Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Doping and tilting on optics in noncentrosymmetric multi-Weyl semimetals

S. P. Mukherjee1 and J. P. Carbotte1, 2


1
Department of Physics and Astronomy, McMaster University, Hamiltion, Ontario, Canada L8S 4M1
2
Canadian Institute for Advanced Research, Toronto, Ontario, Canada M5G 1Z8
Weyl semimetal (WSM) feature tilted Dirac cones and can be type I or II depending on the mag-
nitude of the tilt parameter (C). The boundary between the two types is at C = 1 where the cones
are tipped and there is a Lifshitz transition. The topological charge of a WSM is one. In multi-Weyl
it can be two or more depending on the value of the winding number J. We calculate the absorptive
arXiv:1803.06280v1 [cond-mat.str-el] 16 Mar 2018

part of the AC optical conductivity both along the tilt direction (σzz ) and perpendicular to it (σxx )
as a function of the tilt (C) and chemical potential (µ). For zero tilt there is a discontinuous rise in
both σxx and σzz at photon energy Ω = 2µ followed by the usual linear in Ω law for σxx at J = 1, 2
and σzz at J = 1. For J = 2 and σzz the interband background is constant rather than linear in Ω.
For type I there is a readjustment of optical spectral weight as the tilt is increased. The absorption
starts from zero at 2µ/(1 + C) and then rises in a quasilinear fashion till it merge with the usual
undoped untilted interband background at 2µ/(1 − C). The discontinuous rise at twice the chemical
potential of the untilted case is lost. For type II the interband background of the undoped untilted
case is never recovered. For noncentrosymmetric materials the energies of a pair of opposite chirality
Weyl nodes become shifted by ±Q0 and this leads to two separate absorption edges corresponding
to the effective chemical potential of each of the two nodes at 2(µ + χQ0 ) depending on chirality
χ = ±. We provide analytic expressions for the conductivity in this case which depend only on
the ratio Q0 /µ and tilt when plotted against Ω/µ. The signature of finite energy shift Q0 is more
pronounced for σzz and J = 2 than for the other cases.

PACS numbers: 72.15.Eb, 78.20.-e, 72.10.-d

I. INTRODUCTION paper we considered the AC optical conductivity along


the direction of the tilt ℜσzz (Ω) and perpendicular to it
ℜσxx (Ω) with particular emphasis on the effect of tilt and
Weyl fermions are known to exist in many different of doping away from charge neutrality. We also include
classes of semimetals. Initially suggested in the py- in our continuum Hamiltonian, terms that deal with both
rochlore iridates, Rn2 In2 O7 1 , this was followed with the- TR and inversion symmetry breaking44,45 . The transport
oretical investigations in the noncentrosymmetric tran- properties of mWSM including the effect of anisotropic
sition metal monophosphides2 which were soon verified residual scattering, short range and charged impurities
in experiments for TaAs3–6 and other related materials7,8 within a Boltzmann approximation have been considered
including the time-reversal(TR) symmetry breaking com- by Park et. al.46 and the magnetoconductivity by Sun
pound YbMnBi2 9 . Weyl nodes come in pairs of opposite and Wang47 following closely previous work on ordinary
chirality and result when the degeneracy of a doubly de- Weyl48 .
generate Dirac point is lifted through broken inversion
or TR symmetry. The Weyl cones can be tilted with The paper is structured as follows. In section II we
respect to the energy axis and type I or II Weyl nodes specify the model continuum Hamiltonian on which all
result depending on the magnitude of the tilt (C). For C our work is based. It includes a pair of multi-Weyl node
less than one in units of the appropriate Fermi velocity of opposite chirality and terms which break time reversal
we have type I and for C > 1 type II10 (overtilted). For and inversion symmetries. The first displaces the Weyl
type I in the undoped material the density of state at nodes in momentum space by ±Q and the second dis-
the Fermi surface remains zero but for type II it becomes places them in energy by a shift ±Q0 . The Green’s
finite as electron and hole pockets have formed11–23 . Hy- function is specified and the current Jx (perpendicular to
brid Weyl semimetals24 have also been considered where the tilt direction) and Jz (parallel to the tilt direction)
the tilt of the opposite chirality node can be different and are computed from the Hamiltonian. From this informa-
indeed one type I with the other type II. The topolog- tion the absorptive part of the AC optical conductivity
ical charge of a Weyl node can be greater than one for ℜσxx (Ω) and ℜσzz (Ω) as a function of photon energy are
winding numbers J two and above and these semimet- calculated from a Kubo formula. In section III we reduce
als are referred to as multi-Weyl25–30 . The absorptive the expression for both ℜσxx (Ω) and ℜσzz (Ω) to simple
part of the longitudinal AC optical conductivity ℜσii (Ω) analytic formulas which depend on the tilt and on the
with i = x, y as a function of photon energy provides di- doping through the value of the chemical potential and
rect information on the dynamics of the Dirac and Weyl on the energy shift Q0 of the Weyl nodes. The displace-
fermions31–36 as experiments have confirmed37–42 . The ment in momentum Q of the two nodes, which is known
optical conductivity in multi-Weyl semimetal (mWSM) to play a key role for the real part of the DC anoma-
has been considered by Ahn, Mele and Min43 . In this lous Hall conductivity, drops out of the expressions for
2

the absorptive part of the conductivity. In section IV The electronic energy dispersions corresponding to the
we present results. We start with the centrosymmetric above Hamiltonian are,
case and show results for ℜσzz (Ω) as a function of vari-
able tilt comparing the case J = 1 (Weyl) with J = 2 ǫs,s′ = Cs′ (kz − s′ Q) − s′ Q0 + sv⊥ ×
s
(multi-Weyl) and highlight the difference between type  2 
kx + ky2 J 2
I and type II Weyl. In section V this is followed with k02 + v02 (kz − s′ Q) (3)
a series of results for noncentrosymmetric materials for k02
which Q0 is non-zero. Finite Q0 leads to two steps in both
ℜσxx (Ω) and ℜσzz (Ω) which become modified as the tilt where s = ± stands for conduction(+) and valence(−)
is increased. A comparison of ℜσzz (Ω) with ℜσxx (Ω) is bands and v0 = vz /v⊥ . For a set of values of the param-
presented for a fixed illustrative value of Q0 = 0.5 for eters (v⊥ = 1, k0 = 1, v0 = 0.5, Q = 2.5, Q0 = 0.5) we
the case of J = 1 and both type I and II are consid- plot in Fig.[1] the energy dispersion for different winding
ered. Similarities and differences are emphasized, as is number J. We see that now the conical cross section de-
the low photon energy region. Next a comparison of re- viates from circular and evolves to elongated elliptical at
sults for the ℜσxx (Ω) and ℜσzz (Ω) in the case of J = 1 higher J. The matrix Green’s function corresponding to
with emphasis on the low photon energy regime is pre- the above Hamiltonian is given by,
sented. Only type I case is considered and the value of h i−1
Q0 is varied with a view to understand how the conduc- Ĝs′ (k, z) = I2 z − Ĥs′ (k) , (4)
tivity reflects this parameter. Of particular interest is
the appearance of a particular region of photon energies where I2 is a 2 × 2 unit matrix. It is straight forward
in which only the negative chirality node is contributing. to show that the Green’s function can be written in the
This is followed by a comparison of the J = 1 to the following form,
J = 2 case both for ℜσzz (Ω) for C fixed at 0.5 (type I)
and another at 1.5 (type II) while the energy shift be- 1X 1
Ĝs′ (k, z) =
tween the two Weyl nodes of opposite chirality is varied. 2 s=± z − Cs′ k̃z,s′ + sk̃s′ + s′ Q0
In section VI we provide a summary and conclusion.  
1 − ss′ (vz k̃z,s′ /k̃s′ ) −ss′ v⊥ k0 (k̃−J
/k̃s′ )
(5)
−ss′ v⊥ k0 (k̃+
J
/k̃s′ ) 1 + ss′ (vz k̃z,s′ /k̃s′ )
q
II. FORMALISM where k̃z,s′ = kz − s′ Q and k̃s′ = v⊥ 2 k 2 k̃ J k̃ J + v 2 k̃ 2 .
0 + − z z,s′

We begin with the minimal continuum Hamiltion for


an isolated Weyl node of chirality s′ with both broken
TR and inversion symmetry. Additionally we assume
that the winding number associated with the Weyl node
is J. The broken time reversal symmetry displaces the
Weyl cone in momentum space by an amount ±Q while
the broken inversion symmetry shifts their energy by
±Q0 31,44,45 .
  
′ ′J J
Ĥs′ (k) = Cs′ (kz − s Q) + s v⊥ k0 k̃− σ+ + k̃+ σ− +
 FIG. 1: (Color online) Effect of both broken time-reversal
symmetry as well as inversion symmetry on untilted mWSM
vz (kz − s Q)σz − s′ Q0

(1) with different winding number J. The one on the left is for
J = 1, and on the right for J = 2. We see that the cross
section gets elongated from J = 1 to J = 2 as shown.
Here s′ = ±1 for Weyl nodes of opposite chirality. Cs′ de-
scribe the amount of tilting of the particular chiral node.
The velocity v⊥ is the effective velocity of the quasipar- Now we outline the derivation of the conductivity σi,j
ticles in the plane perpendicular to the z-axis while vz where i, j = x, y, z. For this we need the current-current
is the velocity along it and tilt Cs′ is normalized by this correlation function which in matrix form is given by,
parameter. Here k0 is a system dependent parameter X X Z d3 k
having the dimension of momentum. Also k̃± = k± /k0 Π̂ij (Ωm , q) = T
with k± = kx ± ıky and σ± = 12 (σx ± ıσy ). The Pauli ω ′
(2π)3
n s =±
matrices σi where i = x, y, z are defined as usually by, Jˆi,s′ Ĝs′ (k + q, ωn + Ωm ) × Jˆj,s′ Ĝs′ (k, ωn ), (6)
     
0 1 0 −ı 1 0 where Ωm and ωn are Matsubara frequencies, q is the
σx = , σy = , σz = . (2)
1 0 ı 0 0 −1 internal momentum which we set to zero and Jˆi,s′ is the
3

current operator which is defined as, A. Results for ℜσxx,s′ (Ω)

∂ Ĥs′ (k)
Jˆi,s′ = e . (7) I. For 0 < Cs′ ′ < 1 (type I tilting)
∂ki
Next we have to evaluate the sum over the internal Mat-
subara frequencies ωn and after that we have to analyt- ℜσxx,s′ (Ω) e sL′
= 0, where Ω̃ < Ω
ically continue the result by replacing ıΩm with Ω + ıδ, µ′ e2 /8π
which will give us the retarded current-current correla- e s′ > Ω̃ > Ω
e s′
tion function Πij (Ω, 0). The conductivity is then defined = I 1′ , where Ω
s U L

by the following formula, = Is2′ , where es ,
Ω̃ > Ω (12)
U

Πij (Ω, 0)
σij (Ω) = − . (8)
ıΩ II. For Cs′ ′ > 1 (type II tilting)
Following the above prescription we derive the expres-
sions for the longitudinal conductivity σxx (Ω) and σzz (Ω)
as shown below, ℜσxx,s′ (Ω) e sL′
′ 2
= 0, where Ω̃ < Ω
µ e /8π
Πxx (Ω, 0)
σxx (Ω) = − e s′ > Ω̃ > Ω
= I 1′ , where Ω e s′
ıΩ s U L
2 2 −(2J−1) 2 X Z Λ−s Q Z ∞ 2J−1 e s′ ,
= Is3′ , where Ω̃ > Ω

ıe J k0 v⊥ k⊥ dk⊥ U (13)
= dkz ×
2π 2 Ω ′ k
s′ =± −Λ−s Q 0
  where for any variable a, a′ = a/vz and ã = a′ /µ′ . Also
f (Cs′ kz + k0 k − µs′ ) − f (Cs′ kz − k0 k − µs′ ) × we have used the following shorthands,
 
1
(k02 k 2 + vz2 kz2 ) × + ıπδ(4k0
2 2
k − Ω 2
) , (9) J 3 1 J 1 1
4k02 k 2 − Ω2 Is1′ = (4 + ′ + ′3 )Ω̃ − ( ′ + ′3 )|1 + s′ Q̃0 |
24 Cs′ Cs′ 4 Cs′ Cs′
Πzz (Ω, 0) J J
σzz (Ω) = − + ′3 |1 + s′ Q̃0 |2 − |1 + s′ Q̃0 |3
ıΩ 2Cs′ Ω̃ 3Cs′3′ Ω̃2
ıe2 X Z Λ−s′ Q Z ∞
k⊥ dk⊥ J Ω̃
= 2
dkz × Is2′ =
2π k0 Ω ′ ′ k 3
s =± −Λ−s Q 0
  3 J 3 1 J
Is′ = ( ′ + ′3 )Ω̃ + ′3 |1 + s′ Q̃0 |2
f (Cs′ kz + k0 k − µs′ ) − f (Cs′ kz − k0 k − µs′ ) × 12 Cs′ Cs′ Cs′ Ω̃
  ′ e ′e
2 2
 2 2 2 2 2 2 2 2J −2(J−1) e s′ = 2 1 + s Q0 , Ω
Ω e s′ = 2 1 + s Q0 . (14)
(Cs′ + vz ) k0 k − vz kz − v⊥ (Cs′ − vz )k⊥ k0 L
1 + Cs′ ′ U
1 − Cs′ ′
 
1 2 2 2
× + ıπδ(4k0 k − Ω ) . (10) Note that the displacement in momentum of the two
4k02 k 2 − Ω2
Weyl nodes has dropped out of the absorptive part of
Here Λ is the cutoff, k⊥ is the momentum perpendicular the conductivity in our clean limit calculations. In the
to kz , f (E) = (eE/T + 1)−1 is the Fermi function at inversion symmetric case , with Q0 = 0 our expressions
finite temperature T with µ the chemical potential and (12) to (14) reduce, as they must, to those of reference
µs′ = µ+s′ Q0 is the effective chemical potential for Weyl [35]. In this reference only the case for winding num-
node with chirality s′ . Also we note that k⊥ and kz are ber J = 1 was treated but we have verified here that for
related to each other by the equation, multi-Weyl with J 6= 1, we need to only multiply by J as
q was found in reference [43] for the no tilt case.
k = v⊥ 2 (k /k )2J + v 2 (k /k )2 . (11)
⊥ 0 z z 0

III. DERIVATION OF LONGITUDINAL


CONDUCTIVITIES FOR FINITE Q AND Q0 B. Results for ℜσzz,s′ (Ω)

In this section we derive the expressions for the longi-


tudinal conductivities σxx (Ω) and σzz (Ω) which are the In this subsection we state the final result for
central results of this article. Details of the derivation ℜσzz,s′ (Ω). Essential steps are mentioned in the Ap-
are presented in Appendix A and B so here we only state pendix B. There we see that the same Eq.(12) and (13)
ℜσ ′ (Ω)
the results. hold for ezz,s
2 /8π but with Is′ replaced by Ls′ with,
4

replaced by C2′ because we assume that the tilt has the


  2−J √ same magnitude on both nodes. Tilt inversion symme-
k0 vz Ω J πΓ(1 + J1 ) try guarantees that the conductivity only depends on its
L1s′ = 2 −
2 J Jv⊥ Ω0 2Γ( 23 + J1 ) absolute value. The total contribution to the conduc-
  "  2 # tivity is therefore twice the amount in Eq.(15). This
1 2|µs′ | 1 1 3 1 2|µs′ | also modifies the Ω dependent prefactor in all the L-
− 1 2F 1 ,− ; ; −1 ,   2−J
Cs′ ′ Ω 2 J 2 Cs′2′ Ω J
functions to 2k 0 vz Ω
which can be equivalently
  2−J √ 2 Ω0 2 J Jv⊥
2 k0 vz Ω J πΓ(1 + J1 ) ′ 2e
Ls′ = 2 , written as µ v0 Ω (J
= 1) and k0 v0 (J = 2). This also
2 J Jv⊥ Ω0 Γ( 23 + J1 ) change Ωe s′ to Ω
e L = 2 ′ and Ω e s′ to Ω
eU = 2
L 1+C2 U |1−C2′ | .
  2−J
3 k0 vz Ω J With all this in mind we plot ℜσzz,s′ (Ω) (in appropri-
Ls′ = 2 × ate unit) for type I in Fig.[2] and type II in Fig.[3] for
2 J Jv⊥ Cs′ Ω0

  "  2 # both J = 1 (top frames), 2 (bottom frames).
2|µs′ | 1 1 3 1 2|µs′ |
1+ 2F 1 ,− ; ; 1+ +
Ω 2 J 2 Cs′2′ Ω
  "  2 # 8
2|µs′ | 1 1 3 1 2|µs′ | (a)
1− 2F 1 ,− ; ; 1− , (15) J=1
Ω 2 J 2 Cs′2′ Ω 7

ℜσzz(T=0,Ω)/(e µ′v0/8π)
6
where Ω0 = v⊥ k0 and 2 F 1 is the Hypergeometric func-

2
tion defined as 5

2
X∞
(a)n (b)n z n 4
2 F 1 (a, b; c; z) = , (16)
n=0
(c)n n!
3
with (q)n the (rising) Pochhammer symbol, which is de- 2
fined by C2′ = 0.1
1 = 0.5
(q)n = 1 for n = 0 = 0.9
0
= q(q + 1)...(q + n − 1) for n > 0.(17) 0 1 2 3 4 5 6 7 8 9 10 11 12
Ω/µ
These formulas are central to the current article. In the 0.8
next two sections we will use various special cases of these (b)
formulas to understand the implication of the inversion 0.7
J=2
ℜσzz(T=0,Ω)/(e k0v0/8π)

symmetry breaking, of tilting and of doping on the con-


0.6
ductivity. For zero tilt we get,
  2−J √ 0.5
2

ℜσzz,s′ (Ω) k0 vz Ω J πΓ(1 + J1 )


= L2s′ = 2 (18) 0.4
2
e /8π 2 J Jv⊥ Ω0 Γ( 23 + J1 )
0.3
for Ω > 2µs′ and zero for Ω < 2µs′ , a known result43
in the limit of Q0 = 0. Note that the displacement in 0.2
momentum of the two Weyl nodes (Q) has dropped out C2′ = 0.1
of our clean limit calculations of the absorptive part of 0.1 = 0.5
= 0.9
the conductivity. Of course, as is well known49–51 it plays
0
a crucial role in the DC Hall conductivity which is found 0 1 2 3 4 5 6 7 8 9 10 11 12
to be proportional to Q49–51 . This parameter also enters Ω/µ
prominently in other transport coefficients52,53 .
FIG. 2: (Color online) The real part of the dynamic optical
conductivity in z-direction ℜσzz (T = 0, Ω) at temperature
T = 0 in units of e2 µ′ v02 /8π(for J = 1) and e2 k0 v0 /8π(for
IV. CONDUCTIVITIES WITH INVERSION
J = 2) as a function of photon energy Ω normalized to the
SYMMETRY
chemical potential µ. Top frame (a) corresponds to J = 1
while the bottom frame (b) is for J = 2. They compare results
We begin this section with the case when Q0 = 0 (cen- for different values of tilt namely C2′ = 0.1 (solid red), C2′ =
trosymmetric). We need only consider ℜσzz,s′ (Ω) for 0.5 (dash blue) and C2′ = 0.9 (dash-dot green) all type I. No
J = 1 as ℜσxx,s′ (Ω) has been discussed before35 . For inversion symmetry breaking is included (Q0 = 0). Winding
Q0 = 0 the L-functions defined in Eq.(15) becomes in- number has a large effect on this quantity including the result
dependent of s′ as µs′ is now replaced by µ and Cs′ ′ is of tilting.
5

In Fig.[2] we consider three representative values of tilt, with


C2′ = 0.1 (solid red), C2′ = 0.5 (dash blue) and C2′ = 0.9  
1 1 3 2 x2
(dash-dot green). The solid red curve includes only a 2F 1 ,− ; ;x = 1 − , for J =1
small value of tilt and is very close to the no tilt re- 2 J 2 3

sult. In this case ℜσzz,s′ (Ω) is zero till Ω = 2µ where 1 − x2 arcsin x
= + , for J = 2, (22)
for J = 1 it discontinuously jumps up to merge with 2 2x
the well known linear in photon energy (Ω) law associ-
ated with the interband background absorption of a Dirac
3.2
cone31–33 . The optical spectral weight in the interband (a) J=1
background below 2µ is transferred, due to Pauli block- 2.8
ing, to the Drude interband contribution. In the clean

ℜσzz(T=0,Ω)/(e µ′v0/8π)
limit used here, this is a delta function at Ω = 0 not 2.4

2
shown in our plot. While the sharp jump at Ω = 2µ
2
of the no tilt case remains clearly seen in the red curve

2
of Fig.[2a] this is not so for the other two curves dash 1.6
blue (C2′ = 0.5) and dash-dot green (C2′ = 0.9). These
are qualitatively better described as a roughly quasilinear 1.2
rise out of zero at Ω = 2µ/(1 + C2′ ) and merging with the
0.8 C2′ = 1.5
zero tilt case at Ω = 2µ/(1 − C2′ ). In the units used, the
slope of the interband background in the untilted case =2
0.4 =3
is 2/3, and the conductivity at Ω = 2µ is equal to 4/3 =4
but for the tilted case it is instead half this value namely 0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6
2/3. The quasilinear law in the photon energy interval Ω/µ
2µ/(1 + C2′ ) to 2µ/(1 − C2′ ) passes exactly through the
half way mark of the sharp absorption edge of the un- (b) J=2
tilted case. This follows for type I from Eq.(15) which
0.5
gives,
ℜσzz(T=0,Ω)/(e k0v0/8π)

  2−J √ 0.4
ℜσzz,s′ (2µ) k0 vz 2µ J πΓ(1 + J1 )
= L1s′ (2µ) = 2
2

.
2
e /8π 2 J Jv⊥ Ω0 2Γ( 23 + J1 ) 0.3
(19)
This is exactly half the value of the no-tilt conductivity
0.2
at the absorption edge given by Eq.(18) evaluated at Ω =
2µ. It is completely independent of the tilt so that all the C2′ = 1.5
curves in Fig.[2a] pass through this same point ℜσe2zz/8π
(2µ)
= 0.1 =2
=3
2 2 ′ =4
v
3 0 µ . Similar results hold in the case of winding number
J = 2 (multi Weyl). The interband optical background 0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6
is no longer linear in photon energy for no tilt. Instead Ω/µ
it is constant43 as in graphene54
FIG. 3: (Color online) The real part of the dynamic optical
ℜσzz (Ω) π conductivity in z-direction ℜσzz (T = 0, Ω) at temperature
= k0 v0 (20) T = 0 in units of e2 µ′ v02 /8π (for J = 1) and e2 k0 v0 /8π (for
e2 /8π 4 J = 2) as a function of photon energy Ω normalized to the
chemical potential µ. Top frame (a) corresponds to J = 1
as we see in Fig.[2b] where all curves pass through π/8 while the bottom frame (b) is for J = 2. They compare results
in the units used for the conductivity. As compared with for different values of tilt namely C2′ = 1.5 (solid red), C2′ = 2
the top frame the sharp jump at Ω = 2µ of the no tilt (dash blue), C2′ = 3 (dash-dot green) and C2′ = 4 (dash double
case remains more discernible with increasing tilt. dot magenta) all type II. No inversion symmetry breaking is
included (Q0 = 0). Winding number has a large effect on this
In Fig.[3] we present a series of results for type II Weyl
quantity including the result of tilting.
(C2′ > 1). A first observation is that as long as C2′ < 2
Eq.(19) still gives the value of the conductivity at Ω = 2µ.
However for C2′ > 2 which works out to
"  2 #
ℜσzz,s′ (2µ) 2 ′ 1 1 2
ℜσzz,s′ (2µ) = v0 µ ′ 1 − (for J = 1) (23)
= L3s′ (2µ) e2 /8π C2 3 C2′
e2 /8π  s 
  2−J "  2 #  2  
2k0 vz 2µ J 1 1 3 2 k0 v0  1 2 1 2 
= 2 2F 1 ,− ; ; (21)
, = 1− + arcsin (for J = 2),
2 J Jv⊥ C2′ Ω 0 2 J 2 C ′ 4 C2′ C2′ 2 C2′
2
6

both now depend on the tilt as is verified in the numeri- e0


I. For 0 < C2′ < Q
cal work of Fig.[3]. We present results for C2′ = 1.5 (solid
red), C2′ = 2 (dash blue), C2′ = 3 (dash-dot green) and
C2′ = 4 (dash double dot magenta). As C2′ is increased ℜσxx (Ω) e<Ω
e−
= 0, where Ω L
the curves start at smaller values of photon energy, even- µ′ e2 /8π
tually crossing each other and at the larger values of Ω 1 e− > Ω
e >Ω e−
= I− , where Ω U L
shown, the slope of the quasilinear behavior in the top
frame (J = 1) decreases while in the lower frame (J = 2) 2
= I− , where Ω e+ > Ω
e >Ω e−
L U
they flatten out at ever smaller values. For Ω > C2µ ′ 1 2 e+ > Ω e >Ω
e+
2 −1 = (I+ + I− ), where Ω U L
and C2′ going to infinity Eq.(15) gives, 2 2 e
= (I + I ), where Ω > Ω e +
(29)
+ − U
ℜσzz,s′ (Ω)
= L3s′ (Ω)
e2 /8π
  2−J   e 0 < C2′ < 1
II. For Q
∼ k0 vz Ω J
1 1 3 1
= 2 −1 2F 1 ,− ; ; (24)
2 J Jv⊥ C2′ Ω0 2 J 2 C2′2
ℜσxx (Ω) e<Ω
e−
which is to leading order = 0, where Ω L
µ′ e2 /8π
  2−J 1
= I− , where e+ > Ω
Ω e>Ω e−
k0 vz Ω J L L

= (forJ = 1)
2
Ω0
1
= (I+ 1
+ I− ), where Ωe− > Ωe >Ω
e+
2 J −1 Jv⊥ C2′ U L
  2−J e e
= (I+ + I− ), where Ω > Ω > Ω−
1 2 + e
∼ k0 vz Ω J U U
= (forJ = 2). (25) 2 2 e>Ω
e+
2
2
J −1 Jv C ′
⊥ 2 Ω0 = (I+ + I− ), where Ω U (30)

There is a 1/C2′ decay factor as the tilt gets very large.


More specifically e0 < 1
III. For C2′ > 1 but C2′ Q
ℜσzz,s′ (Ω) v02 Ω
= (forJ = 1) (26)
e2 /8π 2vz C2′ ℜσxx (Ω)
= 0, where e <Ω
Ω e−
L
µ′ e2 /8π
and 1
= I− , where e+ > Ω
Ω e>Ω e−
L L
ℜσzz,s′ (Ω) k0 v0 1 1 e
= (I+ + I− ), where ΩU > Ω− e>Ωe+
= (forJ = 2). (27) L
e2 /8π 2C2′
= (I+1
+ I−3 e+ > Ω
), where Ω e>Ωe−
U U
Consequently the slope of the linear in Ω law of Eq.(26) = (I 3 + I 3 ), where Ω e>Ω e+ (31)
+ − U
decays as the inverse of the tilt parameter for J = 1 and
the magnitude of the constant value in Eq.(27) also de-
cays as one over the tilt. It is interesting to compare these
relations with the case of ℜσxx,s′ (Ω) given in Eq.(13) e0 > 1
IV. For C2′ > 1 and C2′ Q

ℜσxx,s′ (Ω) ∼ JΩ ,
= Is3′ = (28)
µ′ e2 /8π 4C2′ µ
ℜσxx (Ω) e <Ω
e−
= 0, where Ω L
in the limit of interest here. This differs from Eq.(26) by µ′ e2 /8π
a factor of 2v02 for the case J = 1. 1
= I− , where e− > Ω
Ω e>Ωe−
U L
3
= I− , where Ω e+ > Ωe>Ωe−
L U

V. AC CONDUCTIVITY WHEN Q0 6= 0
1
= (I+ 3
+ I− e+ > Ω
), where Ω e>Ω e+
U L
3 3 e
= (I + I ), where Ω > Ω .e +
(32)
+ − U
In this section we describe the effect of the finite Q0
on both the AC conductivities ℜσxx (T = 0, Ω) and
ℜσzz (T = 0, Ω). Here we make the same assumption For ℜσzz (T = 0, Ω) we only need to replace Is′ in Eq.(29)
as before that both the nodes have the same absolute to (32) by the Ls′ of Eq.(15) and drop the µ′ on the left
value of tilt. The tilt direction makes no difference. hand side of these equations.
7

for the negative chirality node ends at Ω/µ = 1.25 and


3.2 J=1, C′2= 0.001 1.66 respectively, both are below the minimum photon
ℜσxx(T=0,Ω)/(e µ′/8π)
2.8 energy for which the positive chirality node contributes
which are 2.5 and 2.14 respectively. In both instances
2.4 we see a linear region with slope exactly equal to its no
tilt value representing the photon energy interval over
2

2(1+Q0/µ)
2(1-Q0/µ) which only the negative chirality node contributes even
1.6 though the tilt has spread out the optical spectral weight
coming from the positive chirality node. In this region
1.2
the contribution from the negative chirality node has also
0.8 recovered its zero tilt slope. In all other curves this no
longer happens and no perfectly linear region remains.
0.4
Q0/µ = 0.5
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Ω/µ
4
(a) J=1, Q0/µ=0.5
FIG. 4: (Color online)The real part of the dynamic optical 3.5
conductivity perpendicular to the z-axis ℜσxx (T = 0, Ω) at

ℜσxx(T=0,Ω)/(e µ′/8π)
zero temperature in units of e2 µ′ /8π as a function of photon 3
C′2= 0.0
energy Ω normalized to the chemical potential µ. This plot = 0.2

2
includes broken inversion symmetry with Q0 /µ = 0.5. The 2.5 = 0.4
winding number is J = 1 and a small tilt C2′ = 0.001 is = 0.8
2 = 1.6
included but has negligible effect on the plot. = 2.4
1.5
2 ′
In Fig.[4] we plot ℜσxx (T = 0, Ω) in units of e µ /8π 1
as a function of photon energy Ω normalized to the chem-
ical potential µ for Q0 = 0.5 and C2′ close to zero. 0.5
The plot follows directly from Eq.(29) and reproduces
0
a similar plot in Ref.[31]. The contribution of the neg- 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6
ative chirality node has the effective chemical potential Ω/µ
µ(1 − Q0 /µ) = µ/2 jumps from zero to a value of 1/3 at 1.1
Ω = µ in the units of Fig.[4]. For the positive chirality (b) J=1, Q0/µ=0.5
1
node the effective chemical potential is instead 3µ and
0.9
the second jump in the conductivity has magnitude one.
ℜσxx(T=0,Ω)/(e µ′/8π)

The slope of the first straight line between Ω/µ equal to 1 0.8
C′2= 0.0
and 3 is 1/3 and involves only the negative chirality node 0.7 = 0.2
2

= 0.4
while in the second interval above Ω/µ = 3 it is twice this 0.6 = 0.8
value and involves both nodes. The double jump behav- 0.5
= 1.6
= 2.4
ior seen in this figure is the hallmarks of broken inversion
0.4
symmetry i.e. a finite value of Q0 . When a finite tilt is
included in the calculation it will modify the contribu- 0.3
tion from each node as shown in Fig.[5] and Fig.[6] for 0.2
ℜσxx (Ω) and ℜσzz (Ω) respectively. Both are for J = 1 0.1
and Q0 /µ = 0.5. The solid black curve in Fig.[5] is the 0
same as shown in Fig.[4] with a first step at Ω/µ = 1 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4
Ω/µ
and a second at Ω/µ = 3. As C2′ is increased dot indigo
C2′ = 0.2, dash red C2′ = 0.4, dash-dot blue C2′ = 0.8,
FIG. 5: (Color online) The real part of the dynamic optical
dash double dot green C2′ = 1.6 and dot double dash
conductivity perpendicular to the z-axis ℜσxx (T = 0, Ω) at
magenta C2′ = 2.4 the steps progressively lose their in- zero temperature in units of e2 µ′ /8π as a function of photon
tegrity as the optical spectral weight is transferred from energy Ω normalized to the chemical potential µ. The top
the region above twice the effective chemical potential frame (a) shows results up to Ω/µ = 6 while the bottom
of a given node to the region below this photon energy. frame (b) is an expanded version of the low energy region to
The amount of transfer and the interval over which this Ω/µ = 2.5 only. In both frames the winding number is J = 1
transfer occurs increase as C2′ is increased out of zero. and the inversion symmetry breaking parameter has been set
As is seen in the lower frame of Fig.[5] where only the to Q0 /µ = 0.5. Six values of tilt are shown (solid black)
region Ω/µ below 2.5 is shown on an expanded scale for C2′ = 0, (dot indigo) C2′ = 0.2, (dash red) C2′ = 0.4, and
the C2′ = 0.2 (dot indigo) and C2′ = 0.4 (dash red) the (dash-dot blue) C2′ = 0.8 (type I), (dash double dot green)
redistribution of optical spectral weight due to the tilt C2′ = 1.6, (dot double dash magenta) C2′ = 2.4.
8

4 an extra factor of v02 and this factor is one in the isotropic


(a) J=1, Q0/µ=0.5 Weyl case (J = 1). For type I, ℜσxx (T = 0, Ω) reduces to
3.5 e2 Ω 31
24π vz consistent with our previous results once an ~2
ℜσzz(T=0,Ω)/(e µ′v0/8π)

3
C′2= 0.0 is restored. For type II (C2′ > 1) we get for large values
2

= 0.2 of Ω,
2.5 = 0.4
2

= 0.8  
= 1.6 e2 Ω 1 1
2 ℜσxx (T = 0, Ω) = 1 + and (33)
= 2.4 8π vz 4C2′ 3C2′2
1.5
 
1 e2 2 Ω 1 1
ℜσzz (T = 0, Ω) = v 1− . (34)
8π 0 vz 2C2′ 3C2′2
0.5

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6 1.6
Ω/µ (a) J=1, C′2=0.5
1.4
1.4 (b) J=1, Q0/µ=0.5

ℜσxx(T=0,Ω)/(e µ′/8π)
1.2
Q0/µ = 0.0
ℜσzz(T=0,Ω)/(e µ′v0/8π)

1.2 = 0.2

2
1 = 0.4
2

= 0.6
1
C′2= 0.0 0.8 = 0.9
2

= 0.2
0.8 = 0.4 0.6
= 0.8
= 1.6
0.6 = 2.4 0.4

0.4 0.2

0.2 0
0 0.5 1 1.5 2 2.5 3
Ω/µ
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3 2
Ω/µ (b)
1.8 J=1, C′2=0.5
FIG. 6: (Color online) The real part of the dynamic optical
ℜσzz(T=0,Ω)/(e µ′v0/8π)

1.6
conductivity along the z-axis ℜσzz (T = 0, Ω) at zero temper-
2

ature in units of e2 µ′ v02 /8π as a function of photon energy 1.4


Q0/µ = 0.0
Ω normalized to the chemical potential µ. This plot is to be = 0.2
2

1.2
compared with Fig.[5]. The top frame (a) shows results up to = 0.4
1 = 0.6
Ω/µ = 6 while the bottom frame (b) is an expanded version = 0.9
of the low energy region to Ω/µ = 3 only. In both frames 0.8
the winding number is J = 1 and the inversion symmetry
0.6
breaking parameter has been set to Q0 /µ = 0.5. Six values
of tilt are shown (solid black) C2′ = 0, (dot indigo) C2′ = 0.2, 0.4
(dash red) C2′ = 0.4, and (dash-dot blue) C2′ = 0.8 (type I), 0.2
(dash double dot green) C2′ = 1.6, (dot double dash magenta)
C2′ = 2.4. 0
0 0.5 1 1.5 2 2.5 3
Ω/µ
Note also that below Ω/µ = 1 the curves for C2′ = 0.2 and
0.4 show concave downward behavior which changes to FIG. 7: (Color online)The real part of the dynamic op-
tical conductivity perpendicular to the z-axis (frame (a))
concave upward as we go through Ω/µ = 1. These effects
ℜσxx (T = 0, Ω) at zero temperature in units of e2 µ′ /8π as
are small and when ignored the underlying behavior is a function of photon energy (normalized to the chemical po-
quasilinear. tential µ) up to Ω/µ = 3. The tilt is C2′ = 0.5 (type I). Five
The results for ℜσzz (T = 0, Ω) in Fig.[6] has much values are chosen for the inversion symmetry breaking param-
the same qualitative behavior as for those in Fig.[5] but eter Q0 /µ namely Q0 /µ = 0 (solid black), Q0 /µ = 0.2 (dash
there are significant quantitative changes. One difference red), Q0 /µ = 0.4 (dash-dot blue), Q0 /µ = 0.6 (dash dou-
is that the curves for C2′ = 0.2 and 0.4 now show concave ble dot green) and Q0 /µ = 0.9 (dot double dash magenta).
upward behavior below Ω/µ = 1 which shift to concave Frame (b) is the real part of the dynamical optical conduc-
downward above this energy. The high energy behavior tivity along the z-axis ℜσzz (T = 0, Ω) at zero temperature in
is also quantitatively different. When there is no tilt units of e2 µ′ v02 /8π as a function of photon energy (normalized
ℜσzz (T = 0, Ω) is the same as ℜσxx (T = 0, Ω) except for to the chemical potential µ) up to Ω/µ = 3.
9

which shows differences between zz and xx conductivities top and bottom curve is the rise out of zero absorp-
beyond a factor of v02 . tion at Ω/µ = 1.33. It is sharper and concave down
in ℜσxx (T = 0, Ω) than for ℜσzz (T = 0, Ω) which is con-
5.5
cave upward, but these are small differences and both
(a) J=1, C′2=0.5 curves are qualitatively quasilinear with slightly different
5
slopes. As Q0 /µ is increased the absorption starts at ever
4.5
decreasing values of photon energies Ω/µ = 2(1−Q 0 /µ)
ℜσzz(T=0,Ω)/(e µ′v0/8π)

1+C2′
4
and a region develops which is entirely due to the nega-
2

3.5 tive chirality node up to 2(1+Q 0 /µ)


. Such a region never
2

1+C2′
3 arises in centrosymmetric materials. Further in the dash
2.5 double dot green curve for Q0 /µ = 0.6 and the dot dou-
2 ble dash magenta for Q0 /µ = 0.9 the negative chirality
1.5 Q0/µ = 0.0 node contribution to the conductivity has recovered its
= 0.2 no-tilt slope of 1/3 as is clearly seen in the figure. This
1 = 0.4
= 0.6
0.5 = 0.9
0 3 (a) J=1, C′2=1.5
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6 6.5 7 7.5 8
Ω/µ

ℜσzz(T=0,Ω)/(e µ′v0/8π)
2.5
0.8

2
(b) J=2, C′ =0.5
0.7 2 2

2
ℜσzz(T=0,Ω)/(e k0v0/8π)

0.6
1.5
0.5
2

1
Q0/µ = 0.0
0.4
= 0.4
0.5 = 0.6
0.3 = 0.7
= 0.9
Q0/µ = 0.0
0.2 0
= 0.2 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6
= 0.4 Ω/µ
0.1 = 0.6
= 0.9
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6 6.5 7 7.5 8 0.55
(b) J=2, C′2=1.5
Ω/µ 0.5
ℜσzz(T=0,Ω)/(e k0v0/8π)

0.45
FIG. 8: (Color online) The real part of the dynamic optical
0.4
conductivity along the z-axis ℜσzz (T = 0, Ω) at zero temper-
2

ature in units of e2 µ′ v02 /8π (for J = 1) and e2 k0 v0 /8π (for 0.35


J = 2) as a function of photon energy (normalized to the 0.3
chemical potential µ) up to Ω/µ = 8. The top frame (a) is 0.25
for J = 1 and the bottom frame (b) is for J = 2. A specific 0.2
value of tilt C2′ = 0.5 (type I) is chosen and the inversion Q0/µ = 0.0
symmetry breaking parameter Q0 /µ is changed. Five values 0.15
= 0.4
are chosen Q0 /µ = 0 (solid black), Q0 /µ = 0.2 (dash red), 0.1 = 0.6
= 0.7
Q0 /µ = 0.4 (dash-dot blue), Q0 /µ = 0.6 (dash double dot 0.05 = 0.9
green) and Q0 /µ = 0.9 (dot double dash magenta). 0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6
Ω/µ
In Fig.[7] we presents the same kind of results as for
FIG. 9: (Color online) The real part of the dynamic optical
the lower frame of Fig.[5] and [6] but now C2′ is fixed conductivity along the z-axis ℜσzz (T = 0, Ω) at zero temper-
at a value of C2′ = 0.5 and Q0 /µ is varied with J = 1. ature in units of e2 µ′ v02 /8π (for J = 1) and e2 k0 v0 /8π (for
The top frame gives results for ℜσxx (T = 0, Ω) and the J = 2) as a function of photon energy (normalized to the
bottom for ℜσzz (T = 0, Ω). Five values of Q0 /µ are chemical potential µ) up to Ω/µ = 6. The top frame (a) is
considered. The solid black is for Q0 /µ = 0, dash red for J = 1 and the bottom frame (b) for J = 2. A specific
for Q0 /µ = 0.2, dash-dot blue for Q0 /µ = 0.4, dash value of tilt C2′ = 1.5 (type II) is chosen and the inversion
double dot green for Q0 /µ = 0.6 and dot double dash symmetry breaking parameter Q0 /µ is changed. Five values
magenta for Q0 /µ = 0.9. For Q0 /µ = 0 solid black are chosen Q0 /µ = 0 (solid black), Q0 /µ = 0.4 (dash red),
both Weyl nodes contribute equally to the absorptive Q0 /µ = 0.6 (dash-dot blue), Q0 /µ = 0.7 (dash double dot
part of the conductivity. The main difference between green) and Q0 /µ = 0.9 (dot double dash magenta).
10

perfectly linear region extends from Ω/µ = 1.6 to 2.13 the momentum space separation of the two Weyl nodes
for Q0 /µ = 0.6 and from 0.4 to 2.53 for Q0 /µ = 0.9. drops out of the absorptive part of the AC conductivity.
In Fig.[8] and [9] we compare results for the ℜσzz (T = This is in sharp contrast to the critical role Q plays in the
0, Ω) in units of e2 µ′ v02 /8π and e2 k0 v0 /8π top and bottom anomalous DC Hall conductivity which is known49–51 to
frames respectively, for two values of tilt C2′ = 0.5 (type be directly proportional to Q. Terms proportional to Q
I) in Fig.[8] and C2′ = 1.5 (type II) in Fig.[9]. The top also enter other transport coefficients51–53 . In the case
frames are for winding number J = 1 and the bottom for of Q0 = 0 (centrosymmetric) our expressions properly
J = 2 (multi-Weyl). The value of the inversion symmetry reduce to known results. For winding number J = 1
breaking parameter Q0 which corresponds to the energy results for ℜσxx (Ω) when both the tilt and the doping
shift ±Q0 associated with the two Weyl nodes is varied. are non-zero are given in Ref.[35]. Ref.[43] gives results
In Fig.[8] Q0 /µ = 0 (solid black), Q0 /µ = 0.2 (dash red), in the multi-Weyl case for both ℜσxx (Ω) and ℜσzz (Ω)
Q0 /µ = 0.4 (dash-dot blue), Q0 /µ = 0.6 (dash double but mainly when doping and tilt are zero. Assuming
dot green) and Q0 /µ = 0.9 (dot double dash magenta). the magnitude of the tilt is the same for both nodes and
In Fig.[9] the line type and color are the same but Q0 /µ = the system has inversion symmetry (Q0 = 0) each Weyl
0, 0.4, 0.6, 0.7 and 0.9. In the top frame of Fig.[8] the point contributes equally to the conductivity. For finite
slope of the high energy linear in Ω/µ law remains 2/3 chemical potential µ, but zero tilt there is no absorption
and all the curves have merged in this region. In the below Ω = 2µ due to the Pauli blocking of the interband
solid black curve for Q0 /µ = 0 the effective chemical optical transitions. The lost optical spectral weight is
potential is the same for both nodes and they contribute transferred to the intraband Drude which, in the clean
equally. For Q0 /µ = 0.9 dot double dash magenta we limit, is a Dirac delta function at Ω = 0. Above the
see that a second region of linear dependence has clearly sharp discontinuous absorption threshold at 2µ the con-
emerged and the slope of this line is 1/3 because only the ductivity takes on its zero doping value. For ℜσxx (Ω)
negative chirality node is contributing and except for very this is a linear law for both J = 1 and J = 2. For
small values of Ω we are in the region Ω > 2µ− /(1 − C2′ ) ℜσzz (Ω) it is again linear in Ω for J = 1 but for J = 2
where the no tilt law applies. This is a clear signature it is instead constant43 independent of Ω. A finite tilt
of broken inversion symmetry. The growth of Q0 has introduces important modifications and these are quali-
provided the mechanism whereby the negative chirality tatively different for type I and type II. The boundary
node is separately revealed as is its characteristic slope. between these two types corresponds to the case where
A similar situation holds for the case J = 2 shown in the Dirac cone is completely tipped over and electron-
the bottom frame of Fig.[8]. In the black curve both hole pockets begin to form at charge neutrality. Tilting
chirality nodes contribute equally. In the dot double dash transfers optical spectral weight from above the Ω = 2µ
magenta curve for Q0 /µ = 0.9 the two plateaus of height threshold to below and the sharp absorption edge of the
π/8 and π/4 respectively are clearly seen. For type II no tilt case is lost. For type I Weyl the changes are con-
Weyl Fig.[9] applies and in that instance the tilt provides fined to the photon range 2µ/(1 + C2′ ) to 2µ/(1 − C2′ ). In
changes even for Ω/µ = 6 where the slope of the high this range the absorption is roughly quasilinear and its
energy quasilinear behavior in the top frame is now less value at Ω = 2µ is exactly half of its no tilt magnitude.
than 2/3 and the various curves do not quit merge. In the Above 2µ/(1 − C2′ ), ℜσxx (Ω) is linear in Ω for both wind-
lower frame two plateaus are seen but both are reduced ing number J = 1 and 2 as is ℜσzz (Ω) for J = 1. For
from their magnitude in Fig.[8]. J = 2, ℜσzz (Ω) is instead constant. For type II Weyl,
the changes due to tilting are more pronounced and they
extend to large values of Ω where the absorption never
VI. SUMMARY AND CONCLUSION returns to its zero tilt value.
For noncentrosymmetric multi-Weyl semimetals the
Within the Kubo formulation for the AC optical con- nodes are no longer at the same energy and effectively
ductivity of a multi-Weyl semimetal we derive analytic each Weyl cone has a different value of chemical poten-
algebraic equations for its absorptive part both perpen- tial (µs′ = µ + s′ Q0 ) with s′ = ± (positive/negative)
dicular ℜσxx (Ω) and parallel ℜσzz (Ω) to the symmetry chirality. For the no tilt case there is now two sharp
axis as a function of photon energy Ω. Our starting absorption edges and a region of photon energy emerges
Hamiltonian contains a multi-Weyl relativistic term with between the two absorption thresholds for which only
winding number J as well as a term that explicitly breaks the negative chirality node contributes. The magnitude
time reversal and another inversion symmetry. An arbi- of the energy shift Q0 controls the size of this region in
trary tilt of the Weyl cones covering both type I and II is photon energy. For finite tilt both absorption edges are
included as well as finite doping away from charge neu- modified as we previously described for the Q0 = 0 case.
trality. Breaking time reversal symmetry splits a doubly The positive chirality node contributes only for photon
degenerate Dirac point into two Weyl nodes of opposite energy greater than 2µ(1 + Q0 /µ)/(1 + C2′ ). For type I
chirality which are displaced in momentum space by ±Q. Weyl this can be larger than the value of Ω at which the
Breaking inversion symmetry further splits the nodes in negative chirality node has recovered its no tilt behavior.
energy by ±Q0 . In the clean limit employed in this work This occurs from Q0 /µ > C2′ . In this situation there is
11

an interval of photon energy where not only is the ab- the magnitude of the constant background for ℜσzz (Ω)
sorption due only to the negative chirality node but it with J = 2.
also takes on its no tilt behavior. For ℜσxx (Ω) this is
e2 Ω e2 Ω 2
8π 3vz J, for ℜσzz (Ω) with J = 1 it is 8π 3vz v0 while with
2
J = 2 it is 8πe
k0 v0 . Above 2µ(1 + Q0 /µ)/(1 − C2′ ) for Acknowledgments
type I, both Weyl nodes contribute equally to the con-
ductivity and its value is exactly twice the value quoted Work supported in part by the Natural Sci-
above. For type II in the limit of very large tilt and Ω ences and Engineering Research Council of Canada
also large, the previous laws still hold but the numerical (NSERC)(Canada) and by the Canadian Institute for
factors are changed and there is an extra factor of 1/C2′ Advanced Research (CIFAR)(Canada). We thanks
which reduces the slopes to zero for the linear laws and A.A.Burkov and D. Xiao for enlightening discussions.

1
X. Wan, A. M. Turner, A. Vishwanath, and S. Y. Savrasov, Nature 527, 495 (2015).
11
“Topological semimetal and Fermi-arc surface states in the L. Huang, T. M. McCormick, M. Ochi, Z. Zhao, M. -
electronic structure of pyrochlore iridates,” Phys. Rev. B T. Suzuki, R. Arita, Y. Wu, D. Mou, H. Cao, J. Yan,
83, 205101 (2011). N. Trivedi, and A. Kaminski,“Spectroscopic evidence for
2
H. Weng, C. Fang, Z. Fang, B. A. Bernevig, and a type II Weyl semimetallic state in MoTe2 ,” Nat.Mater
X. Dai, “Weyl Semimetal Phase in Noncentrosymmet- 15,1155 (2016).
12
ric Transition-Metal Monophosphides,” Phys. Rev. X 5, F. Y. Bruno, A. Tamai, Q. S. Wu, I. Cucchi, C. Bar-
011029 (2015). reteau, A. de la Torre, S. McKeownWalker, S. Riccò, Z.
3
B.Q. Lv, H.M. Weng, B.B. Fu, X.P. Wang, H. Miao, J. Ma, Wang, T. K. Kim, M. Hoesch, M. Shi, N. C. Plumb, E.
P. Richard, X.C. Huang, L.X. Zhao, G.F. Chen, Z. Fang, Giannini, A. A. Soluyanov, and F. Baumberger, “Obser-
X. Dai, T. Qian, and H. Ding, “Experimental Discovery of vation of large topologically trivial Fermi arcs in the can-
Weyl Semimetal TaAs”, Phys. Rev. X 5, 031013 (2015). didate type-II Weyl semimetal WTe2 ,” Phys. Rev. B 94,
4
S. -Y. Xu, I. Belopolski1, N. Alidoust, M. Neupane, G. 121112(R)(2016).
13
Bian, C. Zhang, R. Sankar, G. Chang, Z. Yuan, C. -C. S. Y. Xu, N. Alidoust, G. Chang, H. Lu, B. Singh, I. Be-
Lee, S. -M. Huang, H. Zheng, J. Ma, D. S. Sanchez, B. lopolski, D. Sanchez, X. Zhang, G. Bian, H. Zheng, M. A.
Wang, A. Bansil, F. Chou, P. P. Shibayev, H. Lin, S. Jia, Husanu, Y. Bian, S. M. Huang, C. H. Hsu, T. R. Chang, H.
and M. Z. Hasan, “Discovery of a Weyl fermion semimetal T. Jeng, A. Bansil, V. N. Strocov, S. Jia, and M. Z. Hasan,
and topological Fermi arcs”, Science 349, 613 (2015). “Discovery of Lorentz-violating Weyl fermion semimetal
5
B. Q. Lv, N. Xu, H. M. Weng, J. Z. Ma, P. Richard, X. state in LaAlGe materials,”arXiv:1603.07318(2016).
14
C. Huang, L. X. Zhao, G. F. Chen, C. E. Matt, F. Bisti, K. Deng, G. Wan, P. Deng, K. Zhang, S. Ding, E. Wang,
V. N. Strocov, J. Mesot, Z. Fang, X. Dai, T. Qian, M. M. Yan, H. Huang, H. Zhang, Z. Xu, J. Denlinger, A. Fe-
Shi, and H. Ding, “Observation of Weyl nodes in TaAs”, dorov, H. Yang, W. Duan, H. Yao, Y. Wu, S. Fan, H.
Nature Phys. 11, 724 (2015). Zhang, X. Chen, and S. Zhou, “Experimental observa-
6
L. X. Yang, Z. K. Liu, Y. Sun, H. Peng, H. F. Yang, T. tion of topological Fermi arcs in type-II Weyl semimetal
Zhang, B. Zhou, Y. Zhang, Y. F. Guo, M. Rahn, D. Prab- MoTe2 ,” Nat. Phys.12, 1105 (2016).
15
hakaran, Z. Hussain, S. -K. Mo, C. Felser, B. Yan, and Y. A. Liang, J. Huang, S. Nie, Y. Ding, Q. Gao, C. Hu, S. He,
N. Chen,“Weyl Semimetal phase in non-Centrosymmetric Y.Zhang, C. Wang, B. Shen, J. Liu, P. Ai, L. Yu, X. Sun,
compound TaAs”, Nature. Phys. 11, 728 (2015). W. Zhao, S. Lv, D. Liu, C. Li, Y. Zhang, Y. Hu, Y. Xu, L.
7
N. Xu, H. M. Weng, B. Q. Lv, C. E. Matt, J. Park, F. Zhao, G. Liu, Z. Mao, X. Jia, F. Zhang, S. Zhang, F. Yang,
Bisti, V. N. Strocov, D. Gawryluk, E. Pomjakushina, K. Z. Wang, Q. Peng, H. Weng, X. Dai, Z. Fang, Z. Xu, C.
Conder, N. C. Plumb, M. Radovic, G. Autès, O. V. Yazyev, Chen, and X. J. Zhou,“Electronic evidence for type II Weyl
Z. Fang, X. Dai, T. Qian, J. Mesot, H. Ding, and M. semimetal state in MoTe2 ,”arXiv:1604.01706(2016).
16
Shi,“Observation of Weyl nodes and Fermi arcs in tanta- J. Jiang, Z. K. Liu, Y. Sun, H. F. Yang, R. Rajamathi, Y.
lum phosphide”, Nature Communications 7, 11006 (2016). P. Qi, L. X. Yang, C. Chen, H. Peng, C. -C. Hwang, S. Z.
8
S. -Y. Xu, N. Alidoust, I. Belopolski, Z. Yuan, G. Bian, Sun, S. -K. Mo, I. Vobornik, J. Fujii, S. S. P. Parkin, C.
T. -R. Chang, H. Zheng, V. N. Strocov, D. S. Sanchez, G. Felser, B. H. Yan, and Y. L. Chen, “Signature of type-II
Chang, C. Zhang, D. Mou, Y. Wu, L. Huang, C. -C Lee, S. Weyl semimetal phase in MoTe2 ,”Nat. Commun.8, 13973
-M. Huang, B. Wang, A. Bansil, H. -T. Jeng, T. Neupert, (2017).
17
A. Kaminski, H. Lin, S. Jia, and M. Z. Hasan, “Discov- N. Xu, Z. J. Wang, A. P. Weber, A. Magrez, P. Bugnon,
ery of a Weyl fermion state with Fermi arcs in niobium H. Berger, C. E. Matt, J. Z. Ma, B. B. Fu, B. Q. Lv, N.
arsenide”,Nature Phys. 11, 748 (2015). C. Plumb, M. Radovic, E. Pomjakushina, K. Conder, T.
9
S. Borisenko, D. Evtushinsky, Q. Gibson, A. Yaresko, Qian, J. H. Dil, J. Mesot, H. Ding, and M. Shi, “Discovery
T. Kim, M. N. Ali, B. Buechner, M. Hoesch, and R. J. of Weyl semimetal state violating Lorentz invariance in
Cava,“Time-Reversal Symmetry Breaking Type-II Weyl MoTe2 ,”arXiv:1604.02116 (2016)
18
State in YbMnBi2”, arXiv:1507.04847 (2015). Y. Wu, D. Mou, N. H. Jo, K. Sun, L. Huang, S. L.
10
A. A. Soluyanov, D. Gresch, Z. Wang, Q. Wu, M. Troyer, Bud’ko, P. C. Canfield, and A. Kaminski, “Observation
X. Dai, and B. A. Bernevig, “Type-II Weyl semimetals”, of Fermi arcs in type-II Weyl semimetal candidate WTe2 ,”
12

35
Phys. Rev. B 94, 121113(R) (2016). S. P. Mukherjee, and J. P. Carbotte, “Absorption of circu-
19
S. Khim, K. Koepernik, D. V. Efremov, J. Klotz, T. lar polarized light in tilted Type-I and II Weyl semimetals
Förster, J. Wosnitza, M. I. Sturza, S. Wurmehl, C. Hess, ,” Phys. Rev. B 96, 085114 (2017).
36
J. van den Brink, and B. Büchner, “Magnetotransport J.F. Steiner, A.V. Andreev, and D.A. Pesin, “Anomalous
and de Haasvan Alphen measurements in the type-II Weyl Hall Effect in Type-I Weyl Metals,”Phys. Rev. Lett. 119,
semimetal TaIrTe4 ,”Phys. Rev. B 94, 165145 (2016). 036601 (2017).
20 37
K. Koepernik, D. Kasinathan, D. V. Efremov, S. Khim, S. R. Y. Chen, S. J. Zhang, J. A. Schneeloch, C. Zhang, Q.
Borisenko, B. Büchner, and J. van den Brink, “TaIrTe4 : A Li, G. D. Gu, and N. L. Wang, “Optical spectroscopy
ternary type-II Weyl semimetal,” Phys. Rev. B 93, 201101 study of the three-dimensional Dirac semimetal ZrTe5 ”,
(2016). Phys. Rev. B 92, 075107 (2015).
21 38
E. Haubold, K. Koepernik, D. Efremov, S. Khim, A. Fe- A. B. Sushkov, J. B. Hofmann, G. S. Jenkins, J. Ishikawa,
dorov, Y. Kushnirenko, J. van den Brink, S. Wurmehl, B. S. Nakatsuji, S. Das Sarma, and H. D. Drew, “Optical evi-
Buchner, T. K. Kim, M. Hoesch, K. Sumida, K. Taguchi, dence for a Weyl semimetal state in pyrochlore Eu2 Ir2 O7 ”,
T. Yoshikawa, A. Kimura, T. Okuda, and S. V. Borisenko, Phys. Rev. B 92, 241108(R) (2015).
39
“Experimental realization of type-II Weyl state in non- B. Xu, Y. M. Dai, L. X. Zhao, K. Wang, R. Yang, W.
centrosymmetric TaIrTe4 ,” Phys. Rev. B 95, 241108(R) Zhang, J. Y. Liu, H. Xiao, G. F. Chen, A. J. Taylor, D.
(2017). A. Yarotski, R. P. Prasankumar, and X. G. Qiu, “Optical
22
G. Autès, D. Gresch, M. Troyer, A. A. Soluyanov, and spectroscopy of the Weyl semimetal TaAs”, Phys. Rev. B
O. V. Yazyev, “Robust Type-II Weyl Semimetal Phase 93, 121110(R) (2016).
40
in Transition Metal Diphosphides XP2 (X=Mo, W),” D. Neubauer, J. P. Carbotte, A. A. Nateprov, A. Löhle,
Phys. Rev. Lett. 117, 066402 (2016). M. Dressel, and A. V. Pronin, “Interband optical con-
23
I. Belopolski, S.-Y. Xu, Y. Ishida, X. Pan, P. Yu, D. S. ductivity of the [001]-oriented Dirac semimetal Cd3 As2 ”,
Sanchez, H. Zheng, M. Neupane, N. Alidoust, G. Chang, Phys. Rev. B 93, 121202(R) (2016).
41
T.-R. Chang, Y. Wu , G. Bian, S. -M. Huang, C. -C. Lee, T. Timusk, J. P. Carbotte, C. C. Homes, D. N. Basov, and
D. Mou, L. Huang, Y. Song, B. Wang, G. Wang, Y. -W. S. G. Sharapov, “Three-dimensional Dirac fermions in qua-
Yeh, N. Yao, J. E. Rault, P. LeFèvre, F. Bertran, H. -T. sicrystals as seen via optical conductivity”, Phys. Rev. B
Jeng, T. Kondo, A. Kaminski, H. Lin, Z. Liu, F. Song, S. 87, 235121 (2013).
42
Shin, and M. Z. Hasan,“Fermi arc electronic structure and M. Chinotti, A. Pal, W. J. Ren, C. Petrovic, and L.
Chern numbers in the type-II Weyl semimetal candidate Degiorgi, “Electrodynamic response of the type-II Weyl
Mox W1−x Te2 ,” Phys. Rev. B 94, 085127 (2016). semimetal YbMnBi2 ”,Phys. Rev. B 94, 245101 (2016).
24 43
F. -Y. Li, X. Luo, X. Dai, Y. Yu, F. Zhang, and S. Ahn, E. J. Mele, and H. Min, “Optical conductivity
G. Chen, “Hybrid Weyl semimetal,” Phys. Rev. B 94, of multi-Weyl semimetals,” Phys. Rev. B 95, 161112(R)
121105(R)(2016). (2017).
25 44
G. Xu, H. Weng, Z. Wang, X. Dai, and Z. Fang,“Chern H. -R. Chang, J. Zhou, S. -X. Wang, W. -Y. Shan, and Di
Semimetal and the Quantized Anomalous Hall Effect in Xiao, “RKKY interaction of magnetic impurities in Dirac
HgCr2 Se4 ,” Phys. Rev. Lett. 107, 186806 (2011). and Weyl semimetals,” Phys. Rev. B 92, 241103(R) (2015).
26 45
C. Fang, M. J. Gilbert, X. Dai, and B. A. Bernevig, “Multi- A. A. Zyuzin, S. Wu, and A. A. Burkov,“Weyl semimetal
Weyl Topological Semimetals Stabilized by Point Group with broken time reversal and inversion symmetries,”
Symmetry,” Phys. Rev. Lett. 108, 266802 (2012). Phys. Rev. B 85, 165110 (2012).
27 46
S. -K. Jian and H. Yao, “Correlated double-Weyl semimet- S. Park, S. Woo, E. J. Mele, and H. Min, “Semiclassical
als with Coulomb interactions:Possible applications to Boltzmann transport theory for multi-Weyl semimetals,”
HgCr2 Se4 and SrSi2 ,” Phys. Rev. B 92, 045121 (2015). Phys. Rev. B 95, 161113(R) (2017).
28 47
Z. -M. Huang, J. Zhou, and S. -Q. Shen, “Topo- Y. Sun and A. -M. Wang, “Magneto-optical conductivity of
logical responses from chiral anomaly in multi-Weyl double Weyl semimetals,” Phys. Rev. B 96, 085147 (2017).
48
semimetal,”Phys. Rev. B 96, 085201 (2017). P. E. C. Ashby and J. P. Carbotte, “Magneto-optical con-
29
Y. Sun and A. Wang, “RKKY interaction of ductivity of Weyl semimetals,” Phys. Rev. B 87, 245131
magnetic impurities in multi-Weyl semimetals,” (2013).
49
J. Phys.:Condens. Matter.29, 435306 (2017). A.A. Burkov, “Chiral anomaly and transport in Weyl met-
30
X. Dai, H. -Z. Lu, S. -Q. Shen, and H. Yao, “Detecting als,” J. Phys.:Condens. Matter.27, 113201 (2015).
50
monopole charge in Weyl semimetals via quantum inter- A.A. Burkov, “Anomalous Hall Effect in Weyl Metals,”
ference transport,” Phys. Rev. B 93, 161110(R) (2016). Phys. Rev. Lett. 113, 187202 (2014).
31 51
C. J. Tabert, J. P. Carbotte, and E. J. Nicol,“Optical and A. A. Zyuzin and R. P. Tiwari, “Intrinsic anomalous Hall
transport properties in three-dimensional Dirac and Weyl effect in type-II Weyl semimetals”, JETP Lett.103, 717
semimetals,” Phys. Rev. B 93, 085426 (2016). (2016).
32 52
C. J. Tabert, and J. P. Carbotte,“Optical conductivity of Y. Ferreiros, A. A. Zyuzin, J. H. Bardarson, “Anomalous
Weyl semimetals and signatures of the gapped semimetal Nernst and Thermal Hall Effects in Tilted Weyl Semimet-
phase transition,” Phys. Rev. B 93, 085442 (2016). als”, Phys. Rev. B 96, 115202 (2017).
33 53
J. P. Carbotte,“Dirac cone tilt on interband optical S. Saha, and S. Tewari, “Anomalous Nernst effect in type-
background of type-I and type-II Weyl semimetals”, II Weyl semimetals”, Eur. Phys. J. B 91, 4 (2018).
54
Phys. Rev. B 94,165111 (2016). V. P. Gusynin, S. G. Sharapov, and J. P. Carbotte, “On
34
S. P. Mukherjee and J. P. Carbotte, “Optical response the universal ac optical background in graphene”, New J.
in Weyl semimetal in model with gapped Dirac phase ,” Phys.11, 095013 (2009).
J. Phys.:Condens. Matter.29, 425301 (2017).
13

Appendix A

Following Eq.(9) the real part of σxx (Ω) is written as,


−(2J−1) 2 X Z Λ−s′ Q Z ∞ 2J−1
e2 J 2 k0 v⊥ k⊥ dk⊥
ℜσxx (Ω) = − dkz {f (Cs′ kz + k0 k − µs′ ) − f (Cs′ kz − k0 k − µs′ )} ×
2πΩ ′ −Λ−s′ Q 0 k
s =±

(k02 k 2 2 2 2 2 2
+ vz kz )δ(4k0 k − Ω )
2 2 −2J 2 X Z Λ−s Q Z ∞

2J−1
e J k0 v⊥ k⊥ dk⊥
= − dkz {f (Cs′ kz + k0 k − µs′ ) − f (Cs′ kz − k0 k − µs′ )} ×
8πΩ2 ′ k
s′ =± −Λ−s Q 0
 
2 2 2 2 Ω Ω
(k0 k + vz kz ) δ(k − ) + δ(k + )
2k0 2k0
Z ′ Z ∞
e2 J X Λ−s Q Ω
=− 2
dk z dk{f (Cs′ kz + k0 k − µs′ ) − f (Cs′ kz − k0 k − µs′ )}(k02 k 2 + vz2 kz2 )δ(k − )
8πΩ ′ −Λ−s′ Q vz | kz
k |
2k0
s =± 0

Taking the cut off Λ to be much larger than the momentum space separation between the Weyl nodes and also
much larger than Ω/2k0 we get,
Z 2kΩ  
e2 Jk0 X 0 Cs′ k0 Ω Cs′ k0 Ω
=− dkz f ( kz + − µs′ ) − f ( kz − − µs′ ) (Ω2 + 4k02 kz2 ) (A1)
32πvz Ω2 ′ Ω
− 2k vz 2 vz 2
s =± 0

We note that the variable Q has entirely dropped out of this quantity. Here we can drop the second Dirac delta
function as it clicks at k = − 2kΩ0 which is outside the range of integration. In the second step we have changed the
variable k⊥ to k as shown below,
k 2 = v⊥
2
(k⊥ /k0 )2J + vz2 (kz /k0 )2
2
Jv⊥
kdk = (k⊥ /k0 )2J−1 dk⊥
k0
(A2)
In the limit T → 0 the Fermi functions become theta functions. This gives,
Z Ω  
e2 Jvz2 X 2vz Ω Ω Ω 2 Ω2
ℜσxx (Ω) = − dk z Θ(Cs′ kz − − µ s′ ) − Θ(Cs′ kz + − µ s′ ) − Θ(−Cs′ kz + − µ s′ ) (k +
z )(A3)
8πΩ2 ′ 0 2 2 2 4vz2
s =±

We see that the above expression for ℜσxx (Ω) is tilt-inversion symmetric i.e. if we change Cs′ to −Cs′ then ℜσxx (Ω)
stays same. It only depends on the absolute value of the tilts in two Weyl nodes irrespective of whether the tilt is
clockwise or anticlockwise. The rest of the calculation is straight forward and we state the final result in the main
text in Sec.III.

Appendix B

Here we derive the result for the conductivity σzz (Ω). Following Eq.(10)the real part of σzz (Ω) is written as,
e2 X Z Λ−s′ Q Z ∞
k⊥ dk⊥
ℜσzz (Ω) = − dkz {f (Cs′ kz + k0 k − µs′ ) − f (Cs′ kz − k0 k − µs′ )} ×
8πk02 Ω2 ′ ′ k
s =± −Λ−s Q 0
h  2 2 i Ω Ω

2 2 2 2 2 2 2 2J −2(J−1)
(Cs′ + vz ) k0 k − vz kz − v⊥ (Cs′ − vz )k⊥ k0 δ(k − ) + δ(k + )
2k0 2k0
e2 k02 vz2 X Z Λ−s′ Q Z ∞
1
=− 2 2 1 dk z dk {f (Cs′ kz + k0 k − µs′ ) − f (Cs′ kz − k0 k − µs′ )} k02 k 2 − vz2 kz2 J ×
4πJ(k0 v⊥ ) J Ω2 s′ =± −Λ−s′ Q vz | kz
k0 |


δ(k − )
2k0
XZ Ω  
e2 k02 vz2 2vz Ω Ω 1
=− 1+J 1
dkz f (Cs′ kz + − µs′ ) − f (Cs′ kz − − µs′ ) Ω2 − 4vz2 kz2 J (B1)
4 J πJ(k02 v⊥
2 ) J Ω2 Ω
− 2v 2 2
s′ =± z
14

At zero temperature limit we get,

XZ Ω  
e2 k02 vz2 2vz Ω Ω Ω
ℜσzz (Ω) = − 1+J 1
dkz Θ(C kz − − µs ) − Θ(Cs kz + − µs ) − Θ(−Cs kz + − µs )
s′ ′ ′ ′ ′ ′
4 J πJ(k02 v⊥
2 ) J Ω2
0 2 2 2
s′ =±
1
× Ω2 − 4vz2 kz 2 J
(B2)

You might also like