Dyson series

Expansion of the time evolution operator

In scattering theory, a part of mathematical physics, the Dyson series, formulated by Freeman Dyson, is a perturbative expansion of the time evolution operator in the interaction picture. Each term can be represented by a sum of Feynman diagrams.

This series diverges asymptotically, but in quantum electrodynamics (QED) at the second order the difference from experimental data is in the order of 10−10. This close agreement holds because the coupling constant (also known as the fine-structure constant) of QED is much less than 1.[clarification needed]

Dyson operator

In the interaction picture, a Hamiltonian H, can be split into a free part H0 and an interacting part VS(t) as H = H0 + VS(t).

The potential in the interacting picture is

V I ( t ) = e i H 0 ( t t 0 ) / V S ( t ) e i H 0 ( t t 0 ) / , {\displaystyle V_{\mathrm {I} }(t)=\mathrm {e} ^{\mathrm {i} H_{0}(t-t_{0})/\hbar }V_{\mathrm {S} }(t)\mathrm {e} ^{-\mathrm {i} H_{0}(t-t_{0})/\hbar },}

where H 0 {\displaystyle H_{0}} is time-independent and V S ( t ) {\displaystyle V_{\mathrm {S} }(t)} is the possibly time-dependent interacting part of the Schrödinger picture. To avoid subscripts, V ( t ) {\displaystyle V(t)} stands for V I ( t ) {\displaystyle V_{\mathrm {I} }(t)} in what follows.

In the interaction picture, the evolution operator U is defined by the equation:

Ψ ( t ) = U ( t , t 0 ) Ψ ( t 0 ) {\displaystyle \Psi (t)=U(t,t_{0})\Psi (t_{0})}

This is sometimes called the Dyson operator.

The evolution operator forms a unitary group with respect to the time parameter. It has the group properties:

  • Identity and normalization: U ( t , t ) = 1 , {\displaystyle U(t,t)=1,} [1]
  • Composition: U ( t , t 0 ) = U ( t , t 1 ) U ( t 1 , t 0 ) , {\displaystyle U(t,t_{0})=U(t,t_{1})U(t_{1},t_{0}),} [2]
  • Time Reversal: U 1 ( t , t 0 ) = U ( t 0 , t ) , {\displaystyle U^{-1}(t,t_{0})=U(t_{0},t),} [clarification needed]
  • Unitarity: U ( t , t 0 ) U ( t , t 0 ) = 1 {\displaystyle U^{\dagger }(t,t_{0})U(t,t_{0})=\mathbb {1} } [3]

and from these is possible to derive the time evolution equation of the propagator:[4]

i d d t U ( t , t 0 ) Ψ ( t 0 ) = V ( t ) U ( t , t 0 ) Ψ ( t 0 ) . {\displaystyle i\hbar {\frac {d}{dt}}U(t,t_{0})\Psi (t_{0})=V(t)U(t,t_{0})\Psi (t_{0}).}

In the interaction picture, the Hamiltonian is the same as the interaction potential H i n t = V ( t ) {\displaystyle H_{\rm {int}}=V(t)} and thus the equation can also be written in the interaction picture as

i d d t Ψ ( t ) = H i n t Ψ ( t ) {\displaystyle i\hbar {\frac {d}{dt}}\Psi (t)=H_{\rm {int}}\Psi (t)}

Caution: this time evolution equation is not to be confused with the Tomonaga–Schwinger equation.

The formal solution is

U ( t , t 0 ) = 1 i 1 t 0 t d t 1   V ( t 1 ) U ( t 1 , t 0 ) , {\displaystyle U(t,t_{0})=1-i\hbar ^{-1}\int _{t_{0}}^{t}{dt_{1}\ V(t_{1})U(t_{1},t_{0})},}

which is ultimately a type of Volterra integral.

Derivation of the Dyson series

An iterative solution of the Volterra equation above leads to the following Neumann series:

U ( t , t 0 ) = 1 i 1 t 0 t d t 1 V ( t 1 ) + ( i 1 ) 2 t 0 t d t 1 t 0 t 1 d t 2 V ( t 1 ) V ( t 2 ) + + ( i 1 ) n t 0 t d t 1 t 0 t 1 d t 2 t 0 t n 1 d t n V ( t 1 ) V ( t 2 ) V ( t n ) + . {\displaystyle {\begin{aligned}U(t,t_{0})={}&1-i\hbar ^{-1}\int _{t_{0}}^{t}dt_{1}V(t_{1})+(-i\hbar ^{-1})^{2}\int _{t_{0}}^{t}dt_{1}\int _{t_{0}}^{t_{1}}\,dt_{2}V(t_{1})V(t_{2})+\cdots \\&{}+(-i\hbar ^{-1})^{n}\int _{t_{0}}^{t}dt_{1}\int _{t_{0}}^{t_{1}}dt_{2}\cdots \int _{t_{0}}^{t_{n-1}}dt_{n}V(t_{1})V(t_{2})\cdots V(t_{n})+\cdots .\end{aligned}}}

Here, t 1 > t 2 > > t n {\displaystyle t_{1}>t_{2}>\cdots >t_{n}} , and so the fields are time-ordered. It is useful to introduce an operator T {\displaystyle {\mathcal {T}}} , called the time-ordering operator, and to define

U n ( t , t 0 ) = ( i 1 ) n t 0 t d t 1 t 0 t 1 d t 2 t 0 t n 1 d t n T V ( t 1 ) V ( t 2 ) V ( t n ) . {\displaystyle U_{n}(t,t_{0})=(-i\hbar ^{-1})^{n}\int _{t_{0}}^{t}dt_{1}\int _{t_{0}}^{t_{1}}dt_{2}\cdots \int _{t_{0}}^{t_{n-1}}dt_{n}\,{\mathcal {T}}V(t_{1})V(t_{2})\cdots V(t_{n}).}

The limits of the integration can be simplified. In general, given some symmetric function K ( t 1 , t 2 , , t n ) , {\displaystyle K(t_{1},t_{2},\dots ,t_{n}),} one may define the integrals

S n = t 0 t d t 1 t 0 t 1 d t 2 t 0 t n 1 d t n K ( t 1 , t 2 , , t n ) . {\displaystyle S_{n}=\int _{t_{0}}^{t}dt_{1}\int _{t_{0}}^{t_{1}}dt_{2}\cdots \int _{t_{0}}^{t_{n-1}}dt_{n}\,K(t_{1},t_{2},\dots ,t_{n}).}

and

I n = t 0 t d t 1 t 0 t d t 2 t 0 t d t n K ( t 1 , t 2 , , t n ) . {\displaystyle I_{n}=\int _{t_{0}}^{t}dt_{1}\int _{t_{0}}^{t}dt_{2}\cdots \int _{t_{0}}^{t}dt_{n}K(t_{1},t_{2},\dots ,t_{n}).}

The region of integration of the second integral can be broken in n ! {\displaystyle n!} sub-regions, defined by t 1 > t 2 > > t n {\displaystyle t_{1}>t_{2}>\cdots >t_{n}} . Due to the symmetry of K {\displaystyle K} , the integral in each of these sub-regions is the same and equal to S n {\displaystyle S_{n}} by definition. It follows that

S n = 1 n ! I n . {\displaystyle S_{n}={\frac {1}{n!}}I_{n}.}

Applied to the previous identity, this gives

U n = ( i 1 ) n n ! t 0 t d t 1 t 0 t d t 2 t 0 t d t n T V ( t 1 ) V ( t 2 ) V ( t n ) . {\displaystyle U_{n}={\frac {(-i\hbar ^{-1})^{n}}{n!}}\int _{t_{0}}^{t}dt_{1}\int _{t_{0}}^{t}dt_{2}\cdots \int _{t_{0}}^{t}dt_{n}\,{\mathcal {T}}V(t_{1})V(t_{2})\cdots V(t_{n}).}

Summing up all the terms, the Dyson series is obtained. It is a simplified version of the Neumann series above and which includes the time ordered products; it is the path-ordered exponential:[5]

U ( t , t 0 ) = n = 0 U n ( t , t 0 ) = n = 0 ( i 1 ) n n ! t 0 t d t 1 t 0 t d t 2 t 0 t d t n T V ( t 1 ) V ( t 2 ) V ( t n ) = T exp i 1 t 0 t d τ V ( τ ) {\displaystyle {\begin{aligned}U(t,t_{0})&=\sum _{n=0}^{\infty }U_{n}(t,t_{0})\\&=\sum _{n=0}^{\infty }{\frac {(-i\hbar ^{-1})^{n}}{n!}}\int _{t_{0}}^{t}dt_{1}\int _{t_{0}}^{t}dt_{2}\cdots \int _{t_{0}}^{t}dt_{n}\,{\mathcal {T}}V(t_{1})V(t_{2})\cdots V(t_{n})\\&={\mathcal {T}}\exp {-i\hbar ^{-1}\int _{t_{0}}^{t}{d\tau V(\tau )}}\end{aligned}}}

This result is also called Dyson's formula.[6] The group laws can be derived from this formula.

Application on state vectors

The state vector at time t {\displaystyle t} can be expressed in terms of the state vector at time t 0 {\displaystyle t_{0}} , for t > t 0 , {\displaystyle t>t_{0},} as

| Ψ ( t ) = n = 0 ( i 1 ) n n ! d t 1 d t n t f t 1 t n t i T { k = 1 n e i H 0 t k / V ( t k ) e i H 0 t k / } | Ψ ( t 0 ) . {\displaystyle |\Psi (t)\rangle =\sum _{n=0}^{\infty }{(-i\hbar ^{-1})^{n} \over n!}\underbrace {\int dt_{1}\cdots dt_{n}} _{t_{\rm {f}}\,\geq \,t_{1}\,\geq \,\cdots \,\geq \,t_{n}\,\geq \,t_{\rm {i}}}\,{\mathcal {T}}\left\{\prod _{k=1}^{n}e^{iH_{0}t_{k}/\hbar }V(t_{k})e^{-iH_{0}t_{k}/\hbar }\right\}|\Psi (t_{0})\rangle .}

The inner product of an initial state at t i = t 0 {\displaystyle t_{i}=t_{0}} with a final state at t f = t {\displaystyle t_{f}=t} in the Schrödinger picture, for t f > t i {\displaystyle t_{f}>t_{i}} is:

Ψ ( t i ) Ψ ( t f ) = n = 0 ( i 1 ) n n ! × d t 1 d t n t f t 1 t n t i Ψ ( t i ) e i H 0 ( t f t 1 ) / V S ( t 1 ) e i H 0 ( t 1 t 2 ) / V S ( t n ) e i H 0 ( t n t i ) / Ψ ( t i ) {\displaystyle {\begin{aligned}\langle \Psi (t_{\rm {i}})&\mid \Psi (t_{\rm {f}})\rangle =\sum _{n=0}^{\infty }{(-i\hbar ^{-1})^{n} \over n!}\times \\&\underbrace {\int dt_{1}\cdots dt_{n}} _{t_{\rm {f}}\,\geq \,t_{1}\,\geq \,\cdots \,\geq \,t_{n}\,\geq \,t_{\rm {i}}}\,\langle \Psi (t_{i})\mid e^{-iH_{0}(t_{\rm {f}}-t_{1})/\hbar }V_{\rm {S}}(t_{1})e^{-iH_{0}(t_{1}-t_{2})/\hbar }\cdots V_{\rm {S}}(t_{n})e^{-iH_{0}(t_{n}-t_{\rm {i}})/\hbar }\mid \Psi (t_{i})\rangle \end{aligned}}}

The S-matrix may be obtained by writing this in the Heisenberg picture, taking the in and out states to be at infinity:[7]

Ψ o u t S Ψ i n = Ψ o u t n = 0 ( i 1 ) n n ! d 4 x 1 d 4 x n t o u t t n t 1 t i n T { H i n t ( x 1 ) H i n t ( x 2 ) H i n t ( x n ) } Ψ i n . {\displaystyle \langle \Psi _{\rm {out}}\mid S\mid \Psi _{\rm {in}}\rangle =\langle \Psi _{\rm {out}}\mid \sum _{n=0}^{\infty }{(-i\hbar ^{-1})^{n} \over n!}\underbrace {\int d^{4}x_{1}\cdots d^{4}x_{n}} _{t_{\rm {out}}\,\geq \,t_{n}\,\geq \,\cdots \,\geq \,t_{1}\,\geq \,t_{\rm {in}}}\,{\mathcal {T}}\left\{H_{\rm {int}}(x_{1})H_{\rm {int}}(x_{2})\cdots H_{\rm {int}}(x_{n})\right\}\mid \Psi _{\rm {in}}\rangle .}

Note that the time ordering was reversed in the scalar product.

See also

References

  1. ^ Sakurai, Modern Quantum mechanics, 2.1.10
  2. ^ Sakurai, Modern Quantum mechanics, 2.1.12
  3. ^ Sakurai, Modern Quantum mechanics, 2.1.11
  4. ^ Sakurai, Modern Quantum mechanics, 2.1 pp. 69-71
  5. ^ Sakurai, Modern Quantum Mechanics, 2.1.33, pp. 72
  6. ^ Tong 3.20, http://www.damtp.cam.ac.uk/user/tong/qft/qft.pdf
  7. ^ Dyson (1949), "The S-matrix in quantum electrodynamics", Physical Review, 75 (11): 1736–1755, doi:10.1103/PhysRev.75.1736
  • Charles J. Joachain, Quantum collision theory, North-Holland Publishing, 1975, ISBN 0-444-86773-2 (Elsevier)