Perturbation theory (quantum mechanics)


In quantum mechanics, perturbation theory is a set of approximation schemes directly related to mathematical perturbation for describing a complicated quantum 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. its energy levels and eigenstates) 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 as asymptotic 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, solved system.




Contents





  • 1 Approximate Hamiltonians


  • 2 Applying perturbation theory

    • 2.1 Limitations

      • 2.1.1 Large perturbations


      • 2.1.2 Non-adiabatic states


      • 2.1.3 Difficult computations




  • 3 Time-independent perturbation theory

    • 3.1 First order corrections


    • 3.2 Second-order and higher corrections


    • 3.3 Effects of degeneracy


    • 3.4 Generalization to multi-parameter case

      • 3.4.1 Hamiltonian and force operator


      • 3.4.2 Perturbation theory as power series expansion


      • 3.4.3 Hellmann–Feynman theorems


      • 3.4.4 Correction of energy and state


      • 3.4.5 Effective Hamiltonian




  • 4 Time-dependent perturbation theory

    • 4.1 Method of variation of constants


    • 4.2 Method of Dyson series



  • 5 Strong perturbation theory


  • 6 Examples

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


    • 6.2 Example of first and second order perturbation theory – quantum pendulum



  • 7 References




Approximate Hamiltonians


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



Applying perturbation theory


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 perturbative electric potential to the quantum mechanical model of the hydrogen atom, tiny shifts in the spectral lines of hydrogen caused by the presence of an electric field (the Stark effect) can be calculated. This is only approximate because the sum of a Coulomb potential with a linear potential is unstable (has no true bound states) although the tunneling 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, say α, is very small. Typically, the results are expressed in terms of finite power series in α that seem to converge to the exact values when summed to higher order. After a certain order n ~ 1/α however, the results become increasingly worse since the series are usually divergent (being asymptotic series). There exist ways to convert them into convergent series, which can be evaluated for large-expansion parameters, most efficiently by the variational method.


In the theory of quantum electrodynamics (QED), in which the electron–photon interaction is treated perturbatively, the calculation of the electron's magnetic moment has been found to agree with experiment to eleven decimal places.[1] In QED and other quantum field theories, special calculation techniques known as Feynman diagrams are used to systematically sum the power series terms.



Limitations



Large perturbations


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. In quantum chromodynamics, for instance, the interaction of quarks with the gluon field cannot be treated perturbatively at low energies because the coupling constant (the expansion parameter) becomes too large.[clarification needed]



Non-adiabatic states


Perturbation theory also fails to describe states that are not generated adiabatically from the "free model", including bound states and various collective phenomena such as solitons.[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 conventional superconductivity, in which the phonon-mediated attraction between conduction electrons leads to the formation of correlated electron pairs known as Cooper pairs. When faced with such systems, one usually turns to other approximation schemes, such as the variational method and the WKB approximation. This is because there is no analogue of a bound particle in the unperturbed model and the energy of a soliton typically goes as the inverse of the expansion parameter. However, if we "integrate" over the solitonic phenomena, the nonperturbative corrections in this case will be tiny; of the order of exp(−1/g) or exp(−1/g2) in the perturbation parameter g. 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


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



Time-independent perturbation theory


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 by Erwin Schrödinger in a 1926 paper,[2] shortly after he produced his theories in wave mechanics. In this paper Schrödinger referred to earlier work of Lord Rayleigh,[3] who investigated harmonic vibrations of a string perturbed by small inhomogeneities. This is why this perturbation theory is often referred to as Rayleigh–Schrödinger perturbation theory.[4]



First order corrections


We begin[5] with an unperturbed Hamiltonian H0, which is also assumed to have no time dependence. It has known energy levels and eigenstates, arising from the time-independent Schrödinger equation:


H0|n(0)⟩=En(0)|n(0)⟩,n=1,2,3,⋯displaystyle H_0leftH_0left|n^(0)rightrangle =E_n^(0)left|n^(0)rightrangle ,qquad n=1,2,3,cdots

For simplicity, we have assumed that the energies are discrete. The (0) superscripts denote that these quantities are associated with the unperturbed system. Note the use of bra–ket notation.


We now introduce a perturbation to the Hamiltonian. Let V be a Hamiltonian representing a weak physical disturbance, such as a potential energy produced by an external field. (Thus, V is formally a Hermitian operator.) Let λ 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+λVdisplaystyle H=H_0+lambda VH=H_0+lambda V

The energy levels and eigenstates of the perturbed Hamiltonian are again given by the Schrödinger equation:


(H0+λV)|n⟩=En|n⟩.nrangle .left(H_0+lambda Vright)|nrangle =E_n|nrangle .

Our goal is to express En and |n⟩nrangle |nrangle in terms of the energy levels and eigenstates of the old Hamiltonian. If the perturbation is sufficiently weak, we can write them as a (Maclaurin) power series in λ:


En=En(0)+λEn(1)+λ2En(2)+⋯|n⟩=|n(0)⟩+λ|n(1)⟩+λ2|n(2)⟩+⋯displaystyle n^(0)rightrangle +lambda leftn^(0)rightrangle +lambda left

where


En(k)=1k!dkEndλk|λ=0|n(k)⟩=1k!dk|n⟩dλk|λ=0displaystyle beginalignedE_n^(k)&=frac 1k!frac d^kE_ndlambda ^kbigg _lambda =0\leftdisplaystyle beginalignedE_n^(k)&=frac 1k!frac d^kE_ndlambda ^kbigg _lambda =0\left

When k = 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 we go to higher order.


Substituting the power series expansion into the Schrödinger equation, we obtain


(H0+λV)(|n(0)⟩+λ|n(1)⟩+⋯)=(En(0)+λEn(1)+⋯)(|n(0)⟩+λ|n(1)⟩+⋯)displaystyle left(H_0+lambda Vright)left(leftdisplaystyle left(H_0+lambda Vright)left(left

Expanding this equation and comparing coefficients of each power of λ results in an infinite series of simultaneous equations. The zeroth-order equation is simply the Schrödinger equation for the unperturbed system. The first-order equation is


H0|n(1)⟩+V|n(0)⟩=En(0)|n(1)⟩+En(1)|n(0)⟩n^(0)rightrangle =E_n^(0)leftH_0left|n^(1)rightrangle +Vleft|n^(0)rightrangle =E_n^(0)left|n^(1)rightrangle +E_n^(1)left|n^(0)rightrangle

Operating through by ⟨n(0)|displaystyle langle n^(0)langle n^(0)|, the first term on the left-hand side cancels the first term on the right-hand side. (Recall, the unperturbed Hamiltonian is Hermitian). This leads to the first-order energy shift:


En(1)=⟨n(0)|V|n(0)⟩VleftE_n^(1)=leftlangle n^(0)right|Vleft|n^(0)rightrangle

This is simply the expectation value of the perturbation Hamiltonian while the system is in the unperturbed state. This result can be interpreted in the following way: suppose the perturbation is applied, but we keep the system 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 by ⟨n(0)|V|n(0)⟩Vlangle 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 we compute the corrections to the energy eigenstate, we need to address the issue of normalization. We may suppose


⟨n(0)|n(0)⟩=1,left.n^(0)rightrangle =1,leftlangle n^(0)right|left.n^(0)rightrangle =1,

but perturbation theory assumes we also have ⟨n|n⟩=1displaystyle langle nlangle n|nrangle =1. It follows that at first order in λ, we must have


(⟨n(0)|+λ⟨n(1)|)(|n(0)⟩+λ|n(1)⟩)=1right)left(leftleft(leftlangle n^(0)right|+lambda leftlangle n^(1)right|right)left(left|n^(0)rightrangle +lambda left|n^(1)rightrangle right)=1

⟨n(0)|n(0)⟩+λ⟨n(0)|n(1)⟩+λ⟨n(1)|n(0)⟩+λ2⟨n(1)|n(1)⟩=1left.n^(1)rightrangle +lambda leftlangle n^(1)rightleftlangle n^(0)right|left.n^(0)rightrangle +lambda leftlangle n^(0)right|left.n^(1)rightrangle +lambda leftlangle n^(1)right|left.n^(0)rightrangle +cancel lambda ^2leftlangle n^(1)right=1

⟨n(0)|n(1)⟩+⟨n(1)|n(0)⟩=0.displaystyle leftlangle n^(0)rightleftlangle n^(0)right|left.n^(1)rightrangle +leftlangle n^(1)right|left.n^(0)rightrangle =0.

Since the overall phase is not determined in quantum mechanics, without loss of generality, in time independent theory we may assume ⟨n(0)|n(1)⟩n^(1)rangle n^(1)rangle is purely real. Therefore,


⟨n(0)|n(1)⟩=−⟨n(1)|n(0)⟩,displaystyle leftlangle n^(0)rightdisplaystyle leftlangle n^(0)right

and we deduce


⟨n(0)|n(1)⟩=0.displaystyle leftlangle n^(0)rightleftlangle n^(0)right|left.n^(1)rightrangle =0.

To obtain the first-order correction to the energy eigenstate, we insert our expression for the first-order energy correction back into the result shown above of equating the first-order coefficients of λ. We then make use of the resolution of the identity,


V|n(0)⟩=(∑k≠n|k(0)⟩⟨k(0)|)V|n(0)⟩+(|n(0)⟩⟨n(0)|)V|n(0)⟩=∑k≠n|k(0)⟩⟨k(0)|V|n(0)⟩+En(1)|n(0)⟩,displaystyle right)Vleftright)Vleft

where the |k(0)⟩displaystyle |k^(0)rangle are in the orthogonal complement of |n(0)⟩displaystyle |n^(0)rangle . The first-order equation may thus be expressed as


(En(0)−H0)|n(1)⟩=∑k≠n|k(0)⟩⟨k(0)|V|n(0)⟩Vleftleft(E_n^(0)-H_0right)left|n^(1)rightrangle =sum _kneq nleft|k^(0)rightrangle leftlangle k^(0)right|Vleft|n^(0)rightrangle

For the moment, suppose that the zeroth-order energy level is not degenerate, i.e. there is no eigenstate of H0 in the orthogonal complement of |n(0)⟩displaystyle |n^(0)rangle with the energy En(0)displaystyle E_n^(0)E_n^(0). After renaming the summation dummy index above as k′displaystyle k'k', we can pick any k≠ndisplaystyle kneq nkneq n, and multiply through by ⟨k(0)|langle k^(0)| giving


(En(0)−Ek(0))⟨k(0)|n(1)⟩=⟨k(0)|V|n(0)⟩n^(0)rightrangle left(E_n^(0)-E_k^(0)right)leftlangle k^(0)right.left|n^(1)rightrangle =leftlangle k^(0)right|Vleft|n^(0)rightrangle

We see that the above ⟨k(0)|n(1)⟩n^(1)rangle n^(1)rangle also gives us the component of the first-order correction along |k(0)⟩displaystyle |k^(0)rangle .


Thus in total we get,


|n(1)⟩=∑k≠n⟨k(0)|V|n(0)⟩En(0)−Ek(0)|k(0)⟩k^(0)rightrangle left|n^(1)rightrangle =sum _kneq nfrac n^(0)rightrangle E_n^(0)-E_k^(0)left|k^(0)rightrangle

The first-order change in the n-th energy eigenket has a contribution from each of the energy eigenstates kn. Each term is proportional to the matrix element ⟨k(0)|V|n(0)⟩n^(0)rangle langle k^(0)|V|n^(0)rangle , which is a measure of how much the perturbation mixes eigenstate n with eigenstate k; it is also inversely proportional to the energy difference between eigenstates k and n, which means that the perturbation deforms the eigenstate to a greater extent if there are more eigenstates at nearby energies. We see also that the expression is singular if any of these states have the same energy as state n, which is why we assumed that there is no degeneracy.



Second-order and higher corrections


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


2⟨n(0)|n(2)⟩+⟨n(1)|n(1)⟩=0.left.n^(1)rightrangle =0.2leftlangle n^(0)right|left.n^(2)rightrangle +leftlangle n^(1)right|left.n^(1)rightrangle =0.

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


En(λ)=En(0)+λ⟨n(0)|V|n(0)⟩+λ2∑k≠n|⟨k(0)|V|n(0)⟩|2En(0)−Ek(0)+O(λ3)displaystyle E_n(lambda )=E_n^(0)+lambda leftlangle n^(0)rightE_n(lambda )=E_n^(0)+lambda leftlangle n^(0)right|Vleft|n^(0)rightrangle +lambda ^2sum _kneq nfrac leftlangle k^(0)rightE_n^(0)-E_k^(0)+O(lambda ^3)
|n(λ)⟩=|n(0)⟩+λ∑k≠n|k(0)⟩⟨k(0)|V|n(0)⟩En(0)−Ek(0)+λ2∑k≠n∑ℓ≠n|k(0)⟩⟨k(0)|V|ℓ(0)⟩⟨ℓ(0)|V|n(0)⟩(En(0)−Ek(0))(En(0)−Eℓ(0))−λ2∑k≠n|k(0)⟩⟨n(0)|V|n(0)⟩⟨k(0)|V|n(0)⟩(En(0)−Ek(0))2−12λ2|n(0)⟩∑k≠n⟨n(0)|V|k(0)⟩⟨k(0)|V|n(0)⟩(En(0)−Ek(0))2+O(λ3).displaystyle k^(0)rightrangle frac n^(0)rightrangle E_n^(0)-E_k^(0)+lambda ^2sum _kneq nsum _ell neq nleftk^(0)rightrangle frac n^(0)rightrangle E_n^(0)-E_k^(0)+lambda ^2sum _kneq nsum _ell neq nleft

Extending the process further, the third-order energy correction can be shown to be [6]


En(3)=∑k≠n∑m≠n⟨n(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)⟩∑m≠n|⟨n(0)|V|m(0)⟩|2(En(0)−Em(0))2.displaystyle E_n^(3)=sum _kneq nsum _mneq nfrac m^(0)rangle langle m^(0)left(E_n^(0)-E_m^(0)right)left(E_n^(0)-E_k^(0)right)-langle n^(0)displaystyle E_n^(3)=sum _kneq nsum _mneq nfrac m^(0)rangle langle m^(0)left(E_n^(0)-E_m^(0)right)left(E_n^(0)-E_k^(0)right)-langle n^(0)

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


If we introduce the notation,



Vnm≡⟨n(0)|V|m(0)⟩VV_nmequiv langle n^(0)|V|m^(0)rangle ,


Enm≡En(0)−Em(0)displaystyle E_nmequiv E_n^(0)-E_m^(0)E_nmequiv E_n^(0)-E_m^(0),

then the energy corrections to fifth order can be written


En(1)=VnnEn(2)=|Vnk2|2Enk2En(3)=Vnk3Vk3k2Vk2nEnk2Enk3−Vnn|Vnk3|2Enk32En(4)=Vnk4Vk4k3Vk3k2Vk2nEnk2Enk3Enk4−|Vnk4|2Enk42|Vnk2|2Enk2−VnnVnk4Vk4k3Vk3nEnk32Enk4−VnnVnk4Vk4k2Vk2nEnk2Enk42+Vnn2|Vnk4|2Enk43=Vnk4Vk4k3Vk3k2Vk2nEnk2Enk3Enk4−En(2)|Vnk4|2Enk42−2VnnVnk4Vk4k3Vk3nEnk32Enk4+Vnn2|Vnk4|2Enk43En(5)=Vnk5Vk5k4Vk4k3Vk3k2Vk2nEnk2Enk3Enk4Enk5−Vnk5Vk5k4Vk4nEnk42Enk5|Vnk2|2Enk2−Vnk5Vk5k2Vk2nEnk2Enk52|Vnk2|2Enk2−|Vnk5|2Enk52Vnk3Vk3k2Vk2nEnk2Enk3−VnnVnk5Vk5k4Vk4k3Vk3nEnk32Enk4Enk5−VnnVnk5Vk5k4Vk4k2Vk2nEnk2Enk42Enk5−VnnVnk5Vk5k3Vk3k2Vk2nEnk2Enk3Enk52+Vnn|Vnk5|2Enk52|Vnk3|2Enk32+2Vnn|Vnk5|2Enk53|Vnk2|2Enk2+Vnn2Vnk5Vk5k4Vk4nEnk43Enk5+Vnn2Vnk5Vk5k3Vk3nEnk32Enk52+Vnn2Vnk5Vk5k2Vk2nEnk2Enk53−Vnn3|Vnk5|2Enk54=Vnk5Vk5k4Vk4k3Vk3k2Vk2nEnk2Enk3Enk4Enk5−2En(2)Vnk5Vk5k4Vk4nEnk42Enk5−|Vnk5|2Enk52Vnk3Vk3k2Vk2nEnk2Enk3−2Vnn(Vnk5Vk5k4Vk4k3Vk3nEnk32Enk4Enk5−Vnk5Vk5k4Vk4k2Vk2nEnk2Enk42Enk5+|Vnk5|2Enk52|Vnk3|2Enk32+2En(2)|Vnk5|2Enk53)+Vnn2(2Vnk5Vk5k4Vk4nEnk43Enk5+Vnk5Vk5k3Vk3nEnk32Enk52)−Vnn3|Vnk5|2Enk54displaystyle beginalignedE_n^(1)&=V_nn\E_n^(2)&=frac E_nk_2\E_n^(3)&=frac V_nk_3V_k_3k_2V_k_2nE_nk_2E_nk_3-V_nnfrac V_nk_3E_nk_3^2\E_n^(4)&=frac V_nk_4V_k_4k_3V_k_3k_2V_k_2nE_nk_2E_nk_3E_nk_4-frac ^2E_nk_4^2frac E_nk_2-V_nnfrac V_nk_4V_k_4k_3V_k_3nE_nk_3^2E_nk_4-V_nnfrac V_nk_4V_k_4k_2V_k_2nE_nk_2E_nk_4^2+V_nn^2frac ^2E_nk_4^3\&=frac V_nk_4V_k_4k_3V_k_3k_2V_k_2nE_nk_2E_nk_3E_nk_4-E_n^(2)frac ^2E_nk_4^2-2V_nnfrac V_nk_4V_k_4k_3V_k_3nE_nk_3^2E_nk_4+V_nn^2frac ^2E_nk_4^3\E_n^(5)&=frac V_nk_5V_k_5k_4V_k_4k_3V_k_3k_2V_k_2nE_nk_2E_nk_3E_nk_4E_nk_5-frac V_nk_5V_k_5k_4V_k_4nE_nk_4^2E_nk_5frac E_nk_2-frac V_nk_5V_k_5k_2V_k_2nE_nk_2E_nk_5^2frac E_nk_2-frac E_nk_5^2frac V_nk_3V_k_3k_2V_k_2nE_nk_2E_nk_3\&quad -V_nnfrac V_nk_5V_k_5k_4V_k_4k_3V_k_3nE_nk_3^2E_nk_4E_nk_5-V_nnfrac V_nk_5V_k_5k_4V_k_4k_2V_k_2nE_nk_2E_nk_4^2E_nk_5-V_nnfrac V_nk_5V_k_5k_3V_k_3k_2V_k_2nE_nk_2E_nk_3E_nk_5^2+V_nnfrac E_nk_5^2frac V_nk_3E_nk_3^2+2V_nnfrac E_nk_5^3frac E_nk_2\&quad +V_nn^2frac V_nk_5V_k_5k_4V_k_4nE_nk_4^3E_nk_5+V_nn^2frac V_nk_5V_k_5k_3V_k_3nE_nk_3^2E_nk_5^2+V_nn^2frac V_nk_5V_k_5k_2V_k_2nE_nk_2E_nk_5^3-V_nn^3frac E_nk_5^4\&=frac V_nk_5V_k_5k_4V_k_4k_3V_k_3k_2V_k_2nE_nk_2E_nk_3E_nk_4E_nk_5-2E_n^(2)frac V_nk_5V_k_5k_4V_k_4nE_nk_4^2E_nk_5-frac E_nk_5^2frac V_nk_3V_k_3k_2V_k_2nE_nk_2E_nk_3\&quad -2V_nnleft(frac V_nk_5V_k_5k_4V_k_4k_3V_k_3nE_nk_3^2E_nk_4E_nk_5-frac V_nk_5V_k_5k_4V_k_4k_2V_k_2nE_nk_2E_nk_4^2E_nk_5+frac E_nk_5^2frac V_nk_3E_nk_3^2+2E_n^(2)frac E_nk_5^3right)\&quad +V_nn^2left(2frac V_nk_5V_k_5k_4V_k_4nE_nk_4^3E_nk_5+frac V_nk_5V_k_5k_3V_k_3nE_nk_3^2E_nk_5^2right)-V_nn^3frac E_nk_5^4endalignedbeginalignedE_n^(1)&=V_nn\E_n^(2)&=frac E_nk_2\E_n^(3)&=frac V_nk_3V_k_3k_2V_k_2nE_nk_2E_nk_3-V_nnfrac V_nk_3E_nk_3^2\E_n^(4)&=frac V_nk_4V_k_4k_3V_k_3k_2V_k_2nE_nk_2E_nk_3E_nk_4-frac ^2E_nk_4^2frac E_nk_2-V_nnfrac V_nk_4V_k_4k_3V_k_3nE_nk_3^2E_nk_4-V_nnfrac V_nk_4V_k_4k_2V_k_2nE_nk_2E_nk_4^2+V_nn^2frac ^2E_nk_4^3\&=frac V_nk_4V_k_4k_3V_k_3k_2V_k_2nE_nk_2E_nk_3E_nk_4-E_n^(2)frac ^2E_nk_4^2-2V_nnfrac V_nk_4V_k_4k_3V_k_3nE_nk_3^2E_nk_4+V_nn^2frac ^2E_nk_4^3\E_n^(5)&=frac V_nk_5V_k_5k_4V_k_4k_3V_k_3k_2V_k_2nE_nk_2E_nk_3E_nk_4E_nk_5-frac V_nk_5V_k_5k_4V_k_4nE_nk_4^2E_nk_5frac E_nk_2-frac V_nk_5V_k_5k_2V_k_2nE_nk_2E_nk_5^2frac E_nk_2-frac E_nk_5^2frac V_nk_3V_k_3k_2V_k_2nE_nk_2E_nk_3\&quad -V_nnfrac V_nk_5V_k_5k_4V_k_4k_3V_k_3nE_nk_3^2E_nk_4E_nk_5-V_nnfrac V_nk_5V_k_5k_4V_k_4k_2V_k_2nE_nk_2E_nk_4^2E_nk_5-V_nnfrac V_nk_5V_k_5k_3V_k_3k_2V_k_2nE_nk_2E_nk_3E_nk_5^2+V_nnfrac E_nk_5^2frac V_nk_3E_nk_3^2+2V_nnfrac E_nk_5^3frac E_nk_2\&quad +V_nn^2frac V_nk_5V_k_5k_4V_k_4nE_nk_4^3E_nk_5+V_nn^2frac V_nk_5V_k_5k_3V_k_3nE_nk_3^2E_nk_5^2+V_nn^2frac V_nk_5V_k_5k_2V_k_2nE_nk_2E_nk_5^3-V_nn^3frac E_nk_5^4\&=frac V_nk_5V_k_5k_4V_k_4k_3V_k_3k_2V_k_2nE_nk_2E_nk_3E_nk_4E_nk_5-2E_n^(2)frac V_nk_5V_k_5k_4V_k_4nE_nk_4^2E_nk_5-frac E_nk_5^2frac V_nk_3V_k_3k_2V_k_2nE_nk_2E_nk_3\&quad -2V_nnleft(frac V_nk_5V_k_5k_4V_k_4k_3V_k_3nE_nk_3^2E_nk_4E_nk_5-frac V_nk_5V_k_5k_4V_k_4k_2V_k_2nE_nk_2E_nk_4^2E_nk_5+frac E_nk_5^2frac V_nk_3E_nk_3^2+2E_n^(2)frac E_nk_5^3right)\&quad +V_nn^2left(2frac V_nk_5V_k_5k_4V_k_4nE_nk_4^3E_nk_5+frac V_nk_5V_k_5k_3V_k_3nE_nk_3^2E_nk_5^2right)-V_nn^3frac E_nk_5^4endaligned

and the states to fourth order can be written


|n(1)⟩=Vk1nEnk1|k1(0)⟩|n(2)⟩=(Vk1k2Vk2nEnk1Enk2−VnnVk1nEnk12)|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+Vk3k2Vk1k2Vk4k3Vk2k42Ek1nEk2k32Ek2k4−Vk2k3Vk3k4Vk4nVk1k2Ek1nEk2nEnk3Enk4+Vk1k2Ek1n(|Vk2k3|2Vk2k2Ek2k33−|Vnk3|2Vk2nEk3n2Ek2n)+VnnVk1k2Vk3nVk2k3Ek1nEnk3Ek2n(1Enk3+1Ek2n+1Ek1n)+|Vk2n|2Vk1k3Enk2Ek1n(Vk3nEnk1Enk3−Vk3k1Ek3k12)−Vnn(Vk3k2Vk1k3Vk2k1+Vk3k1Vk2k3Vk1k2)2Ek1nEk1k32Ek1k2+|Vnn|2Ek1n(Vk1nVnnEk1n3+Vk1k2Vk2nEk2n3)−|Vk1k2|2VnnVk1nEk1nEk1k23]|k1(0)⟩+12[Vnk1Vk1k2Enk1Ek2n2(Vk2nVnnEk2n−Vk2k3Vk3nEnk3)−Vk1nVk2k1Ek1n2Enk2(Vk3k2Vnk3Enk3+VnnVnk2Enk2)+|Vnk1|2Ek1n2(3|Vnk2|24Ek2n2−2|Vnn|2Ek1n2)−Vk2k3Vk3k1|Vnk1|2Enk32Enk1Enk2]|n(0)⟩displaystyle k_1^(0)rangle \displaystyle k_1^(0)rangle \

All terms involved kj should be summed over kj such that the denominator does not vanish.





Effects of degeneracy


Suppose that two or more energy eigenstates are degenerate. 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 operator


En(0)−H0displaystyle E_n^(0)-H_0E_n^(0)-H_0

does not have a well-defined inverse.


Let D denote the subspace spanned by these degenerate eigenstates. No matter how small the perturbation is, in the degenerate subspace D the energy differences between the eigenstates of H 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 than D.


The successful perturbations will not be "small" relative to a poorly chosen basis of D. Instead, we consider the perturbation "small" if the new eigenstate is close to the subspace D. The new Hamiltonian must be diagonalized in D, or a slight variation of D, so to speak. These perturbed eigenstates in D are now the basis for the perturbation expansion,


|n⟩=∑k∈Dαnk|k(0)⟩+λ|n(1)⟩.n^(1)rangle .n^(1)rangle .

For the first-order perturbation, we need solve the perturbed Hamiltonian restricted to the degenerate subspace D,


V|k(0)⟩=ϵk|k(0)⟩+small∀|k(0)⟩∈D,k^(0)rangle in D,k^(0)rangle in D,

simultaneously for all the degenerate eigenstates, where ϵkdisplaystyle epsilon _kepsilon _k are first-order corrections to the degenerate energy levels, and "small" is a vector of O(λ)displaystyle O(lambda )displaystyle O(lambda ) orthogonal to D. This amounts to diagonalizing the matrix


⟨k(0)|V|l(0)⟩=Vkl∀|k(0)⟩,|l(0)⟩∈D.k^(0)rangle ,langle k^(0)|V|l^(0)rangle =V_klqquad forall ;|k^(0)rangle ,|l^(0)rangle in D.

This procedure is approximate, since we neglected states outside the D subspace ("small"). The splitting of degenerate energies ϵkdisplaystyle epsilon _kepsilon _k is generally observed. Although the splitting may be small, O(λ)displaystyle O(lambda )displaystyle O(lambda ), compared to the range of energies found in the system, it is crucial in understanding certain details, such as spectral lines in Electron Spin Resonance experiments.


Higher-order corrections due to other eigenstates outside D can be found in the same way as for the non-degenerate case,


(En(0)−H0)|n(1)⟩=∑k∉D(⟨k(0)|V|n(0)⟩)|k(0)⟩.n^(0)rangle right)left(E_n^(0)-H_0right)|n^(1)rangle =sum _knot in Dleft(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 outside D, so we can write


|n(1)⟩=∑k∉D⟨k(0)|V|n(0)⟩En(0)−Ek(0)|k(0)⟩,k^(0)rangle ,|n^(1)rangle =sum _knot in Dfrac langle k^(0)E_n^(0)-E_k^(0)|k^(0)rangle ,

but the effect on the degenerate states is of O(λ)displaystyle O(lambda )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 the nearly 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.



Generalization to multi-parameter case


The generalization of the time-independent perturbation theory to the case where there are multiple small parameters xμ=(x1,x2,⋯)displaystyle x^mu =(x^1,x^2,cdots )x^mu =(x^1,x^2,cdots ) in place of λ can be formulated more systematically using the language of differential 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


From the differential geometric point of view, a parameterized Hamiltonian is considered as a function defined on the parameter manifold that maps each particular set of parameters (x1,x2,⋯)displaystyle (x^1,x^2,cdots )(x^1,x^2,cdots ) to an Hermitian operator H(x μ) that acts on the Hilbert space. The parameters here can be external field, interaction strength, or driving parameters in the quantum phase transition. Let En(x μ) and |n(xμ)⟩n(x^mu )rangle |n(x^mu )rangle be the n-th eigenenergy and eigenstate of H(x μ) respectively. In the language of differential geometry, the states |n(xμ)⟩n(x^mu )rangle |n(x^mu )rangle form a vector bundle over the parameter manifold, on which derivatives of these states can be defined. The perturbation theory is to answer the following question: given En(x0μ)displaystyle E_n(x_0^mu )E_n(x_0^mu ) and |n(x0μ)⟩n(x_0^mu )rangle |n(x_0^mu )rangle at an unperturbed reference point x0μdisplaystyle x_0^mu x_0^mu , how to estimate the En(x μ) and |n(xμ)⟩n(x^mu )rangle |n(x^mu )rangle at x μ close to that reference point.


Without loss of generality, the coordinate system can be shifted, such that the reference point x0μ=0displaystyle x_0^mu =0x_0^mu =0 is set to be the origin. The following linearly parameterized Hamiltonian is frequently used


H(xμ)=H(0)+xμFμ.displaystyle H(x^mu )=H(0)+x^mu F_mu .H(x^mu )=H(0)+x^mu F_mu .

If the parameters x μ are considered as generalized coordinates, then Fμ should be identified as the generalized force operators related to those coordinates. Different indices μ label the different forces along different directions in the parameter manifold. For example, if x μ denotes the external magnetic field in the μ-direction, then Fμ should be the magnetization in the same direction.



Perturbation theory as power series expansion


The validity of the 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 (like Taylor expansion) of the parameters:


En(xμ)=En+xμ∂μEn+12!xμxν∂μ∂νEn+⋯|n(xμ)⟩=|n⟩+xμ|∂μn⟩+12!xμxν|∂μ∂νn⟩+⋯displaystyle nrangle +x^mu nrangle +x^mu

Here μ denotes the derivative with respect to x μ. When applying to the state |∂μn⟩partial _mu nrangle |partial _mu nrangle , it should be understood as the covariant derivative if the vector bundle is equipped with non-vanishing connection. All the terms on the right-hand-side of the series are evaluated at x μ = 0, e.g. EnEn(0) and |n⟩≡|n(0)⟩n(0)rangle |nrangle 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


The above power series expansion can be readily evaluated if there is a systematic approach to calculate the derivates to any order. Using the chain rule, the derivatives can be broken down to the single derivative on either the energy or the state. The Hellmann–Feynman theorems are used to calculate these single derivatives. The first Hellmann–Feynman theorem gives the derivative of the energy,


∂μEn=⟨n|∂μH|n⟩nrangle partial _mu E_n=langle n|partial _mu H|nrangle

The second Hellmann–Feynman theorem gives the derivative of the state (resolved by the complete basis with m ≠ n),


⟨m|∂μn⟩=⟨m|∂μH|n⟩En−Em,⟨∂μm|n⟩=⟨m|∂μH|n⟩Em−En.partial _mu nrangle =frac nrangle E_n-E_m,qquad langle partial _mu mlangle m|partial _mu nrangle =frac nrangle E_n-E_m,qquad langle partial _mu m|nrangle =frac nrangle E_m-E_n.

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


The theorems can be simply derived by applying the differential operator μ to both sides of the Schrödinger equation H|n⟩=En|n⟩,nrangle =E_nH|nrangle =E_n|nrangle , which reads


∂μH|n⟩+H|∂μn⟩=∂μEn|n⟩+En|∂μn⟩.partial _mu nrangle =partial _mu E_npartial _mu H|nrangle +H|partial _mu nrangle =partial _mu E_n|nrangle +E_n|partial _mu nrangle .

Then overlap with the state ⟨m|displaystyle langle mlangle m| from left and make use of the Schrödinger equation ⟨m|H=⟨m|EmH=langle mlangle m|H=langle m|E_m again,


⟨m|∂μH|n⟩+Em⟨m|∂μn⟩=∂μEn⟨m|n⟩+En⟨m|∂μn⟩.partial _mu nrangle .langle m|partial _mu H|nrangle +E_mlangle m|partial _mu nrangle =partial _mu E_nlangle m|nrangle +E_nlangle m|partial _mu nrangle .

Given that the eigenstates of the Hamiltonian always form an orthonormal basis ⟨m|n⟩=δmndisplaystyle langle mlangle m|nrangle =delta _mn, the cases of m = n and mn 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


To the second order, the energy correction reads


En(xμ)=⟨n|H|n⟩+⟨n|∂μH|n⟩xμ+ℜ∑m≠n⟨n|∂νH|m⟩⟨m|∂μH|n⟩En−Emxμxν+⋯,nrangle +langle nnrangle +langle n

where ℜdisplaystyle Re Re denotes the real part function.
The first order derivative μEn is given by the first Hellmann–Feynman theorem directly. To obtain the second order derivative μνEn, simply applying the differential operator μ to the result of the first order derivative ⟨n|∂νH|n⟩displaystyle langle nlangle n|partial _nu H|nrangle , which reads


∂μ∂νEn=⟨∂μn|∂νH|n⟩+⟨n|∂μ∂νH|n⟩+⟨n|∂νH|∂μn⟩.partial _mu partial _nu Hpartial _mu partial _nu E_n=langle partial _mu n|partial _nu H|nrangle +langle n|partial _mu partial _nu H|nrangle +langle n|partial _nu H|partial _mu nrangle .

Note that for linearly parameterized Hamiltonian, there is no second derivative μνH = 0 on the operator level. Resolve the derivative of state by inserting the complete set of basis,


∂μ∂νEn=∑m(⟨∂μn|m⟩⟨m|∂νH|n⟩+⟨n|∂νH|m⟩⟨m|∂μn⟩),partial _nu Hpartial _mu partial _nu E_n=sum _mleft(langle partial _mu n|mrangle langle m|partial _nu H|nrangle +langle n|partial _nu H|mrangle langle m|partial _mu nrangle right),

then all parts can be calculated using the Hellmann–Feynman theorems. In terms of Lie derivatives, ⟨∂μn|n⟩=⟨n|∂μn⟩=0partial _mu nrangle =0langle partial _mu n|nrangle =langle n|partial _mu nrangle =0 according to the definition of the connection for the vector bundle. Therefore, the case m = 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⟩+∑m≠n⟨m|∂μH|n⟩En−Em|m⟩xμ+(∑m≠n∑l≠n⟨m|∂μH|l⟩⟨l|∂νH|n⟩(En−Em)(En−El)|m⟩−∑m≠n⟨m|∂μH|n⟩⟨n|∂νH|n⟩(En−Em)2|m⟩−12∑m≠n⟨n|∂μH|m⟩⟨m|∂νH|n⟩(En−Em)2|m⟩)xμxν+⋯.displaystyle mrangle x^mu \&+left(sum _mneq nsum _lneq nfrac partial _mu H(E_n-E_m)(E_n-E_l)mrangle x^mu \&+left(sum _mneq nsum _lneq nfrac partial _mu H(E_n-E_m)(E_n-E_l)

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 like Mathematica.



Effective Hamiltonian


Let H(0) be the Hamiltonian completely restricted either in the low-energy subspace HLdisplaystyle mathcal H_Lmathcal H_L or in the high-energy subspace HHdisplaystyle mathcal H_Hmathcal H_H, such that there is no matrix element in H(0) connecting the low- and the high-energy subspaces, i.e. ⟨m|H(0)|l⟩=0lrangle =0langle m|H(0)|lrangle =0 if m∈HL,l∈HHdisplaystyle min mathcal H_L,lin mathcal H_Hmin mathcal H_L,lin mathcal H_H. Let Fμ = ∂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[7]


Hmneff(xμ)=⟨m|H|n⟩+δnm⟨m|∂μH|n⟩xμ+12!∑l∈HH(⟨m|∂μH|l⟩⟨l|∂νH|n⟩Em−El+⟨m|∂νH|l⟩⟨l|∂μH|n⟩En−El)xμxν+⋯.displaystyle H_mn^texteffleft(x^mu right)=langle mdisplaystyle H_mn^texteffleft(x^mu right)=langle m

Here m,n∈HLdisplaystyle m,nin mathcal H_Lm,nin mathcal H_L are restricted in the low energy subspace. The above result can be derived by power series expansion of ⟨m|H(xμ)|n⟩H(x^mu )langle m|H(x^mu )|nrangle .


In a formal way it is possible to define an effective Hamiltonian that gives exactly the low-lying energy states and wavefunctions.[8] In practice, some kind of approximation (perturbation theory) is generally required.



Time-dependent perturbation theory



Method of variation of constants


Time-dependent perturbation theory, developed by Paul Dirac, studies the effect of a time-dependent perturbation V(t) applied to a time-independent Hamiltonian H0.[9]


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-dependent expectation value of some observable A, for a given initial state.

  • The time-dependent amplitudes[clarification needed] of those quantum states that are energy eigenkets (eigenvectors) in the unperturbed system.

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


The second quantity looks at the time-dependent probability of occupation for each eigenstate. This is particularly useful in laser 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" of spectral lines (see line broadening) and particle decay in particle physics and nuclear physics.


We will briefly examine the method behind Dirac's formulation of time-dependent perturbation theory. Choose an energy basis |n⟩displaystyle 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⟩jrangle |jrangle at time t = 0, its state at subsequent times varies only by a phase (in the Schrödinger picture, where state vectors evolve in time and operators are constant),


|j(t)⟩=e−iEjt/ℏ|j⟩ .j(t)rangle =e^-iE_jt/hbar |j(t)rangle =e^-iE_jt/hbar |jrangle ~.

Now, introduce a time-dependent perturbing Hamiltonian V(t). The Hamiltonian of the perturbed system is


H=H0+V(t) .displaystyle H=H_0+V(t)~.H=H_0+V(t)~.

Let |ψ(t)⟩displaystyle |psi (t)rangle denote the quantum state of the perturbed system at time t. It obeys the time-dependent Schrödinger equation,


H|ψ(t)⟩=iℏ∂∂t|ψ(t)⟩ .displaystyle HH|psi (t)rangle =ihbar 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⟩nrangle |nrangle :







|ψ(t)⟩=∑ncn(t)e−iEnt/ℏ|n⟩ ,displaystyle |psi (t)rangle =sum _nc_n(t)e^-iE_nt/hbar |nrangle ~,








 



 



 



 




(1)


where the cn(t)s are to be determined complex functions of t which we will refer to as amplitudes (strictly speaking, they are the amplitudes in the Dirac picture).


We have explicitly extracted the exponential phase factors exp⁡(−iEnt/ℏ)displaystyle exp(-iE_nt/hbar )exp(-iE_nt/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⟩jrangle |jrangle and no perturbation is present, the amplitudes have the convenient property that, for all t,
cj(t) = 1 and cn(t) = 0 if n ≠ j.


The square of the absolute amplitude cn(t) is the probability that the system is in state n at time t, since


|cn(t)|2=|⟨n|ψ(t)⟩|2 .displaystyle leftleft|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 a product rule, one obtains


∑n(iℏ∂cn∂t−cn(t)V(t))e−iEnt/ℏ|n⟩=0 .nrangle =0~.sum _nleft(ihbar frac partial c_npartial t-c_n(t)V(t)right)e^-iE_nt/hbar |nrangle =0~.

By resolving the identity in front of V and multiplying through by the bra ⟨n|displaystyle langle ndisplaystyle langle n on the left, this can be reduced to a set of coupled differential equations for the amplitudes,


∂cn∂t=−iℏ∑k⟨n|V(t)|k⟩ck(t)e−i(Ek−En)t/ℏ .krangle ,c_k(t),e^-i(E_k-E_n)t/hbar ~.frac partial c_npartial t=frac -ihbar sum _klangle n|V(t)|krangle ,c_k(t),e^-i(E_k-E_n)t/hbar ~.

where we have used equation (1) to evaluate the sum on n in the second term, then used the fact that ⟨k|Ψ(t)⟩=ck(t)e−iEkt/ℏPsi (t)rangle =c_k(t)e^-iE_kt/hbar Psi (t)rangle =c_k(t)e^-iE_kt/hbar .


The matrix elements of V 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 difference EkEn, the phase winds around 0 several times. If the time-dependence of V is sufficiently slow, this may cause the state amplitudes to oscillate. ( E.g., such oscillations are useful for managing radiative transitions in a laser.)


Up to this point, we have made no approximations, so this set of differential equations is exact. By supplying appropriate initial values cn(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 the ammonia 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)+−iℏ∑k∫0tdt′⟨n|V(t′)|k⟩ck(t′)e−i(Ek−En)t′/ℏ .V(t')c_n(t)=c_n(0)+frac -ihbar sum _kint _0^tdt';langle n|V(t')|krangle ,c_k(t'),e^-i(E_k-E_n)t'/hbar ~.

Repeatedly substituting this expression for cn 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 c_n(t)=c_n^(0)+c_n^(1)+c_n^(2)+cdots

where, for example, the first-order term is


cn(1)(t)=−iℏ∑k∫0tdt′⟨n|V(t′)|k⟩ck(0)e−i(Ek−En)t′/ℏ .displaystyle c_n^(1)(t)=frac -ihbar sum _kint _0^tdt';langle ndisplaystyle c_n^(1)(t)=frac -ihbar sum _kint _0^tdt';langle n

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



Method of Dyson series


Time-dependent perturbations can be reorganized through the technique of the Dyson series. The Schrödinger equation


H(t)|ψ(t)⟩=iℏ∂|ψ(t)⟩∂tdisplaystyle H(t)H(t)|psi (t)rangle =ihbar frac partial partial t

has the formal solution


|ψ(t)⟩=Texp⁡[−iℏ∫t0tdt′H(t′)]|ψ(t0)⟩ ,psi (t_0)rangle ~,|psi (t)rangle =Texp left[-frac ihbar int _t_0^tdt'H(t')right]|psi (t_0)rangle ~,

where T 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)=begincasesA(t_1)A(t_2)&t_1>t_2\A(t_2)A(t_1)&t_2>t_1endcases~.TA(t_1)A(t_2)=begincasesA(t_1)A(t_2)&t_1>t_2\A(t_2)A(t_1)&t_2>t_1endcases~.

Thus, the exponential represents the following Dyson series,


|ψ(t)⟩=[1−iℏ∫t0tdt1H(t1)−1ℏ2∫t0tdt1∫t0t1dt2H(t1)H(t2)+…]|ψ(t0)⟩ .psi (t)rangle =left[1-frac ihbar int _t_0^tdt_1H(t_1)-frac 1hbar ^2int _t_0^tdt_1int _t_0^t_1dt_2H(t_1)H(t_2)+ldots right]|psi (t)rangle =left[1-frac ihbar int _t_0^tdt_1H(t_1)-frac 1hbar ^2int _t_0^tdt_1int _t_0^t_1dt_2H(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)⟩=iℏ∂|ψ(t)⟩∂t ,psi (t)rangle =ihbar frac partial partial t~,[H_0+lambda V(t)]|psi (t)rangle =ihbar frac partial partial t~,

assuming that the parameter λ is small and that the problem H0|n⟩=En|n⟩nrangle H_0|nrangle =E_n|nrangle has been solved.


Perform the following unitary transformation to the interaction picture (or Dirac picture),


|ψ(t)⟩=e−iℏH0(t−t0)|ψI(t)⟩ .psi (t)rangle =e^-frac ihbar H_0(t-t_0)|psi (t)rangle =e^-frac ihbar H_0(t-t_0)|psi _I(t)rangle ~.

Consequently, the Schrödinger equation simplifies to


λeiℏH0(t−t0)V(t)e−iℏH0(t−t0)|ψI(t)⟩=iℏ∂|ψI(t)⟩∂t ,psi _I(t)rangle =ihbar frac partial partial t~,lambda e^frac ihbar H_0(t-t_0)V(t)e^-frac ihbar H_0(t-t_0)|psi _I(t)rangle =ihbar frac partial partial t~,

so it is solved through the above Dyson series,


|ψI(t)⟩=[1−iλℏ∫t0tdt1eiℏH0(t1−t0)V(t1)e−iℏH0(t1−t0)−λ2ℏ2∫t0tdt1∫t0t1dt2eiℏH0(t1−t0)V(t1)e−iℏH0(t1−t0)eiℏH0(t2−t0)V(t2)e−iℏH0(t2−t0)+…]|ψ(t0)⟩ ,psi _I(t)rangle =left[1-frac ilambda hbar int _t_0^tdt_1e^frac ihbar H_0(t_1-t_0)V(t_1)e^-frac ihbar H_0(t_1-t_0)-frac lambda ^2hbar ^2int _t_0^tdt_1int _t_0^t_1dt_2e^frac ihbar H_0(t_1-t_0)V(t_1)e^-frac ihbar H_0(t_1-t_0)e^frac ihbar H_0(t_2-t_0)V(t_2)e^-frac ihbar H_0(t_2-t_0)+ldots right]|psi _I(t)rangle =left[1-frac ilambda hbar int _t_0^tdt_1e^frac ihbar H_0(t_1-t_0)V(t_1)e^-frac ihbar H_0(t_1-t_0)-frac lambda ^2hbar ^2int _t_0^tdt_1int _t_0^t_1dt_2e^frac ihbar H_0(t_1-t_0)V(t_1)e^-frac ihbar H_0(t_1-t_0)e^frac ihbar H_0(t_2-t_0)V(t_2)e^-frac ihbar H_0(t_2-t_0)+ldots right]|psi (t_0)rangle ~,

as a perturbation series with small λ.


Using the solution of the unperturbed problem H0|n⟩=En|n⟩nrangle H_0|nrangle =E_n|nrangle and ∑n|n⟩⟨n|=1displaystyle sum _nsum _n|nrangle langle n|=1 (for the sake of simplicity assume a pure discrete spectrum), yields, to first order,


|ψI(t)⟩=[1−iλℏ∑m∑n∫t0tdt1⟨m|V(t1)|n⟩e−iℏ(En−Em)(t1−t0)|m⟩⟨n|+…]|ψ(t0)⟩ .psi _I(t)rangle =left[1-frac ilambda hbar sum _msum _nint _t_0^tdt_1langle m|psi _I(t)rangle =left[1-frac ilambda hbar sum _msum _nint _t_0^tdt_1langle m|V(t_1)|nrangle e^-frac ihbar (E_n-E_m)(t_1-t_0)|mrangle langle n|+ldots right]|psi (t_0)rangle ~.

Thus, the system, initially in the unperturbed state |α⟩=|ψ(t0)⟩alpha rangle =|alpha rangle =|psi (t_0)rangle , by dint of the perturbation can go into the state |β⟩beta rangle |beta rangle . The corresponding transition probability amplitude to first order is


Aαβ=−iλℏ∫t0tdt1⟨β|V(t1)|α⟩e−iℏ(Eα−Eβ)(t1−t0) ,V(t_1)A_alpha beta =-frac ilambda hbar int _t_0^tdt_1langle beta |V(t_1)|alpha rangle e^-frac ihbar (E_alpha -E_beta )(t_1-t_0)~,

as detailed in the previous section——while the corresponding transition probability to a continuum is furnished by Fermi'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 above Dyson series, as


U(t)=1−iλℏ∫t0tdt1eiℏH0(t1−t0)V(t1)e−iℏH0(t1−t0)−λ2ℏ2∫t0tdt1∫t0t1dt2eiℏH0(t1−t0)V(t1)e−iℏH0(t1−t0)eiℏH0(t2−t0)V(t2)e−iℏH0(t2−t0)+⋯displaystyle U(t)=1-frac ilambda hbar int _t_0^tdt_1e^frac ihbar H_0(t_1-t_0)V(t_1)e^-frac ihbar H_0(t_1-t_0)-frac lambda ^2hbar ^2int _t_0^tdt_1int _t_0^t_1dt_2e^frac ihbar H_0(t_1-t_0)V(t_1)e^-frac ihbar H_0(t_1-t_0)e^frac ihbar H_0(t_2-t_0)V(t_2)e^-frac ihbar H_0(t_2-t_0)+cdots U(t)=1-frac ilambda hbar int _t_0^tdt_1e^frac ihbar H_0(t_1-t_0)V(t_1)e^-frac ihbar H_0(t_1-t_0)-frac lambda ^2hbar ^2int _t_0^tdt_1int _t_0^t_1dt_2e^frac ihbar H_0(t_1-t_0)V(t_1)e^-frac ihbar H_0(t_1-t_0)e^frac ihbar H_0(t_2-t_0)V(t_2)e^-frac ihbar H_0(t_2-t_0)+cdots

and take the perturbation V to be time-independent.


Using the identity resolution


∑n|n⟩⟨n|=1displaystyle sum _nsum _n|nrangle langle n|=1

with H0|n⟩=En|n⟩nrangle H_0|nrangle =E_n|nrangle for a pure discrete spectrum, write


U(t)=1−V−V+⋯displaystyle beginalignedU(t)=1&-leftV\&-leftright+cdots endalignedbeginalignedU(t)=1&-leftV\&-leftright+cdots endaligned

It is evident that, at second order, one must sum on all the intermediate states. Assume t0=0displaystyle t_0=0t_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 factor e−ϵtdisplaystyle e^-epsilon te^-epsilon t in the integrands for ε arbitrarily small. Thus the limit t → ∞ 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 yields


U(t)=1−iλℏ∑n⟨n|V|n⟩t−iλ2ℏ∑m≠n⟨n|V|m⟩⟨m|V|n⟩En−Emt−12λ2ℏ2∑m,n⟨n|V|m⟩⟨m|V|n⟩t2+…+λ∑m≠n⟨m|V|n⟩En−Em|m⟩⟨n|+λ2∑m≠n∑q≠n∑n⟨m|V|n⟩⟨n|V|q⟩(En−Em)(Eq−En)|m⟩⟨q|+…displaystyle nrangle t^2+ldots \&+lambda sum _mneq nfrac langle mE_n-E_mdisplaystyle nrangle t^2+ldots \&+lambda sum _mneq nfrac langle mE_n-E_m

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⟩)nrangle )|n(lambda )rangle =U(0;lambda )|nrangle ).)


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


In a similar way as for small perturbations, it is possible to develop a strong perturbation theory. Let us consider as usual the Schrödinger equation


H(t)|ψ(t)⟩=iℏ∂|ψ(t)⟩∂tdisplaystyle H(t)H(t)|psi (t)rangle =ihbar frac partial 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 [10] and the series is the well-known adiabatic series.[11] This approach is quite general and can be shown in the following way. Let us consider the perturbation problem


[H0+λV(t)]|ψ(t)⟩=iℏ∂|ψ(t)⟩∂tpsi (t)rangle =ihbar frac partial partial t[H_0+lambda V(t)]|psi (t)rangle =ihbar frac partial partial t

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


|ψ⟩=|ψ0⟩+1λ|ψ1⟩+1λ2|ψ2⟩+…displaystyle |psi rangle =|psi _0rangle +frac 1lambda |psi _1rangle +frac 1lambda ^2|psi _2rangle +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 as τ=λtdisplaystyle tau =lambda ttau =lambda t producing the following meaningful equations


V(t)|ψ0⟩=iℏ∂|ψ0⟩∂τdisplaystyle V(t)V(t)|psi _0rangle =ihbar frac partial partial tau

V(t)|ψ1⟩+H0|ψ0⟩=iℏ∂|ψ1⟩∂τpsi _0rangle =ihbar frac psi _1rangle partial tau V(t)|psi _1rangle +H_0|psi _0rangle =ihbar frac psi _1rangle partial tau

⋮displaystyle vdots vdots

that can be solved once we know the solution of the leading order equation. But we know that in this case we can use the adiabatic approximation. When V(t)displaystyle V(t)V(t) does not depend on time one gets the Wigner-Kirkwood series that is often used in statistical mechanics. Indeed, in this case we introduce the unitary transformation


|ψ(t)⟩=e−iℏλV(t−t0)|ψF(t)⟩psi (t)rangle =e^-frac ihbar lambda V(t-t_0)|psi (t)rangle =e^-frac ihbar lambda V(t-t_0)|psi _F(t)rangle

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


eiℏλV(t−t0)H0e−iℏλV(t−t0)|ψF(t)⟩=iℏ∂|ψF(t)⟩∂tpsi _F(t)rangle =ihbar frac partial partial te^frac ihbar lambda V(t-t_0)H_0e^-frac ihbar lambda V(t-t_0)|psi _F(t)rangle =ihbar frac partial partial t

and we see that the expansion parameter λ appears only into the exponential and so, the corresponding Dyson series, a dual Dyson series, is meaningful at large λs and is


|ψF(t)⟩=[1−iℏ∫t0tdt1eiℏλV(t1−t0)H0e−iℏλV(t1−t0)−1ℏ2∫t0tdt1∫t0t1dt2eiℏλV(t1−t0)H0e−iℏλV(t1−t0)eiℏλV(t2−t0)H0e−iℏλV(t2−t0)+…]|ψ(t0)⟩.psi _F(t)rangle =left[1-frac ihbar int _t_0^tdt_1e^frac ihbar lambda V(t_1-t_0)H_0e^-frac ihbar lambda V(t_1-t_0)-frac 1hbar ^2int _t_0^tdt_1int _t_0^t_1dt_2e^frac ihbar lambda V(t_1-t_0)H_0e^-frac ihbar lambda V(t_1-t_0)e^frac ihbar lambda V(t_2-t_0)H_0e^-frac ihbar lambda V(t_2-t_0)+ldots right]|psi _F(t)rangle =left[1-frac ihbar int _t_0^tdt_1e^frac ihbar lambda V(t_1-t_0)H_0e^-frac ihbar lambda V(t_1-t_0)-frac 1hbar ^2int _t_0^tdt_1int _t_0^t_1dt_2e^frac ihbar lambda V(t_1-t_0)H_0e^-frac ihbar lambda V(t_1-t_0)e^frac ihbar lambda V(t_2-t_0)H_0e^-frac ihbar lambda V(t_2-t_0)+ldots right]|psi (t_0)rangle .

After the rescaling in time τ=λtdisplaystyle tau =lambda ttau =lambda t we can see that this is indeed a series in 1/λdisplaystyle 1/lambda 1/lambda justifying in this way the name of dual Dyson series. The reason is that we have obtained this series simply interchanging H0 and V and we can go from one to another applying this exchange. This is called duality principle in perturbation theory. The choice H0=p2/2mdisplaystyle H_0=p^2/2mH_0=p^2/2m yields, as already said, a Wigner-Kirkwood series that is a gradient expansion. The Wigner-Kirkwood series is a semiclassical series with eigenvalues given exactly as for WKB approximation.[12]



Examples



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


Let us consider the quantum harmonic oscillator with the quartic potential perturbation and
the Hamiltonian


H=−ℏ22m∂2∂x2+mω2x22+λx4displaystyle H=-frac hbar ^22mfrac partial ^2partial x^2+frac momega ^2x^22+lambda x^4H=-frac hbar ^22mfrac partial ^2partial x^2+frac momega ^2x^22+lambda x^4

The ground state of the harmonic oscillator is


ψ0=(απ)14e−αx2/2displaystyle psi _0=left(frac alpha pi right)^frac 14e^-alpha x^2/2psi _0=left(frac alpha pi right)^frac 14e^-alpha x^2/2

(α=mω/ℏdisplaystyle alpha =momega /hbar alpha =momega /hbar ) and the energy of unperturbed ground state is


E0(0)=12ℏω.displaystyle E_0^(0)=tfrac 12hbar omega .,displaystyle E_0^(0)=tfrac 12hbar omega .,

Using the first order correction formula we get


E0(1)=λ(απ)12∫e−αx2/2x4e−αx2/2dx=λ(απ)12∂2∂α2∫e−αx2dxdisplaystyle E_0^(1)=lambda left(frac alpha pi right)^frac 12int e^-alpha x^2/2x^4e^-alpha x^2/2dx=lambda left(frac alpha pi right)^frac 12frac partial ^2partial alpha ^2int e^-alpha x^2dxE_0^(1)=lambda left(frac alpha pi right)^frac 12int e^-alpha x^2/2x^4e^-alpha x^2/2dx=lambda left(frac alpha pi right)^frac 12frac partial ^2partial alpha ^2int e^-alpha x^2dx

or


E0(1)=λ(απ)12∂2∂α2(πα)12=λ341α2=34ℏ2λm2ω2displaystyle E_0^(1)=lambda left(frac alpha pi right)^frac 12frac partial ^2partial alpha ^2left(frac pi alpha right)^frac 12=lambda frac 34frac 1alpha ^2=frac 34frac hbar ^2lambda m^2omega ^2E_0^(1)=lambda left(frac alpha pi right)^frac 12frac partial ^2partial alpha ^2left(frac pi alpha right)^frac 12=lambda frac 34frac 1alpha ^2=frac 34frac hbar ^2lambda m^2omega ^2


Example of first and second order perturbation theory – quantum pendulum


Consider the quantum mathematical pendulum with the Hamiltonian


H=−ℏ22ma2∂2∂ϕ2−λcos⁡ϕdisplaystyle H=-frac hbar ^22ma^2frac partial ^2partial phi ^2-lambda cos phi H=-frac hbar ^22ma^2frac partial ^2partial phi ^2-lambda cos phi

with the potential energy −λcos⁡ϕdisplaystyle -lambda cos phi -lambda cos phi taken as the perturbation i.e.


V=−cos⁡ϕdisplaystyle V=-cos phi displaystyle V=-cos phi

The unperturbed normalized quantum wave functions are those of the rigid rotor and are given by


ψn(ϕ)=einϕ2πdisplaystyle psi _n(phi )=frac e^inphi sqrt 2pi psi _n(phi )=frac e^inphi sqrt 2pi

and the energies


En(0)=ℏ2n22ma2displaystyle E_n^(0)=frac hbar ^2n^22ma^2E_n^(0)=frac hbar ^2n^22ma^2

The first order energy correction to the rotor due to the potential energy is


En(1)=−12π∫e−inϕcos⁡ϕeinϕ=−12π∫cos⁡ϕ=0displaystyle E_n^(1)=-frac 12pi int e^-inphi cos phi e^inphi =-frac 12pi int cos phi =0E_n^(1)=-frac 12pi int e^-inphi cos phi e^inphi =-frac 12pi int cos phi =0

Using the formula for the second order correction one gets


En(2)=ma22π2ℏ2∑k|∫e−ikϕcos⁡ϕeinϕ|2n2−k2displaystyle E_n^(2)=frac ma^22pi ^2hbar ^2sum _kfrac int e^-ikphi cos phi e^inphi rightn^2-k^2E_n^(2)=frac ma^22pi ^2hbar ^2sum _kfrac int e^-ikphi cos phi e^inphi rightn^2-k^2

or


En(2)=ma22ℏ2∑k|(δn,1−k+δn,−1−k)|2n2−k2displaystyle E_n^(2)=frac ma^22hbar ^2sum _kfrac left(delta _n,1-k+delta _n,-1-kright)rightn^2-k^2E_n^(2)=frac ma^22hbar ^2sum _kfrac left(delta _n,1-k+delta _n,-1-kright)rightn^2-k^2

or


En(2)=ma22ℏ2(12n−1+1−2n−1)=ma2ℏ214n2−1displaystyle E_n^(2)=frac ma^22hbar ^2left(frac 12n-1+frac 1-2n-1right)=frac ma^2hbar ^2frac 14n^2-1E_n^(2)=frac ma^22hbar ^2left(frac 12n-1+frac 1-2n-1right)=frac ma^2hbar ^2frac 14n^2-1


References




  1. ^ 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..mw-parser-output cite.citationfont-style:inherit.mw-parser-output .citation qquotes:"""""""'""'".mw-parser-output .citation .cs1-lock-free abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/6/65/Lock-green.svg/9px-Lock-green.svg.png")no-repeat;background-position:right .1em center.mw-parser-output .citation .cs1-lock-limited a,.mw-parser-output .citation .cs1-lock-registration abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/d/d6/Lock-gray-alt-2.svg/9px-Lock-gray-alt-2.svg.png")no-repeat;background-position:right .1em center.mw-parser-output .citation .cs1-lock-subscription abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/a/aa/Lock-red-alt-2.svg/9px-Lock-red-alt-2.svg.png")no-repeat;background-position:right .1em center.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registrationcolor:#555.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration spanborder-bottom:1px dotted;cursor:help.mw-parser-output .cs1-ws-icon abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/4/4c/Wikisource-logo.svg/12px-Wikisource-logo.svg.png")no-repeat;background-position:right .1em center.mw-parser-output code.cs1-codecolor:inherit;background:inherit;border:inherit;padding:inherit.mw-parser-output .cs1-hidden-errordisplay:none;font-size:100%.mw-parser-output .cs1-visible-errorfont-size:100%.mw-parser-output .cs1-maintdisplay:none;color:#33aa33;margin-left:0.3em.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-formatfont-size:95%.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-leftpadding-left:0.2em.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-rightpadding-right:0.2em


  2. ^ Schrödinger, E. (1926). "Quantisierung als Eigenwertproblem" [Quantification of the eigen value problem]. Annalen der Physik (in German). 80 (13): 437–490. Bibcode:1926AnP...385..437S. doi:10.1002/andp.19263851302.


  3. ^ Rayleigh, J. W. S. (1894). Theory of Sound. I (2nd ed.). London: Macmillan. pp. 115–118. ISBN 978-1-152-06023-4.


  4. ^ Sulejmanpasic, Tin; Ünsal, Mithat (2018-07-01). "Aspects of perturbation theory in quantum mechanics: The BenderWuMathematica® package". Computer Physics Communications. 228: 273–289. doi:10.1016/j.cpc.2017.11.018. ISSN 0010-4655.


  5. ^ Sakurai, J.J., and Napolitano, J. (1964,2011). Modern quantum mechanics (2nd ed.), Addison Wesley
    ISBN 978-0-8053-8291-4. Chapter 5



  6. ^ Landau, L. D.; Lifschitz, E. M. (1977). Quantum Mechanics: Non-relativistic Theory (3rd ed.). ISBN 978-0-08-019012-9.


  7. ^ Bir, Gennadiĭ Levikovich; Pikus, Grigoriĭ Ezekielevich (1974). "Chapter 15: Perturbation theory for the degenerate case". Symmetry and Strain-induced Effects in Semiconductors. ISBN 978-0-470-07321-6.


  8. ^ 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.


  9. ^ Albert Messiah (1966). Quantum Mechanics, North Holland, John Wiley & Sons.
    ISBN 0486409244; J. J. Sakurai (1994). Modern Quantum Mechanics (Addison-Wesley)
    ISBN 9780201539295.



  10. ^ 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.


  11. ^ 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.


  12. ^ 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.








Popular posts from this blog

How to check contact read email or not when send email to Individual?

Bahrain

Postfix configuration issue with fips on centos 7; mailgun relay