Movatterモバイル変換


[0]ホーム

URL:


Jump to content
WikipediaThe Free Encyclopedia
Search

Perturbation theory (quantum mechanics)

From Wikipedia, the free encyclopedia
Approximate modelling of a quantum system

This article includes a list ofgeneral references, butit lacks sufficient correspondinginline citations. Please help toimprove this article byintroducing more precise citations.(October 2024) (Learn how and when to remove this message)

Inquantum mechanics,perturbation theory is a set of approximation schemes directly related to mathematicalperturbation for describing a complicatedquantum system in terms of a simpler one. The idea is to start with a simple system for which a mathematical solution is known, and add an additional "perturbing"Hamiltonian representing a weak disturbance to the system. If the disturbance is not too large, the various physical quantities associated with the perturbed system (e.g. itsenergy levels andeigenstates) can be expressed as "corrections" to those of the simple system. These corrections, being small compared to the size of the quantities themselves, can be calculated using approximate methods such asasymptotic series. The complicated system can therefore be studied based on knowledge of the simpler one. In effect, it is describing a complicated unsolved system using a simple, solvable system.

Approximate Hamiltonians

[edit]

Perturbation theory is an important tool for describing realquantum systems, as it turns out to be very difficult to find exact solutions to theSchrödinger equation forHamiltonians of even moderate complexity. The Hamiltonians to which we know exact solutions, such as thehydrogen atom, thequantum harmonic oscillator and theparticle in a box, are too idealized to adequately describe most systems. Using perturbation theory, we can use the known solutions of these simpleHamiltonians to generate solutions for a range of more complicated systems.

Applying perturbation theory

[edit]

Perturbation theory is applicable if the problem at hand cannot be solved exactly, but can be formulated by adding a "small" term to the mathematical description of the exactly solvable problem.

For example, by adding a perturbativeelectric potential to the quantum mechanical model of thehydrogen atom, tiny shifts in thespectral lines of hydrogen caused by the presence of anelectric field (theStark effect) can be calculated. This is only approximate because the sum of aCoulomb potential with a linear potential is unstable (has no true bound states) although thetunneling time (decay rate) is very long. This instability shows up as a broadening of the energy spectrum lines, which perturbation theory fails to reproduce entirely.

The expressions produced by perturbation theory are not exact, but they can lead to accurate results as long as the expansion parameter, sayα, is very small. Typically, the results are expressed in terms of finitepower series inα that seem to converge to the exact values when summed to higher order. After a certain ordern ~ 1/α however, the results become increasingly worse since the series are usuallydivergent (beingasymptotic series). There exist ways to convert them into convergent series, which can be evaluated for large-expansion parameters, most efficiently by thevariational method. In practice, convergent perturbation expansions often converge slowly while divergent perturbation expansions sometimes give good results, c.f. the exact solution, at lower order.[1]

In the theory ofquantum electrodynamics (QED), in which theelectronphoton interaction is treated perturbatively, the calculation of the electron'smagnetic moment has been found to agree with experiment to eleven decimal places.[2] In QED and otherquantum field theories, special calculation techniques known asFeynman diagrams are used to systematically sum the power series terms.

Limitations

[edit]

Large perturbations

[edit]

Under some circumstances, perturbation theory is an invalid approach to take. This happens when the system we wish to describe cannot be described by a small perturbation imposed on some simple system. Inquantum chromodynamics, for instance, the interaction ofquarks with thegluon field cannot be treated perturbatively at low energies because thecoupling constant (the expansion parameter) becomes too large, violating the requirement that corrections must be small.

Non-adiabatic states

[edit]

Perturbation theory also fails to describe states that are not generatedadiabatically from the "free model", includingbound states and various collective phenomena such assolitons.[citation needed] Imagine, for example, that we have a system of free (i.e. non-interacting) particles, to which an attractive interaction is introduced. Depending on the form of the interaction, this may create an entirely new set of eigenstates corresponding to groups of particles bound to one another. An example of this phenomenon may be found in conventionalsuperconductivity, in which thephonon-mediated attraction betweenconduction electrons leads to the formation of correlated electron pairs known asCooper pairs. When faced with such systems, one usually turns to other approximation schemes, such as thevariational method and theWKB approximation. This is because there is no analogue of abound particle in the unperturbed model and the energy of a soliton typically goes as theinverse of the expansion parameter. However, if we "integrate" over the solitonic phenomena, the nonperturbative corrections in this case will be tiny; of the order ofexp(−1/g) orexp(−1/g2) in the perturbation parameterg. Perturbation theory can only detect solutions "close" to the unperturbed solution, even if there are other solutions for which the perturbative expansion is not valid.[citation needed]

Difficult computations

[edit]

The problem of non-perturbative systems has been somewhat alleviated by the advent of moderncomputers. It has become practical to obtain numerical non-perturbative solutions for certain problems, using methods such asdensity functional theory. These advances have been of particular benefit to the field ofquantum chemistry.[3] Computers have also been used to carry out perturbation theory calculations to extraordinarily high levels of precision, which has proven important inparticle physics for generating theoretical results that can be compared with experiment.

Time-independent perturbation theory

[edit]

Time-independent perturbation theory is one of two categories of perturbation theory, the other being time-dependent perturbation (see next section). In time-independent perturbation theory, the perturbation Hamiltonian is static (i.e., possesses no time dependence). Time-independent perturbation theory was presented byErwin Schrödinger in a 1926 paper,[4] shortly after he produced his theories in wave mechanics. In this paper Schrödinger referred to earlier work ofLord Rayleigh,[5] who investigated harmonic vibrations of a string perturbed by small inhomogeneities. This is why this perturbation theory is often referred to asRayleigh–Schrödinger perturbation theory.[6]

First order corrections

[edit]

The process begins with an unperturbed HamiltonianH0, which is assumed to have no time dependence.[7] It has known energy levels andeigenstates, arising from the time-independentSchrödinger equation:

H0|n(0)=En(0)|n(0),n=1,2,3,{\displaystyle H_{0}\left|n^{(0)}\right\rangle =E_{n}^{(0)}\left|n^{(0)}\right\rangle ,\qquad n=1,2,3,\cdots }

For simplicity, it is assumed that the energies are discrete. The(0) superscripts denote that these quantities are associated with the unperturbed system. Note the use ofbra–ket notation.

A perturbation is then introduced to the Hamiltonian. LetV be a Hamiltonian representing a weak physical disturbance, such as a potential energy produced by an external field. Thus,V is formally aHermitian operator. Letλ be a dimensionless parameter that can take on values ranging continuously from 0 (no perturbation) to 1 (the full perturbation). The perturbed Hamiltonian is:

H=H0+λV{\displaystyle H=H_{0}+\lambda V}

The energy levels and eigenstates of the perturbed Hamiltonian are again given by the time-independent Schrödinger equation,(H0+λV)|n=En|n.{\displaystyle \left(H_{0}+\lambda V\right)|n\rangle =E_{n}|n\rangle .}

The objective is to expressEn and|n{\displaystyle |n\rangle } in terms of the energy levels and eigenstates of the old Hamiltonian. If the perturbation is sufficiently weak, they can be written as a (Maclaurin)power series inλ,En=En(0)+λEn(1)+λ2En(2)+|n=|n(0)+λ|n(1)+λ2|n(2)+{\displaystyle {\begin{aligned}E_{n}&=E_{n}^{(0)}+\lambda E_{n}^{(1)}+\lambda ^{2}E_{n}^{(2)}+\cdots \\[1ex]|n\rangle &=\left|n^{(0)}\right\rangle +\lambda \left|n^{(1)}\right\rangle +\lambda ^{2}\left|n^{(2)}\right\rangle +\cdots \end{aligned}}}whereEn(k)=1k!dkEndλk|λ=0|n(k)=1k!dk|ndλk|λ=0.{\displaystyle {\begin{aligned}E_{n}^{(k)}&={\frac {1}{k!}}{\frac {d^{k}E_{n}}{d\lambda ^{k}}}{\bigg |}_{\lambda =0}\\[1ex]\left|n^{(k)}\right\rangle &=\left.{\frac {1}{k!}}{\frac {d^{k}|n\rangle }{d\lambda ^{k}}}\right|_{\lambda =0.}\end{aligned}}}

Whenk = 0, these reduce to the unperturbed values, which are the first term in each series. Since the perturbation is weak, the energy levels and eigenstates should not deviate too much from their unperturbed values, and the terms should rapidly become smaller as the order is increased.

Substituting the power series expansion into the Schrödinger equation produces:

(H0+λV)(|n(0)+λ|n(1)+)=(En(0)+λEn(1)+)(|n(0)+λ|n(1)+).{\displaystyle \left(H_{0}+\lambda V\right)\left(\left|n^{(0)}\right\rangle +\lambda \left|n^{(1)}\right\rangle +\cdots \right)=\left(E_{n}^{(0)}+\lambda E_{n}^{(1)}+\cdots \right)\left(\left|n^{(0)}\right\rangle +\lambda \left|n^{(1)}\right\rangle +\cdots \right).}

Expanding this equation and comparing coefficients of each power ofλ results in an infinite series ofsimultaneous equations. The zeroth-order equation is simply the Schrödinger equation for the unperturbed system,H0|n(0)=En(0)|n(0).{\displaystyle H_{0}\left|n^{(0)}\right\rangle =E_{n}^{(0)}\left|n^{(0)}\right\rangle .}

The first-order equation isH0|n(1)+V|n(0)=En(0)|n(1)+En(1)|n(0).{\displaystyle H_{0}\left|n^{(1)}\right\rangle +V\left|n^{(0)}\right\rangle =E_{n}^{(0)}\left|n^{(1)}\right\rangle +E_{n}^{(1)}\left|n^{(0)}\right\rangle .}

Operating through byn(0)|{\displaystyle \langle n^{(0)}|}, the first term on the left-hand side cancels the first term on the right-hand side. (Recall, the unperturbed Hamiltonian isHermitian). This leads to the first-order energy shift,En(1)=n(0)|V|n(0).{\displaystyle E_{n}^{(1)}=\left\langle n^{(0)}\right|V\left|n^{(0)}\right\rangle .}This is simply theexpectation value of the perturbation Hamiltonian while the system is in the unperturbed eigenstate.

This result can be interpreted in the following way: supposing that the perturbation is applied, but the system is kept in the quantum state|n(0){\displaystyle |n^{(0)}\rangle }, which is a valid quantum state though no longer an energy eigenstate. The perturbation causes the average energy of this state to increase byn(0)|V|n(0){\displaystyle \langle n^{(0)}|V|n^{(0)}\rangle }. However, the true energy shift is slightly different, because the perturbed eigenstate is not exactly the same as|n(0){\displaystyle |n^{(0)}\rangle }. These further shifts are given by the second and higher order corrections to the energy.

Before corrections to the energy eigenstate are computed, the issue of normalization must be addressed. Supposing thatn(0)|n(0)=1,{\displaystyle \left\langle n^{(0)}\right|\left.n^{(0)}\right\rangle =1,}but perturbation theory also assumes thatn|n=1{\displaystyle \langle n|n\rangle =1}.

Then at first order inλ, the following must be true:(n(0)|+λn(1)|)(|n(0)+λ|n(1))=1{\displaystyle \left(\left\langle n^{(0)}\right|+\lambda \left\langle n^{(1)}\right|\right)\left(\left|n^{(0)}\right\rangle +\lambda \left|n^{(1)}\right\rangle \right)=1}n(0)|n(0)+λn(0)|n(1)+λn(1)|n(0)+λ2n(1)|n(1)=1{\displaystyle \left\langle n^{(0)}\right|\left.n^{(0)}\right\rangle +\lambda \left\langle n^{(0)}\right|\left.n^{(1)}\right\rangle +\lambda \left\langle n^{(1)}\right|\left.n^{(0)}\right\rangle +{\cancel {\lambda ^{2}\left\langle n^{(1)}\right|\left.n^{(1)}\right\rangle }}=1}n(0)|n(1)+n(1)|n(0)=0.{\displaystyle \left\langle n^{(0)}\right|\left.n^{(1)}\right\rangle +\left\langle n^{(1)}\right|\left.n^{(0)}\right\rangle =0.}

Since the overall phase is not determined in quantum mechanics,without loss of generality, in time-independent theory it can be assumed thatn(0)|n(1){\displaystyle \langle n^{(0)}|n^{(1)}\rangle } is purely real. Therefore,n(0)|n(1)=n(1)|n(0)=n(1)|n(0),{\displaystyle \left\langle n^{(0)}\right|\left.n^{(1)}\right\rangle =\left\langle n^{(1)}\right|\left.n^{(0)}\right\rangle =-\left\langle n^{(1)}\right|\left.n^{(0)}\right\rangle ,}leading ton(0)|n(1)=0.{\displaystyle \left\langle n^{(0)}\right|\left.n^{(1)}\right\rangle =0.}

To obtain the first-order correction to the energy eigenstate, the expression for the first-order energy correction is inserted back into the result shown above, equating the first-order coefficients ofλ. Then by using theresolution of the identity:V|n(0)=(kn|k(0)k(0)|)V|n(0)+(|n(0)n(0)|)V|n(0)=kn|k(0)k(0)|V|n(0)+En(1)|n(0),{\displaystyle {\begin{aligned}V\left|n^{(0)}\right\rangle &=\left(\sum _{k\neq n}\left|k^{(0)}\right\rangle \left\langle k^{(0)}\right|\right)V\left|n^{(0)}\right\rangle +\left(\left|n^{(0)}\right\rangle \left\langle n^{(0)}\right|\right)V\left|n^{(0)}\right\rangle \\&=\sum _{k\neq n}\left|k^{(0)}\right\rangle \left\langle k^{(0)}\right|V\left|n^{(0)}\right\rangle +E_{n}^{(1)}\left|n^{(0)}\right\rangle ,\end{aligned}}}where the|k(0){\displaystyle |k^{(0)}\rangle } are in theorthogonal complement of|n(0){\displaystyle |n^{(0)}\rangle }, i.e., the other eigenvectors.

The first-order equation may thus be expressed as(En(0)H0)|n(1)=kn|k(0)k(0)|V|n(0).{\displaystyle \left(E_{n}^{(0)}-H_{0}\right)\left|n^{(1)}\right\rangle =\sum _{k\neq n}\left|k^{(0)}\right\rangle \left\langle k^{(0)}\right|V\left|n^{(0)}\right\rangle .}

Suppose that the zeroth-order energy level is notdegenerate, i.e. that there is no eigenstate ofH0 in the orthogonal complement of|n(0){\displaystyle |n^{(0)}\rangle } with the energyEn(0){\displaystyle E_{n}^{(0)}}. After renaming the summation dummy index above ask{\displaystyle k'}, anykn{\displaystyle k\neq n} can be chosen and multiplying the first-order equation through byk(0)|{\displaystyle \langle k^{(0)}|} gives(En(0)Ek(0))k(0)|n(1)=k(0)|V|n(0).{\displaystyle \left(E_{n}^{(0)}-E_{k}^{(0)}\right)\left\langle k^{(0)}\right.\left|n^{(1)}\right\rangle =\left\langle k^{(0)}\right|V\left|n^{(0)}\right\rangle .}

The abovek(0)|n(1){\displaystyle \langle k^{(0)}|n^{(1)}\rangle } also gives us the component of the first-order correction along|k(0){\displaystyle |k^{(0)}\rangle }.

Thus, in total, the result is,|n(1)=knk(0)|V|n(0)En(0)Ek(0)|k(0).{\displaystyle \left|n^{(1)}\right\rangle =\sum _{k\neq n}{\frac {\left\langle k^{(0)}\right|V\left|n^{(0)}\right\rangle }{E_{n}^{(0)}-E_{k}^{(0)}}}\left|k^{(0)}\right\rangle .}

The first-order change in then-th energy eigenket has a contribution from each of the energy eigenstateskn. Each term is proportional to the matrix elementk(0)|V|n(0){\displaystyle \langle k^{(0)}|V|n^{(0)}\rangle }, which is a measure of how much the perturbation mixes eigenstaten with eigenstatek; it is also inversely proportional to the energy difference between eigenstatesk andn, which means that the perturbation deforms the eigenstate to a greater extent if there are more eigenstates at nearby energies. The expression is singular if any of these states have the same energy as staten, which is why it was assumed that there is no degeneracy. The above formula for the perturbed eigenstates also implies that the perturbation theory can be legitimately used only when the absolute magnitude of the matrix elements of the perturbation is small compared with the corresponding differences in the unperturbed energy levels, i.e.,|k(0)|V|n(0)||En(0)Ek(0)|.{\displaystyle |\langle k^{(0)}|V|n^{(0)}\rangle |\ll |E_{n}^{(0)}-E_{k}^{(0)}|.}

Second-order and higher-order corrections

[edit]

We can find the higher-order deviations by a similar procedure, though the calculations become quite tedious with our current formulation. Our normalization prescription gives that

2n(0)|n(2)+n(1)|n(1)=0.{\displaystyle 2\left\langle n^{(0)}\right|\left.n^{(2)}\right\rangle +\left\langle n^{(1)}\right|\left.n^{(1)}\right\rangle =0.}

Up to second order, the expressions for the energies and (normalized) eigenstates are:

En(λ)=En(0)+λn(0)|V|n(0)+λ2kn|k(0)|V|n(0)|2En(0)Ek(0)+O(λ3){\displaystyle E_{n}(\lambda )=E_{n}^{(0)}+\lambda \left\langle n^{(0)}\right|V\left|n^{(0)}\right\rangle +\lambda ^{2}\sum _{k\neq n}{\frac {\left|\left\langle k^{(0)}\right|V\left|n^{(0)}\right\rangle \right|^{2}}{E_{n}^{(0)}-E_{k}^{(0)}}}+O(\lambda ^{3})}

|n(λ)=|n(0)+λkn|k(0)k(0)|V|n(0)En(0)Ek(0)+λ2knn|k(0)k(0)|V|(0)(0)|V|n(0)(En(0)Ek(0))(En(0)E(0))λ2kn|k(0)k(0)|V|n(0)n(0)|V|n(0)(En(0)Ek(0))212λ2|n(0)kn|k(0)|V|n(0)|2(En(0)Ek(0))2+O(λ3).{\displaystyle {\begin{aligned}|n(\lambda )\rangle =\left|n^{(0)}\right\rangle &+\lambda \sum _{k\neq n}\left|k^{(0)}\right\rangle {\frac {\left\langle k^{(0)}\right|V\left|n^{(0)}\right\rangle }{E_{n}^{(0)}-E_{k}^{(0)}}}+\lambda ^{2}\sum _{k\neq n}\sum _{\ell \neq n}\left|k^{(0)}\right\rangle {\frac {\left\langle k^{(0)}\right|V\left|\ell ^{(0)}\right\rangle \left\langle \ell ^{(0)}\right|V\left|n^{(0)}\right\rangle }{\left(E_{n}^{(0)}-E_{k}^{(0)}\right)\left(E_{n}^{(0)}-E_{\ell }^{(0)}\right)}}\\[1ex]&-\lambda ^{2}\sum _{k\neq n}\left|k^{(0)}\right\rangle {\frac {\left\langle k^{(0)}\right|V\left|n^{(0)}\right\rangle \left\langle n^{(0)}\right|V\left|n^{(0)}\right\rangle }{\left(E_{n}^{(0)}-E_{k}^{(0)}\right)^{2}}}-{\frac {1}{2}}\lambda ^{2}\left|n^{(0)}\right\rangle \sum _{k\neq n}{\frac {|\left\langle k^{(0)}\right|V\left|n^{(0)}\right\rangle |^{2}}{\left(E_{n}^{(0)}-E_{k}^{(0)}\right)^{2}}}+O(\lambda ^{3}).\end{aligned}}}If an intermediate normalization is taken (in other words, if it is required thatn(0)|n(λ)=1{\displaystyle \langle n^{(0)}|n(\lambda )\rangle =1}), then we obtain a nearly identical expression for the second-order correction to the correction given immediately above. To be precise, for an intermediate normalization, the last term would be omitted.

Extending the process further, the third-order energy correction can be shown to be[8]

En(3)=knmnn(0)|V|m(0)m(0)|V|k(0)k(0)|V|n(0)(En(0)Em(0))(En(0)Ek(0))n(0)|V|n(0)mn|n(0)|V|m(0)|2(En(0)Em(0))2.{\displaystyle E_{n}^{(3)}=\sum _{k\neq n}\sum _{m\neq n}{\frac {\langle n^{(0)}|V|m^{(0)}\rangle \langle m^{(0)}|V|k^{(0)}\rangle \langle k^{(0)}|V|n^{(0)}\rangle }{\left(E_{n}^{(0)}-E_{m}^{(0)}\right)\left(E_{n}^{(0)}-E_{k}^{(0)}\right)}}-\langle n^{(0)}|V|n^{(0)}\rangle \sum _{m\neq n}{\frac {|\langle n^{(0)}|V|m^{(0)}\rangle |^{2}}{\left(E_{n}^{(0)}-E_{m}^{(0)}\right)^{2}}}.}

Corrections to fifth order (energies) and fourth order (states) in compact notation

If we introduce the notation,

Vnmn(0)|V|m(0),{\displaystyle V_{nm}\equiv \langle n^{(0)}|V|m^{(0)}\rangle ,}EnmEn(0)Em(0),{\displaystyle E_{nm}\equiv E_{n}^{(0)}-E_{m}^{(0)},}

then the energy corrections to fifth order can be written

En(1)=VnnEn(2)=|Vnk2|2Enk2En(3)=Vnk3Vk3k2Vk2nEnk2Enk3Vnn|Vnk3|2Enk32En(4)=Vnk4Vk4k3Vk3k2Vk2nEnk2Enk3Enk4|Vnk4|2Enk42|Vnk2|2Enk2VnnVnk4Vk4k3Vk3nEnk32Enk4VnnVnk4Vk4k2Vk2nEnk2Enk42+Vnn2|Vnk4|2Enk43=Vnk4Vk4k3Vk3k2Vk2nEnk2Enk3Enk4En(2)|Vnk4|2Enk422VnnVnk4Vk4k3Vk3nEnk32Enk4+Vnn2|Vnk4|2Enk43En(5)=Vnk5Vk5k4Vk4k3Vk3k2Vk2nEnk2Enk3Enk4Enk5Vnk5Vk5k4Vk4nEnk42Enk5|Vnk2|2Enk2Vnk5Vk5k2Vk2nEnk2Enk52|Vnk2|2Enk2|Vnk5|2Enk52Vnk3Vk3k2Vk2nEnk2Enk3VnnVnk5Vk5k4Vk4k3Vk3nEnk32Enk4Enk5VnnVnk5Vk5k4Vk4k2Vk2nEnk2Enk42Enk5VnnVnk5Vk5k3Vk3k2Vk2nEnk2Enk3Enk52+Vnn|Vnk5|2Enk52|Vnk3|2Enk32+2Vnn|Vnk5|2Enk53|Vnk2|2Enk2+Vnn2Vnk5Vk5k4Vk4nEnk43Enk5+Vnn2Vnk5Vk5k3Vk3nEnk32Enk52+Vnn2Vnk5Vk5k2Vk2nEnk2Enk53Vnn3|Vnk5|2Enk54=Vnk5Vk5k4Vk4k3Vk3k2Vk2nEnk2Enk3Enk4Enk52En(2)Vnk5Vk5k4Vk4nEnk42Enk5|Vnk5|2Enk52Vnk3Vk3k2Vk2nEnk2Enk3+Vnn(2Vnk5Vk5k4Vk4k3Vk3nEnk32Enk4Enk5Vnk5Vk5k4Vk4k2Vk2nEnk2Enk42Enk5+|Vnk5|2Enk52|Vnk3|2Enk32+2En(2)|Vnk5|2Enk53)+Vnn2(2Vnk5Vk5k4Vk4nEnk43Enk5+Vnk5Vk5k3Vk3nEnk32Enk52)Vnn3|Vnk5|2Enk54{\displaystyle {\begin{aligned}E_{n}^{(1)}&=V_{nn}\\E_{n}^{(2)}&={\frac {|V_{nk_{2}}|^{2}}{E_{nk_{2}}}}\\E_{n}^{(3)}&={\frac {V_{nk_{3}}V_{k_{3}k_{2}}V_{k_{2}n}}{E_{nk_{2}}E_{nk_{3}}}}-V_{nn}{\frac {|V_{nk_{3}}|^{2}}{E_{nk_{3}}^{2}}}\\E_{n}^{(4)}&={\frac {V_{nk_{4}}V_{k_{4}k_{3}}V_{k_{3}k_{2}}V_{k_{2}n}}{E_{nk_{2}}E_{nk_{3}}E_{nk_{4}}}}-{\frac {|V_{nk_{4}}|^{2}}{E_{nk_{4}}^{2}}}{\frac {|V_{nk_{2}}|^{2}}{E_{nk_{2}}}}-V_{nn}{\frac {V_{nk_{4}}V_{k_{4}k_{3}}V_{k_{3}n}}{E_{nk_{3}}^{2}E_{nk_{4}}}}-V_{nn}{\frac {V_{nk_{4}}V_{k_{4}k_{2}}V_{k_{2}n}}{E_{nk_{2}}E_{nk_{4}}^{2}}}+V_{nn}^{2}{\frac {|V_{nk_{4}}|^{2}}{E_{nk_{4}}^{3}}}\\&={\frac {V_{nk_{4}}V_{k_{4}k_{3}}V_{k_{3}k_{2}}V_{k_{2}n}}{E_{nk_{2}}E_{nk_{3}}E_{nk_{4}}}}-E_{n}^{(2)}{\frac {|V_{nk_{4}}|^{2}}{E_{nk_{4}}^{2}}}-2V_{nn}{\frac {V_{nk_{4}}V_{k_{4}k_{3}}V_{k_{3}n}}{E_{nk_{3}}^{2}E_{nk_{4}}}}+V_{nn}^{2}{\frac {|V_{nk_{4}}|^{2}}{E_{nk_{4}}^{3}}}\\E_{n}^{(5)}&={\frac {V_{nk_{5}}V_{k_{5}k_{4}}V_{k_{4}k_{3}}V_{k_{3}k_{2}}V_{k_{2}n}}{E_{nk_{2}}E_{nk_{3}}E_{nk_{4}}E_{nk_{5}}}}-{\frac {V_{nk_{5}}V_{k_{5}k_{4}}V_{k_{4}n}}{E_{nk_{4}}^{2}E_{nk_{5}}}}{\frac {|V_{nk_{2}}|^{2}}{E_{nk_{2}}}}-{\frac {V_{nk_{5}}V_{k_{5}k_{2}}V_{k_{2}n}}{E_{nk_{2}}E_{nk_{5}}^{2}}}{\frac {|V_{nk_{2}}|^{2}}{E_{nk_{2}}}}-{\frac {|V_{nk_{5}}|^{2}}{E_{nk_{5}}^{2}}}{\frac {V_{nk_{3}}V_{k_{3}k_{2}}V_{k_{2}n}}{E_{nk_{2}}E_{nk_{3}}}}\\&\quad -V_{nn}{\frac {V_{nk_{5}}V_{k_{5}k_{4}}V_{k_{4}k_{3}}V_{k_{3}n}}{E_{nk_{3}}^{2}E_{nk_{4}}E_{nk_{5}}}}-V_{nn}{\frac {V_{nk_{5}}V_{k_{5}k_{4}}V_{k_{4}k_{2}}V_{k_{2}n}}{E_{nk_{2}}E_{nk_{4}}^{2}E_{nk_{5}}}}-V_{nn}{\frac {V_{nk_{5}}V_{k_{5}k_{3}}V_{k_{3}k_{2}}V_{k_{2}n}}{E_{nk_{2}}E_{nk_{3}}E_{nk_{5}}^{2}}}+V_{nn}{\frac {|V_{nk_{5}}|^{2}}{E_{nk_{5}}^{2}}}{\frac {|V_{nk_{3}}|^{2}}{E_{nk_{3}}^{2}}}+2V_{nn}{\frac {|V_{nk_{5}}|^{2}}{E_{nk_{5}}^{3}}}{\frac {|V_{nk_{2}}|^{2}}{E_{nk_{2}}}}\\&\quad +V_{nn}^{2}{\frac {V_{nk_{5}}V_{k_{5}k_{4}}V_{k_{4}n}}{E_{nk_{4}}^{3}E_{nk_{5}}}}+V_{nn}^{2}{\frac {V_{nk_{5}}V_{k_{5}k_{3}}V_{k_{3}n}}{E_{nk_{3}}^{2}E_{nk_{5}}^{2}}}+V_{nn}^{2}{\frac {V_{nk_{5}}V_{k_{5}k_{2}}V_{k_{2}n}}{E_{nk_{2}}E_{nk_{5}}^{3}}}-V_{nn}^{3}{\frac {|V_{nk_{5}}|^{2}}{E_{nk_{5}}^{4}}}\\&={\frac {V_{nk_{5}}V_{k_{5}k_{4}}V_{k_{4}k_{3}}V_{k_{3}k_{2}}V_{k_{2}n}}{E_{nk_{2}}E_{nk_{3}}E_{nk_{4}}E_{nk_{5}}}}-2E_{n}^{(2)}{\frac {V_{nk_{5}}V_{k_{5}k_{4}}V_{k_{4}n}}{E_{nk_{4}}^{2}E_{nk_{5}}}}-{\frac {|V_{nk_{5}}|^{2}}{E_{nk_{5}}^{2}}}{\frac {V_{nk_{3}}V_{k_{3}k_{2}}V_{k_{2}n}}{E_{nk_{2}}E_{nk_{3}}}}\\&\quad +V_{nn}\left(-2{\frac {V_{nk_{5}}V_{k_{5}k_{4}}V_{k_{4}k_{3}}V_{k_{3}n}}{E_{nk_{3}}^{2}E_{nk_{4}}E_{nk_{5}}}}-{\frac {V_{nk_{5}}V_{k_{5}k_{4}}V_{k_{4}k_{2}}V_{k_{2}n}}{E_{nk_{2}}E_{nk_{4}}^{2}E_{nk_{5}}}}+{\frac {|V_{nk_{5}}|^{2}}{E_{nk_{5}}^{2}}}{\frac {|V_{nk_{3}}|^{2}}{E_{nk_{3}}^{2}}}+2E_{n}^{(2)}{\frac {|V_{nk_{5}}|^{2}}{E_{nk_{5}}^{3}}}\right)\\&\quad +V_{nn}^{2}\left(2{\frac {V_{nk_{5}}V_{k_{5}k_{4}}V_{k_{4}n}}{E_{nk_{4}}^{3}E_{nk_{5}}}}+{\frac {V_{nk_{5}}V_{k_{5}k_{3}}V_{k_{3}n}}{E_{nk_{3}}^{2}E_{nk_{5}}^{2}}}\right)-V_{nn}^{3}{\frac {|V_{nk_{5}}|^{2}}{E_{nk_{5}}^{4}}}\end{aligned}}}and the states to fourth order can be written|n(1)=Vk1nEnk1|k1(0)|n(2)=(Vk1k2Vk2nEnk1Enk2VnnVk1nEnk12)|k1(0)12Vnk1Vk1nEk1n2|n(0)|n(3)=[Vk1k2Vk2k3Vk3nEk1nEnk2Enk3+VnnVk1k2Vk2nEk1nEnk2(1Enk1+1Enk2)|Vnn|2Vk1nEk1n3+|Vnk2|2Vk1nEk1nEnk2(1Enk1+12Enk2)]|k1(0)+[Vnk2Vk2k1Vk1n+Vk2nVk1k2Vnk12Enk22Enk1+|Vnk1|2VnnEnk13]|n(0)|n(4)=[Vk1k2Vk2k3Vk3k4Vk4k2+Vk3k2Vk1k2Vk4k3Vk2k42Ek1nEk2k32Ek2k4Vk2k3Vk3k4Vk4nVk1k2Ek1nEk2nEnk3Enk4+Vk1k2Ek1n(|Vk2k3|2Vk2k2Ek2k33|Vnk3|2Vk2nEk3n2Ek2n)+VnnVk1k2Vk3nVk2k3Ek1nEnk3Ek2n(1Enk3+1Ek2n+1Ek1n)+|Vk2n|2Vk1k3Enk2Ek1n(Vk3nEnk1Enk3Vk3k1Ek3k12)Vnn(Vk3k2Vk1k3Vk2k1+Vk3k1Vk2k3Vk1k2)2Ek1nEk1k32Ek1k2+|Vnn|2Ek1n(Vk1nVnnEk1n3+Vk1k2Vk2nEk2n3)|Vk1k2|2VnnVk1nEk1nEk1k23]|k1(0)+12[Vnk1Vk1k2Enk1Ek2n2(Vk2nVnnEk2nVk2k3Vk3nEnk3)Vk1nVk2k1Ek1n2Enk2(Vk3k2Vnk3Enk3+VnnVnk2Enk2)+|Vnk1|2Ek1n2(3|Vnk2|24Ek2n22|Vnn|2Ek1n2)Vk2k3Vk3k1|Vnk1|2Enk32Enk1Enk2]|n(0){\displaystyle {\begin{aligned}|n^{(1)}\rangle &={\frac {V_{k_{1}n}}{E_{nk_{1}}}}|k_{1}^{(0)}\rangle \\|n^{(2)}\rangle &=\left({\frac {V_{k_{1}k_{2}}V_{k_{2}n}}{E_{nk_{1}}E_{nk_{2}}}}-{\frac {V_{nn}V_{k_{1}n}}{E_{nk_{1}}^{2}}}\right)|k_{1}^{(0)}\rangle -{\frac {1}{2}}{\frac {V_{nk_{1}}V_{k_{1}n}}{E_{k_{1}n}^{2}}}|n^{(0)}\rangle \\|n^{(3)}\rangle &={\Bigg [}-{\frac {V_{k_{1}k_{2}}V_{k_{2}k_{3}}V_{k_{3}n}}{E_{k_{1}n}E_{nk_{2}}E_{nk_{3}}}}+{\frac {V_{nn}V_{k_{1}k_{2}}V_{k_{2}n}}{E_{k_{1}n}E_{nk_{2}}}}\left({\frac {1}{E_{nk_{1}}}}+{\frac {1}{E_{nk_{2}}}}\right)-{\frac {|V_{nn}|^{2}V_{k_{1}n}}{E_{k_{1}n}^{3}}}+{\frac {|V_{nk_{2}}|^{2}V_{k_{1}n}}{E_{k_{1}n}E_{nk_{2}}}}\left({\frac {1}{E_{nk_{1}}}}+{\frac {1}{2E_{nk_{2}}}}\right){\Bigg ]}|k_{1}^{(0)}\rangle \\&\quad +{\Bigg [}-{\frac {V_{nk_{2}}V_{k_{2}k_{1}}V_{k_{1}n}+V_{k_{2}n}V_{k_{1}k_{2}}V_{nk_{1}}}{2E_{nk_{2}}^{2}E_{nk_{1}}}}+{\frac {|V_{nk_{1}}|^{2}V_{nn}}{E_{nk_{1}}^{3}}}{\Bigg ]}|n^{(0)}\rangle \\|n^{(4)}\rangle &={\Bigg [}{\frac {V_{k_{1}k_{2}}V_{k_{2}k_{3}}V_{k_{3}k_{4}}V_{k_{4}k_{2}}+V_{k_{3}k_{2}}V_{k_{1}k_{2}}V_{k_{4}k_{3}}V_{k_{2}k_{4}}}{2E_{k_{1}n}E_{k_{2}k_{3}}^{2}E_{k_{2}k_{4}}}}-{\frac {V_{k_{2}k_{3}}V_{k_{3}k_{4}}V_{k_{4}n}V_{k_{1}k_{2}}}{E_{k_{1}n}E_{k_{2}n}E_{nk_{3}}E_{nk_{4}}}}+{\frac {V_{k_{1}k_{2}}}{E_{k_{1}n}}}\left({\frac {|V_{k_{2}k_{3}}|^{2}V_{k_{2}k_{2}}}{E_{k_{2}k_{3}}^{3}}}-{\frac {|V_{nk_{3}}|^{2}V_{k_{2}n}}{E_{k_{3}n}^{2}E_{k_{2}n}}}\right)\\&\quad +{\frac {V_{nn}V_{k_{1}k_{2}}V_{k_{3}n}V_{k_{2}k_{3}}}{E_{k_{1}n}E_{nk_{3}}E_{k_{2}n}}}\left({\frac {1}{E_{nk_{3}}}}+{\frac {1}{E_{k_{2}n}}}+{\frac {1}{E_{k_{1}n}}}\right)+{\frac {|V_{k_{2}n}|^{2}V_{k_{1}k_{3}}}{E_{nk_{2}}E_{k_{1}n}}}\left({\frac {V_{k_{3}n}}{E_{nk_{1}}E_{nk_{3}}}}-{\frac {V_{k_{3}k_{1}}}{E_{k_{3}k_{1}}^{2}}}\right)-{\frac {V_{nn}\left(V_{k_{3}k_{2}}V_{k_{1}k_{3}}V_{k_{2}k_{1}}+V_{k_{3}k_{1}}V_{k_{2}k_{3}}V_{k_{1}k_{2}}\right)}{2E_{k_{1}n}E_{k_{1}k_{3}}^{2}E_{k_{1}k_{2}}}}\\&\quad +{\frac {|V_{nn}|^{2}}{E_{k_{1}n}}}\left({\frac {V_{k_{1}n}V_{nn}}{E_{k_{1}n}^{3}}}+{\frac {V_{k_{1}k_{2}}V_{k_{2}n}}{E_{k_{2}n}^{3}}}\right)-{\frac {|V_{k_{1}k_{2}}|^{2}V_{nn}V_{k_{1}n}}{E_{k_{1}n}E_{k_{1}k_{2}}^{3}}}{\Bigg ]}|k_{1}^{(0)}\rangle +{\frac {1}{2}}\left[{\frac {V_{nk_{1}}V_{k_{1}k_{2}}}{E_{nk_{1}}E_{k_{2}n}^{2}}}\left({\frac {V_{k_{2}n}V_{nn}}{E_{k_{2}n}}}-{\frac {V_{k_{2}k_{3}}V_{k_{3}n}}{E_{nk_{3}}}}\right)\right.\\&\quad \left.-{\frac {V_{k_{1}n}V_{k_{2}k_{1}}}{E_{k_{1}n}^{2}E_{nk_{2}}}}\left({\frac {V_{k_{3}k_{2}}V_{nk_{3}}}{E_{nk_{3}}}}+{\frac {V_{nn}V_{nk_{2}}}{E_{nk_{2}}}}\right)+{\frac {|V_{nk_{1}}|^{2}}{E_{k_{1}n}^{2}}}\left({\frac {3|V_{nk_{2}}|^{2}}{4E_{k_{2}n}^{2}}}-{\frac {2|V_{nn}|^{2}}{E_{k_{1}n}^{2}}}\right)-{\frac {V_{k_{2}k_{3}}V_{k_{3}k_{1}}|V_{nk_{1}}|^{2}}{E_{nk_{3}}^{2}E_{nk_{1}}E_{nk_{2}}}}\right]|n^{(0)}\rangle \end{aligned}}}

All terms involvedkj should be summed overkj such that the denominator does not vanish.

It is possible to relate thek-th order correction to the energyEn to thek-pointconnected correlation function of the perturbationV in the state|n(0){\displaystyle |n^{(0)}\rangle }. Fork=2{\displaystyle k=2}, one has to consider the inverseLaplace transformρn,2(s){\displaystyle \rho _{n,2}(s)} of the two-point correlator:n(0)|V(τ)V(0)|n(0)n(0)|V|n(0)2=:Rdsρn,2(s)e(sEn(0))τ{\displaystyle \langle n^{(0)}|V(\tau )V(0)|n^{(0)}\rangle -\langle n^{(0)}|V|n^{(0)}\rangle ^{2}=\mathrel {\mathop {:} } \int _{\mathbb {R} }\!ds\;\rho _{n,2}(s)\,e^{-(s-E_{n}^{(0)})\tau }}whereV(τ)=eH0τVeH0τ{\displaystyle V(\tau )=e^{H_{0}\tau }Ve^{-H_{0}\tau }} is the perturbing operatorV in the interaction picture, evolving in Euclidean time. ThenEn(2)=RdssEn(0)ρn,2(s).{\displaystyle E_{n}^{(2)}=-\int _{\mathbb {R} }\!{\frac {ds}{s-E_{n}^{(0)}}}\,\rho _{n,2}(s).}

Similar formulas exist to all orders in perturbation theory, allowing one to expressEn(k){\displaystyle E_{n}^{(k)}} in terms of the inverse Laplace transformρn,k{\displaystyle \rho _{n,k}} of the connected correlation functionn(0)|V(τ1++τk1)V(τ1+τ2)V(τ1)V(0)|n(0)conn=n(0)|V(τ1++τk1)V(τ1+τ2)V(τ1)V(0)|n(0)subtractions.{\displaystyle \langle n^{(0)}|V(\tau _{1}+\ldots +\tau _{k-1})\dotsm V(\tau _{1}+\tau _{2})V(\tau _{1})V(0)|n^{(0)}\rangle _{\text{conn}}=\langle n^{(0)}|V(\tau _{1}+\ldots +\tau _{k-1})\dotsm V(\tau _{1}+\tau _{2})V(\tau _{1})V(0)|n^{(0)}\rangle -{\text{subtractions}}.}

To be precise, if we writen(0)|V(τ1++τk1)V(τ1+τ2)V(τ1)V(0)|n(0)conn=Ri=1k1dsie(siEn(0))τiρn,k(s1,,sk1){\displaystyle \langle n^{(0)}|V(\tau _{1}+\ldots +\tau _{k-1})\dotsm V(\tau _{1}+\tau _{2})V(\tau _{1})V(0)|n^{(0)}\rangle _{\text{conn}}=\int _{\mathbb {R} }\,\prod _{i=1}^{k-1}ds_{i}\,e^{-(s_{i}-E_{n}^{(0)})\tau _{i}}\,\rho _{n,k}(s_{1},\ldots ,s_{k-1})\,}then thek-th order energy shift is given by[9]

En(k)=(1)k1Ri=1k1dsisiEn(0)ρn,k(s1,,sk1).{\displaystyle E_{n}^{(k)}=(-1)^{k-1}\int _{\mathbb {R} }\,\prod _{i=1}^{k-1}{\frac {ds_{i}}{s_{i}-E_{n}^{(0)}}}\,\rho _{n,k}(s_{1},\ldots ,s_{k-1}).}

Effects of degeneracy

[edit]

Suppose that two or more energy eigenstates of the unperturbed Hamiltonian aredegenerate. The first-order energy shift is not well defined, since there is no unique way to choose a basis of eigenstates for the unperturbed system. The various eigenstates for a given energy will perturb with different energies, or may well possess no continuous family of perturbations at all.

This is manifested in the calculation of the perturbed eigenstate via the fact that the operatorEn(0)H0{\displaystyle E_{n}^{(0)}-H_{0}}does not have a well-defined inverse.

LetD denote the subspace spanned by these degenerate eigenstates. No matter how small the perturbation is, in the degenerate subspaceD the energy differences between the eigenstates ofH are non-zero, so complete mixing of at least some of these states is assured. Typically, the eigenvalues will split, and the eigenspaces will become simple (one-dimensional), or at least of smaller dimension thanD.

The successful perturbations will not be "small" relative to a poorly chosen basis ofD. Instead, we consider the perturbation "small" if the new eigenstate is close to the subspaceD. The new Hamiltonian must be diagonalized inD, or a slight variation ofD, so to speak. These perturbed eigenstates inD are now the basis for the perturbation expansion,|n=kDαnk|k(0)+λ|n(1).{\displaystyle |n\rangle =\sum _{k\in D}\alpha _{nk}|k^{(0)}\rangle +\lambda |n^{(1)}\rangle .}

For the first-order perturbation, we need solve the perturbed Hamiltonian restricted to the degenerate subspaceD,V|k(0)=ϵk|k(0)+small|k(0)D,{\displaystyle V|k^{(0)}\rangle =\epsilon _{k}|k^{(0)}\rangle +{\text{small}}\qquad \forall |k^{(0)}\rangle \in D,}simultaneously for all the degenerate eigenstates, whereϵk{\displaystyle \epsilon _{k}} are first-order corrections to the degenerate energy levels, and "small" is a vector ofO(λ){\displaystyle O(\lambda )} orthogonal toD. This amounts to diagonalizing the matrixk(0)|V|l(0)=Vkl|k(0),|l(0)D.{\displaystyle \langle k^{(0)}|V|l^{(0)}\rangle =V_{kl}\qquad \forall \;|k^{(0)}\rangle ,|l^{(0)}\rangle \in D.}

This procedure is approximate, since we neglected states outside theD subspace ("small"). The splitting of degenerate energiesϵk{\displaystyle \epsilon _{k}} is generally observed. Although the splitting may be small,O(λ){\displaystyle O(\lambda )}, compared to the range of energies found in the system, it is crucial in understanding certain details, such as spectral lines inElectron Spin Resonance experiments.

Higher-order corrections due to other eigenstates outsideD can be found in the same way as for the non-degenerate case,(En(0)H0)|n(1)=kD(k(0)|V|n(0))|k(0).{\displaystyle \left(E_{n}^{(0)}-H_{0}\right)|n^{(1)}\rangle =\sum _{k\not \in D}\left(\langle k^{(0)}|V|n^{(0)}\rangle \right)|k^{(0)}\rangle .}

The operator on the left-hand side is not singular when applied to eigenstates outsideD, so we can write|n(1)=kDk(0)|V|n(0)En(0)Ek(0)|k(0),{\displaystyle |n^{(1)}\rangle =\sum _{k\not \in D}{\frac {\langle k^{(0)}|V|n^{(0)}\rangle }{E_{n}^{(0)}-E_{k}^{(0)}}}|k^{(0)}\rangle ,}but the effect on the degenerate states is ofO(λ){\displaystyle O(\lambda )}.

Near-degenerate states should also be treated similarly, when the original Hamiltonian splits aren't larger than the perturbation in the near-degenerate subspace. An application is found in thenearly free electron model, where near-degeneracy, treated properly, gives rise to an energy gap even for small perturbations. Other eigenstates will only shift the absolute energy of all near-degenerate states simultaneously.

Degeneracy lifted to first order

[edit]

Let us consider degenerate energy eigenstates and a perturbation that completely lifts the degeneracy to first order of correction.

The perturbed Hamiltonian is denoted asH^=H^0+λV^,{\displaystyle {\hat {H}}={\hat {H}}_{0}+\lambda {\hat {V}}\,,}whereH^0{\displaystyle {\hat {H}}_{0}} is the unperturbed Hamiltonian,V^{\displaystyle {\hat {V}}} is the perturbation operator, and0<λ<1{\displaystyle 0<\lambda <1} is the parameter of the perturbation.

Let us focus on the degeneracy of then{\displaystyle n}-th unperturbed energyEn(0){\displaystyle E_{n}^{(0)}}. We will denote the unperturbed states in this degenerate subspace as|ψnk(0){\displaystyle \left|\psi _{nk}^{(0)}\right\rangle } and the other unperturbed states as|ψm(0){\displaystyle \left|\psi _{m}^{(0)}\right\rangle }, wherek{\displaystyle k} is the index of the unperturbed state in the degenerate subspace andmn{\displaystyle m\neq n} represents all other energy eigenstates with energies different fromEn(0){\displaystyle E_{n}^{(0)}}. The eventual degeneracy among the other states withmn{\displaystyle \forall m\neq n} does not change our arguments. All states|ψnk(0){\displaystyle \left|\psi _{nk}^{(0)}\right\rangle } with various values ofk{\displaystyle k} share the same energyEn(0){\displaystyle E_{n}^{(0)}} when there is no perturbation, i.e., whenλ=0{\displaystyle \lambda =0}. The energiesEm(0){\displaystyle E_{m}^{(0)}} of the other states|ψm(0){\displaystyle \left|\psi _{m}^{(0)}\right\rangle } withmn{\displaystyle m\neq n} are all different fromEn(0){\displaystyle E_{n}^{(0)}}, but not necessarily unique, i.e. not necessarily always different among themselves.

ByVnl,nk{\displaystyle V_{nl,nk}} andVm,nk{\displaystyle V_{m,nk}}, we denote the matrix elements of the perturbation operatorV^{\displaystyle {\hat {V}}} in the basis of the unperturbed eigenstates. We assume that the basis vectors|ψnk(0){\displaystyle \left|\psi _{nk}^{(0)}\right\rangle } in the degenerate subspace are chosen such that the matrix elementsVnl,nkψnl(0)|V^|ψnk(0){\displaystyle V_{nl,nk}\equiv \left\langle \psi _{nl}^{(0)}\right|{\hat {V}}\left|\psi _{nk}^{(0)}\right\rangle } are diagonal. Assuming also that the degeneracy is completely lifted to the first order, i.e. thatEnl(1)Enk(1){\displaystyle E_{nl}^{(1)}\neq E_{nk}^{(1)}} iflk{\displaystyle l\neq k}, we have the following formulae for the energy correction to the second order inλ{\displaystyle \lambda }Enk=En0+λVnk,nk+λ2mn|Vm,nk|2En(0)Em(0)+O(λ3),{\displaystyle E_{nk}=E_{n}^{0}+\lambda V_{nk,nk}+\lambda ^{2}\sum \limits _{m\neq n}{\frac {\left|V_{m,nk}\right|^{2}}{E_{n}^{(0)}-E_{m}^{(0)}}}+{\mathcal {O}}(\lambda ^{3})\,,}and for the state correction to the first order inλ{\displaystyle \lambda }|ψnk(1)=|ψnk(0)+λmnVm,nkEm(0)En(0)(|ψm(0)+lkVnl,mEnl(1)Enk(1)|ψnl(0))+O(λ2).{\displaystyle \left|\psi _{nk}^{(1)}\right\rangle =\left|\psi _{nk}^{(0)}\right\rangle +\lambda \sum \limits _{m\neq n}{\frac {V_{m,nk}}{E_{m}^{(0)}-E_{n}^{(0)}}}\left(-\left|\psi _{m}^{(0)}\right\rangle +\sum \limits _{l\neq k}{\frac {V_{nl,m}}{E_{nl}^{(1)}-E_{nk}^{(1)}}}\left|\psi _{nl}^{(0)}\right\rangle \right)+{\mathcal {O}}(\lambda ^{2})\,.}

Notice that here the first order correction to the state is orthogonal to the unperturbed state,ψnk(0)|ψnk(1)=0.{\displaystyle \left\langle \psi _{nk}^{(0)}|\psi _{nk}^{(1)}\right\rangle =0\,.}

Generalization to multi-parameter case

[edit]

The generalization of time-independent perturbation theory to the case where there are multiple small parametersxμ=(x1,x2,){\displaystyle x^{\mu }=(x^{1},x^{2},\cdots )} in place of λ can be formulated more systematically using the language ofdifferential geometry, which basically defines the derivatives of the quantum states and calculates the perturbative corrections by taking derivatives iteratively at the unperturbed point.

Hamiltonian and force operator

[edit]

From the differential geometric point of view, a parameterized Hamiltonian is considered as a function defined on the parametermanifold that maps each particular set of parameters(x1,x2,){\displaystyle (x^{1},x^{2},\cdots )} to an Hermitian operatorH(x μ) that acts on the Hilbert space. The parameters here can be external field, interaction strength, or driving parameters in thequantum phase transition. LetEn(x μ) and|n(xμ){\displaystyle |n(x^{\mu })\rangle } be then-th eigenenergy and eigenstate ofH(x μ) respectively. In the language of differential geometry, the states|n(xμ){\displaystyle |n(x^{\mu })\rangle } form avector bundle over the parameter manifold, on which derivatives of these states can be defined. The perturbation theory is to answer the following question: givenEn(x0μ){\displaystyle E_{n}(x_{0}^{\mu })} and|n(x0μ){\displaystyle |n(x_{0}^{\mu })\rangle } at an unperturbed reference pointx0μ{\displaystyle x_{0}^{\mu }}, how to estimate theEn(x μ) and|n(xμ){\displaystyle |n(x^{\mu })\rangle } atx μ close to that reference point.

Without loss of generality, the coordinate system can be shifted, such that the reference pointx0μ=0{\displaystyle x_{0}^{\mu }=0} is set to be the origin. The following linearly parameterized Hamiltonian is frequently usedH(xμ)=H(0)+xμFμ.{\displaystyle H(x^{\mu })=H(0)+x^{\mu }F_{\mu }.}

If the parametersx μ are considered as generalized coordinates, thenFμ should be identified as the generalized force operators related to those coordinates. Different indicesμ label the different forces along different directions in the parameter manifold. For example, ifx μ denotes the external magnetic field in theμ-direction, thenFμ should be the magnetization in the same direction.

Perturbation theory as power series expansion

[edit]

The validity of perturbation theory lies on the adiabatic assumption, which assumes the eigenenergies and eigenstates of the Hamiltonian are smooth functions of parameters such that their values in the vicinity region can be calculated in power series (likeTaylor expansion) of the parameters:

En(xμ)=En+xμμEn+12!xμxνμνEn+|n(xμ)=|n+xμ|μn+12!xμxν|μνn+{\displaystyle {\begin{aligned}E_{n}(x^{\mu })&=E_{n}+x^{\mu }\partial _{\mu }E_{n}+{\frac {1}{2!}}x^{\mu }x^{\nu }\partial _{\mu }\partial _{\nu }E_{n}+\cdots \\[1ex]\left|n(x^{\mu })\right\rangle &=\left|n\right\rangle +x^{\mu }\left|\partial _{\mu }n\right\rangle +{\frac {1}{2!}}x^{\mu }x^{\nu }\left|\partial _{\mu }\partial _{\nu }n\right\rangle +\cdots \end{aligned}}}

Hereμ denotes the derivative with respect tox μ. When applying to the state|μn{\displaystyle |\partial _{\mu }n\rangle }, it should be understood as thecovariant derivative if the vector bundle is equipped with non-vanishingconnection. All the terms on the right-hand-side of the series are evaluated atx μ = 0, e.g.EnEn(0) and|n|n(0){\displaystyle |n\rangle \equiv |n(0)\rangle }. This convention will be adopted throughout this subsection, that all functions without the parameter dependence explicitly stated are assumed to be evaluated at the origin. The power series may converge slowly or even not converge when the energy levels are close to each other. The adiabatic assumption breaks down when there is energy level degeneracy, and hence the perturbation theory is not applicable in that case.

Hellmann–Feynman theorems

[edit]

The above power series expansion can be readily evaluated if there is a systematic approach to calculate the derivates to any order. Using thechain rule, the derivatives can be broken down to the single derivative on either the energy or the state. TheHellmann–Feynman theorems are used to calculate these single derivatives. The first Hellmann–Feynman theorem gives the derivative of the energy,μEn=n|μH|n{\displaystyle \partial _{\mu }E_{n}=\langle n|\partial _{\mu }H|n\rangle }

The second Hellmann–Feynman theorem gives the derivative of the state (resolved by the complete basis withmn),m|μn=m|μH|nEnEm,μm|n=m|μH|nEmEn.{\displaystyle \langle m|\partial _{\mu }n\rangle ={\frac {\langle m|\partial _{\mu }H|n\rangle }{E_{n}-E_{m}}},\qquad \langle \partial _{\mu }m|n\rangle ={\frac {\langle m|\partial _{\mu }H|n\rangle }{E_{m}-E_{n}}}.}

For the linearly parameterized Hamiltonian,μH simply stands for the generalized force operatorFμ.

The theorems can be simply derived by applying the differential operatorμ to both sides of theSchrödinger equationH|n=En|n,{\displaystyle H|n\rangle =E_{n}|n\rangle ,} which reads

μH|n+H|μn=μEn|n+En|μn.{\displaystyle \partial _{\mu }H|n\rangle +H|\partial _{\mu }n\rangle =\partial _{\mu }E_{n}|n\rangle +E_{n}|\partial _{\mu }n\rangle .}

Then overlap with the statem|{\displaystyle \langle m|} from left and make use of the Schrödinger equationm|H=m|Em{\displaystyle \langle m|H=\langle m|E_{m}} again,

m|μH|n+Emm|μn=μEnm|n+Enm|μn.{\displaystyle \langle m|\partial _{\mu }H|n\rangle +E_{m}\langle m|\partial _{\mu }n\rangle =\partial _{\mu }E_{n}\langle m|n\rangle +E_{n}\langle m|\partial _{\mu }n\rangle .}

Given that the eigenstates of the Hamiltonian always form an orthonormal basism|n=δmn{\displaystyle \langle m|n\rangle =\delta _{mn}}, the cases ofm =n andmn can be discussed separately. The first case will lead to the first theorem and the second case to the second theorem, which can be shown immediately by rearranging the terms. With the differential rules given by the Hellmann–Feynman theorems, the perturbative correction to the energies and states can be calculated systematically.

Correction of energy and state

[edit]

To the second order, the energy correction reads

En(xμ)=n|H|n+n|μH|nxμ+mnn|νH|mm|μH|nEnEmxμxν+,{\displaystyle E_{n}(x^{\mu })=\langle n|H|n\rangle +\langle n|\partial _{\mu }H|n\rangle x^{\mu }+\Re \sum _{m\neq n}{\frac {\langle n|\partial _{\nu }H|m\rangle \langle m|\partial _{\mu }H|n\rangle }{E_{n}-E_{m}}}x^{\mu }x^{\nu }+\cdots ,}where{\displaystyle \Re } denotes thereal part function.The first order derivativeμEn is given by the first Hellmann–Feynman theorem directly. To obtain the second order derivativeμνEn, simply applying the differential operatorμ to the result of the first order derivativen|νH|n{\displaystyle \langle n|\partial _{\nu }H|n\rangle }, which reads

μνEn=μn|νH|n+n|μνH|n+n|νH|μn.{\displaystyle \partial _{\mu }\partial _{\nu }E_{n}=\langle \partial _{\mu }n|\partial _{\nu }H|n\rangle +\langle n|\partial _{\mu }\partial _{\nu }H|n\rangle +\langle n|\partial _{\nu }H|\partial _{\mu }n\rangle .}

Note that for a linearly parameterized Hamiltonian, there is no second derivativeμνH = 0 on the operator level. Resolve the derivative of state by inserting the complete set of basis,μνEn=m(μn|mm|νH|n+n|νH|mm|μn),{\displaystyle \partial _{\mu }\partial _{\nu }E_{n}=\sum _{m}\left(\langle \partial _{\mu }n|m\rangle \langle m|\partial _{\nu }H|n\rangle +\langle n|\partial _{\nu }H|m\rangle \langle m|\partial _{\mu }n\rangle \right),}then all parts can be calculated using the Hellmann–Feynman theorems. In terms of Lie derivatives,μn|n=n|μn=0{\displaystyle \langle \partial _{\mu }n|n\rangle =\langle n|\partial _{\mu }n\rangle =0} according to the definition of the connection for the vector bundle. Therefore, the casem =n can be excluded from the summation, which avoids the singularity of the energy denominator. The same procedure can be carried on for higher order derivatives, from which higher order corrections are obtained.

The same computational scheme is applicable for the correction of states. The result to the second order is as follows|n(xμ)=|n+mnm|μH|nEnEm|mxμ+(mnlnm|μH|ll|νH|n(EnEm)(EnEl)|mmnm|μH|nn|νH|n(EnEm)2|m12mnn|μH|mm|νH|n(EnEm)2|n)xμxν+.{\displaystyle {\begin{aligned}\left|n\left(x^{\mu }\right)\right\rangle =|n\rangle &+\sum _{m\neq n}{\frac {\langle m|\partial _{\mu }H|n\rangle }{E_{n}-E_{m}}}|m\rangle x^{\mu }\\&+\left(\sum _{m\neq n}\sum _{l\neq n}{\frac {\langle m|\partial _{\mu }H|l\rangle \langle l|\partial _{\nu }H|n\rangle }{(E_{n}-E_{m})(E_{n}-E_{l})}}|m\rangle -\sum _{m\neq n}{\frac {\langle m|\partial _{\mu }H|n\rangle \langle n|\partial _{\nu }H|n\rangle }{(E_{n}-E_{m})^{2}}}|m\rangle -{\frac {1}{2}}\sum _{m\neq n}{\frac {\langle n|\partial _{\mu }H|m\rangle \langle m|\partial _{\nu }H|n\rangle }{(E_{n}-E_{m})^{2}}}|n\rangle \right)x^{\mu }x^{\nu }+\cdots .\end{aligned}}}

Both energy derivatives and state derivatives will be involved in deduction. Whenever a state derivative is encountered, resolve it by inserting the complete set of basis, then the Hellmann-Feynman theorem is applicable. Because differentiation can be calculated systematically, the series expansion approach to the perturbative corrections can be coded on computers with symbolic processing software likeMathematica.

Effective Hamiltonian

[edit]

LetH(0) be the Hamiltonian completely restricted either in the low-energy subspaceHL{\displaystyle {\mathcal {H}}_{L}} or in the high-energy subspaceHH{\displaystyle {\mathcal {H}}_{H}}, such that there is no matrix element inH(0) connecting the low- and the high-energy subspaces, i.e.m|H(0)|l=0{\displaystyle \langle m|H(0)|l\rangle =0} ifmHL,lHH{\displaystyle m\in {\mathcal {H}}_{L},l\in {\mathcal {H}}_{H}}. LetFμ = ∂μH be the coupling terms connecting the subspaces. Then when the high energy degrees of freedoms are integrated out, the effective Hamiltonian in the low energy subspace reads[10]

Hmneff(xμ)=m|H|n+δnmm|μH|nxμ+12!lHH(m|μH|ll|νH|nEmEl+m|νH|ll|μH|nEnEl)xμxν+.{\displaystyle H_{mn}^{\text{eff}}\left(x^{\mu }\right)=\langle m|H|n\rangle +\delta _{nm}\langle m|\partial _{\mu }H|n\rangle x^{\mu }+{\frac {1}{2!}}\sum _{l\in {\mathcal {H}}_{H}}\left({\frac {\langle m|\partial _{\mu }H|l\rangle \langle l|\partial _{\nu }H|n\rangle }{E_{m}-E_{l}}}+{\frac {\langle m|\partial _{\nu }H|l\rangle \langle l|\partial _{\mu }H|n\rangle }{E_{n}-E_{l}}}\right)x^{\mu }x^{\nu }+\cdots .}

Herem,nHL{\displaystyle m,n\in {\mathcal {H}}_{L}} are restricted in the low energy subspace. The above result can be derived by power series expansion ofm|H(xμ)|n{\displaystyle \langle m|H(x^{\mu })|n\rangle }.

In a formal way it is possible to define an effective Hamiltonian that gives exactly the low-lying energy states and wavefunctions.[11] In practice, some kind of approximation (perturbation theory) is generally required.

Time-dependent perturbation theory

[edit]

Method of variation of constants

[edit]

Time-dependent perturbation theory, initiated byPaul Dirac and further developed byJohn Archibald Wheeler,Richard Feynman, andFreeman Dyson,[12] studies the effect of a time-dependent perturbationV(t) applied to a time-independent HamiltonianH0.[13] It is an extremely valuable tool for calculating the properties of any physical system. It is used for the quantitative description of phenomena as diverse as proton-proton scattering, photo-ionization of materials, scattering of electrons off lattice defects in a conductor, scattering of neutrons off nuclei, electric susceptibilities of materials, neutron absorption cross sections in a nuclear reactor, and much more.[12]

Since the perturbed Hamiltonian is time-dependent, so are its energy levels and eigenstates. Thus, the goals of time-dependent perturbation theory are slightly different from time-independent perturbation theory. One is interested in the following quantities:

  • The time-dependentexpectation value of some observableA, for a given initial state.
  • The time-dependent expansion coefficients (w.r.t. a given time-dependent state) of those basis states that are energy eigenkets (eigenvectors) in the unperturbed system.

The first quantity is important because it gives rise to theclassical result of anA measurement performed on a macroscopic number of copies of the perturbed system. For example, we could takeA to be the displacement in thex-direction of the electron in a hydrogen atom, in which case the expected value, when multiplied by an appropriate coefficient, gives the time-dependentdielectric polarization of a hydrogen gas. With an appropriate choice of perturbation (i.e. an oscillating electric potential), this allows one to calculate the ACpermittivity of the gas.

The second quantity looks at the time-dependent probability of occupation for each eigenstate. This is particularly useful inlaser physics, where one is interested in the populations of different atomic states in a gas when a time-dependent electric field is applied. These probabilities are also useful for calculating the "quantum broadening" ofspectral lines (seeline broadening) andparticle decay inparticle physics andnuclear physics.

We will briefly examine the method behind Dirac's formulation of time-dependent perturbation theory. Choose an energy basis|n{\displaystyle {|n\rangle }} for the unperturbed system. (We drop the (0) superscripts for the eigenstates, because it is not useful to speak of energy levels and eigenstates for the perturbed system.)

If the unperturbed system is an eigenstate (of the Hamiltonian)|j{\displaystyle |j\rangle } at timet = 0, its state at subsequent times varies only by aphase (in theSchrödinger picture, where state vectors evolve in time and operators are constant),|j(t)=eiEjt/|j .{\displaystyle |j(t)\rangle =e^{-iE_{j}t/\hbar }|j\rangle ~.}

Now, introduce a time-dependent perturbing HamiltonianV(t). The Hamiltonian of the perturbed system isH=H0+V(t) .{\displaystyle H=H_{0}+V(t)~.}Let|ψ(t){\displaystyle |\psi (t)\rangle } denote the quantum state of the perturbed system at timet. It obeys the time-dependent Schrödinger equation,H|ψ(t)=it|ψ(t) .{\displaystyle H|\psi (t)\rangle =i\hbar {\frac {\partial }{\partial t}}|\psi (t)\rangle ~.}

The quantum state at each instant can be expressed as a linear combination of the complete eigenbasis of|n{\displaystyle |n\rangle }:

|ψ(t)=ncn(t)eiEnt/|n ,{\displaystyle |\psi (t)\rangle =\sum _{n}c_{n}(t)e^{-iE_{n}t/\hbar }|n\rangle ~,}1

where thecn(t)s are to be determinedcomplex functions oft which we will refer to asamplitudes (strictly speaking, they are the amplitudes in theDirac picture).

We have explicitly extracted the exponential phase factorsexp(iEnt/){\displaystyle \exp(-iE_{n}t/\hbar )} on the right hand side. This is only a matter of convention, and may be done without loss of generality. The reason we go to this trouble is that when the system starts in the state|j{\displaystyle |j\rangle } and no perturbation is present, the amplitudes have the convenient property that, for allt,cj(t) = 1 andcn(t) = 0 ifn ≠ j.

The square of the absolute amplitudecn(t) is the probability that the system is in staten at timet, since|cn(t)|2=|n|ψ(t)|2 .{\displaystyle \left|c_{n}(t)\right|^{2}=\left|\langle n|\psi (t)\rangle \right|^{2}~.}

Plugging into the Schrödinger equation and using the fact that ∂/∂t acts by aproduct rule, one obtainsn(idcndtcn(t)V(t))eiEnt/|n=0 .{\displaystyle \sum _{n}\left(i\hbar {\frac {dc_{n}}{dt}}-c_{n}(t)V(t)\right)e^{-iE_{n}t/\hbar }|n\rangle =0~.}

By resolving the identity in front ofV and multiplying through by thebran|{\displaystyle \langle n|} on the left, this can be reduced to a set of coupleddifferential equations for the amplitudes,dcndt=ikn|V(t)|kck(t)ei(EkEn)t/ .{\displaystyle {\frac {dc_{n}}{dt}}={\frac {-i}{\hbar }}\sum _{k}\langle n|V(t)|k\rangle \,c_{k}(t)\,e^{-i(E_{k}-E_{n})t/\hbar }~.}

where we have used equation (1) to evaluate the sum onn in the second term, then used the fact thatk|Ψ(t)=ck(t)eiEkt/{\displaystyle \langle k|\Psi (t)\rangle =c_{k}(t)e^{-iE_{k}t/\hbar }}.

The matrix elements ofV play a similar role as in time-independent perturbation theory, being proportional to the rate at which amplitudes are shifted between states. Note, however, that the direction of the shift is modified by the exponential phase factor. Over times much longer than the energy differenceEkEn, the phase winds around 0 several times. If the time-dependence ofV is sufficiently slow, this may cause the state amplitudes to oscillate. (For example, such oscillations are useful for managing radiative transitions in alaser.)

Up to this point, we have made no approximations, so this set of differential equations is exact. By supplying appropriate initial valuescn(t), we could in principle find an exact (i.e., non-perturbative) solution. This is easily done when there are only two energy levels (n = 1, 2), and this solution is useful for modelling systems like theammonia molecule.

However, exact solutions are difficult to find when there are many energy levels, and one instead looks for perturbative solutions. These may be obtained by expressing the equations in an integral form,cn(t)=cn(0)ik0tdtn|V(t)|kck(t)ei(EkEn)t/ .{\displaystyle c_{n}(t)=c_{n}(0)-{\frac {i}{\hbar }}\sum _{k}\int _{0}^{t}dt'\;\langle n|V(t')|k\rangle \,c_{k}(t')\,e^{-i(E_{k}-E_{n})t'/\hbar }~.}

Repeatedly substituting this expression forcn back into right hand side, yields an iterative solution,cn(t)=cn(0)+cn(1)+cn(2)+{\displaystyle c_{n}(t)=c_{n}^{(0)}+c_{n}^{(1)}+c_{n}^{(2)}+\cdots }where, for example, the first-order term iscn(1)(t)=ik0tdtn|V(t)|kck(0)ei(EkEn)t/ .{\displaystyle c_{n}^{(1)}(t)={\frac {-i}{\hbar }}\sum _{k}\int _{0}^{t}dt'\;\langle n|V(t')|k\rangle \,c_{k}^{(0)}\,e^{-i(E_{k}-E_{n})t'/\hbar }~.}To the same approximation, the summation in the above expression can be removed since in the unperturbed stateck(0)=δkn{\displaystyle c_{k}^{(0)}=\delta _{kn}} so that we havecn(1)(t)=i0tdtn|V(t)|kei(EkEn)t/ .{\displaystyle c_{n}^{(1)}(t)={\frac {-i}{\hbar }}\int _{0}^{t}dt'\;\langle n|V(t')|k\rangle \,e^{-i(E_{k}-E_{n})t'/\hbar }~.}

Several further results follow from this, such asFermi's golden rule, which relates the rate of transitions between quantum states to the density of states at particular energies; or theDyson series, obtained by applying the iterative method to thetime evolution operator, which is one of the starting points for the method ofFeynman diagrams.

Method of Dyson series

[edit]

Time-dependent perturbations can be reorganized through the technique of theDyson series. TheSchrödinger equationH(t)|ψ(t)=i|ψ(t)t{\displaystyle H(t)|\psi (t)\rangle =i\hbar {\frac {\partial |\psi (t)\rangle }{\partial t}}}has the formal solution|ψ(t)=Texp[it0tdtH(t)]|ψ(t0) ,{\displaystyle |\psi (t)\rangle =T\exp {\left[-{\frac {i}{\hbar }}\int _{t_{0}}^{t}dt'H(t')\right]}|\psi (t_{0})\rangle ~,}whereT is the time ordering operator,TA(t1)A(t2)={A(t1)A(t2)t1>t2A(t2)A(t1)t2>t1 .{\displaystyle TA(t_{1})A(t_{2})={\begin{cases}A(t_{1})A(t_{2})&t_{1}>t_{2}\\A(t_{2})A(t_{1})&t_{2}>t_{1}\end{cases}}~.}Thus, the exponential represents the followingDyson series,|ψ(t)=[1it0tdt1H(t1)12t0tdt1t0t1dt2H(t1)H(t2)+]|ψ(t0) .{\displaystyle |\psi (t)\rangle =\left[1-{\frac {i}{\hbar }}\int _{t_{0}}^{t}dt_{1}H(t_{1})-{\frac {1}{\hbar ^{2}}}\int _{t_{0}}^{t}dt_{1}\int _{t_{0}}^{t_{1}}dt_{2}H(t_{1})H(t_{2})+\ldots \right]|\psi (t_{0})\rangle ~.}Note that in the second term, the 1/2! factor exactly cancels the double contribution due to the time-ordering operator, etc.

Consider the following perturbation problem[H0+λV(t)]|ψ(t)=i|ψ(t)t ,{\displaystyle [H_{0}+\lambda V(t)]|\psi (t)\rangle =i\hbar {\frac {\partial |\psi (t)\rangle }{\partial t}}~,}assuming that the parameterλ is small and that the problemH0|n=En|n{\displaystyle H_{0}|n\rangle =E_{n}|n\rangle } has been solved.

Perform the following unitary transformation to theinteraction picture (or Dirac picture),|ψ(t)=eiH0(tt0)|ψI(t) .{\displaystyle |\psi (t)\rangle =e^{-{\frac {i}{\hbar }}H_{0}(t-t_{0})}|\psi _{I}(t)\rangle ~.}Consequently, theSchrödinger equation simplifies toλeiH0(tt0)V(t)eiH0(tt0)|ψI(t)=i|ψI(t)t ,{\displaystyle \lambda e^{{\frac {i}{\hbar }}H_{0}(t-t_{0})}V(t)e^{-{\frac {i}{\hbar }}H_{0}(t-t_{0})}|\psi _{I}(t)\rangle =i\hbar {\frac {\partial |\psi _{I}(t)\rangle }{\partial t}}~,}so it is solved through the aboveDyson series,|ψI(t)=[1iλt0tdt1eiH0(t1t0)V(t1)eiH0(t1t0)λ22t0tdt1t0t1dt2eiH0(t1t0)V(t1)eiH0(t1t0)eiH0(t2t0)V(t2)eiH0(t2t0)+]|ψ(t0) ,{\displaystyle |\psi _{I}(t)\rangle =\left[1-{\frac {i\lambda }{\hbar }}\int _{t_{0}}^{t}dt_{1}e^{{\frac {i}{\hbar }}H_{0}(t_{1}-t_{0})}V(t_{1})e^{-{\frac {i}{\hbar }}H_{0}(t_{1}-t_{0})}-{\frac {\lambda ^{2}}{\hbar ^{2}}}\int _{t_{0}}^{t}dt_{1}\int _{t_{0}}^{t_{1}}dt_{2}e^{{\frac {i}{\hbar }}H_{0}(t_{1}-t_{0})}V(t_{1})e^{-{\frac {i}{\hbar }}H_{0}(t_{1}-t_{0})}e^{{\frac {i}{\hbar }}H_{0}(t_{2}-t_{0})}V(t_{2})e^{-{\frac {i}{\hbar }}H_{0}(t_{2}-t_{0})}+\ldots \right]|\psi (t_{0})\rangle ~,}as a perturbation series with smallλ.

Using the solution of the unperturbed problemH0|n=En|n{\displaystyle H_{0}|n\rangle =E_{n}|n\rangle } andn|nn|=1{\displaystyle \sum _{n}|n\rangle \langle n|=1} (for the sake of simplicity assume a pure discrete spectrum), yields, to first order,|ψI(t)=[1iλmnt0tdt1m|V(t1)|nei(EnEm)(t1t0)|mn|+]|ψ(t0) .{\displaystyle |\psi _{I}(t)\rangle =\left[1-{\frac {i\lambda }{\hbar }}\sum _{m}\sum _{n}\int _{t_{0}}^{t}dt_{1}\langle m|V(t_{1})|n\rangle e^{-{\frac {i}{\hbar }}(E_{n}-E_{m})(t_{1}-t_{0})}|m\rangle \langle n|+\ldots \right]|\psi (t_{0})\rangle ~.}

Thus, the system, initially in the unperturbed state|α=|ψ(t0){\displaystyle |\alpha \rangle =|\psi (t_{0})\rangle }, by dint of the perturbation can go into the state|β{\displaystyle |\beta \rangle }. The corresponding transition probability amplitude to first order isAαβ=iλt0tdt1β|V(t1)|αei(EαEβ)(t1t0) ,{\displaystyle A_{\alpha \beta }=-{\frac {i\lambda }{\hbar }}\int _{t_{0}}^{t}dt_{1}\langle \beta |V(t_{1})|\alpha \rangle e^{-{\frac {i}{\hbar }}(E_{\alpha }-E_{\beta })(t_{1}-t_{0})}~,}as detailed in the previous section——while the corresponding transition probability to a continuum is furnished byFermi's golden rule.

As an aside, note that time-independent perturbation theory is also organized inside this time-dependent perturbation theory Dyson series. To see this, write the unitary evolution operator, obtained from the aboveDyson series, asU(t)=1iλt0tdt1eiH0(t1t0)V(t1)eiH0(t1t0)λ22t0tdt1t0t1dt2eiH0(t1t0)V(t1)eiH0(t1t0)eiH0(t2t0)V(t2)eiH0(t2t0)+{\displaystyle U(t)=1-{\frac {i\lambda }{\hbar }}\int _{t_{0}}^{t}dt_{1}e^{{\frac {i}{\hbar }}H_{0}(t_{1}-t_{0})}V(t_{1})e^{-{\frac {i}{\hbar }}H_{0}(t_{1}-t_{0})}-{\frac {\lambda ^{2}}{\hbar ^{2}}}\int _{t_{0}}^{t}dt_{1}\int _{t_{0}}^{t_{1}}dt_{2}e^{{\frac {i}{\hbar }}H_{0}(t_{1}-t_{0})}V(t_{1})e^{-{\frac {i}{\hbar }}H_{0}(t_{1}-t_{0})}e^{{\frac {i}{\hbar }}H_{0}(t_{2}-t_{0})}V(t_{2})e^{-{\frac {i}{\hbar }}H_{0}(t_{2}-t_{0})}+\cdots }and take the perturbationV to be time-independent.

Using the identity resolutionn|nn|=1{\displaystyle \sum _{n}|n\rangle \langle n|=1}withH0|n=En|n{\displaystyle H_{0}|n\rangle =E_{n}|n\rangle } for a pure discrete spectrum, writeU(t)=1[iλt0tdt1mnm|V|nei(EnEm)(t1t0)|mn|][λ22t0tdt1t0t1dt2mnqei(EnEm)(t1t0)m|V|nn|V|qei(EqEn)(t2t0)|mq|]+{\displaystyle {\begin{aligned}U(t)=1&-\left[{\frac {i\lambda }{\hbar }}\int _{t_{0}}^{t}dt_{1}\sum _{m}\sum _{n}\langle m|V|n\rangle e^{-{\frac {i}{\hbar }}(E_{n}-E_{m})(t_{1}-t_{0})}|m\rangle \langle n|\right]\\[5mu]&-\left[{\frac {\lambda ^{2}}{\hbar ^{2}}}\int _{t_{0}}^{t}dt_{1}\int _{t_{0}}^{t_{1}}dt_{2}\sum _{m}\sum _{n}\sum _{q}e^{-{\frac {i}{\hbar }}(E_{n}-E_{m})(t_{1}-t_{0})}\langle m|V|n\rangle \langle n|V|q\rangle e^{-{\frac {i}{\hbar }}(E_{q}-E_{n})(t_{2}-t_{0})}|m\rangle \langle q|\right]+\cdots \end{aligned}}}

It is evident that, at second order, one must sum on all the intermediate states. Assumet0=0{\displaystyle t_{0}=0} and the asymptotic limit of larger times. This means that, at each contribution of the perturbation series, one has to add a multiplicative factoreϵt{\displaystyle e^{-\epsilon t}} in the integrands forε arbitrarily small. Thus the limitt → ∞ gives back the final state of the system by eliminating all oscillating terms, but keeping the secular ones. The integrals are thus computable, and, separating the diagonal terms from the others yieldsU(t)=1iλnn|V|ntiλ2mnn|V|mm|V|nEnEmt12λ22m,nn|V|mm|V|nt2++λmnm|V|nEnEm|mn|+λ2mnqnnm|V|nn|V|q(EnEm)(EqEn)|mq|+{\displaystyle {\begin{aligned}U(t)=1&-{\frac {i\lambda }{\hbar }}\sum _{n}\langle n|V|n\rangle t-{\frac {i\lambda ^{2}}{\hbar }}\sum _{m\neq n}{\frac {\langle n|V|m\rangle \langle m|V|n\rangle }{E_{n}-E_{m}}}t-{\frac {1}{2}}{\frac {\lambda ^{2}}{\hbar ^{2}}}\sum _{m,n}\langle n|V|m\rangle \langle m|V|n\rangle t^{2}+\cdots \\&+\lambda \sum _{m\neq n}{\frac {\langle m|V|n\rangle }{E_{n}-E_{m}}}|m\rangle \langle n|+\lambda ^{2}\sum _{m\neq n}\sum _{q\neq n}\sum _{n}{\frac {\langle m|V|n\rangle \langle n|V|q\rangle }{(E_{n}-E_{m})(E_{q}-E_{n})}}|m\rangle \langle q|+\cdots \end{aligned}}}where the time secular series yields the eigenvalues of the perturbed problem specified above, recursively; whereas the remaining time-constant part yields the corrections to the stationary eigenfunctions also given above (|n(λ)=U(0;λ)|n){\displaystyle |n(\lambda )\rangle =U(0;\lambda )|n\rangle )}.)

The unitary evolution operator is applicable to arbitrary eigenstates of the unperturbed problem and, in this case, yields a secular series that holds at small times.

Strong perturbation theory

[edit]

In a similar way as for small perturbations, it is possible to develop a strong perturbation theory. Consider as usual theSchrödinger equation

H(t)|ψ(t)=i|ψ(t)t{\displaystyle H(t)|\psi (t)\rangle =i\hbar {\frac {\partial |\psi (t)\rangle }{\partial t}}}

and we consider the question if a dual Dyson series exists that applies in the limit of a perturbation increasingly large. This question can be answered in an affirmative way[14] and the series is the well-known adiabatic series.[15] This approach is quite general and can be shown in the following way. Consider the perturbation problem

[H0+λV(t)]|ψ(t)=i|ψ(t)t{\displaystyle [H_{0}+\lambda V(t)]|\psi (t)\rangle =i\hbar {\frac {\partial |\psi (t)\rangle }{\partial t}}}

beingλ→ ∞. Our aim is to find a solution in the form

|ψ=|ψ0+1λ|ψ1+1λ2|ψ2+{\displaystyle |\psi \rangle =|\psi _{0}\rangle +{\frac {1}{\lambda }}|\psi _{1}\rangle +{\frac {1}{\lambda ^{2}}}|\psi _{2}\rangle +\ldots }

but a direct substitution into the above equation fails to produce useful results. This situation can be adjusted making a rescaling of the time variable asτ=λt{\displaystyle \tau =\lambda t} producing the following meaningful equations

V(t)|ψ0=i|ψ0τV(t)|ψ1+H0|ψ0=i|ψ1τ{\displaystyle {\begin{aligned}V(t)|\psi _{0}\rangle &=i\hbar {\frac {\partial |\psi _{0}\rangle }{\partial \tau }}\\[1ex]V(t)|\psi _{1}\rangle +H_{0}|\psi _{0}\rangle &=i\hbar {\frac {\partial |\psi _{1}\rangle }{\partial \tau }}\\[1ex]&\;\,\vdots \end{aligned}}}

that can be solved once we know the solution of theleading order equation. But we know that in this case we can use theadiabatic approximation. WhenV(t){\displaystyle V(t)} does not depend on time one gets theWigner-Kirkwood series that is often used instatistical mechanics. Indeed, in this case we introduce the unitary transformation

|ψ(t)=eiλV(tt0)|ψF(t){\displaystyle |\psi (t)\rangle =e^{-{\frac {i}{\hbar }}\lambda V(t-t_{0})}|\psi _{F}(t)\rangle }

that defines afree picture as we are trying to eliminate the interaction term. Now, in dual way with respect to the small perturbations, we have to solve theSchrödinger equation

eiλV(tt0)H0eiλV(tt0)|ψF(t)=i|ψF(t)t{\displaystyle e^{{\frac {i}{\hbar }}\lambda V(t-t_{0})}H_{0}e^{-{\frac {i}{\hbar }}\lambda V(t-t_{0})}|\psi _{F}(t)\rangle =i\hbar {\frac {\partial |\psi _{F}(t)\rangle }{\partial t}}}

and we see that the expansion parameterλ appears only into the exponential and so, the correspondingDyson series, adual Dyson series, is meaningful at largeλs and is

|ψF(t)=[1it0tdt1eiλV(t1t0)H0eiλV(t1t0)12t0tdt1t0t1dt2eiλV(t1t0)H0eiλV(t1t0)eiλV(t2t0)H0eiλV(t2t0)+]|ψ(t0).{\displaystyle |\psi _{F}(t)\rangle =\left[1-{\frac {i}{\hbar }}\int _{t_{0}}^{t}dt_{1}e^{{\frac {i}{\hbar }}\lambda V(t_{1}-t_{0})}H_{0}e^{-{\frac {i}{\hbar }}\lambda V(t_{1}-t_{0})}-{\frac {1}{\hbar ^{2}}}\int _{t_{0}}^{t}dt_{1}\int _{t_{0}}^{t_{1}}dt_{2}e^{{\frac {i}{\hbar }}\lambda V(t_{1}-t_{0})}H_{0}e^{-{\frac {i}{\hbar }}\lambda V(t_{1}-t_{0})}e^{{\frac {i}{\hbar }}\lambda V(t_{2}-t_{0})}H_{0}e^{-{\frac {i}{\hbar }}\lambda V(t_{2}-t_{0})}+\cdots \right]|\psi (t_{0})\rangle .}

After the rescaling in timeτ=λt{\displaystyle \tau =\lambda t} we can see that this is indeed a series in1/λ{\displaystyle 1/\lambda } justifying in this way the name ofdual Dyson series. The reason is that we have obtained this series simply interchangingH0 andV and we can go from one to another applying this exchange. This is calledduality principle in perturbation theory. The choiceH0=p2/2m{\displaystyle H_{0}=p^{2}/2m} yields, as already said, aWigner-Kirkwood series that is a gradient expansion. TheWigner-Kirkwood series is a semiclassical series with eigenvalues given exactly as forWKB approximation.[16]

Examples

[edit]

Example of first-order perturbation theory – ground-state energy of the quartic oscillator

[edit]

Consider the quantum harmonic oscillator with the quartic potential perturbation and the HamiltonianH=22m2x2+mω2x22+λx4.{\displaystyle H=-{\frac {\hbar ^{2}}{2m}}{\frac {\partial ^{2}}{\partial x^{2}}}+{\frac {m\omega ^{2}x^{2}}{2}}+\lambda x^{4}.}

The ground state of the harmonic oscillator isψ0=(απ)14eαx2/2{\displaystyle \psi _{0}=\left({\frac {\alpha }{\pi }}\right)^{\frac {1}{4}}e^{-\alpha x^{2}/2}}(α=mω/{\displaystyle \alpha =m\omega /\hbar }), and the energy of unperturbed ground state isE0(0)=12ω{\displaystyle E_{0}^{(0)}={\tfrac {1}{2}}\hbar \omega }

Using the first-order correction formula, we getE0(1)=λ(απ)12eαx2/2x4eαx2/2dx=λ(απ)122α2eαx2dx,{\displaystyle E_{0}^{(1)}=\lambda \left({\frac {\alpha }{\pi }}\right)^{\frac {1}{2}}\int e^{-\alpha x^{2}/2}x^{4}e^{-\alpha x^{2}/2}dx=\lambda \left({\frac {\alpha }{\pi }}\right)^{\frac {1}{2}}{\frac {\partial ^{2}}{\partial \alpha ^{2}}}\int e^{-\alpha x^{2}}dx,}orE0(1)=λ(απ)122α2(πα)12=λ341α2=342λm2ω2.{\displaystyle E_{0}^{(1)}=\lambda \left({\frac {\alpha }{\pi }}\right)^{\frac {1}{2}}{\frac {\partial ^{2}}{\partial \alpha ^{2}}}\left({\frac {\pi }{\alpha }}\right)^{\frac {1}{2}}=\lambda {\frac {3}{4}}{\frac {1}{\alpha ^{2}}}={\frac {3}{4}}{\frac {\hbar ^{2}\lambda }{m^{2}\omega ^{2}}}.}

Example of first- and second-order perturbation theory – quantum pendulum

[edit]

Consider the quantum-mathematical pendulum with the HamiltonianH=22ma22ϕ2λcosϕ{\displaystyle H=-{\frac {\hbar ^{2}}{2ma^{2}}}{\frac {\partial ^{2}}{\partial \phi ^{2}}}-\lambda \cos \phi }with the potential energyλcosϕ{\displaystyle -\lambda \cos \phi } taken as the perturbation i.e.V=cosϕ.{\displaystyle V=-\cos \phi .}

The unperturbed normalized quantum wave functions are those of the rigid rotor and are given byψn(ϕ)=einϕ2π,{\displaystyle \psi _{n}(\phi )={\frac {e^{in\phi }}{\sqrt {2\pi }}},}and the energiesEn(0)=2n22ma2.{\displaystyle E_{n}^{(0)}={\frac {\hbar ^{2}n^{2}}{2ma^{2}}}.}

The first-order energy correction to the rotor due to the potential energy isEn(1)=12πeinϕcosϕeinϕ=12πcosϕ=0.{\displaystyle E_{n}^{(1)}=-{\frac {1}{2\pi }}\int e^{-in\phi }\cos \phi e^{in\phi }=-{\frac {1}{2\pi }}\int \cos \phi =0.}

Using the formula for the second-order correction, one getsEn(2)=ma22π22k|eikϕcosϕeinϕdϕ|2n2k2,{\displaystyle E_{n}^{(2)}={\frac {ma^{2}}{2\pi ^{2}\hbar ^{2}}}\sum _{k}{\frac {\left|\int e^{-ik\phi }\cos \phi e^{in\phi }\,d\phi \right|^{2}}{n^{2}-k^{2}}},}orEn(2)=ma222k|(δn,1k+δn,1k)|2n2k2,{\displaystyle E_{n}^{(2)}={\frac {ma^{2}}{2\hbar ^{2}}}\sum _{k}{\frac {\left|\left(\delta _{n,1-k}+\delta _{n,-1-k}\right)\right|^{2}}{n^{2}-k^{2}}},}orEn(2)=ma222(12n1+12n1)=ma2214n21.{\displaystyle E_{n}^{(2)}={\frac {ma^{2}}{2\hbar ^{2}}}\left({\frac {1}{2n-1}}+{\frac {1}{-2n-1}}\right)={\frac {ma^{2}}{\hbar ^{2}}}{\frac {1}{4n^{2}-1}}.}

Potential energy as a perturbation

[edit]

When the unperturbed state is a free motion of a particle with kinetic energyE{\displaystyle E}, the solution of theSchrödinger equation2ψ(0)+k2ψ(0)=0{\displaystyle \nabla ^{2}\psi ^{(0)}+k^{2}\psi ^{(0)}=0}corresponds to plane waves with wavenumberk=2mE/2{\textstyle k={\sqrt {2mE/\hbar ^{2}}}}. If there is a weak potential energyU(x,y,z){\displaystyle U(x,y,z)} present in the space, in the first approximation, the perturbed state is described by the equation2ψ(1)+k2ψ(1)=2mU2ψ(0),{\displaystyle \nabla ^{2}\psi ^{(1)}+k^{2}\psi ^{(1)}={\frac {2mU}{\hbar ^{2}}}\psi ^{(0)},}whose particular integral is[17]ψ(1)(x,y,z)=m2π2ψ(0)U(x,y,z)eikrrdxdydz,{\displaystyle \psi ^{(1)}(x,y,z)=-{\frac {m}{2\pi \hbar ^{2}}}\int \psi ^{(0)}U(x',y',z'){\frac {e^{ikr}}{r}}\,dx'dy'dz',}wherer2=(xx)2+(yy)2+(zz)2{\displaystyle r^{2}=(x-x')^{2}+(y-y')^{2}+(z-z')^{2}}. In the two-dimensional case, the solution isψ(1)(x,y)=im22ψ(0)U(x,y)H0(1)(kr)dxdy,{\displaystyle \psi ^{(1)}(x,y)=-{\frac {im}{2\hbar ^{2}}}\int \psi ^{(0)}U(x',y')H_{0}^{(1)}(kr)\,dx'dy',}wherer2=(xx)2+(yy)2{\displaystyle r^{2}=(x-x')^{2}+(y-y')^{2}} andH0(1){\displaystyle H_{0}^{(1)}} is theHankel function of the first kind. In the one-dimensional case, the solution isψ(1)(x)=im2ψ(0)U(x)eikrkdx,{\displaystyle \psi ^{(1)}(x)=-{\frac {im}{\hbar ^{2}}}\int \psi ^{(0)}U(x'){\frac {e^{ikr}}{k}}\,dx',}wherer=|xx|{\displaystyle r=|x-x'|}.

Applications

[edit]

References

[edit]
  1. ^Simon, Barry (1982). "Large orders and summability of eigenvalue perturbation theory: A mathematical overview".International Journal of Quantum Chemistry.21:3–25.doi:10.1002/qua.560210103.
  2. ^Aoyama, Tatsumi; Hayakawa, Masashi; Kinoshita, Toichiro; Nio, Makiko (2012). "Tenth-order QED lepton anomalous magnetic moment: Eighth-order vertices containing a second-order vacuum polarization".Physical Review D.85 (3) 033007.arXiv:1110.2826.Bibcode:2012PhRvD..85c3007A.doi:10.1103/PhysRevD.85.033007.S2CID 119279420.
  3. ^van Mourik, T.; Buhl, M.; Gaigeot, M.-P. (10 February 2014)."Density functional theory across chemistry, physics and biology".Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences.372 (2011) 20120488.Bibcode:2014RSPTA.37220488V.doi:10.1098/rsta.2012.0488.PMC 3928866.PMID 24516181.
  4. ^Schrödinger, E. (1926). "Quantisierung als Eigenwertproblem" [Quantization as an eigenvalue problem].Annalen der Physik (in German).80 (13):437–490.Bibcode:1926AnP...385..437S.doi:10.1002/andp.19263851302.
  5. ^Rayleigh, J. W. S. (1894).Theory of Sound. Vol. I (2nd ed.). London: Macmillan. pp. 115–118.ISBN 978-1-152-06023-4.{{cite book}}:ISBN / Date incompatibility (help)
  6. ^Sulejmanpasic, Tin; Ünsal, Mithat (2018-07-01)."Aspects of perturbation theory in quantum mechanics: The BenderWuMathematica® package".Computer Physics Communications.228:273–289.Bibcode:2018CoPhC.228..273S.doi:10.1016/j.cpc.2017.11.018.ISSN 0010-4655.S2CID 46923647.
  7. ^Sakurai, J.J., and Napolitano, J. (1964,2011).Modern Quantum Mechanics (2nd ed.), Addison WesleyISBN 978-0-8053-8291-4. Chapter 5
  8. ^Landau, L. D.; Lifschitz, E. M. (1977).Quantum Mechanics: Non-relativistic Theory (3rd ed.). Pergamon Press.ISBN 978-0-08-019012-9.
  9. ^Hogervorst M, Meineri M, Penedones J, Salehi Vaziri K (2021). "Hamiltonian truncation in Anti-de Sitter spacetime".Journal of High Energy Physics.2021 (8) 63.arXiv:2104.10689.Bibcode:2021JHEP...08..063H.doi:10.1007/JHEP08(2021)063.S2CID 233346724.
  10. ^Bir, Gennadiĭ Levikovich; Pikus, Grigoriĭ Ezekielevich (1974)."Chapter 15: Perturbation theory for the degenerate case".Symmetry and Strain-induced Effects in Semiconductors. Wiley.ISBN 978-0-470-07321-6.
  11. ^Soliverez, Carlos E. (1981)."General Theory of Effective Hamiltonians".Physical Review A.24 (1):4–9.Bibcode:1981PhRvA..24....4S.doi:10.1103/PhysRevA.24.4 – via Academia.Edu.
  12. ^abDick, Rainer (2020), Dick, Rainer (ed.), "Time-Dependent Perturbations in Quantum Mechanics",Advanced Quantum Mechanics: Materials and Photons, Graduate Texts in Physics, Cham: Springer International Publishing, pp. 265–310,doi:10.1007/978-3-030-57870-1_13,ISBN 978-3-030-57870-1
  13. ^Albert Messiah (1966).Quantum Mechanics, North Holland, John Wiley & Sons.ISBN 0486409244; J. J. Sakurai (1994).Modern Quantum Mechanics (Addison-Wesley)ISBN 9780201539295.
  14. ^Frasca, M. (1998). "Duality in Perturbation Theory and the Quantum Adiabatic Approximation".Physical Review A.58 (5):3439–3442.arXiv:hep-th/9801069.Bibcode:1998PhRvA..58.3439F.doi:10.1103/PhysRevA.58.3439.S2CID 2699775.
  15. ^Mostafazadeh, A. (1997). "Quantum adiabatic approximation and the geometric phase".Physical Review A.55 (3):1653–1664.arXiv:hep-th/9606053.Bibcode:1997PhRvA..55.1653M.doi:10.1103/PhysRevA.55.1653.S2CID 17059815.
  16. ^Frasca, Marco (2007). "A strongly perturbed quantum system is a semiclassical system".Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences.463 (2085):2195–2200.arXiv:hep-th/0603182.Bibcode:2007RSPSA.463.2195F.doi:10.1098/rspa.2007.1879.S2CID 19783654.
  17. ^Lifshitz, E. M., & LD and Sykes Landau (JB). (1965). Quantum Mechanics; Non-relativistic Theory. Pergamon Press.

External links

[edit]
Wikiquote has quotations related toPerturbation theory (quantum mechanics).
Computational
Mathematical
software
Discrete
Analysis
Probability theory
Mathematical
physics
Algebraic
structures
Decision sciences
Other applications
Related
Organizations
National
Other
Retrieved from "https://en.wikipedia.org/w/index.php?title=Perturbation_theory_(quantum_mechanics)&oldid=1314837997"
Categories:
Hidden categories:

[8]ページ先頭

©2009-2025 Movatter.jp