14
Non-Fermi Liquid at (2 + 1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu, 1 Kai Sun, 2 Yoni Schattner, 3 Erez Berg, 3 and Zi Yang Meng 1 1 Beijing National Laboratory for Condensed Matter Physics and Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China 2 Department of Physics, University of Michigan, Ann Arbor, Michigan 48109, USA 3 Department of Condensed Matter Physics, Weizmann Institute of Science, Rehovot, Israel 76100 (Received 30 December 2016; revised manuscript received 17 July 2017; published 27 September 2017) We construct a two-dimensional lattice model of fermions coupled to Ising ferromagnetic critical fluctuations. Using extensive sign-problem-free quantum Monte Carlo simulations, we show that the model realizes a continuous itinerant quantum phase transition. In comparison with other similar itinerant quantum critical points (QCPs), our QCP shows a much weaker superconductivity tendency with no superconducting state down to the lowest temperature investigated, hence making the system a good platform for the exploration of quantum critical fluctuations. Remarkably, clear signatures of non-Fermi liquid behavior in the fermion propagators are observed at the QCP. The critical fluctuations at the QCP partially resemble Hertz-Millis-Moriya behavior. However, careful scaling analysis reveals that the QCP belongs to a different universality class, deviating from both ð2 þ 1ÞD Ising and Hertz-Millis-Moriya predictions. DOI: 10.1103/PhysRevX.7.031058 Subject Areas: Condensed Matter Physics, Strongly Correlated Materials I. INTRODUCTION Understanding the behavior of gapless fermionic liquids in the vicinity of a quantum phase transition is at the heart of strongly correlated electron systems, dating back to the celebrated Hertz-Millis-Moriya framework [13]. In particular, the question of Fermi-liquid instabilities at a magnetic (quantum) phase transition [49] and its applications to heavy-fermion materials [10,11] and transition-metal alloys (such as cuprates and pnictides [12,13]) is of vital importance and of broad interest to the condensed matter and high-energy physics communities. On the other hand, to be able to obtain the understanding of quantum critical phenomena in itinerant electron systems is extremely challenging. Recently, extensive research efforts have been devoted to this question, utilizing advanced renormalization group analysis, including the two-dimensional (2D) Fermi surface coupled to a Uð1Þ gauge field, to Ising-nematic, or to spin-density-wave bosonic fluctuations [1422]. Other approaches include dimensional regularization [23,24] and working in the limit of a large number of boson flavors [25]. Although important insights have been revealed from these studies, many fundamental questions still remain open. For exam- ple, will anomalous dimensions arise at such an itinerant quantum critical point (QCP)? The Hertz-Millis-Moriya theory predicts mean-field scalings. However, utilizing the effective field theory derived in Ref. [17], it was shown that divergence at four-loop order can lead to anomalous dimensions deviating from mean-field exponents [21]. In addition, the stability of these quantum critical points is also an open questions; e.g., will nonanalyticities in the momentum and frequency dependence of the theory lead to an instability towards a fluctuation-induced first-order transition or a fluctuation-induced second-order transition [26,27]? Moreover, various studies have suggested that near such a quantum critical point, critical fluctuations may trigger some other instability, resulting in a new phase that covers the QCP and masks the quantum critical region [28]. Experimentally, a ferromagnetic QCP was reported in a heavy-fermion metal [11]. Since analytical approaches are facing difficulties study- ing itinerant QCPs in a controlled manner, here, we study such a QCP by utilizing unbiased numerical calculations, i.e., the determinant quantum Monte Carlo (QMC) tech- nique, which has been demonstrated as a very effective tool for such problems [2937]. For the numerical studies on QCPs in itinerant fermionic systems, one major challenge lies in the fact that critical quantum fluctuations often trigger strong effective attractions between fermions, resulting in instabilities in the particle-particle channel. For example, in the recent studies on Ising nematic and charge-density-wave (CDW) QCPs, superconducting domes are observed cover- ing the QCPs [2933,35]. Although it is a very intriguing phenomenon that a QCP can induce unconventional super- conductivity, for the study of quantum criticality itself, the induced superconducting dome makes it difficult to obtain direct information about the critical point for the following Published by the American Physical Society under the terms of the Creative Commons Attribution 4.0 International license. Further distribution of this work must maintain attribution to the author(s) and the published articles title, journal citation, and DOI. PHYSICAL REVIEW X 7, 031058 (2017) 2160-3308=17=7(3)=031058(14) 031058-1 Published by the American Physical Society

Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical …Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical …Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and

Non-Fermi Liquid at (2 + 1)D Ferromagnetic Quantum Critical Point

Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and Zi Yang Meng11Beijing National Laboratory for Condensed Matter Physics and Institute of Physics,

Chinese Academy of Sciences, Beijing 100190, China2Department of Physics, University of Michigan, Ann Arbor, Michigan 48109, USA

3Department of Condensed Matter Physics, Weizmann Institute of Science, Rehovot, Israel 76100(Received 30 December 2016; revised manuscript received 17 July 2017; published 27 September 2017)

We construct a two-dimensional lattice model of fermions coupled to Ising ferromagnetic criticalfluctuations. Using extensive sign-problem-free quantum Monte Carlo simulations, we show that themodel realizes a continuous itinerant quantum phase transition. In comparison with other similar itinerantquantum critical points (QCPs), our QCP shows a much weaker superconductivity tendency with nosuperconducting state down to the lowest temperature investigated, hence making the system a goodplatform for the exploration of quantum critical fluctuations. Remarkably, clear signatures of non-Fermiliquid behavior in the fermion propagators are observed at the QCP. The critical fluctuations at the QCPpartially resemble Hertz-Millis-Moriya behavior. However, careful scaling analysis reveals that the QCPbelongs to a different universality class, deviating from both ð2þ 1ÞD Ising and Hertz-Millis-Moriyapredictions.

DOI: 10.1103/PhysRevX.7.031058 Subject Areas: Condensed Matter Physics,Strongly Correlated Materials

I. INTRODUCTION

Understanding the behavior of gapless fermionic liquidsin the vicinity of a quantum phase transition is at theheart of strongly correlated electron systems, dating backto the celebrated Hertz-Millis-Moriya framework [1–3].In particular, the question of Fermi-liquid instabilitiesat a magnetic (quantum) phase transition [4–9] and itsapplications to heavy-fermion materials [10,11] andtransition-metal alloys (such as cuprates and pnictides[12,13]) is of vital importance and of broad interest to thecondensed matter and high-energy physics communities.On the other hand, to be able to obtain the understanding

of quantum critical phenomena in itinerant electron systemsis extremely challenging. Recently, extensive researchefforts have been devoted to this question, utilizingadvanced renormalization group analysis, including thetwo-dimensional (2D) Fermi surface coupled to a Uð1Þgauge field, to Ising-nematic, or to spin-density-wavebosonic fluctuations [14–22]. Other approaches includedimensional regularization [23,24] and working in thelimit of a large number of boson flavors [25]. Althoughimportant insights have been revealed from these studies,many fundamental questions still remain open. For exam-ple, will anomalous dimensions arise at such an itinerantquantum critical point (QCP)? The Hertz-Millis-Moriya

theory predicts mean-field scalings. However, utilizing theeffective field theory derived in Ref. [17], it was shownthat divergence at four-loop order can lead to anomalousdimensions deviating from mean-field exponents [21]. Inaddition, the stability of these quantum critical points isalso an open questions; e.g., will nonanalyticities in themomentum and frequency dependence of the theory lead toan instability towards a fluctuation-induced first-ordertransition or a fluctuation-induced second-order transition[26,27]? Moreover, various studies have suggested thatnear such a quantum critical point, critical fluctuations maytrigger some other instability, resulting in a new phase thatcovers the QCP and masks the quantum critical region [28].Experimentally, a ferromagnetic QCP was reported in aheavy-fermion metal [11].Since analytical approaches are facing difficulties study-

ing itinerant QCPs in a controlled manner, here, we studysuch a QCP by utilizing unbiased numerical calculations,i.e., the determinant quantum Monte Carlo (QMC) tech-nique, which has been demonstrated as a very effective toolfor such problems [29–37]. For the numerical studies onQCPs in itinerant fermionic systems, one major challengelies in the fact that critical quantum fluctuations often triggerstrong effective attractions between fermions, resulting ininstabilities in the particle-particle channel. For example, inthe recent studies on Ising nematic and charge-density-wave(CDW) QCPs, superconducting domes are observed cover-ing the QCPs [29–33,35]. Although it is a very intriguingphenomenon that a QCP can induce unconventional super-conductivity, for the study of quantum criticality itself, theinduced superconducting dome makes it difficult to obtaindirect information about the critical point for the following

Published by the American Physical Society under the terms ofthe Creative Commons Attribution 4.0 International license.Further distribution of this work must maintain attribution tothe author(s) and the published article’s title, journal citation,and DOI.

PHYSICAL REVIEW X 7, 031058 (2017)

2160-3308=17=7(3)=031058(14) 031058-1 Published by the American Physical Society

Page 2: Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical …Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and

two reasons. (1) To accurately measure the critical expo-nents, it is important to examine the close vicinity of theQCP,which becomes challenging if theQCP is buried insidea newquantumphase. (2) It is known that quantumcriticalityis one source for non-Fermi liquid behaviors. In orderto understand and characterize such a non-Fermi liquidinduced by a QCP, it is important to suppress other orderingin the quantum critical region. As a result, obtaining apristine QCP without other induced instabilities becomescrucial for studying theseQCPs,which is onemain objectiveof this paper.In this paper, we construct a model of 2D fermions

interacting with gapless ferromagnetic (Ising) fluctuationsand use the determinant QMC technique to solve thisð2þ 1ÞD problem exactly [29–37]. Our QMC results areconsistent with such a pristine continuous quantum phasetransition in an itinerant ferromagnet in ð2þ 1ÞD, with nosuperconducting ordering at any coupling strengths anddown to the lowest temperature that we can access. Theabsence of superconductivity allows us to study the closevicinity of the QCP, where we found clear signatures ofnon-Fermi liquid behavior in the fermion propagators,induced by critical fluctuations. Furthermore, we find thatbecause of the coupling between fermions and bosonicmodes, the ferromagnetic QCP is different from an ordinaryð2þ 1ÞD Ising transition [38], but it also deviates from thepredictions of the Hertz-Millis-Moriya theory [6]. Hence,our results support the existence of a ferromagnetic QCPwith markedly non-Fermi liquid behavior. These resultsbroaden the theoretical understanding of itinerant quantumcriticality and make connections to the existing experimentphenomena.

II. MODEL AND METHOD

We consider a two-dimensional lattice model of itinerantfermions coupled to an Ising ferromagnet with a transversefield [Fig. 1(a)]. The Hamiltonian is comprised of threeparts,

H ¼ Hf þ Hs þ Hsf; ð1Þ

Hf ¼ −tXhijiλσ

c†iλσ cjλσ þ H:c:; ð2Þ

Hs ¼ −JXhiji

szi szj − h

Xi

sxi ; ð3Þ

Hsf ¼ −ξXi

szi ðσzi1 þ σzi2Þ: ð4Þ

The fermionic part Hf describes spin-1=2 fermions on asquare lattice with two independent orbitals per site(λ ¼ 1, 2). It includes a nearest-neighbor hopping termthat preserves the spin and orbital symmetries, where i andj label the sites, while σ and λ are the spin and orbitalindices, respectively. We work in the grand canonicalensemble, where the fermion density is set by the chemicalpotential μ. All energy scales are measured in units of t. Inaddition, each site of the square lattice has an Ising spinszi ¼ �1, whose quantum dynamics are governed by aferromagnetic transverse-field Ising model Hs. The Isingspins and the fermions are coupled via an on-site Ising termHsf, where σ

ziλ ¼ ðniλ↑ − niλ↓Þ=2 is the z component of the

fermion spin at orbit λ on site i.

i=1

=2Fermion site

Ising site

Coupling -t

Ising spin

Fermion

h

0

0.5

1

1.5

2

2 2.5 3 3.5 4

T

h

TN(h) ~ |h-hc|0.77(4)

hc = 3.270(6)

kx

ky

kx

ky

(a) (b)

FIG. 1. (a) Sketches of our model shown in Eq. (1). Two bands or orbits (λ ¼ 1, 2) of fermions move in a square lattice; at each site, thefermion spin is coupled to an Ising-gauge variable on the same lattice; and the Ising spins are subject to a ferromagnetic interaction Jamong themselves and a transverse magnetic field h. The quantum fluctuations of the transverse-field Ising model furthermore introduceeffective interactions among the fermions. (b) h − T phase diagram of our model in Eq. (1). The phase boundaries (orange data points)for coupling ξ ¼ 1 and μ ¼ −0.5 (fermion density hniλi ≈ 0.8) are the thermal transition points TNðhÞ from the ferromagnetic phase(FM) to the paramagnetic phase (PM). The FM phase is further highlighted by the shaded blue area. For comparison, we also plot thephase boundary without coupling (grey data points), i.e., that of the ð2þ 1ÞD transverse-field Ising model. The coupling changes theposition of the QCP (black dot to red dot along the h axis) as well as the power law of the phase boundary, which is described by criticalexponents ν and z [TNðhÞ ∼ jh − hcjzν] with νz ¼ 0.63 for ð2þ 1ÞD Ising at ξ ¼ 0 and TNðhÞ ∼ jh − hcjc with c ¼ 0.77ð4Þ for FMQCPat ξ ¼ 1. The insets are plots of the low-energy spectral weight,G½k; τ ¼ ðβ=2Þ�, in the FM phase (left) and the PM phase (right), shownfor L ¼ 24. Here, results from several different sets of twisted boundary conditions are superimposed (see Appendix B).

XU, SUN, SCHATTNER, BERG, and MENG PHYS. REV. X 7, 031058 (2017)

031058-2

Page 3: Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical …Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and

While, generically, models describing ferromagnetictransitions in fermionic systems suffer from the signproblem, here, the absence of the sign problem is guaran-teed by the introduction of the two orbitals λ ¼ 1 and 2.The two-orbital model is invariant under the antiunitarysymmetry iτyK, where τy is a Pauli matrix in the orbitalbasis and K is the complex conjugation operator; thus, it isfree of sign problems [39]. The details of the DQMCimplementation are presented in Appendix A.The Hamiltonian in Eq. (1) possesses an SUð2Þ ×

SUð2Þ × Uð1Þ ×Uð1Þ × Z2 symmetry, where the twoSUð2Þ symmetries are independent rotations in the orbitalbasis for spin up and spin down, the two Uð1Þ symmetriescorrespond to conservation of particle number with spin upand spin down, and the Z2 symmetry interchanges spin upand spin down while flipping the Ising spins sz → −sz.As h and T are reduced, the system undergoes a PM-FM

phase transition, spontaneously breaking the Z2 symmetry.In the absence of coupling between the Ising spins and thefermions, the transition belongs to the Ising universalityclass [40]. However, in the presence of the coupling ξbetween the Ising spins and the fermions, the systembecomes strongly correlated near the ferromagnetic QCP.Here, we focus on the properties of this exotic itinerantparamagnetic-ferromagnetic transition.

III. PHASE DIAGRAM

For ξ ¼ 0, the Ising spins are decoupled from thefermions, and the phase transition is governed by thetransverse-field Ising model. The phase diagram is shownin Fig. 1(b), where the phase boundary is marked bygrey data points. Our numerical studies confirm that thephase boundary ends at a QCP at T ¼ 0 with ð2þ 1ÞDIsing universality class [40]. Near the QCP, the transitiontemperature follows the scaling relation, TNðhÞ∼ jh−hcjνzwith hc ¼ 3.04ð2Þ and νz ¼ 0.63, consistent with theliterature [34,38].We now study the itinerant PM-FM transition by turning

on the coupling between the fermions and the Ising spins.We begin by setting the coupling strength ξ ¼ 1 andchemical potential μ ¼ −0.5, which gives rise to a fermiondensity hniλi ≈ 0.8. As shown in Fig. 1(b), turning on thecoupling shifts FM phase boundary (orange data points) tohigher values of T and h. At this coupling strength, down tothe lowest temperature that we have accessed, β ¼ 100(T ¼ 0.01), we observe no signature of any additionalordered phases near the QCP. We identify the finite-temperature ferromagnetic transition by a finite-sizescaling analysis of spin susceptibilities, as explained inAppendix C. Extrapolation towards zero temperature indi-cates that the itinerant PM-FM quantum phase transitionoccurs at hc ¼ 3.270ð6Þ and is of second order, but thescaling behavior near the QCP deviates strongly from thatof the ð2þ 1ÞD Ising universality class. For example, asshown in Fig. 1(b), the transition temperature TN scales as

TNðhÞ ∼ jh − hcjc with hc ¼ 3.270ð6Þ and c ¼ 0.77ð4Þ.Note that because of the itinerant nature of the QCP, theexponent c is no longer expected to obey the relationc ¼ zν [6].Because of the nonzero coupling ξ, the Fermi surface

structure changes across the FM transition. The fermioniclow-energy spectral weight shown in the inset of Fig. 1(b),extracted from the imaginary time Green’s functionGðτ ¼ β=2Þ [30,41], reveals the location of the Fermisurface. In the PM phase (h > hc), fermions with up anddown spins share the same Fermi surface because of thespin degeneracy (Ising symmetry) of the Hamiltonian [rightinset of Fig. 1(b)]. This degeneracy is lifted in the FM phase(h < hc) because of spontaneous symmetry breaking, andthus, the Fermi surface splits [left inset of Fig. 1(b)].At the QCP, low-energy fermionic excitations near the

Fermi surface become strongly coupled with the criticalbosonic spin fluctuations, offering an ideal platform forinvestigating the itinerant FM QCP. Below, we demonstratethat this QCP is stable down to low-energy scales and,more interestingly, that it is characterized by a dramaticbreakdown of Fermi-liquid behavior.

IV. NON-FERMI LIQUID BEHAVIOR

One of the key theoretical predictions for an itinerantQCP is that the critical bosonic fluctuations induce strongdamping of the fermions, resulting in a non-Fermi liquidwhere the low-temperature fermionic quasiparticle weightvanishes at the Fermi surface [14,16–18,22–24,42].Following Ref. [43], we measure the quasiparticle weightfrom the Matsubara-frequency self-energy

ZkF≈

1

1 − ImΣðkF;iω0Þω0

; ð5Þ

where Σ is obtained from our QMC simulations and kF isthe momentum at the Fermi surface (FS). In our finite-temperature simulations, the Matsubara frequencies takediscrete values, ωn ¼ πð2nþ 1ÞT. As an estimator for ZkF

,we use the first Matsubara frequency, ω0 ¼ πT in Eq. (5).The results are shown in Fig. 2(a) for h ¼ hc (squares)

and h > hc (circles). Here, we plot the quasiparticle weightat two different momentum points on the Fermi surface,i.e., kF along the kx direction (θ ¼ 0) and along the kx ¼ kydirection (θ ¼ ðπ=4Þ). In the paramagnetic phase (h > hc),ZðTÞ remains close to unity at low temperatures, indicatingwell-defined quasiparticles on the Fermi surface, asexpected in a Fermi liquid. In contrast, at the QCP(h ¼ hc), ZðTÞ is suppressed with decreasing temperatureand it extrapolates to zero at T → 0, which is the keysignature of a non-Fermi liquid. This is one of the keyfindings of this study.As shown in Fig. 2(a), at the QCP, as T → 0, ZðTÞ

decreases faster for kF along the diagonal direction

NON-FERMI LIQUID AT ð2þ 1ÞD FERROMAGNETIC … PHYS. REV. X 7, 031058 (2017)

031058-3

Page 4: Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical …Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and

(θ ¼ ðπ=4Þ) in comparison with other parts of the FS (e.g.,θ ¼ 0). This anisotropy is due to the anisotropy of theFermi surface. The fermion density in our simulation isaround 0.8, which is not far from perfect nesting (at half-filling). This near-nesting Fermi-surface geometry resultsin soft fermion-bilinear modes around Q ¼ ðπ; πÞ, whichwe have directly observed by measuring the fermion spinsusceptibility as shown in Appendix H. These soft modeshave a finite gap, and thus they are not the origin of the non-Fermi liquid behavior. However, scattering with these softmodes presumably introduces additional dampings for thefermions. Such scattering processes are stronger (weaker)for kF near the diagonal direction θ ¼ ðπ=4Þ (θ ¼ 0),where 2kF is close to (far away from) Q, and thus canlead to the observed anisotropy in ZðTÞ.In Fig. 2(b), we show the imaginary part of the self-

energy, ΣðωnÞ ¼ G−10 −G−1 [where GðωnÞ andG0ðωnÞ are

the fermionic Green’s function, and the Green’s functionof the noninteracting system, respectively] as a functionof the Matsubara frequency ωn. In the paramagnetic phase(circle symbols), −ImðΣÞ approaches zero linearly asωn → 0, as expected for a Fermi liquid. Such a behavioris not seen at the QCP, however, where we observe anincrease of −ImðΣÞ upon decreasing ωn, indicating astrong damping of the fermions at low frequencies. Onepossible mechanism for such surprising frequency depend-ence is related to thermal fluctuations of the FM orderparameter, as shown in Appendix I.The fermion Green’s functions, ImGðkF;ωnÞ, are

presented in Fig. 2(c). Note that −ImGðkF;ωnÞ is relatedto the single-fermion spectral function AðkF;ωÞ through−ImGðωnÞ ¼ R ½ðdωÞ=π�½ωn=ðω2

n þ ω2Þ�AðkF; ωÞ. Thefact that −ImðGÞ increases with decreasing ωn indicatesthat, both for h ¼ hc and h > hc and to the lowesttemperature we have accessed, there is no visible suppres-sion of the fermionic spectral weight at low frequency, i.e.,no sign of an opening of a gap in the fermionic spectrum (atleast down to frequencies of the order of ω ∼ πT). A similarconclusion can be reached by noting that −ImΣ is nevermuch larger than ω0; i.e., the self-energy never dominatesover the bare frequency dependence of the Green’s func-tion. This observation is consistent with the fact that there isno signature of a nearby superconducting phase for anyvalue of h, as shown in Sec. VI and Appendix G. Similarly,there is no signature of any other competing phase thatemerges close to the QCP (see Appendix H).

V. QUANTUM CRITICAL SCALING ANALYSIS

Near the quantum critical point, the bosonic criticalmodes become strongly renormalized by the coupling tothe gapless fermionic degrees of freedom (d.o.f.). As aresult, the universality class of the quantum critical pointis different from that of an ordinary Ising transition in(2þ 1) dimensions. Our QMC results indicate that thebehavior at the QCP resembles the behavior predicted byHertz-Millis theory but also deviates from it in a signifi-cant way. Below, we first summarize the Hertz-Millispredictions and then present a modified Hertz-Millisscaling formula, which fits our QMC data for the Isingspin susceptibility at all the momenta and frequenciessimulated.TheHertz-Millis-Moriya theory isbasedonquantumdynam-

icsobtainedfromtherandomphaseapproximation(RPA)[1–3].Within this approximation, the Ising spin susceptibility,χðh; T; q; ωnÞ ¼ ð1=L2Þ R dτ

Pije

iωnτ−iqrijhszi ðτÞszjð0Þi,takes the following form near the QCP,

χðh; T;q;ωnÞ

¼ 1

ctT2 þ chjh − hcj þ cqq2 þ cωω2 þ Δðq;ωnÞ; ð6Þ

0

0.5

1

0 0.05 0.1 0.15 0.2

0

0.05

0.1

0.15

-6

-4

-2

0

0 0.5 1 1.5 2

0° 45°T=0.1

T=0.07

T=0.05

0°45°

High

Low

Z(T

)

n

-Im

()

Im(G

)(a)

(b)

(c)

kx

ky

FIG. 2. (a) ZkFðTÞ at FM QCP (hc ¼ 3.27, squares) and in the

PM phase (h ¼ 3.60, circles). The left inset is the G½k; ðβ=2Þ� atthe FS for T ¼ 0.05, while the right inset is for T ¼ 0.1. Althoughthere is anisotropy in ZkF

at different parts of the FS, thequasiparticle weight in the kx and kx ¼ ky directions bothapproach zero at FM QCP, indicating a non-Fermi liquid behaviorfor the entire FS. The data in the PM phase show the quasiparticleweight approaching a constant (very close to 1), indicating thesystem is a Fermi liquid. Panel (b) shows −Im(ΣðkF;ωnÞ) at FMQCP (h ¼ hc, square symbol). It increases as ωn → 0 (signifyingthe system at QCP loses its quasiparticleweight with a power law),a non-Fermi liquid behavior, while in the PM phase (h ¼ 3.60,circle symbol), the imaginary part of the self-energy approacheszero linearly as ωn → 0—a Fermi-liquid behavior. (c) Imaginarypart of the single-fermion Green’s function at the FM QCP(h ¼ 3.27, square symbol) and in the PM phase (h ¼ 3.60, circlesymbol). No signature of the gap formation is observed.

XU, SUN, SCHATTNER, BERG, and MENG PHYS. REV. X 7, 031058 (2017)

031058-4

Page 5: Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical …Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and

where ct, ch, cq, cω are constants. Here, cqq2 þ cωω2 comesfrom the bare action of the Ising d.o.f., and theΔðq;ωnÞ term isthe contribution of the fermionic fluctuations. For an isotropic2D Fermi fluid, and for q and ωn much smaller than the Fermimomentum and energy, respectively,

Δðq;ωnÞ ¼ cHMjωnjffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

ω2n þ ðvfqÞ2

q ; ð7Þ

where cHM is a constant and vf is the Fermi velocity.A key property of Δðq;ωnÞ is its singular behavior inthe limit q → 0, ω → 0. Depending on whether onefirst takes the long-wavelength (q → 0) limit or thelow-energy (ω → 0) limit, Δðq;ωnÞ converges to dif-ferent values, limq→0limωn→0Δðq;ωnÞ ∼ ωn=q → 0, whilelimωn→0limq→0Δðq;ωnÞ ¼ cHM. This singularity is of greatimportance for the quantum dynamics at the QCP. It leads tothe following properties:

χðh ¼ hc; T ¼ 0;q;ωn ¼ 0Þ−1 ¼ cqq2; ð8Þ

χðh ¼ hc; T ¼ 0;q ¼ 0;ωnÞ−1 ¼ cHM þ cωω2n: ð9Þ

Beyond the RPA level, the scaling relation above can bemodified by higher-order terms, and the scaling analysis inRef. [2] suggests that the exponent forT shifts from 2 to 1 up tologarithmic corrections.Although these two relations differ by a constant (cHM),

note that the q and ωn dependence has the same scalingexponent. The reason for this behavior is that at q ¼ 0 andsmall ωn (or at ωn ¼ 0 and small q), Δðq;ωnÞ becomesindependent of q or ωn, respectively. Thus, the ωn or qdependence in χ−1 is dominated by the bare action of theIsing d.o.f. The fact that the exponents characterizing the ωand q dependence are the same reflects the emergentLorentz symmetry of the bare Ising action.Our QMC results share some characteristics with

this predicted form but with anomalous scalingdimensions. As shown in Figs. 3(a) and 3(b), weindeed find that limq→0limωn→0χ

−1ðq;ωnÞ differs fromlimωn→0limq→0χ

−1ðq;ωnÞ by a constant cHM ¼ 0.20ð4Þas predicted by the RPA. The ferromagnetic susceptibilityis found to be well described by the following formula:

χðh;T;q;ωnÞ

¼ 1

ctTat þ chjh− hcjγ þ ðcqq2 þ cωω2Þaq=2 þΔðq;ωnÞ;

ð10Þ

with an anomalous exponent aq ¼ 1.85ð3Þ. This is differ-ent both from the exponent for an Ising transition in (2þ 1)dimensions, 1.96, and from the Hertz-Millis value of 2.

The presence of an anomalous exponent is another keyfinding of this study.The functional form in Eq. (10) is analogous to the RPA

prediction [Eq. (6)], but it allows for non-mean-fieldexponents (at, γ, and aq). This form is found to fit allthe numerical data points [Fig. 3(c)].Note that Eq. (10) contains Δðq;ωnÞ, which has the

form of a free-fermion susceptibility. However, as will beshown below, within numerical resolution, as long as Δcaptures the singular behavior at q → 0 and ω → 0, thequality of the fit of χðh; T;q;ωnÞ is not sensitive to thedetailed functional form of Δðq;ωnÞ. More importantly,our key conclusions, e.g., the anomalous dimensionη ¼ 2 − aq ¼ 0.15, are fully independent of the particularchoice of Δðq;ωnÞ. Therefore, here we use the RPA form[Eq. (7)], where, for simplicity, we set vf to 2.By fitting with χðh; TÞ at q ¼ 0 and ωn ¼ 0, we find

that ct ¼ 0.13ð1Þ, at ¼ 1.48ð4Þ, ch ¼ 0.7ð1Þ, γ ¼ 1.18ð4Þ;details can be found in Appendix D. As shown above,by fitting with χ at q ¼ 0 or ω ¼ 0 [Figs. 3(a) and 3(b)],we find aq ¼ 1.85ð3Þ, cq ¼ 1.00ð2Þ, cω ¼ 0.10ð2Þ, andcHM ¼ 0.20ð4Þ. With all the fitting parameters fixed, wecan use Eq. (10) to collapse all χ data for all q, ωn, h, and T.

0

1

2

3

0 2 4 6

T=0.1T=0.025

0

0.5

1

1.5

2

0 0.5 1 1.5 2

-1

0

-0.5 0

F-1(h,T,q, n)

-1(h

,T, q

,n)

log(

-1(q

,0)-

-1(0

,0))

-1(0

,n)

--1

(0,0

)

log(|q|) n

cHM+(c )aq/2( n)aq0.5aqlog(cq)+aqlog(|q|)

(a) (b)

(c)

L=24L=28

FIG. 3. (a) Inverse spin susceptibility at ωn ¼ 0 as a function ofjqj (data points with L ¼ 24, 28 and T ¼ 0.125). The red lineshows the fitting with χ−1 ¼ cqqaq , and we get aq ¼ 1.85ð3Þ. Theblack dashed line shows the slope aq ¼ 2. (b) Inverse spinsusceptibility at q ¼ 0 as a function of ωn (data points withL ¼ 20, 24, and 28 for T ¼ 0.1, with L ¼ 20 for T ¼ 0.025). Thered curve shows the fitting with χ−1 ¼ cHM þ cωω

aωn . (c) Data

collapse for Ising susceptibility against the functional Eq. (10),where F−1¼ctTatþchjh−hcjγþðcqq2þcωω2Þaq=2þΔðq;ωnÞ.The dark violet squares (3946 in total, made up of T ¼ 1.0,0.83, 0.67, 0.5, 0.4, 0.3, 0.25, 0.2, 0.17, 0.13, 0.1;h ¼ 3.27,3.3,3.4,3.5,3.6,3.7,3.8,3.9,4.0; and L ¼ 24) are for datawith zero frequency, while the light blue circles (6096 in total,made up of L ¼ 20with T ¼ 0.1, 0.05, 0.033, 0.025, 0.014, 0.01,and L ¼ 24, 28 with T ¼ 0.1, 0.05) are for data with frequencydependence. The red line with y ¼ x serves as the baseline forcomparison of QMC data with the functional approximation.

NON-FERMI LIQUID AT ð2þ 1ÞD FERROMAGNETIC … PHYS. REV. X 7, 031058 (2017)

031058-5

Page 6: Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical …Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and

The results are presented in Fig. 3(c). Clearly, all the datacollapse onto the same curve, especially at small q, ωn, lowtemperature T, and h ∼ hc.ForΔðq;ωnÞ, we find that as long as limωn→0Δðq;ωnÞ¼0

and limq→0Δðq;ωnÞ ¼ cHM ¼ 0.20ð4Þ, modifying thefunctional form of Δðq;ωnÞ has little impact on χ withinnumerical error bars of the QMC data. This uncertainty inΔðq;ωnÞ prevents us from obtaining detailed informationabout the dynamics at the QCP (e.g., the precise valueof the dynamical critical exponent z). However, since bydefinition Δðq;ωnÞ vanishes at ωn ¼ 0, all the conclusionsregarding the static limit (ωn ¼ 0), including the anoma-lous dimensions η ¼ 2 − aq ¼ 0.15 obtained from the qdependence in χðq;ωn ¼ 0Þ, are independent of the detailsof Δðq;ωnÞ.

VI. SUPERCONDUCTIVITY

We now discuss the superconducting properties close tothe FM QCP. Unlike previous QMC studies of quantumcriticality in metals (such as the Ising-nematic, SDW, andother QCPs [29–33,35]), the FM QCP shows a substantialseparation of scales between the onset of superconductingcorrelations and the phenomena described thus far, i.e.,non-Fermi liquid behavior and quantum critical scaling,thus making this system an ideal platform for the inves-tigation of quantum fluctuations close to criticality.As described in Appendix E, the spin fluctuations induce

an attractive interaction in the spin-triplet channel. The two-band structure of the model allows for a number of distinctsuperconducting order parameters, of which we find thatthe strongest pairing tendencies occur in the orbital-singlet,spin-triplet channel, with the order parameter

Δi;σ ¼ ci1σci2σ; ð11Þwhere σ ¼ ↑, ↓ is the spin index. Indeed, this channel isfound to be the leading instability in a weak-coupling,mean-field analysis (see Appendix E). The two componentsΔ↑ and Δ↓ are related by the Z2 (spin-flip) symmetry of themodel and are therefore of equal magnitude in the mag-netically disordered phase. The order parameter transformsas a scalar under lattice rotations and reflections; hence, weexpect single-fermion excitations to be fully gapped in thesuperconducting state.Because fermions in our model only preserve a Uð1Þ

spin rotational symmetry, a finite-temperature triplet order-ing is, in principle, allowed, in contrast to 2D systems withSUð2Þ spin rotational symmetry where triplet pairing mayonly occur at zero temperature. As shown in Appendix F,the two-component nature of the order parameter allows,in principle, for more exotic phases such as a charge-4esuperconductor (4e SC) and a spin-nematic (SN) phase[44–47]. We have found no numerical evidence for theexistence of these phases, and therefore, we focus on thetriplet, charge 2e superconductor described above.

To probe for superconducting tendencies near theQCP,weexplore three parameter sets: fξ ¼ 1.0; μ ¼ −0.5; J ¼ 1.0g,fξ ¼ 1.5; μ ¼ −2; J ¼ 0.5g, and fξ ¼ 3.0; μ ¼ −2;J ¼ 0.5g. First, the finite-temperature FM-to-PM phasetransitions are identified by finite-size scaling of the Isingspin susceptibilities, as discussed in Appendix D. Then, asdepicted in Fig. 4, the FMQCPs are located by extrapolatingthe finite-temperature phase boundary towards T ¼ 0. Asexpected, the larger the coupling ξ, the higher the criticalfield hc at the FM QCP.Next, we calculate the pairing correlations for all pairing

channels, with on-site and nearest-neighbor form factors(see Appendix G). Among all the available channels, onlythe order parameter defined in Eq. (11) has pairingcorrelations peaked at the QCP. We conclude that it isthe only channel that is substantially enhanced by criticalFM fluctuations.Figure 5 shows the pairing structure factor C ¼

½1=ð2L2Þ�PijσhΔ†iσΔjσi. At all couplings, we find a peak

of the pairing correlations at the QCP. At ξ ¼ 1 andξ ¼ 1.5, the pairing correlations do not grow with thesystem size, indicating we are far from a superconductingtransition. At the strongest coupling, ξ ¼ 3, C grows, albeitvery slowly, with L, indicating that the correlation length ofthe superconducting order parameter is moderately large.Although pairing correlations in this channel are

enhanced close to the QCP, long-range or quasi-long-rangesuperconducting order never develops down to T ¼ 0.025for any of the three parameter sets. This conclusion iscorroborated by an analysis of the superfluid density,shown in Appendix G.

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

2 2.5 3 3.5 4 4.5 5

T

h/J

ξ=1.0 μ=-0.5 J=1.0ξ=1.5 μ=-2.0 J=0.5ξ=3.0 μ=-2.0 J=0.5

FIG. 4. Estimation of the FM QCP for larger coupling. Thereare three parameter sets in total: fξ ¼ 1.0; μ ¼ −0.5; J ¼ 1.0g,fξ ¼ 1.5; μ ¼ −2; J ¼ 0.5g, and fξ ¼ 3.0; μ ¼ −2; J ¼ 0.5g.The finite-temperature phase boundaries are determined fromthe Ising spin susceptibilities, as discussed in Appendix D. TheQCPs are denoted by the colored arrows. Close to each FM QCP,we scan h to measure various superconductivity instabilities. Theresults are shown in Fig. 5.

XU, SUN, SCHATTNER, BERG, and MENG PHYS. REV. X 7, 031058 (2017)

031058-6

Page 7: Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical …Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and

VII. DISCUSSION

In this work, we have constructed a lattice modelthat realizes a pristine ferromagnetic QCP in an itinerantFermi system. At the QCP, clear non-Fermi liquidbehavior is observed. The static (ωn ¼ 0) ferromagneticsusceptibility obeys scaling with an anomalous dimensionη ¼ 2 − aq ¼ 0.15, which deviates from both the ð2þ 1ÞDIsing value (η ¼ 0.036) and the mean-field value expectedfrom Hertz-Millis theory (η ¼ 0).The ωn dependence of the bosonic susceptibility is highly

nontrivial. In particular, the long-wavelength limit (q → 0)and the low-frequency limit (ωn → 0) do not commute. Tofully characterize the quantumdynamics at theQCP,we needto obtain detailed information about Δðq;ωnÞ in Eq. (10),especially in the low-energy limit. The presence of ananomalous dimension in aq and in transition temperatureTN suggests that Δ probably also deviates from the Hertz-Millis form,which has a dynamic critical exponent z ¼ 3 andthus predicts mean-field exponents. At the qualitative level,our conclusions and the observation of anomalous dimen-sions are in good agreement with the four-loop scalinganalysis in Ref. [21]. However, although our studies revealdirect information on the qualitative features of Δðq;ωnÞ(i.e., the singular behavior in the limit q → 0, ωn → 0), thedetailed functional form is beyond the resolution set by finite-size and finite-temperature effects of our QMC study.It is interesting to contrast the results of the present study

with those obtained for a related (but different) problem ofan itinerant Ising-nematic QCP [30,35]. The two problemshave essentially the same description within Hertz-Millistheory. Similarly to our results for an Ising ferromagneticQCP, the Ising nematicQCP displays strong deviations from

Fermi-liquid behavior. A key difference, however, is that inthe FMQCP, Fermi-liquid behavior is lost everywhere alongthe Fermi surface; i.e., there are no “cold spots” with long-lived quasiparticles. Furthermore, near the FM QCP, strongdeviations from Fermi-liquid behavior were found at tem-peratures much larger than any scale associated with super-conductivity. In the Ising-nematic problem, the Ising orderparameter correlations also show a singular behavior in theðq → 0;ωn → 0Þ limit. However, in the FM case presentedhere, the fermionic magnetic susceptibility is exactly con-served, unlike the Ising-nematic case where an approximateconservation was found. An additional key difference is theanomalous exponents detected in the present work.Perhaps the most striking difference between the two

models lies in their superconducting properties. While theIsing-nematic QCP was found to be strongly unstabletowards s-wave superconductivity [35], the leading super-conducting instability in the FM case is described by atwo-component, spin-triplet order parameter, with thepossibility of exhibiting interesting, unconventional phasessuch as a charge-4e superconductor.On the experimental side, our results may shed some new

light on the study about 2D or quasi-2D metallic materialswith Ising magnetic order, like Fe1=4TaS2 [48] andCeCd3As3 [49].

ACKNOWLEDGMENTS

The authors thank F. Assaad and S. Kivelson for helpfuldiscussions. X. Y. X. and Z. Y.M. are supported by theMinistry of Science and Technology of China under GrantNo. 2016YFA0300502, the National Science Foundation ofChina under Grants No. 11421092 and No. 11574359, andtheNational Thousand-Young-Talents ProgramofChina.Wethank the following institutions for allocation of CPU time:the Center for Quantum Simulation Sciences in the Instituteof Physics, Chinese Academy of Sciences; the Tianhe-1Aplatform at the National Supercomputer Center in Tianjin;and the Gauss Centre for Supercomputing e.V. [50] forproviding access to the GCS Supercomputer SuperMUC atLeibniz Supercomputing Centre (LRZ, [51]). K. S. is sup-ported by the National Science Foundation under GrantsNo. PHY-1402971 at the University of Michigan and theAlfredP. SloanFoundation.Y. S. andE. B.were supported bythe Israel Science Foundation under Grant No. 1291/12, bythe US-Israel BSF under Grant No. 2014209, and by aMarieCurie reintegration grant. E. B. was supported by an Alongrant. Y. S. and E. B. thank S. Lederer and S. Kivelson for acollaboration on related topics.

APPENDIX A: DETERMINANTAL QUANTUMMONTE CARLO IMPLEMENTATIONS

The determinantal quantum Monte Carlo (DQMC)formalism starts with the partition function of the originalHamiltonian. To efficiently evaluate the trace in thepartition function, discretized imaginary time is used and

0.38

0.39

0.4

3 3.2 3.4 3.6

C

h/J

ξ=1.0, μ=-0.5, J=1.0

L=12L=14L=16

0.18

0.19

0.2

0.21

3 3.5 4

h/J

ξ=1.5, μ=-2.0, J=0.5

L=12L=14L=16

0.22

0.23

0.24

0.25

3.5 4 4.5 5

h/J

ξ=3.0, μ=-2.0, J=0.5

L=12L=14L=16

FIG. 5. Static pairing-correlation function C ¼ ð1=L2ÞhΔ†Δifor order parameters defined in Eq. (11). For fξ ¼ 1.0; μ ¼ −0.5;J ¼ 1.0g, no enhancement of pairing correlation functionsis observed in any pairing channel down to T ¼ 0.025. Forfξ ¼ 1.5; μ ¼ −2; J ¼ 0.5g and fξ ¼ 3.0; μ ¼ −2; J ¼ 0.5g, thepairing order parameters Δ↑ and Δ↓ show enhanced correlationnear the QCP, in agreement with theoretical analysis. Noenhancement is observed in other pairing channels.

NON-FERMI LIQUID AT ð2þ 1ÞD FERROMAGNETIC … PHYS. REV. X 7, 031058 (2017)

031058-7

Page 8: Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical …Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and

β ¼ MΔτ (Δτ ¼ 0.05). As the studied model contains bothfermion and Ising d.o.f., the trace will involve both a sumover Ising spin configurations and a determinant aftertracing out the fermion d.o.f.

Z ¼ Tr½e−βH�¼

Xsz1���szN¼�1

TrFhsz1 � � � szN jðe−ΔτHÞMjsz1 � � � szNi: ðA1Þ

Let S ¼ ðsz1 � � � szNÞ denote the Ising spins; then,

Z ¼X

S1���SMTrFhS1je−ΔτHjSMihSMje−ΔτHjSM−1i

� � � hS2je−ΔτHjS1i: ðA2ÞNow we can trace out the fermion d.o.f. and obtain theconfigurational weight,

ωC ¼ ωTIC ωF

C ðA3Þwith the Ising part

ωTIC ¼

�Yτ

Yhi;ji

eΔτJszi;τs

zj;τ

��Yi

Yhτ;τ0i

Λeγszτ;is

zτ0 ;i

�; ðA4Þ

where Λ2 ¼ sinhðΔτhÞ coshðΔτhÞ, γ ¼ − 12ln ðtanhðΔτhÞÞ.

For the fermion part, we have

ωFC ¼ det ð1þBM � � �B1Þ: ðA5Þ

As an antiunitary symmetry iτyK (where τy is a Paulimatrix in the orbital basis and K is the complex conjugationoperator) makes the Hamiltonian invariant, the fermion partweight can be further rewritten as

ωFC ¼

����Yσ

detð1þ B1σM � � �B1σ

1 ����2

; ðA6Þ

where

Bλστ ¼ exp ð−ΔτKλσ þ Δτξdiagðsz1;…; szNÞÞ; ðA7Þ

withKλσ the hopping matrix for orbital λ and spin σ. It turnsout that both the fermion weight and the Ising weight arealways positive; thus, there is no sign problem. To system-atically improve the simulation, especially close to the(quantum) critical point, we have implemented both a localupdate in DQMC and a space-time global update [34]. In theglobal update, we use the Wolff algorithm [52] to proposespace-time clusters of the Ising spins and then calculatethe fermion weight to respect the detailed balance as theacceptance rate of the update. Further attempts, with therecently developed self-learning determinantal quantumMonte Carlo scheme [53–55], which can greatly reducethe autocorrelation at the (quantum) critical point and speedup the simulation with theOðNÞ fold, to access larger L andlower T are in progress.

APPENDIX B: Z-DIRECTION FLUX

To reduce spurious finite-size effects, we have used thetechniques introduced in Ref. [56]. The basic idea is tointroduce an effective z-direction flux by multiplying thehopping parameter by Peierls phase factors, i.e.,

−tc†iλσ cjλσ → −teiϕλσij c†iλσ cjλσ;

with ϕλσij ¼ ½ð2πÞ=Φ0�

R rjri A

λσðrÞ · dr.To make sure the model remains free of the sign

problem, we take

ϕ1↑ij ¼ ϕ1↓

ij ¼ −ϕ2↑ij ¼ −ϕ2↓

ij ; ðB1Þ

therefore, the applied flux is not a true magnetic field, as itcouples differently to fermions of different flavors.Translational invariance imposes restrictions on the

magnitude of the effective magnetic field, namely, Bλσ ¼nðΦ0=L2Þ, with n an integer andΦ0 the flux quanta. We usethe Landau gauge AλσðrÞ ¼ −Bλσyx in the bulk of thesystem, whereas special care is taken at the “edges”,

ϕλσiðx;y¼LÞ;jðx;y¼1Þ ¼

Φ0

BλσLx; ðB2Þ

ϕλσiðx;y¼1Þ;jðx;y¼LÞ ¼ −

Φ0

BλσLx: ðB3Þ

While applying a flux in the z direction dramaticallyimproves the convergence to the thermodynamic limit, italso breaks translation symmetry, making it impossible toextract the momentum dependence of fermionic correla-tions. Whenever such information is needed [such as forthe momentum-resolved, singlet-particle Green’s functionGkðωnÞ], we apply a flux in the x or y direction, which isequivalent to twisting the boundary conditions. Just as inthe case of the z-directed flux, choosing Eq. (B1) ensuresthe absence of the sign problem [57]. We have used sixkinds of twisted boundary conditions, (0,0), ðπ=2; 0Þ,ðπ=2; π=2Þ, ðπ; 0Þ, ðπ; π=2Þ, and ðπ; πÞ; thus, we get 16times higher resolution of momentum.

APPENDIX C: THERMAL PHASE TRANSITION

The thermal phase boundary in Fig. 1(b) of the main textis controlled by the 2D Ising critical exponents γ ¼ 7=4and ν ¼ 1, implying that the zero-frequency and zero-momentum Ising spin susceptibility around the finite-temperature critical field hN satisfies

χðh; T; 0; 0Þ ¼ Lγ=νfððh − hNÞL1=νÞ: ðC1Þ

Here, the Ising spin susceptibility is defined as

XU, SUN, SCHATTNER, BERG, and MENG PHYS. REV. X 7, 031058 (2017)

031058-8

Page 9: Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical …Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and

χðh; T;q; iωnÞ ¼1

L2

Xij

0

dτeiωnτ−iq·rijhszi ðτÞszjð0Þi:

ðC2Þ

Figure 6(a) illustrates the behavior of χðh; T; 0; 0Þ atfixed temperature T ¼ 0.5 as a function of transverse field;Fig. 6(b) is the data collapse according to Eq. (C1), fromwhich we can obtain hNðT ¼ 0.5Þ ≈ 3.06. Because of thecoupling between the fermions and the Ising spins, thefermions go through the same finite-temperature phasetransition, as shown in Fig. 7(a). The fermionic spinsusceptibility χFðh; T; 0; 0Þ, where

χFðh; T;q;ωnÞ ¼1

L2

ZdτXijλλ0

eiωnτ−iqrijhσziλðτÞσzjλ0 ð0Þi;

ðC3Þbehaves much like χðh; T; 0; 0Þ, and a data collapse, shownin Fig. 7(b), gives rise to the same hNðT ¼ 0.5Þ ≈ 3.06.

APPENDIX D: QUANTUM CRITICALSCALING ANALYSIS OF χ ðh;T;0;0Þ

In the main text, we have discussed the dynamic Isingspin susceptibility, χðh; T;q;ωnÞ, and performed the quan-tum critical scaling analysis. Here, we reveal more details.According to Eq. (10) in the main text, at q ¼ 0 and

ωn ¼ 0, we have

χðh; T; 0; 0Þ ¼ 1

ctTat þ chjh − hcjγ: ðD1Þ

We detect the power in T by measuring χðh ¼ hc; T; 0; 0Þ,and Fig. 8 shows the fitting of χðh ¼ hc; T; 0; 0Þ−1 ¼ ctTat .

0

100

200

2.8 3 3.2 3.4 0

(h,T

=0.

5,0,

0)

(h,T

=0.

5,0 ,

0)/L

/

h (h/hN-1)L1/

L=10L=12L=14L=16

L=10L=12L=14L=16

(a) (b)

FIG. 6. Finite-temperature FM-to-PM phase transition atT ¼ 0.5. Panel (a) shows the Ising spin susceptibility at q ¼ 0and ωn ¼ 0, and panel (b) shows the data collapse according toEq. (C1). Note that hN is a free fitting parameter, and the best datacollapse gives hNðT ¼ 0.5Þ ≈ 3.06.

0

2

4

6

8

2.8 3 3.2 3.4 0

F(h,

T=

0.5,

0,0)

F(h,

T=

0.5 ,

0,0)

/L

h (h/hN-1)L1/

L=10L=12L=14L=16

L=10L=12L=14L=16

(a) (b)

FIG. 7. Finite-temperature FM-to-PM phase transition atT ¼ 0.5. Panel (a) shows the fermion spin susceptibility atq ¼ 0 and ωn ¼ 0, and panel (b) shows the data collapseaccording to Eq. (C1). Note that hN is a free fitting parameterand the best data collapse gives hNðT ¼ 0.5Þ ≈ 3.06.

-2

-1

-1 -0.5 0

atlog(T)+log(ct)

log[

χ-1(h

c,T

,0,0

)]

log[T]

L=24L=28

fit

FIG. 8. Inverse Ising spin susceptibility at QCP [χ−1ðhc; T;0; 0Þ] as a function of temperature T. The slope of the log-log plotreveals the power at ¼ 1.48ð4Þ, and the intercept gives rise to theprefactor ct ¼ 0.13ð1Þ.

-2

-1

0

-1.5 -1 -0.5 0

γlog(|h-hc|)+log(ch)

log(

χ-1(h

,T,0

,0)-

χ-1(h

c,T

,0,0

))

log(|h-hc|)

L=24L=28

fit

FIG. 9. Inverse Ising spin susceptibility [χðh; T ¼ 0.1; 0; 0Þ]away from QCP (h > hc) as a function of jh − hcj. Here, thetemperature is T ¼ 0.1, and χ−1ðhc; T ¼ 0.1; 0; 0Þ has beendeducted in the plot. The slope of the log-log plot reveals thepower γ ¼ 1.18ð4Þ, and the intercept gives rise to the prefactorch ¼ 0.7ð1Þ.

NON-FERMI LIQUID AT ð2þ 1ÞD FERROMAGNETIC … PHYS. REV. X 7, 031058 (2017)

031058-9

Page 10: Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical …Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and

The fit gives rise to ct ¼ 0.13ð1Þ and at ¼ 1.48ð4Þ. Wefurther detect the power in transverse field h by fittingχðh; T; 0; 0Þ−1 − χðhc; T; 0; 0Þ−1 ¼ chjh − hcjγ , as shownin Fig. 9. The fit gives rise to ch ¼ 0.7ð1Þ and γ ¼ 1.18ð4Þ.

APPENDIX E: EFFECTIVE ATTRACTIONAND BCS ANALYSIS

In this section, we examine the effective attractioninduced by the Ising-field fluctuations. Away from thecritical point, the Ising spins are gapped, and we cantherefore integrate them out, obtaining an effective four-fermion interaction,

Sint ¼ −ξ2

2

Zdτdτ0

Xijλλ0

σziλðτÞχðh; T; ri − rj; τ − τ0Þσzjλ0 ðτ0Þ:

ðE1Þ

This interaction contains two types of terms: (a) attractions,e.g., nλ↑nλ0↑, and (b) repulsions, nλ↑nλ0↓. Focusing oninstabilities in the particle-particle channel, only attractiveinteractions will be considered. We are therefore restrictedto pairing of fermions of equal spin. Within the BCSapproximation, to gain the most free energy, a nodelessorder parameter is preferable. Thus, to satisfy the Pauliprinciple, the order parameter must transform as a singletunder orbital rotations. This implies that the effectiveattraction favors pairing in the channels Δi;σ ¼ ci1σci2σas well as their linear superpositions. These are the pairingchannels that are most favored by the ferromagneticfluctuations. As shown in Sec. VI of the main text, ourQMC results show that, indeed, pairing correlations inthese channels are enhanced near the FM QCP.

APPENDIX F: POSSIBLESUPERCONDUCTING PHASES

The two-component nature of the superconducting orderparameter Eq. (11) may give rise to a number of exoticphases. Although our numerical data do not show evidenceof such phases, in this section we describe how these phasescould be detected. We restrict our attention to the para-magnetic phase since the two components, Δ↑, Δ↓, arerelated by the Z2 symmetry of the model. DefineΔσ ¼ Δeiθσ , and switch to the charge-spin basis,

θ↑ ¼ θc þ θs;

θ↓ ¼ θc − θs: ðF1Þ

Neglecting amplitude fluctuations, the classical phaseaction is therefore S ¼ Sc þ Ss, with

Sc;s ¼1

2

Zd2rKc;sð∇θc;sÞ2: ðF2Þ

The possible phases (apart from a disordered phase) are asfollows:(1) Charge-4e superconductor: θc is quasi-long-range

ordered, while θs is short ranged. An appropriateorder parameter is Δ↑Δ↓ ∝ ei2θc , whose correlationsgo as hðΔ†

↑Δ†↓ÞðrÞðΔ↓Δ↑Þð0Þi ∝ r−½4=ð2πKcÞ�.

(2) Spin-nematic: θs is quasi-long-range ordered,while θcis short ranged. The order parameter isΔ†

↑Δ↓ ∝ e−i2θs ,

whose correlations go as hðΔ†↑Δ↓ÞðrÞðΔ†

↓Δ↑Þð0Þi ∝r−½4=ð2πKsÞ�.

(3) Triplet superconductor: Both sectors have quasi-long-range order. Here, Δ↑ has power-law correla-tions, hΔ†

↑ðrÞΔ↑ð0Þi ∝ r−f½1=ð2πKcÞ�þ½1=ð2πKsÞ�g.The phase diagram of the same model (in different

physical contexts) has been studied in Refs. [44–46]. Todetermine the phase diagram, we consider the criteria forstability against the appearance of a single vortex. In thismodel, there are three kinds of vortices:(1) A vortex of one spin species and an antivortex in

the other. Across the branch cut, θc → θc, θs →θs þ 2π.

(2) A vortex of both spin species. Across the branch cut,θc → θc þ 2π, θs → θs.

(3) A vortex of one of the spin species. Across thebranch cut, θc → θc þ π, θs → θs þ π.

The phase diagram can be derived from considering the freeenergy of a single unpaired vortex, given by F ¼ E − TS.(As in the usual Berezinski-Kosterlitz-Thouless transition,such an analysis reproduces the phase diagram from a morerigorous renormalization group treatment.) The stabilitycondition is F > 0. The energy of a vortex of type 3 is

E3 ¼1

2ðKc þ KsÞ

Zd2r

�π

2πr

�2

¼ π

4ðKc þ KsÞ log

�La

�;

where L is the system size and a is some short-range cutoff.Similarly, E2 ¼ πKs logðL=aÞ and E1 ¼ πKc logðL=aÞ.The entropy is the same in all cases, TS ¼ logðL2=a2Þ.The resulting phase diagram is given in Fig. 10.The stiffnesses Kc and Ks can be extracted from certain

current-current correlation functions [58],

Kc ¼β

4ðδΛ↑↑ þ δΛ↓↓ þ 2δΛ↑↓Þ;

Ks ¼β

4ðδΛ↑↑ þ δΛ↓↓ − 2δΛ↑↓Þ: ðF3Þ

Here,

δΛσ;σ0 ¼ limL→∞

�Λxxσ;σ0

�qx ¼

L; qy ¼ 0

− Λxxσ;σ0

�qx ¼ 0; qy ¼

L

��; ðF4Þ

XU, SUN, SCHATTNER, BERG, and MENG PHYS. REV. X 7, 031058 (2017)

031058-10

Page 11: Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical …Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and

where

Λxxσ;σ0 ðqx; qyÞ ¼

1

L2

ZdτXijλλ0

eiqðri−rjÞhjxiλσðτÞjxjλ0σ0 ð0Þi;

and jxiλσ ¼ iteiϕλσi;iþx c†iλσciþx;λσ þ H:c: is the current density

for fermions of orbital λ and spin σ.

APPENDIX G: SUPERFLUID DENSITYAND PAIRING CORRELATIONS

In this appendix, we provide further details on thenumerical evidence for superconductivity.To identify the leading pairing channel, we considered

all possible on-site and nearest-neighbor pairing orderparameters and computed the pair structure factor for eachchannel. In all channels other than the orbital-singlet, spin-triplet channel defined in Eq. (11) of the main text, we finda weak response with no substantial system-size depend-ence or enhancement close to the FM QCP (not shown).To test whether there are possible superconductivity

instabilities close to the QCP, we measured the superfluiddensities ρc and ρs, which are related to the stiffnesses Kcand Ks defined in Eq. (F3). Here, ρc;s ¼ Kc;s=β. In Fig. 11,we show the temperature dependence of ρc þ ρs. For bothsets of parameters, ρc þ ρs < ð8=πÞT at the largest systemsize, and it decreases with the system size, implying theabsence of quasi-long-range superconducting order downto T ¼ 0.025 (see Fig. 10). Note that upon increasing thecoupling strength and decreasing temperature, the finite-size estimates for the superfluid density grow. It is likelythat there is a transition to a superconducting phase athigher coupling strengths or lower temperatures.For completeness, in Figs. 12 and 13, we show ρc, ρs,

respectively.

FIG. 10. Phase diagram of the phase action Eq. (F2), illustratingthe SN phase, the charge-4e SC phase, and the triplet super-conducting phase.

-0.04

-0.02

0

0.02

0.04

0 0.1 0.2

ρ c+

ρ s

T

ξ=1.0, μ=-0.5, J=1.0

L=12

L=16

L=20

8T/π

-0.04

-0.02

0

0.02

0.04

0.06

0.08

0.1

0.12

0 0.1 0.2

T

ξ=3.0, μ=-2.0, J=0.5

L=12

L=16

L=20

8T/π

FIG. 11. Temperature dependence of ρc þ ρs at QCP fordifferent system sizes and different couplings.

-0.02

-0.01

0

0.01

0.02

0 0.1 0.2

ρ c

T

ξ=1.0, μ=-0.5, J=1.0

L=12

L=16

L=20

2T/π

-0.02

-0.01

0

0.01

0.02

0.03

0.04

0.05

0.06

0 0.1 0.2

T

ξ=3.0, μ=-2.0, J=0.5

L=12

L=16

L=20

2T/π

FIG. 12. Temperature dependence of ρc at QCP for differentsystem sizes and different couplings.

-0.02

-0.01

0

0.01

0.02

0 0.1 0.2

ρ s

T

ξ=1.0, μ=-0.5, J=1.0

L=12

L=16

L=20

2T/π

-0.02

-0.01

0

0.01

0.02

0.03

0.04

0.05

0.06

0 0.1 0.2

T

ξ=3.0, μ=-2.0, J=0.5

L=12

L=16

L=20

2T/π

FIG. 13. Temperature dependence of ρs at QCP for differentsystem sizes and different couplings.

NON-FERMI LIQUID AT ð2þ 1ÞD FERROMAGNETIC … PHYS. REV. X 7, 031058 (2017)

031058-11

Page 12: Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical …Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and

APPENDIX H: FERMION SPIN SUSCEPTIBILITY

In Fig. 2(a) of the main text, we discussed the anisotropyof the quasiparticle fraction ZkF

ðTÞ along the θ ¼ 0 andθ ¼ ðπ=4Þ directions, and we associate such anisotropyto the soft fermion-bilinear mode around Q ¼ ðπ; πÞ inthe fermion spin susceptibility, besides the dominantQ ¼ ð0; 0Þ ferromagnetic fluctuations. Here, we showthe fermion spin susceptibility [Eq. (C3)] in Fig. 14 andreveal that it is indeed the case.Figure 14 shows the fermion spin susceptibility at

h ¼ hc, T ¼ 0.05 from an L ¼ 24 system. The strongestintensity is naturally at Q ¼ ð0; 0Þ; however, nearQ ¼ ðπ; πÞ, there are soft modes (the signal is very weak,and we have to fix the intensity range to less than 1 and takea logarithm of the original data).

APPENDIX I: SELF-ENERGY

As shown in Fig. 2 in the main text, the imaginary part ofthe fermionic self-energy −ImΣðωnÞ increases in magni-tude as the frequency is lowered. While such behavior isexpected in a superconducting state, the superconductingfluctuations were found to be extremely weak. Here, wespeculate about the possible mechanism for such behavior.Close to the classical Ising transition, the dynamics of

the order parameter are slow. We focus on the classicalfluctuations of the Ising spins, i.e., their static configura-tions. For each such static configuration, one may solve thefermionic part of the Hamiltonian and obtain the single-particle Green’s function. Let us further assume that thefluctuations are extremely sharply peaked at q ¼ 0. Then,in a Monte Carlo simulation, the configurations alternatebetween static, nearly spatially uniform configurations ofIsing spins. Most importantly, the sign of the orderparameter changes between these configurations, or else

the system is, by definition, in the ordered phase. Hence,neglecting all other interaction effects, the fermionicGreen’s function takes the form

Gðk; iωnÞ ≈1

2

�1

iωn − ϵk − Δþ 1

iωn − ϵk þ Δ

�; ðI1Þ

where Δ is the magnitude of the local ferromagnetic orderparameter. If ϵk ≪ Δ, ω, at the nominal Fermi surface(ϵk ¼ 0), we find that

−ImΣðiωnÞ ¼ Im½G−1ðk; iωnÞ� − ωn ¼Δ2

ωn;

which increases rapidly as the frequency is lowered. If Δwere large, compared, e.g., to temperature, it would bepossible to measure it by other means. This would alsoimply jImΣðiω0Þj ≫ ω0 ¼ πT. Small gaps, however, aredifficult to detect.

[1] J. A. Hertz, Quantum Critical Phenomena, Phys. Rev. B 14,1165 (1976).

[2] A. J. Millis, Effect of a Nonzero Temperature on QuantumCritical Points in Itinerant Fermion Systems, Phys. Rev. B48, 7183 (1993).

[3] Toru Moriya, Spin Fluctuations in Itinerant ElectronMagnetism (Springer-Verlag, Berlin, Heidelberg, 1985).

[4] G. R. Stewart, Non-Fermi Liquid Behavior in d- andf-Electron Metals, Rev. Mod. Phys. 73, 797 (2001).

[5] A. V. Chubukov, C. Pépin, and J. Rech, Instability of theQuantum-Critical Point of Itinerant Ferromagnets, Phys.Rev. Lett. 92, 147003 (2004).

[6] H. V. Löhneysen, A. Rosch, M. Vojta, and P. Wölfle,Fermi-Liquid Instabilities at Magnetic Quantum PhaseTransitions, Rev. Mod. Phys. 79, 1015 (2007).

[7] A. V. Chubukov and D. L. Maslov, Spin Conservation andFermi Liquid Near a Ferromagnetic Quantum CriticalPoint, Phys. Rev. Lett. 103, 216401 (2009).

[8] Y. Li, E. H. Lieb, and C. Wu, Exact Results forItinerant Ferromagnetism in Multiorbital Systems onSquare and Cubic Lattices, Phys. Rev. Lett. 112, 217201(2014).

[9] S. Xu, Y. Li, and C. Wu, Sign-Problem-Free QuantumMonte Carlo Study on Thermodynamic Properties andMagnetic Phase Transitions in Orbital-Active ItinerantFerromagnets, Phys. Rev. X 5, 021032 (2015).

[10] J. Custers, P. Gegenwart, H. Wilhelm, K. Neumaier,Y. Tokiwa, O. Trovarelli, C. Geibel, F. Steglich, C. Pepin,and P. Coleman, The Break-up of Heavy Electrons at aQuantum Critical Point, Nature (London) 424, 524(2003).

[11] A. Steppke, R. Küchler, S. Lausberg, E. Lengyel,L. Steinke, R. Borth, T. Lühmann, C. Krellner, M. Nicklas,C. Geibel, F. Steglich, and M. Brando, FerromagneticQuantum Critical Point in the Heavy-Fermion MetalYbNi4ðP1−xAsxÞ2, Science 339, 933 (2013).

-π 0 π 2π 3πqx

0

π

q y

e-1.5

e-1

e-0.5

e0

FIG. 14. The intensity distribution of fermion spin susceptibil-ity χFðh ¼ hc; T ¼ 0.05;q;ωn ¼ 0Þ with L ¼ 24. To amplify theweak signal of the soft mode, we have fixed the intensity range tobe less than unity and taken a logarithm of the original data. Theblack line shows the location of 2kF for the noninteracting FS,only a quarter of which are plotted here.

XU, SUN, SCHATTNER, BERG, and MENG PHYS. REV. X 7, 031058 (2017)

031058-12

Page 13: Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical …Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and

[12] Z. Liu, Y. Gu, W. Zhang, D. Gong, W. Zhang, T. Xie, X. Lu,X. Ma, X. Zhang, R. Zhang et al., Nematic QuantumCritical Fluctuations in BaFe2−xNixAs2, Phys. Rev. Lett.117, 157002 (2016).

[13] W. Zhang, J. T. Park, X. Lu, Y. Wei, X. Ma, L. Hao,P. Dai, Z. Y. Meng, Y.-f. Yang, H. Luo, and S. Li, Effect ofNematic Order on the Low-Energy Spin Fluctuations inDetwinned BaFe1.935Ni0.065As2, Phys. Rev. Lett. 117,227003 (2016).

[14] V. Oganesyan, S. A. Kivelson, and E. Fradkin, QuantumTheory of a Nematic Fermi Fluid, Phys. Rev. B 64, 195109(2001).

[15] W. Metzner, D. Rohe, and S. Andergassen, Soft FermiSurfaces and Breakdown of Fermi-Liquid Behavior, Phys.Rev. Lett. 91, 066402 (2003).

[16] S.-S. Lee, Low-Energy Effective Theory of Fermi SurfaceCoupled with U(1) Gauge Field in 2þ 1 Dimensions, Phys.Rev. B 80, 165102 (2009).

[17] M. A. Metlitski and S. Sachdev, Quantum Phase Transi-tions of Metals in Two Spatial Dimensions. I. Ising-NematicOrder, Phys. Rev. B 82, 075127 (2010).

[18] M. A. Metlitski and S. Sachdev, Quantum Phase Transi-tions of Metals in Two Spatial Dimensions. II. Spin DensityWave Order, Phys. Rev. B 82, 075128 (2010).

[19] D. F. Mross, J. McGreevy, H. Liu, and T. Senthil, ControlledExpansion for Certain Non-Fermi Liquid Metals, Phys. Rev.B 82, 045121 (2010).

[20] M. A. Metlitski, D. F. Mross, S. Sachdev, and T. Senthil,Cooper Pairing in Non-Fermi Liquids, Phys. Rev. B 91,115111 (2015).

[21] T. Holder and W. Metzner, Anomalous Dynamical Scalingfrom Nematic and U(1) Gauge Field Fluctuations in Two-Dimensional Metals, Phys. Rev. B 92, 041112 (2015).

[22] A. Schlief, P. Lunts, and S.-S. Lee, Exact Critical Exponentsfor the Antiferromagnetic Quantum Critical Metal in TwoDimensions, Phys. Rev. X 7, 021010 (2017).

[23] D. Dalidovich and S.-S. Lee, Perturbative Non-FermiLiquids from Dimensional Regularization, Phys. Rev. B88, 245106 (2013).

[24] S.-S. Lee, Recent Developments in Non-Fermi LiquidTheory, arXiv:1703.08172.

[25] A. L. Fitzpatrick, S. Kachru, J. Kaplan, and S. Raghu, Non-Fermi Liquid Behavior of Large-NB Quantum CriticalMetals, Phys. Rev. B 89, 165114 (2014).

[26] T. R. Kirkpatrick and D. Belitz, Nature of the QuantumPhase Transition in Clean Itinerant Heisenberg Ferromag-nets, Phys. Rev. B 67, 024419 (2003).

[27] D. Belitz, T. R. Kirkpatrick, and T. Vojta, How GenericScale Invariance Influences Quantum and Classical PhaseTransitions, Rev. Mod. Phys. 77, 579 (2005).

[28] S. Lederer, Y. Schattner, E. Berg, and S. A. Kivelson,Enhancement of Superconductivity Near a NematicQuantum Critical Point, Phys. Rev. Lett. 114, 097001(2015).

[29] E. Berg, M. A. Metlitski, and S. Sachdev, Sign-Problem-Free Quantum Monte Carlo of the Onset of Antiferromag-netism in Metals, Science 338, 1606 (2012).

[30] Y. Schattner, S. Lederer, S. A. Kivelson, and E. Berg, IsingNematic Quantum Critical Point in a Metal: A Monte CarloStudy, Phys. Rev. X 6, 031028 (2016).

[31] Y. Schattner, M. H. Gerlach, S. Trebst, and E. Berg,Competing Orders in a Nearly Antiferromagnetic Metal,Phys. Rev. Lett. 117, 097002 (2016).

[32] Z.-X. Li, F. Wang, H. Yao, and D.-H. Lee, The Nature ofEffective Interaction in Cuprate Superconductors: A Sign-Problem-Free Quantum Monte-Carlo Study, Phys. Rev. B95, 214505 (2017).

[33] Z.-X. Li, F. Wang, H. Yao, and D.-H. Lee, What Makes theTc of Monolayer FeSe on SrTiO3 So High: A Sign-Problem-Free Quantum Monte Carlo Study, Science bulletin 61, 925(2016).

[34] X. Y. Xu, K. S. D. Beach, K. Sun, F. F. Assaad, and Z. Y.Meng, Topological Phase Transitions with SO(4) Symmetryin ð2þ 1ÞD Interacting Dirac Fermions, Phys. Rev. B 95,085110 (2017).

[35] S. Lederer, Y. Schattner, E. Berg, and S. A. Kivelson,Superconductivity and Non-Fermi Liquid Behavior Neara Nematic Quantum Critical Point, Proc. Natl. Acad. Sci.U.S.A. 114, 4905 (2017).

[36] F. F. Assaad and T. Grover, Simple Fermionic Model ofDeconfined Phases and Phase Transitions, Phys. Rev. X 6,041049 (2016).

[37] S. Gazit, M. Randeria, and A. Vishwanath, Emergent DiracFermions and Broken Symmetries in Confined and Decon-fined Phases of Z2 Gauge Theories, Nat. Phys. 13, 484,advance online publication (2017).

[38] P. Pfeuty and R. J. Elliott, The Ising Model with a Trans-verse Field. II. Ground State Properties, J. Phys. C 4, 2370(1971).

[39] C. Wu and S.-C. Zhang, Sufficient Condition for Absence ofthe Sign Problem in the Fermionic Quantum Monte CarloAlgorithm, Phys. Rev. B 71, 155115 (2005).

[40] G. G. Batrouni and R. T. Scalettar, Quantum Phase Tran-sitions, in Ultracold Gases and Quantum InformationLecture Notes of the Les Houches Summer Schoolin Singapore (Oxford University Press, Oxford, 2011),Vol. 91.

[41] N. Trivedi and M. Randeria, Deviations from Fermi-LiquidBehavior Above Tc in 2D Short Coherence Length Super-conductors, Phys. Rev. Lett. 75, 312 (1995).

[42] T. Senthil, Critical Fermi Surfaces and Non-Fermi LiquidMetals, Phys. Rev. B 78, 035103 (2008).

[43] K.-S. Chen, Z. Y. Meng, T. Pruschke, J. Moreno, and M.Jarrell, Lifshitz Transition in the Two-Dimensional HubbardModel, Phys. Rev. B 86, 165136 (2012).

[44] F. Krüger and S. Scheidl, Nonuniversal Ordering of Spinand Charge in Stripe Phases, Phys. Rev. Lett. 89, 095701(2002).

[45] D. Podolsky, S. Chandrasekharan, and A. Vishwanath,Phase Transitions of s ¼ 1 Spinor Condensates in anOptical Lattice, Phys. Rev. B 80, 214513 (2009).

[46] E. Berg, E. Fradkin, and S. A. Kivelson, Charge-4eSuperconductivity from Pair-Density-Wave Order inCertain High-Temperature Superconductors, Nat. Phys.5, 830 (2009).

[47] Y.-F. Jiang, Z.-X. Li, S. A. Kivelson, and H. Yao, Charge-4eSuperconductors: A Majorana Quantum Monte CarloStudy, Phys. Rev. B 95, 241103 (2017).

[48] E. Morosan, H. W. Zandbergen, Lu Li, M. Lee, J. G.Checkelsky, M. Heinrich, T. Siegrist, N. P. Ong, and

NON-FERMI LIQUID AT ð2þ 1ÞD FERROMAGNETIC … PHYS. REV. X 7, 031058 (2017)

031058-13

Page 14: Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical …Non-Fermi Liquid at (2+1)D Ferromagnetic Quantum Critical Point Xiao Yan Xu,1 Kai Sun,2 Yoni Schattner,3 Erez Berg,3 and

R. J. Cava, Sharp Switching of the Magnetization inFe1=4TaS2, Phys. Rev. B 75, 104401 (2007).

[49] Y. Q. Liu, S. J. Zhang, J. L. Lv, S. K. Su, T. Dong, G. Chen,and N. L. Wang, Revealing a Triangular Lattice IsingAntiferromagnet in a Single-Crystal CeCd3As3, arXiv:1612.03720.

[50] www.gauss-centre.eu[51] www.lrz.de[52] U. Wolff, Collective Monte Carlo Updating for Spin

Systems, Phys. Rev. Lett. 62, 361 (1989).[53] J. Liu, Y. Qi, Z. Y. Meng, and L. Fu, Self-Learning

Monte Carlo Method, Phys. Rev. B 95, 041101 (2017).[54] J. Liu, H. Shen, Y. Qi, Z. Y. Meng, and L. Fu, Self-Learning

Monte Carlo in Fermion Systems, Phys. Rev. B 95, 241104(2017).

[55] X. Y. Xu, Y. Qi, J. Liu, L. Fu, and Z. Y. Meng, Self-LearningDeterminantal QuantumMonte Carlo Method, Phys. Rev. B96, 041119 (2017).

[56] F. F. Assaad and H. G. Evertz, World-Line and Determi-nantal Quantum Monte Carlo Methods for Spins, Phononsand Electrons, in Computational Many-Particle Physics,Lecture Notes in Physics, Vol. 739, edited by H. Fehske, R.Schneider, and A. Weiße (Springer, Berlin, Heidelberg,2008), pp. 277–356.

[57] M. H. Gerlach, Y. Schattner, E. Berg, and S. Trebst,Quantum Critical Properties of a Metallic Spin-Density-Wave Transition, Phys. Rev. B 95, 035124 (2017).

[58] D. J. Scalapino, S. R. White, and S. Zhang, Insulator, Metal,or Superconductor: The Criteria, Phys. Rev. B 47, 7995(1993).

XU, SUN, SCHATTNER, BERG, and MENG PHYS. REV. X 7, 031058 (2017)

031058-14