+ All Categories
Home > Documents > arXiv:1602.07055v1 [cond-mat.supr-con] 23 Feb 2016

arXiv:1602.07055v1 [cond-mat.supr-con] 23 Feb 2016

Date post: 21-Feb-2022
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
7
Higgs amplitude mode in massless Dirac fermion systems Ming Lu, 1, 2 Haiwen Liu, 1, 2 Pei Wang, 1, 3 and X. C. Xie 1, 2 1 International Center for Quantum Materials and School of Physics, Peking University, Beijing 100871, China 2 Collaborative Innovation Center of Quantum Matter Beijing 100871, China 3 Department of Physics, Zhejiang Normal University, Jinhua 321004, China (Dated: September 17, 2018) The Higgs amplitude mode in superconductors is the condensed matter analogy of Higgs bosons in particle physics. We investigate the time evolution of Higgs amplitude mode in massless Dirac systems, induced by a weak quench of an attractive interaction. We find that the Higgs amplitude mode in the half-filling honeycomb lattice has a logarithmic decaying behaviour, qualitatively dif- ferent from the 1/ t decay in the normal superconductors. Our study is also extended to the doped cases in honeycomb lattice. As for the 3D Dirac semimetal at half filling, we obtain an undamped oscillation of the amplitude mode. Our finding is not only an important supplement to the previ- ous theoretical studies on normal fermion systems, but also provide an experimental signature to characterize the superconductivity in 2D or 3D Dirac systems. PACS numbers: 74.40.Gh, 74.20.Fg, 74.78.-w, 67.85.-d I. INTRODUCTION A conventional superconductor can be described by a charged complex order parameter Δ(r, t) = |Δ(r, t)|e (x,t) . Its collective fluctuations around equi- librium including the oscillations of the phase and amplitude 1 . The phase mode, being coupled to the elec- tromagnetic field, moves to plasma frequency of the metal as a manifestation of Anderson-Higgs mechanism 2–4 . The amplitude mode oscillates with the angular fre- quency 2|Δ 0 |, analogous to the “vibration” of the lon- gitudinal component of Higgs field in particle physics 5 . In this sense, the amplitude mode in superconductor is sometimes also called Higgs mode or Higgs amplitude mode in the literature 1,5,7–9,13 . Higgs amplitude mode in superconductors, although theoretically predicted many years ago 6 , has only been directly observed recently by the time-resolved Teraherz (THz) pump-probe technique in a clean superconduct- ing film 7,8 , and by measuring the excess sub-gap optical conductance in disordered films near the superconductor- insulator phase transition 9 . The time evolution of the Higgs mode in the collisionless, dissipationless regime was studied intensely. It was revealed that the Higgs mode oscillates at a frequency of 2Δ with a 1/ t de- caying property in the weak coupling limit, where Δ is the asymptotic value of superconducting gap 10–13 . How- ever, previous works all assume that the density of states (DOS) near the Fermi level is almost a constant within the Debye cut-off energy ω D . This assumption obviously fails for honeycomb lattice or Dirac semimetals at half filling. Their DOS is either linear (2D) or quadratic (3D) at low energy, respectively, and vanishes at the Dirac point 15,20 . Since superconductivity is strongly affected by the DOS near the Fermi level, it would be theoret- ically interesting to study the time evolution of Higgs mode in those systems. On the experimental side, the availability of the honeycomb optical lattice 18 and the tunable attractive interaction by Feshbash resonance 19 give a possible test ground for this study. Besides, the expected unique feature of the Higgs mode in supercon- ducting Dirac semimetal can be used as an important experimental characterization to distinguish it from the normal superconductors 21,22 . In this paper we study the quenched dynamics in the weak coupling limit by using the Anderson pseudo-spin formalism 23 . We find that the Higgs mode has a log- decay behaviour in the half-filling honeycomb lattice. To understand this behaviour, we further study the pseudo- spins’ phase dynamics, and analytically solve the lin- earized equations of motion 11–13 . The doped cases is also studied numerically. In the low doping limit, a double- frequency feature is found. The larger frequency in- creases noticeably and its peak broadens with the doping level. In the high doping limit, we are back to the 1/ t decaying property, as in a normal superconductor. When considering the 3D Dirac semimetal at neutral point, we find that the Higgs mode exhibits an undamped oscilla- tion, with all the pseudo-spins precess synchronizely. II. MODEL AND FORMALISM We start by considering the negative-U Hubbard model on honeycomb lattice: ˆ H = - X <ij>,σ ˆ a b + h.c. - U X i ˆ n iˆ n i- μ X ˆ n (1) where ˆ a i ( ˆ b i ) is the on-site annihilation operator on sub- lattice A (B); ˆ n is the number operator on lattice site i with spin index σ; μ is chemical potential and U is the the on-site attractive interaction. We choose the nearest- neighbour hopping as the energy unit throughout this paper. To study the dynamics, we write out the corresponding mean-field Hamiltonian in k-space after a unitary trans- formation: ˆ a kσ = 1 2 (e k ˆ c kσ + ˆ d kσ ), ˆ b kσ = 1 2 (-ˆ c kσ + arXiv:1602.07055v1 [cond-mat.supr-con] 23 Feb 2016
Transcript

Higgs amplitude mode in massless Dirac fermion systems

Ming Lu,1, 2 Haiwen Liu,1, 2 Pei Wang,1, 3 and X. C. Xie1, 2

1International Center for Quantum Materials and School of Physics, Peking University, Beijing 100871, China2Collaborative Innovation Center of Quantum Matter Beijing 100871, China3Department of Physics, Zhejiang Normal University, Jinhua 321004, China

(Dated: September 17, 2018)

The Higgs amplitude mode in superconductors is the condensed matter analogy of Higgs bosonsin particle physics. We investigate the time evolution of Higgs amplitude mode in massless Diracsystems, induced by a weak quench of an attractive interaction. We find that the Higgs amplitudemode in the half-filling honeycomb lattice has a logarithmic decaying behaviour, qualitatively dif-ferent from the 1/

√t decay in the normal superconductors. Our study is also extended to the doped

cases in honeycomb lattice. As for the 3D Dirac semimetal at half filling, we obtain an undampedoscillation of the amplitude mode. Our finding is not only an important supplement to the previ-ous theoretical studies on normal fermion systems, but also provide an experimental signature tocharacterize the superconductivity in 2D or 3D Dirac systems.

PACS numbers: 74.40.Gh, 74.20.Fg, 74.78.-w, 67.85.-d

I. INTRODUCTION

A conventional superconductor can be describedby a charged complex order parameter ∆(r, t) =|∆(r, t)|eiφ(x,t). Its collective fluctuations around equi-librium including the oscillations of the phase andamplitude1. The phase mode, being coupled to the elec-tromagnetic field, moves to plasma frequency of the metalas a manifestation of Anderson-Higgs mechanism2–4.The amplitude mode oscillates with the angular fre-quency 2|∆0|, analogous to the “vibration” of the lon-gitudinal component of Higgs field in particle physics5.In this sense, the amplitude mode in superconductor issometimes also called Higgs mode or Higgs amplitudemode in the literature1,5,7–9,13.

Higgs amplitude mode in superconductors, althoughtheoretically predicted many years ago6, has only beendirectly observed recently by the time-resolved Teraherz(THz) pump-probe technique in a clean superconduct-ing film7,8, and by measuring the excess sub-gap opticalconductance in disordered films near the superconductor-insulator phase transition9. The time evolution of theHiggs mode in the collisionless, dissipationless regimewas studied intensely. It was revealed that the Higgsmode oscillates at a frequency of 2∆∞ with a 1/

√t de-

caying property in the weak coupling limit, where ∆∞ isthe asymptotic value of superconducting gap10–13. How-ever, previous works all assume that the density of states(DOS) near the Fermi level is almost a constant withinthe Debye cut-off energy ωD. This assumption obviouslyfails for honeycomb lattice or Dirac semimetals at halffilling. Their DOS is either linear (2D) or quadratic (3D)at low energy, respectively, and vanishes at the Diracpoint15,20. Since superconductivity is strongly affectedby the DOS near the Fermi level, it would be theoret-ically interesting to study the time evolution of Higgsmode in those systems. On the experimental side, theavailability of the honeycomb optical lattice18 and thetunable attractive interaction by Feshbash resonance19

give a possible test ground for this study. Besides, theexpected unique feature of the Higgs mode in supercon-ducting Dirac semimetal can be used as an importantexperimental characterization to distinguish it from thenormal superconductors21,22.

In this paper we study the quenched dynamics in theweak coupling limit by using the Anderson pseudo-spinformalism23. We find that the Higgs mode has a log-decay behaviour in the half-filling honeycomb lattice. Tounderstand this behaviour, we further study the pseudo-spins’ phase dynamics, and analytically solve the lin-earized equations of motion11–13. The doped cases is alsostudied numerically. In the low doping limit, a double-frequency feature is found. The larger frequency in-creases noticeably and its peak broadens with the dopinglevel. In the high doping limit, we are back to the 1/

√t

decaying property, as in a normal superconductor. Whenconsidering the 3D Dirac semimetal at neutral point, wefind that the Higgs mode exhibits an undamped oscilla-tion, with all the pseudo-spins precess synchronizely.

II. MODEL AND FORMALISM

We start by considering the negative-U Hubbard modelon honeycomb lattice:

H = −∑

<ij>,σ

a†iσbjσ+h.c.−U∑i

ni↑ni↓−µ∑iσ

niσ (1)

where ai (bi) is the on-site annihilation operator on sub-lattice A (B); niσ is the number operator on lattice sitei with spin index σ; µ is chemical potential and U is thethe on-site attractive interaction. We choose the nearest-neighbour hopping as the energy unit throughout thispaper.

To study the dynamics, we write out the correspondingmean-field Hamiltonian in k-space after a unitary trans-

formation: akσ = 1√2(eiθk ckσ + dkσ), bkσ = 1√

2(−ckσ +

arX

iv:1

602.

0705

5v1

[co

nd-m

at.s

upr-

con]

23

Feb

2016

2

e−iθk dkσ):

HMF =−∑k

(µ− |γk|)c†kσckσ −∑k

(µ+ |γk|)d†kσdkσ

−∆∗(t)∑k

(c†k↑c

†−k↓ + d†k↑d

†−k↓

)+ h.c. (2)

where akσ (bkσ) is the Fourier component of ai (bi);eiθk = γk/|γk| with γk =

∑k e

ik·δ and δ being thethree real space nearest-neighbour vectors; the time de-

pendent order parameter ∆(t) = UNc

∑k

⟨a†k↑a

†−k↓

⟩=

UNc

∑k

⟨b†k↑b

†−k↓

⟩, in which Nc is the number of unit

cells and 〈· · · 〉 denotes the time dependent quantum-mechanical expectation value.

We define two set of Anderson pseudo-spins: S(+)k =

12

(c†k↑, c−k↓

)σ( ck↑c†−k↓

), S

(−)k = 1

2

(d†k↑, d−k↓

)σ( dk↑d†−k↓

),

with their corresponding local fields b(±)k (t) =(

∆R(t), ∆I(t), µ∓ |γk|). It is straightforward to check

that the pseudo-spin operators satisfies the commutationrelationship of the angular momentum (with ~ = 1). Us-ing the above definition, the Hamiltonian can be writtenas the sum of the “Zeeman energy” of pseudo-spins intheir corresponding local fields:

HMF = −2∑

k,i=±

b(i)k · S

(i)k (3)

From the Hamiltonian, we can get the equations of mo-

tion of pseudo-spins: ∂∂tS

(i)k (t) = −2b

(i)k × S

(i)k (t), where

i = ± and S(i)k (t) ≡

⟨S

(i)k

⟩are the expectation value

of Anderson pseudo-spin operators. The time depen-dent gap can be written using pseudo-spins as: ∆(t) =U

2Nc

∑k,i=±

(S

(i)xk + iS

(i)yk

).

For simplicity, we can also label the pseudo-spins byenergy state εj rather than k, so that we can combinethe two sets of pseudo-spins as a single set. Explicitly,the equations of motion and time dependent gap can berewritten as:

∂tSj(t) = −2bj(t)× Sj(t) (4)

∆(t) =U

2Nc

∑j

(Sxj (t) + iSyj (t)

)(5)

with:

bj(t) = (∆R(t),∆I(t), εj) (6)

where εj ∈ (−ωD, ωD), and Sj can be view as the classi-cal spin with length 1

2 . Writing like this, the additionalDOS information is needed. It satisfies D(ε) ∝ |ε − µ|,for we have a 2D linear dispersion near the Dirac pointbefore superconducting, see [FIG.(1(c))].

kx

ky

εk

O

kx

ky

εk

O

kx

ky

εk

O

(c) (d) (e)

µ

µ

µ

peudo-spinlocal field

x

z

O

x

z(a) (b)

εj εj

FIG. 1. (Color online) Quenched dynamics illustration andthree doping cases for honeycomb lattice. (a) When t ≤ 0,the system is in the BCS ground state, the pseudo-spins alignin the direction of their local fields. (b) At t = 0+, we changethe interaction strength abruptly to make the system out ofequilibrium. The pseudo-spins start to precess around theirlocal fields, while the local fields also change due to theirdependence on pseudo-spins. (c) The half filling case: µ = 0,where εk ≡ ±|γk|. (d) The high doping limit: µ� ∆0f . (e)The low doping limit: µ ∼ ∆0f .

The quenched dynamics is as follows: at t ≤ 0, the sys-tem is in equilibrium with the initial interacting strengthUi. From the spin Hamiltonian, the initial spins are par-allel to their local fields [Fig. 1(a)]. At t = 0+, wechange the interaction strength to Uf , then the localfields change immediately for the sudden change of ∆(t).Therefore, the current spin configuration is no longer sta-ble. According to equation (4), they will precess aroundtheir local fields[Fig.1(b)], which in turn will change thegap and the local fields simultaneously by equation (5)and (6) . We denote ∆0i and ∆0f as the correspond-ing equilibrium gap when the interaction strength are Uiand Uf , respectively. In the following, they are used todescribe the quenched dynamics for convenience.

III. THREE DOPING CASES FORHONEYCOMB LATTICE

We consider the dynamics of three doping cases forhoneycomb lattice as shown in Fig. 1(c, d, e): half filling,high doping limit and low doping limit.

A. Half filling

Without loss of generality, we choose the initial gap∆0i to be real. The particle-hole symmetry guaran-tees the gap to be real throughout the evolution10.The problem is to solve a system of coupled differen-tial equations (4) with the initial condition: Sj(0) =

3(∆0i

2√

∆20i+ε

2j

, 0,εj

2√

∆20i+ε

2j

), where in this case the gap

and local fields are related with the pseudo-spins as:

∆(t) =Uf

2Nc

∑j S

xj (t) and bj(t) = (∆(t), 0, εj). The DOS

in the half filling case is proportional to |ε|.We numerically simulate equation (4) with N = 50000

energy levels and the Debye cut-off energy ωD = 0.5.The method we use is the Runge-Kutta of the 8-th orderwith an adjustable time step to meet a sufficient highprecision. Other numbers of energy levels are also triedto verify that the results are unaffected by the finite sizeeffect. We also adopt the weak coupling limit(∆0f � ωD)and the weak quench limit(δ∆0 ≡ ∆0i−∆0f � ∆0f ). Tosatisfy this, we quench from ∆0i = 0.013 to ∆0f = 0.012.The result is shown in FIG.(2): the data is well fitted bya log-decay function:

∆(t)

∆0f= a+

2bδ∆0

∆0f

cos(c∆0f t+ d)

ln(e∆0f t)(7)

The envelope functions a±2bδ∆0/∆0f ln(e∆0f t) are usedfor indicating the log-decay behaviour.

The fitted parameter are: a = 0.9975, b = 1.091, c =1.994, d = 0.2554, e = 22.36. We find that c = 2a isalmost exactly satisfied, which means that ∆(t) oscil-lates with the 2∆∞ angular frequency, indicating it isthe Higgs amplitude mode. However, the mode has a log-arithmic decaying property in the present case, while itdecays as 1/

√t in the normal superconductors. This slow

decaying behaviour suggests the Higgs mode in the half-filling superconducting honeycomb lattice has a muchlonger lifetime than that in the usual superconductors24.We also note that a is slightly smaller than 1, meaning∆∞ < ∆0f . Explicitly, we find 1 − a ≈ δ∆2

0/3∆20f . The

similar behaviour has been pointed out in the previousliterature for the normal superconductors, claiming thatthe difference is of order δ∆2

0/6∆20f

12,25.

0 20 40 60 80 100 120 ∆0ft0.92

0.96

1.00

∆(t

)/∆

0f

numerical datalog decay fitenvelope curve

FIG. 2. (color online) Half filling. The numerical data (blue)obtained from simulating N = 50000 energy levels for ∆0i =0.013 and ∆0f = 0.012, with Deybe energy ωD = 0.5. Thered curve is the fit by equation (7), while the green dottedline are the envelope curves.

The slower decaying property compared with nor-mal superconductors can be qualitatively understood by

studying the phase dynamics of the single pseudo-spinson different energy levels10. Explicitly, we numericallycalculate the precession angle φj(t) of pseudo-spin Sjaround the time independent vector b∞j ≡ (∆∞, 0, εj).As shown in FIG.3(a), in the long time limit, the phasebecome linear with respect to time so that we can char-acter the precession frequency by the time averaged fre-quency ωj = 〈ωj(t)〉 = [φj(tmax) − φj(0)]/tmax. InFIG.3(b), we compare ωj for constant and linear DOS.For constant DOS, ωj is equal to the quasi-particle spec-

trum 2√

∆2∞ + ε2j . In the region when εj . 2∆0f , ωj

for both cases coincide with each other. However in thehigher energy region, ωj for linear DOS is much flatterthan that for constant DOS. The decaying of the ampli-tude is due to the dephasing mechanism for the precessionof pseudo-spins. The flatter dispersion of ωj representsa more synchronized precession of the pseudo-spins, re-sulting in a slower decaying of the amplitude.

0 100 200 ∆0ft0

100

200φj/2π

(a)εj =2∆0f

εj =4∆0f

0 2 4 6 εj/∆0f048

12

ωj/∆

0f

(b)const DOSlinear DOS2√

∆2∞ +ε 2

j

FIG. 3. (color online) Phase dynamics for ∆0i = 0.013 and∆0f = 0.012 at half filling. The solid lines are for D(ε) = 1,while the dashed lines are for D(ε) ∝ |ε|. (a). The precessionphases φj for εj = 2∆0f , 4∆0f . They are almost linear forthe large time dynamics and φj for constant DOS has larger“phase slope”. (b). The time averaged precession frequencyωj . For constant DOS, ωj coincides with quasiparticle energyspectrum. For linear DOS case, the flatter ωj ’s dispersiongives rise to in a weaker dephasing, therefore a slower decayof the amplitude.

To quantitatively understand the fitting equation (7),we solve equations of motion (4) by linearizing it around

Sfj ≡(

∆0f

2√

∆20f+ε2j

, 0,εj

2√

∆20f+ε2j

)and bfj ≡ (∆0f , 0, εj):

∂tδSxj (t) = 2εjδS

yj (t)

∂tδSyj (t) =

εj√∆2

0f + ε2j

δ∆(t) + 2∆0fδSzj (t)− 2εjδS

xj (t)

∂tδSzj (t) = −2∆0fδS

yj (t) (8)

where δ∆(t) ≡ ∆(t) − ∆0f and δSj(t) ≡ Sj(t) − Sfj .The above coupled differential equation can be solved byLaplace transform: L [f(t)] → f(s). In the thermody-namic and the weak coupling limit, we arrive at the final

4

form of δ∆(s):

δ∆(s) =δ∆0

2∆0f

1(s

2∆0f

) − 1[(s

2∆0f

)2

+ 1

]tan−1

(s

2∆0f

)

(9)By inverse Laplace transform, we can get the approxi-

mate form of ∆(t) (see Appendix A):

∆(t) ≈ ∆f + 2δ∆0cos 2∆f t

ln 4∆f t(10)

B. Doping cases

In the high doping limit (µ � ∆0f ) as illustrated inFig. 1(d). The system without attractive interaction isbasically a normal metal, therefore we expect the Higgsmode will have the square-root decaying behaviour. Toverify this, we choose µ = 0.12 = 10∆0f and simulateequation (4)-(6) with other parameters equal to those inthe half-filling case. The result is shown in AppendixB. We can see |∆(t)| indeed decays as 1/

√t, with the

oscillation frequency equals to 2∆∞.To see how the mode change from the logarithmic de-

cay to the 1/√t decay, we investigate the low doping limit

where µ ∼ ∆0f [Fig. 1(e)]. By simulating equation (4)-(6) with several different values of µ, we find there aretwo frequencies in the low doping case: one is the Higgsfrequency 2∆∞, the other is slightly larger than the firstone, resulting in a beat pattern as shown in FIG.4 (a).As µ increases, we find both frequencies increase. How-ever, the Higgs frequency increases only slightly, whilethe lager frequency increases more remarkably and thepeak broadens[Fig.4 (b)]. Physically, the decay of theHiggs mode is due to its interaction with the bottom partof the particle-hole continuum11,14. As we doped awayfrom half filling, those states most responsible for thedamping increase, resulting a faster decaying behaviour.When µ is large enough (about 2∆0f ), the second peakcan hardly be discerned and the transform from the log-arithmic decay to square-root decay accomplishes. Wealso find a very interesting empirical formula, associatingthe difference of the two frequencies δω with the chemical

potential µ as: δω∆0f

= 2(

µ∆0i

)2

.

IV. DIRAC SEMIMETAL CASE

We extend our calculation to the 3D Dirac semimetalcase. The DOS is proportional to ε2 when the Fermilevel is on the Dirac point. We numerically solve thecollective motion of pseudo-spins with all the parametersequal to those in the half filling honeycomb lattice case.We find the Higgs amplitude mode in this case exhibits

300 400 500 600 ∆0ft

0.996

1.000

1.004

|∆(t

)|/∆

0f

(a)µ=2∆0i/13

200400600

a.u.

(b)µ=2∆0i/13

1.8 2.0 2.2 ω/∆0f

200400600

a.u. µ=3∆0i/13

0 0.1 0.2 µ/∆0i0

0.1

δω/∆

0f

(c)data pointsδω∆0f

=2( µ

∆0i

)2

FIG. 4. (color online) Low doping case. The quench parame-ters are the same as in FIG.2. (a) The two slightly differentfrequencies give rise to a beat pattern of the amplitude mode.(b) The frequencies obtained by discrete fourier transfrom(DFT) of |∆(t)|. Both frequencies increases as µ increases,while the larger one increase more noticeably. Besides, thelarger frequency peak also broadens and will eventually dis-appear as µ increases, accomplishing the gradual transformfrom logarithimic decay to square root decay. (c) The fre-quencies data (red dots) collect by DFT of different values ofµ, they fit quite well by the empirical formula (blue line).

an undamped oscillation as shown in FIG.5(a). To ex-plain this, we study the phase dynamics φj(t) of eachpseudo-spin Sj(t) that precess around its own time inde-pendent vector b∞j . From FIG.5 (b, c), we can see thatall the pseudo-spins precess with the same angular fre-quency 2∆∞. Therefore, for the two instances of timeseparated by T = π/∆∞, the whole pseudo-spins’ con-figuration is identical. Since ∆(t) depends explicitly onthe sum of x component of all the pseudo-spins, it mustbe periodic and undamped. Compared with 2D case athalf filling, the particle-hole continuum most responsiblefor the damping consist a even smaller fraction of thewhole phase space. Therefore, the damping originatingfrom the interaction with those states is negligible. Wenote that the above discussion is for the singlet pairingcase. However, the triplet pairing is also possible, whichhas three independent Higgs mode17. Studying the timeevolution of these Higgs mode would also be interesting.

V. DISCUSSION AND SUMMARY

For the 2D superconducting Dirac fermion case, thequenched process can be realized on the two-componentcold Fermi gases trapped in a honeycomb opticallattice18, with an attractive Hubbard U tunable by theFeshbach resonance19. The Higgs mode in this case canbe detected with the rf-absorbtion techniques26,27. As forthe Higgs amplitude mode in 3D case, the observation ismade possible by the recent discovery of superconductiv-

5

0 20 40 60 80 100 120 ∆0ft0.90

0.95

∆(t

)/∆

0f

(a)

0 10 20 30 ∆0ft0

4

8

φj/2π

(b)εj =2∆0fεj =4∆0f

0 2 4 6 εj/∆0f

1.9

2.0ωj/∆

0f

(c)

FIG. 5. (color online) 3D Dirac semimetal case. The quenchparameters are equal to those in FIG.2. (a). The Higgsmode shows an undamped oscillation. (b). Precession phaseof single spin on energy levels εj = 2∆0f , 4∆0f . (c). Theprecession of different pseudo-spins synchronize.

ity in Dirac semimetals21,22, together with the develop-ment of the ultrafast THz pump-probe spectroscopy28.In principle, the measurement should be similar to thealready discovered Higgs mode in the clean NbN film7.One can use an intense monocycle THz pump pulse togenerate the Higgs amplitude mode in the superconduct-ing Cd3As2 thin film. Immediately after that, a probepulse also irradiates to the sample. By measuring thepump-probe delay time and the wave form of the trans-mitted probe pulse, one can resolve the time evolution ofthe Higgs mode inside the sample7,8.

In summary, we find the Higgs amplitude mode in half-filling honeycomb lattice has a logarithmic decaying be-haviour. It can be understood by studying its phase dy-namics, and by analytically solving the linearized equa-tions of motion. The dynamics of doped cases in honey-comb lattice is also studied. As for the three dimensionalDirac semimetals case, we find the Higgs mode exhibitsan undamped oscillation when the Fermi level is at theDirac point.

VI. ACKNOWLEDGEMENT

This work was financially supported by NBRP ofChina (2012CB821402 and 2015CB921102) and NSF-China under Grants Nos.11534001, 11504008 and11304280.

Appendix A: INVERSE LAPLACE TRANSFORMOF EQ.9

By doing the Laplace transform of the linearized equa-tions of motion, we get the following equation for δ∆(s)up to the linear order of δ∆0:

δ∆(s)∑j

1(s2 + 4∆2

0f + 4ε2j

)(∆2

0f + ε2j

) 12

=sδ∆0

s2 + 4∆20f

∑j

ε2j(s2 + 4∆2

0f + 4ε2j

)(∆2

0f + ε2j

) 32

(A1)

In the thermodynamic limit and weak coupling limit,we have

∑j f(εj) ∝

∫ ωD

0f(ε)εdε ≈

∫∞0f(ε)εdε. After

the integration, we get equation (9) in the main text.Using the similarity theorem L−1

[ ¯f(s/a)]

= af(at),we need only to find the the inverse Laplace transform off(s) = 1/(s2 + 1) tan−1 s. We achieve this by evaluatingthe Bromwich integral:

f(t) =1

2πi

∫ γ+i∞

γ−i∞dsf(t)est (A2)

where γ should be larger than the real part of any polesin the integrand.

O Re s

Im s

C0

Γ2

Γ1

Γ3

C1

C2

C3

C4

γ1

γ2

γ − i∞

γ + i∞

R

FIG. 6. (color online) The countour of the integral. The redcross represents the pole at s = 0, the red points are branchpoints at s = ±i, the red lines are the two branch cuts.

We choose the contour shown in Fig. 6, and useCauchy’s integral theorem to evaluate the Bromwich in-tegral C0 marking in blue. The Jordan’s lemma tells usthe contributions from big arcs Γ1,Γ2,Γ3 are zero, and itis easy to verify that the integrals along the small arcs γ1

and γ2 have no contributions either. The only remainingparts are the pole at origin and line integrals C1 to C4.So we have:

f(t) = θ(t)− 4I2(t) (A3)

I2(t) = <

eit ∫ ∞0

eixt

(x2 + 2x)

[(ln x

x+2

)2

+ π2

]dx

(A4)

We use the contour in Fig. 7 to evaluate equation(A4),and the only remaining contribution is from the line in-

6

tegral γ1. To the leading order, we have:

I2(t) = <[eit∫ 2a

0

e−2yt

2y(ln y)2dy

](A5)

For large enough t, the above integral can be conductedby using a result by A. Erdlyi29, thus we obtain equa-tion(10) in the main text.

O Re z

Im z

ia

R

R + ia

γ1

γ2

γ3

FIG. 7. (color online)The contour for I2(t), a is a small realpositive number, the integral I2(t) (blue) is replaced by thecontour in red, while integration along γ2 and γ3 are zero.

Appendix B: High doping limit case

We choose µ = 0.12 in this case. Because the ex-act particle-hole symmetry is absent when µ 6= 0, ∆(t)will acquire a time-depended phase during the evolution,thus we plot the amplitude |∆(t)| in the figure. We fit

the data using the following equation provided in manyliteratures11–13:

|∆(t)|∆0f

= a+2bδ∆0

π32 ∆0f

√∆0f t

cos(c∆0f t+ d

π

4

)(B1)

The fitting parameters are: a = 1.0050, b = 0.5142, c =2.0101, d = 0.9827. We see c = 2a is almost exactly sat-isfied, indicating this is the Higgs amplitude mode. How-ever, a is slightly greater than 1, meaning ∆∞ is slightlygreater than ∆0f . This is not so surprising because therelation ∆∞ ≈ ∆0f − δ∆2

0/6∆0f is obtained under thestrictly constant density of state condition. In conclu-sion, in the high doping limit, the system behaves as anormal metal without interaction, resulting the 1/

√t de-

caying property of the amplitude |∆(t)|.

0 20 40 60 80 100 120 ∆0ft0.98

0.99

1.00

|∆(t

)|/∆

0f

numerical data1/√t decay fit

envelope curve

FIG. 8. (color online) High doping limit with µ = 0.12, otherparameters are same as those in FIG. 2 in the main text. Thenumerical data (blue) is well fitted by equation (B1).

1 D. Pekker and C. M. Varma, Annu. Rev. Condens. MatterPhys. 6, 269 (2015).

2 P. W. Anderson, Phys. Rev. 130, 439442 (1963)3 P. W. Higgs, Phys. Lett. 12, 132133 (1964).4 M. Greiter, Ann. Phys. (N.Y.) 319, 217 (2005).5 Pashkin, Alexej, and Alfred Leitenstorfer, Science

345,6201 (2014)6 P. B. Littlewood and C. M. Varma, Phys. Rev. B 26, 4883

(1982).7 R. Matsunaga, Y. I. Hamada, K. Makise, Y. Uzawa, H.

Terai, Z. Wang, and R. Shimano, Phys. Rev. Lett. 111,057002 (2013).

8 R. Matsunaga, N. Tsuji, H. Fujita, A. Sugioka, K. Makise,Y. Uzawa, H. Terai, Z. Wang, H. Aoki, and R. Shimano,Science 345, 1145 (2014).

9 D. Sherman, U. S. Pracht, B. Gorshunov, S. Poran, J.Jesudasan, M. Chand, P. Raychaudhuri, M. Swanson,N. Trivedi, A. Auerbach, M. Scheffler, A. Frydman andM.Dressel, Nat. Phys. 11,188(2015).

10 R. A. Barankov and L. S. Levitov, Phys. Rev. Lett. 96,230403 (2006).

11 A. F. Volkov and S. M. Kogan, Zh. Eksp. Teor. Fiz. 65,2038 (1973) [Sov. Phys. JETP 38, 1018 (1974)]

12 E. A. Yuzbashyan and M. Dzero, Phys. Rev. Lett. 96,230404 (2006).

13 N. Tsuji and H. Aoki, Phys. Rev. B 92, 064508 (2015).14 R. A. Barankov, L. S. Levitov, and B. Z. Spivak, Phys.

Rev. Lett. 93, 160401 (2004)15 P. R. Wallace, Phys. Rev. 71, 622 (1947)16 N. B. Kopnin and E. B. Sonin, Phys. Rev. Lett. 100, 246808

(2008).17 B. Rosenstein, B. Ya. Shapiro, and I. Shapiro Phys. Rev.

B 92, 05450318 L. Tarruell, D. Greif, T. Uehlinger, G. Jotzu, and T.

Esslinger, Nature 483, 302 (2012).19 T. Esslinger, Annual Review of Condensed Matter Physics

1, 129 (2010).20 S. M. Young, S. Zaheer, J. C. Y. Teo, C. L. Kane, E.

J. Mele, and A. M. Rappe, Phys. Rev. Lett. 108, 140405(2012).

21 H. Wang, H. Wang, H. Liu, H. Lu, W. Yang, S. Jia, X.Liu, X. Xie, J. Wei and J. Wang, Nature Materials (2015).

22 L. Aggarwal, A. Gaurav, G. S. Thakur, Z. Haque, A. K.Ganguli, and G. Sheet, Nature Materials (2015).

23 P. W. Anderson, Phys. Rev. 112, 1900 (1958).24 S. Tsuchiya, R. Ganesh, and T. Nikuni, Phys. Rev. B 88,

7

014527 (2013).25 E. A. Yuzbashyan, M. Dzero, V. Gurarie, and M. S. Foster,

Phys. Rev. A 91, 033628 (2015).26 M. Dzero, E. A. Yuzbashyan, B. L. Altshuler, and P. Cole-

man, Phys. Rev. Lett. 99, 160402 (2007).27 C. Chin et al., Science 305, 1128 (2004).

28 T. Kampfrath, K. Tanaka, and K. A. Nelson, Nature Pho-tonics 7, 680 (2013).

29 Erdelyi, A, Archive for Rational Mechanics and Analysis7.1 (1961): 1-20.


Recommended