THERMOMAGNETIC PHENOMENA IN MESOSCOPIC AND PARAMAGNETICALLY LIMITED SUPERCONDUCTORS By Mengling Hettinger A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Physics - Doctor of Philosophy 2015 ABSTRACT THERMOMAGNETIC PHENOMENA IN MESOSCOPIC AND PARAMAGNETICALLY LIMITED SUPERCONDUCTORS By Mengling Hettinger The superconducting fluctuation effect, due to droplets of preformed Cooper pairs above the critical temperature Tc , governs the temperature dependence of kinetic coefficients in superconductors at the onset of the phase transition. The transverse thermoelectric response – Nernst effect – is particularly sensitive to the fluctuations, and the large Nernst signal found in the various superconducting materials has raised much debate on its connection to the origin of unconventional superconductivity. In this thesis, we present a systematic study of the electrical and thermomagnetic transport phenomena in mesoscopic and paramagnetically (Pauli) limited superconductors. In the first chapter of this thesis we concentrate on the study of mesoscopic effects on transport in superconductors. We find that long-range phase coherence developing close to Tc triggers a great amplification of mesoscopic fluctuations due to strong pairing correlations. As a result, mesoscopic conductance fluctuations cease to be universal and exhibit pronounced dependence on temperature. Despite the lack of universality, in the sense of random matrix theory classification, we have discovered a different kind of universality in terms of temperature dependence of fluctuating characteristics. We find that mesoscopic fluctuations of conductivity, transversal thermoelectric coefficient and diamagnetic susceptibility consistently display the same scaling with temperature close to Tc . We connect our results to the existing experimental measurements of conductance fluctuations in superconducting films. Experimental verification of the temperature scaling and the overall magnitude of the mesoscopic fluctuations of Nernst coefficient will provide a powerful tool for a better understanding of thermomagnetic transport phenomena in correlated systems. In the second chapter of this thesis we examine the electrical and thermal transport anomalies in the ultra-thin superconducting films in an external in-plane magnetic field. We concentrate on the Clogston-Chandrasekhar phase transition, i.e., the destruction of superconductivity by a magnetic field by virtue of the Zeeman splitting. Near the quantum critical point of the supercooling line in the phase diagram, we discover highly non-monotonic magnetoresistance. The most remarkable feature of this effect is that fluctuation-induced transport is dominated by the virtual excitations rather than real preformed Cooper pairs. We also carefully study how spin-orbit scattering and other pair-breaking effects modify the fluctuation transport. In the strong spin-orbit scattering regime, we find that the scaling of the thermomagnetic coefficient is the same as conductivity within the classical region of transition, however they are drastically different near the quantum critical point. Even though we primarily focus on the conventional superconductors our result for the Nernst effect may have important implications to the other systems, such as iron-pnictides, and in particular to FeSe compound, which has comparable Zeeman and superconducting gaps. ACKNOWLEDGMENTS First and foremost, I would like to express my sincerest gratitude to my advisor, Dr. Alex Levchenko, who has been patient, supportive, encouraging, caring and understanding since the day I began working with him. His vast knowledge and skills in many areas in condensed matter physics guided me through the process of doing research and writing this thesis. The excellent research atmosphere he provided is simply the best one can possibly ask for. I must thank all the physics and astronomy professors whom I have worked with and learned from over the past five years and whom I have received the time, energy and expertise from. My thesis committee guided me through all these years. I would like to give special thanks to professor Vladimir Zelevinsky, Carlo Piermarocchi, Norman Birge and Scott Pratt for being excellent teachers during my courses of study, and for providing insightful suggestions, constructive feedback and, thoughtful comments at all levels of my research projects. I would never have been able to finish my thesis without the help from my great and cheerful friends and family. I owe a debt of gratitude for the constant support, concern, trust and love to my parents He Zhang, Feng Wang and Rita Murray, and my grandparents Rengsun Zhang and Yunfang Zhang. To my husband, Thomas Hettinger, thank you for listening to my endless complaints and concerns, offering me precious advices, supporting my decisions, and helping me unconditionally through this entire journey. I want to thank all my friends who always be there cheering me up and stood by me through the good times and bad. The research presented in this thesis was supported in part by National Science Foundaiv tion Grants No. DMR-1401908 and ECCS-1407875. TABLE OF CONTENTS LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 Chapter 2 Thermomagnetic transport 2.1 Background and motivation . . . . 2.2 Foundations of UCF . . . . . . . . 2.3 Qualitative considerations . . . . . 2.4 Definitions and assumptions . . . . 2.5 Technical prerequisites . . . . . . . 2.6 Mesoscopic Nernst effect . . . . . . 2.7 Summary . . . . . . . . . . . . . . in mesoscopic superconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 6 8 11 14 17 25 33 Chapter 3 Transport anomalies in Pauli-limited superconductors . 3.1 History of the subject . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Motivation and qualitative picture . . . . . . . . . . . . . . . . . . . 3.3 Fluctuation-induced conductivity near the quantum critical point . 3.4 Effects of pair-breaking scattering . . . . . . . . . . . . . . . . . . . 3.5 Thermomagnetic phenomena . . . . . . . . . . . . . . . . . . . . . . 3.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 36 40 43 49 51 56 APPENDICES . . . . . . . . . . . . . . . . . . . . . . . . . . . Appendix A: Matsubara sums and analytical continuation . . Appendix B: Seebeck (αxx ) and Hall (σxy ) coefficients near Tc Appendix C: Quantum Aslamazov-Larkin terms . . . . . . . . Appendix D: Anomalous Maki-Thompson terms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 61 65 69 71 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 BIBLIOGRAPHY vi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . LIST OF FIGURES Figure 2.1: Figure 2.2: Figure 2.3: Figure 2.4: The layout of the Nernst experiment: by applying a temperature gradient (−∇T ) in the presence of magnetic field, an electric field is generated. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 Building blocks of the diagrammatic technique. Straight lines represent disorder-averaged single-particle Green’s functions. The straightdashed lines represent single impurity lines that carry an overall factor 1/2πνd τ in the diagrams. The vertex function λ(ε, ε , q) is drawn in the ladder approximation, while diagrams with the crossed impurity lines yield parametrically smaller contributions in 1/g 1. The polarization operator Π(ω, q) (“bubble” diagram) is also presented in the main ladder approximation. . . . . . . . . . . . . . . . . . . . . 17 The Dyson equation for the fluctuating propagator L(ω, q) which represented graphically as the wavy line and computed in the ladder approximation. Solid lines represent one-electron Green’s function, greay triangle is the impurity dressed effective vertex, while each cross between two Green’s functions is associated with the electron-electron coupling constant −λsc . . . . . . . . . . . . . . . . . . . . . . . . . . Feynman diagrams for the main fluctuation-induced corrections to the conductivity. In the first row we show the Aslamazov-Larkin diagrams (left) and the Maki-Thompson interference diagram (right). In the second row we show two density of states diagrams. . . . . . 20 23 Figure 2.5: Leading order diagrams for the irreducible correlator of mesoscopic disorder-averaged two pair-propagators. This averaging contains collisions of four diffusion or Cooper modes, and involves forth and sixth order Hikami boxes (internal impurity lines are implicit on diagrams). 28 Figure 2.6: eh The Aslamazov-Larkin diagrams contributing to Kxy (Ω). The wavy lines correspond to the fluctuation propagator L(q, ω); electric current vertices B e and heat current vertices B h are indicated in the figure along with running momenta and frequencies. . . . . . . . . . . . . . vii 30 Figure 3.1: Figure 3.2: Figure A.3: Above the tricritical point T ∗ the second order paramagnet to superconductor transition occurs along the (black) solid line obtained from Eq. 3.1. At T < T ∗ this line becomes a supercooling part of the hysteresis, and the dashed line is its superheating part. The latter is obtained following Ref. [95]. The grey shaded area with the critical point (0, ∆0 ) as its lowest corner bounded by the black dashed line marks the region of quantum fluctuations (QF). . . . . . . . . . . . 37 Phase diagram of a superconducting thin film in a parallel magnetic field parametrized by pair breaking parameter α = D(eHd)2 /6 due to orbital mechanism. Tc0 = Tc (H = 0) is the critical temperature in the absence of a magnetic field. At T = 0 the superconductivity breaks down at the critical value αc = πTc0 /2γE , where ln γE ≈ 0.577 is the Euler constant. . . . . . . . . . . . . . . . . . . . . . . . . . . 39 Integration contour in the plane of complex frequency. The lower part of the contour corresponds to advanced-advanced products of propagators after analytical continuation. The middle section contains mixed causality components of advanced-retarded, while the upper third contains only retarded-retarded products of propagators. . . . 62 viii Chapter 1 Introduction The study of fluctuation effects at the onset of the second order phase transition, which naturally emerged from the Landau’s theory of the phase transformations [1], was instrumental in the development of statistical mechanics and modern condensed matter theory. The seemingly simple question – how accurately does the mean field theory describes the second order phase transitions – lead to the ideas of scaling, universality [2–6], and eventually to the development of the renormalization group [7]. In the context of superconductivity, fluctuations were first studied by Ginzburg [8]. In particular, he analyzed effects of fluctuations above the critical transition temperature, Tc , on the thermodynamic properties of superconductors and demonstrated that, in clean materials with ballistic transport, the fluctuation phenomena become important only in a very narrow temperature region near the transition. This result explained the great level of accuracy of the Bardeen-Cooper-Schrieffer (BCS) theory of superconductivity [9] (which in its essence is a mean field model) in application to various existing experiments of that time. Almost a decade later Aslamazov and Larkin [10], and independently Maki [11] and Thompson [12], realized that the fluctuation region in disordered superconducting films is determined by the resistance per unit square 1 and could be much wider than that in bulk samples. Perhaps even more importantly they demonstrated that within microscopic BCS theory superconducting fluctuations, in fact, play a very important role in explaining temperature dependence of kinetic coefficients, such as conductivity. These authors discovered the phenomenon which nowadays is called the paraconductivity effect – the decrease of the resistance of a superconductor in the normal phase with lowering temperature towards the critical temperature. In a parallel vein this effect was observed in experiments by Glover [13]. Little [14], Langer and Ambegaokar [15], and McCumber and Halperin [16] showed that superconducting fluctuations associated with the phase of the superconducting order parameter, so called phase slips, are also required to explain the residual resistance in the superconducting phase. At very low temperatures, thermal activation of the phase slips is severely suppressed, however resistance may appear due to quantum phase slip tunneling [17]. Abrahams, Redi and Woo [18] demonstrated that fluctuation effects play an important role at the level of single particle properties, namely fluctuation-induced formation of superconducting droplets in the normal state leads to the depletion in the density of states that manifests as a pronounced zero-bias anomaly revealed by the tunneling experiments. The delicate interplay between paraconductivity and density-of-states effects is instrumental in explaining transport anomalies as observed in the granular superconductors. These early studies set the stage for the new and fruitful field of research that spanned over many decades and were recently summarized in the monograph by Larkin and Varlamov [19]. What makes superconducting fluctuations so pronounced in experiments and interesting from the theoretical point of view is their strong dependence on temperature, magnetic field, frequency of external drive parameter, etc. In practice these characteristic features allow one to separate fluctuations from the other competing effects contributing to transport, and use 2 them to extract important information about the microscopic parameters of a material which are not accessible by other means of measurement. In theory, studying fluctuation effects advances our understanding of the underlying origin of superconductivity as a macroscopic quantum phenomenon. This becomes especially critical in the cases when this origin is in fact not completely known, like in the case of high-temperature cuprate superconductors or heavy-fermion superconductors. Furthermore, many ideas of the theory of superconducting fluctuations have been extensively used in other branches of condensed matter physics, e.g. in developing ideas of quantum criticality. For instance, superconducting fluctuations at zero temperature near the upper-critical field, Hc2 , give perhaps the simplest example of a quantum critical metal. A different kind of fluctuation effect that attracted tremendous attention over the span of last several decades is mesoscopic fluctuations [20–23]. This phenomenon occurs due to quantum mechanical interference between coherent electrons backscattering off random disorder potential. It manifests as completely reproducible oscillatory patterns of conductance as a function of magnetic field or the gate voltage with universal amplitude in the units of conductance quantum e2 /h. Since disorder substantially broadens the region of superconducting fluctuations, one may be intrigued by the question of whether there exists a parameter range within which superconducting fluctuations coexist with mesoscopic fluctuations, and what their interplay will give for observable quantities. In particular, one may want to investigate the ultimate fate of universality of conductance fluctuations in the presence of strong superconducting correlations. This set of questions defines the main theme of the present dissertation. Our focus in on thermomagnetic transport in mesoscopic superconductors. Thermomagnetic effects are difficult to measure and difficult to calculate theoretically. The mere difficulty of the problem makes it to be of fundamental importance. 3 In metals, thermomagnetic effects are usually small because of strong cancellation of currents generated by electronic excitations above and below the Fermi level – the electron-hole asymmetry is at the heart of the nonvanishing thermomagnetic response. Because of this compensation property, thermomagnetic effects are very sensitive to the characteristics of the electronic spectrum, presence of impurities, and peculiarities of scattering mechanisms. The inclusion of many-body interaction effects, such as electron-phonon renormalization, electron-electron scattering, drag effects, etc., adds a completely new level of complexity to the problem of calculating thermomagnetic kinetic coefficients. The observation that the collosal Nernst signal – the electric field, Ey , generated as a response to a transverse temperature gradient, ∇x T , in the presence of a perpendicular magnetic field, H – is mediated by superconducting fluctuations in the vicinity of transition triggered much interest to these transport phenomena. On the technical side there is a well defined way how one can compute response functions based on the diagrammatic methods of many-body theory in condensed matter physics [24]. These methods are extremely powerful and, in the context of superconducting and mesoscopic fluctuation effects, can be reduced to the summation of certain classes of diagrams. The diagrammatic technique is especially suited for problems containing a small parameter when the whole treatment simplifies to the summation of ladder type diagrams. Importantly for our study, mesoscopic and superconducting fluctuations are controlled by the same parameter – inverse dimensionless conductance – which allows us to treat these effects on equal footing within the same formalism. This dissertation is organized as follows. In Chapter-II we provide brief but yet sufficient introductory discussion for the foundations of mesoscopic physics of conductance fluctuations. We devote special attention to the mesoscopic effects in supeconducting systems. 4 After reviewing existing literature we place our work in the context of existing studies. Next we elaborate on the technical prerequisites and set the stage for the calculation of conductivity and transverse thermomagnetic power. We demonstrate that mesoscopic fluctuations proliferate in the presence of superconducting pairing correlations and that the universality of this phenomenon breaks down. We then relate our results to available experimental findings. In Chapter-III we concentrate on transport effects in thin superconducting films in the in-plane magnetic field. In such systems, superconductivity is limited either by orbital effects or by spin Zeeman effects. Depending on the film thickness and electron mean free path, one effect dominates the other and we consider both scenarios. In either case, the phase diagram in the field-temperature plane is interesting and we discuss fluctuation effects along the entire transition line including the most interesting quantum critical points. Despite certain similarity of the phase diagram, the underlying nature of fluctuation effects is conceptually different. The most remarkable feature of fluctuations in the Zeeman case (so called paramagnetically or equivalently Pauli limited superconductors) is that fluctuation-induced transport is mediated by virtual Cooper pairs in the quantum limit because of the Zeeman gap in the excitation spectrum at low temperatures. Chapter-III structurally is similar to Chapter-II. We review the history of studies on the subject of Pauli limited systems, discuss more recent tunneling results, and carry out calculation for electrical and thermomagnetic effects. Various technical aspects of this work are delegated to multiple appendices. 5 Chapter 2 Thermomagnetic transport in mesoscopic superconductors 2.1 Background and motivation Universality of conductance fluctuations (UCF) is the hallmark of mesoscopic physics [20– 23]. This phenomenon emerges from the quantum coherence of electron trajectories and is sensitive to changes in external magnetic field or gate voltage. At temperatures below the Thouless energy, T < ET h , which is related to the inverse dwell time for an electron to diffuse across the sample ET h = D/L2 , variance of conductance fluctuations saturates to the universal value of the order of conductance quantum, ∼ e2 /h, as long as characteristic sample size L is smaller than dephasing length, L < Lφ . Interaction effects in normal metals barely change the magnitude and universality of conductance fluctuations although they are crucially important in determining temperature dependence of dephasing effects, and in particular Lφ [25]. Robustness of UCF can be rooted to random matrix theory description of Wigner-Dyson statistics of electron energy levels in disordered conductors. Indeed, in 6 the Landauer picture of transport across a mesoscopic sample, conductance is given by e2 /h times the number of single particle levels within the energy strip of the width of the Thouless energy. While the average number of such levels depends on the dimensionality, random matrix theory predicts that their fluctuations are universally of the order of one [26, 27]. When superconductivity is induced at the boundary of the mesoscopic sample via the proximity effect, UCF remain intact [28, 29]. The magnitude of oscillations changes only by a numerical factor of the order of unity, with a value depending on the underlying symmetry class [30–32]. Things get quantitatively different if superconducting correlations are induced at the bulk of the sample. Experimentally this is achieved by tuning superconducting systems to the vicinity of the critical temperature Tc or the superconductor-insulator transition. There exists compelling evidence from measurements such as those in twodimensional granular arrays [33, 34], sub-micron scale superconducting cylinders [35], and quantum wires [36, 37], that mesoscopic oscillations could become anomalously large, sometimes reaching the level of ∼ 104 × e2 /h. These observations seemingly imply that the role of mesoscopic effects proliferate in the presence of superconducting correlations. Theoretical studies devoted to various aspects of mesoscopic fluctuations in superconductors cover a diverse range of topics. These works include mesoscopic effects on the Josephson current [38–42], upper critical field [43,44], critical temperature [45], condensation energy and glassy phase transitions, [46,47] persistent currents [48–51], density of states, gap fluctuations and level statistics [52–56], and also some transport properties [57, 58]. In the recent years, measurements of the Nernst-Ettingshausen effect and the diamagnetic response in superconductors [59, 60], including high-Tc [61–69] and heavy-fermion systems [70–72], attracted tremendous attention and triggered a flood of theoretical works [73–83]. Our motivation is to study mesoscopic effects on the thermomagnetic transport in superconductors at the onset 7 of Tc where pairing correlations due to preformed Cooper pairs are profoundly important. We find that in a parametrically wide temperature region, g −1 < (T − Tc )/Tc 1, where g is the dimensionless conductance, amplitude of mesoscopic fluctuations in the diagonal component of the conductivity tensor σxx and transversal component of the thermomagnetic tensor αxy become parametrically bigger than their bare values. 2.2 Foundations of UCF In the multiple scattering diffusive regime, the quantum mechanical interference effects associated with coherent backscattering on impurities modify the average value of electrical conductivity. Similarly, correlation functions of the single-particle density of states are affected by the same coherence effect. In what follows, we discuss the second moment of these physical quantities which will lead us to the universality of conductance fluctuations. The point of departure is the Einstein formula (throughout the text we use unites = kB = 1) σαβ (ε) = 2e2 νd (ε)Dαβ (2.1) which relates conductivity σαβ to the diffusion constant Dαβ and density of states νd . We introduce the mean for conductivity σ ¯ and sample specific variance δσ = σ − σ ¯ . Then the product δσ(ε)αβ δσγδ (ε ) = σ02 δαβ δγδ δνd (ε)δνd (ε ) δDαβ (ε)δDγδ (ε ) + 2 ν0 D2 (2.2) contains two contributions related to the density of states fluctuations and the diffusion coefficient fluctuations, while the cross-correlation between ν and D is absent. Here ν0 , D, and 8 σ0 are bare disorder-averaged density of states, diffusion constant, and Drude conductivity respectively, while angular brackets . . . stand for the disorder average of the irreducible correlation function. These correlation functions can be computed by using a conventional Diffuson-Cooperon diagrammatic technique [20–23]. The total conductivity fluctuation can be written in the form δσαβ (ε)δσγδ (ε ) = σ02 [δαβ δγδ Kν (ε − ε ) + (δαγ δβδ + δαδ δβγ )KD (ε − ε )] , Kν (ω) = δνd (ε)δνd (ε ) δ2 = ν02 2βπ 2 Re q 2 KD (ω) = δD(ε)δD(ε ) δ = 2 D 2βπ 2 1 2 Dq − iω 1 q [(Dq 2 )2 + ω2] (2.3) 2 , (2.4) , (2.5) where ω = ε − ε , δ = 1/νd Ld is the mean level spacing and coefficient β captures the symmetry class. For the time reversal invariant system β = 1 (orthogonal ensemble), while β = 2 if it is not (unitary ensemble). To translate conductivity fluctuations into the conductance fluctuations at finite temperature one uses g = σLd−2 (assuming cubic conductor of size L) and g(T ) = g(ε)∂ε f dε, where f (ε) is the Fermi distribution function, so that δgαβ (ε)δgγδ (ε ) = δαβ δγδ G2ν (T ) + (δαγ δβδ + δαδ δβγ )G2D (T ), G2ν (T ) 4 = β e2 ET h π 4 = β e2 ET h π G2D (T ) F (z) = 2 2 dω F (ω/2T ) 2T dω F (ω/2T ) 2T Re q 1 2 Dq − iω 1 q (Dq 2 )2 + ω2 (2.6) 2 , , z coth(z) − 1 . sinh2 (z) (2.7) (2.8) (2.9) The sensitivity of conductance fluctuations to a variation of the Fermi energy shows up in the temperature dependence of the correlations. At low temperatures T 9 ET h , it is sufficient to take ω → 0 limit in the q-summations of G2ν(D) functions since F (ω/2T ) is sharply peaked at zero frequency. In this case both contributions become equal. As an example, suppose that the sample is connected to leads along the x-direction and isolated in the other directions. The leads play the role of a reservoir, and the boundary conditions in that direction correspond to an absorbing wall. The diffusion modes in this direction are thus quantized as qx = nx π/L with nx = 1, 2, . . .. In the other directions, the boundary conditions are those of hard walls which implies the same quantization of diffusion modes with the contribution of the mode ny,z = 0 added. This results in the q-summation in above expression in the form 2 nx =0,ny ,nz (nx + n2y + n2z )−2 . As a result, in the zero temperature limit (restoring Planck’s constant ) 4bd δgαβ (ε)δgγδ (ε ) = βπ 6 e2 2 (δαβ δγδ + δαγ δβδ + δαδ δβγ ), (2.10) with b1 = π 4 /90, b2 ≈ 1.51, and b3 ≈ 2.52. This formula shows that conductance fluctuations do not depend on the strength of disorder since the diffusion constant dropped out from the final expression: the fluctuations are said to be universal. They depend only on the sample geometry and time-reversal symmetry. At high temperatures T ET h , one finds an algebraic decay of fluctuations ∝ (ET h /T )p with a power exponent depending on dimensionality of the system. The significant dependence of the conductance fluctuations on space dimensionality is a consequence of the diffusive nature of electronic transport. 10 2.3 Qualitative considerations We proceed by discussing how results of the previous section will change in superconducting systems close to transition. It has been emphasized early on [43,46] that quantum interference mesoscopic effects may lead to a formation of superconducting droplets that nucleate prior to transition of the whole system. Above Tc there are also thermally induced superconducting fluctuations [19] that are known to be crucially important in describing transport properties. One thus expects that the combined effect of two fluctuation mechanisms may have interesting implications for the kinetic properties of superconductors. Indeed, the probability amplitude of the fluctuations in the pairing gap ∆ is controlled by the competition of Cooper pair condensation energy and entropy, and can be estimated from the Ginzburg-Landau functional. The condensation energy exhibits mesoscopic fluctuations with the amplitude ∝ 1/g and the correlation radius of the order of thermal length ∼ LT = coincides with the superconducting coherence length ξ = D/T . Near Tc the latter D/Tc . On the other hand, ther- mal superconducting fluctuations are susceptible to the Ginzburg-Landau correlation length ξT = ξ Tc /(T − Tc ) LT so that mesoscopic fluctuations are almost local with respect to superconducting fluctuations, and thus should be summed randomly from different blocks of the size ξ. For the sample size L ξ, the above consideration leads to an estimate of the mesoscopic fluctuations of the critical temperature δTc /Tc ∝ (1/g)(ξ/L)(4−d)/2 , where d is the dimensionality of the system. The response functions in superconductors near Tc are governed by the dynamic pair susceptibility propagator L(ω, q) ∝ (Dq 2 + T − Tc + |ω|)−1 for a given mode at finite frequency ω and wave vector q, which acquires mesoscopic fluctuations δL ∝ L2 δTc . Even though the whole effect seems to be small, as it scales inversely proportional to conductance, g 1, the singular nature of L at T − Tc 11 Tc as {q, ω} → 0 translates into the substantial temperature dependence of kinetic coefficients. This is the microscopic reason why the universality of mesoscopic effects breaks down in the case of fluctuating superconductors. Following Ref. [58] we elaborate on this point in the context of conductivity fluctuations and then carry out microscopic verification of this result with further extension to thermomagnetic transport. At the qualitative level the conductivity enhancement near Tc due to superconducting fluctuations can be estimated as δσ ∼ σ ∆q ∆−q τq ∼ q 1 g Tc T − Tc (4−d)/2 , (2.11) where d is dimensionality. In essence this estimate is obtained from the Drude formula for conductivity but applied for fluctuation-induced Cooper pairs. Indeed, the average of pairing gap fluctuations T 1 ν Dq 2 + T − Tc ∆q ∆−q (2.12) measures the average concentration of preformed pairs, while their life-time τq Dq 2 1 + T − Tc (2.13) at a given mode q is nothing but Ginzburg-Landau relaxation time. Because of its mesoscopic origin, discussed above, fluctuations of Tc lead to a giant mesoscopic fluctuations of the order parameter field δ∆q δ∆−q , where double-brackets stand for the thermal and disorder average. Near the critical point δ∆q δ∆−q ∆q ∆−q ∝ 1 g ξ L 12 (4−d)/2 Tc . Dq 2 + T − Tc (2.14) This yields the estimate for mesoscopic fluctuations of conductivity from squaring Eq. 2.11 in the form δσ 2 σ2 ∼ 1 g4 ξT L 4−d Tc T − Tc (8−d)/2 . (2.15) Despite the large factor in the denominator, this expression is substantially more singular than Eq. (2.11) in terms of the temperature dependence near Tc . For concreteness, let’s concentrate on the two-dimensional case d = 2. Recall that perturbative treatment of the fluctuation effect breaks down beyond the Ginzburg region, so that at most we can allow T − Tc ∼ Tc /g when correction to conductivity from Eq. (2.11) becomes of the same order as a bare Drude conductivity. Remarkably, at that temperature scale, and assuming samples size of the order of coherence length L ∼ ξT , the amplitude of mesoscopic fluctuations in conductivity, as estimated from Eq. (2.15), exceeds the bare value by a large parameter √ g 1. From these considerations one can infer an estimate of expected fluctuations in thermo- magnetic response. This can be achieved by the following lines of reasoning. When subject to crossed electric and magnetic fields, the charged carriers acquire a drift velocity v¯x = cEy /H in the x direction (H is in z-direction). That would result in the appearance of a transverse current Jx = en¯ vx with respect to Ey . Under open circuit conditions, no current flows, and the drift of carriers is compensated by the spacial variation of the electric potential ∇x ϕ = −Ex = (enc/σ)(Ey /H). Because of electroneutrality, this generates the gradient of the chemical potential ∇x µ(n, T ) + e∇x ϕ = 0, which ultimately corresponds to the appearance of the temperature gradient ∇x T = (dµ/dT )−1 ∇x µ along the x direction. Hence, the Nernst coefficient ν = Ey /H(−∇x T ) can be expressed in terms of the full temperature derivative of the chemical potential ν = (σ/e2 nc)(dµ/dT ). This result can be checked for a 13 Figure 2.1: The layout of the Nernst experiment: by applying a temperature gradient (−∇T ) in the presence of magnetic field, an electric field is generated. degenerate electron gas; the chemical potential µ(T ) = µ0 − (π 2 T 2 /6)(d ln ν/dµ) reproduces the Sondheimer formula ν = (π 2 T /3mc)(dτ /dµ), where τ is the elastic scattering time. For the preformed Cooper pairs, dµ/dT = −1 so that ν ∝ σ and fluctuations in ν should follow that of conductivity. This conjecture will be verified with a microscopic calculation in Sec 2.6. 2.4 Definitions and assumptions We start with the definition of kinetic coefficients concentrating on the linear response analysis. The electric Jetr and heat Jhtr transport currents are related to the electric field E and temperature gradient ∇T by the following matrix    Jetr Jhtr       σ α  E  =  , α κ −∇T 14 (2.16) where σ is the electric conductivity tensor , α and α are the thermoelectric tensors (α = T α due to Onsager relations), and κ is the thermal conductivity tensor. Applying the open circuit conditions to Eq. (2.16) (see Figure 2.1 for the layout of the experiment), we have E/∇T = σ −1 /α where the inverse of σ is  σ −1 =  1  σyy −σxy   . det |σ| −σyx σxx Multiplying it by α, taking the corresponding element of the matrix, the Nernst coefficient is expressed in terms of the conductivity and thermoelectric tensors ν= Ey 1 αxy σxx − αxx σxy = . 2 + σ2 (−∇x T )H H σxx xy (2.17) An important aspect of the calculation of the transverse thermoelectric response αxy , discussed in detail in Ref. [84], is the need to account for bulk magnetization currents. This issue arises because the microscopic electric and heat currents, as calculated below by the Kubo formula, are composed of transport and magnetization currents Je = Jetr + Jemag , Jh = Jhtr + Jhmag . (2.18) The magnetization currents are currents that circulate in the sample and do not contribute to the net currents which are measured in a transport experiment. On the other hand, they do contribute to the total microscopic currents, and it is thus necessary to subtract them from the total currents to obtain the transport current response. In the presence of an 15 applied electric field, it was shown in Ref. [84] that the magnetization current is given by Jhmag = cM × E, (2.19) where M is the equilibrium magnetization (in the absence of the electric field). It then follows that the transverse thermoelectric response is given by αxy = − Jh cMz cMz + =α ¯ xy + . Ex T T T (2.20) It is apparent from the above expression that αxy is obtained by subtracting the result of two independent calculations: the response of the total current to the applied electric and magnetic fields, and the magnetization currents as derived from the equilibrium magnetization. Therefore, we need to know magnetic susceptibility M = χH, which will be computed diagrammatically along with α ¯ xy . Importance of the magnetization contribution to αxy in the context of superconducting fluctuations was elaborated by Ussishkin [74]. The calculations will be carried out assuming particle-hole symmetry, i.e., neglecting any contributions which arise due to asymmetry around the Fermi surface in properties such as the density of states or scattering time. Particle-hole symmetry implies that σxy = αxx = κxy = 0, and therefore the general expression for the Nernst coefficient Eq. (2.17) greatly simplifies. The conventional result for αxy in the normal metallic state (so-called quasiparticle contribution) also vanishes in this limit. However, it has been emphasized [74] that this result is not required by the symmetry, and will not necessarily hold when additional processes, such as superconducting or mesoscopic fluctuations, are taken into account without breaking the particle-hole symmetry. Another simplification of our analysis comes from assuming s-wave 16 G   1/2πντ   λ   Π   Figure 2.2: Building blocks of the diagrammatic technique. Straight lines represent disorderaveraged single-particle Green’s functions. The straight-dashed lines represent single impurity lines that carry an overall factor 1/2πνd τ in the diagrams. The vertex function λ(ε, ε , q) is drawn in the ladder approximation, while diagrams with the crossed impurity lines yield parametrically smaller contributions in 1/g 1. The polarization operator Π(ω, q) (“bubble” diagram) is also presented in the main ladder approximation. symmetry of the superconducting order parameter. In the context of the high-temperature superconductors, it is of interest to consider also the case of d-wave symmetry in a similar approach. 2.5 Technical prerequisites In our calculations we closely follow Ref. [19] for all the basic definitions and notations of the diagrammatic technique. Graphical rules for constructing Feynman diagrams in the context of transport in the disordered systems are depicted in Fig. 2.2. We start from the 17 disorder averaged single particle Green’s function, which in the energy and momentum (ε, p) representation reads G(ε, p) = 1 , ε − ξp + 2τi sgnε ξp = p2 /2m − εF , (2.21) where τ is the elastic scattering time on impurities. In the parameter T τ one should distinguish three different transport regimes in fluctuating superconductors: the diffusive scattering T τ Tτ 1, the ballistic limit 1 Tτ Tc /(T − Tc ), and the ultra-clean limit Tc /(T − Tc ). We will concentrate on the diffusive case exclusively, which is also mathematically simpler. In the ballistic case, fluctuation effects become strongly non-local in space, while diffusive impurity scattering makes response functions isotropic and local. The basic building block of the Cooper ladder is two Green’s functions connected by one impurity line, which we denote by Γ. Each impurity line brings a factor 1/2πνd τ , where νd is the density of states. In two-dimensional case ν2 = m/2π, so that Γ(ω, q) = 1 2πνd τ dd p R G (ε+ , p+ )G A (ε− , p− ), (2π)d (2.22) where superscript R(A) stands for the retarded (advanced) component of Eq. (2.21). We used shorthand notations ε± = ε ± ω/2 and p± = p ± q/2. Integrating above over ξp and expanding to the first order in ωτ, Dq 2 τ 1, Γ(ω, q) = 1 + τ (iω − Dq 2 ), (2.23) where D = vF2 τ /2 is the diffusion coefficient (here and in the remaining parts of the thesis we concentrate of the two-dimensional case unless otherwise explicitly mentioned). Having 18 Γ one can sum up the whole series of ladder diagrams to obtain the Cooperon propagator C(ω, q) = 1 1 1 1 1 (1 + Γ + Γ2 + . . .) = = . 2 2 2πνd τ 2πνd τ 1 − Γ 2πνd τ Dq − iω (2.24) In practice it is also useful to have the same expression but rewritten in the Matsubara representation with two different energies ε1,2 running through each of the Green’s function lines of the Copper ladder C(ε1 , ε2 , q) = 1 θ(−ε1 ε2 ) , 2 2 2πνd τ Dq + |ε1 − ε2 | (2.25) where θ(x) is the unity step-function. The vertex function dressed with impurities is represented as follows (see Fig. 2.2) λ(ε1 , ε2 , q) = 2πνd τ C(ε1 , ε2 , q). (2.26) The propagator of superconducting fluctuations reads L−1 (ω, q) = −λ−1 sc + Π(ω, q), (2.27) where λsc is the coupling constant in the particle-particle (Cooper) channel and Π(ω, q) = 4πνd T εn 1 2εn + |ω| + Dq 2 >0 (2.28) is the polarization operator (summation goes over the fermionic Matsubara frequency εn = (2n + 1)πT , while ωm = 2πmT corresponds to bosonic frequency). With these ingredients 19 q, k = + = + + p, p q, + n q, k k n Figure 2.3: The Dyson equation for the fluctuating propagator L(ω, q) which represented graphically as the wavy line and computed in the ladder approximation. Solid lines represent one-electron Green’s function, greay triangle is the impurity dressed effective vertex, while each cross between two Green’s functions is associated with the electron-electron coupling constant −λsc . at hand one can explicitly calculate the pair propagator which takes the form (see Fig. 2.3) L−1 (ω, q) = −νd ln T +ψ Tc 1 Dq 2 + |ω| + 2 4πT −ψ 1 2 , (2.29) where ψ(x) is the digamma function, and the critical temperature is expressed through the coupling constant Tc = (2γE ωD /π)e−1/νd λsc with γE = 1.78 being the Euler constant and ωD being the Debye frequency, which cuts logarithmically divergent summation at nmax = ωD /2πT in the polarization operator. At small frequencies and momenta (the most relevant limit for most of the further applications) one can expand digamma functions at small argument {Dq 2 , ω}/T 1 and use ψ (1/2) = π 2 /2 to obtain a simpler expression L(ω, q) = − 1 1 . νd ln(T /Tc ) + πDq 2 /8T + π|ω|/8T (2.30) For the purpose of calculating kinetic coefficients, the crucial objects are current vertices. In particular, the electric current vertex function consists of three Green’s functions with 20 two impurity ladders Bie (ωm , Ωk , q) = 2eT × p εn λ(εn + Ωk , ωm − εn , q)λ(εn , ωm − εn , q) vi (p)G(εn + Ωk , p)G(εn , q)G(ωm − εn , q − p), (2.31) where Ωk is the external frequency (which will be eventually set to zero in the Kubo formula to get dc-transport coefficient). Being a function of three frequencies and momentum, this vertex is fairly complicated, however in the classical region of fluctuations near Tc its evaluation is straightforward. The essential simplification comes from the separation of energy scales. Bosonic modes are pinned to the energy set by the pole structure of superconducting propagator Dq 2 ∼ |ωm | ∼ T − Tc , which can be readily seen from Eq. (2.30). At the same time all fermionic modes are governed by the temperature |εn | ∼ T . Near the transition T − Tc T we can evaluate the vertex function by setting all the bosonic frequencies to zero. As stated above, within the linear response Kubo analysis the external frequency is also set to zero. As a result, we approximate Bie (0, 0, q) = 2eT εn λ2 (εn , −εn , q) p vi (p)G 2 (εn , p)G(−εn , q − p). Transforming now momentum summation into the integral p . . . → νd (2.32) dξp . . . ϑ , where averaging goes along the Fermi surface, and using Eq. (2.21) we get +∞ Bie (0, 0, q) = −2eνd T εn λ2 (q, εn , −εn ) −∞ dξp (i¯ εn − ξp )2 vi (p) i¯ εn + ξp − vj (p)qj , (2.33) ϑ where in addition we used ξq−p ≈ ξp − v · q and abbreviated ε¯n = εn + 1/2τ . From here we need only the leading small-q part of the vertex. Expanding the denominator to the linear 21 order in q and using Eq. (2.26), the above equation transforms into Bie (0, 0, q) = −2eνd T vi vj qj |2¯ εn |2 |εn |2 ϑ εn +∞ −∞ (ξ 2 dξ . + ε¯2n )2 (2.34) The remaining ξ-integration, followed by a εn -summation, can be completed in the closed form in terms of the digamma function Bie = 2eBi (q), Bi (q) = νd Dτ qi ψ Focusing on the diffusive case only T τ 1 1 + 2 4πT τ −ψ 1 2 − 1 ψ 4πT τ 1 2 . (2.35) 1, the above expression simplifies even further Bi (q) = −2νηqi , η = πD/8T . (2.36) We will also need the heat current vertex function Bih (ωm , Ωk , q) = T εn i(2εn + Ωk ) λ(εn + Ωk , ωm − εn , q)λ(εn , ωm − εn , q) 2 × p vi (p)G(εn + Ωk , p)G(εn , p)G(ωm − εn , q − p). (2.37) Under the same approximations as specified above one finds Bih = 2iνωm ηqi = −iωm Bi (q) . (2.38) It should be stressed that such simple expressions for the electrical and heat current vertices are only possible to obtain near Tc . We will see later in the text that in the quantum low temperature regime the above calculation has to be revisited, and frequency dependence of 22 Figure 2.4: Feynman diagrams for the main fluctuation-induced corrections to the conductivity. In the first row we show the Aslamazov-Larkin diagrams (left) and the Maki-Thompson interference diagram (right). In the second row we show two density of states diagrams. both B e and B h (omitted here) will play a crucially important role. The above results however are sufficient to make further progress in addressing the classical region of superconducting fluctuations. As has been discussed in the introductory chapter, the superconducting fluctuations enhance the conductivity above Tc due to so called Aslamazov-Larkin [10] and Maki-Thompson [11,12] contributions as well as density of states effects [18], which are less important for the conductivity unless one studies granular systems. A similar identification of the microscopic contributions applies to other transport coefficients. In the case of the transverse thermomagnetic response, the leading order contribution to αxy is due to the Aslamazov-Larkin (AL) diagrams alone. The contributions of the Maki-Thompson (MT) and density of states (DOS) diagrams are less divergent as T → Tc ; see Fig. 2.4 for the identification of diagrams. Within the linear response analysis, diagonal Aslamazov-Larkin conductivity is deter- 23 ee mined from the following current-current Kubo kernel Kxx : 1 ee (Ω)]R , Im[Kxx Ω→0 Ω σxx = lim ee (Ωm ) = 4e2 T Kxx qωk Bx2 (q)L(ωk , q)L(ωk + Ωm , q) , (2.39) ee ee R (Ωm ) as it is analytically continued ] indicates the retarded component of Kxx where [Kxx from the discrete Matsubara frequencies into the entire complex plane iΩm → Ω + i0. The Aslamazov-Larkin contribution to the transversal thermoelectric coefficient is found from eh : the mixed electric current-heat current Kubo response function Kxy α ¯ xy = H 1 eh lim Re[Kxy (Q, Ωm )]R , cT Ω,Q→0 ΩQ (2.40) where eh Kxy (Ωm , Q) = −4e2 T q Bx (q)By2 (q) (iωn + iΩm /2) (2.41) ωn × [L(ωn , q − Qx )L(ωn , q)L(ωn + Ωm , q) + L(ωn , q)L(ωn + Ωm , q)L(ωn + Ωm , q + Qx )] We finally define magnetic susceptibility from the equilibrium magnetization M = χH. Diagrammatically, it may be calculated to linear order in H by considering the current response to a magnetic field at a finite wavevector Q [85]: χµν = − where αβγ 4e2 c2 αγµ βκν T ω,q xˆγ xˆκ L2 (ω, q)Πα (ω, q)Πβ (ω, q), (2.42) is the anti-symmetric Levi-Civita unity tensor, and xˆ is thecoordinate operator in momentum representation. Below we will consider only the isotropic case χµν = χδµν . With these ingredients at hand we proceed with the calculation of fluctuation-induced corrections 24 2 2 to σxx , αxy , and χ, as well as their mesoscopic correlation functions δσxx , δαxy , and δχ2 along with the possible cross-correlators. 2.6 Mesoscopic Nernst effect The first step in the derivation of desired kinetic coefficients within the Kubo formalism requires consideration of the discrete sums over Matsubara frequencies, such as sums in the response kernels of Eqs. (2.39) and (2.41). Such summations over bosonic frequencies can be conveniently done with the help of closed contour integration in the complex plane by using the following formula f (ωm ) = T ωm 1 4πi dωf (−iω) coth ω . 2T (2.43) Applying this to Eq. (2.39) one finds ee Kxx (Ωm ) = 4e2 q Bx2 (q) 1 4πi dωL(−iω, q)L(−iω + Ωm , q) coth ω . 2T (2.44) The propagators under the integral have breaks of analyticity in the complex plane of ω at Im(ω) = 0 and Im(ω) = −Ωm , so that the integration contour has two branch cuts along these lines. We delegate details of this integration, followed by an analytical continuation, to Appendix-A and present here only the result for the conductivity correction σxx = e2 πT +∞ q Bx2 (q) −∞ dω [ImLR (ω, q)]2 . sinh (ω/2T ) 2 25 (2.45) For the further integrations we define the following dimensionless units: x = ηq 2 , y = πω/8T , and = ln(T /Tc ) ≈ (T − Tc )/Tc . In these units, the interaction propagator and vertex function become ImLR (x, y) = − y 1 , ν ( + x)2 + y 2 Bx2 (x) = 4ν 2 ηx cos2 φ , (2.46) and integrations transform into ∞ 2π q → dφ 0 0 +∞ dx , 8π 2 η −∞ dω πT → 2 2 sinh (ω/2T ) +∞ −∞ dy , y2 (2.47) where we expanded sinh y ≈ y since major contribution comes from the range of parameters {x, y} ∼ 1. Combining these definitions together, rescaling y → ( + x)y first and then x → x, the latter expression transforms into σxx e2 = 2 4π ∞ 2π 2 dφ cos φ 0 0 xdx (x + 1)3 +∞ −∞ dy , (1 + y 2 )2 (2.48) with the three integrals equal to π, 1/2, and π/2, respectively, and thus with the final result σxx = e2 Tc . 16 T − Tc (2.49) So far we have only reproduced the celebrated result of Aslamazov and Larkin [10]. Our immediate task is to generalize this result for mesoscopic effects associated with fluctuations of pair propagator. For that matter we return to Eq. (2.45), take its variation, square the 26 result, and average it over the realization of disorder potential. We thus find 2 = δσxx 4e4 π2T 2 +∞ q1 q2 Bx2 (q1 )Bx2 (q2 ) R −∞ dω1 dω2 sinh (ω1 /2T ) sinh2 (ω2 /2T ) 2 R ImL (ω1 , q1 )ImL (ω2 , q2 ) ImδLR (ω1 , q1 )ImδLR (ω2 , q2 ) . (2.50) In order to calculate the irreducible correlation function of the pairing susceptibility, one has to draw two diagrams for L and connect their diffusive parts by impurity lines. Such construction involves four colliding Diffuson-Cooperon ladders and, on a technical level, requires computation of four- and six-order Hikami boxes [86], see Fig. 2.5 for the illustration. Some of these diagrams have been studied before [43,45–47,58] and we invoke that knowledge for our purposes. In particular we use δL R(A) R(A) (ω1 , q1 )δL Aν 2 (ω2 , q2 ) = 2d g LT L 2 [LR(A) (ω1 , q1 )]2 [LR(A) (ω2 , q2 )]2 . (2.51) Precise calculation of the numerical factor A ∼ 1 is not of principal importance in a view of strong dependence of the whole expression on temperature in the low momentum and frequency limit (ω, q) → 0, which in a way defines T -dependence of transport coefficients. It is then straightforward to show that ImδLR (ω1 , q1 )ImδLR (ω2 , q2 ) = 4Aνd2 g2 LT L 2 ImLR (ω1 , q1 )ReLR (ω1 , q1 )ImLR (ω2 , q2 )ReLR (ω2 , q2 ). (2.52) 2 We take this back into the equation for δσxx , and introduce dimensionless x, y, variables, 27 Figure 2.5: Leading order diagrams for the irreducible correlator of mesoscopic disorderaveraged two pair-propagators. This averaging contains collisions of four diffusion or Cooper modes, and involves forth and sixth order Hikami boxes (internal impurity lines are implicit on diagrams). 28 as in the case considered above, to arrive at 2 = δσxx 2π Ae4 L2T π 4 g 2 L2 ∞ dφ1 dφ2 cos2 φ1 cos2 φ2 0 +∞ dx1 dx2 dy1 dy2 −∞ 0 x1 x2 (x1 + )(x2 + ) , [(x1 + )2 + y12 ]3 [(x2 + )2 + y22 ]3 (2.53) that after rescaling of integration variables gives the final result 2 δσxx = Ae4 32πg 2 ξT L 2 Tc T − Tc 3 . (2.54) Thus by microscopic analysis we have confirmed our earlier result Eq. (2.15) which was based on a qualitative considerations. We have already argued that interplay of superconducting and mesoscopic effects trigger giant fluctuations for samples L ξT at the temperatures T − 2 ∼ (e2 /g)2 (Tc /ET h )3 . ξT , fluctuations saturate to δσxx Tc ∼ Tc /g. For larger samples L We can build on this result to consider emergent mesoscopic fluctuations in the transversal thermoelectric coefficient. We start from Eq. 2.41 where we need only contributions linear in Q, which can be easily extracted by expanding the pair propagator and noticing that ∂L(ω, q) = −Bx (q)L2 (ω, q). ∂qx (2.55) eh Next we have to sum the resulting expression for Kxy (Ωm , Q) over the Matsubara frequency, as in the case of the conductivity calculation by contour integration in the complex plane. We find eh Kxy (Ωm , Q) = −4e2 Q q Bx2 (q)By2 (q) 1 4πi dω coth ω (ω + iΩm /2) 2T ×[L3 (ω, q)L(ω + iΩm , q) − L(ω, q)L3 (ω + iΩm , q)]. 29 (2.56) Qˆ x Bye q + Qˆ x, ω q, ω Qˆ x, Ω Bxe Byh Ω q + Qˆ x, ω + Ω q, ω Qˆ x, Ω Byh q + Qˆ x, ω + Ω q, ω + Ω Bxe Ω Bye Qˆ x eh Figure 2.6: The Aslamazov-Larkin diagrams contributing to Kxy (Ω). The wavy lines correspond to the fluctuation propagator L(q, ω); electric current vertices B e and heat current vertices B h are indicated in the figure along with running momenta and frequencies. 1 30 After further calculation (see analytical continuation that is done carefully in the AppendixA) we arrive at α ¯ xy = 4e2 H cπT R +∞ q Bx2 (q)By2 (q) 3 dω coth −∞ ω 2T R × [ReL (ω, q)] ImL (ω, q) + ReLR (ω, q)[ImLR (ω, q)]3 . (2.57) In the dimensionless units x = ηq 2 and t = πω/2T , and after the rescaling y → (x + )y and x → x, one finds at the intermediate step α ¯ xy = 16e2 H η cπ 3 ∞ 2π dφ cos2 φ sin2 φ 0 0 x2 dx (x + 1)4 +∞ dy −∞ 1 . (1 + y 2 )3 (2.58) After remaining integration, where each of the three integrals yields coefficient π/4, 1/3, and 3π/8 respectively, one finds α ¯ xy = Here H = Tc e ξT2 ∝ 2 2π H T − Tc (2.59) c/eH is the magnetic length and α ¯ xy has the same scaling with temperature as the conductivity. As shown by Ussishkin [74] the magnetization contribution has the same structural form but comes with the coefficient −1/3 instead of 1/2 so that αxy = α ¯ xy +cMz /T has an overall coefficient of 1/6. We can address now the mesoscopic fluctuations of αxy by taking the variation of Eq. (2.57), squaring the result, and averaging over the disorder realization with the help of the correlation function Eq. (2.51). In doing so we encounter quite a cumbersome expression with 2 several contributions to δαxy , but we make an observation that all the emergent terms have exactly the same scaling with temperature and differ from each other only by a numerical 31 coefficient of the order of unity. For brevity we present here one particular such term 2 δαxy =A e2 νd HLT cgT L +∞ dω1 dω2 coth −∞ 2 q1 q2 Bx2 (q1 )Bx2 (q2 )By2 (q1 )By2 (q2 ) (2.60) ω1 ω2 coth [ReLR (ω1 , q1 )]4 [ReLR (ω2 , q2 )]4 ImLR (ω1 , q1 )ImLR (ω2 , q2 ) 2T 2T and carry out the calculation up to a factor modulo one (we will absorb all the numerical factors into the redefinition of coefficient A). Since most relevant frequencies ω ∼ T − Tc are small compared to temperature we can approximate coth(ω/2T ) ≈ 2T /ω. Transforming the above into dimensionless variables 2 δαxy = e2 A g2 ξ 2 LT 2 HL 2 ∞ +∞ dx1 dx2 dy1 dy2 −∞ 0 x21 x22 (x1 + )4 (x2 + )4 , (2.61) ((x1 + )2 + y12 )5 ((x2 + )2 + y22 )5 followed by rescaling and integration, one finds 2 δαxy = e2 A g2 ξT 4 H LT L 2 Tc T − Tc 2 . (2.62) 2 2 , namely ∝ (T −Tc )−4 , has exactly the same temperature scaling as δσxx Interestingly δαxy that we already anticipated based on qualitative considerations in Sec. 2.3, and the above diagrammatic calculations provide the microscopic justification for our results. It remains to consider fluctuation-induced corrections to magnetic susceptibility and its mesoscopic fluctuations. From Eq. (2.42) we get for the Aslamazov-Larkin contribution [85] χ=− 16e2 T 3c2 ωm ,q Πx L3 (ωm , q) Πx Πyy − Πy Πxy , 32 (2.63) where derivatives of the polarization operator can be easily computed from Eq. (2.28) Πx,y = − πνD qx,y , 4T Πyy = − πνD , 4T Πxy = 0. (2.64) Already at this level, by simple power counting of integration variables, one can deduce that χ ∝ Tc /(T − Tc ). Consequently one expects that δχ2 will also scale with T − Tc in the same way as the conductivity and thermomagnetic coefficients. Indeed, 2 δχ =A e2 νd2 ηLT c2 gL 2 q1 q2 Bx2 (q1 )Bx2 (q2 ) dω1 dω2 coth ω1 ω2 coth 2T 2T ×Im[LR (ω1 , q1 )]4 Im[LR (ω2 , q2 )]4 , (2.65) which, as in the previous examples, reduces with standard steps to δχ 2 =A e2 D c2 g 2 LT L 2 Tc . T − Tc 4 (2.66) Finally, we will not delve into detailed calculation of the possible cross-correlation functions between different kinetic coefficients, and merely state here that all such correlations are of the same order and yield the same temperature dependence. 2.7 Summary The main results of this Chapter are expressions Eqs. (2.54), (2.62), and (2.66) for variances of different kinetic coefficients in mesoscopic superconductors. Because of the long-range phase coherence developing close to Tc , sample-specific mesoscopic fluctuations should be observable at large length scales. Similarly to normal samples, these fluctuations are sensi- 33 tive to magnetic field strength, impurity configuration, and gate voltage. However, in sharp contrast to the normal case, where such fluctuations are universal, interaction effects in the Cooper channel trigger a great amplification of fluctuations due to pairing correlations. This interplay of coherent impurity scattering and interactions leads to a spectacular example of quantum mesoscopic phenomena occurring at a macroscopic scale. Despite the fact that mesoscopic fluctuations are no longer universal, in the sense of random matrix theory classification, we have discovered a different kind of universality in the sense of temperature dependence, which was found to be consistently the same for all the considered kinetic coefficients. These calculations have been carried out for homogeneously disordered superconductors. Therefore, our results cannot be directly compared to the existing experimental findings where the samples were granular in their origin [33–36]. Granularity adds another parameter into the model – inter-grain conductance – which leads to a strong competition between Aslamazov-Larkin, Maki-Thompson, and DOS effects [87]. Nonetheless, the main features predicted by the theory should be present for inhomogeneously disordered superconductors as well. Indeed, the predicted sample-specific conductance fluctuations were observed experimentally in samples of macroscopic length, and only in a narrow temperature range in the immediate vicinity of Tc , consistent with the theory. The amplitude of the conductance fluctuations was found to greatly exceed that of the UCF in normal samples. It should be also emphasized that some other features accompanying giant mesoscopic effects, such as suppression of h/2e oscillations in cylindrical samples, negative mangetoresistance, and its asymmetry, can be also addressed within the same theoretical model. As of today we are unaware of experimental measurements of mesoscopic effects in thermomagnetic transport of superconductors, except for the measurements of magnetic susceptibility [88]. The 34 mesoscopic Nernst effect has been studied experimentally only in the non-superconducting systems [89]. Verification of the temperature scaling and the overall magnitude of the effect for mesoscopic fluctuations of the Nernst coefficient predicted here would provide an important test for our understanding of thermomagnetic transport phenomena in correlated systems. 35 Chapter 3 Transport anomalies in Pauli-limited superconductors 3.1 History of the subject According to the microscopic BCS theory [9] a magnetic field H extinguishes superconductivity. In the absence of spin-orbit interaction there are two basic mechanisms. The first one is the diamagnetic effect associated with the action of the field on the orbital motion of electrons forming a Cooper pair. The second, paramagnetic mechanism, is due to Zeeman splitting of the states with the same spatial wave function but opposite spin. In the former case, the estimate for the upper critical field follows from the condition Hc2 ξ 2 Φ0 , where Φ0 = hc/2e is the flux quantum. In contrast, Zeeman splitting destroys superconductivity at a different critical field Hz that follows from the condition Ez ∆, where Ez = gL µB H is the Zeeman energy, µB = e /2mc is the Bohr magneton and gL is the renormalized giro factor, while ∆ is the superconducting gap. The ratio between the two fields is Hz /Hc2 ∼ kF ∼ g kF is Fermi momentum and 1, where is the elastic scattering length. Thus, in bulk systems, the 36 Ez ∆0 2 √ 2 1 superheating 2−nd order QF supercooling T∗ Tc0 1 T Tc0 Figure 3.1: Above the tricritical point T ∗ the second order paramagnet to superconductor transition occurs along the (black) solid line obtained from Eq. 3.1. At T < T ∗ this line becomes a supercooling part of the hysteresis, and the dashed line is its superheating part. The latter is obtained following Ref. [95]. The grey shaded area with the critical point (0, ∆0 ) as its lowest corner bounded by the black dashed line marks the region of quantum fluctuations (QF). suppression of superconductivity is typically governed by the first diamagnetic mechanism. The situation changes in the case of restricted dimensionality. For example, in the case of a thin-film superconductor in a parallel field, the above ratio becomes Hz /Hc2 ∼ (kF )(d/ξ), which can be small provided that the film is thin enough d ξ/kF , such that spin effects dominate. The scenario of paramagnetically limited superconductivity has a long history that goes back to pioneering works by Clogston and Chandrasekhar [90, 91]. The first order phase transition from superconductor to paramagnet was predicted at the critical field approaching √ Ez = 2∆ at low temperatures. In practice, the measured film resistance follows a hysteresis loop [92–96] instead of a sharp first order transition, and the experimental phase diagram is qualitatively as in Fig. 3.1. At low temperatures, the system remains superconducting as 37 the field increases up to the superheating field; above the critical field, the film is trapped in a metastable state. At fields exceeding the superheating threshold the film becomes normal. When the field is reduced, the film stays normal down to the supercooling field Ezsc (T ), which corresponds to the zero binding energy of a Cooper pair. At T = 0, the normal state √ is metastable in the interval ∆ < Ez < 2∆ [97, 98]. One should ote that these papers also √ predicted spatially inhomogeneous state for 2 < Ez /∆ < 1.52. We neglect such possibility in this work. In this Chapter, we study the transport properties near the supercooling field, which is determined by the equation [99] ln Tc Tc0 =ψ 1 2 − Re ψ 1 iE sc + z 2 4πTc (3.1) similar to that in the theory of paramagnetic impurities [100]. Here ψ is the digamma function and Tc0 = Tc (H = 0) is the critical temperature in the absence of a magnetic field. The zero temperature solution of Eq. 3.1, Ezsc (0) = ∆, defines the quantum critical point (QCP), which is the premier interest of our study. Since in this case the critical parameter depends on H, it allows for a well controlled exploration of the QCP and its vicinity by varying the applied magnetic field. In the scenario when orbital effects dominate the phase diagram is determined by a similar equation ln Tc Tc0 =ψ 1 2 −ψ 1 α + 2 4πTc (3.2) where pair-breaking parameter α ∼ τd−1 can be deduced from the time-scale of the loss of phase coherence of the Cooper pairs in the presence of the finite magnetic field τd−1 ∼ DH 2 d2 /Φ20 . This transition is of the second order throughout the entire H − T line, see 38 1.0 α/Tc0   0.8 Normal  Phase   0.6 Superconduc/ng   Phase   0.4 0.2 0.0 0.0 0.2 0.4 0.6 0.8 1.0 T Tc0 Figure 3.2: Phase diagram of a superconducting thin film in a parallel magnetic field parametrized by pair breaking parameter α = D(eHd)2 /6 due to orbital mechanism. Tc0 = Tc (H = 0) is the critical temperature in the absence of a magnetic field. At T = 0 the superconductivity breaks down at the critical value αc = πTc0 /2γE , where ln γE ≈ 0.577 is the Euler constant. Fig. 3.2. Despite apparent similarity between the two cases at the level of the phase diagram, we will demonstrate that the microscopic nature of the fluctuation-induced transport near respective QCPs is conceptually different. Equation (3.2) implicitly defines the critical temperature Tc as a function of the pairbreaking parameter α. A full analytic solution of this equation in terms of Tc (α) is not possible, but asymptotic expressions can be easily extracted. For α Tc0 we can expand the digamma function ψ(x + 1/2) to first order in x and thus obtain ln Tc0 Tc (α) − π2 α = 0. 2 4πTc (α) (3.3) To first order in α this yields Tc (α) ≈ Tc0 − πα/8. The expansion for large values of α is a 39 bit more involved, because there exists a critical value of the pair-breaking parameter αc at which the critical temperature vanishes nonanalytically as a function of α. In order to see this, we rewrite (3.2) in the following form Tc0 = exp ψ Tc (α) 1 α + 2 4πTc (α) exp −ψ 1 2 . (3.4) Since we are in the regime where α is finite but Tc goes to zero, we make use of the asymptotic expansion exp[ψ(x + 1/2)] ≈ x + 1/(4!x) + . . . being valid for large x. This expansion yields Tc0 ≈ 4γE Tc α 1 + 4πTc 4!(α/4πTc ) . (3.5) The critical pair-breaking parameter is defined as the value at which the critical temperature vanishes Tc (αc ) = 0. In this case we can neglect the second term on the right hand side of the above equation and obtain αc = πTc0 /2γE ≈ 0.88Tc0 . Expressed in terms of this quantity √ the critical temperature becomes Tc ≈ ( 6/π) α(αc − α). 3.2 Motivation and qualitative picture The renewed interest in the physics of paramagnetically limited superconductors is motivated by the rapid growth of its experimental realizations. Recent parallel magnetic field studies of two-dimensional superconducting systems were extended to much lower temperatures thus making it feasible to approach the limit of QCP. Tunneling spectroscopy of ultrathin Al and Be films revealed field-induced spin mixing and anomalous resonances in the density of states (DOS) [94, 101, 102]. The latter was successfully explained in theory [103–105], which emphasized the crucial role of superconducting pairing correlations in the paramagnetic 40 state even far from the transition region. A surprising enhancement of superconductivity by a parallel magnetic field, deduced from the transport measurements, was observed in ultrathin, homogeneously disordered amorphous Pb films and in the two-dimensional electron gas realized at the interface of oxide insulators LaAlO3 and SrTiO3 [106]. In addition, pronounced negative magnetoresistance (NMR), concomitant with the enhanced Tc , was reported. Although we do not dwell onto the issue of Tc enhancement in these systems (see Ref. [107] for the recent theoretical proposals), we show that transport anomalies, such as NMR, can be successfully addressed within BCS theory. The issue of NMR in superconductors, either near the QCP or near the parallel field-tuned superconductor-insulator transition, was previously discussed in the literature experimentally [108, 109] and attributed theoretically [110–112] to the proliferation of superconductive fluctuations. These studies emphasized mainly the orbital effect of a magnetic field on the preformed Cooper pairs. In this Chapter we develop transport theory of paramagnetically limited ultrathin superconductors focusing on the quantum regime of zero temperature near the critical Zeeman field. The regime of classical fluctuations was partially discussed in the early papers [113–116]. The conceptual difference of our analysis from the problem of fluctuation-induced transport close to Tc is that unpaired particles, have finite excitation energy Ez . As a result, the activation probability of such pairs is suppressed exponentially ∝ exp(−Ez /T ) at low temperature with the statistical Boltzmann factor. We argue that, while in the standard case the real gapless pairs are only important in the paramagnetically limited case, such pairs are always virtual. Let us illustrate this point by taking the Aslamazov-Larkin correction to the conductivity as an example. Consider first the standard case of near–Tc . In the AL diagram the 41 triangular vertex can be estimated as B(ω, Ω, q) ∝ Dqx ∂Π(ω, q)/∂ω. At small momenta we can take Π(ω, 0) in a clean system. The imaginary part of the polarization operator ImΠ ≈ dξ[n(−ξp + ω)n(ξp ) − f˜(−ξp + ω)f˜(ξp )]δ(ω − 2ξ) = ν(ω/2) tanh(ω/2T ), where the particle and hole occupation numbers are f (ε) = (1 + eε/T )−1 , f˜(ε) = 1 − f (ε). The real 2 part, due to virtual pairs ReΠ(ω, q) ≈ ln |(ω 2 − T 2 )/ωD |, is the familiar Cooper logarithm. The imaginary part contribution B(ω, Ω, q) ∝ Dqx /T . In contrast, the real part contribution vanishes at ω = 0 due to the particle-hole symmetry, ν(ω) = ν. The expansion in ω ∼ T − Tc T yields a correction small in the parameter (T − Tc )/Tc 1. In the presence of a Zeeman field the situation is very different. The pair activation rate, ImΠ(ω, q) ≈ ν(ω)[f (ω/2 − Ez /2) − f (ω/2 + Ez /2)], gives an exponentially suppressed contribution ∝ Dqx exp(−Ez /T )/T . The real part, due to virtual pair excitation, can be 2 |. Its contribution obtained by the Kramers-Kronig relation, Re(ω, q)Π ≈ ln |(ω 2 − Ez2 )/ωD to B(ω, Ω, q) is suppressed only algebraically ∝ Dqx T /Ez2 . Unlike the standard case the virtual quasiparticles make a dominant contribution to the triangular vertex excitations. The algebraic suppression of vertexes is most pronounced in the case of the AL and is manifested in additional factors of Dq 2 and ω in σ AL , which makes it logarithmic in Ez /(Ez − ∆0 ). Note that in the case of the near–Hc2 problem [110] the AL contribution is also suppressed due to the current matrix elements connecting adjacent Landau levels. The regular MT and DOS contributions are proportional to a second derivative of the real part of the polarization operator ReΠ(ω, q). Since the latter is finite at ω = 0, these contributions are as singular as AL terms. These technicalities will be explained in great details in the following section. 42 3.3 Fluctuation-induced conductivity near the quantum critical point We approach the problem of transport in the in-plane magnetic field close to the superconducting transition based on the diagrammatic perturbation theory. Note that the technique based on the time-dependent Ginzburg-Landau formalism applied for studying transport near QCP [117, 118] accounts correctly only for the classical part of AL-type contribution to the conductivity, but it misses completely the quantum zero-temperature corrections. A microscopic approach takes care of all the contributions including the DOS part, resulting from the depletion of the normal state density of states by superconducting fluctuations, and also the MT interference term [110–112]. In fact, at T = 0, where the corrections come from purely quantum fluctuations, these effects turn out to be of the dominant nature. In calculations T {Ez , ∆} τ −1 εF , these conditions are satisfied in many experiments [94, 101]. Our starting point is the current-current response kernel which can be conveniently presented as a sum of three contributions K = K AL + K M T + K DOS . Within this section we will be discussing only longitudinal electrical transport so the subscript Kxx is suppressed in all the expressions for brevity. The general expression for the AL term reads K AL (Ωn ) = −e2 T q,ωk B 2 (ωk , Ωn , q)L(ωk , q)L(ωk + Ωn , q), (3.6) where ωk = 2πkT . Notice here the factor of 4 difference compared to the earlier expression Eq. (2.39). This is due to the fact that, in the finite in-plane magnetic field, different spin projections σ = ± contribute unevenly to the current vertex B. It reads explicitly now as 43 follows B(ωk , Ωn , q) = T σ = JAL p σ , λσ (εm+n , ωk − εm , q)λσ (εn , ωk − εn , q)JAL (3.7) vp G σ (εn+m , p)G σ (εn , p)G −σ (−εn + ωk , −p + q), (3.8) σ,εm with the impurity ladders λσ (εn , εm , q) = θ(−εn εm ) , τ [Dq 2 + |εn − εm | − iσEz sgn(εn − εm )] (3.9) and an integral over the block of three Green’s functions with Gεσn ,p = (iεn − ξp + σEz /2 + isgn(εn )/2τ )−1 . Here we used notations: εm = 2πT (m + 1/2), ξp = p2 /2m − εF , vp = ∂p ξp , θ(ε) is the step-function, and sgn(ε) is the sign-function. The propagator of fluctuating Cooper pairs in Eq. (3.6) is given by L−1 (ωk , q) = −ν ln 1 2 T −ψ Tc0 Ψσ (ωk , q) = ψ + 1 Ψσ (ωk , q) , 2 σ=± 1 Dq 2 + |ωk | + iσEz + 2 4πT . (3.10) (3.11) When calculating the B-vertex one should follow a few basic steps which we already discussed in the previous Chapter: i ) In the leading order in the transferred momentum q one can approximate G −σ (−εn + ωk , −p + q) ≈ G −σ (−εn + ωk , p) + v · q[G σ (−εn + ωk , p)]2 . ii ) Furthermore, one can neglect Zeeman energy as compared to the inverse scattering time in the Green’s functions (provided the condition T {Ez , ∆} τ −1 εF is satisfied) and then complete p integration in a standard way with density of states and angular averaging over the Fermi surface. iii ) Next is the fermionic Matsubara εm sum in Eq. (3.7), which can 44 be found in the closed form with the result B(ωk , Ωn , q) = νd q x D Ωn σ σ [Ψσ (|ωk | + Ωn , q) − Ψσ (|ωk |, q) +Ψ (|ωk+n | + Ωn , q) − Ψσ (|ωk+n |, q)] . (3.12) iv ) The remaining step of the calculation is a bosonic ωk sum followed by an analytical continuation iωn → ω. The latter is accomplished via the contour integration over the circle with two-branch cuts at Im(ω) = 0, −Ωn where the product of propagators in Eq. (3.6) has R breaks of analyticity. After the Ω-expansion of KAL (Ω) to the linear order, one finds for the AL AL AL conductivity correction σ AL = σclAL + σq1 + σq2 , where σclAL = +∞ e2 4πT −∞ q ∞ 2 AL σq1 e = 4π AL σq2 =− e2 4π dω [B RA (ω, q)]2 [ImLR (ω, q)]2 , sinh (ω/2T ) 2 dω coth 0 q ω Re [B RA (ω, q)]2 − [B RR (ω, q)]2 ]∂ω [LR (ω, q)]2 , (3.14) 2T +∞ dω coth q −∞ (3.13) ω 2T × ∂Ω [B RR (ω, Ω, q)]2 [LR (ω, q)]2 − ∂Ω [B AA (ω − Ω, Ω, q)]2 [LA (ω, q)]2 +∂Ω [[B RA (ω − Ω, Ω), q]2 − [B RA (ω, Ω, q)]2 ]|LR (ω, q)|2 . (3.15) The superscript R(A) in the vertex functions and propagators stands for the retarded (advanced) component while subscript cl(q) refers to classical (quantum). This convention comes form the observation that as T → 0 the classical contribution vanishes while the quantum contribution remains finite. 45 We turn now to the derivation of the MT contribution whose response kernel is given by K M T (Ωn ) = e2 T ωk ,q L(ωk , q)ΣM T (ωk , Ωn , q), (3.16) where ΣM T (ωk , Ωn , q) = T σ,εm λσ (εm+n , ωk−n − εm , q)λσsgn[ m m+n ] (εm , ωk − εm , q)JM T (3.17) vp vq−p G (εm+n , q)G −σ (ωk−n − εm , q − p)G σ (εm , p)G −σ (−εm + ωk , q − p). (3.18) σ JM T = p Momentum integration in the block of Green’s functions JM T is done under the same approximations as in the case of the AL term described above. According to the standard convention [19] we now split the MT term into the so-called regular and anomalous contributions: ΣM T (reg) (ωk , Ωn , q) = − ΣM T (an) (ωk , Ωn , q) = − νd D Ωn νd D 2(Dq 2 + Ωn ) [Ψσ (|ωk | + 2Ωn , q) − Ψσ (|ωk |, q)] , (3.19) [Ψσ (−|ωk | + 2Ωn , q) − Ψσ (|ωk |, q)] . (3.20) σ σ After the analytical continuation these translate into the conductivity correction σ M T = MT MT σreg + σan , where MT σreg MT σan = e 2 νd D =− 3 2 8π T e 2 νd D 8πT σq ∞ ω Im[LR (ω, q)[Ψσ (−iω, q)] ], 2T (3.21) LR (ω, q)[Ψσ (iω, q) − Ψσ (−iω, q)] dω . Dq 2 + Γ sinh2 (ω/2T ) (3.22) dω coth σq +∞ −∞ 0 In order to regularize the logarithmically divergent momentum integral in the case of the 46 anomalous contribution we have introduced a pair-breaking cutoff parameter Γ. The microscopic origin of the latter (e.g. spin-orbit scattering) will be discussed in the next section. We finally discuss the DOS contribution to the conductivity. It is given by the similar to Eq. (3.16) expression with K DOS (Ωn ) = e2 T ωk ,q L(ωk , q)ΣDOS (ωk , Ωn , q), (3.23) where ΣDOS (ωk , Ωn , q) = 2T σ,εm JDOS = p [λσ (εm , ωk − εm , q)]2 JDOS , (3.24) vp2 [G σ (εm , p)]2 G σ (εm + Ωn , p)[G −σ (ωk − εm , q − p) + 1 2πνd τ p [G σ (εm , p )]2 G −σ (ωk − εm , q − p )]. (3.25) After completing the standard steps outlined above one arrives at the conductivity correction σ DOS = σclDOS + σqDOS in the form σclDOS = − e 2 νd D 16π 2 T 2 +∞ σq −∞ dω[[Ψσ (iω, q)] − [Ψσ (−iω, q)] ] R L (ω, q), sinh2 (ω/2T ) MT σqDOS = σreg . (3.26) (3.27) The equality between the two contributions in Eq. (3.27) has parallels with the original fluctuation transport considerations at T − Tc T . In the original near–Tc problem, the typical energy of diffusing pairs Dq 2 ∼ T − Tc is smaller than the thermal energy of the quasiparticle ∼ T . In our case, Ez adds to the energy of pairs making it bigger than T . Correspondingly, unlike the near–Tc case, the off-shell energy of a pair, 2ε ∼ T , falls below 47 the pair excitation energy set by Ez . This causes a sign inversion of the energy denominator associated with the unbound intermediate state, and the correction Eq. (3.27) turns out to be positive. In general, the derived above conductivity corrections are applicable at any field H and temperature T above the transition. In the following, we discuss limiting cases of interest. It is convenient to regroup all contributions and present the total conductivity correction as the a sum of zero-temperature (δσq ) and finite-temperature (δσT ) terms, namely δσ(H, T ) = δσq (H) + δσT (H, T ). (3.28) The first term here is determined by the quantum AL [Eqs. (3.14)-(3.15) and DOS [see Eq. (3.27)] contributions, and also the regular part of the MT conductivity [see Eq. (3.21)]. The remaining terms define δσT . The magnitude of δσq decreases monotonically with an increasing field. This leads to a pronounced magnetoresistance at zero temperature. At finite temperature, based on how the quantum critical point is approached, there are several regimes that show different T and H dependencies, which should be experimentally accessible. Below we focus on QCP only and extract the leading singularity in δσq as the function of the Zeeman field. Thermal contribution δσT and various crossover regimes will be discussed in the next section. At zero temperature Ψσ (±iω, q) → ln[(Dq 2 ± iω + iσEz )/4πT ] and the pair propagator can be taken in the leading pole approximation LR(A) (ω, q) ≈ − 2∆20 /νd , Ec2 − (ω ± iDq 2 )2 48 (3.29) which is obtained from Eq. (3.10) under the conditions Dq 2 ∆0 and |Ec ± Ω| ∆0 = πTc0 /2γE where ln γE ≈ 0.57 is the Euler constant, and Ec = ∆0 . Here Ez2 − ∆20 . The branch cut of the propagator (due to the logarithmic structure) also contributes to δσq , but gives the subleading singularity. Within the same accuracy we compute the vertex functions: [B RR(AA) (ω(−Ω), Ω, q)]2 = 8νd2 D 2 Dq (Dq 2 ∓ iω)(Dq 2 ∓ iω − 2iΩ), 4 Ez [B RA(AR) (ω, Ω, q)]2 = 8νd2 D (Dq 2 )2 (Dq 2 − 2iΩ). 4 Ez (3.30) (3.31) Alltogether this leads to the conductivity correction near the Zeeman field-induced quantum critical point δσq (H) = 2e2 ln π2 Ez Ez − ∆0 , (3.32) which is obtained within the logarithmic accuracy. Equation (3.32) is the main result of this section. We have checked explicitly that other contributions, such as the diffusion coefficient renormalization or the contribution with only one or no Cooperon vertex, are either small or nonsingular. Since the temperature can be set to zero in integrations over fast-fermion degrees of freedom, the additional factors of τ results in small prefactors τ Ez , τ DQ2 or τ Ω. 3.4 Effects of pair-breaking scattering In the preceding calculations we assumed that impurity scattering of electrons does not cause spin flips. There are two sources of spin relaxation of conduction electrons: localized spins (magnetic impurities) and spin-orbit (SO) scattering of electrons by nonmagnetic disorder. The latter is characterized by the scattering amplitude ivso ([p×p ]·σ)/p2F , where p and p are the initial and final momenta of an electron, and σ is the spin operator whose components 49 are the Pauli matrices. Let us discuss the effect of SO scattering first starting with the qualitative physical picture [103]. In the absence of both, SO interactions and magnetic field two-spin states, which belong to a given orbital, have the same energy. Magnetic field split this degeneracy. It is important that the splitting energy Ez is exactly the same for all of the orbital states, which is no longer the case for finite SO interaction. Without an external magnetic field the states are still doubly degenerate due to time-reversal invariance (Kramers doublets). A magnetic field splits the Kramers doublets similar to how it splits pure spin states in the absence of SO interactions. The main difference is that this splitting is not exactly uniform anymore [119]. The spin-orbit scattering and finite thickness effects modify the fluctuation transport, due to the finite spectral weight in the particle-particle channel at zero frequency. The addition of a finite spin-orbit scattering introduces a finite lifetime Γ−1 to the Cooperon. At lowest temperatures the superconductivity survives if this scattering is not too strong, Γ Ez with a somewhat lower critical field. While Ez approaches the supercooling transition from above the results obtained in the previous section are expected to crossover to a different regime at Γ ≈ Ec . The finite film thickness affects the crossover in a similar way. Inclusion of these effects was shown to be necessary for quantitative analysis of measurements of the density of states [102]. As was discussed early in Sec. 2.5 [see Eq. (2.25)] the Cooperon is formed by two electron Green’s functions. In the absence of an external magnetic field it is convenient to classify Cooper poles by the total spin of the two electrons S+ = (σ 1 + σ 2 )/2. Spin-orbit scattering does not affect the spin singlet part of the Cooperon (S2+ = 0), however, this scattering leads to total spin relaxation, i.e., the triplet (S2+ = 2) component of the Cooperon decays and, consequently, the pole in the ω plane is shifted from the real axis even for q = 0. An 50 external magnetic field is coupled with the difference S+ = (σ 1 + σ 2 )/2 of two electron spins, and as a result we classified the Cooperon by the eigenvalue of the operator S− · Ez . These eigenvalues for S2− = 2 are 0, ±Ez , corresponding to S−z = 0, ±1 and 0 for S2− = 0. Neither of those two classifications is exact when both a magnetic field and SO scattering take place simultaneously. We assume that the SO effect is weak that allows us to evaluate the Cooperon perturbatively Π(ω, q) = 4πνd T εn 1 , 2+Γ 2ε + |ω| + iσE + Dq n z so >0 σ=± (3.33) −1 2 where Γso = 2/3τso , and τso = 2πνd vso is the time of the spin relaxation by SO scattering [compare Eq. (3.33) to Eq. (2.28)]. This consideration suggests that any physical mechanism of violation of either time-reversal invariance or conservation of spin will have a similar effect on a Cooperon field. In the following discussion we assume that Γ = Γso + Γs + ΓH is the total scattering rate that include spin-orbital, spin-flip and finite film thickness effects. 3.5 Thermomagnetic phenomena Since the Cooperon is no longer a soft mode at (ω, q) → 0 and has a finite gap Γ, due to spin-related scattering processes, it inevitably enters into the pair propagator L(ω, q), shifts its pole and ultimately changes temperature dependence of the kinetic coefficients. Thus with Eq. (3.33) at hand we have to recompute all the basic ingredients of the diagrammatic technique. Most importantly we find the generalized form of the fluctuation propagator LR (ω, q) = − 2∆20 /νd , Ec2 − 2iΓ(ω + iDq 2 ) − (ω + iDq 2 )2 51 (3.34) which is different from the expression that we used before [Eq. (3.29)] in two important aspects. First is the presence of the new term in the denominator which introduces an additional scale Ec2 /Γ to the problem. Second is shifted value of the critical point Ec = Ez2 − ∆2Γ with ∆2Γ = ∆20 + Γ2 . Another important building block of the theory is the vertex function B. We present only its mixed retarded-advanced component [B RA(AR) (ω, Ω, q)]2 = 8νd2 D (Dq 2 + Γ)2 (Dq 2 − 2iΩ), 4 Ez (3.35) which is the most relevant for the transport regime that we will specify next. The subsequent calculations will be carried out assuming Γ critical line Ec2 /Γ T Γ Ec and for temperatures not too close to the Ez , which is quite relevant for the actual experimental realization. Under these conditions the third term in the denominator of Eq. (3.34) can be neglected so that one finds approximately ImLR (ω, q) = − 2Γω 2∆20 . 2 νd (Ec + 2ΓDq 2 )2 + 4Γ2 ω 2 (3.36) In order to determine δσT (H, T ) we have to reexamine all the contributions to fluctuationinduced conductivity. In the course of this analysis we found that the most singular term originates from the classical part of the AL contribution Eq. (3.13), so that in large δσT (H, T ) is governed by σclAL . To calculate this term explicitly in the above discussed regime, we introduce dimensionless variables x = Dq 2 /2T , y = ω/2T , γ = Γ/2T , and = Ec /2T and obtain from Eq. (3.13) δσT (H, T ) = 16e2 π2 ∞ +∞ dx 0 −∞ dy sinh2 y [( 52 γ 2 y 2 x(x + γ)2 , 2 + 2γx)2 + 4γ 2 y 2 ]2 (3.37) where, in addition, we set ∆0 /Ez → 1 assuming that H is tuned sufficiently close to the transition line. The remaining integrations can be done with having small parameter 2 /γ 1, which leads to the result δσT (H, T ) = e2 2π TΓ Ez2 Ez Ez − ∆Γ . (3.38) It is worth mentioning here that, unlike the quantum regime Eq. 3.32, which is logarithmically singular in Ez −∆, the magnetoresistance in the classical region is more pronounced. We have further checked that the other terms remain smaller and scale logarithmically ∝ ln(Ez Γ/Ec2 ) (see Appendices C and D for further details). In a similar spirit we can calculate now αxy . We have verified that, in the regime of classical fluctuations, the relation between current B e and heat B h vertices remains the same B h = (−iω/2e)B e , so that we can proceed immediately to Eq. (2.45) with the vertex and propagator taken from Eqs. (3.35) and (3.34) respectively. At the intermediate step one finds for transverse thermoelectric coefficient αxy 512e L2T = π 2H ∞ +∞ 2 4 dxx γ 0 dy coth(y) −∞ γy( 2 + 2γx) [( 2 + 2γx)2 + 4γ 2 y 2 ]3 (3.39) and after final integrations αxy = 2e where Lz = L2z Γ Ez − ∆Γ 2 H , D/Ez . We reiterate that this result is valid provided Ec2 /Γ (3.40) T Γ Ez and observe that the singularity in αxy is the same as in the conductivity δσT . This resembles similar situation as in the case of near Tc transport at zero field, namely identical scaling of the α and σ with the critical parameter, while the underlying physics is very different. 53 It should be stressed, however, that the field dependence of αxy is extremely sensitive to the pair-breaking scattering. For completeness, we also analyzed behavior of αxy in the limit of negligible Γ max{T, Ec } Ez and find dependence different than that give by Eq. (3.40). For that purpose, we use Eqs. (3.29) and (3.30) in the expression for αxy defined by Eq. (2.57), which leads us to expression αxy = 256e L2T π 2 2H ∞ +∞ dx dy coth(y) −∞ 0 x7 y( 2 + x2 − y 2 ) , [( 2 + x2 − y 2 )2 + 4x2 y 2 ]3 (3.41) which is valid in the thermal region of fluctuations T > Ec . The double integral gives as factor of 9π 2 /256 . This implies that αxy has square-toot singularity √ L2 αxy = 18 2e 2z H Ez . Ez − ∆0 (3.42) We conclude this section by briefly discussing behavior of αxy in the regime when orbital effects of pair-breaking dominate over the spin-related effects. The phase diagram has been already discussed above based on Eq. (3.2), where we analyzed limiting case of classical and quantum fluctuation regimes. At a given pair-breaking strength α, superconductivity is destroyed at T = Tc (α) and at a given temperature T , at α = αc (T ), obtained by solving Eq. (3.2) for T and for α, respectively. In the neighborhood of this classical transition, for α Tc0 we can define the quantity T (α, T ) = [T − Tc (α)]/Tc (α)/ that measures the relative distance from the critical temperature Tc (α). Conversely, in the vicinity of the the pairbreaking quantum phase transition on can define the quantity α (α, T ) = [α − αc (T )]/αc (T ), which can be interpreted as the relative distance from the critical pair-breaking strength αc (T ) at a given T αc0 . In a parametrically broad temperature regime near the transition 54 line (but not too close to the quantum region) it is legitimate to approximate pair-propagator and vertex function by the following simple expressions LR (ω, q) = − 1 1 , νd ln(α/αc (T )) + (Dq 2 − iω)/2αc (T ) Bi = −νd Dqi /αc (T ). (3.43) When combined with Eq. (2.57) this yields αxy = 2e L2T π 2 2H ∞ +∞ dy dx 0 −∞ x2 y coth(αc y/T )( α + x) , [( α + x)2 + y 2 ]3 (3.44) where we introduced dimensionless variables for the momentum x = Dq 2 /2αc and frequency y = ω/2αc . This result is valid for temperatures away from the critical region T αc so that it is legitimate to expand the cotangent at small argument. The double integral gives a factor πT /8αc α so that one arrives at αxy = e L2α αc (T ) , 4π 2H α − αc (T ) where we introduced new length-scale Lα = (3.45) D/αc (T ), and also expanded the logarithm ln(α/αc ) ≈ (α − αc )/αc assuming close vicinity to the transition line. Given the plethora of the different regimes and behaviors, it is desirable to have a systematic experimental study aimed specifically at exploring the physics of a pair-breaking phase transition in superconducting films from the measurements of the conductivity δσ(H, T ) and thermoelectric response αxy . There have been few works using a parallel field as a control parameter to scan across the phase diagram, but they have had other goals mainly focussing on the physics of superconductor-insulator transition. To begin with, it will be useful to observe the finite temperature classical transition and verify the predictions of the fluctuation the55 ory presented in this section in its vicinity. By slowly increasing the pair-breaking strength and lowering the temperature, one could approach the quantum phase transition. Having identified the right films, measurements of the temperature and pair-breaking parameter dependence of the conductivity, would afford an exciting possibility of discovering different regimes in the vicinity of the pair-breaking quantum phase transition. The non-monotonic magnetoresistance due to the presence of superconducting fluctuations that we find, is in stark contrast to the intuitive expectation and is a purely quantum effect. A clear experimental signature of such a characteristically quantum behavior would be an important step forward in the study of quantum phase transitions and low temperature superconductivity. 3.6 Summary In this Chapter we studied electrical and thermal transport anomalies of low dimensional superconducting films in an external in-plane magnetic field. We concentrated on the ClogstonChandrasekhar (CC) phase transition, i.e., the destruction of superconductivity by a magnetic field by virtue of the Zeeman splitting. As a result, a normal paramagnetic state of electrons is formed. The main conclusion we can draw from this study of the CC state is, that despite this state being normal (namely with the mean-field superconducting order parameter vanishing), it is drastically different from a usual normal metal with some attractive interaction. The latter state appears, e.g., in a superconductor at temperatures higher than the transition temperature Tc . The difference becomes apparent when one studies excited states rather than those close to the ground state. In particular, fluctuation-induced transport is dominated by the virtual excitations rather than real preformed Cooper pairs. This leads to a nontrivial magnetoresistance near QCP. The reason why the effects of su- 56 perconducting fluctuations in a CC metal are dramatically enhanced, in comparison with the usual case, is the presence of the pole-like singularity in the correlation function of these fluctuations. This pole at a finite frequency appears due to the fact that the CC transition is of the first order. In contrast, the temperature-driven transition from superconductor to normal metal is of the second order, and in a usual normal state the correlator of the superconducting fluctuations is a smooth function of the frequency, i.e., any superconducting excitations decay very rapidly. Near the QCP of the supercooling line of the phase diagram, magnetoresistance is governed by density of states and regular Maki-Thompson terms. We found complete cancellation of quantum Aslamazov-Larkin corrections, while anomalous Maki-Thomspon and classical Aslamazov-Larkin terms vanish in the zero-temperature limit. Nevertheless, the latter terms are crucially important in describing the crossover regimes at finite temperature where the sign of the magnetoresistance essentially depends on the direction at which the transition line is approached in the field-temperature plane. Surprisingly, we find that, near the transition line, scaling of the conductivity corrections is the same as the scaling of the transversal thermoelectric coefficient, which is analogous to usual transport anomalies near Tc and in the absence of the field. A priori this result is difficult to foresee since the mere mechanism of fluctuation corrections is different in the CC phase. The apparent universality between the singular field dependences of σxx and αxy near the transition is another important observation of this study. We close this Chapter by briefly discussing outstanding problems that remain largely unsolved, where our microscopic approach may give an opportunity to systematically study thermomagnetic phenomena in other various superconducting systems. Since the Nernst effect is highly sensitive to fluctuations, its measurements may shed light on the intimate 57 connection between quantum criticality and unconventional superconductivity with competing or coexisting orders. Perhaps, the most interesting systems from that perspective are heavy-fermion superconductors (e.g. URu2 Si2 ) and iron-pnictide superconductors (e.g. FeSe compounds). It was recently reported [120] that the Nernst coefficient in URu2 Si2 is anomalously enhanced as compared to the theoretically expected value of the standard Gaussian fluctuations. Moreover, contrary to the conventional wisdom, the enhancement is more significant with the reduction of the impurity scattering rate. This unconventional Nernst effect intimately reflects the highly unusual superconducting state embedded in the so-called hiddenorder phase of URu2 Si2 that appears to point to a new type of superconducting fluctuations generated by a degree of freedom which has not been hitherto taken into account. It is tempting to consider that such a degree of freedom is intimately related to the superconducting state with broken time-reversal symmetry. To properly address the data one has to seriously consider possible chiral or Berry-phase fluctuations associated with the broken time-reversal symmetry of the superconducting order parameter [121]. Our theory of Pauli limited superconductivity may be highly relevant for the study of FeSe superconductors. This is a unique solid state system that offers the possibility to enter the previously unexplored realm where the three energies, Fermi energy, superconducting gap, and Zeeman energy, become comparable, and thus access the crossover regime between the weakly coupled Bardeen-Cooper-Schrieffer (BCS) limit and the strongly coupled BoseEinstein-condensate (BEC) limit. The results of the transport properties of FeSe near the BCS-BES crossover reveal intriguing features [122]. What is remarkable is that, for the Nernst effect, the Seebeck coefficient, the temperature derivative of the resistivity, and the Hall coefficient, all show peaks at around 20 K, which is twice as large as the critical tem58 perature (Tc ∼ 10 K). One of the most important subject in the BCS-BEC crossover is the debate concerning the mechanism — preformed pair or pseudogap. It is useful to recall that at the pseudogap temperatures of YBCO and CeCoIn5 , the Nernst effect exhibits its peak. Whether one can draw any conclusion from this similarity for FeSe systems remain to be seen. What is clear, is that theoretical input is urgently needed and so thermomagnetic transport in superconductors will continue to attract tremendous attention from researchers. 59 APPENDICES 60 Appendix A: Matsubara sums and analytical continuation As outlined in the main text we convert the bosonic Matsubara sum into the integral S(Ων ) = 1 4πi dz coth C z L(−iz)L(−iz + Ων ), 2T (A.46) where the contour of integration is a circle which contains two branch-cuts at Imz = 0 and Imz = −Ων , [see Fig. (A.3)], where functions L(−iz) and L(−iz + Ων ) have breaks of analyticity respectively (q-dependence of propagators is suppressed here for brevity). We thus have explicitly +∞ 1 z R S(Ων ) = dz coth L (−iz + Ων )[LR (−iz) − LA (−iz)] + 4πi −∞ 2T +∞−iΩν 1 z A dz coth L (−iz)[LR (−iz + Ων ) − LA (−iz + Ων )] . 4πi −∞−iΩν 2T 61 (A.47) Figure A.3: Integration contour in the plane of complex frequency. The lower part of the contour corresponds to advanced-advanced products of propagators after analytical continuation. The middle section contains mixed causality components of advanced-retarded, while the upper third contains only retarded-retarded products of propagators. In the second integral we shift variable z + iΩν = z which gives us +∞ z R 1 dz coth L (−iz + Ων )[LR (−iz) − LA (−iz)] + S(Ων ) = 4πi −∞ 2T +∞ 1 z − iΩν A dz coth L (−iz − Ων )[LR (−iz ) − LA (−iz )] . 4πi −∞ 2T (A.48) z ν Taking into the account that iΩν is the periodic of cotangent coth z −iΩ = coth 2T , 2T changing back z → z in the second integral, and taking the analytic continuation step Ων → −iΩ, we get 1 S (Ω) = 2π +∞ R dz coth −∞ z R [L (−iz − iΩ, q) + LA (−iz + iΩ, q)]ImLR (−iz, q) . 2T 62 (A.49) Since LR (−iz) = LA (iz) we get for the response kernel 2e2 π ee (Ω)]R = [Kxx +∞ Bx2 (q) q dω coth −∞ ω A [L (ω+Ω, q)+LR (ω−Ω, q)]ImLA (ω, q) . (A.50) 2T We do not write imaginary i in the frequency argument of LR(A) (iz), which is implicit in the definition. To the linear order in external frequency LA (ω + Ω, q) + LR (ω − Ω, q) ≈ Ω∂ω [LA (ω, q) − LR (ω, q)] = 2iΩ∂ω ImLA (ω, q) . (A.51) As a result the retarded component of the current-current response kernel reduces to ee [Kxx (Ω)]R = 2iΩe2 π +∞ q Bx2 (q) dω coth −∞ ω ∂ω ImLA (ω, q) 2T 2 , (A.52) which can be integrated by parts to yield Eq. (2.45). Calculation of the Matsubara sum in the case of thermomagnetic response functions follows the same steps as above but is more involved since it contains three propagators S(Ωm ) = 1 4πi dω coth C ω (ω + iΩm /2) L3 (ω, q)L(ω + iΩm , q) − L3 (ω + iΩm , q)L(ω, q) 2T (A.53) We transform this integral into a contour with branch cuts S(Ωm ) = 1 4πi +∞ dωω+ coth −∞ ω [L3R (ω, q) − L3A (ω, q)]LR (ω + iΩ, q) 2T −[LR (ω, q) − LA (ω, q)]L3R (ω + iΩ, q) + 1 4πi +∞−iΩ dωω+ coth −∞−iΩ ω [LR (ω + iΩ, q) − LA (ω + iΩ, q)]L3A (ω, q) 2T −[L3R (ω + iΩ, q) − L3A (ω + iΩ, q)]LA (ω, q) , 63 (A.54) where we used notation ω± = ω ± iΩm /2. In the second integral we shift ω + iΩm → ω, ω m = coth 2T , and then perform the analytic use the periodicity of the cotangent coth ω−iΩ 2T continuation iΩm → Ω, which gives +∞ ω 1 dωω+ coth LR (ω + Ω, q)ImL3R (ω, q) − L3R (ω + Ω, q)ImLR (ω, q) S(Ω) = 2π −∞ 2T +∞ 1 ω + dωω− coth L3A (ω − Ω, q)ImLR (ω, q) − LA (ω − Ω, q)ImL3R (ω, q) . (A.55) 2π −∞ 2T To the linear order in Ω, there are terms of two kind. First is the direct term proportional to Ω due to the vertex. Second, is the linear term from the expansion of propagators. We focus on the first possibility since it gives the most important contributions. We set Ω → 0 in propagators and get S(Ω) = Ω 4π +∞ dω coth −∞ ω 2T (A.56) LR (ω, q)ImL3R (ω, q) − L3R (ω, q)ImLR (ω, q) − L3A (ω, q)ImLR (ω, q) + LA (ω, q)ImL3R (ω, q) . This can be rewritten as S(Ω) = Ω 2π +∞ dω coth ω ReLR (ω, q)ImL3R (ω, q) − ReL3R (ω, q)ImLR (ω, q) , 2T dω coth ω [ReLR (ω, q)]3 ImLR (ω, q) + ReLR (ω, q)[ImLR (ω, q)]3 , 2T −∞ (A.57) or equivalently Ω S(Ω) = π +∞ −∞ (A.58) which eventually translates into Eq. (2.57). 64 Appendix B: Seebeck (αxx) and Hall (σxy ) coefficients near Tc The Aslamazov-Larkin contribution to the diagonal (Seebeck) thermoelectric coefficient is found from the mixed electric-heat currents Kubo response function AL αxx =− 1 1 eh lim Im[Kxx (Ω)]R , T Ω→0 Ω eh Kxx (Ων ) = 2ieT qω ωn Bx2 (q)L(ωn , q)L(ωn + Ων , q). (B.59) Summation over the Matsubara frequency ωn and analytical continuation follows the same way as in the case of the conductivity calculation, and we obtain AL αxx = e 2πT 2 +∞ q Bx2 (q) −∞ ωdω [ImLR (ω, q)]2 . sinh2 (ω/2T ) (B.60) Without particle-hole asymmetry, αxx is zero. Indeed, [ImLR ]2 is even in frequency while the rest of the integrand is odd. We have to use the generalized form of the pair propagator that explicitly accounts for the particle-hole asymmetry factor Υω , which is essentially dictated 65 by gauge invariance [123] L(ωm , q) = − 1 1 , 2 νd πDq /8T + + π|ωm |/8T + Υω Υω = iωm ∂Tc . 2Tc ∂εF (B.61) Expanding LR to the leading linear in Υω order, ImLR = ImLR |Υ=0 + Υω ∂ ImLR |Υ=0 produces AL = αxx e πT 2 +∞ q Bx2 (q) −∞ ωΥω dω ImLR (q, ω)∂ ImLR (ω, q), 2 sinh (ω/2T ) (B.62) where now both propagators are taken at Υω = 0. After introducing dimensionless variables, at the intermediate step, one has AL αxx 8T e d ln Tc = 3 π εF d ln εF +∞ xmax dx dy −∞ 0 xy 2 (x + ) . [( + x)2 + y 2 ]3 (B.63) Logarithmically divergent momentum (x-integration) has to be regularized so that we introduced upper cut-off xmax 1/ (in the original notations this corresponds to (ξT qmax )2 1. This choice is natural since LR , in the form we use, was obtained from the expansion of the digamma function which works only as long as max{Dq 2 , ω} < T ). After the final integrations one finds AL αxx = 2T e d ln Tc ln π 2 EF d ln EF Tc T − Tc . (B.64) In order to calculate the Hall coefficient, we need to know the transversal component ee of the current-current correlation function Kxy ∼ Bx By LL. In the presence of Landau quantization the vertex in real space becomes an operator Bˆi = −2νd η(−i∇i + 2eAi ), 66 (B.65) where we choose the vector potential in the Landau gauge A = (0, Hx, 0). Different components of the vertex, Bˆx and Bˆy , do not commute and the matrix elements are Bˆαnn  √  a−a ˆ† |n 2 2νd η  i n|ˆ =−  H  n|ˆ a+a ˆ† |n α=x , (B.66) α=y where a ˆ, a ˆ† are oscillator operators. Recalling that n|ˆ a|n = n |ˆ a† |n = √ nδn,n +1 , we see that only transitions between nearest Landau levels n → n ± 1 are allowed. With this at hand we find for the Matsubara response kernel [123] ee Kxy (Ω) = (4eνd η)2 T 8π 4H ∞ (n−1)[Ln+1 (ω, q)Ln (ω−Ω, q)−Ln (ω, q)Ln+1 (ω−Ω, q)] . (B.67) ω n=0 After analytic continuation one gets AL σxy (4eνd η)2 =− 4π 2 4H ∞ +∞ (n + 1) n=0 R ImLn (ω, q)∂ω ReLR n+1 (ω, q) ω 2T dω coth −∞ R − ImLR n+1 (ω, q)∂ω ReLn (ω, q) . (B.68) In the weak field limit, one needs only the first term in the expansion in powers of 1/n and then substitute integral for the n summation (1/ ∂n Ln = 2ν(η/ 2 2 H )Ln H) 2 n → q. Taking into account and, after some algebra, we find AL σxy =− (4eη)2 νd3 η 3πT 2H +∞ q2 q −∞ dω [ImLR (ω, q)]3 , sinh (ω/2T ) 2 (B.69) where we also used integration by parts with respect to the energy variable. Since [ImLR (ω, q)]3 AL is odd in energy without particle-hole asymmetry, σxy vanishes in this case. Expanding to 67 the lowest non-vanishing order we get AL = σxy (4eη)2 νd4 η ∂ ln Tc 2πT 2H ∂εF +∞ q2 q −∞ ωdω [ImLR (ω, q)]2 Im[LR (ω, q)]2 . sinh2 (ω/2T ) Introducing, as usual, dimensionless variables and using integrals and +∞ −∞ +∞ 0 (B.70) xdx/(x + 1)4 = 1/6 y 2 dy/(y 2 + 1)4 = π/16 we finally get AL σxy = ∂ ln Tc e2 (ωc τ ) 48 ∂ ln εF where ωc = eH/m is cyclotron frequency. 68 Tc T − Tc 2 , (B.71) Appendix C: Quantum Aslamazov-Larkin terms Within this section we analyze more carefully the unconventional AL terms in the quantum limit T → 0. We start from Eq. (3.14) and obtain, in the dimensionless variables of the main text AL σq1 =− e2 2π 2 ∞ +∞ dy coth(πy) −∞ dx (−iy)(−iy + 2x)x 0 ∂ ∂y −2 2 + (x − iy)2 2 + c.c. . (B.72) Here c.c. stands for the complex conjugate term. In this expression we can replace the integrand with +∞ −∞ dy coth(πy)[. . .] → 2 ∞ 0 dy coth(πy)Re[. . .]. At 1, coth πy ≈ 1, and using the integral ∞ 4 dy(−iy)(−iy + 2x)∂y 0 4 16 = − [ 2 + (x − iy)2 ]2 x2 + 2 , (B.73) we obtain AL = σq1 8e2 Ez ln π2 Ec 8e2 ln π2 69 Ez . 2(Ez − ∆0 ) (B.74) In the opposite limit, 1, the expansion of coth(πy) ≈ 1/πy gives the vanishing real part of the integrand. We therefore conclude that this contribution is not singular in the high temperature limit when T Ec . The second type of unconventional AL corrections comes from differentiating the triangular vertices B instead of the propagators. In the dimensionless notations, Eq. (3.15) can be reduced to the form AL = σq2 ie2 π2 ∞ +∞ dy coth πy −∞ 0 dxx (−iy + x)[lR (x, y)]2 − (iy + x)[lA (x, y)]2 , (B.75) where lR(A) (x, y) = − In the low-temperature, quantum limit T ∞ dy 0 2 2 . + (x ∓ iy)2 (B.76) Ec , one uses the integral 4(−iy + x) = [ + (x − iy)2 ]2 2 2 2i + x2 (B.77) to obtain AL σq2 =− 8e2 Ez ln . π Ec (B.78) AL In the opposite limit, at higher temperatures away from the critical point, σq2 = −2e2 /π . AL AL Comparing σq1 and σq2 in the quantum limit T → 0, we observe acomplete cancella- tion effect, namely AL terms have no divergent, singular correction in Ez /Ec near QCP. As discussed in the main text, magnetoresistance δσq (H) is determined by DOS and MT contributions. 70 Appendix D: Anomalous Maki-Thompson terms In the main text we introduced the pair-breaking (dephasing) parameter Γ in order to regularize the anomalous MT term. As discussed in details in Sec. (3.4), Γ naturally appears as a result of spin-flip scattering. We thus concentrate on the Γ-dependence of the anomalous MT term from Eq. (3.22) in various temperature regimes. In the dimensionless notations, Eq. (3.22) reduces to MT σan = e2 π +∞ −∞ dy sinh2 (πy) ∞ dx 0 In the high temperature limit MT σan = 1 (or equivalently in original notations T ∞ +∞ e2 π3 dy −∞ dx 0 1 (xy)2 . (Γ/2πT + x) [y 2 − ( − ix)2 ][y 2 − ( + ix)2 ] (B.79) Ec ) 1 x2 , (Γ/2πT + x) [y 2 − ( − ix)2 ][y 2 − ( + ix)2 ] (B.80) which eventually reduces to MT σan = e2 2π 2 ∞ dx 0 1 x 2 (Γ/2πT + x) x + 2 = 71 e2 π /2 − (Γ/2πT ) ln(2πT /Γ) . 2π 2 (Γ/2πT )2 + 2 (B.81) From this expression we can extract various limiting cases. For Γ/2 < 2πEc < T , which is MT possible if Γ/2 < T , one finds σan = e2 /4π . Alternatively, for 2πEc < Γ/2 < T , which is possible if Γ/2 < T , or for 2πEc < T < Γ/2, which is possible if Γ/2 > T , one finds MT σan = (e2 /π)(T /Γ) ln(Γ/Ec ). We proceed with the low temperature limit 1 (or equivalently T Ec ). In this case we have MT σan +∞ e2 = π −∞ ∞ dyy 2 sinh2 (πy) dx 0 1 (Γ/2πT + x) [ x2 . 2 + x2 ]2 (B.82) At T < Ec < Γ, MT σan = e2 2πT 3π 2 Γ ∞ 0 x2 dx e2 2πT = . (x2 + 2 )2 12π Γ (B.83) Alternatively, at T < Γ < Ec or Γ < T < Ec , MT σan = e2 3π 2 ∞ 0 xdx e2 = , (x2 + 2 )2 6π 2 2 (B.84) and consequently, the anomalous MT term has no singular contributions near QCP. The logarithmically divergent correction declared in Eq. (3.32) of the main text originates from the regular part of the MT term. Indeed, from Eq. (3.21) we obtain MT σreg 4e2 = 2 π At high temperatures T MT σreg = 4e2 π3 ∞ ∞ dx 0 dy coth(πy) 0 xy [(y + )2 + x2 ][(y − )2 + x2 ] . (B.85) Ec , ∞ ∞ dx 0 dy 0 x e2 = , [(y + )2 + x2 ][(y − )2 + x2 ] 2π 72 (B.86) while at low temperatures T MT = σreg 4e2 π2 ∞ ∞ dx 0 Ec , near QCP , dy 0 2e2 xy ≈ ln(Ez /Ec ). [(y + )2 + x2 ][(y − )2 + x2 ] π2 73 (B.87) BIBLIOGRAPHY 74 BIBLIOGRAPHY [1] L. D. Landau, JETP 7, 19 (1937) [2] L. P. Kadanoff, Physics 2, 263 (1966). [3] A. Z. Patashinskii and V. L. Pokrovsky, 50, 439 (1966). [4] V. N. Gribov and A. A. Migdal, JETP 55, 1498 (1968). [5] A. M. Polyakov, JETP 55, 1026 (1968). [6] B. I. Halperin and P. C. Hohenberg, Phys. Rev. 168, 898 (1969). [7] K. G. Wilson, Phys. Rev. B 4 3174 (1971); ibid 4, 3184 (1971). [8] V. L. Ginzburg, Sov. Solid State Phys. 2, 61 (1960). [9] J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Phys. Rev. 108, 1175 (1957). [10] L. G. Aslamazov and A. I. Larkin, Phys. Lett. 26A, 238 (1968). [11] K. Maki, Prog. Theor. Phys. 39, 897 (1968). [12] R. S. Thompson, Phys. Rev. B 1, 327 (1970). [13] R. E. Glover, Phys. Rev. Lett. A25, 542 (1697). [14] W. A. Little, Phys. Rev. 156, 396 (1967). [15] J. S. Langer and V. Ambegaokar, Phys. Rev. 164, 498510 (1967). [16] D. E. McCumber and B. I. Halperin, Phys. Rev. B 1, 1054 (1970). [17] N. Giordano, Phys. Rev. Lett. 61, 2137 (1988). 75 [18] E. Abrahams, M. Redi, and J. W. F. Woo, Phys. Rev. B 1, 208 (1970). [19] A. I. Larkin and A. Varlamov, Theory of fluctuations in superconductors (Clarendon Press, Oxford, 2005). [20] B. L. Al’tshuler, JETP Lett. 41, 648 (1985). [21] B. L. Al’tshuler and D. E. Khmel’nitskii, JETP Lett. 42, 359 (1986). [22] P. A. Lee and A. D. Stone, Phys. Rev. Lett. 55, 1622 (1985). [23] P. A. Lee, A. D. Stone, and H. Fukuyama, Phys. Rev. B 35, 1039 (1987). [24] G. D. Mahan, Many-Particle Physics, (Kluwer Academic, 2000). [25] I. L. Aleiner and Ya. M. Blanter, Phys. Rev. B 65, 115317 (2002). [26] B. L. Altshuler and B. I. Shklovskii, Sov. Phys. JETP 64, 127 (1986). [27] M. L. Mehta, Random Matrices, (Elsevier, 2004). [28] S. G. den Hartog, C. M. A. Kapteyn, B. J. van Wees, T. M. Klapwijk, W. van der Graaf, and G. Borghs, Phys. Rev. Lett. 76, 4592 (1996). [29] K. Hecker, H. Hegger, A. Altland, and K. Fiegle, Phys. Rev. Lett. 79, 1547 (1997). [30] P. W. Brouwer and C. W. J. Beenakker, Phys. Rev. B 52, 16772 (1995). [31] A. Altland and M. R. Zirnbauer, Phys. Rev. B 55, 1142 (1997). [32] T. T. Heikkila, M. M. Salomaa, and C. J. Lambert, Phys. Rev. B 60, 9291 (1999). [33] A. Frydman, E. P. Price, and R. C. Dynes, Solid State Comm. 106, 715 (1998). [34] A. Frydman, E. P. Price, and R. C. Dynes, Phys. Uspekhi 41, 220 (1998). [35] Yu. Zadorozhny, D. R. Herman, and Y. Liu, Phys. Rev. B 63, 1444521 (2001). [36] A. V. Herzog, P. Xiong, and R. C. Dynes, Phys. Rev. B 58, 14199 (1998). 76 [37] A. Johansson, G. Sambandamurthy, D. Shahar, N. Jacobson, R. Tenne, Phys. Rev. Lett. 95, 116805 (2005). [38] B. L. Altshuler and B. Z. Spivak, Sov. Phys. JETP 65, 343 (1987). [39] B. Z. Spivak and A. Yu. Zyuzin, JETP Lett. 47, 267 (1988). [40] C. W. J. Beenakker, Phys. Rev. Lett. 67, 3836 (1991). [41] T. Micklitz, Phys. Rev. B 75, 144509 (2007). [42] M. Houzet and M. A. Skvortsov, Phys. Rev. B 77, 024525 (2008). [43] B. Spivak and F. Zhou, Phys. Rev. Lett. 74, 2800 (1995). [44] V. M. Galitski and A. I. Larkin, Phys. Rev. Lett. 87, 087001 (2001). [45] M. A. Skvortsov and M. V. Feigel’man, Phys. Rev. Lett. 95, 057002 (2005). [46] F. Zhou and B. Spivak, Phys. Rev. Lett. 80, 5647 (1998). [47] A. Lamacraft, F. M. Marchetti, J. S. Meyer, R. S. Moir, and B. D. Simons, J. Phys. A 37, L447 (2004). [48] F.-von Oppen and E. K. Riedel, Phys. Rev. B 46, 3203(R) (1992). [49] A. I. Buzdin and A. A. Varlamov, Phys. Rev. Lett. 89, 076601 (2002). [50] G. Schwiete and Y. Oreg, Phys. Rev. Lett. 103, 037001 (2009); Phys. Rev. B 82, 214514 (2010). [51] O. Entin-Wohlamn, H. Bary-Soroker, A. Aharony, Y. Imry, and J. G. E. Harris, Phys. Rev. B 84, 184519 (2011). [52] J. A. Melsen, P. W. Brouwer, K. M. Frahm and C. W. J. Beenakker, Europhys. Lett. 35, 7 (1996). [53] A. Altland, B. D. Simons, and D. Taras-Semchuk, JETP Lett. 67, 22 (1998); Adv. Phys. 49, 321 (2000). 77 [54] M. G. Vavilov, P. W. Brouwer, V. Ambegaokar and C. W. J. Beenakker, Phys. Rev. Lett. 86, 874877 (2001). [55] P. M. Ostrovsky, M. A. Skvortsov, and M. V. Feigel’man Phys. Rev. Lett. 87, 027002 (2001). [56] J. S. Meyer and B. D. Simons, Phys. Rev. B 64, 134516 (2001). [57] B. Z. Spivak and A. Yu. Zyuzin, Europhys. Lett. 8, 669 (1989). [58] F. Zhou and C. Biagini, Phys. Rev. Lett. 81, 4724 (1998). [59] A. Pourret, H. Aubin, J. Lesueur, C. A. Marrache-Kikuchi, L. Berge, L. Dumoulin, and K. Behnia, Nat. Phys. 2, 683 (2006); Phys. Rev. B 76, 214504 (2007). [60] A. Pourret, P. Spathis, H. Aubin, and K. Behnia, New J. Phys. 11, 055071 (2009). [61] Z. A. Xu, N. P. Ong, Y. Wang, T. Kakeshita, and S. Uchida, Nature (London) 406, 486 (2000). [62] Y. Wang, Z. A. Xu, T. Kakeshita, S. Uchida, S. Ono, Y. Ando, and N. P. Ong, Phys. Rev. B 64, 224519 (2001). [63] Y. Wang, N. P. Ong, Z. A. Xu, T. Kakeshita, S. Uchida, D. A. Bonn, R. Liang, and W. N. Hardy, Phys. Rev. Lett. 88, 257003 (2002). [64] C. Capan, K. Behnia, J. Hinderer, A. G. M. Jansen, W. Lang, C. Marcenat, C. Marin, and J. Flouquet, Phys. Rev. Lett. 88, 056601 (2002). [65] H. H. Wen, Z. Y. Liu, Z. A. Xu, Z. Y. Weng, F. Zhou, and Z. X. Zhao, Europhys. Lett. 63, 583 (2003). [66] Z. A. Xu, J. Q. Shen, S. R. Zhao, Y. J. Zhang, and C. K. Ong, Phys. Rev. B 72, 144527 (2005). [67] Y. Wang, L. Li, and N. P. Ong, Phys. Rev. B 73, 024510 (2006). [68] P. Li and R. L. Greene, Phys. Rev. B 76, 174512 (2007). 78 [69] F. F. Tafti, F. Laliberte, M. Dion, J. Gaudet, P. Fournier, and L. Taillefer, Phys. Rev. B 90, (2014). [70] R. Bel, H. Jin, K. Behnia, J. Glouguet, and P. Lejay, Phys. Rev. B 70, 220501(R) (2004). [71] R. Bel, K. Behnia, Y. Nakajima, K. Izawa, Y. Matsuda, H. Shishido, R. Settai, and Y. Onuki, Phys. Rev. Lett. 92, 217002 (2004). [72] Y. Onose, L. Li, C. Petrovic, and N. P. Ong, Europhys. Lett. 79, 17006 (2007). [73] I. Ussishkin, S. L. Sondhi, and D. A. Huse, Phys. Rev. Lett. 89, 287001 (2002). [74] I. Ussishkin, Phys. Rev. B 68, 024517 (2003). [75] V. Oganesyan and I. Ussishkin, Phys. Rev. B 70, 054503 (2004). [76] S. A. Hartnoll, P. K. Kovtun, M. Muller, and S. Sachdev, Phys. Rev. B 76, 144502 (2007). [77] D. Podolsky, S. Raghu, and A. Vishwanath, Phys. Rev. Lett. 99, 117004 (2007). [78] S. Raghu, D. Podolsky, A. Vishwanath, and D. A. Huse, Phys. Rev. B 78, 184520 (2008). [79] A. Sergeev, M. Yu. Reizer, and V. Mitin, Phys. Rev. B 77, 064501 (2008). [80] M. N. Serbyn, M. A. Skvortsov, A. A. Varlamov, and V. Galitski, Phys. Rev. Lett. 102, 067001 (2009). [81] K. Michaeli and A. M. Finkel’stein, Europhys. Lett. 86, 27007 (2009); Phys. Rev. B 80, 214516 (2009). [82] I. Martin and C. Panagopoulos, Europhys. Lett. 91, 67001 (2010). [83] A. Levchenko, M. R. Norman, and A. A. Varlamov, Phys. Rev. B 83, 020506(R) (2011). [84] N. R. Cooper, B. I. Halperin, and I. M. Ruzin, Phys. Rev. B 55, 2344 (1997). [85] L. G. Aslamazov and A. I. Larkin, JETP 40, 321 (1975). 79 [86] S. Hikami, Phys. Rev. B 24, 2671 (1981). [87] I. V. Lerner, A. A. Varlamov, and V. M. Vinokur, Phys. Rev. Lett. 100, 117003 (2008). [88] J. Mosqueira, J. D. Dancausa, and F. Vidal, Phys. Rev. B 84, 174518 (2011). [89] S. Goswami, C. Siegert, M. Pepper, I. Farrer, D. A. Ritchie, and A. Ghosh, Phys. Rev. B 83, 073302 (2011). [90] A. M. Clogston, Phys. Rev. Lett. 9, 266 (1962). [91] B. S. Chandrasekhar, Appl. Phys. Lett. 1, 7 (1962). [92] X. S. Wu, P. W. Adams, Phys. Rev. Lett. 73, 1412 (1994). [93] P. W. Adams, P. Herron, and E. I. Meletis, Phys. Rev. B 58, R2952 (1998). [94] V. Yu. Butko, P. W. Adams, and I. L. Aleiner, Phys. Rev. Lett. 82, 4284 (1999). [95] T. Suzuki, Y. Seguchi and T. Tsuboi, J. Phys. Soc. Jpn. 69, 1462 (2000). [96] X. S. Wu, P. W. Adams, and G. Catelani, Phys. Rev. B. 74, 144519 (2006). [97] P. Fulde and R. A. Ferrel, Phys. Rev. 135, 550 (1964); [98] A. I. Larkin and Yu. N. Ovchinnikov, Sov. Phys. JETP 20, 762 (1965). [99] P. Fulde, Adv. Phys 22, 667 (1973). [100] A. A. Abrikosov and L. P. Gor’kov, Sov. Phys. JETP 12, 1243 (1961). [101] P. W. Adams, Phys. Rev. Lett. 92, 067003 (2004). [102] G. Catelani, Y. M. Xiong, X. S. Wu, and P. W. Adams, Phys. Rev. B 80, 054512 (2009). [103] I. L. Aleiner and B. L. Altshuler, Phys. Rev. Lett. 79, 4242 (1997) [104] H.-Y. Kee, I. L. Aleiner, and B. L. Altshuler, Phys. Rev B 58, 5757 (1998). 80 [105] G. Catelani, Phys. Rev. B 73, 020503(R) (2006). [106] H. J. Gardner, A. Kumar, L. Yu, P. Xiong, M. P. Warusawithana, L. Wang, O. Vafek, and D. G. Schlom, Nat. Phys. 7, 895 (2011). [107] K. Michaeli, A. C. Potter, P. A. Lee, Phys. Rev. Lett. 108, 117003 (2012). [108] V. F. Gantmakher, S. N. Ermolov, G. E. Tsydynzhapov, A. A. Zhukov, T. I. Baturina, JETP Lett. 77, 424 (2003); JETP Lett. 71, 473 (2000). [109] K. A. Parendo, L. M. Hernandez, A. Bhattacharya, and A. M. Goldman, Phys. Rev. B 70, 212510 (2004). [110] V. M. Galitski and A. I. Larkin, Phys. Rev. B 63, 174506 (2001). [111] A. V. Lopatin, N. Shah, and V. M. Vinokur, Phys. Rev. Lett. 94, 037003 (2005). [112] A. Glatz, A. A. Varlamov, and V. M. Vinokur, Phys. Rev. B 84, 104510 (2011). [113] P. Fulde and K. Maki, Z. Physik 238, 233 (1970). [114] K. Aoi, R. Meservey and P. M. Tedrow, Phys. Rev. B 9, 875 (1974). [115] P. M. Tedrow and R. Meservey, Phys. Rev. B 16, 4825 (1977). [116] A. G. Aronov, S. Hikami, and A. I. Larkin, Phys. Rev. Lett. 62, 965 (1989). [117] R. Ramazashvili and P. Coleman, Phys. Rev. Lett. 79, 3752 (1997). [118] V. P. Mineev and M. Sigrist, Phys. Rev. B 63, 172504 (2001). [119] V. E. Kravtsov and M. R. Zirnbauer, Phys. Rev. B 46, 4332 (1992). [120] T. Yamashita, Y. Shimoyama, Y. Haga, T. D. Matsuda, E. Yamamoto, Y. Onuki, H. Sumiyoshi, S. Fujimoto, A. Levchenko, T. Shibauchi, Y. Matsuda, Preprint – arXiv:1411.1233. [121] H. Sumiyoshi and S. Fujimoto, Preprint – arXiv:1403.3977. 81 [122] S. Kasahara, T. Watashige, T. Hanaguri, Y. Kohsaka, T. Yamashita, Y. Shimoyama, Y. Mizukami, R. Endo, H. Ikeda, K. Aoyama, T. Terashima, S. Uji, T. Wolf, H. v. Lohneysenn, T. Shibauchi, Y. Matsuda, Preprint – arXiv:1411.1232. [123] A. G. Aronov, S. Hikami, and A. I. Larkin, Phys. Rev. B 51, 3880 (1995). 82