KONRAD G I L B ERT HE I N R I C H V I E B AHN QUAS I C RY S TA L L I N E OP T I C A L L AT T I C E S FOR U LTRACO LD ATOMS T H I S D I S S ERTAT I ON I S SUBM I T T ED FOR THE DEGREE OF DOCTOR OF PH I L O SOPHY. T R I N I T Y CO L L EGE , OC TOBER 2 0 1 8 CAVEND I S H L A BORATORY, U N I V ERS I T Y O F CAMBR I DGE Copyright © 2018 K.G.H. Viebahn “ T H E MOS T E S S EN T I A L PART O F A L I V I NG CE L L – T H E CHROMOSOME F I - B R E – MAY SU I TA B LY B E CA L L ED AN APER I OD I C CRYS TA L . I N P H Y S I C S WE HAVE DEA LT H I T H ERTO ON LY W I T H P ER I OD I C CRYS TA L S . TO A HUM - B L E PHYS I C I S T ’ S M I N D, T H E S E ARE VERY I N T ERES T I NG AND COMP L I - C AT ED OB J EC T S ; T H E Y CONS T I T U T E ONE OF THE MOS T FA SC I N AT I NG AND COMP L E X MAT ER I A L S TRUC TURES BY WH I CH I N AN I M AT E NATURE PU Z Z L E S H I S W I T S . Y E T, COMPARED W I T H THE APER I OD I C CRYS TA L , T H E Y ARE RATHER P L A I N AND DU L L . T H E D I F F ERENCE I N S TRUC TURE I S O F THE SAME K I N D A S THAT B E TWEEN AN ORD I N ARY WA L L PA P ER I N WH I CH THE SAME PAT T ERN I S R E P E AT ED AGA I N AND AGA I N I N R EGU L AR PER I OD I C I T Y AND A MAS T ERP I E C E O F EMBRO I D ERY, S AY A RAPHAE L TA P E S TRY, WH I C H SHOWS NO DU L L R E P E T I T I O N , BU T AN E L A BORAT E , COHEREN T, M E AN I NG FU L DES I GN TRACED BY THE GREAT MAS T ER .” E RW I N SCHRÖD I NGER , “WHAT I S L I F E ? ” Declaration This dissertation is the result of my own work and includes nothing which is the outcome of work done in collaboration except as declared in the Preface and specified in the text. It is not substantially the same as any that I have submitted, or, is being concurrently submit- ted for a degree or diploma or other qualification at the University of Cambridge or any other University or similar institution except as declared in the Preface and specified in the text. I further state that no substantial part of my dissertation has already been submitted, or, is being concurrently submitted for any such degree, diploma or other qualification at the University of Cambridge or any other University or similar institution except as declared in the Preface and specified in the text. This thesis does not exceed the word limit of sixty thousand words, including abstract, tables, footnotes and appendices, set out by the Faculty of Physics and Chemistry. Abstract Quasicrystals are long-range ordered and yet non-periodic. This interplay results in a wealth of intriguing physical phenomena, such as the inheritance of topological properties from higher dimensions, self-similarity, and the presence of non-trivial structure on all scales. The concept of aperiodic order has been extensively studied in mathematics and geometry, exemplified by the celebrated Penrose tiling. However, the understanding of physical quasicrystals (the vast major- ity of them are intermetallic compounds) is still incomplete owing to their complexity, regarding both growth processes and stability. Ultracold atoms in optical lattices offer an ideal, yet untested environment for investigating quasicrystals. Optical lattices, i.e. standing waves of light, allow the defect-free formation of a large variety of potential landscapes, including quasiperiodic geometries. In recent years, optical lattices have become one of the most successful tools in the large-scale quantum simulation of condensed-matter problems. This study presents the first experimental realisation of a two-dimensional quasicrystalline po- tential for ultracold atoms, based on an eightfold symmetric optical lattice. It is aimed at bring- ing together the fields of ultracold atoms and quasicrystals – and the more general concept of aperiodic order. The first part of this thesis introduces the theoretical aspects of aperiodic order and quasicrystalline structure. The second part comprises a detailed account of the newly de- signed apparatus that has been used to produce quantum-degenerate gases in quasicrystalline lattices. The third and final part summarises the matter-wave diffraction experiments that have been performed in various lattice geometries. These include one- and two-dimensional simple cubic lattices, one-dimensional quasiperiodic lattices, as well as two-dimensional quasicrys- talline lattices. The striking self-similarity of this quasicrystalline structure has been directly observed, in close analogy to Shechtman’s very first discovery of quasicrystals using electron diffraction. In addition, an in-depth study of the diffraction dynamics reveals the fundamental differences between periodic and quasicrystalline lattices, in excellent agreement with ab initio theory. The diffraction dynamics on short timescales constitutes a continuous-time quantum walk on a homogeneous four-dimensional tight-binding lattice. On the one hand, these measurements establish a novel experimental platform for investigating quasicrystals proper. On the other hand, ultracold atoms in quasicrystalline optical lattices are worth studying in their own right: Possible avenues include the observation many-body local- isation and Bose glasses, as well as the creation of topologically non-trivial systems in higher dimensions. Contents Introduction: quasicrystalline potentials for ultracold atoms 11 Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 1 Quasicrystals and aperiodic order 15 1.1 What is aperiodic order? What is a quasicrystal? . . . . . . . . . . . . . . . . . . . . . . 15 1.2 Fibonacci chain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 Self-similarity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 Diffraction patterns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 Cut-and-project . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 Phasons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 1.3 Quasicrystals in two dimensions, the octagonal tiling . . . . . . . . . . . . . . . . . . . 26 Crystallographic restriction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 Self-similarity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 Cut-and-project . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 Phasons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 1.4 Physical quasicrystals and current state of the art . . . . . . . . . . . . . . . . . . . . . 31 2 Design of the apparatus 35 2.1 Suitable constituents for creating quantum many-body systems . . . . . . . . . . . . . 35 2.2 Choice of atomic species . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 2.3 Vacuum apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 2.4 Overview of the experimental sequence . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 3 Laser cooling & trapping of 87Rb & 39K 41 3.1 The alkalis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 Rubidium-87 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 Potassium-39 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 3.2 87Rb & 39K in magnetic fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 Interlude: calibration of magnetic coils . . . . . . . . . . . . . . . . . . . . . . . . 43 3.3 Spectroscopy and absolute frequency stability (laser locking) . . . . . . . . . . . . . . 44 3.4 Relative frequency stability (offset locking) . . . . . . . . . . . . . . . . . . . . . . . . . 46 Rubidium repump lock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46 Potassium imaging lock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 3.5 Magneto-optical trap (MOT) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 3.6 2D MOT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 Fluorescence signal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 3.7 Final MOT stage in potassium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 3.8 Optical molasses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 MOT compensation coils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 3.9 Optical pumping (spin-polarisation) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 10 3.10 Magnetic trapping, transport, and experiment coils . . . . . . . . . . . . . . . . . . . . 53 Technical implementation of transport and experiment coils . . . . . . . . . . . . 55 Compensation coils (offset fields) . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 3.11 Microwave (MW) evaporation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 Majorana losses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 3.12 Absorption imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 High-field imaging in 39K . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 Stern-Gerlach imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 4 Quantum gases 61 4.1 Dipole trap: introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 Choice of trapping geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 Trapping potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 4.2 Trap frequencies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65 4.3 Evaporation and sympathetic cooling . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 4.4 Dipole trap laser setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 4.5 Dipole loading, guide field, and RF state transfer . . . . . . . . . . . . . . . . . . . . . 68 RF antenna . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 4.6 Feshbach fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 4.7 Quantum degeneracy and Bose-Einstein condensation . . . . . . . . . . . . . . . . . . 70 Atom number calibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72 5 Eightfold optical lattice 73 5.1 1D optical lattice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 5.2 Time-evolution and lattice depth calibration . . . . . . . . . . . . . . . . . . . . . . . . 76 5.3 Eightfold optical lattice potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 Incoherent superposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 Relation to octagonal tiling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80 Phase control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 Fully-interfering axes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 Partly-interfering axes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 5.4 Setup and characterisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85 Control loops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 6 Matter-wave diffraction experiment 91 6.1 Basis in momentum space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 6.2 Method: Kapitza-Dirac diffration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 6.3 Diffraction dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 6.4 Quantum walks in high-dimensional homogeneous tight-binding lattices . . . . . . . 97 6.5 Data analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 Summary and outlook 101 Next steps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 A List of components 105 Acknowledgements 109 Bibliography 113 Introduction: quasicrystalline potentials for ultracold atoms Quasicrystals are long-range ordered without being periodic [1–4]. Their order manifests itself in sharp diffraction peaks, exactly as in their periodic counterparts. However, diffraction patterns from quasicrystals often reveal rotational symmetries, most notably fivefold, eightfold, and tenfold, that are incompatible1 with trans- 1 crystallographically ‘forbidden’ lational symmetry. Therefore it immediately follows that long- range order in quasicrystals results not from a periodic arrange- ment of unit cells but from a different paradigm. Quasicrystalline order can be seen as an incommensurate projection of a higher- dimensional periodic lattice; this enables investigation of physics of higher dimensions, in particular in the context of topology [5– 7]. For instance, one-dimensional (1D) quasiperiodic models, such as the Fibonacci chain and the Aubry-Andre model, are closely connected to the celebrated two-dimensional Harper model [6]. An alternative approach to constructing quasicrystals was de- scribed by Penrose [8] who discovered a pentagonal tiling. Using a particular set of tiles and associated matching rules2 it is possible 2 imagine a puzzle game to cover a plane with an infinite tiling, in which non-periodicity is enforced by the matching rules [9]. The fivefold symmetric Pen- rose tiling and the closely related eightfold symmetric Ammann- Beenker tiling [10] (also called ‘octagonal tiling’ [9]) have become paradigms of 2D quasicrystals [9, 11]. In addition to their dis allowed rotational symmetries, these tilings have the remarkable feature of being self-similar in both real and reciprocal space [3]. Self-similarity upon scaling in length by a certain factor (the silver mean 1 + √ 2 in case of the octagonal tiling) implies that non- trivial structure is present on arbitrarily large scales. Correspond- ingly, diffraction patterns from quasicrystals display sharp peaks at arbitrarily small momenta. To date, quasicrystals have been extensively studied in material science, condensed-matter, and crystallography [11, 12]. Various quasicrystalline model systems have been created, most notably in photonic systems [13] and optical lattices [14–17]. Optical lat- tices are particularly adept for emulating quasicrystals since high rotational symmetries can easily be imposed [17]. However, this advantage has yet remained unexplored in the field of quantum gases. 12 Ultracold gases in optical lattices Quantum gases of neutral atoms are one of today’s most ex- citing experimental frontiers in quantum many-body physics. Sparked by the observation of a quantum phase transition us- ing ultracold atoms in an optical lattice [18] there has been much progress towards Feynman’s original proposal of a quantum sim- ulator [19]. A quantum simulator is a machine that solves numer- ically intractable quantum-mechanical problems3 by mapping3 such as the dynamics of the doped fermionic Hubbard model them to another quantum mechanical system with experimen- tally accessible observables. Quantum gases are ideally suited for this application due to the exquisit control of external parame- ters, such as dimensionality, type and strength of interaction, and geometry.4 The current experimental realisations of ‘quantum4 both in lattices and in bulk systems simulators’ [20, 21] have implemented a multitude of theoretically relevant models, in which numerical simulations are either very hard or impossible [22–25], particularly in dimensions higher than one. But there is more: quantum simulation is just one interpreta- tion of quantum gas experiments. Ultracold atoms have recently been used to create previously unseen phases of matter, such as supersolids [26, 27] and unitary Bose gases [28]. Quantum gases in lattices offer a controllable and clean environment to probe novel systems, such as Floquet-engineered hamiltonians [29]. Therefore, these novel quantum systems are worth pursuing in their own right and it is expected that they play an important role in future applications of quantum technology [21, 30]. Scope The experimental apparatus described in this thesis is aimed at bridging the two fields of quantum gases and quasicrystals, thereby creating a new platform to study aperiodic order with ul- tracold atoms. Drawing from established techniques and concepts in both fields, there are numerous directions in which experimen- tation can progress. In particular, the ability to create arbitrarily complex quantum systems by employing aperiodic order [31], and the possibility of inheriting topological properties from higher dimensions [6] are most promising. Outline This thesis offers a comprehensive review of all experimental methods that were used to create the first quasicrystalline optical lattice for ultracold atoms. A very brief overview of the whole ex- perimental apparatus is given in Chapter 2. The methods chapters focus on initial cooling from room temperature to the µK regime (Chapter 3), reaching quantum degeneracy in the optical dipole trap (Chapter 4), and the setup of the eightfold symmetric optical 13 lattice (Chapter 5). Afterwards, the main experimental outcomes of this work are discussed in Chapter 6, namely matter-wave ex- periments. They highlight the self-similarity of our quasicrys- talline optical lattice, and its ability to simulate quantum walks in higher dimensions. Initially, quasicrystals in one and two dimensions will be in- troduced (Chapter 1), thereby summarising key concepts such as aperiodicity, self-similarity, non-periodic diffraction patterns, the cut-and-project scheme, phasonic excitations and the role of ro- tational symmetries. The current state-of-the-art in quasicrystal research is reviewed at the end of Chapter 1. Contributions The apparatus as it is described in this thesis was jointly designed, implemented, and characterised by the author and Matteo Sbroscia, under supervision of Dr. Ulrich Schneider. Major contributions were made by Dr. Stephen Topliss (electronics engineer), Hendrik von Raven, Oliver Brix, Edward Carter, Max Melchner, Michael Höse, and Jr-Chiun Yu. 1 Quasicrystals and aperiodic order The aim of this chapter is to provide an accessible introduction to the topic of quasicrystals and the concept of aperiodic order. It will form the theoretical basis for the experimental results (Chap- ter 6). On the path towards understanding quasicrystals we will follow these steps: 1.1 What is a quasicrystal? The definition of quasicrystals. The notion of aperiodic order. 1.2 A pedagogical example of a one-dimensional quasicrystal: the Fibonacci chain. In this section two important concepts will be introduced: self-similarity, and the cut-and-project method. 1.3 Quasicrystals in two dimensions. Crystallographic restriction and the role of rotational symmetries in quasicrystals. The octagonal aperiodic tiling (or Ammann-Beenker tiling). 1.4 Quasicrystalline materials. What is known about physical quasicrystals? Classification of quasicrystals and open ques- tions in quasicrystal research. What could our experiment con- tribute to these questions? Figure 1.1: Is this a crystal? The image shows a single grain of icosahedral Zn-Mg-Dy. Courtesy M. Feuerbacher, Forschungszentrum Jülich. 1.1 What is aperiodic order? What is a quasicrystal? What is a black hole? What is a magnetic field? What is an atom? Ask these questions to a group of physicists and you will get more or less consistent answers.1 Of course, the depths of research are 1 If you do not have a group of physi- cists at hand, Wikipedia will do. always full of intricacies and, consequently, the answers to some of these questions may depend on context. But overall, a quick look into standard literature will bring clarity. Defining quasicrystals is much trickier. First, this is due to some initial confusion after the discovery of quasicrystals [12]. Second, literature on quasicrystals is scattered over many different fields,2 2 see ref. [9] for a comprehensive list including crystallography [11], condensed-matter physics [12], and mathematics [9]. Third, even after 35 years of research on quasicrystals and their descendants, there are still many open questions [11]. Finally, there has been a paradigm shift in the community from ‘quasicrystals’ to the more general concept of ‘aperiodic order’ [9, 12]. We will get back to this at the end of the chapter. 16 quasicrystalline optical lattices for ultracold atoms Figure 1.2: Crystallographic definition of crystals and quasicrystals [32]. The discovery of quasicrystals contributed the distinction between periodic and aperiodic crystals to this classification. Up until then order had been thought to be synonymous with periodicity. periodic disordered ordered (= sharp diffraction peaks) aperiodic periodic crystals amorphous solids quasicrystals aperiodic crystals incommensurate crystals modulated compound magnetic Let us start by defining the terms ‘crystal’ and ‘aperiodic crys- tal’ according to the International Union of Crystallography (IUCr) [33, 34]. Definition. In the following by ‘crystal’ we mean any solid having an essentially discrete diffraction diagram, and by ‘aperiodic crystal’ we mean any crystal in which three-dimensional lattice periodicity can be considered to be absent. Although the term ‘essentially’ is deliberately vague,3 the over-3 “The word essentially means that most of the intensity of the diffraction is con- centrated in relatively sharp Bragg peaks, besides the always present diffuse scatter- ing.” [34, 35] all message is clear: long-range order – in this case embodied by crystalline solids – is defined by sharp diffraction peaks. The as- sociation of ‘order’ with ‘sharp diffraction peaks’ was introduced in the wake of the discovery of quasicrystals, which proved that the previous paradigm (‘order⇔ periodicity’) had been false. After some initial debate (see ref. [12] for a review) a paradigm shift took place in the community of crystallography. Nowadays there is broad consensus about the definition of crystals by the ex- istence of sharp diffraction peaks. There can be no doubt that the images shown in Fig. 1.3, which were obtained from a single grain of Zn-Mg-Dy (Fig. 1.1), are indeed signatures of a type of crystal. However, the image in Fig. 1.3C displays a tenfold rotational symmetry which is forbidden for periodic crystals.4 Therefore this4 we will prove this later material must be aperiodic and ordered. We will expand more on the notion of aperiodic order in section 1.2. What about quasicrystals then? Are they the same as aperiodic ordered structures? The IUCr dictionary offers two different defi- nitions of a quasicrystal, with a varying degree of overlap with the class of aperiodic ordered materials [32]. Definition A. A quasicrystal is an aperiodic crystal that is not an in- commensurate modulated crystal, nor an incommensurate composite crystal. Often, quasicrystals have crystallographically ‘forbidden’ sym- metries. These are rotations of order different from 2, 3, 4 and 6. This definition explicitly sets quasicrystals apart from two other classes of aperiodic crystals, namely ‘incommensurate modulated crystals’ and ‘incommensurate composite crystals’, as shown in Fig. 1.2. Incommensurate modulated crystals represent crystals with reciprocal lattices that include at least two incommensu- rate wave vectors [36]. Incommensurate composite crystals are quasicrystals and aperiodic order 17 CBA Figure 1.3: Electron diffraction images of icosahedral Zn-Mg-Dy (Fig. 1.1) along a twofold- (A), threefold- (B), and fivefold- (C) rotationally sym- metric axis. Note that the patterns in (B) and (C) appear to be sixfold- and tenfold-symmetric, respectively. This is a consequence of the diffraction intensity being proportional to the square of the modulus of the structure factor, making reflections with oppo- site indices (h, k, l) and (h¯, k¯, l¯) appear equally strong. This effect is known as ‘Friedel’s law’ [38]. Images: courtesy M. Feuerbacher, Forschungszentrum Jülich. structures made of at least two subcrystals with incommensu- rate lattice vectors [37]. According to this definition, the set of quasicrystals and the set of incommensurate crystals (ICs) are disjoint, both being subsets of the set of aperiodic crystals. There are overlaps between the subsets within ICs (and a third kind of ICs, namely ‘incommensurate magnetic structures’ in which the size of the structural unit cell is incommensurate with the size of the magnetic unit cell). However, the distinction between ICs and quasicrystals in the one-dimensional case is non-trivial, as we will see later by studying diffraction patterns of 1D aperiodic lattices. Therefore the IUCr offers an alternative definition [32, 35]: Definition B. The term quasicrystal stems from the property of quasiperi- odicity observed for the first alloys found with forbidden symmetries. Therefore, the alternative definition is: a quasicrystal is an aperiodic crystal with diffraction peaks that may be indexed by n integral indices, where n is a finite number, larger than the dimension of the space (in general). This definition is similar to that of aperiodic crystal. ‘Indexing’ here means that the position of each diffraction peak can be written as a unique set of integers5 times a set of recipro- 5 or a single integer in periodic 1D latticescal lattice vectors. In other words, it means ‘count the minimum number of integers you need to uniquely identify all the diffrac- tion peaks of a D-dimensional crystal. If this number is finite and larger than D, then the crystal is a quasicrystal.’ This classification is summarised in table 1.1. Here there is essentially no distinction between quasicrystals and aperiodic crystals because the index counting method does not distinguish these cases. This definition of quasicrystals is more mathematical and, consequently, slightly # integers needed (n) class of material n uncountably infinite → disordered, amorphous n countably infinite → almost periodic crystal n finite n > D → aperiodic crystal n = D → periodic crystal Table 1.1: Alternative classification of crystals according to the number n of integers needed to index the diffrac- tion peaks, or equivalently, the mini- mum number of (Fourier-)components needed to write the atomic density distribution as a superposition of plane waves. The class of ‘almost peri- odic crystals’ is largely theoretical, in analogy with the mathematical class of ‘almost periodic functions’ [39], since they have not been observed in experiments [35]. 18 quasicrystalline optical lattices for ultracold atoms more tangible, compared to definition A. It has a certain appeal since it lists periodic crystals as natural ‘extension’ of aperiodic crystals in the case n = D (number of integers needed to index the diffraction pattern equal to the spatial dimension of the crystal). Note that there is no distinction between the terms ‘quasicrys- tal’ and ‘quasiperiodic crystal’; the two can be used interchange- ably [35, 40]. However, the adjective ‘quasiperiodic’ often refers to a special class of mathematical functions called ‘almost periodic functions’ [11, 39, 41].66 The definition of almost periodic functions goes back the Danish mathe- matician Harald Bohr, brother of Niels Bohr, and winner of an olympic silver medal in football (1908). The question of rotational symmetries is a subtle one because neither of the definitions list them as a necessary requirement. However, due to crystallographic restriction (section 1.3) high ro- tational symmetries in ordered structures are a sufficient criterion for classification as quasicrystal. For simplicity, we can assume two-dimensional quasicrystals to have ‘forbidden’ rotational sym- metries in this work7 since the eightfold optical lattice obviously7 thereby ignoring the existence of a few aperiodic structures without forbidden symmetry [11, 12] fulfils the criterion. Since the precise definition of quasicrystals is tricky, we shall simply use the term ‘quasicrystal’ for ordered structures with forbidden symmetries,8 whereas the broader con- 8 and also for 1D quasiperiodic struc- tures, for historical reasons cept of ‘aperiodic order’ shall include quasiperiodic structures without forbidden symmetries, as well as quasicrystals. In order to build some intuition on these abstract definitions of quasicrystals, we will now take the Fibonacci chain as an exam- ple. All of the concepts that are introduced in section 1.2 for one dimension will be generalised to two dimensions in section 1.3. 1.2 Fibonacci chain The Fibonacci chain is the archetypical manifestation of aperiodic order9 in one dimension. We construct the Fibonacci chain by9 and quasicrystallinity repeatedly acting with the following substitution rule L → LS (1.1) S → L on a ‘seed’ letter L. Equation 1.1 means that one letter L is to be replaced by a sequence of two letters (LS). An alternative formula- tion of the substitution rule is( L S ) → ( 1 1 1 0 )( L S ) , (1.2) to the same effect that the sum in the matrix multiplication is to be understood as the concatenation of two letters (L + S → LS). This procedure leads to an arbitrarily long sequence of letters L→ LS→ LSL→ LSLLS→ LSLLSLSL→ LSLLSLSLLSLLS→ . . . which is called a Fibonacci chain (or ‘Fibonacci word’). In a dia- grammatic representation the letters L and S can be understood as long and short line segments, respectively; in this representation the vertices between line segments correspond to lattice points. quasicrystals and aperiodic order 19 Table 1.2: Number of letters in the Fibonacci word after m substitution steps. This sequence of numbers has been known for a very long time, at least since the 13th century [42]; nowadays these numbers are known as Fibonacci numbers. m 1 2 3 4 5 6 7 NS(m) 0 1 1 2 3 5 8 NL(m) 1 1 2 3 5 8 13 The occurrence of S and L letters (NS and NL, respectively) after m substitution steps is listed in table 1.2 from which we recognise the famous Fibonacci numbers. We denote the mth Fibonacci num- ber by Fm such that NS(m) = Fm−1 and NL(m) = Fm. Moreover, the total number of letters is then Ntot = NS(m) + NL(m) = Fm+1. In the limit of m → ∞ the ratio between two subsequent Fibonacci numbers Fm and Fm+1 is [12] lim m→∞ Fm+1 Fm = 1+ 1 1+ 1 1+ 11+... = τ , (1.3) where τ is an irrational number called the ‘golden mean’. There- fore we know that in the infinite Fibonacci word the ratio of the overall occurrence of L letters compared to the total number of letters (the ‘frequency’ of L letters NL/Ntot) is irrational. Conversely, if the Fibonacci chain was periodic it would contain (at least) one sub-word that is repeated infinitely often. Therefore the frequency of letters in the limit m → ∞, given by the ratio of letters in this sub-word, would be rational. Thus the Fibonacci chain is aperiodic.10 10 This is true independent of the pre- cise value of L/S (the ratio of lengths in the diagrammatic representation), in contrast to ref. [43]. The cases L/S = 1 and L/S → ∞ (L being finite) are a trivially periodic inasmuch as they contain just one non-zero spacing. At the same time the Fibonacci word is not random, since it was constructed using a deterministic rule (Eq. 1.1). In other words: the Fibonacci chain is long-range ordered.11 11 this will be shown explicitly later by its diffraction pattern The corresponding ‘Fibonacci lattice’ can be constructed from the chain by placing atoms at the vertices in between the line segments. We can write down the position x of the nth atom in a closed form [44] xn = S ( n + 1 ν ⌊ n + 1 ν ⌋) , (1.4) emphasising the presence of long-range order in the system. Here b·c denotes the floor function and ν ≡ L/S the ratio of lengths of the two line segments. To summarise, the Fibonacci chain is both aperiodic and long- range ordered. It is intriguing that as early as 1944, Schrödinger [45] introduced the con- cept of aperiodic order as a possible explanation for life on earth. At the time, the microscopic function of genes and the structure of DNA was un- known, but the question was already recognised on how such a fragile struc- ture (then estimated at around one million atoms) could robustly encode enough information, lasting many generations, to form a living organ- ism and seemingly violate the second law of thermodynamics. Schrödinger proposed an aperiodic crystal as ‘code- script’ in which all the information was encoded by the exact configura- tion of atoms. It is now known that DNA does not represent a determin- istic recipe as Schrödinger envisaged but it functions as a resource that in- teracts with its environment, thereby upholding a constant non-equilibrium of the organism in its surroundings (‘life’) which does not violate the over- all rise in entropy. However, the gist of his original idea that information is carried by aperiodic sequences (the order in which the nucleotides A, C, G, and T appear) has indeed proven correct. Ref. [46] offers a very recent introductory review on Schrödinger’s book. See also ref. [12]. So far, we have considered arbitrary values for L and S, leading to a deterministic and aperiodic arrangement of atoms in the Fibonacci lattice. Next we consider how a particular choice of ν can lead to self-similarity. Self-similarity Self-similarity means invariance under length scaling operations. Therefore, in order to construct a self-similar Fibonacci lattice we simply demand that replacing L with LS and S with L (Eq. 1.1) corresponds to the same (relative) increase in length, namely L + S L ! = L S ≡ ν (1.5) 20 quasicrystalline optical lattices for ultracold atoms Figure 1.4: Self-similarity of the ‘golden’ Fibonacci chain (ν = τ). If the upper Fibonacci chain is scaled down by a factor τ, the resulting structure is again a Fibonacci chain. Figure taken and adapted from [47]. or, equivalently, 1+ 1 ν = ν (1.6) ν2 − ν− 1 = 0 (1.7) leading to ν± = 1± √ 5 2 (1.8) ν+ = 1+ √ 5 2 = 2 cos ( 3pi 10 ) = 1.618033989 · · · ≡ τ (1.9) The value τ is the aforementioned limit of two adjacant Fi- bonacci numbers, the golden mean.12,13 A Fibonacci lattice with12 Another way to get the values ν± is to simply diagonalise the substitution matrix in Eq. 1.2. 13 It is called an ‘algebraic number’ because it is the solution to Eq. 1.7 which only has rational (in this case integers) coefficients. the self-similar property ν = τ is shown in Fig. 1.4. Here self- similarity manifests itself in the appearance of structures on larger and larger lengthscales. Self-similarity is a characteristic feature of quasicrystals which we are able to directly observe in the experiment (Chapter 6). It constrasts the usual notion of coarse-graining in periodic lattices: In the thermodynamic limit (system size V → ∞ and N → ∞, while N/V = constant) the ratio of lattice spacing a to system size tends to zero. Thermodynamic properties such as entropy or magnetisation therefore do not depend on the value of a. In self-similar systems such as quasicrystals this approach fails since, even for arbitrarily big lengthscales, the system still resembles an inflated version of the microscopic model. Apart from being conceptually interesting, self-similarity frequently appears in nature and society in the form of frac- tal structures, for example in network traffic [48], biology and medicine [49], and cosmology [50]. There are also well-known ex- amples in various plants, such as romanesco broccoli, ferns, and sunflowers. The study of ultracold atoms in quasicrystalline lattices might provide a novel approach to studying fractal structures in a well- controlled environment. The use of massive, quantum degener- ate particles – atoms – in this work represents a major advance compared to previous studies of quasicrystals with massless pho- tons [13] and classical particles in optical lattices [14–17]. Having covered the concept of self-similarity we can build some physical intuition on long-range order in quasicrystals by looking at their diffraction pattern, i.e. their structure in reciprocal space. This is particularly relevant for the diffraction experiments that were performed in this work using ultracold atoms. In addition, quasicrystals and aperiodic order 21 diffraction patterns are the only way of fundamentally discrimi- nating between quasicrystals and disordered aperiodic structures. Diffraction pattern of the Fibonacci lattice The density distribution of a 1D discrete lattice can be written as ρ(x) = 1 N N ∑ n=1 δ(x− xn) , (1.10) where the sum goes over all lattice points, xn are the positions of the atoms (taken from Eq. 1.4 in the case of a Fibonacci lat- tice), and δ(x) represents the Dirac δ-distribution. The diffraction pattern arising from ρ(x) is given by its structure factor which corresponds to the modulus squared of its Fourier transform F(q) ≡ |F [ρ(x)]|2 = ∣∣∣∣ 1√2pi ∫ +∞ −∞ ρ(x)eiqx dx ∣∣∣∣2 . (1.11) Here q is the change in wave vector due to the diffraction. Periodic lattices: periodic long-range order 0 1 2 3 4 5 6 7 8 q/(2pi) 0.00 0.05 0.10 0.15 F( q) 0 1 2 3 4 5 6 7 8 q/(2pi) 0.00 0.05 0.10 F( q) A B 0 1 2 3 4 5 6 7 8 q/(2pi) 0.00 0.05 0.10 F( q) C Figure 1.5: Theoretical diffraction patterns (Eq. 1.11) for three different lattices. (A) Periodic lattice with spacing S = 1. (B) Periodic lattice of the type LSLSLS. . . with spacing 1 + τ (two atoms per unit cell). (C) Fibonacci lattice with ν = L/S = τ. The diffraction peaks in (A) and (B) are perfectly regularly spaced (see text), indicating the presence of periodic long-range order. In contrast, (C) consists of a dense set of peaks, highlighting the aperiodic long- range order of the Fibonacci lattice. Although the envelope in (C) looks very similar to that of (B), the main peaks are not spaced regularly as can be seen upon close inspection (see also Fig. 1.7). All simulated lattices in this figure consist of 21 atoms; the value of τ (golden mean) is given by Eq. 1.9. Before we come to the Fibonacci lattice, les us consider the diffrac- tion pattern resulting from a periodic lattice with unity spacing (S = 1). It is simply a comb of δ-peaks with regular spacing q0 = 2pi/S = 2pi, as shown in Fig. 1.5A. There is a weak modu- lation of the height of the peaks and on a much smaller scale than q = 2pi there are ripples in the pattern which are not resolved on the figure. Both of these effects arise from the finite size (21 atoms) of the simulated lattice. This regularly-spaced diffraction pattern is the signature of periodic long-range order. Next, we consider another periodic lattice of the same size (N = 21) but with two incommensurable spacings S = 1 and L = τ, i.e. LSLSLS. . . . The diffraction pattern for this configuration is plotted in Fig. 1.5B. Similar to Fig. 1.5A, the spacing between the peaks is perfectly regular; in this case it is given by q0 ≡ 2piS + L = 2pi 1+ τ . (1.12) As before, the regularly-spaced peaks herald periodic long-range order. However, the height of the peaks is strongly modulated which is a consequence of the presence of two incommensurable lengths S and L. The overall structure is perfectly periodic. Fibonacci lattice: aperiodic long-range order The diffraction pattern of a 21-atom ‘golden’ Fibonacci lattice (ν = τ)14 is plotted in Fig. 1.5C. At first sight it may look simi- 14 Note that any irrational value of ν leads to a quasicrystal, not just ν = τ. But we have chosen the special value ν = τ because it nicely illustrates the self-similarity of the ‘golden’ Fibonacci chain in reciprocal space. This is already hinted at in Fig. 1.5B which – although it is entirely periodic in real space – shows signatures of self-similarity. lar to the periodic case but there are two important differences. Firstly, the main peaks are not spaced periodically (in contrast to Fig. 1.5B) but aperiodically. Secondly, there are many more peaks in between the main peaks and there exists no single fundamental distance q0. By indexing this diffraction pattern we can demon- strate the aperiodic order of the Fibonacci chain. 22 quasicrystalline optical lattices for ultracold atoms Indexing quasiperiodic diffraction patterns: aperiodic order First, let us take a closer look at the diffraction pattern of the Fi- bonacci lattice (Fig. 1.5C). A single reciprocal lattice vector, i.e. a single index (h) where h ∈ Z, is not sufficient to uniquely identify all the peaks since they are not spaced by equal distances. Simi- larly, in the in-plane diffraction image of a physical quasicrystal (Fig. 1.3C) two indices (h, k) are not sufficient to account for all the peak positions since the distances in reciprocal space become smaller and smaller. In fact, the peaks are densely spaced in these cases due to the incommensurability of two (or more) lengths in the quasicrystal. For example, in the ‘golden’ Fibonacci lattice L = τ is incommensurable with S = 1. Consequently, there are as many peaks as there are rational numbers: infinitely many. Does this mean indexing the golden diffraction pattern (Fig. 1.5C) is a lost cause? 1 1 1 2 1 3 1 4 1 5 · · · 2 1 2 3 2 5 3 1 3 2 3 4 4 1 4 3 . . . 5 1 5 2 ... Figure 1.6: Cantor’s proof of Q+ being a countable set. The sketch shows the existence of a bijection from N to Q+ by counting the rational numbers diagonally [51]. Luckily it does not. An intuitive way of counting all positive ra- tional numbers, introduced by Cantor in the 19th century, comes to the rescue. We simply have to write the rational numbers down in a two-dimensional array (Fig. 1.6), ordered by numerators (rows) and denominators (columns). Subsequently we remove all dupli- cates15 and then each rational number along the diagonal can be 15 e.g. 2/4 = 1/2, 2/2 = 1 identified with one non-negative integer, as shown in Fig. 1.6. Inspired by the method of counting Q+ we can index all the diffraction peaks of the 1D quasicrystal by using two integers instead of one, namely ql,m = 2pil a + 2pim b , where l, m ∈ Z and a, b ∈ R . (1.13) At least one of the values a, b has to be irrational. In the case of the golden Fibonacci lattice (ν = τ) the values for a and b can be derived analytically [44] using Eq. 1.4. The result is ql,m = 2pi 1+ 1 τ2 ( l τ + m ) . (1.14) The positive indices for l, m ≤ 8 are shown in Fig. 1.7, together with the simulated diffraction for the 21-atom golden Fibonacci lattice (Fig. 1.5C). How to tell the Fibonacci lattice apart from a random lattice? The dense set of diffraction peaks resulting from the Fibonacci lattice poses an interesting question: How can we be sure that the Fibonacci chain is ordered and not random? On the one hand, the deterministic construction rule should ensure perfect long-range order. On the other hand, long-range order is really defined by the existence of sharp diffraction peaks. Therefore it must be possible to distinguish a random lattice from a Fibonacci lattice just by observing their respective diffraction patterns.1616 In other words: The diffraction process does not know about the rule by which the lattice was created. It only knows about the resulting pattern of lattice sites. For a random lattice, let us consider a non-deterministic con- struction rule in which the length of each bond is randomly cho- sen to be either S = 1 or L = τ. The probability for each line quasicrystals and aperiodic order 23 0 1 2 3 4 5 6 7 8 q/(2p) 0.00 0.05 0.10 F( q) (0 ,1 ) (1 ,0 ) (1 ,1 ) (1 ,2 ) (1 ,3 ) (2 ,2 ) (2 ,3 ) (2 ,4 ) (3 ,4 ) (3 ,5 ) (3 ,6 ) (4 ,5 ) (4 ,6 ) (4 ,7 ) (4 ,8 ) (5 ,7 ) (5 ,8 ) Figure 1.7: Two-dimensional indexing for the 1D golden Fibonacci lattice. The dashed lines are labelled accord- ing to Eq. 1.14. The diffraction pattern was obtained from a 21-atom Fibonacci lattice with ν = τ. 0 1 2 3 4 5 6 7 8 q/(2pi) 0.00 0.05 0.10 F( q) (2 ) (3 ) (5 ) (8 ) (1 0) (1 1) (1 3) (1 6) (1 8) (2 1) 0 1 2 3 4 5 6 7 8 q/(2pi) 0.00 0.05 0.10 0.15 F( q) / 0. 01 analytic numerical A B Figure 1.8: Diffraction patterns of 1D random lattices with ν = L/S = τ (see text). (A) Lattice length N = 21. The most pronounced diffraction peaks are spaced periodically in distances of q0 = 2pi/(S + L) (c.f. Fig. 1.5B) which are labelled according to their position. These peaks arise from the frequently occurring pattern LS. In between the pronounced peaks there are many more sharp peaks which are not spaced periodically. These could be (wrongly) interpreted as the signature of aperiodic order. (B) N = 1000 (light blue). For a larger system size the disorder becomes apparent: the diffraction pattern becomes continuous and smeared out. The numerical curve is overlaid with the analytic results from ref. [9] (black line) which corresponds to an infinite random chain. For even larger system sizes the diffraction pattern would more closely resemble that of the analytic calculation. Some peaks remain sharp due to a frequent occurrence of certain patterns, the same as (A). segment being L shall be 1/τ such that the overall occurrence of L is roughly equal to that in the aperiodic ordered Fibonacci chain.17 The diffraction pattern of such a random lattice (N = 21) 17 this random chain is discussed in more detail in ref. [9]is shown in Fig. 1.8A. At first sight it does not differ greatly from the diffraction pattern of the perfectly ordered Fibonacci lattice (Fig. 1.7); the most pronounced peaks are spaced similarly but not identically to the periodic case. For this random lattice a single index is sufficient to label them, as demonstrated in Fig. 1.8A. The background in the random case shows some noise which could be interpreted as a dense set of peaks. In this system size (21 atoms) there is no qualitative difference between a random lattice and an 24 quasicrystalline optical lattices for ultracold atoms aperiodic ordered lattice. However, the randomness does show up in a sufficiently large system (Fig. 1.8B). For N = 1000 the diffraction pattern becomes smooth18 and tends towards the ana-18 although it retains some sharp peaks which arise from the frequent occurrence of certain combinations such as LS and LLS lytic solution for an infinite random chain [9]. From this we draw the conclusion that system size is crucial for determining true long-range order.19 19 As usual with quasicrystals, it is not always as simple as that. A class of pathological 1D quasicrystals was constructed by Baake and Grimm [52] whose diffraction patterns cannot be distinguished from completely disordered structures. Nevertheless, for all practical purposes the sharp peaks in the diffraction pattern are a clear signature of order. Projection from higher dimensions The indexing method described above uses two indices to describe the diffraction pattern of a 1D quasicrystal. On the one hand, the requirement of two indices can simply be understood as the consequence of the presence of two incommensurable lengths in the system. On the other hand, the double-indexing naturally leads to the notion that dimensions higher than the physical ones are at play in quasicrystals. Therefore one can re-interpret the diffraction pattern in Fig. 1.7 as the diffraction off a (fictitious) two-dimensional periodic lattice, projected down to one dimen- sion. This concept has proven extraordinarily powerful in the determination of quasicrystalline structure and it is nowadays routinely used to describe and classify quasicrystals [11, 53, 54]. It extends naturally to physical dimensions higher than one. In gen- eral, the number of extra dimensions needed to describe a given quasicrystal is equal to the number of its linearly independent lattice vectors over Q [55]. For example, the icosahedral symme- try of the Zn-Mg-Dy quasicrystal (Fig. 1.1) requires six indepen- dent lattice vectors to describe its diffraction pattern (Fig. 1.3) and consequently the periodic ‘image’ lattice is six-dimensional. Our experimental setting involves a planar, eightfold symmetric optical lattice which can be uniquely described by four indepen- dent lattice vectors. Consequently, the experiment is connected to four-dimensional space; Chapter 6 shows the direct simula- tion of a continuous-time quantum walk on a homogeneous 4D tight-binding lattice using this notion. The projection method is not only useful for crystallography and indexing but it is also used to construct models of quasicrys- tals in real space via the so-called ‘cut-and-project’ scheme.2020 also called ‘projection method’ or ‘strip-projection’ Figure 1.9: Schematic of the cut-and- project scheme for a rational slope (m = 2/3) leading to a periodic 1D crystal. The width of the acceptance window is given by the projection of the 2D unit cell (black square) to perpendicular space. Cut-and-project The cut-and-project scheme is widely used to contruct models of quasicrystals, as well as study their physical properties. It is il- lustrated in Figs. 1.9 and 1.10 for one-dimensional structures. The construction involves defining a periodic lattice, for example a simple cubic lattice (∼= Z2) in two dimensions (x, y). Rotation by an angle α defines a new coordinate system (x′, y′) which can be interpreted as parallel space (x′) and perpendicular space (y′). Par- allel space corresponds to the new, physical 1D lattice. In perpen- dicular space a certain ‘acceptance window’ (or ‘slice’) is chosen and points within this window are then projected to parallel space to form the new one-dimensional structure. In one dimension the quasicrystals and aperiodic order 25 description can be written in terms of a slope m = tan α. The width ∆ of the slice is given by the projection of the diagonal of the 2D unit cell (black square in Fig. 1.9) to perpendicular space, namely ∆ = 1+ m√ 1+ m2 . (1.15) Figure 1.10: Schematic of the cut-and- project scheme for an irrational slope (m = 1/τ) leading to an aperiodic 1D crystals, namely the Fibonacci lattice. The crucial parameter in this construction is the value of the slope m. If it is chosen to be rational, the constructed lattice will be periodic (Fig. 1.9). If it is chosen to be irrational, the constructed lattice is aperiodic (Fig. 1.10). In particular, if we choose m = 1/τ the resulting lattice is the golden Fibonacci chain. Note that although the ‘window’ itself looks two-dimensional it is actually just a 1D width in perpendicular space. All points that fall into the width under perpendicular projection are ‘accepted’ and then projected to parallel (physical) space. Phasons Figure 1.11: Phasonic excitation of the Fibonacci lattice. The acceptance window has been displaced by 0.3 lattice spacings compared to Fig. 1.10 (the original Fibonacci lattice is also shown). Small displacements of the acceptance window in the perpen- dicular direction can lead to sudden displacement of some atomic posi- tions. Sequences of LS turn into SL, a process called ‘phason flip’. Figure 1.12: Phasonic excitation with non-zero wavelength (see text). In periodic lattices there are two types of low-energy excitations, namely electronic excitations and phonons (lattice vibrations). Without going into too much detail, the cut-and-project scheme can help us gain some intuition on additional excitations that are unique to quasicrystals, called phasons [56]. Phasons are markedly different from phonons which are characterised by wave packets, propagating through the crystal. In quasicrystals it can be energetically favourable for an atom to perform a sudden ‘phason jump’ to a new equilibrium position, rather than being excited around its original equilibrium posi- tion. This process can be illustrated by phonons in perpendicular space. In Fig. 1.11 the acceptance window in the cut-and-project scheme has been displaced by a small amount (0.3 lattice spac- ings). The resulting lattice is another golden Fibonacci lattice but compared to Fig. 1.10 some atoms have jumped to a new position by a distance 1− τ. These are just the simplest cases of phasonic excitations, corresponding to a centre-of-mass movement of the acceptance window (q = 0 phonon modes). Non-zero q phonon modes, i.e. phonons with finite wavelength, introduce displace- ments that can lead to the appearance of subsequences, such as LSLSL or SLLLS, that do not appear in the Fibonacci lattice (see Fig. 1.12). Moreover, a change in phase of the finite-wavelength acceptance window can lead to the appearance of new spacings, namely L − S and L + S in the resulting 1D lattice, as shown in Fig. 1.13. Figure 1.13: Phasonic excitation with non-zero wavelength and different phase compared to Fig. 1.12 above. New spacings are introduced in this case. These phasonic excitations, in contrast to phonons, are not propagative but diffusive, meaning that their ‘dispersion relation’ is purely imaginary [56]. Phasonic excitations can lead to broad- ening of diffraction peaks and diffuse background scattering. It is expected that phasons contribute considerable entropy to qua- sicrystals, even at low temperatures. Phasons could potentially stabilise them against energetically competing periodic configura- 26 quasicrystalline optical lattices for ultracold atoms tions [53], i.e. minimise the free energy F = E− TS by increasing the TS contribution (T being temperature, S being entropy, and E being the internal energy). Up to this point quasicrystals have been discussed in one di- mension, thereby reviewing their most important characteristics, namly quasiperiodicity, self-similarity, and the projection from higher dimensions. However, one-dimensional space does not support rotations. Therefore, 2D quasicrystals will be covered in the following, beginning with the notion of ‘forbidden’ rotational symmetries. After all, an unexpected tenfold symmetry in the diffraction pattern of Al86Mn14 lead to the discovery of quasicrys- tals [1] in the first place. 1.3 Quasicrystals in two dimensions, the octagonal tiling Before we come to two-dimensional quasicrystals we need to take a step back and consider two-dimensional periodic crystals. A periodic lattice can be uniquely defined by its unit cell and a set of lattice vectors. Translation of a given lattice point by any of the lattice vectors results in an equivalent lattice point. The unit cell describes the arrangement of atoms relative to a lattice point. Note that in any physical system the lattice will be discrete, i.e. the distance between any two lattice points is larger than zero. In the following it will be shown that periodic lattices are incompatible with rotational symmetries higher than sixfold rotational sym- metry and also incompatible with fivefold rotational symmetry, following an original proof of ref. [57]. Crystallographic restriction Theorem. Periodic lattices are not compatible with any n-fold rotational symmetry for n > 6 and n = 5. P1 P2 P3 2�/8 A B P0 Figure 1.14: Sketch of crystallographic restriction for the case n = 8. (A) and (B) represent the two contructions of an eightfold symmetric polygon that are used in the proof. An infinite series of polygons can be constructed by using the edges of (B) to construct yet another polygon, and so on. The fact that the perimeter of (B) is smaller than that of (A) ensures that in the limit of this series the value of the perimeter vanishes, thus contradicting the discreteness of the lattice. Proof. Consider an n-fold rotational symmetry for n > 6 on a discrete lattice (the case n = 5 will considered separately). We start constructing a regular polygon ‘A’ from one lattice point P1 by adding a lattice vector to reach another lattice point P2. Subsequently we add a second lattice vector, rotated by 2pi/n, to the point P2 to reach a third lattice point P3. We keep doing this until we reach again P1, having constructed A in its entirety; an example of this construction is shown in Fig. 1.14 for the case n = 8. Let the perimeter of A be called sA. Now, if we protract all the above vectors −−→ P1P2, −−→ P2P3, . . . from one single lattice point P0 we have constructed another regular polygon ‘B’ with smaller perimeter sB, shown in Fig. 1.14 (blue line). The ratio between the two perimeters of A and B is only dependent on n, namely sB sA = 2 sin (pi n ) < 1 . (1.16) quasicrystals and aperiodic order 27 Taking the edges of the smaller polygon (B) as the new vectors to construct yet another polygon C, and so on, will lead to an infinite series of polygons. Since the ratio of perimeters of two subsequent polygons in this series is strictly smaller than one (Eq. 1.16) the edges of the polygons in the series converge to zero. This contradicts the physical assumption of a discrete lattice. The special case of n = 5 entails the identical construction of the pentagon A but it necessitates a slightly more complicated construction of the pentagon B (and the successive elements of the series). In this case, the smaller pentagon (B) is constructed by successively protracting the vectors −−→ P5P4, −−→ P1P5, . . . from the lattice points P1, P2, . . . , rather than to one central point. The ratio of perimeters is then sB sA = ∣∣∣1− sin2 (pi n )∣∣∣ < 1 . This construction can be used analogously to prove the theorem for all cases including n = 5 and n > 6. The proof of crystallographic restriction completes the discus- sion of periodic lattices. Since sharp diffraction patterns with tenfold rotational sym- metry have been observed [1] (see also Fig. 1.3) it is immediately clear that in an aperiodic lattice higher rotational symmetries must be somehow possible. Furthermore, crystallographic restriction implies that experimental evidence of sharp diffraction peaks with, say, eightfold symmetry is a sufficient criterion to identify a quasicrystal.21 In order to get a more microscopic picture of 21 However, one must avoid various crystallographic pitfalls that can lead to apparent rotational symmetries in the diffraction pattern. For example, a particular arrangement of four atoms in a 2D simple cubic lattice produces an approximately eightfold symmetric diffraction pattern (see ref. [58], section 3, or ref. [9], Chapter 9). This effect is very similar to the diffraction pattern obtained from a 1D simple cubic lattice with two atoms per site which are arranged in a particular way (Fig. 1.5B). This shows signatures of self-similarity although the underlying lattice is entirely periodic. two-dimensional quasicrystal it is instructive to consider aperiodic tilings, a subject within the mathematical field of geometry.22 22 Refs. [3] and [9] give an accessible in- troduction to the topic. An exhaustive treatment can be found in ref. [10]. Aperiodic ordered tilings A tiling, or tessellation, represents a way to cover a plane using a set of geometric shapes (‘tiles’), without any gaps or overlaps [10]. Until the early 60s it was assumed by Hao Wang [59, 60] and others that all coverings of the plane using square tiles with deco- rated edges would repeat themselves after a long distance, i.e. be periodic. However, it was later shown that aperiodic coverings of the plane do indeed exist [61]. Moreover, it was shown that, given a certain set of tiles, it is generally impossible to decide whether or not these tiles support an infinite covering of the plane; this is known as the ‘undecidability of the domino problem’. This proof is another representation of the undecidability of the Turing ma- chine halting problem [62]. It has even been shown that these ‘Wang tiles’ can perform computations in the same way a Turing machine does [63]. Very recently, the undecidability of the domino problem has been mapped to a very general class of many-body hamiltoni- ans, namely two-dimensional spin systems, with the result that is impossible to predict whether these hamiltonians support 28 quasicrystalline optical lattices for ultracold atoms gapless or gapped excitations [31]. It is therefore to be expected that constructing physical systems of quantum particles in ape- riodic potentials such as our quasicrystalline lattice will result in many-body systems of unprecedented complexity. Maybe self- assembly methods could be exploited similar to those that allow DNA-strands to be used as Wang tiles in the construction of crys- tals [64]. It is not unthinkable that one day atoms in quasicrys- talline lattice might act as ‘quantum Turing machine’. Figure 1.15: A patch of the octagonal tiling. The aperiodic tilings considered so far consist of square ‘Wang tiles’ and hence they do not explain the existence of non-crystallographic rotational symmetries. The first pentagonal tiling was found in 1974 by Penrose [8]: I had often doodled by fitting together limited configurations of pentagons and similar shapes but I had never found a good rule for continuing such patterns indefinitely. However, recently I wanted to design something interesting for someone who was in hostipal to look at and I realised that there was a certain definite rule whereby one could continue such a pattern to arbitrary size. This pattern is nowadays widely known as the Penrose tiling. Later, an eightfold analogue to the Penrose tiling was found, the so-called Ammann-Beenker tiling [10] or ‘octagonal tiling’. A patch of the octagonal tiling is shown in Fig. 1.15, consisting of two types of tiles (a square and a rhombus) from which the entire pattern can be pieced together indefinitely. Construction rules How is it possible to construct an aperiodic tiling, such as the octagonal tiling? Similar to the Fibonacci chain there are construc- tion rules that allow deterministic tiling of the plane using two different shapes (in this case a square and a rhombus). In brief, the construction relies on matching rules for adjacent tiles, similar to a puzzle game. This process is illustrated in Fig. 1.16A by ‘decorat- ing’ the vertices and the edges of the tiles with arrows. The arrows on the vertices add local information to the tiles, thereby ensuring ‘validity’ of the infinite tiling, i.e. that there are no gaps. Non- deterministic construction rules can be used to generate random octagonal-like tilings with approximate eightfold symmetry [9]. Another way of constructing the octagonal tiling makes use of its self-similarity. Self-similarity The octagonal tiling is self-similar under scaling the edges of the tiles by the so-called silver mean, λ = 1+ √ 2 . (1.17) Similar to the golden mean (Eq. 1.9) λ is an algebraic number, obeying the equation x2 − 2x− 1 = 0 , (1.18) quasicrystals and aperiodic order 29 CBA Figure 1.16: Octagonal tiling. (A) ‘Dec- orated’ octagonal tiling, illustrating the construction rules. (B) - (C) Self- similarity of the octagonal tiling under ‘silver mean’ scaling. These images were generated using the Mathematica package ‘Aperiodic tilings’ [65], see also ref. [47]. and it can be represented by a continued fraction analogous to that in Eq. 1.3 λ = 2+ 1 2+ 1 2+ 12+... . (1.19) However, the self-similar scaling operation of the octagonal tiling is less trivial than its 1D counterpart, in which applying the scal- ing operation to any patch of the Fibonacci chain results in an- other valid Fibonacci chain (Fig. 1.4). In the two-dimensional case of the octagonal tiling the scaling operation by a factor λ can nev- ertheless be used if certain rules are obeyed which are shown in Fig. 1.17. These allow self-similar scaling even if the position of the origin is unknown. One step of self-similar scaling of the octagonal tiling – also called ‘inflation’ – is shown in Fig. 1.16B-C. Figure 1.17: Scaling by a factor λ (Eq. 1.17) generates a self-similar copy of the octagonal tiling (Fig. 1.16 B and C). The construction of the octagonal tiling based on local rules (even in the case of self-similar inflation) is cumbersome because it is an iterative process. Similar to the Fibonacci chain, the construc- tion via cut-and-project is far easier and much more efficient. In addition, it can provide intuition about phasonic excitations in this lattice. Cut-and-project In the case of the octagonal tiling the cut-and-project scheme starts with constructing a four-dimensional (4D) simple-cubic lattice L ∼= Z4. The basis of L can be written as  1 0 0 0  ,  0 1 0 0  ,  0 0 1 0  ,  0 0 0 1   (1.20) in canonical coordinates (x1, x2, x3, x4), xi ∈ Z. Similar to the Fibonacci case we define a rotation [9] R = 1 2  √ 2 1 0 −1 0 1 √ 2 1√ 2 −1 0 1 0 1 −√2 1  (1.21) 30 quasicrystalline optical lattices for ultracold atoms Figure 1.18: Parallel (A) and perpen- dicular (B) projection (‘shadow’) of a hypercube (unit cell of 4D simple-cubic lattice) using Eq. 1.22. The four basis vectors (Eq. 1.20), i.e. edges of the unit hypercube, are shown in blue (A) and in red (B), respectively. The shadows in both parallel and perpendicular projection are identical. The hypercube has an edge length of unity whereas the shadows have edge lengths 1/ √ 2. (C) The acceptance window in per- pendicular space (grey) for the 4D cut-and-project scheme is the 2D perimeter of the shadow in perpen- dicular projection, i.e. outer green line in (B). The points in (C) are the same elements of Z4 (or, equivalently, the 4D simple-cubic lattice) that are also shown in Fig. 1.19 in physical/parallel space. (A-C) are drawn to the same scale. A B C such that the new coordinates are x y x′ y′  = R  x1 x2 x3 x4  . (1.22) The first two coordinates (x, y) are parallel (physical) space and the last two (x′, y′) are perpendicular space; a projection of (x1, x2, x3, x4) to perpendicular space corresponds to multiplication by the first two rows of R. The projection to parallel space, correspondingly, is given by the multiplication by the last two rows of R. Figure 1.19: Octagonal tiling generated by projection from a Z4 using a 4D hypercube with seven lattice sites per edge and unity lattice spacing. The bond lengths (black lines) resulting from this projection have length 1/ √ 2. As before, the width of the acceptance window is given by the projection of the 4D unit cell to perpendicular space, which means that the acceptance window in this case is two-dimensional. The projections of the 4D unit cell are shown in Fig. 1.18A-B; both perpendicular and parallel projections (‘shadows’) have a 2D octagonal perimeter of the same size. The acceptance window is then chosen to be the perimeter of this octagon (green line in Fig. 1.18B). The points that fall into this window under perpendic- ular projection (Fig. 1.18C) form the octagonal tiling in physical space, simply by performing a parallel projection (Fig. 1.19). Phasons Analogous to the Fibonacci case, phasonic excitations can be in- troduced by shifting the acceptance window by a certain amount. Figure 1.20 shows an example of a phasonic excitation by shifting the acceptance window upwards by 0.2 lattice spacings. Phasonic excitations of the octagonal tiling will play an im- portant role in the description of ‘lattice sites’ in the eightfold rotational optical potential (section 5.3). Figure 1.20: One phasonic excitation (red) introduced to the perfect octag- onal tiling by shifting the acceptance window (Fig. 1.18C) by 0.2 lattice spacings in the vertical direction. The original octagonal tiling (black) was constructed by projection from Z4 us- ing a hypercube of size 24 (three lattice sites per edge, unit lattice spacing). Physical diffraction of aperiodic models The physical significance of aperiodic tilings was realised imme- diately after the discovery of quasicrystals since the calculated diffraction pattern of a Penrose tiling [66] closely resembled the experimental images obtained by Shechtman et al. [1]. Nowa- days aperiodic models such as the Penrose tiling and its three- dimensional icosahedral counterparts [9, 11] are widely accepted quasicrystals and aperiodic order 31 microscopic models for physical icosahedral and decagonal qua- sicrystals whose crystallographic description is essentially com- plete [11]. Although physical quasicrystals with eightfold sym- metry are comparatively less common [54] the octagonal tiling plays a central role in the modern understanding of quasicrystals, particularly (but not only) from a geometric point of view [9, 67]. The diffraction pattern of the octagonal lattice can be calculated analytically [9] and is shown in Fig. 1.21. It can be described by four reciprocal lattice vectors23 {G1, G2, G3, G4} which point in 23 thus, a four-dimensional image lattice is required to generate the octagonal tiling via the cut-and-project scheme the same respective directions as the parallel projection of the 4D basis vectors (red lines in Fig. 1.18). These lattice vectors are also going to be used to describe the experimental diffraction patterns in Chapter 6. Figure 1.21: Theoretical diffraction pattern of the octagonal tiling. The image shows all peaks with at least 10−4 relative intensity (compared to the q = (0, 0) peak), which leads to the discrete appearance (the whole diffraction pattern consists of a dense set of peaks). The pattern is perfectly eightfold symmetric and it shows signatures of self-similarity, i.e. the appearance of diffraction peaks at powers of λ. Image taken from [9] with kind permission of the authors. The main concepts of quasicrystals, namely quasiperiodicity, self-similarity and rotational symmetries have now been discussed in one and two dimensions. In the final section of this chapter the progress that has been made in research on quasicrystals in the past 35 years will be reviewed. 1.4 Physical quasicrystals and current state of the art To date, a huge variety of quasicrystals has been discovered [11, 54]. It is instructive to sort them according to their symmetry and their chemical composition. 32 quasicrystalline optical lattices for ultracold atoms Symmetry The only thermodynamically stable quasicrystals which have been found up to now exhibit either pentagonal (fivefold) or decago- nal (tenfold) or icosahedral (Fig. 1.22) symmetry [11]. Icosahedral quasicrystals are quasiperiodic along all three spacial directions. Conversely, pentagonal and decagonal quasicrystals are quasiperi- odic in the plane of the rotational symmetry but periodic along the orthogonal direction. Therefore these quasicrystals are also sometimes called ‘axial quasicrystals’ [54]. A large variety of axial quasicrystals with varying degrees of structural integrity have been found so far, including pentagonal, octagonal [54, 68], decagonal, and dodecagonal (twelvefold) quasicrystals. However, sevenfold symmetry or any other rotational symmetries not men- tioned above have never been observed [11]. Figure 1.22: A regular convex icosa- hedron. Each of the twenty faces is an equilateral triangle. Around each vertex there are five faces which form a kind of pentagonal pyramid. Chemical composition All physical quasicrystals observed so far are binary or tertiary compounds, the vast majority of them being intermetallic com- pounds. For certain intermetallic materials the relative concentra- tions of the constituents for which stable quasicrystals exist have been mapped out [11]. However, it is currently an open question whether quasicrystals made of a single chemical element can exist. The stability of many quasicrystals seems to depend very strongly on one chemical property, namely the valence electron concentra- tion [69]. This suggests that electronic mechanisms may play an role in stabilising the high rotational symmetries and aperiodic order in general. Transport Electric and thermal transport properties of quasicrystals are gen- erally anisotropic [70]. For example, a high-quality sample of quasicrystalline material can exhibit metallic behaviour along one direction and insulating behaviour along another, despite being an intermetallic compound of aluminium and two other metals (commonly). Similarly, the thermopower or the Hall coefficient can change sign depending on the respective lattice axis it is mea- sured on. These intriguing properties probably can be explained by structural arguments, though it is currently unclear whether local atomic order or long-range aperiodic order is responsible for these unique features [70]. Open questions Despite much effort to understand quasicrystals better there are still many open questions, even on a fundamental level. • Are physical quasicrystals energy- or entropy-stabilised? I.e. is the Helmholtz free energy F = E − TS minimal for S be- quasicrystals and aperiodic order 33 ing maximal, or for E being minimal? Is structural disorder inherent in quasicrystals, or can an ‘ideal’ quasicrystal exist? • How do quasicrystals form? In other words, quoting ref. [11], How does the 1000th atom find its site in a giant unit cell with thousands of atoms? • Finally, why are certain rotational symmetries (fivefold, eight- fold, tenfold) prevalent in quasicrystals, and why have others (sevenfold, elevenfold, thirteenfold) never been observed? Apart from the questions listed above which are primarily condensed- matter- and crystallography-driven, research on quasicrystals also touches on the topics of complexity, information science and biol- ogy [71, 72]. Physical model systems Curiously, the only evidence of naturally occurring quasicrystals has been obtained from samples of meteorite material [73, 74]. All other known quasicrystalline materials have been synthesised in the laboratory, some of them (Fig. 1.1) with similar structural quality as best periodic crystals [11]. In order to tackle the questions mentioned above, a variety of physical ‘model systems’ have been synthesised which aim to emulate quasicrystals with a larger degree of experimental control. Here is a list of physical settings in which quasicrystalline systems have been studied experimentally • aperiodically layered materials [12], e.g. semiconductors [75] • liquid crystals [76] and polymers [77] • self-assembled nanoparticles [78] and colloids [79] • photonic waveguides [5, 13]. In these systems quasiperiodic patterns can be imprinted by utilising a projected optical lat- tice [80–82]. • laser-cooled atoms in dissipative optical lattices [14, 15] • polystyrene spheres in optical lattices [16, 17] • optical diffraction [66, 83] • bilayered graphene [84] Perspectives of ultracold gases in aperiodic lattices This work establishes quasicrystalline optical lattices as a novel platform for ultracold atoms. Several research avenues can be imagined from here. On the one hand, there are many open questions regarding the physics of quasicrystals. It has been proposed by Macé et al. [85] that ultracold atoms in a system very similar to ours (see sec- tion 5.3) could help solve some of these long-standing mysteries. 34 quasicrystalline optical lattices for ultracold atoms This would be particularly interesting from a chemical point of view since up to now only binary and tertiary quasicrystalline compounds have been discovered [11]. Using a single-component quantum gas of potassium, for instance, would resemble this elu- sive ‘unary’ quasicrystal. On the other hand, ultracold atoms in quasicrystalline lattices are interesting and rich systems in their own right. For example, a topological charge pump has been demonstrated experimentally using a 1D photonic quasicrystal [5]. Similary, ultracold atoms in tilted superlattices have recently been shown to represent a 2D topological charge pump, realising in effect a 4D quantum Hall response [86]. These methods may in future be applied also to our quasicrystalline potential thereby realising novel 4D model systems with artificial gauge fields [87]. Another application of aperiodic lattices for ultracold atoms has been the observation of many-body localisation [88]. In particular the role of interactions in quasiperiodic media is poses fundamen- tal questions in physics [89], and experimental results have only recently become available [24, 88, 90, 91]. So far, all experiments employed either 1D [88, 90, 91] or 2D separable [24] potentials. However, more experimental evidence is needed, for instance, to differentiate between two universality classes of localisation phe- nomena, namely in random disorder and in aperiodic order [92]. For the latter class, our experiment is an ideal starting point since the lattice potential closely resembles the octagonal tiling (see section 5.3), a paradigmatic 2D quasicrystal. The eightfold lattice described in this work is a highly flexible system, not only allow- ing 1D quasiperiodic and 2D separable quasiperiodic potentials, but also, for the first time, a 2D non-separable aperiodic lattice, i.e. a 2D quasicrystal. 2 Design of the apparatus This chapter offers a brief overview of the experimental appa- ratus that is used to create quasicrystalline optical potentials for ultracold atoms, before the apparatus is described in detail in Chapters 3-5. First, the choice of atom (rubidium and potas- sium) is motivated with a focus on the ability to perform quantum many-body experiments in future. Then a brief overview of the main vacuum apparatus is given, together with an outline of the experimental sequence that is used to create samples of quantum degenerate gases. For a detailed list of devices, part numbers and components, the reader is referred to Appendix A. 2.1 Suitable constituents for creating quantum many-body sys- tems When setting up a new laboratory to study quantum many-body physics, the choice of quantum constituents – neutral atoms, ions, or molecules – is not an easy one.1 Often there is a trade-off be- 1 There are, of course, many more possible experimental systems but for simplicity only these three are considered here. tween the feasibility to prepare and manipulate ultracold samples on the one hand and the capability to perform interesting, novel experiments on the other. Neutral atoms First to mention are neutral atoms which can be broadly charac- terised by four main properties: the atomic mass, the compound particle flavour2 (boson, fermion), the feasibility to tune contact 2 If the number of neutrons is even, the compound particle (atom) is bosonic, otherwise it is fermionic. Stable isotopes of the elements are more commonly bosonic. interactions [93], and the dominant type of interaction (short- range or long-range). Alkali atoms are relatively easy to cool3 in 3 this will be discussed later in sec- tion 3.1 large quantities but they lack intrinsic long-range interactions. Ions In contrast to neutral atoms, ions are subject to the Coulomb re- pulsion and they offer long coherence times, suitable even for quantum information processing [94]. However, they are hard to cool en masse, the current record being around 200 ions [95]. 36 quasicrystalline optical lattices for ultracold atoms Polar molecules Furthermore, polar molecules are very interesting for quantum simulation applications since they feature long-range dipolar interactions [96], but unfortunately their cooling process involves serious technical challenges due to the plethora of rovibrational sub-levels. 2.2 Choice of atomic species To date, the most successful large-scale quantum many-body ex- periments have been carried out using ultracold atoms in optical lattices, the prime example here being the observation of a quan- tum phase transition using 87Rb-atoms [18]. In the following short review of atomic species we will therefore focus on neutral atoms. The traditional atomic species that are best-suited for laser cooling are 87Rb [97] and 23Na [98] which were the first to be Bose-condensed.4 However, these species lack readily accessible4 together with 7Li [99] Feshbach resonances [93] limiting their feasibility for studying strongly-correlated systems.5 This limitation can be overcome by5 Long-range interactions can never- theless be engineered in few-particle systems by employing the Rydberg- blockade, highlighted by recent ex- periments on few-body quantum systems [100]. using lithium [101] (6Li/7Li) or potassium [102, 103] (39K/40K) which feature broad, accessible Feshbach resonances and are suit- able for investigating strongly correlated physics. In addition, both elements include a stable fermionic isotope (6Li and 40K), a crucial ingredient to study electron-like many-body systems, such as the Fermi-Hubbard model [104], as well as antiferromag- netic quantum magnetism [105, 106]. Going beyond the ‘standard repertoire’ in the alkalis there has been tremendous success in the past decade in reaching quantum degeneracy of species with var- ious special properties. These include elements with two valence electrons, namely ytterbium [107, 108] and strontium [109–111], which possess singlet ground states with with zero electronic an- gular momentum, giving rise to intriguing SU(N)–symmetries and extremely narrow optical transitions. Furthermore, there are elements with very large magnetic moments, most notably dysprosium [112, 113] and erbium [114, 115], which give rise to dipolar interactions. Quantum degeneracy of these species (ytterbium, strontium, dysprosium, and erbium) has been reached with both their bosonic and their fermionic isotopes. However, the advantages of these species often come with some technical cost, for example in ytterbium and strontium: in order to perform laser cooling on narrow optical transitions a laser with a very narrow linewidth is required. This experiment The apparatus described in this thesis is designed to create a va- riety of novel lattice geometries. This requires a large degree of optical access, as well as reliable production of quantum degen- erate samples without too much technical overhead. Therefore a well-established ‘traditional’ combination of atomic species was design of the apparatus 37 chosen, namely rubidium (87Rb) and potassium (39K/40K). The setup employs sympathetic cooling of potassium using 87Rb as coolant; with a suitable laser system it is therefore possible to switch between 39K and 40K on a shot-to-shot basis using only software commands. All the experiments carried out in this work use 87Rb and 39K though we have also demonstrated a magneto- optical trap (MOT) of 40K and the ability to easily switch between the two potassium isotopes. For a more detailed review of the atomic properties of 87Rb and 39K please refer to section 3.1. 2.3 Vacuum apparatus The two most important design criteria for building a cold-atom apparatus today are optical access and cycle time. On the one hand, wide optical access allows flexibility of the type of exper- iments performed. In our case, this is imperative as we require four optical axes in one plane for the optical quasicrystalline lat- tice. On the other hand, a short (< 1 min) cycle time is beneficial for convenient experimentation and efficient optimisation of the apparatus.6 6 and also for the sanity of the gradu- ate students.The ‘Munich design’ pioneered by Greiner et al. [116] combines the two requirements by separating a ‘high-pressure’ (∼ 1× 10−9 mbar) MOT chamber with limited optical access from the science cham- ber (∼ 1 × 10−11 mbar) with excellent optical access. The two chambers are connected by an angled differential pumping section through which the atoms are transported magnetically. This well- established design is now being used in several laboratories, in- cluding LMU Munich [116–120], ETH Zurich [121], Harvard [122], and LPL (Paris) [123]. Our apparatus is shown schematically in Fig. 2.1. It is an adap- tation of the latest iteration [119, 120] of the Munich design which contains an additional high-pressure (∼ 1 × 10−6 mbar) cham- ber for preparing a pre-cooled atomic beam (2D MOT). The main differences between their design and ours can be summarised as follows: • We have two 2D MOT chambers, allowing independent control of the background vapour pressure for rubidium and potas- sium. • The effective differential pumping section between the 2D MOT chambers and the MOT chamber is probably shorter than theirs.7 7 We ordered a titanium part with a 1.5mm clearance over 50mm length but the part we received has a 1.5mm clearence only over 20mm. The rest of the differential pumping section has a wider diameter (∼ 3 mm). However, the pressure differential between 2D MOT chambers (which are not actively being pumped) and the MOT chamber has never been a limiting factor in our experiment and consequently a longer clearence would have been of no avail. • The differential pumping section (20 cm) between the MOT chamber and the ‘knee’ is longer and its inner diameter (10mm) is wider than theirs. • Our science cell is a rectangular cuboid, allowing a more com- pact geometry of experiment coils. It should be noted that the large separation between the MOT chamber (7.2× 10−11 mbar) and the science cell (1.2× 10−11 mbar) 38 quasicrystalline optical lattices for ultracold atoms Ion pump (150 l/s) Ion pump (75 l/s) Knee Science cell Titanium sublimation pump (TSP) Pressure gauge Pressure gauge MOT chamber Magnetic transport Experiment coils 2D MOT chamber (39K/40K) 2D MOT chamber (87Rb) Rubidium ampoule (1g, 27.8% 87Rb) Potassium ampoule (100mg, 2.98 % 40K) Valve Differential pumping section Titanium sublimation pump (TSP)Differentialpumping section Angled valve Valve Cooling block Angled valve Angled valve Angled valve 20 cm Figure 2.1: Schematic of the main experimental apparatus. For clarity, only the lower half of the cooling block and magnetic transport system (brass-coloured) is shown. is now mostly practical and not physically required since the pres- sure difference between the two chambers is only a factor of six. Nevertheless we can maintain efficient MOT loading, facilitated by a very large (∼ 104) pressure differential between the 2D MOT chambers and the MOT chamber. A more comprehensive treatment of the design, implementation and characterisation of the vacuum apparatus can be found in Matteo Sbroscia’s First-Year Report [124]. Table 2.1: Rough overview of the experimental sequence. stage duration section MOT 2.5 s 3.5 final MOT 10ms 3.7 optical molasses 4.5ms 3.8 optical pumping 0.7ms 3.9 magnetic transport 5 s 3.10 MW evaporation 7.5 s 3.11 RF state transfer 20ms 4.5 dipole evaporation 3 s 4.3 total 23 s 2.4 Overview of the experimental sequence Table 2.1 provides a rough overview of the experimental sequence that is used to create quantum degenerate samples of 87Rb and 39K. design of the apparatus 39 Figure 2.2: Completed vacuum setup (4th July 2016). The photo was taken after bake-out had been completed, after the rubidium and potassium ampoules had been broken, and after the turbo-pump had been dis- connected. The heavy cable of the titanium-sublimation pump (TSP) is still connected (top right). Figure 2.3: Science cell and brass cool- ing block seen from main breadboard (see also Fig. 5.17), before any optics have been placed on it (28th February 2017). The ribbon-wire can be seen (black), connecting the upper and the lower experiment coil. The compen- sation coils are wound in red (T-axis), black (D-axis), and blue wire (Z-axis) on 3D-printed mounts. 3 Laser cooling & trapping of 87Rb & 39K This chapter accompanies the atoms on their initial journey from room temperature to the µK-scale before we later present the de- tailed methods to reach quantum degeneracy in the optical dipole trap (Chapter 4). First, each atomic species and their relevant en- ergy scales are introduced. Then the experimental methods to cool and trap these atoms are described. This chapter is focussed on our experimental implementation1; the theoretical introduction 1 A detailed list of optical components can be found in Appendix A.to each method is kept to a minimum. For a general introduction to the topic, the reader is referred to the standard literature [125– 127]. 3.1 The alkalis Table 3.1: Vapour pressures for rubid- ium and potassium for two different temperatures [128]. T Rb K 295K 3.8× 10−7 mbar 1.7× 10−8 mbar 333K 1.6× 10−5 mbar 1.0× 10−6 mbar The inevitable starting point for any experiment involving ul- tracold gases is some kind of atom source at or above2 room- 2 In many cases atoms have to first be heated to increase their vapour pressure. temperature. Thankfully, the vapour pressures of rubidium and potassium are reasonable (see table 3.1) such that a small speci- men (0.1 - 1.0g) of each species in a vacuum chamber is sufficient to start the laser cooling cycle from background vapour. Similar to atomic hydrogen, the ground level of the alkali atoms is the 2S1/2-level of the valence electron. Since here the angular momentum quantum number L is zero, there is no fine structure splitting and consequently only the two hyperfine states F = 1, 2 are thermally populated at ambient temperature; The thermal en- ergy required to significantly populate the excited electronic states is much higher than room temperature, as shown in table 3.4. The hyperfine structure of the alkalis is very favourable for standard laser cooling techniques [126] due to the existence of (almost) closed transitions on the D2-line (2S1/2 → 2P3/2).3 A 3 for a particular choice of F and mF statestransition is said to be ‘closed’ if it can be treated as an effective two-level system, neglecting decay into other (ground) states, e.g. by spontaneous emission. Other advantages include reason- able natural linewidths4 and the commercial availability of lasers 4 wide enough for the transitions to be addressed with a conventional single- mode diode laser and narrow enough to have Doppler temperatures on the order of 100 µK in the relevant wavelength-range (though in past decades laser technology has advanced considerably such that nowadays the availability of lasers is not usually a limiting factor). 42 quasicrystalline optical lattices for ultracold atoms Rubidium-87 -7 9. 9 +1 33 .3 F = 2 F = 1 5 2S1/2 5 2P3/2 F = 0 F = 3 C o o li n g R ep u m p l o ck 6. 83 4 G H z Im ag in g O p ti ca l p u m p in g ( sp in -p o l) 72 .2 13 3. 3 F = 2 D 2- sp ec tr o sc o p y F = 2 t o F ’ = 2 ,3 c ro ss o v er ( 7 8 0 .2 4 1 n m ) 13 3. 3 15 6. 9 F = 1 Δcool R b m as te r la se r em is si o n 53.4 gF = 1/2 gF = -1/2 220 Figure 3.1: Hyperfine structure of the D2-line transition in 87Rb (not to scale) F‘ = 3 F‘ = 2 F‘ = 2, 3 cr os so ve r F‘ = 1 F = 1 F = 2 3.036 GHz 85Rb 52S1/2 6.834 GHz 87Rb 52S1/2 87Rb 52P3/2 Figure 3.2: Rubidium D2-line Doppler- free (blue) and Doppler-broadened (red) spectroscopy. The 87Rb master laser is locked 53.4MHz below (i.e. to the red of) the F′ = 2, 3 crossover line (see also Fig. 3.1). In terms of aptitude for laser cooling 87Rb is one of the most favourable atomic species (see also section 2.2) which is down to three main factors: Firstly, the hyperfine splitting of both the ground and the D2-excited state is much larger than the natural linewidth (6MHz×2pi); therefore the individual hyperfine lev- els (and their crossovers) are nicely resolved by our spectroscopy (Fig. 3.2). Correspondingly the |F = 2, mF = 2〉 → |F = 3, mF = 3〉 (2P3/2) transition, driven by σ+-polarised light, is almost perfectly closed and comparatively few repump photons (F = 1→ F’ = 2) are required to transfer atoms back to the cycling transition. This allows for a near-textbook-like implementation of magneto-optical trapping and optical molasses. Secondly, the atomic mass is rela- tively large, favouring thermalisation and reducing non-adiabatic (‘Majorana’-)losses from magnetic traps. Thirdly, the background scattering length is relatively large and positive (100 a0), favouring fast thermalisation and BEC stability (section 4.6). Potassium-39 In contrast to 87Rb, the hyperfine splitting of the 42P3/2 level in 39K is very narrow (Fig. 3.3, unresolved by our spectroscopy in Fig. 3.4), precluding a ‘properly’ closed transition and necessi- tating many more repump photons to combat this effect. Con- sequently, molasses-cooling is much less efficient than for 87Rb. Additionally, its background scattering length is negative which means that an ultracold cloud of 39K can collapse under its own inter-atomic attractive interactions [129]. However, 39K possesses a readily accessible Feshbach resonance at 402.70(3) G [130] to counteract the negative effects of attractive interactions. This Fesh- bach resonance is indeed the main reason why 39K is widely used for studying strongly interacting or non-interacting bosons [103, 131–136]. The well-established technique of sympathetic cooling with 87Rb [103, 131, 132, 134] allows relatively straightforward production of quantum degenerate samples of 39K. Although the sympathetic production might be in future superseded by all- optical production using the ‘grey molasses’ technique [129, 133, 137, 138] the presence of 87Rb proved to be useful in our lab. It allowed us to benchmark the optical lattice in the early stages of the experiment before we achieved condensation with 39K. More- over, the experiment is designed to perform shot-to-shot switching between 39K and fermionic 40K in future, aided by the abundant 87Rb-atoms for sympathetic cooling of either species. So far the short review of atomic properties considered isolated atoms; in the following the effects of external fields will be briefly discussed. laser cooling & trapping of 87rb & 39k 43 Laser frequencies The relevant optical transitions for laser cooling of 87Rb and 39K are shown in Fig. 3.1 and 3.3, respectively. In the following sec- tions 3.5-3.9 the individual purpose of each laser frequency will be discussed in detail. Note that in 87Rb the reference laser (green line in Fig. 3.1) is emitting 53.4MHz to the red of the spectroscopy line, whereas in 39K the reference laser is locked directly to the spectroscopy signal. 14 .4 6. 7 F = 2 F = 1 4 2S1/2 4 2P3/2 F = 0 F = 3 C o o li n g R ep u m p Im ag in g ( lo w -f ie ld ) O p ti ca l p u m p in g ( sp in -p o l) 3. 3 F = 2 D 2- li n e (7 66 .7 01 n m ) 9. 4 F = 1 K 39 m a st er l a se r em is si o n , l o ck ed t o F = 1 ,2 t o F ’ c ro ss o v er 21 .1 46 1. 7 2 26 .1 23 5. 6 53.0 gF = 1/2 gF = -1/2 Figure 3.3: Hyperfine structure of the D2-line transition in 39K (not to scale). Note the unequal distances between the lock point on the two hyperfine ground levels. F = 1 F = 2 F = 1,2 cr os so ve r 39K 42S1/2 39K D2-line 461.7 MHz 41K 42S1/2 F = 1 F = 2 226.1MHz 235.6MHz Figure 3.4: Potassium D2-line Doppler- free (blue) and Doppler-broadened (red) spectroscopy. The 41K lines are indicated in green. We lock on the F = 1, 2 crossover line which has an assymetric line shape. 3.2 87Rb & 39K in magnetic fields In the presence of a magnetic field B the (2F + 1)-fold degeneracy of mF states within a given hyperfine level F is lifted. In the case of low field strengths the energy shift of an |F, mF〉 state is linear, namely ∆E = gFmFµBB , (3.1) called the anomalous Zeeman effect [127]. Values for gF can be found in Fig. 3.1 and 3.3 and µB is Bohr’s magneton. For strong fields the |F, mF〉 states cease to be good quantum numbers. In this case, the levels of 87Rb or 39K with J = 1/2 (for example) are split into two groups (mJ = ±1/2) of four, i.e. 2I + 1, where I = 3/2 (Paschen- Back effect). The hyperfine and magnetic splittings are plotted in Fig. 3.5 and 3.6 for 87Rb and 39K, respectively, in both the weak- field and the strong-field regimes. For levels with L = 0 and S = 1/2 the energies can be written in an analytic form called the Breit-Rabi formula [127, 139, 140]. Interlude: calibration of magnetic coils The splitting of hyperfine states in magnetic fields represents the most accurate way of calibrating homogeneous magnetic fields in our apparatus. It is a well-understood, single-particle effect with no obvious systematic errors. The levels with J = 1/2 (e.g. the 5 2S1/2 ground level of 87Rb) are particularly well suited for magnetic field calibration as they are unaffected by the electric quadrupole correction [139]. We can experimentally determine the transition frequencies at high magnetic fields (several hundreds of gauss) with a relative uncertainty of 10−4 using Landau-Zener sweeps or Rabi oscillations. By comparing the experimental values to the theoretical predictions (Fig. 3.5 B) we can quantify our mag- netic field to similar precision. Table 3.2 shows some relevant cali- brated field values using RF transitions from the |F = 1, mF = −1〉 to the |F = 1, mF = 0〉 state in the atomic ground level of 87Rb (ar- row in Fig. 3.5 B). The calibration of our guide field coils can be found in section 4.5 using the same method. 44 quasicrystalline optical lattices for ultracold atoms 3.3 Spectroscopy and absolute frequency stability (laser lock- ing) The production and manipulation of ultracold 87Rb & 39K requires an absolute laser frequency stability (long-term and short-term) of roughly 1MHz, corresponding to a relative stability on the order of 10−9. This requirement means that the laser jitters and drifts by much less than the natural linewidth (6MHz×2pi) of the rel- evant optical transitions. In our case the short-term fluctuations (linewidth) of the free-running reference laser is in the 100kHz- range. However, the long-term (thermal) fluctuations are on the order of 100MHz such that a stable reference and a frequency- feedback-loop (‘laser lock’) are necessary. Absolute frequency sta- bility for our 87Rb- and 39K-master lasers is achieved using stan- dard saturated absorption spectroscopy and commercial Pound- Drever-Hall locking electronics which will be briefly described in the following. Relative frequency stability between separate laser diodes and the master lasers is ensured using frequency-offset- locks (see section 3.4 below). Table 3.2: Magnetic field calibration for our experiment coils using RF transitions from the |F = 1, mF = −1〉 state to the |F = 1, mF = 0〉 state in 87Rb. For each current value we take an RF-resonance scan using a Landau-Zener sweep with a half-range of 50kHz and a duration of 20ms. These frequencies are then converted to magnetic field values using the magnetic field hamiltonian (Fig. 3.5 B). Current/A RF res./MHz B-field/G 29.257 228.825 317.68(9) 32.379 253.600 351.44(9) 36.600 287.213 397.19(9) 36.650 287.650 397.78(9) Saturated absorption spectroscopy The saturated absorption spectroscopy technique [141–144] over- comes Doppler-broadening by employing the pump-probe beam configuration. In our apparatus we have a commercial spec- troscopy unit (one for each atomic species) which supplies a fibre-coupled pump-probe setup and a vapour cell. We obtain a spectroscopic absorption signal (y-axes in Figs. 3.2 and 3.4) by scanning the laser piezo (x-axes) together with the diode current (‘feed forward’) using a triangular wave. Laser locking Table 3.3: Spectroscopy, laser locking parameters and lock performance. The linewidths are measured with the ‘EagleEye’ reference cavity. parameter 87Rb 39K vapour cell temp. 50◦C 80◦C optical power/mW ∼ 1 ∼1.4 modul. freq./kHz 28.00 20.00 modulation ampl. 0.012mA laser linewidth/kHz 155 120 The master lasers are locked to spectroscopy lines using top-of- fringe (Pound-Drever-Hall [145]) locking which is insensitive to laser intensity fluctuations. The Pound-Drever-Hall technique uses a weak (typically 5 × 10−4) modulation of the diode cur- rent at a frequency around 20kHz to imprint sidebands onto the laser frequency. The spectroscopy signal is mixed down with the same 20kHz at a fixed phase delay to produce an er- ror signal that represents the derivative of the spectroscopy signal. This is then used to stabilise the laser frequency with a proportional-integral-derivative (PID) loop. We choose the most pronounced spectroscopy features as they promise the largest feedback slopes and the best locking performance; in 87Rb we use the F = 2 → F′ = 2, 3 crossover [127] line (Fig. 3.1), the 39K-laser is locked to the F = 1, 2 → F′ crossover line, as shown in Fig. 3.3. The parameters used for laser locking are itemised in table 3.3, together with the lock performance. laser cooling & trapping of 87rb & 39k 45 0 5000 10000 B-field/G -20 -10 0 10 20 Fr eq ue nc y/ G H z mJ = +12 mJ = -12 F = 2 F = 1 0 500 B-field/G -400 -200 0 200 Fr eq ue nc y/ M H z mF = 0 mF = +1 mF = -1 A B Figure 3.5: Hyperfine and magnetic splitting of electronic levels in 87Rb. (A) 5 2S1/2 level. In the presence of a weak magnetic field the (2F + 1)-fold degeneracy of each hyperfine level F= 1, 2 is lifted. For strong fields the lines are split into two groups (mJ = ±1/2) of four (2I + 1, I = 3/2). (B) 5 2S1/2 F = 1 (ground) level in the low-field regime. The black arrows indicate the magnetic dipole transitions which can be stimulated by RF-radiation in order to calibrate magnetic fields (B), or perform MW evaporation (A), as discussed in section 3.11. 0 500 B-field/G -1000 -500 0 500 1000 Fr eq ue nc y/ M H z mJ = +12 mJ = -12 F = 2 F = 1 0 10 20 B-field/G -10 0 10 Fr eq ue nc y/ M H z mF = 0 mF = +1 mF = -1 A B 0 500 B-field/G -2000 -1000 0 1000 2000 Fr eq ue nc y/ M H z mJ = +32 mJ = +12 mJ = -12 mJ = -32 0 5 10 15 20 B-field/G -40 0 40 Fr eq ue nc y/ M H z F = 0 F = 3 C D Figure 3.6: Hyperfine and magnetic splitting of electronic levels in 39K. (A) 4 2S1/2 level. (B) 4 2S1/2 F = 1 hyperfine (ground) level in the low- field regime. (C) 4 2P3/2 (D2-excited) level. There are 16 mF states which are not resolved on this scale. In the strong-field regime the states are grouped by mJ, each containing four mI states, i.e. 2I + 1, where I = 3/2. (D) 4 2P3/2 (D2-excited) level (same as C) at weak magnetic fields. Each hyperfine level F = 0,1,2,3 is split into mF states (2F+ 1). hierarchy transition λ/nm frequency B-field/G temperature hc/E E/h E/µB E/(kBT) 87Rb ‘gross’ 5 2S1/2 → 5 2P3/2 (D2) 780.241 210 384.230 484THz 3× 108 18 000K fine 5 2P1/2 → 5 2P3/2 (D1-D2) 7.123THz 5× 106 342K hyperfine 5 2S1/2 F = 1→ F′ = 2 6.834 682 611GHz 4900 0.3K 5 2P3/2 F = 2→ F′ = 3 266.650MHz 191 13mK 39K ‘gross’ 4 2S1/2 → 4 2P3/2 (D2) 766.700 921 391.016 170THz 3× 108 19 000K fine 4 2P1/2 → 4 2P3/2 (D1-D2) 1.730THz 1× 106 83K hyperfine 42S1/2 F = 1→ F′ = 2 461.720MHz 330 20mK 42P3/2 F = 2→ F′ = 3 21.1MHz 15 1mK Table 3.4: Hierarchy of some rele- vant energy scales (E) and energy- equivalents in 87Rb [139] and 39K [140]. 46 quasicrystalline optical lattices for ultracold atoms 3.4 Relative frequency stability (offset locking) 510 520 530 540 Rb repump beat note/MHz -90 -60 -30 Si gn al /d Bm Figure 3.7: Frequency scan of the beat note (after mixing) between the Rb repump and the Rb master laser. A lorentzian fit to the peak gives a centre frequency of 526.1MHz. The measurement bandwidth of the spectrum analyser was set to 1 kHz for this measurement. 90 100 110 120 130 Rb repump DDS frequency/MHz 0.0 1.0 2.0 R b at om nu m be r/ 10 7 0.1ms 1.0ms Figure 3.8: Frequency scan of Rb repump flash (during imaging) for two different pulse durations. Note that the DDS unit drives an AOM in double-pass configuration, i.e. the laser frequency is changed by twice the amount shown on the x-axis. The atom numbers are 2d gaussian fits to the atom cloud after loading the dipole trap at a temperature of 6 µK in 12ms ToF. The repump flash is applied 1.5ms before the imaging pulse in order to optically pump the atoms from the |F = 1, mF = −1〉 state to the F = 2 manifold. The lorentzian fit (line) gives a centre frequency of 110.1(1)MHz with a Γ-width of 6.5(1)MHz. Frequency shifts on the order of 100MHz are routinely performed by acousto-optic modulators (AOMs). Their radiofrequency (RF) bandwidths of up to ± 20 % of the central frequency are sufficient to tune the frequency by several natural linewidths in 87Rb or 39K (particularly if used in the convenient double-pass config- uration). However, for the two situations described below we require larger frequency shifts or wider tuning ranges, which can be achieved using an ‘offset lock’ [146]. In this configuration the laser that is to be stabilised is overlapped (on a non-polarising beam-splitter) with a reference laser and the resulting beat note is picked up on a high-bandwidth photodiode. The beat signal can be mixed down with an external RF source, for example a voltage-controlled oscillator (VCO) or a direct-digital-synthesis (DDS) unit if frequency tuning is required. Afterwards the signal is split into two paths with different cable lengths, then recom- bined on a mixer and finally low-passed. The resulting DC voltage can be approximated by a cosine [146], the argument of which is given by the offset frequency times the constant time delay (= length/speed of light) induced by the longer cable. Therefore this setup is sometimes called ‘delay-line lock’. Laser locking is then achieved by using a side-of-fringe lockpoint of the cosine as the error signal for the PID-loop. In our experiment this technique is used for two applications, namely the rubidium repump lock and the potassium imaging lock. Rubidium repump lock The ground state splitting in 87Rb is 6.8GHz (table 3.4) and in our experiment it is bridged by an offset lock. In this setup the raw beat signal is first mixed down with a 7.5GHz reference oscillator to make the RF signal easier to handle. We choose a (constant) lockpoint 220MHz to the blue of the repump transition F = 1 → F′ = 2 in 87Rb , frequency fine-tuning is then achieved by a double-pass 110MHz AOM (-1st order, Fig. 3.8). Since our ru- bidium master laser is locked 53.4MHz to the red of the crossover line (Fig. 3.1) we should set our beat note (after mixing down with the reference oscillator) to 7500.0− (6834.7− 133.3+ 53.4)− 220.0 = 525.2 (3.2) (in MHz). In practice we have a beat note (after mixing) of 526.1MHz, as shown in Fig. 3.7. In Fig. 3.8 we show a spectroscopic measure- ment of the resulting repump laser frequency which is consistent with the above considerations. This measurement was performed using the ‘repump flash’ (see section 3.12) during imaging of the |F = 1, mF = −1〉 state in the dipole trap. Since our main imaging light is only resonant with atoms in the F = 2 level the measured atom number is a good probe of the repump frequency, result- ing from the optical pumping efficiency from |F = 1, mF = −1〉 to laser cooling & trapping of 87rb & 39k 47 F = 2. For longer repump flash pulses we observe a broadening effect similar to that of power broadening. Potassium imaging lock 200 205 210 215 K master-imaging beat note/MHz -80 -60 -40 -20 Si gn al /d Bm Figure 3.9: Spectrum analyser trace of the beat note between the 39K imag- ing and the 39K master laser corre- sponding to the highest data point in Fig. 3.10 below. The measurement bandwidth of the spectrum analyser was set to 1 kHz for this measurement. -220 -200 -180 K master-imaging beat/MHz 0 2 4 K at om nu m be r/ 10 6 σ+ polarised σ− polarised Figure 3.10: Frequency scan of 39K imaging light at a guide field of 2.8(1)G (offset coils at 4.5 A), using σ+ and σ+ polarised light. Only the σ+ polarised light results in a faithful resonance signal at −208.6(1)MHz (lorentzian fit) with a Γ-width of 7.6(5)MHz. The dashed line is a lorentzian fit to the σ− data (wrong polarisation) as a guide to the eye. The frequency is red-detuned with respect to the potassium master laser (indicated by the negative values on the x-axis). The atom numbers (y- axis) are 2d gaussian fits to the atom cloud after loading the dipole trap at a temperature of 9 µK in 3ms ToF. The ground level hyperfine splitting of 39K is only 462MHz (ta- ble 3.4) and that of 40K is 1.3GHz which could, in principle, be bridged by AOMs. However, we require a much larger degree of flexibility than could be achieved by AOMs, since 39K is often im- aged at magnetic fields close to the Feshbach resonance at 402.70G (see section 4.6 on Feshbach resonances). In 39K the corresponding frequency shifts are between 200 and 500MHz to the red of the master laser; in 40K the shifts are even larger and of opposite sign (to the blue, since the ground state hyperfine splitting is inverted). Offset locks allow large tuning ranges of arbitrary sign and are therefore ideally suited for this application. In contrast to rubidium, we mix the beat signal with a DDS signal (which in turn has been mixed with a VCO to cover higher frequencies) such that we can vary the lockpoint in-sequence. Fig- ure 3.9 shows the beat signal between the potassium master laser and the imaging laser. By changing the DDS frequency we are able to obtain a spectroscopic signal of the 39K imaging transition F = 2→ F′ = 3 at low magnetic field, as shown in Fig. 3.10. It cru- cial to use σ+ polarised imaging light in order to address the cor- rect hyperfine state (Fig. 3.10 also shows a frequency scan using the wrong polarisation, in this case the resonance is shifted in a non-trivial manner). The expected imaging transition at 2.8G is at 155.5(2) MHz to the red of the D2-line (fine structure point). Tak- ing into account the position of the 39K master lock point (Fig. 3.3) we expect the imaging transition at −155.5 MHz− 53.0 MHz = −208.5 MHz , (3.3) in excellent agreement with the measured value of 208.6MHz (Fig. 3.10). Up to this point the stage is set for the atoms, and their jour- ney towards quantum degeneracy begins. In the following the individual cooling steps will briefly introduced and subsequently described with the focus on our experimental implementation. However, schematics of large-scale laser setups and lists of optical components, part numbers, etc. have been omitted in favour of readability. 48 quasicrystalline optical lattices for ultracold atoms 3.5 Magneto-optical trap (MOT) Radiation pressure can be used to cool dilute vapours by means of the Doppler effect. This was discovered as early as 1975 by Hänsch and Schawlow [147]: It is well known that light exerts a radiation pressure on any sub- stance which reflects or scatters is. [. . . ] the radiation pressure of laser light has been used to selectively deflect atoms of a chosen isotopic species from a beam. We wish to point out that if the laser radiation is essentially isotropic, but confined to frequencies on the lower half of the Doppler-broadened absorption line of an atomic vapour, the gas can be cooled. That is, the average translational kinetic energy of the atoms can be reduced. If the light is red-detuned with respect to the atomic transition, only atoms moving against the direction of the laser beam will predominantly experience photon momentum kicks and thereby be slowed down. Adding a weak magnetic quadrupole field5 to5 weaker than would be necessary to trap the atoms against gravity this situation leads to atoms that are further away from the field minimum (i.e. the centre of the quadrupole field) reaching the resonance condition more easily than atoms at the field mini- mum. Therefore the atoms experience an effective restoring (i.e. trapping) force towards the centre. This is the principle of the magneto-optical trap (MOT) which nowadays forms the basis of essentially all cold-atom experiments [125–127]. Table 3.5: MOT parameters. Laser powers are given per beam. Detunings are given in MHz from the respective transitions (negative values mean red detunings). parameter 87Rb 39K backgr. pressure 7× 10−11 mbar loading duration 2.5 s 1.0 s gradient (strong ax.) 20G/cm coil current 11.0A gradient @ 1A 1.78G/cm cooling transition F = 2→ F′ = 3 repump transition F = 1→ F′ = 2 cooling power/mW 95-130 40-46 repump power/mW 0.7-1.0 60-87 cool. detuning/MHz -23 -52 rep. detuning/MHz 0 -25 beam waist 13 mm All laser light required for MOT operation and the subsequent near-resonant cooling stages is derived from the respective ref- erence lasers (section 3.3). Frequency shifts are performed by means of double-pass AOMs which are driven by home-built DDS units and RF amplifiers. Subsequently the light is amplified us- ing six separate tapered-amplifiers (TAs). The rubidium and the potassium light is overlapped using ‘magic’ waveplates6 and then 6 which act as λ-plate at 767nm and as λ/2-plate at 780nm distributed via a commercial two-input-by-six-output setup called ‘fibre cluster’. Due to the nature of the polarising beam splitters in Figure 3.11: 2D MOT vacuum cham- ber. The frame is made of titanium and the rectangular windows are made of BK7 glass. The windows are glued to the frame using high-temperature exopy glue, with some kapton foil as spacer material. Inside the chamber (on the right) the mirror-polariser setup with its small aperture (see text) can be seen. All materials were cho- sen such that their respective thermal expansion coefficients match within 15% [124]. laser cooling & trapping of 87rb & 39k 49 the fibre cluster it is impossible to split the light from both input ports into six equal parts. Therefore the waveplates in the fibre cluster have been used to maintain pair-wise-balanced powers (at least within 10%, typically better than 5%) for both inputs. The MOT laser beams are large (table 3.5) and therefore relatively insensitive in terms of alignment, as long as they are roughly cen- tral, collimated and counter-propagating. We aligned one beam of each counter-propagating pair by coupling the light into a fibre plugged into the opposite fibre coupler. A list of relevant parame- ters is given in table 3.5. The weak magnetic field gradient is provided by the same coils that are later used for magnetic transport (section 3.10). In addition, there are three pairs of compensation coils (section 3.8) which counteract stray magnetic fields. 3.6 2D MOT Figure 3.12: Schematic of the 2D MOT chamber showing one of the four coils. The brass block on the outside contains a U-shaped hole which is perfused by cooling water in order to carry away the heat dissipated by the coil. Figure taken and adapted from ref. [124]. Figure 3.13: 2D MOT field configura- tion seen from the axial direction. Fig- ure taken and adapted from ref. [124]. Due to the low pressure in the MOT chamber (7× 10−11 mbar) we observe no spontaneous MOT loading from background vapour. Instead, the MOT is loaded using two 2D MOTs [148–152] which are located on either side of the MOT chamber, each separated by a differential pumping section that maintains a large pressure differential (as described in section 2.3, see also Fig. 2.1). Separate 2D MOTs allow us to control the background vapour pressure for each species independently, leading to improved loading rates compared to refs. [119, 120]. These 2D MOTs are used to precool the atoms and produce a collimated beam7 of 87Rb/39K-atoms 7 this method was also proposed in the original paper by Hänsch and Schawlow [147] that can be directly captured by the main ‘3D MOT’. The basic principle of the 2D MOT is the same as that of the 3D MOT, ex- cept that there is one direction along which less cooling takes place.8 In our 2D MOT design (Fig. 3.11) this manifests itself by an 8 Sometimes this is called a 2D+ MOT since cooling also takes place along the third direction. However, the notion of cooling along the third axis depends on the precise relative power levels between the different axes. In this work the term 2D MOT will be used for simplicity. elongated geometry9 and by an aperture in the axial retro-mirror 9 there is no magnetic field gradient along this axis and hence no restoring force through which atoms leave the chamber. In addition, a polariser is glued to the face of the retro-mirror (with some added spacers, as shown in Fig. 3.11). The purpose of the axial beam is then twofold: On the one hand, it provides some cooling in the axial direction (though this contribution is limited due to the ring-shaped beam profile coming back from the retro-mirror). On the other hand, the central part of the axial beam serves to ‘push’ atoms towards the 3D MOT. The mirror-polariser setup allows us to balance the relative strength of cooling and pushing by changing the (linear) polarisation of the incoming axial beam. For potassium we have an additional laser beam along this axis that is blue-detuned in order to push the atoms through the aperture. In contrast to the 3D MOT all beams are retro-reflected. A summary of relevant 2D MOT parameters is given in table 3.6. The 2D MOT vacuum chambers were assembled in-house, closely following the Munich implementation [153, 154]. The elongated geometry of the 2D MOT is intended to provide a large 50 quasicrystalline optical lattices for ultracold atoms capture volume (though this effect cannot easily be quantified). A detailed account of the design and setup of our 2D MOT chambers can be found in Matteo Sbroscia’s First-year report [124]. Table 3.6: 2D MOT parameters. Laser powers are given per beam. Detunings are given in MHz from the respective transitions (which are identical to those in the MOT). Negative frequency values mean red detunings. Here x and y directions are parallel to the transport axis and along gravity, respectively. parameter 87Rb 39K vapour pr./mbar 4× 10−7 1× 10−6 diff. pumping  1.5mm diff. pump. length 2 cm duration 2.5 s 1.0 s field gradient 21G/cm 22G/cm coil current 8.5A 9.0A gradient @ 1A 2.4G/cm coil windings 63 per coil heat dissipated 30W per coil comp. windings 8 per coil comp. current x 0.0A 2.5A comp. current y 2.5A 0.6A cooling detuning −19 MHz −28 MHz repump detuning 0 MHz −15 MHz push detuning - +42 MHz cooling power x 230mW 190mW cooling power y 320mW 260mW repump power x 48mW 230mW repump power y 39mW 170mW axial cooling 8.5mW ∼1mW push power - 2.0mW beam waist (long) 20 mm beam waist (short) 10 mm axial beam waist 2mm push beam waist - 0.8mm The magnetic gradients in each 2D MOT are provided by four rectangular coils, one on each face of the chamber, as shown in Fig. 3.12. They produce a quadrupole-like field in the plane per- pendicular to the axial direction (Fig. 3.13); field gradients in the axial direction are negligible. Since considerable power is dissi- pated by the 2D MOT coils (table 3.6) they are water-cooled in parallel with the transport coils (section 3.10). The alignment of the axial beam (and its counter-propagating partner) is crucial for proper operation of the 2D MOT. A good alignment technique is to first roughly maximise transmission of the axial beam through the whole apparatus (coming out of the opposite 2D MOT chamber). At this point the reflected, ring-shaped beam should exit the entrance viewport which can be made visible by holding a viewing card with a hole in place. Probably the retro-reflected beam is then not exactly overlapped with the incoming beam. This can be achieved by further beam- walking, whilst keeping the ring-shaped profile approximately constant. The alignment of the transverse beams is not crucial; it has almost no effect on the MOT loading rate provided that the beams are collimated and roughly counter-propagating. Fluorescence signal In order to characterise the loading of the MOT we monitor its fluorescence. Scattered light is collected on a lens at a vacuum viewport on the knee (Fig. 2.1), focussed, split into two parts, and filtered by wavelength before it reaches a photodiode (one for each species).Typical scope traces are shown in Figs. 3.14, 3.15, and 3.16. 0.0 1.0 2.0 Time/s 0 10 20 30 87 R b ph ot od io de /m V 87Rb alone Figure 3.14: Fluorescence signal for MOT loading without potassium (rubidium only). In potassium the optical powers for repump and cooling light are comparable since the upper hyperfine levels are closely spaced (Fig. 3.3) and there is no closed transition. The necessity for higher laser powers leads to increased photon scattering and therefore limits the achievable temperatures in a 39K-MOT to values well above the Doppler-limit [127] of 145 µK. Additionally, we expect light-assisted collisions to play a role in the comparatively modest performance of the 39K-MOT in the presence of 87Rb [155, 156]. Compared to the 39K-only case (Fig. 3.15) we load roughly half as many 39K-atoms if rubidium is present (Fig. 3.16). However, the rubidium is almost unaffected by the potassium; this is impor- tant because we rely on abundant 87Rb-atoms as coolant to reach quantum degeneracy in 39K (section 4.3). laser cooling & trapping of 87rb & 39k 51 3.7 Final MOT stage in potassium 0.0 1.0 2.0 Time/s 0 2 4 6 39 K ph ot od io de /m V 39K alone Figure 3.15: Fluorescence signal for MOT loading without rubidium (potassium only). 0.0 1.0 2.0 Time/s 0 2 4 6 39 K ph ot od io de /m V 87Rb 39K 0 10 20 30 87 R b ph ot od io de /m V Figure 3.16: Fluorescence signal for MOT loading with both species present. (oscilloscope on high- resolution mode for all these mea- surements). Once loading of potassium is saturated (Fig. 3.16) we address the problem of strong photon (re-)scattering (described above) by low- ering the 39K-repump power in the final 10ms of the MOT. During this ‘final MOT’ stage we linearly decrease the 39K-repump power to 2% of its original value. Additionally, we decrease the detuning of the 39K-cooling light to −28 MHz at the beginning of the final MOT stage (the repump detuning remains essentially unchanged). We observe a significant increase in 39K atom number using this procedure. Explaining this effect in detail is non-trivial due to the complexity of the process (dual species MOT, non-equilibrium system, unresolved 39K-hyperfine structure). However, the most plausible explanation seems that of a ‘temporal’ dark MOT [118, 157, 158] in which further heating due to photon (re-)scattering is reduced by 39K-atoms accumulating in the dark F = 1 hyperfine level. Several experiments on 39K have found that increasing the MOT gradient during the final MOT stage (‘compressed MOT’ or cMOT [159]) has a positive effect on the MOT performance [120, 135, 160]. Other experiments (on lithium) have found the op- posite to be true [161]. Yet different experiments (on rubidium) have shown a somewhat chaotic behaviour to produce the best results [162]; they employ an artificial neural network to perform multiparameter optimisation of the final MOT stage. In our case we observe strong fluctuations in the system when increasing the MOT gradient and consequently we leave the MOT gradient un- changed in this stage of the experiment. Similarly, there seems to be no consensus as to whether the detuning of the cooling light should be increased [120, 135] (i.e. shifted away resonance), de- creased [160], or modulated in a non-trivial manner [162]. We find decreasing the detuning of the cooling light (shifting its frequency towards resonance) results in the best performance. 3.8 Optical molasses Table 3.7: Molasses parameters. Laser powers are given per beam. Detunings are given in MHz from the respective transitions, which are the same as for the MOT (negative values mean red detuning). parameter 87Rb 39K duration 4.5ms cooling power 26-35mW 16-18mW repump power 0.7-1.0mW 1.9-2.8mW cooling detuning −63 MHz −4 MHz repump detuning 0 MHz −45 MHz A typical 87Rb-39K-MOT contains on the order of one billion atoms of each species at temperatures in the millikelvin-regime. The standard procedure to reach colder temperatures is the optical molasses technique [163, 164]. Its setup is identical to that of the MOT except that all magnetic field gradients are switched off. On the one hand, this means there is no restoring force towards the centre and consequently the atoms are not trapped any longer. Consequently the duration of this cooling procedure is limited to a couple of milliseconds. On the other hand, polarisation-gradient cooling mechanisms [165] which were previously confined to the trap centre are now happening across the entire atomic cloud. The latter effect is so strong that within 4.5ms the temperature of the cloud drops by almost a factor of ten. Our experimental parame- ters are summarised in table 3.7. 52 quasicrystalline optical lattices for ultracold atoms MOT compensation coilsTable 3.8: MOT compensation coil pa- rameters. Here we use the convention that the x-axis (yellow wire) coincides with the direction of magnetic trans- port, the y-axis (blue) points along the axial direction of the 2D MOTs, and the z-axis (grey) is aligned with gravity. The coils for optical pump- ing (spin-polarisation, s-p) are also wound on the z-axis (black wire). The number of windings is given per coil, the conversion factor between field and current in G/A, and the optimum current values in A. parameter x y z s-p windings 6 5 34 6 15 conversion (G/A) 0.4 0.4 0.37 0.92 current/A 1.5 1.5 0.0 - During optical molasses any residual magnetic field gradients can lead to directional forces and thus inefficient cooling. Therefore we use three pairs of ‘compensation coils’ to counteract stray mag- netic fields caused by ion pumps, optical isolators, and Earth’s magnetic field. The compensation coil parameters are listed in table 3.8; the optimum current values were found by maximising the number of atoms after transport. Changing the compensation fields has negligible effect on the cloud temperature after trans- port. The optical isolators for the lattice are the dominant source of stray fields (by a factor of two) since placing them caused a sizeable shift in the optimum currents. 3.9 Optical pumping (spin-polarisation) Table 3.9: Optimum parameters used for optical pumping of 87Rb. Detunings are given in MHz from the unshifted transitions (at zero field). Negative values mean red detuning. parameter value total duration 0.7ms pumping duration 100 µs repump duration 45 µs offset field 10G pumping detuning −3 MHz repump detuning +4 MHz pumping power 1.5mW repump power 0.4mW beam waist 3mm After optical molasses the atoms need to be magnetically trapped and transported to the science cell for further (evaporative) cool- ing. Only weak-field seeking mF states (i.e. states with a pos- itive zeeman shift, gF × mF positive in Eq. 3.1) can be trapped using magnetic forces.10 In the electronic ground levels of 87Rb 10 since local field maxima are pre- cluded by Earnshaw’s theorem [124] and 39K there are three trappable states: |F = 2, mF = 1, 2〉 and |F = 1, mF = −1〉. After the optical molasses stage we optically pump the 87Rb atoms into the |F = 1, mF = −1〉 state using a flash of σ− polarised pumping (F = 2 → F’ = 2) and repumping light (F = 1 → F’ = 2). Two frequencies are necessary to perform both Zeeman pumping (mF-states) and hyperfine pumping (correct hy- perfine level F = 1). The repump light is only on for the first half of the pulse. An offset field is provided by dedicated coils that are wound around the z-MOT-compensation coils (table 3.8). The cur- rent for the offset field is produced by a capacitor discharge unit; if needed the polarity of the coils can be switched using a current- H-bridge. All relevant parameters can be found in table 3.9; they were found by maximising the number of captured atoms in the magnetic trap. 0 60 120 180 λ/4-waveplate/degrees 0 4 8 C ap tu re d at om s/ 10 8 data no s-p Figure 3.17: Optical pumping effi- ciency. The number of captured atoms is plotted as a function of pumping- polarisation (λ/4-waveplate) in 87Rb. Atom numbers are gauss sum fits to fluorescence images in 5ms ToF after release from the magnetic trap. The minimum (maximum) atom numbers correspond to almost σ+ (σ−) polar- isation, probably with some residual ellipticity. In order to estimate our pumping efficiency we vary the polari- sation using a λ/4-waveplate (Fig. 3.17). The maximum and min- imum numbers of captured atoms in Fig. 3.17 can be interpreted as most atoms being in |F = 1, mF = −1〉 and |F = 1, mF = 1〉 (un-trapped), respectively. Since these data points are roughly symmetric around the situation without any optical pumping (‘no s-p’) we can conclude that our Zeeman- and hyperfine-pumping are both in principle working properly. The deviation from the expected cosine-behaviour leads us to believe that our pumping efficiency is limited by birefringent optical elements between the waveplate and the position of the atoms (possibly the vacuum windows). This birefringence may prohibit perfect σ+ and σ− polarisation and therefore lead to a reduced optical pumping ef- ficiency of roughly 75%. In principle, the birefringence could be laser cooling & trapping of 87rb & 39k 53 accounted for by placing a λ/2-waveplate but this was not found to be crucial for BEC production. There is no general consensus as to which hyperfine state should be used for transport. Some experiments [120, 160] opt for the |F = 2, mF = 2〉 state because its trapping gradient is twice as large compared to states with |mF| = 1. Other experiments [135] report the |F = 1, mF = −1〉 to be beneficial. We use the latter state since it gives the best results (condensate numbers). This could be explained by the improved performance of the MW evaporation (section 3.11) using the lower (F = 1) hyperfine manifold. For 39K we have so far not managed to effectively pump atoms into the |F = 1, mF = −1〉 state11 and consequently we do not perform any 11 however, pumping to the |F = 2, mF = 2〉 state works efficientlyoptical pumping for this species. Potentially the relatively strong offset field of 10G precludes any efficient Zeeman pumping since at this field the upper hyperfine levels (Fig. 3.6 D) are split in a nonlinear way. Moreover, the final MOT stage for 39K leads to a large occupation of the F = 1 manifold which in our case suffices to perform further cooling without any optical pumping. 3.10 Magnetic trapping, transport, and experiment coils Magnetic quadrupole traps [127, 166] are used to capture the atoms in the MOT chamber, transport them to the science cell, and perform MW evaporation (section 3.11). Conceptually the magnetic trap relies on a magnetic field minimum at which atoms with positive Zeeman shift accumulate (gF ×mF has to be positive in Eq. 3.1). The typical quadrupole coil configuration consists of two axially aligned coils with counter-propagating currents; this situation leads to a field minimum in between the two coils that is at exactly zero field with linear gradients in all directions. The main parameter characterising a magnetic trap, which for all practical purposes has infinite trap depth, is its gradient. Due to the second Maxwell equation ∇ · B = 0 the field gradient in the axial direction is twice as big as the gradients in the planar direction, leading to a 2:1-aspect ratio of the trapped atomic cloud. In this work the trap gradient is always quoted along the ‘strong axis’. Table 3.10: Parameters used for mag- netic trapping and transport. parameter value total distance 64.35 cm total duration 4.99 s acceleration 0.2m/s2 max. speed 0.5m/s sampling rate 1kHz MOT gradient @ 1A 1.78G/cm MOT grad. (capture) 53G/cm MOT grad. (transport) 100G/cm exp. gradient @ 1A 3.5(1)G/cm exp. grad. (transport) 100G/cm exp. grad. (compressed) 300G/cm A series of overlapping magnetic traps (Fig. 3.18) can be used to transport atoms over relatively large distances (64 cm in our case) in a ‘conveyor belt’ motion. By choosing appropriate cur- rent ramps in the successive coil pairs the aspect ratio12 can be 12 between the longitudinal gradient along the transport axis and the transverse gradient kept within reasonable bounds in order to avoid heating due to re- peated compression and decompression of the cloud [117]. In our configuration (Fig. 3.19) this aspect ratio varies from 1.0 (when the cloud is at rest) to 2.4 (at the last coil pair before the atoms reach the experiment coils). More information on the design of the coil system can be found in the Master’s thesis of Oliver Brix [167]. 54 quasicrystalline optical lattices for ultracold atoms Figure 3.18: Schematic of magnetic coils used for MOT & magnetic cap- ture (2, green), transport (3/4, blue), MW evaporation and Feshbach fields (5, red). All coils except the experi- ment coils (5) and the push coil (1) are glued into a brass cooling block (only the lower cooling block is shown for clarity). Figure taken from [124]. 2 3 5 1 4 Figure 3.19: Coil currents used for magnetic transport. These were determined by maintaining an ap- proximately constant aspect ratio of the atomic cloud over the travelled distance [117, 167]. The colour code is the same as in Fig. 3.18. 0 10 20 30 40 50 60 Distance/cm 0 20 40 60 80 100 C ur re nt /A MOT push 1 - 16 17 exp m ot 1 2 3 4 5 6 7 8 9 10 12 13 14 15 16 ex p Coil index 0.0 0.5 1.0 Fr ac ti on (r et ur n jo ur ne y) 87Rb alone Figure 3.20: Fraction of rubidium atoms recaptured after transport (return journey). The experiment coils themselves serve a further purpose, in addition to providing a strong magnetic confinement for transport and MW evaporation, namely to create homogeneous Feshbach fields (section 4.6). They are designed such that if the polarity of one experiment coil is flipped the resulting field at the position of the atoms is as homogeneous as possible. In our case we ex- pect residual curvatures on the order to 10−4 mG/µm2 at field strengths of several hundred gauss [167]. If we had a uniform optical trap for the atoms [168] these residual curvatures would result in effective trap frequencies on the order of 1Hz. m ot 1 2 3 4 5 6 7 8 9 10 12 13 14 15 16 ex p Coil index 0.0 0.5 1.0 Fr ac ti on (r et ur n jo ur ne y) 39K alone 39K with 87Rb Figure 3.21: Fraction of potassium atoms recaptured after transport (return journey), with and without rubidium present. The sequence for magnetic transport starts with trapping the atoms after optical pumping in the MOT chamber using a gradient of 53G/cm. Subsequently the MOT field gradient is ramped to 100G/cm and the atoms are transported using the current values shown in Fig. 3.19. In order to convert the distances to timestamps in sequence we use a constant acceleration until we reach a certain max. speed (and vice versa for decelerating). The atoms come to a halt at the 45◦ bend in the transport path (‘knee’ in Fig. 2.1). All relevant transport parameters are summarised in table 3.10. The transport efficiency is characterised by performing a return- journey to the experiment coils and back. Afterwards the number of atoms that survived is counted (using fluorescence imaging in the MOT chamber), as shown in Figs. 3.20 and 3.21. The result- laser cooling & trapping of 87rb & 39k 55 ing transport efficiency is then given by the square root of the fraction of atoms remaining after the return journey (table 3.11). In general, the transport efficiency improves with higher gradi- ents but we are limited to 100A by our power supplies. Another way to enhance efficiency would be to transport the atoms in the |F = 2, mF = 2〉 state which experiences a trap gradient that is twice as large. However, the subsequent MW evaporation stage works better in the |F = 1, mF = −1〉 state and therefore we trans- port the atoms in this state. We observe that magnetic transport is generally more efficient for 39K than for 87Rb (Fig. 3.21). We also observe that lower acceleration improves the efficiency for 87Rb. Table 3.11: Single-journey transport ef- ficiencies for rubidium and potassium. 87Rb alone 64% 39K alone 80% 39K with 87Rb 66% Technical implementation of transport and experiment coils Table 3.12: Experiment coil parameters. Note that the ribbon height has been reduced by taking off < 0.5mm from one side. The temperatures quoted are steady-state values for a 50% duty cycle at 100A; these are reached after about 10minutes while the cooling block is stabilised to 21.5◦C. The temperature values are measured using Pt100 sensors that are directly glued to the coil using thermally conductive epoxy glue. parameter value windings 75 per coil ribbon height 16mm ribbon width 0.5mm resistance upper coil 66.2mΩ resistance lower coil 66.2mΩ temperature upper coil 45◦C temperature lower coil 40◦C All transport coils – except the pair of experiment coils – are glued into a brass block (Fig. 3.18) that is watercooled and actively sta- bilised to 21.5◦C (which is also the set temperature of the laser- table air conditioning). More information on the design of the cooling block can be found in ref. [167]. The experiment coils were wound in-house using copper ribbon wire, kapton insulation be- tween the windings, and thermally conductive but electrically insulating epoxy glue. The side of each coil that faces the brass block was milled down slightly (using a fly-cutter) and lapped in order to ensure flatness and thereby improve thermal contact with the brass block.13 Between the coil and the brass block there is an 13 despite this procedure there was some residual unevenness, probably due to the coil not being fully rigid in itself electrically insulating (but thermally conducting) sheet, as well as heat compound paste that is directly applied to the coil face. Carbon plates are screwed to the cooling block from both sides in order to fix the experiment coils in place. The tightness of the screws was carefully adjusted to optimise cooling performance (table 3.12). Unfortunately, there is a mechanical instability in the upper experiment coil that makes fast switching impossible. If the upper coil is switched off suddenly using a MOSFET switch the effective field strength of the coil is reduced; consequently the position of the magnetic trap moves upwards. Repeated pulses of this kind lead to a degrading of field strength over time that takes several weeks to relax. We observed shifts of several hundred µm, corresponding to 1-2A of current ‘missing’ in the upper coil. We attribute the effectively reduced current to electrical shorts between the windings. These could be potentially introduced by a jerk motion during the sudden switch-off of the coil (rough estimates suggest an outwards force on the order to 10N within one coil caused by the magnetic field of the other coil). The field strength of the lower coil is insensitive to a sudden switch-off. We have avoided fast switching for several months now and the cloud position has been relatively stable since then. 56 quasicrystalline optical lattices for ultracold atoms Compensation coils (offset fields) 20 40 60 80 Experiment coils current/A -40 0 40 80 Ve rt ic al in -s it u po si ti on /µ m 0.0 0.2 0.4 0.6 0.8 1.0 Figure 3.22: Optimisation of compen- sation coil fields using the movement of the quadrupole trap centre as figure of merit. Data is shown for six different current values in amps (colour-coded in different shades of blue) of the Z-compensation coils, the optimum being around 0.2A (least movement). Similar to the MOT chamber we use three pairs of rectangular compensation coils around the science cell to counteract any stray magnetic fields. In an ideal quadrupole configuration the loca- tion of the field minimum does not change with the trap gradient. However, adding an offset field leads to a dependence of the trap centre on its gradient. Consequently we can use the trap move- ment to characterise our ambient magnetic fields and compensate for them by minimising the gradient-dependence of the trap cen- tre (Fig. 3.22). After describing transport and the necessary magnetic hard- ware we will now outline the evaporative cooling sequence in the steep magnetic trap. 3.11 Microwave (MW) evaporation The first stage of evaporative cooling in our experiment takes the atoms from hundreds of µK down to the single-digit µK regime, cold enough to be trapped by the optical dipole trap (Chapter 4). In this technique the highest-lying atoms in the magnetic trap (i.e. the hottest ones) are selectively removed via radio frequency transitions to untrapped states. Evaporative cooling in magnetic traps was first proposed for hydrogen [169] and then demon- strated in 1995 [170, 171] in the wake of the first Bose-Einstein condensates. It remains an important technique to this day due to its conceptual simplicity, particularly in the context of atom-chip microtraps [172]. Future large-scale cold-atom experiments will probably move towards all-optical production of ultracold gases, e.g. loading the dipole trap directly from grey molasses [133], thereby avoiding the technical challenges posed by high-current carrying coils (see above). However, in our setup the possibility of selectively removing 87Rb-atoms (and not the 39K-atoms) is crucial to our sympathetic cooling strategy.1414 For more information on sympa- thetic cooling, please see section 4.3 For evaporation the |F = 1, mF = −1〉 → |F = 2, mF = −2〉 transition in the ground 52S1/2-level of 87Rb (arrow in Fig. 3.5A) is addressed. After transport the 87Rb atoms have a temperature of roughly 500µK in the steep magnetic trap (300G/cm). The initial frequency for MW evaporation is chosen such that only the very hottest atoms in the trap are removed at first. In our case we start at a frequency of 6676MHz which addresses atoms situated at a magnetic field around 76G (Fig. 3.5 A); this magnetic field cor- responds to a thermal energy (kBT) of 5mK. In an ideal thermal gas at 500µK only a tiny fraction (5× 10−5) of atoms have more kinetic energy than this, which justifies our starting frequency. Subsequently the MW frequency is lowered to 6827.2MHz in a 7.5 s ramp. laser cooling & trapping of 87rb & 39k 57 Technical implementation The MW radiation for evaporation is generated by mixing the output of a home-built DDS unit [173] with a reference oscillator at 6500.0MHz. The signal is then amplified several times (10W max. power) and brought to the atoms by a waveguide and a capped horn antenna which is situated around 13 cm from the atoms in direct line of sight. The backreflection from the antenna is dumped via a circulator into a high-power attenuator. Majorana losses The ‘kink’ at the centre of a magnetic quadrupole trap starts to dominate the (quasi-)equilibrium15 in the trap for temperatures 15 Perfect equilibrium for a thermal gas in our system can never be reached since particle number is not conserved. There is always an (exponentially small) fraction of atoms at very high energies that can escape the trap. Subsequent re-equilibration leads to an additional evaporation effect which is negligible for our experimental timescales. in the µK regime. In this region, atoms passing very close to the magnetic field minimum experience a sudden change in direction of the magnetic field, leading to a non-adiabatic spin-flip [125]. In our case this means that trapped atoms in the |F = 1, mF = −1〉 state become un-trapped (|F = 1, mF = 1〉) and lost from the cloud. Even worse, since the atoms in the centre of the trap are the cold- est ones, these atom losses lead to heating due to rethermalisation of the remaining atoms. Therefore we terminate MW evaporation at temperatures around 6.6µK before Majorana losses (Fig. 3.23) and heating (Fig. 3.24) become dominant. 0.0 0.5 1.0 1.5 2.0 Hold time/s 0.0 0.5 1.0 N or m al is ed at om nu m be r 6800.0 GHz 6827.2 GHz 6833.0 GHz Figure 3.23: Majorana losses for 87Rb (no 39K present) for three different fi- nal MW frequencies. The hold time in a steep magnetic trap (300G/cm) after termination of the MW evaporation is plotted on the x-axis. Temperatures and atom numbers are measured after release from the dipole trap in 33ms ToF. 0.0 0.5 1.0 1.5 2.0 Hold time/s 4 6 8 10 12 Te m pe ra tu re /µ K 6800.0 GHz 6827.2 GHz 6833.0 GHz Figure 3.24: Majorana heating for 87Rb in the steep magnetic trap for three different final MW frequencies. Before we turn to the dipole trap and the physics of Bose- Einstein condensation in Chapter 4 the experimental aspects of absorption imaging are briefly discussed in the following. The time-of-flight imaging technique is covered in more detail in sec- tion 6.2 in the context of optical lattices. 3.12 Absorption imaging The absorptive behaviour of a material in the presence light, char- acterised by the Beer-Lambert law, has been known for more than a century. In the case of an atomic vapour and laser light that is resonant with a closed transition in the atom the laser inten- sity is simply reduced exponentially with the distance travelled through the vapour. Thus we can extract the density profile of an atomic cloud by taking a sequence of two images of the laser beam profile, one with the atomic cloud (seen as a shadow) and one without [125]. The optical density of the cloud is then given by OD(x, y) = σ ncolumn(x, y) = − ln ( I(x, y) I0(x, y) ) (3.4) Here σ is the absorption cross-section of a single atom, ncolumn is the column density, I(x, y) is the ‘atom’ picture, and I0(x, y) is the empty picture.16 16 in practice a third image is taken without any laser light (background image) an this image is subtracted from both I(x, y) and I0(x, y) All the measurements on Bose-Einstein condensates (with and without lattices) were performed using this technique. Usually we release the atoms from the dipole trap (see chapter 4 below) and 58 quasicrystalline optical lattices for ultracold atoms perform absorption imaging in 33ms time-of-flight (ToF) on the Z- axis (along gravity). There is resonant imaging and repump light available for both species on all individual lattice axes (named X, Y, D, and T, according to Fig. 5.17 in Chapter 5). In some special cases it can be useful to image the cloud ‘in- situ’, i.e. in the trap, for example to measure the position of the magnetic trap (Fig. 3.22). Note that our capabilities of taking im- ages in ToF from the magnetic trap are limited since we cannot shut down the coil currents quickly enough (section 3.10). Short imaging pulses of 100µs with low intensity17 (I  Isat) are typi-17 such that σ is only weakly depen- dant on I cally used. At low magnetic fields, both 87Rb and 39K are imaged on their respective 2S1/2 |F = 2, mF = 2〉 → 2P3/2 |F = 3, mF = 3〉 transition which comes closest to being a cycling transition. It is crucial for the imaging light to have a definite polarisation with respect to the quantisation axis of the atoms, given by a homogeneous magnetic field. In particular, the low-field imaging scheme requires σ+ polarised light. A wrong polarisation can lead to an effective shift in the imaging transition. Figure 3.10 nicely illustrates this effect to the extent that σ− polarised light causes a shift by several MHz. These imaging resonance scans were taken using a cold atomic cloud released from the dipole trap. Repump flash In order to use the |F = 2, mF = 2〉 → |F = 3, mF = 3〉 cycling tran- sition the atoms have to be prepared in the 2S1/2 |F = 2, mF = 2〉 state, or at least the F = 2 level within 2S1/2, prior to imaging. To this end, a dedicated flash of repump light (see Figs. 3.1 and 3.3) may be necessary, depending on the current stage of the experi- mental sequence. Low-field imaging in 39K is almost never used due to the slowness of our experiment coils (section 3.10), never- theless, 39K repump light is available on two imaging axes in the science cell. In the case of 87Rb, the repump flash is almost always used.18 Figure 3.8 shows a resonance scan of the repump flash18 since the 87Rb atoms spend almost all their lives in the ground F = 1 manifold frequency (which can be tuned using an AOM in double-pass con- figuration) using a cold atomic cloud. In order to maximise the detected atom number, a relatively long flash of 1.0ms is used in the experiment. High-field imaging in 39K Table 3.13: High-field imaging reso- nances for 39K. B-field/G Resonance/MHz 317.38 -301.4 351.44 -352.4 397.19 -419.0 In contrast to 87Rb, the potassium atoms are almost exclusively imaged at strong magnetic fields (see section 4.6 below). The atomic levels in 39K are detailed in Fig. 3.6 for both low and high magnetic fields. In this case, the |F = 1, mF = 1〉 state connects to the |mI = 3/2, mJ = −1/2〉 at high magnetic fields. Imaging is per- formed on a transition from this state to the |mI = 3/2, mJ = −3/2〉 state in 2P3/2. Unfortunately, the mixing of hyperfine states at high fields precludes the existence of a cycling transition and leads a depumping effect [174] of approximately 3% per cycle. The neg- laser cooling & trapping of 87rb & 39k 59 ative effects of depumping can be mitigated by performing an independent measurement of atom number by means of locating the BEC transition (section 4.7). High-field imaging of 39K is routinely performed in our exper- iment and some measured resonance frequencies are shown in table 3.13. The experimental resonances agree excellently with the theoretical prediction from Fig. 3.6, including the electric quadrupole correction [140]. As with low-field imaging, the po- larisation of the imaging light is crucial here. To this end, a setup containing a λ/2 and a λ/4 waveplate is used to optimise the imaging polarisation using atom number as figure of merit. 320 360 400 B-field/G -300 -340 -380 -420Im ag in g re so na nc e/ M H z data theory Figure 3.25: 39Kimaging resonances at strong magnetic fields. The agreement between the experimental values (ta- ble 3.13) and the theoretical prediction (Fig. 3.6) is better than 0.6MHz. Stern-Gerlach imaging In order to discern the population of mF states of a given atomic cloud, the Stern-Gerlach imaging technique is used. The cloud is split according to the magnetic moment of each atom by subject- ing the atoms to a magnetic field gradient while they are in free fall (during time-of-flight). In our apparatus a dedicated gradient coil has been wound in a 3D-printed mount that also contains the RF-antenna (section 4.5), just above the science cell. It is driven by a home-built capacitor discharge unit that is designed to provide a short pulse of high current (table 3.14). This coil provides a vertical field gradient of > 10 G/cm such that the cloud is split vertically into the different mF components. For 87Rb, a 2ms current pulse at the beginning of ToF (15ms) leads to a vertical displacement of 246 µm between each mF com- ponent. Table 3.14: Parameters for Stern- Gerlach imaging. parameter value coil windinds 360 peak current 12A initial rise time 50 µs pulse duration 87Rb 3ms pulse duration 39K 1ms It should be noted that neither the repump flash, nor the Stern- Gerlach imaging technique has been achieved after the magnetic trap but only in the dipole trap. This discussion of absorption imaging, including specific ap- plications thereof, such as high-field imaging in 39K and Stern- Gerlach imaging, concludes the chapter on experimental methods for laser cooling. 4 Quantum gases Up to this point in the experimental sequence (table 2.1) the atoms have been cooled, trapped, transported to the science cell, and cooled down further to µK temperatures. During these stages of the sequence the atom cloud represents a classical, thermal gas. However, at even lower temperatures the collective prop- erties of the gas start to be influenced by the wave-like nature of particles, i.e. quantum mechanics. The first important mani- festation of collective quantum-mechanical behaviour lies in the collisional properties at low momenta. This will be briefly dis- cussed in the context of Feshbach resonances (section 4.6).1 Once 1 Note that the bosonic or fermionic particle symmetry already influences thermodynamic properties, namely via the type of prevalent low-energy col- lision, long before quantum-statistical considerations become relevant. the thermal wavelength of the particles becomes comparable to the inter-particle distance, the gas has to be described fully quantum- statistically, reflecting the strikingly different low-temperature phases of fermions and bosons (section 4.7). The chapter begins with a brief introduction to optical dipole traps. Afterwards our specific setup, namely a crossed-beam dipole trap with elliptical waists, is characterised theoretically and experimentally. Sympathetic cooling is discussed, together with the general properties of Feshbach resonances in 39K. Finally the concept of Bose-Einstein condensation is introduced. 4.1 Dipole trap: introduction The interaction between atoms and monochromatic light can be broadly divided into two separate regimes. On the one hand, near-resonant light leads to dissipative behaviour and, conse- quently, photon scattering. This effect is utilised by laser cooling methods such as the molasses technique (section 3.8). On the other hand, far-detuned light acts dispersively, leading to a con- servative potential for neutral atoms which is often referred to as the dipole potential. Relevant formulae can be derived both semi-classically [127, 175] and purely quantum-mechanically [165, 175, 176]. The latter derivation is based on the Jaynes-Cummings model, describing a two-level atom coupled to a quantised light- field (a coherent state) resulting in a shift of the two atomic levels, called light-shift or AC-Stark shift. Semi-classically, the light-shift can be understood as the average energy of a dipole in the pres- ence of an oscillating electric field, characterised by the real part 62 quasicrystalline optical lattices for ultracold atoms of the polarisability of the atom. In this chapter we will not de- rive the formulae from first principles but instead refer to Grimm et al. [175]. In our experiment we use a standard red-detuned dipole trap at 1064nm such that an intensity maximum results in a trough in the potential. The photon scattering rates in the trap centre are on the order of millihertz and can thus be neglected on experimental timescales. Choice of trapping geometry Both the loading of the dipole trap and the subsequent evapora- tion to quantum degeneracy contain several non-linearities, such that the choice of optimal dipole beam parameters a subtle trade- off between many different variables. Several factors affect the loading efficiency of the dipole trap from a magnetic quadrupole trap. Firstly, a high spatial overlap between the two traps is required. The aspect ratio of an atomic cloud confined to the centre of a radially symmetric magnetic quadrupole field is always approximately 2 : 1 (as discussed in section 3.10). This favours radially symmetric, oblate dipole traps of similar aspect ratio. Secondly, the trapping volume should be large, encouraging large beam waists, so as to transfer as many atoms as possible to the dipole trap. But for the same reason the trap should also be as deep as possible which, given a fixed avail- able laser power, favours small waists. The dipole trap geometry also crucially affects the final stages of the experimental sequence, including optical evaporation and loading into the optical lattice. Optical lattice experiments in two dimensions profit from approximately homogeneous density profiles, favouring oblate traps with high aspect ratios. In addi- tion, anticonfining effects of blue-detuned lattice potentials can be cancelled2 by suitably configured dipole traps. Moreover, the2 In the case of a one-dimensional, blue-detuned, simple-cubic lattice the resulting anticonfining poten- tial is gaussian in both horizontal and vertical directions which can be compensated exactly by an attractive dipole potential with a √ 2 times larger waist. For two-dimensional lattices the cancellation is approximate. Here, instead of an in-plane crossed-beam trap a single dipole beam in the axial direction can be used to cancel the quadratic terms in the anticonfining potential. geometric mean of the trapping frequencies ω¯ = 3 √ ωzω2r should be kept as large as possible in order to avoid very high densities at the centre of the trap which can lead to three-body losses. Finally, the efficiency of sympathetic cooling using a coolant with a larger atomic mass (e.g. cooling 39K with 87Rb) depends in a subtle way on the gravitational sag in the dipole trap (as discussed in the fol- lowing). On the one hand, maximising the spatial overlap between the two species corresponds to having a small differential sag and very small vertical waists. On the other hand, ideal sympathetic cooling of 39K keeps all the potassium atoms in the trap, evap- orating only 87Rb atoms. It turns out that the parameter region of optical powers for which this is the case is larger for smaller vertical waists. Table 4.1: Dipole trap beam param- eters at the position of the atoms, as determined by measuring trap frequencies. axis w0h/µm w0v/µm max. pow. x 285(5) 55(2) >7W y 285(5) 55(2) >7W In our experiment we decided to use a crossed-beam dipole trap with elliptical beam profiles, summarised in table 4.1. The two dipole beams propagate in the plane normal to gravity and their foci intersect at an angle of just under 90◦. For each beam the ratio of horizontal to vertical waist size is about five to one, quantum gases 63 leading to a radially symmetric, oblate trapping geometry. Trapping potential The dipole potential for an idealised two-level (e-g) atom state can be written as [175] Ve-g(r) = −3pic 2 2ω3a ( Γa ωa −ω f + Γa ωa +ω f ) I(r) , (4.1) where ωa is the resonance frequency of the atom, ω f the fre- quency of the light field, Γa the spontaneous scattering rate, and I(r) the laser intensity profile. In a red-detuned dipole trap we have ωa > ω f such that the dipole potential Ve-g is negative, leading to an attractive force towards its centre. The relevant transitions in the alkalis are the D-lines (section 3.1) which both contribute to the total trapping potential, given by Vdip(r) = 1 3 VD1e-g (r) + 2 3 VD2e-g (r) . (4.2) Here we have assumed linearly polarised light for which the trap- ping potential is independent of the given internal hyperfine state of the atom. The intensity profile of a focussed gaussian laser beam, as- sumed to propagate along the z-axis is I(r) = 2P piwh(z)wv(z) exp ( − 2h 2 w2h(z) − 2v 2 w2v(z) ) . (4.3) The paramaters h and v stand for the respective horizontal and vertical coordinate and P is the optical power in watts. The beam waists wi(z) are given by wi(z) = w0i √√√√1+( z piw20i/λ )2 , i ∈ {h, v} (4.4) where w0i are the (single) waists of the elliptical gaussian beam in h, v-direction and λ is the dipole trap wavelength. Now we are going to change into lab coodinates. In the crossed- beam configuration two dipole foci of equal waists are overlapped at just under 90◦ which can be approximated by the sum of two gaussian beams propagating along x and y, respectively. In the lab frame z is chosen to be the direction of gravity and thus the axial direction of the crossed-beam trap. Both beams are supposed to have equal waists such that the trap is radially symmetric in the xy-plane. The dominant trapping effect arises from the respective transverse beam profile of each beam, whereas the beam profile along the axes of propagation only add a weak confinement. In the centre of the trap the potential is harmonic which can be seen by Taylor-expanding Eq. 4.3. The characteristic trapping 64 quasicrystalline optical lattices for ultracold atoms frequencies in the radial (ωr) and axial (ωz) directions for an atom of mass m are then given by ωr = √ 4V0 mw20h (4.5) ωz = √ 8V0 mw20v , (4.6) where V0 is the maximum dipole potential, i.e. Vdip in the centre of the trap. Both dipole beams contribute to the vertical confine- ment which leads to an increase in the vertical trapping frequency by a factor of √ 2 compared to a single-beam trap.360 380 400 Dipole trap power/mW 0 20 40 R b BE C nu m be r/ 10 4 Figure 4.1: Minimal trapping power for 87Rb in the dipole trap. Below a certain threshold (385(5) mW in each beam, marked by the shaded area) the dipole trap cannot hold the rubidium condensate against gravity. 180 190 200 Dipole trap power/mW 0 5 10 39 K BE C nu m be r/ 10 4 Figure 4.2: Same as above, but for 39K. Here the threshold is 190(2) mW. Effects of gravity The expressions 4.5 and 4.6 are good approximations as long as gravity can be neglected. But as soon as the gravitational potential becomes comparable to Vdip it has to tbe included, leading to a total potential of the form Vtot(r) = Vdip(r) + mgz . (4.7) The maximum trap depth Vmax/kB expressed in µK is given by the energy difference between the local minimum and the closest local maximum of Eq. 4.7 which can be determined numerically. In particular, the maximum trap depth vanishes at non-zero dipole trap power, called ‘minimal trapping power’, causing the atoms to spill out of the trap centre. The minimal trapping power is quan- tified experimentally in Figs. 4.1 and 4.2 by measuring the con- densate atom number for various dipole powers. Whereas for low dipole trap powers the vertical trap depth is limiting, the high- power maximum trap depth is limited in the horizontal direction (Vhorz, dashed lines in Fig. 4.3A). Therefore the functional form of Vmax (Fig. 4.3A) exhibits a kink for each atomic species, indicat- ing the cross-over between the vertically and horizontally limited regimes. This cross-over occurs at roughly 0.6W for 39K and 1.2W for 87Rb. In our beam configuration the maximum trap depth for Table 4.2: Measured dipole trap frequencies for 87Rb. The frequencies were experimentally determined by slowly lifting a 87Rb-condensate by means of increasing the dipole power and then suddenly decreasing the power to the precise values listed in the table. This jerk in the position of the dipole trap leads to oscillations (along all axes, though primarily along z) which are fitted to a cosine in order to extract the frequency. power 0.38W 0.40W 0.45W 0.50W 0.55W 0.70W 1.00W ωr/(2pi)(x) 14.2(2) 15.9(2) 17.6(6) 19.2(5) 23.8(5) 28.4(5) ωr/(2pi)(y) 12.2(9) 13.7(9) 14.7(5) 16.6(9) 20.6(6) 23.9(9) ωz/(2pi) 90.8(3) 108.6(5) 120.9(8) 150.7(8) 183.1(7) Table 4.3: Measured dipole trap frequencies for 39K. The measurement protocol is identical to table 4.2 above. power 0.20W 0.25W 0.30W 0.50W 1.00W ωr/(2pi)(x) 16.5(5) 19.1(5) 21.9(5) 29.06(2) 41.1(5) ωr/(2pi)(y) 14.9(5) 17.5(5) 19.4(5) 25.49(2) 35.7(5) ωz/(2pi) 74(1) 115.3(5) 135.9(5) 190(1) 280(5) quantum gases 65 0.0 0.4 0.8 1.2 Dipole trap power/W 0 5 10 Tr ap d ep th / µ K Vmax (Rb) Vhorz Vmax (39K) Vhorz 0.0 0.4 0.8 1.2 Dipole trap power/W -20 -10 0 G ra vi ta ti on al sa g/ µ m Rb 39K diff A B Figure 4.3: (A) Maximum trap depth (Vmax/kB) versus dipole trap power (in each beam). For low powers the ver- tical trap depth is limiting (the atoms leave the trap from the centre along gravity) whereas for higher powers (above 1.2W) the horizontal depth is limiting (the atoms leave the trap to the side). (B) Calculated gravitational sag for 87Rb (red), 39K (blue), and the difference between the two (‘differ- ential sag’, black dashed). The pale blue (red) vertical bars are the mea- sured minimal trapping powers from Fig. 4.1 (4.2) which agree well with the numerical results. 87Rb drops below that of 39K at about 1.0W. Fig. 4.3A also shows the experimental values of minimal trapping power (vertical bars of the same colour as in Figs. 4.1 and 4.2) which agree well with the theoretical prediction. Sympathetic cooling works best if the coolant (87Rb) is predominantly evaporated, such that most of the potassium atoms remain in the trap. It has been argued by Campbell et al. [131] that sympathetic cooling in deep dipole traps does not work well. However, we find decent cooling results with this method despite the limiting 39K trap depth. This could be explained by a different evaporation ramp, higher available 87Rb atom numers, or by the different trap geometry Gravity does not only affect the trap depth but also the vertical position of the trap centre. This effect, also known as gravitational sag,3 is summarised in Fig. 4.3B. 3 Efficient sympathetic cooling relies on a good spatial overlap between the two species. 4.2 Trap frequencies The extent of a cloud of ultracold atoms in a dipole trap is typi- cally much smaller than the beam waists and the expansion of the potential around the trap centre can be restricted to include terms up to quadratic order. For all practical purposes the dipole poten- tial is therefore fully characterised by its vertical and horizontal trap frequencies ωz and ωr = ωx = ωy, respectively. These can be obtained numerically from fitting a parabola to the local minimum of Eq. 4.7 (at the vertical position z given by the gravitational sag, Fig. 4.3B). We find excellent agreement between our measurements (values in tables 4.2 and 4.3) and the numerical predictions, shown in Fig. 4.4. Note that the horizontal trap frequencies are measured on the x/y-axes of the z-axis imaging camera and do not neces- sarily correspond to ωx and ωy (the principal axes of the dipole trap). The mismatch between the two measured horizontal fre- quencies could be either explained by the not-perfectly-90◦ angle between the two dipole axes4 or by slightly different horizontal 4 which should show up as beating in the oscillation signal between the x- and y-axes but was not resolved in our experiments beam waists (the latter seems more likely). 0.0 0.2 0.4 0.6 0.8 1.0 Dipole trap power/W 0 10 20 30 40 H or izo nt al tra p fre qs ./ H z y (Rb)x y (39K) x 0.0 0.2 0.4 0.6 0.8 1.0 Dipole trap power/W 0 100 200 300 Ve rt ic al tr ap fr eq ./ H z Rb 39K A B Figure 4.4: Vertical (axial, A) and horizontal (radial, B) trap frequencies ω/(2pi) versus dipole trap power (per beam). The solid lines and shaded regions show the numerical prediction with a 2.5 µm uncertainty of the vertical beam waist. The dashed lines represent Eqs. 4.5 and 4.6 (neglecting gravity). 66 quasicrystalline optical lattices for ultracold atoms 4.3 Evaporation and sympathetic coolingTable 4.4: Evaporation parameters used to reach quantum degeneracy parameter 87Rb 39K initial power 7W initial V0/kB 43 µK 38 µK final power/mW ∼400 ∼200 final V0/kB 3 µK 1 µK exp. ramp dur. 3 s exp. time const. 0.5 s guide field 2.8(1) G 1st Fesh. (1.7 s) 0G 315.49(2) G 100 a0 141(1) a0 (inter) 2nd Fesh. (1.3 s) 0G 397.8(1) G 100 a0 280 a0 (intra) Quantum degeneracy of 87Rb or 39K is achieved by forced evapo- ration [177] in the dipole trap. In the case of 39K we start evapo- ration at similar temperatures but with far fewer atoms compared to 87Rb, necessitating a more efficient cooling method. We employ the widely used sympathetic cooling technique [131, 178, 179] which utilises the abundant 87Rb atoms as coolant for 39K. The experimental parameters used for evaporation are summarised in table 4.4. They are optimised to maximise condensate atom number, given a fixed ramp duration of 3 s. 4.4 Dipole trap laser setup This section is an interlude about the setup and characterisation of the dipole trap. All dipole beams are derived from a single-frequency solid state laser5 at 1064nm. The two (x and y) dipole trapping beams are5 Mephisto MOPA 55W, Coherent Inc. prepared on a separate laser table6 from which the laser power 6 ‘high-power’ or HP table, which also hosts the lattice laser is distributed via optical fibres to the experiment. The z-dipole axis has been prepared but not implemented yet; it is designed to provide additional radial confinement for the lattice potential once the atoms habe been loaded into the quasicyrstalline potential. Each beam passes through an AOM in order to shift its fre- quency (with respect to the other beams, avoiding cross-interferences) and to stabilise its intensity via PID feedback. After the initial split, the x and y beams both carry 17W of power from which max. 8W is measured after passing through the science cell. The largest losses are caused by the AOM (15%), the DC fibre polar- isation maintaining7 (∼25%), and the glass-cell/air interfaces7 The ‘cold’ fibre coupling efficiency (AOM on pulsed mode with 1% duty cycle) is about 95%. But in DC mode the coupling efficiency is reduced to ∼ 75%, probably due to polarisation drifts which manifest themselves as power drifts after the Glan-Laser cleaning prism. (measured ∼6% each, 45◦ incidence). More information on the preparatory stages of the dipole trap setup can be found in Max Melchner’s Master’s thesis [180]. Table 4.5: Horizontal (w0h) and vertical (w0v) dipole trap beam waists in µm. ‘Design’ refers to the values that were aimed at, ‘theory’ are the expected values using ABCD matrices, given the approximate distances on the bread- board and the measured divergence from the fibre coupler, ‘exp(erimental)’ are the measured values using a cam- era that is focussed to the position of the atoms, and ‘freq(uency)’ refers to inferred values by measuring trap frequencies (table 4.1). Dipole x Dipole y waist w0h w0v w0h w0v design 280 50 280 50 theory 284 53 279 52 exp 284 64 280 47 freq 285(5) 55(2) 285(5) 55(2) After exiting the optical fibre on the experiment laser table each beam passes through a Glan-Laser prim in order to clean the po- larisation. The transmission through two subsequent bending mir- rors is used as PID input signal (45◦ bend) and monitoring signal (22.5◦ bend). At 1064nm, Silicon-based photodiodes are not suit- able for high-bandwidth applications, such as the PID loop, due to the slow diffusion time constant t2 in this wavelength range [181]. Therefore we use high-bandwidth amplified InGaAs photodiodes for PID feedback. The feedback loop is otherwise identical to the lattice intensity stabilisation 5.4. The optical power of each dipole beam (for example, on the x-axis of Fig. 4.1) has been calibrated using a power meter after the beam passed through the science cell. Therefore the true optical power at the position of the atoms is approximately 10% higher than the specified value. Although, in principle, powers up to 8W have been measured after the sci- ence cell, the maximum values used in the sequence are 7W in order to leave some room for long-term drifts. quantum gases 67 After the Glan-Laser prism, the beam first passes through a cylindrical telescope and then a spherical telescope before it is focussed to the atomic cloud via an achromatic lens (125mm fo- cal length), as shown schematically in Fig. 5.17 in the following Chapter. The gaussian beam propagation has been modelled8 for 8 using ABCD matrices the actual physical setup on the experiment table, taking into ac- count the measured beam divergence from the fibre collimation. For elliptical beams, the horizontal and vertical waists propagate independently from each other, making it (in general) impossi- ble to focus both the vertical and the horizontal waist to the same point. The horizontal waist of the dipole beam at the position of the atoms is relatively large, corresponding to a small divergence, whereas the vertical waist is strongly divergent. Therefore, all remaining degrees of freedom9 were optimised to create an es- 9 distances between telescopes and lenses (given the physical constraints on the breadboard) and the available focal lengths of lenses sentially focussed beam at the atoms with specified design-waists. The experimentally measured waists agree reasonably well with the expected values from gaussian propagation (see table 4.5). Note that the theoretical values resulting from the trap frequency measurement assume a radially symmetric dipole trap. The ex- perimental agreement could be improved by modelling the trap frequencies in x and in y separately, accounting for the measured mismatch between the two (Fig. 4.4B). The final achromatic lens to focus each dipole beam to the atoms is shared with the lattice setup by employing the grazing- angle technique. The dipole beam is reflected off a D-shaped mirror and for a short distance (5 cm) it propagates parallel to the corresponding lattice beam. The lattice beam is incident on the final achromat on its optical axis; the dipole beam is incident roughly 5mm away from the optical axis. Since both beams im- pinge on the lens as parallel rays they are focussed to the same spot. Consequently, the dipole beams are overlapped at the po- sition of the atoms under different angles than the lattice, in our case just under 90◦. Due to the relatively large horizontal waists, the trap geometry is insensitive to the exact angle under which the dipole beams are overlapped. This technique avoids the use of expensive dichroic mirrors while maintaining the necessary flexi- bility in the alignment of both lattice and dipole trap. A schematic of the optical setup around the science cell can be found in sec- tion 5.4 below. Alignment All degrees of freedom of alignment, except the relative horizontal angle between the dipole beams (as mentioned above), are impor- tant for the resulting dipole trapping performance. Firstly, the focus of the (vertical) dipole waists should coincide with the atomic cloud, in order to ensure maximum trap depth. This is achieved by first focussing a camera on each dipole axis to the position of the atoms (using in-situ absorption imaging). Subsequently, the last achromat is moved longitudinally (on a 68 quasicrystalline optical lattices for ultracold atoms z-translation stage with µm screws) to focus the dipole beam. Secondly, each dipole beam should propagate perpendicular to gravity, thereby minimising gravitational sag. We achieve this by beam walking using pinholes centred on the achromatic lenses, thereby trusting that our optical table is also aligned perpendicu- lar to gravity. Using the time-of-flight technique on the D-axis1010 one of the lattice axes which does not have a dipole beam we measured the angle of the dipole trap beams with gravity to be 90.0(5)◦. Thirdly and most importantly, the dipole beams have to inter- sect at the same point in space (‘pointing’). This is necessary in order to benefit from the crossed-beam configuration in which the vertical trap depth of both beams add up. If the beams do not intersect vertically, the trap depth is greatly reduced (in the worst case by a factor of two). In addition, the vertical intersection ensures the approximate radial symmetry of the trap (up to the discrepancies in beam waist). To this end, the last kinematic mirror before the final achromat is piezo-driven. Rough pre-alignment is achieved by loading cold atoms into a single-beam dipole trap and maximising the in-situ optical density.11 Fine alignment is achieved by producing a con-11 An optical density of 1.8-2.0 for 87Rb is a good benchmark here. densate and lowering the final evaporation depth until almost no atoms are held against gravity. Then the vertical alignment can be changed in small steps by the piezo-driven mirror to max- imise atom number. This alignment technique can discriminate the optimum position of the mirror to within 10 piezo steps; this corresponds to a mirror-displacement of roughly 170nm and a vertical displacement (at the atoms) of 1.6 µm. The centre of the dipole trap is aligned 61 µm below centre of the magnetic trap for optimum loading efficiency. 4.5 Dipole loading, guide field, and RF state transfer Table 4.6: Magnetic field calibration of the guide field. For each current value we take an RF/MW-resonance scan using a Landau-Zener sweep of 20ms duration. In the case of the RF transfer the transition |F = 1, mF = −1〉 → |F = 1, mF = 0, 1〉 in the ground level of 87Rb is used, in the case of MW we use |F = 2, mF = 2〉 → |F = 1, mF = 1〉. These frequencies are then converted to magnetic field values using the magnetic field hamiltonian (section 3.2). Current/A RF/MW resonance B-field/G 6.0 3.042(8) MHz 4.35(1) 6.0 6843.78(2) MHz 4.33(1) 4.5 1.95(5) MHz 2.8(1) Before the magnetic trap is ramped down and during loading of the dipole trap we apply a homogeneous magnetic field (‘guide field’) using offset coils in the z-direction (Fig. 4.5). This serves three purposes: Firstly, it ensures a definite quantisation axis after the magnetic trap has been switched off and before the (strong) Feshbach field comes on (in the case of 39K).12 Secondly, it com- 12 Though even without a quantisation axis during the entire dipole evapo- ration ramp we achieve condensation with 87Rb. pensates for long-term drifts in the position of the magnetic trap (section 3.10), ensuring good spatial overlap between the magnetic trap and the dipole trap. Our offset coils shift the vertical trap po- sition a little more than 25 µm/A when applied to a high-gradient magnetic trap of 300G/cm (strong axis). Thirdly, we perform a radio-frequency (RF) state transfer, using the guide field to lift the degeneracy of the internal states of 87Rb and 39K. The choice of internal states during evaporative and sympathetic cooling is of crucial importance to the success of these methods [182]. In- elastic collisions between atoms in different internal states can lead to heating and losses. Consequently it is desirable to trans- fer the atoms to their respective absolute ground state, thereby quantum gases 69 precluding inelastic processes. In our case this is achieved simul- taneously for 87Rb and 39K using a Landau-Zener transition from |F = 1, mF = −1〉 to |F = 1, mF = 1〉. The RF transfer is accom- plished by a 20ms-long linear frequency sweep from 1.90MHz to 2.00MHz at a homogeneous magnetic field of 2.8(1) G. -0.2 0.0 0.2 0.4 Time/s 0 3 6 Di po le X ph ot od io de /V dipole trap exp. coil guide field RF transfer -0.4 0.0 0.4 0.8 Cu rre nt cla m p/ V (1 00 A/ V) Figure 4.5: Dipole trap loading pro- cedure (for 39K). Before the magnetic quadrupole field (dark blue) is ramped down we apply a guide field (light blue), as discussed in the text. The polarity of the upper experiment coil (dark blue) is then changed to provide the inter-species Feshbach field during evaporation. In case of 87Rb the dipole loading procedure is identical except that there is no Feshbach field. Figure 4.5 shows a scope trace of our dipole loading procedure, including a photodiode signal of the dipole trap, current clamp measurements of the experiment coil and the offset-z coil, and the timing of the RF transfer. The calibration of our guide field is summarised in table 4.6. RF antenna The RF antenna that is used for the state transfer operates in the near-field regime (λRF = 1.5 m) and consequently the radiation consists mostly of an oscillating B-field. The antenna geometry and its number of windings is therefore largely insignificant (see Fig. 4.6). In addition, the antenna can be approximated by an isotropic emitter, making the coupling efficiency of the radiation independent of the precise alignment (angle) of the antenna with respect to the atoms, as long as it is placed as close as possible (see Fig. 4.6). 1 2 3 4 Distance from glass cell/cm 0 1 2 3 4 5 87 R b at om nu m be r/ 10 8 3 5 7 Figure 4.6: RF antenna coupling effi- ciency. The data represents the surviv- ing 87Rb-atoms after a 1s ramp from 30MHz to 5MHz in a steep-gradient (300G/cm) magnetic trap, i.e. the fewer atoms the larger the coupling efficiency (or Rabi-frequency). Our antenna has circular geometry, consisting of four windings (radius 15mm) and it is wound on a 3D-printed plastic mount that also houses the Stern-Gerlach coil, about 1 cm above the glass cell. The antenna is not impedance-matched to any particular frequency because it is designed to operate in a large range of frequencies (1 - 400MHz).13 13 It was also designed to perform RF evaporation on the mF-states in 87Rb. 4.6 Feshbach fields Feshbach resonances [93, 183] are key to both the production of ultracold atomic gases and the application thereof. The physi- cal origin of this phenomenon lies in the quantum-mechanical scattering process of two particles at low energies [184] which can be characterised by a single parameter called the scattering length.14 This parameter, usually quoted in units of the Bohr- 14 At low energies, s-wave scattering dominates due to the centrifugal barrier [183] which surpresses higher- order scattering. The dominance of s- wave scattering defines the ‘ultracold’ regime. radius a0 = 5.29× 10−11 m, is a measure for the phase shift be- tween the incoming and outgoing wave function of a scattered particle. Its value depends on the respective energy landscape of the inter-atomic molecular potential which, for a given atomic species, is hard to predict. This combined quantum state of the two scattering participants (in the limit of large separation) is called a scattering channel. The energy of a scattering channel de- pends on the composition of internal states and it can be shifted according to the Zeeman effect by applying a magnetic field. If the constituents of two scattering channels have a different composi- tion of internal states then their respective molecular potentials can be tuned relative to each other. In particular, the energy of two free particles (an open scattering channel) can be brought 70 quasicrystalline optical lattices for ultracold atoms into resonance with a bound state of a different channel (a closed channel), leading to a maximum scattering phase shift of pi and a diverging scattering length. This effect is called Feshbach reso- nance. 310 314 318 320 B-field/G 0 1 2 39 K co nd en sa te nu m ./ 10 5 data a(B) -10 0 0 10 0 20 0 Sc at te ri ng le ng th / a 0 Figure 4.7: Experimental optimisation of the inter-species Feshbach field for 39K - 87Rb. The optimum field value (315.49G) produces the largest 39K-condensate. The data points are averages over three identical realisa- tions and the error bars are smaller than the markers. The magnetic field (x-axis) is calibrated using an RF trans- fer in rubidium (table 3.2). The second y-axis shows the scattering length a (red line). The scattering length (a) close to a Feshbach resonance is para metrised by the following function of the magnetic field a(B) = abg ( 1− ∆B B− Bres ) (4.8) where the relevant values for abg, Bres, and ∆B are listed in ta- ble 4.7. When producing condensates of 39K we make use of both Fes- hbach resonances during evaporation, firstly to sympathetically cool 39K with 87Rb and secondly to reach quantum degeneracy with 39K. The field is switched from inter-species to the intra- species value at the point when all 87Rb has fallen out of the trap due to gravity, as detailed in table 4.4. These parameters are op- timised to produce the largest condensates, as shown in Figs. 4.7 and 4.8. 380 400 420 B-field/G 0 1 2 3 39 K co nd en sa te nu m ./ 10 5 data a(B) -20 0 0 20 0 40 0 Sc at te ri ng le ng th / a 0 Figure 4.8: Experimental optimisation of the intra-species Feshbach field for 39K (optimum value 397.8G) using the 39K-condensate number as figure of merit. The second y-axis shows the scattering length a (red line). 4.7 Quantum degeneracy and Bose-Einstein condensation The ultracold temperature regime is defined by the prevalence of s-wave scattering for bosons [183] or p-wave scattering for fermions, which are two-body collisional properties of the relevant atomic species. Such a system may still be thermodynamically described by classical Maxwell-Boltzmann statistics (although the scattering has to be already described quantum-mechanically15). 15 The bosonic or fermionic particle symmetry influences the collective properties of the gas. This can, for instance, show up in the dynamics of the system, such as an anisotropic expansion of a thermal Bose gas after sudden release from the dipole trap [185]. However, if the cloud is cooled down further, the density cru- cially affects the thermodynamics of the system. The phase-space density n λ3dB where λdB = h√ 2pimkBT (4.9) replaces temperature as the relevant thermodynamic parameter because it relates the average inter-particle distance n−1/3 to the thermal de Broglie wavelength λdB. When the phase-space density approaches unity [186] the classical Maxwell-Boltzmann descrip- tion fails and must be replaced by the relevant quantum statistical expression. In this regime the symmetry properties of the individ- ual quantum particles dominate all thermodynamic quantities. In particular, fermionic symmetry – the full many-body wavefunc- tion is anti-symmetric under particle exchange – leads to Pauli exclusion, i.e. the fermionic occupation of a single quantum state Table 4.7: Feshbach resonance param- eters. The values are taken from [93] except the resonance position for 39K which has recently been measured to a much greater precision [130]. species scattering channel Bres/G ∆B/G abg/a0 87Rb/39K 1×87Rb & 1×39K in |F = 1, mF = 1〉 317.9 7.6 34 39K 2×39K in |F = 1, mF = 1〉 402.70(3) -52 -29 quantum gases 71 kx k y T > Tc T << TcT < Tc Figure 4.9: Bose-Einstein condensation in a gas of 39K-atoms. The vertical coordinate in the surface plot cor- responds to the optical density. A thermal gas (T > Tc) is well-described by the Maxwell-Boltzmann distri- bution, corresponding to a gaussian profile in time-of-flight. Below the crit- ical point, a sharp feature at k = (0, 0) starts to appear which grows as the gas is cooled down further. For tem- peratures much lower than Tc, there is no discernible thermal fraction. is bounded from above by 1. Conversely, bosonic symmetry – the many-body wavefunction is symmetric under particle exchange – leads to particle-bunching which allows bosonic systems to lower their energy by occupying a single quantum state. Statistically, these effects are captured by the average occupation 〈ni〉 = 1e(ei−µ)/(kBT) ∓ 1 (4.10) of a given state |i〉 with energy ei. The minus-sign corresponds to bosonic, the plus sign to fermionic quantum statistics. The chem- ical potential µ has been introduced to ensure particle number conservation via N =∑ i 〈ni〉 (4.11) in the grand-canonical ensemble. In this work all lattice exper- iments were carried out with bosons and the remainder of this section will cover the relevant properties of ultracold bosons in a harmonic trap.16 16 However, all necessary hardware has been set up to produce a Fermi-gas of 40K. Bose-Einstein condensation As the temperature of the gas is reduced at constant density, the chemical potential approaches the value of the lowest energy level e0 = 0 in a uniform system3 2 h¯ω in an isotropic 3D harmonic trap (4.12) from below. The occupation of the excited states then tends to a finite value, given by the number equation (Eq. 4.11).17 Therefore, 17 For all excited states (ei > e0) the number equation can be solved analytically by replacing the sum by an integral. all remaining particles (in the thermodynamic limit N → ∞) must occupy the ground state e0. This effect is known as Bose-Einstein condensation. Alternatively, condensation can be achieved by increasing the particle number N at fixed temperature. For an ideal Bose gas in a harmonic trap the critical particle number Nc is given by [186] Nc = 1.202 ( kBT h¯ω¯ )3 . (4.13) 72 quasicrystalline optical lattices for ultracold atoms Here ω¯ denotes the geometric mean of the trapping frequencies. For N ≥ Nc (at fixed T) the excited motional states are ‘satu- rated’ such that adding particles to the system solely increases the ground state occupation.1818 In a realistic scenario, however, at most half the added particles end up in the condensate and the other half is thermal [187]. Experimentally, the condensate can be detected in time-of- flight (see sections 3.12 and 6.2). It appears as a sharply peaked feature above the thermal background [97], as shown in Fig. 4.9. The profile of the condensate peak, a paraboloid, is a consequence of interatomic interactions and trapping geometry. Theoretically, these can be taken into account by the Gross-Pitaevskii equation in the Thomas-Fermi approximation [186].19 The resulting real-space19 by neglecting the kinetic energy term − h¯22m∇2 profile in an anisotropic harmonic trap is then given by n(r) = n0 ( 1− x 2 R2x − y 2 R2y − z 2 R2z ) , (4.14) where Ri are the Thomas-Fermi radii in the three spatial directions and n0 is the central density. Generally, the expansion of a trapped condensate in time-of-flight is anisotropic [97, 186] in contrast to the thermal gas which expands isotropically. Equation 4.14 gives rise to a parabolic density profile of the atomic cloud in time-of- flight which is used extensively in this work as a fit function to measure atom numbers. Thereby the equations quoted in ref. [160] are used. Atom number calibration The observation of a phase transition from a thermal to a Bose- condensed state provides a measurement of atom number which is independent of the specifics of the light-atom interactions of the atomic species in question. For 39K in particular, a faithful atom number measurement via summing the optical density (sec- tion 3.12) is impossible due to the non-existence of a closed imag- ing transition, even at high magnetic fields where no repumping light is required (section 3.12). 5 Eightfold optical lattice This chapter provides the theoretical and experimental ground- work for the experiments carried out in this thesis (Chapter 6). First, the basic equations for non-interacting particles in a 1D op- tical lattice are derived (section 5.1). Second, the single-particle dynamics following a sudden lattice quench is reviewed with a focus on using this method for lattice depth calibration. Third and last, the laser setup for the eightfold rotationally symmetric optical lattice is described (section 5.4). 5.1 1D optical lattice The matter-wave experiments described in Chapter 6 can be treated in the usual Bloch-wave formalism. In this section the matrix hamiltonian for a single atom in a one-dimensional optical lattice is derived which forms the basis for all theoretical predic- tions in Chapter 6. Table 5.1: Relevant recoil energies for 87Rb and 39K using the optical wavelength λlat = 726 nm. species Erec/h 87Rb 4.35kHz 39K 9.71kHz The dipole potential (Eq. 4.2) of a retro-reflected laser beam can be written as V(x) = V0 cos2(klatx) = V0 2 ( 1+ ei2klatx + e−i2klatx 2 ) , (5.1) where any inhomogeneities and (anti-)confining potentials have been neglected.1 Further, it is assumed that the incoming and 1 This assumption is valid for our experimental timescales since the longest lattice pulses are t = 50 µs  1 ωtrap . the retro-reflected beams are perfectly counter-propagating. The lattice wavelength λlat = 2pi/klat = 726 nm is far-blue-detuned with respect to the D-lines in 39K, ensuring that single-photon processes are completely suppressed. The lattice depth2 (V0) can 2 In principle, the value of V0 could be calculated directly from Eq. 4.2. In practice, however, these estimations suffer from systematic uncertainties in the laser powers and beam waists. Therefore the lattice dynamics that are derived in the following will be used to calibrated V0 since they provide a much more precise measure of lattice depth. be expressed conveniently in units of photon recoil energy, i.e. Erec = h2 2mλ2lat . (5.2) Here m denotes the atomic mass and h Planck’s constant; the relevant recoil energies for 87Rb and 39K are listed in table 5.1. Due to the periodicity of the potential3 (Eq. 5.1) we can invoke 3 the lattice spacing a is given by a = λlat 2 = 2pi 2klat (5.3) Bloch’s theorem [188, 189] and introduce the quasimomentum q. The quantity q mod 2klat (5.4) 74 quasicrystalline optical lattices for ultracold atoms -1 0 1 0 5 10 15 -1 0 1 -1 0 1 -1 0 1 Quasimomentum q/klat En er gy /E re c V0 = 0Erec 2 Erec 20 Erec6 Erec Figure 5.1: Band structure for non- interacting particles in a homogeneous, one-dimensional optical lattice. The solutions to Eq. 5.9 for n ∈ {0, 1, 2, 3} (deep blue to light blue) and V0/Erec ∈ {0, 2, 6, 20} (left to right) are shown. For small values of V0 the optical lattice causes only small perturbations on top of the free-particle dispersion p2/(2m). Conversely, in the limit of deep lattices (V0 & 5Erec) the n = 0 Bloch waves become more peaked around each lattice site, the bandwith becomes small, and the dispersion relation can be approximated by a cosine (tight-binding limit). is conserved in a periodic potential. In the Bloch formalism the wave function takes the form ψnq (x)︸ ︷︷ ︸ Bloch wave = unq (x)︸ ︷︷ ︸ Bloch function × eiqx︸︷︷︸ plane wave , (5.5) where the index n has been introduced to account for all possible solutions (bands) of the energy eigenvalue equation Hψ = Eψ (time-independent Schrödinger equation). The ansatz Eq. 5.5 can be inserted to the eigenvalue equation to yield[ − h¯ 2 2m ( ∂2 ∂x2 + q2 ) + V0 2 ( 1+ ei2klatx + e−i2klatx 2 )] unq (x) = E n q u n q (x) . (5.6) Bloch’s theorem states that the functions unq (x) exhibit the same periodicity as V(x), i.e. unq (x) = unq (x + a), and consequently they can be decomposed in a discrete Fourier series: unq (x) = +∞ ∑ l=−∞ cnl (q) e i2lklatx . (5.7) Inserting this expression into Eq. 5.6 results in +∞ ∑ l=−∞ [( − h¯ 2 2m ( (2lklat)2 + q2 ) + V0 2 ) ei2lklatx + V0 4 ( ei2(l+1)klatx + ei2(l−1)klatx )] cnl (q) = E n(q) +∞ ∑ l′=−∞ ei2l ′klatx cnl′(q) , (5.8) which can be easily converted to the following matrix equation for H +lmax ∑ l=−lmax Hl,l′cnl (q) = En(q)cnl′(q) . (5.9) The simplicity of the potential V(x), which contains only two non- trivial Fourier components, namely ±2klat, ensures the straightfor- ward derivation of Eq. 5.8 (no double summation on the left-hand eightfold optical lattice 75 side). Consequently, the matrix Hl,m is sparse, the only non-zero elements being located on the diagonal and the first off-diagonal, i.e. Hl,l′ =  h¯2 2m (q + 2lklat) 2 + V02 for l = l ′ V0/4 for |l − l′| = 1 0 otherwise. (5.10) The truncation of the infinite basis of the coefficients cnl (q) at cer- tain values ±lmax can be justified a posteriori since only a fraction of bands get populated.4. 4 This is not the case for the eightfold optical lattice and consequently the calculation of a ground-state energy in this formulation is usually impossible in that case (section 6.1). The Bloch waves are delocalised over the whole lattice. How- ever, it is possible to recover localised wave functions for each lattice site. These localised functions are called Wannier functions wn(x− xi) = 1√ M ∑ q e−iqxiψnq (x) (5.11) for each lattice site i (M is some normalisation). Bloch waves and Wannier functions are related via a discrete Fourier trans- form. Therefore, a localised particle (i.e. unity occupation of a single Wannier function) corresponds to a flat distribution of Bloch waves (each Bloch wave is equally populated). The Wan- nier functions form an alternative orthonormal basis to solve the Schrödinger equation. The derivation above can be generalised to more sophisticated lattice geometries, as shown in section 6.1. Band structure, tunnelling, and group velocity In the case of periodic lattice geometries, for instance 1D or 2D simple-cubic, or 2D honeycomb geometry, the hamiltonian matrix in Eq. 5.10 can be diagonalised and the corresponding eigenvalues and eigenstates are exact. The eigenenergies are the well-known Bloch bands, depicted in Fig. 5.1 for different values of V0/Erec ∈ {0, 2, 6, 20}. Two regimes emerge here: On the one hand, shallow lattices can be treated as a small perturbation on free-particle behaviour. On the other hand, deep lattices (V0 & 5Erec) lead to peaked Bloch functions around the lattice sites, corresponding to the tight-binding limit. -1 0 1 Quasimomentum q/pia -2 0 2 D is pe rs io n E (q )/ J E(q) vgr -2 0 2 G ro up ve lo ci ty v g r/ ( Ja h¯ ) Figure 5.2: Tight-binding dispersion relation and its derivative, the group velocity. In the tight-binding limit the lowest band of the optical lattice corresponds to a discrete lattice with a fixed tunnelling energy between the sites. A measure for the tunnelling matrix element J between neighbouring lattice sites is the bandwidth of the lowest band, i.e. J = max(E0q)−min(E0q) 4 . (5.12) The dispersion relation in this case (deep blue line in rightmost plot in Fig. 5.1) is the characteristic cosine of the tight-binding model [188]: E(q) = −2J cos(qa) , (5.13) 76 quasicrystalline optical lattices for ultracold atoms where a is the lattice spacing (Eq. 5.3). This dispersion relation is plotted in Fig. 5.2 together with the corresponding group velocity vgr = 1 h¯ ∂E(q) ∂q = 2Ja h¯ sin(qa) . (5.14) Higher dimensions and quantum walks In the case of D-dimensional separable lattices Eq. 5.13 generalises to E(q) = −2J D ∑ i=1 cos(aqi) , (5.15) where q is now a D-component vector. Correspondingly, the indi- vidual components of the group velocity are given by vi(q) = 1 h¯ ∂E(q) ∂qi = 2Ja h¯ sin(aqi) . (5.16) For a given Wannier state that is initially localised to one lattice site all Bloch waves are equally populated, leading to an average root-mean-square group velocity √ v2 = 2Ja h¯ √√√√ D∑ i=1 sin2(aqi) = 2Ja h¯ √ D 2 . (5.17) The √ D scaling will be recovered exactly in the experiment (Chapter 6) for D ∈ {1, 2, 3, 4}. 5.2 Time-evolution and lattice depth calibration Let us now consider the time-evolution of an optical lattice system according to the time-dependent Schrödinger equation ih¯ dψ dt = Hψ . (5.18) As a first example the ballistic expansion of a particle initially localised to one single lattice site will be shown. Figure 5.3: (A) Quantum walk on a 1D homogeneous tight-binding lattice. Initially the particle is localised to one lattice site (deep blue, top). With time, it spreads ballistically in both directions, giving rise to a ‘light-cone’ with a constant group velocity given by Eq. 5.17 for D = 1. Times are given in units of tunnelling times h¯/J. (B) Classical random walk (schematic). Particles initially localised to one lattice site are allowed to hop to neighbouring sites with a certain probability. Contrary to the quantum case, the expansion is diffusive, i.e. gaussian. 0 25 50 Lattice site Pr ob ab ili ty (a .u .) 0 25 50 Lattice site A B t = 0 t = 1 t = 2 t = 3 t = 4 t = 5 t = 6 eightfold optical lattice 77 Homogeneous tight-binding lattice In the limit of deep lattices (V0/Erec  1) it is convenient to work in the Wannier-basis and the system describes only the direct cou- pling of nearest neighbours (homogeneous tight-binding model). The hamiltonian reads Hi,j = J for |i− j′| = 10 otherwise. (5.19) Note that we have changed basis compared to Eq. 5.10 such that the lattice sites i denote the Wannier functions w0(x − xi). The initial state ψ(t = 0) for time-evolution shall simply be unity probability at one lattice site. The time-evolution ψ(t) according to Equation 5.18 can readily be solved, for example using the scipy.integrate package, leading to a linear expansion of the wavepacket over the entire lattice. This light-cone-like behaviour (depicted in Fig. 5.3A) has no classical analogon. Therefore this effect has been coined ‘continuous quantum walk’ [22, 190, 191] in order to differentiate it from a classical random walk (Fig. 5.3B). Time-evolution in momentum space for square lattice pulses Table 5.2: Scheme to group the basis states for a 1D optical lattice. order states 0th {|q = 0〉} 1st {|+2klat〉 , |−2klat〉} 2nd {|+4klat〉 , |−4klat〉} nth {|+2nklat〉 , |−2nklat〉} In our experiment we are primarily interested in the time-evolution of a Bose-Einstein condensate initially at rest, i.e. localised in momentum space. The time-evolution after a sudden switch-on (quench) of the lattice potential can be calculated via integrat- ing the time-dependent Schrödinger equation (Eq. 5.18), as be- fore. In this case it will be beneficial to work in the original basis {|q = 0〉 , |±2klat〉 , |±4klat〉 , . . . } and not in the Wannier basis. Al- though the following considerations apply to reciprocal space, it is instructive to keep the Wannier-picture at the back of one’s mind. It provides a useful analogy in which momentum states sepa- rated by 2klat correspond to ‘neighbouring lattice sites’, coupled by a ‘tunnelling matrix element’ V0/4. This will become clearer in Chapter 6. The kinetic energy part on the diagonal of Hl,l′ (Eq. 5.10) now crucially affects the dynamics of this system. It leads to an effec- tive restoring force that causes the condensate to return to its orig- inal state (|q = 0〉). In preparation for the quasicrystalline lattice the basis shall be labelled according to the number of ±2klat steps in momentum space from |q = 0〉, each step being one ‘diffrac- tion order’, as shown in table 5.2. This is convenient because the hamiltonian (Eq. 5.10) is symmetric in q ↔ −q such that all result- ing states in the time-evolution obey the same symmetry. In a one-dimensional periodic lattice5 the numerical solution 5 any periodic lattice, e.g. 2D and 3D simple-cubic, or 2D triangularto Eq. 5.18 is exact, because only a limited number of orders get populated. This result is shown in Fig. 5.4 (lines), together with experimental data (points). The experimental points are the rela- tive atomic population of each order of q-states.6 6 The measurement and data-analysis procedures are explained in detail in section 6.5 and 6.2, respectively. Agreement between theory and experiment is very good, except for short pulse durations for which a slight ‘lag’ of the experi- 78 quasicrystalline optical lattices for ultracold atoms Figure 5.4: Time-evolution of a Bose- Einstein condensate after a sudden lattice quench (V0 = 15.7Erec) of the D lattice axis (see below for naming con- vention of axes). The plot shows the time evolution grouped by diffraction order; the 0th order corresponds to the condensate, the 1st order to the satellite peaks at ±2klat. For information on image analysis the reader is referred to section 6.5. 0 10 20 30 40 50 Lattice pulse duration/µs 0.0 0.5 1.0 0th Po pu lat io n 1st 2nd 3rd mental data can be observed. This is due to a finite ramping-up time of the lattice pulse on the order of 1µs. This effect can be ac- counted for by making the hamiltonian (Eq. 5.10) time-dependent (see Chapter 6.5) but for now it will be neglected. Lattice depth calibration 15.3 15.5 15.7 15.9 Lattice depth V0/Erec 0.0 0.2 0.4 R es id ua lχ 2 0th 1st 2nd 3rd Figure 5.5: Lattice depth calibration using time-evolution following a sudden quench to a given lattice depth V0. The y-axis shows the squared and summed difference between the experimental values (points in Fig. 5.4) and the numerical calculation using the depths shown on the x-axis. The best fit in this case is obtained with a lattice depth of V0 = 15.7Erec (lines in Fig. 5.4). We use the well-understood process described above to calibrate precisely our lattice depth by comparing experiment and theory. In fact, the lines in Fig. 5.4 were calculated using the calibration method summarised in the following. For a given (unknown) ex- perimental value of V0 a data trace of the type shown in Fig. 5.4 is taken by varying the duration of the (square) lattice pulse. Sub- sequently, the exact time-evolution is calculated according to the Schrödinger equation for a number of lattice depths and the best fit to the experiment is found via the least-squares method (shown in Fig. 5.5). The lattice depths of all four lattice axes (see below) are cali- brated in this way. After discussing the properties and dynamics of 1D optical lat- tice we turn to the description of the full, quasicrystalline lattice. 5.3 Eightfold optical lattice potential The 2D quasicrystalline potential is created using a planar ar- rangement of four mutually incoherent 1D optical lattices, each formed by retro-reflection of a single-frequency laser beam, as shown schematically in Fig. 5.6. The angle between two neigh- bouring lattice axes is 45(1)◦, thereby imposing a global eightfold eightfold optical lattice 79 Figure 5.6: Schematic of the eightfold optical lattice. The four lattice axes are mutually incoherent such that the resulting potential is the sum of four 1D lattices. Each 1D lattice is formed by retro-reflection, as shown in the schematic. However, most of the optics involved in the retro-reflection are not shown. In the apparatus, a cat’s eye configuration is used in which there are two lenses in between the atoms and the retro-mirror. The distance from the atoms to the retro-mirror should be exactly twice the sum of the individual focal lengths of the lenses in order to ensure approximately equal beam waists at the position of the atoms. -4 -2 0 2 4 Position x/λlat -2 0 2 Po si ti on y/ λ la t 0 4 La tt ic e po te nt ia lV (r )/ V 0 Figure 5.7: Lattice potential resulting from four non-interfering axes (blue- detuned). The phases ϕi have been chosen such that the symmetry point is located at r = (0, 0) and coincides with a potential maximum (maximally constructive interference). There are roughly as many maxima as there are minima in this potential. rotational symmetry in close analogy to the octagonal tiling (sec- tion 1.3). In the following section some basic properties of this eight- fold lattice in real space are described. For the momentum-space picture, the reader is referred to Chapter 6. First, we consider the situation of four mutually incoherent lat- tice beams, which also has been used for all experimental results in Chapter 6. Afterwards, alternative schemes that make use of inter-axis interference are considered. -4 -2 0 2 4 Position x/λlat 0 2 4 La tt ic e po te nt ia lV (r )/ V 0 Figure 5.8: Two cuts through the lattice potential shown in Fig. 5.7. The curves are color-coded by the horizontal lines in Fig. 5.7. The distribution of potential maxima and minima is roughly symmetrical around the average potential value (2 V0). Incoherent superposition The sum of four mutually incoherent7 lattice axes of equal depth 7 The terms ‘non-interfering’ and ‘incoherent’ only apply to the inter- axis interference in this context. Of course, the 1D lattices themselves are a result of intra-axis interference. V0 can be written as8 8 neglecting any anti-confinement V(r) = V0 4 ∑ i=1 cos2 (ki · r + ϕi) , (5.20) 80 quasicrystalline optical lattices for ultracold atoms where the individual lattice vectors ki are given by k1 klat = ( 1 0 ) , k2 klat = ( 0 1 ) , k3 klat = 1√ 2 ( 1 1 ) , k4 klat = 1√ 2 ( −1 1 ) (5.21) and the individual phases ϕ1, ϕ2, ϕ3, ϕ4 are determined by the longitudinal position of the retro-reflecting mirror.9 The potential9 The individual lattice axes are created by retro-reflection, resulting in a standing wave. Therefore the phase of the incoming light is exactly cancelled. Eq. 5.20 is shown in Fig. 5.7 for the case that all ϕi = 0. Since our lattice is blue-detuned with respect to the D-lines in both 87Rb and 39K, an intensity maximum results in a potential maximum. In this particular choice of phases (all ϕi = 0), the point r = (0, 0) corresponds to the point of maximally constructive interference (V(0, 0) = 4V0).10 In addition, the whole potential is eightfold10 The irrationality of cos 45◦ = 1/ √ 2 ensures that this value of V(r) occurs exactly once and consequently it is the global maximum of the V(r). However, there will always be values of V(r) that are arbitrarily close to the global maximum. rotationally symmetric around this point.11 Two cuts through the 11 and only around this point potential are shown in Fig. 5.8 which correspond to the horizontal lines of the same colour in Fig. 5.7. If we were to interpret each minimum of V(r) as a ‘lattice site’, it is immediately clear from Fig. 5.8 that the vast majority of lattice sites (or ‘wells’) is located at different energies. Moreover, the ground level h¯ω of each well in harmonic approximation12 varies 12 given by the curvature of the well in the x and y directions strongly. Since deeper wells tend to have stronger curvatures this effect leads to an even larger spread in energies. For these reasons the potential V(r) cannot exhibit a band structure as in the equivalent single-axis case (section 5.1). In addition, tunnelling is inhibited by the energy offset between neighbouring sites. Relation to octagonal tiling It is interesting to compare the resulting potential V(r) to the oc- tagonal tiling (Fig. 1.19). In the easiest case, we choose all phases ϕi to be equal to pi/2 in which case the central point (0, 0) shows maximally destructive interference (V(0, 0) = 0). This potential bears a close resemblance to the ideal octagonal tiling (Fig. 1.19) with a bond length of approximately 0.6λlat. Alternatively, the actual minima of V(r) can be directly connected by lines to repro- duce an approximate version of the octgonal tiling. This process is shown in Fig. 5.9 for ϕi = pi/2. The exact knowledge of all phases ϕi is unrealistic in a real ex- perimental setting (see below). However, the octagonal tiling-effect is relatively robust against a random choice of (constant) phases, as shown in Fig. 5.10 in which all phases ϕi have been chosen at random in between 0 and pi. The random choice of phases leads to the appearance of ‘defects’ or closely-spaced ‘double wells’ which can be seen in Fig. 5.10 upon close inspection. These effects can be interpreted as phasonic excitations, since the manner in which they appear is identical to that of the ideal octagonal tiling (Fig. 1.20). Despite the random choice of phases, the agreement between the potential and the ideal octagonal tiling is very good. Consider, for example, the length of the ‘bonds’ (black lines con- necting the lattice points) and ‘gaps’ (the shorter diagonal of the rhombus tile), the occurrence of which are plotted in Fig. 5.11 for eightfold optical lattice 81 -4 -2 0 2 4 Position x/λlat -2 0 2 Po si ti on y/ λ la t 0 4 La tt ic e po te nt ia lV (r )/ V 0 Figure 5.9: Lattice potential resulting from four non-interfering axes, over- laid with the corresponding minima (black points) and ‘bonds’ (lines). Upon close inspection, the lattice points and bonds can be seen not to exactly represent the octagonal tiling. The phases ϕi have been chosen such that the symmetry point is located at r = (0, 0) and coincides with a poten- tial minimum (maximally destructive interference). The bonds shown here are not necessarily the physical bonds (shortest distance or largest tunnelling matrix element) between lattice sites but they nicely illustrate the similarity to the octagonal tiling. the random-phases potential. The ratio between the mean values of the two is remarkably close to the ideal value of the octagonal tiling: 0.478(6) 0.60(2) = 0.80(4) ' 2 sin (pi 8 ) = 0.765 . . . . (5.22) Sampling from a larger window than Fig. 5.10 would result in an even better agreement. -4 -2 0 2 4 Position x/λlat -2 0 2 Po si ti on y/ λ la t 0 4 La tt ic e po te nt ia lV (r )/ V 0 Figure 5.10: Lattice potential resulting from four non-interfering axes. The phases ϕi have been randomly chosen between 0 and 2pi. 0.45 0.55 0.65 Bond length/λlat 0 100 200 300 O cc ur re nc e Figure 5.11: Histogram of bond lengths (blue) and gap widths (red) for the random-phases lattice patch shown in Fig. 5.10. The gaps are the short diagonals of the rhombus-like tiles. The mean (standard deviation) of this distribution is 0.60(2) λlat and 0.478(6) λlat, respectively. So far we have considered static phases ϕi and the resulting potentials V(r) whose overall structure closely resembles the octagonal tiling in introduced in Chapter 1. Further studies are necessary to decide how exactly the eightfold lattice potential can be used to faithfully model the octagonal tiling. Some of these points will be addressed in the Outlook. Experimentally, one very important question is whether or not phase changes during 82 quasicrystalline optical lattices for ultracold atoms the course of the sequence will result in strong heating in this eightfold lattice, as discussed in the following. Phase control The phase ϕ of a 1D lattice potential that is created by retro- reflection is effectively determined by the path length from the atoms to the retro-reflecting mirror. If this path length varies with time the phase ϕ changes accordingly. Alternatively, the phase of a lattice that is created by the in- terference of running waves can be controlled by changing the incoming phase of the laser light, e.g. by a phase-locked-loop or by deliberately changing the laser frequency. Control over the phase13 of optical lattices have become a cor-13 achived either by stabilisation or deliberate phase changes nerstone in quantum simulation experiments, for example in splitting and combining double-well potentials [192]. However, the application of phase control to quantum gas experiments some- times shows unexpected subtleties. For instance, the powerful method of engineering a variety of hamiltonians by periodically modulating the lattice phases (Floquet engineering [29, 193]) has often been limited to short measurement times and weakly in- teracting gases [194] due to heating. Understanding the role of interactions, in particular, in the heating processes of periodically modulated systems is a field of active research [195–197]; the pre- cise control over the bandwidths of the higher-lying Bloch bands might be a way to mitigate heating [25]. In our experimental setting, the question of phase control is particularly subtle. The fundamental reason for this is precisely the aperiodicity of our lattice potential. In contrast to periodic lattices, in which up to two relative phases can be absorbed into the global centre-of-mass motion of the atoms, in our case there are more than two such phases (ϕi). The non-local effect of changing any of the phases ϕi can be seen by equating an arbitrary-phase potential V(r) with a fixed- phase potential that is displaced by r0 = (x0, y0), i.e. 4 ∑ i=1 cos2 (ki · r + ϕi) = 4 ∑ i=1 cos2 [ki · (r− r0)] which leads to ϕ1 = klatx0 ϕ2 = klaty0 ϕ3 = − klat√ 2 (y0 + x0) ϕ4 = − klat√ 2 (y0 − x0) . Combining these equations yields ϕ3 = − 1√ 2 (ϕ2 + ϕ1) (5.23) ϕ4 = − 1√ 2 (ϕ2 − ϕ1) , (5.24) eightfold optical lattice 83 -4 -2 0 2 4 Position x/λlat -2 0 2 Po si ti on y/ λ la t 0 16 La tt ic e po te nt ia lV (r )/ V 0 Figure 5.12: Lattice potential resulting from four fully-interfering axes (blue- detuned). The phases ϕi have been chosen such that the symmetry point is located at r = (0, 0) and coincides with a potential maximum (maximally con- structive interference). Compared to the non-interfering case, the potential maximum is a factor of four higher in energy. which defines a plane in R4 on which the vector (ϕ1, ϕ2, ϕ3, ϕ4) has to move in order to maintain a simple centre-of-mass motion of the quasicrystalline lattice potential. Any phase changes outside this plane will result in a physical change of the lattice potential. In particular, changing any single phase ϕi without changing the others accordingly, will result in a non-local potential deformation; some sites will disappear and others will appear from nowhere, in close analogy with phasonic excitations mentioned in Chapter 1 that are characteristic for quasicrystals. Therefore it seems likely that the mechanical stability of the retro-reflecting beam path will play a crucial role in the performance of the eightfold optical lattice, particularly once atoms have been loaded into the lattice potential. It would then be very interesting to deliberately excite phasonic modes by slightly displacing one (or several) of the retro- mirrors. Before we turn to the experimental results in momentum space, the situation of fully-interfering lattice beams will be discussed briefly in the following. Fully-interfering axes -4 -2 0 2 4 Position x/λlat 0 4 16 La tt ic e po te nt ia lV (r )/ V 0 Figure 5.13: Two cuts through the interfering lattice potential shown in Fig. 5.12 color-coded accordingly. In contrast to the non-interfering case, potential maxima are steep and isolated.The lattice potential from four interfering lattice axes can be writ- ten as14 14 although the potential should, strictly speaking, include eight in- dividual phases, four incoming and four retro-reflected phases. However, for practical purposes in this work it will be sufficient to consider just four independent phases. V(r) = V0 [ 4 ∑ i=1 cos (ki · r + ϕi) ]2 . (5.25) Experimentally, this can be achieved by simply removing the frequency offsets between each lattice beam. If two or more lattice axes interfere with each other, their respective polarisations also crucially affect the resulting potential. Equation 5.25 only covers the situation of out-of-plane polarisation, i.e. the polarisation of each lattice beam being at right angles to the lattice plane. The opposite case (in-plane polarisation) leads to a modification of the 84 quasicrystalline optical lattices for ultracold atoms potential but the overall structure remains similar to the out-of- plane configuration. In comparison to the non-interfering case discussed above (Figs. 5.7 and 5.8), constructive interference between four axes occurs much less frequently. Consequently, the resulting poten- tial landscape (Eq. 5.25) is dominated by steep, isolated max- ima, which are surrounded by a ‘floor’ of minima, as shown in Fig. 5.12. This effect becomes even more apparent when con- sidering cuts through the potential (Fig. 5.13), emphasising the importance of interference in the eighfold lattice (c.f. Fig. 5.8). 0 2 4 Energy of extremum/V0 0 10 20 30 O cc ur re nc e Figure 5.14: Histogram of minima (blue) and maxima (red) of the exact potential shown in Fig. 5.10 (random phases, non-interfering, blue-detuned lattice). The non-interfering case produces roughly equal amounds of minima and maxima, their distribution is approximately symmetric around 2V0. 0 4 8 12 16 Energy of extremum/V0 0 10 20 O cc ur re nc e Figure 5.15: Histogram of maxima (red) of the exact potential shown in Fig. 5.12 (ϕi = 0, fully-interfering, blue-detuned lattice). The minima would all cluster around zero and have thus been omitted from this plot. In comparison the non-interfering case (Fig. 5.14 above) the maxima are spread out more. The eightfold rotational symmetry is reflected in the histogram by the appearance of maxima in groups of eight. Similar to the non-interfering case, the potential in Eq. 5.25 bears strong similarities with the octagonal tiling, a fact that has been discussed extensively by Jagannathan et al. [85, 198–200]. However, in contrast to the non-interfering case, the octagonal tiling only applies to points of maximally constructive interfer- ence (the isolated maxima discussed above). This means that in order to model the octagonal tiling physically with cold atoms we would have to change the detuning of our lattice from blue- to red-detuned in order to convert the intensity maxima to potential minima. Conversely, the tunnelling between these lattice points (as discussed in ref. [198]) might more closely resemble physical ‘tight-binding’ bonds between lattice sites, compared to the non- interfering case in which the bonds between lattice sites dictated by the octagonal tiling may or may not correspond to the physi- cally most relevant tunnelling matrix elements. Another drawback of the interfering case is that there are numerous shallow ‘local’ intensity maxima surrounding the strong ‘octagonal’ lattice points and it is unclear whether these affect the equilibrium behaviour of a quantum gas in such a lattice. Regarding the ‘bands’ of en- ergy minima, the non-interfering case comes closer to forming a well-defined band of allowed energies (Fig. 5.14), whereas the fully-interfering case shows maxima distributed over the entire range of the potential (Fig. 5.15). Partly-interfering axes The case of partly-interfering lattice axes is particularly interesting since it allows the creation of an separable 2D quasiperiodic po- tential. If two orthogonal lattice axes are chosen to interfere with each other, whereas the other two axes shall be non-interfering, the potential can be written as V(r) = V0 2 ∑ i=1 cos2 (ki · r + ϕi) +V0 [ 4 ∑ i=3 cos (ki · r + ϕi) ]2 . (5.26) A potential of this kind is shown in Fig. 5.16, which is quasiperi- odic in both x and y direction, but without global eightfold ro- tational symmetry. This potential landscape is similar to the 2D bichromatic lattice employed in ref. [24]. It highlights the flexi- bility of our eightfold lattice setup, being able to produce both eightfold optical lattice 85 -4 -2 0 2 4 Position x/λlat -2 0 2 Po si ti on y/ λ la t 0 6 La tt ic e po te nt ia lV (r )/ V 0 Figure 5.16: Lattice potential re- sulting from two fully interfering axes and two non-interfering axes (Eq. 5.26. The phases ϕi are chosen to be (0, 0,pi/2,pi/2). separable and non-separable quasiperiodic potentials in two di- mensions. This completes the discussion of attainable real-space poten- tials. Next, the setup and characterisation of the eightfold lattice (non-interfering case) will be summarised. 5.4 Setup and characterisation The lattice setup is depicted schematically in Fig. 5.17. The axes have been labelled X, Y, D(iagonal), and T(transport). Only the last achromat before the atoms is shown in the schematic. Laser setup The lattice beams are derived from one single-frequency Ti:Sa laser15 at λlat = 726 nm which is far-blue-detuned with respect 15 Matisse, Sirah Lasertechnik GmbH, max. 7W with RefCell locking. The Ti:Sa crystal is pumped by a Millennia laser (Spectra Physics) producing 24W of power at 532nm. to the D-lines in 39K, ensuring suppression of single-photon scat- tering. Cross-interferences between lattice axes are avoided by offsetting their frequencies by more than 10 MHz from each other. Therefore, the corresponding beat notes between the axes oscillate much faster than the atomic kinetic energy scale, given by Erec/h (table 5.1). Beam waists The choice of lattice beam waists is governed by essentially the same trade-off as for the dipole beam waists. • Similar to the dipole trap waists, small lattice waists lead to the highest peak intensity and, consequently, the deepest lattice potential. Therefore, the lattice beam waists should be chosen as small as possible, given a certain maximum optical power (on the ordon the order off 200mW per beam). 86 quasicrystalline optical lattices for ultracold atoms Figure 5.17: Schematic of the science cell optics (lattice in blue and dipole in red). The axes have been labelled X, Y, D(iagonal), and T(transport). The last achromats before the atoms are shown as black double-headed arrows, with their corresponding focal lengths (except for the T-axis for which only one of them is shown). The glass cell is shown in light blue (actual dimensions). f = 125mm f = 125mm f = 125mm f = 125mm f = 125mm f = 100mmf = 100mm D-axis T-axis Y-a xis X-axis LatticeDipoleDip ole Lat tice 25.0mm • Conversely, large in-plane lattice waists (compared to the extent of the atomic cloud) are beneficial in order to ensure homo- geneity of the lattice potential.1616 An overall homogeneous potential is desirable to compare cold-atom quantum simulation experiments to condensed-matter theory, which is often performed in homogeneous lattices. In most lattice experiments the potential envelope is gaussian which necessitates a density-dependent theoretical description of the system. A well-known example is the bosonic or fermionic Mott-insulator. The characteristic ‘wedding-cake’ density profile can be described by the local- density approximation. Elliptical beam profiles are a good compromise between the two. The in-plane waists are chosen as large as possible, whereas the vertical waists can be reduced in order to increase the maximum achievable lattice depth. In addition, the retro-reflecting configu- ration increases the maximum depth (given a certain maximum available beam power) by a factor of two. In the fully-interfering case, the maximum lattice depth is a factor of four higher in the fully-interfering case (see Fig. 5.12, compared to Fig. 5.7), which could support even larger lattice waists (given a maximum power). Blue-detuned lattices lead to an overall anti-confinement re- sulting from the spatially-varying on-site trapping frequencies ωon-site(r). The ground state energy in each well is given by the lowest harmonic oscillator energy E(r) = h¯ωon-site(r) in the two- dimensional case (two lattice beams overlapped in-plane at 90◦). The in-plane envelope E(r) is also approximately gaussian.1717 In a 1D lattice the situation is sep- arable and thus the gaussian anti- confinement is exact. Consequently the anti-confining effect can compensated by weak co-propagating red-detuned dipole beams with larger waists by a factor of √ 2. Alternatively, a single red-detuned laser beam in the vertical direction (also with a factor of √ 2 larger waist) can be used to the same effect. eightfold optical lattice 87 In the eightfold optical lattice these considerations are tricky since the lattice potential is not periodic and it is unclear which of the ‘lattice sites’ – which are all at different energies, even in the homogeneous case – should be taken into account for the anti-confinement. As a guideline, we aimed at a ratio of in-plane waists between lattice and dipole trap of roughly √ 2. However, spatial and physical constraints (gaussian beam-propagation with certain minimal distances in between the components and limited choices of focal lengths) led to slightly smaller waists than initially intended (see table 5.3). In addition, preliminary calculations taking into account the entire spread of ‘lattice sites’ show that the eightfold case leads to an anti-confining waist that is smaller than the lattice waists by a factor of 21/4. We plan to counteract for anti-confining effects using a dedicated dipole trap beam along the z-axis. Table 5.3: Experimental values for the in-plane (wr) and vertical (wz) lattce beam waists (single waists), measured on cameras that are focussed to the position of the atomic cloud. The T- waists are approximately 15% larger than the other waists. Probably this is due to the different optical path length on the incoming beam side (40 cm distance between last lens and atoms). axis wr/ µm wz/ µm D 158 73 Y 153 71 X 161 70 T 184 91 The vertical beam waists (table 5.3) are chosen such that the lattice intensity varies as little as possible across the atomic cloud. Using our maximum power (several hundred mW per beam) we can achieve single-axis lattice depths of at least 40 Erec. Lattice angles One of the limiting factors in observing very high momentum states is the angle between lattice beams. By fitting a straight line to diffraction peaks in time-of-flight the angle of each lattice beam with respect to the absorption image axes is determined. From these measurements (table 5.4) we conclude that the relative angles are 45(1)◦. Table 5.4: Relative lattice beam angles used in this work. axes angle TD 89.5◦ TX 45.6◦ TY 44.0◦ XY 91.6◦ XD 46.2◦ YD 45.4◦ Control loops The intensity of each lattice beam is actively stabilised using a PID feedback loop. Directly after the fibre and the optical isolator (which also acts as a polarisation-cleaning cube) a small (' 4%) proportion of the beam is directed to the ‘PID-photodiode’ using a wedged glass plate. This constitutes the feedback signal for the control loop, the active stabilisation being performed by the RF-power given to the AOM (before the fibre). The lattice pulse signals are created by waveform generators in order to have sub- µs time-resolution and high reproducibility of the DC voltages. In order to maximise the bandwidth of the control loops all cable lengths have been chosen to be as short as possible; the largest contribution comes from the cable connecting the PID-photodiode with the input of the PID-unit (12m). The active control loops have bandwidths on the order of sev- eral hundred kHz and consequently µs-risetimes of lattice pulses can be expected. The PID-parameters have been optimised for a given lattice depth (15Erec) such that a square pulse produces a maximal rise time (0% to 90% set voltage) of 1.1µs, thereby minimising any overshoot. For different pulse shapes and lattice depths the parameters have to be re-optimised for best perfor- 88 quasicrystalline optical lattices for ultracold atoms mance; as usual, there is a trade-off between short rise times and minimal overshoot. -40 -20 0 20 40 In -s it u po si ti on /µ m -600 -300 0 300 600 70 80 90 100 In -s it u cl ou d si ze /µ m Figure 5.18: Lattice sliding method. The upper plot shows the in-situ position of the condensate on the Z-camera, at right angles to the lat- tice beam direction. The dashed line corresponds to the unshifted con- densate position. The lower plot shows the same scan, but using the Thomas-Fermi radius of the cloud as observable. This data was taken using the horizontal knob of the D-axis piezo mirror. Setup and alignment The precise alignment of the lattice beams is similarly critical as that of the dipole trap.18 In brief, the following alignment steps 18 We showed in section 4.4 that the vertical position of the dipole beam can be discrimiated to within 2um. should be observed. • The focus of the incoming beam (beam waist) has to be over- lapping with the atomic cloud. Similar to the dipole trap, the final achromat before the atoms is translated longitudinally (z- translation stage with µm-skrews) to minimise the beam waist. The beam waist can be measured with a dedicated camera on each lattice axis that has been focussed to the position of the atoms via in-situe absorption imaging. In the case of the X and Y axes, which are shared with the dipole trap, this alignment step is omitted since we trust the achromat to be already well- aligned having minimised the dipole trap waists. This assumes firstly that the focal shift due to the different wavelengths is minimal, and secondly, that the expected beam propagation of the lattice beam (using ABCD-matrices) is faithful. The latter point is assured by measuring the actual beam divergence after the fibre coupler for each individual lattice axis. • The pointing of the lattice beam can be aligned via the so-called ‘lattice sliding method’. It entails blocking the retro-path and shining a relatively weak lattice potential onto an atom cloud which is imaged in-situ along an axis that is at right angles to the lattice beam. Since our lattice is blue-detuned, the beam leads to a repulsive potential. If the lattice beam is off-centre in one direction, the atomic cloud is displaced towards the opposite direction. By taking a scan of vertical and horizontal displacement using a piezo-driven mirror, as shown in Fig. 5.18, the lattice beam can be centrally aligned to within a few µm. Note that due to the anti-symmetric response of the cloud, it is crucial to take an entire scan and find the symmetry point which corresponds to the optimum, i.e. unshifted position. In addition, the repulsive potential leads to a broadening of the atom cloud, providing an additional signature of good alignment (lower half of Fig. 5.18). • On the retro-path, the transverse position of both lenses in the cats-eye configuration [201] is important in order to ensure good directional overlap between incoming and retro-reflected beam. This can be achieved by placing a pinhole on each lens mount and maximising the transmission. In addition, the lon- gitudial position of the retro-mirror and the lenses is crucial to ensure approximately equal19 beam waist of incoming and19 The beam emerging from the fi- bre coupler (beam diameter roughly 1.2mm) can never be perfectly col- limated. Hence a 1:1 reproduction of beam waists at the position of the atoms is precluded by the laws of gaussian propagation. retro-reflected lattice beams. It should be noted that for ellipti- cal beams, there are, in principle, no degrees of freedom in the longitudinal position of both the lenses and the retro-mirror, eightfold optical lattice 89 because the cat’s eye requires the distance between atoms and retro-mirror to be exactly twice the sum of the two focal lengths. The achromat on the atoms’ side is assumed to be aligned to the in-situ position of the atoms via absorption imaging. The second cat’s eye lens is placed in good faith by measuring the distances with a ruler. • Finally, the directional pointing of the retro-reflected beam at the position of the atoms is relatively insensitive to the angle of the retro-mirror due to the cat’s eye configuration. In particular, the diffraction angle (Kapitza-Dirac diffraction [202]) at the position of the atoms does not depend on the angle of the retro- mirror. Nevertheless, the angle can be optimised by maximising the effect of 1D Kapitza-Dirac diffraction. 6 Matter-wave diffraction experiment This chapter presents the main experimental results of this thesis, namely the matter-wave diffraction experiments. Short lattice pulses (Kapitza-Dirac diffraction [202, 203]) are used to probe the 2D quasicrystalline structure in reciprocal space. First, the basis of momentum in space of the 2D quasicrystal is described in section 6.1, building on the 1D results from sec- tion 5.1. Section 6.2 covers the experimental methods that are used (short lattice pulses). Afterwards, both the static and the dynamic diffraction patterns are described in theory and experiment. The diffraction dynamics, for very short pulses, are shown to represent quantum walks. Finally, our methods of data-analysis are detailed in section 6.5. 6.1 Basis in momentum space Similar to the indexing method for diffraction pattern of 1D qua- sicrystals (section 1.2) we are going to write down a method to identify all basis states of the 2D quasicrystalline potential in mo- mentum space using four integers. The basis states in momentum space correspond to the matter-wave diffraction peaks that are going to be observed in the experiment. As before, the real-space potential of our optical lattice can be written as VD(r) = V0 D ∑ i=1 cos2 ( Gi 2 · r ) , (6.1) where D = 1, 2, 3, or 4 is the number of mutually incoherent lattice beams and V0 = 14.6(2) Erec is the individual lattice depth.1 The 1 The lattice wavelength λlat = 2pi/klat = 726 nm is far detuned with respect to the D-lines in 39K, ensuring that single-photon processes are completely suppressed. The value of Erec is listed in table 5.1. reciprocal lattice vectors Gi are defined as G1 = ( 1 0 ) , G2 = ( 0 1 ) , G3 = 1√ 2 ( 1 1 ) , G4 = 1√ 2 ( −1 1 ) (6.2) where we have switched to dimensionless units in which 2klat = 1 as shown in Fig. 6.1 (right inset). In contrast to a periodic lattice the combination of several Gi here may give rise to new, smaller momentum scales, as shown the left inset of Fig. 6.1; for example, the combination −G1 + G3 − G4 results in a new k-vector (red arrow) that is shorter than the original G1 by a factor of 1 + √ 2 (the silver mean). This process can be repeated ad infinitum and 92 quasicrystalline optical lattices for ultracold atoms results in a self-similar structure containing arbitrarily small k- vectors, as indicated by the sequence of octagons in Fig. 6.1. Con- sequently, it is impossible to assign a maximum characteristic length to this quasicrystal, heralding the presence of structure on all scales. Furthermore, the momenta that are reachable from k0 = (0, 0) form a dense set in the kx, ky-plane and any element G of this set is determined by four integers (i, j, l, n) ∈ Z4 as G = iG1 + jG2 + lG3 + nG4 . (6.3) While physical momentum remains two-dimensional, all four integers are nonetheless required to describe a given G, since cos(45◦) = sin(45◦) = 1/ √ 2 is irrational and hence incommensu- rable with unity. In fact, Fig. 6.1 can be viewed as an incommensu- rate projection of a 4D simple-cubic ‘image’ lattice to the 2D plane, similar to the ‘cut-and-project’ scheme for constructing the octag- onal tiling (section 1.3). By using fewer than four lattice beams we can control the dimensionality of the image lattice and Z4 reduces to ZD with D ∈ {1, 2, 3, 4}. -6 -4 -2 0 2 4 6 8 kx/klat 0 2 4 k y /k la t 0th 1st 2nd 3rd ≥ 4 Gˆ1 Gˆ2 Gˆ3 Gˆ4 −Gˆ1 +Gˆ3 −Gˆ4 Figure 6.1: Momentum space struc- ture. The first 15 orders of possible diffraction peaks are shown. They are constructed by iteratively adding or subtracting one of the four reciprocal lattice vectors Gi (inset on the right) to the peaks in the previous order, start- ing with k = (0, 0). This results in a fractal structure, whose self-similarity is illustrated by a sequence of oc- tagons, which are each scaled by the silver mean 1+ √ 2 relative to the next. The left inset shows one inflation step (see text). Basis in momentum space We can use Eq. 6.3 to express any accessible k-state as an integer- valued vector bi in ZD, where D is the number of active lattice beams. As described for the 1D case in section 5.1 the nth order of this basis is defined as the set of all elements {bi} with D ∑ j=1 abs ( [bi]j ) = n (6.4) where [a]j denotes the jth component of a vector a and abs() denotes the absolute value. The total number of basis states N up to and including the nth order for D = 4 is listed in table 6.1; it can matter-wave diffraction experiment 93 be written in a closed form as N(n) = min(4,n) ∑ k=0 2k ( 4 k )( n k ) . (6.5) Table 6.1: Number of basis states up to and including the nth order in the 4D case (eightfold lattice). n N(n) 0 1 1 9 2 41 3 129 4 321 5 681 . . . 11 11969 The in table 6.1 are also known to be the coefficients of the Taylor expansion of the following function [204] (1+ x)4 (1− x)5 (6.6) around x = 0. We will later truncate ZD at the nth order, meaning we only take into account states that can be reached by at most n two- photon scattering events. Projection For D = 3 and D = 4 the projection matrices for the states bi onto the kx, ky–plane are given by M3 =  1 0 1√ 2 0 1 1√ 2 0 0 0  (6.7) and M4 =  1 0 1√ 2 − 1√ 2 0 1 1√ 2 1√ 2 0 0 0 0 0 0 0 0  , (6.8) respectively. The third and fourth dimensions are projected to the in-plane diagonals with respect to the first two dimensions. The first two rows of M3 and M4 are simply given by the Gi defined in Eq. 6.2. Note the similarity of M4 to the rotation matrix for the 4D cut-and-project scheme (Eq. 1.21 in section 1.3). Fig. 6.1 shows elements of the basis of order 15 for D = 4, projected onto the kx, ky–plane using the matrix M4. For D = 1(2) the matrices M1(M2) are trivial because here the dimension of bi is the same as the physical dimension. Hamiltonian in momentum space Having constructed the basis we can write down the hamiltonian HD in momentum space HDi,j =  V0/4 for |bi − bj| = 1 4Erec × |MD · bi|2 +V0/2 for i = j 0 otherwise (6.9) Here the norm of a D-dimensional vector a is given by |a| = √√√√ D∑ j=1 ( [a]j )2 . (6.10) The non-zero off-diagonal elements correspond to transition ele- ments for stimulated two-photon scattering events, where atoms scatter photons from one lattice beam into its counterpropagating partner. These transitions connect discrete momentum states sep- arated by ±Gi and effectively realise a tight-binding hamiltonian 94 quasicrystalline optical lattices for ultracold atoms 1D 2D 3D 4D Op tic al de ns ity 0.0 0.3 Figure 6.2: Raw time-of-flight images of 39K resulting from four different lat- tice configurations at fixed lattice pulse duration (t = 3.5 µs). Using just one of the lattice axes results in a regular 1D simple-cubic lattice characterized by G1; adding the perpendicular lattice creates a regular 2D square lattice with G1 and G2. By adding the first diagonal lattice we obtain a lattice of 1D quasiperiodic potentials with dense sets of momentum states along G3 whereas the direction perpendicular to G3 remains regular (labelled 3D). Finally, using all four axes we create the 2D quasicrystal (labelled 4D); its self-similarity is illustrated by the octagons. in momentum space [205]. The matrix elements on the diagonal are given by the kinetic energy term, where the prefactor 22Erec arises from the momentum scale 2h¯klat of the individual lattices and a constant offset of V0/2 which arises from the k = 0 Fourier component of VD(r) (it has no physical significance since it corre- sponds to a global shift in energy). Truncation of basis As in the 1D case (Eq. 5.10), this hamiltonian is infinite-dimensional and, consequently, we need to make it numerically tractable by truncating it. Since the experiment starts with a pure condensate in the |h¯k = 0〉 state, and we apply only short pulses of lattice light, it is sensible to work with a basis of order n = 11. This can be justified a posteriori since even for the longest applied lattice pulses our simulation (using a basis of order 15) shows that the orders n > 11 get populated by less than 15 per cent. As before, the time-evolution can be calculated by numerically integrating the time-dependent Schrödinger equation. This is achieved by writing the hamiltonian matrix HDi,j in the truncated momentum basis; integration is performed using the scipy.integrate library. 6.2 Method: Kapitza-Dirac diffration The experimental sequence starts with the preparation of an al- most pure Bose-Einstein condensate of 39K atoms in a crossed- beam dipole trap, as described in Chapters 3 and 4. Using the Feshbach resonance centred at 402.70(3)G (section 4.6) we tune the contact interaction to zero just before we release the conden- sate from the trap. Then we immediately expose it to the optical lattice for a rectangular pulse of duration t. During this pulse, atoms in the condensate can undergo several stimulated two- photon scattering events (Kapitza-Dirac scattering [206]), which scatter photons from one lattice beam into its counterpropagating partner and transfer quantized momenta of ±2h¯klat, where h¯klat is the momentum of a lattice photon and |Gi| = 2klat. Throughout matter-wave diffraction experiment 95 1D 2D 3D 4D Op tic al de ns ity 0.00 0.25 Figure 6.3: Raw time-of-flight images of 87Rb resulting from four different lattice configurations at fixed lattice pulse duration (t = 5.0 µs). Compared to Fig. 6.2, which was taken with 39K, the resolution in momentum space is worse. In addition, strongly pronounced ‘scattering shells’ (see text) appear for 87Rb (scattering length a = 100 a0) which lead to a diffuse background in the 4D situation. this chapter, the lattice depth of each individual axis is 14.6(2)Erec. Finally, we record the momentum distribution of the atomic cloud by taking an absorption image after tToF = 33 ms (time-of-flight). In a first experiment we fix the lattice pulse duration at t = 3.5 µs and vary the number of lattice beams, as shown in Fig. 6.2. Start- ing from the single-axis (1D) case, we subsequently add lattice axes, finally completing the eightfold symmetric case (4D), rep- resenting the quasicrystalline structure with its striking self- similarity under (1+ √ 2) scaling. Interaction effects For diffraction experiments that are carried out with finite contact interactions (e.g. using 87Rb, Fig. 6.3) one finds that the time-of- flight images feature pronounced ‘scattering shells’ [207, 208] connecting the discrete momentum peaks. These shells appear as characteristic rings on the absorption images and arise from two- body s-wave collisions between parts of the atomic cloud which are moving with respect to each other. In order to eliminate this effect we perform all experiments with 39K and tune the contact interaction to zero by ramping the magnetic field to a value of 351.5(1)G, just before the optical lattice pulse is applied. How- ever, atomic clouds of 39K at vanishing interactions are optically dense enough to absorb essentially all imaging light, preventing any faithful atom number measurement. Therefore we turn inter- actions back on (back to the previous value of roughly 280 a0) once the diffraction orders have separated from each other, such that the individual peaks expand and reveal their atom populations. 6.3 Diffraction dynamics The diffraction dynamics offers one of the most striking signa- ture of the aperiodicity of the eightfold optical lattice. During the lattice pulse the condensate explores reciprocal space in discrete steps of ±Gi, leading to profoundly distinct behaviours in the periodic (2D) and in the quasicrystalline case (4D). Fig. 6.4 shows absorption images for four different values of pulse duration t in 96 quasicrystalline optical lattices for ultracold atoms t = 0.0μs t = 1.0μs t = 3.0μs t = 6.0μs Op tic al de ns ity 0.0 0.3 Figure 6.4: Dynamics of Kapitza-Dirac diffraction in the quasicrystalline optical lattice. The figure shows raw absorption images for four different lattice pulse durations. After 1µs, only the first diffraction order has been populated, while longer pulses lead to populations in successively higher orders as the atoms perform a quantum walk in momentum space. Each octagon has a circumradius of |Gi | = 2klat, illustrating the fun- damental momentum scale due to two-photon processes. the latter configuration, illustrating the occupation of more and more closely spaced momenta. Using individual fits we extract the number of atoms in every k-state up to the seventh diffrac- tion order, i.e. those momenta reachable by seven or fewer two- photon scattering events. In the 2D simple cubic lattice (Fig. 6.5 on the left) only a few momentum states are accessible, since ki- netic energy acts as a harmonic confinement in momentum space. Therefore the dynamics is oscillatory and reminiscent of a sim- ple harmonic oscillator. In the quasicrystalline case, in contrast, the diffraction dynamics is aperiodic (4D, right of Fig. 6.5), since large parts of the population propagate ballistically to progres- sively higher orders, as illustrated by the light blue ‘wave front’. Due to the self-similar momentum space structure, the atoms can access momentum states in successively higher diffraction orders Figure 6.5: Kapitza-Dirac diffrac- tion dynamics in a periodic (2D) and quasicrystalline (4D) lattice. The nor- malized populations (coloured dots) of the condensate (0th order) and the first seven diffraction orders are plot- ted against pulse duration, together with the numerical solution to the Schrödinger equation (lines). The peri- odic case (2D) is oscillatory as kinetic energy limits the accessible momenta. In contrast, the quasicrystalline lattice (4D) contains a dense set of k-states, c.f. Fig.6.1, enabling the population of higher and higher orders. Correspond- ingly, the expansion carries on linearly, indicated by the light blue ‘wave front’ as a guide to the eye. Error bars denote the standard deviations from five re- alisations of the experiment, and are typically smaller than symbol size. 0.0 0.5 1.0 0 5 10 15 20 Lattice pulse duration (µs) 0 5 10 15 20 25 Lattice pulse duration (µs) 2D 4D 1st 2nd 3rd 4th 5th 6th 0th 1st 2nd 3rd 4th 5th 6th 0th 7th7thP op ul at io n matter-wave diffraction experiment 97 0 1 2 3 Lattice pulse duration (µs) 0.0 0.5 1.0 1.5 2.0 rm s m om en tu m (2 h¯ k la t) 1D 2D 3D 4D 0 3 0.0 1.0 rm s/ √ D (2 h¯ k la t) Figure 6.6: Continuous-time quantum walk in momentum space in D dimen- sions, where D is controlled by the number of lattice beams. Dots repre- sent the measured root-mean-square momentum (see text), while lines rep- resent numerical solutions to the full Schrödinger equation. The inset shows the same data, but scaled by √ D. Here the dashed line represents the expan- sion dynamics of a continuous-time quantum walk on a homogeneous D-dimensional tight-binding lattice. The √ D scaling (Eq. 5.17) is a direct consequence of the separability of hy- percubic lattices. Deviations from the linear behaviour at later times are due to kinetic energy, and the lines would differ from each other at long times. Error bars denote standard deviations from five identical realisations of the experiment. that correspond to ever smaller momenta. Our data agrees ex- cellently with exact numerical solutions (lines in Fig. 6.5) of the single-particle time-dependent Schrödinger equation in momen- tum basis. 6.4 Quantum walks in high-dimensional homogeneous tight- binding lattices 0 2 4 6 8 10 12 14 Lattice pulse duration (µs) 0.0 0.5 1.0 1.5 2.0 rm s m om en tu m /√ D (2 h¯ k la t) vp × t 1D 2D 3D 4D Figure 6.7: Root-mean-square mo- mentum for longer times, scaled by a factor √ D as in Fig. 6.6 (inset) in the main text. The dashed black line is the analytic result (Eq. 5.17) in a D-dimensional homogeneous tight- binding model. The dashed blue lines are the solutions to the time-dependent Schrödinger equation, using all states with n ≤ 11. If we take into account that we only detect atoms up the sev- enth diffraction order, the expansion is reduced to slightly lower momentum values (solid lines). In the regime of very short pulses kinetic energy can be neglected for all dimensions and the discrete momentum space structure can be mapped onto a homogeneous tight-binding lattice [205]. A hopping event in this effective lattice corresponds to a two-photon scattering event and connects momenta differing by ±h¯Gi. In this mapping, the diffraction dynamics is equivalent to the expansion of initially localised particles and gives rise to a continuous-time quantum walk with its characteristic light-cone-like propaga- tion [22, 190, 191]. For a hypercubic lattice in D dimensions, the separability of the tight-binding dispersion relation (section 5.1) leads to an average group velocity proportional to √ D (Eq. 5.17). Due to the correspondence between the number of lattice beams and the dimension of the tight-binding hamiltonian, we are able to extend the dynamics to up to four dimensions. Using the appro- priate form of Eq. 6.3 in ZD, we extract the effective root-mean- square momentum in D dimensions, e.g. √〈i2 + j2〉 in the 2D case and √〈i2 + j2 + l2 + n2〉 in the 4D case, from the individual populations of all diffraction peaks, and find excellent agreement between the measurements and the analytic result vp ∝ √ D, as shown in Fig. 6.6. The departure from linear behaviour at longer times is due to kinetic energy and is captured well by the exact numerical solution to the Schrödinger equation (solid lines in 98 quasicrystalline optical lattices for ultracold atoms Fig. 6.6). The extent of the linear region is controlled by the lattice depth. For even longer times, kinetic energy enforces fundamen- tally different behaviours for periodic and quasicrystalline lattices, as shown in Fig. 6.5 (and in Fig. 6.7). 6.5 Data analysis The data analysis procedure for determining the population in each diffraction peak is described in the following. First, we deter- mine the position of the condensate, find the angle of one lattice axis relative to the camera axes, and calibrate the magnification us- ing reference images showing only zeroth and first order diffrac- tion peaks. With this information we can calculate the expected position of each momentum peak. Around each calculated peak position we perform an individual fit to a 2D Thomas-Fermi pro- file (Eq. 4.14) in a square bin of 28 × 28 pixels (56 × 56 for the central condensate). In order to mitigate effects of imaging satu- ration, the fit ignores pixels with optical densities above 2.0. The corresponding atom population of each basis element pi(t) is pro- portional to the integrated Thomas-Fermi profile, the result of which is in Fig. 6.8 as an example. If the population pi(t) is below 0.04% of the total population we ignore it in order to avoid count- ing spurious populations in high diffraction orders, which would otherwise dominate the rms. For Fig. 6.5 we sum all populations in one diffraction order. Figure 6.8: Simplified example of the population count. (A) The raw absorption image (D = 4, t = 3µs). At each calculated peak position with n ≤ 7 we perform an individual fit to a 2D Thomas-Fermi profile. The resulting populations pi(t) are depicted in (B) by the area of the circles. A B Root-mean-square extraction We calculate the root-mean-square momentum in D dimensions as a function of time as √ ∑ i pi(t) ∑j pj(t) |bi|2 , (6.11) where pi(t) are the populations in each diffraction peak bi at a given time t. The sums go over all elements of a basis. matter-wave diffraction experiment 99 0 5 10 15 20 25 Lattice pulse duration [µs] 0 200 400 600 800 1000 1200 1400 1600 1800 S um of po pu la tio n in al lp ea ks (a .u .) 0 5 10 15 20 25 Lattice pulse duration [µs] 0 200 400 600 800 1000 1200 1400 1600 1800 S um of po pu la tio n in al lp ea ks (a .u .) 2D 4D Figure 6.9: Total detected population (a.u.) in all diffraction peaks for the 2D and 4D situations. The solid black lines correspond to the total population and the dashed lines are the simulated populations in peaks up to seventh order and additionally account for peaks that fall below the cutoff. When single peaks (such as the central condensate) are strongly populated we systematically detect too few atoms due to imaging saturation and the finite signal-to-noise ratio of the camera. This effect is reflected by the apparent rise in total detected population during the first few µs. It also explains the subsequent ‘dips’ in atom number in the 2D case. Total population Figure 6.9 shows the total detected population ∑j pj(t) summed over all diffraction peaks for the cases D = 2, 4. For D > 2, the detected population is reduced for longer lattice pulses since more and more peaks are weakly populated and fall below the cut- off. In addition, we underestimate the population of very highly populated peaks, such as the initial condensate, due to the finite signal-to-noise of the camera. Conclusion In conclusion, a quasicrystalline potential for ultracold atoms was realised, which can facilitate the creation of ever more com- plex many-body systems [31]. By observing the occupation of successively closer-spaced momenta, we were able to confirm its self-similar structure in momentum space. In addition, the fun- damentally different diffraction dynamics between periodic and quasicrystalline potentials have been verified, in excellent agree- ment with theory. Finally, we demonstrated the ability to simulate tight-binding models in one to four dimensions, by observing the ballistic spreading of particles in reciprocal space. Summary and outlook In summary, this thesis establishes quasicrystalline potentials as a novel platform for ultracold atoms. The eightfold optical lattice is a highly versatile system, allowing investigation of a large vari- ety of lattice potentials. These include several 1D and 2D periodic geometries, 1D quasiperiodic lattices, and 2D quasiperiodic po- tentials, both separable and non-separable. This work represents the first demonstration of a non-separable, 2D quasicrystalline lattice for ultracold atoms, thereby bringing closer the fields of quasicrystals and ultracold atoms. The focus of this thesis has been on the experimental realisa- tion of above mentioned platform. Experimental details have been given on initial laser cooling (Chapter 3), the production of quan- tum gases (Chapter 4), and the setup and characterisation of the eightfold optical lattice (Chapter 5). In a series of proof-of-principle experiments, two important properties of 2D quasicrystals, namely self-similarity and the connection to higher dimensions, have been demonstrated with ultracold atoms for the first time. These include the direct obser- vation of a quantum walk on a homogeneous tight-binding model in four dimensions. Next steps Having established the eightfold lattice potential, there are many interesting steps to be taken from here. Adiabatic loading 0 10 20 30 Time/µs 0 1 2 3 La tt ic e de pt h/ V 0 Figure 6.10: Triangular lattice pulse shape, duration 20 µs, lattice depth 2.5 Erec. So far, all the experiments described in this thesis have been prob- ing the quasicrystal in momentum space, using transient pulses of laser light. The next obvious step would be to study the sys- tem in equilibrium (if this is possible) and load atoms into the potential for longer times. Here, one of the main challenges will be understanding adiabaticity in this aperiodic potential. Prelim- inary studies show that, for small lattice depths (V0 . 1.5Erec), adiabatic loading of the lattice should indeed be possible. This can be seen by considering triangular lattice pulses as shown in Fig. 6.10. After ramping up the eightfold lattice to a certain value V0 and subsequently ramping it back down, the fraction of atoms remaining in the condensate (0th diffraction order) is a signature 102 quasicrystalline optical lattices for ultracold atoms Figure 6.11: Theoretical final popu- lation in 0th diffraction order after triangular pulse of duration t (x-axis) for various lattice depths V0 in the eightfold optical lattice. After an initial dip, the condensate fraction (0th order) recovers to almost unity for shallow lattice depths (light blues). For higher depths (deep blues), the long-term behaviour shows no recovery of the 0th order, which can be interpreted as non-adiabaticity on these timescales. For t > 100 µs the curves have been fitted with linear functions, the slopes of which are shown in Fig. 6.12. 0 100 200 Triangular pulse duration/µs 0.0 0.5 1.0 Po pu la ti on 0t h or de r 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 Figure 6.12: Slopes from Fig. 6.11, plotted against lattice depth. Non-zero slopes start to appear above 1.5 Erec heralding non-adiabatic behaviour. 1.0 2.0 3.0 Lattice depth V0/Erec 0 10 20 N on -a di ab at ic sl op e ×1 04 /( − 0t h po p/ µs ) of adiabaticity [209]. This has been theoretically investigated in Fig. 6.11 for the full eightfold lattice potential (solutions to time-dependent Schrödinger equation). For shallow lattices (faint blues) adiabaticy can be achieved, even in the aperiodic lattice. However, lattice depths higher than ' 1.5Erec (deep blues) show a non-zero slope for longer lattice pulse durations, heralding non-adiabatic behaviour. The slopes have been fitted for t > 100 µs as a quantitative esti- mate of adiabaticity and they show a cross-over behaviour from adiabadic to non-adiabatic (Fig. 6.12). Experimentally, this behaviour will be very interesting to inves- tigate, in particular as a function of interaction strength. Topology One of the great achievements in the field of quantum gases has been the experimental realisation of topologically non-trivial sys- tems [210–212]. In future, aperiodic lattices may find an appli- cation in this context. For example, a topological charge pump has been demonstrated experimentally using a 1D photonic qua- sicrystal [5]. It would be worth pursuing further studies along this matter-wave diffraction experiment 103 direction in quantum gas systems, particularly in the light of the recent observation of a 4D quantum Hall response in tilted optical superlattices [86]. Many-body localisation Another application of aperiodic lattices for ultracold atoms has been the observation of many-body localisation [88]. In particular the role of interactions in quasiperiodic media is still poses fun- damental open questions in physics [89], and experimental results have only recently become available [24, 88, 90, 91]. So far, all ex- periments employed either 1D [88, 90, 91] or 2D separable [24] potentials. However, more experimental evidence is needed, for instance, to differentiate between two universality classes of local- isation phenomena, namely in random disorder and in aperiodic order [92]. For the latter class, our experiment is an ideal starting point since the lattice potential closely resembles the octagonal tiling (see section 5.3), a paradigmatic 2D quasicrystal. Moreover, our system will be able to investigate the fate of a Mott-insulator in the presence of a weak aperiodic perturbation in two dimensions. In one dimension, there has been evidence of a Bose-glass phase [213, 214], but the two-dimensional case has not been experimentally investigated so far. More ideas In the four-beam lattice the total number of basis states grows very fast with the diffraction order (table 6.1). For example, a basis of order eleven contains almost 12000 states. Therefore it would be interesting to reduce the atom number in the diffraction experi- ment by some orders of magnitude until it becomes comparable to the number of basis states. Will the discreteness of the population then show in the diffraction dynamics? A List of components Near-resonant laser system (‘red table’) identifier description part number manufacturer Lasers K master TA 766.7nm, 2.0W, 760-775nm coarse tuning, 50GHz mode-hop-free TA pro 765-3V0 Toptica Photonics AG Rb master TA 780.24nm, 3.0W, 770-795nm coarse tuning, 50GHz mode-hop-free TA pro 780-3V0 Toptica Photonics AG K imaging 766.7nm, 80mW DL pro Toptica Photonics AG Rb repump 780.24nm, 100mW, 765-805nm coarse tuning, 30GHz mode-hop-free DL pro 780-3V0 Toptica Photonics AG K red loading 767nm, 2.0W BoosTA pro 765-1V0 Toptica Photonics AG K blue loading 767nm, 2.0W BoosTA pro 765-1V0 Toptica Photonics AG K red MOT 767nm, 2.0W BoosTA pro 765-1V0 Toptica Photonics AG K blue MOT 767nm, 2.0W BoosTA pro 765-1V0 Toptica Photonics AG Rb loading MOT 780nm, 3.0W BoosTA pro 780-1V0 Toptica Photonics AG Rb cooling 780nm, 3.0W BoosTA pro 780-1V0 Toptica Photonics AG Rb 780/repump 780nm, 3.0W BoosTA pro 780-1V0 Toptica Photonics AG Optics M45 Laser mirror HR 630-900 nm/0-45◦ s+p, R>99.6% HR 630-900nm Lens-Optics GmbH W2 λ/2 0th order wave plate W2Z15-773 Lens-Optics GmbH W4 λ/4 0th order wave plate W4Z15-773 Lens-Optics GmbH magic-W λ/2 at 780nm and λ at 767nm 0th order wave plate W2Z15-L/2@780&L@767 Lens-Optics GmbH PBS 1/2 inch polarising beam splitter PBC12-600/900 Lens-Optics GmbH BS50 Non-polarising beam splitter 50:50 BS005 Thorlabs, Inc F15 Fibre coupler, f = 15 mm APC, 600-1050nm (used for TA beams) 60FC-4-M15-02 Schäfter-Kirchhoff GmbH F15-A Fibre coupler, f = 15 mm APC, 750-1550nm (used for TA beams) 60FC-4-M15-37 Schäfter-Kirchhoff GmbH F8 Fibre coupler, f = 8.1 mm APC, 630-980nm (for all other beams) 60FC-4-M8-10 Schäfter-Kirchhoff GmbH WW30 Glass plate beam spliter, fused silica, 30min of arc, λ/10 quality WW1Q Lens-Optics GmbH AOMs 80 AOM 80(30) MHz, 700-1100nm, 1×2mm2 active aperture, >2W RF MT80-B30A1-IR AA opto-electronic 110 AOM 110(50) MHz, 700-1100nm, 1×2mm2 active aperture, >2W RF MT110-B50A1-IR AA opto-electronic 200 AOM 200(50) MHz, 700-900nm, 1×2mm2 active aperture, >2W RF MT200-B50A1-800 AA opto-electronic Other Optical fibres Single-mode polarisation-maintaining fibres P3-630PM-FC Thorlabs, Inc CoSy K Fibre-coupled spectroscopy CoSy TEM Messtechnik AG CoSy Rb Fibre-coupled spectroscopy CoSy TEM Messtechnik AG 106 quasicrystalline optical lattices for ultracold atoms Vacuum part description part number manufacturer Vacuum chamber Custom, mostly 316LN (MOT, 20 cm diff. pumping section  10mm, 2 TSPs, knee) - PiNK GmbH Small ion pump Vacion plus 75 l/s starcell pump 9191440 Agilent Technologies LDA Large ion pump Vacion plus 150 l/s starcell pump with installed heaters 220V 9191542 Agilent Technologies LDA Titanium sublimation pump TSP filament cartridge 9160050 Agilent Technologies LDA DN16CF bellow Bellow between 2D MOT and 3D MOT MEW0750251C1 Kurt J. Lesker DN16CF straight valve Valve between 2D MOT and 3D MOT 48124-CE01-0001 VAT Vacuum Products Ltd DN16CF angled valve Valve for potassium and rubidium ampoules 54124-GE02-0001 VAT Vacuum Products Ltd DN40CF angled valve (high-T) Angled valve for high-temperature bakeout section (near science cell) 57132-GE02-0005 VAT Vacuum Products Ltd DN40CF angled valve (low-T) Angled valve for low-temperature bakeout (all others) 54132-GE02-0001 VAT Vacuum Products Ltd Lowest-pressure gauge Pressure gauge for science cell section Barion-extended Vacom GmbH Standard gauge Pressure gauge for MOT chamber and turbopump BAT40C Vacom GmbH Non-magnetic viewport DN40CF 316LN steel, viewport used near science cell, BBAR coating 550-1100nm XTEMP-FT Kurt J. Lesker Standard viewport DN40CF, BBAR coated (two parts broke under vacuum in 2D MOT) VPZL-275BBAR Kurt J. Lesker Non-magnetic viewport DN40CF BBAR coated, replacement for above VPCF40DUVQ-L-BBAR-316L Vacom GmbH MOT setup part description part number manufacturer 2D MOT Vacuum chamber Titanium (grade 2) frame and diff. pumping section (custom made, partly faulty) S2-CF-5045/5044 Vaqtex S.r.l. Window BK7 optical window, 85mm×50mm, thickness 10mm, AR 760-790nm PL.OF.85.50.760 Pegasus Lasersysteme GmbH Polariser 25.0mm  linear glass polariser 510-800nm LPVISC100 Thorlabs, Inc Large λ/4 Mico retarder λ/4 wave plate, 70mm×35mm, 780/767nm RQM7035-C/RQM7036-C Knight optical Ltd MOT Fibre cluster 2 to 6 port fibre cluster, 767nm and 780nm, incl. single-mode fibres and couplers - Schäfter + Kirchhoff GmbH λ/4 wave plate Polymer λ/4 wave plate (custom made) WPQ10E-773-SP Thorlabs, Inc Fluorescence setup Rb filter Laser line filter for 780nm FL05780-10 Thorlabs, Inc K filter Laser line filter for 766nm 65-116 Edmund optics list of components 107 Dipole trap laser system part description part number manufacturer Laser 1064nm, 55W, master diode and four integrated amplifiers Mephisto MOPA 55W Coherent Inc Water-water chiller 1000W at 20◦C (operating) and 15◦C (house), 5.6 l/min P201-16675-WW-S Termotek AG Coolant Cooling mixture P/N 1607-0546L Nalco® CCL100 Nalco Optical isolator 1064nm free-space isolator IO-5-1064-HP Thorlabs, Inc Mirror (M45-1064) High-power mirror, 45◦ s+p R> 99.6%, 500kW/cm2 M1064 Lens-Optics GmbH Wave plate (W2-1064) λ/2 0th order wave plate W2Z15-1064 Lens-Optics GmbH Cube (PBS-1064) 1/2 inch polarising beam splitter PBC12-1064 HP Lens-Optics GmbH AOM 80MHz, 1064nm, 1×2mm active aperture, >2W RF MT80-A1-1064 AA opto-electronic Fibre Photonic crystal fibre, SMA connector, 5m, E324-443-500 aeroGUIDE POWER-15-PM NKT photonics A/S Fibre coupling lens f = 18.40 mm, mounted geltech aspheric lens, AR 1050-1700nm C280TMD-C Thorlabs, Inc Polariser Glan-Laser polariser, 10mm diameter CA, 1064nm V-coating GL10-C26 Thorlabs, Inc Photodiode InGaAs amplified detector 900-1700nm PDA10CS-EC Thorlabs, Inc Cylindrical lens f = −30 mm, C-coating 1050-1700nm LK1982L1-C Thorlabs, Inc Cylindrical lens f = 150 mm, C-coating 1050-1700nm LJ1934L1-C Thorlabs, Inc Lens f = −100 mm, C-coating 1050-1700nm LC1120-C Thorlabs, Inc Lens f = 125 mm, C-coating 1050-1700nm LA1986-C Thorlabs, Inc Piezo mirror mount 1 inch, Picomotor™ piezo clear edge center mount, 2-axis 8821 Newport Spectra-Physics Ltd Lattice laser system part description part number manufacturer Pump laser 532nm, 25W, frequency-doubled laser Millennia EV20S Spectra-Physics (Newport) Ti:Sapph laser 700-790nm, 5.5-8.0W (in this wavelength range, at 25W pump) Matisse 2 TS Sirah Lasertechnik GmbH Water/air chiller Chiller used for Millennia and Ti:Sapph crystal, typ. 20.0◦C, 20psi 6000 Series PolyScience Coolant Cooling mixture P/N 1607-0546L Nalco® CCL100 Nalco Mirror (M45) Laser mirror HR 630-900 nm/0-45◦ s+p, R>99.6% HR 630-900nm Lens-Optics GmbH Wave plate (W2-730) λ/2 0th order wave plate W2Z15-730/W2Z24-730 Lens-Optics GmbH Cube (PBS HP) 1/2 inch polarising beam splitter PBC12-730-HP Lens-Optics GmbH Optical isolator Linos Faraday isolator FI-730-5SV Qioptiq Photonics GmbH & Co KG AOM 110MHz, 700-1100nm, 1×2mm2 active aperture, >2W RF MT110-A1-IR AA opto-electronic Fibre FC/APC to FC/PC connector (custom-made) TS1754306-PM630-HP Thorlabs, Inc Fibre coupler (F8) f = 8.1 mm APC, 630-980nm (fibre in-coupling) 60FC-4-M8-10 Schäfter-Kirchhoff GmbH Fibre coupler (FP5) f = 5.1 mm PC, 630-980nm (fibre out-coupling DYX axes) 60FC-0-M5-10 Schäfter-Kirchhoff GmbH Fibre coupler (FP15) f = 15 mm PC, 600-1050nm (fibre out-coupling T axis) 60FC-0-M15-02 Schäfter-Kirchhoff GmbH Photodiode Si amplified detector, 350-1100nm PDA36A-EC Thorlabs, Inc Cylindrical lens (L -40 cyl) f = −40 mm, N-BK7 B-coated cylindrical lens LK1283L1-B Thorlabs, Inc Cylindrical lens (L 100 cyl) f = 100 mm, N-BK7 B-coated cylindrical lens LJ1567L1-B Thorlabs, Inc Achromat (LA100) 1 inch, f = 100 mm, B-coated achromatic lens (DT axes) AC254-100-B Thorlabs, Inc Achromat (LA125) 1 inch, f = 125 mm, B-coated achromatic lens (YX axes, shared with dipole) AC254-125-B Thorlabs, Inc Achromat (LA400) 1 inch, f = 400 mm, B-coated achromatic lens (T axis) AC254-400-B Thorlabs, Inc Achromats (LA XX) 1 inch, f =XX mm, B-coated achromatic lens (telescopes) AC254-XX-B Thorlabs, Inc Piezo mirror mount 1 inch, Picomotor™ piezo clear edge center mount, 2-axis 8821 Newport Spectra-Physics Ltd. Dichroic mirror (M-DICH) 40mm , HR725-780nm HT1064nm/45◦ (used on YX axes to split off dipole) B-14749/B-14687 S-00929 Laseroptik GmbH (Garbsen) 108 quasicrystalline optical lattices for ultracold atoms Miscellaneous part description part number manufacturer Cameras Andor Zyla Z axis camera, sCMOS 6.5 µm pixel size, 0.5◦ wedged window, 780nm ZYLA-5.5-USB3-9WT Andor Technologies Manta Imaging on D axis, CMOS 5.86 µm pixel size, Gigabit Ethernet AV Manta G-235B POE RCG Allied Vision/Stemmer Imaging IDS uEye Imaging on YXT axes, MOT fluorescence, CMOS 4.5 µm pixel size, USB2.0, global shutter UI-1252LE-M IDS Imaging Development Systems GmbH Beam cam Beam profiling applications, CMOS 4.5 µm pixel size, USB2.0, rolling shutter UI-1542-LE-M IDS Imaging Development Systems GmbH Timing PCIe cards 4× X series multifunction DAQ card, 32 AI, 48 DIO, 4 AO, 1.25MS/s NI PCIe-6353 National Instruments PC Black 4U case, 250GB SSD, Intel Core i5-6500 (3.6GHz), 8GB DDR4, Asus intel LGA1151 various various/World of Computers Ltd Offset lock Photodiode High-bandwidth MSM photodetector G4176-03 Hamamatsu Reference oscillator 7.500GHz, 50Ω SMA, custom made KU LO 7500 RO-587 Kuhne electronic GmbH Amplifier with Bias-T 23dB 1-7000MHz, 50Ω SMA (rubidium offset, used directly after PD) KU LNA BB 0180 A-SMA-BT Kuhne electronic GmbH Amplifier 23dB 1-7000MHz, 50Ω SMA (rubidium offset) KU LNA BB 0180 A-SMA Kuhne electronic GmbH Amplifier 20dB 5-1500MHz, 50Ω SMA (rubidium offset, used after mixer) KU LNA BB 0515 A-SMA Kuhne electronic GmbH MW setup Power amplifier 6400-7200MHz, 40dBm max. output power KU PA 640720-10 A Kuhne electronic GmbH Reference oscillator Programmable oscillator unit MKU LO 8-13 PLL Kuhne electronic GmbH Circulator 4.5-9.0GHz, 3-port circulator CS-20 MCLI MW communications laboratories, Inc Antenna Horn-antenna WG14(WG137), standard gain 14240-20 UAR70 Flann Microwave Ltd Elements Rubidium Rubidium, 99.75% (metals basis) 044214 Alfa Aesar Potassium Potassium-40 enriched 2.98% from potassium chloride, 200mg - - Acknowledgements ‘Experimental science is teamwork.’ I cannot imagine a project for which this statement could be more appropriate than ours. In the past three years I have had the great pleasure and honour to work in a team of clever, supportive, creative, curious, and challenging1 1 ‘But why?’ minds. This team was brought together and led by our supervisor Ulrich Schneider, who managed to convince everyone of us to em- bark on a long journey into the unknown and build a quasicrystal with ultracold atoms. For me, his deep knowledge of physics has been a never-ending source of inspiration and information. I am very grateful for his continuous support and constant availability to answer questions, even in busy weeks, be it in the office or on the phone, at any time of day or night. The first members of the team were Matteo Sbroscia, Oliver Brix, Hendrik von Raven, and myself. Matteo has been the central figure during my PhD. It is not easy to find words that adequately express my feelings towards him, except simply to state that I am extraordinarily grateful for his presence, both as friend and colleague. I could fill pages with anecdotes but still fail to do justice to the role he played in the past three years. Therefore I am going to borrow some phrases from a famous Cavendish alumnus who advised, ‘Work with a teammate who is your intellectual equal’ and, ‘[. . . ] bring together the complementary talents of those involved.’ While I seriously doubt that I have been, or ever will be, Matteo’s intellectual equal, I think we did an O.K. job of profiting from each other’s talents.2 2 his certainly being the more glam- orousOliver was part of the duo, with Hendrik, that had joined us from LMU for a year’s Master’s project. He won the black belt in the disciplines of Autodesk Inventor and Mathematica, developing our magnetic transport system and designing the cooling block. Scientists often dismiss supernatural abilities and black magic for not being compatible with the ‘laws of nature’,3 but Hendrik’s 3 if there is such a thing coding abilities certainly fall into these categories. He single- handedly built our timing system while at the same time plan- ning, developing, and testing one of the most-relied-on compo- nents in the lab, the DDS board. Ed Carter joined within the first year of this project, building a phase-stable interferometer for his Part III project. His arrival marked a new era, him being the first truly English person in the group which consisted, at the time, of four Germans and one 110 quasicrystalline optical lattices for ultracold atoms Italian.4 Ed introduced wit and good humour, which, had it not4 and, rumour has it, two British passports been for Matteo, would have come as a complete novelty to us. Later, having started his PhD in the group, he quickly advanced to playing a vital role in the lab, leading the day-to-day activities and taking the next generation of students under his wing. Among many other accomplishments, Ed perfected the art of writing sequences and lab book entries.5 I am also grateful for him taking5 which sometimes read like crime novels over my position on the grad student committee. In the year 2016 we were joined by Max Melchner (from ETH) and Michael Höse (from LMU) for half-year internships. Max designed our dipole trap laser system and performed useful nu- merical simulations of the adiabatic loading process of the dipole trap. I am grateful for his mathematica scripts, including the trap frequency calculations. Michael, who – by complete coincidence – moved into the same flat share that Oliver had lived in, devel- oped several useful GUIs, completed and characterised our beam shutter design, and implemented the microwave setup. Jr-Chiun Yu started his PhD one year ago and he quickly ab- sorbed a huge amount of information on how to run the machine. His contributions to the experiments are now indispensable and I am confident Jr-Chiun is going to ensure a bright future of the quasicrystal lab. In the course of the experiment, we were joined by a num- ber of hard-working project students, namely Callum Stevens, Joannis Koepsell, Attila Szabó, Robin Croft,6 Shreya Arya, Marius6 who implemented an evolutionary algorithm that we use to optimise RF evaporation sequences Weber, and Elvinas Ribinskas. One year ago, the optical lattice group expanded and qua- sicrystal’s sibling, the Kagome lab, was born. Kagome boasts a fantastic team with Tiffany Harte, Max Melchner, Dan Reed, and Florian Störtz, as well as their former project students Péter Juhász, Greg Chant, and Luca Donini. Now the older sibling is constantly fretting that the younger one will do everything much better and more efficient than herself. Somehow it feels like ‘Team Kagome’ has been around since day one, and I am grateful for all their contributions to the group, regarding both physics and life outside the lab. The optical lattice group is embedded in a wider Cavendish- cold-atom community with a strongly collaborative atmosphere. This is thanks to Zoran Hadzibabic and his team, namely Rob Smith, Christoph Eigen, Jay Man, Jake Glidden, Jinyi Zhang, Panos Christodoulou, Julian Schmitt, Lena Bartha, Timon Hilker, Maciej Galka, Milan Kristajic, and Nathaniel Vilas. I am particu- larly indebted to two former postdocs,7 Nir Navon and7 the lack of postdocs in the quasicrys- tal lab was partly compensated by an abundant supply in Zoran’s group Raphael Lopes, as well as Zoran & Rob, for their continuous sup- port and advice while setting up the apparatus. I owe Zoran a great deal, both professionally and personally, and I am grateful for having been infected by his natural enthusiasm for science. Besides, I believe he is one of the select few8 who truly share my8 together with Nir, and Richard Fletcher limitless appreciation for Trinity’s Full English. Next, I would like to express thanks to Andreas Nunnenkamp, acknowledgements 111 Tijmen Euser, and Mete Atatüre and his team for their support during my time at the Cavendish. I am also grateful for the invaluable contributions to our lab by the Cavendish support staff, including Stephen Topliss, Dan Cross, Marzena Kulma, Graham Cox, Alan Turner, Dave Smith, Huw Prytherch, Nigel Palfrey, Kevin Mott, Gary Large, Chris Summerfield, and others. In addition, I would like to thank Pam Smith and the Rutherford Hub for doing a great job in ad- ministration. Let’s face it, setting up a new laboratory is a mixed bag. It can be an overwhelming, enjoyable, sometimes frustrating, sometimes rewarding, daunting, both stressful and exhilerating9 task, but, 9 The emotions while seeing the first MOT and the first BEC in the lab could be described as a mixture of childish joy and what I imagine zero-g must feel like. more than anything else, it is time-consuming. I am deeply grate- ful to Matteo and our colleagues for sharing so many hours at the lab.10 Despite those long hours, life outside the lab did exist. 10 The total number of hours spent together is comparable to the duration of a typical mission to the ISS. Unfortunately, naming all those wonderful people who made my life worth living (and will hopefully continue to do so) would probably lead to a very long and still incomplete list, so I am not going to attempt it. There must be a few exceptions, though. First and most important, I would like to thank my family, my parents Victoria and Christoph, my brothers Johannes and Emanuel, as well as Emanuel’s wife Mai and their boys Samuel and Adrian. Emanuel’s lot spent half a year in Cambridge, six wonderful months that I will cherish for the rest of my life. Figure A.1: Typical diffraction-limited image of two single bosons in their excited state. 112 quasicrystalline optical lattices for ultracold atoms Then, had Mirjam not saved me, I probably would have de- volved during my PhD to an undernourished eydeet,11 who uses11 a kind of cave man all his energy for pondering physics problems and forgets the an- imate world around him. Thank you, Mirjam, for being there and being you. Another two consistent figures shall not go unnamed here: Daniel Malz and Angela Wittmann. Apart from being brilliant physicists, Daniel and Angela share my passion for music. And while I am happy that all three of us completed our PhDs,12 I will12 within weeks of each other sorely miss their frequent company in the future. The same goes for my dear friends at 4 Portugal St, Pearce A, Burrell’s E, as well as the Barnwell lerds, EBC rowing, CUCO and CUMSSO, cold-atom football, chamber music sessions in the Frazer room, breakfast at Trinity’s, snooker at WT’s, the BA society, Pembroke Choir, CMS lunches, lab breakfasts, Trinity evensongs, astronomy board games, and the dihedral group. I acknowledge financial support from the Cambridge Common- wealth, European & International Trust, as well as the EPSRC. This document was typeset using the tufte-book template1313 tufte-latex.github.io/tufte- latex/ which is licensed under the Apache License, Version 2.0. I would like to thank Michael Feuerbacher, as well as Michael Baake and Uwe Grimm, for providing figures. I am grateful to Anika Bayer for her Adobe Illustrator skills and her advice on formatting. Fi- nally, I would like to thank Ulrich, Matteo, Ed, Angela, Daniel, and Mirjam for proofreading. 114 quasicrystalline optical lattices for ultracold atoms Bibliography 1D. Shechtman, I. Blech, D. Gratias, and J. Cahn, “Metallic Phase with Long-Range Orienta- tional Order and No Translational Symmetry”, Physical review letters 53, 1951 (1984) (cited on pages 11, 26–27, 30). 2D. Levine, and P. J. Steinhardt, “Quasicrystals: a new class of ordered structures”, Physical re- view letters 53, 2477 (1984) (cited on page 11). 3M. Senechal, Quasicrystals and geometry (Cambridge University Press, Cambridge, 1995) (cited on pages 11, 27). 4H.-R. Trebin, Quasicrystals: structure and physical properties (Wiley-VCH, Weinheim; U.K., 2003) (cited on page 11). 5Y. E. Kraus, Y. Lahini, Z. Ringel, M. Verbin, and O. Zilberberg, “Topological States and Adi- abatic Pumping in Quasicrystals”, Physical Review Letters 109, 106402 (2012) (cited on pages 11, 33–34, 102). 6Y. E. Kraus, and O. Zilberberg, “Topological Equivalence between the Fibonacci Quasicrystal and the Harper Model”, Physical Review Letters 109, 116404 (2012) (cited on pages 11–12). 7Y. E. Kraus, Z. Ringel, and O. Zilberberg, “Four-Dimensional Quantum Hall Effect in a Two- Dimensional Quasicrystal”, Physical Review Letters 111, 226401 (2013) (cited on page 11). 8R. Penrose, “The Rôle of Aesthetics in Pure and Applied Mathematical Research”, Bulletin of the Institute of Mathematics and its Applications 10, 266–271 (1974) (cited on pages 11, 28). 9M. Baake, and U. Grimm, editors, Aperiodic Order. Volume 1: A Mathematical Invitation (Cam- bridge University Press, Cambridge, 2013) (cited on pages 11, 15, 23–24, 27–31). 10B. Grünbaum, and G. C. Shephard, Tilings and patterns, A series of books in the mathematical sciences (W.H. Freeman, New York, 1987) (cited on pages 11, 27–28). 11W. Steurer, “Quasicrystals: What do we know? What do we want to know? What can we know?”, Acta Crystallographica Section A Foundations and Advances 74, 1–11 (2018) (cited on pages 11, 15, 18, 24, 30–34). 12E. M. Barber, Aperiodic structures in condensed matter: fundamentals and applications, Series in condensed matter physics (CRC Press, Boca Raton, London, 2009) (cited on pages 11, 15–16, 18–19, 33). 13W. Steurer, and D. Sutter-Widmer, “Photonic and phononic quasicrystals”, Journal of Physics D: Applied Physics 40, R229–R247 (2007) (cited on pages 11, 20, 33). 14L. Guidoni, B. Dépret, A. Di Stefano, and P. Verkerk, “Atomic diffusion in an optical qua- sicrystal with five-fold symmetry”, Physical Review A 60, R4233 (1999) (cited on pages 11, 20, 33). 15L. Guidoni, C. Triché, P. Verkerk, and G. Grynberg, “Quasiperiodic optical lattices”, Physical review letters 79, 3363 (1997) (cited on pages 11, 20, 33). bibliography 115 16J. Mikhael, J. Roth, L. Helden, and C. Bechinger, “Archimedean-like tiling on decagonal qua- sicrystalline surfaces”, Nature 454, 501–504 (2008) (cited on pages 11, 20, 33). 17J. Mikhael, M. Schmiedeberg, S. Rausch, J. Roth, H. Stark, and C. Bechinger, “Proliferation of anomalous symmetries in colloidal monolayers subjected to quasiperiodic light fields”, Pro- ceedings of the National Academy of Sciences 107, 7214–7218 (2010) (cited on pages 11, 20, 33). 18M. Greiner, “Quantum phase transition from a superfluid to a Mott insulator in a gas of ul- tracold atoms”, Nature 415, 39–44 (2002) (cited on pages 12, 36). 19R. P. Feynman, “Simulating physics with computers”, International journal of theoretical physics 21, 467–488 (1982) (cited on page 12). 20I. Bloch, J. Dalibard, and S. Nascimbène, “Quantum simulations with ultracold quantum gases”, Nature Physics 8, 267–276 (2012) (cited on page 12). 21C. Gross, and I. Bloch, “Quantum simulations with ultracold atoms in optical lattices”, Sci- ence 357, 995–1001 (2017) (cited on page 12). 22U. Schneider, L. Hackermüller, J. P. Ronzheimer, S. Will, S. Braun, T. Best, I. Bloch, E. Demler, S. Mandt, D. Rasch, and A. Rosch, “Fermionic transport and out-of-equilibrium dynamics in a homogeneous Hubbard model with ultracold atoms”, Nature Physics 8, 213–218 (2012) (cited on pages 12, 77, 97). 23J.-y. Choi, S. Hild, J. Zeiher, P. Schauss, A. Rubio-Abadal, T. Yefsah, V. Khemani, D. A. Huse, I. Bloch, and C. Gross, “Exploring the many-body localization transition in two dimensions”, Science 352, 1547–1552 (2016) (cited on page 12). 24P. Bordia, H. Lüschen, S. Scherg, S. Gopalakrishnan, M. Knap, U. Schneider, and I. Bloch, “Probing Slow Relaxation and Many-Body Localization in Two-Dimensional Quasiperiodic Systems”, Physical Review X 7, 041047 (2017) (cited on pages 12, 34, 84, 103). 25F. Görg, M. Messer, K. Sandholzer, G. Jotzu, R. Desbuquois, and T. Esslinger, “Enhancement and sign change of magnetic correlations in a driven quantum many-body system”, Nature 553, 481–485 (2018) (cited on pages 12, 82). 26J. Léonard, A. Morales, P. Zupancic, T. Esslinger, and T. Donner, “Supersolid formation in a quantum gas breaking a continuous translational symmetry”, Nature 543, 87–90 (2017) (cited on page 12). 27J.-R. Li, J. Lee, W. Huang, S. Burchesky, B. Shteynas, F. Ç. Top, A. O. Jamison, and W. Ketterle, “A stripe phase with supersolid properties in spin–orbit-coupled Bose–Einstein condensates”, Nature 543, 91–94 (2017) (cited on page 12). 28C. Eigen, J. A. P. Glidden, R. Lopes, E. A. Cornell, R. P. Smith, and Z. Hadzibabic, “Universal Prethermal Dynamics of Bose Gases Quenched to Unitarity”, arXiv:1805.09802 (2018) (cited on page 12). 29A. Eckardt, “Colloquium: Atomic quantum gases in periodically driven optical lattices”, Re- views of Modern Physics 89, 011004 (2017) (cited on pages 12, 82). 30A. Acín, I. Bloch, H. Buhrman, T. Calarco, C. Eichler, J. Eisert, D. Esteve, N. Gisin, S. J. Glaser, F. Jelezko, S. Kuhr, M. Lewenstein, M. F. Riedel, P. O. Schmidt, R. Thew, A. Wallraff, I. Walm- sley, and F. K. Wilhelm, “The European Quantum Technologies Roadmap”, New Journal of Physics 20, arXiv: 1712.03773, 080201 (2018) (cited on page 12). 31T. S. Cubitt, D. Perez-Garcia, and M. M. Wolf, “Undecidability of the spectral gap”, Nature 528, 207–211 (2015) (cited on pages 12, 28, 99). 116 quasicrystalline optical lattices for ultracold atoms 32I. U. of Crystallography, “Quasicrystal”, Online Dictionary of Crystallography (2018) (cited on pages 16–17). 33 IUCr, “International Union of Crystallography Report of the Executive Committee for 1991”, Acta Crystallographica Section A Foundations of Crystallography 48, 922–946 (1991) (cited on page 16). 34I. U. of Crystallography, “Crystal”, Online Dictionary of Crystallography (2018) (cited on page 16). 35R. Lifshitz, “The Definition of Quasicrystals”, arXiv preprint cond-mat/0008152 (2000) (cited on pages 16–18). 36I. U. of Crystallography, “Incommensurate modulated structure”, Online Dictionary of Crys- tallography (2018) (cited on page 16). 37I. U. of Crystallography, “Incommensurate composite crystal”, Online Dictionary of Crystal- lography (2018) (cited on page 17). 38I. U. of Crystallography, “Friedel’s law”, Online Dictionary of Crystallography (2018) (cited on page 17). 39H. A. Bohr, “Die Theorie der fastperiodischen Funktionen”, Ergebnisse der Mathematik und ihrer Grenzgebiete I, 446 (1932) (cited on pages 17–18). 40 common, “Quasicrystal”, Wikipedia (2018) (cited on page 18). 41I. U. of Crystallography, “Quasiperiodicity”, Online Dictionary of Crystallography (2018) (cited on page 18). 42L. Fibonacci, Fibonacci’s Liber abaci: a translation into modern English of Leonardo Pisano’s Book of calculation [translated by Laurence Sigler], Sources and studies in the history of mathematics and physical sciences (Springer, New York, 2002) (cited on page 19). 43C. Janot, Quasicrystals: primer, 2nd ed, Monographs on the Physics & Chemistry of Materials Vol.50 (Oxford University Press, Oxford, 1995) (cited on page 19). 44D. Levine, and P. J. Steinhardt, “Quasicrystals. I. Definition and structure”, Physical Review B 34, 596 (1986) (cited on pages 19, 22). 45E. Schrödinger, What is life? the physical aspect of the living cell, Canto edition., Canto classics (Cambridge University Press, Cambridge, 1992) (cited on page 19). 46P. Ball, “In retrospect: What Is Life?”, Nature 560, 548–550 (2018) (cited on page 19). 47U. Grimm, and M. Schreiber, “Aperiodic Tilings on the Computer”, arXiv:cond-mat/9903010 (1999) (cited on pages 20, 29). 48W. E. Leland, M. S. Taqqu, W. Willinger, and D. V. Wilson, “On the self-similar nature of Eth- ernet traffic (extended version)”, IEEE/ACM Transactions on Networking 2, 1–15 (1994) (cited on page 20). 49S Havlin, S. V. Buldyrev, A. L. Goldberger, R. N. Mantegna, and S. M. Ossadnik, “Fractals in Biology and Medicine”, Chaos, Solitons & Fractals 6, 171–201 (1995) (cited on page 20). 50Y. Baryshev, and P. Teerikorpi, Discovery of Cosmic Fractals (World Scientific Press, 2002) (cited on page 20). 51T. G. Faticoni, The mathematics of infinity: a guide to great ideas, 2nd ed., Pure and applied math- ematics (John Wiley & Sons : Unnumbered) (Wiley, Hoboken, N.J., 2012) (cited on page 22). 52M. Baake, and U. Grimm, “Kinematic diffraction is insufficient to distinguish order from dis- order”, Physical Review B 79, 020203 (2009) (cited on page 24). bibliography 117 53W. Steurer, “Why are quasicrystals quasiperiodic?”, Chemical Society Reviews 41, 6719 (2012) (cited on pages 24, 26). 54W. Steurer, “Twenty years of structure research on quasicrystals. Part I. Pentagonal, octago- nal, decagonal and dodecagonal quasicrystals”, Zeitschrift für Kristallographie - Crystalline Materials 219, 391 (2004) (cited on pages 24, 31–32). 55I. U. of Crystallography, “Aperiodic crystal”, Online Dictionary of Crystallography, Online Dictionary of Crystallography (2018) (cited on page 24). 56M. de Boissieu, “Phonons, phasons and atomic dynamics in quasicrystals”, Chemical Society Reviews 41, 6778 (2012) (cited on page 25). 57W. Scherrer, “Die Einlagerung eines regulären Vielecks in ein Gitter.”, Elemente der Mathe- matik 1, 97–98 (1946) (cited on page 26). 58M. Baake, and U. Grimm, “Mathematical diffraction of aperiodic structures”, Chemical Soci- ety Reviews 41, 6821 (2012) (cited on page 27). 59H. Wang, “Proving theorems by pattern recognition — II”, The Bell System Technical Journal 40, 1–41 (1961) (cited on page 27). 60H. Wang, Games, Logic and Computers. In: Computation, Logic, Philosophy. Mathematics and its Application (China Series), vol 2. (Springer, Dordrecht, 1965) (cited on page 27). 61R. Berger, “The undecidability of the domino problem”, Memoirs of the American Mathemat- ical Society 66, 1–72 (1966) (cited on page 27). 62A. M. Turing, “On Computable Numbers, with an Application to the Entscheidungsprob- lem”, Proceedings of the London Mathematical Society s2-42, 230–265 (1937) (cited on page 27). 63H. Wang, “Notes on a class of tiling problems”, Fundamental Mathematics 82, 295–305 (1975) (cited on page 27). 64E. Winfree, F. Liu, L. A. Wenzler, and N. C. Seeman, “Design and self-assembly of two- dimensional DNA crystals”, Nature 394, 539–544 (1998) (cited on page 28). 65U. Grimm, “Aperiodic tilings on Mathematica”, webpage (2014) (cited on page 29). 66A. Mackay, “Crystallography and the Penrose pattern”, Physica 114A, 609–613 (1982) (cited on pages 30, 33). 67M. Baake, and D. Joseph, “Ideal and defective vertex configurations in the planar octagonal quasilattice”, Physical Review B 42, 8091–8102 (1990) (cited on page 31). 68N. Wang, H. Chen, and K. H. Kuo, “Two-dimensional quasicrystal with eightfold rotational symmetry”, Physical Review Letters 59, 1010–1013 (1987) (cited on page 32). 69A.-P. Tsai, ““Back to the Future”-An Account Discovery of Stable Quasicrystals”, Accounts of Chemical Research 36, 31–38 (2003) (cited on page 32). 70J. Dolinšek, “Electrical and thermal transport properties of icosahedral and decagonal qua- sicrystals”, Chemical Society Reviews 41, 6730 (2012) (cited on page 32). 71D. P. Varn, and J. P. Crutchfield, “Chaotic Crystallography: How the physics of information reveals structural order in materials”, arXiv:1409.5930 (2014) (cited on page 33). 72J. H. E. Cartwright, and A. L. Mackay, “Beyond crystals: the dialectic of materials and infor- mation”, Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engi- neering Sciences 370, 2807–2822 (2012) (cited on page 33). 118 quasicrystalline optical lattices for ultracold atoms 73L. Bindi, P. J. Steinhardt, N. Yao, and P. J. Lu, “Natural Quasicrystals”, Science 324, 1306–1309 (2009) (cited on page 33). 74L. Bindi, C. Lin, C. Ma, and P. J. Steinhardt, “Collisions in outer space produced an icosahe- dral phase in the Khatyrka meteorite never observed previously in the laboratory”, Scientific Reports 6, 38117 (2016) (cited on page 33). 75R. Merlin, K. Bajema, R. Clarke, F. Y. Juang, and P. K. Bhattacharya, “Quasiperiodic GaAs- AlAs Heterostructures”, Physical Review Letters 55, 1768–1770 (1985) (cited on page 33). 76X. Zeng, G. Ungar, Y. Liu, V. Percec, A. E. Dulcey, and J. K. Hobbs, “Supramolecular dendritic liquid quasicrystals”, Nature 428, 157–160 (2004) (cited on page 33). 77K. Hayashida, T. Dotera, A. Takano, and Y. Matsushita, “Polymeric Quasicrystal: Mesoscopic Quasicrystalline Tiling in A B C Star Polymers”, Physical Review Letters 98, 195502 (2007) (cited on page 33). 78D. V. Talapin, E. V. Shevchenko, M. I. Bodnarchuk, X. Ye, J. Chen, and C. B. Murray, “Qua- sicrystalline order in self-assembled binary nanoparticle superlattices”, Nature 461, 964–967 (2009) (cited on page 33). 79S. Fischer, A. Exner, K. Zielske, J. Perlich, S. Deloudi, W. Steurer, P. Lindner, and S. Förster, “Colloidal quasicrystals with 12-fold and 18-fold diffraction symmetry”, Proceedings of the National Academy of Sciences 108, 1810–1814 (2011) (cited on page 33). 80N. K. Efremidis, S. Sears, D. N. Christodoulides, J. W. Fleischer, and M. Segev, “Discrete soli- tons in photorefractive optically induced photonic lattices”, Physical Review E 66 (2002) 10. 1103/PhysRevE.66.046602 (cited on page 33). 81B. Freedman, G. Bartal, M. Segev, R. Lifshitz, D. N. Christodoulides, and J. W. Fleischer, “Wave and defect dynamics in nonlinear photonic quasicrystals”, Nature 440, 1166–1169 (2006) (cited on page 33). 82B. Freedman, R. Lifshitz, J. W. Fleischer, and M. Segev, “Phason dynamics in nonlinear pho- tonic quasicrystals”, Nature Materials 6, 776–781 (2007) (cited on page 33). 83A. Dareau, E. Levy, M. B. Aguilera, R. Bouganne, E. Akkermans, F. Gerbier, and J. Beugnon, “Revealing the Topology of Quasicrystals with a Diffraction Experiment”, Physical Review Letters 119, 215304 (2017) (cited on page 33). 84S. J. Ahn, P. Moon, T.-H. Kim, H.-W. Kim, H.-C. Shin, E. H. Kim, H. W. Cha, S.-J. Kahng, P. Kim, M. Koshino, Y.-W. Son, C.-W. Yang, and J. R. Ahn, “Dirac electrons in a dodecagonal graphene quasicrystal”, Science 361, 782–786 (2018) (cited on page 33). 85N. Macé, A. Jagannathan, and M. Duneau, “Quantum Simulation of a 2d Quasicrystal with Cold Atoms”, Crystals 6, 124 (2016) (cited on pages 33, 84). 86M. Lohse, C. Schweizer, H. M. Price, O. Zilberberg, and I. Bloch, “Exploring 4d quantum Hall physics with a 2d topological charge pump”, Nature 553, 55–58 (2018) (cited on pages 34, 103). 87M. Aidelsburger, “Artificial Gauge Fields with Ultracold Atoms in Optical Lattices”, LMU Munich PhD Thesis (2016) (cited on page 34). 88M. Schreiber, S. S. Hodgman, P. Bordia, H. P. Lüschen, M. H. Fischer, R. Vosk, E. Altman, U. Schneider, and I. Bloch, “Observation of many-body localization of interacting fermions in a quasirandom optical lattice”, Science 349, 842–845 (2015) (cited on pages 34, 103). 89M. Žnidaricˇ, and M. Ljubotina, “Interaction instability of localization in quasiperiodic systems”, Proceedings of the National Academy of Sciences, 201800589 (2018) (cited on pages 34, 103). bibliography 119 90H. P. Lüschen, P. Bordia, S. S. Hodgman, M. Schreiber, S. Sarkar, A. J. Daley, M. H. Fischer, E. Altman, I. Bloch, and U. Schneider, “Signatures of Many-Body Localization in a Controlled Open Quantum System”, Physical Review X 7, 011034 (2017) (cited on pages 34, 103). 91H. P. Lüschen, P. Bordia, S. Scherg, F. Alet, E. Altman, U. Schneider, and I. Bloch, “Obser- vation of Slow Dynamics near the Many-Body Localization Transition in One-Dimensional Quasiperiodic Systems”, Physical Review Letters 119, 260401 (2017) (cited on pages 34, 103). 92V. Khemani, D. N. Sheng, and D. A. Huse, “Two Universality Classes for the Many-Body Lo- calization Transition”, Physical Review Letters 119, 075702 (2017) (cited on pages 34, 103). 93C. Chin, R. Grimm, P. Julienne, and E. Tiesinga, “Feshbach resonances in ultracold gases”, Reviews of Modern Physics 82, 1225–1286 (2010) (cited on pages 35–36, 69–70). 94J. I. Cirac, and P. Zoller, “Quantum Computations with Cold Trapped Ions”, Physical Review Letters 74, 4091–4094 (1995) (cited on page 35). 95J. G. Bohnet, B. C. Sawyer, J. W. Britton, M. L. Wall, A. M. Rey, M. Foss-Feig, and J. J. Bollinger, “Quantum spin dynamics and entanglement generation with hundreds of trapped ions”, Science 352, 1297 (2016) (cited on page 35). 96G. Pupillo, A. Micheli, H. P. Büchler, and P. Zoller, “Condensed Matter Physics with Cold Po- lar Molecules”, arXiv:0805.1896 (2008) (cited on page 36). 97M. H. Anderson, J. R. Ensher, M. R. Matthews, C. E. Wieman, and E. A. Cornell, “Observa- tion of Bose-Einstein Condensation in a Dilute Atomic Vapor”, Science 269, 198–201 (1995) (cited on pages 36, 72). 98K. B. Davis, M.-O. Mewes, M. R. Andrews, N. J. Van Druten, D. S. Durfee, D. M. Kurn, and W. Ketterle, “Bose-Einstein condensation in a gas of sodium atoms”, Physical review letters 75, 3969 (1995) (cited on page 36). 99C. C. Bradley, C. A. Sackett, J. J. Tollett, and R. G. Hulet, “Evidence of Bose-Einstein conden- sation in an atomic gas with attractive interactions”, Physical Review Letters 75, 1687 (1995) (cited on page 36). 100H. Bernien, S. Schwartz, A. Keesling, H. Levine, A. Omran, H. Pichler, S. Choi, A. S. Zibrov, M. Endres, M. Greiner, V. Vuletic´, and M. D. Lukin, “Probing many-body dynamics on a 51-atom quantum simulator”, Nature 551, 579 (2017) (cited on page 36). 101A. G. Truscott, K. E. Strecker, W. I. McAlexander, G. B. Partridge, and R. G. Hulet, “Observa- tion of Fermi Pressure in a Gas of Trapped Atoms”, Science 291, 2570 (2001) (cited on page 36). 102B. DeMarco, and D. S. Jin, “Onset of Fermi degeneracy in a trapped atomic gas”, Science 285, 1703–1706 (1999) (cited on page 36). 103G. Roati, M. Zaccanti, C. D’Errico, J. Catani, M. Modugno, A. Simoni, M. Inguscio, and G. Modugno, “K 39 Bose-Einstein Condensate with Tunable Interactions”, Physical Review Letters 99, 010403 (2007) (cited on pages 36, 42). 104T. Esslinger, “Fermi-Hubbard Physics with Atoms in an Optical Lattice”, Annual Review of Condensed Matter Physics 1, 129–152 (2010) (cited on page 36). 105D. Greif, T. Uehlinger, G. Jotzu, L. Tarruell, and T. Esslinger, “Short-Range Quantum Mag- netism of Ultracold Fermions in an Optical Lattice”, Science 340, 1307–1310 (2013) (cited on page 36). 106A. Mazurenko, C. S. Chiu, G. Ji, M. F. Parsons, M. Kanász-Nagy, R. Schmidt, F. Grusdt, E. Demler, D. Greif, and M. Greiner, “A cold-atom Fermi–Hubbard antiferromagnet”, Nature 545, 462–466 (2017) (cited on page 36). 120 quasicrystalline optical lattices for ultracold atoms 107Y. Takasu, K. Maki, K. Komori, T. Takano, K. Honda, M. Kumakura, T. Yabuzaki, and Y. Taka- hashi, “Spin-Singlet Bose-Einstein Condensation of Two-Electron Atoms”, Physical Review Letters 91, 040404 (2003) (cited on page 36). 108T. Fukuhara, Y. Takasu, M. Kumakura, and Y. Takahashi, “Degenerate Fermi Gases of Ytter- bium”, Physical Review Letters 98, 030401 (2007) (cited on page 36). 109S. Stellmer, M. K. Tey, B. Huang, R. Grimm, and F. Schreck, “Bose-Einstein Condensation of Strontium”, Physical Review Letters 103, 200401 (2009) (cited on page 36). 110Y. N. M. de Escobar, P. G. Mickelson, M. Yan, B. J. DeSalvo, S. B. Nagel, and T. C. Killian, “Bose-Einstein Condensation of Sr 84”, Physical Review Letters 103, 200402 (2009) (cited on page 36). 111S. Stellmer, R. Grimm, and F. Schreck, “Production of quantum-degenerate strontium gases”, Physical Review A 87, 013611 (2013) (cited on page 36). 112M. Lu, N. Q. Burdick, S. H. Youn, and B. L. Lev, “Strongly Dipolar Bose-Einstein Condensate of Dysprosium”, Physical Review Letters 107, 190401 (2011) (cited on page 36). 113M. Lu, N. Q. Burdick, and B. L. Lev, “Quantum Degenerate Dipolar Fermi Gas”, Physical Re- view Letters 108, 215301 (2012) (cited on page 36). 114K. Aikawa, A. Frisch, M. Mark, S. Baier, A. Rietzler, R. Grimm, and F. Ferlaino, “Bose-Einstein Condensation of Erbium”, Physical Review Letters 108, 210401 (2012) (cited on page 36). 115K. Aikawa, A. Frisch, M. Mark, S. Baier, R. Grimm, and F. Ferlaino, “Reaching Fermi Degen- eracy via Universal Dipolar Scattering”, Physical Review Letters 112, 010404 (2014) (cited on page 36). 116M. Greiner, I. Bloch, T. W. Hänsch, and T. Esslinger, “Magnetic transport of trapped cold atoms over a large distance”, Physical Review A 63, 031401 (2001) (cited on page 37). 117M. Greiner, “Magnetischer Transfer von Atomen - ein Weg zur einfachen Bose-Einstein- Kondensation”, Diploma thesis (LMU Munich, Germany, 2000) (cited on pages 37, 53–54). 118T. Rom, “Bosonische und fermionische Quantengase in dreidimensionalen optischen Git- tern”, PhD thesis (LMU Munich and MPQ Garching, Germany, Oct. 2009) (cited on pages 37, 51). 119L. Duca, “Probing topological properties of Bloch bands with ultracold atoms in a honey- comb optical lattice”, PhD thesis (LMU Munich, Germany, July 2015) (cited on pages 37, 49). 120T. Li, “Probing Bloch band geometry with ultracold atoms in optical lattices”, PhD thesis (LMU Munich, Germany, May 2016) (cited on pages 37, 49, 51, 53). 121T. Stöferle, “Exploring Atomic Quantum Gases in Optical Lattices”, PhD thesis (ETH Zurich, Switzerland, 2005) (cited on page 37). 122J. I. Gillen, “The Quantum Gas Microscope”, PhD thesis (Harvard University, Cambridge, Massachusetts, Oct. 2009) (cited on page 37). 123T. Badr, D. B. Ali, J. Seaward, Y. Guo, F. Wiotte, R. Dubessy, H. Perrin, and A. Perrin, “Com- parison of time profiles for the magnetic transport of cold atoms”, arXiv:1809.07096 (2018) (cited on page 37). 124M. Sbroscia, “Towards an optical quasicrystal with ultracold atoms”, First-year report (Uni- versity of Cambridge, Cambridge, Jan. 2016) (cited on pages 38, 48–50, 52, 54). bibliography 121 125W. Ketterle, D. S. Durfee, and D. M. Stamper-Kurn, “Making, probing and understanding Bose-Einstein condensates”, arXiv preprint cond-mat/9904034 5 (1999) (cited on pages 41, 48, 57). 126H. J. Metcalf, and P. Van der Straten, Laser cooling and trapping, Graduate texts in contempo- rary physics (Springer, New York, 1999) (cited on pages 41, 48). 127C. Foot, Atomic physics, Oxford master series in atomic, optical and laser physics (Oxford Uni- versity Press, Oxford, 2005) (cited on pages 41, 43–44, 48, 50, 53, 61). 128C. B. Alcock, V. P. Itkin, and M. K. Horrigan, “Vapour Pressure Equations for the Metallic El- ements: 298–2500k”, Canadian Metallurgical Quarterly 23, 309–313 (1984) (cited on page 41). 129C. Eigen, A. L. Gaunt, A. Suleymanzade, N. Navon, Z. Hadzibabic, and R. P. Smith, “Obser- vation of Weak Collapse in a Bose-Einstein Condensate”, Physical Review X 6, 041058 (2016) (cited on page 42). 130R. J. Fletcher, R. Lopes, J. Man, N. Navon, R. P. Smith, M. W. Zwierlein, and Z. Hadzibabic, “Two- and three-body contacts in the unitary Bose gas”, Science 355, 377–380 (2017) (cited on pages 42, 70). 131R. L. D. Campbell, R. P. Smith, N. Tammuz, S. Beattie, S. Moulder, and Z. Hadzibabic, “Ef- ficient production of large K 39 Bose-Einstein condensates”, Physical Review A 82, 063611 (2010) (cited on pages 42, 65–66). 132J. P. Ronzheimer, M. Schreiber, S. Braun, S. S. Hodgman, S. Langer, I. P. McCulloch, F. Heidrich-Meisner, I. Bloch, and U. Schneider, “Expansion Dynamics of Interacting Bosons in Homogeneous Lattices in One and Two Dimensions”, Physical Review Letters 110, 205301 (2013) (cited on page 42). 133G. Salomon, L. Fouché, S. Lepoutre, A. Aspect, and T. Bourdel, “All-optical cooling of K 39 to Bose-Einstein condensation”, Physical Review A 90, 033405 (2014) (cited on pages 42, 56). 134L. Wacker, N. B. Jørgensen, D. Birkmose, R. Horchani, W. Ertmer, C. Klempt, N. Winter, J. Sherson, and J. J. Arlt, “Tunable dual-species Bose-Einstein condensates of K 39 and Rb 87”, Physical Review A 92, 053602 (2015) (cited on page 42). 135M. Gröbner, P. Weinmann, F. Meinert, K. Lauber, E. Kirilov, and H.-C. Nägerl, “A new quan- tum gas apparatus for ultracold mixtures of K and Cs and KCs ground-state molecules”, Journal of Modern Optics 63, 1829–1839 (2016) (cited on pages 42, 51, 53). 136C. R. Cabrera, L. Tanzi, J. Sanz, B. Naylor, P. Thomas, P. Cheiney, and L. Tarruell, “Quantum liquid droplets in a mixture of Bose-Einstein condensates”, Science 359, 301–304 (2018) (cited on page 42). 137D. Nath, R. K. Easwaran, G. Rajalakshmi, and C. S. Unnikrishnan, “Quantum-interference- enhanced deep sub-Doppler cooling of 39 K atoms in gray molasses”, Physical Review A 88, 053407 (2013) (cited on page 42). 138G. Salomon, L. Fouché, P. Wang, A. Aspect, P. Bouyer, and T. Bourdel, “Gray-molasses cool- ing of 39 K to a high phase-space density”, EPL (Europhysics Letters) 104, 63002 (2013) (cited on page 42). 139D. A. Steck, Rubidium 87 D line data (2001) (cited on pages 43, 45). 140T. G. Tiecke, “Properties of potassium”, University of Amsterdam, The Netherlands, Thesis (2010) (cited on pages 43, 45, 59). 141C. J. Bordé, “Spectroscopie d’absorption saturée de diverses molécules au moyen des lasers à gaz carbonique et à protoxyde d’azote”, CR Acad. Sc. Paris 271, 371–374 (1970) (cited on page 44). 122 quasicrystalline optical lattices for ultracold atoms 142P. Smith, and T. Hänsch, “Cross-Relaxation Effects in the Saturation of the 6328-A Neon-Laser Line”, Physical review letters 26, 740–743 (1971) (cited on page 44). 143T. Hänsch, M. Levenson, and A. Schawlow, “Complete Hyperfine Structure of a Molecular Io- dine Line”, Phys. Rev. Lett. 26, 946–949 (1971) (cited on page 44). 144A. L. Schawlow, “Spectroscopy in a new light”, Reviews of Modern Physics 54, 697 (1982) (cited on page 44). 145E. D. Black, “An introduction to Pound–Drever–Hall laser frequency stabilization”, American Journal of Physics 69, 79–87 (2001) (cited on page 44). 146U. Schünemann, H. Engler, R. Grimm, M. Weidemüller, and M. Zielonkowski, “Simple scheme for tunable frequency offset locking of two lasers”, Review of Scientific Instruments 70, 242–243 (1999) (cited on page 46). 147T. W. Hänsch, and A. L. Schawlow, “Cooling of gases by laser radiation”, Optics Communica- tions 13, 68–69 (1975) (cited on pages 48–49). 148K. Dieckmann, R. Spreeuw, M. Weidemüller, and J. Walraven, “Two-dimensional magneto- optical trap as a source of slow atoms”, Phys. Rev. A 58, 3891 (1998) (cited on page 49). 149J. Schoser, A. Batär, R. Löw, V. Schweikhard, A. Grabowski, Y. B. Ovchinnikov, and T. Pfau, “Intense source of cold Rb atoms from a pure two-dimensional magneto-optical trap”, Physi- cal Review A 66, 023410 (2002) (cited on page 49). 150S. Chaudhuri, S. Roy, and C. S. Unnikrishnan, “Realization of an intense cold Rb atomic beam based on a two-dimensional magneto-optical trap: Experiments and comparison with simulations”, Physical Review A 74, 023406 (2006) (cited on page 49). 151T. G. Tiecke, S. D. Gensemer, A. Ludewig, and J. T. M. Walraven, “High-flux two-dimensional magneto-optical-trap source for cold lithium atoms”, Physical Review A 80, 013409 (2009) (cited on page 49). 152V. Gokhroo, G Rajalakshmi, R. K. Easwaran, and C. S. Unnikrishnan, “Sub-Doppler deep- cooled bosonic and fermionic isotopes of potassium in a compact 2d + –3d MOT set-up”, Journal of Physics B: Atomic, Molecular and Optical Physics 44, 115307 (2011) (cited on page 49). 153T. Y. Li, “An Apparatus for Probing Fermions in Quasi-Two-Dimensional Geometry”, Mas- ter’s thesis (LMU Munich, 2011) (cited on page 49). 154M. Boll, “A 2d+ -3d MOT system for producing laser-cooled K-Rb mixtures”, Diploma the- sis (MPQ Garching and Johannes Gutenberg Universität Mainz, Germany, Nov. 2011) (cited on page 49). 155L. G. Marcassa, G. D. Telles, S. R. Muniz, and V. S. Bagnato, “Collisional losses in a K-Rb cold mixture”, Physical Review A 63, 013413 (2000) (cited on page 50). 156J. Goldwin, S. B. Papp, B. DeMarco, and D. S. Jin, “Two-species magneto-optical trap with 40 K and 87 Rb”, Physical Review A 65, 021402 (2002) (cited on page 50). 157W. Ketterle, K. B. Davis, M. A. Joffe, A. Martin, and D. E. Pritchard, “High densities of cold atoms in a dark spontaneous-force optical trap”, Physical review letters 70, 2253 (1993) (cited on page 51). 158C. S. Adams, H. J. Lee, N. Davidson, M. Kasevich, and S. Chu, “Evaporative Cooling in a Crossed Dipole Trap”, Physical Review Letters 74, 3577–3580 (1995) (cited on page 51). 159W. Petrich, M. H. Anderson, J. R. Ensher, and E. A. Cornell, “Behavior of atoms in a com- pressed magneto-optical trap”, Journal of the Optical Society of America B 11, 1332 (1994) (cited on page 51). bibliography 123 160N. Tammuz, “Thermodynamics of ultracold 39k atomic Bose gases with tuneable interac- tions”, PhD thesis (University of Cambridge, Cambridge, 2011) (cited on pages 51, 53, 72). 161T. Hilker, “Laser Cooling of Bosonic and Fermionic Lithium”, Diploma thesis (MPQ Garching and Technische Universität München, Germany, Sept. 2012) (cited on page 51). 162A. D. Tranter, H. J. Slatyer, M. R. Hush, A. C. Leung, J. L. Everett, K. V. Paul, P. Vernaz-Gris, P. K. Lam, B. C. Buchler, and G. T. Campbell, “Multiparameter optimisation of a magneto- optical trap using deep learning”, arXiv:1805.00654 (2018) (cited on page 51). 163P. J. Ungar, D. S. Weiss, E. Riis, and S. Chu, “Optical molasses and multilevel atoms: theory”, Journal of the Optical Society of America B 6, 2058 (1989) (cited on page 51). 164P. D. Lett, W. D. Phillips, S. L. Rolston, C. E. Tanner, R. N. Watts, and C. I. Westbrook, “Opti- cal molasses”, J. Opt. Soc. Am. B 6, 2084–2107 (1989) (cited on page 51). 165J. Dalibard, and C. Cohen-Tannoudji, “Laser cooling below the Doppler limit by polarization gradients: simple theoretical models”, Journal of the Optical Society of America B 6, 2023 (1989) (cited on pages 51, 61). 166A. L. Migdall, J. V. Prodan, W. D. Phillips, T. H. Bergeman, and H. J. Metcalf, “First Observa- tion of Magnetically Trapped Neutral Atoms”, Physical Review Letters 54, 2596–2599 (1985) (cited on page 53). 167O. Brix, “Towards Bose-Fermi Mixtures in Novel Optical Lattices”, Master’s thesis (LMU Mu- nich, Germany, 2016) (cited on pages 53–55). 168A. L. Gaunt, T. F. Schmidutz, I. Gotlibovych, R. P. Smith, and Z. Hadzibabic, “Bose-Einstein Condensation of Atoms in a Uniform Potential”, Physical Review Letters 110, 200406 (2013) (cited on page 54). 169H. F. Hess, “Evaporative cooling of magnetically trapped and compressed spin-polarized hy- drogen”, Physical Review B 34, 3476–3479 (1986) (cited on page 56). 170W. Petrich, M. H. Anderson, J. R. Ensher, and E. A. Cornell, “Stable, Tightly Confining Mag- netic Trap for Evaporative Cooling of Neutral Atoms”, Physical Review Letters 74, 3352–3355 (1995) (cited on page 56). 171K. B. Davis, M.-O. Mewes, M. A. Joffe, M. R. Andrews, and W. Ketterle, “Evaporative Cooling of Sodium Atoms”, Physical Review Letters 74, 5202–5205 (1995) (cited on page 56). 172J. Fortágh, and C. Zimmermann, “Magnetic microtraps for ultracold atoms”, Reviews of Modern Physics 79, 235–289 (2007) (cited on page 56). 173H. von Raven, “Direct digital synthesis and integrated experiment control for a novel lattice setup”, Master’s thesis (LMU Munich and University of Cambridge, Cambridge, Oct. 2016) (cited on page 57). 174R. J. Fletcher, M. Robert-de Saint-Vincent, J. Man, N. Navon, R. P. Smith, K. G. H. Viebahn, and Z. Hadzibabic, “Connecting Berezinskii-Kosterlitz-Thouless and BEC Phase Transitions by Tuning Interactions in a Trapped Gas”, Physical Review Letters 114, 255302 (2015) (cited on page 58). 175R. Grimm, M. Weidemüller, and Y. B. Ovchinnikov, “Optical dipole traps for neutral atoms”, arXiv:physics/9902072 (1999) (cited on pages 61–63). 176J. Dalibard, and C. Cohen-Tannoudji, “Dressed-atom approach to atomic motion in laser light: the dipole force revisited”, Journal of the Optical Society of America B 2, 1707 (1985) (cited on page 61). 124 quasicrystalline optical lattices for ultracold atoms 177W. Ketterle, and N. V. Druten, “Evaporative Cooling of Trapped Atoms”, Advances In Atomic, Molecular, and Optical Physics, Vol. 37, edited by B. Bederson, and H. Walther, (Academic Press, Jan. 1996), pages 181–236 (cited on page 66). 178C. J. Myatt, E. A. Burt, R. W. Ghrist, E. A. Cornell, and C. E. Wieman, “Production of Two Overlapping Bose-Einstein Condensates by Sympathetic Cooling”, Physical Review Letters 78, 586–589 (1997) (cited on page 66). 179G. Modugno, G. Ferrari, G. Roati, R. J. Brecha, A. Simoni, and M. Inguscio, “Bose-Einstein Condensation of Potassium Atoms by Sympathetic Cooling”, Science 294, 1320 (2001) (cited on page 66). 180M. Melchner von Dydiowa, “An optical dipole trap for ultracold atomic gases”, Master’s the- sis (ETH Zurich and University of Cambridge, Cambridge, UK, Apr. 2017) (cited on page 66). 181 Hamamatsu, “Si photodiodes”, Handbook on phododiodes, url (2018) (cited on page 66). 182Z. Hadzibabic, S. Gupta, C. A. Stan, C. H. Schunck, M. W. Zwierlein, K. Dieckmann, and W. Ketterle, “Fiftyfold Improvement in the Number of Quantum Degenerate Fermionic Atoms”, Phys. Rev. Lett. 91, 160401 (2003) (cited on page 68). 183I. Bloch, W. Zwerger, and J. Dalibard, “Many-body physics with ultracold gases”, Reviews of Modern Physics 80, 885–964 (2008) (cited on pages 69–70). 184L. D. Landau, and E. M. Lifshitz, Quantum mechanics: non-relativistic theory, 3rd ed., rev. and enl., repr. 1991 with corrections, Course of theoretical physics v. 3 (Butterworth-Heinemann, Oxford, Boston, 1991) (cited on page 69). 185R. J. Fletcher, J. Man, R. Lopes, P. Christodoulou, J. Schmitt, M. Sohmen, N. Navon, R. P. Smith, and Z. Hadzibabic, “Elliptic flow in a strongly interacting normal Bose gas”, Physical Review A 98, 011601 (2018) (cited on page 70). 186L. P. Pitaevskii, and S. Stringari, Bose-Einstein condensation, International series of monographs on physics 116 (Clarendon Press, Oxford, 2003) (cited on pages 70–72). 187N. Tammuz, R. P. Smith, R. L. D. Campbell, S. Beattie, S. Moulder, J. Dalibard, and Z. Hadz- ibabic, “Can a Bose Gas Be Saturated?”, Physical Review Letters 106, 230401 (2011) (cited on page 72). 188N. W. Ashcroft, and N. D. Mermin, Solid state physics (Brooks/Cole, Thomson Learning, Lon- don, 1976) (cited on pages 73, 75). 189M. Greiner, “Ultracold quantum gases in three-dimensional optical lattice potentials”, LMU Munich, url (2003) (cited on page 73). 190C. Weitenberg, M. Endres, J. F. Sherson, M. Cheneau, P. Schauß, T. Fukuhara, I. Bloch, and S. Kuhr, “Single-spin addressing in an atomic Mott insulator”, Nature 471, 319–324 (2011) (cited on pages 77, 97). 191P. M. Preiss, M. Ruichao, M. E. Tai, A. Lukin, M. Rispoli, P. Zupancic, Y. Lahini, R. Islam, and M. Greiner, “Strongly correlated quantum walks in optical lattices”, Science 347, 1229–1233 (2015) (cited on pages 77, 97). 192S. Fölling, S. Trotzky, P. Cheinet, M. Feld, R. Saers, A. Widera, T. Müller, and I. Bloch, “Direct observation of second-order atom tunnelling”, Nature 448, 1029–1032 (2007) (cited on page 82). 193N. Goldman, and J. Dalibard, “Periodically Driven Quantum Systems: Effective Hamiltonians and Engineered Gauge Fields”, Physical Review X 4, 031027 (2014) (cited on page 82). bibliography 125 194J. Struck, C. Olschlager, R. Le Targat, P. Soltan-Panahi, A. Eckardt, M. Lewenstein, P. Wind- passinger, and K. Sengstock, “Quantum Simulation of Frustrated Classical Magnetism in Triangular Optical Lattices”, Science 333, 996–999 (2011) (cited on page 82). 195R. Desbuquois, M. Messer, F. Görg, K. Sandholzer, G. Jotzu, and T. Esslinger, “Controlling the Floquet state population and observing micromotion in a periodically driven two-body quantum system”, Physical Review A 96, 053602 (2017) (cited on page 82). 196M. Reitter, J. Näger, K. Wintersperger, C. Sträter, I. Bloch, A. Eckardt, and U. Schneider, “In- teraction Dependent Heating and Atom Loss in a Periodically Driven Optical Lattice”, Physi- cal Review Letters 119, 200402 (2017) (cited on page 82). 197J. Näger, K. Wintersperger, M. Bukov, S. Lellouch, E. Demler, U. Schneider, I. Bloch, N. Gold- man, and M. Aidelsburger, “Parametric instabilities of interacting bosons in periodically- driven 1d optical lattices”, arXiv:1808.07462 (2018) (cited on page 82). 198A. Jagannathan, and M. Duneau, “An eightfold optical quasicrystal with cold atoms”, EPL (Europhysics Letters) 104, 66003 (2013) (cited on page 84). 199A. Jagannathan, and M. Duneau, “The eight-fold way for optical quasicrystals”, The Euro- pean Physical Journal B 87 (2014) 10.1140/epjb/e2014-50164-7 (cited on page 84). 200A. Jagannathan, and M. Duneau, “Tight-Binding Models in a Quasiperiodic Optical Lattice”, Acta Physica Polonica A 126, 490–492 (2014) (cited on page 84). 201J. J. Snyder, “Paraxial ray analysis of a cat’s-eye retroreflector”, Applied optics 14, 1825–1828 (1975) (cited on page 88). 202P. L. Gould, G. A. Ruff, and D. E. Pritchard, “Diffraction of atoms by light: The near-resonant Kapitza-Dirac effect”, Physical review letters 56, 827 (1986) (cited on pages 89, 91). 203P. L. Kapitza, and P. A. M. Dirac, “The reflection of electrons from standing light waves”, Mathematical Proceedings of the Cambridge Philosophical Society 29, 297 (1933) (cited on page 91). 204N. J. A. Sloane, and S. Plouffe, The encyclopedia of integer sequences (San Diego, Calif., 1995) (cited on page 93). 205B. Gadway, “Atom-optics approach to studying transport phenomena”, Physical Review A 92, 043606 (2015) (cited on pages 94, 97). 206S. Gupta, E. Leanhardt, A. Cronin, and D. E. Pritchard, “Coherent manipulation of atoms with standing light waves”, C. R. Acad. Sci. Paris t. 2, 479–495 (2001) (cited on page 94). 207A. P. Chikkatur, A. Görlitz, D. M. Stamper-Kurn, S. Inouye, S. Gupta, and W. Ketterle, “Sup- pression and enhancement of impurity scattering in a Bose-Einstein condensate”, Physical re- view letters 85, 483 (2000) (cited on page 95). 208M. Greiner, I. Bloch, O. Mandel, T. Hänsch, and T. Esslinger, “Exploring Phase Coherence in a 2d Lattice of Bose-Einstein Condensates”, Physical Review Letters 87, 160405 (2001) (cited on page 95). 209J. H. Denschlag, J. E. Simsarian, H Häffner, C McKenzie, A Browaeys, D Cho, K Helmerson, S. L. Rolston, and W. D. Phillips, “A Bose-Einstein condensate in an optical lattice”, Journal of Physics B: Atomic, Molecular and Optical Physics 35, 3095–3110 (2002) (cited on page 102). 210M. Atala, M. Aidelsburger, J. T. Barreiro, D. Abanin, T. Kitagawa, E. Demler, and I. Bloch, “Direct measurement of the Zak phase in topological Bloch bands”, Nature Physics 9, 795– 800 (2013) (cited on page 102). 126 quasicrystalline optical lattices for ultracold atoms 211G. Jotzu, M. Messer, R. Desbuquois, M. Lebrat, T. Uehlinger, D. Greif, and T. Esslinger, “Ex- perimental realization of the topological Haldane model with ultracold fermions”, Nature 515, 237–240 (2014) (cited on page 102). 212L. Duca, T. Li, M. Reitter, I. Bloch, M. Schleier-Smith, and U. Schneider, “An Aharonov-Bohm interferometer for determining Bloch band topology”, Science, 1259052 (2014) (cited on page 102). 213C. D’Errico, E. Lucioni, L. Tanzi, L. Gori, G. Roux, I. P. McCulloch, T. Giamarchi, M. Inguscio, and G. Modugno, “Observation of a Disordered Bosonic Insulator from Weak to Strong In- teractions”, Physical Review Letters 113 (2014) 10.1103/PhysRevLett.113.095301 (cited on page 103). 214A. M. Rey, I. I. Satija, and C. W. Clark, “Quantum coherence of hard-core bosons: Extended, glassy, and Mott phases”, Physical Review A 73 (2006) 10.1103/PhysRevA.73.063610 (cited on page 103).