS depends predominantlyon the orientation of vf,n(kf ;H) relative to the minima of∆n(kf ;H). At ω = 0 the inversion of the SC DOS as afunction of the field for nodal gaps can be obtained in analogywith Refs. 8 and 42.

At higher energies, the second line of Eq. (23) has theBCS form apart from the replacement of the bare energies

and gaps by their impurity renormalized counterparts. There-fore the field-angle variation enters via the anisotropy of theseself-energies as well as via the field dependence of the gaps,∆n(kf ;H), for determining the anisotropy of the upper criti-cal field. The latter effect is only relevant in the vicinity of thetransition to the normal state, where the result is valid for bothnodal and nodeless gaps, including the fully isotropic situa-tion. Crucially, for anisotropic Fermi surfaces the anisotropyin the self-energies and the order parameter is weighted bythe normal-state angle-dependent DOS, Nf,n(kf ), leading toa complex behavior including the switching of the minima and

Page 6: 5 arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013 · 2018. 11. 23. · Field-angle-resolved anisotropy in superconducting CeCoIn 5 using realistic Fermi surfaces Tanmoy Das, 1A.

6

1 1.50.5 kx [π / a]

1

1.5

0.5

k y [π /

a]

kz=π/ckz=0

1 1.50.5 kx [π / a]

1

1.5

0.5

-5 -2.5 0 2.5 5length along FS

0.4

0.6

0.8

1

v FS

-5 -2.5 0 2.5 5length along FS

ϕ0

FIG. 2. (color online) Fermi surfaces (FSs) and Fermi velocities atkz = 0 (left panel) and kz = π/c (right panel). The relative magni-tude of the Fermi velocities (in arbitrary units) is given by the lengthof the (red and green) arrows in the top panels. The bottom panelsshow the Fermi velocities along the Fermi lines of the correspondingkz slice.

maxima found in this work. However, in this regime the en-ergy width of the Fermi weighting factor in the integral ex-ceeds the gap amplitude and a full numerical evaluation isrequired. Our results are consistent with the general obser-vations based on such an expansion.

For each pairing symmetry, the coupled order parametersare computed self-consistently at each temperature T and fora given value of the magnetic field H applied at the angle α tothe (100) direction. We calculate the field-angle oscillations inthe H-T phase diagram for a mesh of 35 field points betweenzero andHc2, 100 temperature points from zero to Tc0, and 31field-angle points α from zero to 90 to extract the anisotropicterms in the heat capacity and thermal conductivity. For allthe calculations, we consider purely intraband impurity scat-tering, u12 = u21 = 0, u11 = u22 = u0, in the clean limit,2Γ1 = 1/τimp,1 = 2nimp/πNf1 = 0.01× 2πTc0, where Tc0is the bare transition temperature, Nf1 is the density of stateson the first band (α-band), and the scattering phase shift ischosen to be δ = arctan(πNf1u0) = π/2 (unitarity limit).

IV. RESULTS

A. Field-induced superconducting DOS anisotropy and therole of Fermi surface topology

As discussed above, an important aspect influencing our re-sults is that, for the realistic band structure, the contributionsfrom different segments of the Fermi surfaces to the net den-sity of states are weighted differently according to both thefactor Nf,n in Eq. (23), and the segment length of the Fermisurface with a given direction of the Fermi velocity. Fig. 2

shows the profiles of the Fermi velocity and the correspond-ing weighting factors. The key point is that, in a tetragonalsystem, vf,n can have a fourfold anisotropy in the plane thateither enhances or competes with the gap anisotropy in deter-mining the contribution to the net DOS in the superconductingstate, see Fig. 3(a). The detailed interplay of the two dependsnot only on the value of the angle-resolved DOS, but also onthe direction of the Fermi velocity.

Indeed, naively one might expect that the relatively largecontribution to Nf,n from the near-45 direction, combinedwith the node of the dx2−y2 at the same angle in Fig. 3(a)should enhance the field-angle anisotropy for that symmetryof the superconducting state relative to the dxy case when thedirection of the greatest Nf,n is fully gapped. In fact, theopposite is true, see Fig. 3(b): the oscillations are enhancedfor dxy symmetry.

This is an indication that the flat parts of the Fermi surfacewith large values of the Fermi velocity, see Fig. 2, contributemore to the total DOS, when the field is at 45 and all fourflat parts are ’active’. When the field is along 0 or 90, onlytwo flat parts contribute. In contrast, the four ’active’ cornerswith smaller velocities, and hence slightly larger Nf,n(kf )give a smaller contribution simply because their arc length isa smaller fraction of the total Fermi surface length in the re-spective kz slice. It follows that dxy-pairing, which has nodesin the flat parts of the FS, exhibits enhanced C(α = 45). Incontrast, the dx2−y2 profile, gaps those regions of the Fermisurface and thus anisotropy of C is suppressed. Hence the ex-act role of the Fermi surface shape and curvature in the field-angle oscillations is highly non-trivial.

At higher temperatures the simple low-T expression inEq. (7) is only qualitatively correct, and both detailedcalculations3,8,42,43 and experiments6,44,45 demonstrated thatthe anisotropy in the heat capacity is reversed relative to thelow-T result. The lower panel in Fig. 3 shows the field-induced SC DOS as a function of quasiparticle energy belowthe SC gap for α = 0 and α = 45 for all three pairingsymmetries considered here. We immediately see that the SCDOS at these angles switch and reverse magnitude, which re-flects in the sign reversal of the oscillations in specific heat asa function of temperature. Note that, due to the presence ofFermi velocities in κ in Eq. (7), a one-to-one correspondencebetween SC DOS and κ is not straightforward for FS that lackcontinuous rotational symmetry in the plane.

We show below that for realistic and material-specificanisotropic FS, we still find the sign reversal of the heat ca-pacity oscillations for d-wave pairing, which was previouslyreported for the rotationally symmetric cases. Hence this signchange is a generic feature of of nodal gaps. However, the keyfinding in this work is that for moderately anisotropic FSs,measurably large field-angle dependence in the heat capac-ity and thermal conductivity is obtained at high fields alreadyfor isotropic gaps, which can lead to misinterpretations if an-alyzed solely in this field range and in terms of simple har-monics of the SC pairing symmetries. We stress that multi-band effects add additional complexity to any analysis, due tocompeting FS anisotropies. For example, we have previouslyshown that if the FS anisotropy in different bands is opposite

Page 7: 5 arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013 · 2018. 11. 23. · Field-angle-resolved anisotropy in superconducting CeCoIn 5 using realistic Fermi surfaces Tanmoy Das, 1A.

7

A(φ)

– A

(min

)

C/T

(α) –

C/T

(min

) FS Angle [φo]

DO

S (E

)

s dxy dx2

-y2

(a) (b)

(c) (d) (e)

ω/Tc0

Field Angle [αo]

α α

FIG. 3. (color online) Fermi surface anisotropy of the normal-state DOS and SC gaps contrasted with the field-angle anisotropyof the Sommerfeld coefficient and the SC DOS. (a) The calculatedFS anisotropy of the normal-state DOS juxtaposed with gap func-tions of three pairing symmetries. All the SC gaps are computedat the FS and all curves are shifted by their corresponding mini-mum value, except for the s-wave gap. (b) Specific heat coeffi-cient γ(α) = C(α)/T , normalized to its value at Tc, calculated atT/Tc0 = 0.1 and H/Hc2 = 0.1 for d-wave gaps and H/Hc2 = 0.5for the s-wave gap. (c)-(e) Field-induced total SC DOS at T = 0 vs.energy at two representative field angles α = 0 and 45 for all threepairing symmetries. Here we used H/Hc2 = 0.5 for s wave and 0.1for both d waves. Note the low- and high-energy crossings in the SCDOS (arrows) are related to the low- and high-T sign reversals in theoscillations of γ and κ in Fig. 4.

(out-of-phase) to each other, then it can lead to additional signreversals in the field-angle dependence of the thermodynamicquantities for s-wave gap, very similar to what was earlier ob-tained for nodal gaps only.11 Furthermore, bands with differ-ent DOS lead to different amplitudes and shapes of the self-consistent value of the SC gaps (i.e., generally ∆1 6= ∆2). Inthis case, the obtained numerical results become less intuitiveto interpret and a simple one-to-one mapping between oscilla-tions and nodes is lost. However, at low temperature and lowfield the generic understanding of the anisotropy as a conse-quence of the nodal structure alone, remains valid. We give adetailed comparison of the different regimes below.

B. Temperature evolution of field-angle-resolved oscillations

We present the full angle-dependent profiles of γ(α) =C(α)/T and κ(α)/T for several temperatures at a representa-tive low field (H/Hc2 = 0.1) for the two nodal gaps, and, forcomparison, for an isotropic s-wave gap at a moderate field

(H/Hc2 = 0.5) in Fig. 4. In accordance with our earlier cal-culation for KyFe2−xSe2 in Ref. 11, we find that a substantialoscillation in γ and κ is present for isotropic s-wave pairing.The amplitude of the oscillation increases with stronger kz-dispersion. As in simple models,8 close to the inversion linethe oscillations are not a simple sum of the twofold and four-fold harmonics, but have a more complex profile.

For nodal dxy and dx2−y2 pairings, the behavior of os-cillations of γ(α) and κ(α) is similar to results obtainedfor quasicylindrical FSs,23 however the amplitude of oscil-lations and, crucially, the location of sign reversals in theH-T phase diagram are modified. Earlier such sign-reversalfeatures were discussed only for highly anisotropic or nodalgap structures.3,6,8,23,44,45 Our material-specific results cautionagainst straightforward interpretation of oscillations at inter-mediate fields as evidence of nodes, emphasizing the need toprobe low energy excitations.

We extract the amplitudes of the fourfold oscillations bydefining

C4α(T ) ≡ ΠC0 −ΠC

45, (24)

where ΠCα = [C(α, T )/T ]/[CN/Tc] and

κ4α(T ) ≡ [Πκ0 + Πκ

90]/2−Πκ45, (25)

where Πκα = [κxx(α, T )/T ]/[κxxN /Tc], and CN and κN are

the corresponding normal-state values at Tc. This defini-tion removes any twofold, sixfold, etc., contribution fromκ(α) originating from the field parallel or perpendicular tothe vortex lines.46 In fact, it is straightforward to show thatfor any function f(α) =

∑Mn=0 a2n cos(2nα) the definition

in Eq. (25) projects out any other harmonic contribution upto M = 5, resulting in κ4α = 2a4. We verified numericallythat the amplitudes of twelvefold and higher order harmonicsare negligible. On the other hand, the definition in Eq. (24) isless robust, but very convenient. It gives C4α = 2a4, whena2 = a6 = a10 = 0, which is sufficient when sample mis-alignment is negligible and when used away from the signreversal line.

C. H-T phase diagram

In figure 5 we compile our results of the thermal quanti-ties into a contour map of the amplitude of the fourfold os-cillations in the H-T phase diagram for γ = C/T (top row)and κ/T (bottom row) for one nodeless and two nodal gaps.Recall that for quasicylindrical (rotationally-invariant in thebasal plane) FSs the specific heat oscillations simply changesign between the dxy and dx2−y2 symmetries.8 While theoverall characteristics of the phase diagram remains qualita-tively the same for material-specific cases, substantial quanti-tative changes result from the inclusion of realistic Fermi sur-faces and the directional- and band-dependent contributionsto the DOS. Important for the comparison with experiment,we find that the location of the sign-reversal lines for nodalgaps shown in Figs. 5(a2) and 5(a3) shifted compared to the

Page 8: 5 arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013 · 2018. 11. 23. · Field-angle-resolved anisotropy in superconducting CeCoIn 5 using realistic Fermi surfaces Tanmoy Das, 1A.

8

0.97

4 -1

0 0.

2 0.

26 0

.32

0.14

-2

0 10

0

0.32

0.

40

0.36

0.

44

0.1

0.2

0.3

0 -6

-2

0

×10-3

2

-20

-4

-10

10

0 ×10-3

0.1

0.3

0.5

0.7

0.7

0.9

0.8

1.0

×10-3

0.97

8 0.

982

[C/T

]/[C

N/T

c0]

0 45 90

0 0.6

0 45 90 0 45 90

T/Tc0 0 0.95 0 0.75 T/Tc0 T/Tc0

-0.02

0.02 0

αo αo αo

0.01

0.3 0.4

0.5

0.6 0.7

0.01

0.1

0.2

0.3

0.475

0.5

0.525

[κ/T

]/ [κ

N/T

c0]

s dxy dx2

-y2

C4α

κ4α

C4α C4α

κ4α κ4α

(a1) (a2) (a3)

(b1) (b2) (b3)

(c1) (c2) (c3)

0.475

0.5

0.525

0.01

0.1

0.2

0.3

0.01

0.3 0.4

0.5

0.6

0.7

0.8

0.8

FIG. 4. (color online) Calculated oscillations of the heat capacity and thermal conductivity as a function of the field direction relative to thex axis. (a1) Sommerfeld coefficient γ = C/T normalized to its normal-state value CN/Tc0 at fixed field H/Hc2 = 0.5 for s wave, plottedfrom low to high T in units of T/Tc0 (bottom to top curves). Each curve is colored by the sign of the fourfold oscillation; a uniform colormap is used for values below−0.025 and above 0.025. (b1) Same as in (a1) but for the normalized thermal conductivity coefficient κ/T . (c1)The fourfold amplitudes of γ (dashed line) and κ/T (solid line) are plotted as a function of T . For direct comparison the results for nodal dxy(H/Hc2 = 0.1) and dx2−y2 (H/Hc2 = 0.33) symmetries are plotted in panels (a2)-(a3), (b2)-(b3), and (c2)-(c3), respectively. Note that thenon-vanishing of κ4α as temperature approaches the phase transition line in panels (c1) and (c3) is a consequence of the in-plane anisotropyof Hc2.

earlier simple models, due to the interplay of the SC order pa-rameter with the FS anisotropies. As a consequence, the signof the fourfold oscillations, C4α and κ4α, may be differentover a wider range of temperatures and fields. This is to becontrasted with the results for rotationally symmetric Fermisurfaces, where the two were found to switch sign almost atthe same temperatures and fields. We also verified that for s-wave pairing the high-T sign reversal is robust and remainsat nearly the same location for a single-band superconductorwith identical FS.

Note also that at intermediate to high temperatures andfields there is very little in the heat capacity oscillation profile

that distinguishes the isotropic gap from that of the dx2−y2

symmetry, see Fig. 5 panels (a1) vs. (a3). However, there isa much more significant difference in transport, Fig. 5 pan-els (b1) vs. (b3), which implies that a simultaneous study ofboth C(α) and κ(α) is highly desirable to gain confidenceabout the underlying pairing symmetry in any multiband sys-tem where the low-temperature, low-field regime is experi-mentally unreachable. Of course, once the low energy sec-tor at low T and low H is accessed, the differences betweendifferent symmetries, and especially between the nodal andisotropic gaps, becomes obvious. Therefore, in general arather detailed comparison between measurements and cal-

Page 9: 5 arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013 · 2018. 11. 23. · Field-angle-resolved anisotropy in superconducting CeCoIn 5 using realistic Fermi surfaces Tanmoy Das, 1A.

9

-0.01

0.01

-0.02

0.02 0

0 -0.01

0.01 0

-0.02

0.02 0

-0.02

0.02 0

-0.01

0.01 0

0 0.

5 1

0 0.

5 1

0 0.5 0 0.5 0 0.5

H/H

c2

T/Tc0

H/H

c2

s dxy dx2-y

2 (a1) (a2) (a3)

(b1) (b2) (b3)

FIG. 5. (color online) Contour maps of fourfold amplitude oscillations of normalized specific heat C4α (row a) and normalized thermalconductivity κ4α (row b) in the H-T phase diagram. Each column denotes a different gap symmetry studied (isotropic nodeless s, nodal dxyand dx2−y2 gaps). All plots use the same color map scale from minimum (red) to maximum (blue). Note that the fourfold amplitude is givenwith respect to field H ‖ (100), i.e., a negative value corresponds to a minimum at α = 0. Here Tc(H) is defined by the vanishing of bothgaps for given symmetry, which determines the line of the (minimum) upper critical field. Since the BPT approximation for isotropic s-wavepairing is not valid at low H , we shaded the corresponding area where our approach is not applicable.

culations of the C4α and κ4α phase diagrams should be em-ployed to draw conclusions about the pairing symmetries.

D. Comparison with experiments

The superconducting Ce-115 compounds are well suited forthe study of field-angle oscillations. Accordingly, there havebeen a number of experiments investigating the anisotropy ofthe thermal conductivity and the heat capacity under the ro-tated field. Here we compare the experimental results withour findings, previously shown in Fig. 5. Since the upper crit-ical field is Pauli-limited, and our calculation does not accountfor the Zeeman splitting, we cannot expect our results to mapdirectly onto the measurements near Hc2. Nevertheless webelieve that a qualitative comparison can be made, especiallyfor systems with strong paramagnetism in the low-field partof the phase diagram, which, when rescaled to the appropriatevalues of the upper critical field, is essentially identical to thatcomputed in the absence of the Zeeman term.50

CeCoIn5: The unconventional nature of superconductiv-ity was recognized early on through the discovery of power-

law dependence in the temperature behavior of the spe-cific heat and thermal conductivity,51,52 magnetic penetrationdepth,53–56 and spin-lattice and muon-spin relaxation rates57,58

consistent with predictions for nodal lines in the gap. On sym-metry grounds the anisotropy of the upper critical field van-ishes near Tc0, Hc2(0) = Hc2(45). In our calculations, wefind that noticeable anisotropy emerges for T/Tc0 . 0.7. Inthis range Hc2(0) > Hc2(45) for both s and dx2−y2 pair-ing symmetries, while the anisotropy is opposite for dxy pair-ing, i.e., the nodal directions have a lower Hc2 value. Theanisotropy for s and dx2−y2 gap is in qualitative agreementwith theHc2 measurements of CeCoIn5 by Settai et al.,59 whoreported Hc2(0) > Hc2(45) at low temperatures. The mea-sured anisotropy is only a few percent, which would be consis-tent with the assumption that the band electron g-factor, andhence the Pauli limiting field is isotropic in the plane, andthe weak anisotropy is due to a residual orbital effect. Re-markably, the opposite relationship, Hc2(0) = 11.8 T <Hc2(45) = 11.9 T was found in Ref. 60. So far the experi-mental discrepancy of the in-plane Hc2 anisotropy remains anopen puzzle.

The original interpretations of the field-angle-resolved ther-

Page 10: 5 arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013 · 2018. 11. 23. · Field-angle-resolved anisotropy in superconducting CeCoIn 5 using realistic Fermi surfaces Tanmoy Das, 1A.

10

-0.01

0.01

-0.02

0.02 0

0 0

0.5

1 0

0.5

1

0 0.5 0 0.5 0 0.5

H/H

c2

T/Tc0

H/H

c2

CeCoIn5 (a1) (a2) (a3)

(b1) (b2) (b3)

CeRhIn5 CeIrIn5

FIG. 6. (color online) Theoretical contour maps of fourfold amplitude oscillations for dx2−y2 gap for specific heat (upper panel), and thermalconductivity (lower panel) of CeCoIn5, reproduced from Figs. 5(a3)-b(3). The same theoretical data is repeated in three columns, but comparedwith three different experimental data for isostructural superconductors within the Ce-115 family. (a1): Specific heat data of CeCoIn5 by Anet al.6 (circles) and Aoki et al.13 (squares). (b1): Thermal conductivity data of CeCoIn5 by Izawa et al.12 (a2): Specific heat data for CeRhIn5

by Park et al.47,48 (a3): Specific heat data for CeIrIn5 by Lu et al.45 (circles) and Kittaka et al.7 (squares). (b3): Thermal conductivity data ofCeIrIn5 by Kashara et al.49 The symbol size gives the corresponding reported amplitude of the oscillation, whereas the filled color depicts itssign. We find reasonable agreement between theory and experiment in both sign and amplitude of oscillations.

mal conductivity12 and specific heat13 measurements contra-dicted each other regarding the location of the d-wave nodallines in CeCoIn5. The controversy was finally settled by theobservation of the inversion in the specific heat oscillations byAn et al.6 In Fig. 6(a1) and (b1) we plot both the C4α and κ4α

experimental data points (symbols). The agreement betweentheory and experiment is quite convincing for dx2−y2 -wavesymmetry and rules out pairing scenarios of either s or dxygap.

CeRhIn5: The high-pressure, angle-resolved specific heatmeasurements of CeRhIn5 by Park et al.47 showed a clearlydelineated fourfold oscillation with C(0) < C(45), whichwas interpreted in favor of d-wave symmetry. The measure-ments were performed down to temperatures as low as 0.3K (0.3T/Tc) and in fields between 0.2 and 0.9 T. At the pres-sure of 1.47 GPa the superconductivity coexists with antiferro-magnetism with superconducting transition Tc = 1.04 K andin-plane Hc2 = 1.2 T at 0.3 K. The measured in-plane Hc2

anisotropy was negligible. As we noted before, in this region

of theH-T phase diagram both s-wave and dx2−y2 -wave gapsare nearly indistinguishable giving rise to fourfold oscillationswith the minimum of C(α) occurring at H ‖ (100). Support-ing the d-wave interpretation, T - and H-dependent measure-ments down to 0.3 K and 0.05 T exhibited power-law behaviorconsistent with unconventional superconductivity with nodes,i.e., C/T ∼ T and C/T ∼

√H , respectively. Additional ex-

periments at higher pressure (2.3 GPa), i.e., in the purely su-perconducting phase, and at T/Tc = 0.14 andH/Hc2 = 0.09showed evidence of fourfold oscillations with a negative am-plitude C4α of order 4%.48 However, to unequivocally ruleout the possibility of s-wave pairing, based on field-angle-resolved measurements alone, experiments would have to beperformed at temperatures significantly below Tc/3, wherethe exponential T -dependence of the fully gapped excitationspectrum becomes visible. Power laws were also seen in otherpressure measurements of the specific heat, spin-lattice andmuon-spin relaxation rates down to T/Tc ≈ 0.15.58,61,62 Inchemically doped CeRh1−xIrxIn5 a T 3 dependence was seen

Page 11: 5 arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013 · 2018. 11. 23. · Field-angle-resolved anisotropy in superconducting CeCoIn 5 using realistic Fermi surfaces Tanmoy Das, 1A.

11

in 1/T1 just below Tc, which tends toward linear in T at lowertemperatures as is typical of dirty d-wave superconductors.63

In Fig. 6(a2) we plot the C4α experimental data points (sym-bols) on top of the phase diagram for dx2−y2 gap. The field-angle-dependent experiments taken by themselves are incon-clusive, though combined with the reported T and H depen-dences are strongly suggestive of dx2−y2 -wave superconduc-tivity in CeRhIn5.

CeIrIn5: There is an ongoing controversy about the pair-ing symmetry in this compound, because of its two differ-ent superconducting domes, namely one as a function of Rhdoping and the other as a function of pressure. In addi-tion, there is disagreement over the interpretation of the ther-mal conductivity data. On one side, the field-angle-resolvedmeasurements49 and power-law dependence in temperaturewere argued as evidence for d-wave gap with vertical linenodes, similar to the sister compound CeCoIn5,52,64 while onthe other side thermal conductivity measurements, in partic-ular the temperature and magnetic field dependence of theresidual value of κ/T along different axes, were interpretedin favor of a three-dimensional hybrid gap with a horizontalline node.65–67 The hybrid gap proposal was inspired by simi-lar gap functions studied some time ago for the heavy-fermionsuperconductor UPt3.37,39,68 To further complicate the inter-pretation, the results by Shakeripour et al. were also arguedto be consistent with vertical line nodes.69 In addition, powerlaws were reported for magnetic penetration depth and spin-lattice-relaxation rate.57,63,70 The temperature behavior of theanisotropic penetration depth was interpreted to be consistentwith vertical line nodes but not with point nodes and a hor-izontal line node of the hybrid gap.70 However, the conclu-sive evidence for the in-plane gap variation comes from veryrecent angle-resolved specific heat measurements at ambientand finite pressure. Lu et al.45 (circles) reported fourfold os-cillations inside the pressure dome of CeIrIn5 with sign rever-sal of the oscillations at high temperatures between 0.4 and0.6Tc. These data taken together with a low-T anisotropyof Hc2(0) > Hc2(45) and the fact that this compound be-longs to the same family of Ce-115s was strongly suggestiveof two-dimensional dx2−y2 -wave pairing with vertical linenodes. Unfortunately, the temperature in Ref. 45 was too highto formally exclude isotropic s-wave pairing, see the phasediagram in Fig. 5(a1) versus (a3). The specific heat data ofRef. 7, on the other hand, were taken down to 80 mK, thatis 0.2Tc. Therefore, the specific heat oscillations are support-ive of the dx2−y2 gap scenario. Data from both experimentsare included in the comparison in Fig. 6(a3). In additionfield-angle-resolved thermal conductivity data were reportedby Kasahara et al.,49 which are shown in Fig. 6(b3). Com-bined with the specific heat oscillations, they provide strongsupport for this pairing symmetry. Hence at present the over-whelming majority of experiments supports the dx2−y2 -wavesuperconductivity with vertical line nodes in CeIrIn5.

V. DISCUSSION AND CONCLUSIONS

We performed realistic model calculations of the field-angle-resolved specific heat and thermal conductivity usinga tight-binding parametrization of the electronic structurewithin a two-band model of superconductivity, which is rel-evant for the Ce-115 heavy fermions. Our systematic analy-sis of field-angle dependence showed that modest anisotropiesin the density of states and the in-plane Fermi velocities of atetragonal crystal contributes significantly to the fourfold os-cillations in the vortex state, when the magnetic field is rotatedin the basal plane. As evidence we showed that such oscilla-tions exist at intermediate to high fields even for an isotropics-wave gap. Remarkably, the sign reversal of fourfold oscil-lations occurs not only for nodal d-wave gaps, but also for anisotropic s-wave gap as the temperature is decreased. This isone of the main findings of this work and implies that awayfrom the low temperature and low field region it may be dif-ficult to distinguish different pairing symmetries based on thefield-anisotropy of a single probe alone.

Finally, we compared the results of the field-angle-resolvedcalculations within our model with recent experimental dataon different members of the Ce-115 family. The behav-ior of the self-consistently determined thermal quantities fornodal dx2−y2 -wave gap is consistent with experimental re-ports for CeCoIn5. The same phase diagram is also consis-tent with specific heat data for CeRhIn5 and CeIrIn5. Sinceboth CeRhIn5 and CeIrIn5 have similar electronic structureas CeCoIn5 near the Fermi energy, we believe that our Fermisurface parametrization is valid for all three compounds. Con-sequently, very similar phase diagrams should result for allthree materials for which material-specific calculations dras-tically improved the agreement between theory and experi-ment. The comparison with experimental data is restrictedto low fields, since the superconductivity in this material isPauli-limited,14,15 and there are indications of a quantum crit-ical point in the vicinity of the upper critical field at zero tem-perature, Hc2(0).16,17 Thus the regime near the upper criti-cal field at low temperatures is beyond the scope of the cur-rent treatment. We find that within our realistic model of theFermi surface parameters the fourfold anisotropy map is inbetter agreement with experiments on CeCoIn5, if we assumea weak dispersion along the kz axis. Note that the relativelysmall anisotropy of the upper critical field in this material doesnot have direct connection with the anisotropy of the electrondispersion, as it likely stems from the Pauli limiting of su-perconductivity. Since the electronic band structure is verysimilar among the Ce-115s near the Fermi energy, we expectthat our findings for CeCoIn5 are also relevant for CeRhIn5

and CeIrIn5 under pressure. Considering that questions re-main about the exact superconducting gap structure and po-tential spin-fluctuation nesting in the Ce-115s,18 a definitivetheoretical account of field-angle-resolved measurements iswarranted.

We conclude with a note of caution for interpreting field-angle-resolved oscillations. Our self-consistent two-bandmodel calculations demonstrated that simple observations ofoscillations and sign reversals in either C(α) or κ(α) are not

Page 12: 5 arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013 · 2018. 11. 23. · Field-angle-resolved anisotropy in superconducting CeCoIn 5 using realistic Fermi surfaces Tanmoy Das, 1A.

12

direct evidence for the presence of nodes or minima in thegap structure. Such conclusions can be drawn either fromlow-energy measurements, or at higher temperatures and fieldfrom a comprehensive simultaneous analysis within the sameframework of bothC(α) and κ(α) measurements. Only a sys-tematic analysis of the fourfold oscillations in the H-T phasediagram can constrain the space of possible pairing scenariosfor a given material.

ACKNOWLEDGMENTS

We thank R. Movshovich, A. V. Balatsky, T. Park, F. Ron-ning, and J. D. Thompson for many discussions and en-couragements. The work at LANL was funded by the U.S.DOE under contract No. DE-AC52-06NA25396 through theLDRD program (T.D.) and the Office of Basic Energy Sci-ences (BES), Division of Materials Sciences and Engineer-ing (M.J.G.). Work at LSU was supported by NSF Grant No.DMR-1105339 (I.V.) and at MSU by NSF Grant No. DMR0954342 (A.B.V.). We are grateful to a NERSC computingallocation by the U.S. DOE through BES with contract No.DE-AC02-05CH11231.

1 I. Vekhter, P. J. Hirschfeld, J. P. Carbotte, and E. J. Nicol, Phys.Rev. B 59, R9023 (1999).

2 P. Miranovic, M. Ichioka, K. Machida, and N. Nakai, J. Phys.:Condens. Matter 17, 7917 (2005).

3 A. B. Vorontsov and I. Vekhter, Phys. Rev. Lett 96, 237001 (2006).4 Y. Matsuda, K. Izawa, and I. Vekhter, J. Phys.: Condens. Matter

18, R705 (2006).5 T. Park, M. B. Salamon, E. M. Choi, H. J. Kim, and S. I. Lee,

Phys. Rev. Lett. 90, 177001 (2003).6 K. An, T. Sakakibara, R. Settai, Y. Onuki, M. Hiragi, M. Ichioka,

and K. Machida, Phys. Rev. Lett. 104, 037002 (2010).7 S. Kittaka, Y. Aoki, T. Sakakibara, A. Sakai, S. Naktsuji, Y. Tsut-

sumi, M. Ichioka, and K. Machida, Phys. Rev. B 85, 060505(2012).

8 A. B. Vorontsov and I. Vekhter, Phys. Rev. B 75, 224501 (2007).9 A. B. Vorontsov and I. Vekhter, Phys. Rev. B 75, 224502 (2007).

10 I. Vekhter and A. B. Vorontsov, Physica B 403, 958(2008)11 T. Das, A. B. Vorontsov, I. Vekhter, and M. J. Graf, Phys. Rev.

Lett. 109, 187006 (2012).12 K. Izawa, H. Yamaguchi, Y. Matsuda, H. Shishido, R. Settai, and

Y. Onuki, Phys. Rev. Lett. 86, 2653 (2001).13 H. Aoki, T. Sakakibara, H. Shishido, R. Settai, Y. Onuki, P. Mira-

novic, and K. Machida, J. Phys.: Condens. Matter 16, L13 (2004).14 A. Bianchi, R. Movshovich, N. Oeschler, P. Gegenwart, F.

Steglich, J. D. Thompson, P. G. Pagliuso, and J. L. Sarrao, Phys.Rev. Lett. 89, 137002 (2002).

15 A. Bianchi, R. Movshovich, C. Capan, P. G. Pagliuso, and J. L.Sarrao, Phys. Rev. Lett. 91, 187004 (2003).

16 A. Bianchi, R. Movshovich, I. Vekhter, P. G. Pagliuso, and J. L.Sarrao, Phys. Rev. Lett. 91, 257001 (2003).

17 J.-P. Paglione, M. A. Tanatar, D. G. Hawthorn, E. Boaknin, R. W.Hill, F. Ronning, M. Sutherland, and L. Taillefer, Phys. Rev. Lett.91 246405 (2003).

18 F. Ronning, J.-X. Zhu, T. Das, M. J. Graf, R. C. Albers, H. Rhee,W. E. Pickett, J. Phys.: Condens. Matter 24, 294206 (2012).

19 T. Maehira, T. Hotta, K. Ueda, and A. Hasegawa, Phys. Rev. Lett.90, 207007 (2003).

20 H. Shishido, T. Ueda, S. Hashimoto, T. Kubo, R. Settai, H.Harima, and Y. Onuki, J. Phys.: Cond. Mat. 15, L499 (2003).

21 T. Maehira, T. Hotta, K. Ueda, and A. Hasegawa, J. Phys. Soc.Jpn. 72, 854 (2003).

22 H. Shishido et al., J. Phys. Soc. Jpn. 71, 162 (2002).23 A. B. Vorontsov and I. Vekhter, Phys. Rev. Lett 105, 187004

(2010).24 U. Brandt, W. Pesch, and L. Tewordt, Z. Phys. 201, 209 (1967).

25 W. Pesch, Z. Phys. B 21, 263 (1975).26 A. Houghton and I. Vekhter, Phys. Rev. B 57, 10831 (1998).27 E. H. Brandt, J. Low Temp. Phys. 24, 409 (1976).28 J. M. Delrieu, J. Low Temp. Phys. 6, 197 (1972).29 I. Vekhter and A. Houghton, Phys. Rev. Lett. 83, 4626 (1999).30 T. Dahm, S. Graser, C. Iniotakis, and N. Schopohl, Phys. Rev. B

66, 144515 (2002).31 J. W. Serene and D. Rainer, Physics Reports 101, 221 (1983).32 Y. Ohashi, Physica C 412-414, 41 (2004).33 V. Mishra, A. Vorontsov, P. J. Hirschfeld, and I. Vekhter, Phys.

Rev. B 80, 224525 (2009).34 G. Seyfarth, J. P. Brison, G. Knebel, D. Aoki, G. Lapertot, and J.

Flouquet, Phys. Rev. Lett. 101, 046401 (2008)35 Y. Sun and K. Maki, Europhys. Lett. 32, 355 (1995).36 M. J. Graf, S.-K. Yip, J. A. Sauls, and D. Rainer, Phys. Rev. B 53,

15147 (1996).37 M. J. Graf, S.-K. Yip, and J. A. Sauls, J. Low Temp. Phys. 102,

367 (1996); Erratum: 106, 727 (1997).38 M. R. Norman and P. J. Hirschfeld, Phys. Rev. B 53, 5706 (1996).39 M. J. Graf, S.-K. Yip, and J. A. Sauls, J. Low Temp. Phys. 114,

257 (1999).40 G. E. Volovik, JETP Lett. 58, 469 (1993); C. Kubert and P. J.

Hirschfeld, Solid State Commun. 105, 459 (1998).41 C. Kubert and P. J. Hirschfeld, Phys. Rev. Lett. 80, 4963 (1998).42 M. Udagawa, Y. Yanase, and M. Ogata, Phys. Rev. B 70, 184515

(2004).43 P. Miranovic, N. Nakai, M. Ichioka, and K. Machida, Phys. Rev.

B 68, 052501 (2003).44 B. Zeng, G. Mu, H. Q. Luo, T. Xiang, I. I. Mazin, H. Yang, L.

Shan, C. Ren, P. C. Dai, and H.-H. Wen, Nat. Comms. 1, 112(2010); doi: 10.1038/ncomms1115.

45 X. Lu, H. Lee, T. Park, F. Ronning, E. D. Bauer, and J. D. Thomp-son, Phys. Rev. Lett. 108, 027001 (2012).

46 The definition of κ4α corrects for the (usually large) twofoldanisotropy between the heat current flowing parallel vs. perpen-dicular to the vortex lines.

47 T. Park, E. D. Bauer, and J. D. Thompson, Phys. Rev. Lett. 101,177002 (2008).

48 T. Park and J. D. Thompson, New Journal Physics 11, 055062(2009).

49 Y. Kasahara et al., Phys. Rev. Lett. 100, 207003 (2008).50 A. Vorontsov and I. Vekhter, Phys. Rev. B 81, 094527 (2010).51 C. Petrovic, P. G. Pagliuso, M. F. Hundley, R. Movshovich, J.

L. Sarrao, J. D. Thompson, Z. Fisk, and P. Monthoux, J. Phys.:Condens. Matter 13, L337 (2001).

Page 13: 5 arXiv:1303.6258v1 [cond-mat.supr-con] 25 Mar 2013 · 2018. 11. 23. · Field-angle-resolved anisotropy in superconducting CeCoIn 5 using realistic Fermi surfaces Tanmoy Das, 1A.

13

52 R. Movshovich, M. Jaime, J. D. Thompson, C. Petrovic, Z. Fisk,P. G. Pagliuso, and J. L. Sarrao, Phys. Rev. Lett. 86, 5152 (2001).

53 R. J. Ormeno, A. Sibley, C. E. Gough, S. Sebastian, and I. R.Fisher, Phys. Rev. Lett. 88, 047005 (2002).

54 E. E. M. Chia, D. J. Van Harlingen, M. B. Salamon, B. D. Yanoff,I. Bonalde, and J. L. Sarrao, Phys. Rev. B 67, 014527 (2003).

55 S. Ozcan, D. M. Broun, B. Morgan, R. K. W. Haselwimmer, J. L.Sarrao, S. Kamal, C. P. Bidinosti, P.J. Turner, M. Raudsepp, andJ. R. Waldram, Europhys. Lett. 62, 412 (2003).

56 E. D. Bauer, F. Ronning, C. Capan, M. J. Graf, D. Vandervelde,H. Q. Yuan, M. B. Salamon, D. J. Mixson, N. O. Moreno, S. R.Brown, J. D. Brown, R. Movshovich, M. F. Hundley, J. L. Sarrao,P. G. Pagliuso, and S. M. Kauzlarich, Phys. Rev. B 73, 245109(2006).

57 Y. Kohori, Y. Yamato, Y. Iwamoto, T. Kohara, E. D. Bauer, M. B.Maple, and J. L. Sarrao, Phys. Rev. B 64, 134526 (2001).

58 W. Higemoto, A. Koda, R. Kadano, Yu Kawasaki, Y. Haga, D.Aoki, R. Settai, H. Shishido, and Y. Onuki, J. Phys. Soc. Jpn. 71,1023 (2002).

59 R. Settai, H. Shishido, S. Ikeda, Y. Murakawa, M. Nakashima,D. Aoki, Y. Haga, H. Harima and Y. Onuki, J. Phys.: Condens.Matter 13, L627 (2001).

60 Murphy et al., Phys. Rev. B 65, 100514 (2002).

61 R. A. Fisher, F. Bouquet, N. E. Phillips, M. F. Hundley, P. G.Pagliuso, J. L. Sarrao, Z. Fisk, and J. D. Thompson, Phys. Rev. B65, 224509 (2002).

62 T. Mito, S. Kawasaki, G.-q. Zheng, Y. Kawasaki, K. Ishida, Y.Kitaoka, D. Aoki, Y. Haga, and Y. Onuki, Phys. Rev. B 63,220507(R) (2001).

63 S. Kawasaki, M. Yashima, Y. Mugino, H. Mukuda, Y. Kitaoka, H.Shishido, and Y. Onuki, Phys. Rev. Lett. 96, 147001 (2006).

64 C. Petrovic, R. Movshovich, M. Jaime, P. G. Pagliuso, M. F.Hundley, J. L. Sarrao, Z. Fisk, and J. D. Thompson, Europhys.Lett. 53, 354 (2001).

65 H. Shakeripour, M. A. Tanatar, S. Y. Li, C. Petrovic, and L. Taille-fer, Phys. Rev. Lett. 99, 187004 (2007).

66 H. Shakeripour, C. Petrovic, and L. Taillefer, New Journal ofPhysics 11, 055065 (2009).

67 H. Shakeripour, M. A. Tanatar, C. Petrovic, and L. Taillefer, Phys.Rev. B 82, 184531 (2010).

68 M. J. Graf, S.-K. Yip, and J. A. Sauls, Phys. Rev. B 62, 14393(2000).

69 I. Vekhter and A. B. Vorontsov, Phys. Rev. B 75, 094512 (2007)70 D. Vandervelde, H. Q. Yuan, Y. Onuki, and M. B. Salamon, Phys.

Rev. B 79, 212505 (2009).


Recommended