Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors (Review Article)
We review the current understanding of superconductivity in the quasi-one-dimensional organic conductors of the Bechgaard and Fabre salt families. We discuss the interplay between superconductivity, antiferromagnetism, and charge-density-wave fluctuations. The connection to recent experimental ob...
Saved in:
| Date: | 2006 |
|---|---|
| Main Authors: | , , |
| Format: | Article |
| Language: | English |
| Published: |
Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України
2006
|
| Series: | Физика низких температур |
| Subjects: | |
| Online Access: | https://nasplib.isofts.kiev.ua/handle/123456789/120192 |
| Tags: |
Add Tag
No Tags, Be the first to tag this record!
|
| Journal Title: | Digital Library of Periodicals of National Academy of Sciences of Ukraine |
| Cite this: | Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors (Review Article) / N. Dupuis, C. Bourbonnais, J.C. Nickel // Физика низких температур. — 2006. — Т. 32, № 4-5. — С. 505–520. — Бібліогр.: 140 назв. — англ. |
Institution
Digital Library of Periodicals of National Academy of Sciences of Ukraine| id |
nasplib_isofts_kiev_ua-123456789-120192 |
|---|---|
| record_format |
dspace |
| spelling |
nasplib_isofts_kiev_ua-123456789-1201922025-02-09T16:32:06Z Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors (Review Article) Dupuis, N. Bourbonnais, C. Nickel, J.C. Spin Models We review the current understanding of superconductivity in the quasi-one-dimensional organic conductors of the Bechgaard and Fabre salt families. We discuss the interplay between superconductivity, antiferromagnetism, and charge-density-wave fluctuations. The connection to recent experimental observations supporting unconventional pairing and the possibility of a triplet–spin order parameter for the superconducting phase is also presented. 2006 Article Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors (Review Article) / N. Dupuis, C. Bourbonnais, J.C. Nickel // Физика низких температур. — 2006. — Т. 32, № 4-5. — С. 505–520. — Бібліогр.: 140 назв. — англ. 0132-6414 PACS: 74.70.Kn, 74.20.Mn, 75.30.Fv https://nasplib.isofts.kiev.ua/handle/123456789/120192 en Физика низких температур application/pdf Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України |
| institution |
Digital Library of Periodicals of National Academy of Sciences of Ukraine |
| collection |
DSpace DC |
| language |
English |
| topic |
Spin Models Spin Models |
| spellingShingle |
Spin Models Spin Models Dupuis, N. Bourbonnais, C. Nickel, J.C. Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors (Review Article) Физика низких температур |
| description |
We review the current understanding of superconductivity in the quasi-one-dimensional organic
conductors of the Bechgaard and Fabre salt families. We discuss the interplay between superconductivity,
antiferromagnetism, and charge-density-wave fluctuations. The connection to recent
experimental observations supporting unconventional pairing and the possibility of a
triplet–spin order parameter for the superconducting phase is also presented. |
| format |
Article |
| author |
Dupuis, N. Bourbonnais, C. Nickel, J.C. |
| author_facet |
Dupuis, N. Bourbonnais, C. Nickel, J.C. |
| author_sort |
Dupuis, N. |
| title |
Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors (Review Article) |
| title_short |
Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors (Review Article) |
| title_full |
Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors (Review Article) |
| title_fullStr |
Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors (Review Article) |
| title_full_unstemmed |
Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors (Review Article) |
| title_sort |
superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors (review article) |
| publisher |
Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України |
| publishDate |
2006 |
| topic_facet |
Spin Models |
| url |
https://nasplib.isofts.kiev.ua/handle/123456789/120192 |
| citation_txt |
Superconductivity and antiferromagnetism in
quasi-one-dimensional organic conductors
(Review Article) / N. Dupuis, C. Bourbonnais, J.C. Nickel // Физика низких температур. — 2006. — Т. 32, № 4-5. — С. 505–520. — Бібліогр.: 140 назв. — англ. |
| series |
Физика низких температур |
| work_keys_str_mv |
AT dupuisn superconductivityandantiferromagnetisminquasionedimensionalorganicconductorsreviewarticle AT bourbonnaisc superconductivityandantiferromagnetisminquasionedimensionalorganicconductorsreviewarticle AT nickeljc superconductivityandantiferromagnetisminquasionedimensionalorganicconductorsreviewarticle |
| first_indexed |
2025-11-27T23:55:35Z |
| last_indexed |
2025-11-27T23:55:35Z |
| _version_ |
1849989780714029056 |
| fulltext |
Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5, p. 505–520
Superconductivity and antiferromagnetism in
quasi-one-dimensional organic conductors
(Review Article)
N. Dupuis1,2, C. Bourbonnais3, and J.C. Nickel2,3
1Department of Mathematics, Imperial College, 180 Queen’s Gate, London SW7 2AZ
United Kingdom
E-mail: n.dupuis@imperial.ac.uk
2Laboratoire de Physique des Solides, CNRS UMR 8502, Universit� Paris-Sud, 91405 Orsay, France
3Regroupement Qu�becois sur les Mat�riaux de Pointe, Universit� de Sherbrooke
Sherbrooke, Qu�bec, Canada J1K-2R1
Received September 15, 2005
We review the current understanding of superconductivity in the quasi-one-dimensional or-
ganic conductors of the Bechgaard and Fabre salt families. We discuss the interplay between su-
perconductivity, antiferromagnetism, and charge-density-wave fluctuations. The connection to re-
cent experimental observations supporting unconventional pairing and the possibility of a
triplet–spin order parameter for the superconducting phase is also presented.
PACS: 74.70.Kn, 74.20.Mn, 75.30.Fv
Keywords: superconductivity, antiferromagnetism, unconventional pairing.
1. Introduction
Superconductivity in organic conductors was first
discovered in the ion radical salt (TMTSF)2PF6 [1].
Later on, it was found in most Bechgaard
[(TMTSF)2X] and Fabre [(TMTTF)2 X] salts. These
salts are based on the organic molecules tetra-
methyltetraselenafulvalene (TMTSF) and tetra-
methyltetrathiafulvalene (TMTTF). The monovalent
anion X can be either a centrosymmetric (PF6, AsF6,
etc.) or a non-centrosymmetric (ClO4, ReO4, NO3,
FSO3, SCN, etc.) inorganic molecule. (See Refs. 2,3
for previous reviews on these compounds.) Although
they are definitely not «high-Tc» superconductors —
the transition temperature is of the order of 1 K –,
these quasi-one-dimensional (quasi-1D) conductors
share several properties of high-Tc superconductors
and other strongly-correlated electron systems such as
layered organic superconductors [4,5] or heavy-fer-
mion materials [6]. The metallic phase of all these
conductors exhibits unusual properties which cannot
be explained within the framework of Landau’s Fermi
liquid theory and remain to a large extent to be under-
stood. The superconducting phase is unconventional
(not s-wave). Magnetism is ubiquitous in these corre-
lated systems and might provide the key to the under-
standing of their behavior.
The quest for superconductivity in organic conduc-
tors was originally motivated by Little’s proposal that
highly polarizable molecules could lead – via an
excitonic pairing mechanism – to tremendously large
transition temperatures. Early efforts towards the
chemical synthesis of such compounds were not suc-
cessful, as far as superconductivity is concerned, but
led to the realization of a 1D charge transfer salt
(TTF–TCNQ) undergoing a Peierls instability at low
temperatures [7]. Attempts to suppress the Peierls
state and stabilize a conducting (and possibly super-
conducting) state by increasing the 3D character of
this 1D conductor proved to be unsuccessful.
Organic superconductivity was eventually discove-
red in the the Bechgaard salt (TMTSF)2PF6 under
9 kbar of pressure [1]. It was subsequently found in
© N. Dupuis, C. Bourbonnais, and J.C. Nickel, 2006
other members of the (TMTSF)2X series. Most of the
Bechgaard salts are insulating at ambient pressure and
low temperatures [8], and it came as a surprise that
the insulating state of these materials is a spin-den-
sity-wave (SDW) rather than an ordinary Peierls
state [2]. The important part played by magnetism in
these compounds was further revealed when it was
found that their phase diagram only shows a part of a
larger sequence of ordered states, which includes the
N�el and the spin-Peierls phases of their sulfur ana-
logs, the Fabre salts (TMTTF)2X series [9].
The charge transfer from the organic molecules to
the anions leads to a commensurate band filling 3/4
coming from the 2:1 stoichiometry. The metallic char-
acter of these compounds at high enough temperature
is due to the delocalization of carriers via the overlap
of �-orbitals between neighboring molecules along the
stacking direction (a axis) (Fig. 1) [2]. The electronic
dispersion relation obtained from quantum chemistry
calculations (extended H�ckel method) is well ap-
proximated by the following tight-binding form
[10–13]
�( ) cos( ) ( ) ( )k � � � �� �2 2 2 2t k a/ t k b t k ca a b b c ccos cos �
� v k k t k b t k bF a F b b b b(| | ) cos( ) cos( )� � � � �� �2 2
� ��2t k cc ccos( ) ,�
(1)
where it is assumed that the underlying lattice is
orthorhombic. This expression is a simplification of
the dispersion relation — the actual crystal lattice
symmetry is triclinic — but it retains the essential
features. The conduction band along the chain direc-
tion has an overall width 4ta ranging between
0.4 and 1.2 eV, depending on the organic molecule
(TMTSF or TMTTF) and the anion. As the electronic
overlaps in the transverse b and c directions are much
weaker than along the organic stacks, the dispersion
law is strongly anisotropic, t /tb a� � 01. and
t /tc b� � � 0 03. , and the Fermi surface consists of two
open warped sheets (Fig. 1). In the second line of
Eq. (1), the electronic dispersion is linearized around
the two 1D Fermi points �kF , with vF the Fermi ve-
locity along the chains (� is the chemical potential).
The next-nearest-chain hopping t t /tb b a� � �� �
2 2
� is
introduced in order to keep the shape of the Fermi
surface unchanged despite the linearization. The an-
ions located in centrosymmetric cavities lie slightly
above or below the molecular planes. This structure
leads to a dimerization of the organic stacks and a
(weak) gap D, thus making the hole-like band ef-
fectively half-filled at sufficiently low energy or tem-
perature [14,15]. (See Refs. 2,7,9 for a detailed dis-
cussion of the structural properties of quasi-1D
organic conductors.) In the presence of interactions,
commensurate band-filling introduces Umklapp scat-
tering, which affects the nature of the possible phases
in these materials.
What is remarkable about these electronic systems
is the variety of ground states that can be achieved ei-
ther by chemical means, namely substituting selenium
by sulfur in the organic molecule or changing the na-
ture of the anion (its size or symmetry), or applying
pressure (Fig. 2). At low pressure, members of the sul-
fur series are Mott insulators (MI) from which either
a lattice distorted spin-Peierls (SP) state — often
preceded by a charge ordered (CO) state — or a com-
mensurate-localized antiferromagnetic state (AF) can
develop. On the other hand, itinerant antiferro-
magnetism (spin-density wave (SDW)) or supercon-
ductivity is found in the selenide series. Under pres-
sure, the properties of the Fabre salts evolve towards
those of the Bechgaard salts. The compound
(TMTTF)2PF6 spans the entire phase diagram as pres-
sure increases up to 50 kbar or so (Fig. 3) [16–18],
thus showing the universality of the phase diagram in
Fig. 2 [19].
A large number of both theoretical and experimen-
tal works have been devoted to the understanding of
the normal phase and the mechanisms leading to
long-range order at low temperature. The presence of
antiferromagnetism over a large pressure range does
506 Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5
N. Dupuis, C. Bourbonnais, and J.C. Nickel
–9.8
–10.0
–10.2
WII
Wb
'
X ZY
X
Y
V
E
,e
V
WI
Wb
�
Wc
Wc�
Wa
a
b
Fig. 1. A side view of the Bechgaard/Fabre salt crystal
structure with the electron orbitals of the organic stacks
(courtesy of J.Ch. Ricquier) (a). Electronic dispersion re-
lation and projected 2D Fermi surface of (TMTTF)2Br
(reprinted with permission from Ref. 13. Copyright 1994
by EDP Sciences) (b).
indicate that repulsive interactions among carriers are
important. The low-dimensionality of the system is
also expected to play a crucial role. On the one hand,
in the presence of repulsive interactions a strongly
anisotropic Fermi surface with good nesting properties
is predominantly unstable against the formation of an
SDW state which is reinforced at low temperature by
commensurate band filling. On the other hand, when
the temperature exceeds the transverse dispersion
�t b, 3D (or 2D) coherent electronic motion is sup-
pressed and the conductor behaves as if it were 1D;
the Fermi liquid picture breaks down and the system
becomes a Luttinger liquid [20,21]. The relevance of
1D physics for the low-temperature properties
(T t b� � ), as well as a detailed description of the cross-
over from the Luttinger liquid to the Fermi liquid, is
one of the most important issues in the debate sur-
rounding the theoretical description of the normal
state of these materials. As far as low-temperature
phases are concerned, a chief objective is to reach a
good description of the superconducting phase — the
symmetry of the order parameter is still under debate
— and the mechanisms leading to superconductivity.
Owing to the close proximity of superconductivity
and magnetism in the phase diagram of Fig. 2, it is es-
sential to first discuss the origin of antiferromag-
netism in both series of compounds.
2. N�el antiferromagnetism and
spin-density wave
2.1. Fabre salts at ambient pressure:
Mott-insulator regime
The Fabre salts (TMTTF)2X at ambient pressure
are located on the left of the phase diagram in Fig. 2.
Both the nature of correlations and the mechanism of
long-range order at low temperature are now rather
well understood. Below the temperature T� � 100 K
(see Fig. 2), the resistivity develops a thermally acti-
vated behavior [22] and the system enters a Mott-insu-
lator regime. The corresponding charge gap � ��� T
can be deduced from T� and turns out to be larger than
the (bare) transverse bandwidth t b� , which in turn
suppresses any possibility of transverse single particle
band motion and makes the system essentially one-di-
mensional. The charge gap 2 � is also directly ob-
served in the optical conductivity [23]. The members
of the (TMTTF)2X series thus behave as typical 1D
Mott insulators below T� with the carriers confined
along the organic stacks — as a result of the Umklapp
scattering due to the commensurability of the elec-
tronic density with the underlying lattice [14,15].
This interpretation agrees with the absence of anomaly
in the spin susceptibility at T� [24], in accordance
with the spin-charge separation characteristic of 1D
systems [21]. It is further confirmed by measurements
of the spin-lattice relaxation rate 1 1/T . The Luttinger
liquid theory predicts [25,26]
1
1
0
2
1T
C T T C Ts
K
� �� �( ) , (2)
where C0 and C1 are temperature independent con-
stants. � s T( ) is the uniform susceptibility and K� the
Luttinger liquid charge stiffness parameter. The two
contributions in (2) correspond to paramagnons or
spinons (q � 0) and AF spin fluctuations (q kF� 2 ).
Both the temperature dependence of � s T( ) and the
presence of AF fluctuations lead to an enhancement of
1 1/T with respect to the Korringa law ( )T T1
1� � const
which holds in higher-dimensional metals. In a
1D Mott insulator K� � 0, which leads to
T C T T Cs1
1
0
2
1
� � �� ( ) in good agreement with experi-
mental measurements of T1 and � s [24].
The low-energy excitations in the Mott-insulator
regime are 1D spin fluctuations. By lowering the tem-
perature, these fluctuations can propagate in the
transverse direction and eventually drive an AF tran-
sition. This transition is not connected to Fermi sur-
face effects. The condition � � �t b precludes a sin-
gle-particle coherent motion in the transverse
direction, and the concept of Fermi surface remains ill
defined in the Fabre salts at ambient pressure. AF
Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors
Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5 507
T, K
100
10
1
10 20 30 50 P, kbar
T� LL
LL
FL
2X
Pm
MI-CO
SP
SC
Pc
(TMTTF)
AF
2X(TMTSF)
X= P
F 6
P
F 6
B
r
C
IO
4
Fig. 2. The generic phase diagram of the Bechgaard/Fabre
salts as a function of pressure or anion X substitution;
Luttinger liquid (LL), Mott insulator (MI), charge order
(CO), spin-peierls (SP), antiferromagnetism (AF), super-
conductivity (SC), Fermi liquid (FL).
long-range order comes from interchain transfer of
bound electron-hole pairs leading to a kinetic exchange
interaction J� between spin densities on neighboring
chains — much in analogy with the exchange interac-
tion between localized spins in the Heisenberg limit.
An effective Hamiltonian can be derived from a
renormalization group (RG) calculation, [27,28]
H J dx x x J
a
t
i
i j
j
b
� � �
�
�
�
� �S S
,
( ) ( ), �
��
�
2
(3)
where t b�
* is the effective interchain hopping at the
energy scale � and a the lattice spacing along the
chain. The sum in Eq. (3) is over nearest-neighbor
chains. The naive value t /b�
*2 � of the exchange inter-
action J� is enhanced by the factor ��/a where
�� �� v /F is the intrachain coherence length in-
duced by the Mott gap along which virtual interchain
hoppings can take place. Within a mean-field treat-
ment of H� , the condition for the onset of long-range
order is given by J k TF� ��1D( , )2 1 where
� �1D( , ) ( )2 1k T T/F
� is the exact power law form
of the 1D AF spin susceptibility. This yields a N�el
temperature
T
t
N
b� �
�2
�
. (4)
Since T� and � decrease under pressure (Fig. 2), Eq.
(4) predicts an increase of TN with pressure — as-
suming a weak pressure dependence of tb
� — as
observed experimentally (see Fig. 2). The relation
T T tN b� � ��
�2 const has been observed in
(TMTTF)2Br [29].
2.2. Bechgaard salts: itinerant magnetism
With increasing pressure, T� drops and finally
merges with the AF transition line at Pm , beyond
which there is no sign of a Mott gap in the normal
phase. The Fabre salts then tend to behave similarly to
the Bechgaard salts (Fig. 2). The change of behavior
at Pm is usually attributed to a deconfinement of carri-
ers, i.e., a crossover from a Mott insulator to a — me-
tallic — Luttinger liquid. At lower temperature, sin-
gle-particle transverse hopping is expected to become
relevant and induces a dimensional crossover at a tem-
perature Tx from the Luttinger liquid to a 2D or 3D
metallic state. With increasing pressure, the AF tran-
sition becomes predominantly driven by the instabil-
ity of the whole warped Fermi surface due to the nest-
ing mechanism. Although there is a general agreement
on this scenario, there is considerable debate on how
the dimensional crossover takes place and the nature
of the low-temperature metallic state.
On the theoretical side, simple RG arguments indi-
cate that the crossover from the Luttinger liquid to
the 2D regime takes place at the temperature [30]
T
t t
tx
b b
a
K
K� � �
�
�
�
��
�
�
���
�
�
1
, (5)
where K� is the Luttinger liquid parameter. For
non-interacting electrons (K� � 1), Eq. (5) would give
T tx b� � : the 2D Fermi surface is irrelevant when tem-
perature is larger than the dispersion in the b direction.
For interacting electrons (K� � 1), the interchain hop-
ping amplitude t b� is reduced to an effective value t b�
*
and the dimensional crossover occurs at a lower tem-
perature T t tx b b� �
�
�� . A detailed theoretical picture
of the dimensional crossover is still lacking. In particu-
lar, whether it is a sharp crossover or rather extends
over a wide temperature range — as shown by the
shaded area in Fig. 2 — is still an open issue.
2.2.1. The strong-correlation picture
Some experiments seem to indicate that correla-
tions still play an important role even in the low-tem-
perature phase of the Bechgaard salts. For instance, a
significant enhancement of 1 1/T T with respect to the
Korringa law — although weaker than in the Fabre
salts at ambient pressure — is still present [24]. This
behavior has been explained in terms of 1D spin fluc-
tuations persisting down to the dimensional crossover
temperature Tx � 10 K, below which the Korringa law
is recovered [24,26].
The restoration of a plasma edge in the transverse b�
direction at low temperature in (TMTSF)2PF6 — ab-
sent in the Fabre salts — suggests the gradual emer-
gence of a coherent motion in the ( )ab planes below
Tx � 100 K [31,32]. (b� is normal to a and c in the ( )ab
plane. It differs from b due to the triclinic structure.)
However, the frequency dependence of the optical
conductivity is inconsistent with a Drude-like metal-
lic state [33,34,23]. The low-energy peak carries only
1% of the total spectral weight and is too narrow to be
interpreted as a Drude peak with a frequency-inde-
pendent scattering time. It has been proposed that this
peak is due to a collective mode that bears some simi-
larities with the sliding of a charge-density wave — an
interpretation supported by the new phonon features
that emerge at low temperature [33]. Furthermore,
99% of the total spectral weight is found in a finite en-
ergy peak around 200 cm �1. It has been suggested that
this peak is a remnant of a (1 4/ )-filled Mott gap � ,
observed in the less metallic Fabre salts at ambient
pressure [35,23]. In this picture, (TMTSF)2PF6 is
close to the border between a Mott insulator and a
Luttinger liquid, and the low-temperature metallic
508 Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5
N. Dupuis, C. Bourbonnais, and J.C. Nickel
behavior is made possible by the interchain coupling
[23,36,37]. A different interpretation has been pro-
posed for the far infrared spectrum in optical conduc-
tivity and is based on the weak half-filling character
of the band for interactions in the Hubbard limit [38].
The longitudinal resistivity in (TMTSF)2PF6 is
found to be metallic, with a T2 law between the SDW
transition and 150 K, crossing over to a sublinear tem-
perature dependence above 150 K with an exponent in
the range 0.5–1 [39,40]. While this observation would
be consistent with a dimensional crossover to a
low-temperature Fermi liquid regime taking place at
Tx � 150 K, the transverse resistance �b along the b
axis apparently fails to show the expected T2 behav-
ior. Given the difficulties of a direct dc measurement,
owing to non-uniform current distributions between
contacts, conflicting results have been published in
the literature [39,41]. Nevertheless, below T � 80 K
�b can be deduced from �a T� 2 and �c T� 15. using a
tunneling argument, which yields � � �c a b
/� ( )1 2 and
therefore �b T� . Moreover, contactless — microwave
— transverse conductivity measurements in the
(TMTSF)2PF6 salt fail to reveal the emergence of a
Fermi liquid T2 temperature dependence of the resis-
tivity in the b direction in this temperature range [42].
As far as �c is concerned, a maximum around
Tmax � 80 K has been observed, with a metallic —
though incoherent — behavior �c T� 15. at lower tem-
perature [43]. Tmax is highly sensitive to pressure,
whereas the interchain hopping t b� is not. Therefore,
Tmax cannot be directly identified with t b� , but could
be related to a — weakly — renormalized value
t Tb x� �* in agreement with predictions of the
Luttinger liquid theory [see Eq. (5)]. The transport
measurements seem to be indicative of a gradual cross-
over between a Luttinger liquid and a Fermi liquid oc-
curring in the temperature range 40–80 K. The onset
of 3D coherence and Fermi liquid behavior would then
be related to the interplane coupling t c� between ( , )a b
planes [43].
The absence of Fermi liquid behavior down to very
low temperatures in the Bechgaard salts seems to be
further supported by photoemission experiments.
ARPES fails to detect quasi-particle features or the
trace of a Fermi surface at 150 K [44]. Similar conclu-
sions were deduced from integrated photoemission at
50 K [45]. However, photoemission results — e.g. the
absence of dispersing structure and a power-law fre-
quency dependence which is spread over a large en-
ergy scale of the order of 1 eV — do not conform with
the predictions of the Luttinger theory and might be
strongly influenced by surface effects.
The existence of strong correlations suggests that
the kinetic interchain exchange J� , which drives the
AF transition in the sulfur series, still plays an impor-
tant role in the Bechgaard salts. In this picture, the
decrease of TN with increasing pressure is due both to
the decrease of J� and the deterioration of the Fermi
surface nesting. This scenario is supported by RG cal-
culations [28].
All the experiments mentioned so far favor diffe-
rent — and sometimes incompatible — scenarios for
the dimensional crossover. However, the high-temper-
ature phase of the Bechgaard salts is always analyzed
on the basis of the Luttinger liquid theory. A consis-
tent interpretation of the experimental results there-
fore requires to find a common K� parameter and to
determine the value of the remnant of the Mott gap
� . NMR [24], dc transport [43,46], and optical mea-
surements [23,36] have been interpreted in terms of
the Luttinger theory with K� � 0 23. and quar-
ter-filled Umklapp scattering [7,37]. This interpreta-
tion, as well as the mere existence of strong correla-
tions, is not without raising a number of unanswered
questions (see the next section). For instance,
K� � 0 23. would lead according to (5) to
T tx b� �
�10 3 , a value much below the experimental
observations.
2.2.2. The weak-correlation picture
On the other hand, there are experiments pointing
to the absence of strong correlations in the Bechgaard
salts. One of the most convincing arguments comes
from the so-called Danner–Chaikin oscillations [47].
Resistance measurements of (TMTSF)2ClO4 in the c
direction show pronounced resonances when an ap-
plied magnetic field is rotated in the ( )ac plane at low
temperature. The complete angular dependence of the
magneto-resistance can be reproduced within a
semiclassical approach. The position of the resonance
peaks is given by the zeros of the Bessel function J0( )�
evaluated at � � �2t cB /v Bb x F z (c is the interchain
spacing in the c direction). This enables a direct mea-
sure of the interchain hopping amplitude in the b di-
rection, yielding t b� � 280 K above the anion order-
ing transition taking place at 24 K, in very good
agreement with values derived from band calculations
[10–12]. These results can hardly be reconciled with
the existence of strong correlations. Sizeable 1D fluc-
tuations should lead to a strong (k w| | , ) dependence of
the self-energy, and in turn to a significant renorma-
lization of k� -dependent quantities like the interchain
hopping amplitudes [28]. This lends support to the
idea that the low-temperature phase of the Bechgaard
salts can be described as a weakly interacting Fermi
liquid subject to spin fluctuations induced by the nest-
ing of the Fermi surface [48,49].
Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors
Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5 509
The weak-coupling approach has been particularly
successful in the framework of the Quantized Nesting
Model [50–52]. The latter explains the cascade
of SDW phases induced by a magnetic field in
(TMTSF)2PF6 and (TMTSF)2ClO4, and provides a
natural explanation for the quantization of the Hall
effect — � xy Ne /h� 2 2 (N integer) per ( )ab plane —
observed in these phases. Furthermore, it reproduces
the experimental phase diagram only for interchain
hopping amplitudes t tb c� �, close to their unre-
normalized values.
Despite the apparent success of the weak-coupling
approach, it has nevertheless become clear that the
SDW phase of the Bechgaard salts is not conven-
tional. Recent experiments have shown that the 2kF
SDW coexists with a 2kF and a — weaker — 4kF
charge-density wave (CDW) in (TMTSF)2PF6
[53,54]. Since there is no 2kF phonon softening associ-
ated to this transition, the emergence of this CDW
state differs from what is usually seen for an ordinary
Peierls state. This unusual ground-state can be ex-
plained on the basis of a quarter-filled 1D model with
dimerization and onsite, nearest-neighbor and
next-nearest-neighbor Coulomb interactions [55–59],
but this explanation remains to be confirmed.
2.2.3. The normal phase above the superconducting
phase
It is remarkable that the superconducting phase lies
next to the SDW phase — which is actually a mixture
SDW–CDW — and reaches its maximum transition
temperature Tc � 1 K at the pressure Pc where TSDW
and Tc join (see Figs. 2 and 3). In the normal phase
above the SDW phase, the resistivity along the a axis
decreases with temperature, reaches a minimum at
Tmin, and then shows an upturn and a strong
enhancement related to the proximity of the SDW
phase transition that occurs at T TSDW � min. The re-
gion of the normal phase where strong AF fluctuations
are present (T T TSDW � � min) extends over the pres-
sure range where the ground state is superconducting
(Fig. 3). Its width in temperature decreases with in-
creasing pressure, so that the superconducting transi-
tion temperature appears to be closely linked to Tmin.
These observations strongly suggest an intimate rela-
tionship between spin fluctuations and superconduc-
tivity in the Bechgaard/Fabre salts [16,17]. The
importance of spin fluctuations above the supercon-
ducting phase is further confirmed by the persistence
of the enhancement of the spin-lattice relaxation rate
1 1/T for P Pc� [24]. Besides the presence of spin fluc-
tuations at low temperature, charge fluctuations have
also been observed in the normal phase via optical
conductivity measurements [33].
3. Superconductivity
Some of the early experiments in the Bechgaard
salts were not in contradiction with a conventional
BCS superconducting state. For instance, the specific
heat in (TMTSF)2ClO4 obeys the standard tempera-
ture dependence C/T T� �� � 2 above the supercon-
ducting transition, and the jump at the transition
C/ Tc� � 167. is close to the BCS value 1.43. The
ratio 2 0 3 33 ( ) .T /Tc� � , obtained from the gap de-
duced from the thermodynamical critical field, is also
in reasonable agreement with the prediction of the
BCS theory (2 3 52 /Tc � . ) [60,61]. Early measure-
ments of H Tc2( ), performed in the vicinity of the
zero-field transition temperature, were also interp-
reted on the basis of the BCS theory [62–65].
Nevertheless, soon after the discovery of organic su-
perconductivity, the high-sensitivity of the supercon-
ducting state to irradiation [66,67] led Abrikosov [68]
to suggest the possibility of an unconventional — trip-
let — pairing, although the non-magnetic nature of
the induced defects is questionable [7]. The sensitivity
to non-magnetic impurities, and thus the existence of
unconventional pairing, was later on clearly estab-
lished by the suppression of the superconducting tran-
sition upon alloying (TMTSF)2ClO4 with a very
small concentration of ReO4 anions [69,70]. A recent
study [71] of the alloy (TMTSF)2(ClO4)x(ReO4)1–x
— with different cooling rates and different values of
x — has confirmed this in remarkable way by showing
510 Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5
N. Dupuis, C. Bourbonnais, and J.C. Nickel
100
10
1
0 2 4 6 8
(TMTTF) PF2 6
M-HI
SP
M
AF
SDW
SC
T, K
P, GPa
Fig. 3. (P,T) phase diagram of (TMTTF)2PF6. The shaded
area above the SDW and SC phase indicates the region of
the normal phase where spin fluctuations are significant.
(Reprinted with permission from Ref. 17. Copyright 2001
by EDP Sciences.)
that the transition temperature Tc is related to the
scattering rate 1/� by
ln
T
T T
c
c c
0 1
2
1
4
1
2
�
�
��
�
�
�� � �
�
�
��
�
�
�� � �
�
�
�
�
�� �
�� (6)
(Tc0 is the transition temperature of the pure system
and � the digamma function), as expected for an
unconventional superconductor in the presence of
non-magnetic impurities [72].
Another indication of a possible unconventional
pairing came from the observation of Gor’kov and
J�rome [73] that the upper critical field H Tc2( ), ex-
trapolated down to T � 0, would exceed the Pauli lim-
ited field [74,75] H T /P c B� �184 20. � T by a factor
of 2. (The value of HP quoted here corresponds to
s-wave pairing.) As spin-orbit interaction is weak in
these systems and cannot provide an explanation for
such a large Hc2, it is tempting to again invoke triplet
pairing. This issue has been revived by recent measure-
ments of the upper critical field in (TMTSF)2PF6
with substantially improved accuracy in angular
alignment and lower temperatures. Lee et al. [76,77]
observed a pronounced upward curvature of H Tc2( )
without saturation — down to T T /c� 60 — for a
field parallel to the a or b� axis, with H Tc
b
2
� ( ) and
H Tc
a
2( ) exceeding the Pauli limited field HP by a fac-
tor of 4. Moreover, H Tc
b
2
� ( ) becomes larger than
H Tc
a
2( ) at low temperatures. Similar results were ob-
tained from simultaneous resistivity and torque mag-
netization experiments in (TMTSF)2ClO4 (Fig. 4)
[78]. The extrapolated value to zero temperature,
Hc2 0 5( ) � T, is at least twice the Pauli limited field.
There are different mechanisms that can greatly in-
crease the orbital critical field H Tc2
orb ( ) in organic con-
ductors. Superconductivity in a weakly-coupled plane
system can survive in a strong parallel magnetic field
if the interplane (zero-field) coherence length �� ( )T
becomes smaller than the interplane spacing d at low
temperature. Vortex cores, with size �� ( )T d� , can
then fit between planes without destroying the super-
conducting order in the planes, and lead to a
Josephson vortex lattice. In the Bechgaard salts, even
for a field parallel to the b� axis, the Josephson limit
�� ( )T d� is however unlikely to be reached, since the
interchain hopping amplitude t c� � 5–10 K is larger
than the transition temperature Tc � 11. K. Neverthe-
less the orbital critical field can be enhanced by a
field-induced dimensional crossover [79–83]. A mag-
netic field parallel to the b� axis tends to localize the
wavefunctions in the ( )ac planes, which in turn weak-
ens the orbital destruction of the superconducting or-
der. When �c ceHc t� �� (which corresponds to a
field of a few Tesla in the Bechgaard salts), the wave
functions are essentially confined in the ( )ac planes
and the orbital effect of the field is completely sup-
pressed. The coexistence between SDW and supercon-
ductivity, as observed in a narrow pressure domain of
the order of 0.8 kbar below the critical pressure Pc
(Fig. 2), can also lead to a large increase of the orbital
upper critical field [84–88].
Regardless of the origin of the large orbital critical
field, another mechanism is required to exceed the
Pauli limited field HP in the Bechgaard salts. For sin-
glet spin pairing, the Pauli limit may be overcome by
a non-uniform Larkin–Ovchinnikov–Fulde–Ferrell
(LOFF) state, where Cooper pairs form with a non-
zero total momentum [89,90]. This mechanism is par-
ticularly efficient in a 1D system [79,81,83,91], due to
the large phase space available for pairing at nonzero
total momentum. For a linearized dispersion law, the
mean-field upper critical field Hc
LOFF diverges as 1/T
in a pure superconductor. Lebed [92] has argued that
the quasi-1D anisotropy reduces Hc
LOFF below the ex-
perimental observations. The only possible explana-
tion for a large upper critical field would then be an
equal-spin triplet pairing. A px -wave triplet state
with a d vector perpendicular to the b� axis was pro-
posed [93] as a possible explanation of the experimen-
tal observations reported in Refs. 76,77.
The triplet scenario in the Bechgaard salts is sup-
ported by recent NMR Knight shift experiments
(Fig. 5) [94,95]. Early NMR experiments by Takiga-
wa et al. already pointed to the unconventional nature
of the superconducting state in (TMTSF)2ClO4 [96].
Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors
Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5 511
T= 25 mK
H // b�
0.3
0.2
0.1
0
0 2 4 6 8 10
�0H, T
2
0
–2
–4
–6
T
o
rq
u
e,
1
0
N
m
–
9
R
e
si
st
iv
ity
,
cm
� zz
Fig. 4. Resistivity (left scale) and torque magnetization
(right) in (TMTSF)2ClO4 at 25 mK for H b| | �. The dotted
line and + symbols on the torque curve represent a tem-
perature-independent normal state contribution. The onsets
of diamagnetism and decreasing resistivity, upon decreasing
field, are indicated by the arrow near Hc2 � 5 T. Arrows
in the low field vortex state indicate field sweep direc-
tions. (Reprinted with permission from Ref. 78. Copyright
2004 by the American Physical Society.)
The proton spin lattice relaxation rate 1 1/T does not
exhibit a Hebel–Slichter peak. It decreases rapidly
just below Tc in contrast to the typical BCS supercon-
ductor where it increases below Tc, reaching a maxi-
mum at T Tc� 0 9. . Furthermore, 1 1
3/T T� for
T / T Tc c2� � — as it is the case for most unconven-
tional superconductors — suggesting zeros or lines of
zeros in the excitation spectrum. Recent experiments
by Lee et al. in (TMTSF)2PF6 show that the Knight
shift, and therefore the spin susceptibility, remains
unchanged at the superconducting transition [94,95].
This indicates triplet spin pairing, since a singlet pair-
ing would inevitably lead to a strong reduction of the
spin susceptibility ( ( ) )� T � �0 0 . It should however
be noticed that the interpretation of the Knight shift
results — due to a possible lack of sample therma-
lization during the time of the experiment — has been
questioned [7,97].
In principle, the symmetry of the order parameter
can be determined from tunneling spectroscopy. Sign
changes of the pairing potential around the Fermi sur-
face lead to zero-energy bound states in the supercon-
ducting gap [98]. These states manifest themselves as
a zero-bias peak in the tunneling conductance into the
corresponding edge [99]. More generally, different
pairing symmetries can be unambiguously distin-
guished by tunneling spectroscopy in a magnetic field
[100–102]. In practice however, the realization of tun-
nel junctions with the TMTSF salts appears to be very
difficult. A large zero-bias conductance peak — sug-
gesting p-wave symmetry — across the junction be-
tween two organic superconductors was observed
[103]. But the absence of temperature broadening
could indicate that this peak is due to disorder rather
than to a midgap state [104].
Information about the symmetry of the order param-
eter can also be obtained from thermal conductivity
measurements. The latter indicate the absence of nodes
in the excitation spectrum of the superconducting state
in (TMTSF)2ClO4 [105], thus suggesting a px -wave
symmetry. However, because of the doubling of the
Fermi surface in the presence of anion ordering, a sin-
glet d- or triplet f-wave order phase would also be
nodeless in (TMTSF)2ClO4 (see Fig. 6 for the different
gap symmetries in a quasi-1D superconductor) [9,106].
4. Microscopic theories of the superconducting
state
The phase diagram of the 1D electron gas within
the g-ology framework [108] is shown in Fig. 7. g1
and g2 denote the backward and forward scattering
amplitudes, respectively, and g3 the strength of the
(half-filling) Umklapp processes. Given the impor-
tance of spin fluctuations in the phase diagram of the
Bechgaard/Fabre salts, as well as the existence of AF
512 Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5
N. Dupuis, C. Bourbonnais, and J.C. Nickel
8
7
6
5
4
3
2
1
0
7
7 S
e
sp
e
ct
ra
,a
rb
. u
n
its
–6000 –3000 0 60003000
Shift from < >� normal
a
bc
d
1.70 K
1.40 K
0.90 K
0.83 K
0.76 K
0.56 K
0.51 K
0.40 K
0.32 K
0.80 K
0.50 K
0.65 K
0.35 K
0.20 K
0.09 K
T/T (H)c
1.0
0.5
/
n
0 0.3 0.6 0.9 1.2
Fig. 5. 77Se NMR spectra Tc (0.81 K at 1.43 T). Each trace
is normalized and offset for clarity. The temperatures
shown in parentheses are the measured equilibrium tem-
peratures before the pulse. In the inset, the spin suscepti-
bility normalized by the normal state � �/ n from measured
first moments are compared with theoretical calculations
[98] for H/Hc2 0 0( ) � (a) and 0.63 (b). Curves c and d
are obtained from the ratio of applied field (1.43 T) to the
measured upper critical field Hc2( )T at which the super-
conducting criteria «onset» and «50 transition» have been
used, respectively, to determine Hc2( )T . (Reprinted with
permission from Ref. 94. Copyright 2002 by the American
Physical Society.)
S
in
g
le
t
T
ri
p
le
t
+
_
_
_
+
_
+
_
+_
+
s
p x p y
d
f x
+
+
+
+
+ _
__
_
+
s d dxy
+
x – y2 2
Fig. 6. Gap symmetries �r k( )� in a quasi-1D superconduc-
tor (after Ref. 107, courtesy of Y. Suzumura). r = + /–
denotes the right/left sheet of the Fermi surface. (Singlet
pairing) s: const, d
x y2 2
�
: cos k� , dxy: r ksin � . (Triplet
pairing) px: r, f: r kcos � , py: sin k� . Next-nearest-neigh-
bor and longer-range pairings are not considered.
ground states, the Bechgaard/Fabre salts should per-
tain to the upper right corner of the 1D phase diagram
(g g1 2 0, � and g g g1 2 32� � | |) where the Umklapp
processes are relevant and the dominant fluctuations
antiferromagnetic. In the Fabre salts, the non-mag-
netic insulating phase observed below T� � 100 K
indicates the importance of Umklapp scattering and
suggests sizable values of g3 for this series. Since the
long-range Coulomb interaction favors g g1 2� , the
Fabre salts are expected to lie to the right of the phase
diagram, i.e. far away from the boundary
g g g1 2 32� � | |. Since the triplet superconducting phase
is lying next to the SDW phase (Fig. 7), it is tempting
to invoke a change of the couplings gi under pressure
to argue in favor of a px -wave triplet superconducting
state [68,109]. Such a drastic change of the couplings,
which would explain why (TMTTF)2PF6 becomes
superconducting above 4.35 GPa [16–18], is however
somewhat unrealistic and has not received any theo-
retical backing so far. The Umklapp scattering being
much weaker in the Bechgaard salts, one cannot ex-
clude that these compounds lie closer to the boundary
between the SDW and the triplet superconducting
phase. A moderate change of the couplings under pres-
sure would then be sufficient to explain the supercon-
ducting phase of (TMTSF)2PF6 observed above
6 kbar or so. However, the destruction of the super-
conducting phase by a weak magnetic field and the ob-
servation of a cascade of SDW phases for slightly
higher fields [50–52] would imply that the interaction
is strongly magnetic-field dependent — again a very
unlikely scenario.
In all probability, the very origin of the supercon-
ducting instability lies in the 3D behavior of these
quasi-1D conductors. Thus the attractive interaction
is a consequence of a low-energy mechanism that be-
comes more effective below the dimensional crossover
temperature Tx . Transverse hopping makes retarded
electron-phonon interactions more effective, since it is
easier for the electrons to avoid the Coulomb repul-
sion [109]. By comparing the sulfur and selenide
series, it can however be argued that, in the pressure
range where superconductivity is observed, the
strength of the electron–phonon interaction is too
weak to explain the origin of the attractive interac-
tion. For narrow tight-binding bands in the organics,
the attraction is strongest for backscattering processes
in which 2kF phonons are exchanged [110,111]. Ac-
cording to the results of x-ray experiments performed
on (TMTSF)2X, however, the electron–phonon ver-
tex at this wave vector does not undergo any signifi-
cant increase in the normal state (Fig. 8). The ampli-
tude of the 2kF lattice susceptibility in (TMTSF)2PF6
— which is directly involved in the strength of the
phonon exchange — is weak. It is instructive to com-
pare with the sulfur analog compound (TMTTF)2PF6,
for which the electron-phonon vertex at 2kF becomes
singular, signaling a lattice instability towards a
spin-Peierls distortion (Fig. 8). This instability pro-
duces a spin gap that is clearly visible in the tempera-
ture dependence of the magnetic susceptibility and nu-
clear relaxation rate [112,113]. These effects are not
seen in (TMTSF)2PF6 close to Pc. The persistent en-
hancement of these quantities indicates that interac-
tions are dominantly repulsive (Sec. 2.2.3.), making
the traditional phonon-mediated source of pairing
inoperant.
Emery [114] pointed out that near an SDW insta-
bility, short-range AF spin fluctuations can give rise
to anisotropic pairing and thus provide a possible
Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors
Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5 513
0
SDWTS
CDWSS
g
g
g – 2g =|g |
1 2 31
2
Fig. 7. Phase diagram (leading fluctuations) of the 1D
electron gas in presence of Umklapp scattering. SS (TS):
singlet (triplet) superconductivity. A gap develops in the
charge sector (Mott insulating behavior) for g g g1 2 32� � | | .
T, K
0
(TMTTF) 2PF6
(TMTSF)2PF6
50
0
100 150 200
I/
T
,a
rb
.u
n
its
Fig. 8. Temperature dependence of the 2kF lattice suscep-
tibility (I/T) as a function of temperature in the normal
phase of (TMTSF)2PF6 (top) and (TMTTF)2PF6 (bot-
tom). (Reprinted with permission from Ref. 53. Copyright
1996 by EDP Sciences.)
explanation of the origin of the superconducting phase
in the Bechgaard salts. Such fluctuations give rise to
an oscillating potential that couples to the electrons.
Carriers can avoid the local Coulomb repulsion and
take advantage of the attractive part of this potential
by moving on different chains. This mechanism, which
can lead to superconductivity at low temperatures,
is the spin-analog of the so-called Kohn–Luttinger
mechanism which assumes the pairing to originate
in the exchange of charge-density excitations pro-
duced by Friedel oscillations [115]. While most theo-
retical results on the spin-fluctuation-induced super-
conductivity are based on RPA-like calculations
[116,117,27,118–124], the existence of such an elec-
tronic pairing mechanism in a quasi-1D conductor has
been recently confirmed by an RG approach [125].
Moreover, it has been recently realized that CDW fluc-
tuations can play an important role in stabilizing a trip-
let phase [126–128,107,129–131]. Below we discuss in
simple terms the link between spin/charge fluctuations
and unconventional pairing [124], and present recent re-
sults obtained from an RG approach [129–131].
4.1. Superconductivity from spin and charge
fluctuations
Considering for the time being only intrachain in-
teractions, the interacting part of the Hamiltonian
within the g-ology framework [108] reads
H g g
q
int ch sp� � � �
�[ ( ) ( ) ( ) ( )]� �q q S q S q (7)
(from now on we neglect the c axis and consider a
2D-model), where �q and Sq are the charge- and
spin-density operators in the Peierls channel
( )q kx F� 2 , g g g /ch � �1 2 2 and g g /sp � � 2 2.
Starting from a half-filled extended Hubbard model,
we obtain g U V1 2� � and g U V2 2� � , where U is
the onsite and V the nearest-neighbor lattice site
(dimer) interaction. For simplicity, we do not con-
sider Umklapp scattering (g3), since it does not play
an important role in the present qualitative discus-
sion. For repulsive interactions g g1 2 0� � ,
short-range spin fluctuations develop at low tempera-
tures due to the nesting of the Fermi surface. They
can be described by an effective Hamiltonian Hint
eff ob-
tained from (7) by replacing the bare coupling con-
stants by their (static) RPA values
g
g
g
g gRPA RPA
ch
ch
ch
ch ch ch( )
( )
( ),q
q
q�
�
� �
1 0
2
�
�
g
g
g
g gRPA RPA
sp
sp
sp
sp sp sp( )
( )
( ),q
q
q�
�
� �
1 0
2
�
� (8)
where �RPA is the static (� � 0) RPA susceptibility.
The bare particle-hole susceptibility diverges at low
temperatures, i.e. �0 0( ) ( ( , ))Q � ��ln maxE / T t b , due
to the Q � ( , )2kF � nesting of the quasi-1D Fermi sur-
face ( � �k k Q� � ��� ). (E0 is a high-energy cut-
off of the order of the bandwidth.) The divergence is
cut off by deviations from perfect nesting, character-
ized by the energy scale ��t b [Eq. (1)]. In the
Bechgaard salts �� �t b 10 K and varies with pressure.
When the nesting of the Fermi surface is good
(small ��t b), the spin susceptibility � sp
RPA( )Q diverges
at low temperatures, thus signaling the formation of
an SDW. A larger value of ��t b frustrates antiferro-
magnetism and, when exceeding a threshold value,
eliminates the transition to the SDW phase [132,133].
In that case, the (remaining) short-range spin fluctua-
tions can lead to pairing between fermions. To see
this, we rewrite the effective Hamiltonian Hint
eff in the
particle-particle (Cooper) channel
H g O Os s sint
eff � � � �
�
��[ ( , ) ( ) ( )
k,k
k k k k
� �
��g O Ot t t( , ) ( ) ( )]k k k k (9)
(we consider only Cooper pairs with zero total mo-
mentum), where
g g gs
RPA RPA( , ) ( ) ( ),k k k k k k� � � � � � � �3 sp ch
g g gt
RPA RPA( , ) ( ) ( )k k k k k k� � � � � � � �sp ch (10)
are the effective interactions in the singlet and triplet
spin pairing channels (Fig. 9). Os ( )k (O kt ( )), is the
annihilation operator of a pair ( , )k k� in a singlet
(triplet) spin state, and Ot t t tO O O� �( , , )1 0 1 denotes
the three components Sz � �1 0 1, , of the triplet state
(total spin S � 1).
On the basis of the effective Hamiltonian (9) the
BCS theory predicts a superconducting transition
whenever the effective interaction g s t, turns out to be
attractive in (at least) one pairing channel. A simple
argument shows that this is indeed always the case in
the presence of short-range spin fluctuations. The spin
susceptibility � sp
RPA( )k k� � exhibits a pronounced
peak around k k Q� � � . Neglecting the unimportant k| |
dependence, its Fourier series expansion reads
� sp
RPA
F n
n
nk k k a n k k( , ) ( ) cos[ ( )]2 1
0
� �
�
�
� �� � � � � � ��
= a nk nk nk nkn
n
n
�
�
� � � �� � � � �
0
1( ) [cos cos sin sin ],
(11)
514 Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5
N. Dupuis, C. Bourbonnais, and J.C. Nickel
where an ! 0. Choosing a an � 0, one obtains a diver-
ging spin susceptibility � sp
RPA
Fk k k( , )2 � �� � �
� � � �� �" �( )k k . The condition a a0 1 0� � !... gives
a broadened peak around k k� �� � �. Eqs. (10, 11)
show that the effective interaction in the singlet
channel contains attractive interactions for any value
of n. In real space, n corresponds to the range of the
pairing interaction in the b direction. The dominant
attractive interaction corresponds to nearest-neigh-
bor-chain pairing (n � 1) and a d
x y2 2
�
-wave order pa-
rameter r k k( ) cos� �� (r kx� sgn( )). The interac-
tion is also attractive in the triplet f-wave channel
( ( ) cos ) r k r k� �� . However, all the three
components of a (spin-one boson) SDW fluctuation
contribute to the superconducting coupling in the sin-
glet channel — hence the factor of 3 in the first of
equations (10). The latter therefore always dominates
over the triplet one when charge fluctuations are not
important. Note that the interaction is repulsive in
the singlet dxy -wave ( ( ) sin ) r k r k� �� and the trip-
let py -wave (sink� ) channels.
Equations (10) show that CDW fluctuations tend
to suppress the singlet pairing, but reinforce the trip-
let one. In the Bechgaard salts, the physical relevance
of CDW fluctuations has been borne out by the puz-
zling observation of a CDW that coexists with the
SDW (Sec. 2.2.) [33,53,54]. Within the framework of
an extended anisotropic Hubbard model, recent RPA
calculations have shown that the triplet f-wave pair-
ing can overcome the singlet d
x y2 2
�
-wave pairing
when the intrachain interactions are chosen such as to
boost the CDW fluctuations with respect to the SDW
ones [126–128]. In a half-filled model, this however
requires the nearest-neighbor (intrachain) interaction
V to exceed U/2. In a quarter-filled model — appro-
priate if one ignores the weak dimerization along the
chains — the condition for f-wave superconductivity
becomes V U/2 2! — V2 is the next-nearest-neighbor
(intrachain) Coulomb interaction — and appears even
more unrealistic. Similar conclusions were reached
within an RG approach [107].
Given that electrons interact through the Coulomb
interaction, not only intrachain but also interchain in-
teractions are present in practice. At large momentum
transfer, the interchain interaction is well known to
favor a CDW ordered state [134–137]. This mecha-
nism is mostly responsible for CDW long-range order
observed in several organic and inorganic low-dimen-
sional solids (e.g. TTF-TCNQ) [138,139]. In the
Bechgaard salts, both the interchain Coulomb interac-
tion and the kinetic interchain coupling (t b� ) are
likely to be important in the temperature range where
superconductivity and SDW instability occur, and
should be considered on equal footing. An RG ap-
proach has recently been used to determine the phase
diagram of an extended quasi-1D electron gas model
that includes interchain hopping, nesting deviations
and both intrachain and interchain interactions
[129–131]. The intrachain interactions turn out to
have a sizeable impact on the structure of the phase di-
agram. Unexpectedly, for reasonably small values of
the interchain interactions, the singlet dx y2 2
�
-wave
superconducting phase is destabilized to the benefit of
the triplet f-wave phase with a similar range of Tc.
The SDW phase is also found to be close in stability to
a CDW phase. Before presenting these results in more
detail (Sec. 4.2.), let us discuss in simple terms the
role of interchain interactions. The interchain back-
ward scattering amplitude g1
� (� 0) contributes to the
effective interaction in the Cooper channel,
g k k g k ks s( , ) ( , )� � � �� � � �
� � � ��
� � � �2 1g k k k k[cos cos sin sin ],
g k k g k kt t( , ) ( , )� � � �� � � �
� � � � ��
� � � �2 1g k k k k[ cos cos sin sin ]. (12)
It thus tends to suppress singlet d
x y2 2
�
pairing, but
favors triplet f-wave pairing. In addition to this «di-
rect» contribution, g1
� reinforces CDW fluctuations,
g q g q g qch ch( ) ( ) cos� �
�
�� � 2 1 (13)
and therefore enhances the f-wave pairing over the
d
x y2 2
�
-wave pairing via the mechanism of fluctuation
exchange [see Eqs. (10)]. As for the interchain for-
ward scattering g2
� , its direct contribution to the DW
channel is negligible, but it has a detrimental effect
on both singlet and triplet nearest-neighbor-chain
pairings. This latter effect, which is neutralized
by the Umklapp scattering processes, can lead to
next-nearest-neighbor-chain pairings when Umklapp
processes are very weak [131].
Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors
Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5 515
g =s,t + + + …
Fig. 9. Diagrammatic representation of the effective inter-
action gs t, in the Cooper channel within the RPA.
4.2. RG calculation of the phase diagram
of quasi-1D conductors
As a systematic and unbiased method with no a pri-
ori assumption, the RG method is perfectly suited to
study competing instabilities. The zero-temperature
phase diagram obtained with this technique is shown
in Fig. 10 [130,131]. In the absence of interchain
interactions (g g1 2 0� �� � ), it confirms the validity of
the qualitative arguments given above. When the
nesting of the Fermi surface is nearly perfect (small
��t ) the ground state is an SDW. Above a threshold
value of ��t , the low-temperature SDW instability
is suppressed and the ground state becomes a
dx y2 2
�
-wave superconducting (SCd) state with an or-
der parameter r k k( ) cos� �� [125]. In the presence
of interchain interactions (g1 0� � ), the region of sta-
bility of the SCd phase shrinks, and a triplet supercon-
ducting f-wave (SCf) phase appears next to the
d-wave phase for ~ .g g / vF1 1 01� �� � � — obtained here
for typical values of intrachain couplings and band pa-
rameters [130,131]. For larger values of the interchain
interactions, the SCd phase disappears and the region
of stability of the f-wave superconducting phase wid-
ens. In addition a CDW phase appears, thus giving
the sequence of phase transitions SDW�CDW�SCf
as a function of ��t . For ~ .g1 012�� , the SDW phase dis-
appears. Note that for ~ .g1 011�
� , the region of stabil-
ity of the CDW phase is very narrow, and there is es-
sentially a direct transition between the SDW and
SCf phases.
The RG calculations yield Tc � 30 K for the SDW
phase in the case of perfect nesting and Tc � 0.6–1.2 K
for the superconducting phase, in reasonable agree-
ment with the experimental observations in the
Bechgaard salts. Fig. 11 shows the transition tempera-
ture Tc as a function of ��t for three different values of
the interchain interactions, ~g1 0� � , 0.11 and 0.14,
corresponding to the three different sequences of
phase transitions as a function of ��t : SDW�SCd,
SDW�(CDW)�SCf and CD�WSCf . The phase di-
agram is unchanged when both g2
� and a weak Umklapp
scattering amplitude g3 are included [130,131].
The RG approach also provides important informa-
tion about the fluctuations in the normal phase. The
dominant fluctuations above the SCd phase are SDW
fluctuations as observed experimentally (Sec. 2.2.).
Although they saturate below T t� �� where the SCd
fluctuations become more and more important, the
latter dominate only in a very narrow temperature
range above the superconducting transition (Fig. 12).
Above the SCf and CDW phases, one expects strong
CDW fluctuations driven by g1
� . Fig. 13 shows that
for ~g1
� � 0.11–0.12, strong SDW and CDW fluctua-
tions coexist above the SCf phase. Remarkably, there
are regions of the phase diagram where the SDW fluc-
tuations remain the dominant ones in the normal
phase above the SCf or CDW phase (Fig. 13,b).
A central result of the RG calculation is the close
proximity of SDW, CDW and SCf phases in the
phase diagram of a quasi-1D conductor within a real-
istic range of values for the repulsive intrachain and
interchain interactions. Although this proximity is
found only in a small range of interchain interactions,
there are several features that suggest that this part of
the phase diagram is the relevant one for the
Bechgaard salts. i) SDW fluctuations remain impor-
tant in the normal phase throughout the whole phase
diagram. They are the dominant fluctuations above
the SCd phase, and remain strong — being sometimes
even dominant — above the SCf phase where they co-
exist with strong CDW fluctuations, in accordance
with observations [24,33]. ii) The SCf and CDW
phases stand nearby in the theoretical phase diagram,
the CDW phase always closely following the SCf
phase when the interchain interactions increase. This
agrees with the experimental finding that both SDW
516 Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5
N. Dupuis, C. Bourbonnais, and J.C. Nickel
0
0.04
0.08
0.12
0 0.04 0.08 0.12 0.16 0.2
t /t
g
1
~ SDW
CDW
SCd
SCf
�
� �
�
Fig. 10. T = 0 phase diagram as a function of t t�� �/ and
~g1
� . Circles: SDW, squares: CDW, triangles: SCd
( ( ) cos )�r k k� �� , crosses: SCf (�r k r k( ) cos� �� ). The
dashed lines indicate two (among many) possible pressure
axes, corresponding to transitions SDW�SCd and
SDW�SCf [130,131].
0.0001
0.001
0.01
0.1
0 0.04 0.08 0.12 0.16 0.2
t /t�
� �
T /tc �
Fig. 11. Transition temperature as a function of �� �t /t for
~g1 0� � , 0.11 and 0.14, corresponding to solid, dotted, and
dashed lines, respectively [130,131].
and CDW coexist in the DW phase of the Bechgaard
salts [53,54] and the existence, besides SDW correla-
tions, of CDW fluctuations in the normal state above
the superconducting phase [33]. iii) Depending how
one moves in practice in the phase diagram as a func-
tion of pressure, these results are compatible with ei-
ther a singlet dx y2 2
�
-wave or a triplet f-wave super-
conducting phase in the Bechgaard salts (see the two
pressure axes in Fig. 10). Moreover, one cannot ex-
clude that both SCd and SCf phases exist in these ma-
terials, with the sequence of phase transitions
SDW�SCd�SCf as a function of pressure. It is also
possible that the SCf phase is stabilized by a magnetic
field [140], since an equal-spin pairing triplet phase is
not sensitive to the Pauli pair breaking effect contrary
to the SCd phase. This would make possible the exis-
tence of large upper critical fields exceeding the Pauli
limit [76,78], and would also provide an explanation
for the temperature independence of the NMR Knight
shift in the superconducting phase [94].
5. Conclusion
Notwithstanding the recent experimental pro-
gresses, many of the basic questions related to super-
conductivity in the Bechgaard and Fabre salts remain
largely open. The very nature of the superconducting
state — the orbital symmetry of the order parameter
and the singlet/triplet character of the pairing — is
still not known without ambiguity even though recent
upper critical field measurements [76–78] and NMR
experiments [94,95] support a triplet pairing.
We argued that the conventional electron–phonon
mechanism is unable to explain the origin of the super-
conducting phase. On the other hand, the proximity
of the SDW phase, as well as the observation of strong
spin fluctuations in the normal state precursor to the
superconducting phase [16,17,24], strongly suggest an
intimate relationship between antiferromagnetism and
superconductivity in the Bechgaard/Fabre salts. The
scenario originally proposed by Emery [114], whereby
short-range AF spin fluctuations can give rise to
anisotropic pairing and thus stabilize a superconduct-
ing phase, is so far the only one that is consistent with
the experimental observations and the repulsive na-
ture of the electron-electron interactions.
Within the framework of the RG approach, it has
recently been shown that when spin and charge fluctu-
ations are taken into account on equal footing, both
singlet dx y2 2
�
- and triplet f-wave superconducting
phases can emerge at low temperatures whenever the
nesting properties of the Fermi surface deteriorate un-
der pressure [126–128,107,129–131]. CDW fluctua-
tions are enhanced by the long-range part of the Cou-
lomb interaction. Remarkably, for a reasonably small
value of the interchain interactions, the singlet
dx y2 2
� -wave phase is destabilized to the benefit of a
triplet f-wave with a similar range of Tc [130,131].
The physical relevance of CDW fluctuations in the
Bechgaard salts has been born out by the observation of
a CDW that actually coexists with the SDW [53,54].
CDW fluctuations were also observed in the normal
state precursor to the superconducting state [33].
As a systematic and unbiased method with no a pri-
ori assumptions, the RG has proven to be a method of
choice to study the physical properties of quasi-1D or-
ganic conductors. An important theoretical issue is
Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors
Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5 517
0.1
1
10
100
1000
0.001 0.01 0.1 1 10 100
T/t
�
�
v F
0.0001
Fig. 12. Temperature dependence of the susceptibilities in
the normal phase above the SCd phase ( � �� �t t0152. and
~ .g1 008� � ). The continuous, dotted, dashed, and
dashed-dotted lines correspond to SDW, CDW, SCd and
SCf correlations, respectively [130,131].
0.1
1
10
100
1000
0.001 0.01 0.1 1 10 100
0.1
1
10
100
1000
0.001 0.01 0.1 1 10 100
T/t
�
�
v F
T/t
�
�
v F
a
b
0.0001
Fig. 13. Temperature dependence of the susceptibilities in
the normal phase above the SCf phase for ~ .g1 012� � ,
� �� �t t0152. (a) and � �� �t t0176. (b).
now to go beyond the instabilities of the normal state.
On the one hand, the RG analysis should be extended
to the low-temperature broken-symmetry phases in or-
der to study the possible coexistence of superconduc-
tivity and antiferromagnetism, as well as CDW and
SDW, as observed in the Bechgaard salts
[86,88,53,54]. On the other hand, the RG technique
might also enable to tackle the unusual properties of
the metallic phase. A recent RG analysis [107] of the
AF spin susceptibility in the normal phase has shown
that below the dimensional crossover temperature, it
differs significantly from the prediction of sin-
gle-channel (RPA) theories. The interplay between
the superconducting and Peierls channels, which is at
the origin of spin-fluctuation induced superconductiv-
ity, might also be responsible for the unusual proper-
ties of the metallic state below the dimensional cross-
over temperature.
1. D. J�rome, A. Mazaud, M. Ribault, and K. Bechgaard,
J. Phys. Lett. (Paris) 41, L95 (1980).
2. D. J�rome and H.J. Schulz, Adv. Phys. 31, 299 (1982).
3. T. Ishiguro and K. Yamaji, Organic Superconductors,
of Springer-Verlag Series in Solid-State Science,
Springer-Verlag, Berlin, Heidelberg (1990), v. 88.
4. R. H. McKenzie, Comments Cond. Matt. Phys. 18, 309
(1998).
5. S. Lefebvre et al., Phys. Rev. Lett. 85, 5420 (2000).
6. J. Flouquet, arXiv: cond-mat/0501602 (unpublished).
7. D. J�rome, Chem. Rev. 104, 5565 (2004).
8. K. Bechgaard et al., Solid State Commun. 33, 1119
(1980).
9. C. Bourbonnais and D. J�rome, in: Advances Synthetic
Metals, Twenty Years of Progress in Science and
Technology, P. Bernier, S. Lefrant, and G. Bidan
(eds.), Elsevier, New York (1999), p. 206, arXiv:
cond-mat/9903101.
10. P M. Grant, J. Phys. (Paris.) 44, 847 (1983).
11. K. Yamaji, J. Phys. Soc. Jpn. 51, 2787 (1982).
12. L. Ducasse et al., J. Phys. C19, 3805 (1986).
13. L. Balicas et al., J. Phys. (France) 4, 1539 (1994).
14. V.J. Emery, R. Bruisma, and S. Bar�ii�, Phys. Rev.
Lett. 48, 1039 (1982).
15. S. Bari�i� and S. Brazovskii, in: Recent Developments
in Condensed Matter Physics, J.T. Devreese (ed.),
Plenum, New York, (1981), v. 1, p. 327.
16. D. Jaccard et al., J. Phys.: Cond. Matt. 13, L89
(2001).
17. H. Wilhelm et al., Eur. Phys. J. B21, 175 (2001).
18. T. Adachi et al., J. Am. Chem. Soc. 122, 3238 (2000).
19. C. Bourbonnais and D. J�rome, Science 281, 1156
(1998).
20. F.D.M. Haldane, J. Phys. C14, 2585 (1981).
21. T. Giamarchi, Quantum Physics in One Dimension,
Oxford University Press, Oxford (2004).
22. C. Coulon et al., J. Phys. (Paris) 43, 1059 (1982).
23. A. Schwartz et al., Phys. Rev. B58, 1261 (1998).
24. P. Wzietek et al., J. Phys. 3, 171 (1993).
25. C. Bourbonnais et al., Phys. Rev. Lett. 62, 1532
(1989).
26. C. Bourbonnais, J. Phys. 3, 143 (1993).
27. C. Bourbonnais and L.G. Caron, Europhys. Lett. 5,
209 (1988).
28. C. Bourbonnais and L.G. Caron, Int. J. Mod. Phys.
B5, 1033 (1991).
29. S.E. Brown et al., Synth. Metals 86, 1937 (1997).
30. C. Bourbonnais, F. Creuzet, D. J�rome, and K.
Bechgaard, J. Phys. Lett. (Paris) 45, L755 (1984).
31. C.S. Jacobsen, D.B. Tanner, and K. Bechgaard, Phys.
Rev. Lett. 46, 1142 (1981).
32. C.S. Jacobsen, D.B. Tanner, and K. Bechgaard, Phys.
Rev. B28, 7019 (1983).
33. N. Cao, T. Timusk, and K. Bechgaard, J. Phys. 6,
1719 (1996).
34. M. Dressel, A. Schwartz, G. Gr�ner, and L. Degiorgi,
Phys. Rev. Lett. 77, 398 (1996).
35. T. Giamarchi, Physica B230–232, 975 (1997).
36. V. Vescoli et al., Science 281, 1181 (1998).
37. T. Giamarchi, Chem. Rev. 104, 5037 (2004).
38. J. Favand and F. Mila, Phys. Rev. B54, 10425
(1996).
39. J. Moser et al., Phys. Rev. Lett. 84, 2674 (2000).
40. P. Auban-Senzier, D. J�rome, and J. Moser, in: Physi-
cal Phenomena at High Magnetic Fields, Z. Fisk, L.
Gor’kov, and L. Schrieffer (eds.), World Scientic,
Singapore (1999).
41. G. Mih�ly, I. K�zsm�rki, F. Z�mborszky, and L. Forro,
Phys. Rev. Lett. 84, 2670 (2000).
42. P. Fertey, M. Poirier, and P. Batail, Eur. Phys. J.
B10, 305 (1999).
43. J. Moser et al., Eur. Phys. J. B1, 39 (1998).
44. F. Zwick et al., Phys. Rev. Lett. 79, 3982 (1997).
45. B. Dardel et al., Europhys. Lett. 24, 687 (1993).
46. A. Georges, T. Giamarchi, and N. Sandler, Phys. Rev.
B61, 16393 (2000).
47. G.M. Danner, W. Kang, and P.M. Chaikin, Phys.
Rev. Lett. 72, 3714 (1994).
48. L.P. Gor’kov, J. Phys. (France) 6, 1697 (1996).
49. A.T. Zheleznyak and V.M. Yakovenko, Eur. Phys. J.
B11, 385 (1999).
50. P.M. Chaikin, J. Phys. (France) 6, 1875 (1996).
51. P. Lederer, J. Phys. (France) 6, 1899 (1996).
52. V.M. Yakovenko and H.S. Goan, J. Phys. 6, 1917
(1996).
53. J.P. Pouget and S. Ravy, J. Phys. 6, 1501 (1996).
54. S. Kagoshima et al., Solid State Commun. 110, 479
(1999).
55. H. Seo and H. Fukuyama, J. Phys. Soc. Jpn. 66, 1249
(1997).
56. N. Kobayashi, M. Ogata, and K. Yonemitsu, J. Phys.
Soc. Jpn. 67, 1098 (1998).
57. S. Mazumdar, S. Ramasesha, R. T. Clay, and D.K.
Campbell, Phys. Rev. Lett. 82, 1522 (1999).
58. Y. Tomio and Y. Suzumura, J. Phys. Soc. Jpn. 69,
796 (2000).
518 Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5
N. Dupuis, C. Bourbonnais, and J.C. Nickel
59. Y. Tomio and Y. Suzumura, J. Phys. Chem. Solids 62,
431 (2001).
60. P. Garoche, R. Brusetti, and K. Bechgaard, Phys.
Rev. Lett. 49, 1346 (1982).
61. P. Garoche, R. Brusetti, and D. J�rome, J. Phys.
Lett. (Paris) 43, L147 (1982).
62. D.L. Gubser et al., Phys. Rev. B24, 478 (1981).
63. K. Murata et al., Mol. Cryst. Liq. Cryst. 79, 639
(1982).
64. R.L. Green, P. Haen, S.Z. Huang, and E.M. Engler,
Mol. Cryst. Liq. Cryst. 79, 183 (1982).
65. R. Brusetti, M. Ribault, D. J�rome, and K.
Bechgaard, J. Phys. (France) 43, 52 (1982).
66. M.Y. Choi et al., Phys. Rev. B25, 6208 (1985).
67. S. Bouffard et al., J. Phys. C15, 2951 (1982).
68. A.A. Abrikosov, JETP Lett. 37, 503 (1983).
69. C. Coulon et al., J. Phys. (France) 43, 1721 (1982).
70. S. Tomic et al., J. Phys. Coll. (Paris) 44, C3 (1983).
71. N. Joo et al., Eur. Phys. J. B40, 43 (2004).
72. Q. Yuan et al., Phys. Rev. B68, 174510 (2003).
73. L.P. Gor’kov and D. J�rome, J. Phys. Lett. (Paris)
46, L643 (1985).
74. A.M. Clogston, Phys. Rev. Lett. 9, 266 (1962).
75. B.S. Chandrasekhar, Appl. Phys. Lett. 1, 7 (1962).
76. I.J. Lee, M.J. Naughton, G.M. Danner, and P.M.
Chaikin, Phys. Rev. Lett. 78, 3555 (1997).
77. I.J. Lee, P.M. Chaikin, and M.J. Naughton, Phys.
Rev. B62, R14669 (2000).
78. J.I. Oh and M.J. Naughton, Phys. Rev. Lett. 92,
067001 (2004).
79. A.G. Lebed, JETP Lett. 44, 114 (1986).
80. L.I. Burlachkov, L.P. Gor’kov, and A.G. Lebed,
Europhys. Lett. 4, 941 (1987).
81. N. Dupuis, G. Montambaux, and C.A.R. S� de Melo,
Phys. Rev. Lett. 70, 2613 (1993).
82. N. Dupuis and G. Montambaux, Phys. Rev. B49, 8993
(1994).
83. N. Dupuis, Phys. Rev. B51, 9074 (1995).
84. R.L. Greene and E.M. Engler, Phys. Rev. Lett. 45,
1587 (1980).
85. R. Brusetti, M. Ribault, D. J�rome, and K.
Bechgaard, J. Phys. (France) 43, 801 (1982).
86. T. Vuleti� et al., Eur. Phys. J. B25, 319 (2002).
87. I.J. Lee, P.M. Chaikin, and M.J. Naughton, Phys.
Rev. Lett. 88, 207002 (2002).
88. I.J. Lee et al., Phys. Rev. Lett. 94, 197001 (2005).
89. A.I. Larkin and Y.N. Ovchinnikov, Sov. Phys. JETP
20, 762 (1965).
90. P. Fulde and R.A. Ferrell, Phys. Rev. 135, A550
(1965).
91. A.I. Buzdin and S.V. Polonskii, Sov. Phys. JETP 66,
422 (1983).
92. A.G. Lebed, Phys. Rev. B59, R721 (1999).
93. A.G. Lebed, K. Machida, and M. Ozaki, Phys. Rev.
B62, R795 (2000).
94. I.J. Lee et al., Phys. Rev. Lett. 88, 017004 (2002).
95. I. J. Lee et al., Phys. Rev. B68, 092510 (2003).
96. M. Takigawa, H. Yasuoka, and G. Saito, J. Phys.
Soc. Jpn. 56, 873 (1987).
97. D. J�rome and C. R. Pasquier, in: Superconductors,
A.V. Narlikar (ed.), Springer-Verlag, Berlin (2005).
98. P. Fulde and K. Maki, Phys. Rev. B139, A788
(1965).
99. K. Sengupta et al., Phys. Rev. B63, 144531 (2001).
100. Y. Tanuma, K. Kuroki, Y. Tanaka, and S.
Kashiwaya, Phys. Rev. B64, 214510 (2001).
101. Y. Tanuma et al., Phys. Rev. B66, 094507 (2002).
102. Y. Tanuma, K. Kuroki, Y. Tanaka, and S.
Kashiwaya, Phys. Rev. B68, 214513 (2003).
103. H.I. Ha, J.I. Oh, J. Moser, and M.J. Naughton,
Synth. Metals 137, 1215 (2003).
104. M. Naughton, private communication (unpublished).
105. S. Belin and K. Behnia, Phys. Rev. Lett. 79, 2125
(1997).
106. H. Shimahara, Phys. Rev. B61, R14936 (2000).
107. Y. Fuseya and Y. Suzumura, J. Phys. Soc. Jpn. 74,
1264 (2005).
108. J. S�lyom, Adv. Phys. 28, 201 (1979).
109. V.J. Emery, J. Phys. Coll. (Paris) 44, C3 (1983).
110. S. Bari�i�, J. Labb�, and J. Friedel, Phys. Rev. Lett.
25, 99 (1970).
111. W.P. Su, J.R. Schrieffer, and A.J. Heeger, Phys.
Rev. Lett. 42, 1698 (1979).
112. F. Creuzet et al., Synth. Metals 19, 289 (1987).
113. C. Bourbonnais and B. Dumoulin, J. Phys. 6, 1727
(1996).
114. V.J. Emery, Synth. Metals 13, 21 (1986).
115. W. Kohn and J.M. Luttinger, Phys. Rev. Lett. 15,
524 (1965).
116. M.T. B�al-Monod, C. Bourbonnais, and V.J. Emery,
Phys. Rev. B34, 7716 (1986).
117. L.G. Caron and C. Bourbonnais, Physica B143, 453
(1986).
118. D.J. Scalapino, E. Loh, and J.E. Hirsch, Phys. Rev.
B34, 8190 (1986).
119. D.J. Scalapino, E. Loh, and J.E. Hirsch, Phys. Rev.
B35, 6694 (1987).
120. K. Miyake, S. Schmitt-Rink, and C.M. Varma, Phys.
Rev. B34, 6554 (1986).
121. H. Shimahara, J. Phys. Soc. Jpn. 58, 1735 (1988).
122. H. Kino and H. Kontani, J. Low Temp. Phys. 117,
317 (1999).
123. K. Kuroki and H. Aoki, Phys. Rev. B60, 3060 (1999).
124. D.J. Scalapino, Phys. Rep. 250, 329 (1995).
125. R. Duprat and C. Bourbonnais, Eur. Phys. J. B21,
219 (2001).
126. K. Kuroki, R. Arita, and H. Aoki, Phys. Rev. B63,
094509 (2001).
127. S. Onari, R. Arita, K. Kuroki, and H. Aoki, Phys.
Rev. B70, 94523 (2004).
128. Y. Tanaka and K. Kuroki, Phys. Rev. B70, R060502
(2004).
129. C. Bourbonnais and R. Duprat, Bull. Am. Phys. Soc.
49, 1:179 (2004).
130. J.C. Nickel, R. Duprat, C. Bourbonnais, and N. Du-
puis, Phys. Rev. Lett. 95, 247001 (2005).
131. J.C. Nickel, R. Duprat, C. Bourbonnais, and N.
Dupuis, arXiv: cond-mat/0510744 (unpublished).
Superconductivity and antiferromagnetism in quasi-one-dimensional organic conductors
Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5 519
132. Y. Hasegawa and H. Fukuyama, J. Phys. Soc. Jpn.
55, 3978 (1986).
133. G. Montambaux, Phys. Rev. B38, 4788 (1988).
134. L.P. Gor’kov and I.E. Dzyaloshinskii, Sov. Phys.
JETP 40, 198 (1974).
135. L. Mih�ly and J. S�lyom, J. Low Temp. Phys. 24,
579 (1976).
136. P.A. Lee, T.M. Rice, and R.A. Klemm, Phys. Rev.
B15, 2984 (1977).
137. N. Menyh�rd, Solidi State Commun. 21, 495 (1977).
138. S. Bari�i� and A. Bjeli�, in: Theoretical Aspects of
Band Structures and Electronic Properties of
Pseudo-One-Dimensional Solids, H. Kaminura (ed.),
D. Reidel, Dordrecht (1985), p. 49.
139. J.P. Pouget and R. Comes, in: Charge Density
Waves in Solids, L.P. Gor’kov and G. Gr�ner (eds.),
Elsevier Science, Amsterdam (1989), p. 85.
140. H. Shimahara, J. Phys. Soc. Jpn. 69, 1966 (2000).
520 Fizika Nizkikh Temperatur, 2006, v. 32, Nos. 4/5
N. Dupuis, C. Bourbonnais, and J.C. Nickel
|