B.4 The Evolution Operator $ \hat{S}$

To solve the equations of motion in the interaction picture (B.7), a unitary operator $ \hat{S}(t,t_0)$ that determines the state vector at time $ t$ in terms of the state vector at time $ t_0$ is introduced

\begin{displaymath}\begin{array}{l} \displaystyle \vert\Psi_\mathrm{I}(t)\rangle...
...\hat{S}(t,t_0) \vert\Psi_\mathrm{I}(t_0)\rangle \ , \end{array}\end{displaymath} (B.15)

$ \hat{S}$ satisfies the initial condition $ \hat{S}(t_0,t_0)=1$. For finite times $ \hat{S}(t,t_0)$ can be constructed explicitly by employing the SCHRÖDINGER picture

\begin{displaymath}\begin{array}{ll} \displaystyle \vert\Psi_\mathrm{I}(t)\rangl...
..._{0}t_0/\hbar} \vert\Psi_\mathrm{I}(t_0)\rangle \ , \end{array}\end{displaymath} (B.16)

which therefore identifies

\begin{displaymath}\begin{array}{l} \hat{S}(t,t_0) \ = \ e^{i\hat{H}_{0}t/\hbar}...
...hat{H}(t-t_0)/\hbar} e^{-i\hat{H}_{0}t_0/\hbar} \ . \end{array}\end{displaymath} (B.17)

Since $ \hat{H}$ and $ \hat{H}_0$ do not commute with each other, the order of the operators must be carefully maintained. Equation (B.17) immediately yields several general properties of $ \hat{S}$ [189] Although (B.17) is the formal solution to the problem posed by (B.15), it is not very useful for computational purposes. Instead one can construct an integral equation for $ \hat{S}$, which can then be solved by iteration. It follows from (B.7) and (B.15) that $ \hat{S}$ satisfies the differential equation

\begin{displaymath}\begin{array}{l} \displaystyle i\hbar \partial_t\hat{S}(t,t_0...
...at{H}^\mathrm{int}_\mathrm{I}(t) \hat{S}(t,t_0) \ . \end{array}\end{displaymath} (B.18)

Integrating both sides of the (B.18) with respect to time with the initial condition $ \hat{S}(t_0,t_0)=1$ yields

\begin{displaymath}\begin{array}{ll} \displaystyle \hat{S}(t,t_0) & \displaystyl...
...athrm{int}_\mathrm{I}(t_{1}) \hat{S}(t_{1},t_0) \ . \end{array}\end{displaymath} (B.19)

By iterating this equation repeatedly one gets

\begin{displaymath}\begin{array}{lll} \displaystyle \hat{S}(t,t_0) &\ = \ & \dis...
...{2})\ldots \hat{H}^\mathrm{int}_\mathrm{I}(t_{n})\ .\end{array}\end{displaymath} (B.20)

Equation (B.20) has the characteristic feature that the operator containing the latest time stands farthest to the left. At this point it is convenient to introduce the time-ordering operator denoted by the symbol $ T_\mathrm{t}$

$\displaystyle \displaystyle T_\mathrm{t}\{\hat{A}(t_{1})\hat{B}(t_{2})\}\ = \ \...
...}(t_{1})\hat{B}(t_{2})\ + \ \theta(t_{2}-t_{1})\hat{B}(t_{2})\hat{A}(t_{1}) \ .$ (B.21)

where $ \theta(t)$ is the step functionB.1. Each time two FERMIons are interchanged, the resulting expression changes its sign. By rearranging the integral using $ T_\mathrm{t}$

\begin{displaymath}\begin{array}{l}\displaystyle \frac{1}{2!}\int_{t_0}^{t} dt_{...
...}(t_{2}) \hat{H}^\mathrm{int}_\mathrm{I}(t_{1}) \ . \end{array}\end{displaymath} (B.22)

The second term on the right hand-side is equal to the first, which is easy to see by just redefining the integration variables $ t_{1}\rightarrow t_{2}$, $ t_{2}\rightarrow t_{1}$. Thus one gets

$\displaystyle \frac{1}{2!}\int_{t_0}^{t} dt_{1} \int_{t_0}^{t} dt_{2}\ T_\mathr...
...at{H}^\mathrm{int}_\mathrm{I}(t_{1}) \hat{H}^\mathrm{int}_\mathrm{I}(t_{2}) \ .$ (B.23)

Thus for the expansion of the $ \hat{S}(t,t_0)$ one obtains

\begin{displaymath}\begin{array}{ll} \displaystyle \hat{S}(t,t_0) & \displaystyl...
...t'\hat{H}^\mathrm{int}_\mathrm{I}(t') \right) \} \ .\end{array}\end{displaymath} (B.24)

M. Pourfath: Numerical Study of Quantum Transport in Carbon Nanotube-Based Transistors