You are on page 1of 13

Fragile superconductivity in the presence of weakly disordered charge density waves

Yue Yu and S. A. Kivelson1


1
Department of Physics, Stanford University, Stanford, CA 94305, USA
(Dated: September 28, 2018)
When superconducting (SC) and charge-density wave (CDW) orders compete, novel low temper-
ature behaviors can result. From an analysis of the Landau-Ginzberg-Wilson theory of competing
orders, we demonstrate the generic occurrence of a “fragile” SC phase at low temperatures and high
fields in the presence of weak disorder. Here, the SC order is largely concentrated in the vicinity of
dilute dislocations in the CDW order, leading to transition temperatures and critical currents that
arXiv:1809.10160v1 [cond-mat.supr-con] 26 Sep 2018

are parametrically smaller than those characterizing the zero field SC phase. This may provide the
outline of an explanation of the recently discovered “resilient” superconducting phase at high fields
in underdoped YBa2 Cu3 O6+δ .

I. INTRODUCTION • Firstly, the magnitude of the primary order must


vanish – or at least must be substantially sup-
There are a variety of microscopic circumstances in pressed - in the core of the defect. This is generic
which correlated electronic materials exhibit compara- in the vortex core of an order parameter with XY
bly strong, and macroscopically competing tendencies to- symmetry, or a domain wall of an order parameter
ward superconducting (SC) and incommensurate charge- with Ising symmetry. It would not be expected, for
density wave (CDW) orders.1 A particularly interesting instance, in a skyrmion of an order parameter with
aspect of the interplay between the two orders is the role Heisenberg symmetry.
of topological defects – vortices in the SC, which can be • Secondly, the elastic “stiffness” of the secondary
induced by the application of an external magnetic field, order parameter, κ, must be sufficiently small, so
H, and dislocations in the CDW, which are produced that the gain in condensation energy from having
by quenched disorder. In cases in which a subdominant the secondary order expressed in the defect core
order is nearly degenerate with the dominant order, the exceeds the elastic cost of having a spatially vary-
subdominant order can be stabilized in the neighbor of ing order parameter. (Note, the critical value of κ
a topological defect where the dominant order vanishes. depends, among other things, on how close is the
The possibility of one or another form of density wave balance is between the primary and secondary or-
order in a vortex core “halo” has been the subject of con- dering tendencies.)
siderable theoretical and experimental interest, especially
in the context of the underdoped cuprate high tempera- Under these circumstances, there exists a second crit-
ture superconductors. Here we explore some qualitative ical value, γc2 (κ) > γc1 such that for γc2 > γ > γc1 ,
new physics that arises in regimes where CDW order is there occurs a finite halo region about any topological
dominant, and in which SC arises in the neighborhood of defect of the primary order parameter in which the sec-
CDW dislocations. ondary order parameter has a non-vanishing amplitude.
There are many examples of this basic physics that have
been discussed. Superconducting cosmic strings2,3 are
an example, in which the superconducting order is the
A. Topological defects and halos secondary order that appears associated with vortices
in a dominant cosmic condensate. Subdominant CDW,
In systems with complex phase diagrams, there is often spin-density wave (SDW), or nematic orders arising in
a range of temperatures, T , in which – in the absence of halos about magnetic field induced vortices in a dom-
competition – two distinct symmetries would be spon- inantly superconducting order have been theoretically
taneously broken, but where the competition between discussed4–13 , and experimentally observed14–17 , in var-
the two orders is strong enough that only the dominant ious cuprate superconductors. Some evidence has been
order parameter develops a non-zero expectation value; presented that in certain Fe-pnictides, superconducting
the subdominant order is quenched by the competition. order can exist in a narrow range of T above the bulk Tc
This state of affairs pertains when the repulsion between in a region about a structural twin boundary - i.e. Ising
the two orders, which corresponds to the term γ in the domain walls of a nematic order parameter.18
Landau-Ginzburg free energy in Eq. 3.1, exceeds a crit- In the present paper, we will focus on the case of com-
ical strength, γc1 . (See Figs. 1(a) and 1(b).) However, petition between CDW and SC order. Where the SC
under some circumstances (that we will derive explicitly), order is dominant, the topological defects in question are
the subdominant order emerges in a “halo” of finite ex- the familiar vortices already mentioned, and the halo is
tent surrounding any topological defect in the primary then a region with local CDW order. Here, at low T
order. Two features of the system are required for this in a strongly type II superconductor, the density of vor-
to occur: tices is controlled by an applied magnetic field, H, and
2

a low density of vortices can be introduced by the appli- Landau-Ginzburg effective field theories we will analyze,
cation of a small field with Hc1 < H  Hc2 . Conversely, J(R) ∼ J0 exp[−R/`], where ` is an appropriate correla-
where the CDW order is dominant, the topological de- tion length; in more microscopically realistic treatments
fects are dislocations. Because there is an emergent XY of metallic systems at low T , this dependence can be
symmetry associated with an incommensurate CDW, at a much more complicated.19,20 Tc can be estimated from
formal level these defects are precisely dual to the super- the implicit solution of the equation χ(Tc )J(R) ∼ 1.
conducting vortices, with the role of CDW and SC orders C. Plan of the paper
interchanged. Now, the halo is a region of local SC or-
der in the vicinity of a dislocation core. No precise dual
Before launching into specific calculations, in Sec. II
relation exists in the presence of a magnetic field, which
we present some representative phase diagrams that can
for all practical purposes couples only to the SC order.
emerge as a consequence of the competition between
However, quenched disorder plays a somewhat analogous
CDW and SC order. Next, in Sec. III, we define a
role; in a quasi-2D system, weak disorder (greater than
model classical effective field theory with fields corre-
a parametrically small value that vanishes as interplane
sponding to unidirectional CDW and SC order. In Sec.
couplings tend to 0) produces dilute, randomly pinned
IV we treat this model in the context of Landau-Ginzberg
dislocations in the CDW order.
theory, meaning that we treat the model as an effective
free energy (with coefficients that are assumed to depend
B. Halos and Long-Range-Order smoothly on T ) and we solve for the field configurations
that minimize it.
Until now, we have focussed on local (or mean-field) In order to explore the effects of oder parameter fluc-
considerations. An isolated vortex or dislocation is a 1D tuations more seriously, in Secs. V and VI we treat
object (or a 0D object in a 2D system), so an isolated vor- the same model as an effective Hamiltonian (i.e. in the
tex halo cannot give rise to a broken symmetry.10 . Unless context of Landau-Ginzberg-Wilson theory), with fixed
inter-halo interactions are taken into account, thermal (temperature independent) parameters. Specifically, in
fluctuations destroy any long-range-order (LRO) that one Sec. V we treat the fluctuations of the fields approx-
would have associated with the order parameter halo. imately using Feynman’s variational approach21 (which
Thus, LRO (if it occurs) is triggered by the coupling be- is exact in a suitable large N limit of the same model22 )
tween halos. As a result, for a system of dilute halos, and in Sec. VI we treat the fluctuations exactly (numer-
there are typically two parametrically distinct character- ically) using classical Monte-Carlo methods.
istic temperatures: a relatively high crossover tempera- Finally, in Sec. VII we generalize the discussion slightly
ture Thal below which the halos are locally well formed to the case of a tetragonal system (where the CDW order
and a lower critical temperature, Tc , at which LRO of the can have its ordering vector in one of two symmetry re-
subdominant order parameter onsets. This lower tem- lated directions) in order to obtain a phase diagram that
perature depends on the nature of the coupling between is suggestive of observed behaviors of the cuprate high
neighboring halos (such as is it frustrated or not) and temperature superconductors. We then comment on in-
thus is less universal. sights one can obtain concerning existing observations in
For the problem at hand, we will treat these issues the cuprates that can be qualitatively well understood
by explicit solution of a simple effective field theory. in terms of the present considerations. In particular, in
However, to develop a general intuition, we can esti- the interest of brevity, we focus on YBCO with doped
mate Tc as follows: 1) Compute the order parameter hole density roughly in the range 0.1 to 0.14, which we
susceptibility of an isolated halo, χ(T ). For instance, identify as a regime in which the principle features of
the susceptibility associated with a CDW halo living the phase diagram are a consequence of the fierce com-
along a vortex line is that of√ an appropriate classi- petition between SC and CDW orders. While in this
cal 1D XY model, χ(T ) ∼ 1/ T ? T (where T ? is the paper, we have considered only problems in which there
mean-field Tc ). 2) Compute the effective coupling be- is a close competition between CDW and SC order, very
tween neighboring halos, J(R), which naturally depends similar considerations apply to other cases in which mul-
strongly on the distance, R, between halos. In the simple tiple orders are intertwined.23–36
II. QUALITATIVE PHASE DIAGRAMS planes. Thus, the CDW and zero field SC phases exhibit
true LRO and the vortices induced by a finite B form an
Abrikosov lattice or (in the presence of disorder) a vortex
Many qualitative aspects of our results are represented glass up to non-zero temperatures.37,38
in the schematic phase diagrams in Fig. 1, which show
a variety of ways the competition between CDW and SC In the following, γ represents the strength of the re-
order plays out; these aspects can be motivated inde- pulsion between the CDW and SC orders, σ is a measure
pendently of the specific method of solution. So as to of the strength of the disorder. Explicit definitions are
avoid subtleties that are special to strictly 2D systems, given when we define the effective field theory in Eqs.
it is assumed that the system in question is a layered, 3.1 and 3.3, below. While at a microscopic level, the
quasi-2D system with weak uniform couplings between electronic structure is changed in a way that affects the
3

Tcdw Tcdw
γc2 > γ > γc1 γ > γ3c
Η σ = 0 Η σ = 0
Charge-Density-Wave Charge-Density-Wave
(CDW) (CDW)

SC +
Fragile
CDW
Thal
SC + Superconductor
CDW SRO Superconductor
In vortex Tc Tc
halos

Τ Τ

(a) (b)

F
γc2 > γ > γc1 Nematic with
Η R CDW SRO F
σ ≠ 0 unidirectional
A Η R
G CDW SRO
A
I SC SRO (nematic near
G
L S - dislocation to smectic)
I
E C L S halos
Weak bidirectional
E C CDW SRO
- disorder pinned


F N

R E
Superconductor
A M Superconductivity
SC +
G A
I T (zero resistance)
CDW SRO L
in vortex E CDW SRO

halos - vortex halos

Τ*(σ) Τ Τ*(σ) Τ

(c) (d)

FIG. 1: Schematic phase diagrams showing competition between CDW and SC order in the absence of disorder (a and b) and
in the presence of weak disorder (c and d). For reference, the dotted lines represent what would be the phase boundaries in the
absence of coupling between the SC and CDW orders (γ = 0). The thin solid lines represent continuous and the heavy purple
line a discontinuous transitions. The dashed lines indicate crossovers. For further explanation Sec. II.

tendency toward both SC and CDW order (especially in appropriate circumstances, there is a second critical value
the case of “unconventional” SC), at the order parame- γc2 > γc1 such that at T = 0, a CDW halo forms about
ter level, gauge invariance precludes any linear (“random an isolated vortex (which is generated by an infinitesimal
field”) coupling between disorder and the SC order pa- non-zero H) for γc2 > γ > γc1 but not for γ > γc2 . We
rameter, so in our model, disorder couples directly only will also encounter a third critical value γc3 > γc2 , which
to the CDW order parameter. (It indirectly affects SC is the smallest value of γ such that for any γ > γc3 there
through the local competition with CDW order.) In this is no range of T and H in which CDW and SC coexist.
section, we will consider explicitly only the case in which γ = σ = 0: The dotted lines in all four panels of Fig.
in the zero field limit, the SC ordering tendency is slightly 1 delineate the phase boundaries in the clean limit and
stronger than the CDW. in the absence of coupling between the order parameters.
As a function of increasing γ, there are a variety of In this limit, at temperatures below the red dotted line
qualitatively distinct regimes possible in the clean limit, (which is vertical since B does not couple significantly to
σ = 0, some of which we will outline here. As already CDW order), there is CDW LRO. The blue dotted line is
mentioned, γc1 is defined such that at H = 0, CDW and the boundary of the SC (Abrikosov lattice) phase. The
SC order coexist at low enough T so long as γc1 > γ, two phase boundaries cross at a decoupled tetracritical
while no such coexistence occurs for γ ≥ γc1 . Under point, and there is a broad region of coexisting SC and
4

CDW order. The assumption that the two orders have would occur robustly. Similarly, no first order transi-
comparable strength is reflected in the relatively small tion is allowed in d ≤ 2, and correspondingly it is likely
magnitude of Tc − Tcdw . that even rather weak disorder eliminates the possibility
γ > 0 with σ = 0: For small γ > 0, the coexistence of such a transition in a quasi 2d system. Thus, this tran-
regime persists, but its area is reduced. In Fig. 1(a) sition is continuous. At long distances, the SC phase is
we show a representative phase diagram for σ = 0 and a “vortex glass,” a phase with non-vanishing long-range
γc2 > γ > γc1 (where γc1 and γc2 were defined in Sec phase rigidity and (to the extent that gauge-field fluctu-
I A, above). Because γ > γc1 , the SC order prevents ations can be ignored) a superconducting version of an
any CDW LRO at H = 0; however, because γ < γc2 , Edwards-Anderson order parameter.37
at small but non-zero H, CDW halos form around each
vortex below the crossover temperature, Thal , shown as The upturn of the SC phase boundary at low T is the
the vertical dashed green line in the figure. CDW LRO in most notable new aspect of this phase diagram. This
the coexistence phase occurs below the solid blue line; it upturn is the promised consequence of the presence of
is highly inhomgeneous at low magnetic fields, triggered topological defects. In the entire range of field and tem-
by the (relatively weak) coupling from one vortex halo to peratures between the blue dotted line and the solid black
the next, and hence is labeled “fragile.” At higher fields, line, it is only the presence of strong, local CDW corre-
the balance between CDW and SC is reversed, with the lations that is quenching the SC order. Thus, so long
result that there arises a pure CDW phase with uniform the superconducting coherence length is not too long
order. Note that at low temperatures, the relatively low (κ < κc ), a locally superconducting halo forms about
value of Hc2 is a consequence of competition with the CDW dislocations where the CDW order vanishes. In-
CDW rather than a property of the SC state itself. evitably, a form of granular SC correlations is generated.
For large enough γ > γc3 > γc2 , the coexistence phase At elevated temperatures, the Josephson coupling be-
is entirely quenched, as shown in Fig. 1(b). The thick tween neighboring dislocation halos is small compared to
purple line delineates a first order field driven SC to CDW T , so there are at best extremely weak macroscopic signa-
transition which ends in a bicritical point at which the tures of these SC correlations. However, global SC phase
phase boundaries of the pure SC (solid blue line) and coherence onsets below a disorder dependent character-
CDW phases (red solid line) meet. istic temperature, T ? (σ), comparable to the Josephson
In fact, there are many other possible topologies of this coupling, J between neighboring dislocations. Note that
phase diagram depending on parameters: For instance, the resulting SC phase is “fragile,” both in the sense that
even for γ > γc3 , rather than a tetracritical point, one it is destroyed by thermal fluctuations at temperatures
can have a first order line that extends to higher tem- far smaller than the zero field Tc , but also in that the
peratures than either of the ordered phases. For smaller critical current of the SC state is set by J(ξcdw ), and so
values of γ, it is possible to have more than one multicrit- is also increasingly small the weaker the disorder.
ical point – for instance, at the high temperature side of
the phase diagram, the SC and CDW phase boundaries Tetragonal crystal with γc2 > γ > γc1 and σc  σ > 0:
could meet at a bicritical point, but then at lower T a In the interest of simplicity, all of the model calculations
region of two-phase coexistence could arise in one of sev- we perform in the present paper are for the case of an
eral possible ways. In the case in which γc1 > γ > 0, the orthorhombic crystal, where the direction of the CDW
phase boundary between the SC phase and the coexis- order is uniquely determined. However, here and in the
tence phase can hit the H = 0 axis at a non-zero critical conclusion, we include a brief discussion of how the phase
temperature, while for γc3 > γ > γc2 (assuming that the diagram differs in a tetragonal crystal where even if the
initial tetracritical point remains stable) this same phase preferred CDW order is unidirectional, it can condense
boundary will hit the T = 0 axis at a non-zero critical in either of at least two symmetry related directions.75
value of the magnetic field. In this case, for σ = 0, CDW LRO breaks not only
γc2 > γ > γc1 and σc  σ > 0: Even when we entirely translation symmetry, but also a discrete rotational (or
ignore the effects of pair-breaking, when the disorder mirror) symmetry of the crystal. For non-zero σ, the
strength exceeds a critical value, σc , the locally pinned translation symmetry aspect of the CDW order is lost,
CDW correlations are sufficiently strong that SC cannot but the discrete point-group symmetry breaking aspect
arise - for σ > σc , no broken symmetry phases exist at survives up to a critical disorder strength. Thus, in this
any T . case, as illustrated in Fig. 1(d), even for non-zero σ there
The solid black curve in Fig. 1(c) shows the only can be a well defined thermodynamic phase transition
true phase boundary in the presence of weak but non- to a state with “vestigial nematic order,” that is to say
vanishing disorder, σc  σ > 0. It is a SC to non- in which there is long-range CDW orientational order
superconducting (“normal”) transition. In any dimen- without positional order.39 Consequently, what appear
sion d ≤ 4, the presence of weak disorder precludes the as CDW crossover lines in the orthorhombic case shown
existence of CDW LRO. A Bragg-glass phase with CDW in Fig. 1(c), become the red solid phase boundaries in
quasi LRO is in principle possible in d = 3, but for a Fig. 1(d) in the tetragonal case. We will return to this
quasi 2d system, it seems unlikely that such a phase case in the final section.
5

III. THE MODEL zero field order, but in which CDW is only weakly sub-
dominant, we will assume 1  1 − αcdw > 0. We will
To make the discussion explicit, and since the qualita- also assume that the repulsion between the two orders is
tive behavior we are exploring is relatively insensitive to reasonably strong, γ ∼ 1.
microscopic details, we begin by considering the prop- Ideally, we would treat the
erties of a minimal classical Landau-Ginzberg-Wilson XZ
(LGW) effective field theory. We consider two complex S=β d~r [H2d + H⊥ ] (3.4)
scalar fields, a charge 2e field ∆ representing the local n
SC order parameter, and φ representing the amplitude as the effective action where e−S gives the Boltzman
of a unidirectional incommensurate CDW. (Generalizing weight for each field configuration. Below we carry this
the considerations to bidirectional CDW order, or mul- out in several approximate ways.
ticomponent SC orders is straightforward, although not
entirely without qualitatively new features.) We will fo-
cus on a quasi-2D limit in which there are only weak IV. SADDLE POINT SOLUTION –
couplings between neighboring 2D layers of a 3D crystal LANDAU-GINZBURG THEORY
- we will discuss explicitly the field theory for a single
plane, and will leave implicit the weak uniform couplings The field configurations that minimize S determine the
between the order parameters on neighboring planes. classical ground-state (T = 0) configurations of the fields.
The effective free energy density of a plane is (More generally, if we treat S as the Landau-Ginzburg
4 (LG) free energy, in which the coefiscients α∆ and αφ and
~ ∆ − α∆ |∆|2 + |∆|
κ  ~  2
H2d = −i∇ − 2eA possibly others are taken to be T dependent, minimizing
2 2 4 S corresponds to LG mean field theory.)
4
κx 2 κ y 2 α φ 2 |φ|
+ |∂x φ| + |∂y φ| − |φ| +
2 2 2 4
γ 2 2 ? ? 1. σ=0
+ |∆| |φ| + h φ + φ h. (3.1)
2
The weak couplings between order parameters (φn and In the following discussion, we focus on γc1 < γ < γc2 .
∆n in neighboring planes (n and n + 1), are taken to be For H = 0, the ground-state has SC order and vanish-
ing CDW order. In this case, the CDW order develops,
Jφ ? J∆ ? and increases in strength with increasing magnetic fields.
H⊥ = − [φ φn +H.C.]− [∆n+1 ∆n +H.C.]. (3.2) However, the magnetic field produces a vortex crystal
2 n+1 2
and suppresses the overall SC order - most strongly in
This is at best an effective Hamiltonian, obtained by inte- vortex cores. Needless to say, the CDW order develops
grating out microscopic degrees of freedom, which means a spatially varying magnitude in response to the field-
that even the parameters that enter the model should induced vortex lattice. No analytic solution of this mod-
properly be taken to have analytic dependences on tem- ulated state exists.
perature, T , and magnetic field, H~ =∇ ~ × A.
~ Moreover, However, the state is increasingly homogeneous as H
in general there should be higher order terms, both in approaches the upper critical field, Hc2 , defined such that
powers of the fields and in powers of derivatives, and at SC order is quenched for H > Hc2 :
low enough temperatures, the quantum dynamics of the (0)
fields can be important. Hc2 = Hc2 [1 − γαφ /α∆ ] (4.1)
Without loss of generality, we have chosen units of the (0)
two order parameters to set the coefiscients of |φ|4 , |∆|4 where Hc2 = (2e)−1 α∆ is the value of Hc2 in the absence
to 1. Moreover, we can chose units of distance and energy of competition with the CDW.
so that α∆ = κ = 1. Because φ is a unidirectional CDW,
κx and κy need not be equal, even in a tetragonal system, (0)
2. Hc2 > H > Hc2 , and σ > 0
but for simplicity we will set κx = κy = 1. In the above,
h is a gaussian random complex field representing the
effect of quenched disorder, with In the presence of weak disorder, dilute pinned dislo-
cations disrupt the CDW order, opening the possibility
h(~r) = 0 h(~r)h(~r0 ) = 2σ 2 δ(~r − ~r0 ), (3.3) of local superconductivity. We now consider the problem
of superconductivity at an isolated CDW dislocation. If
the overline represents the configuration average, and σ there is non-trivial solution for this problem, long-range
characterizes the strength of the disorder. Importantly, superconductivity will always be developed at low enough
the magnetic field couples directly only to the super- temperature up to a critical magnetic field H̃c2 > Hc2 .
conducting order, while the disorder couples only to the Around the critical magnetic field, ∆ is small, so the
CDW order. Since we will in particular focus primarily CDW profile can be firstly solved while neglecting su-
on the case in which superconductivity is the dominant perconductivity. Superconductivity can then be studied
6

treating the CDW as a fixed potential. Neglecting all ∆


1.2
terms, the simplified Hamiltonian for CDW is
κφ αφ 2 1 4 1
H[φ] = |∇φ|2 − |φ| + |φ| . (4.2)
2 2 4
0.8
We assume an ansartz for an isolated dislocation, with
the following boundary condition 0.6

φ(r, θ) = αφ f (r)eiθ 0.4
f (r → ∞) = 1 (4.3)
0.2
f (r → 0) = 0.
0
The solution f (r) can be obtained numerically, as shown 0 0.2 0.4 0.6 0.8 1
in Fig.2.pThe characteristic length of the dislocation is
Rcdw = κφ /αφ .
FIG. 3: Upper critical fields Hc2 (in the absence of disorder)
and H̃c2 (in the presence of dilute, disorder induced disloca-
1 tions) as a function of γ as computed from solution of the
(0)
Landau-Ginzburg equations. Here Hc2 is the value of the
0.8 critical field in the absence of competition with CDW order
(γ = 0). Here we have taken αφ = 0.95 α∆ and κ∆ = κφ = 1.
0.6

0.4
The fact that for non-zero γ, H̃c2 > Hc2 implies a strik-
ingly non-analyticity behavior of Hc2 (σ). Specifically,
0.2 since we expect a non-zero concentration of dislocations
for any non-zero value of σ,
0
0 2 4 6 8
r/R lim Hc2 (σ) = H̃c2 > Hc2 (σ = 0) ≡ Hc2 . (4.6)
σ→0

FIG. 2: The magnitude of CDW order inside a single dislo-


cation as a function of the distance from the center obtained With increasing σ, the concentration of dislocations in-
p minizing H in Eq. 4.2. The characteristic
from numerically creases, and consequently one expects a range in which
length is Rφ = κφ /αφ . Hc2 (σ) is an increasing function of σ as the halos begin to
overlap. However, as we will see below, for large enough
Near the critical field, where SC is weak, the back reac- σ, there is no superconductivity even at H = 0, which
tion of the SC correlations on the form of the dislocation implies the existence of a critical disorder strength, σc
can be ignored. Thus, the saddle point equation for ∆ is such that Hc2 (σ) → 0 as σ → σc from below. (See Fig.
4)
Ô∆ = −(1/2) |∆|2 ∆ (4.4)
2 2
 
Ô ≡ −κ∆ (−i∇ − 2eA) − α∆ + (γαφ )f .

In order to have a nontrivial solution, the operator Ô


must have at least one negative eigenvalue. This problem V. VARIATIONAL SOLUTION
is equivalent to the Schrodinger Eq. for a charged particle
in a magnetic field H and a potential V (r) = γαφ |f (r)|2 .
If we work in symmetric gauge, the solutions can be A first step beyond LG theory is obtained by treating
classified by their out-of-plane angular momentum, m, fluctuations approximately using the Feynman approach,
with the lowest energy solution lying in the m = 0 sector. in which we introduce a quadratic trial Hamiltonian, with
In this sector parameters optimized according to a variational princi-
ple. To treat the disorder averages properly, we introduce
Ôm=0 = −κ∆ ∇2 − α∆ + γαφ f 2 + κ∆ e2 B 2 r2 (4.5) n replicas of the theory, evaluate the trace over disorder
fields h as if they were in equilibrium, and then take the
The critical field H̃c2 can be obtained numerically as the limit as n → 0. (Details of this procedure are reported in
point at which the lowest eigenvalue vanishes. Represen- Ref. 39.) The replicated trial Hamiltonian density in the
tative results are shown in Fig.3; as expected, H̃c2 > Hc2 . normal phase (i.e. in the absence of broken symmetry)
7

is then of the form The variational approach can be extended to the or-
n h
dered phase by adding an explicit symmetry breaking
X κφ µφ field. For instance, for the case of an ordered CDW (in
Htr = |∇φa |2 +|φa |2
2 2 the absence of disorder), a term of the form m ?
2 [φ + φ ]
a
κ∆ µ∆ i must be added to Htr . More about applying this more
+ |(−i∇ − 2eA)∆a |2 + |∆a |2 (5.1) general approach to the present problem is presented in
2 2
Xn the Appendix.
− βσ 2 φ?a φb .
ab
A. Critical value of σ at T = 0
where, after taking the n → 0 limit, the self-consistency
for the variational parameters, µ∆ and µφ , become
As already mentioned, non-zero disorder always pre-
(N + 2) cludes CDW LRO. From Eq. 5.2 it follows that µ∆ de-
µφ = −αφ + h|φ1 |2 i + γh|∆1 |2 i creases monotonically as a function of increasing disorder
N (5.2) for fixed T and H. Thus in the variational solution, CDW
(N + 2)
µ∆ = −α∆ + h|∆1 |2 i + γh|φ1 |2 i disorder is always harmful for superconductivity. (This
N is not unexpected, as the replica trick tends to suppress
and where N = 2. (We have introduced N to signify the the effect of the rare events that are expected to lead
number of components of the order parameter fields, as to the paradoxical enhancement of superconductivity by
there is an interesting large N limit in which the varia- disorder at low T and moderate H.) The dependence of
tional approach becomes exact, but we will always work the T = 0 critical field as a function of disorder, Hc2 (σ),
with N = 2 in the following.) These equations are valid that results from the present analysis is shown in Fig. 4.
so long as no symmetries are broken. The expectation Note that there is a critical disorder strength, σc , where
values that enter these equations are taken with respect even though long-range CDW correlations have been en-
to the trial Hamiltonian, and so depend (in a complicated tirely destroyed, the short-range amplitude of the pinned
non-linear manner) on the variational parameters. CDW order is sufficiently strong that it quenches super-
The mean-square density wave fluctuations are the conductivity, even at T = 0 and H = 0.
sum of two terms - the first reflecting the disorder induced
fluctuations and the latter the thermal fluctuations: 0.6
d3 k
Z
2 2 1
h|φ1 | i = 4σ 0.5
(2π)3 (µφ + κφ kk2 + 2Jφ cos k⊥ )2
(5.3) 0.4
d3 k
Z
1
Hc2 /H (0)
c2

+ 2T
(2π)3 µφ + κφ kk2 + 2Jφ cos k⊥ 0.3

In the absence of disorder, the CDW transition temper- 0.2


ature, Tcdw , is the point at which µφ → 2Jφ . However,
notice that the disorder induced fluctuations diverge as 0.1
µφ → 2Jφ , which (correctly) encodes the fact that for
σ 6= 0, CDW order is precluded in d ≤ 4. The mean 0
0 0.2 0.4 0.6 0.8 1
square superconducting fluctuations do not depend ex-
/
plicitly on the disorder, but are different in the presence c
and absence of a magnetic field. For B = 0,
FIG. 4: Variational result for Hc2 as a function of σ. Here
(0)
d3 k
Z
1 Hc2 is the value of the critical field in the absence of com-
h|∆1 |2 i = 2T 3 2 (5.4) petition with CDW order (γ = 0), σc is the zero field crit-
(2π) µ∆ + κ∆ kk + 2J∆ cos k⊥
ical disorder from Eq. 5.7, and we have set αφ = 0.95 α∆ ,
γ = α∆ = 1 and κ∆ = κφ = 1.
while in the presence of a magnetic field, it can be ex-
pressed as a sum over Landau levels
Specifically, at T = 0, the self-consistent equation sim-
h|∆1 |2 i = plify to
∞ Z
2T Be X dk⊥ 1 µφ = −αφ + 2h|φ1 |2 i
π p=0 (2π) µ∆ + 4eBκ∆ (p + 1/2) + 2J∆ cos k⊥
µ∆ = −α∆ + γh|φ1 |2 i
(5.5)
d3 k
Z
The value of Tc is extracted from the self-consistency 2 2 1
h|φ1 | i = 4σ .
equations as the point at which µ∆ → 2J∆ − 2eκ∆ B. (2π)3 (µφ + κφ kk2 + 2Jφ cos k⊥ )2
8

Since we are working on zero temperature, the thermal it does not give analytic insight, and requires specific
fluctuations can be neglected. The solution for these choices of model parameters, but it does allow us to ver-
equations are ify the qualitative validity of the approximate analytic
results discussed above. In the following we set κ∆ = κφ ,
d3 k
Z
2 1
µφ = −αφ + 8σ 2
αφ = 0.95 α∆ , and γ = α∆ . Since the calculations are
(2π) (µφ + κφ kk + 2Jφ cos k⊥ )2
3
carried out in 2D, this corresponds to setting the inter-
γ plane couplings, J∆ and Jφ = 0.
µ∆ = −α∆ + (µφ + αφ )
2 To permit numerical solution, we discretize the con-
(5.6) tinuum Hamiltonian, but with lattice constant (which is
The critical disorder strength σc can be extracted from not physically meaningful) chosen smaller than the co-
µ∆ → µ∆,c ≡ 2J∆ − 2eκ∆ H. The critical disorder herence lengths. The simulation is set on a square lattice
strength can be calculated with (Lx , Ly ) = (20, 100) and periodic boundary condi-
tions. The vector potential A is set to be (−By, 0). The
π q
Monte Carlo simulation is performed for 400000 steps,
σc2 = κφ (αφ + µφ,c ) µ2φ,c − µ2∆,c
2 and measurement is done on the last 320000 steps. For
2α∆ + µ∆,c (5.7)
µφ,c = − αφ . systems with disorder, the whole process is averaged over
γ 60 different disorder configurations.
At H = 0, the value of σc extracted in this way agrees We measure the following quantities:
well with the results from the Monte Carlo calculations
~ ≡ 1 X
in the next section. GX (R) |hXi Xi0 i|, (6.1)
Lx Ly i

B. Hc2 in the absence of disorder where Xi is the SC or CDW order parameter on each site
i. i0 is chosen for each i, such that the displacement be-
The phase boundaries in the limit σ = 0 are somewhat tween them is R, ~ h i represents the thermal average and
more complicated to derive. In particular, the Hc2 line the overline represents the average over disorder config-
over much of the range of T separates a non SC phase urations. Note that for each value of the magnetic field,
with CDW LRO from a phase with coexisting SC and the lattice constant is chosen to allow periodic bound-
CDW LRO. However, while for H < Hc2 , the coexisting ary conditions. We choose |R| ~ to be the largest possible
phase is a spatially inhomogeneous Abrikosov lattice, the distance between two points, in the system with smallest
Hc2 line itself can be approached from above, where the lattice constant (i.e. largest magnetic field). We first per-
system remains spatially uniform. Thus the only new form thermal averaging to obtain the correlation function
complication is that this involves treating the problem over sites i and i0 . Then we take average of the absolute
in the presence of a broken symmetry, as discussed in value of the correlation function for all the pairs of sites.
the Appendix. Representative results of this analysis are We now compare the phase diagram for zero disorder
shown as in Fig. 5. and σ = 0.05σc , where σc is defined in Eq. 5.7 for H=0.
The nature of the multicritical points that occur in this Here we plot the order parameter h∆i as a function of
limit is still more subtle. Often even when the mean-field temperature and magnetic field, as shown in Fig.5. As
phase diagram suggests a single tetra-critical point, the predicted, disorder on CDW suppresses the superconduc-
variational approach yields weakly first order transitions tivity critical temperature, while increasing the critical
and a still more complex phase diagram. This is likely magnetic field at low T .
unphysical, and in any case affects only a very limited
range around the multicritical point. For this reason,
in presenting the results of the variational calculation in VII. CONCERNING THE CUPRATE PHASE
Fig. 5, we have ignored these subtleties, and have in- DIAGRAM
stead presented only the phase boundaries derived under
the assumption that all transitions are continuous. For In a complex material such as the cuprates, various
reasons discussed in the Appendix, these results are com- microscopic and material specific features always com-
puted in the presence of non-vanishing interplane cou- plicate any theoretical analysis. Even at the level of or-
plings, J∆ = 0.4α∆ and Jφ = 0.01α∆ . der parameter theories, there are more players than the
SC and a single component CDW order. To begin with,
since the Cu-O plane is approximately tetragonal, it is
VI. CLASSICAL MONTE CARLO RESULTS necessary to include at least two CDW orders, with or-
dering vectors related by a C4 rotation. In various parts
Finally, we treat e−S as the Boltzman weight and com- of the phase diagram0 - as well as in serious studies of
pute the phase diagram that results by exact numeri- model problems27,40–45 such as the Hubbard model which
cal Monte-Carlo evaluation of thermodynamic correla- are thought to capture some of the essence of the micro-
tion functions. Of course, the down side of this is that scopic physics of the cuprates - there is also clear evidence
9

Still, in certain cuprates, there is a range of doping in


2 0.03
which the only two orders that have been clearly identi-
0.025
fied are CDW and SC order.50–58 It is thus interesting to
explore to what extent the salient features of the phase
1.5 0.02
diagram in this regime can be understood, following the
above considerations, as an expression of the competition
H/H (0)
c2

0.015 between these two forms of order. To get closer to the


physics of the cuprates, we generalize the above discus-
1 0.01 sion to the case of the competition between CDW and SC
orders in a tetragonal system, where there are two CDW
0.005 order parameters, φx and φy , which are the slowly vary-
ing CDW amplitudes for ordering vectors along the x and
0.5 0 y axes, respectively. Following the analysis in Refs.39,59 ,
0.2 0.4 0.6 0.8 1 1.2
(0)
we assume that there is a strong repulsive interaction
T/Tc between these two components of the CDW order, so
that the preferred ordered state is undirectional (striped)
2 0.03
rather than bidirectional (checkerboard). When we gen-
0.025
eralize the above model to include both φx and φy , the
bulk of the considerations go through in analogous fash-
1.5 0.02 ion. As already mentioned in the introduction, the most
important difference is that in addition to breaking trans-
(0)
H/H c2

0.015 lational symmetry (in one direction), a stripe ordered


CDW also spontaneously breaks C4 rotational symmetry
1 0.01
down to C2 . As shown in Ref. 39, this has a profound
effect on the phase diagram in the presence of weak dis-
0.005
order – while no true incommensurate CDW order exists,
0.5 0 the spontaneous breaking of rotational symmetry persists
0.2 0.4 0.6 0.8 1 1.2 1.4 for disorder strength less than an order 1 critical value
T/T(0)
c - a genuine vestigial nematic phase with a well-defined
critical temperatures is robust in this case.
FIG. 5: Monte Carlo results for the magnitude of the ther- We thus are lead to the schematic phase diagram
mally averaged SC correlations, GSC , defined in Eq. 6.1. The sketched in Fig. 6 for a putative tetragonal cuprate with
bright region is interpretted to be the ordered SC phase, while weak disorder in a regime in which the competition be-
(0)
the dark region is the disordered phase. Hc2 = Hc2 (γ = 1) tween CDW and SC dominate the physics. This is a
is the upper critical magnetic field in the absence of disor-
(0) decorated version of Fig. 1(d). The principle difference
der. Tc is the critical temperature, obtained in Monte Carlo relative to the orthorhombic case is that what was for-
simulation. (Top) Phase diagram for system without disorder
mally a crossover to a high field regime with substantial
with γc2 > γ > γc1 . Phase boundaries for SC (Solid lines) and
CDW (dashed line) are obtained from variational calculation CDW order in Fig. 1(c) now appears as the thermody-
for J∆ = 0.4 and Jφ = 0.01. Phase boundaries for SC un- namic phase boundary of an ordered nematic phase, as
der disordered/ordered CDW have different slope, as verified in Fig. 1(d).76 As before, the “fragile SC” is a state in
in Monte Carlo simulation. A weak first order transition for which global phase coherence is mediated by the Joseph-
CDW may occur due to variational calculation, which leads son coupling between SC halos associated with neighbor-
to a small discontinuity near the critical point (hardly seen). ing dislocations separated by a typical distance of order
(Bottom)Phase diagram for σ = 0.05σc . Disorder clearly in- the CDW correlation length. The “fragile nematic” is a
creases the critical magnetic field, but suppresses the critical dual of this state, in which nematic LRO is mediated by
temperature. In the variational calculation, the weak disor- the interaction between neighboring CDW halos associ-
der suppresses the critical temperature by less than 1%, while
ated with neighboring vortices.
in the Monte Carlo simulation, the suppression is near 20%.
Variational calculation also fails to explain the behaviour at There are other features we have included in Fig. 6
low temperature, while CDW dislocations are essential. that go beyond the considerations of the classical effec-
tive action we have analyzed. In the first place, we have
conjecturally included the effect of quantum fluctuations
of incommensurate SDW order, and suggestive evidence of the SC on the phase diagram at the highest fields and
of PDW order.12,46–49 An additional complication is that lowest T , where they will likely give rise to an “anoma-
at T → 0, the presence of gapless quasiparticles leads lous metallic” regime73 . Here, quantum fluctuations of
to non-local interactions and of course quantum fluctua- the phase of the order parameter on each dislocation halo
tions of the various order parameters need to be included; destroy global phase coherence, even at T = 0, but there
none of these features are captured in the LGW frame- remain substantial SC correlations that extend across
work adopted above. multiple halos. We have also indicated a regime in which
10

PDW order is most likely to arise – associated with vor- the Knight-shift was found to be approximately
tex halos where CDW and SC order coexist.11–13,17 field independent in the regime of the phase dia-
gram which roughly corresponds to the region la-
beled “SC SRO - dislocation halos” in Fig. 6; this
Anomalous was (sensibly) taken as evidence that SC correla-
metal Nematic with tions are essentially absent here.
F
r
unidirectional
Η
a CDW SRO
g (nematic near • It was conjectured in Ref.66 , on the basis high field
i SC SRO to smectic)
l
- dislocation
X-ray diffraction data56,57,66 that the “ideal” CDW
e
halos Weak bidirectional
order - i.e. the order that would arise in a disorder-
S CDW SRO free system at high enough fields to quench SC -
C
- disorder pinned would be unidirectional stripe CDW order with in-
F N
r e
phase ordering ordering in the inter-plane (c-axis)
a m Superconductivity direction. However, to account for the presence of
g a
i t (zero resistance)
CDW SRO
l i
bidirectional short-range CDW correlations (with
e c
(+ possible PDW SRO) the same in-plane ordering vector) with little in
- in vortex halos the way of c-axis correlations, it was conjectured
Τ*(σ) Τ
that the CDW order must be effectively inhomoge-
neous, with more and less ordered regions coexist-
ing. More recent studies of the effect of unidirec-
tional strain on CDW order67 are broadly consis-
FIG. 6: Schematic phase diagrams showing competition be-
tent with this picture.
tween unidirectional CDW and SC order in a weakly disor-
dered tetragonal system as a function of T and B. The solid
black line indicates the superconducting transition and the • Quantum oscillations with all the signatures of a
solid red line is the nematic phase boundary. The various non-superconducting Fermi liquid with a recon-
descriptive text are explained in the text in Sec. VII. structed Fermi surface (presumably due to the pres-
ence of the CDW) is observed64,68 in the same range
These new results potentially give some theoretical ba- of fields and temperatures that the resilient super-
sis for an understanding of observed phenomena in vari- conductivity is detected. In common with the field
ous cuprates. In particular, the fact that many features independence of the Knight shift (discussed above)
of the phase diagram in Fig. 6 correspond to observed and of various thermal transport coefiscients65,69 ,
behaviors in underdoped YBCO in a range of dopings these observations are most readily understood un-
between 0.08 < p < 0.15, supports the conjecture that der the assumption that the SC correlations are
the principal features governing the phase diagram is a entirely quenched by the magnetic field.
fierce competition60 between CDW and SC orders. Some
of the salient features that we have in mind are: Reconciling the conflicting evidence from different
measurements that appear to indicate that the high field
• Recent high field transport and magnetization
state is both entirely metallic and host to strong local
measurements74 have revealed a phase diagram
SC correlations is difficult - conceivably impossible - in
with a sharp upturn in the resistively determined
terms of any uniform electronic structure. However, as
Hc2 at temperatures T < ∼ 2K. This behavior was the materials involved are highly crystalline and believed
presented as evidence of “resilient” SC and cor-
to be structurally homogeneous, there is a natural pred-
related with previous studies61–63 that found evi-
judice against any interpretation that invokes electronic
dence of SC correlations that persist to much higher
inhomogeneities.78 In this context, we reiterate that the
fields than the typical Hc2 . The observation that
electronic inhomogeneities we have invoked are intrinsic
the critical current is anomalously small in this low
features of real systems which always have some degree of
T superconducting phase was adduced as the rea-
disorder, even when that disorder is statistically homoge-
son it had not been previously observed, but the
neous and the high temperature “normal” state shows no
fact that it correlates with magnetic hysteresis64
significant electronic inhomogeneities. Moreover, there is
(vortex pinning) was interpreted as evidence that
an important correlation that is, in principle, experimen-
it is “bulk” effect.77
tally testable. More disordered regions, have stronger,
• A form of the phase diagram that is also reminis- but more short-range correlated CDW order, and thus
cent of our Fig. 6 was proposed in Ref.65 largely on will tend to have lower local values of the SC Tc , but
the basis of NMR studies, although given a some- higher values of Hc2 . So, for example, we would expect
what different interpretation than in Ref.74 . One that light Zn doping of YBCO would produce a small de-
difference is that the fragile SC regime inferred from crease in the zero field Tc , but a broadening of the range
NMR does not extend to as high fields as the “re- of T in which the fragile SC persists at low T and large
silient SC” reported in Ref.74 . More importantly, H.
11

Acknowledgements The self-consistency for the variational parameters,


µ∆ , µφ and m are
We thank J. Tranquada, B. Ramshaw, and P. Ar-
mitage for helpful comments and S. Sebastian for a pre- h i
publication copy of Ref.74 . This work was supported in m2 = µ2φ αφ − hψr2 i − hψi2 i − λh|∆|2 i
part by the Department of Energy, Office of Basic Energy m2 2
Sciences, Division of Materials Sciences and Engineering, µφ = −αφ + + 3hψ1,r i + hψi2 i − λh|∆|2 i
under Contract No. DE-AC02-76SF00515. µ2φ (A3)
" #
2 m2 2 2
µ∆ = −α∆ + 2h|∆| i + γ + hψr i + hψi i
µ2φ
Appendix A: More Concerning the Variational
solution
The value of Tc is extracted from the self-consistency
To obtain the phase boundary of superconductivity in equations as the point at which µ∆ → 2J∆ − 2eκ∆ B.
the CDW ordered phase in the absence of disorder, we This set of equations are valid so long as there is CDW
need to extend the variational treatment to the interior LRO (m 6= 0) but no SC LRO.
of the broken symmetry phase. We thus consider the
following non-replicated trial Hamiltonian density This extension of the variational approach is neces-
sary to compute the phase boundary (Hc2 at low T ) be-
κφ µφ ? m tween the phase with only CDW order and the phase
Htr = |∇φ|2 + [φ + φ]2 − [φ? + φ]
2 8 2 (A1) with coexisting CDW and SC order shown in Fig. 5.
κ∆ 2 µ∆ 2 The phase boundary is identified as the point at which
+ |(−i∇ − 2eA)∆| + |∆| .
2 2 µ∆ → 2J∆ − 2eκ∆ B. It typically occurs that near the
putative multicritical point, the transition becomes (pre-
We have chosen a convention such that the expectation
sumably unphysically) weakly first order. Here, phase
value of φ is real, in which the small amplitude fluctua-
boundaries must be determined by comparing the vari-
tions of the imaginary part is gapless (i.e. is the Gold-
ational free energy of the CDW ordered non SC phase
stone mode). Then, if we express φ in terms of the real
with that of the SC non-CDW and the fully disordered
and imaginary parts, φ ≡ hφi + ψr + iψi , the mean values
phase.
of various quantities are
Another artifact of the variational approach is that
m
hφi = it predicts that all transition temperatures vanish for
µφ J∆ = Jφ = 0, i.e. it fails to incorporate the physics of
d3 k
Z
2 1 the BKT transitions to phases with quasi-long-range or-
h[ψr ] i = T (A2) der. When comparing our variational results with Monte
(2π)3 µφ + κφ kk2 + 2Jφ cos k⊥
Carlo result, we choose J∆ = 0.4α∆ and Jφ = 0.01α∆ .
d3 k
Z
1 The variational phase boundary in Fig. 5 is normalized
h[ψi ]2 i = T
(2π)3 κφ kk2 + 2Jφ cos k⊥ such that it matches the correct critical temperature at
the smallest magnetic field and the critical field at the
while h|∆|2 i is still given by Eq.5.5. lowest temperature.

1
E. Fradkin, S. A. Kivelson, and J. M. Tranquada, Rev. 87, 067202 (2001), URL https://link.aps.org/doi/10.
Mod. Phys. 87, 457 (2015), URL https://link.aps.org/ 1103/PhysRevLett.87.067202.
8
doi/10.1103/RevModPhys.87.457. Y. Zhang, E. Demler, and S. Sachdev, Phys. Rev. B
2
E. Witten, Nucl. Phys. B 249, 557 (1985). 66, 094501 (2002), URL https://link.aps.org/doi/10.
3
S. Zhang, Phys. Rev. Lett. 59, 2111 (1987). 1103/PhysRevB.66.094501.
4 9
S.-C. Zhang, Science 275, 1089 (1997), ISSN 0036-8075, S. A. Kivelson, D.-H. Lee, E. Fradkin, and V. Oganesyan,
http://science.sciencemag.org/content/275/5303/1089.full.pdf, Phys. Rev. B 66, 144516 (2002), URL https://link.aps.
URL http://science.sciencemag.org/content/275/ org/doi/10.1103/PhysRevB.66.144516.
10
5303/1089. S. A. Kivelson, D.-H. Lee, E. Fradkin, and V. Oganesyan,
5
D. P. Arovas, A. J. Berlinsky, C. Kallin, and S.-C. Zhang, Phys. Rev. B 66, 144516 (2002), URL https://link.aps.
Phys. Rev. Lett. 79, 2871 (1997), URL https://link. org/doi/10.1103/PhysRevB.66.144516.
11
aps.org/doi/10.1103/PhysRevLett.79.2871. D. F. Agterberg and J. Garaud, Phys. Rev. B 91,
6
P. A. Lee and X.-G. Wen, Phys. Rev. B 63, 104512 (2015), URL https://link.aps.org/doi/10.
224517 (2001), URL https://link.aps.org/doi/10. 1103/PhysRevB.91.104512.
12
1103/PhysRevB.63.224517. Y. Wang, S. D. Edkins, M. H. Hamidian, J. C. S.
7
E. Demler, S. Sachdev, and Y. Zhang, Phys. Rev. Lett. Davis, E. Fradkin, and S. A. Kivelson, Phys. Rev. B
12

97, 174510 (2018), URL https://link.aps.org/doi/10. http://science.sciencemag.org/content/343/6169/393.full.pdf,


1103/PhysRevB.97.174510. URL http://science.sciencemag.org/content/343/
13
Z. Dai, Y.-H. Zhang, T. Senthil, and P. A. Lee, Phys. Rev. 6169/393.
33
B 97, 174511 (2018), URL https://link.aps.org/doi/ Y. Wang and A. Chubukov, Phys. Rev. B 90,
10.1103/PhysRevB.97.174511. 035149 (2014), URL https://link.aps.org/doi/10.
14
B. Lake, G. Aeppli, K. Clausen, D. McMorrow, K. Lef- 1103/PhysRevB.90.035149.
34
mann, N. Hussey, N. Mangkorntong, M. Nohara, H. Tak- J. D. Sau and S. Sachdev, Phys. Rev. B 89,
agi, T. Mason, et al., Science 291, 1759 (2001). 075129 (2014), URL https://link.aps.org/doi/10.
15
B. Lake, K. Lefmann, N. B. Christensen, G. Aeppli, D. F. 1103/PhysRevB.89.075129.
35
McMorrow, H. M. Rønnow, P. Vorderwisch, P. Smeibidl, H. Meier, M. Einenkel, C. Pépin, and K. B. Efetov, Phys.
N. Mangkorntong, T. Sasagawa, et al., Nature Mat. 4, 658 Rev. B 88, 020506 (2013), URL https://link.aps.org/
(2005). doi/10.1103/PhysRevB.88.020506.
16 36
J. E. Hoffman, E. W. Hudson, K. M. Lang, M. A. Metlitski and S. Sachdev, Phys. Rev. B 82,
V. Madhavan, H. Eisaki, S. Uchida, and J. C. 075128 (2010), URL https://link.aps.org/doi/10.
Davis, Science 295, 466 (2002), ISSN 0036-8075, 1103/PhysRevB.82.075128.
37
http://science.sciencemag.org/content/295/5554/466.full.pdf, D. S. Fisher, M. P. A. Fisher, and D. A. Huse, Phys. Rev.
URL http://science.sciencemag.org/content/295/ B 43, 130 (1991), URL https://link.aps.org/doi/10.
5554/466. 1103/PhysRevB.43.130.
17 38
S. D. Edkins, A. Kostin, K. Fujita, A. P. Mackenzie, G. Blatter, M. V. Feigel’man, V. B. Geshkenbein, A. I.
H. Eisaki, S.-I. Uchida, S. Sachdev, M. J. Lawler, E.-A. Larkin, and V. M. Vinokur, Rev. Mod. Phys. 66,
Kim, J. C. Séamus Davis, et al., ArXiv e-prints (2018), 1125 (1994), URL https://link.aps.org/doi/10.1103/
1802.04673. RevModPhys.66.1125.
18 39
B. Kalisky, J. R. Kirtley, J. G. Analytis, J.-H. Chu, L. Nie, G. Tarjus, and S. A. Kivelson, PROCEEDINGS OF
A. Vailionis, I. R. Fisher, and K. A. Moler, Phys. Rev. THE NATIONAL ACADEMY OF SCIENCES OF THE
B 81, 184513 (2010), URL https://link.aps.org/doi/ UNITED STATES OF AMERICA 111, 7980 (2014), ISSN
10.1103/PhysRevB.81.184513. 0027-8424.
19 40
M. V. Feigel’man, A. I. Larkin, and M. A. Skvortsov, Phys. J. Zaanen and O. Gunnarsson, Phys. Rev. B 40,
Rev. Lett. 86, 1869 (2001), URL https://link.aps.org/ 7391 (1989), URL https://link.aps.org/doi/10.1103/
doi/10.1103/PhysRevLett.86.1869. PhysRevB.40.7391.
20 41
B. Spivak, P. Oreto, and S. A. Kivelson, Phys. Rev. B H. J. Schulz, Phys. Rev. Lett. 64, 1445 (1990), URL
77, 214523 (2008), URL https://link.aps.org/doi/10. https://link.aps.org/doi/10.1103/PhysRevLett.64.
1103/PhysRevB.77.214523. 1445.
21 42
R. Feynman, Statistical Mechanics, A set of Lectures (Ben- S. R. White and D. J. Scalapino, Phys. Rev. Lett. 80,
jamin Cummings Publishing, 1982). 1272 (1998), URL https://link.aps.org/doi/10.1103/
22
S. Sachdev, Statistical Mechanics, A set of Lectures (Cam- PhysRevLett.80.1272.
43
bridge University Press, 1999). C. S. Hellberg and E. Manousakis, Phys. Rev. Lett. 83,
23
C. M. Varma, Phys. Rev. B 55, 14554 (1997), URL https: 132 (1999), URL https://link.aps.org/doi/10.1103/
//link.aps.org/doi/10.1103/PhysRevB.55.14554. PhysRevLett.83.132.
24 44
S. A. Kivelson, E. Fradkin, and V. J. Emery, Nature 393, C. S. Hellberg and E. Manousakis, Phys. Rev. Lett. 78,
550 (1998). 4609 (1997), URL https://link.aps.org/doi/10.1103/
25
C. Castellani, C. Dicastro, and M. grilli, Journal of PhysRevLett.78.4609.
45
Physics and Chemistry of Solids 59, 1694 (1998), M. Vojta, Advances in Physics 58, 699 (2009),
ISSN 0022-3697, URL http://www.sciencedirect.com/ https://doi.org/10.1080/00018730903122242, URL https:
science/article/pii/S0022369798000857. //doi.org/10.1080/00018730903122242.
26 46
S. Chakravarty, R. B. Laughlin, D. K. Morr, and C. Nayak, A. Himeda, T. Kato, and M. Ogata, Phys. Rev. Lett.
Phys. Rev. B 63, 094503 (2001), URL https://link.aps. 88, 117001 (2002), URL https://link.aps.org/doi/10.
org/doi/10.1103/PhysRevB.63.094503. 1103/PhysRevLett.88.117001.
27 47
S. A. Kivelson, I. P. Bindloss, E. Fradkin, V. Oganesyan, K.-Y. Yang, W. Q. Chen, T. M. Rice, M. Sigrist, and F.-C.
J. M. Tranquada, A. Kapitulnik, and C. Howald, Rev. Zhang, NEW JOURNAL OF PHYSICS 11 (2009), ISSN
Mod. Phys. 75, 1201 (2003), URL https://link.aps. 1367-2630.
48
org/doi/10.1103/RevModPhys.75.1201. M. Raczkowski, M. Capello, D. Poilblanc, R. Frésard, and
28
S. Sachdev, Rev. Mod. Phys. 75, 913 (2003), URL https: A. M. Oles, Phys. Rev. B 76, 140505 (2007), URL https:
//link.aps.org/doi/10.1103/RevModPhys.75.913. //link.aps.org/doi/10.1103/PhysRevB.76.140505.
29 49
E. Berg, E. Fradkin, S. A. Kivelson, and J. Tranquada, P. Corboz, T. M. Rice, and M. Troyer, Phys. Rev. Lett.
New Journal of Physics 11, 115004 (2009). 113, 046402 (2014), URL https://link.aps.org/doi/
30
J. S. Davis and D.-H. Lee, Proceedings of the National 10.1103/PhysRevLett.113.046402.
50
Academy of Sciences p. 201316512 (2013). R. Zhou, M. Hirata, T. Wu, I. Vinograd, H. Mayaffre,
31
J. M. Tranquada, G. D. Gu, M. Hücker, Q. Jie, H.-J. Kang, S. Kraemer, A. P. Reyes, P. L. Kuhns, R. Liang, W. N.
R. Klingeler, Q. Li, N. Tristan, J. S. Wen, G. Y. Xu, et al., Hardy, et al., PROCEEDINGS OF THE NATIONAL
Phys. Rev. B 78, 174529 (2008), URL https://link.aps. ACADEMY OF SCIENCES OF THE UNITED STATES
org/doi/10.1103/PhysRevB.78.174529. OF AMERICA 114, 13148 (2017), ISSN 0027-8424.
32 51
E. H. da Silva Neto, P. Aynajian, A. Frano, R. Comin, D. Haug, V. Hinkov, Y. Sidis, P. Bourges, N. B. Chris-
E. Schierle, E. Weschke, A. Gyenis, J. Wen, J. Schneeloch, tensen, A. Ivanov, T. Keller, C. T. Lin, and B. Keimer,
Z. Xu, et al., Science 343, 393 (2014), ISSN 0036-8075, New Journal of Physics 12, 105006 (2010), URL http:
13

66
//stacks.iop.org/1367-2630/12/i=10/a=105006. H. Jang, W.-S. Lee, H. Nojiri, S. Matsuzawa, H. Ya-
52
G. Ghiringhelli, M. Le Tacon, M. Minola, S. Blanco- sumura, L. Nie, A. V. Maharaj, S. Gerber, Y.-J.
Canosa, C. Mazzoli, N. B. Brookes, G. M. Liu, A. Mehta, et al., Proceedings of the National
De Luca, A. Frano, D. G. Hawthorn, F. He, Academy of Sciences 113, 14645 (2016), ISSN 0027-8424,
et al., Science 337, 821 (2012), ISSN 0036-8075, http://www.pnas.org/content/113/51/14645.full.pdf,
http://science.sciencemag.org/content/337/6096/821.full.pdf, URL http://www.pnas.org/content/113/51/14645.
67
URL http://science.sciencemag.org/content/337/ H. Kim, S. Souliou, M. Barber, E. Lefrancois, M. Minola,
6096/821. R. Heid, A. Bosak, A. Mackenzie, B. Keimer, C. Hicks,
53
J. Chang, E. Blackburn, A. Holmes, N. Christensen, et al. (2018), 43.21.01; LK 01.
68
J. Larsen, J. Mesot, R. Liang, D. Bonn, W. Hardy, A. Wa- N. Doiron-Leyraud, C. Proust, D. LeBoeuf, J. Levallois,
tenphul, et al., Nature Physics 8, 871 (2012). J.-B. Bonnemaison, R. Liang, D. A. Bonn, W. N. Hardy,
54
M. Hücker, N. B. Christensen, A. T. Holmes, E. Black- and L. Taillefer, Nature 447, 565 (2007).
69
burn, E. M. Forgan, R. Liang, D. A. Bonn, W. N. Hardy, G. Grissonnanche, O. Cyr-Choiniere, F. Laliberté, S. R.
O. Gutowski, M. v. Zimmermann, et al., Phys. Rev. B De Cotret, A. Juneau-Fecteau, S. Dufour-Beauséjour, M.-
90, 054514 (2014), URL https://link.aps.org/doi/10. E. Delage, D. LeBoeuf, J. Chang, B. Ramshaw, et al.,
1103/PhysRevB.90.054514. Nature communications 5, 3280 (2014).
55 70
R. Comin, R. Sutarto, E. H. da Silva Neto, C. Howald, P. Fournier, and A. Kapitulnik, Phys. Rev. B
L. Chauviere, R. Liang, W. N. Hardy, D. A. 64, 100504 (2001), URL https://link.aps.org/doi/10.
Bonn, F. He, G. A. Sawatzky, and A. Dama- 1103/PhysRevB.64.100504.
71
scelli, Science 347, 1335 (2015), ISSN 0036-8075, K. M. Lang, V. Madhavan, J. E. Hoffman, E. W. Hudson,
http://science.sciencemag.org/content/347/6228/1335.full.pdf, H. Eisaki, S. Uchida, and J. C. Davis, Nature 415, 412 EP
URL http://science.sciencemag.org/content/347/ (2002), URL http://dx.doi.org/10.1038/415412a.
72
6228/1335. K. K Gomes, A. N Pasupathy, A. Pushp, S. Ono, Y. Ando,
56
S. Gerber, H. Jang, H. Nojiri, S. Matsuzawa, H. Ya- and A. Yazdani, 447, 569 (2007).
73
sumura, D. A. Bonn, R. Liang, W. N. Hardy, Z. Islam, A. Kapitulnik, S. A. Kivelson, and B. Spivak, arXiv
A. Mehta, et al., Science 350, 949 (2015), ISSN 0036-8075, preprint arXiv:1712.07215 (2017).
http://science.sciencemag.org/content/350/6263/949.full.pdf, 74 Y.-T. Hsu, H. M, A. J. Davies, M. K. Chan, J. Porras,
URL http://science.sciencemag.org/content/350/ T. Loew, S. V. Taylor, H. Liu, M. L. Tacon, H. Zuo, et al.
6263/949. (????), preprint.
57 75
J. Chang, E. Blackburn, O. Ivashko, A. Holmes, N. B. If the preferred CDW order is not along a symmetry direc-
Christensen, M. Hücker, R. Liang, D. Bonn, W. Hardy, tion, then there are four symmetry related directions the
U. Rütt, et al., Nature communications 7, 11494 (2016). CDW order can chose.
58 76
H. Jang, W.-S. Lee, S. Song, H. Nojiri, S. Matsuzawa, Another potential new feature of the tetragonal case is that
H. Yasumura, H. Huang, Y.-J. Liu, J. Porras, M. Minola, there are a new class of topological defects - domain wall
et al., Phys. Rev. B 97, 224513 (2018), URL https:// defects in the Ising-nematic order parameter, across which
link.aps.org/doi/10.1103/PhysRevB.97.224513. the local orientation of the CDW order rotates. The CDW
59
L. Nie, A. V. Maharaj, E. Fradkin, and S. A. Kivelson, order is generally suppressed (although it need not vanish)
Phys. Rev. B 96, 085142 (2017), URL https://link.aps. at these defects as well, so it is natural to expect enhanced
org/doi/10.1103/PhysRevB.96.085142. superconducting order here as well.
60 77
O. Cyr-Choinière, D. LeBoeuf, S. Badoux, S. Dufour- An obvious possibility that must be considered is that the
Beauséjour, D. A. Bonn, W. N. Hardy, R. Liang, D. Graf, resilient SC reflects some form of chemical inhomogeneity
N. Doiron-Leyraud, and L. Taillefer, Phys. Rev. B in the sample that gives rise to “filamentary” SC. The ob-
98, 064513 (2018), URL https://link.aps.org/doi/10. served magnetic hysteresis is inconsistent with the most
1103/PhysRevB.98.064513. straightforward versions of this interpretation, but a sys-
61
F. Yu, M. Hirschberger, T. Loew, G. Li, B. J. Law- tem of structural inhomogeneities with correlations on ap-
son, T. Asaba, J. B. Kemper, T. Liang, J. Porras, propriate length scales could probably provide an alterna-
G. S. Boebinger, et al., Proceedings of the National tive explanation of all the observed phenomena. There are
Academy of Sciences 113, 12667 (2016), ISSN 0027-8424, various ways to test this, including by comparing proper-
http://www.pnas.org/content/113/45/12667.full.pdf, ties of crystals with similar doped hole concentration grown
URL http://www.pnas.org/content/113/45/12667. under different conditions.
62 78
S. C. Riggs, O. Vafek, J. Kemper, J. Betts, A. Migliori, It has been established beyond dispute in STM studies of
F. Balakirev, W. Hardy, R. Liang, D. Bonn, and G. Boe- BSCCO that the low temperature electronic structure of
binger, Nature Physics 7, 332 (2011). this cuprate is highly inhomogeneous.70–72 Moreover, the
63
J. Corson, R. Mallozzi, J. Orenstein, J. N. Eckstein, and dual behavior - the appearance of enhanced CDW correla-
I. Bozovic, Nature 398, 221 (1999). tions near the SC vortex cores - has been directly visualized
64
S. E. Sebastian, N. Harrison, E. Palm, T. Murphy, in these materials.16,17 By some metrics, this material is
C. Mielke, R. Liang, D. Bonn, W. Hardy, and G. Lon- more disordered than YBCO, so a direct comparison is not
zarich, Nature 454, 200 (2008). possible. However, it seems likely that what is occurring on
65
J. Kacmarcik, I. Vinograd, B. Michon, A. Rydh, A. De- relatively short length scales in BSCCO is likely applicable
muer, R. Zhou, H. Mayaffre, R. Liang, W. Hardy, D. A. in YBCO at somewhat larger scales.
Bonn, et al., ArXiv e-prints (2018), 1805.06853.

You might also like