fundamental studies of superconductors using scanning ... · fundamental studies of superconductors...

37
Fundamental studies of superconductors using scanning magnetic imaging This article has been downloaded from IOPscience. Please scroll down to see the full text article. 2010 Rep. Prog. Phys. 73 126501 (http://iopscience.iop.org/0034-4885/73/12/126501) Download details: IP Address: 171.67.102.110 The article was downloaded on 06/03/2012 at 17:44 Please note that terms and conditions apply. View the table of contents for this issue, or go to the journal homepage for more Home Search Collections Journals About Contact us My IOPscience

Upload: others

Post on 10-Jul-2020

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Fundamental studies of superconductors using scanning magnetic imaging

This article has been downloaded from IOPscience. Please scroll down to see the full text article.

2010 Rep. Prog. Phys. 73 126501

(http://iopscience.iop.org/0034-4885/73/12/126501)

Download details:

IP Address: 171.67.102.110

The article was downloaded on 06/03/2012 at 17:44

Please note that terms and conditions apply.

View the table of contents for this issue, or go to the journal homepage for more

Home Search Collections Journals About Contact us My IOPscience

Page 2: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

IOP PUBLISHING REPORTS ON PROGRESS IN PHYSICS

Rep. Prog. Phys. 73 (2010) 126501 (36pp) doi:10.1088/0034-4885/73/12/126501

Fundamental studies of superconductorsusing scanning magnetic imagingJ R Kirtley

Center for Probing the Nanoscale, Stanford University, Stanford, CA, USA

Received 26 June 2010, in final form 4 August 2010Published 18 November 2010Online at stacks.iop.org/RoPP/73/126501

AbstractIn this review I discuss the application of scanning magnetic imaging to fundamental studies ofsuperconductors, concentrating on three scanning magnetic microscopies—scanning SQUIDmicroscopy (SSM), scanning Hall bar microscopy (SHM) and magnetic force microscopy(MFM). I briefly discuss the history, sensitivity, spatial resolution, invasiveness and potentialfuture developments of each technique. I then discuss a selection of applications of thesemicroscopies. I start with static imaging of magnetic flux: an SSM study provides deeperunderstanding of vortex trapping in narrow strips, which are used to reduce noise insuperconducting circuitry. Studies of vortex trapping in wire lattices, clusters and arrays ofrings and nanoholes show fascinating ordering effects. The cuprate high-Tc superconductorsare shown to have predominantly d-wave pairing symmetry by magnetic imaging of thehalf-integer flux quantum effect. Arrays of superconducting rings act as a physical analog forthe Ising spin model, with the half-integer flux quantum effect helping to eliminate one sourceof disorder in antiferromagnetic arrangements of the ring moments. Tests of the interlayertunneling model show that the condensation energy available from this mechanism cannotaccount for the high critical temperatures observed in the cuprates. The strong divergence inthe magnetic fields of Pearl vortices allows them to be imaged using SSM, even for penetrationdepths of a millimeter. Unusual vortex arrangements occur in samples comparable in size tothe coherence length. Spontaneous magnetization is not observed in Sr2RuO4, which isbelieved to have px ± ipy pairing symmetry, although effects hundreds of times bigger thanthe sensitivity limits had been predicted. However, unusual flux trapping is observed in thissuperconductor. Finally, unusual flux arrangements are also observed in magneticsuperconductors. I then turn to vortex dynamics: imaging of vortices in rings of highlyunderdoped cuprates places limits on spin-charge separation in these materials. Studies ofspontaneous generation of fluxoids upon cooling rings through the superconducting transitionprovide clues to dynamical processes relevant to the early development of the universe, whilestudies of vortex motion in cuprate grain boundaries allow the measurement of current–voltagecharacteristics at the femtovolt scale for these technologically important defects. ScanningSQUID susceptometry allows the measurement of superconducting fluctuations on samplescomparable in size to the coherence length, revealing stripes in susceptibility believed to beassociated with enhanced superfluid density on the twin boundaries in the pnictidesuperconductor Co doped Ba-122, and indicating the presence of spin-like excitations, whichmay be a source of noise in superconducting devices, in a wide variety of materials. Scanningmagnetic microscopies allow the absolute value of penetration depths to be measured locallyover a wide temperature range, providing clues to the symmetry of the order parameter inunconventional superconductors. Finally, MFM tips can be used to manipulate vortices,providing information on flux trapping in superconductors.

(Some figures in this article are in colour only in the electronic version)

This article was invited by Sean Washburn.

0034-4885/10/126501+36$90.00 1 © 2010 IOP Publishing Ltd Printed in the UK & the USA

Page 3: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Contents

1. Introduction 22. Fundamentals 3

2.1. Scanning SQUID microscopy 32.2. Hall bar microscopy 72.3. Magnetic force microscopy 72.4. Summary 10

3. Applications 103.1. Static imaging of magnetic flux 10

3.2. Fluxoid dynamics 19

3.3. Local susceptibility measurements 22

3.4. Penetration depths 25

3.5. Manipulation of individual vortices 274. Conclusions 28Acknowledgments 28References 28

1. Introduction

Superconductivity, the complete absence of electricalresistance in some metals below a critical temperature Tc, isone of the best understood phenomena in solid-state physics.As described in the Bardeen–Cooper–Schrieffer (BCS) theory[1], superconductivity occurs in conventional superconductorsthrough phonon-mediated pair-wise interaction of chargecarriers to form Cooper pairs. The characteristic energy neededto form a single particle excitation from the superconductingcondensate is the energy gap �. The Cooper pairs form amacroscopic quantum state characterized by a wave functionψ(�r) = |ψ(�r)|eiφ(�r), where φ is the Cooper pair phase,with the density of Cooper pairs ns(�r) = |ψ(�r)|2. Themacrosopic quantum nature of the pairing state leads to a widevariety of behaviors: persistent currents, Meissner screening,flux quantization, the Josephson effects and superconductingquantum interference. Superconductivity has become a multi-billion dollar industry, with applications in medical imaging,cell phone filters, electrical motors, power transmission,transportation, magnetic manipulation, magnetic sensing, etc.

However, there is still much to learn about super-conductivity: although superconductivity in the copper-oxidebased perovskites [2, 3] was discovered in 1986, despitealmost 25 years of intense effort a consensus on the pairingmechanism in the cuprates has yet to emerge. The heavyFermion superconductors [4, 5] have large carrier masses,strong interaction between spin and charge degrees offreedom, and potentially a wide variety of Cooper pairingsymmetries. The non-cuprate perovskite superconductorSr2RuO4 [6] is believed to have a p-wave pairing state thatbreaks time-reversal symmetry [7]. Interest in unconventionalsuperconductors has been reawakened with the discovery ofiron-based compounds with high critical temperatures [8].

Scanning magnetic microscopies have played, and willcontinue to play, an important part in our efforts to understandsuperconductivity. There are now a large number of scanningmagnetic microscopies. Although it is beyond the scope of thisreview to cover all of these techniques in detail, it is appropriateto briefly describe some, and provide a few citations as anintroduction to the literature.

In Lorentz microscopy [9–12], a coherent beam of highenergy (300 keV) electrons is passed through a thin sampleand then imaged slightly off the focal plane. Magnetic fieldsintroduce small phase shifts which cause interference in thedefocussed image. Lorentz microscopy can image vorticeswith deep sub-micrometer spatial resolution at video rates and

is sensitive not only to surface fields but also to fields withinthe sample. However, the sample must be thinned.

In magneto-optic microscopy [13–15], the sample isilluminated with linearly polarized light. It passes through amagneto-optically active layer, reflects off the sample surface,passes again through the magneto-optically active layer andthen through a crossed polarizer. Regions of the samplewith magnetic fields produce rotations of the polarization axisof the light, which is detected as a bright spot. Magneto-optic microscopy has the advantage of relative simplicity, buthas limited spatial resolution and sensitivity. Nevertheless,observation of individual superconducting vortices with0.8 µm spatial resolution has recently been reported [15].

In Sagnac interferometry [16–18], a single beam of light issplit into two components which travel along identical paths inopposite directions around a loop. Any effects which break thetime-reversal symmetry, including the magneto-optical effect,will cause interference, which can be sensitively detected.High-spatial resolution can be achieved using scanning near-field optical microscopy [17].

In spin polarized scanning tunneling microscopy [19–21]a scanning tunneling microscope has a tunneling tip witha ferromagnetic or antiferromagnetic atom from which thetunneling electrons emerge. The tunneling probability dependson the spin polarization of the sample, and therefore contrastis achieved if there are spatial variations in the atomic spinpolarizations. Spin polarized tunneling is useful for studyingsuperconductivity in, for example, the high-temperaturecuprate superconductors, since it is believed that magnetismand superconductivity are intimately related in these materials.However, the spin polarized tunneling current is not directlyrelated to the local magnetic fields.

In diamond nitrogen-vacancy microscopy [22–26], amagnetic field dependent shift in the energy levels of a singlenitrogen-vacancy center in diamond is detected optically.There are two basic schemes for detecting this energy shift.In the ‘dc’ technique, microwaves at a fixed frequency as wellas light are incident on the center. The intensity of the resultantfluorescence depends on the local magnetic field. This methodcan detect millitesla fields with, in principle, nanometer scalespatial resolution. In the ‘ac’ technique, spin-echo microwavetechniques are used to detect the modulation of the center’senergy levels by the local magnetic field. The ‘ac’ techniquehas a sensitivity of 60 nT, with in principle a spatial resolutionof a few nanometers, but has reduced sensitivity at zerofrequency. These techniques have the advantages of goodsensitivity, small invasiveness, room temperature operation,

2

Page 4: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

inherently high speed and potentially high-spatial resolution.It remains to be seen whether single nitrogen-vacancies innanometer sized diamond crystallites can be placed at theend of scanning tips while still retaining long lifetimes. Thisis necessary to achieve high-spatial resolution and sensitivitysimultaneously.

A magnetic tunnel junction [27–29] is a planar structurewith two ferromagnetic electrodes. The tunneling currentthrough the junction depends on the relative alignment of themagnetic moments of the electrodes, and therefore dependson the magnetic field environment. Magnetic tunnel junctionssuitable for scanning microscopy are commercially available.They currently have lateral dimensions (and therefore spatialresolutions) of about 4 µm, and field noises at 100 Hz of about5 × 10−8 T Hz−1/2 [29]. They have the advantage of roomtemperature operation, but are not currently as sensitive asHall bars or SQUIDs.

There are already excellent reviews covering the topicof magnetic imaging of superconductors [30–32], and therehas been an enormous amount of work in this area. I willtherefore not attempt to survey the field exhaustively, but willonly cover recent work in three scanning magnetic microscopytechniques: scanning SQUID, scanning Hall bar and magneticforce microscopies. I will present a short introduction to thebasics, advantages and shortcomings of each technique, andthen present a selection of fundamental applications of thesemagnetic microscopies in the area of superconductivity.

2. Fundamentals

2.1. Scanning SQUID microscopy

Josephson [33] predicted in 1962 the possibility of coherenttunneling of Cooper pairs between two weakly coupledsuperconductors, and wrote down the fundamental relations:

Is = I0 sin ϕ,

V = h

2e

dt, (1)

where Is is the supercurrent through the weak link, ϕ is thedifference between the Cooper pair phases on the two sides ofthe weak link, V is the voltage across the weak link and I0 is thecritical current, the maximum current that can pass through theweak link before a voltage develops. A physical Josephsonjunction can be modeled as an ideal Josephson element inparallel with a capacitor and a resistor in the resistively shuntedjunction (RSJ) model [34, 35]. This model obeys the equation

IB = h

2eR

dt+ I0 sin ϕ +

hC

2e

d2ϕ

dt2+ In, (2)

where IB is the bias current through the weak link, R is theresistance, C is the capacitance and In is a noise current, givenfor the case of an ideal Johnson noise source resistor by

〈In〉2 = 4kBT �f/R, (3)

where 〈In〉2 is the time-averaged current noise squared and�f is the frequency band width over which the current noise

is measured. This equation has the form of the equation ofmotion of a driven, damped harmonic oscillator. The dynamicsof a Josephson weak link can be visualized as that of a particlemoving in a sinusoidal ‘washboard’ potential with oscillationamplitude I0�0/2π , where �0 = h/2e is the superconductingflux quantum, and the average slope of the potential is given by〈dE/dϕ〉 = �0(IB−In)/2π . In the absence of a noise current,the particle will escape from a local potential minimum and rundown the washboard potential if the bias current IB is greaterthan I0, in the process developing a voltage given by the secondof equations (1). The solutions to equation (2) can be dividedinto two classes, according to whether the Stewart–McCumberparameter βc = 2πI0R

2C/�0 is greater or less than 1. Ifβc < 1, a running particle retraps into a local potentialminimum (V = 0), as soon as IB = I0, and the current–voltage characteristic is non-hysteretic. On the other hand, ifβc > 1, a running particle is not retrapped until IB < I0, andthe current–voltage characteristic is hysteretic. In the presenceof noise, the SQUID can be thermally excited out of the zero-voltage state at bias current IB < I0 [36]. In addition, the weaklink can macroscopically quantum tunnel through the barrierfrom the zero-voltage to the voltage state [37].

A superconducting quantum interference device (SQUID)is a superconducting ring interrupted by at least one Josephsonweak link. For SQUID microscope applications a SQUID withtwo weak links is most often used. Such a SQUID is labelleda dc SQUID, since the modulation of the SQUID criticalcurrent with applied magnetic field can be observed usingtime-independent bias currents. Consider a SQUID with twoweak links 1,2 with critical currents Ic1, Ic2, phase drops acrossthe weak links ϕ1, ϕ2 and inductances in the two arms of theSQUID L1, L2. Neglecting for the moment the current throughthe resistive and capacitive elements of the junctions, the biascurrent across the SQUID IB and the magnetic flux throughthe SQUID � are given by

IB = I1 sin ϕ1 + I2 sin ϕ2,

� = �a − L1I1 + L2I2, (4)

where �a is the externally applied flux. Integrating allthe changes in phase around the SQUID loop along a pathsufficiently deep inside the superconductor that there are nosupercurrents along the integration path (i.e. they are shieldedout by the bulk of the superconductor), and using the factthat the canonical momentum of the Cooper pairs is given by�pcanonical = �p + q �A, where q = 2e is the charge on the Cooperpairs and �A is the vector potential, and finally by insisting thatthe Cooper pair wavefunction be single valued, leads to

2πn = ϕ2 − ϕ1 +2π

�0(�a + L2I2 − L1I1), (5)

where n an integer. Solutions to equations (4) and (5) areshown in figure 1 for the case of symmetric weak-link criticalcurrents (Ic1 = Ic2) and arm inductances (L1 = L2). Theparameter γ = 2πLI0/�0 (L = L1 + L2) determines thedepth of the modulation of the SQUID critical current with flux:larger SQUID inductances lead to smaller modulations. The

3

Page 5: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 1. Plot of the critical current Ic of a symmetric(L1 = L2 = L/2, I1 = I2 = I0) 2-weak link SQUID as afunction of applied flux, for different values of the parameterγ = 2πLI0/�0. I0 is the critical current of one of the weak linksand L is the total inductance of the SQUID.

solutions for non-symmetric SQUIDs are more complicated,but qualitatively similar [38] to those shown here.

There are four relevant sources of noise in dc SQUIDs:Johnson noise arises from thermal fluctuations, assumed to bedominated by the shunt resistor in the RSJ model [39]; shotnoise arises from the discrete charge of the quasiparticles [40]and Cooper pairs [41] traversing the weak links; quantumnoise arises from zero point motion in the shunt resistors [42]and 1/f noise can arise from a number of sources. Tescheand Clarke calculate [39] that the signal-to-noise ratio in dcSQUIDs is optimized when βL = 2LI0/�0 ≈ 1. In addition,the Stewart–McCumber parameter βc should be about 1, sothat the SQUID is just non-hysteretic. If these two conditionsare met, the optimal flux noise power S� from the first threesources is given by

S� =

4kBT L(πLC)1/2 Johnson noise,hL shot noise,hL quantum noise.

(6)

Typical values for scanning SQUID sensors using aNb–Al2O3–Nb trilayer process are L = 100 pH, I0 =6 µA, C = 8.36 × 10−13 F, R = 5 and T = 4.2 K.These parameters lead to �Johnson = √

S� = 3 ×10−7�0 Hz−1/2 for the flux noise due to Johnson noise; �shot =1.2 × 10−7�0 Hz−1/2 for shot noise and �quantum = 5 ×10−8�0 Hz−1/2 for noise due to zero point motion.

In the frequency range of interest Johnson noise, shot noiseand quantum noise are ‘white’: the frequency distribution ofthe noise is independent of frequency f . 1/f noise, as its nameimplies, has a noise amplitude that follows a 1/f dependence.It almost always dominates the noise in SQUIDs at sufficientlylow frequencies. In the Dutta, Dimon and Horn model [43] 1/f

noise results from the superposition of ‘shot-like’ contributionsfrom two-level traps with a large distribution of trap energies.These traps could be, for example, defects in the oxide in atunnel junction, or trap sites for superconducting vortices in thesuperconducting bulk making up the SQUID. Koch et al [44]have suggested that 1/f noise in SQUIDs could result from

coupling of flux into the SQUID from electronic spins with awide distribution of characteristic times.

As can be seen in figure 1, the critical current of theSQUID is periodic in flux, with a period given by �0. Themost commonly used scheme for determining this magneticflux uses ac flux modulation in a flux locked loop [45, 46].If the SQUID is current biased just above the critical currentthe voltage varies sinusoidally with flux. If the dc componentof the total (external+feedback) flux is such that the SQUIDvoltage is at an extremum, the first harmonic voltage responsewill be zero and any flux change will give a linear response.Negative feedback on the dc component of the flux is then usedto keep the response zeroed: this feedback signal is linearlyproportional to the flux threading the SQUID, with a constantof proportionality that can easily be determined with greatprecision. Traditionally the voltage signal from the SQUID hasbeen amplified at low temperatures using a tuned LC circuit ora transformer, and at room temperature using phase-sensitivelock-in detection. However, SQUID array amplifiers can alsobe used [47, 48]. These have the advantage of bandwidths ofover 100 MHz, and no requirement for ac flux modulation.

For this review I will concentrate on high-spatialresolution magnetic microscopies. High-spatial resolution isachieved by placing a small sensor close to the sample. Inthe case of SQUID microscopy this means making the SQUIDsmall, or making a pickup loop integrated into the SQUIDsmall. The scaling of the flux signal and spatial resolutionwith sensor size depends on the type of field source. For thisreview I will concentrate on three basic field sources: a pointdipole �m generates the magnetic induction

�B = µ0

3r(r · �m) − �mr3

(dipole field), (7)

where r is the direction and r is the magnitude of the vectorbetween the dipole source and the point of interest. The mag-netic flux �s through a square area of side d oriented parallel tothe xy plane and centered at �r = zz above a dipole at the originwith moment m oriented parallel to the z-axis is given by

�s = 2µ0md2

π√

d2/2 + z2(d2 + 4z2)(dipole flux). (8)

In the limit z → 0 �s → 2√

2µ0m/πd.A superconducting vortex generates the induction

�B = �0

2πr2r (monopole field), (9)

where �0 = h/2e is the total magnetic flux generated by thevortex. The magnetic flux �s through a square area of side d

oriented parallel to the xy plane and centered at �r = zz abovea vortex at the origin is given by

�s = 2�0

πtan−1

(d2

2z√

2d2 + 4z2

)(monopole flux).

(10)

In the limit z → 0 �s → �0. Finally, an infinitely long andnarrow line of current I in the y direction at x = 0, z = 0

4

Page 6: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 2. Plot of the flux through a square area of side d a height z above a (a) dipole source with moment m oriented normal to the scanplane (b) a monopole source with flux �0 and (c) a line source with current I . For a given value of z/d the flux signal scales like d−1 for adipole source, is independent of d for a monopole source and scales like d for a current line source.

0.0 0.5 1.0 1.5 2.010–2

10–1

100

z/d

p/p

)1.0=d/z(

Current lineMonopoleDipole

0.0 0.5 1.0 1.5 2.01

2

3

z/d

d/x(a) (b)

Figure 3. (a) Plot of the ratio of the flux through a square area of side d at a height z to that at a height z/d = 0.1 for various sources offield. (b) Full width at half-maximum for the dipole and monopole field sources, and the peak-to-peak distance for the current line source, ofthe flux through a square area of side d as a function of height z.

generates the field

�B = µ0I

2πrθ (current field), (11)

where θ = (xz−zx)/r . The magnetic flux �s through a squarearea of side d oriented parallel to the xy plane at a height z

above this line has peaks at x = ±√d2 + z2/2 with magnitude

�s = µ0Id

4πln

[(d − √

d2 + z2)2 + 4z2

(d +√

d2 + z2)2 + 4z2

](current flux).

(12)

In the limit z → 0 the magnetic fluxes at the peaks divergelogarithmically.

Figure 2 plots the calculated flux through a square area ofside d, as a function of x, the lateral position of the SQUIDrelative to the field source, for various spacings z betweenpickup loop and sample, for these three different sources offield. For a given value of z/d for a dipole source the flux signalthrough the SQUID gets larger as d gets smaller, scaling like1/d. For a monopole source, such as a superconducting vortex,the peak flux is independent of d at constant z/d. For a line

of current, the peak flux signal increases like d as the pickuploop diameter increases. In all cases the spatial resolution isabout d for z � d.

Figure 3(a) plots the ratio of the peak flux at a sensorheight z to the peak flux at z/d = 0.1 for the three different fieldsources, while figure 3(b) plots the full width at half-maximum�x for the dipole and monopole field sources, and the peak-to-peak distance for the current line source, as a function of sensorheight z. The peak amplitude falls off nearly exponentiallywith sensor height for a dipole source, but falls off less quicklyfor a monopole or current line source. Similarly, the spatialresolution falls off quickly with sensor height for the dipolesource, but less quickly for the monopole and current linesources.

The same qualitative conclusions can be made for morecomplex sources of field [49]. Consider a geometry in whichthe sample takes up a half-space z < 0. For the static case themagnetic fields in the half-space z > 0 can be written as thegradient of a scalar potential ϕm: �B = ��ϕm, where �2ϕm = 0.If we resolve the fields at the surface of the sample into Fourier

5

Page 7: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

components

�bk(z = 0) = 1

(2π)2

∫ ∞

−∞

∫ ∞

−∞�B(x, y, z = 0)ei(kxx+kyy) dx dy,

(13)

each of the components of the field for z > 0 will be of the form�bk(z) = �bk(z = 0)e−kz, where k =

√k2x + k2

y : The fields will

decay exponentially as z gets larger, with the higher Fouriercomponents decaying more rapidly. Therefore to get optimalsensitivity and spatial resolution, the sensor must be placedclose to the sample.

It was recognized almost immediately after the demon-stration of the Josephson effect [50] and superconductingquantum interference effects [51] that local magnetic fieldscould be imaged by scanning samples and SQUIDs relativeto each other [52]. However, it was not until 1983 that thefirst two-dimensional scanning SQUID microscope was builtby Rogers and Bermon at IBM Research [53]. This instrumentwas used to image superconducting vortices in association withthe IBM Josephson computer program. Other notable earlyefforts in SQUID microscopy were by the Wellstood group atthe University of Maryland [54–56], the van Harlingen groupat the University of Illinois [57, 58], the Wikswo group atVanderbilt University [59], the Clarke group at U.C. Berkeley[60] and the Kirtley group at IBM Research [61, 62]. Kirtleyand Wikswo [46] reviewed some of fundamentals and earlywork concerning scanning SQUID microscopy.

There are several competing strategies for achieving goodspatial resolution in a SQUID microscope sensor. The firststrategy is to make the SQUID very small (µ-SQUID), withnarrow and thin constrictions for the Josephson weak links[63, 64]. This strategy has the advantage of simplicity,since only one level of lithography is required. The originalµ-SQUIDs had hysteretic current–voltage characteristics.This meant that flux-locked-loop feedback schemes could notbe used, and the flux sensitivity of these SQUIDs was relativelypoor. However, recently non-hysteretic, sub-micrometer sizedSQUIDs have been made [65–67]. Hao et al [67] madetheir SQUIDs non-hysteretic using a second layer of tungstento shunt them, and reported white noise floor levels of0.2µ�0 Hz−1/2 in a SQUID with a diameter of about 370 nm.

A second strategy is a self-aligned SQUID recentlyreported by Finkler et al [68]. In this work three aluminumevaporations are made onto a quartz tube that has been pulledinto a sharp tip with apex diameter between 100 and 400 nm.The first two evaporations, performed at an angle of 100◦

degrees relative to the axis of the tube, form the leads. Athird evaporation along the tip axis forms a ring with two weaklinks to the leads, forming a dc SQUID. Finkler et al report non-hysteretic current–voltage characteristics and a flux sensitivityof 1.8×10−6�0 Hz−1/2 for SQUIDs with an effective area of0.34 µm2, operating at fields up to 0.6 T.

A third strategy is to make a more conventional SQUID,but to have well shielded superconducting leads to a smallpickup loop integrated into it [62, 69, 70] (see figure 4).This has the disadvantage of complexity, since multiplelevels of metal are required to shield the pickup loop leads,but has the advantages of reduced interaction between the

Figure 4. Advanced integrated scanning SQUID susceptometerdesign. (a) Diagram of a counterwound SQUID susceptometer.(b) Optical micrograph of pickup loop region of an optically definedSQUID susceptometer. The outer ring is the field coil and the innerring is the flux pickup loop of the SQUID. The edge of the siliconsubstrate is just visible on the left side. The inset shows an atomicforce microscopy cross-section of the structure along a horizontalline through the center of the pickup loop. The pickup loop isclosest to the sample when the tip is aligned at a 2.5◦ angle.(c) Design for a SQUID sensor in which the pickup loop is definedusing focussed ion beam (FIB) etching. In this case the pickup loopcan touch down first if the alignment angle is between 2◦ and 5◦.(d) Scanning electron microscopy images of the sensor after FIBdefinition of the pickup loop. Figure reprinted with permissionfrom [71]. Copyright 2008 by the American Institute of Physics.

SQUID and the sample and ease of using flux feedbackschemes. Figure 4 illustrates the properties of current advancedscanning SQUID sensor design using the integrated pickuploop strategy [71]. This SQUID sensor uses a Nb–AlOx–Nb trilayer for the junctions, two levels of Nb for wiring andshielding and SiO2 for insulation between the various levels,as well as local field coils integrated into the pickup loopregion, to allow local magnetic susceptibility measurements[69, 72]. High symmetry in the SQUID and field coil layout(see figure 4(a)) and tapered terminations are desirable todiscriminate against background fields and reduce unwantedelectromagnetic resonances [73]. Figure 4(b) is an opticalmicrograph of a scanning SQUID susceptometer in whichboth the field coil and pickup loop were defined using opticallithography. Figure 4(c) shows the layout for a device inwhich the pickup loop area is left as a ‘tab’ in the opticallithography step, and then patterned into a loop using focussedion beam lithography [74]. The completed device is shownin a scanning electron micrograph in figure 4(d). This deviceuses a non-planarized process, meaning that each successivelayer conforms to the topography of the previous one. Thisleads to constraints on the alignment angles that can be used toget the minimum spacing between pickup loop and sample, asillustrated in the insets in figures 4(b) and (c). These constraintscan be reduced using a planarized process [75], in which

6

Page 8: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 5. Scanning electron microscope image of a 100 nm Hallprobe. The four leads are separated by narrow etch lines. The gateshields the Hall cross from stray electrical charges and allows themodulation of the 2DEG beneath. The Hall probe will touch thesample surface at the contact tip. Figure reprinted with permissionfrom [87]. Copyright 2007 by the American Institute of Physics.

the insulating layers are chemically–mechanically polished toplanes before successive Nb wiring levels are added.

If we assume a SQUID or SQUID pickup loop diameterof 1 µm, a SQUID-sample spacing of 0.1 µm, and a SQUIDnoise level of 10−6�0 Hz−1/2, and using the peak signalsfrom figure 2, we find a minimum detectable dipole signal of139 electron spins Hz−1/2, a minimum flux signal of 1.25 ×10−6�0 Hz−1/2 and a minimum detectable line of currentsignal of 1.65 nA Hz−1/2.

2.2. Hall bar microscopy

A Hall bar develops a transverse voltage Vx = −IyBz/n2de

when a current Iy passes through it in the presence of amagnetic induction Bz perpendicular to the plane of the Hallbar [30–32], where n2d is the carrier density per unit area of theHall bar. Therefore, materials with small carrier densities suchas semi-metals, semi-conductors or two-dimensional electrongases at the interface between semi-conductors with differentbandgaps develop larger Hall voltages. Early scanning Hallbar systems used evaporated films of bismuth [76], InSb[77] or GaAs [78]. More recently high sensitivity andspatial resolution have been achieved with GaAs/AlxGa1−xAsheterojunction structures [79–84]. In addition, 250 nm sizedscanning probes using GaSb/InAs/GaSb [83] and micrometer-scale Si/SiGe and InGaAs/InP Hall crosses have beencharacterized at low temperatures [85, 86]. Hicks et al [87]have fabricated Hall bars as small as 85 nm on a side (seefigure 5) from GaAs/AlxGa1−xAs heterojunction electron gasmaterial, and estimate a field noise of 500 µT Hz−1/2 and aspin sensitivity of 1.2 × 104µB Hz−1/2 at 3 Hz and 9 K forsensors 100 nm on a side. Sandhu et al have fabricated 50 nmbismuth Hall bar sensors with a noise of 0.8G Hz−1/2 [88].At present Hall bars from GaAs/AlxGa1−xAs heterojunctionsare less sensitive above about 100 K because of thermalexcitations. Scanning Hall bar sensors made of InAs have

been used to image single magnetic biomolecular labels atroom temperature [89].

Magnetic tunnel junctions and Hall bars in the ballisticregime [90–92] produce a signal which is proportional to 〈Bz〉,the magnetic field perpendicular to the sensor, averaged overthe sensor. The situation is more complicated for a Hall barin the diffusive regime [93], but may be qualitatively similar.The average fields 〈Bz〉 for the three sources of field canbe inferred from equations (8), (10) and (12) (figure 2), bysubstituting �s → 〈Bz〉d2 and in figure 3 by substituting�p → 〈Bz〉p, where 〈Bz〉p is the peak value of the fieldaveraged over the sensor area. For a given value of z/d, 〈Bz〉is proportional to 1/d3 for a dipole source, proportional to1/d2 for a monopole source and proportional to 1/d for a lineof current source. Boero et al [94] estimate that the optimalJohnson noise limited minimum detectable magnetic field in amicro-Hall cross geometry is given by

Bmin ≈√

4kBT R0

vsatw(thermal), (14)

where kB is Boltzmann’s constant, T is the operatingtemperature, R0 is the output resistance at zero magneticfield, vsat is the saturation carrier drift velocity and w is thewidth of the cross. This results in Bmin ∼ 2nT Hz−1/2

for w = 1 µm at 300 K for doped InSb. Thermal noiseis expected to be independent of sensor size, and thereforethe scaling for sensitivity is the same as for 〈Bz〉. At thelow temperatures used for imaging superconductors, transportin the Hall sensor is often quasi-ballistic and the signal tonoise is better characterized by a mobility and an optimumprobe current. The latter is typically defined as the maximumcurrent before the onset of pronounced 1/f noise due to one ofseveral possible sources, e.g. generation–recombination noiseat deep traps or across heterointerfaces, heating or impactionization. However, noise in micrometer-scale Hall barsis often dominated by 1/f noise [87, 92]. The minimumdetectable field can then be written as [89]

Bmin ≈ 1

µd

√αHGN�f

nf1/f, (15)

where µ is the mobility, d is the sensor size, αH is Hooge’s1/f noise parameter, GN ∼ 0.325 is a constant, n is thecarrier density, �f is the measurement bandwidth, and f is thefrequency. Equation (15) combined with equations (8), (10)and (12) imply that for a Hall bar dominated by 1/f noise,for a given z/d the minimum detectable dipole moment scaleslike d2, the minimum detectable flux scales like d, and theminimum detectable current is independent of d. It has beenreported that low frequency noise in Hall bar devices can besignificantly reduced by optimizing the voltage on a gate overthe Hall cross [87, 95] and that sufficiently small devices havenoise which is composed of a single Lorentzian spectrum [96].

2.3. Magnetic force microscopy

A magnetic force microscope [97–99] images small changes inthe resonance frequency of a cantilever due to the interactionbetween magnetic material on a sharp tip at the end of the

7

Page 9: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

zt

zt

zt

Figure 6. MFM modeling using the point dipole tip approximation: force derivative curves for (a) a dipole source located at the origin withmagnetic moment ms oriented parallel to z (solid line) or parallel to x (dashed line), equation (7), (b) a monopole source located at the originwith magnetic flux �0, equation (9) and (c) a line of current I along the y-axis, equation (11). The tip is modeled as a point dipole withmoment mt oriented parallel to the z-axis scanning a height z above the xy plane along the x-axis (equations (17)).

cantilever and local sample magnetic fields. The signal from amagnetic force microscope is proportional to dFtip,z/dz, thederivative of the z-component of the force on the tip withrespect to its z-position. Since this force sums contributionsfrom all the magnetic material in the tip, the sensitivity andspatial resolution of MFM depends on the shape of the tip.For some applications and tip materials the tip can be modeledas having a magnetic monopole at its apex, with the forceapproximated by Ftip,z = qBz, where q is the effectivemagnetic monopole moment of the tip [32, 100]. However,Lohau et al have shown that their experimental MFM data canbe modeled either using a monopole or a dipole approximationfor the end of the tip, but that in either case the effective positionof the monopole or dipole is a function of the field gradient fromthe sample [101].

It is instructive to model the ultimate spatial resolutionand sensitivity of an MFM in the point monopole and dipoleapproximations for the tip. The z-component of the forcederivative on the tip from a field source can be written as [101]

∂Fz

∂z= − q

µ0

∂Bz

dz+ mt,x

∂2Bx

∂z2+ mt,y

∂2By

∂z2+ mt,z

∂2Bz

∂z2, (16)

where q is the magnetic monopole flux and mt,i are the dipolemoments of the tip in the x, y, z directions. Assume forsimplicity that the tip is magnetized only in the z-direction.Then mt,x = mt,y = 0 and the force derivatives due to thedipole, monopole and current line sources of equations (7),(9), and (11) are given by

∂Fz

∂z= 3mt,zµ0

4πr9[−5(ms,xx + ms,yy)z(3r2 − 7z2)

+ ms,z(3r4 − 30r2z2 + 35z4)] (dipole–dipole)

= 3�0mt,zz

5z2 − 3r2

r7(dipole–monopole)

= −mt,zµ0Ix

π

x2 − 3z2

(x2 + z2)3(dipole–current line),

(17)where mt,z is the dipole moment of the tip.

Similarly, the expressions for the force derivative in themonopole approximation are given by

∂Fz

∂z= − 3q

4πr7[(xms,x + yms,y)(r

2 − 5z2)

+ ms,zz(3r2 − 5z2)] (monopole–dipole),

= − q�0

2πµ0

r2 − 3z2

r5(monopole–monopole),

= qI

π

xz

(x2 + z2)2(monopole–current line). (18)

The force derivative relations given by equations (17) and(18) are plotted in figures 6 and 7, respectively.

Within these approximations the force derivative signalstrength increases strongly with decreasing z—like 1/z5 fora dipole source, 1/z4 for a monopole source and 1/z3 for acurrent line source, in the point dipole tip approximation. In themonopole tip approximation the powers are 1/z4, 1/z3 and 1/z2

for the three sources, respectively. The widths of the predictedforce derivative features are proportional to the height z of thetip above the source. The constant of proportionality does notdepend strongly on source: the full width at half-maximum (ordistance between the maximum and minimum force derivativefor a line of current source) is �x/z=0.6, 0.66 and 0.74 for thedipole tip, and 0.74, 1 and 1.16 for the monopole tip, for thethree sources respectively. Of course, these conclusions wouldbe modified for a real tip with a finite size. It has been estimatedthat the effective tip–sample spacing and radius of curvaturefor conventional MFM tips are both about 10 nm [100].

In order to estimate the sensitivity of MFM, one needs toknow the minimum detectable force derivative, which dependson a number of characteristics of the MFM, including those ofthe cantilever. Using the lumped mass model for an MFMcantilever, the equation of motion of the tip end position z isthat of a driven, damped harmonic oscillator

md2z

dt2+

dz

dt+ kz = Fsignal(t) + Fthermal(t), (19)

8

Page 10: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 7. MFM modeling using the monopole tip approximation: MFM force derivative curves for (a) a dipole source located at the originwith magnetic moment ms oriented parallel to z (solid line) or parallel to x (dashed line), equation (7), (b) a monopole source located at theorigin with magnetic flux �0, equation (9) and (c) a line of current I along the y-axis, equation (11). The tip is modeled as a point monopolewith magnetic flux q scanning along the x-axis a height z above the xy plane (equations (18)).

where m is the effective mass of the cantilever, = k/ω0Q isthe damping constant, k is the cantilever spring constant, ω0 isthe angular resonant frequency, Q is the quality factor, Fsignal

is the signal force (in this case the magnetic interaction betweenthe tip and sample) and Fthermal is the force due to thermalfluctuations. The equipartition theorem implies that the powerspectral density of thermal fluctuations SF = 4 kBT . Thenthe thermally limited minimum detectable force derivative isgiven by

∂F

∂z min= 1

A

√4kkBT BW

ω0Q, (20)

where BW is the measurement bandwidth and A is theamplitude of oscillation of the cantilever.

A major thrust in research in MFM is to reduce the spatialextent of the magnetic material in the cantilever tip to improvespatial resolution and spin sensitivity. This is often done byshaping previously deposited material using focussed ion beametching [102, 103]. Sharply defined magnetic tips can alsobe produced by evaporating magnetic material on an electronbeam deposited carbon and ion etched needle [104] or onto acarbon nanotube [105] at the end of a conventional Si cantilevertip (see figure 8). Small amounts of magnetic material can alsobe deposited directly onto cantilever tips through nanoscaleholes fabricated in a stencil mask [106]. One approach wouldbe to place nano-magnets directly on the apex of cantilevertips. This would serve three goals: improved spatial resolution,improved spin sensitivity and improved interpretability, sincethe magnetic moment of the deposited nano-magnets wouldhave a narrow distribution [107]. However, such tips couldbe especially susceptible to switching of the tip momentin sufficiently strong magnetic fields, reversing the contrastof the MFM images [108]. Although this complicates theinterpretation of the images, the tip magnetization state is alsoknown better, since the tip is generally close to saturationeither up or down except for a narrow window around theswitching points. The tip in MFM exerts a relatively strong

Figure 8. Scanning electron microscope (SEM) image of a MFMcantilever. Inset upper left: high-resolution SEM image of the apexof the pyramid, where a coated carbon nanotube tip (CCNT) isvisible. The arrow shows the direction of the metal evaporation.Figure reprinted with permission from [105]. Copyright 2004 by theAmerican Institute of Physics.

magnetic field, comparable to the saturation magnetization ofthe magnetic material used in the tip, on the sample. Thiscan cause switching of the local magnetic moment of thesample [109], but can also be taken advantage of for vortexmanipulation studies (see section 3.5).

Assume for the moment that it will be possible to locatesmall nanoparticles at the end of an MFM cantilever tip. A 7 nmdiameter cobalt nanoparticle will have a total dipole momentof mt = 2.5 × 10−19A − m2, assuming a magnetization ofM = 1.4 × 106 A m−1 [110]. Taking values of Q =50 000,ω0 = 2π × 50 kHz, A = 1 nm, k = 2 N m−1, and T =4 K, the thermally limited force gradient is ∂F/∂z|min =1.7 × 10−7 N m−1. Using the dipole approximation for thetip, the peak force due to a sample dipole ms with both tipand sample moments oriented parallel to the z-axis is givenby (equation (17), figure 6) ∂F/∂zpeak = 24µ0msmt/4πz5.Using z = 15 nm results in a minimum detectible samplemoment of 23 Bohr magnetons. From figure 6 the full widthat half-maximum of the force derivative response would be�x = 0.6z = 9 nm.

9

Page 11: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Table 1. Sensitivities and spatial resolutions.

Dipole Monopole Line of current

N (µB Hz−1/2) �x (nm) � (�0 Hz−1/2) �x (nm) I (A Hz−1/2) �x (nm)

MFM 23 9 3.4 × 10−5 10 2.6 × 10−5 11SHM 4.5 × 104 100 3 × 10−3 108 1.14 × 10−4 118SSM 78 600 8.8 × 10−7 650 5.5 × 10−9 710

2.4. Summary

Each of the three techniques described here has advantagesand disadvantages: MFM has the best spatial resolution(∼10–100 nm), SSM the worst (∼0.3–10 µm), with Hall barsintermediate (∼0.1–5 µm). The relative sensitivities of thethree techniques depend on the type of field source. As in allmicroscopies there are tradeoffs between sensitivity and spatialresolution. The spatial resolution for SQUIDs and Hall bars isroughly given by �x ≈ d for heights z � d , where d is thesensor size (figures 2 and 3). For MFM the spatial resolutionis roughly set by the tip–sample spacing z, in the limit wherethe radius of curvature of the tip end is smaller than z (figures 7and 6). The flux sensitivity of SQUIDs is nearly independentof SQUID or pickup loop size, so the field sensitivity is roughlyproportional to the inverse of the pickup loop or SQUID area.With Hall bars limited by thermal noise the field sensitivityis independent of sensor size, but Hall bars dominated by 1/f

noise are expected to have a field sensitivity that scales like 1/d.Current MFM’s measured field gradients, rather than fields,and the tradeoff between sensitivity and spatial resolution areless clear than for SSM and SHM, although larger amountsof magnetic material will produce larger sensitivity but willrequire a larger tip volume, reducing resolution. Table 1compares sensitivity and spatial resolution for a state of theart SQUID, with a pickup loop 0.6 µm in diameter and aflux noise of 0.7 × 10−6�0 Hz−1/2 (figure 4 [71]), a stateof the art Hall bar with sensor size 100 nm and a field noiseof 500 µT Hz−1/2 at 3 Hz (figure 5 [87], and a hypotheticalMFM with a 7 nm Co nanoparticle at the tip as discussed insection 2.3. These numbers should be treated with a great dealof caution. For example, I have assumed that z/d = 0.1 for theSQUID and SHM, and z = 15 nm for the MFM. These maybe unreasonably optimistic estimates. However, table 1 doesindicate how relative sensitivities depend on the field source.As an example, SSM and MFM have roughly comparablespin sensitivities, but SSM has nearly two orders of magnitudebetter sensitivity for a monopole source, and nearly 4 ordersof magnitude better sensitivity for a line of current source.

Table 2 provides a summary of the scaling exponents n

for the minimum detectable field sources for SSM, SHM, andMFM. In this table the minimum detectable dipole moment,flux and current are proportional todn for SSM and SHM, whilefor MFM these quantities are proportional to zn. For SSM andSHM it is assumed that z/d is held constant. For SSM it isassumed that thermal noise dominates. There are two columnsfor SHM, for noise dominated by thermal fluctuations or by1/f noise. For MFM there are two columns, for the monopoleand the dipole tip approximations.

Table 2. Scaling exponents n.

Hall Hall MFM MFMSSM 1/f thermal monopole dipole

Line of current −1 0 1 2 3Monopole 0 1 2 3 4Dipole 1 2 3 4 5

There are further considerations in deciding whichmagnetic microscopy is best for a given application: SSMrequires a cooled sensor, about 9 K for the most technicallyadvanced Nb SQUIDs, although Hall bars and MFM also havethe highest sensitivities at low temperatures. SSM is the moststraightforward to calibrate absolutely. Hall bars tend to bemechanically delicate and the most sensitive to destructionby electrostatic discharge. MFM applies the largest fieldsto the sample. In general Hall bars and MFM can toleratelarger applied magnetic fields than SQUIDs. As we shallsee in section 3, each has been very useful in the study ofsuperconductors.

3. Applications

3.1. Static imaging of magnetic flux

Understanding the trapping of vortices in superconductors is ofcrucial technological importance, both because vortex pinningis the primary mechanism for enhanced critical currents intype II superconductors [111], and also because trappedvortices are an important source of noise in superconductingelectronic devices. Superconducting vortices are especiallyeasy to image using magnetic imaging because they arehighly localized, quantized and have relatively large magneticfields. A number of different techniques have been used forimaging superconducting vortices, including SSM, SHM andMFM. MFM has the advantage for this application that it canimage both the vortex magnetic fields and sample topographysimultaneously, so that macroscopic pinning sites can beidentified. It has been used for a number of years to imagesuperconducting vortices [112–119]. Care must be taken,however, because the large magnetic fields exerted by the tipcan distort and dislodge vortices. As we will see in section 3.5,this can also be an advantage. An particularly striking exampleof simultaneous magnetic and topographic imaging of vorticesin Nb and YBCO is shown in figure 9 [113].

3.1.1. Narrow strips. The sensitivity of high-Tc

superconducting sensors such as SQUIDs and hybridmagnetometers based on high-Tc flux concentrators is limited

10

Page 12: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 9. (a) Magnetic force microscope image of 0.3 µm × 0.3 µm areas for a Nb film and the corresponding topographic image (b).(d) MFM image of a single vortex (2 × 2 µm−2) in a YBCO film and (e) the corresponding topographic image. The white circles in (b) and(e) correspond to the distance at which the stray field emanating from the vortex has decreased to 1/e of its maximum value. The smalldotted circle in (b) defines an area with diameter 2ξ ≈ 22 nm. The lower panels (c) and ( f ) are cross-sections of the MFM data and thetopography profiles along the white lines indicated in (a), (b), (d) and (e). The dotted curves correspond to a Gaussian profile fitting of theMFM signal after filtering with a low-pass filter. Figure reprinted with permission from [113]. Copyright 2002 by Elsevier.

by 1/f noise. One source of this noise is the movementof vortices trapped in the sensor. One technique forreducing this source of noise is to trap vortices at a distancefrom the sensitive regions of the superconducting circuitryusing holes or motes [121, 122]. Jeffery et al [123] useda scanning SQUID microscope to image flux trapping insuperconducting electronic circuitry with motes. Veauvy et al[124] studied vortex trapping in regular arrays of nanoholesin superconducting aluminum. Noise from trapped vorticesin high-Tc SQUID washers and flux concentrators can beeliminated by dividing the high-Tc body into thin strips[125, 126]. Below a critical magnetic induction no vortex

trapping occurs in these strips for a given strip width. Anumber of models for the critical inductions of thin-film stripshave been proposed [120, 127–129]. Indirect experimentaltesting of these models was done by observing noise inhigh-Tc SQUIDs as a function of strip width and induction[125, 126, 130]. More direct experimental verification of these

models was presented by Stan et al [131] using scanningHall probe microscopy on Nb strips and Suzuki et al [132]using SQUID microscopy on NdBa2Cu3Oy thin-film patternswith slots. Both experiment and theory found that thecritical magnetic induction varied roughly as 1/W 2, whereW is the strip width. However, the experimental [131]and theoretical [120, 127] pre-factors multiplying this 1/W 2

dependence differ significantly.Figures 10(a)–(d) show SQUID microscope images

of 35 µm wide YBCO strips cooled in various magneticinductions. Figure 10(e) plots Bc+, the lowest induction atwhich vortices are trapped in the strips (squares), and Bc−, thehighest induction at which vortices are not trapped (dots) as afunction of strip width. The lines represent various theoreticalpredictions.

The Gibbs free energy for a vortex in a superconductingstrip of width W in an applied induction Ba perpendicular to

11

Page 13: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

(a)

(c) (d)

(b) (e)

Figure 10. SQUID microscope images of 35 µm wide YBCO thin-film strips cooled in magnetic inductions of (a) 5, (b) 10, (c) 20 and(d) 50 µT. (e) Critical inductions for vortex trapping in YBCO thin films as a function of strip width. The squares (Bc+) represent the lowestinduction in which vortices were observed, and the dots (Bc−) are the highest inductions for which vortex trapping was not observed. Thedashed–dotted line is the metastable critical induction B0 (equation (23)) [120], the short-dashed and long-dashed lines are BL

(equation (22)), the critical induction calculated using an absolute stability criterion, and the solid line is BK (equation (24)), calculatedusing a dynamic equilibrium criterion between thermally activation and escape for vortices. Figure reprinted with permission from [133].Copyright 2008 by the American Physical Society.

Figure 11. Top row: scanning Hall probe microscopy images of vortex configurations in square areas 19.6 µm on a side in asuperconducting Nb wire grid of 0.95 µm × 0.95 µm square holes, at the filling fractions f indicated. The second row shows maps of thepositions in the grid occupied by vortices. Note the ordering observed near f = 1/3 (c), f = 1/2 (e) and f = 4/3 (h). Figure reprintedwith permission from [143]. Copyright 1993 by the American Physical Society.

the strip plane can be written as [133]

G(x) = �20

2πµ0�ln

[αW

ξsin

(πx

W

)]

∓ �0(Ba − n�0)

µ0�x(W − x), (21)

where � = 2λ2/d is the Pearl length of the strip, λ is theLondon penetration depth, α is a constant of order 1, n isthe areal density of vortices, ξ is the coherence length and x

is the lateral position of the vortex in the strip. The criticalinduction model of Likharev [127] states that in order to trapa vortex in a strip the vortex should be absolutely stable: thecritical induction then happens when the Gibbs free energy in

the middle of the strip is equal to zero, and leads to

BL = 2�0

πW 2ln

(αW

ξ

). (22)

This is plotted as the dashed and dotted lines in figure 10for two assumed values of α. A second model, proposedby Clem [120], considers a metastable condition, in whichthe induction is just large enough to cause a minimum in theGibbs energy at the center of the strip, d2G(W/2)/dx2 = 0,leading to

B0 = π�0

4W 2. (23)

This is plotted as the dotted–dashed line in figure 10. Finally,Kuit et al proposed that the critical induction should result from

12

Page 14: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 12. (a) Schematic diagram for the tricrystal (1 0 0) SrTiO3 substrate used in the phase-sensitive pairing symmetry tests of Tsueiet al [61]. Four epitaxial YBa2Cu3O7−δ thin-film rings are interrupted by 0, 2 or 3 grain boundary Josephson weak links. (b) ScanningSQUID microscopy image of the four superconducting rings in (a), cooled in a field <5 mG. The central ring has �0/2 magnetic flux, where�0 is the superconducting flux quantum, spontaneously generated in it. The other 3 rings, which have no spontaneously generated flux, arevisible through a small change in the self-inductance of the SQUID when passing over the superconducting walls of the rings.

a dynamic equilibrium between vortex thermal generation andescape. This leads to the critical induction

BK = 1.65�0

W 2. (24)

It can be seen in figure 10 that the dynamic equilibriummodel (solid line) fits the data the best of the three models,with no fitting parameters. Kuit et al also found that n,the areal density of trapped vortices, increased linearly withapplied induction above the critical value, in agreement withthe dynamic equilibrium model, and that the trapping positionsshowed lateral ordering, with a critical field for formation of asecond row that was consistent with a numerical predictionof Bronson et al [134]. Similar good agreement betweenexperiment and the dynamic equilibrium model was found bythe same authors for Nb thin film strips.

3.1.2. Superconducting wire lattices, clusters, and nano-hole arrays. Two-dimensional arrays of Josephson junctions[135–140] have been studied extensively, not only because theyoffer the possibility of studying Kosterlitz–Thouless effects[141] in an ordered two-dimensional system, but also becauseof interesting effects that were predicted [142] and observed[135, 136] to occur when the applied field is a rational fractionof the field required to populate each cell with a quantum ofmagnetic flux. Scanning magnetic microscopy has been usedextensively for studying vortex trapping in two-dimensionalsuperconducting wire networks [58, 143], clusters [58] andring arrays [144, 145]. Superconducting wire networks are ofinterest because they are a physical realization of a frustratedxy model [142, 146, 147], where the variable is the phaseof the superconducting order parameter. The filling fractionf of vortices/site can be varied by cooling the network invarious fields, and detailed predictions have been made forthe ground state vortex configurations at rational f fractions[148]. Studies of these systems were at first performed using

transport [149–152], but direct imaging of the vortex positionsprovides more detailed information. Runge et al did magneticdecoration experiments on superconducting wire lattices [153].Although these experiments provided direct information on thevortex positions, only one set of conditions, such as magneticfield and temperature, can be explored for each sample usingthis technique. An example of scanning Hall bar microscopyon a superconducting wire network is shown in figure 11 [143].In this figure ordered vortex arrangements can be seen nearthe fractional filling factors of f = 1/3, 1/2 and 4/3, aspredicted by Teitel and Jayaprakash [146, 142], along withgrain boundaries between ordered domains and vacancies.

Field et al [154] studied vortex trapping in Nb films withlarge arrays of 0.3 µm holes on a square lattice. They foundstrong matching effects when the cooling fields correspondedto one or two vortices per site, as well as at several fractionalmultiples of the matching field. They also observed strikingdomain structure and grain boundaries between the domains.

3.1.3. Half-integer flux quantum effect. The discovery in1986 of superconductivity at high temperatures in the cuprateperovskites [2, 3] generated enormous excitement. Althoughthere was a great deal of evidence that the superconductinggap in the cuprates was highly anisotropic, with a significantdensity of states at low energies [155, 156], phase-sensitivemeasurements [157, 158] were required to distinguish, forexample, d-wave from highly anisotropic s-wave pairingsymmetry. The first such experiments [159, 160] used themagnetic field dependence of the critical current in singlejunctions and 2-junction SQUIDs between single crystals ofYBa2Cu3O7−δ (YBCO) and Pb to demonstrate a π -phase shiftbetween the component of the superconducting pairing orderparameter perpendicular to adjacent in-plane crystal faces (e.g.ignoring the effects of twinning, between the a- versus b-axisnormal faces) of the YBCO. Similar phase-sensitive Josephsoninterference experiments were performed by Brawner and Ott

13

Page 15: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

[161], Iguchi and Wen [162], Miller et al [163] and Kouznetsovet al [164].

A phase-sensitive technique for determining pairingsymmetry that is complementary to Josephson interferenceis to image magnetic fields generated by spontaneoussupercurrents in superconducting rings that have an oddnumber of intrinsic π -phase shifts in circling them or inJosephson junctions that have intrinsic π -shifts along them[165, 166]. The first such experiments were done by Tseuiet al [61], using YBCO films grown epitaxially on tricrystalSrTiO3 substrates. A scanning SQUID microscope wasused to image the magnetic fields generated by supercurrentscirculating rings photolithographically patterned in the YBCOfilms. In the original experiments the central ring (figure 12),which has 3 grain boundary Josephson junctions, has either1 or 3 intrinsic π -phase shifts for a superconductor withpredominantly dx2−y2 pairing symmetry, and should thereforehave states of local energy minima, when no external magneticfield is applied, with the total flux through the ring of� = (n + 1/2)�0, n an integer, for sufficiently large LIc

products, where L is the inductance of the ring and Ic isthe smallest critical current of the junctions in the ring.This is the half-integer flux quantum effect. The two outerrings that cross the tricrystal grain boundaries have 0 or2 intrinsic π -phase shifts for a dx2−y2 superconductor, andshould therefore have local energy minima at � = n�0—conventional flux quantization. Similarly, the ring that doesnot cross any grain boundaries has no intrinsic phase drop,and also has local minimum energies for � = n�0. Theseexpectations were confirmed using a SQUID microscope usinga number of techniques [61], consistent with YBCO havingdx2−y2 pairing symmetry. Subsequent tricrystal experimentseliminated the possibility of a non-symmetry related origin forthe half-integer flux quantization in the 3-junction ring [167],ruled out extended s-wave symmetry [168], demonstratedthe existence of half-flux quantum Josephson vortices [169],showed results consistent with dx2−y2 pairing symmetry for thehole-doped high-Tc cuprate superconductors Tl2Ba2CuO6+δ

[170], Bi2Sr2CaCu2O8+δ [171] and La2−xSrxCuO4−y [172],for a range of doping concentrations [172], and theelectron-doped cuprates Nd1.85Ce0.15CuO4−y (NCCO) andPr1.85Ce0.15CuO4−y (PCCO) [173], and for a large range intemperatures [174]. SQUID magnetometry experiments ontwo-junction YBCO–Pb thin-film SQUIDs were performed byMathai et al [175], and on tricrystals by Sugimoto et al [176].Josephson interference, SQUID magnetometry and microwavemeasurements were made on biepitaxial all high-Tc YBCOjunctions by Cedergren et al [177].

Grain boundary junctions in the high-Tc cupratesuperconductors are Josephson weak links. As such they areused for fundamental and device applications. In addition,grain boundaries play an important role in limiting the criticalcurrent density in superconducting cables made from thecuprates [178]. Anomalies in the dependence on magnetic fieldof the critical current of cuprate grain boundary junctions withmisorientation angles near 45◦ can be understood in terms oftheir dx2−y2 pairing symmetry combined with faceting, whichnaturally occurs on a 10–100 nm scale in these grain boundaries

Figure 13. Scanning SQUID microscope image of an area of1024 × 256 µm2 along an asymmetric 45◦ [0 0 1] tilt YBa2Cu3O7−δ

bicrystal grain boundary (arrows). There are some ten bulk vorticesin the grains (film thickness ∼180 nm), but there is also flux of bothsigns spontaneously generated in the grain boundary. The bottomsection shows a cross-section through the data along the grainboundary measured in units of �0 penetrating the SQUID pickuploop. Figure reprinted with permission from [180] Copyright 1996by the American Physical Society.

Figure 14. Generation of half-flux quanta in connected andunconnected YBa2Cu3O7−δ–Au–Nb zigzag structures. Shown arescanning SQUID micrographs of (a) 16 antiferromagneticallyordered half-integer flux quanta at the corners of a connected zigzagstructure and (b) 16 ferromagnetically ordered half-flux quanta at thecorners of an unconnected zigzag structure (T = 4.2 K). The cornerto corner distance is 40 µm in both (a) and (b). Figure reprinted withpermission from [184]. Copyright 2003 Macmillan Publishers Ltd.

[179]. This combination produces a series of intrinsic π -phaseshifts along such grain boundaries. These π -shifts havespontaneous supercurrents and magnetic fields associated withthem [180]. An example of SQUID microscope imaging ofthese spontaneous fields is shown in figure 13. If the distancebetween intrinsic π -shifts is shorter than the Josephsonpenetration depth, the total magnetic flux associated with eachwill be less than �0/2 [181]; a high density of facets canresult in ‘splintered’ Josephson vortices with fractions of �0

of flux [182].Facetted (zigzag) junctions can also be fabricated by

design, using for example a ramp-edge YBCO/Nb junctiontechnology [183]. An example is shown in figure 14, whichshows scanning SQUID microscope images of two facettedjunctions [184]. In figure 14(a) the corner junctions areconnected by superconducting material. In this case there isa strong interaction between the half-flux quantum Josephsonvortices upon cooling, so that they have a strong tendency toalign with alternating signs of the circulating supercurrentsand spontaneous magnetizations. In figure 14(b) the corner

14

Page 16: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 15. (a) Schematic of YBCO–Nb two-junction rings fabricated to phase sensitively determine the momentum dependence of thein-plane gap in YBCO. (b) SQUID microscope image, taken with a square pickup loop 8 µm on a side, of a square area 150 µm on a sidecentered on one of the rings, cooled and imaged in zero field at 4.2 K. (c) The outer circle of images, taken after cooling in zero field, has afull-scale variation of 0.04 �0 flux through the SQUID: the inner ring, taken after the sample was cooled in a field of 0.2 µT, has a full-scalevariation of 0.09 �0. The rings cooled in zero field had either 0 or �0/2 of flux in them; the rings cooled in 0.2 µT had �0/2, �0 or 2�0 offlux, as labelled. Figure reprinted with permission from [187]. Copyright 2006 Macmillan Publishing Ltd.

junctions have been isolated from each other by etchingchannels between them. In this case the half-flux quantumJosephson vortices all have the same alignment, even whencooled in a relatively small field.

Lombardi et al [185] and Smilde et al [186] havemapped out the in-plane momentum dependence of thesuperconducting gap amplitude of YBCO by measuring thecritical current as a function of junction normal angle fora series of biepitaxial YBCO grain boundary junctions,and YBCO/Nb ramp junctions, respectively. The latterexperiments, performed in part on untwinned epitaxial YBCOfilms, were able to determine the anisotropy of the gapbetween the a and b in-plane crystalline directions. However,these experiments were only sensitive to the amplitude ofthe superconducting order parameter, not its phase. Phase-sensitive measurements of the momentum dependence of thein-plane gap in YBCO were performed by Kirtley et al [187]by imaging the spontaneously generated magnetic flux in aseries of YBCO–Nb two-junction rings, with one junctionangle normal relative to the YBCO crystalline axes heldfixed from ring to ring, while the other junction normalangle varied. Some results from this study are shown infigure 15. The rings either had n�0 (integer flux quantization)or (n + 1/2)�0 (half-integer flux quantization) of flux in them,n an integer. The transition between integer and half-integerflux quantization happened at junction angles slightly differentfrom 45◦ + n × 90◦, as would be expected for a pure dx2−y2

superconductor, because of the difference in gap amplitudesbetween the crystalline a and b phases. In addition, carefulintegration of the total flux in the rings showed that anyimaginary component to the order parameter, if present, mustbe small.

3.1.4. Ring arrays. Arrays of superconducting rings areof interest as a physical analog of a spin: for conventionalsuperconducting rings the states with shielding supercurrentflowing clockwise or counter-clockwise are degenerate ata magnetic flux bias of �0/2 (half of a flux quantumthreading each ring). Two neighboring rings can interactmagnetically, making them a physical analog of the Ising spin.Frustration can be introduced into the system by choosingthe arrangement of the rings: square and hexagonal latticesare geometrically unfrustrated, while triangular and Kagomelattices are frustrated.

Davidovic et al used scanning Hall bar microscopy tostudy two-dimensional arrays of closely spaced Nb rings(see figure 16) [144, 145]. In these experiments Davidovicet al found that it was possible to find small islands ofantiferromagnetically ordered ‘spins’, but no perfect ordering,even in arrays with no geometrical frustration. One sourceof disorder in arrays of conventional superconducting rings isvariation in area from ring to ring from lithographic variations.Conventional rings must be flux biased to near �0/2 flux withhigh precision (see figure 16), and area variations lead to fluxvariations.

The same technology used for the pairing symmetry testsof figure 15 was used to form two-dimensional arrays ofπ -rings, superconducting rings with an intrinsic phase shiftof π [184]. An example of SQUID microscope imaging ofseveral π -ring arrays is shown in figure 17 [188]. π -rings havetwo degenerate states at zero flux bias, therefore eliminatingthe need for precise flux biasing of the rings, and therebyeliminating slight differences in the sizes of the rings as asource of disorder [188]. Although π -ring arrays showed morenegative values of the bond-order σ , corresponding to greater

15

Page 17: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 16. A sequence of field-cooled images of a honeycomb lattice of 1 µm diameter hexagonal Nb thin-film rings, taken in increasingapplied fluxes near an applied flux per ring of �0/2. The flux increases from left to right starting at the upper left corner with 0.4913�0 andending in the lower right corner at 0.5066�0. Figure reprinted with permission from [145]. Copyright 1997 by the American PhysicalSociety.

antiferromagnetic ordering, than arrays of conventional rings,perfect ordering was never observed in either conventional orπ -ring arrays, possibly because of the large number of nearlydegenerate states in this system [189].

3.1.5. Tests of the interlayer tunneling model. One candidatemechanism for superconductivity at high temperatures in thecuprate perovskite superconductors is the interlayer tunnelingmodel, in which the superconductivity results from anincreased coupling between the layers in the superconductingstate [190–194]. An essential test of this theory is thestrength of the interlayer Josephson tunneling in layeredsuperconductors. For the interlayer tunneling model tosucceed, the interlayer coupling in the superconducting

state must be sufficiently strong to account for the largecondensation energy of the cuprate superconductors. Thebest materials for testing this requirement are Tl2Ba2CuO6+δ

(Tl-2201) and HgBa2CuO4+δ (Hg-1201), which have highcritical temperatures (Tc ≈ 90 K) and a single copper-oxideplane per unit cell. Scanning SQUID microscope images weremade of interlayer Josephson vortices emerging from the a−c

face of Tl-2201 [195] (figure 18) and Hg-1201 [196]. Theextent of these vortices in the a-axis direction is set by theinterlayer penetration depth λc. Some spreading of the vortexfields occurs as the crystal surface is approached [197], but notenough to affect the qualitative conclusions from these studies:the experimentally determined interlayer penetration depth λc

is an order of magnitude larger than required by the interlayertunneling model. These conclusions are in agreement with

16

Page 18: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 17. SQUID microscopy images of four electricallydisconnected arrays of π -rings with 11.5 µm nearest neighbordistances. These images were taken at 4.2 K with a 4 µm diameterpickup loop after cooling in nominally zero field at 1–10 mK s−1.The bond-order σ is a measure of the antiferromagnetic order in thearrays. The geometrically unfrustrated square and honeycombarrays could have perfect antiferromagnetic correlation, whichwould correspond to σ = −1. The minimum possible bond-orderfor the frustrated Kagome and triangle lattices is σ = −1/3. Figurereprinted with permission from [188]. Copyright 2005 by theAmerican Physical Society.

measurements of the Josephson plasma frequency in thesematerials [198].

3.1.6. Pearl vortices. Vortices in thin superconductorssatisfying d � λL, where d is the thickness and λL isthe London penetration depth, first described by Pearl [199],have several interesting attributes. The field strengths Hz

perpendicular to the films diverge as 1/r at distances r � �,where � = 2λ2

L/d is the Pearl length, in Pearl vortices,whereas in Abrikosov (bulk) vortices the fields diverge asln(r/λL) [200]. Since in the Pearl vortex much of thevortex energy is associated with the fields outside of thesuperconductor, the interaction potential Vint(r) between Pearlvortices has a long range component Vint ∼ �/r for r �

[199], unlike Abrikosov vortices, which have only short rangeinteractions. Although it was originally believed that theinteraction between Pearl vortices Vint ∼ ln(�/r) for r � �

leads to a Berezinskii–Kosterlitz–Thouless (BKT) transitionwhich is cut off due to screening on a scale � [111], Kogan[201] reports that the BKT transition could not happen inthin superconducting films of any size on insulating substratesbecause of boundary conditions at the film edges that turnthe interaction into a near exponential decay. Grigorenkoet al, using SHM, observed domains of commensurate vortex

Figure 18. (a) Scanning SQUID microscope image of an interlayerJosephson vortex emerging from the ac face of a single crystal of thesingle-layer cuprate superconductor Tl2Ba2CuO6+δ , imaged at 4.2 Kwith a square pickup loop 8.2 µm on a side. (b) A two-dimensionalfit to the vortex in (a) with λc, the penetration depth perpendicular tothe CuO2 planes, equal to 18 µm and z0, the spacing between theSQUID pickup loop and the sample surface, equal to 3.6 µm.(c) The difference between the data and the fit. (d) Cross-sectionsperpendicular and parallel to the long axis of the vortex, offset forclarity. Figure reprinted with permission from [195]. Copyright1998 AAAS.

patterns near rational fractional matching fields of a periodicpinning array in a thin-film superconductor, which could onlyform when the vortex–vortex interactions are long range [202].

Figure 19 shows scanning SQUID microscope images ofPearl vortices in ultra-thin [Ba0.9Nd0.1CuO2+x]m/[CaCuO2]n(CBCO) high-temperature superconductor films. Pearlvortices can be magnetically imaged in such thin films,although the Pearl length can be as large as 1 mm, because ofthe strong 1/r divergence of the Pearl vortex magnetic fieldsat the vortex core: the thin-film limit for the two-dimensionalFourier transform of the z-component of the field from anisolated vortex trapped in a thin film is given by [199, 203]

Bz(k, z) = �0e−kz

1 + k�, (25)

where z is the height above the film, k =√

k2x + k2

y , � =2λ2

ab/d, λab is the in-plane penetration depth, d is the filmthickness and �0 = h/2e.

The open circles in figures 19(b) and (d) are cross-sectionsthrough the data of figures 19(a) and (c). The solid lines are

17

Page 19: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 19. SQUID microscope image and cross-sectional data along the positions indicated by the dashed line in (a), (c) of Pearl vorticestrapped in two artificially layered (Ba0.9Nd0.1CuO2+x)m/(CaCuO2)n superconducting films. The SQUID pickup loops were a square 7.5 µmon a side (a),(b), and an octagon 4 µm on a side (c),(d) (schematics superimposed on image). The open symbols in (b),(d) are thecross-sectional data; the solid lines in (b), (d) are fits. Figure reprinted with permission from [299]. Copyright 2004 by the AmericanPhysical Society.

fits to the data using equation (25), using the Pearl length �

as a fitting parameter. The in-plane London penetration depthλab of these thin films can be inferred from � if d is known.

3.1.7. Mesoscopic superconductors. Vortices in super-conductors with dimensions comparable to the super-conducting coherence length ξ have been predicted to nucleatespontaneously in a number of interesting configurations, suchas shells [204], giant vortices [205] and antivortices, so asto preserve the symmetry of the sample [206, 207]. Vortextrapping patterns in mesoscopic superconducting samples havebeen studied using Bitter decoration [208], scanning SQUIDmicroscopy [209], ballistic Hall magnetometry [91], transport[210] and Hall bar microscopy. An example of Hall barmicroscopy results is shown in figure 20. Nishio et al [211]report results in agreement with Ginzburg–Landau calculationsfor temperatures considerably lower than the superconductingtransition temperature.

3.1.8. Sr2RuO4. Sr2RuO4, with the same crystal structureas La2CuO4, the parent compound for the first class

of cuprate perovskite compounds shown to exhibit high-temperature superconductivity [2], was discovered in 1994to be superconducting at about 1 K [6]. It is thought tobe a px ± ipy-wave, triplet pairing superconductor withbroken time-reversal symmetry. One of the consequencesof this pairing symmetry state is spontaneously generatedsupercurrents, and consequent large magnetic fields at surfacesand boundaries between px + ipy and px − ipy domains [212].Although there is evidence for characteristic magnetic fieldsof 0.5 G in Sr2RuO4 from muon spin resonance experiments[213], and evidence for broken time-reversal symmetry atthe surface of Sr2RuO4 has been obtained using Sagnacinterferometry [18], no evidence for spontaneously generatedmagnetic fields at the surfaces of Sr2RuO4 samples has beenfound in scanning Hall bar and scanning SQUID experiments,with much higher sensitivity than the magnetic fields originallypredicted [214–216]. A possible explanation for this failure toobserve spontaneous magnetization directly is that the px ±ipy

domains are small, so that the fields from closely spaceddomain boundaries nearly cancel each other. The experimentalinformation on px ± ipy domain sizes is not consistent, with

18

Page 20: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 20. Scanning Hall probe microscope images of a4 µm × 4 µm square Pb film at applied fields of (a) 4.6 G, (c) 6.6 G,(e) 8.6 G and (g) 10.6 G. ((b), (d), ( f ) and (h)). The results of 2Dmonopole fits to (a), (c), (e) and (g), respectively. The broken linesdenote the shape of the square. The locations of the vortices areshown as white points. Figure reprinted with permission from [211].Copyright 2008 by the American Physical Society.

the first phase-sensitive pairing symmetry experiments onSr2RuO4 being consistent with domain sizes of order 1 mm[217], while later phase-sensitive results are consistent withdomain sizes 1 µm or less [218], and the Sagnac interferometrybeing consistent with domain sizes intermediate betweenthese two sizes [18]. Recently Raghu et al have suggestedthat the superconductivity in Sr2RuO4 arises primarily inquasi-one-dimensional bands, and is therefore not expectedto generate experimentally detectable edge currents [219].

Dolocan et al [220, 221] have magnetically imaged singlecrystals of Sr2RuO4 using a scanning µSQUID mountedon a tuning fork in a dilution refrigerator. This allowedcombined magnetic and topographic information on a lengthscale of 1 µm, at temperatures below 400 mK. They foundcoalescence of vortices and the formation of flux domains(see figure 21). The formation of lines of vortices in a fieldtilted relative to the crystalline c-axis [222] can be attributedin this case to the large anisotropy in Sr2RuO4, resulting inan attractive interaction between vortices in a plane definedby the c-axis and the magnetic field direction [221]. Thevortex coalescence observed when the applied magnetic fieldis parallel to the c-axis may be due to weak intrinsic pinning atdomain walls, and the existence of a mechanism for bringing

vortices together that overcomes the conventional repulsivevortex–vortex interaction.

3.1.9. Magnetic superconductors. Magnetic order andsuperconductivity are competing orders since the Meissnerstate excludes a magnetic field from the bulk andsuperconductivity is destroyed at sufficiently high fields.However, superconductivity and magnetism can coexist[223–226] if the orientation of the local magnetic momentsvaries on a length scale shorter than the superconductingpenetration depth λ, or if the field generated by themagnetization is carried by a so-called spontaneous vortexlattice [227]. The former order has been observed inErRh4B4 and HoMo6S8 [228, 229], and indirect evidence forthe latter has been reported in UCoGe [230]. Several Hallbar studies have been performed on artificial ferromagnetic-superconducting hybrid structures [231–234]. figure 22[235] shows scanning Hall bar microscopy measurementsof ErNi2B2C, which has a superconducting Tc of ≈11 K,becomes antiferromagnetic below TN ≈ 6 K, and exhibitsweak ferromagnetism below TWFM ≈ 2.3 K [236]. Bluhm et al[235] found that in this material the superconducting vorticesspontaneously rearranged to pin on twin boundaries uponcooling through the antiferromagnetic transition temperature,and that a weak random magnetic signal appears in theferromagnetic phase.

3.2. Fluxoid dynamics

3.2.1. Limits on spin-charge separation. One explanation forthe peculiar normal state properties and high superconductingtransition temperatures of the high-Tc cuprate perovskites isthat there exists a new state of matter in which the elementaryexcitations are not electron-like, as in conventional metals, butrather the carriers ‘fractionalize’ into ‘spinons’, chargeless spin1/2 fermions, and ‘chargons’, bosons with charge +e [237].Senthil and Fisher proposed a model for such a separationin two dimensions with sharp experimental tests [238]: inconventional superconductors a Cooper pair with charge 2e

circling around a vortex with h/2e of total magnetic fluxexperiences a phase shift of 2π . However, a chargon wouldpick up only a phase shift of π . In order for the chargon wavefunction to remain single-valued, it must be accompanied bya ‘vison’ that provides an additional π phase shift. In theSenthil-Fisher test, a superconducting ring with flux trappedin it is warmed through the superconducting transition. If thenumber of vortices in the ring is odd, then it will also containa vison that could persist above Tc. If the ring is then recooledin zero field before the vison escapes, it will spontaneouslygenerate a single flux quantum with random sign. On theother hand, if the number of vortices in the ring is even, nosuch ‘vortex memory’ effect would exist. Bonn et al [239](figure 23) tested for this effect in highly underdoped singlecrystals of YBa2Cu3O6+x . These crystals have a very sharpbut low transition temperature favorable for such a test. Theyfound no evidence for a vortex memory effect, and were ableto put stringent upper limits on the energy associated with avison. Some models in which superconductivity results from

19

Page 21: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 21. Scanning SQUID microscope images of flux domains in Sr2RuO4 at T = 0.36 K after field cooling at various fields. In all cases,the magnetic field amplitude applied along the c-axis (H⊥) was kept constant at 2 G while the in-plane field (Hab) was (a) 0 G, (b) 5 G,(c) 10 G, (d) 50 G. The imaging area is 31 µm × 17 µm. The field scale in G is shown on the right; dark regions are superconductingvortex-free regions. Figure reprinted with permission from [220]. Copyright 2005 by the American Physical Society.

Figure 22. Scanning Hall bar image of a single-crystal sample of ErNi2B2C. After cooling below Tc in a weak field (2.4 Oe in (a), (b), (d),(e), 18 Oe in (c)), a vortex distribution consistent with random pinning is observed on the ab face ((a) 7.3 K). Upon reducing the temperaturebelow TN, the vortices spontaneously organize along twin domain walls along the [1 1 0] ((b) 5.3 K) or [1 1 0] ((c) 4.2 K) direction. Thepattern gradually disappears as the temperature is raised again ((d) 5.7 K, (e) 6.0 K, different cycle than (a) and (b)). Figure reprinted withpermission from [235]. Copyright 2006 by the American Physical Society.

spin-charge separation predicts h/e fluxoids in materials withlow superfluid density [240–242]. Wynn et al [243], usingscanning SQUID and Hall bar microscopies, found no evidenceforh/e vortices in strongly underdoped YBa2Cu3O6+x crystals,also placing limits on spin-charge separation in these materials.

3.2.2. Superconducting rings. As we have seen in previoussections, the dynamics of fluxoids in a ring configuration areof interest as a model for the Ising spin system [144, 145, 188]section 3.1.4, and for placing limits on spin-charge separationin underdoped cuprate superconductors [239] section 3.2.1.Kirtley et al studied fluxoid dynamics in photolithographicallypatterned thin-film rings of the underdoped high-temperature

superconductor Bi2Sr2CaCu2O8+δ using a scanning SQUIDmicroscope, and concluded that their results could beunderstood in terms of thermally activated nucleation of a Pearlvortex in, and transport of the Pearl vortex across, the ringwall [244].

Fluxoid dynamics in superconducting rings may alsoprovide clues to physics on an entirely different scale: the earlydevelopment of the universe. Immediately after the Big Banga fundamental symmetry related the particles that mediatedthe electromagnetic, strong and weak forces. However,as the universe cooled down this symmetry was broken insuch a way that, for example, the W and Z bosons, whichmediate the weak interaction, have mass, while photons, which

20

Page 22: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 23. Cross-sections and images of single (bottom rightimage, lower cross-section) and double (middle right image, highercross section) flux quanta trapped in a 50µm × 50 µm square ofsingle-crystal YBa2Cu3O6.35 (Tc = 6.0 K) with a 10 µm hole drilledusing a focussed ion beam (top right image). When the sample washeated to 5.6 K for 1 s and recooled to 2 K, these flux quanta escapedwith no sign of the vortex memory associated with visons. Figurereprinted with permission from [239]. Copyright 2001 MacmillanPublishing Ltd.

mediate the electromagnetic force, do not. Kibble proposed anintuitive picture of this symmetry breaking of the early universe[245, 246], with the underlying idea that causality governedthe number of defects because a finite time delay is needed forinformation to be transferred between different regions of asystem. Zurek pointed out that Kibble’s ideas could be testedby studying vortices in superfluids and superconductors [247].As a superconductor or superfluid cools through its transitiontemperature different regions can nucleate into a state withthe same order parameter amplitude ψ but different phasesφ. This will produce ‘kinks’ or ‘defects’ where φ wouldeventually (at low temperature) change by 2π . These kinks canbe removed by annihilation with anti-kinks if the coherencelength ξ is long enough and the cooling rate is sufficientlyslow. However, as the system cools this annihilation processproceeds less rapidly, until at some point the kinks ‘freeze-in’ because part of the system cannot communicate with otherparts quickly enough. In a superconducting ring a phase shiftof 2π around the ring corresponds to a fluxoid trapped in thering, so that the process of forming and freezing in kinks isalso called spontaneous fluxoid generation. Kibble and Zurekshowed that causality implies that the probability P of finding aspontaneous fluxoid in the ring (when cooled in zero externalfield) depends on the cooling rate τ−1

Q as P ∝ (τQ/τ0)−σ ,

where τ0 is a characteristic time and σ depends on howthe correlation length ξ and the relaxation time τ vary withtemperature close to the transition. For a superconducting ringand assuming a mean-field approximation σ = 1/4.

Tests of the Kibble–Zurek prediction have been madeby looking for vortices in superfluid 4He [248, 249] and3He [250, 251], in superconducting films [15, 252], andsuperconducting rings interrupted by Josephson junctions[253–255]. However, verification of the ideas of Kibble andZurek using arguably the simplest system, superconductingrings, was not made until recently. Kirtley et al [256]

Figure 24. SQUID microscope images of a 12 × 12 array of 20 µminside diameter, 30 µm outside diameter thin-film rings of Mo3Si,cooled in zero field through the superconducting Tc at various rates.The resulting ring vortex numbers were consistent with thermallyactivated spontaneous fluxoid formation. Figure reprinted withpermission from [256]. Copyright 2003 by the American PhysicalSociety.

(figure 24) used a SQUID microscope to image an array ofrings after repeated cooling in various magnetic fields andcooling rates to determine the probability of spontaneousfluxoid generation. They showed that a second mechanismprevailed, in which the final density of fluxoids depended on abalance between thermal generation and the relaxation rates offluxoids [257–259]. They argued that the thermal fluctuationmechanism is complementary to the ‘causal’ mechanismand should be considered in attempts to understand phasetransitions, both in the laboratory and in the early universe.The Kirtley et al experiments [256] were done on rings withparameters which favored the thermal activation mechanismover the Kibble–Zurek (causal) mechanism. Recently Monacoet al [260] have studied superconducting rings with parametersmore favorable to the causal mechanism, using a geometry inwhich the flux state of a single ring can be manipulated andsensed rapidly, and find results consistent with the Kibble–Zurek prediction, taking into account the fact that the ring’scircumference was much smaller than the coherence length atthe temperature at which the fluctuations were frozen in.

3.2.3. High-Tc grain boundaries. Grain boundaries in thecuprate high-Tc superconductors are widely used for devicesand fundamental studies, and govern the transport propertiesof superconductors with technological applications [178]. It istherefore important to understand the mechanism of dissipationin transport across grain boundaries. Transport across grain

21

Page 23: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 25. Scanning Hall bar measurements of vortex noise in a 24◦

symmetric YBCO grain boundary (GB). (a) Sketch of the sample.The bridge width is 50 µm. (b), (c) Magnetic image of the GB area,(b) at 10 K and applied current I = 29 mA and (c) at 10 K afterthermal cycling to 70 K with I = 0, showing individual vortices (redonline) and antivortices (blue online). The regions imaged areoutlined by the square dashed line (orange online) in (a). (d), (e)The Hall probe signals versus time at 10 K for different currentsapplied to the GB measured at two positions marked in figure 1(b).The switching events are vortices passing under the probe. Thecircles (red online) indicate three-level switching events. Figurereprinted with permission from [261]. Copyright 2009 by theAmerican Institute of Physics.

boundaries is usually studied with a voltage threshold on theorder of nanovolts. At these voltages millions of vorticestraverse the boundary per second. Much smaller voltages canbe measured by magnetically imaging the vortices in the grainboundaries. Kalisky et al [261] performed such measurementson grain boundaries produced by epitaxial growth of thecuprate high-Tc superconductor YBa2Cu3O7−δ on bicrystals ofSrTiO3 using a large scanning area Hall bar microscope [84].They observed (figure 25) telegraph noise in the Hall barsignal when the sensor was directly above the grain boundary,which they attributed to the motion of vortices. They inferredthe voltage across the grain boundary from the frequencyof the telegraph noise, with voltage sensitivity as small as2 × 10−16 V. The dependence of the inferred grain boundaryvoltage on current followed either an exponential V ∝ emI/Ic

or a power law V ∝ In dependence with high values ofm or n. These results were qualitatively different fromgrain boundary transport measurements at higher voltages

Figure 26. SQUID signal (left axis) and ring current (right axis) asa function of applied flux �a for two aluminum rings, both withthickness d = 60 nm and annulus width w = 110 nm. Thefluctuation theory (dashed, red online) was fit to the data (noisysolid, blue online) through the dependence of Iring on �a andtemperature. (a) to (c) ring radius R = 0.35 µm, fittedTc(�a = 0) = 1.247 K, and γ = 0.075. The parameter γcharacterizes the size of the ring. The smooth solid line (greenonline) is the theoretical mean-field response for T = 1.22 K andshows the characteristic Little–Parks line shape, in which the ring isnot superconducting near �a = �0/2. The excess persistent currentin this region indicates the large fluctuations in the Little–Parksregime. (d ) R = 2 µm, fitted Tc(�a = 0) = 1.252 K and γ = 13.Figure reprinted with permission from [265]. Copyright 2007 byAAAS.

[178, 262, 263], and could not be fully explained using existingmodels.

3.3. Local susceptibility measurements

3.3.1. Superconducting fluctuations. Experimental knowl-edge of superconducting fluctuations in one dimensionhas largely been derived from transport measurements[264], which require electrical contacts and an externallyapplied current. The SQUID susceptometer, along witha ring geometry for the superconductor, allows the studyof fluctuation effects, without contacts, in isolated, quasione-dimensional rings in the temperature range wherethe circumference is comparable to the temperature-dependent Ginzburg–Landau coherence length ξ(T ). In suchmeasurements, an example of which is shown in figure 26[265], a magnetic field is applied to the ring using a single-turn field coil, co-planar and concentric with the pickup loop,both of which are integrated into the SQUID sensor [72]. Theresponse of the ring to this applied flux is measured by thepickup loop. The SQUID sensor is in a carefully balancedgradiometer configuration, such that it is insensitive to the

22

Page 24: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 27. Local susceptibility image in underdoped Ba(Fe1−xCox)2As2, indicating increased diamagnetic shielding on twin boundaries.(a) Local diamagnetic susceptibility, at T = 17 K, of the ab face of a sample with x = 0.051 and Tc = 18.25 K, showing stripes of enhanceddiamagnetic response (white). In addition there is a mottled background associated with local Tc variations that becomes more pronouncedas T → Tc. Overlay: sketch of the scanning SQUID sensor. The size of the pickup loop sets the spatial resolution of the susceptibilityimages. (b) and (c) Images of the same region at (b) T = 17.5 K and (c) at T = 18.5 K show that the stripes disappear above Tc. Atopographic feature (scratch) appears in (b) and (c). (d) Crystal structure of a unit cell of the parent compound BaFe2As2. (e) Top view ofthe FeAs layer with tetragonal symmetry, and ( f ) an exaggerated view of the orthorhombic distortion that occurs at low temperatures.(g) Schematic of a possible arrangement of spins across a twin boundary in the antiferromagnetic state. Figure reprinted with permissionfrom [285]. Copyright 2010 by the American Physical Society.

fields applied by the field coil [73]. Nevertheless, furtherbackground subtraction, by comparing the mutual inductancebetween the field coil and SQUID with the pickup loop closeto the ring with that with the pickup loop at a distance, isrequired to measure the response field due to the ring, whichcan be 7 orders of magnitude smaller than the applied field.A scanned sensor allows multiple rings to be measured in asingle cooldown. The experimental results agree with a fullnumeric solution of thermal fluctuations in a Ginzburg–Landauframework that includes non-Gaussian effects [266, 267] for allof the rings for which calculations were numerically tractable.This is in contrast to previous work on a single Al ring [268],which disagreed strongly with theory.

Similar techniques have allowed the measurement ofspontaneous persistent currents in normal metal rings [269].

3.3.2. Stripes. The pnictide class of superconductors [8]have the highest critical temperatures (57 K) observed for anon-cuprate superconductor, multiband Fermi surfaces, and atcertain doping levels a paramagnetic to antiferromagnetic aswell as a tetragonal to orthorhombic transition above the super-conducting transition temperature. In Ba(Fe1−xCox)2As2

(Co doped Ba-122) and other members of the 122 family

(AFe2As2 with A = Ca, Sr, and Ba), doping causes thespin-density-wave transition temperature and the structuraltransition temperature to decrease [270, 271] falling to zero ator near the doping where the highest Tc occurs, suggestingthe importance of lattice changes in determining transportproperties. Both experiment [272–274] and theory [275]have suggested a close relationship between structural andmagnetic properties, leading some authors to describe thelattice and spin-density-wave transition by a single orderparameter [276, 277]. In addition, structural strain appears toplay a significant role in the superconductivity. For example, inthe 122 compounds, small amounts of non-hydrostatic pressurecan induce superconductivity [278–282]. In addition, there isevidence that the structural perfection of the Fe-As tetrahedronis important for the high critical temperatures observed in theFe pnictides [283, 284].

Kalisky et al [285] have shown that Co doped Ba-122 hasstripes of enhanced susceptibility below the superconductingtransition temperature (see figure 27). These stripes areresolution limited using a SQUID pickup loop with an effectivediameter of 4 µm, and are believed to be associated withboundaries between twins- crystallites with their in-plane a, b

axes rotated by 90◦. The amplitude of the susceptibility

23

Page 25: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 28. (a) Geometry of a SQUID susceptometer. (b) Mutual inductance between the field coil and the pickup loop, as a function of thespacing z between the SQUID substrate and a thin CBCO film. The symbols are data, taken with various alternating currents through thefield coil. The solid line is modeling using equation (25), with � = 11.8 µm. (b) Comparison of the Pearl length � for a number of CBCOsamples using fitting of SQUID magnetometry images of vortices (vertical axis) versus susceptibility measurements (horizontal axis).Figure reprinted with permission from [299]. Copyright 2004 by the American Physical Society.

stripes becomes larger as the superconducting transitiontemperature is approached from below. Since the susceptibilitysignal for a homogeneous superconductor becomes largeras the penetration depth becomes shorter and the superfluiddensity becomes larger, it seems reasonable to associatethe susceptibility stripes with shorter penetration depths andenhanced superfluid densities on the twin boundaries. Twinboundaries have been associated with enhanced superfluiddensity in conventional superconductors [286], but theinfluence of twin planes on superconductivity in, for example,the cuprates is less clear [287–291].

It is difficult to calculate analytically the effect of a planeof reduced penetration depth (enhanced superfluid density) onthe susceptibility observed in scanning SQUID measurements.Kirtley et al [292] have done a finite-element calculation inthe appropriate geometry. Because of the uncertainty in thewidth of the region with enhanced superfluid density, it is alsodifficult to estimate the size of the enhancement in the Cooperpair density from the scanning susceptibility measurements.Nevertheless, Kirtley et al estimate an enhancement in thetwo-dimensional Cooper pair density in the range between1019 and 1020 m−2. For comparison, the two-dimensionalelectron liquid [293] at the LaAlO3–SiTiO3 interface, whichexhibits superconductivity at about 0.2 K [294], has a carrierconcentration of about 1017 m−2. The temperature dependenceof the stripe amplitude can be best fit by assuming thatthe twin boundary has a different critical temperature thanthe bulk, although no stripes were observed above thebulk Tc.

Kalisky et al also observed that vortices tended to avoidpinning on the twin boundaries, and that they were difficultto drag over the twin boundaries using either a SQUIDsusceptometer sensor or a magnetic force microscope tip [292].The enhancement of supercurrent density, as well as the barrierto vortex motion presented by the twin boundaries, providestwo mechanisms for enhanced critical currents in twinnedsamples. Prozorov et al [295] have noted an enhancementof the critical current of slightly underdoped single crystals ofBa(Fe1−xCox)2As2, which they associate with twinning.

3.3.3. Spinlike susceptibility. There appears to be acomponent of 1/f α (α ∼ 1) noise in superconductingdevices, which is apparently due to magnetic flux noise, and isremarkably universal [296, 297]. Recently Koch, DiVincenzoand Clarke (KDC) have proposed that this noise is due tounpaired, metastable spin states [44]. The proposed spin statesare bistable, with a broad distribution of activation energies forchanging their spin orientation. When the spin state changesit changes the amount of magnetic flux coupling into thesuperconducting device. Each contributes a Lorentzian witha particular cutoff frequency to the total noise; adding upa large number of Lorentzians with a broad distribution ofcutoff frequencies results in noise with 1/f distribution [43].KDC estimate that a density of unpaired spin states of about5 × 1017 m−2 is required to explain universal 1/f noise.

Bluhm et al [298] have found a signal in scanningSQUID susceptometry measurements at low temperatures in anumber of materials, both metals and insulators, even including

24

Page 26: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 29. (a), (b) Optical micrographs of two single crystals of the pnictide superconductor LaFePO. (c) Susceptibility scan of #1 atT = 0.4 K. (d) Change in heff = h + λ between 0.4 and 3 K over sample #2. (e), ( f ) Maps of local Tc over the same areas as in (c) and (d).The crosses indicate points where �λ(T ) data were collected. (g) �λ of samples #1 and #2. The dashed lines are fits between0.7 K < T < 1.6 K. (h) Black lines are possible superfluid densities for LaFePO sample #1, point 2, with different λ(0). Shaded area:superfluid density of YBa2Cu3O6.99 from [306, 307]. The width of the shaded area reflects uncertainty in λ(0). Figure reprinted withpermission from [304]. Copyright 2009 by the American Physical Society.

Au, that has a paramagnetic response with a temperaturedependence consistent with unpaired spins. This susceptibilityhas a component that is out of phase with the applied field,implying that it could contribute to 1/f -like magnetic noise.The density of these spin states is estimated to be in the range of1017 m−2, consistent with the KDC estimate. This implies thatscanning SQUID susceptometry could be used as a diagnosticfor determining which materials and processes contribute moststrongly to 1/f noise in superconducting devices.

3.4. Penetration depths

Penetration depths have been inferred from fitting the magneticimages of superconducting vortices for a number of years[32, 79, 80, 299–302]. An alternate means is to measure themutual inductance between the SQUID pickup loop and a fieldcoil integrated into the sensor in SSM [72] (figure 28). Thetwo-dimensional Fourier transform of the z-component of theresponse field in the pickup loop, with a current I in a circularring of radius R oriented parallel to, and a height z above ahomogeneous bulk superconductor with penetration depth λ is

given by [303]

Bz(k) = −πµ0IR

(q − k

q + k

)J1(kR)e−2kz, (26)

where q =√

k2 + 1/λ2 and J1 is the order 1 Bessel functionof the first kind. Similarly the response field for a thin-filmsuperconducting sample with Pearl length � is given by [303]

Bz(k) = −πµ0IRJ1(kR)e−2kz

1 + k�. (27)

To a good approximation, if z λ the magnetic fields resultingfrom the screening of the field coil fields by the sample actas if they are due to an image coil spaced by a distance2heff = 2(z + λ) from the real coil, where z is the physicalspacing between the sample surface and the field coil. ASQUID sensor with a single circular field coil of radius R anda co-planar, coaxial circular pickup loop of radius a orientedparallel to a homogeneous bulk sample surface will have amutual inductance [304]

M = µ0

2πa2

(1

R− R2

(R2 + 4h2eff)

3/2

). (28)

25

Page 27: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

The solid line in figure 28(b) is obtained by numericallyintegrating the 2D Fourier transform of equation (27) over thearea of the pickup loop, for various values of z, and fit to thedata by varying �. Figure 28(c) compares the values obtainedfor the Pearl lengths for a number of the CBCO samples usingmagnetometry and susceptometry methods. The two methodsagree within experimental error over the range of Pearl lengthspresent. Since the fitting to the Pearl vortex images was doneassuming each vortex has �0 of total flux threading through it,rather than a fractional value [305], this agreement means thatthe superconducting layers are sufficiently strongly Josephson-coupled that it is energetically favorable for the vortex fluxto thread vertically through the superconducting layers, asopposed to escaping between the layers.

Figure 29 shows results from a study using scanningSQUID susceptometry of the penetration depth of the pnictidesuperconductor LaFePO [304]. This study used the techniqueof measuring the mutual inductance between the SQUIDfield coil and pickup loop as a function of spacing z, andthen repeating the measurement at fixed z while varying thetemperature. Since the mutual inductance is a function of z + λ,the temperature dependence of λ can be inferred from suchmeasurements, even without a detailed model for the mutualinductance. Spatially resolved susceptibility measurements(figures 29(c)–( f)) showed that the effective height dependedon the topography of the sample, with scratches, bumpsand pits effecting the measurements strongly. However, themeasured penetration depth and the temperature dependenceof the penetration depth was reproducible for positions morethan 10 µm from surface irregularities. In this case thepenetration depth had a linear temperature dependence atlow temperatures, similar to the cuprate high-temperaturesuperconductors, indicative of well formed nodes in the energygap. It should be noted that in this geometry it is λab, thein-plane penetration depth, that controls the mutual inductance.

Perhaps the most promising application of MFM tothe study of superconductivity is for high-spatial resolution,absolute measurements of penetration depths. There are twoways of inferring superconducting penetration depths fromMFM measurements. The first is, as mentioned above, to fitMFM vortex images to a model with the penetration depth asa fitting parameter [301, 302]. While Nazaretski et al modeledtheir tip numerically [301], Luan et al used an analyticalapproach which is easy to fit to experimental data and notstrongly affected by differences between tips [302]. A secondmethod is to consider the interaction between the magneticmaterial in the tip and its image in the superconducting materialdue to Meissner screening of the tip fields. Following Xuet al [308], Lu et al [309] wrote the force due to the interactionbetween the magnetic tip, approximated as a magnetic dipole,and its image in the superconductor as

F(z) = 3µ0m20

32π [z + λ(T )]4, (29)

where m0 is the tip’s magnetic moment and z is the physicaldistance between tip and sample. They fit their measured MFMforce derivative curves to the z λ approximation

dF(z)

dz= − 3µ0m

20

8π [z + λ(T )]5(30)

Figure 30. Normalized superfluid densityρs(T )/ρs(0) ≡ λab(0)2/λab(T )2 versus T for the pnictidesuperconductor Ba(Fe0.95Co0.05)2As2. �λab(T ) is determined fromMFM (squares) and SSS (diamonds) experiments by measuring thechange in the diamagnetic response at fixed height. These values areoffset to match the absolute value of λab(T ) by fitting the MFM datato a truncated cone model (circles). The green solid line shows a fitto a two-band s-wave model. The width of the dashed band reflectsthe uncertainty in λab(0). Inset: �λab(T ) versus T at low T . Blackdashed line: one gap s-wave model with a = 1.5 and �0 = 1.95Tc.Magenta dashed line: �λ(T ) = cT 2.2 (c = 0.14 nm K2.2). Figurereprinted with permission from [302]. Copyright 2010 by theAmerican Physical Society.

to obtain the temperature dependence of the penetration depthλ

for a single crystal of YBCO. Luan et al [302] used a truncatedcone tip model to write

∂Fz(z, T )

∂z− ∂Fz(z, T )

∂z

∣∣∣∣z=∞

= A

{1

z + λab(T )+

h0

[z + λab(T )]2+

h20

2[z + λab(T )]3

}(31)

for the difference between the force derivative at a givenheight z and that at large distances, where A, which dependson the tip shape and coating, is determined from fittingexperimental data at low temperatures, λab(T ) is the in-planecomponent of the penetration depth and h0 is the truncationheight of the tip. Model independent values for �λ(T )

were obtained by mapping a change in λ to a change inz since they are equivalent. Absolute values for λ(T )

were obtained by fitting to equation (31). Figure 30 showsexperimental results for the superfluid density ρs of thepnictide superconductor Ba(Fe0.95Co0.05)2As2 obtained byMFM and SSM measurements. The �λ(T ) results from MFMmeasurements were later confirmed by a bulk technique onsimilar samples [310]. The advantage of MFM is that both theabsolute value and high sensitivity relative values of λ can beobtained in the same measurement, so that superfluid densityis obtained over the full temperature range and from it the gapstructure of the order parameter. Luan et al found a two full-gap model model fit the data well, consistent with the s+− orderparameter most frequently discussed in the pnictides. Luan

26

Page 28: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Figure 31. MFM imaging and manipulation of individual vortices in YBCO at T = 22.3 K. (a), (b) Schematic drawings of an MFM tip(triangle) that attracts a vortex (thick lines) in a sample with randomly distributed pinning sites (dots): at large heights (a) the force betweentip and vortex is too weak to move the vortex; at small heights (b) the vortex moves right and then left as the tip rasters over it. (c) MFMimage taken at z = 420 nm (maximum applied lateral force F lat

max ≈ 6 nN). (d) z = 170 nm (F latmax ≈ 12 nm). Inset: scan taken at a

comparable height at T = 5.2 K. (e) Line cut through the data in (c) along the dashed line, showing the signal from a stationary vortex (solidcurve with peak to the left, blue online). Overlapping it is a line cut from the reverse scan (solid curve with peak to the right, green online).( f ) Line cut through the data in (d) along the dashed line, showing a typical signal from a dragged vortex. Figure reprinted by permissionfrom [312]. Copyright 2008 Macmillan Publishers Ltd.

et al also found that ρs was uniform on the sub-micrometerscale despite highly disordered vortex pinning.

3.5. Manipulation of individual vortices

While there is a vast literature on the properties ofsuperconducting vortices [111], there has been relatively littlework on manipulation of individual vortices by scanningmagnetic probes. Such manipulation, performed with MFM[112, 116, 119, 311–313] or SSM [314], can directly measure

the interaction of a moving vortex with the local disorderpotential [312]. Keay et al used MFM to observe sequentialvortex hopping between sites in an array of artificial pinningcenters [313]. An example of another such MFM studyis shown in Figure 31 [312]. Here vortices in a highquality, detwinned sample of YBa2Cu3O6.991 are fixed at lowtemperatures and high scan heights, but at higher temperaturesand lower scan heights the tops of the vortices can be draggedby an attractive magnetic interaction with the tip. The vorticesare dragged along the fast scan direction until a combination

27

Page 29: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

of pinning and elastic forces overcomes the attractive tip–vortex force. The displacement of the vortex in the slowscan direction is much longer than in the fast scan direction,demonstrating a ‘vortex wiggling’ effect: alternating transverseforces markedly enhance vortex dragging. Vortex motion inboth the slow and fast scan directions is stochastic. From suchmeasurements Auslaender et al [312] were able to infer themaximum lateral dragging force before the top of the vorticeswere depinned. They also found a marked in-plane anisotropyin the distance w the top of the vortex was dragged by thetip in the fast scan direction, larger than could be accountedfor by the in-plane anisotropy of the vortex core and pointpinning. They speculated that a likely source for extra pinninganisotropy is nanoscale clustering of oxygen vacancies alongthe Cu–O chains in YBCO.

4. Conclusions

In summary, there are now many types of scanning magneticmicroscopies, each with their own strengths and weaknesses.Of the three covered here, SSM is the most sensitive for lowresolution applications, but requires a cooled sensor. MFM hasthe highest spatial resolution, but also applies the largest fieldsto the sample. SHM can be used over a broader temperaturerange than SSM. A major thrust in development has beentoward higher spatial resolution, which requires smallersensors scanned closer to the sample. The scaling of resolutionand sensitivity with sensor size and spacing depends both onthe technique and the source of field. Many of the applicationsdiscussed here have involved the imaging of magnetic flux inspecial geometries, unconventional superconductors or both.Others have involved time dependence, through, e.g., multiplecooldowns, cooldowns at varying cooling rates or noisemeasurements. Some exciting new developments have beenscanning susceptibility, which allows the measurement of localpenetration depths, spins and superconducting fluctuationswith unprecedented sensitivity and resolution; MFM imagingof Meissner force derivatives, which allow spatially resolvedabsolute measurements of penetration depths; and controlledmanipulation of superconducting vortices. I expect to seemany other exciting developments in the future.

Acknowledgments

I would like to thank Lan Luan, Beena Kalisky, Cliff Hicks,Danny Hykel and Klaus Hasselbach for carefully readingthis review. Any errors are mine, not theirs. I would alsolike to acknowledge support from the NSF under Grant No.PHY-0425897, by the German Humboldt Foundation, and bythe French NanoSciences Fondation, during this work.

References

[1] Bardeen J, Cooper L N and Schrieffer J R 1957 Microscopictheory of superconductivity Phys. Rev. 106 162–4

[2] Bednorz J G and Muller K A 1986 Possible high Tc

superconductivity in the Ba–La–Cu–O system Z. Phys. B64 189–93

[3] Wu M K, Ashburn J R, Torng C J, Hor P H, Meng R L,Gao L, Huang Z J, Wang Y Q and Chu C W 1987Superconductivity at 93 K in a new mixed-phaseY–Ba–Cu–O compound system at ambient pressure Phys.Rev. Lett. 58 908–10

[4] Steglich F, Aarts J, Bredl C D, Lieke W, Meschede D,Franz W and Schafer H 1979 Superconductivity in thepresence of strong Pauli paramagnetism: CeCu2Si2 Phys.Rev. Lett. 43 1892–96

[5] Stewart G R 1984 Heavy-fermion systems Rev. Mod. Phys.56 755–87

[6] Maeno Y, Hashimoto H, Yoshida K, Nishizaki S, Fujita T,Bednorz J G and Lichtenberg F 1994 Superconductivity ina layered perovskite without copper Nature 372 532–4

[7] Mackenzie A P and Maeno Y 2003 The superconductivity ofSr2RuO4 and the physics of spin-triplet pairing Rev. Mod.Phys. 75 657–712

[8] Takahashi H, Igawa K, Arii K, Kamihara Y, Hirano M andHosono H 2008 Superconductivity at 43 K in an iron-basedlayered compound LaO1−xFxFeAs Nature 453 376–8

[9] Harada K, Matsuda T, Bonevich J, Igarashi M, Kondo S,Pozzi G, Kawabe U and Tonomura A 1992 Real-timeobservation of vortex lattices in a superconductor byelectron microscopy Nature 360 51–3

[10] Harada K, Matsuda T, Kasai H, Bonevich J E, Yoshida T,Kawabe U and Tonomura A 1993 Vortex configuration anddynamics in Bi2Sr1.8CaCu2Ox thin films by Lorentzmicroscopy Phys. Rev. Lett. 71 3371–4

[11] Matsuda T, Harada K, Kasai H, Kamimura O andTonomura A 1996 Observation of dynamic interaction ofvortices with pinning centers by Lorentz microscopyScience 271 1393

[12] Harada K, Kamimura O, Kasai H, Matsuda T, Tonomura Aand Moshchalkov V V 1996 Direct observation of vortexdynamics in superconducting films with regular arrays ofdefects Science 274 1167–70

[13] Koblischka M R and Wijngaarden R J 1995 Magneto-opticalinvestigations of superconductors Supercond. Sci. Technol.8 199–213

[14] Larbalestier D C et al 2001 Strongly linked current flow inpolycrystalline forms of the superconductor MgB2 Nature410 186–9

[15] Golubchik D, Polturak E and Koren G 2009 Correlationsbeyond the horizon arXiv:0911.5613

[16] Kapitulnik A, Dodge J S and Fejer M M 1994High-resolution magneto-optic measurements with aSagnac interferometer (invited) J. Appl. Phys. 75 6872

[17] Petersen B L, Bauer A, Meyer G, Crecelius T and Kaindl G1998 Kerr-rotation imaging in scanning near-field opticalmicroscopy using a modified Sagnac interferometer Appl.Phys. Lett. 73 538

[18] Xia J et al 2008 Polar Kerr-effect measurements of thehigh-temperature YBa2Cu3O6+x superconductor: evidencefor broken symmetry near the pseudogap temperaturePhys. Rev. Lett. 100 127002

[19] Wei J Y T , Yeh N C, Fu C C and Vasquez R P 1999Tunneling spectroscopy study of spin-polarizedquasiparticle injection effects in cuprate/manganiteheterostructures J. Appl. Phys. 85 5350

[20] Wachowiak A, Wiebe J, Bode M, Pietzsch O, Morgenstern Mand Wiesendanger R 2002 Direct observation of internalspin structure of magnetic vortex cores Science 298 577

[21] Wiebe J, Wachowiak A, Meier F, Haude D, Foster T,Morgenstern M and Wiesendanger R 2004 A 300 mKultra-high vacuum scanning tunneling microscope forspin-resolved spectroscopy at high energy resolution Rev.Sci. Instrum. 75 4871

[22] Lifshitz E, Dag I, Litvitn I D and Hodes G 1998 Opticallydetected magnetic resonance study of electron/hole Traps

28

Page 30: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

on CdSe quantum dot surfaces J. Phys. Chem. B102 9245–50

[23] Chernobrod B M and Berman G P 2005 Spin microscopebased on optically detected magnetic resonanc J. Appl.Phys. 97 014903

[24] Degen C L 2008 Scanning magnetic field microscope with adiamond single-spin sensor Appl. Phys. Lett. 92 243111

[25] Balasubramanian G et al 2008 Nanoscale imagingmagnetometry with diamond spins under ambientconditions Nature 455 648–51

[26] Maze J R, Stanwix P L, Hodges J S, Hong S, Taylor J M,Cappellaro P, Jiang L, Dutt M V G, Togan E andZibrov A S 2008 Nanoscale magnetic sensing withan individual electronic spin in diamond Nature455 644–7

[27] Gallagher W J et al 1997 Microstructured magnetic tunneljunctions (invited) J. Appl. Phys. 81 3741

[28] Schrag B D, Liu X, Shen W and Xiao G 2004 Current densitymapping and pinhole imaging in magnetic tunnel junctionsvia scanning magnetic microscopy Appl. Phys. Lett.84 2937–9

[29] Schrag B D et al 2006 Magnetic current imaging withmagnetic tunnel junction sensors: case study and analysisConf. Proc. 32nd Int. Symp. for Testing and FailureAnalysis (Austin, TX) (Metals Park, OH: ASMInternational) p 13

[30] Hartmann U 1994 Scanning probe microscopy onsuperconductors: achievements and challenges Appl.Phys. A 59 41–8

[31] De Lozanne A 1999 Scanning probe microscopy ofhigh-temperature superconductors Supercond. Sci.Technol. 12 R43–56

[32] Bending S J 1999 Local magnetic probes of superconductorsAdv. in Phys. 48 449–535

[33] Josephson B D 1962 Possible new effects in superconductivetunnelling Phys. Lett. 1 251–3

[34] McCumber D E 1968 Tunneling and weak-linksuperconductor phenomena having potential deviceapplications J. Appl. Phys. 39 2503–8

[35] Stewart W C 1968 Current–voltage characteristics ofJosephson junctions Appl. Phys. Lett. 12 277–80

[36] Fulton T A and Dunkleberger L N 1974 Lifetime of thezero-voltage state in Josephson tunnel junctions Phys.Rev. B 9 4760–8

[37] Voss R F and Webb R A 1981 Macroscopic quantumtunneling in 1 µm Nb Josephson junctions Phys. Rev. Lett.47 265–8

[38] Smilde H J H, Ariando, Blank D A A, Hilgenkamp H andRogalla H 2004 π -SQUIDs based on Josephson contactsbetween high-Tc and low-Tc superconductors Phys. Rev. B70 024519

[39] Tesche C D and Clarke J 1977 Dc SQUID: Noise andoptimization J. Low. Temp. Phys. 29 301–31

[40] Dahm A J, Denenstein A, Langenberg D N, Parker W H,Rogovin D and Scalapino D J 1969 Linewidth of theradiation emitted by a Josephson junction Phys. Rev. Lett.22 1416–20

[41] Stephen M J 1968 Theory of a Josephson oscillator Phys.Rev. Lett. 21 1629–32

[42] Koch R H, Van Harlingen D J and Clarke J 1981 Quantumnoise theory for the dc SQUID Appl. Phys. Lett.38 380–2

[43] Dutta P, Dimon P and Horn P M 1979 Energy scales for noiseprocesses in metals Phys. Rev. Lett. 43 646–9

[44] Koch R H, Divincenzo D P and Clarke J 2007 Model for1/f flux noise in SQUIDs and qubits Phys. Rev. Lett.98 26

[45] Clarke J 1989 Principles and applications of SQUIDs Proc.IEEE 77 1208–23

[46] Kirtley J R and Wikswo J R Jr 1999 Scanning SQUIDmicroscopy Ann. Rev. Mater. Sci. 29 117–48

[47] Welty R P and Martinis J M 1991 A series array of dcSQUIDs IEEE Trans. Magn. 27 2924–6

[48] Huber M E, Neil P A, Benson R G, Burns D A, Corey A M,Flynn C S, Kitaygorodskaya Y, Massihzadeh O,Martinis J M and Hilton G C 2001 DC SQUID series arrayamplifiers with 120 MHz bandwidth IEEE Trans. Appl.Supercond. 11 1251–6

[49] Roth B J, Sepulveda N G and Wikswo J P Jr 1989 Using amagnetometer to image a two-dimensional currentdistribution J. Appl. Phys. 65 361–72

[50] Anderson P W and Rowell J M 1963 Probable observation ofthe Josephson superconducting tunneling effect Phys. Rev.Lett. 10 230–2

[51] Jaklevic R C, Lambe J, Silver A H and Mercereau J E 1964Quantum interference effects in Josephson tunneling Phys.Rev. Lett. 12 159–60

[52] Zimmerman J E and Mercereau J E 1964 Quantized fluxpinning in superconducting niobium Phys. Rev. Lett.13 125–6

[53] Rogers F P 1983 A device for experimental observation offlux vortices trapped in superconducting thin filmsMaster’s Thesis Massachusetts Institute of Technology,Cambridge, MA

[54] Mathai A, Song D, Gim Y and Wellstood F C 1992One-dimensional magnetic flux microscope based on thedc superconducting quantum interference device Appl.Phys. Lett. 61 598–600

[55] Black R C, Mathai A, Wellstood F C, Dantsker E,Miklich A H, Nemeth D T, Kingston J J and Clarke J 1993Magnetic microscopy using a liquid nitrogen cooledYBa2Cu3O7 superconducting quantum interference deviceApp. Phys. Lett. 62 2128–30

[56] Black R C, Wellstood F C, Dantsker E, Miklich A H,Koelle D, Ludwig F and Clarke J 1995 Imagingradio-frequency fields using a scanning SQUIDmicroscope Appl. Phys. Lett. 66 1267

[57] Vu L N and Van Harlingen D J 1993 Design andimplementation of a scanning SQUID microscope IEEETrans. Appl. Supercond. 3 1918–21

[58] Vu L N, Wistrom M S and van Harlingen D J 1993 Imagingof magnetic vortices in superconducting networks andclusters by scanning SQUID microscopy Appl. Phys. Lett.63 1693–5

[59] Ma Y P, Thomas I M, Lauder A and Wikswo J P Jr 1993 Ahigh resolution imaging susceptometer IEEE Trans. Appl.Supercond. 3 1941–4

[60] Chemla Y R, Grossman H L, Poon Y, McDermott R,Stevens R, Alper M D and Clarke J 2000 Ultrasensitivemagnetic biosensor for homogeneous immunoassay Proc.Natl Acad. Sci. 97 14268–72

[61] Tsuei C C, Kirtley J R, Chi C C, Lock See Yu-Jahnes,Gupta A, Shaw T, Sun J Z and Ketchen M B 1994 Pairingsymmetry and flux quantization in a tricrystalsuperconducting ring of YBa2Cu3O7−δ Phys. Rev. Lett.73 593–6

[62] Kirtley J R, Ketchen M B, Stawiasz K G, Sun J Z,Gallagher W J, Blanton S H and Wind S J 1995High-resolution scanning squid microscope Appl. Phys.Lett. 66 1138–40

[63] Wernsdorfer W, Bonet Orozco E, Hasselbach K, Benoit A,Barbara B, Demoncy N, Loiseau A, Pascard H andMailly D 1997 Experimental evidence of the Neel-Brownmodel of magnetization reversal Phys. Rev. Lett.78 1791–4

[64] Veauvy C, Hasselbach K and Mailly D 2002 Scanningµ-superconduction quantum interference device forcemicroscope Rev. Sci. Instrum. 73 3825

29

Page 31: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

[65] Troeman A G P, Derking H, Borger B, Pleikies J, Veldhuis Dand Hilgenkamp H 2007 NanoSQUIDs based on niobiumconstrictions Nano Lett. 7 2152–6

[66] Granata C, Esposito E, Vettoliere A, Petti L and Russo M2008 An integrated superconductive magnetic nanosensorfor high-sensitivity nanoscale applicationsNanotechnology 19 275501–600

[67] Hao L, Macfarlane J C, Gallop J C, Cox D, Beyer J, Drung Dand Schurig T 2008 Measurement and noise performanceof nano-superconducting-quantum-interference devicesfabricated by focused ion beam Appl. Phys. Lett.92 192507

[68] Finkler A, Segev Y, Myasoedov Y, Rappaport M L,Neeman L, Vasyukov D, Zeldov E, Huber M E, Martin Jand Yacoby A 2010 Self-aligned nanoscale SQUID on atip Nano Lett. 10 1046–9

[69] Ketchen M B, Awschalom D D, Gallagher W J,Kleinsasser A W, Sandstrom R L, Rozen J R andBumble B 1989 Design, fabrication, and performance ofintegrated miniature SQUID susceptometers IEEE Trans.Magn. 25 1212–5

[70] Ketchen M B and Kirtley J R 1995 Design andperformance aspects of pickup loop structures forminiature SQUID magnetometry IEEE Trans. Appl.Supercond. 5 2133–6

[71] Koshnick N C, Huber M E, Bert J A, Hicks C W, Large J,Edwards H, and Moler K A 2008 A terraced scanningsuperconducting quantum interference devicesusceptometer with sub-micron pickup loops Appl. Phys.Lett. 93 243101

[72] Gardner B W, Wynn J C, Bjornsson P G, Straver E W J,Moler K A, Kirtley J R and Ketchen M B 2001 Scanningsuperconducting quantum interference devicesusceptometry Rev. Sci. Instrum. 72 2361–4

[73] Huber M E, Koshnick N C, Bluhm H, Archuleta L J, Azua T,Bjornsson P G, Gardner B W, Halloran S T, Lucero E Aand Moler K A 2008 Gradiometric micro-SQUIDsusceptometer for scanning measurements of mesoscopicsamples Rev. Sci. Instrum. 79 053704

[74] Freitag M, Tsang J C, Kirtley J R, Carlsen A, Chen J,Troeman A, Hilgenkamp H and Avouris P 2006Electrically excited, localized infrared emission fromsingle carbon nanotubes Nano Lett. 6 1425–33

[75] Ketchen M B, Pearson D, Kleinsasser A W, Hu C K,Smyth M, Logan J, Stawiasz K, Baran E, Jaso M andRoss T 1991 Sub-µm, planarized, Nb–AlO–NbJosephson process for 125 mm wafers developed inpartnership with Si technology Appl. Phys. Lett.59 2609–11

[76] Broom R F and Rhoderick E H 1962 Studies of theintermediate state in thin superconducting films Proc.Phys. Soc. 79 586–93

[77] Weber H W and Riegler R 1973 Measurement of the fluxdensity distribution in type-II superconductors with severalmicro field-probes Solid-State Commun. 12 121–4

[78] Tamegai T, Krusin-Elbaum L, Civale L, Santhanam P,Brady M J, Masselink W T, Holtzberg F and Feild C 1992Direct observation of the critical state field profile in aYBa2Cu3O7−y single crystal Phys. Rev. B 45 8201–4

[79] Chang A M, Hallen H D, Harriott L, Hess H F, Kao H L,Kwo J, Miller R E, Wolfe R, Van der Ziel J and Chang T Y1992 Scanning Hall probe microscopy Appl. Phys. Lett.61 1974–6

[80] Chang A M, Hallen H D, Hess H F, Kao H L, Kwo J andSudbo A 1992 Scanning Hall-probe microscopy of avortex and field fluctuations in La1.85Sr0.15CuO4 filmsEurophys. Lett. 20 645–50

[81] Oral A, Bending S J and Henini M 1996 Real-time scanningHall probe microscopy Appl. Phys. Lett. 69 1324–6

[82] Oral A, Barnard J C, Bending S J, Kaya I I, Ooi S, Tamegai Tand Henini M 1998 Direct Observation of melting of thevortex solid in Bi2Sr2CaCu2O8 single crystals Phys. Rev.Lett. 80 3610–3

[83] Grigorenko A N, Bending S J, Gregory J K andHumphreys R G 2001 Scanning Hall probe microscopy offlux penetration into a superconducting YBaCuO thin filmstrip Appl. Phys. Lett. 78 1586–8

[84] Dinner R B, Beasley M R and Moler K A 2005 Cryogenicscanning Hall-probe microscope with centimeter scanrange and submicron resolution Rev. Sci. Instrum.76 103702

[85] Van Veen R G, Verbruggen A H, Van der Drift E, Radelaar S,Anders S and Jaeger H M 1999 Micron-sized Hall probeson a Si/SiGe heterostructure as a tool to study vortexdynamics in high-temperature superconducting crystalsRev. Sci. Instrum. 70 1767–70

[86] Cambel V, Karapetrov G, Elias P, Hasenohrl S, Kwok W K,Krause J and Manka J 2000 Approaching the pT rangewith a 2DEG InGaAs/InP Hall sensor at 77 KMicroelectron. Eng. 51 333–42

[87] Hicks C W, Luan L, Moler K A, Zeldov E and Shtrikman H2007 Noise characteristics of 100 nm scale GaAs/AlGaAsscanning Hall probes Appl. Phys. Lett. 90 133512

[88] Sandhu A, Kurosawa K, Dede M and Oral A 200450 nm Hall sensors for room temperature scanningHall probe microscopy Japan. J. Appl. Phys.43 777–8

[89] Mihajlovic G, Xiong P, von Molnar S, Field M andSullivan G J 2007 InAs quantum well Hall devices forroom-temperature detection of single magneticbiomolecular labels J. Appl. Phys. 102 034506

[90] Peeters F M and Li X Q 1998 Hall magnetometer in theballistic regime Appl. Phys. Lett. 72 572

[91] Geim A K, Dubonos S V, Lok J G S, Grigorieva I V,Maan J C, Hansen L T and Lindelof P E 1997 BallisticHall micromagnetometry Appl. Phys. Lett. 71 2379

[92] Novoselov K S, Morozov S V, Dubonos S V, Missous M,Volkov A O, Christian D A and Geim A 2003 Submicronprobes for Hall magnetometry over the extendedtemperature range from helium to room temperatureJ. Appl. Phys. 93 10053

[93] Bending S J and Oral A 1997 Hall effect in a highlyinhomogeneous magnetic field distribution J. Appl. Phys.81 3721–5

[94] Boero G, Demierre M, Besse P A and Popovic R S 2003Micro-Hall devices: performance, technologies andapplications Sensors Actuators A 106 314–20

[95] Li Y, Ren C, Xiong P, von Molnar S, Ohno Y and Ohno H2004 Modulation of noise in submicron GaAs/AlGaAshall devices by gating Phys. Rev. Lett. 93 246602

[96] Muller J, von Molnar S, Ohno Y and Ohno H 2006Decomposition of 1/f noise in AlxGa1−xAs/GaAs Halldevices Phys. Rev. Lett. 96 186601

[97] Martin Y and Wickramasinghe H K 1987 Magnetic imagingby force microscopy with 1000 Å resolution Appl. Phys.Lett. 50 1455

[98] Rugar D, Mamin H J, Guethner P, Lambert S E, Stern J E,McFadyen I and Yogi T 1990 Magnetic force microscopy:general principles and application to longitudinalrecording media J. Appl. Phys. 68 1169–83

[99] Hartmann U 1999 Magnetic force microscopy Ann. Rev.Mater. Sci. 29 53–87

[100] Schonenberger C and Alvarado S F 1990 Understandingmagnetic force microscopy Z. Phys. B 80 373–83

[101] Lohau J, Kirsch J, Carl A, Dumpich G and Wassermann E F1999 Quantitative determination of effective dipole andmonopole moments of magnetic force microscopy tipsJ. Appl. Phys. 86 3410–7

30

Page 32: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

[102] Phillips G N, Siekman M, Abelmann L and Lodder J C 2002High resolution magnetic force microscopy using focusedion beam modified tips Appl. Phys. Lett. 81 865–7

[103] Litvinov D and Khizroev S 2002 Orientation-sensitivemagnetic force microscopy for future probe storageapplications Appl. Phys. Lett. 81 1878

[104] Koblischka M R, Hartmann U and Sulzbach T 2003Improvements of the lateral resolution of the MFMtechnique Thin Solid Films 428 93–7

[105] Deng Z, Yenilmez E, Leu J, Hoffman J E, Straver E W J,Dai H and Moler K A 2004 Metal-coated carbon nanotubetips for magnetic force microscopy Appl. Phys. Lett.85 6263–5

[106] Champagne A R, Couture A J, Kuemmeth F and Ralph D C2003 Nanometer-scale scanning sensors fabricated usingstencil lithography Appl. Phys. Lett. 82 1111

[107] Sun S, Murray C B, Weller D, Folks L and Moser A 2000Monodisperse FePt nanoparticles and ferromagnetic FePtnanocrystal superlattices Science 287 1989

[108] Kirtley J R, Deng Z, Luan L, Yenilmez E, Dai H andMoler K A 2007 Moment switching in nanotube magneticforce probes Nanotechnology 18 465506–600

[109] Abraham D W and McDonald F A 1990 Theory of magneticforce microscope images Appl. Phys. Lett. 56 1181–3

[110] Otani Y, Kohda T, Novosad V, Fukamichi K, Yuasa S andKatayama T 2000 Shape induced in-plane magneticanisotropy reorientation in epitaxial hexagonal closepacked cobalt dots J. Appl. Phys. 87 5621–3

[111] Blatter G, Feigel’man M V, Geshkenbein V B, Larkin A I andVinokur V M 1994 Vortices in high-temperaturesuperconductors Rev. Mod. Phys. 66 1125–388

[112] Hug H J, Moser A, Parashikov I, Stiefel B, Fritz O,Guntherodt H J and Thomas H 1994 Observation andmanipulation of vortices in a YBa2Cu3O7 thin film with alow temperature magnetic force microscope Physica C235 2695–6

[113] Volodin A, Temst K, Bruynseraede Y, Van Haesendonck C,Montero M I, Schuller I K, Dam B, Huijbregtse J M andGriessen R 2002 Magnetic force microscopy of vortexpinning at grain boundaries in superconducting thin filmsPhysica C 369 165–70

[114] Moser A, Hug H J, Parashikov I, Stiefel B, Fritz O,Thomas H, Baratoff A, Guntherodt H-J and Chaudhari P1995 Observation of single vortices condensed into avortex-glass phase by magnetic force microscopy Phys.Rev. Lett. 74 1847–50

[115] Yuan C W, Zheng Z, De Lozanne A L, Tortonese M,Rudman D A and Eckstein J N 1996 Vortex images in thinfilms of YBa2Cu3O7−x and Bi2Sr2Ca1Cu2O8+x obtained bylow-temperature magnetic force microscopy J. Vac. Sci.Technol. B 14 1210–3

[116] Moser A, Hug H J, Stiefel B and Guntherodt H J 1998 Lowtemperature magnetic force microscopy onYBa2Cu3O7-delta thin films J. Magn. Magn. Mater.190 114–23

[117] Volodin A, Temst K, Van Haesendonck C andBruynseraede Y 2000 Imaging of vortices in conventionalsuperconductors by magnetic force microscopy Physica C332 156–9

[118] Volodin A, Temst K, Haesendonck C V and Bruynseraede Y2000 Magnetic force microscopy on superconductingNbSe2 and Nb surfaces Physica B 284 815–6

[119] Roseman M and Grutter P 2002 Magnetic imaging anddissipation force microscopy of vortices onsuperconducting Nb films Appl. Surf. Sci. 188 416–20

[120] Clem J R 1998 Vortex exclusion from superconducting stripsand SQUIDs in weak perpendicular ambient magneticfields Bull. Am. Phys. Soc. 43 401 (paper K36.06 andunpublished)

[121] Ketchen M B, Herrell D J and Anderson C J 1985 Josephsoncross-sectional model experiment J. Appl. Phys.57 2550–74

[122] Bermon S and Gheewala T 1983 Moat-guarded JosephsonSQUIDs IEEE Trans. Magn. MAG-19 1160–4

[123] Jeffery M, Van Duzer T, Kirtley J R and Ketchen M B 1995Magnetic imaging of moat-guarded superconductingelectronic circuits Appl. Phys. Lett. 67 1769–71

[124] Veauvy C, Hasselbach K and Mailly D 2004 Micro-SQUIDmicroscopy of vortices in a perforated superconducting Alfilm Phys. Rev. B 70 214513

[125] Dantsker E, Tanaka S and Clarke J 1997 High-Tc superconducting quantum interference devices with slots orholes: low 1/f noise in ambient magnetic fields Appl.Phys. Lett. 70 2037–9

[126] Jansman A B M, Izquierdo M, Flokstra J and Rogalla H 1999Slotted high-Tc dc SQUID magnetometers IEEE Trans.Appl. Supercond. 9 3290

[127] Likharev K K 1972 Sov. Radiophys. 14 722[128] Kuznetsov A V, Eremenko D V and Trofimov V N 1999

Onset of flux penetration into a thin superconducting filmstrip Phys. Rev. B 59 1507–13

[129] Maksimova G M 1998 Mixed state and critical current innarrow semiconducting films Phys. Solid State 40 1607–10

[130] Dantsker E, Tanaka S, Nilsson P-A, Kleiner R and Clarke J1996 Reduction of 1/f noise in high-Tc dcsuperconducting quantum interference devices cooled inan ambient magnetic field Appl. Phys. Lett. 69 4099–101

[131] Stan G, Field S B and Martinis J M 2004 Critical field forcomplete vortex expulsion from narrow superconductingstrips Phys. Rev. Lett. 92 097003

[132] Suzuki K, Yijie Li, Utagawa T and Tanabe K 2000 Magneticimaging of NdBa2Cu3Oy thin-film patterns with slotsAppl. Phys. Lett. 76 3615–7

[133] Kuit K H, Kirtley J R, van der Veur W, Molenaar C G,Roesthuis F J G, Troeman A G P, Clem J R,Hilgenkamp H, Rogalla H and Flokstra J 2008 Vortextrapping and expulsion in thin-film YBa2Cu3O7−δ stripsPhys. Rev. B 77 134504

[134] Bronson E, Gelfand M P and Field S B 2006 Equilibriumconfigurations of Pearl vortices in narrow strips Phys.Rev. B 73 144501

[135] Webb R A, Voss R F, Grinstein G and Horn P M 1983Magnetic field behavior of a Josephson-junction array:two-dimensional flux transport on a periodic substratePhys. Rev. Lett. 51 690–3

[136] Voss R F and Webb R A 1982 Phase coherence in a weaklycoupled array of 20 000 Nb Josephson junctions Phys.Rev. B 25 3446–9

[137] Sanchez D and Berchier J L 1981 Properties of n × n squarearrays of proximity effect bridges J. Low Temp. Phys.43 65–89

[138] Resnick D J, Garland J C, Boyd J T, Shoemaker S andNewrock R S 1981 Kosterlitz–Thouless transition inproximity-coupled superconducting arrays Phys. Rev. Lett.47 1542–5

[139] Abraham D W, Lobb C J, Tinkham M and Klapwijk T M1982 Resistive transition in two-dimensional arrays ofsuperconducting weak links Phys. Rev. B 26 5268–71

[140] Fazio R and Van Der Zant H 2001 Quantum phase transitionsand vortex dynamics in superconducting networks Phys.Rep. 355 235–334

[141] Kosterlitz J M and Thouless D J 1973 Ordering, metastabilityand phase transitions in two-dimensional systemsJ. Phys. C 6 1181

[142] Teitel S and Jayaprakash C 1983 Josephson-junction arrays intransverse magnetic fields Phys. Rev. Lett. 51 1999–2002

[143] Hallen H D, Seshadri R, Chang A M, Miller R E,Pfeiffer L N, West W, Murray C A and Hess H F 1993

31

Page 33: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

Direct spatial imaging of vortices in a superconductingwire network Phys. Rev. Lett. 71 3007–10

[144] Davidovic D, Kumar S, Reich D H, Siegel J, Field S B,Tiberio R C, Hey R and Ploog K 1996 Correlations anddisorder in arrays of magnetically coupledsuperconducting rings Phys. Rev. Lett. 76 815–8

[145] Davidovic D, Kumar S, Reich D H, Siegel J, Field S B,Tiberio R C, Hey R and Ploog K 1997 Magneticcorrelations, geometrical frustration, and tunable disorderin arrays of superconducting rings Phys. Rev. B55 6518–40

[146] Teitel S and Jayaprakash C 1983 Phase transitions infrustrated two-dimensional xy models Phys. Rev. B27 598–601

[147] Kolahchi M R and Straley J P 1991 Ground state of theuniformly frustrated two-dimensional xy model nearf = 1/2 Phys. Rev. B 43 7651–4

[148] Halsey T C 1985 Josephson-junction arrays in transversemagnetic fields: Ground states and critical currents Phys.Rev. B 31 5728–45

[149] Pannetier B, Chaussy J, Rammal R and Villegier J C 1984Experimental fine tuning of frustration: two-dimensionalsuperconducting network in a magnetic field Phys. Rev.Lett. 53 1845–8

[150] Itzler M A, Behrooz A M, Wilks C W, Bojko R andChaikin P M 1990 Commensurate states in disorderednetworks Phys. Rev. B 42 8319–31

[151] Rzchowski M S, Benz S P, Tinkham M and Lobb C J 1990Vortex pinning in Josephson-junction arrays Phys. Rev. B42 2041–50

[152] Lobb C J, Abraham D W and Tinkham M 1983 Theoreticalinterpretation of resistive transition data from arrays ofsuperconducting weak links Phys. Rev. B 27 150–7

[153] Runge K and Pannetier B 1993 First decoration ofsuperconducting networks Europhys. Lett. 24 737–42

[154] Field S B, James S S, Barentine J, Metlushko V, Crabtree G,Shtrikman H, Ilic B and Brueck S R J 2002 Vortexconfigurations, matching, and domain structure in largearrays of artificial pinning centers Phys. Rev. Lett.88 067003

[155] Scalapino D J 1995 The case for dx2−y2 pairing in the cupratesuperconductors Phys. Rep. 250 329–65

[156] Annett J F, Goldenfeld N D and Leggett A J 1996 PhysicalProperties of High Temperature Superconductors vol 5, edD M Ginsberg (Singapore: World Scientific) chapter 6

[157] Van Harlingen D J 1995 Phase-sensitive tests of thesymmetry of the pairing state in the high-temperaturesuperconductors—evidence for dx2−y2 symmetry Rev.Mod. Phys. 67 515–35

[158] Tsuei C C and Kirtley J R 2000 Pairing symmetry in cupratesuperconductors Rev. Mod. Phys. 72 969

[159] Wollman D A, Van Harlingen D J, Lee W C, Ginsberg D Mand Leggett A J 1993 Experimental determination of thesuperconducting pairing state in YBCO from the phasecoherence of YBCO–Pb dc SQUIDs Phys. Rev. Lett.71 2134–7

[160] Wollman D A, Van Harlingen D J, Giapintzakis J andGinsberg D M 1995 Evidence for dx2−y2 Pairing from themagnetic field modulation of YBa2Cu3O7−δ–Pb Josephsonjunctions Phys. Rev. Lett. 74 797–800

[161] Brawner D A and Ott H R 1994 Evidence for anunconventional superconducting order parameter inYBa2Cu3O6.9 Phys. Rev. B 50 6530–3

[162] Iguchi I and Wen Z 1994 Experimental evidence for a d-wavepairing state in YBa2Cu3O7−y from a study ofYBa2Cu3O7−y /insulator/Pb Josephson tunnel junctionsPhys. Rev. B 49 12388–91

[163] Miller J H Jr, Ying Q Y, Zou Z G, Fan N Q, Xu J H,Davis M F and Wolfe J C 1995 Use of tricrystal junctions

to probe the pairing state symmetry of YBa2Cu3O7−δ

Phys. Rev. Lett. 74 2347–50[164] Kouznetsov K A et al 1997 c-axis Josephson tunneling

between YBa2Cu3O7−δ and Pb: direct evidence for mixedorder parameter symmetry in a high-Tc superconductorPhys. Rev. Lett. 79 3050–3

[165] Geshkenbein V B and Larkin A I 1986 Josephson effect insuperconductors with heavy fermions JETP Lett. 43 395

[166] Geshkenbein V B, Larkin A I and Barone A 1987 Vorticeswith half magnetic flux quanta in ‘heavy-fermion’superconductors Phys. Rev. B 36 235–8

[167] Kirtley J R, Tsuei C C, Sun J Z, Chi C C, Yu-Jahnes L S,Gupta A, Rupp M, and Ketchen M B 1995 Symmetry ofthe order parameter in the high-Tc superconductorYBa2Cu3O7−δ Nature 373 225

[168] Tsuei C C, Kirtley J R, Rupp M, Sun J Z, Yu-Jahnes L S,Chi C C, Gupta A and Ketchen M B 1995 Fluxquantization in tricrystal cuprate rings—a newprobe of pairing symmetry J. Phys. Chem. Solids56 1787–95

[169] Kirtley J R, Tsuei C C, Rupp M, Sun J Z, Yu-Jahnes L S,Gupta A, Ketchen M B, Moler K A and Bhushan M 1996Direct imaging of integer and half-integer Josephsonvortices in high-Tc grain boundaries Phys. Rev. Lett.76 1336

[170] Tsuei C C, Kirtley J R, Rupp M, Sun J Z, Gupta A,Ketchen‘M B, Wang C A, Ren Z F, Wang J H andBhushan M 1996 Half-integer flux quantum effect andpairing symmetry in high-Tc tetragonal Tl2Ba2CuO6+δ

films Science 271 329[171] Kirtley J R, Tsuei C C, Raffy H, Li Z Z, Gupta A, Sun J Z

and Megert S 1996 Half-integer flux quantum effect intricrystal Bi2Sr2CaCu2O8+δ Europhys. Lett. 36 707

[172] Tsuei C C, Kirtley J R, Hammerl G, Mannhart J, Raffy H andLi Z Z 2004 Robust dx2−y2 Pairing symmetry in hole-dopedcuprate superconductors Phys. Rev. Lett. 93 187004

[173] Tsuei C C and Kirtley J R 2000 Phase-sensitive evidence ford-wave pairing symmetry in electron-doped cupratesuperconductors Phys. Rev. Lett. 85 182–5

[174] Kirtley J R, Tsuei C C and Moler K A 1999 Temperaturedependence of the half-integer magnetic flux quantumScience 285 1373

[175] Mathai A, Gim Y, Black R C, Amar A and Wellstood F C1995 Experimental proof of a time-reversal-invariant orderparameter with a π shift in YBa2Cu3O7−δ Phys. Rev. Lett.74 4523–6

[176] Sugimoto A, Yamaguchi T and Iguchi I 2002 Temperaturedependence of half flux quantum in YBa2Cu3O7−y

tricrystal thin film observed by scanning SQUIDmicroscopy Physica C 367 28–32

[177] Cedergren K, Kirtley J R, Bauch T, Rotoli G, Troeman A,Hilgenkamp H, Tafuri F and Lombardi F 2010 Interplaybetween static and dynamic properties of semifluxons inYBa2Cu3O7−δ 0-π Josephson junctions Phys. Rev. Lett.104 177003

[178] Hilgenkamp H and Mannhart J 2002 Grain boundariesin high-Tc superconductors Rev. Mod. Phys.74 485–549

[179] Hilgenkamp H, Mannhart J and Mayer B 1996 Implicationsof dx2−y2 symmetry and faceting for the transportproperties of grain boundaries in high-Tc superconductorsPhys. Rev. B 53 14586–93

[180] Mannhart J, Hilgenkamp H, Mayer B, Gerber C, Kirtley J R,Moler K A and Sigrist M 1996 Generation of magneticflux by single grain boundaries of YBa2Cu3O7−x Phys.Rev. Lett. 77 2782–5

[181] Kirtley J R, Moler K A and Scalapino D J 1997 Spontaneousflux and magnetic-interference patterns in 0–π Josephsonjunctions Phys. Rev. B 56 886–91

32

Page 34: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

[182] Mints R G, Papiashvili I, Kirtley J R, Hilgenkamp H,Hammerl G and Mannhart J 2002 Observation ofsplintered Josephson vortices at grain boundaries inYBa2Cu3O7−δ Phys. Rev. Lett. 89 067004

[183] Smilde H-J H, Hilgenkamp H, Rijnders G, Rogalla H andBlank D H A 2002 Enhanced transparency ramp-typeJosephson contacts through interlayer deposition Appl.Phys. Lett. 80 4579–81

[184] Hilgenkamp H, Ariando, Smilde H J, Blank D H A,Rogalla H, Kirtley J R and Tsuei C C 2003 Ordering andmanipulation of the magnetic moments in large-scalesuperconducting π -loop arrays Nature 422 50

[185] Lombardi F, Tafuri F, Ricci F, Miletto Granozio F, Barone A,Testa G, Sarnelli E, Kirtley J R and Tsuei C C 2002Intrinsic d-wave effects in YBa2Cu3O7−δ grain boundaryJosephson junctions Phys. Rev. Lett. 89 207001

[186] Smilde H J H, Golubov A A, Ariando, Rijnders G,Dekkers J M, Harkema S, Blank D H A, Rogalla H andHilgenkamp H 2005 Admixtures to d-wave gap symmetryin untwinned YBa2Cu3O7 superconducting films measuredby angle-resolved electron tunneling Phys. Rev. Lett.95 257001

[187] Kirtley J R, Tsuei C C, Ariando V, Harkema S andHilgenkamp H 2006 Angle-resolved phase-sensitivedetermination of the in-plane symmetry in YBa2Cu3O7−δ

Nature Phys. 2 190–4[188] Kirtley J R, Tsuei C C, Ariando, Smilde H J H and

Hilgenkamp H 2005 Antiferromagnetic ordering in arraysof superconducting π -rings Phys. Rev. B 72 214521

[189] O’Hare A, Kusmartsev F V, Laad M S and Kugel K I 2006Evidence of superstructures at low temperatures infrustrated spin systems Physica C 437–438 230–3 Proc.4th Int. Conf. on Vortex Matter in NanostructuredSuperconductors, VORTEX IV (Crete, Greece)

[190] Wheatley J M, Hsu T C and Anderson P W 1988 Interlayereffects in high-Tc superconductors Nature 333 121

[191] Anderson P W 1991 Comment on ‘Anomalous spectralweight transfer at the superconducting transition ofBi2Sr2CaCu2O8+δ’ Phys. Rev. Lett. 67 660

[192] Anderson P W 1992 Experimental constraints on the theoryof high-Tc superconductivity Science 256 1526–31

[193] Chakravarty S, Sudbo A, Anderson P W and Strong S 1993Interlayer tunneling and gap anisotropy inhigh-temperature superconductors Science261 337–40

[194] Anderson P W 1995 Interlayer tunneling mechanism forhigh-Tc superconductivity: comparison with c axisinfrared experiments Science 268 1154–5

[195] Moler K A, Kirtley J R, Hinks D G, Li T W and Xu M 1998Images of interlayer Josephson vortices in Tl2Ba2CuO6+δ

Science 279 1193–6[196] Kirtley J R, Moler K A, Villard G and Maignan A 1998

c-axis penetration depth of Hg-1201 single crystals Phys.Rev. Lett. 81 2140–3

[197] Kirtley J R, Kogan V G, Clem J R and Moler K A 1999Magnetic field of an in-plane vortex outside a layeredsuperconductor Phys. Rev. B 59 4343–8

[198] Tsvetkov A A et al 1998 Global and local measures of theintrinsic Josephson coupling in Tl2Ba2CuO6 as a test ofthe interlayer tunneling model Nature 395 360–2

[199] Pearl K 1964 Current distribution in superconducting filmscarrying quantized fluxoids Appl. Phys. Lett. 5 65–6

[200] Kogan V G, Dobrovitski V V, Clem J R, Mawatari Y andMints R G 2001 Josephson junction in a thin film Phys.Rev. B 63 144501

[201] Kogan V G 2007 Interaction of vortices in thinsuperconducting films and theBerezinskii–Kosterlitz–Thouless transition Phys. Rev. B75 064514

[202] Grigorenko A N, Bending S J, Van Bael M J, Lange M,Moshchalkov V V, Fangohr H and PAJ de Groot 2003Symmetry locking and commensurate vortex domainformation in periodic pinning arrays Phys. Rev. Lett.90 237001

[203] Kogan V G, Simonov A Yu and Ledvij M 1993 Magneticfield of vortices crossing a superconductor surface Phys.Rev. B 48 392–7

[204] Lozovik Yu E and Rakoch E A 1998 Energy barriers,structure, and two-stage melting of microclusters ofvortices Phys. Rev. B 57 1214–25

[205] Fink H J and Presson A G 1966 Magnetic irreversiblesolution of the Ginzburg–Landau equations Phys. Rev.151 219–28

[206] Chibotaru L F, Ceulemans A, Bruyndoncx V andMoshchalkov V V 2000 Symmetry-induced formation ofantivortices in mesoscopic superconductors Nature408 833–5

[207] Chibotaru L F, Ceulemans A, Bruyndoncx V andMoshchalkov V V 2001 Vortex entry and nucleation ofantivortices in a mesoscopic superconducting trianglePhys. Rev. Lett. 86 1323–6

[208] Grigorieva I V, Escoffier W, Richardson J, Vinnikov L Y,Dubonos S and Oboznov V 2006 Direct observation ofvortex shells and magic numbers in mesoscopicsuperconducting disks Phys. Rev. Lett. 96 077005

[209] Nishio T, Okayasu S, Suzuki J and Kadowaki K 2004Penetration of vortices into micro-superconductorsobserved with a scanning SQUID microscope Physica C412 379–84

[210] Moshchalkov V V, Gielen L, Strunk C, Jonckheere R, Qiu X,Van Haesendonck C and Bruynseraede Y 1995 Effect ofsample topology on the critical fields of mesoscopicsuperconductors Nature 373 319–22

[211] Nishio T, Chen Q, Gillijns W, De Keyser K, Vervaeke K andMoshchalkov V V 2008 Scanning Hall probe microscopyof vortex patterns in a superconducting microsquare Phys.Rev. B 77 012502

[212] Matsumoto M and Sigrist M 1999 Quasiparticle states nearthe surface and the domain wall in a px + ipy-wavesuperconductor J. Phys. Soc. Japan 68 994–1007

[213] Luke G M, Fudamoto Y, Kojima K M, Larkin M I, Merrin J,Nachumi B, Maeno Y, Mao Z Q and Mori Y 1998Time-reversal symmetry-breaking superconductivity inSr2RuO4 Nature 394 558–61

[214] Bjornsson P G, Maeno Y, Huber M E and Moler K A 2005Scanning magnetic imaging of Sr2RuO4 Phys. Rev. B72 12504

[215] Kirtley J R, Kallin C, Hicks C W, Kim E A, Liu Y,Moler K A, Maeno Y and Nelson K D 2007 Upper limiton spontaneous supercurrents in Sr2RuO4 Phys. Rev. B76 14526

[216] Hicks C W, Kirtley J R, Lippman T M, Koshnick N C,Huber M E, Maeno Y, Yuhasz W M, Maple M B andMoler K A 2010 Limits on superconductivity-relatedmagnetization in Sr2RuO4 and PrOs4Sb12 from scanningSQUID microscopy Phys. Rev. B 81 214501

[217] Nelson K D, Mao Z Q, Maeno Y and Liu Y 2004 Odd-paritysuperconductivity in Sr2RuO4 Science 306 1151

[218] Kidwingira F, Strand J D, Van Harlingen D J and Maeno Y2006 Dynamical superconducting order parameterdomains in Sr2RuO4 Science 314 1267

[219] Raghu S, Kapitulnik A and Kivelson S A 2010 Hidden quasione-dimensional superconductivity in Sr2RuO4

arXiv:1003.2266[220] Dolocan V O, Veauvy C, Servant F, Lejay P, Hasselbach K,

Liu Y and Mailly D 2005 Observation of vortexcoalescence in the anisotropic spin-triplet superconductorSr2RuO4 Phys. Rev. Lett. 95 097004

33

Page 35: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

[221] Dolocan V O, Lejay P, Mailly D and Hasselbach K 2006Observation of two species of vortices in the anisotropicspin-triplet superconductor Sr2RuO4 Phys. Rev. B74 144505

[222] Grigorenko A, Bending S, Tamegai T, Ooi S and Henini M2001 A one-dimensional chain state of vortex matterNature 414 728–31

[223] Saxena S S et al 2000 Superconductivity on the border ofitinerant-electron ferromagnetism in UGe2 Nature406 587–92

[224] Aoki D, Huxley A, Ressouche E, Braithwaite D, Flouquet J,Brison J P, Lhotel E and Paulsen C 2001 Coexistence ofsuperconductivity and ferromagnetism in URhGe Nature413 613–6

[225] Huy N T, Gasparini A, De Nijs D E, Huang Y, Klaasse J C P,Gortenmulder T, De Visser A, Hamann A, Gorlach T andLohneysen H 2007 Superconductivity on the border ofweak itinerant ferromagnetism in UCoGe Phys. Rev. Lett.99 067006

[226] Slooten E, Naka T, Gasparini A, Huang Y K and De Visser A2009 Enhancement of superconductivity near theferromagnetic quantum critical point in UCoGe Phys. Rev.Lett. 103 97003

[227] Tachiki M, Matsumoto H, Koyama T and Umezawa H 1980Self-induced vortices in magnetic superconductors SolidState Commun. 34 19–23

[228] Moncton D E, McWhan D B, Schmidt P H, Shirane G,Thomlinson W, Maple M B, MacKay H B, Woolf L D,Fisk Z and Johnston D C 1980 Oscillatory magneticfluctuations near the superconductor-to-ferromagnettransition in ErRh4B4 Phys. Rev. Lett. 45 2060–3

[229] Lynn J W, Shirane G, Thomlinson W and Shelton R N1981 Competition between ferromagnetism andsuperconductivity in HoMo6S8 Phys. Rev. Lett.46 368–71

[230] Ohta T, Hattori T, Ishida K, Nakai Y, Osaki E, Deguchi K,Noriaki K and Satoh I 2010 Microscopic coexistence offerromagnetism and superconductivity in single-crystalUCoGe J. Phys. Soc. Japan 79 023707

[231] Fritzsche J, Kramer R B G and Moshchalkov V V 2009Visualization of the vortex-mediated pinning offerromagnetic domains in superconductor-ferromagnethybrids Phys. Rev. B 79 132501

[232] Van Bael M J, Bekaert J, Temst K, Van Look L,Moshchalkov V V, Bruynseraede Y, Howells G D,Grigorenko A N, Bending S J and Borghs G 2001 Localobservation of field polarity dependent flux pinning bymagnetic dipoles Phys. Rev. Lett. 86 155–8

[233] Menghini M, Kramer R B G, Silhanek A V, Sautner J,Metlushko V, De Keyser K, Fritzsche J, Verellen N andMoshchalkov V V 2009 Direct visualization of magneticvortex pinning in superconductors Phys. Rev. B 79 144501

[234] Silevitch D M, Reich D H, Chien C L, Field S B andShtrikman H 2001 Imaging and magnetotransport insuperconductor/magnetic dot arrays J. Appl. Phys.89 7478

[235] Bluhm H, Sebastian S E, Guikema J W, Fisher I R andMoler K A 2006 Scanning Hall probe imaging ofErNi2B2C Phys. Rev. B 73 014514

[236] Canfield P C, Bud’Ko S L and Cho B K 1996 Possibleco-existence of superconductivity and weakferromagnetism in ErNi2B2C Physica C 262 249–54

[237] Anderson P W 1987 The resonating valence bond state inLa2CuO4 and superconductivity Science 235 1196–8

[238] Senthil T and Fisher M P A 2001 Fractionalization in thecuprates: Detecting the topological order Phys. Rev. Lett.86 292–5

[239] Bonn D A, Wynn J C, Gardner B W, Lin Yu-Ju, Liang R,Hardy W N, Kirtley J R and Moler K A 2001 A limit on

spin-charge separation in high-Tc superconductors fromthe absence of a vortex-memory effect Nature 414 887–9

[240] Sachdev S 1992 Stable hc/e vortices in a gauge theory ofsuperconductivity in strongly correlated systems Phys.Rev. B 45 389–99

[241] Nagaosa N and Lee P A 1992 Ginzburg-landau theoryof the spin-charge-separated system Phys. Rev. B45 966–70

[242] Nagaoka N 1994 e versus 2e Quantization in the spin gapphase of high-Tc superconductors J. Phys. Soc. Japan63 2835–6

[243] Wynn J C, Bonn D A, Gardner B W, Lin Y-J, Liang R,Hardy W N, Kirtley J R and Moler K A 2001 Limits onspin-charge separation from h/2e fluxoids in veryunderdoped YBa2Cu3O6+x Phys. Rev. Lett. 87 197002

[244] Kirtley J R, Tsuei C C, Kogan V G, Clem J R, Raffy H andLi Z Z 2003 Fluxoid dynamics in superconducting thinfilm rings Phys. Rev. B 68 214505

[245] Kibble T W B 1976 Topology of cosmic domains and stringsJ. Phys. A 9 1387–98

[246] Kibble T 2007 Phase-transition dynamics in the lab and theuniverse Phys. Today 60 (9) 47–52

[247] Zurek W H 1985 Cosmological experiments in superfluidhelium? Nature 317 505–8

[248] Hendry P C, Lawson N S, Lee R A M, McClintock V E andWilliams C D H 1994 Generation of defects in superfluid4He as an analogue of the formation of cosmic stringsNature 368 315–7

[249] Dodd M E, Hendry P C, Lawson N S, McClintock P V E andWilliams C D H 1998 Nonappearance of vortices in fastmechanical expansions of Liquid 4He through the lambdatransition Phys. Rev. Lett. 81 3703–6

[250] Bauerle C, Bunkov Yu M, Fisher S N, Godfrin H andPickett G R 1996 Laboratory simulation of cosmic stringformation in the early Universe using superfluid 3HeNature 382 332–4

[251] Ruutu V M H, Eltsov V B, Gill A J, Kibble T W B,Krusius M, Makhlin Yu G, Blacais B, Volovik G E andWen Xu 1996 Vortex formation in neutron-irradiatedsuperfluid 3He as an analogue of cosmological defectformation Nature 382 332–4

[252] Carmi R and Polturak E 1999 Search for spontaneousnucleation of magnetic flux during rapid cooling ofYBa2Cu3O7−δ films through Tc Phys. Rev. B60 7595–600

[253] Carmi R, Polturak E and Koren G 2000 Observation ofspontaneous flux generation in a multi-Josephson-junctionloop Phys. Rev. Lett. 84 4966–9

[254] Kavoussanaki E, Monaco R and Rivers R J 2000 Testing theKibble–Zurek scenario with annular Josephson tunneljunctions Phys. Rev. Lett. 85 3452–5

[255] Monaco R, Mygind J and Rivers R J 2002 Zurek–Kibbledomain structures: the dynamics of spontaneous vortexformation in annular Josephson tunnel junctions Phys. Rev.Lett. 89 080603

[256] Kirtley J R, Tsuei C C and Tafuri F 2003 Thermally activatedspontaneous fluxoid formation in superconducting thinfilm rings Phys. Rev. Lett. 90 257001

[257] Hindmarsh M and Rajantie A 2000 Defect formation andlocal gauge invariance Phys. Rev. Lett. 85 4660–3

[258] Ghinovker M, Shapiro B Ya and Shapiro I 2001 Spontaneousmagnetic-flux generation in superconducting ringEurophys. Lett. 53 240

[259] Donaire M, Kibble T W B and Rajantie A 2007 Spontaneousvortex formation on a superconducting film New J. Phys.9 148

[260] Monaco R, Mygind J, Rivers R J and Koshelets V P 2009Spontaneous fluxoid formation in superconducting loopsPhys. Rev. B 80 180501

34

Page 36: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

[261] Kalisky B, Kirtley J R, Nowadnick E A, Dinner R B,Zeldov E, Ariando, Wenderich S, Hilgenkamp H,Feldmann D M and Moler K A 2009 Dynamics of singlevortices in grain boundaries: I–V characteristics on thefermo-volt scale Appl. Phys. Lett. 94 202504

[262] Dimos D, Chaudhari P, Mannhart J and LeGoues F K 1988Orientation dependence of grain-boundary critical currentsin YBa2Cu3O7−δ bicrystals Phys. Rev. Lett. 61 219–22

[263] Heinig N F, Redwing R D, Nordman J E and Larbelestier D C1999 Strong to weak coupling transition in lowmisorientation angle thin film YBa2Cu3O7−x bicrystalsPhys. Rev. B 60 1409–17

[264] Lau C N, Markovic N, Bockrath M, Bezryadin A andTinkham M 2001 Quantum phase slips in superconductingnanowires Phys. Rev. Lett. 87 217003

[265] Koshnick N C, Bluhm H, Huber M E and Moler K A 2007Fluctuation superconductivity in mesoscopic aluminumrings Science 318 1440

[266] von Oppen F and Riedel E K 1992 Flux-periodic persistentcurrent in mesoscopic superconducting rings close to Tc

Phys. Rev. B 46 3203–6[267] Scalapino D J, Sears M and Ferrell R A 1972 Statistical

mechanics of one-dimensional Ginzburg–Landau fieldsPhys. Rev. B 6 3409–16

[268] Zhang X and Price J C 1997 Susceptibility of a mesoscopicsuperconducting ring Phys. Rev. B 55 3128–40

[269] Bluhm H, Koshnick N C, Bert J A, Huber M E andMoler K A 2009 Persistent currents in normal metal ringsPhys. Rev. Lett. 102 136802

[270] Chu J-H, Analytis J G, Kucharczyk C and Fisher I R 2009Determination of the phase diagram of the electron-dopedsuperconductor Ba(Fe1−xCox)2As2 Phys. Rev. B79 014506

[271] Ni N, Tillman M E, Yan J-Q, Kracher A, Hannahs S T,Bud’ko S L and Canfield P C 2008 Effects of Cosubstitution on thermodynamic and transport propertiesand anisotropic Hc2 in a(Fe1−xCox)2As2 single crystalsPhys. Rev. B 78 214515

[272] Kimber S A J et al 2009 Similarities between structuraldistortions under pressure and chemical doping insuperconducting BaFe2As2 Nature Mater. 8 471–5

[273] Liu R H et al 2009 A large iron isotope effect inSmFeAsO1−xFx and Ba1−xKxFe2As2 Nature459 64–7

[274] Pickett W E 2009 Iron-based superconductors: timing iscrucial Nature Phys. 5 87–8

[275] Yildirim T 2009 Strong coupling of the Fe-spin state and theAs–As hybridization in iron-pnictide superconductorsfrom first-principle calculations Phys. Rev. Lett.102 037003

[276] Fang C, Yao H, Tsai W-F, Hu J P and Kivelson S A 2008Theory of electron nematic order in LaFeAsO Phys. Rev. B77 224509

[277] Xu C, Muller M and Sachdev S 2008 Ising and spin orders inthe iron-based superconductors Phys. Rev. B 78 020501

[278] Colombier E, Bud’ko S L, Ni N and Canfield P C 2009Complete pressure-dependent phase diagrams forSrFe2As2 and BaFe2As2 Phys. Rev. B 79 224518

[279] Goldman A I et al 2009 Lattice collapse and quenching ofmagnetism in CaFe2As2 under pressure: a single-crystalneutron and x-ray diffraction investigation Phys. Rev. B79 024513

[280] Kreyssig A et al 2008 Pressure-induced volume-collapsedtetragonal phase of CaFe2As2 as seen via neutronscattering Phys. Rev. B 78 184517

[281] Yu W, Aczel A A, Williams T J, Bud’ko S L, Ni N,Canfield P C and Luke G M 2009 Absence ofsuperconductivity in single-phase CaFe2As2 underhydrostatic pressure Phys. Rev. B 79 020511

[282] Alireza P L, YT Ko, Gillett J, Petrone C M, Cole J M,Lonzarich G G and Sebastian S E 2009 Superconductivityup to 29 K in SrFe2As2 and BaFe2As2 at high pressures J.Phys.: Condens. Matter 21 012208

[283] Zhao J et al 2008 Structural and magnetic phase diagram ofCeFeAsO1−xFx and its relation to high-temperaturesuperconductivity Nature Mater. 7 953–9

[284] Lee C H, Iyo A, Eisaki H, Kito H, Fernandez-Diaz M T,Ito T, Kihou K, Matsuhata H, Braden M and Yamada K2008 Effect of structural parameters on superconductivityin fluorine-free LnFeAsO1−y (Ln = La, Nd) J. Phys. Soc.Japan 77 3704

[285] Kalisky B, Kirtley J R, Analytis J G, Chu J-H, Vailionis A,Fisher I R and Moler K A 2010 Stripes of increaseddiamagnetic susceptibility in underdoped superconductingBa(Fe1−xCox)2As2 single crystals: evidence for anenhanced superfluid density at twin boundaries Phys.Rev. B 81 184513

[286] Khlyustikov I N and Buzdin A I 1987 Twinning-planesuperconductivity Adv. Phys. 36 271–330

[287] Fang M M, Kogan V G, Finnemore D K, Clem J R,Chumbley L S and Farrell D E 1988 Possibletwin-boundary effect upon the properties of high-Tc

superconductors Phys. Rev. B 37 2334–7[288] Dalidovich D, Berlinsky A J and Kallin C 2008 Superfluid

density near the critical temperature in the presence ofrandom planar defects Phys. Rev. B 78 214508

[289] Sonier J E et al 1997 Measurement of the fundamental lengthscales in the vortex state of YBa2Cu3O6.60 Phys. Rev. Lett.79 2875–8

[290] Abrikosov A A, Buzdin A I, Kulic M L and Kuptsov D A1989 Phenomenological theory for twinning-planesuperconductivity in YBa2Cu3O7−x Supercond. Sci.Technol. 1 260

[291] Deutscher G and Muller K A 1987 Origin of superconductiveglassy state and extrinsic critical currents in high-Tc oxidesPhys. Rev. Lett. 59 1745–7

[292] Kirtley J R, Kalisky B, Luan L and Moler K A 2010 Meissnerresponse of a bulk superconductor with an embedded sheetof reduced penetration depth Phys. Rev. B 81 184514

[293] Breitschaft M et al 2010 Two-dimensional electronliquid state at LaAlO3–SrTiO3 interfaces Phys. Rev. B81 153414

[294] Reyren N et al 2007 Superconducting interfaces betweeninsulating oxides Science 317 1196

[295] Prozorov R, Tanatar M A, Ni N, Kreyssig A, Nandi S,Bud’ko S L, Goldman A I and Canfield P C 2009 Intrinsicpinning on structural domains in underdoped singlecrystals of Ba(Fe1−xCox)2As2 Phys. Rev. B 80 174517

[296] Wellstood F C, Urbina C and Clarke J 1987 Low-frequencynoise in dc superconducting quantum interference devicesbelow 1 K Appl. Phys. Lett. 50 772

[297] Koch R H, Clarke J, Goubau W M, Martinis J M,Pegrum C M and Harlingen D J 1983 Flicker (1/f )noise in tunnel junction DC squids J. Low Temp. Phys.51 207–24

[298] Bluhm H, Bert J A, Koshnick N C, Huber M E and Moler K A2009 Spinlike susceptibility of metallic and insulating thinfilms at low temperature Phys. Rev. Lett. 103 26805

[299] Tafuri F, Kirtley J R, Medaglia P G, Orgiani P andBalestrino G 2004 Magnetic imaging of Pearl vortices inartificially layered (Ba0.9Nd0.1CuO2+x)m/(CaCuO2)nsystems Phys. Rev. Lett. 92 157006

[300] Kirtley J R, Tsuei C C, Moler K A, Kogan V G, Clem J R andTurberfield A J 1999 Variable sample temperaturescanning superconducting quantum interference devicemicroscope Appl. Phys. Lett. 74 4011

[301] Nazaretski E, Thibodaux J P, Vekhter I, Civale L,Thompson J D and Movshovich R 2009 Direct

35

Page 37: Fundamental studies of superconductors using scanning ... · Fundamental studies of superconductors using scanning magnetic imaging ... using scanning magnetic imaging J R Kirtley

Rep. Prog. Phys. 73 (2010) 126501 J R Kirtley

measurements of the penetration depth in asuperconducting film using magnetic force microscopyAppl. Phys. Lett. 95 262502

[302] Luan L, Auslaender O M, Lippman T M, Hicks C W,Kalisky B, Chu J-H, Analytis J G, Fisher I R, Kirtley J Rand Moler K A 2010 Local measurement of thepenetration depth in the pnictide superconductorBa(Fe0.95Co0.05)2As2 Phys. Rev. B 81 100501

[303] Kogan V G 2003 Meissner response of anisotropicsuperconductors Phys. Rev. B 68 104511

[304] Hicks C W, Lippman T M, Huber M E, Analytis J G,Chu J-H, Erickson A S, Fisher I R and Moler K A 2009Evidence for a nodal energy gap in the iron-pnictidesuperconductor LaFePO from Penetration depthmeasurements by scanning SQUID susceptometry Phys.Rev. Lett. 103 127003

[305] Mints R G, Kogan V G and Clem J R 2000 Vortices inmagnetically coupled superconducting layered systemsPhys. Rev. B 61 1623–9

[306] Bonn D A and Hardy W N 2007 Handbook of HighTemperature Superconductivity ed J R Schrieffer(New York: Springer)

[307] Pereg-Barnea T, Turner P J, Harris R, Mullins G K,Bobowski J S, Raudsepp M, Liang R, Bonn D A andHardy W N 2004 Absolute values of the Londonpenetration depth in YBa2Cu3O6+y measured by zero fieldESR spectroscopy on Gd doped single crystals Phys.Rev. B 69 184513

[308] Xu J H, MillerJ H Jr and Ting C S 1995 Magnetic levitationforce and penetration depth in type-II superconductorsPhys. Rev. B 51 424–34

[309] Lu Q, Mochizuki K, Markert J T and De Lozanne A 2002Localized measurement of penetration depth for a high-Tc

superconductor single crystal using a magnetic forcemicroscope Physica C 371 146–50

[310] Kim H et al 2010 London penetration depth inBa(Fe1−xTx)2As2 (T = Co,Ni) superconductors irradiatedwith heavy ions arXiv:1003.2959

[311] Straver E W J, Hoffman J E, Auslaender O M, Rugar D andMoler K A 2008 Controlled manipulation of individualvortices in a superconductor Appl. Phys. Lett. 93 172514

[312] Auslaender O M, Luan L, Straver E W J, Hoffman J E,Koshnick N C, Zeldov E, Bonn D A, Liang R,Hardy W N and Moler K A 2008 Mechanics of individualisolated vortices in a cuprate superconductor Nature Phys.5 35–9

[313] Keay J C, Larson P R, Hobbs K L, Johnson M B, Kirtley J R,Auslaender O M and Moler K A 2009 Sequential vortexhopping in an array of artificial pinning centers Phys. Rev.B 80 165421

[314] Gardner B W, Wynn J C, Bonn D A, Liang R, Hardy W N,Kirtley J R, Kogan V G and Moler K A 2002 Manipulationof single vortices in YBa2Cu3O6.354 with a locally appliedmagnetic field Appl. Phys. Lett. 80 1010

36