You are on page 1of 13

PHYSICAL REVIEW X 13, 021008 (2023)

Superconducting Fluctuations Observed Far above T c in the Isotropic Superconductor K3 C60


Gregor Jotzu ,1,2,* Guido Meier ,1 Alice Cantaluppi,1 Andrea Cavalleri ,1,3 Daniele Pontiroli ,4
Mauro Riccò ,4 Arzhang Ardavan ,3 and Moon-Sun Nam 3
1
Max Planck Institute for the Structure and Dynamics of Matter, 22761 Hamburg, Germany
2
The Hamburg Centre for Ultrafast Imaging, 22761 Hamburg, Germany
3
Department of Physics, Clarendon Laboratory, University of Oxford, Oxford OX1 3PU, United Kingdom
4
Dipartimento di Scienze Matematiche, Fisiche e Informatiche, Università degli Studi di Parma,
Parma 43124, Italy

(Received 20 September 2021; revised 21 December 2022; accepted 9 February 2023; published 17 April 2023)

Alkali-doped fullerides are strongly correlated organic superconductors that exhibit high transition
temperatures, exceptionally large critical magnetic fields, and a number of other unusual properties. The
proximity to a Mott insulating phase is thought to be a crucial ingredient of the underlying physics and may
also affect precursors of superconductivity in the normal state above T c . We report on the observation of a
sizable magneto-thermoelectric (Nernst) effect in the normal state of K3 C60 , which displays the
characteristics of superconducting fluctuations. This nonquasiparticle Nernst effect emerges from an
ordinary quasiparticle background below a temperature of 80 K, far above T c ¼ 20 K. At the lowest fields
and close to T c , the scaling of the effect is captured by a model based on Gaussian fluctuations. The
behavior at higher magnetic fields displays a symmetry between the magnetic length and the correlation
length of the system. The temperature up to which we observe fluctuations is exceptionally high for a three-
dimensional isotropic system, where fluctuation effects are expected to be suppressed.
DOI: 10.1103/PhysRevX.13.021008 Subject Areas: Condensed Matter Physics,
Strongly Correlated Materials,
Superconductivity

Recent work suggests that the exceptional properties of phase coherence may be responsible for the disappearance
alkali-doped fulleride superconductors, A3 C60 , result from of superconductivity at T c . Yet, some uncertainty remains on
an unusual cooperation between electron-phonon and this assignment because of the large spread of experimen-
electron-electron coupling [1–4]. The former is primarily tally determined superfluid densities in A3 C60 [9,10].
governed by a dynamical Jahn-Teller distortion of the C60 In K3 C60 [see Fig. 1(a)], observables such as the specific
molecules, leading to an inverted Hund’s coupling between heat and the pressure dependence of T c suggest that the
electrons, while the latter contributes to a suppression of the material may be well described by weak-coupling BCS
effective bandwidth. With increasing lattice spacing, super- theory [9,10], but discrepancies in the size and temperature
conductivity in A3 C60 acquires a “domelike” T c , eventually dependence of the superconducting gap remain [11]. Very
evolving into a Mott insulator with an antiferromagnetic recent measurements in few-layer thin films of K3 C60 have
ground state [5,6]. Unlike other high-temperature super- reported the appearance of a pseudogap up to about twice
conductors, however, A3 C60 features no anisotropy and T c [12].
displays the characteristics of an s-wave superconductor. Finally, upon illumination with midinfrared laser pulses,
Additionally, A3 C60 seems to follow the Uemura relation optical properties compatible with superconductivity have
[7,8], with a transition temperature T c proportional to the been observed in K3 C60 at temperatures that exceed T c by
superfluid density, suggesting that the loss of long-range an order of magnitude [13–16], further underscoring a highly
unusual normal state. Recent experiments have also provided
* suggestive magnetic anomalies when Rb3 C60 interacts with
Corresponding author.
gregor.jotzu@mpsd.mpg.de electromagnetic vacuum modes in an optical cavity [17]. One
of the proposed mechanisms for these phenomena suggests
Published by the American Physical Society under the terms of that the effect of the light field consists in synchronizing
the Creative Commons Attribution 4.0 International license. preexisting, but phase-incoherent, Cooper pairs [15,18,19].
Further distribution of this work must maintain attribution to
the author(s) and the published article’s title, journal citation,
The magneto-thermoelectric effect known as the Nernst
and DOI. Open access publication funded by the Max Planck effect offers a powerful probe of the presence and nature
Society. of Cooper pairs above T c [20].

2160-3308=23=13(2)=021008(13) 021008-1 Published by the American Physical Society


GREGOR JOTZU et al. PHYS. REV. X 13, 021008 (2023)

above T c . However, even for a transition driven by a


thermal breakdown of Cooper pairing, the thermal diffusion
of short-lived Cooper pairs may also contribute to the
Nernst signal above T c [20,22,23].
In a number of materials, the superconducting contri-
bution to the Nernst effect is much larger than its quasi-
particle counterpart. Even precursors of superconductivity
may exceed the quasiparticle background [20,23–28].
However, this cannot be assumed to be the case, in general
FIG. 1. Probing the Nernst effect in K3 C60 . (a) The fcc lattice [29,30]: The largest Nernst effect observed so far originates
structure of K3 C60 , with potassium atoms shown in gray and from quasiparticles in bismuth [20,31,32]. A careful
carbon buckyballs in green. (b) Schematic of the measurement analysis of the field and temperature dependence of the
configuration. A temperature gradient ∇T is applied orthogonally Nernst effect is therefore required to distinguish different
to the external magnetic field B. The voltage V is then measured contributions [27,33].
orthogonally to both. Cartoons of various possible contributions In general, the presence of such precursors of super-
to the Nernst signal, such as quasiparticles (left), short-lived conductivity is expected to be suppressed as the dimension-
Cooper pairs (center), and mobile vortices (right), are included. ality of the system increases. Indeed, Nernst signals that could
be related to superconducting precursors have been reported
The Nernst effect describes the appearance of an electric in layered materials, where the interlayer superconducting
field, Ey ¼ −N∂x T, transverse to an applied temperature coherence length is short [24,25,27,33–39], and in thin-
gradient, ∂x T, and to a magnetic field Bz pointing along the film samples with a thickness smaller than or comparable to
third spatial direction. The Nernst signal N is related to the the coherence length [23,26,28]. However, we are not aware
conductivity and thermoelectric tensors, σ and α, via of any previous observations in a microscopically isotropic
and macroscopically three-dimensional system.
αxy σ xx − αxx σ xy αxy Interestingly, K3 C60 and Rb3 C60 were the first fully
N¼ ≈ − SμH Bz ð1Þ three-dimensional materials where slight deviations of the
σ 2xx þ σ 2xy σ xx
normal state conductivity in the immediate vicinity of T c
could be attributed to paraconductivity [40].
for an isotropic system. Here, μH ¼ σ xy =ðσ xx Bz Þ denotes
In the experiments reported in this paper, air-sensitive
the Hall mobility and S ¼ αxx =σ xx the Seebeck coefficient. K3 C60 powders were compressed into pellets (see
The approximate equality holds for small Hall angles and is Appendix A) and incorporated into a circuit board printed
an excellent approximation for the parameters used in on an FR4 substrate, which features low thermal conduc-
this work. tivity. Embedded heaters and temperature sensors, as well
For a metal, the effect can be seen as a combination of a as indium-coated contacts with contact resistances below
flow of charges carrying entropy along a temperature 1 Ohm, were used to optimize these measurements (see
gradient (the Seebeck effect) and the deflection of moving Appendix B and Fig. 6). Four-probe resistance measure-
charges in the presence of a magnetic field (the Hall effect) ments (see Appendix C) were used to identify the super-
[see Fig. 1(b)]. However, with exact particle-hole sym- conducting transition, as shown in the inset of Fig. 2(a). In
metry, the two terms in Eq. (1) would cancel exactly [21]. the zero field, we found T c ðB ¼ 0Þ ¼ 19.8 K, in good
The overall sign and amplitude of the effect depend on the agreement with previous reports and with magnetic mea-
details of the quasiparticle band structure, to which the surements on the same batch of samples [9,10,14]. A
Nernst signal is very sensitive. The expected magnetic field Werthamer-Helfand-Hohenberg (WHH) theory [41] was
dependence is linear. In the free-electron approximation, used to extrapolate the zero-temperature upper critical field,
this signal is linear in temperature as well. μ0 H c2 ð0Þ, from the field dependence of the resistive
In a superconductor, a different contribution to the transition via μ0 H c2 ð0Þ ¼ 0.69μ0 T c ð∂Hc2 =∂TÞjT c , where
Nernst effect arises from the movement of superconducting μ0 denotes the vacuum permeability. This yielded a value of
vortices. When these mobile vortices carry entropy along about 39 T, corresponding to a zero-temperature coherence
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
the applied temperature gradient [see Fig. 1(b)], they also length of ξ0 ¼ Φ0 =2πμ0 H c2 ð0Þ ¼ 2.9 nm (here, Φ0 is
carry magnetic flux, which induces a voltage in the trans- the magnetic flux quantum). This lies within the range of
verse direction [20]. This effect is characterized by a highly values extrapolated from other experiments [9,10] and
nonlinear dependence on Bz , perhaps owing to a competi- closely matches the value recently determined using pulsed
tion between vortex density and vortex mobility. fields in a range exceeding Hc2 [42]. Importantly, this value
If, at T c , superconducting long-range order breaks down is also orders of magnitude smaller than the thickness of the
because of fluctuations of the phase of the order parameter sample or the powder grain size (see Appendix A), meaning
while a finite Cooper pair amplitude remains, this vortex that the description of the sample remains fully three
Nernst effect would be expected to survive at temperatures dimensional throughout.

021008-2
SUPERCONDUCTING FLUCTUATIONS OBSERVED FAR ABOVE … PHYS. REV. X 13, 021008 (2023)

superconductors spanning several orders of magnitude in


critical fields and temperatures show signals that peak close
to this bound [28]. Although the amplitude of N does not
yet saturate at 15 T, the largest magnetic field achieved in
our measurement, the value it seems to approach agrees
with this upper bound—a quantitative discussion is found
in Appendix D.
As a function of magnetic field, the Nernst signal in the
superconducting phase is strongly nonlinear [see Fig. 7(a)]
and peaks at a field Bmax that reduces on approaching the
critical temperature. This reduction fits very well to a linear
function with a slope of −2.27ð3Þ T=K; see Fig. 4(a). It
displays a zero-field intersect at T ¼ 19.5ð1Þ K, close to
the resistive critical temperature T c ðB ¼ 0Þ ¼ 19.8 K.
Note that Bmax has previously been shown to bear some
similarity to a softening mode [20,29], and its vanishing
value when approaching T c ð0Þ suggests that the Nernst
effect changes in nature when crossing T c . Above T c , this
nonlinear magnetic field dependence persists and is com-
patible with the existence of a “ghost critical field”—more
details are provided in Appendix E.
Above T c , precursors of superconductivity may contrib-
ute to the Nernst signal, but quasiparticles can also play a
role. The latter contribution is expected to scale linearly
with magnetic field and temperature (at constant volume)
for a simple metallic state [21].
We therefore used the temperature- and magnetic-field-
normalized Nernst signal, N ¼ N=BT ¼ ν=T, in order
FIG. 2. Nernst signal in the superconducting regime. (a) Mea- to distinguish superconducting and quasiparticle con-
sured transverse thermoelectric effect N in the presence of tributions. Here, ν ¼ N=B denotes the magnetic-field-
magnetic fields ranging from 1 T (gray line) to 15 T (blue line), normalized Nernst signal, generally referred to as the
in steps of 1 T. The red dashed line shows the critical temperature “Nernst coefficient.”
in the zero field. Inset: sample resistance near the critical In Fig. 3(a), N is shown to evolve smoothly across T c ,
temperature, 0 T (black) to 15 T (blue) range. (b) Contour plot remaining positive and retaining a strong field dependence,
of N as a function of temperature and magnetic field. The black as expected for a signal that is primarily caused by
dotted line shows the resistive T c ðBÞ. A Gaussian convolution
superconducting fluctuations. Strikingly, the inset in
was used for smoothing, and in the low-T, low-B region, the data
are sparse and interpolated. See Fig. 7(b) for raw data and the Fig. 3(a) shows that N displays near-universal behavior
field-normalized N=B. upon approaching the field-dependent critical temperature
T c ðBÞ [see also Figs. 4(b) and 7(d)]. Theoretical work on
two-dimensional systems [44–46] predicted such behavior
As expected, no Nernst signal (N) is observed for close to the phase transition given by T c ðBÞ, and our results
temperatures far below T c, likely due to freezing of vortex suggest that this also holds in three dimensions.
motion. For higher temperatures, entering the temperature At higher temperatures, we observed two characteristic
range where vortices become mobile, the signal is seen to features in the data [see Fig. 3(b)]. First, at T 0 ðBÞ ≈ 50 K,
increase [see Fig. 2(a)]. For a given magnetic field, the the signal changed from positive to negative. The temper-
number of vortices in the system remains nearly constant ature of the zero crossing, T 0 ðBÞ, hence denotes the point
(as all fields in our measurements are much larger than the where the superconducting and the quasiparticle contribu-
lower critical field), but their mobility increases rapidly. tions to the Nernst effect have the same magnitude. Note
Near T c, the Nernst signal reduces again. A detailed that T 0 ðBÞ shows a field dependence similar to T c ðBÞ [see
quantitative understanding of this well-known phenomenon Fig. 4(a)]. This suggests that the aforementioned observa-
is still lacking—both the increasing vortex-vortex inter- tion, that the Nernst effect close to the phase transition
actions that appear as vortex length scales increase and a tracks the field-dependent critical temperature T c ðBÞ, still
change in the entropy per vortex are relevant [20,43]. remains visible at much higher temperatures.
However, very recent results suggest that an upper bound Second, a minimum in N appears at T min ðBÞ ≈ 80 K,
for this vortex Nernst effect may exist, and various above which N shows a linear and positive slope. In order

021008-3
GREGOR JOTZU et al. PHYS. REV. X 13, 021008 (2023)

FIG. 3. Nernst effect above T c . (a) Temperature and magnetic-


field normalized Nernst signal N close to T c (red dashed line).
Open symbols show the raw data; lines are 10-point boxcar
averages for 5 T, 10 T, and 15 T (gray to blue). Inset: same data FIG. 4. Superconducting Nernst effect vs temperature and B
rescaled with the field-dependent T c ðBÞ. (b) N at higher temper- field. (a) Characteristic quantities extracted from the Nernst
atures, after 10-point boxcar averaging (note the reduced y scale). signal: Bmax denotes the fitted peak of NðBÞ [see Fig. 2(b)];
The red curve corresponds to the quasiparticle contribution T 0 and T min are the zero crossing and minimum of N ,
−2.7μH S=T. respectively; T c denotes the resistive superconducting transition
temperature. All lines are linear fits. Error bars indicate fit
uncertainties. (b) Quasiparticle subtracted contribution to the
to disentangle the superconducting contribution to N , we Nernst coefficient νS as a function of dimensionless distance to
first focus on the high-temperature limit, where the quasi- the critical temperature ϵ and dimensionless field h. The black
particle contribution should become dominant. We can dotted line shows T c ðBÞ; the gray dashed line is h ¼ ϵ. (c) νS as a
compare the behavior of N in this regime to the expected function of correlation length ξ and magnetic length l̃B . The
signal in a single-band free-electron model of a metal, dashed line shows l̃B ¼ ξ.
jN j ∼ μH S=T, with a prefactor of order unity [21]. This
relationship implies that the Nernst effect is proportional to
the ratio of the mobility and the Fermi energy in the metal, dependence of N . Given the complex band structure of
which was found to hold in a variety of materials, with N K3 C60 , the agreement with such a simple scaling is
ranging from 1 mV=K2 T down to 1 nV=K2 T [20]. We use remarkable. It extends its range of validity to values of
previously determined values for the T-linear S and μH N as low as 20 pV=K2 T, which are 2 orders of magnitude
[47,48], where the latter shows a linear dependence on smaller than those reported so far [20].
temperature that has been shown to scale with the expan- The rapid change in the slope of N around 80 K
sion of the lattice. As shown in Fig. 3(b), we find excellent indicates a significant change in the electronic properties
agreement with our data above T min , using a quasiparticle of the material. Such a change could, in principle, be caused
contribution of −2.7μH S=T, and no discernible field by the appearance of charge-density wave order [49], and in

021008-4
SUPERCONDUCTING FLUCTUATIONS OBSERVED FAR ABOVE … PHYS. REV. X 13, 021008 (2023)

a number of cuprate superconductors, a very similar feature For the latter scenario, a theory based on Gaussian
was found to coincide closely with the pseudogap temper- fluctuations in a Ginzburg-Landau model [22] predicts a
ature T  [27]. In K3 C60 , a transition to a frozen orientational superconducting contribution to the transverse Peltier
disorder of the C60 molecules is known to occur but at a coefficient αxy , which is proportional to the magnetic field
temperature very close to 200 K [50,51]. There are no and otherwise depends only on the correlation length ξ and
observations pointing at the appearance of competing orders fundamental constants. Together with the temperature-
around 80 K, although it is worth noting that a certain dependent conductivity of the sample, this allows for a
deviation from linearity has been observed in the Seebeck prediction of the superconducting contribution to the
effect, which has been attributed to electron-phonon cou- Nernst coefficient, νSCG , given by
pling or precursors of superconductivity [47,52,53].
An effect related to the superconducting ground state N SCG αSCG
xy k e2 ξ
νSCG ¼ ¼ ¼ B 2 ð2Þ
seems much more likely, given that the magnetic field Bz σ xx Bz 12πℏ σ xx
dependence of the Nernst effect in this regime tracks T c ðBÞ.
In order to further explore the magnetic-field dependence of with
the effect, in Fig. 4(b) we plot the quasiparticle-subtracted
Nernst coefficient νS ¼ N S =B ¼ N=B þ 2.7μH S as a func- ξ ξ0
ξ ¼ p0ffiffiffi ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi ð3Þ
tion of the dimensionless temperature distance to the ϵ ðT − T c Þ=T c
critical temperature ϵ ¼ T=T c ð0Þ − 1 and the dimension-
for a three-dimensional system. Here, kB denotes the
less magnetic field h ¼ B=μ0 H̃c2 . Here, we use the
Boltzmann constant. The two-dimensional version of this
Ginzburg-Landau critical field, which is based on the
theory is in good agreement with measurements on conven-
low-field extrapolation H̃ c2 ¼ μ0 T c ð∂Hc2 =∂TÞjT c . It differs tional [20,23,28] and some unconventional [29,33] super-
from the WHH extrapolation by a factor of 0.69 (see conductors. Its prediction of a field-independent Nernst
Ref. [46] for details and for the expected universal behavior coefficient applies to the low-field regime, where the
of the two-dimensional Nernst effect in these units). Close correlation length ξ is short compared to the magnetic
to the critical line (given by h ¼ −ϵ for h ≪ 1), the Nernst length lB , and T c ðBÞ ≈ T c ð0Þ. As it is a continuum theory, it
coefficient depends only on the dimensionless distance may also become invalid once ξ becomes as short as the
normal to the critical line. Above T c ð0Þ (i.e., for ϵ > 0), we lattice spacing, and generally speaking, Ginzburg-Landau
observe an approximate symmetry with respect to the gray theory is only applicable in the vicinity of T c .
dashed line where h ¼ ϵ. In Fig. 5(a), we compare our measurements for fields up
A physical explanation for this symmetry has been to 6 T to this theory by subtracting the fitted quasiparticle
suggested in previous work on two-dimensional systems contribution determined above [red line in Fig. 3(b)] from
[44–46,54]. In the normal state, the superconducting the measured Nernst signal. See Appendix F and Fig. 9 for
contribution to the Nernst effect is determined by a single other possible subtraction schemes. We find that
function of two length scales: the (temperature-dependent) pffiffiffi the data
pffiffiffi are overall well described by the simple 1= ϵ scaling
Ginzburg-Landau correlation length ξ p
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ffiffiffiξ0 = ϵ and the predicted by the theory if we use a constant conductivity
magnetic length l̃B ¼ ℏ=ð2eBÞ ¼ lB = 2, where e is the of 6 ðmΩ cmÞ−1 .
electron charge and ℏ the reduced Planck constant [55]. In The temperature dependence of the conductivity of the
the low-field limit (where the magnetic length is large), the sample is small enough that it has a negligible effect in the
Nernst coefficient is determined only by the correlation regime where superconducting fluctuations are visible,
length (see below), and it increases as the correlation length both for the intrinsic conductivity and the effective con-
becomes longer when approaching T c . However, at larger ductivity of the compressed powder sample we study (see
fields or very close to the critical point, the correlation Appendix G and Figs. 9 and 10). Although we expect our
length can become comparable to the magnetic length, measurement to be sensitive to the intrinsic conductivity
which then limits the size of the Nernst coefficient. rather than grain boundary effects [39,57], this value is still
In thin films of Nb0.15 Si0.85 , the Nernst coefficient is about 3 times larger than the conductivity of high-purity
found to be approximately symmetric with respect to the single crystals [9]. Interestingly, however, a very similar
line where the correlation length ξ equals the magnetic discrepancy was found in measurements of the paracon-
length l̃B [54]. In Fig. 4(c), we present the Nernst ductivity [40], suggesting that the residual conductivity
coefficient in K3 C60 , showing that this approximate sym- describing the transport properties of superconducting
metry can also be found in a three-dimensional system. fluctuations may not be identical to the one extracted from
In the following, we consider two scenarios through direct measurements.
which precursors of superconductivity could cause a Nernst At higher temperatures, beyond the expected regime
effect far above T c : a vortex-based Nernst signal surviving of validity of this model, deviations from the simple
in a phase-fluctuating regime above T c , or a signal caused scaling occur, which might be captured by perturbative
by short-lived Cooper pairs. expansions [44,45], but further work would be needed to

021008-5
GREGOR JOTZU et al. PHYS. REV. X 13, 021008 (2023)

deviates from the thermodynamic critical point, which is a


characteristic of unconventional superconductors [59].
Note that the theory presented above predicts a very
different result for the Nernst coefficient in a (quasi) two-
dimensional system [22]. There, α2D 2 2 2
xy =Bz ¼ ξ kB e =6πℏ ,
and a sheet resistance must be used to compute ν. The blue
dotted line in Figs. 5(a) and 5(b) shows the expected signal
under the assumption that a dimensional reduction of the
system has taken place (see Appendix H), which was
discussed in the context of alkali-doped fullerides [60]. Our
results clearly deviate from this prediction, confirming the
three-dimensional nature of the state we observe.
In order to gauge the plausibility of a phase-fluctuating
scenario, we use the framework proposed by Emery and
Kivelson [61] to estimate the temperature T θ at which global
phase coherence in the superconductor would be destroyed
by thermal fluctuations—even if pairing were to survive up
to a higher “mean field temperature” T MF . Taking the most
recent (and largest) value for the penetration depth in K3 C60 ,
λ ¼ 890 nm [62], we find a temperature T θ as low as 80 K.
This is only 4 times larger than T c , whereas for conventional
superconductors, T θ =T MF can be on the order of 105 .
Additionally, by taking into account some degree of quan-
tum fluctuations (as expected given the relative proximity of
a Mott-insulating state), it is possible that the superconduct-
ing transition is somewhat suppressed below T MF.
We are not aware of a quantitative prediction of the
vortex Nernst signal above T c in a three-dimensional
system. Results in two dimensions [43] suggest an impor-
tant role of the lattice geometry and therefore cannot simply
be extrapolated to our system. As the difference between
mean-field models and fully quantum-mechanical descrip-
tions can become less pronounced in higher dimensions,
FIG. 5. Scaling of the superconducting Nernst coefficient above
T c . (a) Quasiparticle-subtracted contribution to the Nernst co-
good agreement of our data with a theory based on
efficient, νS ¼ N S =B ¼ N=B þ 2.7μH S, for magnetic fields be- Gaussian fluctuations does not necessarily rule out agree-
tween 3 T and 6 T. Color scale as in Fig. 1. The data are shown as ment with a theory based on phase fluctuations.
a function of the distance to the field-dependent critical temper- Further theoretical work is therefore required for a
ature at zero field. (b) Data for higher fields, up to 15 T. The quantitative distinction between these two theoretical
dashed red line shows the expected low-field value from a model scenarios. In particular, capturing the behavior at high
based on three-dimensional Gaussian fluctuations, which scales temperatures and magnetic fields observed in our data
with ϵ−1=2 . A constant conductivity of 6 ðmΩ cmÞ−1 is used (see could serve as an important benchmark for this comparison.
text). The dashed blue line shows the equivalent signal for a two- On a qualitative level, the presence of an additional
dimensional system, which scales with ϵ−1 . The top axes indicate temperature scale (and hence an additional length scale)
where the correlation length ξ becomes equal to the lattice is expected to be visible in a phase-fluctuating scenario
spacing aLattice or to the magnetic length lB at the highest field [63], but it does not seem to be observable in our data.
shown in the respective panel. In the context of light-induced superconductivity in
K3 C60 , our data provide an important input for any
extend those theories to isotropic systems [58]. However, theoretical framework based on the synchronization of
the data presented in Figs. 4 and 5 already show some of the preexisting (stable or short-lived) but globally phase-
qualitative features that would be expected within such a incoherent Cooper pairs, which would also need to cor-
framework. In particular, the downturn at high temper- rectly describe the initial static state.
atures, which was also observed in two-dimensional It would be highly interesting to extend our work to
systems [20], may be an indication of the quantum nature Rb3 C60 and especially Rbx Cs3−x C60, where quantum phase
of fluctuations [44]. Interestingly, at higher fields [see fluctuations caused by the proximity of the Mott-insulating
Fig. 5(b)], the behavior does not follow a single power law. state will be enhanced. For the latter family, a suppression
This suggests that in a high field, the resistive T c ðBÞ in T c upon approaching the quantum phase transition has

021008-6
SUPERCONDUCTING FLUCTUATIONS OBSERVED FAR ABOVE … PHYS. REV. X 13, 021008 (2023)

been observed [5,6], but it could not be reproduced in an


otherwise quantitatively successful theoretical model [1].
Studying the Nernst effect in this regime, which should be
possible using the experimental framework presented here,
would provide new insights concerning the nature of the
superconducting transition in the fullerides, and of phase-
incoherent superconductivity in general.
We thank Michele Buzzi, Dante Kennes, Daniel
Podolsky, Dharmalingam Prabhakaran, and Maksym
Serbyn for insightful discussions, and Boris Fiedler for
superb technical assistance. This project was funded by the
Cluster of Excellence “CUI: Advanced Imaging of Matter”
(EXC 2056, Project No. 390715994) of the Deutsche
Forschungsgemeinschaft (DFG, German Research
Foundation), by the DFG Collaborative Research Center
SFB 925 (Project No. 170620586), and by the Engineering
and Physical Sciences Research Council (EPSRC, Grant
code EP/P000479/1)
FIG. 6. Experimental setup. (a) Sample (red) pressed on top of a
APPENDIX A: SAMPLE PREPARATION circuit board (green) using a PMMA piston (beige). The sample
space is encapsulated using nonconductive epoxy glue. The base
The K3 C60 powder used in this work was prepared and temperature is measured with a sensor (black) pressed against the
characterized as previously reported in Refs. [13–15]. bottom of the circuit board. The sample and circuit board are
In brief, finely ground C60 powder and metallic potassium mounted on the cold finger of a cryostat (copper) using titanium
were placed in a vessel inside a Pyrex vial in stoichiometric screws (not shown), with spacers made of PMMA (gray).
amounts, evacuated to 10−6 mbar, and sealed. The two (b) Schematic of the circuit board. The sample compartment is
materials were heated at 523 K for 72 h and then at 623 K indicated by a gray rectangle. A temperature sensor is placed on
each side of the sample in a milled pocket, encapsulated by
for 28 h, and kept separated to ensure that the C60 was only
thermally conductive epoxy glue. Resistive heaters are placed
exposed to clean potassium vapor. After regrinding and on each side. The Nernst signal is measured with the indicated
pelletizing in an inert Ar atmosphere, the sample was indium-coated contacts in the center of the sample compartment.
annealed at 623 K for 5 days. Powder x-ray diffraction The other contacts are used for four-point resistance measurements.
measurements confirmed the purity of K3 C60 and indicated (c) Transverse voltage at þ15 T (blue) and −15 T (gray). For the
a domain size between 100 nm and 400 nm. Optical data shown in the main text, the Nernst signal is evaluated as half of
microscopy measurements show grain sizes on the order the difference between signals measured at opposite fields.
of 10 μm. Magnetic susceptibility measurements yielded a
T c of 19.8 K [13]. For the Nernst effect and four-point sample. They were used to monitor the temperature gradient
resistivity measurements, the sample was handled inside an across the sample, which was induced using a resistive
Ar glove box with less than 0.2 ppm O2 and H2 O. It was heater. An additional Cernox sensor was attached to the
placed inside an FR4 frame, which had been glued to the bottom of the circuit board to monitor the base temperature.
circuit board described below using thermally and electri- The circuit board was mounted to the cold finger of a cryostat
cally insulating, minimally outgassing glue (Epo-TeK 301- using nonmagnetic (titanium) screws and spring washers,
2FL-T); see Fig. 6. The powder was then compressed with and PMMA spacers were used for additional thermal
an FR4 piston, hence creating a pellet of around 150-μm insulation. In the sample compartment, the copper contacts
thickness, and sealed with the same glue. The resistance of were coated with indium, yielding contact resistances below
the sample was monitored to ensure that no contamination 1 Ohm. The transverse voltage was measured while slowly
occurred during the sealing process and the subsequent cooling the sample. Data for opposite magnetic fields [see
transfer to the cryostat. Fig. 6(c)] were then subtracted to compute the Nernst signal.

APPENDIX B: NERNST EFFECT APPENDIX C: RESISTANCE MEASUREMENTS


MEASUREMENT SETUP
The resistance of the sample was determined using a low-
We used a printed copper circuit board on an FR4 substrate frequency lock-in measurement in a linear four-contact
(which featured a low thermal conductivity, about configuration, with contacts as shown in Fig. 6(b). Above
0.1 W=Km at 10 K [64]), as shown in Fig. 6(b). Cernox T c , the sample showed an increase in resistance
temperature sensors were embedded in thermally conductive upon cooling, as previously observed in granular K3 C60
glue (Stycast 2850FT) in milled pockets on each side of the samples [9,10]. In order to determine T c , the point at which

021008-7
GREGOR JOTZU et al. PHYS. REV. X 13, 021008 (2023)

the resistance changes slope was used, which yielded a zero- resistivity of the sample ρ via SV ¼ Φ0 aLattice N=ρ, where
field T c consistent with magnetic susceptibility measure- aLattice ¼ 1.42 nm denotes the lattice spacing. Taking this
ments on the same batch of sample. We verified that the width ratio at the temperature and magnetic field corresponding to
of the transition is not sensitive to reducing the probe current the peak in N, values for SV very close to kB ln 2 ¼ 0.96 ×
below the value of 2 μA that we used. 10−23 J=K were found for a range of materials. As the
resistivity of our samples is dominated by grain boundary
effects, we use the intrinsic low-temperature normal resis-
APPENDIX D: MAXIMUM VORTEX tivity of around ρ0 ¼ 0.5 mΩ cm [9] to estimate whether S
NERNST SIGNAL may exceed kB ln 2. Taking a resistivity between ρ0 and
ρ0 =3 as a low guess for the resistivity at the peak of N
Very recent results suggested that a universal upper (which lies below T c, but at high fields, so that flux-flow
bound for the vortex Nernst signal below T c may exist [28]. resistance appears [28]) then yields values between
By looking at Figs. 2(a) or 7(b), it is clear that for K3 C60 at 0.1 × 10−23 J=K and 0.4 × 10−23 J=K. This result is below,
15 T, the signal has not reached its maximum possible value but reaches the same order of magnitude as, the upper
yet. However, a sublinear dependence of the peak in NðTÞ bound identified in Ref. [28].
as a function of magnetic field is already visible. Given the
upper critical field of around 39 T, we can therefore
APPENDIX E: VISIBILITY OF THE GHOST
estimate the largest value of NðT; BÞ to be around 2 μV=K.
CRITICAL FIELD
This value in itself is comparable to the values compiled
by Rischau et al. [28]. The upper bound proposed in Previous theoretical and experimental work on the
that work relates to the entropy per vortex per layer SV , Nernst effect in two-dimensional and layered supercon-
which can be determined from the Nernst signal N and the ductors identified a “ghost critical field” at which the

FIG. 7. B-T maps of the Nernst signal. (a) N as a function of magnetic field for different temperatures below T c ðBÞ. (b) Raw data of N
as a function of temperature and magnetic field (see Fig. 2 for a smoothed contour map). Gray dots indicate the B-T values of each
measurement; the data are interpolated to the nearest available point. The black dotted line shows T c ðBÞ. (c) Same as (b), but showing
the field-normalized value N=B. (d) Smoothed contour plot of (c).

021008-8
SUPERCONDUCTING FLUCTUATIONS OBSERVED FAR ABOVE … PHYS. REV. X 13, 021008 (2023)

Nernst signal N (B) peaks, for a given temperature above as the temperature is larger than about 1.07 T c (i.e., about
T c [26,29,33,46,54]. This happens when the magnetic 21 K for our system). For temperatures closer to T c , the
length lB becomes comparable to the correlation length ghost field rapidly approaches zero. Although the approxi-
ξðTÞ. This ghost critical field is typically close to, but mate solution found in 2D is unlikely to hold exactly in 3D,
higher than, the “mirror field” given by the reflection we can use it to estimate where the ghost critical field may
around T ¼ T c of the linear low-field limit of H c2 , i.e., appear in K3 C60 . We find that, for K3 C60, the 2D expression
given by BMirror ¼ μ0 ðT − T c Þð∂H c2 =∂TÞjT c . Expressed in for the ghost critical field reaches 15 T (the maximum field
terms of the dimensionless quantities defined in the main accessible in our experiment) at 21.3 K already. The mirror
text, this mirror field hence occurs where h ¼ t − 1. field reaches a value of 15 T at a temperature of 25 K (i.e.,
No exact analytical expression for the ghost field has about 1.25 T c ). This means that our experiment is mostly
been found so far, but numerical work reported in Ref. [46] carried out in a regime where the magnetic field acts as a
has shown that, for a two-dimensional system, it can be linear probe of the intrinsic properties of the sample and
approximated by the expression h ¼ 1.12t − 0.94 as soon does not impose an additional length scale for the thermo-
dynamics of the Nernst effect.
In the immediate vicinity of T c , however, signatures of
the ghost critical field may appear. Figure 8 shows the
Nernst signal as a function of magnetic field, at different
temperatures close to T c ¼ 19.8 K. Well below T c, a sharp
peak can be observed in the (quasiparticle-subtracted)
Nernst signal (see also Fig. 7), which is the quantity shown
as Bmax in Fig. 3(c). Upon approaching T c , the last range
where a sharp peak can be identified is 17.5–18 K (see blue
curve on the right). Above that range, N S ðBÞ becomes more
flat (similar to what was observed in NbSi [46]). Given the
flatness of this feature, it is difficult to unequivocally
determine the peak position; however, the data are com-
patible with a peak that moves towards even lower fields.
For temperatures above T c [Fig. 8(b)], our data are
compatible with a peak in N S ðBÞ that very rapidly moves
towards higher fields. As outlined above, we expect that
this peak may exceed 15 T at around 21 K, and indeed our
data in the 20–21 K range already show that the peak has
shifted to 8 T or higher, but sublinear behavior is still
clearly visible.

APPENDIX F: SUBTRACTION OF THE


QUASIPARTICLE SIGNAL
By comparing different subtraction schemes, we verify
that the details of how the quasiparticle contribution to N is
subtracted do not strongly affect the comparison to the
Gaussian fluctuation model shown in Fig. 5: In Fig. 9(a),
instead of subtracting the temperature-dependent quasipar-
ticle function (−2.7μH S), we subtract its fixed value at
100 K. Here, the simple theoretical model seems to capture
the data even at higher temperatures. In Fig. 9(b), we plot N
without any quasiparticle subtraction. This leads to a
deviation at high temperatures (as expected given that
the signal changes sign there), but within the expected
range of validity of the theoretical model, it still captures
FIG. 8. Magnetic field dependence of the Nernst signal close to
the critical temperature. (a) Quasiparticle-subtracted Nernst the data well. We have also verified that using a temper-
signal as a function of magnetic field for different temperatures ature-dependent value of the conductivity in Eq. (2) (where
just below T c and just above T c (b). Note the different vertical we have used the quadratic dependence found in Ref. [65]
scales. A peak in N s above T c ðBÞ may indicate the presence of a as a comparison) has a negligible effect in the rel-
ghost critical field (see text). evant range.

021008-9
GREGOR JOTZU et al. PHYS. REV. X 13, 021008 (2023)

expected scaling (dotted gray line vs dashed dark red


line). We additionally investigate the effect of temper-
ature-dependent conductivity by dividing our measured
Nernst coefficient νS ¼ N S =B by either the temperature-
dependent intrinsic resistivity, ρI , based on measurements on
single-crystal samples [9], or the measured temperature-
dependent resistivity of our compressed powder sample, ρP
(see Fig. 10). In both cases, we only find very small changes
with respect to the scaling of the Nernst coefficient itself.
Note that we do not expect the measured resistivity of the
powder sample to be the correct quantity in order to
determine the expected size of the Nernst effect: In a
resistance measurement, current flows through grain boun-
daries, leading to a voltage drop that will provide the
dominant contribution to the measured resistance. In a
measurement of the Nernst signal, on the other hand, no
current is flowing in the steady state (where the Nernst
voltage is measured), meaning that the electrical resistance
caused by grain boundaries does not contribute, and only the
intrinsic resistivity of the sample determines the signal. See
also Refs. [39,57] for theoretical and experimental work
corroborating this statement.

APPENDIX H: COMPARISON TO
TWO-DIMENSIONAL THEORY
In order to illustrate the distinct three-dimensional nature
of our observations, there are two possible scenarios to take
into account, in which the effect could become effectively
two dimensional: First, one could consider the limit of a
sample that is thin enough (in the direction of the magnetic
field and transverse to the thermal gradient) that the
FIG. 9. Effect of subtracting the quasiparticle signal on the correlation length ξ exceeds the thickness d. In that case,
scaling analysis. (a) Same as Fig. 5(a), but subtracting the fixed one would use the resistivity multiplied by the sample
value of N=B at 100 K instead of a temperature-dependent thickness as a sheet resistance. The expression for the
function. The gray dotted line shows the theoretical prediction of Nernst coefficient for Gaussian superconducting fluctuations
Eq. (2) but using a quadratic temperature dependence for the [Eq. (2)], νSCG ¼ ðkB e2 =12πℏ2 Þðξ=σ xx Þ, would become
conductivity. (b) Nernst coefficient N=B without any subtraction ðkB e2 =6πℏ2 Þðξ2 =σ xx aÞ; i.e., it would be multiplied by a
of the quasiparticle contribution. Note that this results in some (temperature-dependent) factor of 2ξ=d. In the range
negative values for N at higher temperatures, which do not appear explored in Fig. 5, this factor would range from 1 × 10−4
in this logarithmic plot.
to 6 × 10−6 ; i.e., the signal would be several orders of
magnitude smaller than the signal predicted by the three-
APPENDIX G: RATIO OF NERNST dimensional theory and observed in our data, in addition to
COEFFICIENT AND RESISTIVITY having a different temperature dependence. Furthermore, as
the Ginzburg-Landau correlation length remains below
The theory for a Nernst effect caused by Gaussian super- 10 nm even at the data closest to T c and our sample has a
conducting fluctuations [22] predicts a result for the Nernst thickness of around 150 μm (and even the smallest sample-
coefficient νSCG that depends only on the (temperature- related length scale, the size of the crystal domains, is
dependent) correlation length ξ and the conductivity of the between 100 nm and 400 nm), this scenario can be ruled out.
sample. As the conductivity of the sample is generally Second, a more realistic scenario would be that the sample
temperature dependent, this will affect the scaling of the might undergo spontaneous dimensional reduction in the
Nernst coefficient with temperature. For K3 C60, this effect is applied magnetic field, as has been considered in the context
expected to be very small, as the change in the Nernst of alkali-doped fullerides before [60]. In this case, the
coefficient is much larger than the change of the conductivity. Lawrence-Doniach model could be used to describe such
Figure 9 shows how taking into account the temperature an effectively layered sample [see Eq. (14) in Ref. [22] ]. The
dependence of the intrinsic conductivity would change the expected contribution to the Nernst coefficient would then be

021008-10
SUPERCONDUCTING FLUCTUATIONS OBSERVED FAR ABOVE … PHYS. REV. X 13, 021008 (2023)

FIG. 10. Ratio of the Nernst signal and the resistivity. Same as Figs. 5(a) and 5(b), but divided by the intrinsic resistivity ρI of the
sample (a,b) or the powder resistivity ρP (c,d).

kB e2 ξxy [2] M. Capone, M. Fabrizio, C. Castellani, and E. Tosatti,


νLD
SCG ¼ 2
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi; ðH1Þ Strongly Correlated Superconductivity, Science 296, 2364
6πℏ σ xx a 1 þ ð2ξz =aÞ2
(2002).
[3] S. Chakravarty, M. P. Gelfand, and S. Kivelson, Electronic
where a ¼ aLattice ¼ 1.42 nm is the lattice spacing, and ξxy Correlation Effects and Superconductivity in Doped Full-
(ξz ) denotes the correlation length transverse to (along) the B erenes, Science 254, 970 (1991).
field. In the limit of a vanishing ξz (i.e., for completely [4] Phonon effects may not necessarily be required for an
decoupled layers), this essentially corresponds to taking the inverted Hund’s coupling, see H.-C. Jiang and S. Kivelson,
Electronic Pair Binding and Hund’s Rule Violations in
two-dimensional result for αxy from Ref. [22] and using the
Doped C60 , Phys. Rev. B 93, 165406 (2016).
resistivity multiplied by the lattice spacing as a sheet [5] Y. Takabayashi, A. Y. Ganin, P. Jeglič, D. Arčon, T. Takano,
resistance. Compared to the 3D result for the Nernst Y. Iwasa, Y. Ohishi, M. Takata, N. Takeshita, K. Prassides,
coefficient ν, this corresponds to multiplying by a temper- and M. J. Rosseinsky, The Disorder-Free Non-BCS Super-
ature-dependent coefficient of 2ξ=a. We have included a plot conductor Cs3 C60 Emerges from an Antiferromagnetic
of the expected signal in this scenario in Fig. 5 as a dashed Insulator Parent State, Science 323, 1585 (2009).
blue line. [6] R. H. Zadik, Y. Takabayashi, G. Klupp, R. H. Colman, A. Y.
Ganin, A. Potočnik, P. Jeglič, D. Arčon, P. Matus, K.
Kamarás, Y. Kasahara, Y. Iwasa, A. N. Fitch, Y. Ohishi, G.
Garbarino, K. Kato, M. J. Rosseinsky, and K. Prassides,
Optimized Unconventional Superconductivity in a Molecu-
[1] Y. Nomura, S. Sakai, M. Capone, and R. Arita, Exotic lar Jahn-Teller Metal, Sci. Adv. 1, e1500059 (2015).
s-Wave Superconductivity in Alkali-Doped Fullerides, J. [7] Y. Uemura, A. Keren, L. Le, G. Luke, W. Wu, J. Tsai, K.
Phys. Condens. Matter 28, 153001 (2016). Tanigaki, K. Holczer, S. Donovan, and R. Whetten, System

021008-11
GREGOR JOTZU et al. PHYS. REV. X 13, 021008 (2023)

Dependence of the Magnetic-Field Penetration Depth in [23] A. Pourret, H. Aubin, J. Lesueur, C. A. Marrache-Kikuchi,
C60 Superconductors, Physica (Amsterdam) 235–240C, L. Bergé, L. Dumoulin, and K. Behnia, Observation of the
2501 (1994). Nernst Signal Generated by Fluctuating Cooper Pairs, Nat.
[8] Y. J. Uemura et al., Universal Correlations between T c and Phys. 2, 683 (2006).
ns =m (Carrier Density over Effective Mass) in High-T c [24] Y. Wang, L. Li, and N. P. Ong, Nernst Effect in High-T c
Cuprate Superconductors, Phys. Rev. Lett. 62, 2317 Superconductors, Phys. Rev. B 73, 024510 (2006).
(1989). [25] M.-S. Nam, A. Ardavan, S. J. Blundell, and J. A. Schlueter,
[9] O. Gunnarsson, Alkali-Doped Fullerides: Narrow-Band Fluctuating Superconductivity in Organic Molecular Metals
Solids with Unusual Properties (World Scientific, Singapore, Close to the Mott Transition, Nature (London) 449, 584
2004). (2007).
[10] K. M. Kadish and R. S. Ruoff, Fullerenes: Chemistry, [26] P. Spathis, H. Aubin, A. Pourret, and K. Behnia, Nernst
Physics, and Technology (John Wiley & Sons, New York, Effect in the Phase-Fluctuating Superconductor InOx, Euro-
2000). phys. Lett. 83, 57005 (2008).
[11] C. M. Lieber and Z. Zhang, Physical Properties of Metal- [27] O. Cyr-Choinière, R. Daou, F. Laliberté, C. Collignon, S.
Doped Fullerene Superconductors, Solid State Phys. 48, Badoux, D. LeBoeuf, J. Chang, B. J. Ramshaw, D. A. Bonn,
349 (1994). W. N. Hardy, R. Liang, J.-Q. Yan, J.-G. Cheng, J.-S. Zhou,
[12] M.-Q. Ren, S. Han, S.-Z. Wang, J.-Q. Fan, C.-L. Song, J. B. Goodenough, S. Pyon, T. Takayama, H. Takagi, N.
X.-C. Ma, and Q.-K. Xue, Direct Observation of Full-Gap Doiron-Leyraud, and L. Taillefer, Pseudogap Temperature
Superconductivity and Pseudogap in Two-Dimensional T  of Cuprate Superconductors from the Nernst Effect,
Fullerides, Phys. Rev. Lett. 124, 187001 (2020). Phys. Rev. B 97, 064502 (2018).
[13] M. Mitrano, A. Cantaluppi, D. Nicoletti, S. Kaiser, A. [28] C. W. Rischau, Y. Li, B. Fauqué, H. Inoue, M. Kim,
Perucchi, S. Lupi, P. Di Pietro, D. Pontiroli, M. Riccò, S. R. C. Bell, H. Y. Hwang, A. Kapitulnik, and K. Behnia,
Clark, D. Jaksch, and A. Cavalleri, Possible Light-Induced Universal Bound to the Amplitude of the Vortex Nernst
Superconductivity in K3 C60 at High Temperature, Nature Signal in Superconductors, Phys. Rev. Lett. 126, 077001
(London) 530, 461 (2016). (2021).
[14] A. Cantaluppi, M. Buzzi, G. Jotzu, D. Nicoletti, M. Mitrano, [29] F. F. Tafti, F. Laliberté, M. Dion, J. Gaudet, P. Fournier, and
D. Pontiroli, M. Riccò, A. Perucchi, P. Di Pietro, and A. L. Taillefer, Nernst Effect in the Electron-Doped Cuprate
Cavalleri, Pressure Tuning of Light-Induced Superconduc- Superconductor Pr2−x Cex CuO4 : Superconducting Fluctu-
tivity in K3 C60 , Nat. Phys. 14, 837 (2018). ations, Upper Critical Field Hc2 , and the Origin of the T c
[15] M. Budden, T. Gebert, M. Buzzi, G. Jotzu, E. Wang, T. Dome, Phys. Rev. B 90, 024519 (2014).
Matsuyama, G. Meier, Y. Laplace, D. Pontiroli, M. Riccò [30] T. Yamashita, Y. Shimoyama, Y. Haga, T. Matsuda, E.
et al., Evidence for Metastable Photo-induced Supercon- Yamamoto, Y. Onuki, H. Sumiyoshi, S. Fujimoto, A.
ductivity in K3 C60 , Nat. Phys. 17, 611 (2021). Levchenko, T. Shibauchi et al., Colossal Thermomagnetic
[16] M. Buzzi, G. Jotzu, A. Cavalleri, J. I. Cirac, E. A. Demler, Response in the Exotic Superconductor URu2 Si2, Nat. Phys.
B. I. Halperin, M. D. Lukin, T. Shi, Y. Wang, and D. 11, 17 (2015).
Podolsky, Higgs-Mediated Optical Amplification in a Non- [31] A. v. Ettingshausen and W. Nernst, Ueber das Auftreten
equilibrium Superconductor, Phys. Rev. X 11, 011055 electromotorischer Kräfte in Metallplatten, welche von
(2021). einem Wärmestrome durchflossen werden und sich im
[17] A. Thomas, E. Devaux, K. Nagarajan, T. Chervy, M. Seidel, magnetischen Felde befinden, Ann. Phys. (Berlin) 265, 343
D. Hagenmüller, S. Schütz, J. Schachenmayer, C. Genet, G. (1886).
Pupillo, and T. W. Ebbesen, Exploring Superconductivity [32] K. Behnia, M.-A. Méasson, and Y. Kopelevich, Nernst
under Strong Coupling with the Vacuum Electromagnetic Effect in Semimetals: The Effective Mass and the Figure of
Field, arXiv:1911.01459. Merit, Phys. Rev. Lett. 98, 076603 (2007).
[18] Y. J. Uemura, Dynamic Superconductivity Responses in [33] J. Chang, N. Doiron-Leyraud, O. Cyr-Choiniére, G.
Photoexcited Optical Conductivity and Nernst Effect, Phys. Grissonnanche, F. Laliberté, E. Hassinger, J.-P. Reid, R.
Rev. Mater. 3, 104801 (2019). Daou, S. Pyon, T. Takayama, H. Takagi, and L. Taillefer,
[19] M. Buzzi, D. Nicoletti, S. Fava, G. Jotzu, K. Miyagawa, K. Decrease of Upper Critical Field with Underdoping in
Kanoda, A. Henderson, T. Siegrist, J. A. Schlueter, M. S. Cuprate Superconductors, Nat. Phys. 8, 751 (2012).
Nam, A. Ardavan, and A. Cavalleri, Phase Diagram for [34] N. Ong, Y. Wang, S. Ono, Y. Ando, and S. Uchida, Vorticity
Light-Induced Superconductivity in κ-ðETÞ2 -X, Phys. Rev. and the Nernst Effect in Cuprate Superconductors, Ann.
Lett. 127, 197002 (2021). Phys. (Leipzig) 516, 9 (2004).
[20] K. Behnia and H. Aubin, Nernst Effect in Metals and [35] A. Kondrat, G. Behr, B. Büchner, and C. Hess, Unusual
Superconductors: A Review of Concepts and Experiments, Nernst Effect and Spin Density Wave Precursors in Super-
Rep. Prog. Phys. 79, 046502 (2016). conducting LaFeAsO1−x Fx , Phys. Rev. B 83, 092507
[21] K. Behnia, The Nernst Effect and the Boundaries of the (2011).
Fermi Liquid Picture, J. Phys. Condens. Matter 21, 113101 [36] A. Pourret, L. Malone, A. B. Antunes, C. S. Yadav, P. L.
(2009). Paulose, B. Fauqué, and K. Behnia, Strong Correlation and
[22] I. Ussishkin, S. L. Sondhi, and D. A. Huse, Gaussian Low Carrier Density in Fe1þy Te0.6 Se0.4 as Seen from Its
Superconducting Fluctuations, Thermal Transport, and Thermoelectric Response, Phys. Rev. B 83, 020504(R)
the Nernst Effect, Phys. Rev. Lett. 89, 287001 (2002). (2011).

021008-12
SUPERCONDUCTING FLUCTUATIONS OBSERVED FAR ABOVE … PHYS. REV. X 13, 021008 (2023)

[37] M. S. Nam, C. Mézière, P. Batail, L. Zorina, S. Simonov, [51] A. Goldoni, L. Sangaletti, F. Parmagiani, S. L. Friedmann,
and A. Ardavan, Superconducting Fluctuations in Organic Z.-X. Shen, M. Peloi, G. Comelli, and G. Paolucci, Phase
Molecular Metals Enhanced by Mott Criticality, Sci. Rep. 3, Transition, Molecular Motions, and Inequivalent Carbon
3390 (2013). Atoms in K3 C60 (111) Single-Phase Ordered Films, Phys.
[38] S. Kasahara, T. Yamashita, A. Shi, R. Kobayashi, Y. Rev. B 59, 16071 (1999).
Shimoyama, T. Watashige, K. Ishida, T. Terashima, T. [52] K. Sugihara, T. Inabe, Y. Maruyama, and Y. Achiba,
Wolf, F. Hardy, C. Meingast, H. v. Löhneysen, A. Thermoelectric Power of Alkali Doped C60 , J. Phys. Soc.
Levchenko, T. Shibauchi, and Y. Matsuda, Giant Super- Jpn. 62, 2757 (1993).
conducting Fluctuations in the Compensated Semimetal FeSe [53] D. T. Morelli, Thermoelectric Power of Superconducting
at the BCS–BEC Crossover, Nat. Commun. 7, 12843 (2016). Fullerenes, Phys. Rev. B 49, 655 (1994).
[39] S. D. Kang and G. J. Snyder, Charge-Transport Model for [54] A. Pourret, H. Aubin, J. Lesueur, C. A. Marrache-Kikuchi,
Conducting Polymers, Nat. Mater. 16, 252 (2017). L. Bergé, L. Dumoulin, and K. Behnia, Length Scale for the
[40] X.-D. Xiang, J. G. Hou, V. H. Crespi, A. Zettl, and M. L. Superconducting Nernst Signal above T c in Nb0.15 Si0.85 ,
Cohen, Three-Dimensional Fluctuation Conductivity in Phys. Rev. B 76, 214504 (2007).
Superconducting Single Crystal K3 C60 and Rb3 C60 , Nature [55] Note that thepdefinition
ffiffiffi of the magnetic length includes a
(London) 361, 54 (1993). factor of 1= 2. This definition is chosen such that a
[41] N. R. Werthamer, E. Helfand, and P. C. Hohenberg, Temper- comparison of the magnetic length and the correlation
ature and Purity Dependence of the Superconducting length is equivalent to a comparison of the dimensionless
Critical Field, Hc2 . III. Electron Spin and Spin-Orbit field and dimensionless amplitude. See the experimental
Effects, Phys. Rev. 147, 295 (1966). work [54] and theoretical work [46,56] confirming this.
[42] Y. Kasahara, Y. Takeuchi, R. H. Zadik, Y. Takabayashi, [56] W. J. Skocpol and M. Tinkham, Fluctuations Near Super-
R. H. Colman, R. D. McDonald, M. J. Rosseinsky, K. conducting Phase Transitions, Rep. Prog. Phys. 38, 1049
Prassides, and Y. Iwasa, Upper Critical Field Reaches (1975).
90 Tesla Near the Mott Transition in Fulleride Super- [57] M. Soroka, K. Knížek, Z. Jirák, P. Levinský, M. Jarošová, J.
conductors, Nat. Commun. 8, 14467 (2017). Buršik, and J. Hejtmánek, Anomalous Nernst Effect in the
[43] D. Podolsky, S. Raghu, and A. Vishwanath, Nernst Effect Ceramic and Thin Film Samples of La0.7 Sr0.3 CoO3 Per-
and Diamagnetism in Phase Fluctuating Superconductors, ovskite, Phys. Rev. Mater. 5, 035401 (2021).
Phys. Rev. Lett. 99, 117004 (2007). [58] M. Serbyn (private communication).
[44] K. Michaeli and A. M. Finkel’stein, Fluctuations of the [59] G. Blatter, M. V. Feigel’man, V. B. Geshkenbein, A. I.
Superconducting Order Parameter as an Origin of the Larkin, and V. M. Vinokur, Vortices in High-Temperature
Nernst Effect, Europhys. Lett. 86, 27007 (2009). Superconductors, Rev. Mod. Phys. 66, 1125 (1994).
[45] M. N. Serbyn, M. A. Skvortsov, A. A. Varlamov, and V. [60] S. Hoshino, P. Werner, and R. Arita, Unconventional
Galitski, Giant Nernst Effect Due to Fluctuating Cooper Orbital Ordering and Emergent Dimensional Reduction
Pairs in Superconductors, Phys. Rev. Lett. 102, 067001 in Fulleride Superconductors, Phys. Rev. B 99, 235133
(2009). (2019).
[46] A. Glatz, A. Pourret, and A. A. Varlamov, Analysis of the [61] V. J. Emery and S. A. Kivelson, Importance of Phase
Ghost and Mirror Fields in the Nernst Signal Induced by Fluctuations in Superconductors with Small Superfluid
Superconducting Fluctuations, Phys. Rev. B 102, 174507 Density, Nature (London) 374, 434 (1995).
(2020). [62] V. Buntar, F. M. Sauerzopf, and H. W. Weber, Lower
[47] T. Inabe, H. Ogata, Y. Maruyama, Y. Achiba, S. Suzuki, K. Critical Fields of Alkali-Metal-Doped Fullerene Super-
Kikuchi, and I. Ikemoto, Electronic Structure of Alkali conductors, Phys. Rev. B 54, R9651 (1996).
Metal Doped C60 Derived from Thermoelectric Power [63] I. Ussishkin and S. L. Sondhi, On the Interpretation of the
Measurements, Phys. Rev. Lett. 69, 3797 (1992). Nernst Effect Measurements in the Cuprates, Int. J. Mod.
[48] L. Lu, V. H. Crespi, M. S. Fuhrer, A. Zettl, and M. L. Cohen, Phys. B 18, 3315 (2004).
Universal Form of Hall Coefficient in K and Rb Doped [64] A. L. Woodcraft and A. Gray, A Low Temperature Thermal
Single Crystal C60 , Phys. Rev. Lett. 74, 1637 (1995). Conductivity Database, AIP Conf. Proc. 1185, 681
[49] R. Bel, K. Behnia, and H. Berger, Ambipolar Nernst Effect (2009).
in NbSe2 , Phys. Rev. Lett. 91, 066602 (2003). [65] O. Klein, G. Grüner, S.-M. Huang, J. B. Wiley, and R. B.
[50] Y. Yoshinari, H. Alloul, G. Kriza, and K. Holczer, Molecu- Kaner, Electrical Resistivity of K3 C60 , Phys. Rev. B 46,
lar Dynamics in K3 C60 : A C13 NMR Study, Phys. Rev. Lett. 11247 (1992).
71, 2413 (1993).

021008-13

You might also like