Optical and Magnetization Studies on Europium Based Iron Pnictides Von der Fakultät Mathematik und Physik der Universität Stuttgart zur Erlangung der Würde eines Doktors der Naturwissenschaften (Dr. rer. nat.) genehmigte Abhandlung Vorgelegt von Sina Maria Ute Zapf aus Offenbach am Main Hauptberichter: Prof. Dr. M. Dressel Erster Mitberichter: Prof. Dr. B. Keimer Zweiter Mitberichter: Prof. Dr. D. N. Basov Tag der mündlichen Prüfung: 18.12.2014 1. Physikalisches Institut der Universität Stuttgart 2015 Expect problems and eat them for breakfast. A. A. Montapert ABSTRACT Understanding the interplay of magnetism and superconductivity is one of the central top- ics in contemporary condensed matter research: while those are traditionally antagonistic phenomena, there is growing experimental evidence that unconventional superconductiv- ity - for instance in cuprates, heavy fermion or organic materials - is closely linked to magnetism [1–3]. Among all these material classes, the cuprates garnered the most at- tention because for more than two decades, they were the only known “high-temperature superconductors”. Only recently, they lost this monopoly position due to the discovery of relatively high critical temperatures in iron pnictides [4, 5]. One obvious common feature of cuprates and iron-based superconductors is an antifer- romagnetic parent phase. While undoped cuprates are Mott insulators, the magnetism in iron pnictides is usually considered as rather itinerant with a metallic spin density wave ground state. Chemical substitution or external pressure continuously suppress the magnetic transition temperature, until superconductivity emerges. Thus, magnetic fluc- tuations are widely debated as a key ingredient for high-temperature superconductivity. Recently, however, another similarity has moved into the focus of interest: at the nematic phase transition, electronic interactions break the rotational symmetry of the crystal, accompanied by strong fluctuations [6]. Compared to the antiferromagnetic phase, this transition usually takes place at slightly higher temperatures. Unfortunately, the exper- imental access to this exotic phase is challenging, as crystals usually appear twinned, meaning that domains mask the intrinsic anisotropy. The investigations carried out in the framework of this thesis mainly concentrate on europium based iron pnictides. These are a peculiar member of the 122 family as they develop at low temperatures (∼ 20K) an additional magnetic order of the local rare earth moments (see Fig. 0.1). The Eu2+ spin structure is a so-called “A-type” antiferromag- netism, meaning that spins of one layer are ferromagnetically aligned (along the a-axes), while neighbouring planes are coupled antiferromagnetically. Furthermore, neutron mea- surements revealed the possibility to detwin those compounds (i.e. to align the twinned domains along one direction) at low temperatures in laboratory-scale magnetic fields [7], a feat which is not possible in other iron pnictides and thus probably related to the rare iv Abstract earth magnetism. Therefore, europium based iron pnictides provide a unique platform to study the interplay of structural, magnetic and electronic effects in high-temperature superconductors. For this challenging purpose, we have employed SQUID magnetometry and Fourier-transform infrared spectroscopy on EuFe2(As1−xPx)2 single crystals; the main results are summarized as followed. Figure 0.1: Magnetic struc- ture of EuFe2As2 at 2.5K [8]. Fe2+ moments (brown ar- rows) align antiferromagnet- ically along the a- and the c-axis, but ferromagnetically along the b-axis. Eu2+ mo- ments (blue arrows) align within one europium layer ferromagnetically along the a-axis, and neighbouring ab- planes are coupled antiferro- magnetically. By systematic studies of the in- and out-of-plane magnetic properties of a series of EuFe2(As1−xPx)2 single crystals, we derived the complex magnetic phase diagram of eu- ropium based iron pnictides (see chapter 7, Ref. [9] and Ref. [10]). As depicted in Fig. 0.2, the Eu2+ magnetic ordering involves two steps for all phosphorus substituted compounds. The transition at a higher temperature TN is characterized throughout the whole phase di- agram by an antiferromagnetic interlayer coupling. One should note, however, that phos- phorus substitution induces a ferromagnetic out-of-plane component, leading to canted A-type antiferromagnetism. This canting develops rather quickly at low substitution lev- els, similar to the suppression of TN and the spin density wave transition temperature. These observations imply that the Eu2+ spin orientation is notably influenced by the spin density wave anisotropy. Thus, our measurements demonstrate significant interactions between the local Eu2+ and the (quasi-) itinerant spin density wave magnetism, which is consistent with theoretical calculations [11]. The second magnetic transition in the europium subsystem displays glass-like dynamics, affecting the coupling between the planes. As its transition temperature Tglass is lower than the onset of magnetic order, it is referred to as “re-entrant” spin glass phase. With increasing phosphorus substitution, Tglass rises quickly until it approaches TN, leading to successive transitions that are separated by ∼ 2K. Thus, only measurements with rather small temperature steps are able to resolve both phases. Nevertheless, we find indications for similar dynamics when reviewing published data of other europium based Abstract v compounds (see chapter 6). Therefore, we suggest that the re-entrant spin glass phase is the key to understand how bulk superconductivity can coexist with the usually strong Eu2+ magnetism. Figure 0.2: Phase Dia- gram of EuFe2(As1−xPx)2. TN (red dots) indicates a canted A-type antiferro- magnetic transition, Tglass (blue triangles) a spin glass transition, and Tc,on the onset of supercon- ductivity (grey squares). Closed symbols correspond to transition temperatures deduced from magneti- zation, open ones from resistivity measurements. Shadowed lines are guides to the eye. The light grey area indicates the onset of superconductivity, while bulk superconductivity is fully developed in the dark grey regime. Despite the interactions of local and itinerant magnetism revealed by our magnetization studies, the electrodynamic response of EuFe2As2 is surprisingly unaffected by the Eu2+ magnetic transition - the infrared spectra basically do not change when passing TN (see chapter 8). By comparing several fitting approaches, we find that the most realistic results are obtained with a model that contains two Drude components (probably representing electron and hole bands) and a Lorentzian around 1000 cm−1, which can be attributed to interband transitions involving Fe hole pockets [12]. Importantly, we are able to extract trends that are independent from the fitting method. For instance, we conclude that the spin density wave transition affects not only both Drude components, but also the spectral weight transfer, which appears in iron pnictides usually up to very high energies. These observations provide interesting insights into the relation of this transfer to the localization of itinerant carriers driven by correlation effects. Combining our expertise obtained from magnetization and optical studies on europium compounds, we have also performed magneto-optical measurements on EuFe2As2 (see chapter 9, [13]), which revealed a much more complex detwinning process than expected. vi Abstract At temperatures below the local Eu2+ ordering, the detwinning occurs in two steps, mean- ing that the crystal gets already partially detwinned with b ‖ H at ∼ 0.1T, followed by a reverse reorientation of structural domains with a ‖ H at ∼ 0.6T. Above TN, only the latter process takes place. We ascribe these observations to an indirect magneto-elastic coupling of the Eu2+ ions via the Fe2+ spins, which was - to our best knowledge - not pro- posed before. Moreover, we give a simple model for the two-step detwinning process based on the competition between magneto-crystalline anisotropy, antiferromagnetic exchange coupling and Zeeman energy (see Fig. 0.3). a a a a a a a b b b bbbb H twin BII twin AII (d) H ≥ H2 (c) H = HSF (b) H ≈ H1 (a) H = 0T Figure 0.3: Sketch of the twin distribution of EuFe2As2 dependent on an (from left to right increasing) external magnetic field H ‖ [110]T at T < TEu. (a) For H = 0T, the zero-field crystal is twinned and the domains are equally distributed. The Eu2+ spins are ordered A- type antiferromagnetically with the spin direction along the a-axis. (b) With external field, twin variant B‖ (red, bottom) with b ‖ H gets energetically favoured and therefore grows on the expense of variant A‖ (green, top), as soon as the energy difference exceeds the twin boundary pinning energy. (c) With further increasing field, Eu2+ spins in the remaining type A‖ twins flip along the field direction. Energetically, twin B‖ is still more favourable. (d) At slightly higher fields, twin A‖ is favoured and the crystal is detwinned with a ‖ H. Most importantly, the induced detwinning is persistent, meaning that a significant im- balance of twin domains remains when the magnetic field is switched off and the tempera- ture is raised up to the nematic and spin density wave transition. Such strongly hysteretic behaviour is surprising from a very fundamental point of view, as the detwinning that is induced at low temperatures with very small magnetic fields persists up to much higher temperatures, i.e. energy scales. Thus, our whole studies on europium based pnictides reveal a remarkable interdependence between magnetic, electronic and structural effects that might be very important to understand the unconventional superconductivity in these fascinating materials. ZUSAMMENFASSUNG Eines der zentralen Themen der heutigen Festkörperforschung ist es, die Wechselwirkungen zwischen Supraleitung und Magnetismus besser zu verstehen. Obwohl diese Phänomene traditionell nicht miteinander vereinbar sind, mehren sich die Hinweise, dass unkonventio- nelle Supraleitung, die beispielsweise in Kupferoxiden, Schweren Fermionen oder organi- schen Materialien vorkommt, überraschenderweise sehr eng mit Magnetismus verbunden ist [1–3]. Dabei stellen die Kupferoxide vermutlich die am meisten untersuchte Material- klasse dar, da sie mehr als 20 Jahre die einzigen bekannten Vertreter der “Hochtempera- tursupraleiter” waren. Tatsächlich ging diese Monopolstellung erst kürzlich verloren, als Supraleitung mit überraschend hohen Übergangstemperaturen in Eisenpniktiden entdeckt wurde [4, 5]. Vergleicht man verschiedene Hochtemperatursupraleiter, so ist eine offensichtliche Ge- meinsamkeit, dass Antiferromagnetismus eine wichtige Rolle im Phasendiagramm spielt: während undotierte Kupferoxide Mott-Isolatoren sind, geht man bei Eisenpniktiden in den meisten Fällen von einem metallischen Spindichtewellen-Grundzustand aus. Durch chemische Substitution oder das Anlegen von externem Druck wird dessen Übergangs- temperatur kontinuierlich verringert, bis schließlich Supraleitung einsetzt. Deshalb liegt es nahe, antiferromagnetische Fluktuationen mit dem Auftreten von Hochtemperatursupra- leitung zu assoziieren. Aktuell rückt allerdings eine weitere Gemeinsamkeit immer mehr in den Fokus der Forschung: bei dem sogenannten “nematischen” Übergang wird die Ro- tationssymmetrie des Kristalls durch elektronische Wechselwirkungen gebrochen; dabei treten sehr starke Fluktuationen auf [6]. Verglichen mit der antiferromagnetischen Phase findet dieser Übergang im Normalfall bei einer leicht höheren Temperatur statt. Leider ist seine Erforschung relativ kompliziert, da die resultierende strukturelle und elektronische Anisotropie durch Domänenbildung maskiert wird. Die im Rahmen dieser Arbeit durchgeführten Untersuchungen konzentrieren sich auf Europium-haltige Eisenpniktide. Diese entwickeln zusätzlich zur Spindichtewelle bei tie- fen Temperaturen (ca. 20K) eine lokale magnetische Ordnung. Hierbei sind die Eu2+ Spins innerhalb einer Ebene ferromagnetisch, aber benachbarte Ebenen antiferromagnetisch ge- koppelt - auch “A-typ” Antiferromagnetismus genannt (vgl. Abbildung 0.4). Desweiteren viii Zusammenfassung ging aus Neutronenstreuungs-Experimenten hervor, dass in EuFe2As2 relativ niedrige Ma- gnetfelder ∼ 1T ausreichen, um bei Temperaturen unterhalb der Eu2+-Ordnung alle Do- mänen gleich auszurichten [7]; in anderen Eisenpniktiden ist dies nicht möglich. Aus diesem Grund bieten Europium-haltige Eisenpniktide die außergewöhnliche Möglichkeit, struktu- relle, magnetische und elektronische Effekte, sowie deren Wechselwirkungen in Hochtem- peratursupraleitern zu untersuchen. Dafür wurden im Rahmen dieser Arbeit Magnetisie- rungsmessungen und Fourier-Transform-Infrarotspektroskopie an EuFe2(As1−xPx)2 Ein- kristallen durchgeführt. Die wichtigsten Ergebnisse werden im Folgenden vorgestellt. Abbildung 0.4: Magnetische Struktur von EuFe2As2 bei 2.5K [8]. Die Fe2+ Spins (brau- ne Pfeile) sind antiferroma- gnetisch entlang der a- und c-Achse, aber ferromagnetisch entlang der b-Achse ausgerich- tet. Die Eu2+ Spins (blaue Pfei- le) einer Ebene sind ferroma- gnetisch gekoppelt und zeigen in Richtung der a-Achse; benach- barte Ebenen sind antiferroma- gnetisch gekoppelt. Auf der Basis von richtungsabhängigen Untersuchungen der magnetischen Eigenschaften von EuFe2(As1−xPx)2 Einkristallen konnte das komplexe magnetische Phasendiagramm von Europium-haltigen Eisenpniktiden bestimmt werden; dieses ist in Abb. 0.5 zu sehen (s. Kapitel 7, Ref.[9] und Ref. [10]). Man stellt fest, dass die lokale Eu2+-Ordnung tatsächlich schrittweise abläuft. Betrachtet man zunächst den Übergang bei höherer Temperatur TN, so bleibt die Kopplung zwischen den Ebenen durch das gesamte Phasendiagramm antifer- romagnetisch. Allerdings entwickelt sich mit steigender Phosphor-Substitution rasch eine ferromagnetische Komponente entlang der c-Richtung; man spricht von “gekantetem” A- typ Antiferromagnetismus. Da die Zunahme dieser Komponente mit einer Unterdrückung der Ordnungstemperatur von Spindichtewelle und lokalem Eu2+-Magnetismus verbunden ist, liegt es nahe, dass die beiden magnetischen Ordnungen sich gegenseitig beeinflussen; dies wird auch durch theoretische Berechnungen unterstützt [11]. Der zweite magnetische Übergang bei tieferen Temperaturen Tglass ist glasartig und be- trifft die Kopplung zwischen den Ebenen. Da er einer magnetischen Ordnung folgt, spricht man auch von einem “re-entrant” Spin-Glas. Betrachtet man den Verlauf der Übergangs- temperatur Tglass, so steigt sie bei niedrigen Phosphor-Konzentrationen schnell an und Zusammenfassung ix Abbildung 0.5: Pha- sendiagramm von Eu- Fe2(As1−xPx)2. TN (rote Punkte) ist die Übergangs- temperatur des gekanteten A-typ Antiferromagnetismus, Tglass (blaue Dreiecke) die des Spin-Glas Übergangs, und Tc,on die der Supraleitung (graue Quadrate). Geschlos- sene Symbole wurden aus Magnetisierungsmessungen, offene aus Widerstandskur- ven extrahiert. Transparente Linien zeigen die wichtigen Trends; während die hell- graue Fläche das Einsetzen von Supraleitung markiert, entspricht die dunkelgraue Region der voll entwickelten Supraleitung. folgt dann TN im Abstand von etwa 2K. Aus diesem Grund kann dieses ungewöhnli- che Verhalten nur in Messungen mit sehr kleinen Temperaturschritten aufgelöst werden. Nichtsdestrotrotz wurden ähnliche Beobachtungen auch bei anderen Europium-haltigen Eisenpniktiden gemacht (s. Kapitel 6). Deshalb ist dieses glasartige Verhalten vermut- lich der Schlüssel, um die Koexistenz von Europium-Magnetismus und Supraleitung zu erklären. Trotz der Wechselwirkungen zwischen den magnetischen und elektronischen Subsyste- men, die durch diese Magnetisierungsmessungen nachgewiesen werden konnten, ist das optische Spektrum von EuFe2As2 überraschend unbeeinflusst von der lokalen Eu2+ Ord- nung (s. Kapitel 8). Der Vergleich mehrerer Modellierungsansätze ergibt, dass die rea- listischsten Ergebnisse durch eine Analyse mit zwei Drude- und einer Lorentzkomponen- ten (bei 1000 cm−1) erhalten werden. Dabei repräsentieren erstere vermutlich Elektronen- und Lochbänder, letztere Interbandübergänge in die Eisen-Lochtaschen [12]. Vor allem aber können durch diesen Vergleich Trends bestimmt werden, die unabhängig vom Mo- dellierungsansatz sind: der Spindichtewellen-Übergang beeinflusst nicht nur beide Drude- Komponenten, sondern auch den Transfer von spektralem Gewicht, der in Eisenpniktiden üblicherweise bis zu sehr hohen Energien auftritt. Diese Beobachtungen liefern interessante Informationen, inwiefern dieser Transfer mit der Lokalisierung von freien Ladungsträgern aufgrund von Korrelationseffekten zusammenhängt. x Zusammenfassung Die Erfahrungen mit magnetischen und optischen Messungen wurden schließlich ge- nutzt, um eine Reihe von magneto-optischen Untersuchungen an EuFe2As2 durchzuführen (s. Kapitel 9 und Ref. [13]). Dabei konnte aufgedeckt werden, dass dem Ausrichten der Domänen im Magnetfeld ein wesentlich komplexerer Prozess zu Grunde liegt, der jedoch durch ein relativ einfaches Modell erklärt werden kann (s. Abb. 0.6). Bei Temperaturen unterhalb von TN verläuft diese Ausrichtung in zwei Schritten: zunächst wird ein Großteil der Domänen schon bei sehr niedrigen Feldern von ∼ 0.1T mit der b-Achse parallel zum Magnetfeld H ausgerichtet; dem folgt eine Ausrichtung mit der a-Achse parallel zu H bei Feldern von etwa 0.6T. Oberhalb von TN findet nur letzterer Prozess statt. Dieses Verhal- ten deutet auf eine bis dahin unbekannte indirekte magneto-elastische Kopplung hin, bei denen die Eu2+-Spins über die Fe2+-Momente mit dem Gitter wechselwirken können. a a a a a a a b b b bbbb H twin BII twin AII (d) H ≥ H2 (c) H = HSF (b) H ≈ H1 (a) H = 0T Abbildung 0.6: Schematische Darstellung der Domänenverteilung in EuFe2As2 in Abhän- gigkeit eines (von links nach rechts zunehmenden) externen magnetischen Feldes H ‖ [110]T bei Temperaturen T < TEu. (a) Bei H = 0T sind die Domänen (“twins”) gleichmäßig verteilt, wenn der Kristall im Nullfeld abgekühlt wurde. (b) In einem externen Feld werden zunächst die Domänen B‖ (rot, unten) mit b ‖ H energetisch begünstigt, weswegen ihr Anteil auf Kosten der Domänen A‖ (grün, oben) steigt, sobald die Energie die der Haftung der Domänenwände übersteigt. (c) Wird das externe Feld weiter erhöht, kommt es zu einem Eu2+ Spin-Flip in den verbleibenden A‖-Domänen; energetisch sind jedoch noch die anderen bevorzugt. (d) Dies ändert sich erst bei leicht höheren Feldern und alle Domänen werden mit der a-Achse entlang von H ausgerichtet. Die vermutlich wichtigste Beobachtung bei diesen Studien ist, dass das magnetfeldin- duzierte Ungleichgewicht der Domänen bestehen bleibt, auch wenn das Feld ausgeschaltet und die Temperatur bis zum Spindichtewellen- und nematischen Übergang erhöht wird. Berücksichtigt man die relevanten Energieskalen, so ist eine solch robuste Hysterese tat- sächlich sehr überraschend. Somit konnten die im Rahmen dieser Arbeit durchgeführten Studien an Europium-haltigen Eisenpniktiden ein bemerkenswertes Zusammenspiel von magnetischen, elektronischen und strukturellen Effekten aufdecken, welches einen wichti- gen Einblick in die faszinierenden Zusammenhänge unkonventioneller Supraleiter gibt. PUBLICATIONS Some results of this work have been already published: • S. Zapf, D. Wu, L. Bogani, H. S. Jeevan, P. Gegenwart, and M. Dressel Varying Eu2+ magnetic order by chemical pressure in EuFe2(As1−xPx)2 Phys. Rev. B 84, 140503(R) (2011). • S. Zapf, B. Gorshunov, D. Wu, E. Zhukova, V. S. Nozdrin, S. Haindl, K. Iida, and M. Dressel Intra-gap absorption in superconducting Ba(Fe1−xCox)2As2 thin films studied by a Fabry-Pérot resonant technique J. Supercond. Nov. Magn. 21, 1557 (2013). • S. Zapf, H. S. Jeevan, T. Ivek, F. Pfister, F. Klingert, S. Jiang, D. Wu, P. Gegenwart, R. K. Kremer, and M. Dressel EuFe2(As1−xPx)2: reentrant spin glass and superconductivity Phys. Rev. Lett. 110, 237002 (2013). • A. S. Prokhorov, E. S. Zhukova, A. A. Boris, I. E. Spektor, B. P. Gorshunov, V. S. Nozdrin, E. A. Motovilova, L. S. Kadyrov, S. Zapf, S. Haindl, K. Iida, M. Dressel, K. V. Chizh, M. S. Storozhevykh, L. V. Arapkina, V. A. Chapnin, O. V. Uvarov, V. P. Kalinushkin, and V. A. Yuryev Submillimeter quasioptical spectroscopy of multilayer conducting and superconduct- ing systems Radiophys. Quant. Electr. 56, 620 (2014). • A. E. Karakozov, S. Zapf, B. Gorshunov, Y. G. Ponomarev, M. V. Magnitskaya, E. Zhukova, A. S. Prokhorov, V. B. Anzin, and S. Haindl Temperature dependence of the superfluid density as a probe for multiple gaps in Ba(Fe0.9Co0.1)2As2: Manifestation of three weakly interacting condensates Phys. Rev. B 90, 014506 (2014). • A. E. Karakozov, B. P. Gorshunov, Y. G. Ponomarev, A. S. Prokhorov, M. V. Magnitskaya, E. Zhukova, K. Iida, M. Dressel, S. Zapf, and S. Haindl xii Publications Studying the structure of the superconducting state in Co doped multiband BaFe2As2 compounds JETP Letters 100, 328 (2014). • S. Zapf, C. Stingl, K. W. Post, J. Maiwald, N. Bach, I. Pietsch, D. Neubauer, A. Löhle, C. Clauss, S. Jiang, H. S. Jeevan, D. N. Basov, P. Gegenwart, and M. Dressel Persistent detwinning of iron pnictides by small magnetic fields Phys. Rev. Lett. 113, 227001 (2014). • S. Zapf, D. Neubauer, A. Kadau, K. Post, H. S. Jeevan, P. Gegenwart, D. N. Basov, and M. Dressel Electronic scattering effects in Europium based iron pnictides in preparation. Additional scientific publications which are not presented in this thesis: • B. Gorshunov, E. Zhukova, V. I. Torgashev, L. Motovilova, L. S. Kadyrov, F. Fis- chgrabe, V. Moshnyaga, T. Zhang, R. Kremer, U. Pracht, S. Zapf, and M. Dressel Boson peak in overdoped manganites La1−xCaxMnO3 (0.5 ≤ x ≤ 1) Phys. Rev. B 87, 245124 (2013). • B. P. Gorshunov, E. S. Zhukova, V. I. Torgashev, L. S. Kadyrov, E. A. Motovilova, F. Fischgrabe, V. Moshnyaga, T. Zhang, R. Kremer, U. Pracht, S. Zapf, and M. Dressel Boson peak in manganites J. Phys.: Conf. Ser. 486, 012001 (2014). • S. Haas, E. Heintze, S. Zapf, B. Gorshunov, M. Dressel, and L. Bogani Direct observation of the discrete energy spectrum of two lanthanide-based single- chain magnets by far-infrared spectroscopy Phys. Rev. B 89, 174409 (2014). • L. S. Kadyrov, B. P. Gorshunov, E. S. Zhukova, V. I. Torgashev, A. S. Prokhorov, E. A. Motovilova, F. Fischgrabe, V. T. Moshnyaga, T. Zhang, R. K. Kremer, U. Pracht, S. Zapf, J. Pokorný, G. Untereiner, S. Kamba, and M. Dressel Terahertz-infrared electrodynamics of overdoped manganites La1−xCaxMnO3 Phase Transitions 87, 1050 (2014). Contents 1 Introduction 1 2 Magnetism 5 2.1 Basic definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 2.2 Magnetic interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 2.3 Types of magnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 3 Superconductivity 17 3.1 BCS theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 3.2 Unconventional high-temperature superconductivity . . . . . . . . . . . . . 18 3.2.1 Cuprates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 3.2.2 Iron-based superconductors . . . . . . . . . . . . . . . . . . . . . . . 22 3.2.3 Magnetic fluctuations . . . . . . . . . . . . . . . . . . . . . . . . . . 26 3.3 Superconductivity and Magnetism . . . . . . . . . . . . . . . . . . . . . . . 27 3.3.1 Type I and II superconductors . . . . . . . . . . . . . . . . . . . . . 27 3.3.2 Magnetization of a superconductor . . . . . . . . . . . . . . . . . . . 28 3.3.3 Upper critical field(s) . . . . . . . . . . . . . . . . . . . . . . . . . . 30 3.3.4 Exotic phases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 4 Electrodynamic properties 33 4.1 Basic definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 4.2 Metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 4.3 Broken symmetry states in metals . . . . . . . . . . . . . . . . . . . . . . . 39 4.3.1 General quantum mechanical description . . . . . . . . . . . . . . . . 39 4.3.2 Superconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 4.4 Iron-based superconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 4.4.1 Spectral features of the normal state . . . . . . . . . . . . . . . . . . 46 4.4.2 Analysis methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 4.4.3 Antiferromagnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 4.4.4 Superconductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 4.4.5 Unconventional normal state properties . . . . . . . . . . . . . . . . 53 5 Materials and methods 59 5.1 FTIR spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 5.2 SQUID magnetometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 5.3 Investigated samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 xiv Contents 6 Eu based iron pnictides 67 6.1 EuFe2As2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 6.2 Phase diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 6.2.1 Pressurized systems . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 6.2.2 Electron doping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 6.2.3 Eu dilution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 6.3 Electronic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86 6.3.1 Normal state properties . . . . . . . . . . . . . . . . . . . . . . . . . 86 6.3.2 Superconducting properties . . . . . . . . . . . . . . . . . . . . . . . 91 7 Magnetization studies on EuFe2(As1−xPx)2 95 7.1 Canted A-type antiferromagnetism . . . . . . . . . . . . . . . . . . . . . . . 95 7.2 Re-entrant spin glass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 7.3 Phase diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 7.4 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 8 Optical properties of EuFe2As2 117 8.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117 8.2 Multi-component analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120 9 Magnetic detwinning of EuFe2As2 131 9.1 Magneto-optical measurements . . . . . . . . . . . . . . . . . . . . . . . . . 132 9.2 Magnetization measurements . . . . . . . . . . . . . . . . . . . . . . . . . . 139 9.3 Detwinning mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143 9.4 Other techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145 9.5 Detwinning fraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 9.6 Further infrared studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149 9.7 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152 10 Summary and outlook 153 A THz investigations on Ba 122 thin films 157 A.1 Fabry-Pérot resonant technique . . . . . . . . . . . . . . . . . . . . . . . . . 158 A.2 Superfluid density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161 A.3 Irradiation studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162 B Acknowledgements 167 Bibliography 171 1 INTRODUCTION The discovery of high temperature superconductivity in cuprates as well as iron-based pnictides and chalcogenides unified two conventionally antagonistic fields of solid state physics: magnetism and superconductivity. While magnetic interactions were fascinat- ing already the ancient Greeks, superconductivity is a relatively “young” phenomenon, discovered only in 1911 [2, 14]. Conventional superconductors exhibit two characteristic properties below a critical tem- perature Tc: zero-resistivity and the expulsion of any external magnetic field. Both charac- teristics have a huge potential for applications, such as non-dissipative power lines, energy storages and levitating transport vehicles. Accordingly, enormous efforts were made to un- derstand the microscopic origin of superconductivity. Nevertheless, it took almost 50 years until Bardeen, Cooper and Schrieffer solved this puzzle by introducing a coherent ground state formed out of electron pairs with opposite spin and momentum [15]. As internal or external magnetic fields directly act on those properties, it is evident that magnetism and superconductivity are traditionally competing phenomena. One of the main drawbacks for applications is that the “glue” which holds those electron pairs together consists conventionally of electron-phonon interactions [16, 17], which re- quire very low temperatures to manifest [18]. Therefore, the discovery of high temperature superconductivity (up to 164K [19]) in cuprates [20] and iron-based superconductors [4] shook the scientific world and attracted tremendous interest even in the general public. 2 1 Introduction Interestingly, the phase diagram of both material classes contains a magnetically ordered phase that can be suppressed upon doping, chemical substitution or mechanical pressure, until superconductivity emerges [3]. While the parent compounds of cuprates are antifer- romagnetic Mott insulators, those of iron pnictides are metals that exhibit a spin density wave phase. A similar situation can be found in a variety of other superconductors with unconventional properties, such as heavy fermion or organic materials [1]. Therefore, the idea developed that magnetic fluctuations might be the “pairing glue” of unconventional superconductors. It turned out, however, that a uniform description of high-temperature superconduc- tors is much more complicated than expected. One reason is that a close inspection of their phase diagrams reveals a number of exotic phases that appear in the vicinity of the superconducting dome; it is not clear whether these additional phases support or com- pete with superconductivity [21]. One example that attracted recently a lot of attention is the nematic phase, which is the breaking of the rotational C4 symmetry driven by electronic interactions [6]. This means that additionally to a tetragonal-to-orthorhombic structural transition, an electronic in-plane anisotropy sets in, whose magnitude is much larger than one would expect from the lattice distortion. Unfortunately, the experimental investigation of this exotic phase is challenging, as the induced orthorhombic distortion is so small that crystals usually appear twinned, meaning that domains mask the intrinsic anisotropy. While in cuprates, mechanical clamps and magnetic fields were found to detwin single crystals [22, 23], in the case of iron pnictides, the effect of laboratory magnetic fields on the twin domain distribution is typically rather weak [24] and thus, only mechanical clamps are commonly used [6]. However, this introduces an explicit symmetry breaking by uniaxial pressure, meaning that one disturbs the anisotropy which one actually wants to measure [25]. EuFe2As2 is a peculiar parent compound of the iron pnictides, since, in addition to the spin density wave in the FeAs layers (∼ 190K), it exhibits long-range magnetic or- der of the Eu2+ local moments at low temperatures (∼ 19K) [26–28]. Nevertheless, bulk superconductivity up to ∼ 30K can be induced by mechanical pressure [29], isovalent substitution [30, 31] and doping [32–34]. Furthermore, neutron studies revealed that mag- netic fields in the range of 1T are able to detwin EuFe2As2 [7], allowing the study of the electronic in-plane anisotropy at typical laboratory-scale fields. Thus, Eu based iron pnictides offer the extraordinary possibility to investigate the interplay of unconventional electronic properties, superconductivity, itinerant and local magnetism. Therefore, sys- tematic (magneto-)optical and magnetization studies of EuFe2(As1−xPx)2 single crystals were performed in the framework of this thesis. 3The following text is organized as follows: first, the basic concepts of magnetic and superconducting materials are summarized (see chapters 2 and 3), focusing on high- temperature superconductivity and the interplay of magnetism and superconductivity. Subsequently, the electrodynamic properties of metals, broken symmetry phases, and in particular iron pnictides are reviewed (see chapter 4), which is followed by a brief sum- mary of the investigated materials and chosen methods (see chapter 5). The experimental part of this thesis is introduced in chapter 6 with a review of Eu based iron pnictides. This paves the way for the magnetic phase diagram of EuFe2(As1−xPx)2 that is gained by our magnetization studies presented in chapter 7; from the comparison with literature, we discuss further implications for other Eu based materials. Despite the peculiar local magnetic order, EuFe2As2 displays typical electrodynamic properties of correlated materials. Therefore, we use it in chapter 8 as a model system to compare various fitting approaches that are commonly applied to iron pnictides. Fi- nally, magneto-optical measurements on EuFe2As2 are presented that reveal a much more complicated magnetic detwinning mechanism than expected (see chapter 9), uncovering a unique possibility to study the intrinsic in-plane anisotropy of iron pnictides without the application of any symmetry-breaking external force. All those findings are concluded in chapter 10, evidencing a remarkable interplay of magnetic, electronic and structural effects in Eu based iron pnictides. 2 MAGNETISM Magnetism in condensed matter contains a wide variety of phenomena: the response of a solid on an external magnetic field depends delicately on whether its magnetic moments are interacting or basically isolated from each other. In this chapter, a sound basis should be established to understand the complex magnetic order(s) in Eu based iron pnictides. Therefore, important magnetic quantities are introduced, followed by a brief introduction of magnetic interactions in solids. Finally, different types of magnetism are categorized and phenomenologically described. If not declared explicitly, explanations follow Ref. [14]. 2.1 Basic definitions In free space, a magnetic field with strength H leads to a magnetic flux density B of: B = µ0H (2.1) with the permeability of free space µ0. 6 2 Magnetism However, if H acts on a magnetic solid, one has to account also for the magnetization M of this solid and equation 2.1 transfers into1: B = µ0(H + M). (2.2) The magnetic susceptibility χ describes, how much a material gets magnetized by a magnetic field and is generally defined (along a certain direction) as: χ = dM dH . (2.3) For a so-called “linear material”, equation 2.2 can be rewritten as: B = µ0(1 + χ)H = µ0µrH (2.4) with the relative permeability of the material µr = 1 + χ. When analysing experimental data, one further has to take into account that in magnetic solids, the internal magnetic field strength Hint is reduced with respect to the external one, Hext: Hint = Hext −NM (2.5) with the demagnetization factor N , which depends on the geometry of the sample and varies 0 ≤ N ≤ 1. An intuitive picture for this field reduction is the following [35]: in a magnetized sam- ple, the magnetic moments are preferentially aligned along one direction. Imagining these magnetic moments as miniature magnets with north and south poles (see Fig. 2.1a), neigh- bouring poles will cancel each other out; only poles at the surface will be left (see Fig. 2.1b). The latter will cause an internal magnetic field, which is opposed to the magnetization and therefore called “demagnetization field”. Its strength Hd will be very sensitive to the shape of the material. For a very thin sample and an in-plane magnetic field, there is basi- cally no place for the formation of these surface poles, causing negligible demagnetization effects (N → 0). For the out-of-plane direction, however, the situation is vice versa and the demagnetization factor approaches unity. Thus, the experimentally determined susceptibility of a linear material is actually: χexp = M Hext = M Hint +NM = χintrinsic1 +Nχintrinsic (2.6) 1 With a relative permeability of µr ≈ 1. 2.2 Magnetic interactions 7 Figure 2.1: Schematic drawing to explain the demagnetization field (adapted from Ref. [35]). (a) In a magnetized sample, magnetic moments depicted as miniature magnets with north (N) and south (S) poles are preferentially aligned. (b) As neighbouring poles will cancel each other out, only poles at the surface will be left. These cause an internal magnetic field (“demagnetization field” with strength Hd) opposed to the magnetization M. with the intrinsic susceptibility: χintrinsic = M Hint (2.7) As usually χintrinsic  1, demagnetization effects are often negligible. However, sample shape effects might determine the measured susceptibility in ferromagnets (see section 2.3), causing χexp → 1/N . 2.2 Magnetic interactions In order to describe magnetic long range order, one has to understand the underlying interactions between the magnetic moments. Usually, magnetic dipolar interactions are negligible, and the dominant mechanism can be described by exchange interactions via the Heisenberg Hamiltonian: H = −2 ∑ i>j JijSiSj (2.8) with the exchange constant Jij between the ith and jth spin, whose sign determines the magnetic ground state. If the overlap between neighbouring magnetic atoms is high enough, direct exchange interaction is possible. However, this is often not the case, for instance for the 4f -orbitals of rare earth elements, and indirect exchange has to be considered. In metals, conduction electrons can mediate the exchange between local moments via spin-polarization. This is known as “RKKY-exchange”, named after the discoverers Ruderman, Kittel, Kasuya and Yosida [36]. Depending on the details of the Fermi surface and the distance between the magnetic moments, the RKKY-exchange constant oscillates between positive and negative sign. 8 2 Magnetism 2.3 Types of magnetism In this section, different types of magnetism are phenomenologically described. Starting with the behaviour of non-interacting moments in a magnetic field, diamagnetism and paramagnetism are introduced. In contrary, ferro-, antiferro-, and helimagnetic materials show long-range magnetic order, where the involved magnetic moments act cooperatively. Finally, also glassy short-range dynamics are discussed. Diamagnetism All materials exhibit diamagnetism: due to the orbital motion of the electrons, an external magnetic field induces a magnetic moment opposed to the direction of the applied field. This process is rather temperature-independent (see Fig. 2.2a). However, the induced negative susceptibility (see Fig. 2.2b) is usually much smaller than other magnetic contri- butions. One exception are superconductors which are “perfect diamagnets”, as will be discussed in more detail in chapter 3.3.1. Figure 2.2: Typical (a) temperature- and (b) field-dependent magnetization of a diamagnetic material. The susceptibility is temperature-independent and negative. Paramagnetism The key ingredient for paramagnetism are unpaired, non-interacting electrons: while their magnetic moments do not have any preferred orientation in zero field, they can be aligned by an external magnetic field. The degree of alignment depends on the strength of the external field as well as on the temperature, as thermal fluctuations counteract the field- enforced order. 2.3 Types of magnetism 9 Figure 2.3: Typical (a) temperature- and (b) field-dependent magnetization of a paramag- netic material. (a) The temperature-dependent magnetization can be described by the Curie law χ ∝ 1/T . (b) The magnetization follows with increasing field a Brillouin function, which is almost linear at low fields (as can be seen by the red line, a guide to the eye). From a quantum mechanical description of non-interacting electrons with total angular momentum J = L + S (L: orbital momentum, S: spin momentum), one can derive the magnetic field dependence of a paramagnet as (see Fig. 2.3b): M = MsBJ ( gJµBJB kBT ) (2.9) with the saturation magnetization: Ms = ngJµBJ, (2.10) the Brillouin function: BJ(x) = 2J + 1 2J coth (2J + 1 2J x ) − 12J coth ( x 2J ) , (2.11) the Landé g-factor: gJ = 3 2 + S(S + 1)− L(L+ 1) 2J(J + 1) , (2.12) the number of spins per unit volume n, the Bohr magneton µB, the Boltzmann constant kB, and the temperature T . In typical paramagnetic materials, except for very high fields and very low tempera- tures, the magnetization shows a linear field dependence and the temperature-dependent susceptibility follows the Curie law (see Fig. 2.3a): χ = nµ0µ 2 eff 3kB 1 T (2.13) 10 2 Magnetism with the effective magnetic moment: µeff = gJµB √ J(J + 1). (2.14) Ferromagnetism Due to exchange interactions, a ferromagnet exhibits below the Curie temperature TC - even in the absence of any external field - a spontaneous magnetization, as all magnetic moments align along one direction. The corresponding Hamiltonian (including an applied magnetic field) can be written as: H = −2 ∑ i>j JijSiSj + gµB ∑ j SjB (2.15) with the first term expressing the Heisenberg exchange energy with Jij > 0 (see equa- tion 2.8) and the second term giving the Zeeman energy in an external field (g: the electron’s g-factor). This Hamiltonian can be solved using a mean field approach. It follows that at low temperatures, the magnetic order is even in the absence of any external field self-sustaining, meaning that the alignment of moments causes an internal (molecular) field, that in turn favours the parallel alignment of the spins. This results in a spontaneous magnetization. At elevated temperatures, thermal fluctuations counteract this process until any order is destroyed at the Curie temperature. The transition itself is a second-order phase transition, as the derivative of the spontaneous magnetization (and thus the second derivative of the system’s free energy) shows a discontinuous jump at TC. However, the transition smears out for a finite external magnetic field and thus, no more “true” phase transition takes place. In magnetization measurements, three characteristic observations are commonly used to identify ferromagnetic order (see Fig. 2.4): • For temperatures T > TC, the susceptibility follows at small external fields the Curie-Weiss law: χ ∝ 1 T − TC . (2.16) • At T < TC, the field-dependent magnetization resembles that of a paramagnet, however, the saturation fields are much lower. • Most remarkably, irreversibility effects appear in the temperature- and field-dependent magnetization at temperatures T < TC. One has to distinguish whether the crystal was cooled down through TC with or without external magnetic field. 2.3 Types of magnetism 11 Figure 2.4: Typical (a) temperature- and (b) field-dependent magnetization of a ferro- magnetic material. (a) For T > TC, the temperature-dependent magnetization follows the Curie-Weiss law χ ∝ 1/(T − TC). For T < TC, the magnetization increases with increasing temperature if the crystal was cooled down in zero field (ZFC), and decreases if the crystal was cooled down in an external field (FC). (b) The field-dependent magnetization shows at temperatures T < TC a hysteresis, defined by the virgin curve, saturation magnetization Ms, remanent magnetization Mr, and coercive field Hc. These irreversibility effects are also known as “magnetic hysteresis”. They result, as ferromagnets typically consist of many domains, which are energetically favourable as they reduce the demagnetization fields introduced in section 2.1. Whereas in each domain, the magnetic moments are uniformly aligned, different domains (separated by domain walls) are randomly oriented with respect to each other. Thus, the total net magnetization in the absence of any external field is zero. An external magnetic field is in principle able to move domain walls. However, they can be pinned due to strains or impurities. Thus, only if the magnetic field overcomes the associated pinning energy, the domain wall can move further. At higher fields, it is also possible that all spins of one domain rotate at once. In any case, one has to consider also the magneto-crystalline anisotropy, meaning that crystals posses crystallographic directions that are preferred (“easy axis”) or unpreferred (“hard axis”) in a magnetization process. Thus, with increasing field, one single domain will be created where the magnetic moments point along the easy axis closest to the direction of the external field. Only when the magnetic field energy gets high enough, the moments will rotate away from the easy axis to the direction of the external field. As in all those dynamics certain energy thresholds need to be overcome, a magnetic hysteresis results. The hysteresis in field-dependent magnetization measurements is defined by (see Fig. 2.4b): • the initial magnetization curve after the crystal was cooled down in zero magnetic field to temperatures T < TC, called “virgin curve” • the saturation magnetization Ms 12 2 Magnetism • the remanent magnetization Mr which stays after the sample was once fully magne- tized and the field was switched off • the coercive field Hc which needs to be applied to reduce the magnetization from the remanent value to zero. For temperature-dependent magnetization measurements (in a small magnetic field), one has to distinguish below TC (see Fig. 2.4b): • zero-field cooled (ZFC) curve: the magnetization increases (from zero) with increas- ing temperature, as thermal fluctuations allow more and more domains to be oriented along the field • field-cooled (FC) curve: as the symmetry of the system was broken externally while cooling through TC, a preferred spin direction was initialized; thermal fluctuations are weakening the order and lead to a decrease of the magnetization with increasing temperature. Antiferromagnetism For an exchange constant Jij < 0, neighbouring spins tend to align antiparallel. This leads below the Néel temperature TN to antiferromagnetic order. For a theoretical description following equation 2.15, an antiferromagnet can be considered to consist of two opposite, interpenetrating ferromagnetic sublattices. Antiferromagnetic behaviour can be usually identified by (see Fig. 2.5): • For temperatures T > TN, the susceptibility follows at small external fields the Curie-Weiss law χ ∝ 1 T + TN . (2.17) • At T < TN and relatively small fields, the temperature-dependent magnetization is strongly direction-dependent. • At T < TN, the field-dependent magnetization along the spin direction shows a step before achieving saturation. • The temperature- and field-dependent magnetization does not display hysteretic behaviour. In order to understand the direction-dependence of the magnetization below the Néel temperature, one should first consider the magnetization at small fields applied at 0K 2.3 Types of magnetism 13 Figure 2.5: Typical (a) temperature- and (b) field-dependent magnetization of an antifer- romagnetic material. (a) For T > TN, the temperature-dependent magnetization follows the Curie-Weiss law χ ∝ 1/(T + TN). For T < TN, the magnetization increases for the direction parallel to the external field, and stays constant for the perpendicular direction. (b) The field- dependent magnetization along the spin direction rises (at 0K) abruptly at the spin-flop field HSF and increases further until saturation is reached at HS. For finite temperatures below TN, this curve smears out due to thermal fluctuations. Brown arrows display schematically the field-dependent spin alignment. parallel (M‖) and perpendicular (M⊥) to the spin direction (see Fig. 2.5a): while for the former case, small magnetic fields will not affect the antiparallel aligned spins and therefore the net magnetization is zero, in the latter case, the spins will be tilted slightly out of the balance position, leading to non-zero magnetization. With increasing temperature, the thermal fluctuations counteract the exchange interactions. This does not notably influence the effect perpendicular to the spin direction. Therefore, M⊥ is almost flat for temperatures T < TN. However, for the field parallel to the spin direction, the field can act on the thermally excited spins and thus M‖ increases with increasing temperature up to TN. In the case of high magnetic fields [37], it is clear that at one point the external field must dominate over the internal exchange interaction, leading to a parallel alignment of spins along the field direction. If the field acts perpendicular to the spin direction, this will be a smooth process. However, the situation gets more complicated if the external field is aligned parallel to the spin direction (see Fig. 2.5b). In this case, Zeeman energy, exchange energy and magneto-crystalline anisotropy compete with each other, leading (at T = 0K) to the total energy: E = −MB(cosθ + cosφ) + JM2cos(θ + φ)− 12∆(cos 2θ + cos2φ) (2.18) 14 2 Magnetism with the exchange coupling constant J and the magneto-crystalline anisotropy ∆. Thus, the ground state energy of the antiferromagnetic configuration is independent of the ex- ternal field: E0 = −JM2 −∆. (2.19) For higher fields, a “spin-flop” transition takes place: the spins rotate abruptly to a config- uration where the angle between spins and external field is < 90◦ and the same (counted clockwise and counter-clockwise) for both sublattices (see Fig. 2.5b). The energy of the spin-flop phase is therefore: E = −2MBcosθ + JM2cos2θ −∆cos2θ. (2.20) Minimizing this energy with respect to θ and re-inserting the result in equation 2.20, yields the minimum energy of the spin-flop phase: E0,SF = −JM2 − M 2B2 2JM2 −∆. (2.21) Comparing equations 2.19 and 2.21, the spin-flop takes place at: BSF = √ 2J∆− (∆/M)2. (2.22) The corresponding field-dependent magnetization is depicted in Fig. 2.5b. The magneti- zation rises abruptly at the spin-flop field and increases afterwards steadily, as the spins align gradually along the external field, until they are fully aligned. One should also note that if the magneto-crystalline anisotropy is very strong, the spins can directly rotate from the antiparallel to the parallel alignment (with respect to the external field). This process is called “spin-flip”. Helimagnetism There exists a wide variety of other magnetic phases despite anti- und ferromagnetism, often related to one of them or even both. One possible ground state of particular interest for layered materials (such as several rare earth systems) is helimagnetism. In this case, the spins are ordered ferromagnetically within one layer, however the magnetic moments of neighbouring layers differ by a finite angle that is neither equal to zero (which would cor- respond to ferromagnetism) nor equal to pi (which would correspond to antiferromagnetic interlayer coupling). Thus, the spins will rotate along a spiral, which is not necessarily commensurate with the lattice. 2.3 Types of magnetism 15 Spin glass In materials where frustration and disorder play an important role, magnetic long-range order is prohibited [38]. In particular the RKKY-interaction, which oscillates between antiferromagnetic and ferromagnetic coupling, often leads to competing interactions. In these materials, one can distinguish in principle four temperature ranges: at very high temperatures, the spins are independent due to strong thermal fluctuations. When low- ering the temperature, first local and then also long-range interactions will occur: the system starts seeking its ground state. However, due to the competing interactions, it gets stuck at the so-called “freezing temperature” Tf in a metastable state. As this “spin glass” phase is only metastable, hysteretic and time-dependent behaviour can be observed. There exist several types of spin glasses. Systems with dominant short range interactions are sometimes called “mictomagnetic” or “cluster glass”, as the short range interactions lead to magnetic clusters that freeze in at Tf. It is also possible that a magnetically ordered state transforms at lower temperatures to a spin glass - although this is counterintuitive as one would expect a more ordered state at lower temperatures. The low-temperature phase resembles then a cluster glass, however, in order to make clear that this phase originates from another state than paramagnetism, it is referred to as “re-entrant spin glass”. One theoretical explanation for this behaviour is that only the transverse spin components freeze at the glass transition [39]. 3 SUPERCONDUCTIVITY In the beginning of the 20th century, cryogenics was still in its early infancy. It was not before 1908 that H. K. Onnes managed to liquify 4He, pushing the experimentally accessible temperature range down to a few Kelvin. This opened a completely new world in condensed matter research. Only three years later, he observed during his studies on Hg at a critical temperature Tc ∼ 4.2K an abrupt resistivity drop to zero - superconductivity, until then an unknown state of matter, was discovered. More than twenty years later, W. Meissner and R. Ochsenfeld found out that a super- conductor is not just an ideal conductor: it is also a perfect diamagnet, meaning that magnetic fields can not penetrate into its bulk. Dubbed after its discoverers, this second characteristic of a superconductor is known today as “Meissner Ochsenfeld effect”. Although great progress was achieved in experiments and more and more superconduct- ing compounds were found, a microscopic theoretical description was only developed in 1957 by J. Bardeen, L. N. Cooper and J. R. Schrieffer. In the following chapter, a brief introduction in their “BCS theory” will pave the way to distinguish between conventional and unconventional superconductivity. Particular focus will be put on the discussion of high-temperature superconductors. Although magnetism and superconductivity are usually antagonistic phenomena, magnetic fluctuations might be the key to understand their extraordinary properties. This motivates some further insights into the interplay of magnetism and superconductivity. If not given explicitly, explanations follow Ref. [2] and [40] or references given at the beginning of a section. 18 3 Superconductivity 3.1 BCS theory Early theoretical descriptions of superconductivity were based on macroscopic, phenomeno- logical models. The success of those approaches can be still seen today, as hereby defined characteristic quantities of the superconducting state such as the London penetration depth and the Ginzburg-Landau coherence length are still widely used1. In 1950, the observation of the so-called “isotope effect” introduced a relationship be- tween the onset of superconductivity and the atomic mass of a superconducting com- pound [41, 42], indicating that superconductivity arises from interactions between elec- trons and lattice vibrations. A theoretical description of those interactions [16, 17] paved the way for the microscopic BCS theory [15]. The fundamental idea behind BCS theory is that electrons of opposite spin and momen- tum form pairs (so-called “Cooper pairs”), assuming an attractive potential due to weak electron-phonon-coupling that exceeds the repulsive Coulomb interactions. Those Cooper pairs condense in a coherent ground state, leading to a gap in the density of states. This energy gap is the order parameter of the superconducting state; its squared absolute value proportional to the Cooper pair density. Originally, the effective attractive potential between electrons was assumed to be inde- pendent of momentum, leading to an isotropic s-wave gap: 2∆0(T = 0) = 3.5kBTc (3.1) with a temperature-dependence that can be well approximated by: ∆(T ) ∆(0) = tanh [ Tc T ∆(T ) ∆(0) ] . (3.2) The experimental verification of this energy gap, demonstrated by optical spectroscopy [43], was one of the big successes of BCS theory. 3.2 Unconventional high-temperature superconductivity Although BCS theory describes successfully a huge variety of superconductors, there exist also some materials where it fails. Even in some elemental superconductors such as Pb and 1 The London penetration depth gives the distance where an (electro-) magnetic field inside a super- conductor is screened to 1/e of its initial value; the Ginzburg-Landau coherence length describes the length scale over which the superconducting order parameter varies due to a local distortion, such as a vortex. 3.2 Unconventional high-temperature superconductivity 19 Hg, the observed behaviour slightly differs from BCS predictions. It turns out that the latter compounds are still “conventional superconductors” with a strong electron-phonon coupling and Eliashberg theory can achieve a proper description [44]. However, if the coupling between electrons is not mediated via phonons, total spin or orbital momentum are non-zero, superconductivity is inhomogeneous or magnetic interactions are observed, the term “unconventional superconductivity” is used. Famous examples are heavy fermion superconductors, ruthenates, cuprates, organic and iron-based superconductors [1]. One subclass of unconventional superconductors are high-temperature superconductors, which have in common that (as their name already suggests) their critical temperature exceeds any limit accessible by electron-phonon interactions [18, 45]. Up to now, only two classes of high-temperature superconductors are known: the cuprates and the iron-based superconductors, whose high critical temperatures were discovered in 1986 by G. Bednorz and K. A. Müller [20] and in 2008 by the group of H. Hosono [4], respectively. 3.2.1 Cuprates Cuprates are layered materials, where superconducting CuO2 planes alternate along the crystallographic c-axis with spacer layers, resembling in total perovskite materials. Prob- ably the most studied compounds are doped YBa2Cu3O7 (“YBCO”, [46]) and Bi2Sr2- CaCu2O8 (“BSCCO”, [47]) which show maximum critical temperatures around ∼ 90K. At present, the highest observed Tc is 164K for HgBa2Ca2Cu3O8 under pressure [19]. In the following, the generic phase diagram of cuprates is described (see Fig. 3.1), giving a basis to elaborate similarities and differences to iron pnictides. Tuning through the phase diagram results from electron or hole doping1; the spacer layers serve as charge reservoir. Antiferromagnetic Mott insulator The parent phase of cuprates is an antiferromagnetic Mott insulator that can be well described by the Hubbard model: although the conduction band is half filled and thus metallic behaviour would be expected, correlations (i.e. the Coulomb repulsion) between electrons are strong enough to localize the electrons. With electron or hole doping, the Mott phase gets suppressed. d-wave superconductivity While the Cooper pairs are conventional singlets, the superconducting order parameter has unconventional dx2−y2-wave symmetry. This means that the gap changes sign along 1 The phase diagrams for electron and hole doping are slightly asymmetric: for electron doping, the Mott phase is usually more robust and the superconducting dome narrower. 20 3 Superconductivity Figure 3.1: Generic phase diagram of cuprates. The parent compounds are antiferromagnetic Mott insulators (green). With doping, the magnetic long range order is suppressed and a pseudogap phase (grey) appears, separating the Mott phase and the superconducting dome (SC: blue). Above the superconducting dome, strange metal behaviour (red) is observed, with a crossover to a Fermi liquid (yellow) in the overdoped range. This might indicate a quantum critical point (QCP: red) buried by the superconducting dome. the Fermi surface, leading to nodes in the order parameter at the diagonals of the Brillouin zone, where the superconducting gap vanishes (see Fig. 3.2a,b). The maximum gap value does not follow the BCS relation given in equation 3.1 and is typically much higher. There is now broad consensus that the main pairing interactions take place between car- riers within the CuO2 layers. As the coherence length is rather small, a three-dimensional superconducting state appears due to Cooper pair tunneling along the c-axis. Thus, cuprates display natural Josephson junctions. Unconventional normal state properties One of the main problems about a theoretical description of superconductivity in cuprates is that the normal state from which superconductivity arises is still not understood: when passing at high temperatures through the phase diagram, the Mott insulator gets replaced by a “pseudogap” phase, which is followed by “strange metal” behaviour; conventional Fermi liquid1 behaviour sets in only in the overdoped region. In the pseudogap phase, a momentum-dependent gap appears [48]. While its absolute value does not vary strongly with temperature, the gap opens for different momenta at 1 Landau’s Fermi liquid theory describes a system of interacting fermions by introducing long living excitations. Those quasiparticles possess renormalized dynamical properties (such as an effective mass), but can be described like a free electron gas. This results at sufficiently low temperature in a T 2- and ω2-dependence of the scattering rate. 3.2 Unconventional high-temperature superconductivity 21 Figure 3.2: (a) Schematic Fermi surface of the cuprates. The grey shaded area marks the occupied electron states. (b) Fermi surface superimposed with the dx2−y2 -wave symmetry of the superconducting order parameter. Red and blue correspond to different signs of the order parameter, which has nodes (white) along the diagonals of the Fermi surface. (c) Development of the Fermi surface: the normal state Fermi surface (T > T ∗) is first reduced to Fermi arcs in the pseudogap phase (Tc < T < T ∗), which develop until superconductivity sets in (T < Tc) to nodes. different temperatures. This leads to a reduction of the Fermi surface to arcs1, which shrink further with decreasing temperature (see Fig. 3.2c). When superconductivity sets in, almost the entire Fermi surface is already gapped: only the region at the nodes is still occupied. Those nodes remain in the superconducting phase, consistent with the dx2−y2- wave symmetry of the superconducting order parameter. Due to those (and many more) similarities of the two gaps, the idea was developed whether the pseudogap phase is a precursor of superconductivity, where non-coherent Cooper pairs form. In contrast, recent results point more towards another explanation: the pseudogap seems to be a separate phase. Indeed, ARPES and high-resolution STM measurements found a small feature which they identified as a superconducting gap, with a smaller absolute value than the pseudogap [52–54]. However, there is not yet consensus about the nature of the pseudogap. To complicate this issue further, there are more and more additional phases found by careful studies on cuprates such as (fluctuating) charge density waves [55–58] and nematicity [59, 60]. Whether all those orders are competing with superconductivity, 1 Recent quantum oscillation measurements indicate that those arcs could be indeed small electron pockets; charge order wave vectors that reconnect the remaining arcs might play a crucial role [49–51]. 22 3 Superconductivity a necessary key ingredient, or just appear due to an unknown “mother phase” needs to be seen [21]. Another mystery in cuprates is the strange metal behaviour right above the supercon- ducting dome. In this regime, the Mott-Ioffe-Regel limit is violated, meaning that the mean free path of carriers is less than the interatomic distance [62, 63]. Thus, coherent transport should not be possible and incoherent dynamics set in, as will be further dis- cussed on the basis of optical measurements in chapter 4. Furthermore, the resistivity displays over a wide temperature range linear behaviour, which might stem from spin fluctuations or a quantum critical point1 [64]. 3.2.2 Iron-based superconductors Similar to cuprates, iron-based superconductors are layered materials [3, 5, 65]. Their common ingredient are layers where Fe2+ is tetrahedrally surrounded by pnictogen or chalcogen atoms. Thus, one differs between iron pnictides and iron chalcogenides. Further classification results from the stoichiometric composition of the parent compounds. According to this classification scheme, LaFeAsO1−xFx (which originally caused the “iron age of superconductivity” [69]) belongs to the “1111”-family. Replacing La by other rare earth elements increases Tc up to 56K in bulk materials; recent work on FeSe (“11”- chalcogenide) monolayers indicated even higher accessible critical temperatures [67]. How- ever, most research was done up to now on “122” iron pnictides due to the availability of high quality single crystals, focusing on materials with BaFe2As2 as parent compound (see Fig. 3.3a). Eu based iron pnictides also belong to this family; more detailed discussions concerning their magnetic and electronic properties can be found in chapter 6. Here, the focus is on the generic phase diagram of 122 iron pnictides (see Fig. 3.3b); tuning through the phase diagram can be reached by electron or hole doping2, mechanical pressure, as well as isovalent substitution. Antiferromagnetism Iron pnictides are multiband systems, where up to five Fe 3d-bands cross the Fermi energy. This leads to a Fermi surface that consists usually of multiple quasi-two-dimensional hole and electron pockets centered around the (0,0)-“Γ”- and (pi,0)/(0,pi)-“X”-points of the 1 A quantum critical point is a second order phase transition that takes place at absolute zero tempera- ture. It leads to a divergence of several physical quantities. 2 In literature there is an ongoing debate, whether superconductivity is induced in iron pnictides and chalcogenides due to doping or disorder [68]. However, this should not be in the focus of this thesis and for simplicity, the expression “doping” is used. 3.2 Unconventional high-temperature superconductivity 23 Figure 3.3: Crystal structure (adapted from Ref. [8]) and generic phase diagram of 122 iron pnictides. (a) Conducting Fe (beige) - pnictogen (green) planes are separated by spacer layers, consisting of another element such as Eu, Ba, Sr or Ca (blue). In the antiferromagnetic phase, the itinerant electrons of Fe order in a stripe-like configuration (brown arrows). (b) The an- tiferromagnetic (AFM) phase (green) is found in the parent compounds. At slightly higher temperatures, a nematic phase transition (grey) takes place, accompanied by a tetragonal-to- orthorhombic structural transition. With doping, mechanical pressure or isovalent substitution (“x”), those orders are suppressed and superconductivity sets in (SC: blue). Above the su- perconducting dome, strange metal behaviour (red) is observed, with a crossover to a Fermi liquid (yellow) in the overdoped range. This might indicate a quantum critical point (QCP: red) buried by the superconducting dome. unfolded1 Brillouin zone (see Fig. 3.4a). In general, the electronic system is more three- dimensional than in the case of cuprates, visible in Fermi sheets that slightly warp along the kz-direction and a less pronounced electronic anisotropy. The ground state of the parent phase is an antiferromagnetic metal with a magnetic moment per Fe atom of typically < 1µB. This value is much smaller than the possible 6µB for Fe2+. Indeed, it was found that the antiferromagnetic ordering vector is usu- ally (pi,0) or (0,pi), matching the nesting vector between electron and hole pockets (see Fig. 3.4a); the corresponding stripe-like spin order which is ferromagnetic in one direction and antiferromagnetic in the others is depicted in Fig 3.3a. This points (at least for the 122 iron pnictides) towards a rather itinerant magnetism with a spin density wave ground state. However, one should note that in some compounds, there are also indications for a more local character of magnetism; one example are 11 chalcogenides, where the magnetic order vector differs from the nesting vector. Therefore, magnetism in iron-based super- 1 The folded Brillouin zone is that of the actual lattice, considering Fe and pnictogen atoms; the unfolded Brillouin zone arises when only the Fe atoms are considered. While the Γ -point is the same in both, the X-point in the unfolded scheme corresponds to the M -point in the folded one. 24 3 Superconductivity Figure 3.4: Schematic Fermi surface as well as superconducting gap symmetry and structure of iron pnictides (adapted from Ref. [65, 71]). (a) The Fermi surface usually consists of multiple hole (dark blue) and electron pockets (dark green) centered around the (0,0)-“Γ”- and (pi,0)/(0,pi)-“X”-points of the unfolded Brillouin zone. Those pockets are nested by the antiferromagnetic ordering vectors (dark red). (b) A superconducting order parameter with s±-symmetry is isotropic and changes sign between electron and hole pockets. (c) For the extended s±-symmetry, the order parameter might contain accidental nodes which can be lifted by scattering [72]. conductors has probably a dual character with signatures of both, local and itinerant magnetism. Superconductivity While it is consensus that the Cooper pairs are singlets, the symmetry of the supercon- ducting order parameter of iron pnictides is still under debate. Mostly favoured is the so-called “s±-symmetry”, where s-wave gaps at the different Fermi surface sheets have op- posite sign (see Fig. 3.4b, [70, 71]). In the “extended s±-wave” scenario, those gaps may have accidental nodes (see Fig. 3.4c) which are not given by symmetry and can be therefore lifted by scattering [72]. This possibly explains the variety of contradictory measurements concerning nodes in the order parameter, as their occurrence would be strongly affected by scattering and thus the sample quality. Any analysis about gap sizes and symmetries are further hampered by the multiband character of the iron pnictides, which causes the opening of several superconducting gaps. Further details on this topic and a possible smoking gun experiment to prove the s±-symmetry with the help of optical spectroscopy are given in the attachment. Despite the gap symmetry, another interesting difference between cuprates and iron- based superconductors concerns the separation of superconductivity and long-range mag- netic order in the phase diagram. In iron pnictides, the maximum Tc appears usually close to the (extrapolated) end point of the antiferromagnetic phase transition. In some com- pounds, even a coexistence region of spin density wave and superconducting order emerges 3.2 Unconventional high-temperature superconductivity 25 (in contrast to cuprates, where the Mott insulating phase and the superconducting dome are separated by the pseudogap phase). The observation of competing order parameters suggests that this coexistence appears on the microscopic scale [73, 74]. Unconventional normal state properties Although the pseudogap phase governs the phase diagram of cuprates, it is either absent or at least much less pronounced in the case of iron pnictides (see chapter 4.4.5). Never- theless, the normal state above the superconducting dome is anything but conventional. There is growing experimental and theoretical evidence for a “nematic phase” transition that precedes or accompanies the antiferromagnetic ordering [6, 25, 75–80, 240]. The term “nematic” is borrowed from liquid crystal theory and describes the breaking of the C4 rota- tional symmetry. Interestingly, this does not only mean that a tetragonal-to-orthorhombic phase transition takes place: an electronic in-plane anisotropy sets in, whose magnitude is much larger than the orthorhombic lattice distortion. Furthermore, the b-axis resistivity is higher than that along the a-direction, although one would expect the opposite1. In other words, the itinerant electrons do not just follow the lattice anisotropy. Therefore, the transition is generally believed to be driven electronically either by spin or orbital fluctuations. Unfortunately, no matter which of those interactions is responsible, at the transition one will always observe an anisotropy of all - structural, orbital and spin de- grees of freedom - as they are strongly entangled. Thus, it is extremely difficult to decide what is actually driving the nematic order. However, exactly this question is of particular interest: the nematic phase in the iron pnictides phase diagram extends far above the superconducting dome and appears also in cuprates. Therefore, it might be the essential key ingredient for high-temperature superconductivity. The experimental investigation of nematicity is challenging. In order to probe the fluctuations above the transition, piezo elements can be used that induce some small amount of strain on the sample [25, 81]. Those experiments uncovered a huge nematic susceptibility that diverges at the transition temperature, independent of crystal type or quality [81]. Below the nematic transition, as the orthorhombic distortion is only small, the forma- tion of twin domains usually obscures the intrinsic anisotropy. In cuprates, mechanical clamps and magnetic fields were found to detwin single crystals [22, 23]. In the case of iron pnictides, the effect of typical laboratory magnetic fields on the Fe spins is rather weak: 14T detwin BaFe2As2 only by ∼ 10% [24]. Thus, mechanical clamps are commonly 1 A stronger orbital overlap along the shorter b-axis as well as the ferromagnetic spin alignment should both lead to a higher conductivity with respect to the a-axis. 26 3 Superconductivity used [6]. However, this introduces an explicit symmetry breaking by uniaxial pressure, which must be considered carefully [25]: similar to ferromagnets, where the magnetization depends on the external magnetic field, the intrinsic nematic response can only be mea- sured in the limit of zero external symmetry-breaking force. Indeed, mechanical clamps were found to significantly enhance the transition temperatures in iron pnictides and even induce additional anisotropy above [78]. In the framework of this thesis, it was uncovered that small magnetic fields are enough to persistently detwin EuFe2As2. This provides the unique possibility to study the low tem- perature electronic in-plane anisotropy of iron pnictides without applying any symmetry- breaking external force. The corresponding results are discussed in chapter 9 and Ref. [13]. Despite the nematic phase, another similarity can be found between cuprates and iron pnictides when looking at their normal state behaviour: a strange metal phase extends far above the superconducting dome, accompanied by linear resistivity. In BaFe2(As1−xPx)2 - which is usually considered as a very clean material as P substitution does not induce appreciable scattering - a diverging effective mass around the optimum substitution level was verified [83, 84]. Therefore, similar to cuprates the existence of a quantum critical point and its relation to unconventional superconductivity is discussed. 3.2.3 Magnetic fluctuations Due to the proximity of high-temperature superconductivity to a magnetic phase in both, cuprates and iron-based superconductors, it is intriguing to consider magnetic spin fluc- tuations as a possible pairing glue for Cooper pairs. However, those interactions lead to a repulsive pairing potential, opposite to the basic assumptions of BCS theory (see sec- tion 3.1). Further insight how such a potential can still support superconductivity can be found when looking at the BCS self-consistency gap equation for singlet pairing [85]: ∆(k) = − ∑ k′ V eff(k− k′) 2 √ ∆2(k) + 2(k) ∆(k′) (3.3) with the momenta k and k′, the energy gap∆, the effective two-particle pairing interaction V eff, as well as the dispersion of the electrons in the normal state (k). Obviously, an attractive pairing potential V eff < 0 (as assumed by Bardeen, Cooper and Schrieffer) might support superconductivity, as left and right sides of equation 3.3 can obtain the same sign. For V eff > 0, equation 3.3 can be only fulfilled for a strong momentum-dependent potential that leads to gaps with opposite sign for the momenta k and k′. The simplest solution for this problem is a d-wave gap, as indeed observed in cuprates. However, also the s±-symmetry favoured in iron-based superconductors fulfils those conditions. 3.3 Superconductivity and Magnetism 27 In order to get experimental information about the importance of spin fluctuations, inelastic neutron scattering has turned out to be a very powerful tool. It was found that a resonance peak in the superconducting spin excitation spectrum is a unique feature of unconventional superconductors [5, 65, 88]. This magnetic resonance appears at the same wave vector as the static magnetism in the parent compounds: at temperatures T < Tc, the antiferromagnetic neutron scattering signal gets depleted at lowest energies and a peak appears at energies smaller than 2∆. Theory predicts such a peak if the superconducting energy gap takes opposite values on different parts of the Fermi surface. Thus, it indicates unusual repulsive electron pairing. In iron pnictides, such a resonance was demonstrated for the first time for Ba0.6K0.4Fe2As2 [89]. One can further compare the spectrum of spin excitations with the Eliashberg spectral function α2F (ω), where α denotes the electron-boson coupling strength and F (ω) the frequency-dependent bosonic density of states [86, 87]. While the former can be measured with inelastic neutron scattering, the latter can be extracted from tunneling or infrared spectroscopy data. If spin fluctuations are responsible for Cooper pair formation, α2F (ω) should be proportional to the imaginary part of the spin susceptibility. Indeed, both - iron pnictides and cuprates superconductors - show that the resonance mode coincides with features in α2F (ω) [5], further supporting the idea of magnetic fluctuations as pairing glue. 3.3 Superconductivity and Magnetism As the electrons forming a Cooper pair contain within the BCS theory opposite spin and momentum, it is very clear that magnetism and superconductivity are conventionally antagonistic phenomena. Their interplay is in the focus of the following sections. 3.3.1 Type I and II superconductors A “type I” superconductor is a perfect diamagnet, meaning that an external magnetic field gets shielded completely by supercurrents in a thin surface layer. Importantly, this holds no matter whether the magnetic field is switched on below or above Tc, if the field does not exceed a critical value Bc2 where superconductivity breaks down. However, the detailed shielding mechanism depends on the cooling procedure. In the case of zero-field cooling, shielding currents are induced that counteract the application of the magnetic field; those currents do not decay, because the resistance is zero. As this is nothing else than Lenz’s law, the same behaviour would also occur in an ideal conductor. The difference between a superconductor and an ideal conductor lies in the field-cooled behaviour: at temperatures T > Tc, magnetic flux can penetrate inside both specimen; for 28 3 Superconductivity T < Tc, however, this flux is expelled only in the case of a superconductor. This additional effect that can not be explained by Lenz’s law was discovered by Meissner and Ochsenfeld. Therefore, the corresponding phase is called the “Meissner phase”. High-temperature superconductors belong to another class of superconductors, the “type II” superconductors, where an intermediate “Shubnikov phase” appears: above the critical field Bc1, the external field can penetrate inside the material within “Abrikosov vortices”. The Cooper pair density is zero at the core of such a vortex; it increases away from the core within the Ginzburg-Landau coherence length. Correspondingly, the magnetic field has a maximum at the core of the vortex; it decreases away from the core within the London penetration depth. The ground state of the system is a triangular vortex lattice, where each vortex contains a single flux quantum; higher magnetic fields lead to a higher density of vortices, until superconductivity breaks down at Bc2. 3.3.2 Magnetization of a superconductor Magnetization measurements are often used to characterize a superconductor, as a dia- magnetic signal indicates bulk superconductivity. However, the interpretation of those magnetization measurements can be very complex, as will be shown below. Temperature-dependent magnetization The first case that should be discussed are temperature-dependent magnetization mea- surements [90, 91]. In the strictest sense, only a susceptibility of -1 that was gained in a field-cooled measurement proves 100% bulk superconductivity. This can be understood intuitively when considering a type I superconducting cylinder which has an axial hole, where a magnetic field is applied parallel to the cylinder axis. For zero-field cooling, sur- face currents will shield everything “inside”, also the hole - the whole specimen acts as a perfect diamagnet. For field-cooling, however, the result is different: the magnetix flux that could penetrate inside the specimen in the normal conducting phase needs to be ex- pelled. This takes place only in the superconducting volume; the flux in the hole remains. Therefore, only the superconducting volume acts as a perfect diamagnet. In summary, the field-cooled susceptibility will be less negative than the zero-field cooled curve; only the field-cooled response will give the true superconducting volume fraction. Unfortunately, a less negative field-cooled signal does not mean necessarily a bad sam- ple quality or a smaller superconducting volume fraction. For instance, granular super- conductors have intergranular spaces that can trap flux. Another example are type II superconductors, where vortex dynamics can lead to very complex hysteretic behaviour, which will be discussed in more detail in the next section. 3.3 Superconductivity and Magnetism 29 Field-dependent magnetization Real type II superconductors contain imperfections. Indeed, the system can win energy when the vortices run through normal-conducting defects, as this effectively reduces the vortex length. Thus, the vortices are “pinned” to those defects. If this pinning is strong, the superconductor is characterized as “hard”, leading to interesting hysteresis effects due to flux trapping. The magnetization curves of such hard superconductors can be well described by Bean’s critical state model [92]. While at B < Bc1, the penetration of the magnetic field is given by the London penetration depth, at B > Bc1, vortices will start penetrating into the material. Bean assumed that those vortices get pinned when entering through the surface, leading below the surface to a broadened shielding volume with a homogeneous critical current density jc. For external fields much smaller than the upper critical field, jc is assumed as constant. Figure 3.5: Development of an internal magnetic field Bint inside a hollow superconducting cylinder (wall thickness d) when an external magnetic field Bext is applied (Bext > Bc1), following Bean’s critical state model. The field distribution is schematically depicted in (a) for increasing Bext and in (b) for decreasing Bext. (c) Bint dependent on Bext; fields displayed in (a) and (b) are marked by symbols with the same colour. A distinct hysteresis results, as vortices get directly pinned when entering the superconductor: with increasing field, the cylinder gets gradually “filled” with vortices from the outside to the inside; with decreasing field, opposite supercurrents cancel the vortices - again from the outside to the inside. Thus, vortices are trapped inside the walls for a certain range of Bext. 30 3 Superconductivity In order to understand the implications of those assumptions, a hollow superconducting cylinder with thin walls of thickness d and neglectable small Bc1 can be considered, where a magnetic field Bext is applied along the cylinder axis (see Fig. 3.5a). For small fields, Bext gets completely shielded within the cylinder walls and the magnetic field at the center of the cylinder Bint stays zero (see Fig. 3.5c). With increasing Bext, the volume where vortices penetrate the superconductor increases until the cylinder walls are completely filled. From this point, an increasing external field also leads to an increase of Bint. The hysteretic effects get obvious when the external field is decreased (see Fig. 3.5c): due to the decrease, an opposed critical current is induced at the surface, meaning that the density of vortices decreases where this current flows. When decreasing Bext further, Bint does not change until a homogeneous state is reached in the whole cylinder. At zero external field, a finite internal field is trapped inside the hollow cylinder. Indeed, the processes in a bulk sample are very similar [93], leading to characteristic strongly hysteretic magnetization curves that appear also in iron pnictides [94]. Consid- ering that jc should decrease when approaching the upper critical field, magnetization curves of hard superconductors can be well described by Bean’s critical state model [95]. 3.3.3 Upper critical field(s) As already mentioned, superconductivity gets destroyed at the upper critical field Bc2. In principle, there are two processes responsible for breaking up the Cooper pairs: spin and orbital pair-breaking [96]. The first case only happens in spin singulett supercon- ductors, where the Cooper pairs consist of electrons with opposite spin: As the magnetic field induces an energy splitting between those (Zeeman effect), superconductivity gets destroyed in the so-called “Pauli” or “Chandrasekhar-Clogston limit” [97, 98] as soon as this difference exceeds the condensation energy of the Cooper pairs. In the case of orbital or diamagnetic pair-breaking, the Cooper pairs are literally torn apart, as they possess opposite momentum and thus the Lorentz force in a magnetic field will act in opposite directions. While the upper critical field of most of the conventional superconductors is determined by the orbital pair-breaking (described by the Werthamer-Helfand-Hohenberg theory [99]), the Pauli limit plays a more important role in unconventional superconduc- tors. 3.3.4 Exotic phases Intuitively, one would not expect superconductivity in magnetic materials, as the magnetic exchange field between spins creates usually very high internal fields. Nonetheless, in some cases superconductivity is still energetically favourable, leading to exotic behaviour. 3.3 Superconductivity and Magnetism 31 Here, some of the effects that were discussed for Eu based 122 iron pnictides are shortly introduced. The first phenomenon that is observed in many of the Eu based iron pnictides is the “re-entrant superconductivity”: for temperatures T < Tc, superconductivity is first de- stroyed (Tn,re) and then re-enters at lower temperatures (Tc,re). This behaviour results from temperature-dependent magnetic interactions and was initially predicted in 1971 for materials that exhibit the Kondo effect [100]; in this case the pair-breaking first increases with decreasing temperature due to the Kondo effect, until it decreases again at the coher- ence temperature. Such a full cascade of transitions was experimentally verified in 1977 on La-Y-Ce alloys [101] with Tc = 0.55K, Tn,re = 0.27K and Tc,re = 0.05K. Figure 3.6: Typical tem- perature-dependent resistivity of a re-entrant superconduc- tor. Superconductivity that sets in at Tc gets suppressed at slightly lower temperatures Tn,re due to competing mag- netism. If the pair-breaking ef- fects get weaker with decreas- ing temperature, superconduc- tivity can eventually re-enter at Tc,re. Nowadays, several rare earth materials are known to show re-entrant superconductivity. One example are borocarbides [102], which have a similar crystal and magnetic structure as EuFe2As2. In Ho(Ni2Co2)B2C, three different magnetic phases exist below Tc. In the low temperature phase, spins within one layer are ferromagnetically coupled (with a small out-of-plane component) and neighbouring planes are coupled antiferromagnetically. The higher temperature phases are similar, but show slightly incommensurate spirals. It turns out that the incommensurable magnetism is harmful to superconductivity, in contrast to the commensurable phase. Theoretically, this behaviour is yet not well understood, but leads to the characteristic temperature-dependence of a re-entrant superconductor, where superconductivity is recovered when the commensurable magnetism sets in. The situation is more clear in superconducting cluster glasses such as (Gd,La)Ru2 and (Nd,Th)Ru2 [102], where the short-range ferromagnetic interactions compete with superconductivity, lead- ing to re-entrant superconductivity. In this case, the superconducting state is probably recovered, when larger ferromagnetic clusters become unstable at low temperatures. 32 3 Superconductivity While the phenomenon of re-entrant superconductivity directly shows that superconduc- tivity and magnetism compete, the question of how ferromagnetism and superconductivity can coexist microscopically is more challenging to reply and several theoretical predictions have been made. In 1964, Fulde and Ferrell as well as Larkin and Ovchinnikov considered independently from each other the effect of ferromagnetic moments on conventional singlet supercon- ductors [103–105]. They found a stable phase, where the Cooper pairs contain non-zero momentum, corresponding to the momentum difference of the Fermi surfaces of spin-up and down electrons. This leads to a spatial modulation of the superconducting order parameter, meaning that the Cooper pair density varies periodically in space. Although it was yet not possible to verify this “FFLO-state” in ferromagnetic superconductors, there are strong indications that such a spatially modulated state exists at high fields in unconventional superconductors [106]. Another approach to this topic is that conventional ferromagnetism could be even ham- pered to occur in a superconducting phase. In this spirit, it was suggested already in 1959, that in the superconducting phase only very small ferromagnetic domains will ap- pear, which cancel each other out on the scale of the superconducting coherence length in a “cryptoferromagnetic” alignment [107]. Furthermore, it was also predicted that in a ferromagnetic superconductor, spontaneous vortices can occur, lowering the system’s energy similar to the Shubnikov phase in conven- tional type II superconductors [108, 109]. At the moment, this effect is favoured to occur in Eu based iron pnictides [31], although a direct experimental proof is still missing. Finally, one should note that an external magnetic field can in principle also enhance superconductivity if it acts against the internal magnetism. In rare earth materials such as the Chevral-phase compound EuxSn1-xMo6S8 [111, 112], the appearance of a second super- conducting dome at high magnetic fields was observed, while the original superconducting dome was already suppressed at lower fields. This effect was theoretically explained in 1962 [110]: in those superconductors, the exchange field between conduction electrons and local magnetic moments can be positive or negative, i.e. parallel or antiparallel to the local moment direction. In the latter case, the exchange field is opposed to an external magnetic field. Thus, for a certain field range, the external and internal magnetic fields almost cancel each other out, significantly reducing the conduction electron polarization. This leads to a re-appearance of superconductivity. 4 ELECTRODYNAMIC PROPERTIES Optical spectroscopy is a powerful method to gather information about the electrodynamic properties of materials, as it is able to resolve the energy scales of electronic interactions. In this chapter, basic definitions and concepts to describe metals and broken symmetry states are introduced, following Ref. [113] and [114]. Those concepts are then employed to review the electrodynamic properties of 122 iron pnictides [115]. 4.1 Basic definitions In analogy to the relation of magnetic flux density and field strength discussed in chap- ter 2.1, an electric displacement field D can be defined in free space as: D = 0E (4.1) with the permeability of free space 0 and the electric field E. In a linear and isotropic dielectric material, an external electric field creates a polariza- tion P and equation 4.1 needs to be extended: D = 0E + P = 0rE (4.2) 34 4 Electrodynamic properties with the relative dielectric constant (i.e. permittivity) r, which can be related to the refractive index n˜ of the medium via: n˜ = √r. (4.3) In a dispersive medium, n˜ and r are both complex frequency-dependent quantities: n˜(ω) = n(ω) + ik(ω) (4.4) r(ω) = 1(ω) + i2(ω). (4.5) Despite {n(ω), k(ω)} and {1(ω), 2(ω)}, there exist also two other pairs of optical constants that are commonly used for describing the electrodynamic properties of solids. The first is the frequency-dependent conductivity: σ(ω) = σ1(ω) + iσ2(ω) (4.6) that relates to the relative dielectric constant as: r(ω) = ∞ + i σ(ω) 0ω (4.7) with ∞ being the relative permittivity at ω →∞. A common way to excess those optical constants are reflection measurements, where the reflected wave displays a phase shift θ(ω) and an amplitude change r(ω) with respect to the incident wave. In an experiment, the measured quantity is usually the reflectivity: R(ω) = |r(ω)|2 = (n− 1) 2 + k2 (n+ 1)2 + k2 . (4.8) Unfortunately, the phase shift θ = arctan−1 [ 2k n2 + k2 − 1 ] (4.9) is accessible only by some measurement techniques such as ellipsometry or time domain spectroscopy. However, as for all optical constants the causality condition holds, the Kramers-Kronig transformation can be used to calculate the phase shift: θ(ω) = −ω pi P ∫ ∞ 0 lnR(x) x2 − ω2dx (4.10) 4.2 Metals 35 where P denotes the Cauchy principle value. As usually the reflectivity is measured only over a limited frequency range, R(ω) needs to be extrapolated properly to low and high frequencies. Once this is done, one has obtained the last pair of optical constants from which all the others: {n(ω), k(ω)}, {1(ω), 2(ω)} or {σ1(ω), σ2(ω)} can be calculated. Indeed, Kramers-Kronig transformations connect all real and imaginary parts of optical constants, and can be utilized to derive “sum rules”, which are powerful tools in the analysis of spectroscopic measurements. It should be noted that within this thesis, uniformly the SI system is used. However, in optical spectroscopy often the CGS notation or even a mixture of unit systems can be found. Most importantly, frequency and scattering rate are commonly given in [cm−1]. By using the conversion: [cm−1]CGS ∗2pic−−−→ [1/s]SI (4.11) the corresponding SI unit [1/s] can be obtained. 4.2 Metals The electrodynamic properties of metals can be well understood considering the phe- nomenological Drude and Lorentz model. While the former describes intraband interac- tions, the latter accounts for interband transitions. In order to model the electrodynamics in a real material, contributions from both models are included. While the reflectivity is not additive, the conductivity is (according to Matthiessen’s rule). Drude model In the Drude model, a metal is described as a classical gas of electrons. Such an electron interacts with its environment only in the form of collisions; the average time between subsequent collisions is given by the relaxation time τ or the scattering rate Γ = 1/τ . This leads in the presence of a time-dependent external field E(t) to the equation of motion: m dr2 dt2 + m τ dr dt = qE(t) (4.12) with mass m and charge q of the free carriers. By solving this equation, one obtains for the real and imaginary part of the frequency-dependent conductivity: σ1(ω) = σdc 1 1 + ω2τ2 = 0ω 2 pτ 1 1 + ω2τ2 (4.13) σ2(ω) = σdc ωτ 1 + ω2τ2 = 0ω 2 pτ ωτ 1 + ω2τ2 (4.14) 36 4 Electrodynamic properties with the frequency-independent prefactor given either by the dc conductivity σdc = Nq2τ m (4.15) or the plasma frequency ωp = √ Nq2 0m (4.16) that depend on the density N of the free carriers. The Sommerfeld extension of the Drude model accounts for the existence of a Fermi surface, leading to a more realistic description with an effective massm∗ in equations 4.12 - 4.15 and a mean free path l given by the Fermi velocity vF: l = vFτ. (4.17) In Fig. 4.1a, the reflectivity R(ω) of a typical Drude response is displayed. For ω  1/τ , the electrodynamic properties of a metal depend mainly on the dc conductivity and the reflectivity can be well described by the Hagen-Rubens relation: R(ω) ≈ 1− 2 √ 20ω σdc , (4.18) as can be seen in the inset of Fig. 4.1a. With increasing frequencies, the reflectivity stays close to unity, until it drops rapidly at ωp and the metal gets transparent. Fig. 4.1b displays the corresponding complex conductivity. σ1(ω) is almost flat at low frequencies with σ1 ≈ σdc and decreases at higher frequencies rather rapidly to zero, with the inflection point of the curve at the scattering rate. In contrast, σ2(ω) vanishes at low frequencies and has a maximum at 1/τ . One should note that the area below the real part of the conductivity is a measure of the carrier density and effective mass. This can be directly seen when integrating: ∫ ∞ 0 σ1(ω)dω = 0pi 2 ω 2 p = Nq2pi 2m (4.19) Indeed, this so-called “f -sum rule” is valid independent from the Drude model. Extended Drude model In the Drude model, the scattering rate is assumed to be frequency-independent. However, this assumption is invalid for many correlated materials, where the carriers couple to a 4.2 Metals 37 Figure 4.1: Typical (a) reflectivity and (b) conductivity of a metal according to the Drude model. (a) R(ω) (brown line) is close to unity until it drops rapidly at ωp. The inset shows that at low frequencies, the reflectivity can be well described by the Hagen-Rubens equation (red dotted line). (b) σ1(ω) (green line) is close to σdc (red dotted horizontal line) and decreases at higher frequencies with the inflection point of the curve at 1/τ , where σ2(ω) (blue dashed line) displays a maximum. bosonic mode. Therefore, an extended Drude model was introduced [116], where 1/τ got replaced by a complex damping function: M(ω) = 1 τ(ω) − iωλ(ω) (4.20) with the frequency-dependent scattering rate 1/τ(ω) and the optical mass enhancement factor λ(ω). Those parameters can be derived from the optical conductivity: 1 τ(ω) = 0ω 2 p 1 σ1(ω) (4.21) 1 + λ(ω) = m ∗ mb = −0ω2p 1 ω 1 σ2(ω) (4.22) with the band mass mb. This is of particular interest for unconventional superconductors, as from 1/τ(ω) the Eliashberg spectral function α2F (ω) (introduced in chapter 3.2.3) and the coupling strength to the bosonic mode λ (also called “mass renormalization factor”) can be calculated. 38 4 Electrodynamic properties Lorentz model In many cases, resonant interaction of light with matter can be phenomenologically de- scribed by the Lorentz model of damped harmonic oscillators. Typical examples are ionic or phononic vibrations, as well as interband transitions. It follows for the real and imagi- nary part of the frequency-dependent conductivity: σ1(ω) = 0ω2p ω2/τ (ω20 − ω2)2 + ω2/τ2 (4.23) σ2(ω) = −0ω2p ω(ω20 − ω2) (ω20 − ω2)2 + ω2/τ2 (4.24) with the same definition of the plasma frequency ωp as in the Drude model; the damping 1/τ accounts for the energy loss of the oscillating dipoles due to collisions. In Fig. 4.2a, the reflectivity R(ω) of a Lorentz oscillator is displayed. While the reflec- tivity is rather flat at low frequencies, it increases strongly around ω0 due to the onset of interband transitions. For higher frequencies, the reflectivity stays again rather constant, until it drops similar to the Drude model at ωp. Figure 4.2: Typical (a) reflectivity and (b) conductivity of a Lorentz oscillator. While R(ω) (brown line) is rather flat at low frequencies, it increases around ω0, flattens again, and drops finally similar to the Drude model at ωp. While σ1(ω) (green line) displays a maximum at ω0 with the FWHM = 1/τ , σ2(ω) (blue dashed line) has an inflection point at ω0 and extrema at ω0 ± 1/2τ . 4.3 Broken symmetry states in metals 39 The corresponding conductivity can be seen in Fig. 4.2b. As the interband transitions lead to strong absorption, σ1(ω) displays a maximum at ω0 with the full width at half maximum (FWHM) given by 1/τ ; σ2(ω) is negative below ω0 and positive at higher frequencies, with extrema at ω0 ± 1/2τ . 4.3 Broken symmetry states in metals Electron-electron and electron-boson interactions might trigger in a metal a phase transi- tion to an ordered, coherent ground state with broken symmetry. Examples are supercon- ductivity and density wave states. While in the former case, gauge symmetry is broken via the Higg’s mechanism, in the latter case, translational symmetry is broken due to a periodic variation of spin or charge. Further differences can be found when considering the pairing conditions: while for conventional superconductors1, electrons with opposite momentum pair, in density waves electrons and holes pair with a total momentum of Q = 2kF. Therefore, the latter process requires a substantial nesting of the Fermi surface, meaning that parts of the Fermi surface can be mapped on top of each other by translation with Q. 4.3.1 General quantum mechanical description In the metallic phase, the interaction of light with matter can be described sufficiently in the framework of the classical Drude and Lorentz model, as discussed in section 4.2. For superconductors and density waves, however, a quantum mechanical description is necessary. According to Fermi’s golden rule, the net transition rate from an initial (occupied) to a final (unoccupied) state (with energies E and E + ~ω, respectively) depends on: • the energy of the external probe ~ω • the number of occupied quasiparticle states, given by f(E)D(E) (with the energy- dependent Fermi distribution f(E) and the density of states D(E)) • the number of unoccupied states, given by [1− f(E + ~ω)]D(E + ~ω) • the transition matrix elementM 1 This section will focus on conventional superconductors with an isotropic energy gap. 40 4 Electrodynamic properties and can be calculated in general as: W ∝ ∫ |M|2D(E)D(E + ~ω) [f(E)− f(E + ~ω)]dE. (4.25) For broken symmetry states, one needs to consider some peculiarities about the density of states and the transition matrix element that will be discussed below. Single particle density of states In the broken symmetry phase, a temperature-dependent energy gap appears in the single particle excitation spectrum, as already discussed for a (conventional) superconductor in chapter 3.1; the gap equation is the same for a density wave. The corresponding single particle density of states is depicted in Fig. 4.3. Above the critical temperature where the symmetry is broken (T > Tc), the density of states is basically flat and the occupation of states follows the Fermi distribution. In the broken symmetry phase, states within the energy gap can not be occupied. As the total number of states needs to be conserved, the gapped states are reshuffled, leading to a divergent density of states at EF ±∆. Figure 4.3: Single particle density of states for a metal that exhibits at Tc a transition to a broken symmetry phase for (a) T > Tc, (b) 0 < T < Tc, and (c) T = 0; dark areas mark occupied states, light colours unoccupied ones. (a) For T > Tc, the density of states follows the Fermi distribution. (b,c) Below Tc, a temperature-dependent gap develops around EF, leading to a divergence of the density of states at EF ± ∆. The absorption of photons with ~ω ≥ 2∆(T ) can break Cooper pairs (black solid arrow). (b) Only for T > 0, thermally activated electrons are available to absorb photons (dashed black arrow). The formation of an energy gap has important implications for the absorption of pho- tons: while Cooper pairs can be broken only for energies ~ω ≥ 2∆(T ), thermally excited 4.3 Broken symmetry states in metals 41 electrons can also absorb photons with lower energies. This leads to two independent absorption channels, whose properties will be discussed in the next sections. Coherence effects While transitions between different states are independent for a metal, the situation is more complicated in the case of a broken symmetry phase where carriers are condensed into a coherent ground state [40, 61]. As a quasiparticle of such a ground state consists of two carriers with well-defined momentum- and spin-relation, it can be shown that one will always find pairs of transitions that involve the same excitation states, although the scattering process is different (for a typical example, see table 4.1). Process State (k1, ↑) (−k1, ↓) (k2, ↑) (−k2, ↓) 1 initial 1 0 0 0 1 final 0 0 1 0 2 initial 1 0 1 1 2 final 1 1 1 0 Table 4.1: Two processes in a superconductor that lead to a scattering of an electronic excitation from the initial state (k1, ↑) to the final state (k2, ↑). States that are occupied by an electron are noted as “1”, unoccupied ones by “0”. Pairs of opposite spin and momentum that form a Cooper pair are marked in blue; if in such a pair only one electronic state is occupied, this is called electronic excitation (red). While the first process involves a scattering of an electron from (k1, ↑) to (k2, ↑) (marked in bold signs), the second process involves a scattering from (−k2, ↓) to (−k1, ↓). However, the initial and final states concerning the electronic excitation are the same. This needs to be considered when calculating the transition matrix elements, as equiva- lent matrix elements have to be added coherently before taking the squared absolute value in equation 4.25. It turns out that a proper mathematical description can be achieved via so-called “coherence factors” F , transforming the transition matrix element as: |M|2 → F|M|2 (4.26) in the broken symmetry state. The absolute value of F depends on the type of perturbation and varies between zero and one. Fig. 4.4a displays the frequency-dependence (at T = 0K) of the electromagnetic absorp- tion for a superconductor and a spin density wave system, calculated from equation 4.25 with considering the appropriate coherence factors. While for the latter case, the singu- larity in the density of states is directly visible in the absorption spectrum, for a supercon- ductor, the coherence factor approaches zero at the gap energy and cancels the singularity 42 4 Electrodynamic properties in the density of states. This leads to a smooth increase of the absorption rate for ener- gies above 2∆. In Fig. 4.4b, also the temperature-dependence of the absorption rate is depicted for electromagnetic irradiation with small energies ~ω  2∆. In the case of a superconductor, one can observe just below Tc a surprising enhancement of the absorption rate. As this enhancement rises again from the coherence factor, this feature is usually referred to as “coherence peak”. Figure 4.4: (a) Frequency- and (b) temperature-dependent absorption of photons for a superconductor (blue) and a spin density wave material (red) at T = 0 and ~ω  2∆, respectively, as resulting from Fermi’s golden rule (adapted from Ref. [113]). (a) In both cases, an energy gap appears below 2∆. While for the spin density wave, the singularity in the density of states is directly visible in the absorption spectrum, for a superconductor, coherence effects lead to a smooth increase of the absorption. (b) A coherence peak slightly below Tc is only visible for a superconductor. 4.3.2 Superconductors In the so-called “two-fluid” model, the electrodynamic properties of a superconductor are approximated as a superposition of the response of superconducting and normal- conducting electrons, neglecting any superconducting gap and coherence effects. For the condensed electrons, such a classical descriptions transforms the Drude conductivity (see equation 4.13) in the limit τ →∞ to: σ1(ω) = pi 2 Nq2 m δ(ω = 0) = pi2 0 λ2L δ(ω = 0) (4.27) σ2(ω) = Nq2 m 1 ω = 0 λ2L 1 ω (4.28) 4.3 Broken symmetry states in metals 43 with a delta-function in σ1(ω) at zero frequency δ(ω = 0), the London penetration depth: λL = √ m µ0Nq2 = c ωp,c (4.29) and charge and mass of a single electron, if N is defined as the density of superconduct- ing electrons. When all electrons are condensed in the superconducting ground state, the plasma frequency in the normal state ωp corresponds to that of the superconduct- ing condensate ωp,c. One should note that despite λL, sometimes also a more general, frequency-dependent penetration depth is defined: λ = √ 0 σ2(ω)ω . (4.30) The 1/ω-divergence in σ2(ω) leads to a rapid, characteristic increase of the reflectivity towards unity. FGT sum rule As the f -sum rule must hold also when passing from the normal to the superconducting state, it follows that the missing spectral weight that appears in the superconducting phase due to the opening of the gap is transferred to the delta-function at zero frequency. This directly leads to the “FGT sum rule” (named after Ferrell, Glover and Tinkham [117–119]), relating the conductivity in the normal state σn1 (ω), the conductivity in the superconducting state σs1(ω), and the London penetration depth by:∫ ∞ 0 [σn1 − σs1]dω = 0pi 2 c2 λ2L . (4.31) Dirty vs. clean limit For a more detailed description of the electrodynamics, one has to consider the different characteristic length scales; their relative magnitude determines the nature of the super- conducting state. In particular, one has to distinguish between the dirty (local) and the clean limit, 1/τ > ∆ and 1/τ < ∆, respectively. Thus, in the clean limit most of the spectral weight of the Drude component lies below 2∆ and for T < Tc, almost all free carriers collapse into the condensate. In contrast, in the dirty limit a significant amount of Drude weight lies above 2∆. Therefore, only a fraction of the Drude spectral weight collapses into the condensate, leading to a more pronounced gap feature in the optical spectrum. 44 4 Electrodynamic properties Mattis-Bardeen description Based on the BCS theory, the electrodynamic properties of a superconductor were first analytically calculated by Mattis and Bardeen [120], assuming τ → ∞, and extended for finite scattering rates by Zimmermann et al. [121]. Fig. 4.5 displays the real and the imaginary part of the conductivity for a superconductor in the dirty limit at T = 0.8 Tc, decomposed into contributions from Cooper pairs (broken by absorbed phonons) and thermally activated electrons; Fig. 4.6 adds the temperature-dependence of σ1. The general features that were discussed up to now, i.e. the non-diverging behaviour of σ1(ω ≤ 2∆), the coherence peak in σ1(T ) (for ω  2∆) and the 1/ω-dependence of σ2(ω) are all nicely visible. Figure 4.5: Frequency-dependence of the (a) real and (b) imaginary part of σ(ω) (black line) for a superconductor in the dirty limit at T/Tc = 0.8, decomposed into contributions from Cooper pairs (blue dashed line) and thermally activated electrons (red dashed line) (adapted from Ref. [122]). (a) The photon activated electrons contribute only for energies ~ω > 2∆(T ) to σ1(ω); coherence effects cancel the singularity in the density of states, leading to a continuous increase of their response. The contribution of thermal electrons decreases with increasing frequency, following the behaviour of the density of states. The sum of both contributions displays a characteristic minimum at 2∆(T ). (b) The 1/ω-behaviour of the phonon activated electrons dominates the response of σ2(ω); the contribution of the thermally activated electrons is opposite, but negligible. At 2∆(T ), only a very weak kink is visible. 4.4 Iron-based superconductors Optical spectroscopy revealed valuable information about correlations [123] and broken symmetry states [124–127, 266] already within the first year after the discovery of iron- based superconductors. Most of the relevant charge dynamics take place in the infrared 4.4 Iron-based superconductors 45 Figure 4.6: (a) Frequency- and (b) temperature-dependence of σ1(ω) for a superconductor in the dirty limit for several temperatures and photon energies, respectively. (a) The pronounced gap feature vanishes with increasing temperature. However, the response of thermally acti- vated electrons at very low energies gets strongly enhanced close to Tc. (b) The latter is consistent with a pronounced coherence peak at energies ~ω  2∆. frequency range, as summarized in Fig. 4.7. This section provides an overview of great achievements and open questions provoked by optical investigations on 122 iron pnictides. Figure 4.7: Characteristic energy scales in iron-based superconductors. In order to study the superconducting gap, terahertz (THz) and far-far-infrared (FFIR) frequencies have to be covered. While itinerant dynamics are relevant in the far-infrared (FIR) energy range, incoherent transport and interband transitions govern the mid- and near-infrared (MIR and NIR, respectively). Ellipsometry measurements can be extended to very high energies. 46 4 Electrodynamic properties 4.4.1 Spectral features of the normal state Similar to cuprates, the normal state in-plane reflectivity and optical conductivity of 122 iron pnictides do not show well-separated components, leading despite a zero-energy Drude peak to a broad mid-infrared plateau that is rather untypical for a metal and often referred to as “incoherent” region; no clear plasma edge is visible (see Fig. 4.8b, [128, 129, 266]). At higher frequencies, humps characteristic for interband transitions appear. Figure 4.8: Characteristic (a) reflectivity and (b) conductivity of 122 iron pnictides in the normal state (adapted from Ref. [266]). (a) R(ω) (brown line) typically displays in the FIR an overdamped metallic frequency-dependence; at higher frequencies, no clear plasma edge is visible - only humps typical for interband transitions. (b) Correspondingly, σ1(ω) is slowly decreasing at low frequencies with a long tail extending in the MIR region; it is followed by a pronounced peak structure around 5000 cm−1. A cut-off separating itinerant Drude-like behaviour (red area) from local and interband (Lorentz-like) dynamics (blue area) is often set around 1500 cm−1. However, in reality a significant overlap exists. With decreasing temperature, the response gets more metallic, meaning that spectral weight is shifted within the mid- and far-infrared region to lower frequencies. However, spectral weight shifts occur also at higher frequencies up to the visible spectral range, indi- cating that correlation effects are of importance [130], as discussed further in section 4.4.5. 4.4 Iron-based superconductors 47 Along the c-direction, the flat MIR background dominates, which decreases with de- creasing temperature1 [131]. However, lowering the temperature also uncovers a weak Drude component, indicating interlayer coherence. Thus, iron pnictides display much less electronic anisotropy than cuprates. Nevertheless, both families of high-temperature su- perconductors share the broad plateau-like background in the mid-infrared, that is not a simple Drude response. 4.4.2 Analysis methods In the case of cuprates, two principal approaches exist that are commonly used to describe the mid-infrared charge dynamics: one-component and multi-component models [116]. In the former case, it is assumed that the low-frequency response originates from itinerant carriers with a frequency-dependent scattering rate and mass. The formalism of this extended Drude model was already introduced in section 4.2. In the multi-component approach, the conductivity spectrum is described by a free-carrier Drude term and a set of Lorentzian oscillators, accounting for example for interband transitions or impurity states. The debate which model describes better the charge dynamics in cuprates is yet not solved. On the one hand, well-separated absorptions in lightly doped cuprates favour the multi-component approach. On the other hand, the extended Drude formalism could uncover a frequency-dependent scattering rate over a broad energy range and many inter- esting aspects like the violation of Fermi liquid theory or the coupling to a temperature- dependent bosonic mode, both consistent with other measurement techniques. In the case of iron pnictides, the multiband character further complicates both, single- and multi-component analysis. In principle, hole and electron bands with different scat- tering rates should be considered. Furthermore, as five Fe d-bands are situated close to the Fermi energy, low-lying interband transitions are plausible. Indeed, recent infrared studies uncovered for Ba, Sr and Ca compounds two well-defined features around 1000 and 2300 cm−1 [12]. Due to its temperature- and doping-dependence, the former one can be attributed (via a comparison to ARPES measurements) to interband transitions in- volving Fe hole pockets. The origin of the second feature is yet not clear; the same holds for another hump in the optical conductivity, that was observed in several compounds at even lower energies (around 150 cm−1) [132–137]. Despite that, Dirac physics might also play a certain role [139, 140]. 1 Moon et al. demonstrate that a true c-direction response can be only obtained for cleaved samples, not for polished specimen [131]. 48 4 Electrodynamic properties Disentangling all those processes in a multi-component analysis is almost impossi- ble. Therefore, a “broad Drude” approach was suggested by Wu et al.: they modelled the mostly temperature-independent background of eight different 122 compounds by a very broad Drude and two high-frequency Lorentzians that account for interband tran- sitions [128, 130]. The remaining temperature-dependent part of the spectrum can be described by one narrow Drude. There is no spectral weight transfer between the different subsystems. This model has the advantage that the coherent dynamics are described by only a few parameters. Nakajima et al. extended this approach, allowing the broad Drude component to vary with temperature [129]. However, fitting the spectra was still possible without spectral weight transfer between the different subsystems. In further publications, the broad back- ground narrowed and was more and more attributed to the hole bands [141]. There exist also advocates of the conventional multi-component approach known from cuprates, where one or a series of Lorentzians is used to model the mid-infrared plateau. They argue that the mean free path of a very broad Drude would violate the Mott-Ioffe- Regel limit, indicating that the conductivity in such a band is no longer metallic and therefore better described by a bound excitation [142]. It turns out that in this case, enough free parameters exist in order to fit the spectrum with only one Drude. In order to account for electron and hole bands, one needs additional information, such as the relative spectral weights of the Drude components which can be extracted from photoemission experiments [143, 144]. The ambiguity of different modelling approaches is a serious drawback of optical spec- troscopy on iron pnictides. Maybe one of the biggest problems is that usually one group only publishes results using a single modelling approach. Thus, it gets not clear whether (on the first view) contradictory studies result from different fitting approaches, sample quality or have a physical meaning. As a direct comparison of different modelling ap- proaches is still lacking, such a study was performed in the framework of this thesis and is presented in chapter 9. 4.4.3 Antiferromagnetism The early observation of energy gaps opening at the high-temperature phase transition in iron pnictide parent compounds is one of the great achievements of optical spectroscopy: already in 2008, Dong et al. proposed on the basis of optical measurements on polycrys- talline LaOFeAs a spin density wave, driven by a nesting instability between electron and hole Fermi surfaces [124]. Only six months later, Hu et al. reported more detailed studies on single crystals of Ba- and SrFe2As2 [266]. 4.4 Iron-based superconductors 49 At the spin density wave transition, the reflectivity develops a rather sharp edge in the far-infrared, above which R(ω) gets substantially suppressed (see Fig. 4.9a). This is a typi- cal signature of the formation of an energy gap. However, with decreasing temperature the low-frequency reflectivity still increases, indicating that the Fermi surface is only partially gapped. The remaining Drude response is dramatically narrowed, with the scattering rate being decreased by ∼ 90%. Figure 4.9: Characteristic (a) reflectivity R(ω) and (b) conductivity σ1(ω) of 122 iron pnic- tides in the spin density wave state (solid curves) compared to the normal state (grey light area) (adapted from Ref. [266]). The Fermi surface gets only partially gapped, visible in an increase of R(ω) and σ1(ω) at low frequencies (red area). Spectral weight is shifted to higher frequencies; σ1(ω) displays two bump-like features corresponding to two different gaps (purple area). An additional spectral weight shift appears even in the range of interband transitions (blue area), indicating correlations. The sharp peak around 260 cm−1 is the signature of an FeAs phonon [145]. In the real part of the optical conductivity, the spin density wave becomes noticeable as spectral weight is reshuffled from the far-infrared to higher frequencies, where it piles up usually in the form of two well-separated peaks, corresponding to two different energy gaps (see Fig. 4.9b). The peak position marks the gap energy; it is around 2∆/kBTSDW ∼ 3.5 for the smaller and ∼ 10 for the bigger gap. As only spectral weight and peak position of the lower gap follow the mean field temperature-dependence, it was suggested in accordance with band structure calculations that this one has itinerant character, while the higher gap 50 4 Electrodynamic properties appears due to local exchange splitting [146]. A different interpretation results from c-axis measurements, where exclusively the smaller gap is visible [147]. From this, the authors proposed that the larger gap is due to the nesting of 2D Fermi surfaces and the smaller gap results from the backfolding of the magnetic Brillouin zone that cuts the 3D Fermi sheet. Finally, also a magnon-assisted indirect spin density wave excitation was discussed that could lead to the high energy gap [139]. 4.4.4 Superconductivity Optical signatures of a clear superconducting gap were first seen in 1111 LaFeAsO0.9F0.1−δ and NdFeAsO0.82F0.18 polycrystalline samples [125, 126]. A more quantitative analysis could be only achieved after 122 single crystals were available [127]. Since then, most optical studies were carried out on (nearly) optimally Co doped Ba 122 compounds due to the availability of high quality single crystals [128–130, 133, 134, 142, 154, 155] and thin films [148–153]. Dirty vs. clean limit It turned out that the in-plane optical response of Ba1−xKxFe2As2 and Ba(Fe1−xCox)2As2 is quite different in the superconducting regime: while the former displays over a broad frequency range a quasilinear dependence of σ1(ω), the latter shows a steep onset of absorption at energies corresponding to (one of) the smaller superconducting gap(s) [135]. Charnukha et al. demonstrated that both behaviours can be explained by using an effective two-band Eliashberg theory with a strong coupling to spin fluctuations, if K substitution leads to effective bands in the clean limit, while Co doped compounds are in the dirty limit. Superconducting gaps The same calculations yield for Ba0.68K0.32Fe2As2 isotropic superconducting gaps with 2∆/kBTc ∼ 2.2 and ∼ 6, in quantitative agreement with other measurement techniques such as ARPES, STM and specific heat. The situation is much more complicated in the case of Ba(Fe1−xCox)2As2, where several reports exist which show neither consensus about the gap sizes, nor about the number of gaps (see Fig. 4.10a). While most reports agree on one directly visible gap at around the BCS mean field value, fitting approaches usually need at least one more gap at significantly higher or lower energies for a satisfactory description. Similar to cuprates [116], an unexpected high intra-gap absorption was evidenced in the superconducting state of Ba(Fe1−xCox)2As2, which cannot be accounted for by thermally 4.4 Iron-based superconductors 51 Figure 4.10: (a) Extracted superconducting gap sizes [128–130, 133, 134, 142–144, 148–155] and (b) typical temperature-dependent conductivity in the THz frequency range for nearly optimally doped Ba(Fe1−xCox)2As2 (adapted from Ref. [151]). (a) While most reports agree on one directly visible gap at around the BCS mean field value (green dashed line / green area), fitting approaches usually need at least one more gap at significantly higher or lower energies for a satisfactory description (light blue areas). Black squares denote measurements on single crystals, red points on thin films. (b) A coherence peak is commonly observed below Tc at very low frequencies. However, the conductivity does not approach zero at lowest temperatures, indicating an anisotropic gap or pair-breaking interband scattering. broken pairs [134, 148, 150, 151]. It appears in films as well as single crystals. As it can be fitted with a Drude peak (with an overall conserved spectral weight), a low-lying impurity band is unlikely the reason [134]. Other explanations are an anisotropic gap [151] or pair-breaking interband scattering1 [150, 156]. An important difference between K and Co doping is that in the former case, atoms out of the FeAs planes are substituted, while in the latter case, the FeAs planes are directly affected [156]. Therefore, stronger impurity scattering is expected for Co doping, consistent with smaller residual resistivity ratios and superconductivity in the dirty limit. It needs to be seen, whether this enhanced scattering is also responsible for the strong intra-gap absorption. Proving this would be another very strong indication for the s±- symmetry of the superconducting order parameter. Measurements on Ba(Fe1−xCox)2As2 1 In an s± superconductor, interband scattering by non-magnetic impurities is pair-breaking, as it mixes hole and electron states that have superconducting order parameters of opposite phase [156]. 52 4 Electrodynamic properties thin films, where the impurity scattering gets step-wise enhanced by proton irradiation, were performed and supervised within this thesis and are presented in the attachment. Coherence peak Optical spectroscopy on Ba(Fe1−xCox)2As2 thin films was able to resolve a well-pronounced coherence peak in the THz frequency range (see Fig. 4.10, [148, 150, 151]). In general, a coherence peak can arise when the portions of Fermi surface that are coupled by an experimental probe have gaps of the same sign and magnitude [150]. As the momentum transfer of light at THz frequencies is negligible compared to the reciprocal lattice momen- tum, only parts of the same Fermi sheets are coupled. Therefore, the observed coherence peak is consistent with an s±-symmetry of the superconducting order parameter. Further- more, the position of the coherence peak can be used in multiband superconductors as a quantitative measure of the lower superconducting gap, demonstrated by Dahm et al. for MgB2 [157]. FGT sum rule All infrared studies agree on a rapid recovery of the missing spectral weight in the FIR frequency range, yielding in-plane London penetration depths (in accordance with the extraction from σ2(ω)) between 200-400 nm. However, one very careful ellipsometry study performed by Charnukha et al. on Ba0.68K0.32Fe2As2 uncovered a superconductivity- induced suppression of an interband transition at energies well above the plasma edge [158]. Although the transferred spectral weight is only ∼ 0.5% and therefore too small to be detected in infrared studies, it could have a big effect on reducing the kinetic energy of the system, pointing towards an unconventional pairing mechanism. Such an effect was already proposed for cuprates, however, is there still under debate [116]. Bosonic spectrum As the extended Drude analysis accounts only for itinerant carriers, the contribution of all interband transitions needs to be eliminated before any further data processing. Moreover, one has to keep in mind that due to the multiband nature of iron pnictides, only an average spectrum for bosonic excitations can be extracted. Up to now, an extended Drude analysis was carried out mainly on K and Co doped iron pnictides [130, 135, 142, 159–161]. In all cases, a frequency-dependent scattering rate is observed at low frequencies, which increases linearly with frequency (at least) up to ∼ 50meV. A further Eliashberg analysis reveals a strongly temperature-dependent bosonic spectrum, which is inconsistent with phonon coupling. The mass renormalization 4.4 Iron-based superconductors 53 factor increases with decreasing temperature, reaching values between 3-41, much above the electron-phonon coupling constant λ = 0.21. Both, the temperature-dependence of the bosonic spectrum as well as the high coupling constant would be consistent with spin fluctuations playing a key role for bosonic excitations in pnictides. This situation is very similar to that in cuprates [116]. c-axis coherence Measurements on cleaved Ba(Fe0.926Co0.074)2As2 crystals reveal not only interlayer coher- ence in the normal state, but also a robust interlayer superconducting condensate with a penetration depth of ∼ 1000nm [131]. This is in stark contrast to underdoped cuprates, where normal state transport is incoherent along the c-axis, leading to Josephson coupling of CuO2 layers in the superconducting phase [116]. 4.4.5 Unconventional normal state properties Optical spectroscopy is able to reveal important information on the unconventional nor- mal state charge dynamics of high-temperature superconductors, as already demonstrated impressively for cuprates [116]. Therefore, similar approaches were applied to iron-based superconductors in order to identify (via a comparison to cuprates) key ingredients for high-temperature superconductivity. It turned out that iron pnictides are moderately cor- related materials, where (instead of Mott physics) Hund’s coupling plays an important role. While the appearance of a pseudogap or a quantum critical point is yet not well established, the nematic phase drew enormous attention - even giving a feedback effect for research on cuprates. Important insights gained from infrared spectroscopy on iron pnictides are summarized below. Correlations A comparison between the kinetic energy of electrons obtained by band structure calcu- lations and optical measurements revealed that iron pnictides are moderately correlated materials2 [123]. This indicates that coherent and incoherent processes play both a cer- tain role in transport properties, in good agreement with the broad mid-infrared plateau discussed above. 1 One has to keep in mind that the absolute value of the mass renormalization factor strongly depends on the plasma frequency and thus on the model chosen for the analysis. This is probably the reason why Tu et al. report smaller values than those given here [142]. 2 A similar result can be obtained experimentally by comparing the spectral weight of coherent and incoherent processes [162]. However, this method is very sensitive to the fitting approach. 54 4 Electrodynamic properties In particular, LaFeAsO was found to be more correlated than LaFePO [165]. This observation fits well with systematic room temperature studies performed by Nakajima et al. on electron doped Ba(Fe1−xCox)2As2, hole doped Ba1−xKxFe2As2, and isovalently substituted BaFe2(As1−xPx)2 [129, 141, 166]. They conclude that coherent transport is enhanced by Co and P substitution, and slightly suppressed by K doping; incoherent trans- port is basically not affected by Co doping, reduced by P substitution and enhanced by K doping. Therefore, it seems that chemical substitution controls the Drude spectral weight in quite different ways depending on dopant sites or type. The commonality found in those studies is that superconductivity seems to be suppressed when the charge dynamics become too coherent. Hund’s coupling The renormalization of electronic bands due to correlations does not only reduce the Drude response (compared to band theory predictions), it also implies that additional spectral weight is present at higher energies [163]. This can lead to unusual temperature- dependent effects at high energies, which can not be attributed to interband transitions. Such temperature-dependent spectral weight transfer around 5000 cm−1 was already re- ported in 2008 for the ab-plane of Ba- and SrFe2As2 [266], and later also confirmed along the c-axis of Ba(Fe0.926Co0.074)2As2 [131]. The direction of spectral weight transfer, its independence of Co doping and its energy scale are not compatible with a Mott-like cor- relation mechanism, but can be explained with Hund’s rule coupling effects [163, 164]. Due to Hund’s coupling, electron spins tend to align parallel. This can lead on the one hand to the localization of electrons, if their kinetic energy is smaller than the coupling energy; on the other hand, the thus induced local and itinerant electrons are not isolated from each other, but still coupled via Hund’s coupling [164]. As at low temperatures, thermal excitations become weak, more and more itinerant electrons get localized by this interaction, explaining an increasing spectral weight transfer to high energies. Pseudogap In cuprates, the pseudogap in the density of states is directly visible in the out-of-plane conductivity: while σc(ω) is rather flat at T > T ∗ (indicating incoherent transport), the low-frequency conductivity gets suppressed below T ∗. Surprisingly, the gap value is basically temperature-independent. An extended Drude analysis indicates that this process is intimately related to a gap in the excitation spectrum, visible as the onset of in-plane coherence and a gap in the frequency-dependent in-plane scattering rate (due to the coupling to a bosonic resonance mode). 4.4 Iron-based superconductors 55 In iron pnictides, the appearance of a pseudogap is yet not well established. Up to now, two different “types” of pseudogaps were proposed from optical measurements. In underdoped Ba1−xKxFe2As2, spin density wave and superconducting order coexist. Addi- tionally, a gap-like hump structure appears well above Tc for frequencies around 150 cm−1, corresponding to the energy range of the superconducting gap [136, 137]. Dai et al. argued via a spectral weight analysis that while the spin density wave is a competitive order to superconductivity, this pseudogap acts as a precursor (Tc < TPG < TSDW) [137]. One should note, however, that a similar feature was already observed in BaFe2As2 [132], as well as optimally electron and hole doped compounds [133–135]. In those cases, alter- native interpretations were given, involving a low-lying interband transition [133, 134] or an impurity band [134]. Moon et al. investigated single crystals of the Ba 122 family [138]. For optimally sub- stituted Co and P compounds, they observed a shallow dip in the conductivity around 650 cm−1 that is accompanied by a threshold in 1/τ(ω), typically for the pseudogap in cuprates. This feature does not appear in the parent compound and the overdoped re- gion, coinciding with the development of antiferromagnetic correlations. Therefore, they suggested that in iron pnictides, antiferromagnetic fluctuations related to the spin density wave instability of the parent compound cause a pseudogap. Nematic phase and in-plane anisotropy Optical studies on mechanically detwinned Ba(Fe1−xCox)2As2 as-grown and annealed sam- ples were performed by Dusza [167, 170, 172] and Nakajima et al. [169, 171], respectively. In both cases, a pronounced anisotropy at finite frequencies appears that can be attributed to an anisotropic spin density wave gap formation (see Fig. 4.11). This anisotropy extends up to very high frequencies, well exceeding the energy scale of kBTs,SDW. Moreover, the FeAs phonon around 260 cm−1 gets highly anisotropic; in BaFe2As2, it basically disap- pears along the a-axis. Detailed measurements on twinned BaFe2As2 evidenced a phonon splitting below Ts,SDW, with the minority contribution (along the a-axis) having a higher center frequency [174]. Using local density approximation combined with dynamical mean field theory, Yin et al. [168] calculated the anisotropic optical conductivity for BaFe2As2 resulting from orbital polarization. They found three extra excitations appearing in the spin density wave phase around 1000 cm−1. While two lower energy excitations are much more pronounced along the a-direction, the third peak is more pronounced along the b-direction. All this is in good agreement with the results shown in Fig. 4.11. However, one should note that the 56 4 Electrodynamic properties Figure 4.11: In-plane conductivity of mechanically detwinned BaFe2As2 (adapted from Ref. [169]) for (a) E ‖ a and (b) E ‖ b-axis and temperatures below (blue line) and above (red line) Ts,SDW. At low energies and temperatures, σa > σb is observed, consistent with resistivity measurements. A pronounced anisotropy appears in particular at finite frequencies that can be attributed to an anisotropic spin density wave gap formation. Furthermore, the FeAs phonon around 260 cm−1 disappears in σa(ω). A cusp-like shape for the lowest energy spin density wave feature in σb is only observed by Nakajima et al.. Along the b-direction, spectral weight is transferred up to very high energies. experiments reveal spectral weight transfer to even higher energies than predicted, in particular along the b-direction (∼ 17000 cm−1). From a multi-component analysis one should be able to distinguish whether the un- conventional behaviour ρb > ρa results from the anisotropy in the scattering rate or the Fermi surface parameters n/m∗ (determined by the Drude spectral weight), as 1/ρ = σdc = nq2τ/m∗. Unfortunately, there is no consensus concerning this topic: Dusza et al. claim that the anisotropy results from the Fermi surface parameters, outweighing the conventionally expected 1/τa > 1/τb [167, 170]. In contrast, Nakajima et al. conclude 1/τb > 1/τa, while the Drude weight stays rather isotropic [171]. In chapter 9, we will discuss our results for magnetically detwinned EuFe2As2. While those measurements were obtained under constant external stress due to a me- chanical clamp, Mirri et al. investigated the optical anisotropy that is induced in BaFe2As2 4.4 Iron-based superconductors 57 by a tunable Helium gas pressure cell [173]. While optical anisotropy can be induced by pressure below and above Ts,SDW (consistent with resistivity measurements), a hysteretic behaviour and a persistent detwinning (after the pressure is reduced to zero) can be only detected below Ts,SDW. Furthermore, they observe at Ts,SDW a divergence of the reflectiv- ity anisotropy normalized to the applied pressure. This behaviour is strikingly similar to resistivity measurements under tunable strain, indicating a nematic transition [25]. Quantum criticality In the proximity of an antiferromagnetic quantum critical point, spin fluctuations should be strong enough to modify the scattering process of quasiparticles and induce non-Fermi liquid behaviour. Optical measurements are in principle an ideal tool to get further insight into this topic: different conduction channels can be disentangled, uncovering the dc conductivity and scattering rate of coherent carriers. In Ba(Fe0.92Co0.08)2As2 and Ba(Fe0.95Ni0.05)2As2, the normal state resistivity can be well fitted up to high temperatures with R(T ) ∝ Tn, using the exponents n = 1.25 and n = 1.5, respectively [128, 130]. However, a multi-component analysis by Wu et al. uncovered in both cases a “hidden” Fermi liquid behaviour 1/σdc(T ) ∝ 1/τ(T ) ∝ T 2 of the narrow Drude, being masked by a temperature-independent incoherent background. In contrast to those compounds, the resistivity in Ba0.6K0.4Fe2As2 is rather linear at low temperatures, but shows a tendency towards saturation above 175K [161]. Here, Dai et al. concluded from a multi-component analysis that the incoherent background masks non-Fermi liquid behaviour of the narrow Drude (1/σdc(T ) ∝ 1/τ(T ) ∝ T ), extending up to room temperature. This holds also for BaFe2(As0.7P0.3)2. Most importantly, the authors demonstrated that the spectral weight at a certain cut-off frequency follows the same temperature-dependence as the inverse scattering rate, giving a model-independent proof of their findings. As already mentioned in chapter 3.2.2, a diverging effective mass around the optimum substitution level was up to now only uncovered for BaFe2(As1−xPx)2, but not for Co doped compounds [83, 84]. Therefore, those on the first view contradictory results of Wu and Dai et al. could be another indication how sensitively the electronic properties of iron pnictides dependent on the substitution position. A model-independent spectral weight analysis on electron doped compounds could shed further light on this topic. 5 MATERIALS AND METHODS In this chapter, the basic concepts of Fourier-ransform Intfrared (FTIR) spectroscopy and Superconducting Quantum Interference Device (SQUID) magnetometry are introduced. As the corresponding setups used for the measurements in this thesis are either commercial machines or cryostats constructed in the framework of another thesis, the description concentrates on the main principles. The chapter concludes with an overview about the investigated samples. 5.1 FTIR spectroscopy Fourier-transform spectroscopy is a powerful tool to study the electrodynamic proper- ties of solids [113]. Instead of time-consuming frequency sweeps needed for example in conventional grating spectrometers, the energy resolution is gained via a Fourier transfor- mation of interfered light. Today, this method is widely used in the infrared, as here light sources, beamsplitters, mirrors, filters, windows and detectors are available that cover a broad frequency range; the commercially available setups contain usually modifications of a Michelson interferometer (see Fig. 5.1). In a conventional Michelson interferometer, monochromatic light is split into two ap- proximately equal parts, which follow different beam paths, before they interfere again at the beamsplitter and pass (or get reflected by) a sample. The interference pattern depends on the optical path difference. By modulating the length of one of the light paths with 60 5 Materials and methods Figure 5.1: Schematic drawing of the beampath in a Michelson interferometer. Light from a source passes a semi-transparent beamsplitter; half of the light gets reflected by a fixed mirror, the other half by a movable one. After the reflection on these mirrors, the two beams meet again at the beamsplitter and interfere. The interfered light passes the sample (or gets reflected on it) and finally is collected by the detector that records the intensity dependent on the position of the movable mirror (i.e. on the optical path difference). a movable mirror, one obtains the intensity as a function of mirror position; this spatial dependence can be translated to a frequency-dependence via Fourier transformation, no matter whether the light is mono- or polychromatic. This allows rapid recording of spectra over a broad energy range. As an obtained spectrum contains both, information about the sample and the light source, one has to normalize it properly. For a transmission spectrum, this normalization is done by a comparison measurement without the sample; for a reflection spectrum, the response of a metallic mirror (usually gold, silver or aluminium) that possesses a reflection close to unity is taken. In order to obtain reliable absolute values, one has to make sure that sample and ref- erence have the same size and surface consistence. There are basically three widely used technical implementations in order to satisfy those requirements: • The same size of sample and reference can be guaranteed by taking the same aperture in both measurements. This is commonly done for transmission measurements, as well as for reflection measurements on films or other homogeneous, rather flat and big samples. • A microscope can be used to focus infrared light on very small, flat and homoge- neous areas of both, sample and mirrors. The spot size is much smaller than the dimensions of sample and reference, guaranteeing the same amount of light hitting both specimens. 5.1 FTIR spectroscopy 61 • Alternatively, after measuring the reflection of a sample, one can coat its surface in-situ with a thin gold layer, creating the corresponding “perfect” mirror with the same size and surface consistence; all measurements are then repeated on the gold coated samples (under the same experimental conditions). In order to account for any time-dependent influences (for example caused by source or detector), one can take additionally the spectrum of a standard reference mirror. Table 5.1 summarizes which approaches were taken for the infrared reflection measure- ments performed in the framework of this thesis, as well as important technical param- eters. Microscope measurements were carried out on a Bruker Vertex spectrometer with attached Hyperion microscope at the 1st Physical Institute, Stuttgart; gold evaporation measurements on a Bruker IFS 66v/s in the group of D. N. Basov at the UC San Diego. In gold evaporation measurements, the sample gets mounted point-like on a copper cone. This does not only avoid back-reflections from the sample holder, but also minimizes the glued area and thus induced strain on the sample, being important for the magnetic de- twinnnig experiments described in chapter 9. The interesting temperature range between 15K and 290K was accessible by Helium flow cryostats; all cryostats were designed to con- tain a mechanically movable holder that allows precise positioning of sample and reference in the focus of the light (see Ref. [175] and [176] for the detailed designs). Range Magnet Technique Detector Beamsplitter Source Window FFIR no gold 1.6K Bolometer Mylar 23µm Xenon PP FIR no gold 4.2K Bolometer Mylar 6µm Globar PP MIR no microscope 77K Photovoltaic KBr Globar KBr NIR no microscope 77K InSb CaF2 Tungsten KBr FFIR yes gold 1.6K Bolometer Mylar 23µm Xenon PP FIR yes gold 4.2K Bolometer Mylar 6µm Globar PP MIR yes gold 4.2K Photovoltaic KBr Globar PP Table 5.1: Technical parameters of the FTIR-measurements performed in the framework of this thesis for the different frequency ranges. One can distinguish between measurements with and without magnetic field, as well as measurements, where absolute values were obtained via in-situ gold evaporation or the availability of small spot sizes in a microscope setup. Furthermore, the used detectors, beamsplitters, sources and windows (PP: polypropylene) are given. One should note that the mid-infrared spectral range was not recorded by a conventional MCT-detector, as it displays strong non-linearity effects non-negligible in the case of high reflecting iron pnictides. Fig. 5.2 displays the advanced design of the magneto-optical setup [177]. A helium flow cryostat is directly connected to a 9T Oxford Spectromag, meaning that only one polypropylene window has to be transmitted in reflection measurements. This reduces 62 5 Materials and methods intensity losses caused by the window; however, this window can not be changed, meaning that most of the light in the mid-infrared frequency range gets absorbed. A linear transla- tor allows the positioning of sample and reference mirror in the focus of the light, as well as the movement of the sample to the gold evaporation unit. Figure 5.2: Magneto-optical setup as designed and constructed in the group of D. N. Basov, UC San Diego (adapted from Ref. [177]). A helium flow cryostat is directly connected to the vacuum chamber of a 9T Oxford Spectromag magnet. A linear translator allows the positioning of sample and reference in the focus of the light, as well as the movement of the sample to the gold evaporation unit. 5.2 SQUID magnetometry A Superconducting Quantum Interference Device is able to convert magnetic flux to a voltage with an extremely high sensitivity that reaches the quantum limit [2, 178]. It consists of a superconducting loop that is interrupted by one (rf or ac SQUID) or two (dc SQUID) Josephson junctions. In an external magnetic field, the combination of Josephson effect and flux quantization causes a voltage drop across the loop that varies periodically with the external field (if the bias current through the loop is fixed); the periodicity is one elemental flux quantum Φ0. In order to enhance the sensitivity, the SQUID sensor is usually not directly exposed to the magnetic field that should be measured, but shielded from the environment. The magnetic field of interest is then inductively coupled to the SQUID via a flux transformer 5.2 SQUID magnetometry 63 that consists in the easiest case of a pickup coil (exposed to the environment) and an input coil (close to the SQUID). In order to avoid background signals, one can also use a gradiometer; it consists of coils wound in opposite directions. As a magnetic field rapidly decreases spatially, one can thus distinguish between local and distant magnetic field sources: while the latter induces equally, but opposite currents within the two coils, local fields of interest will cause a net signal. Magnetic Property Measurement System Magnetization measurements performed in the framework of this thesis were obtained on a 7T MPMS XL (Magnetic Property Measurement System) built by Quantum Design [91]. It consists of an rf SQUID sensor, protected by a superconducting shield that guarantees a stable environment. The detection coil is a second-order gradiometer, consisting of three sets of coils, where the outer coils are wound opposite to the inner ones. In the dc mode, the sample is moved through those detection coils, changing the flux. This flux change is then detected by the SQUID sensor; the overall output signal is the absolute magnetization in “emu” (electromagnetic unit), which is usually normalized to the mass of the sample. Furthermore, one can also perform ac measurements, where a small ac drive magnetic field is applied. This causes a time-dependent moment in the sample that allows the mea- surement of the “dynamic susceptibility” without motion [179]. At very low frequencies, this quantity corresponds to the static susceptibility χ that can be extracted from the slope of field-dependent dc magnetization measurements. At higher frequencies, however, the signal might gain a phase shift ϕ. In this case, the susceptibility is conventionally described by the real and imaginary parts: χ′ = χ cosϕ (5.1) χ′′ = χ sinϕ (5.2) where the imaginary component χ′′ indicates dissipative processes in the sample; those appear for example in ferromagnets, spin glasses or type II superconductors. The MPMS includes computer controlled temperature and magnetic field stabilization; temperatures between 1.6K and 400K, as well as magnetic fields between -7T and 7T are accessible. In order to reveal hysteretic behaviour, special care has to be taken concerning the temperature and field setting mode [10]. When measuring at stabilized temperatures T , the stabilization process usually also includes oscillating around the target temperature and its duration depends on the setpoint temperature. For a material with time-dependent magnetization M , both lead to non-reproducibile M(T ) curves. Therefore, we typically 64 5 Materials and methods chose to sweep the temperature in M(T ) measurements1. For the field setting, the “oscil- lating mode” is the most accurate, as the magnetic field over- and undershoots the target value in cycles with decreasing relative amplitude in order to acquire an optimal result. However, for hysteretic samples one should take the “no overshoot” mode, meaning that the magnetic field is changed monotonously to the desired value. Furthermore, one has to consider the effect of the remanent magnetization captured in the superconducting coils. Usually, we therefore determine the zero field by searching at temperatures above any hysteretic phase the magnetic field (in the range of a few G), which leads to zero signal at the SQUID. However, as the cooling process in a homogeneous zero field can be crucial (especially for superconductors), some measurements were also done using the Quantum Design “ultra low field” option which guarantees field nulling better than 0.05G remanent field2. We did not find any qualitative difference of the magnetization compared to measurements with our standard procedure to determine the zero field. Samples are typically mounted with the help of a plastic straw. In the case of low-signal measurements, the mounting should be as symmetric as possible around the sample in order to avoid that the background signal has a different center than the sample position; in this case, an asymmetric signal would be generated that can not be analysed by the MPMS software. For measurements with the magnetic field applied along the cleaved surface of a plate-like sample, the sample can be fixed between two stripes of plastic foil by wrapping Teflon tape around them. For the perpendicular direction, both halfs of a gelatine capsule are padded by Teflon tape or cotton to form a plateau where in between the sample can be placed. Both, plastic foil and gelatine capsule, can be sewed in the straw. 5.3 Investigated samples In the framework of this thesis, several EuFe2(As1−xPx)2 high quality single crystals syn- thesized in the group of P. Gegenwart (University of Göttingen / Augsburg) by the Bridge- 1 Typical measurements take less than 10 s with the settings: 4 cm scan length, 2 scans per measurements and 24 points, i.e. for a temperature sweep with 0.2K/min, the temperature error is less than 0.05K and thus negligible (visible also in the absent thermal hysteresis in the paramagnetic regime). For ac measurements covering a big temperature range, we typically chose the sweep velocity 0.1K/min, as here measurements usually are slower. With the parameters: 1G drive amplitude, 7Hz wave frequency, amplifier gain 1, 1 block to average, 1 scan per measurement and 1 s settling time, measurements are faster than 1min, i.e. in this case the temperature error is less than 0.1K. 2 Those measurements were performed in the group of R. Kremer, MPI FKF, Stuttgart 5.3 Investigated samples 65 man method [230] were investigated by magnetization measurements (see chapter 7); Table 5.2 summarizes their compositions x and masses m. x 0 0.055 0.09 0.12 0.16 0.165 m(mg) 6.1 9.7 0.45 1.1 2.3 7.0 x 0.17 0.26 0.35 0.39 1 m(mg) 1.2 0.44 0.59 4.0 0.16 Table 5.2: Compositions x and masses m of EuFe2(As1−xPx)2 single crystals investigated in magnetization measurements (see chapter 7). For optical measurements in the infrared frequency range (see chapter 8), further freshly cleaved EuFe2As2 and BaFe2As2 single crystals with shiny surfaces> 1mm2 were available; for terahertz measurements (see attachment A), thin Ba(Fe1−xCox)2As2 films grown by pulsed laser deposition [180] were supplied by S. Haindl and K. Iida (IFW Dresden). 6 EU BASED IRON PNICTIDES EuFe2As2 is exceptional among the parent compounds of the iron pnictides, as it exhibits - additional to the Fe spin density wave - long-range magnetic order of the Eu2+ local moments at low temperatures. Nevertheless, bulk superconductivity can be induced by mechanical pressure, isovalent substitution and by electron or hole doping. Thus, Eu based iron pnictides offer the extraordinary possibility to study the interplay of unconventional superconductivity, itinerant and local magnetism. In this chapter, first the magnetic properties of EuFe2As2 are introduced. Their de- velopment under pressure, doping and Eu substitution are in the focus of the following discussion, providing a basis for a unified phase diagram which will be presented in chap- ter 7. Afterwards, the electronic properties of Eu based iron pnictides are reviewed, focusing on the influence of the Eu2+ magnetic moments. Additionally, measurements on Eu 122 compounds that revealed important general properties for iron pnictides are shortly presented. 6.1 EuFe2As2 Measurements using a commercial SQUID magnetometer are ideal to get first insight into the magnetic properties of a material. The standard procedures commonly employed to Eu 122 iron pnictides are summarized in Table 6.1. One should note that further investigations 68 6 Eu based iron pnictides such as (resonant) x-ray scattering, Mössbauer studies and neutron diffraction1 are usually needed to completely understand the magnetic ordering of such layered materials, in which often inter- and intralayer interactions compete, causing “exotic” types of magnetism. Measurement Information M(T ) at T = Ts,SDW Feature at spin density wave transition. M(T ) at T > TEu Curie-Weiss fit provides information about the predom- inant magnetic interactions between Eu2+ spins and the valency of Eu. M(T ) and M(H) at T ≤ TEu Shape of curves provides information about the ordering. M(T ) at T = TEu TEu shifts with increasing field to lower temperatures for antiferromagnetic order, to higher temperatures for ferro- magnetism. M(T ) at T < TEu Hysteresis often results from ferromagnetism, spin glass behaviour or superconductivity. M(H) at T < TEu Saturated moment gives evidence about the Eu valency. Table 6.1: Summary of standard procedures to get information about the magnetism in Eu 122 iron pnictides from temperature- (T ) and field- (H) dependent magnetization (M) measurements. Combined magnetization and Mössbauer studies revealed already in 1993 that EuFe2As2 exhibits two magnetic transitions: an itinerant ordering in the Fe subsystem at ∼ 190K is followed at lower temperatures (∼ 19K) by the local ordering of magnetic Eu2+ mo- ments [216]. The itinerant order can be identified as a spin density wave that is accom- panied by a tetragonal-to-orthorhombic structural phase transition [26]. Resistivity and specific heat show only one sharp feature at Ts,SDW, indicating a first-order transition sim- ilar to SrFe2As2 (see Fig. 6.1a-b, [27, 28, 181]). The spin density wave is also visible as a jump in the magnetization (see Fig. 6.1c, [182]). However, as divalent Eu2+ magnetic mo- ments (S = 7/2, L = 0) dominate the magnetization, usually a paramagnetic background has to be subtracted before identifying the tiny feature. Below TEu ∼ 19K, the temperature-dependent magnetization is basically constant along the c-axis, while a clear cusp appears for magnetic fields applied in the ab-plane (see Fig. 6.1d-e). Both evidences antiferromagnetic behaviour. However, the Curie-Weiss fit indicates predominant ferromagnetic interactions among the Eu2+ spins [28, 182, 216]. Therefore, in analogy to other layered systems, A-type antiferromagnetism was proposed, meaning that spins order ferromagnetically within the ab-plane plane, but neighbouring layers are coupled antiferromagnetically. This suggestion was confirmed by neutron and 1 Unfortunately, Eu is a strong neutron absorber, complicating those investigations [8]. 6.1 EuFe2As2 69 1 0 2 0 3 00 2 1 6 0 2 0 0 0 4 8 0 1 0 0 2 0 0 0 . 2 0 . 4 0 2 0 1 2 0 3 6 0 3 6 1 5 0 2 0 0 ( e ) H | | c , 0 . 1 T χ d c (e mu / m ol) T ( K )  (10 -4 emu / mol) T ( K ) ( c ) ( b )  a b (m Ω cm ) T s , S D W T E u T ( K ) ( a ) ( g ) H | | c ( f ) H | | a b χ d c (e mu / m ol) ( d ) H | | a b 0 . 1 T M (µB / f.u.) µB H ( T ) 2 K 1 5 K 2 5 K M (µB / f.u.) 2 K 1 5 K 2 5 K C (J / mol K) Figure 6.1: Standard characterization measurements on EuFe2As2. (a) The resistivity ρab(T ) shows kinks at Ts,SDW and TEu [224]. (b) While the specific heat C(T ) exhibits a pronounced peak at Ts,SDW, (c) for magnetization measurements the paramagnetic background first has to be extracted to identify a tiny feature caused by the spin density wave transition [28]. (d) The temperature-dependent dc susceptibility χdc(T ) = M(T )/µ0H shows for an in-plane magnetic field (H ‖ ab) a cusp at TEu and (e) flattens at T < TEu for H ‖ c. Both is typical for an antiferromagnet with the easy axis in the ab-plane. (f) At 2K (blue line), M(H) saturates for H ‖ ab at ∼ 1.0T, (g) for H ‖ c at ∼ 1.6T. The saturation fields increase for 15K (purple line) and 25K (magenta line). A step-like increase for H ‖ ab indicates a spin-flip or flop (also called “metamagnetic transition”), which is typical for an antiferromagnet [182]. resonant x-ray diffraction measurements [8, 183, 189]; the detailed magnetic structure of EuFe2As2 is shown in Fig. 6.2. Both, itinerant Fe and local Eu2+ order are commensurate, with the spins pointing parallel or antiparallel along the a-axis, consistent with theoretical predictions [181]. One should note that Mössbauer and magnetic torque measurements indicate a slight canting (∼ 10◦) of the Eu2+ moments out of the ab-plane [234, 253], which was not observed by neutron and resonant x-ray measurements [8, 183, 189, 195]. While the Eu2+ magnetic order is already suppressed at fields of only ∼ 1T [7], the spin density wave is still robust in pulsed magnetic fields of 55T [190]. 70 6 Eu based iron pnictides Figure 6.2: Magnetic struc- ture of EuFe2As2 at 2.5K [8]. Fe2+ moments (brown ar- rows) align antiferromagnet- ically along the a- and the c-axis, but ferromagnetically along the b-axis. Eu2+ mo- ments (blue arrows) align within one Eu layer ferro- magnetically along the a- axis, and neighbouring ab- planes are coupled antiferro- magnetically. Table 6.2 summarizes the magnetic properties of related compounds that are located at the “end” of Eu 122 phase diagrams. Only EuFe2As2 exhibits a spin density wave transition. Furthermore, the Eu2+ magnetic order varies tremendously among the different compounds, although Eu is in all cases divalent1. The origin of the huge diversity of different phases is quite simple: as different layers are too far separated for a direct overlap of the Eu 4f -orbitals, the magnetic interactions are probably mediated via conduction electrons by the indirect RKKY-exchange [36]. However, this indirect coupling oscillates between ferro- and antiferromagnetic regimes depending on the details of the Fermi surface. Theoretical calculations by Akbari et al. revealed even an influence of the spin density wave on the RKKY-exchange (see also chapter 7.1, [11, 219]). Thus, when passing through the Eu 122 phase diagrams, strong changes in the Eu2+ magnetic order can be expected. Compound TEu Weiss Proposed order Direction Ref. EuFe2As2 19K FM A-type AFM a-axis [8, 216] EuFe2P2 29K FM FM c-axis [210, 211] EuRu2As2 17K FM FM c-axis [215] EuCo2As2 39K FM A-type AFM vs. helical ab-plane vs. almost along c-axis [214, 251] EuNi2As2 14K AFM [216] Table 6.2: Characteristic information about the Eu2+ magnetism in Eu 122 parent com- pounds: magnetic ordering temperature TEu, predominant spin exchange given by the sign of the Curie or Néel temperature, proposed order and spin direction. 1 In the case of EuCo2P2 and EuRh2As2, mixed Eu valencies between 2 and 3 appear [217, 218]. For both, incommensurate helimagnetism was observed. Bandstructure calculations indicate a strong sensitivity of the magnetic configuration on the Eu valency [218]. 6.2 Phase diagrams 71 6.2 Phase diagrams In order to compare several compounds, the transition temperatures are consistently de- termined as followed: superconductivity is defined by its onset and the zero resistivity temperature, Tc,on and Tc,0, respectively. The Eu2+ magnetic ordering temperature TEu corresponds to a kink in resistivity and a peak in magnetization measurements. Standard measurements on Eu 122 pnictides often do not show clearly distinguishable features that would allow a distinct separation of structural and spin density wave transition tempera- tures. Thus, for the sake of simplicity, those transitions will be treated as one and referred to as Ts,SDW. 6.2.1 Pressurized systems In the case of isovalent substitution, a certain element is replaced by a smaller (larger) one to shrink (expand) the unit cell similar to applied mechanical pressure. However, one should keep in mind that even isovalent substitution induces additional chemical disorder. It is still under debate, whether such disorder could be the actual tuning parameter through the iron pnictide phase diagrams [68]. Mechanical pressure In 2009, Miclea et al. reported that pressures around 2GPa induce in EuFe2As2 supercon- ductivity up to ∼ 30K [29]. However, zero resistivity was prohibited by a re-entrance of resistivity at the Eu2+ magnetic ordering temperature (see Fig. 6.3a). Detailed investiga- tions by Matsubayashi and Kurita et al. revealed that bulk superconductivity is achieved only if hydrostatic pressure is transmitted ([222, 224], see Fig. 6.3b), reconciling several pressure studies [188, 220, 221, 223, 225–229, 232]. Fig. 6.4 displays the phase diagram of EuFe2As2 under hydrostatic pressure [222, 224]. At low pressures, the spin density wave gets continuously suppressed and the Eu2+ mag- netic ordering temperature increases slightly (see also Fig. 6.3c). Around ∼2GPa, TEu starts to rise strongly. Surprisingly, this change in Eu2+ magnetism coincides with the pressure, where the onset of superconductivity - without reaching zero resistance - can be observed. Thus, in a narrow pressure range, Eu2+ magnetism, spin density wave order and traces of superconductivity coexist. Zero resistance and bulk superconductivity oc- curs only between ∼ 2.5 − 3GPa. It disappears as soon as the Eu2+ magnetic ordering temperature is as high as Tc,on. Magnetization measurements indicate that A-type antiferromagnetic Eu2+ order coex- ists with superconductivity. Above ∼ 6GPa, a ferromagnetic signal is observed, visible in a clear out-of-phase signal of the ac susceptibility (see Fig. 6.3d, [222]). TEu increases fur- 72 6 Eu based iron pnictides 0 2 0 4 00 . 0 0 . 2 0 . 4 0 5 0 1 0 0 1 5 0 2 0 0 0 . 0 0 . 2 0 . 4 T c , o n T E u ( b ) 2 . 3 0 G P a ( a ) T ( K )  (m Ω cm ) 2 . 0 3 G P a T c , o n T s , S D W T E u 0 G P a 1 . 1 G P a 1 . 9 G P a 2 . 2 G P a 2 . 6 G P a T ( K )  (m Ω cm) 1 5 2 0 2 51 2 3 4 2 0 4 0 6 0 1 . 1 G P a 0 G P a χ d c (e mu / m ol) T E u , A F M 0 . 8 G P a T ( K ) ( c ) T E u , F M8 G P a ( d ) χ'ac (a.u.)6 G P a 4 G P a T E u , A F M χ''ac (a.u.) T ( K ) Figure 6.3: Characteristic temperature-dependent in-plane resistivity and magnetization of EuFe2As2 under pressure. Grey arrows highlight the development of the transition tempera- tures with pressure. (a) Using silicone oil as pressure transmitter, superconductivity sets in at Tc,on ∼ 30K, but due to a resistivity re-entrance at TEu, zero resistivity is not achieved [29]. (b) Using Daphne 7474 as transmitting medium, Ts,SDW is continuously suppressed with pres- sure. Around 2.6GPa, zero resistance without re-entrance and non-Fermi liquid behavior (R(T ) ∼ T for T > Tc) is observed [224]. (c) DC susceptibility: at low pressures, TEu in- creases slightly. The cusp is typical for antiferromagnetic order [222]. (d) AC susceptibility: antiferromagnetism does not induce an out-of-phase signal (χ′′ ≈ 0 at 4GPa). However, at ∼ 6GPa a second feature appears with χ′′ > 0 that strongly shifts to higher temperature and gets dominant with increasing pressure. The strong out-of-phase signal is typical for ferromagnetism [222]. 6.2 Phase diagrams 73 0 2 4 6 8 1 0 1 20 2 0 4 0 1 0 0 1 5 0 E u F e 2 A s 2 T s , S D W T c , o n T c , 0 T E u P r e s s u r e ( G P a ) T ( K) Figure 6.4: Phase diagram of EuFe2As2 under pressure (open symbols: Ref. [222], closed symbols: Ref. [224]). TEu (blue diamonds) slightly increases at low pressures and starts to rise stronger just before Ts,SDW (black squares) gets suppressed and the superconducting dome is positioned (Tc,on: light red circles, Tc,0: red circles). ther up to pressures around 10GPa, where a collapsed tetragonal phase transition takes place [221, 223]. As the ionic radius of Eu2+ is larger than that of Eu3+, this struc- tural phase transition coincides with an increase of the Eu valency. As trivalent Eu is nonmagnetic, TEu decreases for higher pressures [222]. Isovalent substitution: EuFe2(As1-xPx)2 In 2009, Ren et al. discovered in EuFe2(As0.7P0.3)2 superconductivity that sets in at ∼ 26K and coexists with Eu2+ ferromagnetism below ∼ 20K [30]. Both transitions are visible in the specific heat, proving their bulk character. This observation attracted tremendous interest in the condensed matter community, as superconductivity and ferro- magnetism are usually antagonistic phenomena (see chapter 3.3) and “chemical” pressure was discovered as a tuning parameter through the iron pnictide phase diagram. It turned out that the electronic properties of EuFe2(As1-xPx)2 are extremely sensitive to the synthesis, possibly due to procedure-dependent internal strain: for single crystals grown by the Bridgeman method, the superconducting dome appears around x ∼ 0.2 [9, 10, 230, 233, 235, 238, 240, 241, 272, 280], while for polycrystals as well as single crystals grown by solid state reaction the dome is shifted to x ∼ 0.3 [30, 231, 232, 234, 236, 237, 243]. However, the Eu2+ magnetic order is not affected by the growth process. Fig. 6.5 compares the resulting phase diagrams, which resemble a lot EuFe2As2 under pressure. Most remarkably, TEu starts to increase above x ∼ 0.12, which again coincides with the onset of superconductivity. This reveals a significant interplay between electronic and magnetic properties in those compounds. 74 6 Eu based iron pnictides However, in contrast to EuFe2As2 under pressure, TEu decreases at lower substitution levels. It is intriguing to assign the minimum around x ∼ 0.12 to the change from anti- to ferromagnetic Eu2+ ordering, as signatures of ferromagnetism are observed at high sub- stitution levels, consistent also with theoretical calculations [207, 239]. However, several measurements indicated antiferromagnetic Eu2+ order up to significantly higher P concen- trations [230, 243]. Furthermore, even results for the pure P compound are contradictory: while Mössbauer and magnetization measurements indicate a successive helical ordering around 26K with the spins tilted 15 ◦ from the c-axis, neutron measurements state pure ferromagnetism. It should be noted that indications for successive magnetic orderings were also reported for mixed compounds [232, 234, 236, 237, 243]; in Ref. [242], the lower transition was attributed to a minority phase. One contribution of this thesis was to rec- oncile those - on the first view contradictory - reports (see Table 6.3), which is presented in more detail in chapter 7, as well as in Ref. [9] and [10]. 0 . 0 0 . 4 0 . 80 1 0 2 0 1 0 0 2 0 0 0 . 0 0 . 4 0 . 8 ( a ) B r i d g e m a n ( b ) S o l i d s t a t e r e a c t i o n T (K ) T s , S D W T c , o n T c , 0 T E u P c o n t e n t i n E u F e 2 ( A s 1 - x P x ) 2 Figure 6.5: Phase Diagram of EuFe2(As1-xPx)2 for samples grown by (a) the Bridgeman method [9, 10, 230, 233, 235, 238, 240, 241, 272, 280] and (b) solid state reaction [30, 231, 234, 236, 237]. In both cases, TEu (blue diamonds) shows a minimum around x ∼ 0.12. This minimum is close to the composition, where Ts,SDW gets suppressed (black squares) and superconductivity sets in (Tc,on: light red circles, Tc,0: red circles). Isovalent substitution: Eu(Fe1-xRux)2As2 An alternative way of isovalent substitution is the replacement of Fe by Ru. It should be noted that Ru influences the volume of the unit cell much less than P substitu- tion [246]. Nevertheless, Eu(Fe1-xRux)2As2 also shows bulk superconductivity in a narrow range around x ∼ 0.2 (see Fig. 6.6, [31, 245]). However, the onset of superconductivity can be observed at least up to x = 0.55, which is in stark contrast to the mechanically pressurized or P substituted case. One obvious difference to those systems is that the 6.2 Phase diagrams 75 Technique Sample Observation Eu2+ Ref. Specific heat Poly Increasing magnetic field broad- ens and shifts anomaly to higher temperatures. FM [30] Magnetic compton scattering Poly Enhanced spin polarization of the Fe-3d orbital. FM [231] Mössbauer Poly Increased Eu hyperfine field, Eu2+ spins almost aligned along c-axis (20◦ tilted). FM [234] Neutron Single No antiferromagnetic reflec- tions, Eu2+ spins aligned along c-axis. FM [241] Neutron Single Eu2+ spins aligned along c-axis, minority phase with lower or- dering temperature. FM [242] Magneti- zation Single Shape and field-dependence of magnetization. AFM [230, 243] Muon, Möss- bauer Single Eu2+ spins almost aligned along c-axis (12◦ tilted). AFM [243] Table 6.3: Arguments for Eu2+ ferromagnetism (FM) vs. antiferromagnetism (AFM) in superconducting EuFe2(As1-xPx)2 polycrystalline samples (poly) or single crystals (single). Eu2+ magnetic ordering temperatures of EuFe2As2 and EuRu2As2 are quite similar and therefore, TEu basically stays constant in the whole phase diagram. Thus, it is intriguing to conclude that superconductivity gets suppressed, if the Eu2+ magnetic ordering ex- ceeds Tc. However, this conclusion does not explain why Eu(Fe1-xRux)2As2 exhibits bulk superconductivity only in a very narrow range. Despite the lower ordering temperatures than in the case of EuFe2P2, EuRu2As2 is also suggested to order ferromagnetically along the c-axis (see Table 6.2). Therefore, it is not surprising that for Eu(Fe1-xRux)2As2, successive anti- and ferromagnetic transitions were observed [31, 245], too, and Mössbauer studies on a superconducting sample indicate a reorientation of the Eu2+ spins towards the c-axis [31]. Mechanical vs. chemical pressure In order to compare the influence of mechanical and chemical pressure, Sun and Tokiwa et al. investigated EuFe2As2 and EuFe2(As1-xPx)2 under pressure, concluding that the 76 6 Eu based iron pnictides 0 . 0 0 . 5 1 . 00 1 0 2 0 1 0 0 2 0 0 1 0 2 0 3 0 0 . 0 0 . 2 0 . 4( b ) T s , S D W T c , o n T c , 0 T E u R u c o n t e n t i n E u ( F e 1 - x R u x ) 2 A s 2 T (K ) ( a ) T c , o n x = 0 . 2 5 T ( K )  / 300K x = 0 . 5 5 x = 0 . 4 8 Figure 6.6: (a) Phase diagram of Eu(Fe1-xRux)2As2 [245]. TEu (blue diamonds) stays almost constant. Superconductivity sets in (light red circles), as soon as Ts,SDW (black squares) is suppressed. While zero resistivity (red circles) is achieved only around x ∼ 0.2, traces of superconductivity exist up to x = 0.55, (b) visible in resistivity measurements for x = 0.25 (magenta line), 0.48 (purple line) and 0.55 (pink line), which also exhibit a resistivity re- entrance at TEu. overall behaviour is very similar [232, 238]1. However, when incorporating also the Eu2+ magnetic order in the phase diagram, as it is done in Fig. 6.7, it gets obvious that the actual processes are more complex: choosing the equivalent pressure for complete P substi- tution x =ˆ 15.4GPa, one can map reasonable well the electronic properties on top of each other, however, not the Eu2+ magnetic ordering temperature. Moreover, when compar- ing the influence on the crystallographic c- and a-axes (see Fig. 6.8), mechanical pressure compresses the c-axis significantly stronger than P substitution if one uses the above cho- sen equivalent. Thus, assuming that isovalent substitution just acts as chemical pressure is clearly oversimplified: actually superconductivity is induced much faster compared to mechanical pressure. The most plausible scenario is that the disorder induced by isovalent substitution causes the differences in the phase diagram. Disorder could also weaken the RKKY-exchange, explaining why TEu first decreases in EuFe2(As1-xPx)2. 1 Sun et al. proposed that a pressure-induced valence change of Eu sets in already at the left side of the superconducting dome for both, EuFe2As2 under mechanical pressure and EuFe2(As1-xPx)2 [232]. This could lead to additional electron doping, possibly even favouring superconductivity. However, previous Mössbauer studies on superconducting EuFe2(As0.7P0.3)2 did not reveal any Eu3+ foreign phase [243] and Matsubayashi et al. [222] revealed valency changes in EuFe2As2 only at much higher pressures. Intriguingly, changes in the valency are reported in both cases close to the solidification of the pressure transmitting medium. It is not clear how this should induce a measurement error, nevertheless further investigations are necessary to clarify this issue. 6.2 Phase diagrams 77 0 . 0 0 . 2 0 . 4 0 . 6 0 . 8 1 . 00 2 0 4 0 1 0 0 2 0 0 0 5 1 0 1 5 T s , S D W T c , 0 T E u C h e m i c a l p r e s s u r e T (K ) E u F e 2 A s 2 M e c h a n i c a l p r e s s u r e ( G P a ) Figure 6.7: Phase diagrams of EuFe2As2 single crystals under chemical (P substitution, open dark symbols, [9, 10, 230, 233, 235, 238, 240, 241, 272, 280]) and mechanical pressure (closed bright symbols, [222, 224]) mapped on top of each other, using the relation: P substitution x =ˆ 15.4GPa. While the electronic properties are quite similar, TEu (blue diamonds) develops differently. However, in both cases TEu starts to increase stronger close to the pressure, where Ts,SDW (black squares) gets suppressed and superconductivity appears (Tc,0: red circles). 0 . 0 0 . 4 0 . 8 1 0 1 1 1 2 0 . 0 0 . 4 0 . 8 3 4 5 0 5 1 0 1 5 0 5 1 0 1 5 c - axis (Å) C h e m i c a l p r e s s u r e ( P s u b s t i t u t i o n ) c o l l a p s e d t e t r a g o n a l M e c h a n i c a l C h e m i c a l a - axis (Å) M e c h a n i c a l p r e s s u r e ( G P a ) Figure 6.8: Crystallographic c-axis (left) and a-axis (right) of EuFe2As2 under chemical (P substitution, blue open points, [230, 236]) and mechanical pressure (green squares, [221]). Using the equivalent: P substitution x =ˆ 15.4GPa that maps the electronic properties on top of each other, reveals that mechanical pressure contracts the c-axis much stronger than P substitution. 78 6 Eu based iron pnictides Summary In summary, when comparing the phase diagram of mechanically pressurized EuFe2As2 with those of EuFe2(As1-xPx)2 and Eu(Fe1-xRux)2As2, we can conclude for the interplay of superconductivity and magnetism: • Microscopic coexistence of spin density wave ordering and superconductivity was yet not evidenced. • TEu does not behave universal. However, for mechanically pressurized EuFe2As2 and EuFe2(As1-xPx)2, it increases strongly as soon as the spin density wave is suppressed and superconductivity sets in. • Considering only bulk superconductivity, the superconducting dome is rather nar- row in all cases. For mechanical pressure and P substitution, superconductivity is suppressed as soon as TEu ≈ Tc,on. • In all phase diagrams, the Eu2+ magnetic order changes with substitution. However, contradictory reports exist which kind of magnetism coexists with superconductivity. 6.2.2 Electron doping In the case of 122 iron pnictides, electron doping can be realized by the substitution of Fe by an atom that prefers a lower valence state. Indeed, superconductivity up to ∼ 25K was found in Co doped BaFe2As2 [247]. However, similar approaches were much less successful for Eu based 122 iron pnictides. Eu(Fe1-xCox)2As2 In 2009, Jiang et al. presented results on Eu(Fe0.89Co0.11)2As2 single crystals grown by using (Fe,Co)As as self-flux, which showed the onset of superconductivity at ∼ 21K [32]. However, with this synthesis technique only polycrystals where produced that exhibit zero resistivity at low temperatures [249], but no single crystals [248–251, 258, 260].1 The synthesis of single crystals with zero resistance could be achieved only by using Sn-flux, however, with very low transition temperatures (significantly below 10K) [252–257, 259]. Fig. 6.9 compares the resulting phase diagrams2. One should note that here, the growth 1 Chen et al. argued that in their polycrystals, the coupling between conduction electrons and local Eu2+ moments is reduced, visible in resistivity measurements with less obvious features at TEu and the absence of re-entrant superconductivity [249]. 2 Only the study by Ying et al. [248] does not fit to the others. The reason is probably that only nominal sample compositions are given. Due to this uncertainty, it is excluded from the further discussions. 6.2 Phase diagrams 79 technique influences all properties, also the Eu2+ magnetic transition. In both cases, the Eu2+ magnetic ordering temperature shows a minimum, where structural as well as spin density wave gets suppressed and the rather narrow superconducting dome is positioned around this composition. Mössbauer experiments on superconducting compounds indi- cate that the spin density wave does not take place in the whole sample volume [254]. Thus, again rather phase separation than microscopic coexistence of superconductivity and itinerant Fe magnetism is observed. 0 . 0 0 . 2 0 . 40 1 0 2 0 1 0 0 2 0 0 0 . 0 0 . 2 0 . 4 ( b ) S n - f l u x i n E u ( F e 1 - x C o x ) 2 A s 2 T (K ) ( a ) S e l f - f l u x E l e c t r o n d o p i n g T s , S D W T c , o n T c , 0 T E u Figure 6.9: Phase diagram of Eu(Fe1-xCox)2As2 samples grown using (a) self-flux [249– 251, 258, 260] and (b) Sn-flux [252–257, 259]. In both cases, the Eu2+ magnetic ordering temperature (blue diamonds) shows a minimum around the composition, where Ts,SDW (black squares) gets suppressed and the superconducting dome is positioned (Tc,on: light red circles, Tc,0: light red circles). Zero resistivity for self-flux grown samples was only observed in polycrystalline samples [249]. More similarities to mechanically pressurized and isovalently substituted EuFe2As2 can be found when comparing the evolution of Eu2+ magnetic order throughout the phase diagram: as soon as the spin density wave gets suppressed, the Eu2+ spins cant more and more out of the ab-plane. Although magnetization measurements show up to high Co contents indications for antiferromagnetic interlayer coupling (often interpreted in terms of canted antiferromagnetism [253, 254, 257] or helimagnetism [32, 251]) and multiple magnetic transitions [250, 251, 256, 257], neutron diffraction measurements on a super- conducting sample concluded ferromagnetic ordering along the c-direction with a moment size of 6.2µB, slightly lower than the theoretical expected value of ∼ 7µB. Those contra- dictory reports get even more mysterious, when considering that EuCo2As2 shows very similar magnetic behaviour to EuFe2As2: even signatures of a metamagnetic transition were observed in M(H)-curves around ∼ 3T [214]. 80 6 Eu based iron pnictides Eu(Fe1-xNix)2As2 Surprisingly, Ni substitution on the Fe site does not induce superconductivity at all [251, 261, 262]. Mössbauer measurements on doped compounds reveal a canting of the Eu2+ spins out of the ab-plane and multiple magnetic transitions appear that were interpreted in terms of a crossover from anti- to ferromagnetism. Ren et al. concluded that this ferromagnetism suppresses superconductivity [261]. However, this argumentation has two weak points: first of all, the end member EuNi2As2 is antiferromagnetic, and not ferro- magnetic [216]. Secondly, when comparing the development of TEu and Ts,SDW for Co and Ni, both can be mapped perfectly on top of each other at low doping level1 (see Fig. 6.10). 0 . 0 0 . 1 0 . 21 0 2 0 1 0 0 2 0 0 E l e c t r o n d o p i n g T s , S D W T E u c l o s e d s y m b o l s : C oo p e n s y m b o l s : N i T (K ) Figure 6.10: Phase di- agram of Eu(Fe1-xYx)2As2 with Y = Co [249, 258] (closed symbols) or Ni [261] (open symbols). The x- axis of the Ni phase diagram was expanded by a factor of two, as Ni can provide twice as much electrons as Co. TEu (blue diamonds) and Ts,SDW (black squares) perfectly match. Eu(Fe1-xIrx)2As2 Up to now, the most “successful” approach to induce superconductivity by electron doping has turned out to be Ir substitution [33, 263, 264]. In 2013, Paramanik et al. reported for Eu(Fe0.86Ir0.14)2As2 polycrystals an onset of superconductivity at 22.6K and zero- resistivity as well as diamagnetic shielding at 19K, which is disturbed by a resistivity re- entrance due to Eu2+ magnetic ordering around 14K [33]. Only two months later, Jiao et al. published results on Eu(Fe0.88Ir0.12)2As2 single crystals with a sharp superconducting transition between 22K and 21K (without any resistivity re-entrance) [263]. Both groups proposed a canting of the Eu2+ spins out of the ab-plane. However, whereas the first group suggested canted A-type antiferromagnetic ordering of Eu2+, the latter favoured a ferromagnetic ordering. 1 In order to compare them correctly, one has to consider that Ni substitution provides two electrons per substituent, whereas Co provides only one. Thus, the x-scale of the Ni phase diagram should be expanded by a factor of two. Indeed, it is quite surprising that the mapping works out so well, as the end members of the phase diagrams, EuNi2As2 and EuCo2As2, have quite different Eu2+ magnetic ordering temperatures (see Table 6.2). 6.2 Phase diagrams 81 Summary Summing up, electron doping induces superconductivity in EuFe2As2 when substituting Fe by Co or Ir. While in the latter case, zero resistivity and a clear diamagnetic shielding was observed around 20K, the results for Co doping are less convincing: the diamagnetic shielding is significantly lower and the phase diagram strongly depends on the growth technique. Ni substitution does not induce superconductivity at all, although Ts,SDW and TEu show exactly the same behaviour as for Co doping. When comparing electron doping to isovalent substitution and mechanical pressure, the overall behaviour is very similar: superconductivity probably does not coexist with the spin density wave order on a microscopic scale and TEu shows an almost identical behaviour as in the case of P substitution. The reports about the type of Eu2+ magnetic order are as contradictory as for isovalent substitution. Despite those similarities, one new aspect concerning the interplay of superconductivity and magnetism should be emphasized: the results for Co doping clearly show that superconductivity with zero resistivity can appear, even when the onset temperature is much lower than TEu. Thus, one could speculate that superconductivity is only suppressed by the Eu2+ magnetic order, if the two antag- onistic phenomena would like to set in at similar temperatures. This might happen for Ni substitution, as Ni doping usually leads to slightly lower critical temperatures than Co [247, 265]. 6.2.3 Eu dilution Chemical substitution at the Eu site does not only act as doping or isovalent pressure, it also weakens the Eu2+ magnetism. Up to now, poly- and single crystals were successfully grown for Ca, Sr, Ba, K, Na and La substitution: Ca, Sr and Ba are isovalent substitu- tions; K and Na substitution leads to hole doping, La to electron doping. In some cases, those approaches were further combined with physical pressure or additional substitutions at another element site (“codoping”). Unfortunately, although the magnetic and super- conducting properties are directly linked to each other, detailed studies on the evolution of the Eu2+ magnetism dependent on dilution level are still rare. Isovalent substitution Isovalent substitution on the Eu site is not expected to induce superconductivity, as cor- responding 122 parent compounds exhibit “only” structural as well as spin density wave transitions. Those appear for Ca, Eu, Sr and Ba approximately at 170K, 190K, 205K and 140K, respectively [5]. One should note that the transition temperatures do not just 82 6 Eu based iron pnictides follow the ionic radii, which are increasing from Ca to Ba [66].1 Thus, only Ca substitution will induce positive chemical pressure and shrink the unit cell volume, while Ba and Sr will have the opposite effect. The dilution effect on the local Eu2+ magnetic order can be summarized as followed: • As expected, Ba substitution continuously suppresses Ts,SDW [269]. While the lowest detectable TEu = 0.5K was found for x = 0.45, all samples with higher Ba content (except BaFe2As2) still show at low temperatures an enhanced response of the spe- cific heat. This indicates that Eu2+ magnetic moments significantly influence the material’s properties even at very high dilution levels. • In the case of Ca substitution, the Eu2+ magnetic transition is completely suppressed at x = 0.75, but for 50% dilution still as high as 9K [267, 268]. Additional physi- cal pressure on Eu0.5Ca0.5Fe2As2 induces superconductivity with a maximum Tc of 24K and possibly a coexistence region of spin density wave ordering, local Eu2+ magnetism and superconductivity (around ∼ 1.4GPa). • Several studies investigated the superconductivity in Sr diluted samples with addi- tional 11-14% Co-doping [270–274]. While Eu(Fe1-xCox)2As2 often only shows traces of superconductivity, in Eu diluted specimen bulk superconductivity was observed up to 15K. It was suggested that superconductivity disappears as soon as Tc < TEu (see Fig. 6.11, [271]). Specific heat measurements under external magnetic fields showed that an antiferromagnetic Eu2+ order at 2K persists up to a Eu dilution of as high as 80% [273]. Thus, the Eu2+ magnetism in EuFe2As2 iron pnictides is surprisingly robust against iso- valent dilution. Hole doping In 2008, Jeevan et al. reported for hole doped Eu0.5K0.5Fe2As2 polycrystals a super- conducting transition at 32K [34]. The high quality of the samples was indicated by a rather sharp resistivity transition with a width of less than 2K. Bulk superconductiv- ity was evidenced by a well-defined anomaly in the specific heat and a full diamagnetic shielding. However, those measurements revealed only very weak features below 10K 1 The reason why BaFe2As2 shows significantly lower structural and spin density wave transition tem- peratures as the others is yet not clear. Previous quantum oscillation measurements indicated effective masses systematically smaller than in the case of EuFe2As2, meaning that the latter is much more correlated [209]. 6.2 Phase diagrams 83 0 . 0 0 . 2 0 . 4 0 . 6 0 . 8 1 . 00 1 0 2 0 3 0 T c , o n T c , 0 T E u E u 1 - x S r x ( F e 0 . 8 4 C o 0 . 1 4 ) 2 A s 2 T (K ) S r c o n t e n t Figure 6.11: Phase diagram of Eu1-xSrx(Fe0.86Co0.14)2As2 [271]. The Eu2+ magnetic or- der (blue diamonds) is continuously suppressed with Sr substitution. Superconductivity and magnetism coexist, as long as the onset of superconductivity (Tc,on: light red circle) and zero resistance temperature (Tc,0: light red circle) are well above TEu. that could be ascribed to Eu2+ magnetism. As dilution typically leads to the suppres- sion of magnetic order, it is very tentative to assign weak signatures of Eu2+ ordering to short-range interactions. However, one has to keep in mind that even in non-diluted compounds, the superconducting signal can be strong enough to mask Eu2+ magnetism. Indeed, systematic magnetization studies on x = 0.38 polycrystals revealed that a small peak buried by the diamagnetic signal shifts with increasing magnetic field towards lower temperatures, indicating a long-range antiferromagnetic ordering [279]. Measurements on codoped Eu0.5K0.5(Fe1-yNiy)2As2 give further evidence for such a long range magnetic or- der at 50% dilution level: as soon as Ni doping has suppressed superconductivity, a strong antiferromagnetic Eu2+ signal can be detected around 8K [286]. Fig. 6.12 shows the phase diagram of Eu1-xKxFe2As2. Further experimental studies [275– 282] showed that x = 0.5 actually represents the optimal doping level for poly- and single crystals. It should be noted that the synthesis of single crystals was successfully performed only in 2012, as K is highly reactive and hard to be controlled. Therefore, the single crystal phase diagram is still incomplete, especially concerning higher substitution levels. However, those doping levels are of special interest, as KFe2As2 is a particular end member of the iron pnictides: it is superconducting with a Tc of 3.8K [283]. In Fig. 6.12, also studies on the effect of Na doping are included [284, 285]. First measurements indicate a rather broad superconducting dome similar to K doping (see Fig. 6.12). Additional P substitution on a 30% diluted sample was able to increase Tc by 5K [287]. In summary, hole doping combined with Eu dilution leads to a rather broad super- conducting dome with high Tc, which is consistent with other 122 iron pnictides [80]. 84 6 Eu based iron pnictides 0 . 0 0 . 2 0 . 4 0 . 6 0 . 8 1 . 00 2 0 4 0 1 2 0 1 6 0 2 0 0 E u 1 - x N a x F e 2 A s 2 T s , S D W T c , 0 E u 1 - x K x F e 2 A s 2 T s , S D W T c , o n T c , 0 T E u T (K ) H o l e d o p i n g Figure 6.12: Phase diagram of hole doped Eu1-xYxFe2As2 (Y = Na [286], K [279, 280]). The superconducting dome and the corresponding transition temperatures are marked in red (Na: dark red stars for Tc,0; K: light red circles for Tc,on and red circles for Tc,0; closed symbols: single crystals [280], open symbols: polycrystalline samples [279]). The underdoped part of the dome is depicted more transparent, as here evidence for bulk superconductivity is not yet provided. Ts,SDW is faster suppressed for Na (grey hexagons) than for K substitution (black squares). TEu (blue diamonds) is continuously suppressed and disappears around 60% substitution level. Although Na substitution leads to the faster suppression of the structual and spin density wave transition, the superconducting domes are very similar. Experimental studies on the Eu2+ magnetism are complicated, as the diamagnetic shielding typical for a bulk supercon- ductor masks the Eu2+ signal. However, there is growing experimental evidence indicating a coexistence of Eu2+ long-range magnetic order and superconductivity, consistent with the results from isovalent Eu dilution. Electron doping Substituting Eu by the rare earth ion La leads to the suppression of the structural and spin density wave transition, probably due to electron doping [288–291]. However, only traces of superconductivity were found for the maximum possible solid-solution of x = 0.27. Zhang et al. applied additional pressure on x = 0.22 and 0.27 samples, successfully inducing superconductivity with sharp transitions and zero resistivity (see Fig. 6.13, [291]). Interestingly, neither La doping nor additional pressure up to 2.2GPa do notably influence the Eu2+ magnetic transition temperature. Indeed, TEu slightly increases under pressure. Surprisingly, one can observe the coexistence of spin density wave magnetism, superconductivity and local Eu2+ magnetism, while Tc is always smaller than TEu. 6.2 Phase diagrams 85 0 . 0 0 . 1 0 . 20 5 0 1 0 0 1 5 0 2 0 0 0 1 2 T (K ) L a c o n t e n t E u 1 - x L a x F e 2 A s 2 T s , S D W T E u T c , o n T c , 0 E u 0 . 7 3 L a 0 . 2 7 F e 2 A s 2 u n d e r p r e s s u r e P r e s s u r e ( G P a ) Figure 6.13: Phase diagram of Eu1-xLaxFe2As2 (left) and Eu0.73La0.27Fe2As2 under pressure (right). Black squares correspond to Ts,SDW and blue diamonds to TEu. While zero resistivity Tc,0 (red circles) appears only in a narrow dome for Eu0.73La0.27Fe2As2 under pressure, the onset of superconductivity Tc,on (light red circles) appears in a much broader range. A coex- istence region of spin density wave magnetism, superconductivity and local Eu2+ magnetism shows up. Summary: Eu dilution When summarizing all those studies on Eu diluted compounds, the most important find- ings concerning the interplay of magnetism and superconductivity are: • There is growing experimental evidence that the Eu2+ antiferromagnetic interlayer coupling is quite robust against dilution [273, 279, 286], as can be seen in Fig. 6.14. Overall, a rather smooth decrease of TEu with dilution level takes place. This is typical for an intermetallic compound with local antiferromagnetism that is diluted with nonmagnetic ions [271]. Ba substitution seems to suppress the Eu2+ magnetic ordering the fastest. Interestingly, ionic radii as well as Ts,SDW changes the most in the case of Eu substitution by Ba. This indicates a complex interplay between structural, electronic and magnetic properties. • It is quite surprising that when comparing EuFe2As2 and Eu0.73La0.27Fe2As2 un- der pressure, the latter diluted compound exhibits significantly lower critical tem- peratures [291]. This implies that in some cases, disorder might be even more harmful to superconductivity than Eu2+ magnetism. In contrast, Tc increases in Eu0.7Na0.3Fe2As2 by additional P substitution on the As site [287] and in Eu1-xSrx- (Fe0.86Co0.14)2As2 by increasing Eu dilution [271]. • Indications for phase coexistence of (zero-resistance) superconductivity with spin density wave as well as diluted Eu2+ magnetism was reported in several compounds 86 6 Eu based iron pnictides 0 . 0 0 . 2 0 . 4 0 . 6 0 . 8 1 . 00 1 0 2 0 E u 1 - x B a x F e 2 A s 2 E u 1 - x C a x F e 2 A s 2 E u 1 - x S r x ( F e 0 . 8 4 C o 0 . 1 4 ) 2 A s 2 E u 1 - x K x F e 2 A s 2 T (K ) E u d i l u t i o n l e v e l T E u : Figure 6.14: Suppression of the Eu2+ magnetic ordering temperature TEu with increasing Eu dilution for Eu1-xYxFe2As2. Y = Ba (pink open squares, [269]), Ca (beige stars, [268]), Sr (codoped with 14% Co) (blue open diamonds, [271]) and K (dark cyan closed hexagons, [279]. [267, 268, 285, 291], which is in stark contrast to non-diluted systems. Further systematic studies on these compounds is highly desirable in order to shed new light on the complex interplay between superconductivity and magnetism. 6.3 Electronic properties 6.3.1 Normal state properties Band structure Early angle resolved photoemission studies (ARPES) on EuFe2As2 [191, 192] revealed a band structure typical for 122 iron pnictides: in the paramagnetic phase, electron pock- ets around the X-point ((pi,pi)-point) and hole pockets around the Γ -point ((0,0)-point) could be resolved. Below the spin density wave transition, the electron-like bands (at the X-point) are folded back to the Γ -point, where they hybridize with the hole-like bands, leading to the opening of a spin density wave gap and a droplet-like Fermi surface. Such rather three-dimensional Fermi surface sections are also consistent with quantum oscilla- tion measurements [209]. Those ARPES studies did not resolve any change when crossing the Eu2+ magnetic ordering temperature, indicating a weak coupling between the Eu and FeAs subsystems. Therefore, Eu based compounds were examined in further studies in order to get character- istic information about 122 iron pnictides, and the Eu2+ magnetic ordering was completely neglected [199, 205, 235, 290, 293]. For example, EuFe2(As1-xPx)2 was the first chemically pressurized superconductor investigated by ARPES [235]. It was found that P substitution 6.3 Electronic properties 87 mainly influences the hole pockets in the center of the Brillouin zone, leading to a strongly three-dimensional Fermi surface with reduced nesting and (at least at low P substitution) to a charge-neutral development, which can not be described by a simple rigid-band shift1. In contrast to the early investigations on EuFe2As2, recent results show that the coupling between Eu and FeAs layers can not be neglected [204]. While former studies were carried out with energies, where the photoemission cross section of the d-orbitals is dominant, Adhikary et al. also used photon energies, which are sensitive to p-orbitals. This revealed that the electronic states near the Fermi energy level possess large p-orbital contributions due to Eu-As and Fe-As hybridizations, and are sensitive to the Eu2+ ordering. Electronic transport The electronic transport properties of EuFe2As2 are similar to other iron pnictides (see chapter 3.2.2, [188]): although a spin density wave takes place at elevated temperatures Ts,SDW ∼ 190K, the resistivity (of twinned samples) stays metallic down to lowest temper- atures (see also Fig. 6.1). At TEu, the resistivity exhibits only a slight kink; the influence on the scattering rate and thus also on the optical properties is negligible [184]. While Hall and thermopower measurements indicate that the carriers dominating the electronic transport above Ts,SDW are holes [28, 280], also the influence of Dirac fermions was sug- gested [255]. In order to study the spin density wave, besides ARPES also optical spectroscopy [146, 184, 186] and scanning tunneling microscope (STM) measurements [203] were performed on EuFe2As2. While ARPES studies reported a spin density wave gap 2∆ ≈ 50 − 70meV [191], STM results indicate a gap value of 2∆ ≈ 90meV, with a small suppression of the density of states at the Fermi energy that persists even up to higher temperatures than Ts,SDW. Further contradictory results were obtained from optical measurements in the infrared frequency range: while Wu et al. found only one spin density wave gap below 161meV [184], Moon et al. resolved two energy gaps below 56 and 178meV, consistent with BaFe2As2 and SrFe2As2 [146, 266]. 1 It was found that an electron pocket with yz-orbital character and a hole pocket with xy-character do not change with P substitution. Both stay basically two-dimensional. However, a strongly dispersive hole pocket with z2-orbital character increases with increasing P content, and the almost degenerate hole pockets with xz- and yz-orbital character are strongly suppressed at the Γ -point (and even disappear close to optimal doping), while they remain almost unchanged at the Z-point. A reduction from three to only one hole pocket at the Γ -point is consistent with DFT calculations [207]. 88 6 Eu based iron pnictides Coupling between magnetic and electronic subsystems Probably the most interesting question about Eu 122 pnictides is, how strong the local Eu2+ magnetic moments are coupled to the itinerant electrons in the FeAs layers. In the case of superconducting compounds, the influence of the Eu2+ magnetism on the elec- tronic properties is clearly visible in transferred hyperfine fields from the Eu2+ moments on Fe of about 1T [234, 251], causing phenomena such as the re-entrance of resistivity in the superconducting phase and field-enhanced superconductivity (see chapter 6.3.2). For EuFe2As2, however, the situation is more complicated: it was concluded that electron scattering on the local Eu2+ moments plays only a minor role for the transport proper- ties [188] and the reflectivity basically does not change at the Eu2+ magnetic ordering temperature [184]. Only recently, ARPES [204], Mössbauer [200] and resonant x-ray mea- surements [195] could resolve signatures of coupled FeAs and Eu subsystems. In order to get more information about the coupling strength, electron and nuclear magnetic spin resonance studies are discussed in the following. Electron spin resonance (ESR) is in principle a powerful tool to investigate the spin dynamics in Eu based iron pnictides, as one probes the relaxation of the local Eu2+ spins which is dominated by the interaction with conduction electrons [187, 202, 248, 258, 269, 272, 277, 289]. Therefore, the resonance field and linewidth of the ESR signal both depend on the exchange coupling between Eu2+ spins and itinerant electrons, as well as on the electronic density of states at the Fermi energy. In a common metal, the ESR linewidth increases linearly with temperature (“Korringa relaxation”); the g-factor, which is determined from the resonance field, is temperature-independent and shifted compared to its vacuum value (“Pauli susceptibility”). However, if the conduction electrons couple stronger to the Eu2+ spins than to the lattice, the microwave energy absorbed by the Eu2+ moments is not emitted fast enough to the lattice (“bottleneck scenario”) [258, 272]. Instead, it is kept in the magnetic system of Eu2+ spins and conduction electrons. As a consequence, the ESR response depends on experimental parameters such as microwave frequency and Eu concentration. Unfortunately, there is up to now no consensus whether such a bottleneck scenario is observed in Eu based iron pnictides or not [258, 269, 272].1 1 Garcia et al. argued that an almost unchanged g-factor through the phase diagram (which does thus not reflect the evolution of the Fermi surface), the measured magnetic field-dependence of the linewidth and opposing evolutions of the calculated density of states at the Fermi energy and the actual linewidth slope, all suggest a bottleneck scenario for Eu(Fe1-xCox)2As2 [258]. However, Krug von Nidda et al. did neither observe any field dependence of the linewidth, nor that Eu dilution changes its slope [272]. Furthermore, Rosa et al. reported that when replacing Eu by Ba, the spin density wave is suppressed and the Korringa slope decreases simultaneously; opposite to what one would expect in the bottleneck scenario [269]. 6.3 Electronic properties 89 A much clearer picture can be drawn by nuclear magnetic resonance (NMR) spec- troscopy, which probes the spin polarization of 75As nuclei that takes place via the Fermi contact interaction with delocalized carriers. If the Eu2+ local moments are interacting with the conduction electrons, this should be visible in the signal. Thus, NMR spec- troscopy is ideal to investigate the coupling between Eu2+ magnetic moments and the FeAs layers. Up to now, 75As NMR investigations were performed on Eu(Fe0.995Co0.005)2- As2 [252] and Eu diluted compounds [273, 281]. It was found that the typical measurement parameters - spin-lattice relaxation rate, spin-spin relaxation rate and the shift of the cen- tral line - all follow a Curie-Weiss law. Furthermore, even in diluted compounds, the spin-lattice relaxation is more than two times larger than those of Ba 122 pnictides, and it increases with the Eu content; the latter is also true for the Curie-Weiss constant. All this proves that the 75As nuclei interact with the localized Eu2+ moments. In order to get an estimate about the coupling between the FeAs layers and Eu subsystems, one can calculate the hyperfine coupling by relating the central line shift to the dc magnetic sus- ceptibility. For Eu(Fe0.995Co0.005)2As2, the hyperfine coupling is 60 times larger than for NdFeAsO0.85F0.15, providing direct experimental evidence for a strong coupling between the Eu2+ localized moments and the conduction electrons in the FeAs layers [252]. Electronic in-plane anisotropy Up to now, most of the investigations on the in-plane anisotropy of iron pnictides con- centrated on Ba(Fe1-xCox)2As2. Recently, magnetic torque measurements on very small samples [295], as well as resistivity and thermopower measurements on mechanically de- twinned crystals were also performed on Eu compounds [197, 240]. In those studies, the focus was on investigating the general in-plane anisotropy and possible nematic behaviour of iron pnictides, and the Eu2+ magnetic order was neglected. In order to examine whether the Eu2+ magnetism leads to an additional anisotropic response, Fig. 6.15 depicts resistivity data for Eu(Fe1-xCox)2As2 [197]. The overall be- haviour is similar to Ba 122-compounds (see chapter 3.2.2); the Eu2+ magnetic order does not notably induce additional anisotropy. Magnetic detwinning Although magnetic fields are usually very ineffective to detwin iron pnictides [24], EuFe2- As2 can be completely detwinned below TEu by small magnetic fields (∼ 1T) applied along one of the orthorhombic axes [7]. In the corresponding experiment, Xiao et al. observed already at 0T a strong imbalance of twin domains, possibly due to strain induced by 90 6 Eu based iron pnictides 0 . 6 0 . 8 1 . 0 0 1 0 0 2 0 0 1 . 0 1 . 5 0 1 0 0 2 0 0 T E uT E u T S D W  a  b  (T ) /  (30 0K) T S D W T ( K )  b /  a ( b ) x = 0 . 0 3 4( a ) x = 0 T s T ( K ) Figure 6.15: Temperature-dependent resistivity ρ(T ) [197] and its anisotropy for mechan- ically detwinned Eu(Fe1-xCox)2As2 with (a) x = 0 and (b) x = 0.034. Vertical lines mark the transition temperatures TEu (orange; kink in resistivity), TSDW (red; maximum in specific heat, minimum in dρ/dT ) and Ts (brown; onset of hump in specific heat). The anisotropy with ρb > ρa arises already slightly above Ts and increases with doping. The Eu2+ magnetic order does not notably induce additional anisotropy. the mounting of the crystal1, and a magnetic field was applied along the orthorhombic a-axis of the minority domain type (defined as (100)o-direction). Fig. 6.16 summarizes the main results. At low fields, the intensity of the (400)o-reflection (representing the minority twins) first decreases, before it abruptly increases around 0.7T and the (040)o-reflection completely disappears. Only a slight hysteresis was revealed. The authors attributed their observations to a giant spin-lattice coupling that leads at high fields to a detwinning of the crystal with its longer orthorhombic a-axis aligned parallel to the external magnetic field; the changes at low fields were interpreted in terms of a tilting of the minority domains out of their balance position due to internal strain induced by the detwinning process. Xiao et al. also performed magneto-resistance measurements on EuFe2As2, concluding that the main observations can be well explained by Eu2+ spin scattering [198]. For currents and magnetic fields applied parallel to one of the orthorhombic in-plane axes, a giant, unconventional magneto-resistance effect takes place: at temperatures T < TEu, the resistance first increases with increasing magnetic field, until it abruptly decreases at the Eu2+ spin-flip field (∼ 0.7T). The low field behaviour resembles the positive magneto- resistance that appears when an external magnetic field is applied parallel to the sublattice magnetization of an antiferromagnet; the abrupt decrease can be understood in terms of 1 Proposed in private communications by the main author Y. Xiao. 6.3 Electronic properties 91 3.9 4.0 4.1 0 200 400 600 Hs||s(100)O 0.5T H ||s(100)O 1T 0T (040)O In te ns ity s(a .su .) H(r.sl.su.) (400)O 3.9 4.1 4.0 3.9 4.0 4.1 4.2 0 1 2 3 Magneticsfields(T) (r. sl. su .) H (a) (b) (400)O (040)O (400)O (040)O Figure 6.16: Selected Bragg reflections of EuFe2As2 at 2K with the magnetic field applied along one of the orthorhombic axes [7]. Momentum transfer Q-scans around the (400)o and (040)o reflections (a) under three typical magnetic fields 0T (black), 0.5T (green) and 1T (red line) as well as (b) for decreasing and increasing magnetic fields displayed in a contour plot. HD1Crit marks the field, where twins tilt out of their balance position, caused by internal strain due to the detwinning at HD2Crit. superzone boundary theory, which predicts an energy gap in the antiferromagnetic state; as soon as the Eu2+ spins flip along the external magnetic field, the antiferromagnetic state and thus the energy gap gets destroyed and the resistivity decreases. In contrast, at temperatures slightly above the Eu2+ magnetic ordering, the resistance directly decreases with increasing field. This can be explained in terms of suppressed spin fluctuations in a magnetic field. Chapter 9 will show measurements done in the framework of this thesis, demonstrating that the magnetic detwinning is actually more complex than presented by Xiao et al.. Furthermore, those studies prove that the magneto-resistance is mainly governed by twin domain dynamics (and not Eu2+ spin scattering), causing distinct hysteretic behaviour which can be observed when well-defined cooling procedures are obeyed. 6.3.2 Superconducting properties Quantum criticality Non-Fermi liquid behaviour, indicated for example by a linear temperature-dependence of the resistivity, was discovered in several unconventional superconductors around optimal doping [292]. Its appearance is widely discussed as due to an underlying quantum critical point or strong spin fluctuations in the vicinity of a suppressed magnetic order. While non-Fermi liquid behaviour above the superconducting dome is also commonly observed in iron pnictides, a divergent quasiparticle mass evidencing a quantum critical point was 92 6 Eu based iron pnictides (up to now) only confirmed for BaFe2(As1-xPx)2, possibly because this is a very clean system [83, 84]. In order to elucidate the origin of non-Fermi liquid behaviour in iron pnictides in more detail, Kurita et al. investigated pressurized EuFe2As2, which shows above the super- conducting dome a linear temperature-dependence of the resistivity [229]. For optimum pressure, while resistivity and Hall measurements at 0T ruled out any divergent quasipar- ticle mass, Fermi liquid behaviour could be restored in magnetic fields around 10T. From those studies, the following conclusions were proposed: on the one hand, as EuFe2As2 under pressure can be considered as an even cleaner system as BaFe2(As1-xPx)2, the non- divergent quasiparticles mass demonstrates that quantum criticality is not the key ingre- dient for high temperature superconductivity in iron pnictides. On the other hand, the latter results under high magnetic fields indicate that spin fluctuations drive the uncon- ventional scattering above the superconducting dome: probably, the external fields which are further enhanced by the Eu2+ moments lead to a high enough internal field to suppress the spin fluctuations. This is an impressive example of how important information about unconventional superconductivity can be revealed by Eu based iron pnictides. Influence of Eu magnetism Understanding the interplay between superconductivity and magnetism is one of the great challenges of today’s solid state physics. Indeed, the coexistence of Eu2+ magnetic order and superconductivity is surprising from a very fundamental point of view: how can the Fe electrons contribute on the one hand to superconductivity and mediate on the other hand the Eu2+ magnetic RKKY-exchange? In order to understand this interplay in more detail, here the main observations are summarized. • The competition of the two orders gets very obvious in the case of the resistivity re-entrance which appears in several superconducting samples at TEu. Magnetic Compton scattering experiments on P and Co substituted Eu compounds indicated that the Fe 3d-contribution is strongly enhanced around TEu [231, 260]. This en- hancement can be quickly suppressed with relatively small magnetic fields in the ab-plane, consistent with the suppression of the re-entrance or even an enhance- ment of Tc found in magneto-resistance measurements [32, 220, 256, 270]. Early work suggested that those observations are due to the suppression of magnon ex- citations [32, 270]. However, as previous experiments demonstrated below TEu a spin canting of the Eu2+ moments and thus a ferromagnetic component along the c-direction, it was proposed that the suppression of the canting by in-plane magnetic fields likely causes the field enhanced superconductivity [9, 256]. Interestingly, the re-entrance only appears when TEu and Tc are very close to each other [238]. This 6.3 Electronic properties 93 is consistent with observations made in chapter 6.2: superconductivity can appear much below TEu (for example in the case of Co doping), however, if TEu and Tc directly compete with each other, superconductivity disappears (for example for P substitution). • Strikingly, the Eu2+ magnetic order was also found to influence the superconducting gap: while in BaFe2(As1-xPx)2 nodal superconductivity was observed, for P substi- tuted Eu compounds infrared and ESR measurements indicated an isotropic s-wave gap [233, 272]. One possible explanation for the different behaviour is that the nodes in iron-based superconductors are not given by symmetry and thus can be lifted by Eu2+ spin scattering [72, 233]. However, a differing band structure due to the short- ened c-axis (compared to Ba 122 compounds) might be also the reason [294]. • Microscopic coexistence of superconductivity and Eu2+ magnetism was demonstrated recently by muon spin relaxation and Mössbauer spectroscopy on EuFe2(As1-xPx)2 [243]. Upper critical fields Compared to other iron pnictides, the Eu 122 family shows significantly reduced critical fields Bc1 and Bc2. In Eu diluted compounds, lower critical fields of only 10 -100mT are reported [272, 276, 277, 279]; for non-diluted compounds, the Meissner state can be even completely absent [263]. While on the one hand, determining those very small lower critical fields is experimentally challenging, on the other hand, the reduction of Bc2 to an experimentally accessible scale opens new perspectives: in the case of EuFe2As2 under pressure, it was found that the upper critical fields (∼ 20T at 0K) are slightly anisotropic, and their temperature-dependence revealed a concave curvature around 1T, similar to the Jaccarino-Peter compensation effect (see chapter 3.3.3, [220, 225]). The results could be nicely described by an internal exchange field of ∼ 75T due to the magnetic Eu2+ moments, explaining the reduced upper critical fields compared to other iron pnictides. For diluted Eu0.5K0.5Fe2As2 [282] and EuxSr1-x(Fe0.89Co0.11)2As2 (x = 0.20 and 0.46) [274], the upper critical fields (∼ 60T for K and 20-30T for Sr substitution) are consistent with a dominant Pauli paramagnetic effect that is proportional to the critical temperature Tc. Interestingly, a concave curvature similar to EuFe2As2 under pressure was also observed for Sr substitution. Although the authors ascribed the unconventional slope to multiband effects, the analysis in terms of internal compensation fields would be interesting; especially as the effect was observed to weaken with Eu dilution. 7 MAGNETIZATION STUDIES ON EuFe2(As1−xPx)2 Although high quality single- and polycrystalline EuFe2(As1−xPx)2 samples were soon available after the discovery of superconductivity in EuFe2(As0.7P0.3)2 [30], investigations on a series of samples are still rare (see chapter 6). In our opinion, however, it needs exactly such systematic studies in order to understand the complex dynamics in Eu based iron pnictides. Therefore, we have performed in- and out-of-plane magnetization mea- surements on a complete set of high quality EuFe2(As1−xPx)2 (x = 0, 0.055, 0.09, 0.12, 0.16, 0.165, 0.17, 0.26, 0.35, 0.39, 1) single crystals (see chapter 5.3), which elucidated the complex magnetic ordering in those compounds and contributed important insights into the interplay of superconductivity and magnetism. The main results can be found in Ref. [9] and Ref. [10], as well as are described in the following in more detail. 7.1 Canted A-type antiferromagnetism In 2011, the first phase diagrams were published for EuFe2(As1−xPx)2 [230, 234]. Despite the availability of single- and polycrystalline samples over the whole phase diagram, they could not deliver a clear picture of how the Eu2+ order changes with chemical pressure and a serious debate was launched which kind of magnetism coexists with superconductivity. We approached this question by low-field in- and out-of-plane magnetization measurements 96 7 Magnetization studies on EuFe2(As1−xPx)2 on a series of EuFe2(As1−xPx)2 single crystals (x = 0, 0.055, 0.12 and 0.35) that should be non-superconducting according to the phase diagram of Ref. [230]; the results which are published partially in Ref. [9] are presented in this section. EuFe2(As0.88P0.12)2 Fig. 7.1 shows the temperature-dependent dc susceptibility χdc(T ) = M(T )/µ0H for x = 0.12, measured parallel and perpendicular to the ab-plane. The shape of the field- cooled curves (see Fig. 7.1a,b) is characteristic for the A-type antiferromagnetism identified in the parent compound: while at TN ∼ 16K, a cusp appears for H ‖ ab, the out- of-plane magnetization flattens. With increasing field, the transition temperature TN is suppressed and a broad shoulder appears that is typical for field-induced ferromagnetism; the suppression is faster for H ‖ ab, consistent with the anisotropy of EuFe2As2. Figure 7.1: Temperature-dependent dc susceptibility χdc(T ) = M(T )/µ0H of EuFe2(As0.88P0.12)2: (a) FC curves with H ‖ ab, (b) FC curves with H ‖ c and (c) ZFC (black open squares) as well as FC (orange closed circles) curves with H ‖ c for various applied fields. The shape of the FC curves is characteristic for antiferromagnetic interlayer coupling. However, a distinct hysteresis appears for H ‖ c at small magnetic fields. 7.1 Canted A-type antiferromagnetism 97 Nevertheless, we find at low magnetic fields a distinct difference between the field-cooled and zero-field cooled curves for H ‖ c that is quickly suppressed with increasing field (see Fig. 7.1c). In order to study this hysteresis, we have also measured the field-dependence of the magnetization, concentrating on low fields (see Fig. 7.2a). We observe a clear hysteresis for H ‖ c; for H ‖ ab, only a much smaller hysteresis appears, which is likely caused by our limited precision in crystal alignment. To follow the opening of the hysteresis for H ‖ c, we plot in Fig. 7.2b the difference ∆Mc between the magnetization curves acquired by sweeping the field down and up (i.e. the hysteresis height). ∆Mc decreases on raising the temperature, until the hysteresis vanishes at TN and it is instructive to compare this trend with the difference between the ZFC and FC curves at the same field. The two trends coincide, demonstrating for T < TN the development of a ferromagnetic component along the crystallographic c-direction. Figure 7.2: (a) Field-dependent isothermal magnetization M(H) of EuFe2(As0.88P0.12)2 for H ‖ ab (brown) and H ‖ c (orange) at T = 2K. (b) Height of the H ‖ c hysteresis ∆Mc (difference between the magnetization curves acquired by sweeping the field down and up, left panel), compared to the temperature-dependence of the difference between the ZFC and FC curves at 20G (right panel); the dotted lines are a guide to the eye to stress the correlation between the two trends. Other compositions Further temperature-dependent out-of-plane measurements on a x = 0.35 single crystal reveal a similar behaviour as for EuFe2(As0.88P0.12)2, while signatures of hysteretic be- haviour in the M(T ) curves of EuFe2As2 and EuFe2(As0.945P0.055)2 are much less clear; corresponding low-field hysteresis measurements at 2K with H ‖ c can not resolve a hysteresis for x = 0, while for x = 0.055 the hysteresis height is slightly and for x = 0.35 significantly positive (see Fig. 7.3). Therefore, we conclude that the ferromagnetic 98 7 Magnetization studies on EuFe2(As1−xPx)2 component develops already at P concentrations much smaller than those where super- conductivity sets in, and persists further in the overdoped range. Figure 7.3: Magnetization curves of EuFe2As2, EuFe2(As0.945P0.055)2, and EuFe2(As0.65P0.35)2: temperature-dependent dc susceptibility χdc(T ) = M(T )/µ0H (upper panels) measured with H = 20G (H ‖ c, ZFC: black open squares, FC: orange closed circles) and hysteresis height ∆Mc (lower panels, 2K: blue closed circles, 10K: red open circles); for x = 0.35, also ∆Mab is given. The hysteresis height for H ‖ c reveals already a small hysteresis for x = 0.055, which is much more evident for x = 0.35. In the latter case, a distinct hysteresis opening is also visible for in-plane measurements, which has a different temperature-dependence. Fig. 7.3d also displays a comparison between isothermal in- and out-of-plane magneti- zation loops of EuFe2(As0.65P0.35)2. One observes a distinct hysteresis opening for H ‖ ab as well as a small opening for H ‖ c. The different temperature-dependence of ∆Mab and ∆Mc indicates two unequal hysteretic mechanisms. Interpretation By analysing measurements of the magnetic properties of various non-superconducting Eu 122 compounds, we identify a ferromagnetic contribution along the c-direction that 7.1 Canted A-type antiferromagnetism 99 becomes stronger with increasing P substitution. Strikingly, we are able to verify a small out-of-plane ferromagnetic contribution already for x = 0.055, a concentration that is much smaller than those where superconductivity sets in. We have also investigated a single crystal of the overdoped substitution range (x = 0.35), where we find hysteretic behaviour for H ‖ c and H ‖ ab. As in- and out-of-plane hysteresis show a different temperature-dependence, we conclude unequal hysteretic mechanisms for the two crystal- lographic directions. In the following discussion, we want to concentrate on the out-of- plane ferromagnetic component. As already discussed in chapter 6, Mössbauer measurements indicate even for EuFe2As2 a small canting of the Eu2+ moments out of the ab-plane; it increases strongly with increasing P substitution until the Eu2+ spins are aligned almost along the c-direction when superconductivity sets in [234]. We show that this canting leads in magnetic sus- ceptibility measurements to a ferromagnetic signal along the c-axis, i.e. hysteretic be- haviour. The fact that one cannot resolve any hysteresis in the parent compound is consistent with an extremely small spin canting, which becomes already more pronounced for EuFe2(As0.945P0.055)2. We believe that the hysteresis loops observed in our single crystal measurements for H ‖ c constitute the key to reconcile the various phase diagrams of Eu 122 pnictides proposed in literature [230, 234]. Since contributions from the c-direction are unavoidable in measurements on polycrystalline samples, a hysteresis is always seen as soon as the spin canting is sufficiently pronounced. In the following we compare the phase diagram of Ref. [230] (based on single crystal measurements with H ‖ ab) with that of Ref. [234] (where polycrystalline samples were investigated). While the former group observes sig- natures of antiferromagnetic Eu2+ order in the superconducting phase, the latter claims the coexistence of ferromagnetism and superconductivity. Indeed, those statements do not contradict each other, if one considers that Ref. [230] is only sensitive to the order- ing of the Eu2+ moments in the ab-plane, while Ref. [234] detects a superposition of in- and out-of-plane coupling: both measurements are consistent with superconductivity that coexists with canted antiferromagnetism, meaning that the interlayer coupling is antifer- romagnetically, but the spins are canted with a ferromagnetic net component along the c-direction. It is intriguing that the canting develops with the suppression of the spin density wave, implying that the Eu2+ spin orientation is defined by the spin density wave anisotropy. Calculations by Akbari et al. on multiband iron pnictides confirm an influence of the itin- erant spin density wave phase on the RKKY-interaction between localized moments [11]. They show that the spin rotational symmetry is broken by the two-dimensional spin den- 100 7 Magnetization studies on EuFe2(As1−xPx)2 sity wave which results in an anisotropic RKKY-interaction described by an anisotropic XXZ-type Heisenberg exchange. This anisotropy is only present in the spin density wave state and vanishes in the normal state. As this exchange is a general feature of rare earth based systems, such a ferromagnetic out-of-plane component that develops with the suppression of the spin density wave can be also expected for other Eu 122 pnictides. 7.2 Re-entrant spin glass Although the in-plane magnetization measurements on EuFe2(As0.65P0.35)2 presented in Fig. 7.3 imply for high P concentrations - in accordance with many other publications (see chapter 6) - ferromagnetic interlayer coupling, our detailed investigations reveal a more complex, time-dependent behaviour. Therefore, we have performed a systematic study of the superconducting and magnetic properties of a complete set of EuFe2(As1−xPx)2 (x = 0, 0.055, 0.09, 0.12, 0.16, 0.165, 0.17, 0.26, 0.35, 0.39, 1) single crystals (see chapter 5.3) using dc and ac magnetization (see chapter 5.2), dc resistivity and heat capacity mea- surements1. In order to account for the time-dependencies, the temperature is usually swept continuously in M(T ) cycles (see chapter 5.2); furthermore we distinguish between field-cooled curves that were obtained while decreasing (“field-cooled cooled”: FCC) and increasing (“field-cooled heat”: FCH) the temperature. Results are published in Ref. [10]2. EuFe2P2 Fig. 7.4 and Fig. 7.5 show the in-plane (H ‖ ab) magnetic behaviour of a EuFe2P2 single crystal. In a very small probing field of 2G two consecutive magnetic transitions can be clearly resolved in the zero-field cooled and field-cooled dc magnetic susceptibility. A sharp peak at T1 ∼ 29.0K is followed by an upturn starting at T2 ≈ 27.7K, leading to a second peak at ∼ 27K in the ZFC magnetization. Whereas the transition at T1 exhibits no thermal hysteresis, the second transition at T2 is characterized by a pronounced ZFC-FCC hysteresis which vanishes if a larger dc magnetic field (H ‖ ab) is applied (see Fig. 7.5). It can be finally suppressed for fields around 500G. Increasing the field even higher shifts the peak at T1 down (see Fig. 7.4c) and at around 1T, the peak has completely disappeared. Note that a broad shoulder develops at ∼ 0.3T, which was interpreted for EuFe2As2 as due to a metamagnetic transition [182]. Furthermore, specific heat measurements on the single 1 Magnetization data are complemented with four point dc resistivity measurements with a current of the typical value 1mA. In the case of small resistances (especially for the c-direction measurements), lock- in technique with low frequencies (77Hz) was used. Transition temperatures derived from resistivity measurements are indicated with “∗”. Heat capacities were measured with a relaxation-type calorimeter. 2 Preliminary studies can be also found in Ref. [296]. 7.2 Re-entrant spin glass 101 crystal show a sharp peak at T1 with a broad shoulder at lower temperatures, consistent with the width of the anomaly for polycrystalline samples reported in Ref. [211] and prove that both transitions are bulk properties (see Fig. 7.4b). Figure 7.4: Magnetization curves of EuFe2P2 (H ‖ ab), together with specific heat data. (a) Temperature-dependent FCC curves χdc(T ) = M(T )/µ0H at small fields between 2G (orange) and 100G (purple) show two consecutive magnetic transitions at T1 = 29.0K and T2 = 27.7K; the transition at lower temperature T2 is quickly suppressed in the external field. (b) Temperature-dependent ZFC (black open squares) and FCC (orange filled circles) dc susceptibility χdc(T ) (upper panels) measured with H = 2G, revealing that a strong hysteresis sets in at T < T2. The lower panel adds the temperature-dependent heat capacity Cp for our single crystal (purple filled diamonds) and for polycrystalline EuFe2P2 (open dark blue stars, [234]). The single crystal heat capacity shows a sharp feature at T1 that is followed by a broad shoulder (visible also in the data of polycrystalline EuFe2P2). (c) ZFC magnetization for 500G (dark blue), 0.1T (blue), 0.3T (cyan), 0.5T (dark cyan) and 1T (green). T1 decreases with increasing external field, also depicted in the inset. 102 7 Magnetization studies on EuFe2(As1−xPx)2 Figure 7.5: (a) ZFC (black open squares) and FCC (orange filled circles) in-plane magneti- zation curves of EuFe2P2 at H = 2G, 5G, 20G, 100G, 500G and 0.1T (H ‖ ab), as well as (b) ZFC (black open squares) and FCC (brown filled circles) out-of-plane magnetization at H = 2G. The in-plane low-temperature transition and any irreversibility has disappeared at 500G; out-of-plane, a small hysteresis due to spin canting can be identified at low fields. In order to get more insight into the character of the second transition at T2 we studied in detail its thermal hysteresis and frequency-dependence by ac magnetization measurements. Fig. 7.6a displays the differences between the ZFC, FCC and FCH susceptibilities for an in-plane field of 2G. The thermal hysteresis is visible in the ZFC-FCC splitting at T < T2. Repeated FCC and FCH cycles (heating / cooling rates 0.2 K/min) revealed a very slow time-dependence of the magnetization below T2, leading to a growth of the magnetization and consequently a negative difference of FCH - FCC. A time-dependence of the magnetization below T2 is also visible in a time-dependence of the ZFC magnetization which after some rapid initial increase grows almost linearly in time with a rate of ∼1%/h (see Fig. 7.6b). The time-dependence of the magnetization becomes also apparent in a frequency-dependence of the real and imaginary component of the ac susceptibility, χ′ac(T ) and χ′′ac(T ), as depicted in Fig. 7.6c. Below T2 a peak appears in both components (χ′ac ≈ 50×χ′′ac) which shifts to higher temperature with increasing frequency following a Vogel-Fulcher behaviour (see Fig. 7.6d). We can rule out any relation of the time and frequency-dependence to flux line lattice dynamics, since EuFe2P2 is far off from any proposed superconducting phase [230, 236]. 7.2 Re-entrant spin glass 103 Figure 7.6: Magnetization curves of EuFe2P2 (H ‖ ab). (a) Magnetic hysteresis sets in at T < T2, visible in χdc,FCC −χdc,ZFC > 0 (green stars) (µ0H = 2G); ZFC and FCC curves are depicted in the upper panel for a direct comparison of the characteristic temperatures T1 and T2. Time-dependence for T < T2 is revealed by FC cycling, visible in χdc,FCH − χdc,FCC < 0 (pink open diamonds) and (b) time-dependent magnetization after ZFC cooling (µ0H = 2G, 25K: blue closed squares, 28.5K: pink open circles). (c) A frequency-dependence in the ac susceptibility χ′ac(T ) and χ′′ac(T ) (no dc field, ac drive amplitude 1G, 0.1Hz: dark grey open triangles, 100Hz: brown closed circles) sets in also below T2. (d) Vogel-Fulcher-Fit of the peak below T2 in χ′′ac(T ). Until now, magnetic ordering at higher P concentrations in polycrystalline and single crystalline samples of EuFe2(As1−xPx)2 was assigned to ferromagnetism [230, 234, 236]. Measurements on polycrystalline samples [234, 236], however, are not able to allow conclu- sions about a possible antiferromagnetic interlayer coupling, and measurements on single crystals of EuFe2(As1−xPx)2 [230] failed to reveal two separate magnetic transitions and their different thermal hysteretic behaviour because of too coarse temperature steps. In view of the shape of the M(T ) anomaly at T1 (see Fig. 7.4c) we suggest that the Eu2+ moments in EuFe2P2 order in a canted A-type antiferromagnetic structure with 104 7 Magnetization studies on EuFe2(As1−xPx)2 the spin components being ferromagnetically aligned along the c-axis1. Additionally, we detect below T2 a second phase transition with glassy character which we associate to the ordering of the in-plane components of the Eu2+ moments. The development of a glassy phase below a magnetic phase transition, commonly referred to as re-entrant spin glass [38, 297–299], indicates a competition between antiferromagnetic and ferromagnetic spin exchange interactions in the system. In the case of EuFe2P2, the antiferromagnetic RKKY interlayer coupling competes with the ferromagnetic intralayer interactions of the spins. In fact, DFT based calculations re- vealed a very small energy difference of antiferromagnetic and ferromagnetic ground states for EuFe2(As1−xPx)2 [230]. We therefore suggest that in EuFe2P2, competition between ferro- and antiferromagnetism causes glassy freezing of spin components in the ab-plane at T2 and a decoupling of the magnetic Eu layers. Such a freezing of transverse magnetic components following long range magnetic order (which has already set in) is consistent with mean field theoretical calculations for a re-entrant spin glass [39]. Our conclusions are not only supported by the time-dependent magnetization behavior at T < T2, but also by the development of M(T ) with external fields: by application of a magnetic field of ∼ 500G along the ab-plane, the energy barrier between different equilibrium states can be overcome, the glass transition is suppressed and the temperature-dependent magneti- zation resembles that of EuFe2As2. This interpretation is consistent with neutron studies by Ryan et al. [212], as those are not sensitive to the freezing of the small in-plane spin component, as long as the ferromagnetic spin component along the c-axis still exists. EuFe2(As0.835P0.165)2 In order to study the complex interplay of magnetism and superconductivity in mixed As - P samples we have investigated in detail the magnetic and superconducting properties of a single crystal of EuFe2(As0.835P0.165)2. The in-plane electrical resistivity (see Fig. 7.7) proves the onset of superconductivity at T ∗c,on ∼ 22K indicated by a steep initial decrease of the resistivity. Zooming into the transition reveals re-entrant behaviour at about 19K followed by a smooth decrease towards zero resistivity which is achieved only below T ∗c,0 ≈ 9K. Fig. 7.9 displays the corresponding zero-field and field-cooled cooled temperature-de- pendent magnetization. Both, in- and out-of-plane magnetization are very similar to EuFe2P2: for H ‖ ab, two peaks are visible that are shifted to slightly lower temperatures (T1 ∼ 19.0K, T2 ∼ 16.8K); for H ‖ c, a slight hysteresis below T1 indicates the spin 1 Due to sample shape effects we can not distinguish between A-type antiferromagnetism and helimag- netism. 7.2 Re-entrant spin glass 105 Figure 7.7: In-plane resistivity curves of superconducting EuFe2(As0.835P0.165)2 with T ∗c,on ∼ 22K. At T > T ∗c,on, R(T ) is linear; below ∼ 19K, a re-entrance of resistivity appears. Zero resistivity is reached only below T ∗c,0 ∼ 9K canting along the c-axis. Additionally, a pronounced downturn occurs below 15K for both field directions, which ends up in a diamagnetic signal for H ‖ c. While the in-plane low-temperature transition at T2 has already disappeared at 500G, irreversibilities due to superconductivity remain at least up to 0.5T. Further evidence for a spin glass transition can be found in Fig. 7.8, which compares the low-temperature resistivity behaviour to selected ac and dc magnetization data. A difference between ZFC and FCH curves sets in at T < Tc,on. However, FC cycling reveals two time-dependent glassy transitions, as χdc,FCH − χdc,FCC > 0 at T < Tc,on and χdc,FCH − χdc,FCC < 0 at T < T2 (pink open diamonds). We therefore conclude that superconducting EuFe2(As0.835P0.165)2 shows an analogous re-entrant spin glass behaviour as EuFe2P2. The additional positive peak in the FCH-FCC curve between 17K and 21K could be ascribed to vortex dynamics, as it coincides with the steep initial decrease of the resistivity marking the onset of superconductivity. Having this interpretation in mind, we can also interpret all the other features in the ac susceptibility: the resistivity re-entrance and T ∗c,0 are both accompanied by a sharp peak in χ′′ac(T ). Such a feature, which comes along with a dip in χ′ac (slightly visible for the 100G curve) was reported for several high-Tc SC and is known as “peak effect” [300, 301] arising due to a maximum in the critical current density. The broad peak in χ′′ac(T ) at T < T2 could be a sign of granular SC or of flux reaching the center of the sample according to Bean’s critical state model [302, 303]. However, as χdc,FCH − χdc,FCC < 0 exactly at T < T2, it probably arises due to the spin glass transition. Furthermore, one should note that in-plane ac susceptibility measurements with applied high dc fields H ‖ ab are able to reveal a diamagnetic in-plane shielding signal if the dc field is large enough (∼ 1T) to saturate the Eu2+ magnetism. 106 7 Magnetization studies on EuFe2(As1−xPx)2 Figure 7.8: (a) Details of the resistivity curves as well as (b) in-plane magnetization curves (H ‖ ab) of EuFe2(As0.835P0.165)2, in order to determine the transition temperatures. Dashed lines are guides to the eye, highlighting T ∗c,50 = Tc,on ∼ 20.8K (purple), T1 ∼ 19.5K (red), T2 ∼ 16.8K (blue) and T ∗c,0 ∼ 9.0K (purple). (b) In-plane ZFC (black open squares) and FCC (orange filled circles) magnetization shows atH = 2G two magnetic transitions similar to x = 1.0 with a steep drop in the ZFC curve below ∼ 15K (measured with the Quantum Design ultra low field option, see chapter 5.2). A difference between ZFC and FCH curves sets in at T < Tc,on; FC cycling reveals two time-dependent glassy transitions, as χdc,FCH−χdc,FCC > 0 at T < Tc,on and χdc,FCH − χdc,FCC < 0 at T < T2 (pink open diamonds). Combining ac susceptibility measurements (drive 1G, frequency 7Hz) with dc fields (0G: grey open triangles, 100G: violet closed circles, 0.3T: pink open diamonds, 1T: purple closed circles), negative susceptibility χ′ac < 0 is revealed for H = 1T below 15K. χ′′ac(T ) shows a non-zero- signal below Tc,on, a sharp peak at the maximum re-entrant resistivity (that is accompanied by a small dip in χ′ac(T )), a broad hump below T2 and again a sharp peak at Tc,0. Increasing the dc field from 0 to 100G only suppresses the broad hump below T2. 7.2 Re-entrant spin glass 107 Figure 7.9: (a) ZFC (black open squares) and FCC (orange filled circles) in-plane mag- netization curves of EuFe2(As0.835P0.165)2 at H = 2G, 100G, 500G, 0.3T, 0.5T and 1.0T (H ‖ ab), as well as (b) ZFC (black open squares) and FCC (brown filled circles) out-of-plane magnetization at H = 2G (measured with the Quantum Design ultra low field option, see chapter 5.2). The in-plane low-temperature transition at T2 (marked by a shoulder) has dis- appeared at 500G, however, irreversibilities due to superconductivity remain at least up to 0.5T; out-of-plane, a small hysteresis due to spin canting can be identified, as well as a steep drop in the ZFC curve below ∼ 15K, reaching negative magnetization. Figure 7.10: Field-dependent magnetization curves of EuFe2(As0.835P0.165)2 for H ‖ ab (left) and H ‖ c (right), at several temperatures between 2K (blue) and 25K (red). At 2K, a distinct hysteresis that remains up to highest applied fields (7T) is clearly visible (see also inset), indicating superconductivity. 108 7 Magnetization studies on EuFe2(As1−xPx)2 Finally, we have also measured the in- and out-of-plane field-dependent magnetization curves at various temperatures (see Fig. 7.10). Most remarkably, we find at 2K a distinct hysteresis that remains up to highest applied fields (7T), typical for a hard superconductor according to Bean’s critical state model (see chapter 3.3.2). EuFe2(As0.65P0.35)2 We have also investigated a single crystal of EuFe2(As0.65P0.35)2 in more detail, in order to address the open question about the width of the superconducting dome. Fig. 7.11a displays the corresponding in-plane resistivity, showing a sharp drop around T1 ∼ 20.7K, which fades in a broad shoulder at T2 ∼ 18.8K. Such a resistivity behaviour was interpreted by Cao et al. [236] as re-entrant superconductivity. However, in-plane ZFC magnetization measurements reveal two magnetic transitions similar to x = 1.0, indicating the onset of canted antiferromagnetism and re-entrant spin glass behaviour at T1 and T2, respectively (see Fig. 7.11 and Fig. 7.12). Indeed, any out- of-phase susceptibility signal is already suppressed at 100G, the same field, where the T2 transition gets barely visible in ZFC and FC measurements. Furthermore, no diamagnetic signal is visible in any dc or ac susceptibility measurement. One should note that the double-feature in M(T ) curves for superconducting samples was in some cases assigned to result from the Paramagnetic Meissner Effect [234]. However, this effect would only appear in the FC curve [304], whereas we observe features also in ZFC data. Furthermore, we have also measured the in- and out-of-plane field-dependent magne- tization curves at various temperatures (see Fig. 7.13). In contrast to superconducting EuFe2(As0.835P0.165)2 (see Fig. 7.10), we find at 2K no high-field hysteresis. Therefore, we conclude that EuFe2(As0.65P0.35)2 is not superconducting; it displays similar to EuFe2P2 canted antiferromagnetism and re-entrant spin glass behaviour. 7.2 Re-entrant spin glass 109 Figure 7.11: (a) Details of the resistivity curves as well as (b) in-plane magnetization curves (H ‖ ab) of EuFe2(As0.65P0.35)2, in order to determine the transition temperatures. Dashed lines are guides to the eye, highlighting T1 ∼ 20.7K (red) and T2 ∼ 18.8K (blue). (a) The resistivity approaches 22K rather linear. A sharp drop is visible around T1, which fades in a broad shoulder at ∼ T2. (b) The in-plane ZFC (black open squares) and FCC (orange filled circles) magnetization shows atH = 2G two magnetic transitions similar to x = 1.0 (measured with the Quantum Design ultra low field option, see chapter 5.2). At T < T2, a difference between ZFC and FCH curves sets in (green filled stars), and FC cycling reveals a glassy transition with χdc,FCH − χdc,FCC < 0 (pink open diamonds). Combining ac susceptibility measurements (drive 1G, frequency 7Hz) with dc fields (0G: grey open triangles, 100G: violet closed circles, 1T: purple open circles), suppresses in χ′ac(T ) already at 100G the T2 transition and at 1T also the T1 transition; however, there is no signature of negative susceptibility. χ′′ac(T ) shows a non-zero-signal below T2, which is completely suppressed at 100G. 110 7 Magnetization studies on EuFe2(As1−xPx)2 Figure 7.12: (a) ZFC (black open squares) and FCC (orange filled circles) in-plane magne- tization curves of EuFe2(As0.65P0.35)2 at H = 2G, 5G, 20G, 100G, 500G, 0.3T, 0.5T and 1.0T (H ‖ ab, 2G curves were measured with the Quantum Design ultra low field option, see chapter 5.2). The distinct in-plane low-temperature transition at T2 (marked by a shoulder) has almost vanished at 100G. Figure 7.13: Field-dependent magnetization curves of EuFe2(As0.65P0.35)2 for H ‖ ab (left) and H ‖ c (right), at several temperatures between 2K (blue) and 25K (red). No high-field hysteresis can be observed (see also inset). 7.3 Phase diagram 111 7.3 Phase diagram In order to follow the compositional dependence of the two magnetic transitions con- sistently found in EuFe2P2, EuFe2(As0.835P0.165)2, and EuFe2(As0.65P0.35)2, we have ex- tended our studies to EuFe2(As1−xPx)2 single crystals with x = 0, 0.055, 0.09, 0.12, 0.16, 0.17, 0.26, 0.35 and 0.39; corresponding in-plane dc magnetization curves measured at 2G are depicted in Fig. 7.14. In all P substituted specimen we observe two consecutive mag- netic transitions, which we ascribe, in analogy to the previous sections, to a canted A-type antiferromagnetic transition at T1 = TN and to a glassy freezing of the spin components in the ab-plane at T2 = Tglass, with TN > Tglass. Figure 7.14: ZFC (black open squares), FCC (brown filled circles) and FCH (orange filled circles) in-plane magnetization curves of EuFe2(As1−xPx)2 with x = 0.09, 0.16, 0.17, 0.26 and 0.39 at H = 2G (H ‖ ab). Arrows are guides to the eye, highlighting Tc,on (purple), T1 (red) and T2 (blue). In Fig. 7.15 we have compiled the resulting magnetic phase diagram together with the superconducting dome. According to our investigations on single crystals the re-entrant spin glass transition appears for all P substituted specimen. The effect of chemical disorder of the As and P anions on the RKKY-exchange must be ruled out as the origin of the glass transition since it occurs also in well-ordered EuFe2P2 crystals. We rather ascribe the glass transition to a competition of ferromagnetic interactions within a layer with antiferromagnetic RKKY-interactions between neighbouring layers. 112 7 Magnetization studies on EuFe2(As1−xPx)2 Figure 7.15: Phase Diagram of EuFe2(As1−xPx)2. TN = T1 (red dots) indicates a canted A-type antiferromagnetic transition, with a ferromagnetic net component of the Eu2+ spins along the c-direction, Tglass = T2 (blue triangles) a spin glass transition due to the freezing of the spins in the ab-plane and Tc,on the onset of superconductivity (grey squares). Closed sym- bols indicate transition temperatures deduced from magnetization, open ones from resistivity measurements [305]. Shadowed lines are guides to the eye. The light grey area indicates the onset of superconductivity, while bulk superconductivity is fully developed in the dark grey regime [230, 238]. The transition temperatures exhibit a nonmonotonic behaviour with P substitution. At low P concentration, 0 < x . 0.12, the antiferromagnetic Eu2+ transition temperature follows the transition temperature of the spin density wave. Coupling of the itinerant Fe magnetism and the Eu2+ local spin moments was theoretically predicted and experimen- tally confirmed by the increasing canting of spins out of the ab-plane concomitant with the suppression of the spin density wave (see section 7.1). With increasing canting, the ferromagnetic component of the Eu2+ spins along the c-direction increases, and the compe- tition with the antiferromagnetic RKKY-interaction between the layers is enhanced which finally leads to the development of the spin glass phase. In the superconducting regime, the transition temperatures vary only slightly with P concentration. When superconduc- 7.4 Outlook 113 tivity is finally suppressed, both transition temperatures, T1 and T2, increase markedly, probably due to a Lifshitz transition [235, 238, 280] which affects the RKKY-exchange. Antiferromagnetic interlayer coupling developing up to high P concentrations, as well as a rather narrow superconducting dome, are consistent with experiments on EuFe2As2 under pressure [222]. Between the concentrations x ≈ 0.12 and x ≈ 0.26 the onset of a superconducting transition is found, while fully developed bulk superconductivity occurs in an even narrower regime [230, 238]. As concluded previously, a Lifshitz transition near x ≈ 0.23 coincides with the upper limit of superconductivity [235, 238, 280]. Investigations of polycrystalline samples, however, resulted in a somewhat broader dome extending to an upper limit of x ≈ 0.4 [234, 236]. The assignment of the upper limit was based on the assumption that two subsequent resistivity anomalies seen in samples with x ≈ 0.4 indicate the onset of superconductivity succeeded by a resistivity re-entrance due to Eu2+ magnetic ordering. Our experiments on single crystals rather indicate that these two anomalies are purely of magnetic origin as we do not see any signature of superconductivity in our x=0.35 crystal. 7.4 Outlook The main results of this chapter were already published in 2011 [9] and 2013 [10]. As the research on Eu based iron pnictides was pretty active since then, this section summarizes the response of the community on our results, as well as general conclusions we can draw for Eu based iron pnictides. Eu magnetism The idea of a ferromagnetic component along the c-axis due to the Eu2+ spin canting was directly accepted by the community for EuFe2(As1−xPx)2 [236], as well as evidenced for Co and Ir doped compounds [33, 253, 254, 257, 263]. We want to note, however, that we would be very careful to interpret the different absolute values of ab-plane and c-axis magnetization as being connected to this canting [253]: demagnetization effects (see chap- ter 2.1) play a very important role for the rare earth systems - even in the paramagnetic phase close to the magnetic transition(s) - as can be seen in Fig. 7.16 for the parent com- pound1. When recalculating the susceptibility by considering a demagnetization factor close to that of an infinite flat plate (Nc → 1; the sample thickness is two orders of mag- 1 The volume of the sample was calculated by its mass, molar mass, and the lattice constants published in Ref. [230]. 114 7 Magnetization studies on EuFe2(As1−xPx)2 nitude smaller than the in-plane dimensions), one can basically suppress and even reverse the anisotropy. Figure 7.16: (a) Temperature-dependent volume (V ) susceptibility χdc(T ) = M(T )/µ0H/V of EuFe2As2 for H ‖ ab (orange) and H ‖ c (brown) at H = 20G: (a) as measured; (b) cor- rected with the demagnetization factor Nc = 1, Nab = 0; (c) corrected with Nc = 0.65, Nab = 0.175. Demagnetization effects are strong enough to reduce or even reverse the anisotropy of χdc(T ). Furthermore, we want to stress that antiferromagnetic interlayer coupling developing up to high P concentrations is consistent with experiments on EuFe2As2 under pressure [222]. Recent neutron measurements suggest for superconducting Co as well as P substi- tuted compounds that the Eu2+ moments are basically aligned along the c-axis; a second transition was either not detected [241, 259] or attributed to a minority phase [241]. Im- portantly, this is not contradictory to our results of the re-entrant spin glass phase: first of all, the neutron measurements can not exclude a small spin component staying in the ab-plane; secondly, we do not expect any typical antiferromagnetic signal in the re-entrant spin glass phase. For the future, we suggest neutron studies under an in-plane magnetic field that suppresses the spin glass phase; in this case, possibly antiferromagnetic reflec- tions could be detected. Finally, we want to note that successive magnetic orderings can be identified also in the case of Ru substitution [31], Co doping [250, 251, 256, 257] and Ir doping [263]. This implies the appearance of a re-entrant spin glass phase in all those compounds. Therefore, we conclude that our magnetic phase diagram derived for EuFe2(As1−xPx)2 (see Fig. 7.15) is characteristic for (non-diluted) Eu based iron pnictides. In order to confirm this, we suggest time-dependent measurements also on other compounds. 7.4 Outlook 115 Interplay of superconductivity and magnetism The question how bulk superconductivity can coexist with Eu2+ magnetic ordering is quite fundamental and requires the exact knowledge of the magnetic structure. Consistent with Mössbauer and neutron powder diffraction [212, 234, 241], the results of our experiments imply that in superconducting samples, a large net component of the Eu2+ spins is fer- romagnetically aligned perpendicular to the layers. In addition, we find that glass-like dynamics and freezing of the in-plane component develops below T2 which destroys coher- ence between the Eu layers. Superconductivity in the iron pnictides is commonly believed to take place mainly in the FeAs layers. In this scenario, the inner field resulting from the Eu2+ ferromagnetic component along the c-axis could be screened by the formation of spontaneous vortices perpendicular to the layers [31]. Together with the destroyed coher- ence between the Eu layers due to the glass dynamics, this scenario might be the key to understand how superconductivity can coexist with the usually strong Eu2+ magnetism. 8 OPTICAL PROPERTIES OF EuFe2As2 In this chapter, the in-plane optical response of EuFe2As2 is analysed. Particular focus is set on the applicability of different modelling approaches, and the determination of model-independent trends. 8.1 Overview The temperature-dependent in-plane optical reflectivity of EuFe2As2 was measured in a wide frequency range from 40 to 10000 cm−1 (see chapter 5.1), using an infrared Fourier transform spectrometer; the corresponding optical conductivity was calculated via the Kramer’s Kronig analysis. For the low-frequency extrapolation, in a first iteration the Hagen Rubens relation was used, which got subsequently replaced in the fitting process; the thus determined temperature-dependence of σdc follows nicely the resistivity curve (as will be shown below). At high frequencies, the optical response was extended up to 30000 cm−1 by room temperature measurements with a Woollam spectroscopic ellipsome- ter1; higher frequencies were extrapolated following measurements up to 32 eV carried out for BaFe2As2 [169]. 1 Ellipsometry measurements were performed by J. Braun at the 1. Physical Institute, Stuttgart. 118 8 Optical properties of EuFe2As2 Fig. 8.1a,b displays the thus obtained spectra at 200K and 290K, focusing on the high frequency range. There is considerable spectral weight transfer up to very high energies, which probably can be ascribed to Hund’s coupling (see chapter 4.4.5); the total spectral weight is not recovered in the infrared frequency range. Figure 8.1: (a-b) Frequency-dependent in-plane reflectivity and conductivity of EuFe2As2 at T =200K (purple) and 290K (red), as well as (c-d) spectral weight (SW) analysis. While (c) displays the frequency-dependent spectral weight for 15K (blue), 30K (light blue), 180K (dark blue), 200K (purple) and 220K (violet) normalized to 290K, (d) shows the total spectral weight at 10000 cm−1; the linear fit (green dashed line) acts as a guide to the eye. With decreasing temperature, spectral weight is shifted to high energies (above ∼ 6000 cm−1); the total spectral weight is not recovered at 10000 cm−1. At T < Ts,SDW = 190K, a gap develops in the optical spectrum, and additional spec- tral weight transfer appears at low energies (see Fig. 8.1f and Fig. 8.2). We find only one clearly visible gap, leading at 15K to a peak in the optical conductivity around 1400 cm−1 ≈ 10.8 kBTs,SDW. Its asymmetric shape is typical for a spin density wave gap (see chapter 4.3), but could indicate also the existence of two neighbouring features (as theoretically proposed by Yin et al. [168]). One should further note the FeAs phonon at around 260 cm−1, which gains at T < Ts,SDW considerably oscillator strength and shifts to higher frequencies. Our finding of a single spin density wave gap is consistent with Ref. [184], however, we are also aware that an additional smaller gap around the BCS weak coupling limit was 8.1 Overview 119 reported by Moon et al. [146]. As we investigated a sample of the same batch as Ref. [184], one can speculate that the occurrence of the low-energy feature might strongly depend on the sample quality1, but we are not aware of any similar reports for other compounds. Therefore, it is intriguing to relate this untypical behaviour to the influence of the Eu2+ magnetism. Unfortunately, while there are predictions for the influence of the spin density wave on the Eu2+ magnetism [11], the reverse case is - to our best knowledge - not yet considered theoretically. Furthermore, one should note that the spectra at 15K and 30K are basically the same within the experimental error, meaning that Eu2+ spin scattering seems to have a negligible influence on the optical properties in the infrared frequency range. Figure 8.2: Frequency-dependent in-plane reflectivity and conductivity of EuFe2As2 at T =15K (blue), 30K (light cyan), 180K (dark blue) and 290K (red). At T < Ts,SDW, a gap de- velops in the optical spectrum and spectral weight is shifted to ∼ 1150 cm−1 and ∼ 1400 cm−1 at 180K and 15K, respectively; furthermore, the FeAs phonon at around 260 cm−1 gains con- siderably oscillator strength. One should note that the spectra at 15K and 30K are the same within the resolution. The noise visible in the conductivity spectrum at low frequencies is typical for a reflectivity close to unity; data are not smoothed. 1 While our sample has a residual resistivity ratio R(300K)/R(5K) of ∼ 4.1, the one of Moon et al. has only a ratio of ∼ 2.4. 120 8 Optical properties of EuFe2As2 8.2 Multi-component analysis As already discussed in chapter 4.4.2, the multiband character of iron pnictides complicates any analysis of optical spectra. Different multi-component models exist, however, it is still not clear which one should be favoured and how much the modelling influences the outcome of the analysis. Therefore, we have analyzed the optical response of twinned EuFe2As2, using three different modeling approaches and discuss the commonly found features. Importantly, we always show (semi-)logarithmic and linear plots, emphasizing the low- and high-energy range, respectively. Broad Drude approach In 2010, Wu et al. proposed that the spectra of iron pnictides contain a temperature- independent background in the mid-infrared energy range that can be fitted by a very broad Drude term [128, 130]. However, a closer inspection reveals that such a temperature- independent approximation is too rough due to the considerable spectral weight transfer up to very high energies (see Fig. 8.1). In order to investigate, whether such a broad Drude approach with temperature-depen- dent parameters can be used to nicely describe the infrared response of EuFe2As2, Fig. 8.3 and Fig. 8.4 display the thus fitted spectra at 290K and 15K, respectively. At 290K, the overall fit quality is good; however, one can note a slight discrepancy in the far-infrared, where the fit first over- and then underestimates the response. Indeed, this is exactly the energy range, where the model fails at low temperatures: at 15K, an upturn of the conductivity above ∼ 350 cm−1 implies an additional, broad Lorentzian at frequencies above 1000 cm−1. Fig. 8.4 excludes that a broad Lorentzian around 3.4kBTs,SDW (the smaller gap found by Moon et al. [146]) can describe this feature. Therefore, we conclude that the broad Drude approach is too oversimplified in order to model nicely the far-infrared response of iron pnictides. 8.2 Multi-component analysis 121 Figure 8.3: Frequency-dependent in-plane (a-b) reflectivity and (c-d) conductivity of EuFe2As2 at T =290K in different logarithmic and linear plot styles. Grey squares rep- resent data points, red lines the result of a broad Drude fit; the composition of the single components is displayed in (c). The overall fit quality of the broad Drude approach is good; slight discrepancies can be found in the far-infrared. 122 8 Optical properties of EuFe2As2 Figure 8.4: Frequency-dependent in-plane (a-b) reflectivity and (c-d) conductivity of EuFe2As2 at T =15K in different logarithmic and linear plot styles. Grey squares represent data points, red lines the result of a broad Drude fit; the composition of the single components is displayed in (c). A spin density wave gap leads to the piling up of an asymmetric peak at ∼ 1400 cm−1 ≈ 10.8kBTs,SDW. The broad Drude approach fails to fit the spectra between ∼ 350 cm−1 and 1000 cm−1, even if we assume that the broader Drude is gapped with a gap value of 3.4kBTs,SDW. 8.2 Multi-component analysis 123 MIR band approach As already noted by Tu et al. [142], a multi-component analysis with only one Drude and a series of Lorentzians that model incoherent transport in the mid-infrared is able to describe the carrier dynamics of iron pnictides. Fig. 8.5 and Fig. 8.6 prove that such a fitting approach is able to nicely describe spectra at high and low temperatures, respectively. Figure 8.5: Frequency-dependent in-plane (a-b) reflectivity and (c-d) conductivity of EuFe2As2 at T =290K in different logarithmic and linear plot styles. Grey squares represent data points, red lines the result of a MIR band fit; the composition of the single components is displayed in (c). The MIR band approach nicely describes the spectra above Ts,SDW. The corresponding temperature-dependent fit parameters are displayed in Fig. 8.7. Typ- ically for a metal, the Drude component narrows smoothly with decreasing temperature, until the scattering rate rapidly decreases for T < Ts,SDW (see Fig. 8.7a). One should note that the obtained σdc values perfectly follow the temperature-dependence of the resistivity. 124 8 Optical properties of EuFe2As2 Figure 8.6: Frequency-dependent in-plane (a-b) reflectivity and (c-d) conductivity of EuFe2As2 at T =15K in different logarithmic and linear plot styles. Grey squares represent data points, red lines the result of a MIR band fit; the composition of the single components is displayed in (c). The MIR band approach nicely describes the spectra below Ts,SDW. However, the increase of σdc with lowering temperature does not outweigh the decrease of the scattering rate: Drude spectral weight is already lost above the spin density wave transition. As can be seen in Fig. 8.7c, at T > Ts,SDW, the spectral weight is transferred to the hump-like structure around 6000 cm−1 (and even higher energies), which shows a significant blueshift with decreasing temperature; any spectral weight transfer is stopped below the spin density wave transition. Interestingly, the width of this feature is constant above Ts,SDW, but decreases rapidly at T < Ts,SDW - directly reflecting the behaviour of the Drude component. Those observations indicate a strong relationship to the dynamics 8.2 Multi-component analysis 125 of the itinerant carriers, consistent with the suggestion that the bump originates from Hund’s coupling. Figure 8.7: Temperature-dependent fit parameters for the MIR band approach: (a) Scatter- ing rate Γ , σdc compared to results from resistivity measurements (normalized to σdc(300K), cyan curve), and spectral weight (SW) of the Drude component; (b) Center frequency ν, damping Γ and spectral weight (SW) of the feature that is attributed to Hund’s coupling; (c) spectral weight of the temperature-dependent components. The grey, vertical dashed line denotes Ts,SDW; other dashed lines are a guide to the eye. Fig. 8.7d also displays the spectral weight of the other temperature-dependent fit compo- nents. Their total spectral weight decreases with decreasing temperature, as was already discussed in the general spectral weight analysis (see Fig. 8.1). As the MIR band fol- lows the behaviour of the Drude component, it probably contains a strong contribution from itinerant carriers; indeed, the spectral weight gapped in the spin density wave phase mostly corresponds to changes in the mid-infrared frequency range. 126 8 Optical properties of EuFe2As2 Subtraction approach As in general, a model that includes two Drude components as well as a mid-infrared band is overdetermined, one needs additional information for a reasonable fitting approach. Here, we focus on the result of recent infrared studies: by subtracting low temperature conductivity spectra from that at room temperature, Marsik et al. found for Ba, Sr and Ca compounds a well-defined feature around 1000 cm−1 that could be attributed to interband transitions involving Fe hole pockets [12]. Figure 8.8: Frequency-dependent in-plane (a-b) reflectivity and (c-d) conductivity of EuFe2As2 at T =290K in different logarithmic and linear plot styles, as well as (e) the conductivity difference ∆σ1 = σ1(200K) − σ1(290K). Grey squares represent data points, red lines the result of a fit with two Drude components and a Lorentzian at 1000 cm−1 that can be attributed to interband transitions involving Fe hole pockets; the composition of the single components is displayed in (c). The fitting approach nicely describes the spectra above Ts,SDW. 8.2 Multi-component analysis 127 We believe that such a subtraction feature can be nicely described by a Lorentzian at 1000 cm−1 that broadens with increasing temperature. Indeed, EuFe2As2 shows the same feature (see Fig. 8.8e). Therefore, we have fitted the spectra with two Drude components and one Lorentzian at 1000 cm−1 that broadens with increasing temperature; the other fit parameters of the Lorentzian are held constant in first approximation. As can be seen in Fig. 8.8 and Fig. 8.9, the spectra can be nicely described by this approach. Figure 8.9: Frequency-dependent in-plane (a-b) reflectivity and (c-d) conductivity of EuFe2As2 at T =15K in different logarithmic and linear plot styles. Grey squares repre- sent data points, red lines the result of a fit with two Drude components and a Lorentzian at 1000 cm−1 that can be attributed to interband transitions involving Fe hole pockets; the composition of the single components is displayed in (c). The fitting approach nicely describes the spectra below Ts,SDW. However, one should note that other fitting variants are possible, too; for instance, we have also fitted the high-temperature spectra under the condition that the broader 128 8 Optical properties of EuFe2As2 Drude should not get too broad (not to violate the Mott-Ioffe-Regel limit), allowing a changing spectral weight of the Lorentzian at 1000 cm−1. The corresponding temperature- dependent fit parameters of both fitting approaches are displayed in Fig. 8.10. The results for the narrow Drude and the bump around 6000 cm−1 are independent from the fitting approach. However, we can not disentangle the parameters for the broad Drude and the Lorentzian at 1000 cm−1; a realistic description probably lies in between or even contains more temperature-dependent Lorentzians (for example another one at 2300 cm−1 [12]). However, we want to point out that with this model, we only see gap-like features in the spectral weight of the Drude components, proving that this description is more realistic than a general MIR band approach with only one Drude. Figure 8.10: Temperature-dependent fit parameters for the model with two Drude compo- nents and a Lorentzian at 1000 cm−1 under two different constraints: in the one case (filled squares), only the damping of the Lorentzian was changed; in the other case (open points), the scattering rate of the broad Drude was forced significantly below 1000 cm−1. (a) Scatter- ing rate Γ of the narrow Drude; (b) scattering rate Γ of the broad Drude; (c) σdc (sum of both Drude components) compared to results from resistivity measurements (normalized to σdc(300K), cyan curve); (d) Center frequency ν and damping Γ of the feature that is attributed to Hund’s coupling; (e) damping Γ of the Lorentzian at 1000 cm−1; (f) spectral weight of the temperature-dependent components. The grey, vertical dashed line denotes Ts,SDW; other dashed lines are a guide to the eye. 8.2 Multi-component analysis 129 Summary We have shown exemplarily on EuFe2As2 that a broad Drude model without any additional interband transitions in the mid-infrared is too oversimplified to describe the temperature- dependent spectra of iron pnictides. In contrast, a more general MIR band approach yields better results. The most realistic results are obtained with a model that contains two Drude components (probably representing electron and hole bands) and a Lorentzian at 1000 cm−1, which can be attributed to interband transitions involving Fe hole pockets [12]. In particular, we find that some fitting results are independent from the model: • The spin density wave gap affects the spectral weight and the scattering rate of both Drude components. • For temperatures above Ts,SDW, spectral weight is transferred from the narrow Drude, but also from higher frequencies in the mid-infrared (and thus either from the broad Drude or the interband transition at 1000 cm−1), to the bump-like structure around 6000 cm−1 and even higher energies. At T < Ts,SDW, this spectral weight transfer stops and the bump considerably narrows - directly reflecting the behaviour of the Drude component(s). Those observations indicate a strong relationship to the dynamics of the itinerant carriers, consistent with the suggestion that the bump originates from Hund’s coupling. 9 MAGNETIC DETWINNING OF EuFe2As2 As it was known that EuFe2As2 can be completely detwinned by applying relatively small magnetic fields (∼ 1T) at temperatures T < TEu (see chapter 6.3.1), we performed a series of magneto-optical measurements in order to study the electronic in-plane anisotropy of iron pnictides without the need of complex mechanical detwinning devices. In this chapter, we concentrate on the results from magneto-optical FIR measurements, which revealed a much more complicated magnetic detwinning mechanism than expected from previous neutron studies [7]. Combined with magnetization measurements, the observations can be explained by an indirect magnetoelastic coupling of the Eu2+ spins to the lattice. Those findings are further confirmed by resistivity, magneto-resistance, thermal expansion and magnetostriction measurements and are published in Ref. [13]. Finally, preliminary results on infrared measurements over a broad frequency range on persistently detwinned EuFe2As2 are presented. For all measurements, a well-defined cooling procedure was obeyed: first, the sample was cooled from T > Ts,SDW to low temperatures in zero magnetic field and the zero- field cooled (ZFC) response was measured. Afterwards, when an in-plane magnetic field 132 9 Magnetic detwinning of EuFe2As2 parallel to the [110]T-direction1 was first applied and then removed, we call this “field treatment” (FT). 9.1 Magneto-optical measurements The far-infrared spectral range is ideal to investigate the detwinning mechanism, as it is governed by the anisotropic spin density wave gap as well as the FeAs phonon, and the zero-field reflectivity is not notably influenced by the Eu2+ spin scattering (see chapter 8). Detwinning at T < TEu Fig. 9.1 displays the FIR reflectivity of EuFe2As2 at T = 15K for H ‖ [110]T with H = 0T (ZFC), 1.0T, and 0T after field treatment (with 1T). The main observations can be summarized as: • The zero-field cooled 0T curves perfectly merge, indicating an initially balanced twin domain ratio2. • At a magnetic field of 1T, a notable difference is induced between the two polar- izations: while the reflectivity is enhanced for E ‖ H, it is suppressed for the per- pendicular polarization. At the same time, the FeAs phonon gets strongly reduced for E ‖ H and enhanced for the perpendicular direction. In accordance with optical studies on BaFe2As2 using mechanical clamps (see chapter 4.4.5), those observations indicate that the crystal is detwinned with the orthorhombic a-axis being aligned along the external field: a ‖ H. This is consistent with the results from neutron studies [7]. • Surprisingly, we also observe a notable difference between the two polarizations after the magnetic field is switched off, opposite to that at 1T. This indicates a second detwinning process with b ‖ H at even lower magnetic fields, which stays persistent after the field is removed. In order to investigate this two-step detwinning process in more detail, we have per- formed a series of magneto-optical FIR measurements with small field steps of ≤ 0.1T. The results for E ⊥ H are shown in Fig. 9.2 and 9.3; for clarity, the relative reflectivity 1 We adapt the notation of the tetragonal system [hkl]T which is rotated by 45 ◦ along the c-axis with respect to the orthorhombic system [hkl]O, e.g. [110]T = [100]O 2 The sample was mounted point-like on a copper cone. This avoids not only back-reflections from the sample holder, but also minimizes the glued area and therefore external stress on the sample. 9.1 Magneto-optical measurements 133 Figure 9.1: Frequency-dependent reflectivity of EuFe2As2 at T =15K, for H = 0T (ZFC), 1T and 0T after FT (H ‖ [110]T); while the upper panel displays the whole FIR range, the lower panel focuses on the FeAs phonon mode at ∼ 260 cm−1. The magnetic field induces anisotropy between R(E ‖ H) (red) and R(E ⊥ H) (black) as well as changes the FeAs phonon mode, both corresponding to the magnetic detwinning. A distinct anisotropy remains after the magnetic field is switched off; it is opposite to that at 1T. is displayed, i.e. the reflectivity normalized to its initial zero-field value. While Fig. 9.2 is a false-colour plot of the field and frequency-dependent relative reflectivity, revealing on the first view the main dynamics, Fig. 9.3 resolves the details in a conventional plot. It can be concluded: • For increasing magnetic fields, the reflectivity rises rapidly at around 0.1T for fre- quencies1 above ∼ 300 cm−1, afterwards staying almost constant. Thus, H1 ∼ 0.1T can be identified as the critical field where twin domains preferentially align with b ‖ H. A sharp drop in the reflectivity at H2 ∼ 0.65T marks the second, opposite detwinning process with a ‖ H. For higher fields, the reflectivity stays again almost constant. • For decreasing fields, the latter process is reversible with a slightly lower critical field ∼ 0.60T. However, the detwinning at low fields is persistent. 1 The behaviour at lower frequencies is more complex, probably as Eu2+ spin scattering plays here a more important role. 134 9 Magnetic detwinning of EuFe2As2 • One should note the distinct changes of the FeAs phonon, which is strongly reduced at low and enhanced at high fields (see Fig. 9.3b). Those strong phonon changes (which follow the intensity changes of the reflectivity) support that a structural detwinning causes the electronic anisotropy. Figure 9.2: False colour plot of the frequency- and field-dependent FIR relative reflectivity R(H)/R(0T) of EuFe2As2 at T =15K (H ‖ [110]T, E ⊥ H). The detwinning fields (where the reflectivity abruptly de- and increases) can be identified as H1 ∼ 0.1T and H2 ∼ 0.65T for increasing, and H2 ∼ 0.6T for decreasing fields. 9.1 Magneto-optical measurements 135 Figure 9.3: Frequency-dependent relative reflectivity (normalized to the ZFC reflectivity) of EuFe2As2 at T =15K, for H = 0T (ZFC), 1T and 0T after FT (H ‖ [110]T, E ⊥ H); while the upper panel displays the whole FIR range, the lower panel focuses on the FeAs phonon mode at ∼ 260 cm−1. For increasing fields (left panels), the reflectivity first decreases and then increases abruptly at ∼ 0.1T and ∼ 0.65T, respectively. For decreasing fields only the upper process takes place at ∼ 0.60T; the lower process is irreversible. The low-field detwinning is accompanied by a reduction of the FeAs phonon oscillator strength; the high-field process by an enhancement. Grey dashed arrows mark the direction of the spectral changes with sweeping the magnetic field. 136 9 Magnetic detwinning of EuFe2As2 Detwinning at T > TEu While the neutron studies concentrated on temperatures T < TEu, we have also investi- gated whether any magnetic detwinning can be observed above TEu. Fig. 9.4 displays the FIR reflectivity of EuFe2As2 at T = 30K for H ‖ [110]T with H = 0T (ZFC), 1.0T, and 0T after field treatment (with 1T): • Similar to the measurements at 15K, the zero-field cooled 0T curves perfectly merge, proving an initially balanced twin domain ratio as well as the high quality of the measurements. • At a magnetic field of 1T, a notable difference is induced between the two polar- izations, similar to that at 15K. This indicates that the crystal gets sufficiently detwinned also at temperatures T > TEu by small magnetic fields with a ‖ H. • Again, we also observe a notable difference between the two polarizations after the magnetic field is switched off. However, now the direction of the anisotropy is the same as at 1T, indicating only one (persistent) detwinning process. Figure 9.4: Frequency-dependent FIR reflectivity of EuFe2As2 at T =30K, for H = 0T (ZFC), 1T and 0T after FT (H ‖ [110]T). The magnetic field induces anisotropy between R(E ‖ H) (red) and R(E ⊥ H) (black) as well as changes the FeAs phonon mode, both corresponding to the magnetic detwinning. A distinct anisotropy remains after the magnetic field is switched off. A detailed series of measurements reveals that the detwinning process at 30K is rather continuous, with the main changes arising between 0.3T and 0.7T (see Fig. 9.5). The induced detwinning shows an almost complete persistence. 9.1 Magneto-optical measurements 137 Figure 9.5: False colour (upper panel) and conventional plot (lower panel) of the frequency and field-dependent FIR relative reflectivity R(H)/R(0T) of EuFe2As2 at T =30K (H ‖ [110]T, E ⊥ H). The detwinning is rather continuous, with the main changes arising between 0.3T and 0.7T. A distinct anisotropy remains after the magnetic field is switched off. 138 9 Magnetic detwinning of EuFe2As2 Control measurements In order to verify that the induced anisotropy corresponds to the magnetic detwinning, measurements were repeated on EuFe2As2 with H ‖ [100]T, i.e. with the magnetic field at a 45◦ angle to the orthorhombic axes and the easy axes of the Eu2+ magnetic moments. In this case, no twin is favoured and thus, no detwinning should appear (limited by the accuracy of aligning the crystal perfectly along the [100]T-direction). Additionally, comparison measurements were performed on BaFe2As2 single crystals with H ‖ [110]T. As can be seen in Fig. 9.5, in both control experiments no anisotropy is induced within the limit of our resolution. This proves that the observed effects discussed above must be related (i) to a redistribution of twins and (ii) to the presence of Eu2+ moments. Figure 9.6: Frequency-dependent relative reflectivities of BaFe2As2 (T =15K) and EuFe2As2 (T =15K, 30K) for different magnetic field directions. Relative reflectivities of (a) BaFe2As2 with H ‖ [110]T = 0.5T (grey) and 1T (black) (T =15K, E ⊥ H) as well as of (b,c) EuFe2As2 with H ‖ [100]T = 1T ((b) T =15K and (c) 30K, E ‖ H: red, E ⊥ H: black) do not show magnetic field induced anisotropy. Heating persistence As our experiments reveal a persistent detwinning induced by small magnetic fields at low temperatures (T < Ts,SDW = 190K), a central question that arises is what happens when a persistently detwinned crystal is heated up. Fig. 9.7 displays a series of reflectivity measurements, where EuFe2As2 was subject of the following cooling and field treatment procedure: • Reflectivity spectra were taken after initial zero-field cooling at 30K and 15K, re- vealing an isotropic response (see Fig. 9.7a,b). • Field treatment at 15K (with 1T) induces a distinct anisotropy with R(E ‖ H) < R(E ⊥ H), corresponding to a persistent detwinning for temperatures T < TEu with 9.2 Magnetization measurements 139 b ‖ H and a more pronounced spin density wave gap along the b-direction (compare Fig. 9.7c with Fig. 8.2). • This anisotropy persists when heating up to 30K (see Fig. 9.7d). • Subsequent field treatment at 30K (with 1T) reverses the anisotropy, consistent with a persistent detwinning for temperatures T > TEu with a ‖ H. This proves that the persistent detwinning induced at temperatures T < TEu is robust against heating up above the Eu2+ magnetic ordering temperature. However, subsequent field treatment at T > TEu is able to inverse the anisotropy. Figure 9.7: Frequency-dependent reflectivity of EuFe2As2 after ZFC at (a) T = 30K and (b) 15K; the reflectivity is isotropic. (c) Field treatment at 15K (with 1T) induces a distinct anisotropy, (d) which persists when heating (in 0T) to 30K. (e) Subsequent field treatment at 30K (with 1T) reverses the anisotropy. Indeed, the thus induced anisotropy persists up to the structural and spin density wave transition at Ts,SDW, as demonstrated in Fig. 9.8. However, if the crystal was once heated above Ts,SDW and cooled down afterwards in zero field, this anisotropy has completely vanished (see Fig. 9.8d). 9.2 Magnetization measurements Since the magneto-optical measurements indicate that the Eu2+ spins drive the detwin- ning, magnetization measurements1 should provide further information about the complex detwinning mechanism. 1 For the magnetization measurements, the sample was fixed between two stripes of plastic foil by wrap- ping Teflon tape around them. This minimize the external stress on the sample. 140 9 Magnetic detwinning of EuFe2As2 Figure 9.8: Frequency-dependent reflectivity of EuFe2As2 after field treatment at T = 30K with 1T (H ‖ [110]T) for (a) 30K, (b) 100K and (c) 180K. Although the magnetic field is switched off during the measurement, the reflectivity along the orthorhombic axes is anisotropic. (d) After the crystal was heated to 220K (well above Ts,SDW) and cooled down again to 180K, any anisotropy disappeared. Detwinning at T < TEu Fig. 9.9a displays the field-dependent magnetization at 2K for H ‖ [110]T. In the first field cycling after zero-field cooling,M(H) exhibits two transitions for increasing, but only the upper one for decreasing field. The transition at higher fields can be identified as a spin-flip due to the abrupt increase of the magnetization, which saturates afterwards (see chapter 2.3). In further field sweeps, only the spin-flip is visible. Moreover, the low-field transition is completely invisible for field sweeps with H ‖ [100]T, suggesting a relationship to the magnetic detwinning with b ‖ H observed in the magneto-optical measurements. Figure 9.9: Field-dependent magnetization of EuFe2As2 at T = 2K for different magnetic field directions: (a) H ‖ [110]T and (b) H ‖ [100]T. (a) At ∼ 0.7T, M(H) sharply increases, afterwards reaching saturation. Irreversible behaviour at around 0.1T is only observed during the first field cycling and when the field is parallel to the crystal’s orthorhombic axes (black points: increasing H after ZFC, red points: decreasing H within the first cycling, blue stars: increasing H in the second cycling). 9.2 Magnetization measurements 141 Temperature-dependent magnetization measurements with H ‖ [110]T further support this interpretation (see Figure 9.10): • Zero-field cooled M(T ) curves obtained at very low fields show a peak at TEu, how- ever, the magnetization does not drop to zero for lower temperatures (see Fig. 9.10a). Indeed, the curve can be described as a superposition of antiferromagnetic suscepti- bilities parallel and perpendicular to the easy axes (see chapter 2.3), corresponding to equally distributed twin domains. • For intermediate fields, M(T < TEu) shows a flat curvature, indicating an antifer- romagnetic spin alignment with the easy axes perpendicular to the external field (see Fig. 9.10b). As the easy axes of EuFe2As2 corresponds to its crystallographic a-direction, this is consistent with the low-field magnetic detwinning (that leads to b ‖ H). • A similar flat curvature below TEu can be observed for very low fields after field treatment (see Fig. 9.10d), indicating that the detwinning with b ‖ H is persistent. • As expected from the field-dependent magnetization measurements, the transition at TEu smears out above the spin-flip field, as field enforced ferromagnetism sets in (see Fig. 9.10c). Figure 9.10: Temperature-dependent magnetization for EuFe2As2 at certain magnetic fields H ‖ [110]T with and without field treatment. (a) 0.1T (ZFC), (b) 0.3T (ZFC), (c) 0.9T (ZFC), and (d) 0.1T (FT with 1T). (a) While the ZFC 0.1T curve displays a cusp at TEu, (d) the corresponding FT curve displays a rather flat magnetization below TEu. (b) The latter curvature is also observed at 0.3T (after ZFC). (c) M(T ) at 0.9T increases smoothly with decreasing temperature without any distinct anomaly at TEu. The question remains how exactly the spin-flip is related to the second detwinning process with a ‖ H. Therefore, field-dependent magnetization measurements at 15K are 142 9 Magnetic detwinning of EuFe2As2 displayed together with the results from magneto-optical measurements in Fig. 9.11a. The overall magnetization behaviour is very similar to that at 2K, despite a smooth increase of the magnetization above the spin-flip field due to thermal fluctuations. While the irreversible behaviour in optical and magnetization measurements appears in both cases around H1 ∼ 0.1T, the spin-flip field HSF ∼ 0.5T slightly precedes the detwinning process observed in the reflectivity by about 0.1T. Figure 9.11: Field-dependent magnetizationM(H) (grey open circles) and false colour plot of the frequency and field-dependent FIR relative reflectivity R(H)/R(0T) (E ⊥ H) of EuFe2As2 at (a) T =15K and (b) 30K (H ‖ [110]T). The detwinning fields H1 and H2 as well as the spin-flip field HSF of Eu2+ can be identified (H1 < HSF < H2). Detwinning at T > TEu As can be seen in Fig. 9.11b, no transition is visible in M(H) for T = 30K. It is in- triguing to associate the smooth increase of the magnetization with the rather continuous detwinning process at T > TEu. 9.3 Detwinning mechanism 143 9.3 Detwinning mechanism In summary, we have to explain how the magnetic detwinning takes place below and above the local Eu2+ magnetic ordering. While at T < TEu, the crystal gets detwinned with b ‖ H at very low fields (∼ 0.1T at 15K) and with a ‖ H at slightly higher magnetic fields (∼ 0.6T at 15K), at T > TEu, only the latter takes place. Most strikingly, a significant imbalance of twin domains remains even when the field is switched off, and the temperature is raised up to T > Ts,SDW. Indirect magneto-elastic coupling The first question one should address is how the magnetic detwinning fields can be reduced in EuFe2As2 by more than two orders of magnitude compared to other iron pnictides. Magneto-elastic coupling usually arises due to spin-orbit interactions. However, the orbital momentum of Eu2+ is zero; thus the magnetic anisotropy induced by spin-orbit interactions is negligible. Another possibility to induce magnetic anisotropy is by dipole- dipole interactions. However, the resulting anisotropy is much weaker [306]. Therefore, other unconventional interactions must cause our observations. From the doped or (chemically) pressurized Eu compounds, it is well known that the Eu2+ and Fe2+ magnetic orders are strongly intertwined (see chapter 6). Furthermore, the orbital moment of Fe2+ is non-zero, leading in BaFe2As2 to a significant magneto-elastic coupling [24]. Hence, we suggest that the Eu2+ spins couple indirectly to the lattice via the Fe2+ spins. Such an indirect magneto-elastic coupling was, to the best of our knowledge, not yet observed before. Detwinning at T < TEu In order to explain the two-step detwinning process at temperatures T < TEu, a simple model based on the competition between magneto-crystalline anisotropy ∆, antiferromag- netic exchange coupling J and Zeeman energy is sufficient. Two twin domains have to be considered, one with the easy a-axis perpendicular (type B‖) and one with it parallel (type A‖) to the external field (see Fig. 9.12a). 144 9 Magnetic detwinning of EuFe2As2 (a) H = 0T (b) H ≈ H1 (c) H = HSF (d) H ≥ H2 Detwinning: a || H a a a a a a a b b b bbbb a b c a b c H H H (e) HSF H E E0 twin BII twin AII Detwinning: b || H twin BII twin AII H1 H2 Δ Figure 9.12: Sketch of the twin distribution and spin configuration of EuFe2As2 dependent on an (from left to right increasing) external magnetic field H ‖ [110]T at T < TEu. Top and bottom rows show a detail of the EuFe2As2 crystal and magnetic structure (Eu atoms and spins (blue), Fe atoms (yellow) and spins (brown), As atoms (green)) that correspond to the twin distribution sketched in the middle row. (a) For H = 0T, the ZFC crystal is twinned and the domains are equally distributed. The Eu2+ spins are ordered A-type antiferromagnetically with the spin direction along the a-axis. (b) With external field, twin variant B‖ (red, bottom) with b ‖ H gets energetically favoured and therefore grows on the expense of variant A‖ (green, top), as soon as the energy difference exceeds the twin boundary pinning energy. (c) With further increasing field, Eu2+ spins in the remaining type A‖ twins flip along the field direction. Energetically, twin B‖ is still more favourable. (d) At slightly higher fields, twin A‖ is favoured and the crystal is detwinned with a ‖ H. (e) Corresponding energy curves (E0: ground state). While at low fields twin B‖ is energetically favourable (red area), twin A‖ gets favourable at higher fields (green area). Detwinning takes place at H1 and H2, when the energy gain exceeds the pinning energy of the twin boundary. 9.4 Other techniques 145 After cooling in zero magnetic field, the crystal is twinned with equally distributed variants. The simplest form for the energy is then given by (see chapter 2.3): E = −2Mµ0H(cos θ + cosϕ) + JM2 cos(θ + ϕ) −12∆  sin2 θ + sin2 ϕ (B‖) cos2 θ + cos2 ϕ (A‖) , (9.1) where θ and ϕ are the angles between spins and magnetic field and E0 = −JM2 −∆ is the 0T ground state energy. Minimizing the energy for θ = ϕ yields (see Fig. 9.12e): E B‖ min = E0 − M2(µ0H)2 2JM2 +∆ (9.2) E A‖ min = E0 +∆− M2(µ0H)2 2JM2 −∆ (9.3) Thus, at low fields, the Eu2+ spins of variant B‖ gradually rotate towards the direction of the external magnetic field, lowering the system’s energy. When EB‖min−E A‖ min exceeds the twin boundary pinning energy, variant B‖ grows irreversibly on the expense of variant A‖, and the crystal gets detwinned with b ‖ H (see Fig. 9.12b). Increasing the magnetic field further induces a spin-flip in variant A‖ (see Fig. 9.12c), but this twin variant is energetically favourable only at slightly higher fields, when the crystal gets detwinned with a ‖ H (see Fig. 9.12d). As characteristic for domain dynamics, these processes are strongly irreversible, leading to a significant persistent detwinning. Detwinning at T > TEu At T > TEu, the unordered Eu2+ spins gradually align along the direction of the external field, leading to a net magnetic moment along H. This favours for H ‖ [110]T the domain type, where the Fe2+ spins are aligned along the field direction. Therefore, a detwinning with a ‖ H can be observed, as soon as the energy gain exceeds the twin boundary pinning energy. 9.4 Other techniques Additional resistivity, magneto-resistance, thermal expansion and magnetostriction mea- surements were performed in the group of P. Gegenwart (by J. Maiwald, C. Stingl, N. Bach, and I. Pietsch, University of Göttingen / Augsburg). The main results are depicted in Fig. 9.13 and 9.14 and described in detail in Ref. [13]. 146 9 Magnetic detwinning of EuFe2As2 Resistivity and magneto-resistance Resistivity measurements after field treatment impressively demonstrate the persistent de- twinning, as a strong anisotropy along the orthorhombic axes remains after the magnetic field was switched off (see Fig. 9.13a,b). Consistent with the magneto-optical measure- ments, the induced anisotropy is opposite for field treatment below and above TEu. One should note that its magnitude agrees well with values obtained for mechanically detwinned Eu compounds [197, 240], and remains virtually constant up to Ts,SDW. Figure 9.13: Resistivity and magneto-resistance of ZFC EuFe2As2 (H ‖ [110]T, data taken by J. Maiwald, University of Göttingen / Augsburg). (a,b) Temperature-dependent resistivity ρFT(T ) after field treatment with 4T for currents parallel (red) and perpendicular (black) to H, normalized to the corresponding zero-field cooled value. A strong anisotropy along the orthorhombic axes remains after the magnetic field was switched off. For field treatment at (a) T = 4K and (b) at 30K the opposite behaviour is observed. (c,d) Magneto-resistance ρ‖(H)/ρ‖(0T) − 1 at (c) 5K and (d) 30K. (c) Below TEu, ρ‖(H) first increases and then decreases with increasing H; (d) above TEu, it decreases with increasing H. The initial ZFC value is not recovered after field sweeping. Magneto-resistance measurements further resemble the results obtained by the magneto- optical measurements, however, the critical fields seem to be different (Fig. 9.13c,d). The main reason for this discrepancy is that all electronic contributions add up at zero fre- quency, meaning that other field-dependent processes such as Eu2+ spin scattering also play an important role [198]; this masks the response due to the magnetic detwinning. Moreover, one should note that the electric contacts needed in resistivity measurements 9.4 Other techniques 147 induce significant stress on the sample. Indeed, already this force is often strong enough to partially detwin a sample1 and might influence transition temperatures and fields. Thermal expansion and magnetostriction Concurrent evidence for a persistent structural detwinning is also found in thermal expan- sion and magnetostriction measurements (see Fig. 9.14), techniques which directly probe the induced length changes. Relative length changes are of the order of 10−3, exceeding the typical magnetostriction of Eu based materials by at least two orders of magnitude. Figure 9.14: Thermal expansion and magnetostriction ∆L(T )/L0 of ZFC EuFe2As2 (H ‖ [110]T, ∆L ‖ H, data taken by C. Stingl, N. Bach, and I. Pietsch, University of Göttingen). (a) Thermal expansion after zero-field cooling (black) and field treatment with 2T at 4K (cyan). Apart from transitions at TEu and Ts,SDW, the sample is shorter after field treatment. (b,c) Magnetostriction at (b) 30K and (c) 5K for increasing (black, brown) and decreasing (green) H. (c) While at T > TEu, the sample expands along H, (b) at T < TEu, it first contracts and then expands. The initial ZFC value is not recovered after field sweeping. By comparing those results with experimentally determined lattice constants [26], one can yield information about the detwinning fraction. Assuming a temperature-independent fraction n of twins with b ‖ H, the mean length of a unit cell in field direction follows as L¯ = n · b(T ) + (1− n) · a(T ). (9.4) 1 Therefore, Fig. 9.13 only displays relative values of the resistivity. 148 9 Magnetic detwinning of EuFe2As2 The relative length change compared to aT (the tetragonal lattice constant at 300K) then reads as: ∆L L0 = L¯− aT aT . (9.5) By setting n = 66%, we can reproduce the ZFC curve in Fig. 9.14 reasonably well, indi- cating that the sample is already partially detwinned due to the uniaxial pressure from the dilatometer, which favours twins with the shorter axis b ‖ ∆L. Nevertheless, the twin configuration can be changed to n = 70% by field treatment. Importantly, the induced imbalance of twin domains stays constant up to Ts,SDW 9.5 Detwinning fraction From neutron measurements, we know that the crystal gets completely detwinned at 1T [7]. However, the question remains whether we can quantize the detwinning fraction after field treatment. Unfortunately, our methods do not provide a key for a quantitative estimate: • As in BaFe2As2 the FeAs phonon is only visible along the b-direction, it is intriguing to use this phonon in order to determine the detwinning fraction. In our measure- ments on EuFe2As2, however, even for 1T (where the crystal should be completely detwinned according to the neutron studies), the phonon is still visible in σa (see Fig. 9.15). Its presence could be attributed to the imperfect polarizer and the im- proper alignment of polarizer and sample with respect to the external magnetic field, however, its suppression along the a-direction might be also strongly composition- dependent - similar to the appearance of the smaller spin density wave gap (see also section 9.6). Therefore, we are not able to extract reliable information about the detwinning fraction by our magneto-optical measurements. • Resistivity measurements show that the anisotropy induced by field treatment agrees well with values obtained for mechanically detwinned Eu compounds [197, 240]. Therefore, the magnetic detwinning provides a good alternative to mechanical de- vices; however, the absolute value of the detwinning fraction stays unclear. • Thermal expansion and magnetostriction measurements can not be used to deter- mine the detwinning fraction, as the uniaxial pressure from the dilatometer already detwins the sample partially. However, we speculate that more information can be extracted from already published neutron diffraction data: in the first manuscript of Xiao et al., where no magnetic fields were used at all, equally distributed twin domains were reported [8]. However, in a later 9.6 Further infrared studies 149 Figure 9.15: Polarization-dependent optical conductivity of EuFe2As2 around the FeAs phonon mode at ∼ 260 cm−1 for 15K with and without magnetic field; 0T data were used for an appropriate extrapolation before performing the Kramers-Kronig transformation. (a) 0T (ZFC), (b) 1T (H ‖ [110]T, resolution 1 cm−1). In the magnetic field, the oscillator strength of the phonon is enhanced for σb (E ⊥ H, black) and reduced for σa (E ‖ H, red). study about the influence of magnetic fields on the structure and Eu2+ magnetism, the same authors claimed at 0T a domain ratio of 1:3 [7]; the low-field detwinning process was not observed at all. We suspect that the authors did not obey a well-defined cooling process, causing at 0T an imbalanced twin domain ratio due to the persistent detwinning. This would imply that field treatment partially detwins EuFe2As2 with the domain ratio of 1:3. One possible explanation why the crystal does not get completely detwinned at H1 and after field treatment is that we have to consider the three-dimensional nature of the A-type antiferromagnetism, as presented in Fig. 9.16. Within one layer, Eu2+ spins of next-nearest structural domains can be aligned parallel (see Fig. 9.16a,b) or antiparallel (see Fig. 9.16c,d). The reorientation of the intermediate domain will be less favourable for the latter case and might be prohibited. This would lead to a detwinning fraction of 75% (consistent with our reinterpretation of the neutron data): 50% of the domains are initially oriented with b ‖ H, additional 25% can be reoriented by the field treatment (see Fig. 9.16a,b). 9.6 Further infrared studies We have also performed first test measurements over a broad energy range on magnetically detwinned samples. However, the detectors available in the FFIR and MIR energy range could not guarantee (in conjunction with the complex magneto-optical setup, see chap- ter 5.1) a high enough signal-to-noise ratio as well as a sufficient time-stability for a com- 150 9 Magnetic detwinning of EuFe2As2 a b a b b a a b a bb a a b b a a b a b b a a b (a) (b) (c) (d) Figure 9.16: Considering the three-dimensionality of A-type antiferromagnetism could ex- plain, why the crystal gets not completely detwinned at H1. Within one layer, Eu2+ spins (blue) of next-nearest structural domains can be aligned (a,b) parallel or (c,d) antiparallel. The reorientation of the intermediate domain will be less favourable for the latter case. plete gold evaporation run. Therefore, we present here only preliminary results at 30K1. While FIR data are measured completely in the magneto-optical setup, the reflectivity of the other energy ranges were obtained as followed: absolute values achieved in the conven- tional zero-field cryostat were normalized by the ratio of the reflectivities before and after field treatment, both obtained in the magneto-optical setup. FFIR data are smoothed; due to strong absorptions in the MIR range originating from the non-exchangeable polypropy- lene windows of the cryostat, the mid-infrared spectrum was strongly smoothed, as well as cutted above 3000 cm−1; higher frequencies were extrapolated with the 0T data, low frequencies according to the σdc-values. We note that all spectral ranges perfectly merge (see Fig. 9.17). Similar to the reports on BaFe2As2 (see chapter 4.4.5), we observe a strong anisotropy in the mid-infrared frequency range. By fitting with the “subtraction approach” intro- duced in chapter 8.2, we find 1/τb > 1/τa and slightly higher Drude weights for the a-direction, similar to Ref. [171]. Furthermore, we confirm that the spin density wave gap around ≈ 10.8kBTs,SDW consists of two features, with the higher feature stronger visible in the b-direction. This is consistent with measurements and theoretical calculations for BaFe2As2 [168, 169]. However, we do not observe any evidence for a lower spin density wave gap along one of the main crystallographic directions, consistent with our results on the twinned crystal (see section 8.1). Furthermore, our observations concerning the FeAs phonon differ from the behaviour reported for BaFe2As2, as already discussed in chapter 9.5. Along the orthorhombic a- 1 Further measurements on a newly constructed magneto-optical setup at the 1st Physical Institute, Stuttgart are scheduled. 9.6 Further infrared studies 151 Figure 9.17: Frequency-dependent in-plane (a-b) reflectivity and (c-d) conductivity of EuFe2As2 at T =30K in different logarithmic and linear plot styles. Grey (light cyan) squares represent data points for E ‖ a (E ‖ b), black (dark cyan) lines the corresponding fit with two Drude components and a Lorentzian at 1000 cm−1. (d) The phonon is narrower along the b-direction and slightly shifted to lower frequencies. direction, it has a reduced oscillator strength, but does not disappear. Indeed, we are able to identify for σa(ω) a slight blueshift of the phonon (see Fig. 9.17d). This blueshift is consistent with a weak higher-energy a-axis phonon contribution revealed for BaFe2As2 by Schafgans et al. [174]. We suggest that this contribution is more pronounced in the case of EuFe2As2. Therefore, high-resolution far-infrared measurements on mechanically detwinned EuFe2As2 crystals could give important information on the anisotropic phonon dynamics in 122 iron pnictides, which are yet not well understood [171]. 152 9 Magnetic detwinning of EuFe2As2 9.7 Outlook Our magneto-optical and magnetization measurements reveal a dramatic reduction of magnetic detwinning fields compared to other AFe2As2 (A = Ba, Sr, Ca) iron pnictides by indirect magneto-elastic coupling of the Eu2+ ions. Such an indirect magneto-elastic coupling was, to our best knowledge, not demonstrated before. We find that only ∼ 0.1T are sufficient for persistent detwinning below the local Eu2+ ordering, accessible by typical laboratory magnetic fields; a significant imbalance of twin domains remains constant up to the structural and electronic phase transition (190K). This is surprising from a very fundamental point of view, as the detwinning induced at low temperatures with very small magnetic fields persists up to much higher temperatures and thus energy scales. This may even lead to technical applications. With respect to the research on high temperature superconductors, the whole effect un- covers a remarkable interdependence between magnetic, electronic and structural effects. It is surprising that the Eu2+ magnetic order strongly affects the lattice, although it does not destroy superconductivity in substituted compounds. This opens up unprecedented possibilities for fundamental research concerning the interplay of superconductivity and magnetism. Finally, our findings provide a unique possibility to study macroscopically the intrinsic in-plane anisotropy of iron pnictides without the application of any symmetry- breaking external force. 10 SUMMARY AND OUTLOOK The investigations carried out in the framework of this thesis reveal that Eu based iron pnictides offer a unique playground to examine the interplay of unconventional electronic properties, superconductivity, itinerant and local magnetism. Through systematic magnetization studies on a complete set of EuFe2(As1−xPx)2 single crystals, we have derived the complex magnetic phase diagram of Eu 122 pnictides (see chapter 7). As a first step, we identified a ferromagnetic out-of-plane component that develops in the underdoped, non superconducting region of the phase diagram, consis- tent with canted A-type antiferromagnetism. In addition, the Eu2+ magnetic transition temperature TN is suppressed similar to the structural and spin density wave transition Ts,SDW and rises again as soon as superconductivity sets in. This implies that the Eu2+ spin orientation in the parent compound is defined by the spin density wave anisotropy, consistent with theoretical calculations by Akbari et al. [11]. Therefore, we conclude sig- nificant interactions between the local Eu2+ and the (quasi-) itinerant Fe magnetism. Furthermore, we identified the development of a re-entrant spin glass phase at lower temperatures, characterized by time-dependent processes that appear due to the freezing of the in-plane spin components. The corresponding transition temperature Tglass rises quickly in the underdoped region until it is almost as high as TN for superconducting compounds. For higher P substitution levels, both transition temperatures continuously increase, separated only by ∼ 2K. One should further note the rather narrow supercon- 154 10 Summary and outlook ducting dome, which might be related to a Lifshitz transition [235] that also affects the RKKY-exchange. However, our studies rule out any change to ferromagnetic interlayer coupling at high P substitutions. All these findings are probably common features of (non-diluted) Eu based iron pnic- tides, as discussed on the basis of a review given in chapter 6. It can be speculated that the Eu2+ ferromagnetic component along the c-axis is screened by the formation of spontaneous vortices perpendicular to the layers [31]. Most importantly, the glass-like dy- namics indicate a decoupling of the Eu layers, therefore implying that superconductivity in the intermediate FeAs layers is mostly unaffected by any magnetic interlayer coupling. Thus, our findings elucidate how bulk superconductivity can coexist with the strong Eu2+ magnetic ordering. Despite the interactions of local and itinerant magnetism revealed by our magnetization studies, the electrodynamic response of EuFe2As2 is surprisingly unaffected by the Eu2+ magnetic transition, evidenced by infrared spectra that basically do not change when passing TN (see chapter 8). However, we are able to confirm earlier results of Wu et al. [184], who reported the absence of any spin density wave gap feature around the BCS weak-coupling limit. Although it is intriguing to relate this absence to the local Eu2+ magnetism, such a relationship has (yet) no theoretical background. In order to obtain more information about the general electrodynamic properties of iron pnictides, we have further compared typical fitting approaches. We find that real- istic results are obtained with a model that contains two Drude components (probably representing electron and hole bands) and a Lorentzian around 1000 cm−1, which can be attributed to interband transitions involving Fe hole pockets [12]. Importantly, we find that regardless of the fitting method, the spin density wave transition affects not only both Drude components, but also the spectral weight transfer to a bump-like structure in the near-infrared and even higher energies. This supports a strong relationship of the latter feature to the dynamics of the itinerant carriers, consistent with correlation effects due to Hund’s rule coupling. Finally, it was demonstrated in chapter 9 that magneto-optical measurements in the far-infrared frequency range are ideal to investigate the magnetic detwinning mechanism in EuFe2As2. We find that only ∼ 0.1T are sufficient for a persistent detwinning below the local Eu2+ ordering temperature, which is accessible by typical laboratory magnetic fields; above TN, only slightly higher fields (around 1T) are needed. We ascribe these observations to an indirect magneto-elastic coupling of the Eu2+ ions via the Fe2+ spins, which was - to our knowledge - not observed before, and give a simple model based on the competition between magneto-crystalline anisotropy, antiferromagnetic exchange coupling and Zeeman 155 energy. Our findings are further confirmed by magnetization, magneto-striction, thermal expanison and (magneto-)resistance measurements. Most importantly, the term persistent detwinning implies that a significant imbalance of twin domains remains when the magnetic field is switched off and the temperature is raised up to Ts,SDW. Such strongly hysteretic behaviour is surprising from a very funda- mental point of view, as the detwinning that is induced at low temperatures with very small magnetic fields persists up to much higher temperatures and thus energy scales. In summary, the whole effect uncovers a remarkable interdependence between magnetic, electronic and structural effects in Eu based iron pnictides. While all studies performed in the framework of this thesis concentrated on EuFe2- (As1−xPx)2, we suggest that many of our findings have further implications for other materials, especially for rare earth containing high-temperature superconductors such as related Eu 122 pnictides, Ce 1111 pnictides and rutheno-cuprates [82, 307]. In chapter 6 and 7 we have already presented evidence for the canted antiferromagnetic and re-entrant spin glass phase in other Eu 122 pnictides. One should note further that glass-like be- haviour was also reported for rutheno-cuprates [308, 309] and CeFePO [310]; this implies that glassy dynamics may be the key to how high-temperature superconductivity can coexist with rare earth magnetism. Furthermore, we are confident that the low-field persistent detwinning discovered in EuFe2As2 actually appears also in other rare earth containing materials. Therefore, we are planning further investigations on diluted Eu compounds. Moreover, the persistent mag- netic detwinning is also interesting for other experimental techniques such as ARPES [311], where the implementation of additional mechanical devices is challenging, and for theo- reticians [312], since it provides the unique possibility to study the macroscopic in-plane anisotropy close to the phase transition without the application of any symmetry-breaking force. Finally, one should note that the work presented here also opens up very concrete ques- tions concerning the research on Eu based 122 superconductors that we want to address in the near future. In this regard, we suggest that studies on diluted compounds may uncover the source of the anomalies related to the spin density wave (see chapter 8) and the superconducting gap [233]. Likewise, further optical investigations on superconducting EuFe2(As1−xPx)2 compounds might reveal more information about the relationship of the Lifshitz transition to the rather narrow superconducting dome. Furthermore, it will be interesting to examine whether the indirect magneto-elastic coupling also influences the nematic fluctuations [313]. Therefore, this thesis can be concluded safe in the knowledge 156 10 Summary and outlook that it has not only contributed important insights, but also inspired future work in a very interesting field of nowadays solid state research. A THZ INVESTIGATIONS ON BA 122 THIN FILMS As already presented in chapter 4.4, many contradictory reports exist about the optical properties of Ba(Fe1−xCox)2As2. In particular, the number of superconducting gaps as well as the origin of the unexpected strong intra-gap absorption is highly debated. In the following, our approaches to contribute to this topic are shortly summarized. In all cases, a coherent source Terahertz spectrometer (4 - 45 cm−1) with a Mach-Zehnder interferometer was used in order to measure the transmission and the corresponding phase shift of Ba(Fe0.9Co0.1)2As2 thin films on dielectric (La,Sr)(Al,Ta)O3 (LSAT) or CaF2 sub- strates; additional transmission measurements were performed to gain the optical proper- ties of the substrates. Due to multiple reflections at the surfaces, the measured properties oscillate with a periodicity dependent on the optical thickness of the sample (see Fig. A.1). In the normal state, the optical conductivity of the film can be modeled with a single Drude term, accounting for its metallic nature. In the superconducting state, each maxi- mum is evaluated individually with two Drude terms: one with a very low scattering rate that models the superconducting δ-function, the other one with a high scattering rate accounting for the quasi-particle absorption [148]. The advantage of such an approach is that it is independent of any model, i.e. for example of the gap symmetry. A more detailed 158 A THz investigations on Ba 122 thin films description of the measurement technique and the analysis can be found in Ref. [148] and Ref. [314]. A.1 Fabry-Pérot resonant technique One of the main problems of such measurements is that the optical intra-gap conduc- tivity of metallic superconductors can be extracted experimentally only with rather large uncertainty, caused by the huge negative value of the dielectric constant that mainly deter- mines the transmission. Therefore, we have performed systematic studies of Fabry-Pérot resonators, which consist of two identical superconducting iron pnictide thin films on LSAT substrates that are positioned face-to-face to each other, separated by a spacer. The cor- responding results that are already published in Ref [315] and Ref [316] are summarized in this section. Enhancement principle Fig. A.1a shows exemplarily the calculated transmission of such a Fabry-Pérot arrange- ment compared to the spectrum of a usual sample (single film on substrate) and the spectrum that would be caused by light passing two identical free standing films. For the Fabry-Pérot arrangement, a complex transmission pattern develops due to multiple reflec- tions at numerous interfaces: the highest peaks arise at frequencies where the transmission of both, the single film on a substrate and the free standing films resonator, is maximized. In the minima of the substrate-film arrangement, the spectrum of the Fabry-Pérot res- onator corresponds to the free standing films system. The resonances of the Fabry-Pérot arrangement enhance in amplitude over several orders of magnitude and exceed at some points the transmission of the single film on a substrate. As the maximum sensitivity to σ1(ω) of the films is achieved when the whole peak structure is experimentally resolved, one is restricted to rather thin films to achieve transmissivity at least of the order 10−5, which is the resolution limit of the spectrometer. The enhancement of sensitivity with respect to σ1(ω) can be demonstrated by nomo- grams, showing lines of constant transmission coefficient and phase shift plotted in the coordinates of real and imaginary parts of σ(ω) (or other optical constants) of the films. Fig. A.2 compares the nomograms of a single film and a Fabry-Pérot resonator on the basis of 30 nm Ba(Fe0.9Co0.1)2As2 film data that is representative for iron pnictides. The line of constant transmission coefficient is noticeably less steep for the Fabry-Pérot arrangement, meaning that the accuracy to σ1(ω) is higher, whereas the accuracy to the phase decreases. When fitting the spectra, the accuracy to σ1(ω) also benefits from “double-structures” in the THz spectra of Fabry-Pérot resonators, which emerge when the resonances between A.1 Fabry-Pérot resonant technique 159 Figure A.1: (a) Calculated transmission at temperatures T > Tc (left panel) and T < Tc (right panel) for a single film on substrate (black dashed line), two free standing films (red dashed-dotted line) and the Fabry-Pérot (FP) arrangement (green line). The Fabry-Pérot resonator is shown schematically in the inset of (a). Model parameters are chosen close to measurements (dsub = 0.495mm, dgap = 1.934mm, dfilm = 30nm, σ1(T > Tc) = 6000Ωcm−1, σ1(T < Tc) = 2000Ωcm−1, ωp = 3464cm−1). (b) Nomogram: calculated constant values of transmission coefficient Θ (solid lines) and phase shift Φ (dashed lines) in optical constant co- ordinates (vertical: σ1, horizontal: 1) with the values typical for a 30 nm Ba(Fe0.9Co0.1)2As2 film in the superconducting state (T = 5K, ν = 5.13 cm−1, Θ = 0.17, Φ = 3.7 rad/cm−1). Black lines show data for a single film on a substrate (LSAT, dsub = 0.5mm), red lines for the Fabry-Pérot resonator (LSAT, dgap = 2mm). The line of constant transmission coefficient is noticeably less steep for the Fabry-Pérot arrangement, meaning that the accuracy to σ1 is higher, whereas the phase measurement (dashed lines) is more sensitive to σ1 in the case of the single film on substrate. different interfaces coincide. These resonances show different dependence on σ1(ω) and σ2(ω) of the films and the shape of the double-structure is thus very characteristic for a distinct ratio of these quantities. Thus, in a narrow frequency range, more information is available than from a spectrum of a single film on a substrate. Fitting procedure In order to account for the complex geometry of the Fabry-Pérot resonator, we expanded our fitting procedure to a five-layer model. Furthermore, we measured the single films for a temperature T1 > Tc and determine their normal state conductivity σ(T1) (we take T1 = 30K slightly above Tc ∼ 22K). The found value of σ(T1) is then used to precisely determine the geometrical parameters of the Fabry-Pérot system: substrate thickness dsub, space between the films dgap (“gap width”) and “gap absorption”. The latter parameter 160 A THz investigations on Ba 122 thin films accounts for the non-perfect geometry of the resonator (for example due to the non- parallelism of interfaces), which can be modelled in first approximation by a finite (but small) absorption in the gap between the films. Importantly, all spectra are processed with the same set of geometrical parameters. Experimental results We have determined the optical response of three Fabry-Pérot resonators, consisting of pairs of two identical films (dfilm = 25nm, 30 nm and 50 nm) that are glued via a spacer (metal or Plexiglas ring, diameter: ∼ 1 cm) face-to-face to each other. In order to empirically check, whether it is realistic to obtain data of such a Fabry-Pérot arrangement despite experimental challenges, first, a resonator composed of relatively thin films (dfilm = 25nm, Tc = 22K, dfilm ∼ 1mm, dgap ∼ 1mm) was examined. We found that the five-layer model describes the spectra nicely, as depicted in Fig A.2a. However, as far as physics is concerned, the film quality was rather poor, as indicated by a very high London penetration depth λL ∼ 1.46 µm (at 5K). As we expected a higher film quality for thicker samples, further measurements were carried out on 50 nm films (Tc = 21K, dsub ∼ 1mm, dgap ∼ 1mm). However, for these thick films the transmission coefficient of the Fabry-Pérot resonator in the superconducting state was too low, i.e. beyond the resolution limit of the spectrometer (see Fig A.2b). 30 nm films (Tc = 23K, dsub ∼ 0.5mm, dgap ∼ 2mm) were found to satisfy the re- quirements given by film quality and resolution limit of the spectrometer (see Fig A.2c). However, we found that an additional factor - the plane-parallelism of the substrates - plays a crucial role for providing the high quality factor of the Fabry-Pérot resonator. In that case, the thickness of the substrates was non-uniform in the range 0.487mm to 0.505mm. Thickness measurements of the complete resonator gave values between 2.934mm and 2.958mm, meaning that the biggest effect causing a non-parallelism of the whole system came from non-plane-parallel substrates. Since such pronounced non-perfection that leads to the broadening of the Fabry-Pérot resonance peaks in the spectra is not included in the model based on Fresnel optics, we were not able to describe the experimental spectra consistently over a broad frequency and temperature range; the disturbance is too large to be taken into account by the “gap absorption” mentioned above. In summary, we have shown that our Fabry-Pérot resonators made out of superconduct- ing thin films (on substrates) have theoretically the potential to enhance the sensitivity to the real part of the optical conductivity at very low frequencies. In the experiment, the res- olution limit of our setup restrains the film thickness in the case of Ba(Fe0.9Co0.1)2As2 to approximately 30 nm. However, our measurements on 25 nm films proof that the thin film A.2 Superfluid density 161 Figure A.2: Transmission coefficient spectra (dots) of Ba(Fe0.9Co0.1)2As2 Fabry-Pérot res- onators measured at 5K, fitted (line) with a five-layer-model for the film thicknesses (a) dfilm = 25nm, (b) 30 nm and (c) 50 nm. quality and homogeneity is still a serious issue for iron pnictides that has to be solved be- fore proceeding with our work. We further note that the success of this approach depends delicately on the uniform thickness of the substrate. One solution would be to deposit two films on both sides of a substrate, which is, however, experimentally challenging. A.2 Superfluid density As already mentioned in the previous section A.1, a common problem of THz investigations on superconductors is the rather huge error of σ1(ω) at low frequencies. However, the 162 A THz investigations on Ba 122 thin films corresponding imaginary part σ2(ω) (which determines the penetration depth) can usually be extracted with very high accuracy. Therefore, our collaborator A. Karakozov has analysed the temperature-dependence of the normalized superfluid density: ρs(t) = λL(0)2/λL(t)2 (A.1) with the relative temperature t = T/Tc, using the α-model that describes multiband superconductors via BCS-like equations with a reduced critical temperature [317, 318]. This model was then applied to fit the optical properties of one of our Ba(Fe0.9Co0.1)2As2 films on LSAT substrate (Tc = 22K, dfilm = 50 nm). Considering the existence of a small energy gap with 2∆/kBTc ∼ 2 (see Fig. 4.10), the superfluid density can be only described within a model of at least three weakly interacting condensates with 2∆/kBTc ∼ 2, 3 and 4, as depicted in Fig. A.3. This is an important refinement of the commonly used minimal model which describes the electrodynamic properties of iron pnictides with two effective bands. Figure A.3: Experimental values for the normalized superfluid density ρs(t) = λL(0)2/λL(t)2 (red circles) for a 50 nm Ba(Fe0.9Co0.1)2As2 film with Tc = 22K. Additionally, calculations for a two-band (blue dotted line) and a three-band fit (red solid line) with a minimum gap 2∆/kBTc ∼ 2 are compared to a BCS one-band approach (adapted from Ref. [317]). The best description can be obtained by the three-band approach. A.3 Irradiation studies In 2011, Efremov et al. predicted that two-band superconductors with s±-symmetry might undergo a disorder-induced transition to s++-symmetry, with both gaps tending to the same, finite value [319, 320]. As optical spectroscopy is not sensitive to the phase A.3 Irradiation studies 163 of the superconducting order parameter, the corresponding zero-crossing of the smaller energy gap would appear as if the gap first closes and then re-emerges. Such a unique behaviour constitutes a smoking gun experiment to proof the s±-symmetry for iron-based superconductors. Furthermore, manipulating the scattering rate has also the potential to deliver impor- tant information about the origin of the strong intra-gap absorption: while nodes can be lifted by increasing scattering [72], pair-breaking should be rather enhanced (as long as superconductivity is not destroyed too much). Therefore, we have investigated the effect of proton and α-particle irradiation on the optical properties of 50 nm thin Ba(Fe0.9Co0.1)2As2 films on CaF2 substrate. The irradi- ation was performed on the DanFysik-350 at the P. N. Lebedev Physical Institute of the Russian Academy of Sciences (Moscow) with an energy of 200 keV, which is high enough to guarantee that all particles pass the film. The dose for one irradiation is 2 · 1014 cm−2, corresponding to a decrease of Tc by 2K per step for protons (see Fig. A.5c), and a de- crease of 3.5K per step for α-particles. The latter stronger suppression of Tc is consistent with the creation of magnetic scattering centres [321]; therefore, we concentrate here on the results obtained via proton irradiation. The corresponding investigations were super- vised in the framework of this thesis, and can be found in more detail in Ref. [322] and Ref. [323]. Intra-gap absorption The unusual strong non-BCS-like intra-gap absorption can be best visualized, when com- paring the frequency-dependent conductivity at T > Tc with that at T  Tc. As displayed in Fig. A.4, even after several irradiation steps that suppress Tc by more than 10K, the conductivity at lowest frequencies and 5.6K still exceeds the normal state values. Thus, the intra-gap absorption must originate - at least to some extent - from pair-breaking interband scattering. Crossover from s±- to s++-symmetry In order to observe the crossover from s±- to s++-symmetry of the superconducting order parameter, we track the position of the coherence peak Tmax for eight irradiation steps1; the corresponding Tc value is determined by the abrupt onset of frequency-dependent changes in the measured phase, typical for a superconductor at T < Tc. For the 7th and 1 If its shape resembles rather a plateau than a peak, we consider the temperature where the plateau sets in as Tmax. 164 A THz investigations on Ba 122 thin films Figure A.4: Frequency-dependent optical conductivity σ1(ω) at 30K (closed dark symbols) and 5.6K (open bright symbols) for (a) the unirradiated film (Tc = 26K, blue), (b) the film after five irradiation steps with protons (Tc = 16K, red), as well as (c) the film after eight irradiation steps (Tc ∼ 10K, green, adapted from Ref. [322]). In all cases, a strong intra-gap absorption can be observed below Tc. 8th irradiation, however, the transition is already very weak and possibly smeared out too much, in order to determine Tc in this way. As can be seen in Fig. A.5c, the irradiation suppresses Tc and the residual resistivity ratio (σ(30K)/σ(295K)) at least for the first six irradiation steps linearly, meaning that the scattering rate is enhanced proportional to the irradiation dose. In parallel, the coherence peak is quickly suppressed, until it is shifted out of the measured temperature window for the fifth irradiation (see Fig. A.5a and Fig. A.5c). Interestingly, Tmax is quicker suppressed than Tc, meaning that the BCS relation between gap value and critical temperature is violated; this is an important condition to observe the s±- to s++ crossover. Furthermore, the permittivity is reduced significantly with irradiation, while the penetration depth λL raises almost linearly, consistent with a smooth decrease of the superfluid density. For the 7th irradiation, however, the behaviour starts to change. The conductivity in the normal state, as well as the residual resistivity ratio increase slightly. This might indicate that the number of defects got so high that an impurity band has formed, contributing to the overall conductivity. We want to stress, however, that this does not affect our interpretations concerning the coherence peak, as the conductivity in the normal state is still perfectly flat in the THz range (see Fig. A.4c). Our remarkable observation is that at the same time, the coherence peak shifts back to higher frequencies, clearly visible A.3 Irradiation studies 165 Figure A.5: Optical properties of a 50 nm Ba(Fe0.9Co0.1)2As2 film (Tc = 26K) depen- dent on the irradiation with protons (dose: 2 · 1014 cm−2 per step; adapted from Ref. [322]). (a,b) Temperature-dependent (a) optical conductivity σ1(T ) and (b) permittivity 1(T ) for selected irradiation steps (unirradiated: green closed diamonds, 2nd irradiation: dark cyan open squares, 5th irradiation: blue closed hexagons, 7th irradiation: purple open circles, 8th irradiation: red closed stars). (c,d) Residual resistivity ratio RRR σ(30K)/σ(295K) (grey squares), compared to (c) the suppression of Tc (orange squares) and the maximum of the coherence peak Tmax, as well as (d) London’s penetration depth at 5 cm−1 (green squares). Lines and arrows are guides to the eye, stressing that the 8th irradiation clearly deviates from the trend expected from the first irradiation steps. 166 A THz investigations on Ba 122 thin films when comparing in Fig. A.5a σ1(T ) of the 7th and 8th irradiation step. Furthermore, the penetration depth saturates, or even tends to decrease. Those observations may indicate a re-appearance of the lower superconducting gap in the measured temperature and fre- quency window, i.e. a crossover from s±- to s++-symmetry of the superconducting order parameter. B ACKNOWLEDGEMENTS Writing this thesis sometimes felt like a time travel through the last four years and I got aware once more how lucky I was to have so many people who supported me. • First of all, I would like to thank my supervisor Prof. Dressel. You always supported and trusted me - probably in some situations more than I did. Most of all, I want to thank you for the warm and familiar atmosphere at your institute. • I would also thank Prof. Keimer and Prof. Basov for kindly accepting to co-referee this thesis. I always enjoyed discussions with members of your groups. • Furthermore, I would like to thank Dan Wu. You introduced me with incredible patience in the field of iron pnictides and infrared spectroscopy. We really miss you in our team. Thanks also to Neven Barišić. I learnt a lot from our discussions. • Work at our institute like we know it would not be possible without the great techni- cal support by Gabi Untereiner, the low temperature and the mechanical workshop, as well as the motherly care of Agni Schmidt. 168 B Acknowledgements Furthermore, I want to acknowledge all my collaboration partners, especially: • The sample growers: H. S. Jeevan in the group of Prof. P. Gegenwart, S. Haindl and K. Iida in the group of Prof. B. Holzapfel, as well as S. Aswartham in the group of S. Wurmehl. Without your support, this thesis would not have been possible. • Boris Gorshunov and Elena Zhukova, our frequent visitors from Moscow. Boris, I learnt so much more from you than just to “press c”. Elena, I always enjoyed the “Hühnerstall” with you. • R. Kremer, as well as Mrs. Siegle and Mrs. Brücher for the technical support with the specific heat measurements, your confidence concerning the SQUID and the introduction to the ultra low field option. Mr. Kremer, I really learnt a lot while revising the first version of our PRL and I always enjoy(ed) our discussions. • Prof. D. N. Basov and Kirk Post, as well as all the group members and friends in San Diego: thanks for your great hospitality. Dimitri, I was always inspired, motivated and full of ideas when I came back from your group to Stuttgart. Kirk, you are the most positive person I ever met in my life (“... but look, how stable the temperature is...”) and I really appreciate the work with you, especially your russian-american cowboy Helium filling technique as well as the surfing lessons. • My collaborators in Göttingen (now Augsburg). I learnt a lot from our long discus- sions. • Prof. I. Eremin and A. Akbari for the theoretical support and the mutual exchange, as well as Y. Xiao for the discussion of the neutron data; you helped me to survive my first PRB experiences. Thanks also to Prof. J. Fink, Prof. J. Schmalian, Prof. R. Hackl, and Prof. I. Fisher for fruitful discussions. • D. Efremov and O. Dolgov for a lot of fruitful discussions concerning our thin film measurements. Furthermore, I have to thank the German Academic Exchange Service, the DFG, as well as the Baden-Württembergische Landesförderung for financial support. However, the whole thesis would not have been possible without my great colleagues at the 1st Physical Institute: • First of all, I would like to thank all the Infra-Rats, it was a pleasure to work with all of you. Special thanks for accepting me as “IR Mom” and all the help with the official stuff in the lab. 169 • Rebecca and Tobi, you were at my side from the first day. Becci, I learnt a lot when passing your “lessons” (which started on my first day at the institute with changing the pump oil). Tobi, no one is a more talented “lab dancer” than you... • David, Andy and Anja, you enriched our team with new energy and always good mood. David, thanks for the great support in the last two yours, especially while writing the thesis. Andy, I was really grateful when someone finally took my “Küken” burden. Anja, please never ever drink again two energy drinks in a row. • Uwe, although you chose the other side (without breakfast), thanks for working hardly on re-discovering my optimism. • Sabrina, Conrad and Eric, thanks for the nice inclusion in your “Team Columbus”. The worst lab day got better with a good mineral water in the evening... Sabrina, we should have thrown this cryostat out of the window. • Kirk, David, Anja and Conrad, the whole magnetic detwinning project would not have been possible without you and your night shifts. • My diploma student Felix, my master student Micha, as well as my bachelor students: Kilian, Fabi, Andreas and Alina - I hope that you could learn from me as much as I learnt from you and your questions. • Tomislav, Eva, Julia and Stefano, thanks a lot for the help in the dc and ellipsometry lab. The work atmosphere was always great. Special thanks also to my old colleagues from Würzburg, in particular to Prof. Geurts, Prof. Ossau, Prof. Umbach and Prof. Claessen for the reports that supported me in getting my scholarships. Although it was quite some work, I would also like to thank Prof. Honerkamp to suggest me for organizing the SPP summer school. Finally, I want to thank my family and friends for all the support during the years. I can not find the words to express how much this means to me. Christian, I will never forget when you came to my lab with a freshly baked pizza in an improvised delivery box (and it was still warm). Thanks for all your patience. Bibliography [1] M. R. Norman, Science 332, 196 (2011). [2] W. Buckel and R. Kleiner, Supraleitung, Wiley, Weihnheim (2013). [3] D. N. Basov and A. V. Chubukov, Nature Physics 7, 272 (2011). [4] Y. Kamihara, T. Watanabe, M. Hirano, and H. Hosono, J. Am. Chem. Soc. 130, 3296 (2008). [5] D. C. Johnston, Adv. Phys. 59, 6 (2010). [6] I. R. Fisher, L. Degiorgi, and Z. X. Shen, Rep. Prog. Phys. 74, 124506 (2011). [7] Y. Xiao, Y. Su, W. Schmidt, K. Schmalzl, C. M. N. Kumar, S. Price, T. Chatterji, R. Mittal, L. J. Chang, S. Nandi, N. Kumar, S. K. Dhar, A. Thamizhavel, and T. Brückel, Phys. Rev. B 81, 220406(R) (2010). [8] Y. Xiao, Y. Su, M. Meven, R. Mittal, C.M.N. Kumar, T. Chatterji, S. Price, J. Persson, N. Kumar, S.K. Dhar, A. Thamizhavel, and T. Brückel, Phys. Rev. B 80, 174424 (2009). [9] S. Zapf, D. Wu, L. Bogani, H. S. Jeevan, P. Gegenwart, and M. Dressel, Phys. Rev. B 84, 140503 (2011). [10] S. Zapf, H. S. Jeevan, T. Ivek, F. Pfister, F. Klingert, S. Jiang, D. Wu, P. Gegenwart, R. K. Kremer, and M. Dressel, Phys. Rev. Lett. 110, 237002 (2013). [11] A. Akbari, I. Eremin, and P. Thalmeier, Phys. Rev. B 84, 134513 (2011). [12] P. Marsik, C. N. Wang, M. Rössle, M. Yazdi-Rizi, R. Schuster, K. W. Kim, A. Dubroka, D. Munzar, T. Wolf, X. H. Chen, and C. Bernhard, Phys. Rev. B 88, 180508(R) (2013). [13] S. Zapf, C. Stingl, K. W. Post, J. Maiwald, N. Bach, I. Pietsch, D. Neubauer, A. Löhle, C. Clauss, S. Jiang, H. S. Jeevan, D. N. Basov, P. Gegenwart, and M. Dressel, Phys. Rev. Lett. 113, 227001 (2014). 172 Bibliography [14] S. Blundell, Magnetism in Condensed Matter, Oxford University Press Inc., New York (2001). [15] J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Phys. Rev. 106, 162 (1957). [16] H. Fröhlich, Phys. Rev. 79, 845 (1950). [17] J. Bardeen, Phys. Rev. 80, 567 (1950). [18] W. L. McMillan, Phys. Rev. 167, 331 (1968). [19] L. Gao, Y. Y. Xue, F. Chen, Q. Xiong, R. L. Meng, D. Ramirez, C. W. Chu, J. H. Eggert, and H. K. Mao, Phys. Rev. B 50, 4260 (1994). [20] J. G. Bednorz and K. A. Müller, Z. Phys. B 64, 189 (1986). [21] E. Fradkin and S. A. Kivelson, Nature Phys. 8, 864 (2012). [22] N. Lavrov, Y. Ando, S. Komiya, and I. Tsukada, Phys. Rev. Lett. 87, 017007 (2001). [23] A. Lavrov, S. Komiya, and Y. Ando, Nature 418, 385-386 (2002). [24] J.-H. Chu, J. G. Analytis, D. Press, T. D. Ladd, Y. Yamamoto, and I. R. Fisher, Phys. Rev. B 81, 214502 (2010). [25] J.-H. Chu, H.-K. Kuo, J. G. Analytis, and I. R. Fisher, Science 337, 710 (2012). [26] M. Tegel, M. Rotter, V. Wieß, F. M. Schappacher, R. Pöttgen, and D. Johrendt, J. Phys.: Condens. Matter 20, 452201 (2008). [27] H. S. Jeevan, Z. Hossain, D. Kasinathan, H. Rosner, C. Geibel, and P. Gegenwart, Phys. Rev. B 78, 052502 (2008). [28] Z. Ren, Z. Zhu, S. Jiang, X. Xu, Q. Tao, C. Wang, C. Feng, G. H. Cao, and Z. A. Xu, Phys. Rev. B 78, 052501 (2008). [29] C. F. Miclea, M. Nicklas, H. S. Jeevan, D. Kasinathan, Z. Hossain, H. Rosner, P. Gegenwart, C. Geibel, and F. Steglich, Phys. Rev. B 79, 212509 (2009). [30] Z. Ren, Q. Tao, S. Jiang, C. Feng, C. Wang, J. Dai, G. Cao, and Z. Xu, Phys. Rev. Lett. 102, 137002 (2009). [31] W. H. Jiao, Q. Tao, J. K. Bao, Y. L. Sun, C. M. Feng, Z. A. Xu, I. Nowik, I. Felner, and G. H. Cao, Europhys. Lett. 95, 67007 (2011). Bibliography 173 [32] S. Jiang, H. Xing, G. Xuan, Z. Ren, C. Wang, Z. A. Xu, and G. Cao, Phys. Rev. B 80, 184514 (2009). [33] U. B. Paramanik, D. Das, R. Prasad, and Z. Hossain, J. Phys.: Condens. Matter 25, 265701 (2013). [34] H. S. Jeevan, Z. Hossain, D. Kasinathan, H. Rosner, C. Geibel, and P. Gegenwart, Phys. Rev. B 78, 092406 (2008). [35] M. Reis, Fundamentals of Magnetism, Academic Press, Amsterdam (2013). [36] M. A. Ruderman and C. Kittel, Phys. Rev. 96, 99 (1954). [37] M. Getzlaff, Fundamentals of Magnetism, Springer, Berlin (2008). [38] J. A. Mydosh, Spin Glasses: An Experimental Introduction, Taylor & Francis, London (1993). [39] M. Gabay and G. Toulouse, Phys. Rev. Lett. 47, 201 (1981). [40] M. Tinkham, Introduction to Superconductivity, Dover Publications (2012). [41] C. A. Reynolds, B. Serin, W. H. Wright, and L. B. Nesbitt, Phys. Rev. 78, 487 (1950). [42] E. Mawell, Phys. Rev. 78, 477 (1950). [43] R. E. Glover III and M. Tinkham, Phys. Rev. 108, 243 (1957). [44] G. M. Eliashberg, Sov. Phys. JETP 11, 696 (1960). [45] L. Boeri, O.V. Dolgov, and A. A. Golubov, Phys. Rev. Lett. 101, 026403 (2008). [46] M. K. Wu, J. R. Ashburn, C. J. Torng, P. H. Hor, R. L. Meng, L. Gao, Z. J. Huang, Y. Q. Wang, and C. W. Chu, Phys. Rev. Lett. 58, 908 (1987). [47] H. Maeda, Y. Tanaka, M. Fukutomi, and T. Asano, Jpn. J. Appl. Phys. 27, L209 (1988). [48] M. R. Norman, D. Pines, and C. Kallin, Adv. Phys. 54, 8 (2005). [49] N. Doiron-Leyraud, C. Proust, D. LeBoeuf, J. Levallois, J.-B. Bonnemaison, R. Liang, D. A. Bonn, W. N. Hardy, and L. Taillefer, Nature 447, 565 (2007). [50] H.-B. Yang, J. D. Rameau, Z.-H. Pan, G. D. Gu, P. D. Johnson, H. Claus, D. G. Hinks, and T. E. Kidd, Phys. Rev. Lett. 107, 047003 (2011). 174 Bibliography [51] N. Harrison and S. E. Sebastian, New J. Phys. 16, 063025 (2014). [52] M. C. Boyer, W. D. Wise, K. Chatterjee, M. Yi, T. Kondo, T. Takeuchi, H. Ikuta, and E. W. Hudson, Nature Phys. 3, 802 (2007). [53] R.-H. He, M. Hashimoto, H. Karapetyan, J. D. Koralek, J. P. Hinton, J. P. Testaud, V. Nathan, Y. Yoshida, H. Yao, K. Tanaka, W. Meevasana, R. G. Moore, D. H. Lu, S.-K. Mo, M. Ishikado, H. Eisaki, Z. Hussain, T. P. Devereaux, S. A. Kivelson, J. Orenstein, A. Kapitulnik, and Z.-X. Shen, Science 331, 1579 (2011). [54] Y. He, Y. Yin, M. Zech, A. Soumyanarayanan, M. M. Yee, T. Williams, M. C. Boyer, K. Chatterjee, W. D. Wise, I. Zeljkovic, T. Kondo, T. Takeuchi, H. Ikuta, P. Mistark, R. S. Markiewicz, A. Bansil, S. Sachdev, E. W. Hudson, and J. E. Hoffman, Science 344, 608 (2014). [55] C. V. Parker, P. Aynajian, E. H. da Silva Neto, A. Pushp, S. Ono, Jinsheng Wen, Z. Xu, G. Gu, and A. Yazdani, Nature 468, 677 (2010). [56] D. H. Torchinsky, F. Mahmood, A. T. Bollinger, I. Boz˘ović, and N. Gedik, Nature Mater. 12, 387 (2013). [57] R. Comin, A. Frano, M. M. Yee, Y. Yoshida, H. Eisaki, E. Schierle, E. Weschke, R. Sutarto, F. He, A. Soumyanarayanan, Yang He, M. Le Tacon, I. S. Elfimov, Jennifer E. Hoffman, G. A. Sawatzky, B. Keimer, and A. Damascelli, Science 343, 390 (2014). [58] E. H. da Silva Neto, P. Aynajian, Al. Frano, R. Comin, E. Schierle, E. Weschke, A. Gyenis, J. Wen, J. Schneeloch, Z. Xu, S. Ono, G. Gu, M. Le Tacon, and A. Yazdani, Science 343, 393 (2014). [59] V. Hinkov, D. Haug, B. Fauqué, P. Bourges, Y. Sidis, A. Ivanov, C. Bernhard, C. T. Lin, and B. Keimer, Science 319, 597 (2008). [60] M. J. Lawler, K. Fujita, J. Lee, A. R. Schmidt, Y. Kohsaka, C. K. Kim, H. Eisaki, S. Uchida, J. C. Davis, J. P. Sethna, and E.-A. Kim, Nature 466, 347 (2010). [61] J. R. Waldram, Superconductivity in metals and cuprates, Institute of Physics Pub- lishing, London (1996). [62] N. F. Mott, Philos. Mag. 26, 1015 (1972). [63] A. F. Yoffe and A. R. Regel, Prog. Semicond. 4, 237 (1960). [64] T. Moriya and K. Ueda, Rep. Prog. Phys. 66, 1299 (2003). Bibliography 175 [65] N. L. Wang, H. Hosono, and P. Dai, Iron-based superconductors, Pan Stanford, Sin- gapore (2013). [66] F. Ronning, T. Klimczuk, E. D. Bauer, H. Volz, and J. D. Thompson, J. Phys.: Condens. Matter 20, 322201 (2008). [67] Q.-Y. Wang, Z. Li, W.-H. Zhang, Z.-C. Zhang, J.-S. Zhang, W. Li, H. Ding, Y.-B. Ou, P. Deng, K. Chang, J. Wen, C.-L. Song, K. He, J.-F. Jia, S.-H. Ji, Y.-Y. Wang, L.-L. Wang, X. Chen, X.-C. Ma, and Q.-K. Xue, Chin. Phys. Lett. 29, 037402 (2012). [68] H. Wadati, I. Elfimov, and G. A. Sawatzky, Phys. Rev. Lett. 105, 157004 (2010). [69] I. Mazin, Nature 464, 183 (2010). [70] I. I. Mazin, D. J. Singh, M. D. Johannes, and M.-H. Du, Phys. Rev. Lett. 101, 057003 (2008). [71] P. J. Hirschfeld, M. M. Korshunov, and I. I. Mazin, Rep. Prog. Phys. 74, 124508 (2011). [72] V. Mishra, G. Boyd, S. Graser, T. Maier, P. J. Hirschfeld, and D. J. Scalapino, Phys. Rev. B 79, 094512 (2009). [73] E. Wiesenmayer, H. Luetkens, G. Pascua, R. Khasanov, A. Amato, H. Potts, B. Banusch, H.-H. Klauss, and D. Johrendt, Phys. Rev. Lett. 107, 237001 (2011). [74] P. Cai, X. Zhou, W. Ruan, A. Wang, X. Chen, D.-H. Lee, and Y. Wang, Nature Commun. 4, 1596 (2012). [75] M. A. Tanatar, E. C. Blomberg, A. Kreyssig, M. G. Kim, N. Ni, A. Thaler, S. L. Bud’ko, P. C. Canfield, A. I. Goldman, I. I. Mazin, and R. Prozorov, Phys. Rev. B 81, 184508 (2010). [76] M. Yi, D. Luc, J.-H. Chu, J. G. Analytis, A. P. Sorini, A. F. Kemper, B. Moritz, S.-K. Mo, R. G. Moore, M. Hashimoto, W.-S. Lee, Z. Hussain, T. P. Devereaux, I. R. Fisher, and Z.-X. Shen, PNAS 108, 6878 (2011). [77] S. Kasahara, H. J. Shi, K. Hashimoto, S. Tonegawa, Y. Mizukami, T. Shibauchi, K. Sugimoto, T. Fukuda, T. Terashima, A. H. Nevidomskyy, and Y. Matsuda, Nature Phys. 486, 382 (2012). [78] C. Dhital, Z. Yamani, W. Tian, J. Zeretsky, A. S. Sefat, Z. Wang, R. J. Birgeneau, and S. D. Wilson, Phys. Rev. Lett. 108, 087001 (2012). 176 Bibliography [79] E. P. Rosenthal, E. F. Andrade, C. J. Arguello, R. M. Fernandes, L. Y. Xing, X. C. Wang, C. Q. Jin, A. J. Millis, and A. N. Pasupathy, Nature Phys. 10, 225 (2014). [80] R. M. Fernandes, A. V. Chubukov, and J. Schmalian, Nature Phys. 10, 97 (2014). [81] H. H. Kuo and I. R. Fisher, Phys. Rev. Lett. 112, 227001 (2014). [82] G. F. Chen, Z. Li, D. Wu, G. Li, W. Z. Hu, J. Dong, P. Zheng, J. L. Luo, and N. L. Wang, Phys. Rev. Lett. 100, 247002 (2008). [83] K. Hashimoto, K. Cho, T. Shibauchi, S. Kasahara, Y. Mizukami, R. Katsumata, Y. Tsuruhara, T. Terashima, H. Ikeda, M. A. Tanatar, H. Kitano, N. Salovich, R. W. Giannetta, P. Walmsley, A. Carrington, R. Prozorov, and Y. Matsuda, Science 22, 1554 (2012). [84] T. Shibauchi, A. Carrington, and Y. Matsuda, Annu. Rev. Condens. Matter Phys. 5, 113 (2014). [85] D. Manske, Theory of unconventional superconductors, Springer, Berlin (2004). [86] J. P. Carbotte, E. Schachinger, and D. N. Basov, Nature 401, 354 (1999). [87] E. G. Maksimov, M. L. Kulić, and O. V. Dolgov, Adv. Cond. Matt. Phys., 423725 (2010). [88] D. S. Inosov, J. T. Park, P. Bourges, D. L. Sun, Y. Sidis, A. Schneidewind, K. Hradil, D. Haug, C. T. Lin, B. Keimer, and V. Hinko, Nature Phys. 6, 178 (2010). [89] A. D. Christianson, E. A. Goremychkin, R. Osborn, S. Rosenkranz, M. D. Lumsden, C. D. Malliakas, I. S. Todorov, H. Claus, D. Y. Chung, M. G. Kanatzidis, R. I. Bewley, and T. Guidi, Nature 456, 930 (2008). [90] C. P. Poole Jr, R. Prozorov, H. A. Farach, and R. J. Creswick, Superconductivity, Elsevier, Waltham (2014). [91] M. McElfresh, Fundamentals of magnetism and magnetic materials, Quantum Design (1994). [92] C. P. Bean, Phys. Rev. Lett. 8, 250 (1962). [93] Y. Yeshrun, A. P. Malozemoff, and A. Shaulov, Rev. Mod. Phys. 68, 911 (1996). [94] S.V. Chong, S. Hashimoto, and K. Kadowaki, Solid State Comm. 150, 1178 (2010). [95] Y. B. Kim, C. F. Hempstead, and A. R. Strnad, Phys. Rev. Lett. 129, 528 (1963). Bibliography 177 [96] F. Herlach and N. Miur, High Magnetic Fields - Science And Technology, Vol. 3, World Scientific, Singapore (2006). [97] A. M. Clogston, Phys. Rev. Lett. 9, 266 (1962). [98] B. S. Chandrasekhar, Appl. Phys. Lett. 1, 7 (1962). [99] N. R. Werthamer, E. Helfand, and P. C. Hohenberg, Phys. Rev. 147, 295 (1966). [100] E. Müller-Hartmann and J. Zittartz, Phys. Rev. Lett. 26, 428 (1971). [101] K. Winzer, Solid State Comm. 24, 551 (1977). [102] K.-H. Müller and V. N. Narozhnyi, Rep. Prog. Phys. 64, 943 (2001). [103] P. Fulde and R. A. Ferrell, Phys. Rev. 135, A550 (1964). [104] A. I. Larkin and Y. N. Ovchinnikov, Zh. Eksp. Teor. Fiz. 47, 1136 (1964). [105] A. I. Larkin and Y. N. Ovchinnikov, Sov. Phys. JETP 20, (1965). [106] H. A. Radovan, N. A. Fortune, T. P. Murphy, S. T. Hannahs, E. C. Palm, S. W. Tozer, and D. Hall, Nature 425, 51 (2003). [107] P. W. Anderson and H. Suhl, Phys. Rev. 116, 898 (1959). [108] M. Tachiki, H. Matsumoto, T. Koyama, and H. Umezawa, Solid State Comm. 34, 19 (1980). [109] H. S. Greenside, E. I. Blount, and C. M. Varma, Phys. Rev. Lett. 46, 49 (1981). [110] V. Jaccarino and M. Peter, Phys. Rev. Lett. 9, 290 (1962). [111] O. Fischer, M. Decroux, S. Roth, R. Chevrel, and M. Sergent, J. Phys. C 8, L474 (1975). [112] H. W. Meul, C. Rossel, M. Decroux, O. Fischer, G. Remenyi, and A. Briggs, Phys. Rev. Lett. 53, 497 (1984). [113] M. Dressel and G. Grüner, Electrodynamics of Solids, Cambridge University Press, Cambridge (2002). [114] M. Fox, Optical properties of solids, Oxford University Press, Oxford (2001). [115] N. L. Wang, H. Hosono, and P. Dai, Iron-based Superconductors, CRC Press, Boca Raton (2013). 178 Bibliography [116] D. N. Basov and T. Timusk, Rev. Mod. Phys. 77, 721 (2005). [117] M. Tinkham, Phys. Rev. 104, 845 (1956). [118] R. A. Ferrell and R. E. Glover, Phys. Rev. 109, 1398 (1958). [119] M. Tinkham and R. A. Ferrell, Phys. Rev. Lett. 2, 331 (1959). [120] D. C. Mattis and J. Bardeen, Phys. Rev. 111, 412 (1958). [121] W. Zimmermann, E. H. Brandt, M. Bauer, E. Seider, and L. Genzel, Physica C 183, 99 (1991). [122] U. S. Pracht, E. Heintze, C. Clauss, D. Hafner, R. Bek, D. Werner, S. Gelhorn, M. Scheffler, M. Dressel, D. Sherman, B. Gorshunov, K. S. Il’in, D. Henrich, and M. Siegel, IEEE Trans. THz Sci. Technol. 3, 269 (2013). [123] M. M. Qazilbash, J. J. Hamlin, R. E. Baumbach, L. Zhang, D. J. Singh, M. B. Maple, and D. N. Basov, Nature Phys. 5, 647 (2009). [124] J. Dong, H. J. Zhang, G. Xu, Z. Li, G. Li, W. Z. Hu, D. Wu, G. F. Chen, X. Dai, J. L. Luo, Z. Fang, and N. L. Wang, Europhys. Lett. 83, 27006 (2008). [125] G. F. Chen, Z. Li, G. Li, J. Zhou, D. Wu, J. Dong, W. Z. Hu, P. Zheng, Z. J. Chen, H. Q. Yuan, J. Singleton, J. L. Luo, and N. L. Wang, Phys. Rev. Lett. 101, 057007 (2008). [126] A. Dubroka, K. W. Kim, M. Rössle, V. K. Malik, A. J. Drew, R. H. Liu, G. Wu, X. H. Chen, and C. Bernhard, Phys. Rev. Lett. 101, 097011 (2008). [127] G. Li, W. Z. Hu, J. Dong, Z. Li, P. Zheng, G. F. Chen, J. L. Luo, and N. L. Wang, Phys. Rev. Lett. 101, 107004 (2008). [128] D. Wu, N. Barišić, P. Kallina, A. Faridian, B. Gorshunov, N. Drichko, L. J. Li, X. Lin, G. H. Cao, Z. A. Xu, N. L. Wang, and M. Dressel, Phys. Rev. B 81, 100512(R) (2010). [129] M. Nakajima, S. Ishida, K. Kihou, Y. Tomioka, T. Ito, Y. Yoshida, C. H. Lee, H. Kito, A. Iyo, H. Eisaki, K. M. Kojima, and S. Uchida, Phys. Rev. B 81, 104528 (2010). [130] N. Barišić, D. Wu, M. Dressel, L. J. Li, G. H. Cao, and Z. A. Xu, Phys. Rev. B 82, 054518 (2010). Bibliography 179 [131] S. J. Moon, A. A. Schafgans, M. A. Tanatar, R. Prozorov, A. Thaler, P. C. Canfield, A. S. Sefat, D. Mandrus, and D. N. Basov, Phys. Rev. Lett. 110, 097003 (2013). [132] F. Pfuner, J. G. Analytis, J.-H. Chu, I. R. Fisher, and L. Degiorgi, Eur. Phys. J. B 67, 513 (2009). [133] E. van Heumen, Y. Huang, S. de Jong, A. B. Kuzmenko, M. S. Golden, and D. van der Marel, Europhys. Lett. 90, 37005 (2010). [134] R. P. S. M. Lobo, Y. M. Dai, U. Nagel, T. Rõõm, J. P. Carbotte, T. Timusk, A. Forget, and D. Colson, Phys. Rev. B 82, 100506(R) (2010). [135] A. Charnukha, O. V. Dolgov, A. A. Golubov, Y. Matiks, D. L. Sun, C. T. Lin, B. Keimer, and A. V. Boris, Phys. Rev. B 84, 174511 (2011). [136] Y. S. Kwon, J. B. Hong, Y. R. Jang, H. J. Oh, Y. Y. Song, B. H. Min, T. Iizuka, S. Kimura, A. V. Balatsky, and Y. Bang, New J. Phys. 14, 063009 (2012). [137] Y. M. Dai, B. Xu, B. Shen, H. H. Wen, J. P. Hu, X. G. Qiu, and R. P. S. M. Lobo, Phys. Rev. B 86, 100501(R) (2012). [138] S. J. Moon, A. A. Schafgans, S. Kasahara, T. Shibauchi, T. Terashima, Y. Matsuda, M. A. Tanatar, R. Prozorov, A. Thaler, P. C. Canfield, A. S. Sefat, D. Mandrus, and D. N. Basov, Phys. Rev. Lett. 109, 027006 (2012). [139] J. N. Hancock, S. I. Mirzaei, J. Gillett, S. E. Sebastian, J. Teyssier, R. Viennois, E. Giannini, and D. van der Marel, Phys. Rev. B 82, 014523 (2010). [140] Y. Imai, F. Nabeshima, D. Nakamura, T. Katase, H. Hiramatsu, H. Hosono, and A. Maeda, J. Phys. Soc. Jpn. 82, 043709 (2013). [141] M. Nakajima, S. Ishida, T. Tanaka, K. Kihou, Y. Tomioka, T. Saito, C. H. Lee, H. Fukazawa, Y. Kohori, T. Kakeshita, A. Iyo, T. Ito, H. Eisaki, and S. Uchida, Sci Rep. 4, 5873 (2014). [142] J. J. Tu, J. Li, W. Liu, A. Punnoose, Y. Gong, Y. H. Ren, L. J. Li, G. H. Cao, Z. A. Xu, and C. C. Homes, Phys. Rev. B 82, 174509 (2010). [143] E. G. Maksimov, A. E. Karakozov, B. P. Gorshunov, A. S. Prokhorov, A. A. Voronkov, E. S. Zhukova, V. S. Nozdrin, S. S. Zhukov, D. Wu, M. Dressel, S. Haindl, K. Iida, and B. Holzapfel, Phys. Rev. B 83, 140502(R) (2011). 180 Bibliography [144] E. G. Maksimov, A. E. Karakozov, A. A. Voronkov, B. P. Gorshunov, S. S. Zhukov, E. S. Zhukova, V. S. Nozdrin, A. S. Prokhorov, S. Haindl, B. Holzapfel, L. Schultz, D. Wu, M. Dressel, K. Iida, P. Kallina, and F. Kurth, JETP 93, 736 (2011). [145] A. Akrap, J. J. Tu, L. J. Li, G. H. Cao, Z. A. Xu, and C. C. Homes, Phys. Rev. B 80, 180502(R) (2009). [146] S. J. Moon, J. H. Shin, D. Parker, W. S. Choi, I. I. Mazin, Y. S. Lee, J. Y. Kim, N. H. Sung, B. K. Cho, S. H. Khim, J. S. Kim, K. H. Kim, and T. W. Noh, Phys. Rev. B 81, 205114 (2010). [147] Z. G. Chen, T. Dong, R. H. Ruan, B. F. Hu, B. Cheng, W. Z. Hu, P. Zheng, Z. Fang, X. Dai, and N. L. Wang, Phys. Rev. Lett. 105, 097003 (2010). [148] B. Gorshunov, D. Wu, A. A. Voronkov, P. Kallina, K. Iida, S. Haindl, F. Kurth, L. Schultz, B. Holzapfel, and M. Dressel, Phys. Rev. B 81, 060509(R) (2010). [149] A. Perucchi, L. Baldassarre, S. Lupi, J. Y. Jiang, J. D.Weiss, E. E. Hellstrom, S. Lee, C. W. Bark, C. B. Eom, M. Putti, I. Pallecchi, C. Marini, and P. Dore, Eur. Phys. J. B 77, 25 (2010). [150] R. V. Aguilar, L. S. Bilbro, S. Lee, C. W. Bark, J. Jiang, J. D. Weiss, E. E. Hellstrom, D. C. Larbalestier, C. B. Eom, and N. P. Armitage, Phys. Rev. B 82, 180514 (R) (2010). [151] T. Fischer, A. V. Pronin, J. Wosnitza, K. Iida, F. Kurth, S. Haindl, L. Schultz, B. Holzapfel, and E. Schachinger, Phys. Rev. B 82, 224507 (2010). [152] Y. A. Aleshchenko, A. V. Muratov, V. M. Pudalov, E. S. Zhukova, B. P. Gorshunov, F. Kurth, and K. Iida, JETP 94, 719 (2012). [153] A. Perucchi, L. Baldassarre, B. Joseph, S. Lupi, S. Lee, C. B. Eom, J. Jiang, J. D. Weiss, E. E. Hellstrom, and P. Dore, Eur. Phys. J. B 86, 274 (2013). [154] K. W. Kim, M. Rössle, A. Dubroka, V. K. Malik, T. Wolf, and C. Bernhard, Phys. Rev. B 81, 214508 (2010). [155] D. Wu, N. Barišić, M. Dressel, G. H. Cao, Z-A. Xu, J. P. Carbotte, and E. Schachinger, Phys. Rev. B 82, 184527 (2010). [156] Y. M. Dai, B. Xu, B. Shen, H. H.Wen, X. G. Qiu, and R. P. S. M. Lobo, Europhys. Lett. 104, 47006 (2013). Bibliography 181 [157] B. B. Jin, T. Dahm, A. I. Gubin, Eun-Mi Choi, Hyun Jung Kim, Sung-IK Lee, W. N. Kang, and N. Klein, Phys. Rev. Lett. 91, 127006 (2003). [158] A. Charnukha , P. Popovich , Y. Matiks, D. L. Sun, C. T. Lin, A. N. Yaresko, B. Keimer, and A. V. Boris, Nature Commun. 2, 219 (2011). [159] J. Yang, D. Hüvonen, U. Nagel, T. Rõõm, N. Ni, P. C. Canfield, S. L. Bud’ko, J. P. Carbotte, and T. Timusk, Phys. Rev. Lett. 102, 187003 (2009). [160] D. Wu, N. Barišić, M. Dressel, G. H. Cao, Z-A. Xu, E. Schachinger, and J. P. Carbotte, Phys. Rev. B 82, 144519 (2010). [161] Y. M. Dai, B. Xu, B. Shen, H. Xiao, H. H. Wen, X. G. Qiu, C. C. Homes, and R. P. S. M. Lobo, Phys. Rev. Lett. 111, 117001 (2013). [162] A. Lucarelli, A. Dusza, F. Pfuner, P. Lerch, J. G. Analytis, J.-H. Chu, I. R. Fisher, and L. Degiorgi, New J. Phys. 12, 073036 (2010). [163] A. A. Schafgans, S. J. Moon, B. C. Pursley, A. D. LaForge, M. M. Qazilbash, A. S. Sefat, D. Mandrus, K. Haule, G. Kotliar, and D. N. Basov, Phys. Rev. Lett. 108, 147002 (2012). [164] N. L. Wang, W. Z. Hu, Z. G. Chen, R. H. Yuan, G. Li, G. F. Chen, and T. Xiang, J. Phys.: Condens. Matter 24, 294202 (2012). [165] Z. G. Chen, R. H. Yuan, T. Dong, and N. L. Wang, Phys. Rev. B 81, 100502(R) (2010). [166] M. Nakajima, T. Tanaka, S. Ishida, K. Kihou, C. H. Lee, A. Iyo, T. Kakeshita, H. Eisaki, and S. Uchida, Phys. Rev. B 88, 094501 (2013). [167] A. Dusza, A. Lucarelli, F. Pruner, J.-H. Chu, I.R. Fisher, and L. Degiorgi, Europhys. Lett. 93, 37002 (2011). [168] Z. P. Yin, K. Haule, and G. Kotliar, Nature Phys. 7, 294 (2011). [169] M. Nakajima, T. Liang, S. Ishida, Y. Tomioka, K. Kihou, C. H. Lee, A. Iyo, H. Eisaki, T. Kakeshita, T. Ito, and S. Uchida, PNAS 108, 12238 (2011). [170] A. Dusza, A. Lucarelli, A. Sanna, J.-H. Chu, I.R. Fisher, and L. Degiorgi, New J. of Phys. 14, 023020 (2012). [171] M. Nakajima, S. Ishida, Y. Tomioka, K. Kihou, C. H. Lee, A. Iyo, T. Ito, T. Kakeshita, H. Eisaki, and S. Uchida, Phys. Rev. Lett. 109, 217003 (2012). 182 Bibliography [172] A. Dusza, A. Lucarelli, J.-H. Chu, I.R. Fisher, and L. Degiorgi, J. Supercond. Nov. Magn. 26, 2603 (2013). [173] C. Mirri, A. Dusza, S. Bastelberger, J.-H. Chu, H.-H. Kuo, I. R. Fisher, and L. Degiorgi, Phys. Rev. B 89, 060501(R) (2014). [174] A. A. Schafgans, B. C. Pursley, A. D. LaForge, A. S. Sefat, D. Mandrus, and D. N. Basov, Phys. Rev. B 84, 052501 (2011). [175] C. Clauß, Optical investigations on the charge ordered state of two dimensional con- ductors, Diplomarbeit am 1. Physikalischen Institut, Stuttgart (2008). [176] J. Merz, Infrarotspektroskopie an Eisenpniktiden, Bachelorarbeit am 1. Physikalis- chen Institut, Stuttgart (2014). [177] A. D. LaForge, A. Frenzel, B. C. Pursley, T. Lin, X. Liu, J. Shi, and D. N. Basov, Phys. Rev. B 81, 125120 (2010). [178] W. G. Jenks, I. M. Thomas, and J. P. Wikswo Jr., EAP 19, 457 (1997). [179] D. Martien, Introduction to AC susceptibility, Quantum Design (2002). [180] K. Iida, J. Hänisch, R. Hühne, F. Kurth, M. Kidszun, S. Haindl, J. Werner, L. Schultz, and B. Holzapfel, Appl. Phys. Lett. 95, 192501 (2009). [181] C. Krellner, N. Caroca-Canales, A. Jesche, H. Rosner, A. Ormeci, and C. Geibel, Phys. Rev. B 78, 100504(R) (2008). [182] S. Jiang, Y. Luo, Z. Ren, Z. Zhu, C. Wang, X. Xu, Q. Tao, G. Cao, and Z. A. Xu, New J. Phys. 11 025007 (2009). [183] J. Herrero-Martín, V. Scagnoli, C. Mazzoli, Y. Su, R. Mittal, Y. Xiao, T. Brückel, N. Kumar, S.K. Dhar, A. Thamizhavel, and L. Paolasini, Phys. Rev. B. 80, 134411 (2009). [184] D. Wu, N. Barišić, N. Drichko, S. Kaiser, A. Faridian, M. Dressel, S. Jiang, Z. Ren, L. J. Li, G. H. Cao, Z. A. Xu, H. S. Jeevan, and P. Gegenwart, Phys. Rev. B 79, 155103 (2009). [185] D. Kasinathan, A. Ormeci, K. Koch, U. Burkhardt, W. Schnelle, A. Leithe-Jasper, and H. Rosner, New J. Phys. 11 025023 (2009). [186] W. Z. Hu, J. Dong, G. Li, Z. Li, P. Zheng, G. F. Chen, J. L. Luo, and N. L. Wang, Front. Phys. China 4, 459 (2009). Bibliography 183 [187] E. Dengler, J. Deisenhofer, H. A. Krug von Nidda, S. Khim, J. S. Kim, K. H. Kim, F. Casper, C. Felser, and A. Loidl, Phys. Rev. B 81, 024406 (2010). [188] T. Terashima, N. Kurita, A. Kikkawa, H. S. Suzuki, T. Matsumoto, K. Murata, and S. Uji, J. Phys. Soc. Jpn. 79, 103706 (2010). [189] J. Koo, J. Park, S. K. Cho, K. D. Kim, S.-Y. Park, Y. H. Jeong, Y. J. Park T. Y. Koo, K.-P. Hong, C.-H. Lee, J.-Y. Kim, B.-K. Cho, K. B. Lee, and H.-J. Kim, J. Phys. Soc. Jpn. 79, 114708 (2010). [190] M. Tokunaga, I. Katakura, N. Katayama, and K. Ohgushi, J. Low Temp. Phys. 159, 601) (2010). [191] S. de Jong, E. van Heumen, S. Thirupathaiah, R. Huisman, F. Massee, J. B. Goed- koop, R. Ovsyannikov, J. Fink, H. A. Dürr, A. Gloskovskii, H. S. Jeevan, P. Gegen- wart, A. Erb, L. Patthey, M. Shi, R. Follath, A. Varykhalov, and M. S. Golden, Europhys. Lett. 89, 27007 (2010). [192] B. Zhou, Y. Zhang, L. X. Yang, M. Xu, C. He, F. Chen, J. F. Zhao, H. W. Ou, J. Wei, B. P. Xie, T. Wu, G. Wu, M. Arita, K. Shimada, H. Namatame, M. Taniguchi, X. H. Chen, and D. L. Feng, Phys. Rev. B 81, 155124 (2010). [193] M. S. Kim, N. H. Sung, Y. Son, M. S. Ko, and B. K. Cho, Appl. Phys. Lett. 98, 172509 (2011). [194] H. Kobayashi, S. Ikeda, Y. Yoda, H. Nakamura, and M. Machida, Phys. Rev. B 84, 184304 (2011). [195] J. Herrero-Martín, C. Mazzoli, V. Scagnoli, L. Paolasini, H. Walker, Y. Xiao, T. Brückel, R. Mittal, N. Kumar, S. K. Dhar, A. Thamizhavel, and Y. Su, J. Supercond. Nov. Mag. 24, 705 (2011). [196] A. Blachowski, K. Ruebenbauer, J. Zukrowski, K. Rogacki, Z. Bukowski, and J. Karpinski, Phys. Rev. B 83, 134410 (2011). [197] J. J. Ying, X. F. Wang, T. Wu, Z. J. Xiang, R. H. Liu, Y. J. Yan, A. F. Wang, M. Zhang, G. J. Ye, P. Cheng, J. P. Hu, and X. H. Chen, Phys. Rev. Lett. 107, 067001 (2011). [198] Y. Xiao, Y. Su, S. Nandi, S. Price, B. Schmitz, C. M. N. Kumar, R. Mittal, T. Chatterji, N. Kumar, S. K. Dhar, A. Thamizhavel, and T. Brückel, Phys. Rev. B 85, 094504 (2012). 184 Bibliography [199] L. Rettig, R. Cortés, S. Thirupathaiah, P. Gegenwart, H. S. Jeevan, M. Wolf, J. Fink, and U. Bovensiepen, Phys. Rev. Lett. 108, 097002 (2012). [200] M. Ikeda, K. Yoshida, and T. Kobayashi, J. Phys. Soc. Jpn.. 81, 033703 (2012). [201] S. K. Mohanta, S. N. Mishra, S. M. Davane, S. Layek, and Z. Hossain, Mod. Phys. Lett. B 27, 1350234 (2013). [202] M. Ikeda, M. Hagiwara, T. Kobayashi, W. Hirata, S. Miyasaka, and S. Tajima, J. Korean Phys. Soc. 62, 2007 (2013). [203] A. Dutta, Anupam, Z. Hossain, and A. K. Gupta, J. Phys.: Condens. Matter 25, 375602 (2013). [204] G. Adhikary, N. Sahadev, D. Biswas, R. Bindu, N. Kumar, A. Thamizhavel, S. K. Dhar, and K. Maiti, J. Phys.: Condens. Matter 25, 225701 (2013). [205] I. Avigo, R. Cortés, L. Rettig, S. Thirupathaiah, H. S. Jeevan, P. Gegenwart, T. Wolf, M. Ligges, M. Wolf, J. Fink, and U. Bovensiepen, J. Phys.: Condens. Matter 25, 094003 (2013). [206] A. V. Tsvyashchenko, V. A. Sidorov, L. N. Fomicheva, K. Gofryk, R. A. Sadykov, and J. D. Thompson, Phys. Status Solidi B 250, 589 (2013). [207] F. Drief, S. Kacimi, and A. Zaoui, Comput. Mater. Sci. 84, 97 (2013). [208] C. W. Kwang-Hua, Chemical Physics 415, 186 (2013). [209] P. F. S. Rosa, B. Zeng, C. Adriano, T. M. Garitezi, T. Grant, Z. Fisk, L. Balicas, R. R. Urbano, and P. G. Pagliuso, arXiv:1402.2412. [210] H. Raffius, E. Mörsen, B. D. Mosel, and W. Müller-Warmuth, J. Phys. Chem. Solids 49, 785 (1988). [211] C. Feng, Z. Ren, S. Xu, S. Jiang, Z. A. Xu, G. Cao, I. Nowik, I. Felner, K. Matsub- ayashi, and Y. Uwatoko, Phys. Rev. B 82, 094426 (2010). [212] D. H. Ryan, J. M. Cardogan, S. Xu, Z. A. Xu, and G. H. Cao, Phys. Rev. B 83, 132403 (2011). [213] L. Hua, J. M. Shen, Q. L. Zhu, and L. Chen, Physica B 406, 47 (2011). [214] J. Ballinger, L. E. Wenger, Y. K. Vohra, and A. S. Sefat, J. Appl. Phys. 111, 07E106 (2012). Bibliography 185 [215] W. H. Jiao, I. Felner, I. Nowik, and G. H. Cao, J. Supercond. Nov. Magn. 25, 441 (2012). [216] H. Raffius, E. Mörsen, B. D. Mosel, W. Müller-Warmuth, W. Jeitschko, L. Ter- büchte, and T. Vomhof, J. Phys. Chem. Solids 54, 135 (1993). [217] M. Reehuis, W. Jeitschko, M. H. Möller, and P. J. Brown, J. Phys. Chem. Solids 53, 687 (1992) . [218] S. Nandi, A. Kreyssig, Y. Lee, Yogesh Singh, J. W. Kim, D. C. Johnston, B. N. Harmon, and A. I. Goldman, Phys. Rev. B 79, 100407 (2009). [219] A. Akbari, P. Thalmeier, and I. Eremin New J. Phys. 15 033034 (2013). [220] T. Terashima, M. Kimata, H. Satsukawa, A. Harada, K. Hazama, S. Uji, H. S. Suzuki, T. Matsumoto, and K. Murata, J. Phys. Soc. Jpn. 78, 083701 (2009). [221] W. Uhoya, G. Tsoi, Y. K. Vohra, M. A. McGuire, A. S. Sefat, B. C. Sales, D. Mandrus, and S. T. Weir, J. Phys.: Condens. Matter 22 292202 (2010). [222] K. Matsubayashi, K. Munakata, N. Katayama, M. Isobe, K. Ohgushi, Y. Ueda, N. Kawamura, M. Mizumaki, N. Ishimatsu, M. Hedo, I. Umehara, and Y. Uwatoko, Phys. Rev. B 84, 024502 (2011). [223] W. Uhoya, G. Tsoi, Y. K. Vohra, M. A. McGuire, and A. S. Sefat, J. Phys.: Condens. Matter 22 292202 (2010). [224] N. Kurita, M. Kimata, K. Kodama, A. Harada, M. Tomita, H. S. Suzuki, T. Mat- sumoto, K. Murata, S. Uji, and T. Terashima, Phys. Rev. B 83, 214513 (2011). [225] N. Kurita, M. Kimata, K. Kodama, A. Harada, M. Tomita, H. S. Suzuki, T. Mat- sumoto, K. Murata, S. Uji, and T. Terashima, Phys. Rev. B 83, 100501(R) (2011). [226] H. B. Banks, W. Bi, L. Sun, G. F. Chen, X. H. Chen, and J. S. Schilling, Physica C 471, 476 (2011). [227] N. Kurita, M. Kimata, K. Kodama, A. Harada, M. Tomita, H. S. Suzuki, T. Mat- sumoto, K. Murata, S. Uji, and T. Terashima, J. Phys.: Conf. Ser. 273, 012098 (2011). [228] N. Kurita, M. Kimata, K. Kodama, A. Harada, M. Tomita, H. S. Suzuki, T. Mat- sumoto, K. Murata, S. Uji, and T. Terashima, J. Phys.: Conf. Ser. 391, 012132 (2012). 186 Bibliography [229] N. Kurita, M. Kimata, K. Kodama, A. Harada, M. Tomita, H. S. Suzuki, T. Mat- sumoto, K. Murata, S. Uji, and T. Terashima, Phys. Rev. B 88, 224510 (2013). [230] H. S. Jeevan, D. Kasinathan, H. Rosner, and P. Gegenwart, Phys. Rev. B 83, 054511 (2011). [231] A. Ahmed, M. Itou, S. Xu, Z. Xu, G. Cao, Y. Sakurai, J. Penner-Hahn, and A. Deb, Phys. Rev. Lett. 105, 207003 (2010). [232] L. Sun, J. Guo, G. Chen, X. Chen, X. Dong, W. Lu, C. Zhang, Z. Jiang, Y. Zou, S. Zhang, Y. Huang, Q. Wu, X. Dai, Y. Li, J. Liu, and Z. Zhao, Phys. Rev. B 82, 134509 (2010). [233] D. Wu, G. Chanda, H. S. Jeevan, P. Gegenwart, and M. Dressel, Phys. Rev. B 83, 100503(R) (2011). [234] I. Nowik, I. Felner, Z. Ren, G. H. Cao, and Z. A. Xu, J. Phys.: Condens. Matter 23, 065701 (2011). [235] S. Thirupathaiah, E. D. L. Rienks, H. S. Jeevan, R. Ovsyannikov, E. Slooten, J. Kaas, E. van Heumen, S. de Jong, H. A. Duerr, K. Siemensmeyer, R. Follath, P. Gegenwart, M. S. Golden, and J. Fink, Phys. Rev. B 84, 014531 (2011). [236] G. Cao, S. Xu, Z. Ren, S. Jiang, C. Feng, and Z. A. Xu, J. Phys.: Condens. Matter 23, 464204 (2011) [237] Z. Guguchia, A. Shengelaya, A. Maisuradze, L. Howald, Z. Bukowski, M. Chikovani, H. Luetkens, S. Katrych, J. Karpinski, and H. Keller, J. Supercon. Nov. Magn. 26, 285 (2012). [238] Y. Tokiwa, S.-H. Hübner, O. Beck, H. S. Jeevan, and P. Gegenwart, Phys. Rev. B 86, 220505(R) (2012). [239] W. Li, J. X. Zhu, Y. Chen, and C. S. Ting, Phys. Rev. B 86, 155119 (2012). [240] S. Jiang, H. S. Jeevan, J. Dong, and P. Gegenwart, Phys. Rev. Lett. 110, 067001 (2013). [241] S. Nandi, W. T. Jin, Y. Xiao, Y. Su, S. Price, D. K. Shukla, J. Strempfer, H. S. Jeevan, P. Gegenwart, and T. Brückel, Phys. Rev. B 89, 014512 (2014). [242] S. Nandi, W. T. Jin, Y. Xiao, Y. Su, S. Price, W. Schmidt, K. Schmalzl, T. Chatterji, H. S. Jeevan, P. Gegenwart, and T. Brückel, Phys. Rev. B 90, 094407 (2014). Bibliography 187 [243] J. Munevar, H. Micklitz, M. Alzamora, C. Argüello, T. Goko, F. L. Ning, T. Munsie, T. J. Williams, A. A. Aczel, G. M. Luke, G.F. Chen, and W. Yu, Solid State Commun. 187, 18 (2014). [244] X. Xu, W. H. Jiao, N. Zhou, Y. K. Li, B. Chen, C. Cao, J. Dai, A. F. Bangura, and G. Cao, Phys. Rev. B. 89, 104517 (2014). [245] W. H. Jiao, J. K. Bao, Q. Tao, H. Jiang, C. M. Feng, Z. A. Xu, and G. H. Cao, J. Phys.: Conf. Ser. 400, 022038 (2012). [246] S. Sharma, A. Bharathi, K. Vinod, C. S. Sundar, V. Srihari, S. Sen, H. Ghosh, A. K. Sinha, and S. K. Deb, arXiv:1312.7055 [247] S. Nandi, M. G. Kim, A. Kreyssig, R. M. Fernandes, D. K. Pratt, A. Thaler, N. Ni, S. L. Bud’ko, P. C. Canfield, J. Schmalian, R. J. McQueeney, and A. I. Goldman, Phys. Rev. Lett. 104, 057006 (2010). [248] J. J. Ying, T. Wu, Q. J. Zheng, Y. He, G. Wu, Q. J. Li, Y. J. Yan, Y. L. Xie, R. H. Liu, X. F. Wang, and X. H. Chen, Phys. Rev. B 81, 052503 (2010). [249] X. B. Chen, Z. A. Ren, H. S. Ding, and L. H. Liu, Sci. China 53, 1212 (2010). [250] M. Nicklas, M. Kumar, E. Lengyel, W. Schnelle, and A. Leithe-Jasper, J. Phys.: Conf. Ser. 273, 012101 (2011). [251] I. Nowik, I. Felner, Z. Ren, G. H. Cao, and Z. A. Xu, New J. Phys. 13, 023033 (2011). [252] Z. Guguchia, J. Roos, A. Shengelaya, S. Katrych, Z. Bukowski, S. Weyeneth, F. Murányi, S. Strässle, A. Maisuradze, J. Karpinski, and H. Keller, Phys. Rev. B 83, 144516 (2011). [253] Z. Guguchia, S. Bosma, S. Weyeneth, A. Shengelaya, R. Puzniak, Z. Bukowski, J. Karpinski, and H. Keller, Phys. Rev. B 84, 144506 (2011). [254] A. Blachowski, K. Ruebenbauer, J. Zukrowski, Z. Bukowski, K. Rogacki, P. J. W. Moll, and J. Karpinski, Phys. Rev. B 84, 174503 (2011). [255] M. Matusiak, Z. Bukowski, and J. Karpinski, Phys. Rev. B 83, 224505 (2011). [256] V. H. Tran, T. A. Zaleski, Z. Bukowski, L. M. Tran, and A. J. Zaleski, Phys. Rev. B 85, 052502 (2012). [257] V. H. Tran, Z. Bukowski, L. M. Tran, and A. J. Zaleski, New J. Phys. 14, 073052 (2012). 188 Bibliography [258] F. A. Garcia, A. Leithe-Jasper, W. Schnelle, M. Nicklas, H. Rosner, and J. Sichelschmidt, New J. Phys. 14, 063005 (2012). [259] W. T. Jin, S. Nandi, Y. Xiao, Y. Su, O. Zaharko, Z. Guguchia, Z. Bukowski, S. Price, W. H. Jiao, G. H. Cao, and T. Brückel, Phys. Rev. B 88, 214516 (2013). [260] C. Rumble, M. Itou, S. Jiang, Z. A. Xu, G. Cao, Y. Sakurai, J. Penner-Hahn, and A. Deb, J. Appl. Phys. 113, 013907 (2013). [261] Z. Ren, X. Lin, Q. Tao, S. Jiang, Z. Zhu, C. Wang, G. Cao, and Z. Xu, Phys. Rev. B 79, 094426 (2009). [262] I. Nowik, I. Felner, Z. Ren, Z. A. Xu, and G. H. Cao, J. Phys.: Conf. Ser. 217, 012121 (2010). [263] W. H. Jiao, H. F. Zhai, J. K. Bao, Y. K. Luo, Q. Tao, C. M. Feng, Z. A. Xu, and G. H. Cao, New J. Phys. 15, 113002 (2013). [264] U. B. Paramanik, P. L. Paulose, S. Ramakrishnan, A. K. Nigam, C. Geibel, and Z. Hossain, Supercond. Sci. Technol. 27, 075012 (2014). [265] L. J. Li, Y. K. Luo, Q. B. Wang, H. Chen, Z. Ren, Q. Tao, Y. K. Li, X. Lin, M. He, Z. W. Zhu, G. H. Cao, and Z. A. Xu, New J. Phys. 11, 025008 (2008). [266] W. Z. Hu, J. Dong, G. Li, Z. Li, P. Zheng, G. F. Chen, J. L. Luo, and N. L. Wang, Phys. Rev. Lett. 101, 257005 (2008). [267] A. Mitsuda, T. Matoba, F. Ishikawa, Y. Yamada, and H. Wada, J. Phys. Soc. Jpn. 79, 073704 (2010). [268] A. Mitsuda, S. Seike, T. Matoba, H. Wada, F. Ishikawa, and Y. Yamada, J. Phys.: Conf. Ser. 273, 012100 (2011). [269] P. F. S. Rosa, C. Adriano, W. Iwamoto, T. M. Garitezi, T. Grant, Z. Fisk, and P. G. Pagliuso, Phys. Rev. B 86, 165131 (2012) [270] Y. He, T. Wu, G. Wu, Q. J. Zheng, Y. Z. Liu, H. Chen, J. J. Ying, R. H. Liu, X. F. Wang, Y. L. Xie, Y. J. Yan, J. K. Dong, S. Y. Li, and X. H. Chen, J. Phys.: Condens. Matter 22, 235701 (2010). [271] R. Hu, S. L. Bud’ko, W. E. Straszheim, and P. C. Canfield, Phys. Rev. B 83, 094520 (2011). Bibliography 189 [272] H. A. Krug von Nidda, S. Kraus, S. Schaile, E. Dengler, N. Pascher, M. Hemmida, M. J. Eom, J. S. Kim, H. S. Jeevan, P. Gegenwart, J. Deisenhofer, and A. Loidl, Phys. Rev. B 86, 094411 (2012). [273] R. Sarkar, R. Nath, P. Khuntia, H. S. Jeevan, P. Gegenwart, and M. Baenitz, J. Phys.: Condens. Matter 24, 045702 (2012). [274] R. Hu, E. D. Mun, M. M. Altarawneh, C. H. Mielke, V. S. Zapf, S. L. Bud’ko, and P. C. Canfield, Phys. Rev. B 85, 064511 (2012). [275] Anupam, P. L. Paulose, H. S. Jeevan, C. Geibel, and Z. Hossain, J. Phys.: Condens. Matter 21, 265701 (2009). [276] V. A. Gasparov, H. S. Jeevan, and P. Gegenwart, JETP Letters 89, 294 (2009). [277] N. Pascher, J. Deisenhofer, H.-A. Krug von Nidda, M. Hemmida, H. S. Jeevan, P. Gegenwart, and A. Loidl, Phys. Rev. B 82, 054525 (2010). [278] H. S. Jeevan and P. Gegenwart, J. Phys.: Conf. Ser. 200, 012060 (2010). [279] Anupam, P. L. Paulose, S. Ramakrishnan, and Z. Hossain, J. Phys.: Condens. Matter 23, 455702 (2011). [280] J. Maiwald, H. S. Jeevan, and P. Gegenwart, Phys. Rev. B 85, 024511 (2012). [281] T. Dey, P. Khuntia, A. V. Mahajan, Anupam, and Z. Hossain, Europ. Phys. J. B 86, 458 (2013). [282] V. A. Gasparov, A. Audouard, L. Drigo, A. I. Rodigin, C. T. Lin, W. P. Liu, M. Zhang, A. F. Wang, X. H. Chen, H. S. Jeevan, J. Maiwald, and P. Gegenwart, Phys. Rev. B 87, 094508 (2013). [283] K. Sasmal, B. Lv, B. Lorenz, A. M. Guloy, F. Chen, Y.-Y. Xue, and C.-W. Chu, Phys. Rev. Lett. 101, 107007 (2008). [284] Y. Qi, Z. Gao, L. Wang, D. Wang, X. Zhang, and Y. Ma, New J. Phys. 10 123003 (2008). [285] Y. Qi, L. Wang, Z. Gao, X. Zhang, D. Wang, C. Yao, C. Wang, C. Wang, and Y. Ma, New J. Phys. 14, 033011 (2012). [286] Anupam, V. K. Anand, P. L. Paulose, S. Ramakrishnan, C. Geibel, and Z. Hossain, Phys. Rev. B 85, 144513 (2012). 190 Bibliography [287] Y. R. Jang, J. B. Hong, B. H. Min, M. A. Jung, Y. Y. Song, H. J. Oh, K. J. Lee, M. H. Jung, S. Kanematsu, K. Imura, K. Deguchi, N. K. Sato, and Y. S. Kwon, Supercond. Sci. Technol. 24, 085017 (2011). [288] T. Wu, G. Wu, H. Chen, Y. L. Xie, R. H. Liu, X. F. Wang, and X. H. Chen, J. Magn. Magn. Mater 321, 3870 (2009). [289] F. A. Garcia, E. M. Bittar, C. Adriano, T. M. Garitezi, C. Rettori, and P. G. Pagliuso, J. Phys.: Conf. Ser. 273, 012093 (2011). [290] B. Zhou, L. X. Yang, F. Chen, M. Xu, T. Wu, G. Wu, X. H. Chen, and D. L. Feng, J. Phys. Chem. Solids 72, 474 (2011). [291] M. Zhang, J. J. Ying, Y. J. Yan, A. F. Wang, X. F. Wang, Z. J. Xiang, G. J. Ye, P. Cheng, X. G. Luo, J. Hu, and X. H. Chen, Phys. Rev. B 85, 092503 (2012). [292] T. Moriya and K. Ueda, Adv. in Phys. 49, 555 (2000). [293] P. Richard, C. Capan, J. Ma, P. Zhang, N. Xu, T. Qian, J. D. Denlinger, G.-F. Chen, A. S. Sefat, Z. Fisk, and H. Ding, J. Phys.: Condens. Matter 26, 035702 (2014). [294] M. Xia, W. H. Jiao, Z. R. Ye, Q. Q. Ge, Y. Zhang, J. Jiang, R. Peng, X. P. Shen, Q. Fan, G. H. Cao, T. Zhang, B. P. Xie, and D. L. Feng, J. Phys.: Condens. Matter 26, 265701 (2014). [295] X. Xu, W. H. Jiao, N. Zhou, Y. K. Li, B. Chen, C. Cao, J. Dai, A. F. Bangura, and G. Cao, Phys. Rev. B 89, 104517 (2014). [296] F. Klingert, Magnetic Investigations on EuFe2As2 - Superconductors, Diplomarbeit am 1. Physikalischen Institut, Stuttgart (2012). [297] B. Maji, K. G. Suresh and A. K. Nigam, J. Phys.: Condens. Matter 23, 506002 (2011). [298] A. C. D. van Enter and J. L. van Hemmen, Phys. Rev. B 31, 603 (1985). [299] K. Jonason, J. Mattsson, and P. Nordblad, Phys. Rev. B 53, 6507 (1996). [300] A. M. Campbell and J. E. Evetts, Adv. i. Phys. 21, 199 (1972). [301] J. Giapintzakis, R. L. Neiman, D. M. Ginsberg, and M. A. Kirk, Phys. Rev. B 50, 16001 (1994). [302] M. C. Frischherzt, F. M. Sauerzopft, H. W. Webert, M. Murakamit, and G. A. Emel’chenko, Supercond. Sci. Technd. 8, 485 (1995). Bibliography 191 [303] G. Prando, P. Carretta, R. De Renzi, S. Sanna, A. Palenzona, M. Putti, and M. Tropeano, Phys. Rev. B 83, 174514 (2011). [304] W. Braunisch, N. Knauf, V. Kataev, S. Neuhausen, A. Grütz, A. Kock, B. Roden, D. Khomskii, and D. Wohlleben, Phys. Rev. Lett. 68, 1908 (1992). [305] F. Pfister, Anisotroper Transport in eisenbasierten Supraleitern, Bachelorarbeit am 1. Physikalischen Institut, Stuttgart (2013). [306] M. Colarieti-Tosti, S. I. Simak, R. Ahuja, L. Nordström, O. Eriksson, D. Åberg, S. Edvardsson, and M. S. S. Brooks, Phys. Rev. Lett. 91, 157201 (2003). [307] L. Bauernfeind, W. Widder, and H. F. Braun, Physica C 254, 151 (1995). [308] C. A. Cardoso, F. M. Araujo-Moreira, V. P. S. Awana, E. Takayama-Muromachi, O. F. de Lima, H. Yamauchi, and M. Karppinen, Phys. Rev. B 67, 020407(R) (2003). [309] S. Garcia, L. Ghivelder, and I. Felner, Phys. Rev. B 79, 054506 (2009). [310] S. Lausberg, J. Spehling, A. Steppke, A. Jesche, H. Luetkens, A. Amato, C. Baines, C. Krellner, M. Brando, C. Geibel, H.-H. Klauss, and F. Steglich, Phys. Rev. Lett. 109, 216402 (2012). [311] Private communication with Prof. J. Fink, IFW Dresden. [312] Private communication with Prof. J. Schmalian, KIT. [313] Private communication with Prof. I. Fisher, Stanford. [314] U. Pracht, THz Spectroscopy on Quantum Phase Transition in Ultra-Thin Films of Titanium Nitride, Masterarbeit am 1. Physikalischen Institut, Stuttgart (2012). [315] S. Zapf, B. Gorshunov, D. Wu, E. Zhukova, V. S. Nozdrin, S. Haindl, K. Iida, and M. Dressel, J. Supercond. Nov. Magn. 21, 1557 (2013). [316] A. S. Prokhorov, E. S. Zhukova, A. A. Boris, I. E. Spektor, B. P. Gorshunov, V. S. Nozdrin, E. A. Motovilova, L. S. Kadyrov, S. Zapf, S. Haindl, K. Iida, M. Dressel, K. V. Chizh, M. S. Storozhevykh, L. V. Arapkina, V. A. Chapnin, O. V. Uvarov, V. P. Kalinushkin, and V. A. Yuryev, Radiophys. Quant. Electr. 56, 620 (2014). [317] A. E. Karakozov, S. Zapf, B. Gorshunov, Y. G. Ponomarev, M. V. Magnitskaya, E. Zhukova, A. S. Prokhorov, V. B. Anzin, and S. Haindl, Phys. Rev. B 90, 014506 (2014). 192 Bibliography [318] A. E. Karakozov, B. P. Gorshunov, Y. G. Ponomarev, A. S. Prokhorov, M. V. Magnitskaya, E. Zhukova, K. Iida, M. Dressel, S. Zapf, and S. Haindl, JETP Letters 100, to be translated (2014). [319] D. V. Efremov, M. M. Korshunov, O. V. Dolgov, A. A. Golubov, and P. J. Hirschfeld, Phys. Rev. B 84, 180512(R) (2011). [320] D. V. Efremov, A. A. Golubov, and O. V. Dolgov, New J. Phys. 15, 013002 (2013). [321] C. Tarantini, M. Putti, A. Gurevich, Y. Shen, R. K. Singh, J. M. Rowell, N. Newman, D. C. Larbalestier, P. Cheng, Y. Jia, and H.-H. Wen, Phys. Rev. Lett. 104, 087002 (2010). [322] M. Schilling, THz investigation on iron-based superconductors, Masterarbeit am 1. Physikalischen Institut, Stuttgart (2014). [323] A. Baumgartner, THz Untersuchungen an eisenbasierten Supraleitern, Bachelorar- beit am 1. Physikalischen Institut, Stuttgart (2013). CURRICULUM VITAE Education 1996–2005 Friedrich-Dessauer-Gymnasium Aschaffenburg 2005–2010 Julius-Maximilians-Universität Würzburg, elite program “FOKUS Phy- sik” since 06/2010 PhD project, Prof. Dressel, University of Stuttgart Experience 09/2007–10/2007 MPI FKF Stuttgart, Department Kern, Dr. Vogelgesang, Bache- lor thesis: Wavelength Dependent Optoelectronic Characterization of Carbon Nanotube Based Field Effect Transistors 08/2008–09/2008 University of Würzburg, EP3 Molenkamp, Prof. Geurts, internship: Infrared-Spectroscopy on GdMnO3 01/2009–04/2009 UC Berkeley, Department of Chemistry, Prof. Pines, co-worker on the project: Imaging a Temperature Gradient via NMR 04/2009-04/2010 University of Würzburg, EP3 Molenkamp, Prof. Geurts, Master thesis: Spinflip-Raman-Spectroscopy on CdMnTe / CdMgZnTe - Quantum Well Structures since 06/2010 University of Stuttgart, 1st Physical Institute, Prof. Dressel: infrared and THz spectroscopy, SQUID magnetometry, dc resistivity, teach- ing (laboratory courses), supervision of four Bachelor and two Master students 04/2013-09/2013 Several visits at the UC San Diego, Department of Physics, Prof. Basov: magneto-optical investigations on EuFe2As2 194 Curriculum Vitae Selected scholarships and Awards 09/2010–04/2012 Scholarship of the Postgraduate Research Grants Program of Baden- Württemberg 03/2012 Best Poster Award at the International Conference on Magnetism and Superconductivity, Istanbul 03/2014 Düker Award for the best scientific presentation of a research project for school students DECLARATION OF ORIGINALITY I hereby declare that this thesis and the work reported herein was composed by and originated entirely from me. Information derived from the published and unpublished work of others has been acknowledged in the text and references.