ABSTRACT Title of Dissertation: JOSEPHSON EFFECTS IN THE IRON-BASED SUPERCONDUCTOR FETE1−XSEX Samuel Deitemyer Doctor of Philosophy, 2025 Dissertation Directed by: Professor Steve Rolston Department of Physics Professor Steven M. Anlage Department of Physics The iron-based superconductor FeTe1−xSex has emerged as a promising platform for combining superconductivity and topology in a single system, for the realization of topolog- ical quantum computing. Besides this, FeTe1−xSex hosts rich physical phenomena such as S± superconductivity, Majorana bound states, and higher-order topological superconduc- tivity, among others. Despite the interest in superconducting devices based on FeTe1−xSex, there have been relatively few demonstrations of Josephson junctions in FeTe1−xSex-based systems. In this dissertation we measured Josephson effects in a FeTe1−xSex-based device and found three signatures of unconventional Josephson junction behavior. This first signature was the existence of two distinct Josephson diffraction patterns under applied RF irradi- ation, which likely arises from flux flow and a phase slip line in FeTe1−xSex. The second signature was the emergence of sudden jumps in the DC current at which Shapiro steps arise, as a function of applied RF power. This was measured by mapping dV dI vs DC current and RF power. We provide two potential explanations for this phenomenon based on non- equilibrium superconductivity. The third signature is a minimum critical current at zero magnetic field when RF irradiation is present which resembles the π-Josephson junctions formed as a consequence of multiband superconductivity. JOSEPHSON EFFECTS IN THE IRON-BASED SUPERCONDUCTOR FETE1−XSEX by Samuel Deitemyer Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2025 Advisory Committee: Professor Steve Rolston, Chair Professor Steven M. Anlage, Chair Professor Alicia J. Kollár Professor Kasra Sardashti Professor Christopher Jarzynski, Dean’s Representative © Copyright by Samuel Deitemyer 2025 Dedication To my wife Tienne, your love and support truly made this possible. I couldn’t have done it without your constant encouragement and positivity. ii Acknowledgements I am deeply indebted to so many people who have helped me along the way out of the goodness of their hearts. Firstly, I would like to thank Professor Rolston and Professor Anlage for their support and guidance throughout my graduation. I am sincerely grateful for all the effort and time they have spent helping me. I would also like to thank Professor Anlage for sharing his expertise in superconduc- tivity inside and outside the classroom. Your interest in the topics always came through and made learning fun. I would like to thank Dr. Jimmy Williams for introducing me to research as an undergraduate student and for mentoring me as a graduate student. I learned so much from your mentorship, and for that, I am truly grateful. I would like to extend my sincere thanks to Professor Kollár. You have been so generous with your time, and I truly can not express how appreciative I am of your support and kindness. On top of this, your coaching helped me land a job after graduation. I would also like to thank Professor Cohen for his support and open-door policy. Unfortunately, I abused that policy, but it was incredibly valuable to me to have chats when I needed to. I would like to thank Rodney Snyder for teaching me nearly everything I know about measurements. You have been an excellent mentor and friend to so many. I would also like to thank Professor Gong for supporting me during my time in his group and for developing my written communication skills significantly. There are so many others I want to thank, such as Doug Benson, John Abrahams, Tom Loughran, Mark Lecates, Jon Hummel, Karen Gaskell, and many others, for their technical support. Additionally, I would like to thank all of my friends throughout my PhD who made the experience enjoyable. iii Contents Dedication ii Acknowledgements iii Table of Contents iv List of Figures vi 1 Introduction 1 1.1 Thesis Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 2 Superconductivity 3 2.1 Introduction to Superconducting Phenomena . . . . . . . . . . . . . . . . . 3 3 Review of FeTeSe 9 3.1 Crystal Structure and Parent Compounds . . . . . . . . . . . . . . . . . . 9 3.2 Superconductivity in FeTeSe . . . . . . . . . . . . . . . . . . . . . . . . . . 12 3.2.1 Notable Attempts to Increasing Tc . . . . . . . . . . . . . . . . . . 12 3.2.2 Fermi Surface/Pairing Mechanism/Pairing Symmetry . . . . . . . . 16 3.3 Topology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 3.3.1 Topology In FeTeSe . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 3.4 Topological Superconductivity . . . . . . . . . . . . . . . . . . . . . . . . . 23 3.4.1 Topological Superconductivity in FeTeSe . . . . . . . . . . . . . . . 25 3.4.2 Higher Order Topological Superconductivity in FeTeSe . . . . . . . 28 3.5 Josephson Effects in FeTeSe-Based Devices . . . . . . . . . . . . . . . . . . 29 4 Josephson Junctions 32 4.1 Josephson Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 4.1.1 Derivation of Josephson Equations . . . . . . . . . . . . . . . . . . 32 4.1.2 Shapiro Steps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 4.1.3 Magnetic Field Dependence of Josephson Junctions . . . . . . . . . 39 5 FeTeSe-Al Josephson Junction 44 5.1 Fabrication and Measurements . . . . . . . . . . . . . . . . . . . . . . . . . 44 5.1.1 Fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 5.1.2 Measurement Procedures . . . . . . . . . . . . . . . . . . . . . . . . 46 iv 5.2 Inner and outer junction effect . . . . . . . . . . . . . . . . . . . . . . . . . 50 5.2.1 Discussion of Inner/Outer Junction . . . . . . . . . . . . . . . . . . 57 5.3 Jumps in Shapiro Mapping . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 5.3.1 Discussion of Jumps in Shapiro Mapping . . . . . . . . . . . . . . . 71 5.4 Magnetic Field dependence . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 5.4.1 Fraunhofer Measurements of Inner Junction and the Outer Junction 76 5.4.2 Magnetic Field and RF Irradiation . . . . . . . . . . . . . . . . . . 78 5.4.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 5.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 6 Conclusion and Outlook 87 A Appendix 89 A.1 Lock-in, DC, RF, and B-Field in a Dilution Refrigerator . . . . . . . . . . 89 A.1.1 Lock-in Measurements . . . . . . . . . . . . . . . . . . . . . . . . . 89 A.1.2 Current-Bias Measurements . . . . . . . . . . . . . . . . . . . . . . 90 A.1.3 Measurements in a dilution refrigerator . . . . . . . . . . . . . . . . 92 A.1.4 Magnetic Field and Superconducting Magnets . . . . . . . . . . . . 97 A.2 Failed Devices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97 A.3 Wire Bonding Tips and Tricks . . . . . . . . . . . . . . . . . . . . . . . . . 99 A.4 Data dump . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 A.5 Depositions System SOPs . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 References 139 v List of Figures 2.1 Resistance vs Temperature curve for a superconductor. . . . . . . . . . . . 4 2.2 Diagram of the Meissner effect and the London penetration depth. . . . . 5 2.3 Diagram of relationship between critical temperature and critical field. . . 5 2.4 Diagram showing the critical current of a superconductor. . . . . . . . . . 6 2.5 Diagram showing the electron density of states around the Fermi energy in a superconductor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 3.1 a,b) The 2D tetragonal crystal structure of FeTe1−xSex. The larger spacing between layers bonded by weak vdW forces and smaller spacing in the layer due to strong covalent bonds is visible in (a). The tetragonal structure is most clear from the square structure visible along the c-axis of the material (b). Adapted from Ref. [4]. . . . . . . . . . . . . . . . . . . . . . . . . . . 11 3.2 Doping dependence of FeTe1−xSex, showing the antiferromagnetic state when Se = 0.0 and the superconducting state from Se = 0.1 to Se = 1.0. The references in the figure are; ‘Bulk-Ref [4]’: Ref. [15], ‘Bulk-Ref [11]’: Ref. [16], ‘Crystal-Ref [3]’: Ref. [17]. Adapted from Ref. [18]. . . . . . . . . . 12 3.3 Resistance vs temperature that we measured for a FeTe0.55Se0.45 flake, show- ing a Tc around 14 K. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 3.4 Tc enhancement by removing excess Fe atoms. The positions of excess Fe atoms are shown to be at an interstitial location within the crystal. The magnetic susceptibility and resistivity are shown as a function of temperature for the higher Fe concentration (SC2, 11% excess Fe) and for the lower Fe concentration (SC1, 3% excess Fe). The lower Fe concentration improves the diamagnetic response and the Tc of the crystal. Adapted from Ref. [14]. 15 3.5 The effect of oxygen annealing on removing excess Fe and improving the Tc. The diamagnetic response, critical temperature, and critical current are all enhanced due to the effective removal of excess Fe by oxygen annealing. Adapted from Ref. [19] . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 3.6 A phase diagram showing the pressure and temperature-dependent behavior of FeSe. Specifically, the pressure-dependent enhancement of Tc can be seen until 10 GPa. Afterward, this FeSe begins to undergo a transition to the hexagonal phase. Adapted from Ref. [23] . . . . . . . . . . . . . . . . . . 17 3.7 The calculated band structure (a) and Fermi surface (b) of LaFeAsO1−xFx. The key properties here are two electron cylinders around the M point and two hole cylinders around the Γ point. Adapted from Ref. [33] . . . . . . 18 vi 3.8 Quasiparticle interference probes of the unconventional superconductivity of FeTeSe. a) The tunneling current as a function of position. b) The tunneling current as a function of sample bias, showing the reduced tunneling current due to the quasiparticle gap. c) The ratio of the conductance at positive and negative voltage bias as a function of position. d) The Fourier transform of (c) showing the quasiparticle intensity vs the wavevector. e) The quasiparticle interference under a magnetic field of 10 T showing the enhanced or suppressed scattering based on the phase difference between the pockets, confirming the S± pairing symmetry. Adapted from Ref. [37] . . 19 3.9 Illustration of the band inversion effect giving rise to topological surface states. The strong spin-orbit coupling opens a bandgap and causes twisting of the bands (i.e., band inversion). At the interface with materials without band inversion, the bands must be reverted to an untwisted non-inverted state and must cross to do so. This closes the bandgap locally, giving rise to topologically protected surface states. . . . . . . . . . . . . . . . . . . . . 21 3.10 First principles calculation of the band structure of FeTeSe, revealing band inversion and a non-zero topological invariant. a) The band structure of FeSe. b) the band structure of FeTeSe, neglecting the spin-orbit interaction. The main difference from FeSe (a) is that the Γ2 − band (bolded in red) is pushed down in energy and crosses the Fermi energy along the Γ-Z direction. c) The band structure of FeTeSe after adding the spin-orbit interaction. d) A zoom-in on the Γ-Z direction showing the avoided crossing which gives rise to the band inversion and topology in FeTeSe. Adapted from Ref. [44] 23 3.11 Illustration of braiding operations on Majorana bound states. The Majorana wavefunction starts in the ground states defined by pairs of Majorana par- ticles. By exchanging the particles’ positions, the wavefunction will now be in a superposition of the ground and excited states. For successive braiding operations, the order of exchanges affects the final state. Adapted from Ref. [43] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 3.12 Illustration of a vortex in a superconductor. The order parameter (ψ) is re- duced inside a radius equal to the coherence length (ξ), eventually becoming zero at the core of the vortex. The magnetic field intensity (H) decays over the penetration depth of the superconductor (λ). . . . . . . . . . . . . . . 26 3.13 A direct probe of Majorana bound states at the core of a superconducting vortex using STM. A peak in the conductance can be seen at the core of the vortex using STM (a). This peak in the conductance can be shown to be centered on zero energy, as well as at the core of the vortex (b). An insensitivity to magnetic field was also demonstrated, ruling out CdGM states and Kondo resonances (c,d). Adapted from Ref. [51] . . . . . . . . 27 3.14 Visualization of the S± superconductivity on the surfaces of FeTeSe, and the resulting helical hinge zero modes. The closing of the superconducting gap at an angle between the two surfaces of FeTeSe and the topological surface states give rise to the higher order topology in this material. Adapted from Ref. [58]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 vii 3.15 a-c) A transport study on FeTeSe (a) revealing a zero bias conductance peak when the electrode is in contact with the edge of the sample (c), which is absent when just contacting the top surface (b), confirming the presence of helical hinge zero modes. The small peak in (b) is likely due to tunneling through the hBN into the helical hinge zero mode, which exists at zero energy. Adapted from Ref. [58]. . . . . . . . . . . . . . . . . . . . . . . . 29 3.16 a) A constriction Josephson junction fabricated from FeTeSe crystals. b) Shapiro steps were reliable in this system, confirming the realization of a Josephson junction in this device. Adapted from Ref. [59]. . . . . . . . . . 30 3.17 a) A FeTeSe-FeTeSe homojunction where the vdW gap between he crystals allows for tunneling between the FeTeSe flakes. b) The complicated mag- netic field dependence of the critical current, showing asymmetry in B and a minimum close to zero. c) The description for the unusual magnetic field dependence is a combination between 0-junction behavior (the typical case) and π-junction behavior (arising in Josephson junctions with multiband su- perconductors or magnetic insulating layers). Adapted from Ref. [61]. . . 31 4.1 Schematic of a Josephson junction showing the macroscopic wavefunction of each superconducting electrode. . . . . . . . . . . . . . . . . . . . . . . . 33 4.2 A typical current-voltage curve, where current is shown on the x-axis and voltage is shown on the y-axis, adapted from Ref. [63]. . . . . . . . . . . . 36 4.3 Diagram of Shapiro step behavior of a Josephson junction. . . . . . . . . . 38 4.4 Diagram showing the magnetic field incident on a Josephson junction and the closed loop utilized to determine the magnetic field dependence. . . . 40 4.5 Plot of the Fraunhofer pattern showing the normalized critical current as a function of the magnetic flux. . . . . . . . . . . . . . . . . . . . . . . . . . 43 5.1 a-f) Overview of the fabrication process of the Al-FeTeSe devices. First exfoliation of the FeTeSe was performed (a), followed by spinning of an e- beam resist (b). A device pattern was then written into the e-beam resist and developed (c). Titanium/Aluminum 5 nm/50 nm was deposited by magnetron sputtering (d), and the excess Al was removed during the liftoff stage (e). This process was then repeated for Ti/Au 5 nm/ 50 nm, using electron beam deposition instead of sputtering. . . . . . . . . . . . . . . . 45 5.2 Schematic and dimensions of FeTeSe device showing the side view and top view. The FeTeSe flake is ∼1.8 µm wide and the Al electrode is ∼1.0 µm wide. The intended device was at the Al-FeTeSe interface, which has an area of 1.8 µm2, however we will see that other regions in the device can potentially contribute as well. . . . . . . . . . . . . . . . . . . . . . . . . . 47 5.3 Optical image of the FeTeSe device structure showing the electrodes used for the four-terminal measurement. This four-terminal measurement should isolate the voltage drop across the Al-FeTeSe interface, and a small amount of the FeTeSe and Al materials as well, which we will see can potentially have a significant impact. This measured region is circled in red. . . . . . 48 viii 5.4 Circuit schematic of the measurement technique used to investigate the de- vice; the 4-terminal current-bias measurement with a DC voltage source and an RF source. The lock in amplifier produced a low frequency AC voltage excitation which we convert into a current excitation I+ using a bias resis- tor. The lock-in then measures the differential resistance dV dI in Ohms using a phase locked technique. This resistance was then measured as a function of DC current using a voltage source and a bias resistor, and as a function of RF frequency and power (using a SMB100A microwave signal generator). See Appendix A.1.2 for a more detailed discussion of the measurement methods. 49 5.5 a,b) Initial dV dI vs I measurements of the FeTeSe device. An initial jump from a 0 Ω state to a 3 Ω resistance state occurs around 0.25 µA (a). Two other distinct jumps in resistance can be seen at 6 µA and 13 µA. We will later find Josephson effects emerginf for both the jump at 0.25 µA (the inner junction) and at 6 µA (the outer junction). The resistance jump at 13 µA likely corresponds to the bulk Al electrode, which will be shown later. . . 50 5.6 dV dI vs I measurements at a frequency of 0.5 GHz and a RF power of -54 dBm. This RF power corresponds to the output of the RF generator, not the RF power incident on the device. At this power, there are clear oscillations of the differential resistance. The peaks correspond to the jump between Shapiro steps and the minima correspond to the Shapiro step plateaus. . . . . . . 52 5.7 The result of numerically integrating Fig. 5.6 to reveal the voltage as a function of current and the associated Shapiro steps. The voltage is shown in units of the characteristic voltage step V = hf/2e, which for 0.5 GHz is 1.0 µV . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 5.8 a,b) Map of dV dI vs I vs RF power, revealing the Bessel function-like depen- dence of the Shapiro steps (a). See Refs. [68, 73] for more details on the dependence of the Shapiro steps in the current-bias, which follow a Bessel function-like oscillation. When the scale is modified, additional features are also visible in the range of -35 dBm to -25 dBm (b), which resemble Shapiro steps (in the outer junction), but are not clearly resolvable in this sweep. . 54 5.9 a,b) Map of dV dI vs I vs RF power at 3.97 GHz. The minimum in the dV dI is still clearly visible for the inner junction (a). However, when the scale is adjusted, additional peaks are visible, which strongly resemble a Shapiro step map for the outer junction (b). . . . . . . . . . . . . . . . . . . . . . 55 5.10 a) Linear cuts of the dV dI vs I vs RF power map at 3.97 GHz. b) The I-V curves resulting from numerically integrating the linear cuts in (a). The Shapiro step height can not be clearly resolved due to the finite slope where the plateaus in voltage would normally be. This arises from the 3 Ω background from the normal state resistance of the inner junction. . . . . . . . . . . . 57 5.11 a) Linear cuts of the dV dI vs I vs RF power map at 3.97 GHz minus the 3 Ω background from the normal state of the inner junction. b) The I-V curves resulting from numerically integrating the linear cuts in (a), revealing quantized Shapiro steps at the expected voltages of 8.2 µV . This suggests there are indeed Josephson effects for both the inner junction and outer junction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 ix 5.12 a) Schematic of vortex flux flow, whereby vortices and antivortices nucleate at weak points in the superconductor, and flow perpendicular to the su- percurrent by the Magnus effect. The vortex contains quasiparticles states, which, when moved, will result in finite dissipation and therefore a finite re- sistance/voltage across the device. b) Schematic of a phase slip line, whereby the vortices moving in the wake of the following vortex become sufficiently deformed such that they are described by a normal dissipating line through- out the device. See Refs. [78, 79] for a discussion of phase slip lines and Refs. [80–82] for a discussion of the kinematic vortices in phase slip lines. 60 5.13 Diagram showing the voltage vs current curves arising due to the successive nucleation of phase slip lines, adapted from Ref. [87]. Specifically, when the current is increased, additional phase slip lines are generated, which con- tribute the same amount of resistance to the device. This would result in steps in the resistance vs current curves, which can also be seen from the steadily increasing slope observed in the voltage vs current curves corre- sponding to jumps in the resistance by ∼ 0.25Ω. . . . . . . . . . . . . . . 62 5.14 dV dI vs I measurements at a frequency of 3.97 GHz at a power of -6 dBm (a) and 0.5 GHz at an RF power of -50 dBm (b). The absence of a clear, consistent step height in resistance suggests that the generation of additional phase slip lines does not have to do with the jumps measured in dV dI . This suggests that if the Josephson effects are indeed generated by phase slip lines, then there is likely a single flux flow region or phase slip line as opposed to the generation of multiple phase slip lines. . . . . . . . . . . . . . . . . . . 63 5.15 A dV dI vs I vs RF power map at 3.45 GHz. A discontinuous jump in the mapping can be seen at ∼-4 dBm. This jump is most evident in the jump in the envelope of the Shapiro steps, which represents the transition to the fully normal state. Another unusual feature of this jump is that, despite the RF power being increased (which would typically decrease the critical current due to heating effects), the current defining the envelope of the Shapiro steps is actually enhanced after increasing the RF power beyond this jump. . . 65 5.16 A fine resolution mapping of dV dI vs I at a frequency of 3.55 GHz. In or- der to determine if the jump is an artifact of the measurement and simply corresponds to a shift in the applied power, we can see if the pattern after the jump can be mapped to a region before the jump. The jump does not appear to be an artifact of the measurement, as the pattern after the jump (highlighted in blue) does not align with the pattern before the jump (high- lighted in white). The envelope of the Shapiro steps was used as a guide to decide how much to shift the pattern down. . . . . . . . . . . . . . . . . . 66 5.17 a-d) dV dI vs I vs RF power mapping at a frequency of 3.55 GHz, repeated 4 times. This demonstrates that the jumps are a real, repeatable phenomenon and not due to a random occurrence, such as someone bumping the system. However, it should be noted that this does seem to be a dynamic effect, because in (d) there are two additional jumps at -3 dBm and at -2 dBm not seen in (a), (b), and (c). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 x 5.18 dV dI vs I vs RF power mapping at a frequency of 3.98 GHz. In this measure- ment, there is minimal change in the envelope of the Shapiro steps; however, there is a clear change in the behavior of the junction. Specifically, there is a discontinuous jump in the peaks in dV dI as a function of current, and the slope of these peaks vs RF power and DC current changes after the jump. Most interesting is the fact that the jump happens over a range of powers as opposed to at a specific power. This likely rules out frequency infidelity as causing the jumps, as there is unlikely to be a shift in the frequency that oc- curs differently for positive and negative DC currents, unless it is a physical property of the Josephson junction. . . . . . . . . . . . . . . . . . . . . . 68 5.19 a,b) dV dI vs I vs RF power mapping at a frequency of 3.35 GHz (a) and 3.5 GHz (b). At these frequencies, a small dip can be seen in the differential resistance, which occurs at progressively lower currents as the RF power is increased. These lines seem to precipitate the occurrence of jumps, and in the case of the 3.5 GHz mapping (b), a reemergence of the Shapiro steps after they had nearly disappeared. These precipitating lines further support that the jumps in the Shapiro mapping are a physical phenomenon within the device and not an artifact of the measurement, as they link the lower power behavior before the jump to the higher power behavior where the jump takes place. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 5.20 dV dI vs I vs RF power mapping at a frequency of 5 GHz. The precipitating lines are visible within this mapping, but do not intersect the envelope of the Shapiro steps. Consequently, there is no jump in the behavior of the Josephson junction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 5.21 The Wyatt-Dayem effect demonstrates the enhanced critical current in su- perconducting strips of Tin under RF irradiation. Specifically, Wyatt mea- sured the critical current as a function of RF power at a range of tempera- tures, reprinted from Ref. [90]. They found that for small RF powers, there was an enhancement of the critical current with RF irradiation, and that at large RF powers, the critical current was reduced, which is the typical behavior and is due to heating effects. . . . . . . . . . . . . . . . . . . . . 74 5.22 Diagram showing the potential regions giving rise to the Josephson effects in this device. The supercurrent is perpendicular to the magnetic field for the phase slip lines and parallel to the magnetic field for the Al-FeTeSe interface. In order to see the Fraunhofer effects usually present in Josephson junctions, the Magnetic field would need to be perpendicular to the supercurrent. . . 75 5.23 The initial sweep of dV dI vs I vs B up to a magnetic field of 200 mT. In this sweep, only the outer junction critical current is visible, and it appears to oscillate before dropping to zero. . . . . . . . . . . . . . . . . . . . . . . . 76 5.24 A finer sweep of dV dI vs I vs B up from 0 mT to 30 mT. In this sweep, oscillations of the critical current are visible, which resemble the Fraunhofer pattern typical of Josephson junctions. A Fraunhofer pattern is shown next to the data for reference. . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 xi 5.25 A fine sweep of dV dI vs I vs B from -2 mT to 2 mT showing the magnetic field oscillations of the inner junction. Concurrent with the typical case of Fraunhofer diffraction, there is a maximum of the critical current at B=0 and the critical current decays to zero at ∼ 2mT . . . . . . . . . . . . . . . 79 5.26 A fine sweep of dV dI vs I vs B from -2 mT to 2 mT, taken under 2 GHz RF irradiation at a power of -45 dBm. Similar to the case of no RF irradiation, the critical current decays over ∼ 2mT . This is also the case for the Shapiro steps. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80 5.27 a,b) A fine sweep of dV dI vs I vs B from -2 mT to 2 mT, taken under 3.95 GHz RF irradiation at a power of -9.5 dBm. A minimum in the critical current of the outer junction can be found at B=0 (a). This unusual phenomenon is typically only found in Josephson junctions with magnetic tunnel barriers or multiband superconductor Josephson junctions. In contrast, at the same RF power and frequency, the inner junction retains the critical current maximum at B=0. This maximum is visible for the critical current as well as for the first Shapiro step of the inner junction. . . . . . . . . . . . . . . . . . . . 81 5.28 a) dV dI vs I measurements at a frequency of 3.98 GHz, highlighting the RF powers where the subsequent magnetic fields will be performed. b) dV dI vs I and B at a frequency of 3.98 GHz and a power of -3 dBm (before the jump). Here, a minimum of the critical current is observed at B=0; however, the envelope is relatively unmodified and retains a maximum at B=0. c) dV dI vs I and B at a frequency of 3.98 GHz and a power of -1 dBm (after the jump). Here, both the Shapiro steps and the envelope of the Shapiro steps display a minimum at B=0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85 5.29 Magnetic diffraction pattern for a π-Josephson junction formed at a YBCO corner junction, as was investigated by Wollman et al. [97]. The unique aspects of this junction are the anisotropic multiband superconductivity, such that there will be a relative phase difference of π between the Josephson junction on the different edges. The primary consequences are a minimum of the critical current at B=0 and oscillations of the critical current over 2Φ0/Φ as opposed to Φ0/Φ for a typical junction. This could explain the roughly doubled scale of the magnetic oscillations in our outer junction (5 mT) vs our inner junction (2 mT). . . . . . . . . . . . . . . . . . . . . . . 86 A.1 A typical current-voltage curve, where current is shown on the x-axis and voltage is shown on the y-axis, adapted from Ref. [63]. . . . . . . . . . . . 91 A.2 Lock-in current-bias measurements utilizing a 1 MΩ resistor to set the cur- rent bias from the Lock-in voltage, resulting in the measurement of dV dI . A Lock-in frequency of ∼ 13 Hz is typically used. . . . . . . . . . . . . . . . 93 A.3 Shapiro Steps plotted as V vs I and I vs V, showing that the current bias measurement (V vs I) results in a well defined single value function which is more easily measurable experimentally. . . . . . . . . . . . . . . . . . . . 94 A.4 Lock-in current-bias measurements utilizing a 1 MΩ resistor to set the cur- rent bias from the Lock-in voltage and the DC voltage source, resulting in the measurement of dV dI as a function of the DC current. . . . . . . . . . . 94 xii A.5 Lock-in current-bias measurements utilizing a 1 MΩ resistor to set the cur- rent bias from the Lock-in voltage and the DC voltage source, resulting in the measurement of dV dI as a function of the DC current. Low-pass-filters and a sapphire board are used to thermalize the electrons. . . . . . . . . . . . 95 A.6 Lock-in and DC current-bias measurements with the addition of a RF mi- crowave source, resulting in the measurement of dV dI as a function of the DC current and RF irradiation (frequency and power). Low-pass-filters and a sapphire board are used to thermalize the electrons in the DC lines. RF at- tenuators are used to reduce the heating from the higher temperature plates through the RF lines. A DC block is used to stop the DC signal from being shunted to the attenuator grounds. . . . . . . . . . . . . . . . . . . . . . . 96 A.7 a) An attempted FeTeSe-Au-Al Josephson junction. b) An attempted FeTeSe- Gr-Al Josephson junction. . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 A.8 a) An attempted FeTeSe-FeTeSe crossbar Josephson junction, which dis- played no superconductivity in the top FeTeSe flake, potentially due to strain. b) A FeTeSe flake suspended over a SiO2 trench, with the intent of applying strain through electrostatic gating. c) The resulting data from the strain device, showing no control effect at 1.7 K. If this experiment were to be repeated, the strain control near Tc should be easier to prove as a proof of concept. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99 A.9 Wirebonder V-shaped path of the tip. . . . . . . . . . . . . . . . . . . . . . 101 A.10 Wirebonder metal stuck to the bond pads which can be used for subsequent bonds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 A.11 A table of the parameters corresponding to the measurements displayed in the Data Dump Section. All data is from the same device measured in the main results section, and is taken over a current range of -6.6 µA to 6.6 µA. 103 A.12 dV dI vs I vs RF power at 100 MHz. . . . . . . . . . . . . . . . . . . . . . . . 104 A.13 dV dI vs I vs RF power at 200 MHz. . . . . . . . . . . . . . . . . . . . . . . . 104 A.14 dV dI vs I vs RF power at 350 MHz. . . . . . . . . . . . . . . . . . . . . . . . 105 A.15 dV dI vs I vs RF power at 400 MHz. . . . . . . . . . . . . . . . . . . . . . . . 105 A.16 dV dI vs I vs RF power at 700 MHz. . . . . . . . . . . . . . . . . . . . . . . . 106 A.17 dV dI vs I vs RF power at 800 MHz. . . . . . . . . . . . . . . . . . . . . . . . 106 A.18 dV dI vs I vs RF power at 900 MHz. . . . . . . . . . . . . . . . . . . . . . . . 107 A.19 dV dI vs I vs RF power at 1.2 GHz. . . . . . . . . . . . . . . . . . . . . . . . 107 A.20 dV dI vs I vs RF power at 2 GHz. . . . . . . . . . . . . . . . . . . . . . . . . 108 A.21 dV dI vs I vs RF power at 3 GHz. . . . . . . . . . . . . . . . . . . . . . . . . 108 A.22 dV dI vs I vs RF power at 3.25 GHz. . . . . . . . . . . . . . . . . . . . . . . . 109 A.23 dV dI vs I vs RF power at 3.45 GHz. . . . . . . . . . . . . . . . . . . . . . . . 109 A.24 dV dI vs I vs RF power at 3.5 GHz. In this map the jump shown in the main text is missing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 A.25 dV dI vs I vs RF power at 3.7 GHz. . . . . . . . . . . . . . . . . . . . . . . . 110 A.26 dV dI vs I vs RF power at 3.8 GHz. . . . . . . . . . . . . . . . . . . . . . . . 111 A.27 dV dI vs I vs RF power at 3.9 GHz. . . . . . . . . . . . . . . . . . . . . . . . 111 A.28 dV dI vs I vs RF power at 3.95 GHz. . . . . . . . . . . . . . . . . . . . . . . . 112 A.29 dV dI vs I vs RF power at 4 GHz. . . . . . . . . . . . . . . . . . . . . . . . . 112 xiii A.30 dV dI vs I vs RF power at 7 GHz. . . . . . . . . . . . . . . . . . . . . . . . . 113 A.31 dV dI vs I vs Magnetic Field at 500 MHz. . . . . . . . . . . . . . . . . . . . . 113 A.32 dV dI vs I vs Magnetic Field at 800 MHz. . . . . . . . . . . . . . . . . . . . . 114 A.33 dV dI vs I vs Magnetic Field at 3.55 GHz. . . . . . . . . . . . . . . . . . . . . 114 A.34 dV dI vs I vs RF power at 3.55 GHz and 2 mT. . . . . . . . . . . . . . . . . . 115 A.35 dV dI vs I vs RF power at 6 GHz. . . . . . . . . . . . . . . . . . . . . . . . . 115 xiv Chapter 1 Introduction 1.1 Thesis Overview In this thesis, I hope to give readers some familiarity with Josephson effects, highlight the development and state of research into the unconventional superconductor FeTe1−xSex, and discuss some of the unusual results we found when performing cryogenic measurements of FeTe1−xSex. In Chapter 2, I will give a very brief overview of the basic phenomena of Super- conductivity to prepare unfamiliar readers for discussions of FeTe1−xSex and Josephson junctions. In Chapter 3, I will review the development and current state of research regarding the unconventional superconductor FeTe1−xSex. This section may be of interest for those concerned with Majorana fermions in FeTe1−xSex. In Chapter 4, I will give a basic description of Josephson junctions and the related phenomena. Specifically, I will describe the Josephson equations, the AC Josephson effects, and the Fraunhofer pattern arising in the presence of a magnetic field. In Chapter 5, I will discuss the main results of our study, the Josephson effects in a FeTe1−xSex-Aluminum Josephson junction. I will discuss the presence of two Josephson 1 effects at different energy scales, discontinuous jumps in the AC Josephson diffraction pattern as RF power is increased, and the unusual magnetic diffraction pattern arising under RF irradiation. In Chapter 6, I will conclude by summarizing our findings and discussing some of the future avenues for research on the unconventional superconductor FeTe1−xSex. The Appendix contains information I hope will be useful for readers interested in some specific details. This section covers the details of experimental methods, details of the deposition systems, various technical methods (tips and tricks), and finally, a data dump of the remaining data measured in the FeTe1−xSex-Al Josephson junction, not already shown in Chapter 5. 2 Chapter 2 Superconductivity 2.1 Introduction to Superconducting Phenomena In the following, I will give a brief introduction to the relevant physical phenomena of superconductors, which I learned from the excellent Introduction to Superconductivity by Michael Tinkham (Ref. [1]), as have many before me. This phenomenological description will serve as a basis for understanding superconducting phenomena in typical superconduct- ing systems, as well as provide a basis for the unconventional superconductor FeTe1−xSex, which we utilized in our study. Superconductivity was first discovered by Kamerlingh Onnes in 1911 (Ref. [2]). This discovery was made possible by Onnes’s successful liquefaction of helium in 1908, for which he would later receive the Nobel Prize. The successful liquification of helium opened the door for generations of low-temperature experiments. However, only 3 years later, in 1911, Onnes was researching the resistance of metals as a function of temperature, when he dis- covered that, in Mercury, below a temperature of ∼4.2 K, the electric resistance of Mercury became negligible. A schematic representation is shown in Fig. 2.1. This is a hallmark behavior of superconducting materials , and shortly thereafter led to demonstrations of persistent dissipationless current flow in a superconducting ring. 3 Figure 2.1: Resistance vs Temperature curve for a superconductor. Another early phenomenon discovered in superconducting materials was the Meissner effect, discovered by Meissner and Ochsenfeld in 1933 (Ref. [3]). The Meissner effect is the demonstration of perfect diamagnetism by superconducting materials. Therefore, superconductors expel magnetic fields from their interior, as shown in Fig. 2.2. As noted by Tinkham (Ref. [1]), this would seem to be described by the perfect screening of the magnetic fields by dissipationless currents in the superconductor. However, according to Faraday’s law, a perfect conductor would not require that the magnetic flux equal to zero, and would instead require that the time rate of change of the magnetic flux equal to zero, dB dt = 0. This would trap the magnetic flux when the material entered the superconducting state. In 1935 the London brothers formulated two equations, dJ dt = ne2 m E and ∇ × J = −ne2 m B which result in the exponential screening and expulsion of magnetic fields over a length known as the London penetration depth λL = √ m µ0nse2 . An extension of this phenomenon, which flows naturally from the idea that super- conductors and magnetic fields are antagonistic to each other, is the idea of the critical field of a superconductor. This is a field large enough that the free energy cost to expel 4 Figure 2.2: Diagram of the Meissner effect and the London penetration depth. the field is greater than the free energy gained by forming the superconducting condensate, and therefore, superconductivity is destroyed. The critical field follows from a maximum value at zero temperature to zero field at the critical temperature of the superconductor, see Fig. 2.3. This relation is approximately described by Bc = Bc(0)(1− (T/Tc)2). Figure 2.3: Diagram of relationship between critical temperature and critical field. 5 A related phenomenon in superconductors is the critical current density. This is the maximum current density that a superconductor can maintain before the superconductivity breaks down and transitions to the normal state, see Fig. 2.4. This breakdown of the superconductivity is a direct consequence of the critical field of the superconductor. As the current flows through a superconducting material, there will be a ‘self-field’ generated at the surface of the superconductor. If this surpasses the critical field of the superconductor, superconductivity is destroyed. For the simple case of a cylindrical superconducting wire, the relation Jc = Bc µ2 0λ can be derived. Where Jc is the critical current density in the superconductor defined as the critical current per unit area. This relation is modified in samples with different geometries or when one of the length scales of the superconductor is smaller than the London penetration depth. Figure 2.4: Diagram showing the critical current of a superconductor. Lastly, I will briefly touch on the two fundamental theories that describe the behav- ior of superconductors. In 1957, Bardeen, Cooper, and Schrieffer established a microscopic theory to describe how, below the superconducting transition temperature, an attractive 6 electron-phonon interaction can cause pairs of electrons to condense into a bosonic quan- tum state. It would later be found that other interactions can give rise to superconduc- tivity besides electron-phonon coupling, in so-called unconventional superconductors. This condensate consists of Cooper pairs and is separated from electronic states by the super- conducting gap, 2∆. Above this gap is an enhanced quasiparticle density of states, which was ‘pushed’ out of the superconducting gap, see Fig. 2.5. These peaks in the density of states can be used to identify the superconducting gap in a multitude of measurements. The BCS theory is particularly useful in its description of quasiparticles states. These Bo- goliubov quasiparticles are unpaired superpositions of electron and hole states which exists outside of the superconducting gap. The existence of quasiparticles in superconductors has a number of significant impacts on devices, which we will discuss in Chapter 5. Figure 2.5: Diagram showing the electron density of states around the Fermi energy in a superconductor. 7 In order to describe the spatially-dependent phenomena in superconductors, the Ginzburg-Landau theory was developed, which defines a macroscopic order parameter based on the theory of second-order phase transitions. This order parameter turns out to be the macroscopic wavefunction of the superconductor and is defined as ψ = ψ0e iφ. Where ψ0 = √ ns, and ns is the Cooper pair density. φ is the macroscopic phase of the superconductor which is crucial for understanding the dynamics of systems involving su- perconductors, as we will discuss in Chapter 4. It was later shown by Gor’kov that this theory could be derived from the BCS theory. The wavefunction of a macroscopic su- perconductor represented by a single phase leads to a number of interesting phenomena, such as flux quantization, Type II superconductors, superconducting vortices, and most importantly for this study, the Josephson effect. The Josephson effect occurs when two superconductors are separated by a non-superconducting layer that is sufficiently thin to allow for phase-coherent transport. In these devices, the phase difference between the two superconductors evolves according to the Josephson equations, which will be discussed in detail in Chapter 4.1. 8 Chapter 3 Review of FeTeSe FeTe1−xSex is a class of materials derived from the parent compounds FeTe and FeSe, which demonstrates varying material properties based on the ratio of Te to Se. Under certain sub- stitutional compositions, the class of FeTe1−xSex crystals demonstrate superconductivity and topological surface states, which has motivated significant research efforts aimed at the realization of Majorana fermions, since the discovery of superconductivity in FeTe1−xSex in 2008 (Ref. [4]). This came less than 1 year after the discovery of the first iron-based superconductor LaOFeAs (Ref. [5]), which shocked the scientific community because su- perconductivity and magnetism are considered antagonistic electronic phases. 3.1 Crystal Structure and Parent Compounds FeTeSe and its parent compounds FeSe and FeTe are two-dimensional van der Waals (vdW) materials. This means that they are composed of atomically thin 2D layers of strongly co- valently bonded atoms, which are weakly bonded by van der Waals forces to subsequent layers, as shown in Fig. 3.1. The prototypical example of a 2D material is graphene. In 2004, single-layer graphene was produced using scotch tape to separate an individual atomic layer from a bulk piece of graphite (Ref. [6]), which is known as graphene (Gr). This demonstration of a free-standing 2D material challenged previous assumptions based 9 on early works by Pierls in 1935 (Ref. [7]) and Landau in 1937 (Ref. [8]), which implied that divergent out-of-plane fluctuations would cause 2D materials to be unstable. Furthermore, in 1966 (Ref. [9]), works by Mermin-Wagner suggested that there could be no long-range order in one-dimensional or two-dimensional systems due to the enhancement of long-range fluctuations. It should be noted that Honenburg is sometimes credited alongside Mermin and Wagner for his earlier unpublished proof, and Berezinskii is sometimes credited as well for his independent proof. It is now widely accepted that Gr was realizable because it is not a perfect 2D system in the way the theoretical treatments assumed; the finite size of graphene monolayers and out-of-plane displacements (ripples) of the 2D structure relax the strict requirements of these theoretical works. The discovery of two-dimensional layers of graphene led to intense research over the past 20 years. Many interesting phenom- ena have now been demonstrated in 2D materials, and most significantly for this study, superconductivity. Besides being a 2D material, the layers of FeTeSe form a tetragonal PbO-type struc- ture at room temperature, as can be seen from the diagram of FeSe shown in Fig. 3.1, where a square structure can be seen along the c-axis. The central Fe atoms in vdW layer are bonded to Se atoms, alternating above or below the Fe plane. In FeTeSe crystals, a per- centage of Se atoms will be substituted with Te atoms according to the overall composition. At lower temperatures, certain compositions of FeTeSe crystals undergo a structural tran- sition to the orthorhombic phase, breaking the C4 rotational symmetry and resulting in C2 rotational symmetry. This occurs at ∼90 K in FeSe crystals and is related to an electron- nematic phase transition (Ref. [10]). In FeTe, which is not superconducting, there is no electronic-nematic transition. Instead, there is a phase transition to an antiferromagnetic phase below ∼70 K (Ref. [11]). 10 Figure 3.1: a,b) The 2D tetragonal crystal structure of FeTe1−xSex. The larger spacing between layers bonded by weak vdW forces and smaller spacing in the layer due to strong covalent bonds is visible in (a). The tetragonal structure is most clear from the square structure visible along the c-axis of the material (b). Adapted from Ref. [4]. 11 3.2 Superconductivity in FeTeSe 3.2.1 Notable Attempts to Increasing Tc As can be seen in the phase diagrams of FeTe1−xSex, the parent compound FeSe is a superconductor with a Tc of ∼9 K in bulk crystals (Refs. [4, 12, 13]) and FeTe is an antifer- romagnet which demonstrates no superconductivity (Ref. [11]). FeTe1−xSex demonstrates superconductivity within a range of x=0.1 to x=1, depending on the growth process. De- spite the non-superconducting nature of FeTe, when FeSe is substituted with Te, the Tc can be enhanced to 14 K. For our FeTe0.55Se0.45 crystals, we measured the Tc in the range of 13 K to 14 K, see Fig. 3.3, which is in the typical range for this composition. The highest Tc is typically measured in samples of FeTe0.6Se0.4 which can reach ∼15 K (Ref. [14]). Figure 3.2: Doping dependence of FeTe1−xSex, showing the antiferromagnetic state when Se = 0.0 and the superconducting state from Se = 0.1 to Se = 1.0. The references in the figure are; ‘Bulk-Ref [4]’: Ref. [15], ‘Bulk-Ref [11]’: Ref. [16], ‘Crystal-Ref [3]’: Ref. [17]. Adapted from Ref. [18]. As with all superconductors, there have been substantial efforts to study and increase the Tc of the family of FeTeSe materials, with the idea of understanding high-temperature 12 Figure 3.3: Resistance vs temperature that we measured for a FeTe0.55Se0.45 flake, showing a Tc around 14 K. superconductivity or realizing room-temperature superconductivity. Below, I give a sum- mary of the efforts to enhance superconductivity in the class of FeTeSe superconductors. Initial improvements to the Tc were mostly realized by improving the composition of the crystals. As was mentioned previously, excess iron helps stabilize the growth of FeTeSe crystals (Ref. [19]); however, it is detrimental to the material performance. This was demonstrated in Ref. [14], where they showed an enhancement of the Tc (measured from the onset of superconductivity) from 11.6 K to 14.8 K by reducing the Fe content from Fe1.11Te0.6Se0.4 to Fe1.04Te0.6Se0.4, as is shown in Fig. 3.4. Further, magnetic susceptibility measurements showed a sharp transition to a diamagnetic state in Fe1.04Te0.6Se0.4 suggesting bulk superconductivity, whereas Fe1.11Te0.6Se0.4 samples showed a broad transition which 13 would suggest a reduction in the superconducting volume fraction. They suggested that these excess Fe atoms, which sit between the vdW layers of FeTeSe, are coupled to the Fe in the vdW layers and may lead to the localization of the superconductivity. In order to solve the issue of the beneficial role of excess Fe for growth and the detrimental effect on superconductivity, a number of groups investigated oxygen annealing methods to remove excess Fe after the growth of FeTeSe crystals (Refs. [19–22]). Specifically, it was shown that the excess Fe could be reduced from 1.5% to 0.1% by annealing at 400◦ C at ∼1.5% molar ratio of oxygen to Fe for more than 1 hour (Ref. [19]). The Tc and Jc (critical current density) were enhanced as the O2 percentage was increased, until around 1.5% when the Jc started to slightly decrease, as is shown in Fig. 3.5. Another method of increasing the Tc in the family of FeTeSe crystals was pressure. In studies on FeSe crystals, the Tc was enhanced to ∼37 K by the application of ∼7-9 GPa of hydrostatic pressure (Refs. [23, 24]). At higher pressures, the Tc decreases, and a hexagonal structure arises (Fig. 3.6). The initial increase in Tc up to ∼7-9 GPa coincides with the significant reduction of the unit cell volume (∼10% at 1.5 GPa). This increases the interaction strength between the atoms, which likely drives the increase in Tc. We will discuss the unconventional mechanism of superconductivity in the family of FeTeSe crystals in Section 3.2.2. Finally, the largest enhancement in Tc was seen in monolayer FeSe films grown on SrTiO3 substrates. In these devices, a superconducting gap was observed above 65 K (Refs. [25–27]) and later confirmed by studies utilizing Angle Resolved Photoemission Spectroscopy (ARPES) (Refs. [28, 29]) as well as in-situ four-probe measurements (Ref. [30]). The physical origin of the drastic increase in Tc was the subject of intense investi- gation, and, ultimately, it was suggested that the unusual pairing mechanism of FeSe is enhanced by forward scattering electron-phonon interactions between FeSe and the SrTiO3 substrate, despite not being a typical BCS superconductor (Ref. [31]). This assertion was backed by growing SrTiO3 with different oxygen isotopes, which modifies the strength 14 Figure 3.4: Tc enhancement by removing excess Fe atoms. The positions of excess Fe atoms are shown to be at an interstitial location within the crystal. The magnetic susceptibility and resistivity are shown as a function of temperature for the higher Fe concentration (SC2, 11% excess Fe) and for the lower Fe concentration (SC1, 3% excess Fe). The lower Fe concentration improves the diamagnetic response and the Tc of the crystal. Adapted from Ref. [14]. of the electron-phonon interaction (extracted from ARPES measurements), resulting in modifications to the size of the bandgap of FeSe. 15 Figure 3.5: The effect of oxygen annealing on removing excess Fe and improving the Tc. The diamagnetic response, critical temperature, and critical current are all enhanced due to the effective removal of excess Fe by oxygen annealing. Adapted from Ref. [19] 3.2.2 Fermi Surface/Pairing Mechanism/Pairing Symmetry To understand the pairing symmetry in FeTeSe and the larger class of iron-based super- conductors, I highly recommend the excellent perspective by Hoffman (Ref. [32]), which I will follow here. In order to discuss the pairing symmetry in FeTeSe, it is instructive to 16 Figure 3.6: A phase diagram showing the pressure and temperature-dependent behavior of FeSe. Specifically, the pressure-dependent enhancement of Tc can be seen until 10 GPa. Afterward, this FeSe begins to undergo a transition to the hexagonal phase. Adapted from Ref. [23] discuss the Fermi surface in the larger class of iron-based superconductors. Specifically, in the case of LaFeAsO1−xFx, Mazin et al. calculated the Fermi surface in the case of x = 0 (Fig. 3.7) and found that there were two electron cylinders centered on the M point and two hole cylinders centered on the Γ point (Ref. [33]). There is also a heavy hole band at the Γ point; however, Mazin showed that this went away at the relatively small doping of x = 0.04 - 0.05. This Fermi surface turns out to be representative of a large number of iron- based superconductors. It was further deduced by Mazin that these Fermi surfaces may be similar enough to be linked by a small range of K vectors in the Brillouin zone. Two Fermi surfaces linked by a common wavevector result in significant Fermi surface nesting, which may give rise to a number of emergent phenomena. If superconductivity were mediated by spin-fluctuations (a proposed pairing mechanism for the high Tc cuprate superconductors), this would require the order parameter on the electron and hole pockets to have opposite signs. This is denoted as S± pairing, as there is a change in the sign of the order parameter between the two pockets (hence ±), but the order parameter is isotropic and nonzero in each of the pockets (hence S). This is in contrast to the nodal order parameter in the d-wave 17 superconductivity of cuprate superconductors, which has multiple lobes. Figure 3.7: The calculated band structure (a) and Fermi surface (b) of LaFeAsO1−xFx. The key properties here are two electron cylinders around the M point and two hole cylinders around the Γ point. Adapted from Ref. [33] In the case of FeTeSe, the Fermi surface was experimentally verified by orbital- polarization resolved ARPES (Ref. [34]). Further, the spin-fluctuation mediated super- conductivity was confirmed using neutron scattering to measure the spin resonance in the superconducting state of FeTe0.6Se0.4, which was found to be characterized by a wavevector matching the Fermi surface nesting wavevector (Ref. [35]). Finally, the pairing symme- try was confirmed by scanning tunneling microscopy (STM) (Ref. [36] using a technique known as quasiparticle interference imaging (Ref. [37]). Typical STM measurements either measure the tunneling current as a function of applied bias voltage (Fig. 3.8b), which can reveal the superconducting gap, or map the tunneling current as a function of position (Fig. 3.8a), which can reveal structures within the superconductor. In order to measure quasiparticle interference, the ratio of the conductance at positive bias and negative bias was mapped in order to probe particle-hole symmetric quasiparticles (Fig. 3.8c). A Fourier transform was then applied to this mapping, giving rise to a map of the quasiparticle in- tensity as a function of wavevector (Fig. 3.8d). Because the different wavevectors scatter between pockets of the same sign or opposite sign, and the magnetic field induces vortices, which may provide additional time-reversal-odd scattering, the intensity of scattering for 18 the different wavevectors should differ. This is seen in the experimental results, as q3, which connects pockets of the same sign, is enhanced, and q2, which connects pockets of opposite sign, is suppressed (Fig. 3.8e). This confirms the sign reversal of the order param- eter between electron and hole pockets and, therefore, the S± pairing symmetry in FeTeSe superconductors. Figure 3.8: Quasiparticle interference probes of the unconventional superconductivity of FeTeSe. a) The tunneling current as a function of position. b) The tunneling current as a function of sample bias, showing the reduced tunneling current due to the quasiparticle gap. c) The ratio of the conductance at positive and negative voltage bias as a function of position. d) The Fourier transform of (c) showing the quasiparticle intensity vs the wavevector. e) The quasiparticle interference under a magnetic field of 10 T showing the enhanced or suppressed scattering based on the phase difference between the pockets, confirming the S± pairing symmetry. Adapted from Ref. [37] 3.3 Topology One aspect of FeTeSe crystals I have not yet touched on is the topological nature of the electronic states in these crystals. This is central to the interest in FeTeSe as combined 19 topology and superconductivity give rise to Majorana bound states, which have been posed as an attractive platform for quantum computing. I will begin by providing a brief histor- ical and phenomenological survey of topological materials, which will follow the excellent perspective by Moore (Ref. [38]). At their simplest level, topological insulators are materials that possess linear-dispersion (Dirac-like, i.e., graphene) electronic states at their surfaces when interfacing with non- topological materials or media. Topological insulators get their name from the mathemati- cal field of topology, which covers the properties of objects or spaces that can be smoothly transformed between one another. Associated with this ‘smoothness’ is the topological in- variant of the system. There is an old joke that a topologist can’t tell the difference between a coffee mug and a doughnut. This classic example of topology is based on the idea that a coffee mug can be smoothly deformed into a doughnut without tearing the surface, as both objects have one hole, which serves as the topological invariant in this system. However, transitioning from a ball to a doughnut would require an abrupt transition, i.e., tearing the surface. In crystals, there are many different types of topological invariants, so I will briefly cover the most straightforward examples. The discovery of the quantum Hall effect, where 1D conductive edge states flow around the edge of a 2D electron gas under external magnetic fields (Ref. [39]), led Haldane to suggest that such an effect could be realized in crystals without an externally applied magnetic field. This led to the idea of topological insulators, which were predicted (Ref. [40]) and later shown to be realizable in a 2D quantum well, characterized by edge states with quantized conductance, shown in Ref. [41]. In this system, the spin-orbit interaction takes the place of the external magnetic field in producing the edge states. After this, a 3D topological insulator was first realized in BixSb1−x, where the dispersion relation of the surface states was revealed and a linear Dirac point was identified (Ref. [42]). These topological surface states arose due to the opening of a band gap by the spin-orbit interactions in these crystals, which induce band inversion. 20 I will now give a simplified description of what gives rise to topological surface states. In a simple case of a topological crystal, spin-orbit coupling modifies energies in the band structure of the material, and when spin-orbit coupling is strong, these bands can be inverted in the bulk of the material. To match the symmetry of the bands at the interface between a topological insulator and a normal insulator (or vacuum), the inverted bands must revert to the uninverted case. In the process, these bands must cross the Fermi energy, and therefore, electronic states are generated at the surface (Fig. 3.9). In reality, band inversion is not the only requirement for a topological insulator, and the Berry curvature should be calculated around a closed surface in the Brillouin zone, which classifies the topological invariant of the material. Additionally, I have not touched on the symmetries in these systems; the topological states in these systems can be protected by a number of symmetries, such as time reversal symmetry (quantum Hall effect), as well as various crystal symmetries. For a detailed discussion of topological materials, refer to Ref. [43]. Figure 3.9: Illustration of the band inversion effect giving rise to topological surface states. The strong spin-orbit coupling opens a bandgap and causes twisting of the bands (i.e., band inversion). At the interface with materials without band inversion, the bands must be reverted to an untwisted non-inverted state and must cross to do so. This closes the bandgap locally, giving rise to topologically protected surface states. Now that a simple understanding of topological materials has been established, we 21 can touch on the suggested topology of FeTeSe, which is still being debated. Unlike in the above examples, FeTeSe is not an insulating material in the bulk. Although the insulating states in the bulk of topological insulators make the isolated measurement of the surface states easier, the topological nature of the surface states is not dependent on an insulating bulk of the material. Additionally, because both the topological states and the current flow in superconductors happen at the surface, probes of the topology of FeTeSe are still realizable, as will be shown next. 3.3.1 Topology In FeTeSe The topological states in FeTeSe were first intensively investigated by Wang et al. (Ref. [44]), who carried out first principle calculations of the band structure comparing the band structure of FeSe and FeTe0.5Se0.5, as well as ARPES measurements of FeTe0.5Se0.5 crystals to support their findings. Intuitively, the heavier Te atoms, as opposed to the lighter Se atoms, would suggest an increased magnitude of spin-orbit coupling, due to the relativistic nature of this effect. Therefore, one may imagine that as the content of Te is increased, the topologically trivial FeSe may transition to a topologically non-trivial state in FeTe1−xSex. Calculations of the band structure of FeSe and FeTeSe revealed similar structures, with electron and hole pockets around the M and Γ points, respectively, as is typical for most Fe-based superconductors. However, there is one clear difference between these two band structures. Relative to the case of FeSe (Fig. 3.10a), in FeTeSe (Fig. 3.10b) the Γ2 − band (bolded in red) is pushed down in energy, and along the Γ-Z direction now crosses the Fermi energy, as well as other bands near the Fermi level. After this, the authors added the spin-orbit interaction to their calculation (Fig. 3.10c), which reveals that one of the crossings is avoided due to the spin-orbit interaction and a bandgap is opened (Fig. 3.10d). This avoided crossing mixes (often referred to as twists) the bands, similar to the case of traditional 3D topological insulators, resulting in band inversion. They confirmed a non-zero topological invariant, assuming a curved chemical potential represented by a 22 dashed line. Topological surface states can arise in non-insulating materials, and therefore, utilization of a curved Fermi surface is necessary to exclude any effects of the chemical potential crossing a band on the calculated Berry curvature. After this, the Γ2 − band crossing the Fermi energy was confirmed by ARPES measurements taken as a function of successive K doping of the FeTe0.5Se0.5 crystals. Figure 3.10: First principles calculation of the band structure of FeTeSe, revealing band inversion and a non-zero topological invariant. a) The band structure of FeSe. b) the band structure of FeTeSe, neglecting the spin-orbit interaction. The main difference from FeSe (a) is that the Γ2 − band (bolded in red) is pushed down in energy and crosses the Fermi energy along the Γ-Z direction. c) The band structure of FeTeSe after adding the spin-orbit interaction. d) A zoom-in on the Γ-Z direction showing the avoided crossing which gives rise to the band inversion and topology in FeTeSe. Adapted from Ref. [44] 3.4 Topological Superconductivity So, what is the significance of topological surface states in FeTeSe? This significance is related to Majorana bound states, which can emerge when superconducting and topologi- cal systems are combined, and may have applications in topologically protected quantum computation due to their unique exchange statistics. Below, I will give a brief overview 23 of Majorana Fermions, Majorana-bound states, and a simple description of the exchange statistics. For a more in-depth discussion, please refer to the referenced articles. The term Majorana Fermion refers to a Fermion which is its own antiparticle, and was originally proposed in the context of particle physics. Because particles and antipar- ticles must have opposite charge, a Majorana Fermion, which is its own antiparticle, must have zero charge. In superconducting systems, electron-hole symmetry is imposed by the superconducting condensate, and quasiparticle excitations that are combinations of elec- tron and hole creation operators can be generated (Ref. [45]). These quasiparticles are commonly referred to as Bogoliubov quasiparticles. Among the Bogoulibov quasiparticles in a typical superconductor, a Majorana Fermion may be generated if a quasiparticle is its own antiparticle. In order to realize such a quasiparticle, the normal electron states in the superconductor must follow a linear Dirac-like dispersion, according to Ref. [46]. Based on this idea, in 2008, breakthrough theoretical works by Fu and Kane (Refs. [47, 48]) suggested that a superconductor coupled to the linear dispersion of electron states in a topological insulator could realize Majorana bound states. They demonstrated that these Majorana bound states would exhibit non-Abelian exchange statistics that would make them useful for quantum computation. A brief note on exchange statistics: Abelian exchange statistics are what describe typical Fermions and Bosons, which accrue a complex phase eiϕ in the wavefunction when two particles positions are exchanged, where eiϕ will be 1 for Bosons (due to the symmetric wavefunction) and -1 for Fermions (due to the antisymmetric wavefunction). Non-Abelian exchange statistics imply that when two particles are exchanged, a phase is accrued and, critically, the mode of the wavefunction is modified as well. Further, the non-abelian exchange statistics imply the wavefunction depends on the order in which particle exchanges are performed and the rotation direction of the exchange. Since the wavefunction can be controlled this way, a method of entanglement known as braiding was proposed as a basis for qubits. Figure 3.11 shows a schematic of Majorana braiding where the qubit is initially 24 in the ground state (|012034⟩) and finally in a superposition of the ground and excited state ((|012034⟩+ |112134⟩)/ √ 2). Figure 3.11: Illustration of braiding operations on Majorana bound states. The Majo- rana wavefunction starts in the ground states defined by pairs of Majorana particles. By exchanging the particles’ positions, the wavefunction will now be in a superposition of the ground and excited states. For successive braiding operations, the order of exchanges af- fects the final state. Adapted from Ref. [43] 3.4.1 Topological Superconductivity in FeTeSe Along with the prediction of surface states in FeTeSe by Wang et al. (Ref. [44]), the author also noted that the combination of topological surface states and superconductivity in a single material (FeTeSe) could give rise to Majorana bound states on the surface of FeTeSe. The superconducting quasiparticle gap ∆, however, will ‘gap out’ these surface states, making them inaccessible to experimental probes. Fortunately, the order param- eter (and therefore the quasiparticle gap) becomes zero at the core of Abrikosov vortices (Fig. 3.12). Therefore, it was predicted that Majorana bound states could be bound to 25 the cores of Abrikosov vortices in topological superconductors (Refs. [47, 49]). ARPES measurements were able to prove topological superconductivity in FeTeSe by imaging the spin-polarized Dirac-like surface states of FeTe0.55Se0.45 as well as proving the s-wave nature of the superconductivity (Ref. [50]). Figure 3.12: Illustration of a vortex in a superconductor. The order parameter (ψ) is reduced inside a radius equal to the coherence length (ξ), eventually becoming zero at the core of the vortex. The magnetic field intensity (H) decays over the penetration depth of the superconductor (λ). That same year, the first direct experimental probe of Majorana bound states in FeTeSe was achieved (Ref. [51]). This was achieved using STM, which we discussed above for the determination of the pairing symmetry. STM locally probes the density of states as a function of energy by measuring the tunneling conductance from an atomically sharp tip as a function of bias voltage. Additionally, STM is able to resolve spatial features with a resolution smaller than 0.1 nm, which allows spatial probing on the atomic scale, and is much smaller than the nm scale vortices. With these advantageous properties Wang et al. (Ref. [51]) were able to image vortices on the surface of FeTe0.55Se0.45 under an applied magnetic field of 0.5 T. The tunneling conductance was then measured as a function of energy at the center of the vortex and at the edge of the vortex (Fig. 3.13). These measurements reveal a zero-bias peak in the conductance at the center of the vortex, a signature of tunneling through the Majorana bound states, which are predicted to occur at zero energy (Refs. [47–49]). The author also ruled out Caroli de Gennes Matricron (CdGM) 26 Figure 3.13: A direct probe of Majorana bound states at the core of a superconducting vortex using STM. A peak in the conductance can be seen at the core of the vortex using STM (a). This peak in the conductance can be shown to be centered on zero energy, as well as at the core of the vortex (b). An insensitivity to magnetic field was also demonstrated, ruling out CdGM states and Kondo resonances (c,d). Adapted from Ref. [51] states (Ref. [52]) and Kondo resonances (Ref. [53]), which can occur at low energies and masquerade as Majorana bound states. In the case of CdGM states, these are predicted to occur at discrete energy levels of µ∆2/EF for µ = 1/2, 3/2, 5/2, ... (Ref. [52, 54]). However, the most convincing evidence is that CdGM states should split as a function of space when moving away from the vortex center (Ref. [55]), this behavior was not observed, see Fig. 3.13b. In regard to Kondo resonances, these tunneling states would typically take place at a magnetic impurity, and as a function of applied magnetic field, should split in energy (Ref. [56]), which was also not observed (Fig. 3.13c,d). 27 3.4.2 Higher Order Topological Superconductivity in FeTeSe So far, we have discussed topological materials in which the topological states exist with dimensions N − 1, where N is the dimension of the parent material (1D edge states in 2D Electron gases, 2D surface states in 3D topological insulators). However, it is possible to have what is known as a higher-order topological insulator, where topological states can exist in N − 2 (1D edge modes in a 3D platform) or N − 3 dimensions (0D bound states in a 3D platform). Zhang, Cole, and Das Sarma (an early pioneer of topological supercon- ductivity) predicted that higher order topological superconductivity could potentially be realized in FeTeSe, due to the combination of 2D topological surface states and anisotropic S± pairing symmetry (Ref. [57]). They showed that the anisotropy would result in surface states on the top surface with a sign reversal of the pairing potential compared to the side surfaces. Due to the sign reversal of the pairing potential, there must be an angle where the pairing potential passes through zero, and therefore, the superconducting gap is closed, resulting in 1D helical hinge Majorana zero modes. Fig. 3.14 from Ref. [58] schematically shows how topology and S± superconductivity come together to realize this phenomenon. Figure 3.14: Visualization of the S± superconductivity on the surfaces of FeTeSe, and the resulting helical hinge zero modes. The closing of the superconducting gap at an angle between the two surfaces of FeTeSe and the topological surface states give rise to the higher order topology in this material. Adapted from Ref. [58]. That same year, Gray et al. confirmed the existence of these helical hinge zero modes (Ref. [58]). They fabricated two types of contacts to FeTe0.55Se0.45, (1) using a hexagonal 28 boron nitride insulating layer to block the hinge so the electrode only touches the top surface, and (2) contacting both the hinge and the top surface. They found a sharp peak in conductance at zero bias in the sample contacting the edge, which was absent in the contact to the top surface (Fig. 3.15). Additionally, they demonstrated that this peak likely does not arise from Andreev bound states by showing that the temperature dependence differed from what would be expected of Andreev bound states (∝ 1 T ). The discovery of helical hinge Majorana zero modes in FeTe0.55Se0.45 provides further evidence of the S± pairing symmetry, as gap anisotropy is necessary to realize these higher-order topological states. Figure 3.15: a-c) A transport study on FeTeSe (a) revealing a zero bias conductance peak when the electrode is in contact with the edge of the sample (c), which is absent when just contacting the top surface (b), confirming the presence of helical hinge zero modes. The small peak in (b) is likely due to tunneling through the hBN into the helical hinge zero mode, which exists at zero energy. Adapted from Ref. [58]. 3.5 Josephson Effects in FeTeSe-Based Devices Although we will discuss the specifics of Josephson junctions in the following Chapters, I will now give an overview of the current research that has been conducted on FeTeSe-based Josephson junctions. The first study revealing Josephson effects in FeTeSe was performed by Wu et al. (Ref. [59]). In their study, they fabricated FeTeSe-based constriction junctions using a focused ion beam to etch the FeTeSe down to widths ranging from ∼250 nm to 29 ∼650 nm (Fig. 3.16a). In these devices they measured Shapiro steps matching the predicted height (Fig. 3.16b), as well as large induced gaps of up to 2 meV, determined by the IcRn product and the Ambegaokar-Baratoff relation, described in Ref. [60] (see chapter 5 for a more detailed discussion of the Ambegaokar-Baratoff relation). Another FeTeSe-based Josephson junctions was produced by Qiu et al. by stacking two FeTeSe crystals on top of one another (Fig. 3.17a). In this device they measured a number of novel properties, but most relevant to this study was the unusual dependence of the critical current on the magnetic field (Fig. 3.17b). This unusual dependence appears to be a mix of a 0-Junction (the typical case) and a π-Junction (typically only in systems with magnetic materials or multiband superconductors). They suggested that these effects arise due to ferromagnetism at the interface between the two FeTeSe flakes. However, they pointed out that excess Fe, which would be the most obvious cause of the interfacial ferromagnetism, is not likely driving the ferromagnetic behavior in these devices, as the hysteresis effects disappeared above the superconducting critical temperature of FeTeSe. Figure 3.16: a) A constriction Josephson junction fabricated from FeTeSe crystals. b) Shapiro steps were reliable in this system, confirming the realization of a Josephson junction in this device. Adapted from Ref. [59]. 30 Figure 3.17: a) A FeTeSe-FeTeSe homojunction where the vdW gap between he crystals allows for tunneling between the FeTeSe flakes. b) The complicated magnetic field depen- dence of the critical current, showing asymmetry in B and a minimum close to zero. c) The description for the unusual magnetic field dependence is a combination between 0-junction behavior (the typical case) and π-junction behavior (arising in Josephson junctions with multiband superconductors or magnetic insulating layers). Adapted from Ref. [61]. 31 Chapter 4 Josephson Junctions 4.1 Josephson Physics Josephson junctions are devices where two superconducting electrodes are separated by a thin non-superconducting region (Ref. 4.1). When the non-superconducting region is thinner than the relevant coherence length (the superconductor coherence length in a superconductor-insulator-superconductor junction and the normal metal coherence length in a superconductor-normal metal-superconductor junction), there will be tunneling of cooper pairs between the superconducting banks. As a consequence of the phase-coherent transport, Josephson junctions demonstrate zero resistance for pair tunneling at zero bias voltage, but finite resistance (in general) at finite bias. Brian Josephson first described the governing equations of Josephson junctions in 1962, which are known as the Josephson equations (Ref. [62]). Below, we will discuss the derivation of the Josephson equations as well as the primary phenomena observable in this system. 4.1.1 Derivation of Josephson Equations As Ginzburg-Landau theory demonstrates, each of the superconducting electrodes in a Josephson junction can be described by a macroscopic wave function, respectively. 32 Figure 4.1: Schematic of a Josephson junction showing the macroscopic wavefunction of each superconducting electrode. ψ1 = √ n1e iφ1 (4.1) ψ2 = √ n2e iφ2 , (4.2) where φ1 and φ2 are the macroscopic phases of the first and second superconducting electrodes, respectively, and n1 and n2 are the densities of Cooper pairs. We will later assume the electrodes are identical for simplicity; however, the electron densities must first be handled independently, as the charge passing from one electrode to the other will be represented by ṅ1 or ṅ2. This does not violate n1 = n2, because our system is attached to a battery, so the electrodes will not become charged by a finite ṅ1 or ṅ2. These equations do not explicitly account for the battery as a charge source or sink; therefore, this must be handled through a reasonable assumption which I will cover later. The time-independent Schrodinger equation for a two-level system can be used to describe the coupling between the two electrodes, where the ground state energy U char- acterizes zero coupling between the two superconductors and the excited state energy K characterizes the coupling energy between the two electrodes. Because we will apply a voltage V between the two superconducting electrodes, we can assume U1 = +V/2 and U2 = −V/2. For now, we exclude the contribution of magnetic fields. iℏ dψ1 dt = qV 2 ψ1 +Kψ2. (4.3) 33 iℏ dψ2 dt = −qV 2 ψ2 +Kψ1. (4.4) Substituting the wavefunction of each electrode into the time-dependent Schrodinger equa- tion results in the equations iℏ( √ ṅ1 + iφ̇1 √ n1)e iφ1 = qV 2 √ n1e iφ1 +K √ n2e iφ2 , (4.5) iℏ( √ ṅ2 + iφ̇2 √ n2)e iφ2 = −qV 2 √ n2e iφ2 +K √ n1e iφ1 . (4.6) By equating the real and imaginary parts of each equation, we arrive at four equations. Where we have defined the phase difference between the junction as δ = φ1 − φ2. ṅ1 = 2K ℏ √ n1n2 sin(δ) , ṅ2 = −2K ℏ √ n1n2 sin(δ), (4.7) φ̇1 = K ℏ √ n2 n1 cos(δ)− qV 2ℏ , φ̇2 = K ℏ √ n1 n2 cos(δ) + qV 2ℏ . (4.8) Where we can see that ṅ1 = −ṅ2. (4.9) This represents the transport of Copper pairs between the electrodes, and as such, we can write J = 2K ℏ √ n1n2 sin(δ) = ṅ, (4.10) where J is the current density. Now we can make the reasonable assumption that n1 = n2, based on the idea that the Josephson junction is connected to a battery that will source or sink the charge in each electrode as the current flows through the junction. With this, we can additionally define Jo = 2K ℏ n1. This is the critical current density of the Josephson junction, and is analogous to the critical current density of a superconductor on a simple 34 level. Using this definition, the first Josephson equation can be written as J = Jo sin(δ). (4.11) By subtracting the equations in Eq. 4.8, and using our previous definition for the phase difference between the superconductors, we can write the second Josephson equation δ̇ = 2eV ℏ , (4.12) where we have substituted q = 2e for the Cooper pair. And its time-integrated form. δ(t) = δ0 + 2e ℏ ∫ V (t)dt (4.13) The result of this equation is that if a finite DC voltage is applied across the junction, the phase will change at a constant rate, given in Eq. 4.12, causing oscillations of the current (Eq. 4.11), and there will be no net current flow from the superconducting state. In real devices, there will be conduction due to quasiparticles at this point, however, they are not treated within this model. A typical current-voltage curve, as measured by Shapiro et al., is shown in Fig. 4.2, where the voltage was swept (y-axis) and the tunneling current was measured (x-axis). In comparison, if a small current less than Jo is driven through the junction, then there will be a constant finite phase difference between the superconducting electrodes, given by Eq. 4.11. From Eq. 4.12, this would mean that the voltage across the junction must be zero. Driving a current instead of a voltage is referred to as a current-bias measurement and is discussed in detail in Sec. A.1.2. 35 Figure 4.2: A typical current-voltage curve, where current is shown on the x-axis and voltage is shown on the y-axis, adapted from Ref. [63]. 4.1.2 Shapiro Steps Only one year after Josephson’s landmark publication, Sidney Shapiro measured the ex- istence of finite jumps in the current-voltage curve in a Josephson junction irradiated by microwave photons (Ref. [63]). We will next discuss this phenomenon in detail. Due to the microwave photons irradiating the sample, the voltage across the junction can be expressed as the sum of the DC voltage V0 and the AC voltage V1, V = V0 + V1 cos(ω1t). (4.14) Substituting this into the integral form of the second Josephson equation ( Eq. 4.13) results 36 in the equation δ(t) = δ0 + 2eV0 ℏ t+ 2eV1 ℏω1 sin(ω1t). (4.15) Substituting this into the first Josephson equation (Eq. 4.11), results in the equation J = Jo sin ( δ0 + 2eV0 ℏ t+ 2eV1 ℏω1 sin(ω1t) ) . (4.16) Equations of the type sin(sin(x)) can be expanded as a sum of Bessel functions, resulting in the equation J = Jo ∑ (−1)nJn ( 2eV1 ℏω1 ) sin ( γ0 + 2eV0 ℏ t− nω1t ) , (4.17) where Jn is the nth Bessel function of the first type. The current will oscillate with time and average out to zero unless nω1 = 2eV0 ℏ is satisfied. At which point, there will be a DC contribution for the nth term in the series. As a result, if the voltage is increased, there will be no increase in current until V0 = nℏω1 2e = nhf 2e , (4.18) where the applied frequency f is typically in the range of GHz, and the voltage steps correspond to ∼ 2µV/GHz. Similarly, if a current is applied to the junction, the voltage will remain constant until sufficient current exists to drive the voltage to jump to the nth step, shown in Fig. 4.3. This results in finite jumps in the current-voltage curves known as Shapiro steps, shown in Fig. 4.3. Deviations from the typical Shapiro step behavior can be used as a probe of the properties of Josephson junctions. For example, voltage jumps occurring at fractional values of the Shapiro steps can indicate deviations from the typical current-phase relation (δ̇ = 2eV ℏ ). Specifically, higher frequency sinusoidal contributions to the current-phase relation can arise in highly transparent Josephson junctions, as is shown in Ref. [64]. 37 Another example is the missing odd-value Shapiro steps, which are predicted to arise in the presence of Majorana bound due to the 4π period current phase relationship (Refs. [65, 66]). However, missing odd Shapiro steps have been observed in topologically trivial Josephson junctions as well (Ref. [67]), and therefore this is not direct evidence of Majorana bound states. Figure 4.3: Diagram of Shapiro step behavior of a Josephson junction. An important note about the current-voltage curves: As Tinkham notes in Chap- ter 6.3.4 of Introduction to Superconductivity (Ref. [1]), the above model assumes that a voltage is applied to the sample and that the current is measured. This is known as a voltage-bias measurement and gives rise to the relatively simple solutions shown above. However, in realistic cases, due to the zero resistance of the superconductor, the finite 38 resistance in the rest of the circuit sets the current through the Josephson junction, ef- fectively rendering the measurement a current-bias measurement (where current is applied and the voltage is measured). For this reason, current bias measurements are often used to measure Josephson junctions (see the Appendix A.1.2 for a more in-depth discussion of the advantages of current-bias measurements). In the case of current bias measurements, an analytical solution similar to Eq. 4.17 can not be obtained and, therefore, numerical solutions are required. Numerical solutions show that finite DC currents can exist for volt- ages other than the Shapiro step heights (Refs. [68, 69]), such as between the Shapiro step plateaus, as is shown schematically in Fig. 4.3. 4.1.3 Magnetic Field Dependence of Josephson Junctions I will now address the contribution of the external magnetic fields on the behavior of the Josephson junction. A brief summary of the derivation I will follow: I will look at the modification of the critical current in a short Josephson junction (relative to the penetration depth) in an applied magnetic field. In order to calculate the critical current, I use the gauge invariant phase difference across the junction. The dependence of the gauge invariant phase difference on the magnetic field can be derived by setting the sum of the order parameter around the loop equal to zero. After this, the order parameter is eliminated from the expression and a relationship between the gauge invariant phase and the magnetic field si obtained. We can imagine a Josephson junction comprising two superconductors separated by a thin non-superconducting material of thickness d, shown in Fig. 4.4. A spatially uniform magnetic field, B⃗ = Bŷ, is then applied in the y-direction, perpendicular to the direction of the current flow (in the z-direction). Due to the applied magnetic field, the phase difference across the junction, which will determine the supercurrent, is now the gauge invariant phase difference γ. Unlike the order parameter, γ will vary across the length of the junction (x-direction) under an applied magnetic field, which we will see later. The gauge invariant phase difference across the 39 Figure 4.4: Diagram showing the magnetic field incident on a Josephson junction and the closed loop utilized to determine the magnetic field dependence. Josephson junction is γ(x) = φ2 − φ1 − 2π Φ0 ∫ 2 1 A · dl (4.19) −γ(x+∆x) = φ4 − φ3 − 2π Φ0 ∫ 4 3 A · dl, (4.20) where A is the vector potential due to the external magnetic field, and the minus sign on γ in the second equation is because the integral is in the opposite direction as the gauge invariant phase difference, which is defined from the left electrode to the right electrode. Φ0 is the magnetic flux quantum and is defined by Φ0 = h 2e . According to the London equations, in the electrodes ∇φ = 2π Φ0 (ΛJ + A). (4.21) However, because we choose to take our path deep within the electrodes J = 0. Finally, in order to calculate the phase difference between the superconducting electrodes, we can first consider the phase around the entire loop, which must be single valued mod(2π) as this is the order parameter characterizing the superconductor. This can be represented by 40 the integral of the phase gradient around the loop ∫ ∇φ · dl = (φ2 − φ1) + (φ3 − φ2) + (φ4 − φ3) + (φ1 − φ4) = 0. (4.22) Which can also be written as ∫ ∇φ · dl = (φ2 − φ1) + ∫ 3 2 ∇φ · dl + (φ4 − φ3) + ∫ 1 4 ∇φ · dl = 0. (4.23) Substituting Eqs. 4.19, 4.20 and 4.21 into Eq. 4.32 the order parameter can be eliminated to yield γ(x) + 2π Φ0 ∫ 2 1 A · dl+ 2π Φ0 ∫ 3 2 A · dl− γ(x+∆x) + 2π Φ0 ∫ 4 3 A · dl+ 2π Φ0 ∫ 1 4 A · dl = 0, (4.24) which is equivalent to γ(x+∆x)− γ(x) = 2π Φ0 ∮ A · dl. (4.25) Resulting in the equation, γ(x+∆x)− γ(x) = 2πΦ Φ0 , (4.26) where the flux can be represented by the area of penetration of the magnetic field Φ = BA, with A = (d + 2λ)∆x. The penetration depth is incorporated in the thickness of the junction as the magnetic field will also penetrate this region. This results in the equation γ(x+∆x)− γ(x) ∆x = 2π Φ0 B(d+ 2λ), (4.27) which as ∆x goes to 0 becomes dγ dx = 2π Φ0 B(d+ 2λ). (4.28) 41 Integrating this equation yields the gauge invariant phase difference as a function of x γ(x) = 2π Φ0 B(d+ 2λ)x+ γ0. (4.29) Therefore, the current density in the junction will have a spatial dependence according to J = J0 sin ( 2π Φ0 B(d+ 2λ)x+ γ0 ) , (4.30) which, under the assumption of a homogeneous critical current J0, can be integrated over the length of the junction L to get the total current flowing through the junction I = I0 sin(γ0) sin ( πΦ Φ0 ) πΦ Φ0 , (4.31) where Φ = B(d + 2λ)L. The current has a maximum value at γ0 = π/2, resulting in the usual Fraunhofer diffraction pattern for the maximum current, shown in Fig. 4.5, given by Imax = I0 ∣∣∣∣∣∣ sin ( πΦ Φ0 ) πΦ Φ0 ∣∣∣∣∣∣ . (4.32) It should be noted that this calculation was done neglecting self-field effects, which become significant when L > λJ , where λJ = √ Φ0 2πµ0(d+2λ)Jc , and is defined as the Josephson penetration depth. 42 Figure 4.5: Plot of the Fraunhofer pattern showing the normalized critical current as a function of the magnetic flux. 43 Chapter 5 FeTeSe-Al Josephson Junction We fabricated devices based on the interface between Al and FeTe0.55Se0.45, a material which provides a compelling platform for investigating phenomena associated with uncon- ventional superconductivity. In this system, Josephson effects were observed, which deviate from the typical behavior of Josephson junctions shown in Section 4.1. For each of the un- usual features seen, I will present the data step by step to highlight the atypical features demonstrated. For each of the unusual phenomena, I will discuss potential mechanisms driving the behavior. Ultimately, we find that the Josephson effects present in this system may arise from a phase slip line induced in FeTe0.55Se0.45 instead of the expected Josephson junctions at the Al-FeTe0.55Se0.45 interface. 5.1 Fabrication and Measurements 5.1.1 Fabrication Before diving into the results, I will give a brief description of the fabrication and mea- surements of our devices, shown schematically in Fig. 5.1. In order to produce flakes of FeTe0.55Se0.45, which we can process using standard chip processing techniques, we first take bulk crystals of FeTe0.55Se0.45 and exfoliate them by the scotch tape method. The 44 Figure 5.1: a-f) Overview of the fabrication process of the Al-FeTeSe devices. First exfo- liation of the FeTeSe was performed (a), followed by spinning of an e-beam resist (b). A device pattern was then written into the e-beam resist and developed (c). Titanium/A- luminum 5 nm/50 nm was deposited by magnetron sputtering (d), and the excess Al was removed during the liftoff stage (e). This process was then repeated for Ti/Au 5 nm/ 50 nm, using electron beam deposition instead of sputtering. adhesive force of the scotch tape is stronger than the weak inter-layer vdW force holding the 2D layers of FeTe0.55Se0.45 together. As a result, when the tape is peeled apart, the FeTe0.55Se0.45 crystal are cleaved. This tape is then applied to a clean substrate composed of conductive p-doped silicon capped with 285 nm SiO2. Moderate pressure was applied and the sample was given approximately 5 minutes to ensure the FeTe0.55Se0.45 interfaces 45 the substrate well. After this, the tape is peeled back from the substrate and flakes of FeTe0.55Se0.45 will be cleaved by the adhesive force of the substrate and the tape (Fig. 5.1 a), leaving flakes of FeTe0.55Se0.45 approximately 1-10 µm wide and 10-100 nm thick. After flakes have been produced, electron beam resist (PMMA 950A4) can be spun onto the substrate and baked at 180 ◦C for 2 mins (Fig. 5.1b). Alignment marks are then written by electron beam lithography, and subsequently developed in a 3:1 mixture of isopropanol (IPA):methyl isobutyl ketone (MIBK). Optical images of the sample and alignment marks are then used to locate the position of the flake on the chip exactly, and then the intended electrodes can be patterned by electron beam lithography as well (Fig. 5.1c). In the case of our devices, there were two lithography and deposition steps. Aluminum electrodes were deposited by Magnetron Sputtering and Au electrodes were deposited by Electron Beam Deposition, where, in both cases, a thin adhesion layer of ∼5 nm Ti was used (Fig. 5.1d). See Appendix A.5 for standard operating procedures, maintenance, and modifications to the Magnetron Sputtering system and the Electron Beam Deposition system. After metal deposition, the PMMA is removed by acetone, and therefore the metal on the surface of the PMMA is removed as well (Fig. 5.1e). This process is repeated for Au electrode (Fig. 5.1f). The resulting device is shown schematically in Fig. 5.2 and the device image is shown in Fig. 5.3. The area of the electrode-FeTeSe interface is ∼ 1.8 µm2. 5.1.2 Measurement Procedures After device fabrication, the substrate was affixed to a small sample holder using PMMA glue, which was then affixed to a larger sample holder by vacuum grease. At this point, the electrodes on the device were connected to the electrodes on the sample holder by wire bonding (See Appendix A.3 for details of wire bonding). One of the electrodes is also connected to the bonding pad corresponding to the RF lines. This allows measurement of the samples using a Lock-in 4-terminal current-bias measurement scheme (shown in Fig. 5.4), while irradiating the sample with RF photons using Rhode and Schwarz SMB100A 46 Figure 5.2: Schematic and dimensions of FeTeSe device showing the side view and top view. The FeTeSe flake is ∼1.8 µm wide and the Al electrode is ∼1.0 µm wide. The intended device was at the Al-FeTeSe interface, which has an area of 1.8 µm2, however we will see that other regions in the device can potentially contribute as well. microwave signal generator. Frequency ranging from 100 MHz to 7 GHz and an applied RF power ranging from -70 dBm to 5 dBm were used throughout the measurements. The real incident RF power on the device was likely reduced by greater than 10 dB due to attenuators in the lines. The specifics of 4-terminal measurements are discussed in Ap- pendix A.1.2. The four terminal measurement scheme isolates the voltage drop across the Al-FeTe0.55Se0.45 interface (nominally including a Ti layer) from the voltage drop across the lines and other interfaces in the system (voltage drops across the small strips of Al and FeTe0.55Se0.45 comprising the electrodes could still be measured). After wire bonding, the samples are then loaded into either a Leiden Dilution refrigerator or an Oxford Instru- ments Dilution Refrigerator, which were utilized to achieve temperatures of ∼50 mK and ∼25 mK, respectively. See Appendix A.1.3 for details of DC and RF wiring. 47 Figure 5.3: Optical image of the FeTeSe device structure showing the electrodes used for the four-terminal measurement. This four-terminal measurement should isolate the voltage drop across the Al-FeTeSe interface, and a small amount of the FeTeSe and Al materials as well, which we will see can potentially have a significant impact. This measured region is circled in red. 48 Figure 5.4: Circuit schematic of the measurement technique used to investigate the device; the 4-terminal current-bias measurement with a DC voltage source and an RF source. The lock in amplifier produced a low frequency AC voltage excitation which we convert into a current excitation I+ using a bias resistor. The lock-in then measures the differential resistance dV dI in Ohms using a phase locked technique. This resistance was then measured as a function of DC current using a voltage source and a bias resistor, and as a function of RF frequency and power (using a SMB100A microwave signal generator). See Appendix A.1.2 for a more detailed discussion of the measurement methods. 49 5.2 Inner and outer junction effect When measuring a Josephson Junction, the differential resistance (dV/dI) vs current mea- surement resembles measurements of a superconducting flake, but with a significantly re- duced critical current. We therefore expect to measure zero resistance at zero bias current and a resistive state emerging at higher currents. We measured this phenomenon in our Al-FeTeSe Josephson junction at a base temperature of ∼50 mK, shown in Fig. 5.5a. Figure 5.5: a,b) Initial dV dI vs I measurements of the FeTeSe device. An initial jump from a 0 Ω state to a 3 Ω resistance state occurs around 0.25 µA (a). Two other distinct jumps in resistance can be seen at 6 µA and 13 µA. We will later find Josephson effects emerginf for both the jump at 0.25 µA (the inner junction) and at 6 µA (the outer junction). The resistance jump at 13 µA likely corresponds to the bulk Al electrode, which will be shown later. We can clearly see there is a nearly zero resistance state around zero bias current, and a critical current of approximately 0.25 µA, beyond which the differential resistance jumps to approximately 3 Ω (Fig. 5.5a). Expanding the range of our DC current sweep to 20 µA reveals that there is another jump in resistance to approximately 16 Ω occurring at a current of 7.5 µA (Fig. 5.5b). This is typical for Josephson junctions, as the electrodes 50 also have a critical current beyond which there will be a jump in the differential resistance. We will see later see, however, that other effects can also give rise to multiple jumps in resistance. There is a large jump in the resistance around 13 µA likely arising from the transition to the