02.02.2014 Views

Computational Approaches to Superconductivity, Research Report ...

Computational Approaches to Superconductivity, Research Report ...

Computational Approaches to Superconductivity, Research Report ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Computational</strong> <strong>Approaches</strong> <strong>to</strong> <strong>Superconductivity</strong>, <strong>Research</strong> <strong>Report</strong><br />

2006-2012<br />

The main activity of our Minerva <strong>Research</strong> Group, estabilished in July 2009, is the study of complex solids, in<br />

particular superconduc<strong>to</strong>rs, using ab-initio DFT calculations. In collaboration with other groups, we combine<br />

these techniques with more traditional many-body methods.<br />

1 Understanding the electronic structure, magnetism and superconductivity<br />

of iron superconduc<strong>to</strong>rs<br />

<strong>Superconductivity</strong> in compounds containing iron (Fe) was considered impossible until 2008, when a critical<br />

temperature (T c ) of 26 K was reported in F-doped LaOFeAs. After this first report, dozens of related compounds<br />

were synthesized, with critical temperatures up <strong>to</strong> 56 K in SmOFeAs.[2] Fig. 1 shows the basic features that<br />

characterize these Fe superconduc<strong>to</strong>rs (FeSC):<br />

(a)-(b) A common structural motive, consisting of square planes of iron a<strong>to</strong>ms, andX 4 tetrahedra;X is either a<br />

pnic<strong>to</strong>gen - As,P - or a chalchogen - Se,Te.<br />

(c) A quasi-two-dimensional Fermi surface, comprising two hole and two electron sheets, strongly nested with<br />

each other. 1<br />

(d) A phase diagram, in which superconductivity sets in after suppressing a spin density wave (SDW) state, by<br />

means of doping or pressure.<br />

a)<br />

b)<br />

c)<br />

Γ<br />

Y<br />

M<br />

X<br />

d)<br />

Figure 1: Common features of Fe-based superconduc<strong>to</strong>rs<br />

(FeSC): FeX layers, seen from<br />

the <strong>to</strong>p (a) and side (b); Fe andX a<strong>to</strong>ms are red<br />

and green, respectively. (c) Two-dimensional<br />

Fermi surface of LaOFeAs, from Ref. [6], with<br />

the hole and electron pockets centered at ¯X, Ȳ<br />

and ¯M respectively. Here a third hole pocket at<br />

¯Γ is also present. (d) Phase diagram, showing<br />

the transition from a spin density wave (SDW)<br />

<strong>to</strong> a superconducting (SC) state;xis an external<br />

tuning parameter (doping, pressure).<br />

These properties suggest that FeSC may form <strong>to</strong>gether with the high-T c cuprates a wide class of compounds,<br />

in which the superconducting pairing is mediated by magnetic excitations (spin fluctuations). In contrast <strong>to</strong> the<br />

d-wave of the cuprates, the <strong>to</strong>pology of the Fermi surface of FeSC favours in most compounds a superconducting<br />

gap with opposite signs on the hole and electron sheets of the Fermi Surface (s ± ). [4]<br />

Performing and understanding electronic structure calculations, [5] has been the main focus of our activity in<br />

this field. We have studied in detail the formation of the non-magnetic electronic structure, explaining the weak<br />

electron-phonon (ep) interaction - Refs. [6, 7, 8, 9]. We have shown that the description of the SDW state is<br />

problematic in the standard local spin density functional theory (LSDFT) approach, due <strong>to</strong> the cohexistence<br />

of itinerant and localized magnetic moments - Refs. [6, 10]. This lead us <strong>to</strong> two different approaches <strong>to</strong> go<br />

beyond standard LSDFT calculations: combining self-consistent antiferromagnetic (AFM) and non-magnetic<br />

(NM) calculations <strong>to</strong> estimate the ep coupling in the paramagnetic (PM) state, [11] and using the Gutzwiller<br />

approximation <strong>to</strong> include strong local correlations. [12] Finally, we have used LSDFT calculations <strong>to</strong> interpret<br />

experimental data, i.e. optics, [13] phonons [14], spin susceptibilty and specific heat. [15]<br />

1.1 Non-magnetic Electronic Structure<br />

In the region of the phase diagram where superconductivity is observed, Angle Resolved Pho<strong>to</strong>emission (ARPES)<br />

and de-Haas-van-Alphen experiments [16, 17] find the typical Fermi surface <strong>to</strong>pology sketched in Fig. 1(c), with<br />

1 Here and in the following, we do not consider the 122 Fe chalchogenides, [3] which are anomalous in many respects.<br />

1


hole and electron pockets at the center and edges of the Brillouin zone. Theoretical models for superconductivity<br />

mediated by spin-fluctuations (SF) rely on this <strong>to</strong>pology, and on the modifications induced by structural deformation<br />

and three dimensional dispersions, <strong>to</strong> explain the trends in critical temperatures and gap symmetry in<br />

different FeSC. [18, 19] Non-magnetic DFT calculations reproduce the experimental Fermi surfaces quite well,<br />

except for a moderate <strong>to</strong> strong quasi-particle renormalization, and a sistematic relative shift of the hole and<br />

electron pockets. [20]<br />

Understanding the formation of this peculiar electronic structure is a highly non-trivial task. [6] We have used<br />

2<br />

xy xy xy<br />

2<br />

xz xz Xz<br />

2<br />

yz yz Yz<br />

2<br />

XY XY XY<br />

1<br />

1<br />

1<br />

1<br />

0<br />

0<br />

0<br />

0<br />

-1<br />

-1<br />

-1<br />

-1<br />

-2<br />

-2<br />

-2<br />

-2<br />

-3<br />

-3<br />

-3<br />

-3<br />

-4<br />

-4<br />

-4<br />

-4<br />

Γ X M Γ<br />

Γ X M Γ<br />

Γ X M Γ<br />

Γ X M Γ<br />

2<br />

z z z<br />

2<br />

x x X<br />

2<br />

y y Y<br />

2<br />

zz zz zz<br />

1<br />

1<br />

1<br />

1<br />

0<br />

0<br />

0<br />

0<br />

-1<br />

-1<br />

-1<br />

-1<br />

-2<br />

-2<br />

-2<br />

-2<br />

-3<br />

-3<br />

-3<br />

-3<br />

-4<br />

-4<br />

-4<br />

-4<br />

Γ X M Γ<br />

Γ X M Γ<br />

Γ X M Γ<br />

Γ X M Γ<br />

Figure 2: Band structure of the two-dimensional tigh-binding model for a single FeAs layer of<br />

LaOFeAs (2D LaOFeAs). The five Fed(xy,xz,yz,XY =x 2 −y 2 andzz=3z 2 −1) and three<br />

Asp(x,y,z) NMTO’s are shown in the insets. Thexandy axes are directed along the shortest<br />

Fe-Fe distance in Fig. 1 – From Ref. [6].<br />

the NMTO downfolding method [21] <strong>to</strong> derive an accurate model of the two-dimensional band structure of<br />

LaOFeAs (2D LaOFeAs). Using the symmetry properties of the single FeAs layer, we could exactly reduce the<br />

problem <strong>to</strong> a single Fe unit cell, and derive an accurate analytical tight-binding model, with five Fe d and three<br />

As p orbitals. Including the As p orbitals explicitely allowed us <strong>to</strong> study the material trends with tetrahedral<br />

angle [18, 22] and the effect of three-dimensional dispersion, also for those compounds in which As is replaced<br />

by a different ligandX, such as P, Se or Te.<br />

Fig. 2 shows the corresponding bands, decorated with a fatness proportional <strong>to</strong> the partial character of the orbitals<br />

shows in the small insets. This complicated electronic structure derives from a subtle interplay between the<br />

covalent bonding of Fe t 2 (xz,yz andxy) and Asporbitals, and direct metallic bonding for the Fe e states. The<br />

dashed line marks the position of the Fermi level (E F ) for the nominal electron countp 6 d 6 of most undoped Fe<br />

based superconduc<strong>to</strong>rs. The Fermi surface is shown in Fig. 1 (c); the states at E F are all oft 2 character, and the<br />

curvature of the bands that form the hole and electron pockets is strongly affected by Fe-As hybridization.<br />

One of the properties of this electronic structure, around the filling p 6 d 6 , is a weak coupling of the electronic<br />

states <strong>to</strong> phonons – the <strong>to</strong>tal electron-phonon coupling constantλ=0.2, is even lower than in elemental aluminum<br />

(λ = 0.44), which has a T c of ∼ 1 K, as we have shown in Ref. [7]. The <strong>to</strong>p panel of Fig. 3 shows the<br />

results of a linear-response calculation for LaOFeAs. [23] The phonon spectrum extends up <strong>to</strong> ∼ 500 cm −1 ,<br />

with a weak ep coupling, distributed among three peaks of Fe-As character at ω 100, 200 and 300 cm −1 . The<br />

distribution of theep spectral functionα 2 F(ω) (Eliashberg function) is uniform, indicating no strong dormantep<br />

interaction. [24, 25]. The lower panel of the same figure shows a similar calculation for the analogous compound,<br />

in which Fe is replaced by nickel - LaONiAs. This compound is also superconducting, but with a much lower<br />

T c 5 K, and the nominal filling is p 6 d 8 . Here, the Fermi level sits ∼ 1 eV higher in Fig. 2, crossing the<br />

states derived from the directed x 2 − y 2 orbitals. Single-band Migdal Eliashberg theory for phonon-mediated<br />

superconductivity is sufficient <strong>to</strong> explain the experimental T c ’s, and several experimental observations such as<br />

the specific heat jump at T c and the size of the superconducting gap – Refs. [7, 8]. Similar results are found for<br />

other iron and nickel compounds. [26]<br />

2


DOS<br />

α 2 F(ω)<br />

500<br />

400<br />

λ q<br />

ν<br />

La<br />

Fe<br />

As<br />

O<br />

ω (cm -1 )<br />

300<br />

200<br />

100<br />

0<br />

Γ X M Γ Z R A<br />

0 0.2<br />

DOS<br />

0.2<br />

α 2 F(ω)<br />

500<br />

400<br />

λ q<br />

ν<br />

La<br />

Ni<br />

As<br />

O<br />

ω (cm -1 )<br />

300<br />

200<br />

100<br />

0<br />

Γ X M Γ Z R A 0 0.2<br />

0.4 0.8<br />

Figure 3: Phonon dispersion, density of states<br />

(DOS) andep spectral function –α 2 F(ω) – calculated<br />

in density functional perturbation theory<br />

[23] for LaOFeAs (<strong>to</strong>p) and LaONiAs (bot<strong>to</strong>m).<br />

The dashed line represents λ(ω) =<br />

2 ∫ ω<br />

0 α2 F(Ω)dΩ – From Refs. [7, 8].<br />

If E F is shifted further up, until an electron count p 6 d 10−x , the ep matrix elements are even higher, due <strong>to</strong> the<br />

strong hybridization with the X p states. This is realized in BiOCu 1−x S, where the Cu-S layers play the role<br />

of the Fe-X layers – Ref. [9]. For this material, we have studied the competition of spin fluctuations and ep<br />

coupling, using linear response calculations and the local spin-density functional version of the random-phase<br />

approximation (RPA) for spin fluctuations. We have shown that BiOCu 1−x S is a quite unique compound where<br />

both a conventional phonon-driven and an unconventional triplet superconductivity are possible, and compete<br />

with each other. We argued that, in this case, it may be possible <strong>to</strong> switch from conventional <strong>to</strong> unconventional<br />

superconductivity by varying such parameters as doping or pressure.<br />

1.2 The SDW state: problems and hints from DFT calculations<br />

In a large part of their phase diagram – red in Fig. 1 (d) – many FeSC exhibit long-range magnetic order; the most<br />

common pattern is aQ = (0,π) SDW, in which the the Fe spins are aligned aligned ferromagnetically along one<br />

of the edges of the Fe squares, and antiferromagnetically along the other, forming parallel stripes. For this reason,<br />

this pattern is usually called AFM stripe. In this AFM state, the lattice often displays a small orthorhombic<br />

dis<strong>to</strong>rtion. 11 and 122 chalchogenides exhibit a double-stripe and a block-AFM pattern respectively. Local spin<br />

density functional calculations encounter several problems in describing the formation and suppression of this<br />

magnetic order. In Ref. [10] we have shown for the first time that it is impossible <strong>to</strong> reproduce, within a single<br />

DFT calculation, the correct value of electronic (fermiology, magnetic moment) and structural (tetrahedral angle,<br />

B 1g phonon frequency) properties. We have interpreted this as an indication that in actual FeSC two types of<br />

magnetic moments cohexist. [30]<br />

The problem is the following: The magnetic moments measured byµSR and inelastic neutron scattering typically<br />

range from 0.5-0.6 – 1.0µ B in the pnictides, <strong>to</strong>∼2.0µ B for the 11 chalchogenides. The magnetic ground state<br />

in LSDFT is the same SDW order observed by experiments, but the magnitude of the magnetic moment (∼ 2.0<br />

µ B ) is almost independent on the compound, and often much larger than the experimental one. Even worse,<br />

for a large range of dopings for which experiments see no trace of long-range magnetic order, LSDFT equally<br />

predicts an AFM ground state, with m ∼ 2.0 µ B . The puzzling thing is that suppressing this large ordered<br />

moment in the calculations spoils the almost perfect agreement with experiment for the structural and vibrational<br />

properties, introducing errors as large as 25 % for some phonon frequencies, and for the internal position of the<br />

X a<strong>to</strong>ms. [11, 28, 29]<br />

In order <strong>to</strong> understand the origin and implications of this surprising resut, in Ref. [6] we have studied a simplified<br />

model for AFM in FeSC. This is an extension of the so-called extended S<strong>to</strong>ner theory for itinerant ferromagnets<br />

[31], which we have recently used <strong>to</strong> study spin fluctuations in Ni 3 Al, [32] and is usually a very good<br />

approximation <strong>to</strong> the full self-consistent DFT result. [31, 33] A staggered magnetic field ∆ Q , with the same<br />

3


m/∆(µ B<br />

eV -1 )<br />

2<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

Q=(0,π)<br />

I=0.59 eV<br />

I=0.82 eV<br />

x=0.00<br />

x=0.05<br />

x=0.10<br />

x=0.15<br />

x=0.20<br />

0 0.5 1 1.5 2 2.5<br />

m(µ B<br />

)<br />

Y<br />

Γ<br />

Y<br />

Γ<br />

M<br />

X<br />

M<br />

X<br />

Figure 4: Non-interacting, static spin susceptibilities<br />

χ(m)=m/∆ for 2D LaOFeAs from a<br />

S<strong>to</strong>ner model for AFM stripes m is the magnetization,<br />

I the S<strong>to</strong>ner parameter, ∆ the applied<br />

staggered magnetic field, andxis doping in the<br />

rigid band approximation. For a given value<br />

of the S<strong>to</strong>ner interaction parameter, I, the selfconsistent<br />

moment is given by χ(m) = 1/I.<br />

I = 0.59 and I = 0.82 yield the experimental<br />

and LSDFT value of the magnetic moment,<br />

respectively. The two corresponding Fermi surfaces<br />

are also shown. From Ref. [6].<br />

periodicityQof the SDW, is applied <strong>to</strong> a system, coupling the k and k+Q blocks of the corresponding tightbinding<br />

Hamil<strong>to</strong>nian. The resulting magnetizationm is related <strong>to</strong>∆via the self-consistency condition∆ = mI,<br />

where I is the S<strong>to</strong>ner interaction parameter, which in LSDFT calculations is set by the choice of the exchange<br />

and correlation functional. Using the 2D tight-binding Hamil<strong>to</strong>nian for LaOFeAs of Ref. [6] and an orbitalindependent<br />

exchange field for the Fedorbitals, we obtained the non-interacting static spin susceptibilityχ(m)<br />

plotted in Fig. 4, as: χ(m) = m/∆, for several rigid-band dopings x. The curve shown in the figure is for an<br />

AFM stripe – Q = (0,π). This function is uniquely determined by the non-interacting band structure of Fig. 2.<br />

Its shape reflects two different regimes for magnetism: (a) a small moment regime (m ≤ 1.0), dominated by the<br />

strong nesting between hole and electron Fermi pockets, where magnetism is very feeble, and rapidly suppressed<br />

by doping; and a more robust large moment regime (m ≥ 1.0µ B ), which persists up <strong>to</strong> very high dopings. The<br />

two values ofI represented by dashed lines were chosen <strong>to</strong> reproduce the ordered moment of LaOFeAs detected<br />

by experiments (m ∼ 0.4µ B ,I = 0.59eV ), [27] and the value found in self-consistent LSDFT calculations (m<br />

∼ 2.0 µ B , I = 0.82 eV). At the side of the figure, we also show the two corresponding Fermi surfaces. These<br />

are qualitatively very different: the small-moment one is reminiscent of the non-magnetic (NM) Fermi surface<br />

shown in Fig. 1 (c), with small SDW gaps; in the large moment one, there is an almost complete reconstruction,<br />

due <strong>to</strong> a 10 times larger exchange splitting (∆ ∼ 2 eV). ARPES experiments in magnetic samples show Fermi<br />

surfaces similar <strong>to</strong> the small-moment one; the second <strong>to</strong>pology is never observed. The two Fermi surfaces give a<br />

pic<strong>to</strong>rial representation of the problems encountered by LSDFT in the pnictides: the fermiology and the suppression<br />

of the magnetic moment with doping is well reproduced by small moment calculations, which correspond<br />

<strong>to</strong> an exchange field of ∼ 2 − 300 meV; in order <strong>to</strong> reproduce the structural properties, instead, much larger<br />

exchange fields∼ 2 eV are required.<br />

What we proposed in Ref. [10] is that these two regimes cohexist in actual FeSC samples, with large, localized<br />

moments that extend in the whole phase diagram, even the regions where experiments see no trace of static<br />

magnetism. [30] The superconducting samples are thus strongly paramagnetic, i.e. very distant from the nonmagnetic<br />

DFT picture.<br />

1.3 Magnetism, going beyond Density Functional Theory<br />

We proposed a model the paramagnetic state in DFT in Ref. [11], where we studied the effect of magnetism<br />

on phonon dispersions and electron-phonon coupling. The three <strong>to</strong>p panels of fig. 5 show the results of<br />

three self-consistent calculations of phonon dispersions and ep coupling for BaFe 2 As 2 , with a ground state<br />

which is non-magnetic (NM), anti-ferromagnetic with a checkerboard pattern (AFMc), and anti-ferromagnetic<br />

stripe (AFMs). The thick, colored lines represent the phonon densities of states (DOS), the dashed lines give<br />

the same-spin component of the Eliashberg function: α 2 F σσ (ω). The corresponding ep coupling constants<br />

λ σσ = 2 ∫ dωα 2 F σσ (ω)/ω are 0.18, 0.33 and 0.18 respectively. The average ep matrix elements of the two<br />

AFM calculations (λ σσ /N σ (0)=0.25) are about 1.5 time larger than the NM result (λ/N(0)=0.15), due <strong>to</strong> magne<strong>to</strong>elastic<br />

coupling. We obtained an estimate of the coupling in the paramagnetic phase, combining the phonon<br />

frequencies andep matrix elements of the AFM solution, with the Fermi surface given by the non-magnetic calculati<br />

ons. The resultingep spectal functionα 2 F(ω) is plotted in blue in Fig. 5 <strong>to</strong>gether with the non-magnetic<br />

one; the increased weight around 20 meV yields a 50 % increase in the <strong>to</strong>tal ep coupling. The corresponding<br />

phonons are the B 1g Raman-active modes that dis<strong>to</strong>rt the tetrahedra. From this result, <strong>to</strong>gether with a rigid-band<br />

approximation for doping, we estimated a new upper bound for the ep coupling in the paramagnetic phase of<br />

FeSC, λ PM = 0.35. This value, which is twice as large as the one estimated in the NM phase [7], is still <strong>to</strong>o<br />

4


F(ω)<br />

0.2<br />

0.1<br />

0<br />

0.2<br />

0.1<br />

0<br />

0.2<br />

0.1<br />

NM<br />

λ=0.18<br />

N σ<br />

(0)=1.18<br />

AFMc<br />

λ=0.33<br />

N σ<br />

(0)=1.36<br />

AFMs<br />

λ=0.18<br />

N σ (0)=0.68<br />

<strong>to</strong>t<br />

Fe<br />

As<br />

0<br />

0 10 ω (meV) 20 30<br />

10 ω (meV) 20 30<br />

0.3<br />

α 2 F(ω)<br />

0.2 λ(ω)<br />

0.1<br />

PM<br />

NM<br />

Figure 5: Top panel: Phonon DOS (full lines)<br />

and same-spin ep spectral function α 2 F σσ(ω)<br />

(dashed line) for non-magnetic (NM), antiferromagnetic<br />

checkerboard (AFMc) and<br />

antiferromagnetic stripe (AFMs) BaFe 2As 2.<br />

Lower panel: α 2 F σσ(ω) and λ σσ(ω) for<br />

non-magnetic (NM) and paramagnetic (PM)<br />

BaFe 2As 2. – From Ref. [11].<br />

m/µ B<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

J=0.10U<br />

J=0.075U<br />

J=0.05U<br />

S<strong>to</strong>ner<br />

0.0<br />

0.4 0.5 0.6 0.7 0.8 0.9<br />

I/eV, χDFT −1 /eV<br />

Figure 6: Ordered stripe magnetic moment<br />

m of undoped, 2D LaOFeAs as a function of<br />

I eff = U −0.017J, in the Gutzwiller approximation<br />

(red, green, blue symbols), <strong>to</strong>gether<br />

with the inverse of the non-interacting spin susceptibility<br />

χ −1 (m), shown in fig. 4. – From<br />

Ref. [12].<br />

low <strong>to</strong> explain the superconducting T c alone, but is high enough <strong>to</strong> yield visible effects in realistic models for<br />

the pairing.<br />

The cohexistence of localized and itinerant moments in FeSC can only be captured by methods that include local<br />

fluctuations, such as the dynamical mean field theory (DMFT). The solution of multiband Hubbard models, in<br />

which the kinetic part is given by realistic models of the DFT band structure (LDA+DMFT), reproduces a state<br />

with small ordered moments and large local fluctuating moments. [34] A very similar picture is given by the<br />

simpler Gutzwiller approximation. [35]. In Ref. [12], we solved a multiband Hubbard Hamil<strong>to</strong>nian, in which<br />

the kinetic part is given by the 8-band model of Ref. [6], and the full a<strong>to</strong>mic interaction between the electrons<br />

in the iron orbitals is parametrized by the Hubbard interactionU and an average Hund’s-rule interaction J. We<br />

found that for a wide range of (U,J) parameters the ground state is a metallic spin density wave (SDW), with<br />

a small magnetic moment. The local moment (< S 2 >∼ 2.0µ B ) is much larger than the ordered one, but still<br />

far from the fully polarized a<strong>to</strong>mic value (m = 4.0µ B ). Consistently with the DMFT results, we found that the<br />

value of the magnetic moment is mostly determined by the Hund’s coupling J. Indeed, Fig. 6 shows that the<br />

results for different sets of calculations collapse on<strong>to</strong> a single curve, if the moment is plotted as a function of the<br />

effective S<strong>to</strong>ner parameterI = J −0.017U. For comparison, the inverse of the static susceptibility at x = 0 is<br />

also shown; the good agreement between the two curves indicates that the magnetically ordered phase is a stripe<br />

spin-density wave of quasiparticles.<br />

1.4 Interpretation of experimental spectra<br />

In Ref. [13] we compared model calculations of the intraband optical spectra, treated in a multi-band Eliashberg<br />

model, experimental data [36] and first-principles calculations, <strong>to</strong> show that interband transitions give a non<br />

negligible contribution already in the infrared region of the spectrum. This leads <strong>to</strong> a substantial failure of the<br />

extended-Drude-model analysis on the measured optical data without subtraction of interband contributions.<br />

5


As a consequence the electron-boson coupling gets dramatically overestimated. This effect was also recently<br />

measured by the Keimer Department. [37]<br />

In Ref. [14], in collaboration with the group of Bernhard Keimer, we have analyzed the anomalous dependence<br />

of the c-axis B 1g phonon in FeSe/Te samples, using Raman scattering and first-principles calculations.<br />

The local density calculations, including magnetic polarization and Fe nons<strong>to</strong>ichiometry in the virtual crystal<br />

approximation, reproduce well the experimental data in the Fe-deficient samples, while the effects of Fe excess<br />

are poorly reproduced. This may be due <strong>to</strong> excess Fe-induced local magnetism and low-energy magnetic<br />

fluctuations that cannot be treated accurately within these approaches. In Ref. [15], we have investigated the<br />

specific heat and upper critical fields of KFe 2 As 2 . The measured Sommerfeld coefficient γ n =94(3) mJ/mol K 2<br />

of KFe 2 As 2 is anomalously large, but this is due <strong>to</strong> extrinsic effects. Taking this in<strong>to</strong> account, KFe 2 As 2 should<br />

be classified as a weakly or intermediately coupled superconduc<strong>to</strong>r with a <strong>to</strong>tal electron-boson coupling constant<br />

λ <strong>to</strong>t =1 (including a calculated weak electron-phonon couplingλ ph =0.17).<br />

2 <strong>Superconductivity</strong> in intercalated picene<br />

Studying carbon, with its many allotropes and compounds, is one of the most fascinating problems in science.<br />

To mention only physics, before graphene, which was awarded the Nobel prize in 2010 for opening exciting perspectives<br />

for fundamental and applied research, many other systems such as nanotubes, fullerenes, intercalated<br />

graphites have played an important role in several fields, including superconductivity.<br />

5<br />

0 1 2 3 800 4 1600 2400 6 3200<br />

ω (cm -1 )<br />

1 2 3<br />

4 5 6<br />

Figure 7: Top: Structure of a picene<br />

molecule, and of the corresponding crystal.<br />

Bot<strong>to</strong>m: Phonon spectrum of crystalline picene<br />

in DFPT [23], showing representative vibration<br />

patterns. The modes 4 and 5, in red, give the<br />

highest contribution <strong>to</strong> ep coupling in our rigid<br />

band calculations. – From Ref. [45].<br />

The most recent discovery is a new class of aromatic superconduc<strong>to</strong>rs, comprising molecular crystals doped<br />

with alkali or alkaline earths metals. These are polycyclic aromatic hydrocarbons, i.e. planar molecules formed<br />

by a number of juxtaposed hexagonal benzene rings (C 6 H 6 ). The first report of superconductivity in this class of<br />

materials concerned picene doped with potassium (K) and rubidium (Rb) with a maximum critical temperature<br />

(T c ) of 18 K [38]. After that, superconductivity was also found in phenanthrene and coronene doped with K<br />

(T c =5K and 15 K), [39, 40] phenanthrene doped with strontium (Sr) and barium (Ba) (T c =5.6-5.8 K) [41], and<br />

1,2:8,9-dibenzopentacene(C 30 H 18 ) doped with K (T c =33 K) [42]. Intercalated hydrocarbons are strongly related<br />

<strong>to</strong> two families of carbon superconduc<strong>to</strong>rs: graphite intercalation compounds (GICs), which are conventional,<br />

BCS-like superconduc<strong>to</strong>rs [43] and alkali-doped fullerenes (A 3 C 60 ), which instead have a rich phase diagram<br />

determined by a non-trivial interplay between strong localep interactions and on-site Coulomb correlations in a<br />

highly non-adiabatic regime [44].<br />

In Ref. [45] we performed first-principles calculations of the electronic structure, phonon spectra and electronphonon<br />

(ep) interaction of doped picene in the rigid band approximation (RBA), using density functional perturbation<br />

theory (DFPT). Based on our results for the vibrational properties, and on previous studies on electronic<br />

properties and correlations, we showed that picene, and most likely other intercalated hydrocarbons, belong <strong>to</strong><br />

a class of strongly correlatedep superconduc<strong>to</strong>rs, for which a local approach that includes both the ep coupling<br />

6


and Coulomb correlations on an equal footing may be more appropriate. The main results of our study are summarized<br />

in Fig. 7. The <strong>to</strong>p panel shows the structure of the isolated picene molecules, and their arrangement in<br />

a crystal; the lower panel shows the spectrum of crystalline picene calculated in LDA; in the inset, typical displacements<br />

are shown. The modes indicated in red are those which give a larger contribution <strong>to</strong> theep coupling.<br />

The molecularep matrix element, isV ep = 150±20 meV for holes andV ep = 110±5 meV for electrons. These<br />

values are comparable <strong>to</strong> the bandwidth of the conduction band (W∼300 meV) and typical phonon frequencies<br />

(ω ph ∼ 200 meV).<br />

For a rigid-band doping of picene with ∼ 3 electrons per molecule, N(E F ) = 7.06 states spin −1 meV −1 , λ =<br />

0.78 andω ln = 1021 cm −1 . Usingµ ∗ = 0.12 in the McMillan formula, we obtain a critical temperatureT c = 56.5<br />

K. This is more than three times the experimental value of T c , and we need <strong>to</strong> use a larger value of µ ∗ = 0.23<br />

<strong>to</strong> reproduce the experimental T c . However, it is important <strong>to</strong> stress that the estimates of T c based on Migdal-<br />

Eliashberg theory must be taken with care in the case of picene. In fact, the bandwidth of the conduction bands<br />

(W ∼ 300 meV for electrons and ∼ 1 eV for holes), the frequencies of strongly coupled phonons (ω ph ∼ 200<br />

meV), theep coupling strength (V ep ∼ 110-150 eV), all have similar magnitudes, close also <strong>to</strong> that of Coulomb<br />

repulsion (U ∼ 1.2 eV), estimated by other authors [46]. In this range of parameters, the two most important<br />

approximations in Migdal-Eliashberg theory of superconductivity, Migdal’s theorem and the Morel-Anderson<br />

scheme for the screening of the Coulomb repulsion, are invalid. In fullerenes, this regime of parameters gives<br />

rise <strong>to</strong> a variety of interesting phenomena, the most spectacular being the occurrence of ep superconductivity<br />

near a Mott insulating phase, well captured by theoretical studies of local models of interacting phonons and<br />

electrons in the non-adiabatic regime. [44]<br />

This first study motivated us <strong>to</strong> start a collaboration on electronic and vibrational properties of intercalated<br />

hydrocarbons, with the high-pressure and optics group of Università la Sapienza, in Rome. [47]<br />

3 Iridates<br />

Recently, Ohgushi et al. reported resonant x-ray diffraction study of CaIrO 3 that indicates this material exhibits<br />

a Mott insulating J eff = 1/2 state [48]. Previously, it was found that spin-orbit coupling can play an important<br />

role in the electronic properties even for a 4d 5 system such as Sr 2 RhO 4 [49, 50]. Since CaIrO 3 is a 5d 5 system,<br />

this material provides a more oppurtune platform <strong>to</strong> study the interplay between spin-orbit coupling and onsite<br />

Coulomb repulsion. Furthermore, this material proffers a playground for a recent theory that suggests<br />

that systems in the J eff = 1/2 state, depending on bond geometry, lead <strong>to</strong> interesting varieties of low-energy<br />

Hamil<strong>to</strong>nians [51]. The density functional calculations within the local density approximation showed that the<br />

spin-orbit coupling splits the Ir t 2g bands in<strong>to</strong> fully filled quartet of J eff = 3/2 bands and half-filled doublet of<br />

J eff = 1/2 bands [52]. This metallic state is contrary <strong>to</strong> the experiments that show that CaIrO 3 is an insula<strong>to</strong>r.<br />

However, the half-filled J eff = 1/2 bands have a band width of ∼1 eV and a value for the on-site Coulomb<br />

repulsion ofU = 2.75 eV yields a Mott insulating state. The LDA+SO+U calculations give an antiferromagnetic<br />

ground state for CaIrO 3 along the c axis with <strong>to</strong>tal moments aligning antiparallel along the c axis, in agreement<br />

with the experiment. For U = 2.75 eV, the <strong>to</strong>tal magnetic moment is 0.67 µ B , with an orbital contribution of<br />

0.28 µ B and a spin contribution of 0.38 µ B . These values differ from what is expected for the ideal J eff = 1/2<br />

state, and this deviation might be explained by the mixing of J eff =1/2 bands with Ir e g bands due <strong>to</strong> the tilting<br />

of IrO 6 octahedra.<br />

Note: Our Minerva <strong>Research</strong> Group was estabilished in July 2009; before this, L. Boeri was a staff member in<br />

the Andersen Department (Electronic Structure Theory). For consistency, all the activity on fe-based superconduc<strong>to</strong>rs<br />

is reported here.<br />

References<br />

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

[2] J. Paglione and R. L. Greene, Nature Phys. 6, 645 (2010); D. C. Johns<strong>to</strong>n, Adv. Phys. 59, 803 (2010);<br />

special issue of Rep. Prog. Phys. , vol. 74, 2011.<br />

[3] J. Guo, S. Jin, G. Wang, S. Wang, K Zhu, T. Zhou, M. He, and X. Chen, Phys. Rev. B 82, 180520 (2010).<br />

[4] P. J. Hirschfeld, M. M. Korshunov, I. I. Mazin, Rep. Prog. Phys. 74, 124508 (2011).<br />

7


[5] <strong>to</strong> mention only a few early works: S. Lebegue, Phys. Rev. B 75 035110 (2007); D. J. Singh and M.-H. Du,<br />

Phys. Rev. Lett. 100, 237003 (2008); D. J. Singh, Phys. Rev. B 78, 094511 (2008); Alaska Subedi, Lijun<br />

Zhang, D. J. Singh, and M. H. Du, Phys. Rev. B 78, 134514 (2008); Alaska Subedi, Lijun Zhang, D. J.<br />

Singh, and M. H. Du, Phys. Rev. B 78, 134514 (2008); I. I. Mazin, D. J. Singh, M. D. Johannes, and M.<br />

H. Du, Phys. Rev. Lett. 101, 057003 (2008). K. Kuroki, Seiichiro Onari, Ryotaro Arita, Hide<strong>to</strong>mo Usui,<br />

Yukio Tanaka, Hiroshi Kontani, and Hideo Aoki, Phys. Rev. Lett. 101, 087004 (2008), Verónica Vildosola,<br />

Leonid Pourovskii, Ryotaro Arita, Silke Biermann, and An<strong>to</strong>ine Georges, Phys. Rev. B 78, 064518 (2008),<br />

for the basic electronic structure; For magnetism: T. Yildirim, Phys. Rev. Lett. 101, 057010 (2008); S.<br />

Ishibashi, K. Terakura, and H. Hosono, J. Phys. Soc. Jpn. 77 (2008) 053709; Z. P. Yin,et al, Phys. Rev.<br />

Lett. 101, 047001 (2008).<br />

[6] O. K. Andersen and L. Boeri, Ann. Phys. (Leipzig) 523, 8 (2011).<br />

[7] L. Boeri, O. V. Dolgov, and A. A. Golubov, Phys. Rev. Lett. 101, 026403 (2008).<br />

[8] L. Boeri, O. V. Dolgov, and A. A. Golubov, Physica C 469, 628 (2009).<br />

[9] L. Ortenzi, S. Biermann, O. K. Andersen, I. I. Mazin, and L. Boeri, Phys. Rev. B 83, 100505(R) (2011).<br />

[10] I. I. Mazin, M. D. Johannes, L. Boeri, K. Koepernik, and D. J. Singh, Phys. Rev. B 78, 085104 (2008).<br />

[11] L. Boeri, M. Calandra, I. I. Mazin, O. V. Dolgov, and F. Mauri, Phys. Rev. B 82, 020506 (2010).<br />

[12] T. Schickling, F. Gebhard, J. Bünemann, L. Boeri, O. K. Andersen, and W. Weber, Phys. Rev. Lett. 108,<br />

036406 (2012).<br />

[13] L. Benfat<strong>to</strong>, E. Cappelluti, L. Ortenzi, and L. Boeri, Phys. Rev. B 83, 224514 (2011).<br />

[14] Y. J. Um, A. Subedi, P. Toulemonde, A. Y. Ganin, L. Boeri, M. Rahlenbeck, Y. Liu, C. T. Lin, S. J. E.<br />

Carlsson, A. Sulpice, M. J. Rosseinsky, B. Keimer, and M. Le Tacon, Phys. Rev. B 85, 064519 (2012).<br />

[15] M. Abdel-Hafiez, S. Aswartham, S. Wurmehl, V. Grinenko, C. Hess, S.-L. Drechsler, S. Johns<strong>to</strong>n, A. U. B.<br />

Wolter, B. BÃijchner, H. Rosner, and L. Boeri, Phys. Rev. B 85, 134533 (2012).<br />

[16] P Richard, T Sa<strong>to</strong>, K Nakayama, T Takahashi and H Ding, Rep. Prog. Phys. 74, 12451 (2011).<br />

[17] A Carring<strong>to</strong>n, Rep. Prog. Phys. 74, 124507 (2011).<br />

[18] Kazuhiko Kuroki, Hide<strong>to</strong>mo Usui, Seiichiro Onari, Ryotaro Arita, and Hideo Aoki Phys. Rev. B 79, 224511<br />

(2009).<br />

[19] S. Graser et al., New J. Phys. 11, 025016 (2009); Ronny Thomale, Christian Platt, Werner Hanke, and B.<br />

Andrei Bernevig, Phys. Rev. Lett. 106, 187003 (2011), and many others.<br />

[20] L. Ortenzi, E. Cappelluti, L. Benfat<strong>to</strong>, and L. Pietronero, Phys. Rev. Lett. 103, 046404 (2009).<br />

[21] O. K. Andersen and T. Saha-Dasgupta, Phys. Rev. B 62, R16219 (2000).<br />

[22] C.-H. Lee, A. Iyo, H. Eisaki, H. Ki<strong>to</strong>, M.T. Fernandez-Diaz, T. I<strong>to</strong>, K. Kihou, H. Matsuhata, M. Braden,<br />

and K. Yamada, J. Phys. Soc. Jpn. 77, 083704 (2008).<br />

[23] Stefano Baroni, Stefano de Gironcoli, Andrea Dal Corso, and Paolo Giannozzi, Rev. Mod. Phys. 73, 515<br />

(2001).<br />

[24] W. Weber, L. F. Mattheis, Phys. Rev. B 25, 2248 and 2270 (1982).<br />

[25] L. Boeri, J. Kortus, and O. K. Andersen, Phys. Rev. Lett. 93, 237002 (2004).<br />

[26] Alaska Subedi and David J. Singh, Phys. Rev. B 78, 132511 (2008); Alaska Subedi, Lijun Zhang, D. J.<br />

Singh, and M. H. Du, ibid 134514; Alaska Subedi, David J. Singh, and Mao-Hua Du, ibid 060506 (2008).<br />

[27] C. de la Cruz, et al., Nature (London) 453, 899 (2008); (poly) m=0.36µ B ; more recent reports yield larger<br />

values of the magnetic moment. See: N. Qureshi et al. ,Phys. Rev. B 82, 184521 (2010). (polycrystals)<br />

m=0.63µ B ; H.F. Li et al., Phys. Rev. B 82, 064409 (2010) m=0.8µ B (single + poly-crystals).<br />

[28] M. Zbiri, et al, Phys. Rev. B 79, 064511 (2009);<br />

[29] D. Reznik, et al, Phys. Rev. B 80, 214534 (2009).<br />

8


[30] I. I. Mazin and M. D. Johannes, Nature Physics 5, 141 - 145 (2009).<br />

[31] O.K. Andersen, J. Madsen, U.K. Poulsen, O. Jepsen, and J. Kollár Physica B&C 86-88, 249 (1977).<br />

[32] L. Ortenzi, L. Boeri, P. Blaha and I.I. Mazin, submitted.<br />

[33] O. Gunnarsson, J. Phys. F: Metal Phys. 6, 587 (1976).<br />

[34] M. Aichhorn et al., Phys. Rev. B 80, 085101 (2009) and ibid., 054529 (2011); P. Hansmann et al., Phys.<br />

Rev. Lett. 104, 197002 (2010); Z.P. Yin, K. Haule, and G. Kotliar, Nature Phys. 7, 294 (2011).<br />

[35] M. C. Gutzwiller, Phys. Rev. 137, A1726 (1965).<br />

[36] M.M. Qazilbash, J.J. Hamlin, R.E. Baumbach, L. Zhang, D.J. Singh, M.B. Maple, and D.N. Basov, Nature<br />

Physics 5, 647 (2009).<br />

[37] A. Charnukha, P. Popovich, Y. Matiks, D.L. Sun, C.T. Lin, A.N. Yaresko, B. Keimer, and A. V. Boris, Nat.<br />

Comm. 2, 219 (2011).<br />

[38] R. Mitsuhashi, Y. Suzuki, Y. Yamanari, H. Mitamura. T. Kambe, N. Ikeda, H. Okamo<strong>to</strong>, A. Fujiwara, M.<br />

Yamaji, N. Kawasaki, Y. Maniwa, and Y. Kubozono, Nature (London) 464, 76 (2010).<br />

[39] X. F. Wang, R. H. Liu, Z. Gui, Y. L. Xie, Y. J. Yan, J. J. Ying, X. G. Luo, and X.H. Chen, Nat. Commun. 2,<br />

507 (2011).<br />

[40] Y. Kubozono, H. Mitamura, X. Lee, X. He, Y. Yamanari, Y. Takahashi, Y. Suzuki, Y. Kaji, R. Eguchi, K.<br />

Akaike, T. Kambe, H. Okamo<strong>to</strong>, A. Fujiwara, T. Ka<strong>to</strong>, T. Kosugi and H. Aoki, Phys. Chem. Chem. Phys.<br />

13, 16476 (2011).<br />

[41] X. F. Wang, Y. J. Yan, Z. Gui, R. H. Liu, J. J. Ying, X. G. Luo, and X. H. Chen, Phys. Rev. B 84, 214523<br />

(2011).<br />

[42] M. Xue, T. Cao, D. Wang, Y. Wu, H. Yang, X. Dong, J. He, F. Li, G. F. Chen, arXiv:1111.0820.<br />

[43] T. E. Weller et al., Nature Phys. 1, 39 (2005); N. Emery et al., Phys. Rev. Lett. 95, 087003 (2005); J. S.<br />

Kim et al., Phys. Rev. Lett. 99, 027001 (2007).<br />

[44] A. F. Hebard et al., Nature 350, 600 (1991); O. Gunnarsson, Alkali-Doped Fullerides (World Scientific,<br />

Singapore 2004); A. Y. Ganin et al., Nature Mater. 7, 367 (2008); M. Capone et al., Rev. Mod. Phys. 81,<br />

943 (2009).<br />

[45] A. Subedi and L. Boeri, Phys. Rev. B 84, 020508(R) (2011).<br />

[46] G. Giovannetti and M. Capone, Phys. Rev. B 83, 134508 (2011).<br />

[47] B. Joseph, L. Boeri et al., Journal of Physics Cond. Matter (2012).<br />

[48] K. Ohgushi, J.-I. Yamaura, H. Ohsumi, K. Sugimo<strong>to</strong>, S. Takeshita, A. Tokuda, H. Takagi, M. Takata, and<br />

T.-H. Arima, e-print arXiv:1108.4523.<br />

[49] M. W. Haverkort, I. S. Elfimov, L. H. Tjeng, G. A. Sawatzky, and A. Damascelli, Phys. Rev. Lett. 101,<br />

026406 (2008).<br />

[50] G.-Q. Liu, V. N. An<strong>to</strong>nov, O. Jepsen, and O. K. Andersen, Phys. Rev. Lett. 101, 026408 (2008).<br />

[51] G. Jackeli and G. Khaliullin, Phys. Rev. Lett. 102, 017205 (2009).<br />

[52] A. Subedi, Phys. Rev. B 85, 020408(R) (2012).<br />

9

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!