You are on page 1of 11

Physica C 514 (2015) 279–289

Contents lists available at ScienceDirect

Physica C
journal homepage: www.elsevier.com/locate/physc

Organic superconductors: The Bechgaard salts and relatives


S.E. Brown ⇑
Department of Physics and Astronomy, UCLA, Los Angeles, CA 90095-1547 USA

a r t i c l e i n f o a b s t r a c t

Article history: Organic conductors were originally considered a route to achieving high temperature superconductivity.
Received 14 January 2015 While that goal could not be met, what came to be was a class of materials in which the interplay
Received in revised form 13 February 2015 between correlations and dimensionality, and sometimes geometric frustration, lead to a spectacular
Accepted 15 February 2015
diversity of phases and phenomena that are tuned by magnetic field, pressure, and temperature.
Available online 24 February 2015
Highlighted here are the physical properties of the superconducting and normal states of the first family
of organic superconductors, the quasi-one dimensional Bechgaard salts (TMTSF)2X, as well as the
Keywords:
quasi-two dimensional compounds j-(BEDT-TTF)2X. In both cases, the preponderance of experiments
Molecular conductors
Superconductivity
indicate that the superconductivity is nodal. As well, the importance of correlations is evident in the
Organic superconductors temperature/pressure phase diagrams, and the influence of low-energy magnetic fluctuations over the
normal state properties above the superconducting transition temperature is substantial.
Ó 2015 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
2. Some molecules and crystal structures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
2.1. The Bechgaard salts, (TMTSF)2X, and sulfur analogs (TMTTF)2X . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
2.2. Salts based on the BEDT-TTF donor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
3. Superconductivity and normal state properties of the q1d materials, (TMTSF)2X . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
3.1. Superconducting properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285
3.2. Normal state properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 286
4. Superconductivity and normal state properties of the q2d superconductors j-(ET)2X . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
4.1. Superconducting properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
4.2. Normal state properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
4.3. High-field phases of the q2d superconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
5. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 288
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 288
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 288

1. Introduction In this model, carriers propagate freely along a polymeric backbone,


which is augmented by a highly polarizable sidegroup in each unit,
Following acceptance of the BCS theory of superconductivity Fig. 1. The ability to adjust the constituents would allow for
came interest in exploiting that understanding toward creating optimizing parameters so as to maximize T c . In the following years,
materials with higher transition temperatures T c . Commonly associ- there was substantial interest in synthetic metals, not just for
ated with the push for synthesizing organic superconductors was high temperature superconductivity, but also for the range of appli-
Little’s 1964 proposal for a superconducting polymeric chain [1]. cations a lightweight, flexible conductor could offer. Of course,
research into polymer conductors exploded following the 1980 dis-
covery that polyacetylene could be doped and conduct electricity.
⇑ Tel.: +1 310 825 4234.
Unfortunately, doped polymers were never found to superconduct,
E-mail address: brown@physics.ucla.edu

http://dx.doi.org/10.1016/j.physc.2015.02.030
0921-4534/Ó 2015 Elsevier B.V. All rights reserved.
280 S.E. Brown / Physica C 514 (2015) 279–289

polarizable transfer between the stacks. Pressure broadens both bands,


shifting the wavevector toward the Brillouin zone center due to
sidechain
the increased charge transfer [10].
Since constructing synthetic metals in this way leads to
relatively narrow bandwidths, the influence of Coulomb interac-
tions would ordinarily be expected, and indeed 4kF fluctuations
were detected in the case of TTF-TCNQ [11]. Nevertheless, a
polymer striking success of this example is that the bandwidths are
spine sufficient to overcome the repulsions, perhaps assisted by the
polarizability of the TTF HOMO. Going forward, the results invited
Fig. 1. Model of Little’s proposal [1] for a superconducting synthetic metal. The a strategy by which the nesting conditions are worsened by
polarizable sidechains are intended to provide a means for screening the repulsive increasing the overlap orthogonal to the stack direction.
interaction for delocalized carriers moving along the spine.
The Bechgaard salts, (TMTSF)2X, with X = ClO4, PF6, AsF6, etc.,
were a natural direction in following this strategy. As for TTF, the
whereas an alternative route, based on charge transfer salts of molecules stack and the p–p overlap is the most important for
molecular constituents, is widely known for just that. the electronic properties. There is also some interstack overlap,
In the 1970s, considerable effort went into building organic since the packing arrangement for the donors is different from that
conductors from structures incorporating a molecular donor and of TTF-TCNQ in a crucial way. Specifically, The TMTSF layers align
acceptor, charge transfer salts (CTS). Among the most significant the molecules so that there is greater interstack contact, see
advances was the synthesis of the donor tetrathiafulvalene (TTF) Fig. 4 (though there is some cancelation due to overlap con-
molecule [2], and subsequent crystallization with the electron tributions of opposite sign). Further, the bandwidths are larger
acceptor tetracyanoquinodimethane (TCNQ) [3]: not only was it due to greater Se–Se overlap within the stacks. Noteworthy also
conducting at room temperature, it was also specifically metallic. is that the acceptors are singly charged, such that there is a large
Single crystals of TTF itself, in fact, reveal the structural aspect potential to further ionization. Thus, the Bechgaard salts and the
crucial to the CTS route to synthetic metals. Namely, the highly vast majority of well-studied organic CTS crystallize in a 2:1 ratio,
symmetric and planar TTF molecules stack with considerable inter- with one electron donated for every two donors.
molecular p–p alignment and overlap. The oxidation potential is In 1979, Jérome et al., reported superconductivity in
small, and a wide range of salts were synthesized. Further, owing (TMTSF)2PF6 under about 0.9 GPa of pressure, and with that dis-
to the high molecular symmetry the crystal quality can be quite covery came the new field of organic superconductivity [12]. In
good. Thus, the necessary ingredients for metallic properties were the particular case of (TMTSF)2PF6, a metal–insulator transition
all present: delocalized electronic states from p–p overlap, high occurs at 12.1 K in the absence of applied pressure. It was rea-
degree of crystalline order, and partial band filling from oxidation, lized from NMR experiments that the resulting ground state is
setting the Fermi energy inside the bands associated with the TTF a linearly-polarized spin-density wave (SDW), with lineshapes
HOMO and TCNQ LUMO [4,5]. A further consideration was that consistent with incommensurate wavevector components [13].
with sufficient intermolecular orbital overlap, Coulomb energies Analysis of 1H NMR lineshapes revealed the wavevector compo-
would not be so great so as to localize the carriers. nents as [14], in spite of the integral hole count. In the following
The structure of TTF-TCNQ with its segregated stacks, arranged decades, many dozens of organic CTS superconductors were dis-
in layers, appears in Fig. 2 [6]. Two bands cross the Fermi energy EF covered [15]. A prevalent theme is the proximity of the super-
and correspond to the anti-bonding dispersion of predominant TTF conducting phase to antiferromagnetic insulating states, which
character, and the bonding TCNQ states. Since the intrastack p–p are considered either Mott insulators, or SDW insulators. The lat-
overlap is strongest, the Fermi surface consists of two pairs of open ter are epitomized by the Bechgaard salts; in this family of con-
sheets, warped by interchain coupling resulting in substantial con- ductors, the Fermi surface consists of open sheets with very
ductivity anisotropy. good, but nevertheless imperfect nesting that drives the spin-
While TTF-TCNQ was known for its metallic properties, the high density wave incommensurability, and consequently often
degree of Fermi surface nesting makes it unstable to charge- referred to as quasi-one dimensional (q1d). Typically, the
density wave formation below T P ¼ 54 K [7], which structural superconductivity emerges once the SDW ground state is
studies associate with distortions primarily on the TCNQ stacks suppressed. The compounds exhibiting direct transitions
[8]. Additional transitions occur at lower temperature (49 K, Mott ? superconductor are characterized by stronger interstack
38 K). Variants on this theme, with TSF (tetraselenafulvalene), or contacts and consequently q2d. In almost all cases, the donor
HMTSF, lead to similar results [9]. The wavevector of the 54 K and counterion sublattices are arranged in layers, with strongly
transition phase is incommensurate, due to the incomplete charge reduced interlayer hopping.
Below, we discuss the materials in Section 2. This is followed by
the physical properties and understandings of the normal and
superconducting states in the Bechgaard salts and related TMTTF
salts in Section 2.1. The best known q2d superconductors are also
2:1 salts of the BEDT-TTF (ET) donor, which are discussed in
Section 2.2. By now, it is widely accepted that the superconductors
are singlet paired. Nevertheless, in a number of compounds, the
superconducting state is stable beyond the paramagnetic limiting
field, so long as orbital suppression is avoided by applying the
magnetic field precisely in the plane of the layers. By now,
Fig. 2. Crystal structure of TTF-TCNQ, projected along the stacking (b) axis (left), experimental evidence for inhomogeneous superconductivity,
and the a-axis. Yellow corresponds to sulfur, blue to nitrogen, and black to carbon.
(For interpretation of the references to colour in this figure legend, the reader is
sometimes referred to as LOFF or FFLO states, is convincing. And
referred to the web version of this article.) since the organic superconductors are so far unique in this regard,
S.E. Brown / Physica C 514 (2015) 279–289 281

we include some brief remarks pertaining to the novel supercon-


ductivity observed beyond the Pauli field.
Unfortunately, brevity forces the exclusion of many subtopics
and contributions of great interest. For example, organic
conductors have served well as prototypes for exploring quantum
oscillations and angular-dependent magnetoresistance (AMRO) in
high magnetic fields, and exploiting those to extract Fermi surface
parameters in quasi-one dimensional and quasi-two dimensional
cases. These explorations led to the discovery of field-induced
spin density waves [16–18], first seen in (TMTSF)2PF6 and
(TMTSF)2ClO4, which are also accompanied by quantized Hall
plateaus in a bulk system [19]. Field-induced superconductivity,
maximally stable at 30 T in k-(BETS)2FeCl4, is overlooked [20]. Fig. 4. A representative crystal structure for the Bechgaard salts. Shown here is the
structure for (TMTSF)2PF6, with views projected along the b-axis and with the
Also omitted is the subject of quantum spin liquids, which became
stacking (a) axis vertical (left), and along the a-axis. The hash marks correspond to
a subject for direct experimentation only after the report for the unit cell.
absence of magnetic order to T=J  1 in j-(ET)2Cu2(CN)3 [21].
Expansive reviews and collections related to these subjects and
others, as well as the material presented here, can be found, for HOMO, and high degree of crystallinity when forming salts by the
example, in Refs. [15,22–25]. standard electrolysis method. The highly symmetric TMTTF/TMTSF,
It is perhaps useful to contrast the opportunities presented by and ET donors are associated with more than 10, and greater than
the organic superconductors to inorganic materials such as the 50 independent superconducting compounds, respectively, and
cuprates. From the transport perspective, the organic materials more than 50 have been identified in compounds made from
are relatively cleaner owing to the fixed stoichiometry (e.g., 2:1) other organic molecules. In this article, we will discuss physical
and good crystalline quality, and hence the utility of high-field phenomena associated with the superconductors made from
research. However, that constraint goes along with fixed carrier TMTSF and BEDT-TTF donor molecules. Here, we introduce the
concentration. Historically, ground state tuning has instead been materials, their crystal and electronic structures, in order to pro-
accomplished through control over the lattice constants either by vide some context for the more detailed descriptions that follow.
counterion substitution or mechanical pressure. A lot has been
learned from a few measurement techniques, such as transport 2.1. The Bechgaard salts, (TMTSF)2X, and sulfur analogs (TMTTF)2X
and optics, magnetotransport, thermodynamic measurements,
and magnetic resonance probes. Unfortunately, some of the probes Just as for TTF-TCNQ, the stacking arrangement of the donor
that saw tremendous developments recently have not been molecules is the predominant feature of the crystal structure for
successfully applied to the organic CTS. For example, the crystals the Bechgaard salts (TMTSF)2X and their isostructural TMTTF
have proven too small, or result in too much diffuse scattering analogs. We will refer to the broader family as (TM)2X, where the
for measurements of the magnetic Bragg peaks in the AF ordered choices for anions X = PF6, ClO4, etc., are singly charged and elec-
materials, let alone inelastic scattering measurements. Also, the
tronically inert. The (high-temperature) triclinic unit cell (P1) con-
crystals are particularly prone to photon damage in ARPES
sists of two donor molecules per unit cell [26]. An example is
experiments, and to date there is a notable scarcity of scanning
shown in Fig. 4, which depicts two views of (TMTSF)2PF6. Within
tunneling microscopy/spectroscopy.
the stacks, the separation between the molecules is roughly twice
the Van der Waals radius, which for TMTSF is 3.8 Å. The coun-
2. Some molecules and crystal structures terions for the TM salts are falling into two classes, centrosymmet-
ric and non-centrosymmetric. In either case, they are confined to
In Fig. 3 is shown a number of organic molecules used for crys- the well-separated spaces between the donors. The centrosymmet-
tallization in conducting charge transfer complexes, and including ric anions are mostly hexafluorides, including the octahedral PF6,
many known superconductors. Many of the favorable properties of As6, and SbF6, but include also Br. The non-centrosymmetric anions
TTF are transferred from one to the other, namely relatively low include the tetrahedral ClO4, ReO4, and BF4, and the distorted,
oxidation potential of the chalcogen-centered pz dominated polarized tetrahedron FSO3. In all cases, at high temperature, the
counterions are undergoing rotations and positioned at locations
of inversion symmetry, which has at least two notable
consequences. First, the inter-donor distance within the stacks is
slightly dimerized, which formally sets the formal band-filling to
1/2. Second, for the non-centrosymmetric case, an order–disorder
symmetry breaking transition is expected upon cooling.
The electronic band structures are calculated in the tight-bind-
ing approximation by assuming that Se–Se contact is the dominant
source of overlap, and that the molecular orbital approximation
applies [27–29]. Thus,

Eðk ¼ 2t a cosðka aÞ  2t b cosðkb bÞ  2tc cosðkc cÞÞ; ð1Þ

with some characteristic band parameters for the a–b plane are
summarized in Table 1, where the intrastack dimerization is
ignored. As for TTF, the HOMO for TMTSF and TMTTF are the anti-
Fig. 3. Selected donor molecules used as building blocks in organic superconduc-
tors and conductors. While there are no known superconductors with TTF itself,
bonding out-of-plane p-orbitals, resulting in much larger intrastack
there are more than 100 using the TTF derivatives TMTSF or TMTTF, BEDT-TTF, and bandwidths so that ta is negative. Due to partial cancelation of
BETS. interstack hopping integrals in the b direction, one-dimensional
282 S.E. Brown / Physica C 514 (2015) 279–289

Table 1
Characteristic band parameters for selected TM salts, in the molecular orbital
104 (TMTSF)2PF6
approximation. Energies displayed are in meV. (a) = Ref. [30], (b) = Ref. [31].

Salt ta tb Ref. 103


S–SbF6 179 3 a
S–PF6 179 24 a c*
102
Se–PF6 230 58 b

ρdc (Ω cm)
101
b'
physics remains important. Since the inert anions also leave the lay-
100
ers well-separated, the overlap in the third direction is considerably
smaller. The resulting open Fermi surface with exaggerated overlap
is depicted in Fig. 5, for an orthorhombic unit cell comprised of two 10-1
donors and relatively weaker t c . a
The experimental evidence for the proposed electronic struc- 10-2
ture comes, for example, by way of anisotropic conductivity mea-
surements and angle-dependent magnetoresistance. In the first 10-3
case, a large anisotropy, qb =qa  100 over temperatures ranging
from the SDW transition at 12–300 K [32], which suggests a ratio 10-4
1 10 100
for the transfer integrals of ta =tb  10 (Fig. 6). The ratio qc =qa is Temperature (K)
even larger, albeit temperature dependent. This latter effect has
been interpreted as a dimensionality crossover [33]. The angular Fig. 6. Temperature dependent resistivity measured along the orthogonal axes
0
dependence of the magnetoresistance of the Bechgaard salts has a; b ; c in (TMTSF)2PF6. After Ref. [32].
been studied thoroughly as the prototypical quasi-one dimensional
conductor [34,35]. For example, Fermi surface parameters have clamp pressure cell is sufficient to move considerably beyond the
been inferred from prominent features in the c-axis magnetoresis- suppression of the spin-density wave (SDW) ground state of
tance qzz using a–c rotations of the field with results proportionally (TMTSF)2PF6, and well into the superconducting phase. Likewise,
consistent. the spin-Peierls (SP) ground state of the insulating (TMTTF)2PF6
Typically, however, the correlations are more significant than is replaced by an antiferromagnetic phase with comparable
those producing the SDW state for (TMTSF)2PF6. A more complete pressures. Thus, the phase diagram was pieced together from the
phase diagram for the series of salts (TMTSF)2X, (TMTTF)2X appears pressure effect on a series of salts. In fact, however, much larger
in Fig. 7 [36,37]. The principal method for achieving ground-state pressures have been applied to TMTTF salts with PF6 [38,39], and
control is through the lattice spacing, which in turn influences later with SbF6 [40], AsF6 [41] counterions, and driven the systems
the ratio of orbital overlap integrals t b =t a , and their relationship through the expected sequence of phases and finally terminating
to the on-site and near neighbor repulsive interactions U; V. The with superconducting ground states. Thus, a broader perspective
insulators are found on the low-pressure side of the phase diagram, of the physics of the Bechgaard salts and the TMTTF analogs
where the overlap is smaller and correlation effects are more effec- includes not just relatively weak correlations and resulting
tive in localizing carriers. More precisely, that control is exercised spin-density waves, but also the Mott states of the TMTTF salts.
by the application of chemical or mechanical pressure. In the first Ground states and ordering temperatures for a few salts are
case, substitution of the counterion, such as PF6 for SbF6 leads to included in Table 2.
reduced lattice constants. Likewise, choosing the TMTSF molecule In fact, an important aspect for the insulators is the prevalence
in place of the TMTTF molecule with its larger chalcogen pz orbitals of charge-ordering (CO) instabilities [42–44]. That is, there are two
leads to greater overlap. To illustrate, we consider the calculated
band parameters and lattice constants for (TMTTF)2SbF6,
(TMTTF)2PF6, and (TMTSF)2PF6 in Table 1. Modest mechanical
pressure produces similar effects, since molecular crystals are also
relatively soft in all directions. Even 1 GPa from a standard BeCu
T
TM2X

CO
1

0.5 SP
AF AF/SDW
SC
k /π

0
y

P
(TMTSF) 2PF6
(TMTTF) 2PF6
(TMTTF) 2SbF6

(TMTTF) 2Br

−0.5

−1
0.5
1
0
0
Fig. 7. Generic phase diagram for the TMTSF/TMTTF salts with centrosymmetric
kx/π −0.5 −1 anions. Pressure increases the interchain overlap tb , influencing directly the ground
kz/π states. Ambient-pressure properties for selected salts are shown by their position-
ing along the pressure axis. The ground states are labeled as CO = charge ordered,
Fig. 5. Model quasi-one dimensional Fermi surface for an orthorhombic cell. The SP = spin-Peierls, AF = antiferromagnet, SC = superconductivity. The effective pres-
warping in the b; c directions is exaggerated for emphasis. sure span is of order 5.0 GPa.
S.E. Brown / Physica C 514 (2015) 279–289 283

Table 2 which are observed when forming salts with a remarkably wide
Transition temperatures to broken-symmetry ground states for TM2X. The designa- variation of counterion species. The polytypes are each designated
tions are: AF = antiferromagnetic, SP = spin-Peierls, SDW = spin-density wave,
SC = superconducting.
with a Greek letter; an example is the j-phase arrangement,
j-ET2Cu(NCS)2, shown in Fig. 8, with its highly dimerized arrange-
Salt Gnd. state T c (K) P ment on a parquet-like pattern. Some of the other polytype
S–SbF6 AF 8 Ambient arrangements are depicted in Fig. 9. Quite generally, the ethylene
S–AsF6 SP 13 Ambient endgroups are located adjacent to the neighboring counterion
S–PF6 SP 20 Ambient
Se–PF6 SDW 12.1 Ambient
spacer layers, so the bar-like entities shown are intended to sug-
Se–PF6 SC 1.4 0.6 GPa gest a projected view orthogonal to the long molecular axis.
Se–ClO4 SC 1.4 Ambient The molecular stacks in ET salts tend to have a shifted overlap,
due to steric effects originating with the ethylene endgroups. This
weakens the p-orbital overlap and reduced conductivities along
routes to the insulating state. In the first, the CO results from an the stacking direction when compared to the TMTSF salts or TTF-
effective 1/4-filling and significant near-neighbor repulsion V. TCNQ. On the other hand, the staggered interstack S–S contacts
The starting point here is the extended Hubbard model [45]. are significant and provide sufficient overlap along with higher
Perhaps, also, it is assisted by coupling to the counterion displace- relatively higher conductivity in the ET salts perpendicular to the
ments [46–48]. A second, distinct route to an insulating state stacks, and hence are commonly grouped as quasi-two dimen-
occurs even when V is less important, because bond alternation sional organic conductors and superconductors [22, see Ch. 11].
along the stacks leads to the so-called Dimer-Mott state. When discussing the ET salts, a consideration sometimes
neglected is also a source of disorder. The molecular conformation
2.2. Salts based on the BEDT-TTF donor shown in the figure includes a notable twist at each end, resulting
from ethylene group ordering. This is common to the case of ET
The large number of conductors and superconductors based on donors, which are nonplanar, and in this way different from TTF
the BEDT-TTF donor can be linked to the considerable variety of or TMTSF. Where the twists are in the identical direction, the
intralayer molecular arrangements for the ET-based compounds conformation is said to be eclipsed, and in opposite directions they
are staggered. Considerable conformational disorder is not
uncommon, is often quenched upon cooling, and can lead to thermal
hysteresis in transport properties [49].
The interlayer overlap is extremely weak in most cases. For
some materials, such as the familiar j phase conductors, there
has been sufficient effort in determining whether coherent inter-
layer transport occurs at all, and in other cases it is very weak
[23,50]. Consequently, the band structure is usually discussed in
terms of the intralayer overlaps and Fermi surface. Then, for many
of the polytypes, there are four donors per unit cell. This would
apply to the j; a, and b00 cases, for example. With two holes dis-
tributed among the four donors, the projected FS area is precisely
one (2d) Brillouin zone. Typically, two bands cross the Fermi
energy producing a FS featuring a q2d hole pocket and a q1d sheet,
such as that shown in Fig. 10, and reproduced from Ref. [51].
Again, there is ample experimental evidence supporting this
general picture, particularly when it comes to magnetotransport
results. The quantum oscillations set in at relatively modest mag-
netic fields for the q2d pockets, with magnetic breakdown giving
the expected area of one BZ at higher fields [23,50, and references
therein].
The picture is incomplete, however, since correlations play a
significant role in many of the ET salt families, with the relative
importance again tuned with chemical or mechanical pressure.
The j-phase materials illustrate this point very simply. In that

Fig. 8. Crystallographic arrangement of donor molecules and counterions in j-


(BEDT-TTF)2Cu(NCS)2. (Top:) view along b-axis, and (bottom:) along the a-axis, Fig. 9. Intralayer BEDT-TTF arrangements for representative polymorphs of BEDT-
showing the strongly dimerized arrangement of the donors for the j phase. TTF, designated as a; b; b00 .
284 S.E. Brown / Physica C 514 (2015) 279–289

Fig. 10. Example of quasi-two dimensional band structure and Fermi surface, containing a closed pocket and an open sheet, from j-(ET)2Cu(NCS)2 (after Ref. [51]). Many of
the ET salts containing four donors in the unit cell have electronic structure with these features.

case, the minimal model frequently adopted is that where the only weakly first order, an interesting question to ask is, what is
overlap integral associated with intradimer pair (Fig. 8) effectively the influence of the collective charge fluctuations at low tempera-
distributes the hole over both donors [52], such that the on-site ture and nearby to the CO phase boundary? In Ref. [61], it was sug-
Coulomb repulsion U also pertains to the dimer. Then, the gested that they are associated with an effective attraction
interdimer overlaps are relevant for carrier delocalization. In this sufficient for pairing. We return to this point at the conclusion of
case, the T=P phase diagram is somewhat simpler than for the q1d the article.
materials. In Fig. 11 is a depiction of the tuning by pressure for
j-(ET)2Cu[N(CN)2]Cl (j-Cl) [53]. Effectively, pressure influences
the U=t ratio, which was also manipulated indirectly by deuterating 3. Superconductivity and normal state properties of the q1d
the ethylene groups in the ambient-pressure superconductor materials, (TMTSF)2X
j-(ET)2Cu[N(CN)2]Br (j-Br) [49]. Unlike the TM salts, the q1d insta-
bilities (SP, SDW) associated with the quasi-one dimensionality are With the phase transition to a fully-gapped density wave state
missing. Instead, there is a line of first-order phase transitions occurring only at T SDW  12 K in (TMTSF)2PF6, the uniaxially-ori-
separating the conducting and insulating sides, which terminates ented SDW is suppressed with only modest pressure by worsening
at a critical point [53,54]. The ground state on the conducting side the nesting conditions. Superconductivity is stabilized at pressures
is superconducting, where the low-temperature resistivities follow greater than Pc ’ 0:6 GPa. Antiferromagnetic spin fluctuations (SF)
the form characteristic of a Fermi Liquid, dqðTÞ  T 2 [55,56]. The nevertheless persist to much greater pressures, and consequently,
ground state on the insulating side is apparently determined by the Bechgaard salts are among the first known examples where
the details of the frustration on the anisotropic triangular lattice superconducting pairing is often linked to magnetic interactions.
(on which the dimer pairs are situated): it is either AF (j-Cl), or In the same year, superconductivity was discovered in the heavy
Quantum Spin Liquid (QSL, j-(ET)2Cu2(CN)3) [21,57,58]. fermion compound CeCu2Si2 [62].
In those polytypes absent a strong dimerization, CO instabilities The temperature/pressure phase diagram for (TMTSF)2PF6
often prevail [59,60]. In principle, the transition separating appears in Fig. 12. NMR experiments indicate that the SDW ground
conducting and insulating phases can be continuous or first-order, state is incommensurate, arising from imperfect nesting condi-
with the symmetry-breaking leading to horizontal or diagonal tions. Pressure increases overlap, in turn worsening the nesting
stripes. However, generic expectations would have the transition conditions. The Peierls temperature TSDW = 12.1 K is reduced
be first order at sufficiently low temperature that a magnetic order smoothly to T = 0 at a critical value of Pc  0:6 GPa. For P > Pc ,
parameter develops on top of the CO. Should the transition remain superconductivity is observed at temperatures less than

14
T
12
crossover (TMTSF)2PF6
10
temperature (K)

insulating 8
CP
6
SDW SC
4
R~T2
2
AFI
SC 0
0 0.2 0.4 0.6 0.8 1 1.2
P
κ-Br
κ-Cl

κ-NCS

pressure (GPa)

Fig. 12. Temperature/pressure phase diagram for (TMTSF)2PF6, showing the


suppression of the spin density wave ground state with pressure, and the
Fig. 11. Depiction of temperature/pressure phase diagram for j-(ET)2Cu[N(CN)2]Cl. emergence of a superconducting state. The ground states of (TMTSF)2AsF6 evolve
See Ref. [53]. with pressure similarly [63].
S.E. Brown / Physica C 514 (2015) 279–289 285

T c K 1:5 K, with the transition temperature slowly and monotoni- 6


cally decreasing upon increasing the pressure beyond Pc . The
superconducting transition temperature is maximum close to P c , 5 μ0H//b’
and decreases with greater pressures [64]. Replacing PF 6 for other

magnetic field μ0H (T)



counterions, such as AsF 6 or ClO4 gives similar results, with the 4 (TMTSF) PF
2 6
perchlorate salt already superconducting at ambient pressure.
P = 0.6GPa)
3
3.1. Superconducting properties
2 μ0H//a μ0HP
At the time of discovery, the physical properties of the super-
conducting state were expected conventional, fully independent 1
of the spin-density wave instability and with a fully-gapped μ0H//c*
s-wave order parameter. In the 3–4 years that followed, that pic- 0
ture was very much in doubt for several reasons. On the experi- 0 0.5 1 1.5
mental side, counter-evidence accumulated quickly. For instance, temperature T (K)
a rapid suppression of the critical temperature results from the
introduction of non-magnetic impurities [65,66] was regarded as Fig. 13. l0 Hc2 for (TMTSF)2PF6, determined resistively and with the field parallel to
0
the conducting planes (a-, b -directions), and orthogonal to the planes (c ). (After
evidence for a sign change of the order parameter over the Fermi Ref. [74]).
surface. This, and the quasi-one-dimensionality of the electronic
structure led to the suggestion that intrastack triplet (px -wave)
pairing was a possibility [67]. At about the same time, a model length is shorter than the mean-free-path. That leaves other
for interstack pairing by antiferromagnetic spin fluctuations [68] possibilities: that the superconductor is driven unstable by the
was introduced for the first time [69]. These two proposals lead magnetic field, but does not revert directly to the normal state.
to distinctly different properties in the superconducting state. In Instead, there is an intermediate state which is superconducting
the first case, the FS is fully-gapped, and therefore the low tem- but characterized by a nonzero spin susceptibility. The inhomoge-
perature thermodynamic quantities are activated. In addition, a neous state first proposed Fulde and Ferrell [78] and independently
Hebel–Slichter (HS) enhancement of the spin-lattice relaxation by Larkin and Ovchinnikov [79], follows that prescription. Such
rate is expected for T 6 T c , and quasiparticle Zeeman coupling does LOFF (or FFLO) formation is a consequence of the quasiparticle
not suppress the superconconductivity. On the other hand, the Zeeman interaction in clean, Pauli-limited superconductors, and
interstack AF spin fluctuation picture leads to a d-wave-like order should exhibit a greater range of stability in layered systems
parameter with lines of gap zeroes. Thus, the thermodynamic because orbital suppression is quenched for in-plane fields. Even
properties in the limit T ! 0 are power laws in temperature and then, however, Pauli-limiting behavior would be expected, and
the HS peak in the relaxation rate should be suppressed. then the LOFF state is accessed through a line of phase transitions
1
H was the target nucleus in the first exploration of the relax- at a magnetic field of the order of the paramagnetic limit. However,
ation rate in the ambient-pressure superconductor (TMTSF)2ClO4 0
with Hkb , there is no signature for Pauli-limiting behavior, and
[70]. To avoid orbital suppression of the superconductivity with therefore, (TMTSF)2PF6 was instead considered a candidate for a
high magnetic field, Takigawa and co-workers followed the origi- spin triplet pairing state [76,80].
nal 27Al example of Hebel and Slichter [71], by cycling the mag- Knight shift measurements on (TMTSF)2ClO4 make the possibil-
netic field to zero for the evolution time. The magnetic field was ity for triplet pairing very unlikely, at least at low magnetic field
raised again to the resonance condition only to read the resulting strengths. In Ref. [81] is reported 77Se NMR spectroscopy at
magnetization decay. The relaxation rate was found to follow temperatures less than 100 mK, and magnetic fields less than 1 T.
roughly a T 3 power law from T c down to  T c /2, and interpreted Under these conditions, the spin polarization was found to be less
as evidence for line nodes in the SC order parameter [70,72]. Not than half that expected for the normal state. The remaining
long afterward, similar behavior was observed for cuprate super- response could be attributed to triplet pairing [82], but then it is
conductor YBa2Cu3O6.91 [73]. important to note that even Zeeman shifts of the quasiparticle
At that time, the idea of a nodal superconductor, for TMTSF as states near to gap zeroes assure non-zero spin polarization in a
well as the cuprates was still controversial. With regard to the for- magnetic field [83]. Therefore, the experimental results are
mer, the highly anisotropic field/temperature phase diagram
reported for the PF6 salt regenerated interest in the problem [74].
Later, similar results were reported for the perchlorate salt [75]. 40
−1 −1
0.4
(T1T) (s−K) , H//b’
The upper critical field shown in Fig. 13 was determined using
−1 −1
[ T T] −1 (s−K)−1

resistivity measurements, which revealed stable superconductivity 30 6×(T1T) (s−K) , H//a 0.3
to magnetic fields significantly greater than the paramagnetic- Rzz
Rzz (Ω)

limiting field HP  2:5 T [76]. The robustness against magnetic 20 0.2


fields applies only when the field is aligned to the molecular layers,
0
that is, in the a—b plane. Tilting of the field quickly leads to orbital
1

10 0.1
suppression of the SC state. Avoiding orbital suppression is particu-
77

0
larly strong for fields aligned orthogonal to the stacks, along b . 0 0
A superconductor surviving to fields exceeding the para- 0 1 2 3 4 5
magnetic limit can occur under several circumstances. From a
magnetic field μ H (T)
thermodynamic perspective, if the superconducting state acquires 0
a spin susceptibility of the order of the normal state, there is noth-
Fig. 14. 77Se ½T T 1 vs. l0 H, in the limit T ! 0, normalized to the normal state, and
ing to be gained by reverting to the normal state. The most familiar contrasted to the interlayer resistance Rzz ðl0 HÞ. The recovery of the NMR relaxation
means for doing that is strong spin-orbit scattering [77]. For that to toward the normal state is observed to occur at much smaller fields than the
apply, the superconductor is in the dirty limit, where the coherence resistance.
286 S.E. Brown / Physica C 514 (2015) 279–289

5 Similar behavior was documented in the field-dependent speci-


fic heat. The phase diagram obtained from those results, for the
H // a
4
field applied along the stacking axis, is shown in Fig. 15. Shown
C onset are the onset temperatures in different fields for both specific heat,
R onset and transport. In the former case, the field beyond which
μ 0Hc2 (T)

3 R zero superconductivity is undetectable appears paramagnetic limited.


Whereas, the transport results include signatures for the persis-
2 tence of an SC state to much larger fields. Therefore, despite the
transport properties, evidence for a superconductivity state at high
fields, in the bulk, remains elusive for the Bechgaard salts. On the
1
other hand, we discuss briefly later that the Zeeman-driven
inhomogeneous (LOFF) phase is stabilized in some quasi-two
0 dimensional systems, such as j-(ET)2CuðNCSÞ2 ; k-(BETS)2FeCl4,
0 0.5 1 1.5
and b00 -(ET)2SF5CH2CF2SO3.
T (K)

Fig. 15. Comparison of the phase diagram obtained from specific heat and transport 3.2. Normal state properties
measurements. The magnetoresistance is measured along the c-axis. After [84].

The normal state properties of the TMTSF salts lend credibility


considered as strong evidence for a singlet pairing state with lines to the scenario for pairing mediated by AF spin fluctuations. For
of zeroes over the Fermi surface. instance, signatures are evident in the 77Se nuclear spin lattice
The argument is only strengthened by specific heat measure- relaxation rate (SLRR), and furthermore, what is interpreted as
ments performed on (TMTSF)2ClO4. In Ref. [84] are reported non-Fermi Liquid (NFL) behavior is detected in the resistivities.
detailed measurements using a vector magnet to apply in-plane With respect to the SLRR, the manifestation of the fluctuations is
fields, the results of which were used to map the gap structure, in an enhancement over what could be expected from the
which resembles a d-wave-like order parameter. With the excep- Korringa relaxation of an ordinary metal. In Fig. 16 is T 1 vs. T
1
tion of the loose end of the indication for a full gap from thermal for (TMTSF)2PF6 over a range of pressures covering both sides of
conductivity measurements [85], all of these features, as well as
the critical pressure at Pc ¼ 0:6 GPa. At lower pressures, T 1 1
the normal state properties, are consistent with superconducting
increases and tends to diverge as the temperature is lowered
pairing mediated by interstack antiferromagnetic spin fluctuations
toward T SDW . Such behavior is expected for critical slowing of the
[86].
dynamics. For P > Pc , the divergence is suppressed. Nevertheless,
What remains unclear is the nature of the superconducting
an enhanced rate, largely temperature independent and therefore
state at high magnetic fields, for field strength of the order of
very different from Korringa, remains. Clearly, however, in the
and exceeding the paramagnetic limiting field of 2–2.5 T. Both
the specific heat [84] and the NMR relaxation [81] rates would limit that T ! 0, ½T 1 T1 ! constant, which we take to be the
appear inconsistent with interesting proposals, such as the hallmark of a Fermi Liquid ground state.
possibility for a field-induced transition to a triplet order parame- A Curie–Weiss (CW) relation T 1 eT þ H, with H increasing
ter [87–89], or the transition to the inhomogeneous LOFF state monotonically with pressure [81,92] describes the SLRR extremely
[90,91]. In the case of the relaxation rates, significant field-depen- well for pressure P > P c (Fig. 17). The behavior matches what is
dence of the SC state relaxation rate were reported. At lower fields, expected for 2d overdamped spin fluctuations, provided that the
temperature is not low when compared to the spin fluctuation
½T 1 T1 is a field-dependent constant in the limit T ! 0. However,
damping scale [93,94]. Following that phenomenology,
as shown in Fig. 14, ½T 1 T1 approaches the normal state result very
rapidly at fields in the range 1.5–2 T, while the interlayer ½T 1 T1  vðQ Þ; ð2Þ
resistance Rzz , measured in situ, remains negligible.
and vðQ Þ  T þ H. Notably, similar observations have been made
for the cuprates, e.g., Ref. [95], and the pnictides e.g., Ref. [96]. In
150
P(GPa) 0.45
0.53
0.60
0.68 45
0.78
0.95 40 (TMTSF)2PF6
100
1.15 P=0.9GPa μ H=4.91T//b’
T −1 (s−1)

35 0
Se T −1 (s−1)

30

25
0.8 T1T (s−K)
1

50
20
0.6
Θ∼11K
77

15 0.4

10 0.2
T (K)
0 0
0 5 10 15 20 25 30 5 0 10 20 30

temperature T (K) 0
0 5 10 15 20 25 30
77
Fig. 16. Se NMR spin lattice relaxation rate vs. temperature for (TMTSF)2PF6, for temperature T (K)
applied pressures P <> P c . T 1
1 tends to diverge on approaching T SDW ðPÞ. For higher
pressures the ground state is superconducting, coincident with large enhancements Fig. 17. Normal state 77Se T 1
1 vs. T in the regime at P = 0.9 GPa. The red line follows
1
of T 1 from spin fluctuations. a Curie–Weiss form, T 1 T  ðT þ HÞ. The data is replotted as T 1 T vs. T in the inset.
S.E. Brown / Physica C 514 (2015) 279–289 287

the case of the Bechgaard salts, however, there is no indication for 4.2. Normal state properties
pseudogap behavior, and significantly, the Curie–Weiss behavior
is uninterrupted to temperatures significantly less than T c [81], Following the discussion on the Bechgaard salts, it is natural to
whereas the expectation would be a crossover to constant T 1 T. To question whether overdamped AF spin fluctuations are providing
our knowledge, the temperature regime T  T c has not been exam- the pairing for the j-phase materials, as well as others. On that
ined in other systems such as the pnictides. basis, we wish to recount the evidence for spin fluctuations in
Coincident with the Curie–Weiss relaxation rate is a low-tem- the normal state. First, we note that in Ref. [118] is a model cal-
perature resistivity of the form qðTÞ ¼ q0 þ aT. In a magnetotrans- culation based on the fluctuation–exchange approximation which
port of several families of correlated superconductors including supports a superconducting state with T c  10 K, with 4 gap nodes
Bechgaard salts, cuprates, and pnictides, the magnitude of the over the FS, two each for the closed pocket and the q1d sheets. The
intrinsic linear-T contribution was correlated to the superconduct- gap symmetry is dx2 y2 . Incorporating the Mott physics requires a
ing transition temperature [97,98]. As for the relaxation rate, the strong-coupling approach; there, as well, d-wave superconductiv-
linear-T resistivity persists to T  T c . ity is found to emerge [119]. Specific heat experiments in an in-
These observations are provided some context by the results of plane magnetic field were carried out to probe the gap symmetry
renormalization group calculations [86,99]. Scattering amplitudes in j-NCS Xand j-Br, with results consistent with the gap nodes
from particle-hole and particle–particle channels are included; identified by the spin fluctuation model [120].
the effect of the pairing interaction on the SF is to suppress the In Fig. 18 appears the temperature dependence of the 13C relax-
low-temperature saturation of the correlation length. T-linear ation rate for j-NCS and j-Br [121]. For both, ½T 1 T1 increases upon
resistivity implied for temperatures greater than the SF energy cooling to about 50 K, in a manner consistent with the development
scale. However, the behavior is linked directly to the highly-nested
of AF spin correlations. For T < 50 K, ½T 1 T1 sharply but continu-
Fermi surface, proximity to a quantum critical point at Pc , and the
ously decreases. The change in behavior coincides with the metal–
interference between electron–electron and electron–hole scatter-
insulator crossover line, with Fermi Liquid behavior on the conduct-
ing. The suggestion is that the interference modifies the physical
ing side, depicted in Fig. 11. An important question then is how to
properties within some range of the SDW critical point, leading
evaluate the role of spin fluctuations at temperatures of the order
to the unexpected observations.
of the superconducting transition temperature T c . The decrease in
½T 1 T1 has been discussed as possibly one of a pseudogap phe-
4. Superconductivity and normal state properties of the q2d nomenon, associated with the onset of short-range spin fluctuations
superconductors j-(ET)2X and coherence of the intralayer transport [122,123]. Of course, the
notion of a pseudogap is attractive in that it is suggestive of related
4.1. Superconducting properties phenomena in systems such as the cuprates. However, there are
many reasons to question whether such a relationship exists.
The first ET superconductor identified was b-(ET)2ReO4, with Perhaps most clear is this: contrary to underdoped cuprates, there
T c  2 K [100]. The highest transition temperatures for any of the is no evidence for a broken symmetry and subsequent folding of
ET-based compounds are in the j-phase materials [22] [and refer- the Brillouin zone, with the exception of those effects associated
ences therein], with T c = 13.4 K, 11.5 K, and 10.5 K, with coun- with a pressure-dependent lattice superstructure that is reported
terions X = Cu[N(CN)2]Cl [101], Cu[N(CN)2]Br [102], and Cu(NCS)2 to develop at quite high temperatures in j-Br [124]. Specifically,
[103]. With regard to the pairing, 13C Knight shift measurements the quantum oscillations observed in magnetotransport experi-
in the superconducting state of j-Br show that the spin part of ments in j-NCS are attributed to expected sections of the Fermi sur-
the wave function is a singlet [104], and this carries over to j- face. The Korringa ratio K 2s T 1 T is shown to increase some marginally
(BEDT-TTF)2Cu(NCS)2 (j-NCS) as well [105]. Measurements of below 50 K [106,122]. Nevertheless, the enhancement relative to
the 13C spin lattice relaxation are consistent with the presence of high temperatures is still present, so spin fluctuations remain
nodal quasiparticles, in that ½T 1 T1 scales as T 3 in low fields and important in the normal state properties at T  T c .
T  T c , and moreover, there is no evidence for a Hebel–Slichter
enhancement of the relaxation rate for T ! T  c [104,106,107]. 4.3. High-field phases of the q2d superconductors
Both aspects are similar to what is observed for the cuprates. The
conclusion has carried with it considerable controversy. For exam- We divert the discussion here to comment briefly on the unu-
ple, different methods for determining the superfluid density, lþ SR sual results obtained in some organic superconductors in high
and microwave conductivity experiments, produced evidence for
both line nodes [108,109] and a full gap [110], respectively. 0.2
Specific heat experiments were interpreted differently as well; κ−NCS
these experiments are described in Ref. [111] (gap zeroes) and in κ−Br
Ref. [112] full gap. Recent measurements and interpretation of
0.15
T T] −1 (s−K) −1

superfluid density from microwave conductivity measurements


on j-NCS were found consistent with line nodes [113].
Regrettably, we have overlooked other important work dedicated
0.1
to this issue, and for that we apologize. Nevertheless, it would
appear that most of the evidence appears to support a nodal super-
1

conducting state, though notably missing from the discussion are


13

0.05
[

phase-sensitive experiments as were performed on the cuprates


[114,115]. We make a final note regarding an investigation of the
effects of defects on T c . With the Bechgaard salts [65,66] and with
Sr2RuO4 [116], T c is sharply suppressed by introducing defects or 0
0 50 100 150 200 250 300
impurities. In a similar investigation of j-NCS, the transition tem- temperature (K)
perature was found to decrease only marginally [117], and to our
knowledge the apparent discrepancy has not yet been resolved. Fig. 18. ½T 1 T1 vs. T for j-NCS and j-Br, after Ref. [121].
288 S.E. Brown / Physica C 514 (2015) 279–289

magnetic fields. A consequence of the very weak interplane Acknowledgements


coupling, to the degree that the interlayer Ginzburg–Landau
correlation length n? is considerably less than the separation of The author is grateful for many discussions and insights learned
the conducting planes for T  T c [125,126]. Therefore, for in-plane from colleagues and collaborators on the subject of organic
fields, the flux penetrates in the form of Josephson vortices. An superconductors. Here, special recognition is given to our friend,
important consequence is that in larger fields, where the inter- the late Professor James Brooks. Support from NSF under Grant
Josephson vortex separation falls below the Josephson penetration No. DMR-1410343 is also acknowledged.
depth, intralayer orbital currents vary as the inverse field. That is,
orbital energy shifts become inconsequential relative to Zeeman
References
shifts, making selected organic superconductors ideal cases for a
high-field transition to a high-field phase as discussed by Larkin [1] W. Little, Phys. Rev. 134 (1964) A1416–A1424.
and Ovchinnikov [79] and Fulde and Ferrell [78]. Such LOFF (or [2] F. Wudl, G.M. Smith, E.J. Hufnagel, Chem. Commun. 1 (1970) 1435.
[3] J. Ferraris, D.O. Cowan, V. Walatka, J.H. Perstein, J. Am. Chem. Soc. 95 (1973)
FFLO) states could be considered intermediate phases, between
948–949.
uniform superconducting and normal states, in which the Cooper [4] A.J. Berlinsky, J.F. Carolan, L. Weiler, Solid State Commun. 15 (5) (1974) 795–
pairs acquire momentum and associated spin response. In the 801.
Larkin/Ovchinnikov manifestation, the order parameter changes [5] S. Shitzkovsky, M. Weger, H. Gutfreund, J. Phys. 39 (1978) 711–717.
[6] T.J. Kistenmacher, T.E. Phillips, D.O. Cowan, Acta Cryst. B 30 (3) (1974) 763–
sign in real space. The best known system for LOFF physics is j- 768.
NCS, and by now the evidence is convincing. It has been long [7] R. Comès, G. Shirane, S.M. Shapiro, A.F. Garito, A.J. Heeger, Phys. Rev. B 14
appreciated that the system remains in the SC state well beyond (1976) 2376–2383.
[8] R.H. Friend, M. Miljak, D. Jérome, Phys. Rev. Lett. 40 (1978) 1048–1051.
the paramagnetic limiting field [112,127]. A transition to a high [9] D. Jerome, H.J. Shulz, Adv. Phys. 31 (1982) 299–490.
field superconducting phase (normal ? LOFF) has been seen [10] S. Megtert, R. Coms, C. Vettier, R. Pynn, A. Garito, Solid State Commun. 31 (12)
independently in the electrodynamic response [128] and torque (1979) 977–980.
[11] S. Khanna, J. Pouget, R. Comes, A. Garito, A. Heeger, Phys. Rev. B 16 (1977)
measurements [129]. NMR gives clear evidence for a continuous 1468–1479.
transition connected to an inhomogeneous spin polarization in a [12] D. Jérome, A. Mazaud, M. Ribault, K. Bechgaard, J. Phys. Lett. 41 (1980) L95–
bulk superconductor [105] as well as evidence in the relaxation L98.
[13] A. Andrieux, D. Jerome, K. Bechgaard, J. Phys. Lett. 42 (4) (1981) L87–L90.
rate for spin-polarized bound states at the gap zeroes [130]. This [14] T. Takahashi, Y. Maniwa, H. Kawamura, G. Saito, J. Phys. Soc. Jpn. 55 (4) (1986)
example is not unique: observations consistent with LOFF physics 1364–1373.
have been reported in k-(BETS)2FeCl4 [131], as well as the [15] A.G. Lebed (Ed.), The Physics of Organic Superconductors and Conductors,
Springer Series in Materials Science, vol. 110, Springer, Heidelberg, 2008.
all-organic superconductor b00 -(ET)2SF5CH2CF2SO3 [132]. [16] J.F. Kwak, J.E. Schirber, R.L. Greene, E.M. Engler, Phys. Rev. Lett. 46 (19) (1981)
1296–1299.
[17] P.M. Chaikin, M.Y. Choi, R.L. Greene, J. Mag. Mag. Mat. 31–34 (1983) 1268–
5. Concluding remarks 1272.
[18] L.P. Gor’kov, A.G. Lebed, J. Phys. Lett. 45 (1984) L433–L440.
[19] W. Kang, S.T. Hannahs, P.M. Chaikin, Phys. Rev. Lett. 70 (20) (1993) 3091–
In the j-phase superconductors, as for TMTSF, the preponder- 3094.
ance of evidence would support AF spin fluctuations as responsible [20] S. Uji, H. Shinagawa, T. Terashima, T. Yakabe, Y. Terai, M. Tokumoto, A.
Kobayashi, H. Tanaka, H. Kobayashi, Nature 410 (2001) 908–910.
for superconducting pairing and which leads to sign changes and [21] Y. Shimizu, K. Miyagawa, K. Kanoda, M. Maesato, G. Saito, Phys. Rev. Lett. 91
nodal structures in the gap function. This seems most clear for (2003) 107001.
the case of TMTSF, where the nuclear spin lattice relaxation rate [22] T. Ishiguro, K. Yamaji, G. Saito, in: Organic Superconductors, Springer Series in
Solid State Sciences, Springer Verlag, Berlin, Heidelberg, 1998.
is seen greatly enhanced prior to the transformation to the SC state,
[23] J. Singleton, C. Mielke, Contemp. Phys. 43 (2002) 63–96.
and close to the critical pressure the resistivity varies linearly in [24] P. Batail (Ed.), Molecular Conductors, Chemical Reviews, vol. 104, 2004
temperature. Thus, the 1979 discovery of superconductivity in (special issue).
the TMTSF salts, coinciding with the first observations of heavy fer- [25] R. Kato (Ed.), Molecular Conductors, Crystals, vol. 2, 2012 (special issue).
[26] N. Thorup, G. Rindorf, H. Soling, K. Bechgaard, Acta Cryst. B 37 (1981) 1236–
mion superconductivity, marked the beginning of an era in which 1240.
electronic correlations were recognized as an indispensable ingre- [27] P.M. Grant, Phys. Rev. B 26 (12) (1982) 6888–6895.
dient in many superconducting compounds. In q2d organics such [28] P.M. Grant, Phys. Rev. Lett. 50 (13) (1983) 1005–1008.
[29] A. Jacko, H. Feldner, E. Rose, F. Lissner, M. Dressel, R. Valentí, H. Jeschke, Phys.
as the j phases, magnetic fluctuations are important for T > T c , Rev. B 87 (2013) 155139.
though the anamolous drop in the Korringa constant below about [30] K. Yoshimi, H. Seo, S. Ishibashi, S. Brown, Phys. Rev. Lett. 108 (2012) 096402.
50 K remains an open question. In all, the case for the sign change [31] L. Ducasse, M. Abderrabba, J. Hoarau, M. Pesquer, B. Gallois, J. Gaultier, J. Phys.
C 19 (20) (1986) 3805.
of the gap is strong but perhaps not bullet-proof. Unfortunately, [32] M. Dressel, K. Petukhov, B. Salameh, P. Zornoza, T. Giamarchi, Phys. Rev. B 71
the deeper understanding that would come to both the (7) (2005) 075104.
Bechgaards and the j systems from inelastic neutron scattering [33] H. Wilhelm, D. Jaccard, R. Duprat, C. Bourbonnais, D. Jérome, J. Moser, C.
Carcel, J.M. Fabre, Eur. Phys. J. B 21 (2) (2001) 175–183.
measurements, let alone elastic scattering, is still missing. With
[34] G.M. Danner, W. Kang, P.M. Chaikin, Phys. Rev. Lett. 72 (23) (1994) 3714–
that backdrop, we emphasize that our goal here was not to be all 3717.
inclusive, even regarding the ET salts. Instead, we focused on those [35] S.E. Brown, M.J. Naughton, P.M. Chaikin, in: The Physics of Organic
superconductors and Conductors, Springer Series in Materials Science, 110,
two systems studied most thoroughly, and moving away from
Springer, Heidelberg, 2008. Ch. La tour des sels Bechgaard, pp. 49–87.
them may change the story. For example, there remains another [36] W. Yu, F. Zhang, F. Zamborszky, B. Alavi, A. Baur, C.A. Merlic, S.E. Brown, Phys.
proposed route to superconductivity, which nevertheless depends Rev. B 70 (12) (2004) 121101.
on correlations. In Ref. [61], it is proposed that collective charge [37] F. Iwase, K. Sugiura, K. Furukawa, T. Nakamura, Phys. Rev. B 84 (2011)
115140.
fluctuations related to nearby CO instabilities can induce pairing, [38] T. Adachi, E. Ojima, K. Kato, H. Kobayashi, T. Miyazaki, M. Tokumoto, A.
and several candidate systems are suggested among the b00 or h Kobayashi, J. Am. Chem. Soc. 122 (13) (2000) 3238–3239.
polytypes. (Note that this would ordinarily not apply to the j [39] D. Jaccard, H. Wilhelm, D. Jérome, J. Moser, C. Carcel, J.M. Fabre, J. Phys.:
Condens. Matter 13 (4) (2001) L89.
materials where CO instabilities are less likely.) Optics measure- [40] M. Itoi, C. Araki, M. Hedo, Y. Uwatoko, T. Nakamura, J. Phys. Soc. Jpn. 77 (2)
ments, both Raman and infrared, have resulted in some unusual (2008) 023701.
observations in the superconductor b00 -(ET)2SF5CH2CF2SO3, includ- [41] M. Itoi, M. Kano, N. Kurita, M. Hedo, Y. Uwatoko, T. Nakamura, J. Phys. Soc.
Jpn. 76 (5) (2007) 053703.
ing the temperature-dependent oscillator strength at low energies, [42] D.S. Chow, F. Zamborszky, B. Alavi, D.J. Tantillo, A. Baur, C.A. Merlic, S.E.
which are attributed to CO fluctuations [133]. Brown, Phys. Rev. Lett. 85 (8) (2000) 1698–1701.
S.E. Brown / Physica C 514 (2015) 279–289 289

[43] P. Monceau, F.Y. Nad, S. Brazovskii, Phys. Rev. Lett. 86 (18) (2001) 4080–4083. [97] N. Doiron-Leyraud, P. Auban-Senzier, S. René de Cotret, C. Bourbonnais, D.
[44] F. Zamborszky, W. Yu, W. Raas, S.E. Brown, B. Alavi, C.A. Merlic, A. Baur, Phys. Jérome, K. Bechgaard, L. Taillefer, Phys. Rev. B 80 (21) (2009) 214531.
Rev. B 66 (8) (2002) 081103. [98] N. Doiron-Leyraud, S. René de Cotret, A. Sedeki, C. Bourbonnais, L. Taillefer, P.
[45] H. Seo, H. Fukuyama, J. Phys. Soc. Jpn. 66 (1997) 1249–1252. Auban-Senzier, D. Jérome, K. Bechgaard, Europhys. J. B 78 (2010) 23–36.
[46] J. Riera, D. Poilblanc, Phys. Rev. B 63 (2001) 241102. [99] A. Sedeki, D. Bergeron, C. Bourbonnais, Physica B 405 (11, Supplement 1)
[47] P. Auban-Senzier, C.R. Pasquier, D. Jérome, S. Suh, S.E. Brown, C. Mézière, P. (2010) S89–S91;
Batail, Phys. Rev. Lett. 102 (2009) 257001. A. Sedeki, D. Bergeron, C. Bourbonnais, in: Proceeding of the 8th International
[48] L. Zorina, S. Simonov, C. Méziére, E. Canadell, S. Suh, S.E. Brown, P. Foury- Symposium on Crystalline Organic Metals, Superconductors and
Leylekian, P. Fertey, J.P. Pouget, P. Batail, J. Mater. Chem. 19 (2009) 6980– Ferromagnets; Yamada Conference LXIV – ISCOM, 2009.
6994. [100] S.S.P. Parkin, E.M. Engler, R.R. Schumaker, R. Lagier, V.Y. Lee, J.C. Scott, R.L.
[49] Y. Nakazawa, H. Taniguchi, A. Kawamoto, K. Kanoda, Phys. Rev. B 61 (2000) Greene, Phys. Rev. Lett. 50 (1983) 270–273.
R16295–R16298. [101] J.M. Williams, A.M. Kini, H.H. Wang, K.D. Carlson, U. Geiser, L.K. Montgomery,
[50] J. Wosnitza, J. Low Temp. Phys. 146 (2007) 641–667. G.J. Pyrka, D.M. Watkins, J.M. Kommers, S.J. Boryschuk, A.V.S. Crouch, W.K.
[51] K. Oshima, T. Mori, H. Inokuchi, H. Urayama, H. Yamochi, G. Saito, Phys. Rev. B Kwok, J.E. Schirber, D.L. Overmyer, D. Jung, M.-H. Whangbo, Inorg. Chem. 29
38 (1988) 938–941. (18) (1990) 3272–3274.
[52] H. Kino, H. Fukuyama, J. Phys. Soc. Jpn. 64 (8) (1995) 2726–2729. [102] A.M. Kini, U. Geiser, H.H. Wang, K.D. Carlson, J.M. Williams, W.K. Kwok, K.G.
[53] S. Lefebvre, P. Wzietek, S. Brown, C. Bourbonnais, D. Jérome, C. Mézière, M. Vandervoort, J.E. Thompson, D.L.A. Stupka, Inorg. Chem. 29 (14) (1990) 2555–
Fourmigué, P. Batail, Phys. Rev. Lett. 85 (25) (2000) 5420–5423. 2557.
[54] F. Kagawa, K. Miyagawa, K. Kanoda, Nature 436 (2005) 534–537. [103] H. Urayama, H. Yamochi, G. Saito, K. Nozawa, T. Sugano, M. Kinoshita, S. Sato,
[55] K. Kanoda, Physica C 282–287 (Part 1) (1997) 299–302. K. Oshima, A. Kawamoto, J. Tanaka, Chem. Lett. 17 (1) (1988) 55–58.
[56] B.J. Powell, R.H. McKenzie, J. Phys.: Condens. Matter 18 (45) (2006) R827. [104] H. Mayaffre, P. Wzietek, D. Jérome, C. Lenoir, P. Batail, Phys. Rev. Lett. 75 (22)
[57] Y. Kurosaki, Y. Shimizu, K. Miyagawa, K. Kanoda, G. Saito, Phys. Rev. Lett. 95 (1995) 4122–4125.
(2005) 177001. [105] J.A. Wright, E. Green, P. Kuhns, A. Reyes, J. Brooks, J. Schlueter, R. Kato, H.
[58] B.J. Powell, R.H. McKenzie, Rep. Prog. Phys. 74 (5) (2011) 056501. Yamamoto, M. Kobayashi, S.E. Brown, Phys. Rev. Lett. 107 (2011) 087002.
[59] T. Takahashi, Y. Nogami, K. Yakushi, J. Phys. Soc. Jpn. 75 (5) (2006) 051008. [106] S.M. De Soto, C.P. Slichter, A.M. Kini, H.H. Wang, U. Geiser, J.M. Williams,
[60] H. Seo, C. Hotta, H. Fukuyama, Chem. Rev. 104 (11) (2004) 5005–5036. Phys. Rev. B 52 (14) (1995) 10364–10368.
[61] J. Merino, R.H. McKenzie, Phys. Rev. Lett. 87 (2001) 237002. [107] K. Kanoda, K. Miyagawa, A. Kawamoto, Y. Nakazawa, Phys. Rev. B 54 (1)
[62] F. Steglich, J. Aarts, C.D. Bredl, W. Lieke, D. Meschede, W. Franz, H. Schäfer, (1996) 76–79.
Phys. Rev. Lett. 43 (25) (1979) 1892–1896. [108] L.P. Le, G.M. Luke, B.J. Sternlieb, W.D. Wu, Y.J. Uemura, J.H. Brewer, T.M.
[63] R. Brusetti, M. Ribault, D. Jérome, K. Bechgaard, Journal de Physique 43 (Paris) Riseman, C.E. Stronach, G. Saito, H. Yamochi, H.H. Wang, A.M. Kini, K.D.
(1982) 801. Carlson, J.M. Williams, Phys. Rev. Lett. 68 (1992) 1923–1926.
[64] R.L. Greene, E.M. Engler, Phys. Rev. Lett. 45 (19) (1980) 1587–1590. [109] D. Achkir, M. Poirier, C. Bourbonnais, G. Quirion, C. Lenoir, P. Batail, D. Jérome,
[65] M.Y. Choi, P.M. Chaikin, S.Z. Huang, P. Haen, E.M. Engler, R.L. Greene, Phys. Phys. Rev. B 47 (1993) 11595–11598.
Rev. B 25 (10) (1982) 6208–6217. [110] M. Dressel, O. Klein, G. Grüner, K.D. Carlson, H.H. Wang, J.M. Williams, Phys.
[66] N. Joo, P. Auban-Senzier, C.R. Pasquier, D. Jérome, K. Bechgaard, Europhys. Rev. B 50 (1994) 13603–13615.
Lett. 72 (2005) 645–651. [111] O.J. Taylor, A. Carrington, J.A. Schlueter, Phys. Rev. Lett. 99 (2007) 057001,
[67] A.A. Abrikosov, J. Low Temp. Phys. 53 (1983) 359–374. http://dx.doi.org/10.1103/PhysRevLett. 99.057001.
[68] C. Bourbonnais, D. Jérome, in: The Physics of Organic superconductors and [112] R. Lortz, Y. Wang, A. Demuer, P.H.M. Böttger, B. Bergk, G. Zwicknagl, Y.
Conductors, Springer Series in Materials Science, 110, Springer, Heidelberg, Nakazawa, J. Wosnitza, Phys. Rev. Lett. 99 (2007) 187002.
2008. Ch. Interacting Electrons in Quasi-One-Dimensional Organic [113] S. Milbradt, A.A. Bardin, C.J.S. Truncik, W.A. Huttema, A.C. Jacko, P.L. Burn, S.-
Superconductors, pp. 357–412. C. Lo, B.J. Powell, D.M. Broun, Phys. Rev. B 88 (2013) 064501.
[69] V.J. Emery, J. Phys. 44 (NC-3) (1983) 977–982. [114] D.A. Wollman, D.J. Van Harlingen, W.C. Lee, D.M. Ginsberg, A.J. Leggett, Phys. Rev.
[70] M. Takigawa, H. Yasuoka, G. Saito, J. Phys. Soc. Jpn. 56 (1) (1987) 873–876. Lett. 71 (1993) 2134–2137, http://dx.doi.org/10.1103/PhysRevLett.71.2134.
[71] L. Hebel, C. Slichter, Phys. Rev. 107 (1957) 901–902. [115] C.C. Tsuei, J.R. Kirtley, C.C. Chi, L.S. Yu-Jahnes, A. Gupta, T. Shaw, J.Z. Sun, M.B.
[72] Y. Hasegawa, H. Fukuyama, J. Phys. Soc. Jpn. 56 (1987) 877–880. Ketchen, Phys. Rev. Lett. 73 (4) (1994) 593–596.
[73] T. Imai, T. Shimizu, H. Yasuoka, Y. Ueda, K. Kosuge, J. Phys. Soc. Jpn. 57 (1988) [116] A.P. Mackenzie, R.K.W. Haselwimmer, A.W. Tyler, G.G. Lonzarich, Y. Mori, S.
2280–2283. Nishizaki, Y. Maeno, Phys. Rev. Lett. 80 (1998) 161–164.
[74] I.J. Lee, M.J. Naughton, G.M. Danner, P.M. Chaikin, Phys. Rev. Lett. 78 (18) [117] J.G. Analytis, A. Ardavan, S.J. Blundell, R.L. Owen, E.F. Garman, C. Jeynes, B.J.
(1997) 3555–3558. Powell, Phys. Rev. Lett. 96 (2006) 177002.
[75] J.I. Oh, M.J. Naughton, Phys. Rev. Lett. 92 (6) (2004) 67001. [118] J. Schmalian, Phys. Rev. Lett. 81 (1998) 4232–4235.
[76] A.G. Lebed, K. Yamaji, Phys. Rev. Lett. 80 (12) (1998) 2697–2700. [119] G. Sordi, P. Sémon, K. Haule, A.-M.S. Tremblay, Phys. Rev. Lett. 108 (2012)
[77] J. Vicent, S. Hillenius, R. Coleman, Phys. Rev. Lett. 44 (1980) 892–895. 216401.
[78] P. Fulde, R.A. Ferrell, Phys. Rev. 135 (3A) (1964) A550–A563. [120] L. Malone, O.J. Taylor, J.A. Schlueter, A. Carrington, Phys. Rev. B 82 (2010)
[79] A.I. Larkin, Y.N. Ovchinnikov, JETP 20 (3) (1965) 762–768. 014522.
[80] A.G. Lebed, Phys. Rev. B 59 (2) (1999) R721–R724. [121] A. Kawamoto, K. Miyagawa, Y. Nakazawa, K. Kanoda, Phys. Rev. Lett. 74
[81] J. Shinagawa, Y. Kurosaki, F. Zhang, C. Parker, S.E. Brown, D. Jérome, J.B. (1995) 3455–3458.
Christensen, K. Bechgaard, Phys. Rev. Lett. 98 (14) (2007) 147002. [122] E. Yusuf, B.J. Powell, R.H. McKenzie, Phys. Rev. B 75 (2007) 214515.
[82] B.J. Powell, J. Phys.: Condens. Matter 20 (34) (2008) 345234. [123] B.J. Powell, E. Yusuf, R.H. McKenzie, Phys. Rev. B 80 (2009) 054505.
[83] K. Yang, S.L. Sondhi, Phys. Rev. B 57 (14) (1998) 8566–8570. [124] H. Weiss, M.V. Kartsovnik, W. Biberacher, E. Steep, E. Balthes, A.G.M. Jansen,
[84] S. Yonezawa, Y. Maeno, K. Bechgaard, D. Jérome, Phys. Rev. B 85 (2012) K. Andres, N.D. Kushch, Phys. Rev. B 59 (1999) 12370–12378.
140502. [125] P.A. Mansky, G. Danner, P.M. Chaikin, Phys. Rev. B 52 (10) (1995) 7554–7563.
[85] S. Belin, K. Behnia, Phys. Rev. Lett. 79 (11) (1997) 2125–2128. [126] J.R. Kirtley, K.A. Moler, J.A. Schlueter, J.M. Williams, J. Phys.: Condens. Matter
[86] C. Bourbonnais, A. Sedeki, Phys. Rev. B 80 (8) (2009) 085105. 11 (8) (1999) 2007.
[87] H. Shimahara, J. Phys. Soc. Jpn. 69 (7) (2000) 1966–1969. [127] J. Singleton, J.A. Symington, M.-S. Nam, A. Ardavan, M. Kurmoo, P. Day, J.
[88] N. Belmechri, G. Abramovici, M. Héritier, Europhys. Lett. 82 (4) (2008) 47009. Phys.: Condens. Matter 12 (40) (2000) L641–L648.
[89] K. Kajiwara, M. Tsuchiizu, Y. Suzumura, C. Bourbonnais, J. Phys. Soc. Jpn. 78 [128] C.C. Agosta, J. Jin, W.A. Coniglio, B.E. Smith, K. Cho, I. Stroe, C. Martin, S.W.
(10) (2009) 104702. Tozer, T.P. Murphy, E.C. Palm, J.A. Schlueter, M. Kurmoo, Phys. Rev. B 85
[90] S. Yonezawa, S. Kusaba, Y. Maeno, P. Auban-Senzier, C. Pasquier, K. Bechgaard, (2012) 214514.
D. Jérome, Phys. Rev. Lett. 100 (11) (2008) 117002. [129] B. Bergk, A. Demuer, I. Sheikin, Y. Wang, J. Wosnitza, Y. Nakazawa, R. Lortz,
[91] Y. Fuseya, C. Bourbonnais, K. Miyake, Europhys. Lett. 100 (5) (2012) 57008. Phys. Rev. B 83 (2011) 064506.
[92] W. Wu, P.M. Chaikin, W. Kang, J. Shinagawa, W. Yu, S.E. Brown, Phys. Rev. [130] H. Mayaffre, S. Krämer, M. Horvatić, C. Berthier, K. Miyagawa, K. Kanoda, V.F.
Lett. 94 (9) (2005) 097004. Mitrović, Nat. Phys. 10 (2014) 928–932.
[93] A.J. Millis, H. Monien, D. Pines, Phys. Rev. B 42 (1) (1990) 167–178. [131] S. Uji, T. Terashima, M. Nishimura, Y. Takahide, T. Konoike, K. Enomoto, H.
[94] T. Moriya, Adv. Phys. 49 (2000) 555–606. Cui, H. Kobayashi, A. Kobayashi, H. Tanaka, M. Tokumoto, E.S. Choi, T.
[95] Y. Kitaoka, K. Ishida, G.Q. Zheng, S. Ohsugi, K. Asayama, J. Phys. Chem. Solids Tokumoto, D. Graf, J.S. Brooks, Phys. Rev. Lett. 97 (2006) 157001.
54 (10) (1993) 1385–1392. special Issue Spectroscopies in Novel [132] R. Beyer, B. Bergk, S. Yasin, J.A. Schlueter, J. Wosnitza, Phys. Rev. Lett. 109
Superconductors. (2012) 027003.
[96] F. Ning, K. Ahilan, T. Imai, A. Sefat, M. McGuire, B. Sales, D. Mandrus, P. Cheng, [133] A. Girlando, M. Masino, J.A. Schlueter, N. Drichko, S. Kaiser, M. Dressel, Phys.
B. Shen, H.-H. Wen, Phys. Rev. Lett. 104 (2010) 037001. Rev. B 89 (2014) 174503.

You might also like