Notes on Physics¶
Welcome! This is the place for me to write some notes on physics. Contents will be added indefinitely.
Classical Mechanics:¶
Phase space Lagrangian¶
Regular derivation¶
The climax part of classical mechanics lies in the Lagrangian and Hamiltonian form. It starts with the extremal principle, the real motion of a mechanical system is the one makes the variation of the action \(S\) vanish, i.e.,
When the Lagrangian is not depend on time explicitly, we get
which gives us the Lagrangian equation:
Define the canonical momentum \(p=\frac{\partial L}{\partial \dot{q}}\), we have
We can easily prove that energy is an integral of motion (which is a conserved quantity in motion) based on the homogeneity of time. When the Lagrangian does not depend on time explicitly, we found that its total derivative is
We have used Lagrangian equation in the above derivation,and we have now
indicating that \(H=p\dot{q}-L\) is conserved in the motion, which is called Hamiltonian with physical meaning of energy.
On the other hand, from:
which means \(dH\) is the total differential with respect to \(q\) and \(p\). And from:
we get
This is the Hamilton canonical equations.
The total derivative with time is:
where Hamilton’s equations are used. It also indicates that if H is not depend on time explicitly, we have the conservation of energy.
Note
From Lagragian to Hamiltonian, we did a Legendre transformation, makes the dependence \(L(q,\dot{q},t)\) to \(H(q,p,t)\), i.e., \(L\) is a function in configuration space \((q,\dot{q})\), but \(H\) is in phase space \((q,p)\).
Dynamic system¶
Mathematically, a continuous-time dynamical system is defined to be a system of first order differential equations
where \(\mathbf{f}\) is known as the vector field and \(\mathbb{R}\) is the set of real numbers. The space in which \(\mathbf{z}\) is defined is called phase space.
Lagrange’s equations do not form a dynamical system, because they implicitly contain second-order derivatives, \(\ddot{q}\). However, there is a standard way to obtain a system of first-order equations from a second-order system, which is to double the size of the space of time-dependent variables by treating the generalized velocities \(u\) as independent of the generalized coordinates, so that the dynamical system is \(\dot{q} = u, \dot{u} = \ddot{q}(q,u,t)\). Then the phase space is of dimension \(2n\). This trick is used very frequently in numerical problems, because the standard numerical integrators require the problem to be posed in terms of systems of first-order differential equations.
In the particular case of Lagrangian mechanics, expanding out \(\frac{d}{dt}\frac{\partial L}{\partial \dot{q}}\) using the chain rule and moving all but the highest-order time derivatives to the right-hand side,
The matrix \(H\) acting on \(\ddot{q}\), whose elements are given by \(H_{i,j}\equiv\frac{\partial^2 L}{\partial \dot{q}_i \partial \dot{q}_j}\), is called the Hessian matrix . It is a kind of generalized mass tensor, and for our method to work we require it to be nonsingular, so that its inverse, \(H^{-1}\), exists and we can find \(\ddot{q}\). Then our dynamical system becomes:
Momentum instead of velocity
We can achieve our aim of finding 2n first-order differential equations by using many choices of auxiliary variables other than \(u\). These will be more complicated functions of the generalized velocities, but the extra freedom of choice may also bring advantages.
In particular, Hamilton realized that it is very natural to use as the new auxiliary variables the set \(p=\left\{p_i|i=1,\cdots,n\right\}\) defined by
where \(p_i\) is called the canonical momentum conjugate to \(q_i\).
At this moment, we shall assume that the above equation can be solved to give \(\dot{q}\) as a function of \(q\) and \(p\)
The Lagrange’s equations immediately give us
The above two equations do indeed form a dynamical system, but so far it looks rather unsatisfactory: now \(u\) is defined only implicitly as a function of the phase-space variables \(q\) and \(p\), yet the right-hand side of above equation involves a partial derivative in which the \(q\)-dependence of \(u\) is ignored!
We can fix the latter problem by holding \(p\) fixed in partial derivatives with respect to \(q\) (because it is an independent phase-space variable) but then subtracting a correction term to cancel the contribution coming from the \(q\)-dependence of \(u\). Applying the chain rule, we get
where we have already defined the Hamiltonian \(H(q, p, t)= p\cdot u − L(q,u, t)\) is a function of \((q, p)\).
Given the importance of \(\partial H/ \partial q\) it is natural to investigate whether \(\partial H/\partial p\) plays a significant role as well. Differentiating \(H(q, p, t)= p\cdot u − L(q,u, t)\) we get
the above two equations are just the Hamilton’s equations we derived before.
Phase space Lagrangian¶
The equation \(H(q,p,t)\equiv p\cdot\dot{q}-L(q,\dot{q},t)\) suggest we define \(L_{ph}(q,\dot{q},p,t)\equiv p\cdot\dot{q}-H(q,p,t)\). If \(\dot{q}=u(q,p,t)\) were identically satisfied, even on arbitrarily varied phase-space paths, then \(L_{ph}\) would simply be \(L\) expressed in phase-space coordinates.
However, one can easily construct a counter example to show that this is not the case: consider a variation of the path in which we can vary the direction of its tangent vector, at some point \(z\equiv (q, p)\), while keeping this point fixed. Then \(\dot{q}\) changes, but \(u\) remains the same. Thus \(L_{ph}\) and \(L\) are the same value only on the subset of paths (which includes the physical paths) for which \(p\cdot\dot{q} = p\cdot u(q, p, t)\) .
Replacing \(L\) by \(L_{ph}\) in \(S = \int L dt\) we define the phase-space action integral
We know from variational calculus that \(S_{ph}\) is stationary under arbitrary variations of the phase-space path (with endpoints fixed), explicitly, we get:
which gives us the Hamilton’s equations:
Note
More generally, the phase-space Lagrangian can depend on \(\dot{p}\) as well, and makes the Lagrange’s equations become:
References
- L.D. Landau & E.M. Lifshitz, Mechanics.
- Lecture notes by Bob Dewar on website Classical Mechanics.
Topological Insulator:¶
Lecture notes¶
Here is the lecture notes on TI, which is partly based on the lecture notes given by Janos Asboth, Laszlo Oroszlany, and Andras Palyi, Here is the resource, and you can also download a copy here.
Lecture 1 : 1-d SSH model¶
The Su-Schrieffer–Heeger (SSH) model¶
The simplest non-trivial topology : 1-d lattice.
Peierls instability makes the atoms dimerize.
Polyacetylene Structure:
Tight-binding method: first quantization¶
Tight-binding method: Single electron total Hamiltonian in atom chain:
with periodical potential: \(U(x+na)=U(x)\)
Assume single atom potential \(V(x)\) with Hamiltonian
and well solved eigen-value and eigen-wave-function:(consider only one state)
Do the combination
with \(\phi_n=\phi(x-x_n), x_n=na\). Define \(\Delta U(x)=U(x)-V(x)\), and substitute following equation
we get
Define: \(\langle \phi(x-ma+na)|\Delta U(x) |\phi(x)\rangle=-J(x_m-x_n)\)
We get:
Because of the Tranformation symmetry of the Hamiltonian, the resulting wavefunction \(\psi(x)\) should take Bloch form, which means we should have the solution \(a_m=e^{ikx_m}\), then, we get
Consider only the nearest-hopping interaction, define \(J_0=J(0), J=J(\pm a)\), then we have:
Second quantization¶
In the second quantization language, the expectation value of energy becomes a operator, set \(\mathscr{H}=\frac{p^2}{2m}+U(x)\), we have
with \(\psi \to \hat{\psi}=\sum\limits_n \hat{c}_n \phi_n, H_{mn}=\langle \phi|\mathscr{H}|\phi \rangle\) \(\phi_n\) is a orthonormal and complete basis in Hilbert space, like plane-waves \(e^{ikx}\) or energy eigen-states of \(H_0\), \(\mathscr{H}\) is the energy operator in single particle first quantization picture, which can only act on Hilbert space, while the second quantization energy operator \(\hat{H}\) acts on Fock space. Here, in tight-binding method, \(\phi_n\) is the wave-function of site \(n\) of the energy eigen-state \(H_0\).
Consider only the nearest interaction, we have:
\(h.c.\) means hermitian conjugation. Rewrite it in matrix form: \(\hat{H}=\sum\limits_{mn}\hat{c}_m \tilde{H}_{mn}c_n\), we have
In the case when \(t=t_n\), \(\hat{c}_n\) satisfy the Bloch condition, we can transform it into momentum space, with \(\hat{c}_n=\frac{1}{\sqrt{M}}\sum\limits_k \hat{c}_k e^{ikx_n}\),
we can easily get
which gives us the dispersion relation:
On the other hand, keep in mind that we would get \(c_n=e^{ik(n-1)a}c_1\), that is
which also gives us \(E(k)=te^{ika}+t^*e^{-ika}\).
More generally, \(t_n\) can be different from each other, for example, if they are all different up to \(4\), but have a super-periodicity with \(t_{5}=t_1\), then there will have 4 sub-bands, in the example we will consider below, we have two \(t\), \(t_1\neq t_2\), and we have two sub-bands.
If each atom have a valance electron, then the above mentioned energy band structure \(E(k)=2tcoska\) is not the stable fundamental mode, it will dimerizes to lower the total energy, that means we’ll get following coupling case:
with the Hamiltonian:
with \(M=2N\).
Review¶
In 1-d atom chain, we have got the Hamiltonian in second-quantization frame as:
\(h.c.\) means hermitian conjugation. Rewrite it in matrix form: \(\hat{H}=\sum\limits_{mn}\hat{c}_m {H}_{mn}\hat{c}_n\), we have
Energy band and topology¶
\(t_n\) can differ from each other.
- For a open chain with \(M\) atoms, we have \(t_M=0\), and this matrix will give us \(M\) eigen-values and eigen-functions.
- Possess translational invariance with \(c_{n+1}=c_1e^{ikna}\), it will be diagonalized with \(H(k)=te^{ika}+t^*e^{-ika}\).
- Staggered hopping parameters with \(t_1\neq t_2\), but have property \(c_{2n+1}=c_1e^{iknb},c_{2n+2}=c_2e^{iknb}\). We can block the Hamiltonian up in \(2\times 2\) blocks and also pair up \(c_{2n+1},c_{2n+2}\).
The Hamiltonian:
with \(M=2N\).
For a more beautiful notation, define \(\mathbf{c}_n^\dagger=(c_{n,1}^\dagger,c_{n,2}^\dagger)=(c_{2n-1}^\dagger,c_{2n}^\dagger)\), then we have
with
and
when \(|m-n|>1\), we have \(H_{mn}=0\).
For example, for 6 cells (12 sites), we have
Open chain: \(T_6=0\).
Closed chain with translational symmetry, \(T_n=T,U_n=U\), with
\[\begin{split}U=\begin{pmatrix}0&v\\v^*&0\end{pmatrix},T=\begin{pmatrix}0&0\\w&0\end{pmatrix}\end{split}\]Using three Pauli matrices
\[\begin{split}\sigma_x=\begin{pmatrix}0&1\\1&0\end{pmatrix},\sigma_y=\begin{pmatrix}0&-i\\i&0\end{pmatrix},\sigma_z=\begin{pmatrix}1&0\\0&-1\end{pmatrix}\end{split}\]
We get
Using \(\mathbf{c}_n=e^{ik(n-1)b}\mathbf{c}_1\), we have
which gives us \(H=H(k)\oplus H(k)\cdots \oplus H(k)=\oplus_{n=1}^N H(k)\) with
with
with \(w=|w|e^{i arg(w)}\).
with
with eigen-energy
and eigen-wavefunctions
with \(tan\phi=h_y/h_x\).
For example, set \(arg(v)=arg(w)=0\), we have
Can not tell the difference \(|v|-|w|=\pm\delta\).
Energy-band description is not completed, it can give us many information, but not the whole, others are hidden in the wave-function. Alternatively, recalling \(H(k)=\mathbf{h}(k)\cdot \mathbf{\sigma}\), the Hamiltonian should contain the whole information, but we have only used \(|h|\), in topological aspect, \(\mathbf{h}(k)\) will suffices.
Set \(arg(v)=0\),\(kb=[0,2\pi]\), we have two cases
- \(|w|<|v|, \mathbf{inter}<\mathbf{intra}\)
- \(|w|>|v|, \mathbf{inter}>\mathbf{intra}\)
Winding number¶
\(H(k)=\mathbf{h}(k)\cdot \mathbf{\sigma}\), \(\mathbf{h}(k)=0\) is a degenerate point with \(|v|=|w|\), two bands cross, define \(h(k)=h_x(k)+ih_y(k)\), we have
define
When
- \(|w|>|v|, \nu=1, \mathbf{inter}>\mathbf{intra}\)
- \(|w|<|v|, \nu=0, \mathbf{inter}<\mathbf{intra}\)
A example, \(N=20, M=2N=40, w=1, v=0.5\), we get eigen-energys:
Edge-states:
Chiral symmetry¶
Recalling the Hamiltonian:
Define projector operators:
and the chiral operator \(\Sigma_z=P_A-P_B\), The matrix elements of \(\Sigma_z\) vanish, \(\langle 0|c_r \Sigma c^\dagger_s|0\rangle=0\) if sites \(r\) and \(s\) are in different unit cells.
In first-quantization, we have
and
In matrix form, we have
\(\Sigma_z\) is local, it does not mix site between unit cells, and inherits the algebra from \(\sigma_z\):
Recalling
There are no onsite terms in the Hamiltonian, so we have
This is the chiral symmetry.
Note
Actually, here, \(H\) is defined in momentum space, but \(\Sigma_z\) in real space, we should write \(\tilde{H}=U^\dagger H U\) for some unitary matrix \(U\), but the property survives!
Consequences: For eigenstates \(|\psi_n\rangle\) of \(H\), we have
\(E_n\neq 0\), two orthonormal states \(|\psi_n\rangle, \Sigma_z|\psi_n\rangle\), which gives
\[\begin{split}\begin{pmatrix}\alpha^* &\beta^*\end{pmatrix}\begin{pmatrix}\alpha\\-\beta\end{pmatrix}=0\end{split}\]\[\Rightarrow |\alpha|^2=|\beta|^2\]\(E_n=0\), we particularly have \(\Sigma_z=\Sigma_zH\), \(|\psi_n\rangle\) can be eigen-states of \(H\) and \(\Sigma\) simultaneously, \(\Sigma_z|\psi_n\rangle=\pm|\psi_n\rangle\), gives \(|\psi_n\rangle=\begin{pmatrix}1 \\0\end{pmatrix}\) or \(|\psi_n\rangle=\begin{pmatrix}0 \\1\end{pmatrix}\)
Review 2¶
In the last class, we have solved 1-d atom chain with staggered hopping parameters and got the Hamiltonian:
with \(tan\phi=h_y/h_x\).
Eigen-values are \(E(k)=\pm|h(k)|\), with eigen-functions: \(|\pm\rangle=\begin{pmatrix}\pm e^{-i\phi(k)}\\ 1\end{pmatrix}\).
\(\mathbf{h}(k)\) depicted as follow:
Topological transition¶
We have already seen, for \(|w|>|v|\) or \(|w|<|v|\), we got different winding number, means there is a topological transition at \(|w|=|v|\). In the energy-band point of view, it means the gap between two energy bands closes (across each other) and reopens. There are two ways to change the winding number and get a topological transition:
- Pull the path through the origin in the \(h_x-h_y\) plane.
- Lift it out of the plane (breaking the chiral symmetry).
Case 1(a): \(v=0.5, w=0\to 1\)
Case 1(b): \(w=1, v=2.5\to 0\)
Case 2: Sublattice potential \(H_{sublattice}=u\sigma_z\). Recalling
with \(U=\begin{pmatrix}0&v\\v^*&0\end{pmatrix},U=\begin{pmatrix}0&0\\w&0\end{pmatrix}\), now \(U\) changes to \(U=\begin{pmatrix}u&v\\v^*&-u\end{pmatrix}\), and \(H(k)=\mathbf{h}(k)\cdot\mathbf{\sigma}\) with \(h_z(k)=u\).
Set \(w=1, v=1.5-cos\theta, u=sin\theta, \theta=0\to \pi\),
Case 3: Charge Pump: Using the Hamiltonian \(H=u(t)\sum\limits_{i=1}^N(-1)^{i-1}c_i^{\dagger}c_i+\sum\limits_{i=1}^N[t_0+\delta(t)(-1)^i](c_i^{\dagger}c_{i+1}+h.c.)\), with \(u(t)=u_0 sin\theta, \delta(t)=\delta_0 cos\theta\), add time variation with \(\theta=2\pi t/T\), in previous notation, \(v=t_0-\delta(t), w=t_0+\delta(t)\). set \(u=sin\theta, v=0.75-0.25cos\theta, w=0.75+0.25cos\theta\), the vector \(\vec{h(k)}\) is
The energy band evolution is
Winding number v.s. Zak phase¶
From
We have defined the winding number
From the Zak phase definition:
Recalling \(|\psi\rangle=\frac{1}{\sqrt{2}}\begin{pmatrix}\pm e^{-i\phi(k)}\\ 1\end{pmatrix}\Rightarrow\gamma=\frac{1}{{2}}\oint dk\nabla_k\phi=\pm\pi \ \textit{or}\ 0\).
Lecture 2 : Berry Phase and Chern number¶
Berry Phase review¶
Assuming a physical system is depended on some parameters \(\mathbf{R}=(R_1,R_2,\cdots,R_N)\), we have the snapshot Hamiltonian \(H(\mathbf{R})\), its eigen-values and eigen-states:
where \(|n(\mathbf{R})\rangle\) can have an arbitrary phase prefactor.
The parameters \(\mathbf{R(t)}\) are slowly changed with time \(t\), then the adiabatic evolution of time-dependent Schrodinger equation:
Take the Ansatz: \(|\psi(t)\rangle=e^{i\gamma_n(t)}e^{-i\int_0^tE_n(\mathbf{R(t')})dt'}|n(\mathbf{R(t)})\rangle\), we have
That gives the Berry phase expression:
Define Berry connection:
Gauge transformation:
\(\gamma=\oint\mathbf{A}(\mathbf{R})d\mathbf{R}\) is gauge invariant.
Gauge and Parallel transportation: recalling the arbitrary phase
why shouldn’t we choose one which makes
from
then we have
There is no Berry Phase in this frame, which is called inertial frame, the condition \(\frac{d}{dt}|n\rangle\equiv 0\) is called parallel transportation. All the information resorted to \(|n(\mathbf{R})\rangle\), similar to a transformation from active frame to passive frame.
Berry curvature¶
Define the Berry curvature:
Using Stokes theorem, we have for the Berry Phase:
where \(\mathcal{S}\) is any surface whose boundary is the loop \(\mathcal{C}\).
Two useful formula:
- \(B_j=\epsilon_{jkl}\partial_kA_l=-Im\epsilon_{jkl}\partial_k\langle n|\partial_ln\rangle=-Im\epsilon_{jkl}\langle\partial_kn|\partial_ln\rangle\), that is \(\mathbf{B}^{(n)}=-Im\sum\limits_{n'\neq n}\langle\nabla n|n'\rangle\times\langle n'|\nabla n\rangle\).
\(\mathbf{B}^{(n)}=-Im\sum\limits_{n'\neq n}\langle\nabla n|n'\rangle\times\langle n'|\nabla n\rangle\), to calculate \(\langle n'|\nabla n\rangle\), start from:
\[H(\mathbf{R})|n\rangle=E_n|n\rangle\]\[\Rightarrow (\nabla H)|n\rangle+H|\nabla n\rangle=(\nabla E_n)|n\rangle+E_n|\nabla n\rangle\]\[\Rightarrow \langle n'|\nabla H|n\rangle+\langle n'|H|\nabla n\rangle=E_n\langle n'|\nabla n\rangle\]\[\Rightarrow \langle n'|\nabla n\rangle=\frac{\langle n'|\nabla H|n\rangle}{E_n-E_{n'}}\]
then we get:
\[\mathbf{B}^{(n)}=-Im\sum\limits_{n'\neq n}\langle\nabla n|n'\rangle\times\langle n'|\nabla n\rangle=-Im\sum\limits_{n'\neq n}\frac{\langle n|\nabla H|n'\rangle\times\langle n'|\nabla H|n\rangle}{(E_n-E_{n'})^2}\]
which is gauge invariant!
Berry curvature from perturbation theory¶
We can use time-independent perturbation theory to derive the changes of instant snapshot basis:
we have
We see that \(\langle n|\Delta n(\mathbf{R})\rangle=0\), which means we have used parallel transport gauge, more general, we should add a arbitrary phase factor in the above equation for \(|n(\mathbf{R+\Delta R})\rangle\).
From \(\mathbf{B}^{(n)}=-Im\sum\limits_{n'\neq n}\langle\nabla n|n'\rangle\times\langle n'|\nabla n\rangle\), we also get:
Also notice:
Which gives:
Benchmark: Spin-1/2¶
Gauge!Gauge!Gauge!
2-level Hamiltonian \(H(\mathbf{R})=h_0(\mathbf{R})\sigma_0+\mathbf{h}(\mathbf{R})\cdot\mathbf{\sigma}\), we can set \(h_0=0\), because it does not affect the eigenstates, eigen-energy are \(\pm|\mathbf{h}|\), introduce the unit vector:\(\hat{\mathbf{h}}=\mathbf{h}/|\mathbf{h}|\), the endpoints of \(\hat{\mathbf{h}}\) map out the surface of a unit sphere, called the Bloch sphere shows below:
Lecture 3 : Chern Insulator¶
Quantum Hall Insulator¶
The first Topological Insulator is shown in Integer quantum Hall effect. For 2D electron gas (2DEG), by applying a B field perpendicular to the 2D plane, we get plateaus of Hall conductance \(\sigma_{xy}=1/\rho_{xy}\) which is quantized in \(\nu e^2/h\). In the same time, \(\rho_{xx}=\sigma_{xx}=0\) (which may seems counterintuitive, but noticing the tensor nature of \(\rho\) and \(\sigma\)). The quantization of \(\sigma_{xy}\) is very accurate, it is the second best way to measure the fine structure constant, and it is inert to interactions and impurities. For a charge neutral particle in 2D:
with \(H_0=\frac{\hat{p}^2}{2m}\).
In a magnetic field with a vector potential \(\vec{A}\), we have \(H=\frac{(\hat{p}-e\vec{A})^2}{2m}\).
For a constant uniform magnetic field, choose Landau gauge: \(A_x=0, A_y=Bx\), we have
\(H\) is invariant under translations along the y axis, so \([p_y,H]=0\), we can write the eigen states of \(H\): \(\psi=f(x)exp(ik_yy)\) and \(H\psi=\epsilon\psi\). which gives:
\(\phi_n(x)\) satisfies the harmonic oscillator potential \(V(x)=\frac{k}{2}x^2\) with \(k=e^2B^2/m, x_0=\frac{\hbar k_y}{eB}\). The eigen energies are the famous harmonic one:
Some comments
- Landau levels are highly degenerated, they form energy bands with \(n\) denotes band index and \(k_y\) the momentum, all the states with the same \(n\) (but different \(k_y\)) have the same energy (flat bands).
- The degeneracy of each Landau level is \(D=(2S+1)\frac{\Phi}{\Phi_0}\), where \(S\) denotes spin, and \(\Phi_0\) is the magnetic flux quanta with \(\Phi_0=h/e\).
- We can also use symmetry gauge: \(A_x=-\frac{By}{2}, A_y=\frac{Bx}{2}\), though it is more cumbersome, the physics is quite insightful. More on symmetry gauge, see[1].
Edge states with Quantization of the Hall conductivity¶
Electrons at the edge can move along the edge in one direction, this edge states gives quantized Hall conductivity.
At top edge: \(N=\int \frac{dpdr}{2\pi\hbar}=L\int \frac{dp}{2\pi\hbar}\), we get
We have use the relation \(v_F=\frac{\partial E(k)}{\hbar \partial k}=\frac{\mu_{top}-\mu}{\hbar(k_{Ftop}-k_F)}\)
Net current for top edge:
We have used \(\mu_1-\mu_2=e(V_1-V_2)\). And, similarly, we have \(I_{bottom}=I_{top}\), that gives
If one has \(n\) copies of edge states, then total Hall conductivity is:
Half BHZ model¶
See the lattice below: Consider at each site, there are two
orbital modes coupled together, in the tight-binding approximation, we
can write the Hamiltonian:
Assuming translational invariance along x and y, i.e. \(U(n)=U,T_x(n)=T_x,T_y(n)=T_y\), after Fourier transform: \(c_{n_x,n_y}=\frac{1}{\sqrt{N}}\sum_{k_x,k_y}c_{k_x,k_y}e^{ik_xn_x+ik_yn_y}\) with \(N=N_xN_y, k_xN_x=2\pi m_1,k_yN_y=2\pi m_2\), we get
with
\(U=\Delta\sigma_z, T_x=\frac{1}{2}\sigma_z-\frac{iA}{2}\sigma_x, T_y=\frac{1}{2}\sigma_z-\frac{iA}{2}\sigma_y\), we get
Chern number using \(\mathbf{h}\)¶
Written the Hamiltonian as: \(H(k)=\mathbf{h}\cdot\mathbf{\sigma}\), we get
and we get the eigen-energies:
Set \(A=1\), for some special parameter \(\Delta\), the
dispersion relation shows below:
From the dispersions also the analysis we can see that Dirac cone
appears at
\(k_x,k_y=0,\Delta=-2; k_x,k_y=0,\pm\pi,\Delta=0; k_x,k_y=\pm\pi,\pm\pi,\Delta=2\).
It means that the monopole seats at \(\mathbf{h}=0\), at each region, we
can calculate the Chern number of the valence band in investigating how
many times does the torus formed by the image of the Brillouin zone in
the space of \(\mathbf{h}\) contail the origin. The Torus for different
\(\Delta=-2.5,-1,1,2.5\) shown below (for clarity, only half of the
torus is shown):
From \(\mathbf{h}\), we get the Chern number as follow: (Set \(A=1\))
- \(|\Delta|>2, Q=0\)
- \(-2<\Delta<0, Q=-1\)
- \(0<\Delta<2, Q=1\)
As long as the whole band is gapped, we have the same topology and the same Chern number. In the region \(|\Delta|>2\), reminding the Hamiltonian (In what follows, we always consider \(A=1\))
It represent Hamiltonians topologically equivalent to the atomic limit, the limit in which all hoppings are set to zero, as if the lattice constant were infinity. The \(|\Delta|>2\) region is topologically the same phase as the phase \(|\Delta|\to\infty\), which has trivial (momentum-independent) eigenstates and zero Chern number.
Chern number as a Skyrmion number¶
Looking at the vector \(\mathbf{h}\), the
\((h_x(k_x,k_y),h_y(k_x,k_y))\) shows below
To understand the Chern number, we can see if the \(h_z\) component
change sign or not in the BZ. If it changes sign, it corresponds to a
skyrmion, and the Chern number is non-zero, it is the same as Skyrmion
number. If it does not change sign, then the topology is trivial. The
value of \(h_z(\mathbf{k})\) at high-symmetry points in the BZ are shown
below: From which we can get the Chern number.
Low-energy Hamiltonian¶
We can also calculate the Chern number using the low-energy Hamiltonian. At \(\Delta=-2\), the energy gap collapses at the \(\Gamma=(0,0)\) point, near this point, we have
For the Hamiltonian \(H(k)=k_x\sigma_x+k_y\sigma_y+m\sigma_z\), we can get the monopole field for \(E_-\)state is
Then we get
Which gives:
Alternatively, from \(\mathbf{B}=\frac{1}{2}\frac{(k_x,k_y,m)}{(k_x^2+k_y^2+m^2)^{3/2}}\),
For \(m>0\) and \(m<0\), the integral region and direction shows below:

Bfield
Note here the Chern number is not an integer, because we used continuous model, and \(k_x,k_y\) can range from \(-\infty\) to \(\infty\), so the band structure \(\pm\sqrt{k_x^2+k_y^2+m^2}\) run riot and never turns back, which should give us another half of the Chern number (though we have no idea is we should add or subtract half-values). But as we can see below, the differences of Chern number between two states gives the right answer.
So from \(\Delta<-2\to \Delta>-2\), \(m<0\to m>0\), we get
Also we know \(\Delta<-2\), \(C=0\), so for \(-2<\Delta<0\), \(C=-1\).
Chern number as an obstruction¶
Write \(\mathbf{h}=|\mathbf{h}|(sin\theta cos\phi, sin\theta sin\phi, cos\theta)\). In principle, for the whole Bloch sphere, we must at least use two different bases which are well defined. If for the whole band, we have \(h_z<0\) or \(h_z>0A\), then, we can use one set of eigen base, which gives us zero Chern number. But for the region \(-2<\Delta<0\), at the point \(\Gamma\), we have \(h_z(0,0)=\Delta+2>0\), \(h_z/h=1=cos\theta\), \(\theta=0\); at the point \((0,\pm\pi)\) or \((\pm\pi,\pm\pi)\), we have \(h_z<0\), \(\theta=\pi\), means we must use two set of eigen bases, denotes no-zero Chern number. The resulting \(\hat{h}=\mathbf{h}/|\mathbf{h}|\) for different\(\Delta=-2.5\to 2.5\) shown below:
Haldane Model¶
Haldane model describe the model of Graphene with real nearest-neighbor-hopping parameters but complex next-nearest-neighbor-hopping parameters resulting from a nonzero magnetic field. But the field is periodic and have zero net flux per plaquette. The real space and reciprocal space shows below (with unit length 1):

lattice
The vectors are:
Set \(\mathbf{b_1},\mathbf{b_2}\) as two base vectors, then the reciprocal lattice spanned by \(\mathbf{b_1^*},\mathbf{b_2^*}\) with \(\mathbf{b_i}\cdot\mathbf{b_j^*}=2\pi\delta_{ij}\), that is
Define \(\mathbf{K}=\frac{1}{3}(\mathbf{b_1^*}+\mathbf{b_2^*})=\begin{pmatrix}-4\sqrt{3}\pi/9\\0\end{pmatrix}, \mathbf{K'}=-\mathbf{K}\).
We get the Hamiltonian:
\(M\) breaks inversion symmetry of \(A\) and \(B\) sublattice, and \(t_2\) is the next-nearest-neighbor coupling, for a staggered magnetic field shows below, it changes \(t_{ij}\to t_{ij}e^{i\frac{e}{\hbar}\int_{j\to i}\mathbf{\mathcal{A}}\cdot d\mathbf{l}}=t_{ij}e^{i\frac{e}{\hbar}\chi\Phi_0}=t_{ij}e^{i\frac{e}{\hbar}\chi\frac{h}{e}}=t_{ij}e^{i\phi}\).

mag
Use \(a_{iA}=\frac{1}{\sqrt{N}}\sum_ka_{kA}e^{i\mathbf{k}\cdot\mathbf{R}_{iA}}, a_{iB}=\frac{1}{\sqrt{N}}\sum_ka_{kB}e^{i\mathbf{k}\cdot\mathbf{R}_{iB}}\), \(N\) is the number of unit cell. Then we can get:
We arrived at
with
Because of the \(C_3\) symmetry, the Dirac cone can only happen at \(K,K'\). At the vicinity of \(K\), we have
At the vicinity of \(K'\), we have
where \(-\pi<\phi<\pi\).
Start from \(M\to \pm\infty\), we have zero Chern number, assuming \(t_2>0, \phi>0\), from the previous results, when \(M\) from \(+\infty\) decreases to \(M=3\sqrt{3}t_2sin\phi\), \(K'\) band close and reopen, the Chern number changes to
Then when \(M>-3\sqrt{3}t_2sin\phi\) to \(M<-3\sqrt{3}t_2sin\phi\), we have
We can do the same calculation for \(\phi<0\), we get the phase chart below:

phase chart
Because \(k_x,k_y\) have no dependence on \(M,t_2,\phi\), \(h_z\) in the range of \((M-3\sqrt{3}t_2sin\phi\to M+3\sqrt{3}t_2sin\phi)\), we can draw a spheroid \(\Sigma'\) with height \(6\sqrt{3}sin\phi\) and center \((0,0,M/t_2)\). We can get the Chern number using following pictures:

spheroid
Also, we can get the Chern number using the whole 3-d picture (set \(\phi=\pi/2, M/t_2=6,0,-6\)):
Other good Books and review papers on this subject:
- Topological Insulators and Topological Superconductors by Bernevig, B..
- Topological Insulators: Dirac Equation in Condensed Matters by Shun-Qing Shen.
- Fruchart and D. Carpentier, Comptes Rendus Physique 14, 779 (2013).
- Kohmoto, Annals of Physics 160, 343 (1985).
- Xiao, M.-C. Chang, and Q. Niu, Reviews of Modern Physics 82, 1959 (2010).
- Mokrousov and F. Freimuth, arXiv:1407.2847 [cond-Mat] (2014).
More geometrical staffs at this one Geometry, Topology and Physics, by Nakahara.
Weyl Semi-metal¶
Graphene¶
In the tight-binding approximation, when only nearest neighborhood couplings are considered, the Hamiltonian of Graphene can be written as:
where \(h(\vec{k})=\sum\limits_{\vec{\delta}_i}e^{i\vec{k}\cdot\vec{\delta}_i}=|h(\vec{k})|e^{i\phi(\vec{k})},\vec{\delta}_i\) are three position vectors shown in the following diagram.

Fig. 1 Crystal Structure of Graphene: \(\vec{a}_{1}\) and \(\vec{a}_{2}\) are Bravais crystal vectors for a Graphene unit cell. Each primitive unit cell has two atomic sties, A and B. \(\vec{\delta}_{i}\) specifies B-s’ position around A site. (b) Brillouin Zone for Graphene \(\vec{b}_{1}\) and \(\vec{b}_{2}\) are reciprocal vector basis for intrinsic Graphene. Its corners are known as K and K’ points.
Then the Hamiltonian is:
with the eigen-value: \(E(\vec{k})=-t|h(\vec{k})|\) and eigen-function (only show one of the two):
We should notice that the \(1/2\) factor is quite important here, when \(\phi\) changes \(2\pi\), the wave-function does not return to its original value, but with a minus sign. If instead, we want the wave-function to be single valued, the function :math:`psi(vec{k})`should change accordingly.
At the vicinity of Dirac point (K or K’, here we expand the things near K), we have:
We can see that the general \(\phi\) turn out to be the angle of \(\vec{q}\) with the x axis. Then, wind a circle around the Dirac point K at some energy in the band structure shown below (Fig. 2(a)), the corresponding \(\phi\) (Fig. 2(c))winds one round too.
So, if \(\vec{q}\) circles around the Dirac point one turn, \(\phi\) changes from 0 to \(2\pi\), in order to keep the basic wave-function \(u(\vec{q})\) single-valued, \(\psi\) must changes \(\pi\).
More explicitly, we calculate the vector potential in momentum space with:
Then we get:
We got the Berry’s phase \(\gamma=\pm\pi\). It is this non-trivial phase of the equal-frequency surface that makes us call it Weyl semi-metal, and the Dirac points called Weyl nodes.
Three dimension: Weyl semi-metal and Chern number¶
In three dimension, things can goes the same way. Using a simple model with Hamiltonian:
It has two Weyl nodes: \(\vec{K}=\pm\left(k_0,0,0\right)\), which means if we treat \(k_x\) as a variable, only when \(k_x=\pm k_0\), the corresponding energy band \(E_{k_x}(k_y,k_z)\) is crossing at the point \(k_y,k_z=(0,0)\).
Also, at the Weyl node (say \(k_x=k_0\)), we have:
with \(v_x=2t_xsink_0,v_y=2t_y,v_z=2t_z\). Without loss of generality, we can set \(v_x=v_y=v_z\) (the only effect is the shape changing from sphere to ellipsoid, which has no effect on the topology), then we get:
with eigen-value: \(E(\vec{k})=v|\vec{q}|\) and eigen-function (only show one of the two):
It is easy to find that this wave-function will give us a magnetic field with a monopole at \(\vec{K}\), which will give us non-trivial equal-frequency surface Chern number \(C=1\).
Bulk-boundary corresponding¶
In order to see things clearer, also, to see the connection of Weyl semi-metal with Topological insulator, we treat \(k_x\) as a variable, the Hamiltonian is:
For example, we set \(t_x=t_y=t_z=1,m=2,k_0=1\), three typical energy band dispersions shown below:
To see if the system with \(k_x\neq k_0\) is a topological insulator or not, we can plot the diagram of \(\vec{h}(k_y,k_z)\), and see how many times the resulting torus incorporates the origin point. Typical shape of the torus shows below:
We found that at the region \(k_x=[-k_0,k_0]\), the origin point is in the torus once with Chern number \(C=1\), outside of that, we got \(C=1\) (Noticing we have \(k_x=[-\pi,\pi]\)) . This is why we plot the edge state in Fig. 4(a), but not in Fig. 4(c). In the non-trivial case, for any energy inside the gap, we get a edge state, so different \(k_x\) will give us a edge-state line, which is called Fermi-arc, especially, when we look at the case of Fermi surface with energy \(E=0\), the Fermi-arc stretch from one Weyl node to another, like the picture shown below:
Alternatively, we can look at the Bulk-boundary corresponding another way. The Weyl points behave like “magnetic” monopoles in momentum space whose charge is given by the chirality; they are actually a source of “Berry flux” rather than magnetic flux.
Consider a curve in the surface Brillouin zone encircling the projection of the bulk Weyl point, which is traversed counterclockwise as we vary the parameter \(\lambda: 0 \to 2\pi; \mathbf{k}_{\lambda} = [k_x (\lambda),k_y (\lambda)]\) [see Fig. 6(a)].We show that the energy \(\lambda\) of a surface state at momentum \(\mathbf{k}_{\lambda}\) crosses \(E = 0\) at some value of \(\lambda\). Consider \(H(\lambda,k_z) = H(\mathbf{k}_{\lambda},k_z)\), which can be interpreted as the gapped Hamiltonian of a two-dimensional system (with \(\lambda\) and \(k_z\) as the two momenta). The two periodic parameters \(\lambda\) , \(k_z\) define the surface of a torus in momentum space. The Chern number of this two-dimensional band structure is given by the Berry curvature integration: \(\frac{1}{2\pi}\int\mathscr{F}dk_zd\lambda\), which, by the Stokes theorem, simply corresponds to the net monopole density enclosed within the torus. This is obtained by summing the chiralities of the enclosed Weyl nodes. Consider the case when the net chirality is unity, corresponding to a single enclosed Dirac node. Then, the two-dimensional subsystem is a quantum Hall insulator with unit Chern number. When the system is given a boundary at \(z = 0\), we expect a chiral edge state for this subsystem [see Fig. 6(b)]. Hence, this surface state crosses zero energy somewhere on the surface Brillouin zone \(\mathbf{k}_{\lambda_0}\) . Such a state can be obtained for every curve enclosing the Weyl point. Thus, at zero energy, there is a Fermi line in the surface Brillouin zone, that terminates at the Weyl point momenta [see Fig. 6(c)]. An arc beginning on a Weyl point of chirality c has to terminate on a Weyl point of the opposite chirality. Clearly, the net chirality of the Weyl points within the \((\lambda,k_z)\) torus was a key input in determining the number of these states. If Weyl points of opposite chirality line up along the \(k_z\) direction, then there is a cancellation and no surface states are expected.
References:
- Wan, A.M. Turner, A. Vishwanath, and S.Y. Savrasov, Physical Review B 83, 205101 (2011).
- A.M. Turner, A. Vishwanath, and C.O. Head, Topological Insulators 6, 293 (2013).
- Haldane, Physical Review Letters 93, 206602 (2004).
Condensed Matter Physics:¶
Linear response theory¶
Kubo formula¶
Many times, the world shows us as a black box, we can only get limited knowledge about it and the rest are filled with our theories, in this way, theories can always changing and are never completed. we always use a probe to probe a system and check its response to this probe, and naturally, we expect when the perturbation of the probe is ignorable to the system, the response should be linear to the probe, which is called linear response theory. The most important result of linear response theory is consummated by Kubo formula, we are now going to derive it.
Zero temperature¶
Let \(\hat{H}_0\) be the full many-body Hamiltonian for some isolated system that we are interested in, and assume the existence of a set of eigenkets \(\{|n\rangle\}\) that diagonalize \(\hat{H}_0\) with associated eigenvalues (energies) \(\varepsilon_n\).
In addition to \(\hat{H}_0\), we now turn on an external probe potential \(\hat{V}(t)\), such that the total Hamiltonian \(\hat{H}(t)\) satisfies:
Note
The additional factor \(e^{\eta t}\) means we switch on the external potential adiabatically from \(t\to -\infty\), we’ll see later that it is this factor gives us the way to detour the singular points, it is an analytical continuation which is a reflection of causality.
The Schrödinger equation of the system now reads:
Warning
us the way to detour the singular points, it is an analytical continuation which is a reflection of causality.
Important
us the way to detour the singular points, it is an analytical continuation which is a reflection of causality.
Hint
us the way to detour the singular points, it is an analytical continuation which is a reflection of causality.
Fermi’s Golden rule¶
Some topics want to explore¶
- Green’s Function
- Dirac,Majorana,Weyl fermions
Something about Python¶
Python 学习¶
- 数组初始化
#初始化一维数组
p1=[1]*100
#初始化二维数组100*10
p2=[[1] * 100 for i in xrange(10)]
#Below does not work!
p2=[[1] * 100]*10
#Or we can use:
p2=[[None for col in range(100)] for row in range(10)]
#初始化三维维数组2*3*4
p3=[[[1] * 2 for i in xrange(3)] for j in xrange(4)]
#Or
p3=[[[1 for i in xrange(2)] for j in xrange(3)] for k in xrange(4)]
#Or
p3=[]
for i in range(2):
p3.append([])
for j in range(3):
p3[i].append([])
for k in range(4):
p3[i][j].append(0)
- python 中的内积
dot
是没有复共轭的,vdot
有,而matlabdot(a,b)
是对a取复共轭的。
Download the Latest PDF Version , contact me through Gmail.