You are on page 1of 9

This article was downloaded by: 10.3.97.

143
On: 20 Dec 2023
Access details: subscription number
Publisher: CRC Press
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: 5 Howick Place, London SW1P 1WG, UK

Handbook of Superconductivity
Fundamentals and Materials, Volume One
David A. Cardwell, David C. Larbalestier, I. Braginski Aleksander

Flux Quantization

Publication details
https://www.routledgehandbooks.com/doi/10.1201/9780429179181-13
Colin Gough
Published online on: 05 Jul 2022

How to cite :- Colin Gough. 05 Jul 2022, Flux Quantization from: Handbook of Superconductivity,
Fundamentals and Materials, Volume One CRC Press
Accessed on: 20 Dec 2023
https://www.routledgehandbooks.com/doi/10.1201/9780429179181-13

PLEASE SCROLL DOWN FOR DOCUMENT

Full terms and conditions of use: https://www.routledgehandbooks.com/legal-notices/terms

This Document PDF may be used for research, teaching and private study purposes. Any substantial or systematic reproductions,
re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contents will be complete or
accurate or up to date. The publisher shall not be liable for an loss, actions, claims, proceedings, demand or costs or damages
whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.
A2.7
Flux Quantization

Colin Gough

A2.7.1 Introduction c. the primary voltage standard, where the volt is now
defined in terms of a frequency, f = V/φ0 , equiva-
The quantization of magnetic flux is a direct manifestation lent to the number of flux quanta per second pass-
of the macroscopic quantum description of the supercon- ing through a Josephson junction, where V is the
ducting state. Flux quantization is a defining property of any voltage across the junction. This is known as the ac
superconductor, as is the closely related Meissner effect (see Josephson relation [3], where the frequency f = 484
Chapter A2.4). MHz μV−1. Using an array of superconducting junc-
Almost as an afterthought, while discussing the quantum- tions in series, the frequency and, hence, the voltage
mechanical, wave-like properties of superconductors, London can be measured to very high precision (see Chapters
(1950), in his monograph on Superfluids [1], added a foot- A2.8 and H5.1).
note predicting that flux in a superconducting ring would d. flux quanta as the Q-bits used in current embodiments
be quantized in units of h/q, where q was the charge of the of quantum computing (see Chapter H3.7).
superconducting electrons. London assumed that q was equal
to the single electron charge. The subsequent BCS theory [2] The observation of quantized flux in units of h/2e provided
showed that superconductivity involved paired electron states the first convincing evidence for the pairing of electrons in
(Cooper pairs). Flux is therefore quantized in units of conventional superconductors [4, 5]. A similar experiment
using an early ceramic YBa 2Cu3O8 sample [6] confirmed
(
φ0 = h / 2e 2.07 × 10−15 Tm 2 or in equivalent units of Vs ) electron pairing in the cuprate high-temperature supercon-
(A2.7.1) ductors. Because such superconductors are believed to have
d-wave symmetry, superconducting circuits that result in
The flux Φ within a thick superconducting ring or cylinder is quantized flux trapping with Φ = (n + 1/2)φ0 can be devised.
therefore given by The observation of trapped states involving half a flux quan-
tum provides the most convincing demonstration to date for
Φ=
∫ B.ds = nφ
area
0 (A2.7.2) d-wave superconductivity in the cuprate high-temperature
superconductors [7–9].
In 1955, Abrikosov showed that the magnetic state of type
where n is an integer.
II superconductors involved singly quantized flux lines pen-
Although London assumed that the quantization of flux
etrating the bulk of the material [10]. At the centre of the
was of purely academic interest, worthy only of a footnote,
flux line, the superconducting order parameter is reduced to
this property now underpins many of the most important
zero, so that states involving a circulating vortex of current
device applications of superconductors. These include:
can exist within the bulk, the flux line core essentially act-
a. superconducting quantum interference devices ing as line singularity around which the quantized current
(SQUIDs), where the response to a magnetic field is can flow. The flux associated with the circulating current is
periodic in the flux quantum (see Chapter H3.2); ϕ0, so that their number density is B/ϕ0m−2. In many cuprate
b. various versions of single flux quantum logic (SFQ), high-temperature superconductors, the coupling between the
where the quantum of flux in a superconducting superconducting CuO2 planes is so weak that they are essen-
ring is used as the elementary bit of information (see tially two-dimensional. Flux penetrates such layers in the
Chapters H3.5 and H3.6); and form of two-dimensional vortices, or flux pancakes, again

69
70 Handbook of Superconductivity

involving a single flux quantum ϕ0. As the interlayer coupling


is increased, the flux pancakes align along the field direction
to form the equivalent of a conventional flux line in a three-
dimensional, but strongly anisotropic type II superconductor
(see Chapter A3.1).
Single electron flux quantum states were later found in nor-
mal metals via the Aaronov–Bohm effect [11].

A2.7.2 Theory
A2.7.2.1 Quantization of Flux in a Ring or
Cylinder FIGURE A2.7.1 (Left) A superconducting ring or cylinder in an
external field B, illustrating the path of integration well inside the
Flux quantization is a direct consequence of a macroscopic
superconductor along which Js = 0; (Right) an allowed wave solution
wave-mechanical description of the superconducting state.
with n = 6 waves around the loop, so that the enclosed flux is 6ϕ0.
This was first suggested by London and was further devel-
oped by Abrikosov [10] within the context of the Ginsburg–
since ψ(r) must be single-valued, as shown by a typical wave
Landau (GL) theory [12], based on Landau’s earlier model of
solution with n = 6 in Figure A2.7.1(Right). Furthermore,
second-order phase transitions. In such a model, the super-
conducting state is described by a macroscopic wave function
ψ (r )e iθ(r ) where |ψ|2 = ns is the number density of supercon-
ducting electrons, and θ(r) is a position-dependent phase. By
∫ A.ds = ∫ curl A . ds = ∫ B. ds
path area area
(A2.7.8)

analogy with the conventional wave-mechanical treatment of


is simply the magnetic flux Φ enclosed within the integra-
particles in a magnetic field, the supercurrent density Js can
tion path. Combining these two results, we have the principal
be written as
result that
e*
Js =  Ψ * ( −i∇ − e * A ) Ψ + complex conjugate  (A2.7.3) n 2 π h
2m*  Φ=
e*
=n
2e
(A2.7.9)

where the term involving the vector potential A arises from


where we have assumed Cooper pairing with e* = 2e. The
the generalized momentum, p → p – e* A, of a particle with
experimental confirmation of flux quantization in units of
effective mass m* and charge e* in a magnetic field B = curl A.
h/2e therefore provides a direct confirmation of electron pair-
The above expression can be written more simply as
ing in the superconducting state.
e * nd The Meissner state with Φ = 0 is a special case of the above
Js =  ∇θ ( r ) − e * A  . (A2.7.4) relationship with n = 0. This is the only allowed solution for
m* 
a ‘singly connected’ superconductor—a superconductor with
We now consider a superconducting ring or cylinder and no hole or line singularity, such as a flux line core, passing
evaluate the line integral of the above expression well inside through it.
the superconductor, where the screening currents associ- If the thickness of the superconducting ring or cylinder is
ated with the Meissner effect are effectively zero, as shown in less than the penetration depth, or if we take an integration
Figure A2.7.1(Left) (see Chapter A2.4). Along this path Js = 0, path close to the inner surface, the screening current will not
so that be zero. It is then not the total flux that is quantized, but what
is termed the fluxoid. In this more general case, we can write

A= ∇θ ( r ) . (A2.7.5)
∫ B.ds + µ λ ∫ j.ds = nφ
e* 2
0 0 (A2.7.10)
area path
On integrating around the path, we have
2 2
 where we have substituted the relationship m* /nse * = µ 0λ L ,
∫
path
A. ds =
e* ∫ ∇θ(r ).d s.
path
(A2.7.6) where λL is the London penetration depth over which the sur-
face screening currents decay.
Fundamental to the operation of any SQUID is the period-
Now,
icity of its properties with the flux in units of the flux quantum
ϕ0. This may easily be shown by considering the energy of a
∂θ ( r )
∫ ∇θ(r ). ds = ∫
path path
∂s
ds = [θ ( r )]BA = n2π (A2.7.7) superconducting ring intersected by a single Josephson junc-
tion, as shown in Figure A2.7.2.
Flux Quantization 71

Minimizing the energy with respect to Φenclosed, we obtain

Φ − Φenclosed = LI c sin ( 2πΦenclosed / φ0 ) . (A2.7.13)

The supercurrent flowing around the ring, (Φ – Φenclosed)/L, is


therefore periodic in Φenclosed in units of the flux quantum ϕ0.
This is the basis of operation of both the rf and dc SQUID. For
a dc SQUID, the critical current is periodic in the applied flux
Φ such that

I c = 2I 0 cos(πΦ / φ0 ) (A2.7.14)

as discussed in more detail in the section on Josephson junc-


tion devices (Chapter H3.2).
FIGURE A2.7.2 A superconducting ring intersected by a
Josephson junction.
A2.7.3 Experimental
The stored energy E in a ring with inductance L can be A2.7.3.1 Flux Quantization in Conventional
written as Superconductors

(
E = Φ − Φenclosed )2 / 2L − EJ cos(θA − θB . ) (A2.7.11) The confirmation of flux quantization in conventional super-
conductors was published independently by two groups in
The first term is the stored magnetic energy (1/2)Li 2 = (1/2) 1961 [4, 5]. The same journal issue also included a number of
(Φ – Φenclosed)2/L, where (Φ – Φenclosed) is the difference related theoretical papers. Both experiments involved the use
between the applied flux Φ (Bapplied × area of ring) and of very small-diameter superconducting cylinders coated on
the flux enclosed within the superconducting ring, while the surface of a nonmagnetic core.
the second term is the Josephson energy –EJcos(θA – θB), In the measurements of Deaver and Fairbank [4], a thin cyl-
where θA – θB is the difference in phase of the supercon- inder of tin was deposited on a 1 cm length of 13-μm-diameter
ducting order parameter across the Josephson junction and copper wire. This was cooled in a small axial field, and the
EJ = I cφ0 2π (see Chapter A2.8). If we again consider a line magnetic moment trapped was deduced from the voltage
integral well within the superconducting ring, where Js = 0, induced when the sample was vibrated inside a pick-up coil.
we obtain θA – θB = 2πΦenclosed /φ0 . Many successive measurements were made on cooling in
The total energy E can therefore be written as different fields. The trapped flux was shown to be a stepped
function of applied field consistent with flux quantization in
E = ( Φ − Φenclosed ) /2L − EJ cos ( 2πΦenclosed / φ0 ). (A2.7.12)
2
units of h/2e, as shown in Figure A2.7.3. Doll and Näbauer

FIGURE A2.7.3 The first experimental confirmation of flux quantization in units of h/2e, demonstrated by the discrete values of flux
trapped in a small tin cylinder when cooled in a small axial magnetic field [4].
72 Handbook of Superconductivity

FIGURE A2.7.4 Demonstration of flux quantization in units of h/2e in a ceramic YBCO ring [6]. The lines are spaced exactly h/2e apart and
the regular transitions between the quantum states were induced by short bursts of electromagnetic radiation.

[5] performed a slightly different experiment, trapping field In one such experiment, Wellstood and co-workers used a
in a lead cylinder coated on the surface of a 10-mm quartz conventional s-wave superconductor to make a superconducting
fibre. This was supported on a very light torsion fibre and the loop connecting adjacent edges of a square HTS crystal. Because
trapped magnetic moment deduced from the torsional oscilla-
tions of the sample in the applied field. To their surprise, they
observed the flux trapped in units of about half the value pre-
dicted by London, but consistent with later ideas of electron
pairing and the quantization of flux in units of h/2e.

A2.7.3.2 Flux Quantization in the Cuprate


High-Temperature Superconductors
Many exotic theories were initially proposed to account for
the very high transition temperatures of the cuprate super-
conductors, including superconducting states involving 2, 4,
8 or even 16 electrons. At the same time, several experienced
researchers remained highly sceptical of the ‘superconduct-
ing’ transition, believing it could simply be a transition to a
very much lower-resistance state. Early measurements [6] on a
multiphase ceramic ring of Y-Ba-Cu-O demonstrated the exis-
tence of very long-lived, trapped flux states differing in flux by
h/2e, as shown in Figure A2.7.4. These measurements not only
confirmed the existence of a truly superconducting, zero-resis-
tance, quantum mechanical state, but also confirmed electron
pairing, just as in conventional superconductors. However,
such measurements cannot reveal anything about the nature
of the microscopic mechanisms leading to such pairing.

A2.7.3.3 Flux Trapping in the Cuprate


HTc Superconductors FIGURE A2.7.5 Scanning flux microscope measurements illus-
It is now firmly established that the cuprate HTc supercon- trating the d-wave symmetry of cuprate superconductors by Tsuei
ductors have a predominantly d-wave symmetry, with nodes and co-workers. (a) the orientations of the tricrystal and epitaxially
grown high Tc superconductors and the position of the lithographi-
in the superconducting wave function at 45° to the in-plane
cally patterned rings, which are intersected by weak-link, Josephson
CuO bonds. The sign of the wave function therefore reverses
junctions, on crossing the substrate grain boundaries. (b) The trap-
on changing between the a- and b-directions, as illustrated ping of half a flux quantum in the central ring crossing three grain
schematically in Figure A2.7.5(a). Several experiments have boundaries, with no flux quanta trapped in the two rings crossing
confirmed this sign change, consistent with d-wave symmetry single grain boundaries, and one flux quantum (though it could have
of the HTS wave function [7–9]. been any integer) in the isolated ring. (Courtesy of Tsuei.)
Flux Quantization 73

of the symmetry of the d-wave function, the weak-link Josephson cuprate superconductors. Such trapping has been confirmed
junctions formed at the boundaries between the d-wave and for all cuprate superconductors investigated to date, but has
s-wave superconductors have to accommodate the change in never been observed for conventional superconductors.
sign with direction of the superconducting d-wave function. This
implies that the flux trapped in such a loop is given by
A2.7.3.4 The SQUID and Flux Quantization
Φenclosed = (n + 1/ 2 ) φ0 . (A2.7.15) As indicated in the previous section, the flux enclosed by a
superconducting ring containing one or more Josephson junc-
The smallest flux that can be trapped is, therefore, half a flux tions is periodic in the flux quantum φ0 . SQUIDs (described in
quantum, and all trapped flux states are offset from the origin Chapter H3.2) can measure changes in applied flux (B × area
by half a flux quantum, as Mathai et al. [8] confirmed in a of SQUID) to an accuracy approaching 10 −7 φ0 H/z−1/2, making
series of careful measurements. a SQUID the most sensitive of any broadband magnetic sen-
Half quantum flux trapping was also observed by Kirtley et sor. A single Josephson junction has a Fraunhofer pattern with
al. [9] in a series of measurements in which thin film HTc super- a strong flux quantum dependence, which can also be used to
conductors were deposited on a tricrystal substrate (three crys- measure magnetic fields, see Schilling et al. [13].
tals cut in specific crystalline orientations and fused together Figure A2.7.6 illustrates the voltage response of a HTS rf
along their carefully cut edges). As shown in Figures A2.7.5(a) SQUID as a function of the externally applied flux ϕe, illustrat-
and (b), a series of thin film rings were lithographically pat- ing the response periodic in φ0 [14]. The curves are obtained
terned, one enclosing the intersection of all three interfaces, for different values of rf bias.
two rings crossing a single interface and another ring well away
from the interface areas. The tricrystal was carefully designed
with crystalline orientations chosen so that, over the three junc- A2.7.4 Flux Quanta and the
tions crossing the central ring, symmetry dictates that there has Mixed State of Type II
to be a π phase change in the superconducting phase. The lowest Superconductors
flux that such a ring can trap is therefore ± 1/2ϕ0, as shown in
Figure A2.7.5(b), taken from the cover of Science (Vol 271, 1995). Unlike type I superconductors, which exhibit the Meissner
Experiments demonstrating 1/2 quantum flux trapping effect excluding the penetrations of flux into the bulk, flux
provide the strongest evidence for the d-wave symmetry of the enters type II superconductors in a semi-regular array of

FIGURE A2.7.6 The periodic response of an rf HTS SQUID as a function of applied field for various levels of rf bias. The response can be
shown to be periodic in ϕ 0 over many thousands of flux quanta. (Courtesy of Gross.)
74 Handbook of Superconductivity

quantized Abrikosov vortices each carrying a single quantum


of flux.
Although quantized flux lines in the bulk of a supercon-
ductor cannot be visualized directly, their existence can be
inferred from small-angle neutron diffraction. In such mea-
surements, the magnetic moment of the neutron probes the
spatial variations of the internal magnetic field associated
with the circulating currents around each flux line. Such
measurements by Cribier et al. [15] provided the first direct
evidence for the flux line lattice in the mixed state of type II
superconductors.
Figure A2.7.7 shows a more recent example of a very high-
resolution neutron diffraction pattern obtained from a trian-
gular flux lattice in the mixed state of a large single crystal of
niobium. Note the similarity with an X-ray diffraction pat-
tern. Neutron scattering provides detailed information on
the spatial variations of the local flux density B(x), whereas
X-rays probe the variation in electron density ρ(x). Neutron
diffraction has proved to be a very powerful tool for looking
at the phase transitions and rich flux lattice states of type II
and high-Tc superconductors as a function of temperature and
applied magnetic field. FIGURE A2.7.8 Magnetic flux decoration measurements illustrat-
ing a near-perfect triangular lattice [14].

A2.7.5 Flux Decoration Measurements


the regions where flux lines leave the surface of a supercon-
It is possible to decorate individual flux lines emerging from ductor. This technique was pioneered by Essmann and Trauble
the surface of a superconductor with very fine magnetic pow- [16] in the 1970s. For a more recent review of the Bitter deco-
der, which can then be visualized using a scanning electron ration technique, see Fasano et al. [17]. Figure A2.7.8 shows
microscopy (SEM). The magnetic grains tend to segregate in an example of such a measurement, which illustrates a near-
perfect triangular lattice of singly quantized flux lines emerg-
ing from the surface of a superconductor. Occasional defects,
such as dislocations in the flux line lattice, can be identified.

A2.7.6 Scanning SQUID and Hall


Probe Microscopes
More recently, microscopes using dc SQUIDs with small
pick-up coils (< 10 μm) as magnetic pick-up coils have been
used to map out the fields produced by quantized flux lines
emerging from the surface of thin films and single crystals.
Figure A2.7.5(b) shows the use of such a microscope to mea-
sure the half flux quantum of flux in a HTS thin film ring on a
tricrystal substrates (see Figure A2.7.5[a]). A small Hall probe
can also be used to probe flux line structures with almost the
same sensitivity [18].

A2.7.7 Direct Measurement of
Quantized Flux by Electron
Holography
Another exciting modern advance has been the development
FIGURE A2.7.7 Small-angle neutron diffraction from the lattice by Tonomura’s group of transmission electron microscopy
of quantized flux lines in a niobium single crystal. (ILL Grenoble, by using interfering coherent electron beams to image quantum
courtesy of Forgan.) vortices in thin cross-sections of superconducting materials
Flux Quantization 75

FIGURE A2.7.9 Image of individual flux lines in a thinned single


crystal of niobium. Individual vortices are images as ‘dimples’ with
one side darker than the other. The larger streaks crossing the image
are interference effects associated with crystal bending and differ-
ences in thickness. (Courtesy of Tonomura.)

[19]. An example of such an image in a thin section of a Nb FIGURE A2.7.10 Magneto-optic images of vortices in NbSe2 at
superconductor is shown in figure A2.7.9, where every ‘dim- 4.2 K in a field of 8 gauss. The white marker represents 10 μm.
ple’ represents an individual flux line.

A2.7.8 Scanning Tunneling A2.7.10 Superconducting Logic


Microscopy Circuits and Quantum
Computing
Hess et al. [20] used Lorentz scanning tunneling microscopy
(STM) to investigate the electronic core states of Abrikosov Electronic circuits use Josephson junctions with single flux
flux lines. See also the review of such measurement by quanta as the quanta of information form the basis of vari-
Suderow et al. [21]. ous ultra-fast, low-power, computer circuits; see Chapters
STM/STS measurements can also be used to investigate H3.5 and H3.6. Currently, there is a great interest in the use of
flux lines. See, for example, Cottet et al. [22] investigation of superconducting circuits using the switching of flux quantum
flux vortices in the type II superconductor MgB2. states to act as Q-bits to carry information; see Chapter H3.7.

A2.7.9 Magneto-Optics References


Researchers from the University of Oslo and Bell Laboratories [1] London F 1950 Superfluids Vol 1, (New York: Wiley)
were the first to observe flux lines using an optical microscope p 152
[23]. Flux lines in NbSe2 were imaged by Hess et al. with ~ 1 μm [2] Bardeen J, Cooper L N and Schrieffer J R 1957 Phys. Rev.
resolution making use of the magneto-optic Faraday effect, 108 1175
involving rotation of the plane of polarisation of light passing [3] Josephson B D 1962 Phys. Lett. 1 251
through a ferrite garnet film placed in close proximity to the [4] Deaver B S and Fairbank W M 1961 Phys. Rev. Lett. 7 43
surface (Figure A2.7.10). [5] Doll R and Nabauer M 1961 Phys. Rev. Lett. 7 51
Magneto-optic methods are reviewed by Jooss et al. [24]. [6] Gough C E, Colclough M S, Forgan E M, Jordan R G,
Even real-time magneto-optical imaging of the vortex lattice Keene M, Muirhead C M, Rae AIM, Thomas N, Abell J S
and individual vortices is now possible [25]. Such techniques and Sutton S 1987 Nature 326 855
enable the dynamics of both correlated groups of vortices and [7] Wollman D A, Van Harlingen D J, Giapintzakis J and
individual flux quanta to be investigated as a function of time, Ginsberg D M 1995 Phys. Rev. Lett. 74 797
temperature and applied field. It is important to be able to [8] Mathai A, Gim Y, Black R C, Amar A and Wellstood F C
visualize such events, because such processes are almost cer- 1995 Phys. Rev. Lett. 74 4523
tainly present though often ignored in theoretical models for [9] Kirtley J R, Tsuei C C, Rupp M, Sun J Z, Yu-Jahnes L -S,
thermally activated and quantum creep. For a more detailed Gupta A, Ketchen M B, Moler K A and Bhushan M 1996
review of magneto-optic methods, see Chapter G3.5. Phys. Rev. Lett. 76 1336
76 Handbook of Superconductivity

[10] Abrikosov A A 1957 Sov. Phys. JETP 5 1174 [23] Goa P E, Haughlin H, Baziljevich M, Il’yashenko E,
[11] Aaronov Y and Bohm D 1959 Phys. Rev. 115 485 Gammel P and Johansen T H 2001 Supercond. Sci.
[12] Ginsburg V L and Landau L D 1950 Zh. Eksperim. I. Technol. 14 729
Teor. Fiz. 20 1064 (in Russian) [24] Jooss Ch, Albrecht J, Kuhn H, Leonhardt s and
[13] Gross R, Chaudhari P, Kawasaki M, Ketchen M B and Kronmüller H 2002 Rep. Prof. Phys. I65 651
Gupta A 1990 Appl. Phys. Lett. 57 727 [25] Gao P, Hauglin H, Olson Å A, Baziljevich M and
[14] Schilling M, Barthelmess H J, Krey S and Ludwig F 2007 Johansen T H 2003 Rev. Sc. Instrum. 74 141
Advances in Solid State Physics 40 769
[15] Cribier D, Jacort B, Rao L M and Farnoux B 1964 Phys. Further Reading
Lett. 9 106
[16] Essmann U and Trauble H 1967 Phys. Lett. A24 526 Fujita S et al Quantum Theory of Conducting Matter –
[17] Fasano F and Menghini M 2008 Supercond. Sci. Technol. Superconductivity (2009 New York: Springer)
2 (2) 023001 Kleiner R and Godoy K Supercondutivity - An Introduction
[18] Oral A, Bending S J, Humphreys R G and Heneni M (3rd edition) (2015 Berlin: Wiley-VCH)
1996 J. Low Temp. Phys. 105 1135 Rose-Innes A C and Rhoderick E H Introduction to
[19] Harada K, Matsuda T, Bonevich J, Igarisho M, Kopndo Superconductivity (1978 Oxford: Pergamon)
S, Pozzi G, Kawabe U and Tomura A 1992 Nature 360 51 Tinkham M Introduction to Superconductivity (2nd edition)
[20] Hess H F, Robinson R B, Dynes R C, Valles J M and (1996 Singapore: McGraw Hill)
Waszczak J V 1989 Phys. Rev. Lett. 62 214 Waldram JR Superconductivity of Metals and Cuprates (1990
[21] Suderow H, Guillamon I, Rodrigo J G and Viera S 2014 Bristol: IOP Publishing)
Supercond. Sci. Technol. 27(6) 063001
[22] Cottet M J G, Cantoni M, Mansart B, Alexander D T
L, Hebert C, Zhigadlo N D, Karpinski J and Carbone F
2013 Phys. Rev. B 88 014505

You might also like