<<

PHYS 90507. Topology and Dirac in condensed matter Assaf

Section 1: A. The Dirac Hamiltonian in 2D Starting from the time dependent : 휕 퐻 = 푖ℏ 훽 − 푐(푝⃗. 훾⃗) + 푚푐2퐼 푡 휕푡 4

Here m is the ‘rest-mass’, c the velocity, 푝⃗ = 푝푥푥̂ + 푝푦푦̂ is the momentum in 2D and 훾⃗ = 훾푥푥̂ + 퐼 0 훾 푦̂ + 훾 푧̂ are the Dirac gamma (γ) matrices. 훽 = ( ) where I is the 2 identity matrix. And 퐼 = 푦 푧 0 −퐼 4 퐼 0 ( ) 0 퐼 Keeping in mind the γ-matrices in the Dirac representation:

0 휎푥,푦,푧 훾푥,푦,푧 = ( ) −휎푥,푦,푧 0 0 1 0 −푖 1 0 With, 휎 = ( ), 휎 = ( ), 휎 = ( ) 푥 1 0 푦 푖 0 푧 0 −1 −푖휔푡 We will solve 퐻푡Ψ = 0 for Ψ = ϕe with ℏ휔 = 퐸 We get:

2 (퐸 + 푚푐 )퐼 −푐푝⃗. 휎⃗ 휙1 ( 2 ) ( ) = 0 푐푝⃗. 휎⃗ (−퐸 + 푚푐 )퐼 휙2

2 (퐸 + 푚푐 )퐼휙1 − 푐푝⃗. 휎⃗휙2 = 0 Or equivalently, { 2 (−퐸 + 푚푐 )퐼휙2 + 푐푝⃗. 휎⃗휙1 = 0 This can be rearranged into:

2 (−퐸 − 푚푐 )퐼휙1 + 푐푝⃗. 휎⃗휙2 = 0 { 2 (−퐸 + 푚푐 )퐼휙2 + 푐푝⃗. 휎⃗휙1 = 0 To finally get the equation in a slightly different form and a different basis:

2 (−퐸 + 푚푐 )퐼 푐푝⃗. 휎⃗ 휙2 ( 2 ) ( ) = 0 푐푝⃗. 휎⃗ (−퐸 − 푚푐 )퐼 휙1 From this we can define the time-independent Dirac Hamiltonian to be:

퐻 = 푐(푝⃗. 훼⃗) + 푚푐2훽 With the Dirac α matrices given by:

0 휎푥,푦,푧 훼푥,푦,푧 = ( ) 휎푥,푦,푧 0 Since we are interested in the problem in 2D, we can drop out z terms to get the following matrix Hamiltonian:

푚푐2 0 0 푐(푝푥 − 푖푝푦) 2 0 푚푐 푐(푝푥 + 푖푝푦) 0 퐻 = 0 푐(푝푥 − 푖푝푦) −푚푐2 0 푐(푝 + 푖푝 ) 0 2 ( 푥 푦 0 −푚푐 )

1

PHYS 90507. Topology and Dirac fermions in condensed matter Assaf

For a particle with spin-1/2, the wavefunction decomposed into a position dependent part and a , takes the form:

푎1 푖 푎 푝⃗.푟⃗ 휓(푟⃗) = ( 2) 푒ℏ 푎3 푎4 So that we have the following eigenvalue problem that results from the Dirac equation 퐻휓= 퐸휓 can be written as:

2 0 푐(푝 − 푖푝 ) 푚푐 − 퐸 0 푥 푦 푎1 2 0 푚푐 − 퐸 푐(푝푥 + 푖푝푦) 0 푎2 (푎 ) = 0 0 푐(푝푥 − 푖푝푦) −푚푐2 − 퐸 0 3 푐(푝 + 푖푝 ) 0 2 푎4 ( 푥 푦 0 −푚푐 − 퐸) Eigenvalue It is easy to see that the eigenvalue problem decouples into two independent determinant equations so that we have:

2 푚푐 − 퐸 푐(푝푥 − 푖푝푦) | 2 | = 0 푐(푝푥 + 푖푝푦) −푚푐 − 퐸 Or

± 2 4 2 2 2 2 4 2 2 퐸 = ±√푚 푐 + 푐 (푝푥 + 푝푦) = ±√푚 푐 + 푐 푝

In terms of the wavevector k:

± 2 4 2 2 2 퐸 = ±√푚 푐 + (ℏ푐) (푘푥 + 푘푦)

2 Let us define Δ ≡ 푚푐 and try to plot E(px,py) for different characteristic values of Δ while keeping c constant. For simplicity let us use the following parameters: Δ = 0meV, 10meV and 100meV for Fig. 1(a-c) respectively, and ℏ푐 = 6푒푉Å.

FIG 1. (a-c) Three Dirac energy-momentum spectra with Δ = 0meV, 10meV and 100meV respectively and ℏ푐 = 6푒푉Å.

2

PHYS 90507. Topology and Dirac fermions in condensed matter Assaf

The solid-state physicist’s take on this result: While in high-energy physics the Dirac equation first described the behavior of free electrons and their the positron that has an opposite energy, in a solid, the Dirac equation can describe a semiconductor with an energy gap equal to 2Δ and a band structure that is non-parabolic away from the band gap:

Since 퐸(푝) = ±√푚2푐4 + 푐2푝2

퐸±(0) = ±푚2푐4

Therefore, the energy gap, 퐸+(0) − 퐸−(0) = 2푚2푐4 = 2Δ Limiting case 1:

Also notice the limiting case, for 푚2푐4 ≪ 푐2푝2, 퐸 → 푐푝. The bands are linear at large p. Limiting case 2:

푝2 푝2 When 푚2푐4 ≫ 푐2푝2, 퐸(푝) = ±푚2푐4√1 + = ±푚2푐4 (1 + ). The bands are parabolic at 푚2푐2 2푚2푐2 small p. In assignment 1, you will also show that the ‘effective mass’ of Dirac particles is energy dependent. Limiting case 3: Using ℏ푐푘 = 푐푝

푑퐸 ℏ푐2푘 푣푓 = = ℏ푑푘 √푚2푐4 + (ℏ푐푘)2

When 푘 → ∞

ℏ푐2 푣푓 = → 푐 푚2푐4 √ + (ℏ푐)2 푘2

Eigenvector Let us now proceed and extract the eigenvectors resulting from the Dirac eigenvalue problem.

2 ± 0 푐(푝 − 푖푝 ) 푚푐 − 퐸 0 푥 푦 푎1 2 ± 0 푚푐 − 퐸 푐(푝푥 + 푖푝푦) 0 푎2 (푎 ) = 0 0 푐(푝푥 − 푖푝푦) −푚푐2 − 퐸± 0 3 푐(푝 + 푖푝 ) 0 2 ± 푎4 ( 푥 푦 0 −푚푐 − 퐸 ) The problem decoupled into two independent 2x2 eigenvector equations so that for each energy we get two independent 2x1 eigen-:

2 ± (푚푐 − 퐸 )푎1 + 푐(푝푥 − 푖푝푦)푎4 = 0

2 ± (푚푐 − 퐸 )푎2 + 푐(푝푥 + 푖푝푦)푎3 = 0

2 ± 푐(푝푥 − 푖푝푦)푎2 − (푚푐 + 퐸 )푎3 = 0

2 ± 푐(푝푥 + 푖푝푦)푎1 − (푚푐 + 퐸 )푎4 = 0

3

PHYS 90507. Topology and Dirac fermions in condensed matter Assaf

Or equivalently:

푐(푝푥 + 푖푝푦) 푐(푝푥 − 푖푝푦) 푎 = 푎 푎푛푑 푎 = 푎 4 (푚푐2 + 퐸±) 1 3 (푚푐2 − 퐸±) 2 So that we have two possible eigenvector solutions given by:

1 0 0 1 휙1 = 0 and 휙2 = 푐(푝푥−푖푝푦) 푐(푝푥+푖푝푦) (푚푐2−퐸±) 2 ± ((푚푐 +퐸 )) ( 0 ) Normalized;

1 0 (푚푐2+퐸±)2 0 (푚푐2+퐸±)2 1 휙 = √ and 휙 = √ 푐(푝푥−푖푝푦) 1 (푚푐2+퐸±)2+(푐푝)2 0 2 (푚푐2−퐸±)2+(푐푝)2 푐(푝푥+푖푝푦) (푚푐2−퐸±) 2 ± ((푚푐 +퐸 )) ( 0 ) We have 2 solutions for each energy eigenvalue. Section 1: B. Weyl fermions Going back to Klein-Gordon Hamiltonian (in 3D) we have:

ℏ2휕2 [− + ℏ2∇2] Ψ = 푚푐2Ψ 푐2휕푡2 In a seminal paper in 1929 Hermann Weyl noticed a special solution to the Dirac equation when m=0. In this case, the Hamiltonian above reduced to:

ℏ2휕2 [− + ℏ2∇2] Ψ = 0 푐2휕푡2

So that if Ψ = 휓e−푖휔푡, we have,

[퐸2 + 푐2ℏ2∇2]Ψ = 0

2 Keeping in mind the properties of the Pauli matrices, in particular that σi =I, we can write this as:

퐸Ψ = ±푖푐ℏ∇⃗⃗⃗. 휎⃗ Or, Using the Weyl representation for the 훾 matrices, one can decouple the system to get two that would have the form.

±푐(푝⃗. 휎⃗)휓 = 퐸휓 We’re going to call these two equations the Weyl equations.

Let’s work out the Weyl problem in the 3D case. Starting from the 3D (퐻푊), and assuming the wavefunction is a planewave that multiplies a spinor we get: − 퐸 − 푐푝푧 −푐푝 푎1 ( + ) ( ) = 0 −푐푝 퐸 + 푐푝푧 푎2

4

PHYS 90507. Topology and Dirac fermions in condensed matter Assaf

− 퐸 + 푐푝푧 푐푝 푎3 ( + ) ( ) = 0 푐푝 퐸 − 푐푝푧 푎4 A small parenthesis: if we were two consolidate these two equations in a single 4x4 system we get the Weyl system: − 0 0 퐸 − 푐푝푧 −푐푝 푎1 + 0 0 −푐푝 퐸 + 푐푝푧 푎2 ( − ) ( ) = 0 퐸 + 푐푝푧 푐푝 0 0 푎3 + 푎 푐푝 퐸 − 푐푝푧 0 0 4 The problem remains decoupled of course, but we get a sense of why Weyl introduced a different basis with different α matrices (the Weyl representation), in which: 0 0 1 0 훽 = (0 0 0 1) 1 0 0 0 0 1 0 0 If we start from the conventional Dirac equation 퐻 = 푐(푝⃗. 훼⃗) + 푚푐2훽

Using the 훽 matrix given above, we recover the 4x4 Weyl system. Let us know go back to the 2x2 system and solve it: − 퐸 − 푐푝푧 −푐푝 푎1 ( + ) ( ) = 0 −푐푝 퐸 + 푐푝푧 푎2 − 퐸 + 푐푝푧 푐푝 푎3 ( + ) ( ) = 0 푐푝 퐸 − 푐푝푧 푎4 Eigenvalue The eigenvalue is simple to obtain in both cases:

2 2 2 2 So that, 퐸 − 푐 (푝푥 + 푖푝푦)(푝푥 − 푖푝푦) − 푐 푝푧 = 0 Or,

± 2 2 2 퐸 = ±푐√푝푥 + 푝푦 + 푝푧 = ±푐푝

We get a 3D massless Dirac cone with a linear dispersion in all momentum directions Eigenvector For the first system of equations:

퐸± − 푐푝 −푐푝− 푎 푧 1 ( + ± ) (푎 ) = 0 −푐푝 퐸 + 푐푝푧 2 We get:

± − (퐸 − 푐푝푧)푎1 − 푐푝 푎2 = 0 푝− ± 휓1 = (±푝 − 푝푧) 1

5

PHYS 90507. Topology and Dirac fermions in condensed matter Assaf

For the second system we get: − 퐸 + 푐푝푧 푐푝 푎3 ( + ) ( ) = 0 푐푝 퐸 − 푐푝푧 푎4 ± − (퐸 + 푐푝푧)푎3 + 푐푝 푎4 = 0 푝− ± − 휓2 = ( ±푝 + 푝푧) 1 It is convenient to work in 2D in order to get a sense of the implications of the Weyl treatment. Let us then set pz=0 without any loss of generality. We get the following normalized solutions: 푝− ± 1 ± 휓 = ( 푝 ) 1 2 √ 1 푝− ± 1 ∓ 휓 = ( 푝 ) 2 2 √ 1

Spin expectation value and helicity The handedness of the solutions is set by their helicity. The helicity can be computed by looking at the ⃗⃗ Σ. 푝⃗⁄ ⃗⃗ ⃗⃗ expectation value of the operator |푝| with Σ given by: Σ = 휎푥푥̂ + 휎푦푦̂. Before we examine helicity and its meaning, let us simply compute the spin (x, y) expectation values for each one of the E+ branch For the (1) eigenvector: 푝− 1 푝+ 0 1 〈휎 〉 = 〈휓+휎 휓+〉 = ( 1) ( ) ( 푝 ) 푥 1 푥 1 2 푝 1 0 1 풑+ + 풑− 풑 〈흈 〉 = = 풙 풙 ퟐ풑 풑

푝− 1 푝+ 0 −푖 〈휎 〉 = 〈휓+휎 휓+〉 = ( 1) ( ) ( 푝 ) 푦 1 푦 1 2 푝 푖 0 1 + − −풊풑 + 풊풑 풑풚 〈흈 〉 = = 풚 ퟐ풑 풑

For the (2) eigenvector: 푝− 1 푝+ 0 1 − 〈휎 〉 = 〈휓+휎 휓+〉 = (− 1) ( ) ( 푝 ) 푥 2 푥 2 2 푝 1 0 1

6

PHYS 90507. Topology and Dirac fermions in condensed matter Assaf

풑+ + 풑− 풑 〈흈 〉 = − = − 풙 풙 ퟐ풑 풑

푝+ 푝− 1 − 0 −푖 − 〈휎 〉 = 〈휓+휎 휓+〉 = ( ) ( ) ( 푝 ) 푦 2 푦 2 2 푝 푖 0 1 1

− + −풊풑 + 풊풑 풑풚 〈흈 〉 = = − 풚 ퟐ풑 풑

We can now take a look at the helicity of the spinor in each respective case. The helicity is the sum of the spin expectation values projected onto the momenta directions. It is thus a measure of the spin ‘handed-ness’ with respect to the momentum.

Σ⃗⃗. 푝⃗ 휎푥푝푥 + 휎푦푝푦 〈 〉 = 〈 〉 |푝| |푝|

+ For 휓1 : Σ⃗⃗. 푝⃗ 푝2 푝2 〈 〉 = 푥 + 푥 = +1 |푝| |푝|2 |푝|2

+ For 휓2 :

2 2 Σ⃗⃗. 푝⃗ 푝 푝푦 〈 〉 = − 푥 − = −1 |푝| |푝|2 |푝|2

+ + Thus, helicity defines the ‘winding’ direction of the spin expectation value. It is clear the 휓1 and 휓2 + wind in opposite directions, hence the term right-handed (R) solution to refer to 휓1 and left-handed (L) + solution to refer to 휓2 . This is illustrated in FIG. 3

FIG 3. Fermi surface contours showing spin angular momentum direction for the L- and R- handed solutions. Degeneracy and absence of spin-momentum locking Note that the solution ends up being a superposition of two massless Dirac cones (or Weyl cones) of opposite helicity, meaning the overall helicity of the fermions is zero. A recent major accomplishment in condensed-matter physics demonstrated that in certain materials having inversion asymmetry the two Weyl pairs can be separated in k-space yielding real helical Weyl fermions at different k-points. One

7

PHYS 90507. Topology and Dirac fermions in condensed matter Assaf can also think of the surface state of topological insulators as 2D Weyl fermions that are separated in real space.

8