+ All Categories
Home > Documents > Concepts relating magnetic interactions, intertwined ...electron interactions, IPs, and correlated...

Concepts relating magnetic interactions, intertwined ...electron interactions, IPs, and correlated...

Date post: 10-Sep-2020
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
8
Concepts relating magnetic interactions, intertwined electronic orders, and strongly correlated superconductivity J. C. Séamus Davis a,b,c,d,1 and Dung-Hai Lee e,f,1 a Department of Physics, Cornell University, Ithaca, NY 14853; b Condensed Matter Physics and Material Science Department, Brookhaven National Laboratory, Upton, NY 11973; c School of Physics and Astronomy, University of St. Andrews, Fife KY16 9SS, Scotland; d Kavli Institute at Cornell for Nanoscale Science, Cornell University, Ithaca, NY 14853; e Department of Physics, University of California, Berkeley, CA 94720; and f Materials Science Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720 This contribution is part of the special series of Inaugural Articles by members of the National Academy of Sciences elected in 2010. Contributed by J. C. Séamus Davis, September 9, 2013 (sent for review July 13, 2013) Unconventional superconductivity (SC) is said to occur when Cooper pair formation is dominated by repulsive electronelectron interactions, so that the symmetry of the pair wave function is other than an isotropic s-wave. The strong, on-site, repulsive elec- tronelectron interactions that are the proximate cause of such SC are more typically drivers of commensurate magnetism. Indeed, it is the suppression of commensurate antiferromagnetism (AF) that usually allows this type of unconventional superconductivity to emerge. Importantly, however, intervening between these AF and SC phases, intertwined electronic ordered phases (IP) of an unex- pected nature are frequently discovered. For this reason, it has been extremely difcult to distinguish the microscopic essence of the correlated superconductivity from the often spectacular phenome- nology of the IPs. Here we introduce a model conceptual framework within which to understand the relationship between AF electronelectron interactions, IPs, and correlated SC. We demonstrate its effectiveness in simultaneously explaining the consequences of AF interactions for the copper-based, iron-based, and heavy-fermion superconductors, as well as for their quite distinct IPs. high-Tc superconductivity | strong correlations | fermiology E mergence, the coming into being through evolution, is an important concept in modern condensed matter physics (1). Superconductivity is a classic example of emergence in the realm of quantum matter: as the energy scale decreases, the effective electronelectron interactions responsible for Cooper pairing and thus superconductivity evolves from the elementary microscopic Hamiltonian through unanticipated modications (2). This evo- lution is why it is so difcult to derive superconductivity (SC) from rst principles. Finding the microscopic mechanism of Cooper pairing means discovering the nature of the ultimate effective electronelectron interaction at the lowest energy scales. In the last three decades, unconventional (35) forms of SC have been discovered in many strongly correlated (repulsive electronelectron interaction) systems. These materials fascinate a lay per- son for their high superconducting transition temperatures and therefore the potential for revolutionary applications in power generation/transmission, transport, information technology, sci- ence, and medicine. They intrigue (and challenge) physicists to identify the mechanism of their high pairing-energy scale and be- cause of the many intertwined (6, 7) electronic phases (IPs) that have been discovered in juxtaposition with the unconventional superconductivity. These IPs have been hypothesized to arise together from one parent statesuch that the various order parameters are intertwined rather than simply competing with each other(7). The best known and most widely studied exam- ples of such materials include the copper-based (811) and iron- based (1214) high-temperature superconductors, the heavy fer- mion superconductors (1517), and the organic superconductors (18). One thing commonly noted in these systems is that SC nor- mally borders antiferromagnetism (AF): in the phase diagram spanned by temperature and a certain control parameter (chem- ical-doping, pressure, etc.), an SC dome stands adjacent to the AF phase (Fig.1). However, the precise way the two phases are con- nected varies greatly from system to system. Another very common observation is the appearance of other ordered phases of electronic matter that intertwine with the SC. These exotic intertwined phases (IPs) occur in the terra incognita between the SC and the AF (Fig. 1, gray). Examples include the charge/spin density wave (1921) and intra-unit-cell symmetry breaking (2123) orders in the copper-based superconductors and the nematic order (24, 25) in the iron-based superconductors. A key long-term objective for this eld has therefore been to identify a simple framework within which to consider the re- lationship between the antiferromagnetic interactions, the inter- twined electronic orders that appear at its suppression, and the correlated superconductivity. Because in all of the systems considered here SC emerges from the extinction of AF, it is widely believed that the effective elec- tronelectron interaction triggering the Cooper pairing could be AF in form. In that case, of course, the same argument could apply to the other intertwined electronic phases. These ideas motivate the assertion that AF effective electronelectron interactions may drive both the correlated SC and the other IPs. Until recently, however, there has been little consensus on this issue. One reason is that the experimental evidence for many such intertwined states has only been rmly established in recent years. Another reason is that, although magnetism in proximity to unconventional SC appears universal, the nature of the IPs changes from system to system for reasons that appear mysterious. In this paper, we therefore explore the plausibility that an AF effective interaction could be the driving force for both the unconventional SC and the intertwined orders in the copper- based, iron-based, and heavy fermion superconductors. [We omit discussion of organic superconductors (26, 27) for the sake of brevity.] Here we will not try to rigorously solve for the ground state under different conditions. Our goal is to ask whether the known IPs are the locally stable mean-eld phases when the sole Signicance This study describes a unied theory explaining the rich ordering phenomena, each associated with a different symmetry breaking, that often accompany high-temperature superconductivity. The essence of this theory is an antiferromagnetic interaction,the interaction that favors the development of magnetic order where the magnetic moments reverse direction from one crystal unit cell to the next. We apply this theory to explain the superconductivity, as well as all observed accompanying ordering phenomena in the copper-oxide superconductors, the iron-based superconductors, and the heavy fermion superconductors. Author contributions: J.C.S.D. and D.-H.L. designed research, performed research, analyzed data, and wrote the paper. The authors declare no conict of interest. Freely available online through the PNAS open access option. 1 To whom correspondence may be addressed. E-mail: [email protected] and jcdavis@ ccmr.cornell.edu. www.pnas.org/cgi/doi/10.1073/pnas.1316512110 PNAS | October 29, 2013 | vol. 110 | no. 44 | 1762317630 PHYSICS INAUGURAL ARTICLE Downloaded by guest on December 30, 2020
Transcript
Page 1: Concepts relating magnetic interactions, intertwined ...electron interactions, IPs, and correlated SC. We demonstrate its effectiveness in simultaneously explaining the consequences

Concepts relating magnetic interactions, intertwinedelectronic orders, and strongly correlated superconductivityJ. C. Séamus Davisa,b,c,d,1 and Dung-Hai Leee,f,1

aDepartment of Physics, Cornell University, Ithaca, NY 14853; bCondensed Matter Physics and Material Science Department, Brookhaven National Laboratory,Upton, NY 11973; cSchool of Physics and Astronomy, University of St. Andrews, Fife KY16 9SS, Scotland; dKavli Institute at Cornell for Nanoscale Science, CornellUniversity, Ithaca, NY 14853; eDepartment of Physics, University of California, Berkeley, CA 94720; and fMaterials Science Division, Lawrence Berkeley NationalLaboratory, Berkeley, CA 94720

This contribution is part of the special series of Inaugural Articles by members of the National Academy of Sciences elected in 2010.

Contributed by J. C. Séamus Davis, September 9, 2013 (sent for review July 13, 2013)

Unconventional superconductivity (SC) is said to occur whenCooper pair formation is dominated by repulsive electron–electroninteractions, so that the symmetry of the pair wave function isother than an isotropic s-wave. The strong, on-site, repulsive elec-tron–electron interactions that are the proximate cause of such SCare more typically drivers of commensurate magnetism. Indeed, itis the suppression of commensurate antiferromagnetism (AF) thatusually allows this type of unconventional superconductivity toemerge. Importantly, however, intervening between these AF andSC phases, intertwined electronic ordered phases (IP) of an unex-pected nature are frequently discovered. For this reason, it has beenextremely difficult to distinguish the microscopic essence of thecorrelated superconductivity from the often spectacular phenome-nology of the IPs. Herewe introduce amodel conceptual frameworkwithin which to understand the relationship between AF electron–electron interactions, IPs, and correlated SC. We demonstrate itseffectiveness in simultaneously explaining the consequences of AFinteractions for the copper-based, iron-based, and heavy-fermionsuperconductors, as well as for their quite distinct IPs.

high-Tc superconductivity | strong correlations | fermiology

Emergence, the coming into being through evolution, is animportant concept in modern condensed matter physics (1).

Superconductivity is a classic example of emergence in the realmof quantum matter: as the energy scale decreases, the effectiveelectron–electron interactions responsible for Cooper pairing andthus superconductivity evolves from the elementary microscopicHamiltonian through unanticipated modifications (2). This evo-lution is why it is so difficult to derive superconductivity (SC) fromfirst principles. Finding the microscopic mechanism of Cooperpairing means discovering the nature of the ultimate effectiveelectron–electron interaction at the lowest energy scales.In the last three decades, unconventional (3–5) forms of SC have

been discovered in many strongly correlated (repulsive electron–electron interaction) systems. These materials fascinate a lay per-son for their high superconducting transition temperatures andtherefore the potential for revolutionary applications in powergeneration/transmission, transport, information technology, sci-ence, and medicine. They intrigue (and challenge) physicists toidentify the mechanism of their high pairing-energy scale and be-cause of the many intertwined (6, 7) electronic phases (IPs) thathave been discovered in juxtaposition with the unconventionalsuperconductivity. These IPs have been hypothesized to “arisetogether from one parent state” such that “the various orderparameters are intertwined rather than simply competing witheach other” (7). The best known and most widely studied exam-ples of such materials include the copper-based (8–11) and iron-based (12–14) high-temperature superconductors, the heavy fer-mion superconductors (15–17), and the organic superconductors(18). One thing commonly noted in these systems is that SC nor-mally borders antiferromagnetism (AF): in the phase diagramspanned by temperature and a certain control parameter (chem-ical-doping, pressure, etc.), an SC dome stands adjacent to the AF

phase (Fig.1). However, the precise way the two phases are con-nected varies greatly from system to system.Another very common observation is the appearance of other

ordered phases of electronic matter that intertwine with the SC.These exotic intertwined phases (IPs) occur in the terra incognitabetween the SC and the AF (Fig. 1, gray). Examples include thecharge/spin density wave (19–21) and intra-unit-cell symmetrybreaking (21–23) orders in the copper-based superconductorsand the nematic order (24, 25) in the iron-based superconductors.A key long-term objective for this field has therefore been toidentify a simple framework within which to consider the re-lationship between the antiferromagnetic interactions, the inter-twined electronic orders that appear at its suppression, and thecorrelated superconductivity.Because in all of the systems considered here SC emerges from

the extinction of AF, it is widely believed that the effective elec-tron–electron interaction triggering the Cooper pairing could beAF in form. In that case, of course, the same argument could applyto the other intertwined electronic phases. These ideas motivatethe assertion that AF effective electron–electron interactions maydrive both the correlated SC and the other IPs. Until recently,however, there has been little consensus on this issue. One reasonis that the experimental evidence for many such intertwined stateshas only been firmly established in recent years. Another reasonis that, although magnetism in proximity to unconventional SCappears universal, the nature of the IPs changes from system tosystem for reasons that appear mysterious.In this paper, we therefore explore the plausibility that an AF

effective interaction could be the driving force for both theunconventional SC and the intertwined orders in the copper-based, iron-based, and heavy fermion superconductors. [We omitdiscussion of organic superconductors (26, 27) for the sake ofbrevity.] Here we will not try to rigorously solve for the groundstate under different conditions. Our goal is to ask whether theknown IPs are the locally stable mean-field phases when the sole

Significance

This study describes a unified theory explaining the rich orderingphenomena, each associatedwith a different symmetry breaking,that often accompany high-temperature superconductivity. Theessence of this theory is an ”antiferromagnetic interaction,” theinteraction that favors the development of magnetic order wherethe magnetic moments reverse direction from one crystal unit cellto the next.We apply this theory to explain the superconductivity,as well as all observed accompanying ordering phenomena in thecopper-oxide superconductors, the iron-based superconductors,and the heavy fermion superconductors.

Author contributions: J.C.S.D. and D.-H.L. designed research, performed research, analyzeddata, and wrote the paper.

The authors declare no conflict of interest.

Freely available online through the PNAS open access option.1To whom correspondence may be addressed. E-mail: [email protected] and [email protected].

www.pnas.org/cgi/doi/10.1073/pnas.1316512110 PNAS | October 29, 2013 | vol. 110 | no. 44 | 17623–17630

PHYS

ICS

INAUGURA

LART

ICLE

Dow

nloa

ded

by g

uest

on

Dec

embe

r 30

, 202

0

Page 2: Concepts relating magnetic interactions, intertwined ...electron interactions, IPs, and correlated SC. We demonstrate its effectiveness in simultaneously explaining the consequences

effective electron–electron interaction is AF. We understandthat the actual effective interactions may be more complex thanthis simplest AF form; we deliberately omit these details with thegoal of identifying a simple framework within which all of therelevant phenomena can be considered. Two very recent pre-prints based on a related approach, but focusing only on thecopper-oxide superconductors, have been published (28, 29).

Effective HamiltonianThus, we start by studying the assertion that fermiology (the Fermisurface topology) + AF effective electron–electron interaction cangenerate the known IPs in different types of correlated SC mate-rials. Our effective Hamiltonian is viewed as evolved from the bareHamiltonian for strongCoulomb interactions, and our strategy is toexplore, in different ordering channels, which order dominates asthe exchange constants of the effective interactions increase fromzero. Under these circumstances, it is the AF interaction that isuniversal, whereas it is the fermiology that is not. The effectiveHamiltonian corresponding to our assertion is

Heff =Xk

′Xs

eðkÞnsðkÞ+Xi; j

JijSi ·Sj: [1]

Here, k and s are the momentum and spin labels, respectively. Sirepresents the total spin operator in the ith unit cell and is given

by12P

τ;sc†iτsσ

→ss′ciτs, where τ labels the degrees of freedom in each

unit cell (e.g., orbital, different sites). In addition, in Eq. 1, Σ′k is

a sum restricted to the neighborhood of the Fermi surface (whichcan have several disconnected pieces), eðkÞ is the dispersion ofthe relevant band in the vicinity of the Fermi surface, and JijSi ·Sjshould be understood as an electron–electron scattering term.Although we write it in real space, it should be converted tothe band eigen-basis and projected to the neighborhood of theFermi surface for each different type of system.Obviously, many simplifying assumptions have already been

made here. Note that, aside from the fact that Σ′k restricts states

to low single-particle excitation energies, there is no furtherconstraint on the Hilbert space, there is no gauge field, and theparticle statistics are the usual Fermi statistics. The only effectof interactions is captured by the Jij term. This simplification assertsthat the low energy physics, even nonfermi liquid behavior, can bethe result of the AF effective interaction. Therefore, althoughin the absence of Jij Eq. 1 describes a Fermi liquid, in its presence,the system may behave otherwise precisely because Jij can drivemany intertwined instabilities, and strong (critical) fluctuationsbetween these instabilities can then drive non-Fermi liquid be-havior. Thus, writing down Eq. 1 is not equivalent to assuminga nearly AF Fermi liquid (30, 31). This statement is particularly

true near, for example, the AF quantum critical point where Jijcan exhibit a strong dependence on the energy cutoff down tothe lowest energy.Of course, we do understand that many learned readers may

question our starting point of Eq. 1. However, in the search fora simple conceptual framework within which to understand quitedifferent correlated superconductors along with their distinct andcomplex IPs in multiple material systems (8–17), such a simplestarting point can have many advantages.

Copper-Based SuperconductorsFor the case of the copper-based SCs (8–11), we use a simple one-band model to describe the first term of Eq. 1; the relevant Fermisurface is shown in Fig. 2A. For Jij, we use the simplest nearestneighbor interaction to emulate the AF correlations. The utilityof Eq. 1 is validated in part by the Fermi liquid quasi-particleLandau quantization observed by high field quantum oscillationexperiments (32, 33); it is theoretically plausible (34) that suchFermi liquid behavior can be regained when the strong magneticfield quenches the relevant fluctuations.It has been known since the early days of cuprate SC that AF

fluctuations can induce d-wave Cooper pairing (35–37). We beginby reproducing what is known. Using the effective Hamiltonianspecified above, we obtain (Methods) the leading and subleadingSC gap functions shown in Fig. 2 C and D. (The idea to mean-field decouple the magnetic interaction to obtain Cooper pairingoriginates from refs. 38 and 39.) These two gap functions are ap-proximately described by cos  kx − cos  ky and cos  kx + cos  ky, respec-tively. This result indicates that the cuprates can have extendeds-wave pairing tendency (39) after all. Furthermore, Fig. 2B showshow this Fermi surface exhibits eight special hot spots (the reddots) where the AF Brillouin zone (BZ; dashed lines) crosses it.These spots are hypothesized to play a leading role in the interplayof intertwined phases and superconductivity in cuprates (21, 28).

Tem

pera

ture

Control Parameter

AF

IPSC

Fig. 1. Schematic phase diagramof unconventional superconductors.Starting from a robust commensu-rate AF, a control parameter, such ascarrier density or pressure, is variedso that the critical temperature TAFof the AF phase diminishes. Eventu-ally, an unconventional SC phaseappears at higher values of thecontrol parameter, and its criticaltemperature Tc is usually domeshaped. The intervening gray regionis where the AF phase and the SCphase connect. It is here that theintertwined phases of electronic matter have typically been discovered. Thecharacteristics of the IPs are highly distinctive to each system, as is the preciseway (e.g., first order, coexistence, quantum critical) that the AF–SC connectionoccurs. By contrast, the appearance of unconventional SC phase on suppressionof an AF state is virtually universal.

A B

DC

Fig. 2. Fermi surface and unconventional superconducting states of cuprates.(A) The cuprate first BZ, within which all of the momentum-space (k-space)electronic states of the system are described when not in the AF state. It spansa range −π=a< kx ≤ π=a; −π=a< ky ≤ π=a, where a is the unit cell dimension. Thedimensions of the BZ in A are in units of π=a. The model Fermi surface of thecuprates, constructed using a tight-binding single band model with first (t),second (t′), and third (t″) neighbor hopping, where t′=t = 0:3 and t″=t= 0:2, isshown. (B) This Fermi surface exhibits eight special hot spots (the red dots) wherethe AF BZ (dashed lines) crosses it. They appear to play a leading role in theinterplay of intertwined phases and superconductivity (21, 28). The black andgray arrows are the wavevectors of the leading and subleading charge densitywave instability. (C) The leading spin-singlet superconducting gap function de-rived from Eq. 1. The hatch size is proportional to the magnitude and the colorindicates the sign (red, −; blue, +). (D) The subleading singlet superconductinggap function derived from Eq. 1. The hatch size is proportional to the magnitudeand the color indicates the sign (red, −; blue, +). The gap functions in C and D arewell described by cos  kx − cos  ky and cos  kx + cos  ky , respectively.

17624 | www.pnas.org/cgi/doi/10.1073/pnas.1316512110 Davis and Lee

Dow

nloa

ded

by g

uest

on

Dec

embe

r 30

, 202

0

Page 3: Concepts relating magnetic interactions, intertwined ...electron interactions, IPs, and correlated SC. We demonstrate its effectiveness in simultaneously explaining the consequences

In the particle–hole interaction channel, Eq. 1 predicts (Methods)two types of instabilities: one preserves translation invariance (aQ=0 instability) and the other (a finite Q instability) does not. Withinour approach, the leadingQ = 0 instability is to a nematic state. Theorder parameter and the associated Fermi surface distortion areshown in Fig. 3A. This instability leads to the breaking of the crystal90° rotation symmetry, such as has been reported within the CuO2unit cell (21). The fact that the cuprate Fermi surface has sucha tendency to Pomeranchuk distort has been widely discussed (40).The subleading Q = 0 instability is twofold degenerate. The

order parameters and the associated Fermi surface distortions areshown in Fig. 3B andC. (Similar instabilities in hexagonal systemswere discussed recently in ref. 41.) Because the distorted Fermisurfaces are not k↔− k symmetric, these instabilities lead to timereversal symmetry breaking. In Fig. 4A, we show the ground statecurrent distribution produced by these order parameters in Fig.3 B and C. Depending on the quartic term in the Landau freeenergy expansion, the order parameters of Fig. 3 B and C cancoexist. In Fig. 4B, we show the ground state intra-unit-cell cur-rents associated with the symmetric and antisymmetric combi-nation of the order parameters in Fig. 3 B and C, respectively.Clearly this subleading time reversal breaking Pomeranchuk in-stability leads to states with the same broken symmetry as the loopcurrent states proposed in ref. 42 and not inconsistent withreported time reversal symmetry breaking in cuprates (43–45).However, it is important to stress that our Q = 0 instability doesnot lead to a pseudogap. Moreover, although this instability issubleading here, it is possible that material dependent detailsomitted in our simple effective action can change that.

The leading Q ≠ 0 instability in the particle-hole channel is acharge density wave (CDW) instability [in a recent preprint (28),a related idea was discussed]. The subject of CDWorder in cupratesuperconductors has a long history. An apparently bidirectionalmodulated CDW with only short range order is widely observedusing spectroscopic imaging scanning tunneling microscopy (21),but it was difficult to be certain these were true bulk phenomena.Therefore, for a long time the only bulk charge density wave orderthat was firmly established experimentally was the unidirectionalcharge density wave (stripes) (19) in the La2BaCuO4 family ofcompounds (20). Recently, however, signatures of apparent bi-directional CDW order have been observed by X-ray scattering inbulk YBa2Cu3O7 crystals (46–48).In Fig. 5 A–D, we present the leading CDW order parameters

that are generated by Eq. 1. Fig. 5 A–D represents the CDW orderparameters whose ordering wavevectors are the four horizontaland vertical black arrows connecting the hot spots in Fig. 2B. (Thegray arrows are the ordering wavevector of the subleading chargedensity wave order that we find. This result is different from theresult of ref. 28 where the gray arrows are the leading CDWwavevectors, perhaps due to the difference in the details of effec-tive interaction and bandstructure used in the two approaches.) Atthe quadratic level in a Landau free energy expansion, the orderparameters in Fig. 5 A and B are degenerate with those in panelsFig. 5 C and D. Depending on the coefficients of the fourth-orderterms, they can be either mutually exclusive (which results in aunidirectional CDW) or coexist (which results in a bidirectionalCDW). In Fig. 5E, we show the energy gap of a bidirectional CDWthat corresponds to the out-of-phase coexistence of the orderparameters in Fig. 5 A–D (48).Finally, the fact that there are both strong nematic (Fig. 3A)

and CDW (Fig. 5 A–D) susceptibilities implies that, in the pres-ence of disorder, which can serve as localized external orderingfields, locally nematic and CDW ordering can be induced to co-exist. This statement is consistent with the scanning tunnellingmicroscopy (STM) experiments (21). Such short-range disorderedinduced ordering can exist even when in the clean limit the systemis not yet long-range ordered.Obviously there is another key issue requiring discussion here: the

pseudogap of the cuprates. This unexplained gap to single-electronexcitations is anisotropic in k-space and appears at Tp � Tc forunderdoped cuprates (4, 8, 9, 11, 49). We hypothesize that theconsequences of an effective Hamiltonian as described in Eq. 1could also account for such a pseudogap. The various instabilities

A

B

C

Fig. 3. The Q = 0 intertwined particle-hole instabilities of the cuprates (Meth-ods). (A) The order parameter of the leadingQ= 0 (Pomeranchuk) instability andthe associated Fermi surface distortion of cuprates derived from Eq. 1. Here thehatch size is proportional to the magnitude, and the color indicates the sign(red, −; blue, +), and the dashed line marks the undistorted Fermi surface. Thisinstability breaks the 90° rotation symmetry and leads to nematicity. (B andC) Theorder parameter of a degenerate pair of subleading Q = 0 instabilities and theassociated Fermi surface distortions. Because the distorted Fermi surfaces are notk↔−k symmetric, these instabilities lead to time reversal symmetry breaking.

Fig. 4. The ground state current of the T-breaking Q = 0 particle-holeinstabilities in the cuprates. (A) The ground state current distribution asso-ciated with the order parameters in Fig. 3 B and C. (B) The ground statecurrent distribution produced by the symmetric and antisymmetric linearcombination of the order parameters in Fig. 3 B and C. In A and B, thethickness of the arrow is proportional to the magnitude of the current.

Davis and Lee PNAS | October 29, 2013 | vol. 110 | no. 44 | 17625

PHYS

ICS

INAUGURA

LART

ICLE

Dow

nloa

ded

by g

uest

on

Dec

embe

r 30

, 202

0

Page 4: Concepts relating magnetic interactions, intertwined ...electron interactions, IPs, and correlated SC. We demonstrate its effectiveness in simultaneously explaining the consequences

(except those at Q = 0) discussed here can all gap out, at leastpartially, the single-electron excitation spectrum. However, due tothe intertwining of these instabilities, the order parameter mayfluctuate from one type to another. This fluctuation would preventthe system from becoming long range ordered without eliminatingthe actual pseudogap for the single-electron excitations.Thus, we consider the order parameters of different inter-

twined orders to form a multicomponent supervector. The mag-nitude of the supervector is then responsible for the single-particle gap. The direction of the supervector is the soft degree offreedom, which ultimately determines the long ranged order ofthe system. However, although this direction fluctuates, the sin-gle-electron excitation spectrum remains gapped. In our case,when the gap is partial, the low energy excitations include boththe directional fluctuations of the supervector and the remaininggapless single particle excitations. Of course because of the cou-pling with the collective excitations, these single-particle excita-tions can have unusual, e.g., non-Fermi liquid, properties.Now it remains to show with a supervector formed using the

AF, SC, and CDW order parameters that there is a pseudogap inthe single electron excitation spectrum, no matter where thesupervector points. The results are shown in Fig. 6. In Fig. 6A,the supervector points in the AF direction. This supervector is shownby the red arrow on the order parameter sphere on the right. Such anorder has the biggest effect at the hot spots (Fig. 2B) where the gap ismaximal (Fig. 6A,Left). Here and in other panels, a vanishing single-particle gap at any point on the Fermi surface means that, along thenormal direction, there remains Fermi crossing, i.e., the Fermi sur-face has either moved or reconstructed. The supervector in Fig. 6Bpoints in the SC direction. The gap spectrum shown on the left is thefamiliar d-wave gap. In Fig. 6C, the supervector points in the CDWdirection. The energy gap spectrumon the left shows a nodal feature.The supervector in Fig. 6D lies in the plane spanned by the SC andthe CDW but directionally between the two. Finally, in Fig. 6E, thesupervector points in a generic direction. Obviously in reality, dif-ferent components of the supervector do not have to have the samenorm so that the fluctuations of the supervector actually occur ona spheroid; hence, there is no enlarged symmetry.Wehope thisfiguremakes the heuristic case that a pseudogap can also be a consequenceof the effective interaction in Eq. 1, when the effects of fluctuatingintertwined order parameters are dominant. Many of the anomalousphysical properties in the pseudogap state could then be attributed toorientational fluctuations of this intertwined supervector.

Iron-Based SuperconductorsNext we carry out the equivalent exercise for the iron-basedsuperconductors (12–14). The first term of Eq. 1 is studied here

using a five-band tight-bindingmodel with the Fermi surface shownin Fig. 7A. The blue and red lines mark the hole and electron Fermisurfaces, respectively. To simulate the magnetic correlation in iron-based superconductors, we include both the first (J1) and second(J2) neighbor interaction in Jij. (A similar Hamiltonian, with adoped Mott insulator basis, was proposed in ref. 50, 51.) This ef-fective interaction has been derived from the functional renorm-alization group calculation (52). Phenomenologically, there ismounting evidence that themagnetic correlations in the iron-basedmaterials are not due to Fermi surface nesting (53). It is then moreappropriate to view the second term in Eq. 1 as being generated byexcitations over the entire bandwidth. The essential differencefrom a Mott insulator here is the absence of a charge gap. There-fore, the generation of the effective magnetic interactions is moregradual. Using these inputs for J2=J1 ≥ 0:7, we find (Methods) theleading and subleading SC order parameter shown in Fig. 7 B andC, respectively. The leading gap function has the S± symmetry (54)and the subleading one has dx2−y2 symmetry (13, 14).In the particle-hole interaction channel, we find that (Methods)

the iron-based superconductors also have strong Q = 0 instabilities.In Fig. 8A andB, we show the leading and subleading Fermi surfacedistortions that we determine from Eq. 1 when the distortion am-plitude is small. Here the undistorted Fermi surface is shown usingdashed lines. This result agrees with the functional renormalizationgroup findings (52). The leading Fermi surface distortion preservesthe point group symmetry of the crystal. Note that because bothelectron and hole pockets expand (or shrink), it preserves the totalcharge density. (We note that a large amplitude distortion of thistype can drive the system to undergo a semimetal to insulatortransition.) The subleading Q = 0 instability breaks the 90° rotationsymmetry. Although it is subleading at the quadratic level of theLandau free energy expansion, it can become leading once thecubic coupling with the (strong) AF fluctuation is taken into ac-count [note that the AF order in the iron-based materials also

E

C D

A B

Fig. 5. The leading intertwined Q ≠ 0 particle-hole instabilities of the cuprates(Methods). The ordering wavevector is ð±δ, 0Þ in A and B and ð0,±δÞ in C and D.The black arrows in Fig. 2B are approximately the ordering wavevectors. (E)The energy gap of CDW produced by the equal amplitude superposition of thecharge density wave depicted in A–D. The phase of the superposition is taken tobe +, +, −, −, and hence corresponds to a d-wave symmetry. (The energy gapassociated with the +, +, +, + superposition of A–D is similar.) The charge densitywave gap plotted in E is defined as the minimum energy gap of the mean-fieldHamiltonian in Methods along the momentum cut normal to the Fermi surfacebut within the energy thin shell. The smallness of this combined order param-eter near the nodes can give rise to the effect of Fermi arcs.

A B

DC

E

Fig. 6. Intertwined instabilities and the cuprate pseudogap. We represent thethree instabilities contained in Eq. 1 (AF, SC, and CDW) using a supervectorrepresenting the combination of the three order parameters. The admixtureof AF, SC, and CDW phases can then be indicated by using the location ofthe supervector on a sphere. In each panel from A to E, the direction of thesupervector is shown as the red arrow on the order parameter sphere on theright of the same panel. The gap shown on the left of each panel is the min-imum energy gap of the mean-field Hamiltonian in Methods along the mo-mentum cut normal to the Fermi surface but within the energy thin shell. Thesize of the hatch is proportional to the value of the single particle energy gap.

17626 | www.pnas.org/cgi/doi/10.1073/pnas.1316512110 Davis and Lee

Dow

nloa

ded

by g

uest

on

Dec

embe

r 30

, 202

0

Page 5: Concepts relating magnetic interactions, intertwined ...electron interactions, IPs, and correlated SC. We demonstrate its effectiveness in simultaneously explaining the consequences

breaks the 90° rotation symmetry (55)]. In Fig. 8C, we show theeffect of the symmetry breaking distortion we find on the orbitaloccupation nxzðkÞ− nyzðkÞ. The fact that one needs magneticfluctuations to stabilize the C4-breaking Fermi surface distortion isconsistent with the arguments presented previously (12, 56, 57).Thus, the result in Fig. 8 B andC can explain the ubiquitous nematicordering found in the iron-based superconductors (24, 25). It alsoaccounts for the photoemission observation of the substantial dif-ference in the dxz and dyz orbital occupation in the nematic-distortedstate (58).Within our approach of Eq. 1, iron-pnictides show a very weak

Q ≠ 0 CDW instability. The ordering wavevector of the leadingCDW is approximately (π, π). However, due to the poor overlapsbetween the Fermi surfaces on the (π, π) displacement and thefact that (π, π) only approximately connects electron with electronor hole with hole pockets, a weak CDWwill not gap out the Fermisurfaces. Therefore, we will not devote more space to consider-ation of the CDW instability in pnictides.

Heavy Fermion SuperconductorsFinally, we use this same conceptual framework to consider the ca-nonical heavy fermion superconductor CeCoIn5. The band structurewe used for this material is the one given in ref. 59. The tight-bindingmodel consists of two orbitals per unit cell: the Wannier orbitalsassociated with the light and heavy bands, respectively. The Fermisurface is shown in Fig. 9A. The Jij we use to emulate the AF cor-relation in CeCoIn5 is the simple nearest neighbor interaction. Withthese inputs, we determine fromEq. 1 the leading and subleading SCorder parameter; the results are shown in Fig. 9B andC. The leadingsuperconducting gap function, with dx2−y2 symmetry (Methods), isin excellent agreement with that recently determined by the STMquasiparticle interference spectroscopy (59). The reason that the SCgap primarily opens on the large Fermi surface centered at (π, π) is

because the hot spots associated with the AF scattering all reside onthat Fermi surface. This fact is shown in Fig. 9D. Like the cuprates,the subleading SC gap function has extended s-wave symmetry.In the particle-hole channel, our general approach in Eq. 1 also

predicts that CeCoIn5 has Q = 0 and Q ≠ 0 instabilities (Methods).The order parameter and the Fermi surface distortion associatedwith the leadingQ= 0 instability is shown in Fig. 10A. This distortionbreaks the crystal 90° rotation symmetry and leads to nematicity. It isvery interesting that, like the cuprates, the subleadingQ= 0 instabilityis also to a degenerate pair of time reversal symmetry breaking states.The order parameter and the distorted Fermi surfaces are shown inFig. 10B andC. The obvious similarity between theQ= 0 instabilitiesin the heavy fermions and the cuprates is quite striking.The order parameter of the leading (weak) Q ≠ 0 charge density

wave instability is shown in Fig. 11 A–D. The energy gap producedby the in-phase coexistence of the order parameters in Fig. 11 A andB with those in Fig. 11 C and D is shown in Fig. 11E. Experimentalsearches of the signatures of these instabilities are under way.Searching for instabilities intertwined with superconductivity

in heavy fermion compounds now seems an important futuredirection. However, one must bear in mind that the equivalentchemical pressure places CeCoIn5 near the optimal pressure,where the SC transition temperature is the highest. The intertwinedinstabilities tend to occur near the junction between AF and SC.Therefore, unless negative pressure can somehow be applied, theycan remain out of reach for CeCoIn5. A better system for realizingintertwined instabilities is CeRhIn5 which is AF at ambient pressure.

A B

C

Fig. 7. Model Fermi surface and superconducting states of iron pnic-tides. (A) The pnictide-first BZ when not in the AF state. It spans a range−π=a< kx ≤ kx=a;−π=a< ky ≤ π=a, where a is the dimension of a unit cell con-taining only one Fe atom (we neglect the effects on unit cell definition of theout of plane As atoms). Our model Fermi surface of iron-pnictides using a five-band tight-binding model is shown as five closed contours: two red (outlinethe electron pocket) and three blue (outline the hole pocket). (B) The leadingspin singlet superconducting gap function derived from Eq. 1. The symmetry isS±. The hatch size is proportional to the magnitude, and the color indicates thesign (red, −; blue, +). (C) The subleading singlet superconducting gap functionderived from Eq. 1. The symmetry is dx2−y2 . The hatch size is proportional tothe magnitude, and the color indicates the sign (red, −; blue, +). The resultsin B and C are obtained using J2=J1 = tan  0:3π.

A

B

C

Fig. 8. Leading intertwined instabilities of iron-pnictides. (A) The leading Q =0 instability and the associated Fermi surface distortions in the iron-basedsuperconductors derived from Eq. 1. The hatch size is proportional to the mag-nitude, and the color indicates sign (red, −; blue, +). The dashed lines mark theundistorted Fermi surface. The area of both electron and hole pockets expand(or shrink) so that the total charge density is kept constant. This distortion doesnot break any symmetry and hence is difficult to pin down. (B) The subleadingQ = 0 instability and the associated Fermi surface distortion derived from Eq. 1.This distortion breaks the 90° rotation symmetry and couples to the stripe-likeunidirectional AF correlation strongly. Although the instability in B is sub-leading at the quadratic level of the Landau free energy expansion, it canbecome leading once the cubic coupling with the AF order parameter is takeninto account. (C) The effect of the Fermi surface distortion in B on the orbitaloccupation nxzðkÞ−nyzðkÞ.

Davis and Lee PNAS | October 29, 2013 | vol. 110 | no. 44 | 17627

PHYS

ICS

INAUGURA

LART

ICLE

Dow

nloa

ded

by g

uest

on

Dec

embe

r 30

, 202

0

Page 6: Concepts relating magnetic interactions, intertwined ...electron interactions, IPs, and correlated SC. We demonstrate its effectiveness in simultaneously explaining the consequences

By carefully studying the pressure–temperature phase diagram, onemight be able to find similar phenomena as in underdopedcuprates. If so, this will give additional support for applicabilityof the simple theory envisioned in Eq. 1.

ConclusionFrom the above studies, using the simple concept of a controllinginfluence of AF electron–electron interactions, it seems fair to saythat the low-energy effective Hamiltonian given by Eq. 1 can bevery useful in achieving an elementary understanding of the SC andthe intertwined instabilities in several canonical classes of un-conventional superconductors. Specifically we note that thesestudies demonstrate why, although superconductivity is universal,the nature of the Fermi surface distortion and/or the density waveinstabilities depends so much on the details of the fermiology. Suchdependence is the reason why the intertwined electronic orderedstates in correlated SC compounds are so strongly material de-pendent. Moreover, due to these distinct intertwined orders, Eq. 1does not describe a “nearly antiferromagnetic Fermi liquid.” Thus,our approach indicates that many of the anomalous properties ofthe cuprates and the pnictides may be due to the fluctuations of theorder parameter among the relevant intertwined orders, whereasthe severity of these fluctuations can be material dependent. Weunderstand that the point of view presented here is simplified.However, with a goal of identifying concepts that can simplyrelate strong AF electron–electron interactions, intertwinedelectronic ordered phases, and strongly correlated SC in distinctmaterial types, this is perhaps a good thing. We hope that the ap-proach presented here can help to distill the essence of theunconventional pairing mechanism from the impressive phenom-enology of the IPs in present and future strongly correlated hightemperature superconductors.

MethodsWe used the following procedures to determine the favored competingorders from the effective Hamiltonian, starting with

Heff =Xk

′Xs

eðkÞnsðkÞ+Xi,j

JijS→

i · S→

j , [2]

where nsðkÞ=ψ†k,sψk,s, and ψ†

k,s creates an electron in the single-particleeigenstate at momentum k with spin s. As mentioned in the text, Σk restricts

the sum to single particle eigenstates whose energy is within a thin shellfrom the Fermi energy.

First, we reexpress the second term in terms of the band eigen basis:

Xi,j

JijS→

i · S→

j =1A

Xk,p,q

Xs1,2,3,4

Vqðk;pÞψ†k+q,s1 σ

→s1 ,s2ψk,s2 ·ψ

†p−q,s3σ

→s3 ,s4ψp,s4 , [3]

where

Vqðk;pÞ= JðqÞnϕ*αðk+qÞðk+qÞ ·ϕαðkÞðkÞ

o×nϕ*αðp−qÞðp−qÞ ·ϕαðpÞðpÞ

o:

Here, A is the total area, ϕ is the band eigen wavefunctions in the orbitalbasis, and JðqÞ is the Fourier transform of Jij . For the copper-based, iron-based, and heavy fermion superconductors, JðqÞ is taken to be an overallcoupling strength Jeff times the following form factors:

cos  kx + cos kyðcopper−basedÞcos  θ

�cos  kx + cos  ky

�+ sin  θ

�2  cos kxcos ky

�ðiron−basedÞcos kx + cos kyðheavy− fermionÞ:

[4]

Jeff is a renormalized coupling strength that is unknown a priori. The resultfor the iron-based superconductors were generated with θ= 0:3π. In Vqðk;pÞ,the band index, e.g., αðk+qÞ, is defined to be the index of the band that isclosest to the Fermi energy at momentum k + q. If the corresponding singleparticle state has energy beyond the energy shell, ϕ is set to zero. In Eq. 4,ϕ is unity, a two-component vector, and a five-component vector for thecuprates, CeCoIn5, and pnictides, respectively. For CeCoIn5, if onedecides to include the magnetic interaction between the f electronsonly, one needs to replace ϕ*

αðpÞðpÞ ·ϕαðqÞðqÞ in Eq. 4 by ϕ*2,αðpÞðpÞϕ2,αðqÞðqÞ,

where 2 labels the f electron Wannier orbitals. The results for CeCoIn5

remain qualitatively unchanged using either formula.The next step is to decouple Eq. 3 in the particle-particle (for Cooper

pairing) and particle-hole (for charge and spin density wave and Pomer-anchuk distortion). The first-instability-mode analysis described here allowsus to determine the functional form of the order parameter. However, itdoes not fix the overall magnitude. Once the functional form is determined,we use the mean-field Hamiltonians described here to determine the energygaps, Fermi surface distortions, etc. The overall magnitude of the order

A B

DC

Fig. 9. Fermi surface and unconventional superconducting states of the heavyfermion compound CeCoIn5. (A) The first BZ and Fermi surface associated witha two-band band structure in ref. 58. The BZ spans a range −π=a< kx ≤ π=a;−π=a< ky ≤ π=a, where a is the dimension of a unit cell. The leading (B) andsubleading (C) spin singlet SC gap functions. The leading gap function has dx2−y2

symmetry and the subleading one has extended S symmetry. In B andC, the hatchsize is proportional to the magnitude of the gap, and the color indicates the sign(red, −; blue, +). (D) The Fermi surface and hot spots (the pink dots) of CeCoIn5.

A

B

C

Fig. 10. The intertwined Q = 0 particle-hole Instabilities of CeCoIn5. (A) Theleading Pomeranchuk instability and the associated Fermi surface distortion. Thehatch size is proportional to themagnitude of the order parameter, and the colorindicates the sign (red, −; blue, +). The dashed line marks the undistorted Fermisurface. This Fermi surface distortion leads to the breaking of the 90° rotationsymmetry. (B and C) The degenerate pair of subleading Pomeranchuk instabilitiesand their Fermi surface distortions. In both panels, the distorted Fermi surfaces donot respect thek↔− k symmetry. Consequently time reversal symmetry is broken.

17628 | www.pnas.org/cgi/doi/10.1073/pnas.1316512110 Davis and Lee

Dow

nloa

ded

by g

uest

on

Dec

embe

r 30

, 202

0

Page 7: Concepts relating magnetic interactions, intertwined ...electron interactions, IPs, and correlated SC. We demonstrate its effectiveness in simultaneously explaining the consequences

parameter is chosen to yield approximately the same maximal energy gapwhen each order parameter exists alone. The purpose is to convey thequalitative features not to make quantitative comparative predictions.

Cooper Pairing. In the particle-particle channel, we focused on the spin singletCooper pairing. This restriction leads to the following factorization of Eq. 3:

HMF =Xk

′Xs

eðkÞnsðkÞ+ 3A

Xp,k

Xa,b

Vscðp;kÞ

×nψ+−k,aψ

+k,bΔðpÞeab +Δ* ðkÞebaψp,aψ−p,b +Δ* ðkÞΔðpÞe2ab

o:

[5]

Here a and b label the spin and e↑↓ = − e↓↑ = 1 and e↑↑ = e↓↓ = 0. In Eq. 5

Vscðp;kÞ= Jðp− kÞ�ϕ*αðkÞð−kÞ ·ϕαðpÞð−pÞ

�×�ϕ*αðkÞðkÞ ·ϕαðpÞðpÞ

�: [6]

We then integrate out the electrons and keep up to the quadratic terms in Δ′s.The result is the following free energy form

1A

Xk,p′ ΔðpÞKT ðp;kÞΔ* ðkÞ, [7]

where

KT ðp;kÞ= 6Vscðp;kÞ− 36Z

d2q

ð2πÞ2Vscðp;qÞχT ðqÞVscðq;kÞ, [8]

where the temperature (T)-dependent free fermion pair susceptibility is given by

χT ðkÞ∝1− 2fðeðkÞÞ

eðkÞ : [9]

Here the proportionality constant is unimportant for our purposes because itcan be absorbed into the unknown Jeff (see below).

The leading (subleading) gap functions are the eigenfunctions ofMT ðq;kÞ=χT ðqÞVscðq;kÞ with the largest (second largest) eigenvalue. (The proportionalityconstant in χT changesall eigenvaluesby the samemultiplicative constantbutnotthe eigenfunctions.) These eigenfunctions are the order parameters that willfirst (second) become unstable as Jeff increases (at a temperature T much lessthan the thickness of the energy shell). These eigenfunctions are obtained nu-merically after discretizing the momentum space enclosed by the energy shell[under such discretization,MT ðq, kÞ becomes a matrix]. We diagonalize theMT

matrix and then average the eigenfunctions along the direction perpendicularto the Fermi surface. This procedure leads to the results presented in the text.

CDW and Pomeranchuk Instability. CDW and Pomeranchuk instability occurs inthe spin singlet particle-hole channel. Decoupling Eq. 3 in this channel leadsto the following mean-field Hamiltonian:

HMF =Xk

′Xs

eðkÞnsðkÞ

−3A

Xk,p,Q

Vcdw ðp; kÞ×nΔQðpÞψ†

k,aψk+Q,a +ψ†p+Q,aψp,aΔ*

QðkÞ− 2ΔQðpÞΔ*QðkÞ

o:

[10]

Here

Vcdw ðp; kÞ= Jðp− kÞ�ϕ*αðp+QÞðp+QÞ ·ϕαðk+QÞðk+QÞ

�×�ϕ*αðkÞðkÞ ·ϕαðpÞðpÞ

�:

[11]

Again, we integrate out the fermions to arrive at the following quadratic freeenergy form:

1A

Xk,p,Q

ΔQðpÞ~KQ,T ðp; kÞΔ*QðkÞ, [12]

where

~KQ,T ðp; kÞ= 6Vcdw ðk;pÞ− 9Z

d2q

ð2πÞ2 Vcdw ðp;qÞ~χQ,T ðqÞVcdw ðq; kÞ: [13]

Here the free fermion particle-hole susceptibility is given by

~χQ,T ðqÞ∝f�eðq+QÞ�− f

�eðqÞ�

eðqÞ− eðq+QÞ : [14]

The leading order parameter is the eigenfunction of ~MQ,T ðq, kÞ=~χQ,T ðqÞVcdw ðq;kÞ with the largest eigenvalue. Here we have to search boththe ordering wavevector Q as well as the leading form factor. This task isagain achieved numerically after discretizing the momentum space withinthe energy shell and diagonalizing the resulting matrix ~MQ,T . As before, weperform an average of the eigenvector along the direction perpendicular tothe Fermi surface, which leads to the results presented in the text.

The Pomerahnchuk distortion is determined as the leading order pa-rameter in the Q→ 0 limit of MQ,T . In our calculation, we always find bothQ = 0 and Q ≠ 0 instabilities.

Spin Density Wave. A spin density wave is a spin triplet particle-hole instability.Decoupling Eq. 3 in this channel leads to the followingmean-field Hamiltonian:

HMF =Xk

′Xs

eðkÞnsðkÞ+ 1A

Xp,k,Q

Vsdwðp; kÞ

×�mQðpÞ ·ψ†

k,cσ→

cdψk+Q,d +m*QðkÞ ·ψ†

p+Q,aσ→

abψp,b −mQðpÞ ·m*QðkÞ

�,

[15]

where

Vsdwðp; kÞ= JðQÞ�ϕ*αðp+QÞðp+QÞ ·ϕαðpÞðpÞ

�×�ϕ*αðkÞðkÞ ·ϕαðk+QÞðk+QÞ

+12Jðp− kÞ×

�ϕ*αðp+QÞðp+QÞ ·ϕαðk+QÞðk+QÞ

�×�ϕ*αðkÞðkÞ ·ϕαðpÞðpÞ

�:

[16]

Like before, we integrate out the fermions. The resulting quadratic freeenergy form reads to

1A

Xk,p

Ksdw,Q,T ðp; kÞmQðpÞ ·m*QðkÞ, [17]

where

Ksdw,Q,T ðp; kÞ= −Vsdwðk;pÞ− 2Z

d2q

ð2πÞ2Vsdwðp;qÞ~χQ,T ðqÞVsdwðq; kÞ: [18]

Here the free fermion particle-hole susceptibility is given by Eq. 14. The leadingorder parameter is the eigenfunction of ~MQ,T ðq, kÞ=~χQ,T ðqÞVsdw ðq; kÞ with theminimum eigenvalue. As before, we search the leading order parameter nu-merically after discretizing the momentum space within the energy shell.

Pseudogap of the Cuprates. Fig. 6 is generated by superposing the orderparameter terms in Eq. 5 (SC), Eq. 10 (CDW), and Eq. 15 (spin density wave)to form a grand mean-field Hamiltonian. If we include all necessary com-ponents, the supervector with AF, SC, and CDW order as components willhave 3 + 2 + 2 + 2 = 9 components. [The last 2 + 2 is the number of com-ponents of the CDW, associated with the ð± δ, 0Þ and ð0,± δÞ order.] Thislarge number is too complex to handle and impossible to present theresults. We simplify the situation to a supervector with only three compo-nents. The first component is the superconducting order. Here we restrict thephase of the superconducting order parameter to be real. The second compo-nent is the CDW order. For this order parameter we choose the bidirectional

A B

DC

E

Fig. 11. The leading intertwined Q ≠ 0 particle-hole Instabilities of CeCoIn5.(A–D) The leading charge density wave order parameter. (A and B) The or-dering wavevectors are ± ð0:56π,0:26πÞ. (C and D) The ordering wavevectorsare ± ð0:26π,0:56πÞ. (E) The energy gap produced by the in-phase coexistenceof order parameters in A–D. The hatch size is proportional to the magnitude,and the color indicates the sign (red, −; black and blue, +).

Davis and Lee PNAS | October 29, 2013 | vol. 110 | no. 44 | 17629

PHYS

ICS

INAUGURA

LART

ICLE

Dow

nloa

ded

by g

uest

on

Dec

embe

r 30

, 202

0

Page 8: Concepts relating magnetic interactions, intertwined ...electron interactions, IPs, and correlated SC. We demonstrate its effectiveness in simultaneously explaining the consequences

CDW corresponding to the in-phase superposition of the fundamental densitywave in two orthogonal directions. Moreover we set the the overall (sliding)phase of the CDW order parameter to be real. The third component is the AF.Here we restrict the order parameter to a point in a particular, say the z di-rection. In addition to the above simplification, we also tune the magnitudes ofthe order parameters so that they each gives rise to an energy gap of ap-proximately equal magnitude. This procedure leads to the following mean-field Hamiltonian:

HMF =Xk

′Xs

eðkÞnsðkÞ

+n1

(1A

Xp,k

Vsdwðp; kÞhfsdw,Qs ðpÞψ†

k,cσzcdψk+Qs ,d + f*sdw,Qs

ðkÞψ†p+Qs ,aσ

zabψp,b

i

+n2

(1A

Xp,k

Xa,b

3ea,bVscðp; kÞ×hfscðpÞψ+

−k,aψ+k,b + f*scðkÞψp,bψ−p,a

i)

+n3

(1A

Xk,p,Qc

ð−3ÞVcdw ðp;kÞhfcdw,Qc

ðpÞψ†k,aψk+Qc ,a + f*cdw,Qc

ðkÞψ†p+Qc ,aψp,a

i):

[19]

Here, Qs = ðπ, πÞ, Qc = ð±δ, 0Þ,ð0,±δÞ, and fsdw,Qs,fsc ,fcdwQc

are the form factorsof the leading order parameters determined previously, properly scaledto produce a similar maximum gap when each order parameter existsalone. The n1, n2, and n3 are the components of the supervector shownin Fig. 6. In general for incommensurate δ, the above mean-field Ham-iltonian couples infinite many k points together. The result presentedpreviously is obtained by truncating this infinite set to the following 10-element set: fk,k ± ðδ,0Þ,k ± ð0,δÞ,k + ðπ,πÞ,k+ ðπ,πÞ± ðδ, 0Þ,k+ ðπ,πÞ± ð0,δÞg.This truncation leads to a 20 × 20 Nambu matrix for each k. This matrix isdiagonalized numerically to determine the energy gap. The minimumenergy gap among all k (within the energy thin shell) for each directionnormal to the Fermi surface is plotted in Fig. 6.

ACKNOWLEDGMENTS. We thank S. A. Kivelson for a most useful discussionon theQ = 0 instabilities in the cuprates. We also thank D. K.Morr, M. Norman,and S. Sachdev for helpful discussions and communications. J.C.S.D. is sup-ported by the Center for Emergent Superconductivity, an Energy Frontier Re-search Center, headquartered at Brookhaven National Laboratory and fundedby the US Department of Energy (DOE), under Grant DE-2009-BNL-PM015.D.-H.L. is supported by the DOE Office of Basic Energy Sciences, Division ofMaterials Science, under Grant DE-AC02-05CH11231.

1. Anderson PW (1995) Physics: The opening to complexity. Proc Natl Acad Sci USA 92(15):6653–6654.

2. Schrieffer JR (1964) Theory of Superconductivity (W. A. Benjamin, New York).3. Monthoux P, Pines D, Lonzarich GG (2007) Superconductivity without phonons. Nature

450(7173):1177–1183.4. Norman MR (2011) The challenge of unconventional superconductivity. Science 332(6026):

196–200.5. Scalapino DJ (2012) A common thread: The pairing interaction for unconventional super-

conductors. Rev Mod Phys 84(4):1383–1417.6. Berg E, et al. (2009) Striped superconductors: How the cuprates intertwine spin, charge and

superconducting orders. New J Phys 11(11):115004.7. Fradkin E, Kivelson SA (2012) Ineluctable complexity. Nat Phys 8(12):864–866.8. Orenstein J, Millis AJ (2000) Advances in the physics of high-temperature superconductivity.

Science 288(5465):468–475.9. Damascelli A, Hussain Z, Shen ZX (2003) Angle-resolved photoemission studies of the cuprate

superconductors. Rev Mod Phys 75(2):473–541.10. Norman MR and Pepin C (2003) The electronic nature of high-temperature cuprate super-

conductors. Rep Prog Phys 66(10):1547–1610.11. Lee PA, Nagosa N, Wen X-G (2006) Physics of high-temperature superconductivity. Rev Mod

Phys 78(1):17–85.12. Paglione J, Greene R (2010) High-temperature superconductivity in iron-based materials.

Nat Phys 6:645–658.13. Wang F, Lee D-H (2011) The electron-pairing mechanism of iron-based superconductors.

Science 332(6026):200–204.14. Hirschfeld PJ, Korshunov MM, Mazin I (2011) Gap symmetry and structure of Fe-based su-

perconductors. Rep Prog Phys 74:124508.15. Heffner RH, Norman MR (1996) Heavy fermion superconductivity. Comments on Condensed

Matter Physics 17(6):361–408.16. Pfleiderer C (2009) Superconducting phases of f-electron compounds. Rev Mod Phys 81(4):

1551–1624.17. Coleman P (2007) Handbook of Magnetism and Advanced Magnetic Materials, ed Kronmuller

H, Parkin S. (John Wiley and Sons, Hoboken, NJ), Vol 1, pp 95–148.18. Saito G, Yoshida Y (2011) Organic superconductors. Chem Rec 11(3):124–145.19. Kivelson SA, et al. (2003) How to detect fluctuating stripes in the high-temperature super-

conductors. Rev Mod Phys 75(4):1201–1241.20. Fujita M, et al. (2012) Progress in neutron scattering studies of spin excitations

in high-Tc cuprates. J Phys Soc Jpn 81:011007.21. Fujita K, et al. (2012) Spectroscopic imaging scanning tunneling microscopy studies of elec-

tronic structure in the superconducting and pseudogap phases of cuprate high-Tc super-conductors. J Phys Soc Jpn 81:011005.

22. Bourges P, Sidis Y (2011) Novel magnetic order in the pseudogap state of high-Tc copperoxides superconductors. C R Phys 12(5-6):461–479.

23. Lawler MJ, et al. (2010) Intra-unit-cell electronic nematicity of the high-T(c) copper-oxidepseudogap states. Nature 466(7304):347–351.

24. Chuang TM, et al. (2010) Nematic electronic structure in the “parent” state of the iron-basedsuperconductor Ca(Fe(1-x)Co(x))2As2. Science 327(5962):181–184.

25. Chu JH, et al. (2010) In-plane resistivity anisotropy in an underdoped iron arsenide super-conductor. Science 329(5993):824–826.

26. Taillefer L (2010) Scattering and pairing in cuprate superconductors. Annu Rev CondensedMatter Phys 1:51–70.

27. Kanoda K (1997) Electron correlation, metal-insulator transition and superconductivity inquasi-2D organic systems, (ET)(2)X. Physica C 282:299–302.

28. Sachdev S, La Placa R (2013) Charge ordering in metals with antiferromagnetic spin correla-tions. Preprint at arXiv:1303.2114v4.

29. Laughlin RB (2013) Hartree-Fock computation of the high-Tc cuprate phase diagram. Preprintat arXiv:1306.5359.

30. Monthoux P, Pines D (1993) YBa2Cu3O7: A nearly antiferromagnetic Fermi liquid. Phys Rev BCondens Matter 47(10):6069–6081.

31. Sachdev S (2003) Colloquium: Order and quantum phase transitions in the cuprate super-conductors. Rev Mod Phys 75(3):913–932.

32. Doiron-Leyraud N, et al. (2007) Quantum oscillations and the Fermi surface in an underdopedhigh-Tc superconductor. Nature 447(7144):565–568.

33. Sebastian SE, Harrison N, Lonzarich GG (2012) Towards resolution of the Fermi surface inunderdoped high-Tc superconductors. Rep Prog Phys 75(10):102501.

34. Senthil T, Lee PA (2009) Synthesis of the phenomenology of the underdoped cuprates. PhysRev B 24:245116.

35. Scalapino DJ, Loh E, Jr., Hirsch JE (1986) d -wave pairing near a spin-density-wave instability.Phys Rev B Condens Matter 34(11):8190–8192.

36. Gros C, Joynt R, Rice TM (1987) Superconducting instability in the large-U limit of the two-dimensional Hubbard model. Zeitschrift Physik B Condensed Matter 68(4):425–432.

37. Kotliar G, Liu JL (1988) Superexchange mechanism and d-wave superconductivity. Phys Rev BCondens Matter 38(7):5142–5145.

38. Anderson PW (1987) The Resonating Valence Bond State in La2CuO4 and Superconductivity.Science 235(4793):1196–1198.

39. Baskaran G, Zhou Z, Anderson PW (1987) The resonating valence bond state and high Tcsuperconductivity: A mean field theory. Solid State Commun 63(11-12):973–976.

40. Fradkin E, et al. (2010) Nematic Fermi fluids in condensed matter physics. Annu Rev Con-densed Matter Phys 1:153–178.

41. Maharaj AV, Thomale R, Raghu D (2013) Particle-hole condensates of higher angular mo-mentum in hexagonal systems. Preprint at arXiv:1303.2361v1.

42. Varma CM (1999) Pseudogap phase and the quantum-critical point in copper-oxide metals.Phys Rev Lett 83(17):3538–3541.

43. Kaminski A, et al. (2002) Spontaneous breaking of time-reversal symmetry in the pseudogapstate of a high-Tc superconductor. Nature 416(6881):610–613.

44. Fauqué B, et al. (2006) Magnetic order in the pseudogap phase of high-Tc superconductors.Phys Rev Lett 96(19):197001.

45. Xia J, et al. (2008) Polar Kerr-effect measurements of the high-temperature YBa2Cu3O6+x

superconductor: evidence for broken symmetry near the pseudogap temperature. Phys RevLett 100(12):127002.

46. Chang J, et al. (2012) Direct observation of competition between superconductivity andcharge density wave order in YBa2Cu3O6.67. Nat Phys 8(12):871–876.

47. Ghiringhelli G, et al. (2012) Long-range incommensurate charge fluctuations in (Y,Nd)Ba2Cu3O(6+x).Science 337(6096):821–825.

48. Li JX, Wu CQ, Lee D-H (2006) Checkerboard charge density wave and pseudogap in high-Tccuprates. Phys Rev B 74:184515.

49. Timusk T, Statt B (1999) The pseudogap in high-temperature superconductors: An experi-mental survey. Rep Prog Phys 62(1):61–122.

50. Seo K, Bernevig BA, Hu JP (2008) Pairing symmetry in a two-orbital exchange coupling modelof oxypnictides. Phys Rev Lett 101(20):206404.

51. Si Q, Abrahams E (2008) Strong correlations and magnetic frustration in the high Tc ironpnictides. Phys Rev Lett 101(7):076401.

52. Zhai H, Wang F, Lee D-H (2009) Antiferromagnetically driven electronic correlations in ironpnictides and cuprates. Phys Rev B 80(6):064517.

53. Zhou KJ, et al. (2013) Persistent high-energy spin excitations in iron-pnictide superconductors.Nat Commun 4:1470.

54. Mazin II, Singh DJ, Johannes MD, Du MH (2008) Unconventional superconductivity with a signreversal in the order parameter of LaFeAsO1-xFx. Phys Rev Lett 101(5):057003.

55. Lynn JW, Dai P (2009) Neutron studies of the iron-based family of high Tc magnetic super-conductors. Physica C 469(9-12):469–476.

56. Chen F, et al. (2008) Theory of electron nematic order in LaFeAsO. Phys Rev B77(22):224509.

57. Xu C, Mueller M, Sachdev S (2008) Ising and spin orders in the iron-based superconductors.Phys Rev B 78(2):020501.

58. Yi M, et al. (2011) Symmetry-breaking orbital anisotropy observed for detwinned Ba(Fe1−xCox)2 As2above the spin density wave transition. Proc Natl Acad Sci USA 108(17):6878–6883.

59. Allan MP, et al. (2013) Imaging Cooper pairing of heavy fermions in CeCoIn5. Nat Phys 9(8):468–473.

17630 | www.pnas.org/cgi/doi/10.1073/pnas.1316512110 Davis and Lee

Dow

nloa

ded

by g

uest

on

Dec

embe

r 30

, 202

0


Recommended