DOC PREVIEW
MIT 8 512 - Linear Response Theory

This preview shows page 1-2-3-4 out of 11 pages.

Save
View full document
View full document
Premium Document
Do you want full access? Go Premium and unlock all 11 pages.
Access to all documents
Download any document
Ad free experience
View full document
Premium Document
Do you want full access? Go Premium and unlock all 11 pages.
Access to all documents
Download any document
Ad free experience
View full document
Premium Document
Do you want full access? Go Premium and unlock all 11 pages.
Access to all documents
Download any document
Ad free experience
View full document
Premium Document
Do you want full access? Go Premium and unlock all 11 pages.
Access to all documents
Download any document
Ad free experience
Premium Document
Do you want full access? Go Premium and unlock all 11 pages.
Access to all documents
Download any document
Ad free experience

Unformatted text preview:

Lecture 1: Linear Response TheoryResponse Functions and the Interaction RepresentationResponse FunctionsElectron Density Response to an Applied Electric PotentialSanity Check: Free FermionsThe Correlation Function S("017Er, t)Measuring S("017Eq, )MIT OpenCourseWare http://ocw.mit.edu 8.512 Theory of Solids IISpring 2009For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms.Lecture 1: Linear Response Theory Last semester in 8.511, we discussed linear response theory in the context of charge screening and the free-fermion polarization function. This theory c an be extended to a much wider range of areas, however, and is a very useful tool in solid state physics. We’ll begin this semester by going back and studying linear response theory again with a more formal approach, and then returning to this like superconductivity a bit later. 1.1 Response Functions and the Interaction Representation In solid state physics, we ordinarily think about many-body systems, with something on the order of 1023 particles. With so many particles, it is usually impossible to even think about a wave function for the whole system. As a result, it is often more useful for us to think in terms of the macroscopic observable behaviors of systems rather than their particular microscopic states. One example of such a macroscopic property is the magnetic susceptibility χH ≡ ∂M , which ∂H is a measure of the response of the net magnetization M of a system to an applied magnetic field H (�r, t). This is the type of behavior we will be thinking about: we can mathematically probe the system with some perturbing external probe or field (e.g. H (�r, t)), and try to predict what the system’s response will be in terms of the expectation values of some observable quantities. Let ˆH be the full many-body Hamiltonian for some isolated system that we are interested in. We spent most of 8.511 thinking about how to solve for the behavior of a system governed by ˆH. As interesting as that behavior may be, we will now consider that to be a solved problem. That ˆis, we will assume the e xistence of a set of eigenkets {|n�} that diagonalize H with associated eigenvalues (energies) En. In addition to ˆH, we now turn on an external probe potential Vˆ, such that the total Hamil-tonian HT ot satisfies: ˆ ˆHT ot = H + Vˆ(1.1) In particular, we are interested in probe potentials that arise from the coupling of some external scalar or vector field to s ome sort of “density” in the sample. For example, the external field can be an electric potential U (�r, t), which couples to the electronic charge density ρˆ(�r) such that � ˆV = d�r ˆρ (�r) U (�r, t) (1.2) V where the electron density operator ˆρ (�r) is given by N� ˆρ (�r) = δ (�r − �ri) (1.3) i=1 1� �� �� �� ���Response Functions and the Interaction Representation 2 In first quantized language, with r�ibeing the position of electron i the N-electron system. In second quantized notation, recall ρˆ(�r) = Ψ† (�r) Ψ (�r) (1.4) where Ψ† (�r) and Ψ (�r) are the electron field creation and annihilation operators, respectively. The momentum space version of the electron density operator, ˆρ (q�), is related to ρˆ(�r) through the Fourier transforms: i� �ρˆ(�r) = e q· rρˆ(�q) (1.5) q i��rΨ (�r) = e k·c�(1.6)k k such that e−i� �rρˆ(q�) = q·(1.7) r = c†c�(1.8)k−�kq k Equation (1.7) is the first quantized form of ˆρ (�q), and equation (1.8) is the second quantized form with c†the creation ope rator for an electron with momentum1 �k−q� and c�the destruction k−�kq operator for an electron with momentum �k. Returning to equation (1.1), we’d like to think about Vˆas a perturbation on the external field-free system Hamiltonian ˆH as the unperturbed Hamil-H. This leads us naturally to consider ˆˆtonian within the interaction picture representation. Recall that this H is a very complicated beast with all of the electron-electron repulsions included, but for our purposes we just take as a given that there are a set of eigenstates and energies that diagonalize this Hamiltonian. Recall the formulation of the interaction representation: ∂ h∂t|φ (t)� = (ˆi¯ H + Vˆ) φ (t)� (1.9)|ˆWe can “unwind” the natural time dependence due to H from the state ket φ (t)� to form an interaction representation state ket φ˜(t)�I by || ˜Ht|φ (t)�I = e i ˆφ (t)� (1.10)|Htφ˜(t)� (1.11)|φ (t)�I = e−i ˆ| Note that in the absence of Vˆ, these interaction picture state kets are actually the Heisenberg picture state kets of the system. Also, we have now officially set ¯h = 1. After substituting (1.11) into (1.9), we obtain i¯Ht ˆHt˜h∂ = e i ˆV e−i ˆφ (t)� (1.12)∂t | ˆ= VIφ˜(t)� (1.13)| 1 �k and q� are actually wavevectors, which differ from momenta by a factor of ¯ h = 1. h. When in doubt, assume ¯Response Functions and the Interaction Representation 3 where we have set Ht ˆHt VˆI = e i ˆV e−i ˆ(1.14) Thus the interaction picture state ket evolves simply according to the dynamics governed solely by the interaction picture perturbing potential VˆI. More generally, we c an write any observable (operator) in the interaction picture as ˆHt ˆHt AI = e i ˆAe−i ˆ(1.15) We can integrate equation (1.12) with respect to t to get � t ˆdt� VI (t�) φ˜(t�)� (1.16)|φ˜(t)� = |φ0� − i |−∞ At first it seems like we have not done much to benefit ourselves, since all we have done is to convert the ordinary Schrodinger equation, a PDE, into an integral equation. However, if VˆI is small, then we can iterate equation (1.16): � t dt� VˆI (t�) |φ0� + · · · (1.17)|φ˜(t)� ≈ |φ0� − i −∞ The essence of linear response theory is that we focus ourselves on cases where VˆI is suffi-ciently weak that the perturbation series represented by equation (1.17) has essentially converged after including just the first non-trivial term listed ab ove. This term is linear in VˆI. Throughout this discussion, we will be working at T = 0, so φ0� is simply the ground state ˆ|of the non-perturbe d total system Hamiltonian H. Note that we have taken our initial time, i.e. the lower limit of


View Full Document
Download Linear Response Theory
Our administrator received your request to download this document. We will send you the file to your email shortly.
Loading Unlocking...
Login

Join to view Linear Response Theory and access 3M+ class-specific study document.

or
We will never post anything without your permission.
Don't have an account?
Sign Up

Join to view Linear Response Theory 2 2 and access 3M+ class-specific study document.

or

By creating an account you agree to our Privacy Policy and Terms Of Use

Already a member?