Lecture 2 : Quantum Integrable Systems
Total Page:16
File Type:pdf, Size:1020Kb
Lecture 2 : Quantum Integrable Systems Maxime Fairon 2 December 2020 1 Motivation and definition Let us recall the classical picture which is at the basis of integrability. First, we are given a real phase space M of dimension 2n, which we assume for simplicity to admit global Darboux coordinates in terms of n positions q and n momenta p. In convenient situations, we may be able to split M with respect to these two sets of coordinates as M ' Mq × Mp. Then, we are also given a Hamiltonian H(p; q) 2 C1(M), which could take the suitable form n 1 X H = p2 + V;V = V (q) 2 C1(M ) : (1.1) 2 j q j=1 Finally, we have a canonically defined Poisson bracket which allows us to consider the following equations of motion : df = ff; Hg ; (1.2) dt for any function f 2 C1(M) whose evolution we are interested in studying. The quantum picture that we are going to introduce is motivated by quantum mechanics, and we need to reinterpret the classical construction. We are still considering the space Mq where we can measure positions, but quantum mechanics now tells us that we should be interested in the probability of finding a given system in a particular configuration. We are led to consider a complex 2 n Hilbert space S of states depending on Mq, for example the Hilbert space L (R ; dq) of complex- n valued square-integrable functions on Mq = R . What will be important from the point of view of 2 n physics is to consider the set S1 of states of unit norm, such as the unit-length vectors in L (R ; dq), which can be used to compute the probability of finding a particular system in a region of Mq or to make \measurements". To be more precise, let us introduce an algebra A of operators, called observables, acting on S. Given an element F 2 A, we can compute the expectation value of F in a state 2 S1 which is given by E [F] := h ; F i ; (1.3) where h−; −i denotes the inner product on S. A state will evolve over time according to Schr¨odinger'sequation : @ 2 i = − ~ ∆ + V ; V = V (q) 2 C1(M ) : (1.4) ~ @t 2m q Pn 2 2 1 Here, we use the Laplacian ∆f = j=1 @ f=@qj for f 2 C (Mq), while ~ is the Planck constant. From a more modest point of view, these considerations lead us to the problem of solving (1.4) with 1 m = 1 and ~ being seen as a formal parameter, from which we get the following simplified version n 2 @ 1 X @ i = H ; H = −i + V: (1.5) ~ @t 2 ~@q j=1 j The element H, called the quantum Hamiltonian, is an element of the algebra A of operators acting on S. If the operator H is time-independent, we can formally solve (1.5) by introducing the one- parameter group of operators i U(t) = exp − tH ; t 2 R : (1.6) ~ @U(t) It suffices to note that i~ @t = U(t)H = HU(t), in order to get the solution (t) = U(t) (0) for the initial condition (0). Furthermore, we can use U to obtain the evolution of an observable F 2 A. Mathematically, we note that1 −1 E (t)[F] = h (t); F (t)i = h (0); (U(t) FU(t)) (0)i = E (0)[F(t)] ; (1.7) for the one-parameter group of operators F(t) = U(t)−1FU(t). This can be reformulated as saying that the expectation value of the observable F in the state (t) coincides with the expectation value of the evolution operator F(t) in the initial state (0). We obtain in that way the Heisenberg picture of dynamics for observables n 2 dF(t) −i 1 X @ = [F(t); H] ; H = −i + V; (1.8) dt 2 ~@q ~ j=1 j where [−; −] denotes the commutator [A; B] = AB − BA. The classical (1.2) and quantum (1.8) equations of motion are quite similar. The role of the Pois- son bracket {−; −} is now played by the (rescaled) commutator, and both operations are bideriva- tions. Comparing the quantum Hamiltonian H in (1.5) with its classical counterpart (1.1) suggests that each momentum pj has its role now being played by the derivation −i~@=@qj. Hence we are \replacing" the Poisson algebra C1(M) by the algebra of operators A acting on S. Put together, we can give a rough correspondence between the classical and the quantum settings as in Table1. In order to understand what is the analogue of an integrable system in the quantum case, let us try to solve (1.5) by decomposing this equation with respect to an eigenbasis for H of normalised 2 eigenfunctions (φ`) corresponding to a discrete set of complex eigenvalues (E`), i.e. φ` 2 S1 is such that Hφ` = E`φ` : (1.9) P Then the evolution of some state (q; 0) comes from its decomposition as (q; 0) = ` c`φ`(q) for some constants (c`). Indeed, it is simply given by X −iE`t=~ (q; t) = c` e φ`(q) : (1.10) ` 1Here we require that H is a self-adjoint operator, which means that hHφ, i = hφ, H i holds for all φ, 2 S. This implies that U(t) is unitary, that is hU(t)φ, i = hφ, U(t)−1 i. 2The eigenvalues are real if H is self-adjoint. 2 Classical Quantum Poisson algebra of functions C1(M) A Algebra of operators Classical Hamiltonian H H Quantum Hamiltonian Poisson bracket {−; −} −i [−; −] Commutator ~ Classical equation of motion df = ff; Hg dF = −i [F; H] Heisenberg equation dt dt ~ Table 1: correspondence between classical and quantum notions. So now the question reduces to find an eigenbasis for H. By definition, an operator F 2 A commutes with H if [F; H] = FH − HF = 0 : (1.11) The key remark is to realise that if F commutes with H, then HFφ` = FHφ` = E`Fφ` ; (1.12) so that Fφ` is itself an eigenvector of H of eigenvalue E`. Therefore, we can consider the more precise problem of finding a common normalised eigenbasis of H and F, that is we want 0 0 φ` 2 S1 such that Hφ` = E`φ`; Fφ` = E`φ`;E`;E` 2 C : (1.13) This strategy can be repeated with an operator commuting with both H and F, then a fourth operator commuting with the first three ones, and so on and so forth. But when should we stop? 1 In the classical case, we limited our choice of Poisson commuting elements to n = 2 dim(M) = dim(Mq) functions which are functionally independent, so we shall introduce a similar property of independence. Definition 1.1. A set of operators H1;:::; Hn 2 A is a quantum (completely) integrable system if these operators pairwise commute (i.e. [Hi; Hj] = 0 for all 1 ≤ i; j ≤ n) and there exists no relation between them. A word of warning here : the existence of a relation usually means that we can write the identity P (H1;:::; Hn) = 0 ; (1.14) for some nonzero polynomial P 2 C[z1; : : : ; zn] in n variables. (Since the operators (Hi) pair- wise commute, replacing the variables (zi) by them yields a well-defined operator.) Depending on the context, P may have a slightly different form, e.g. we could seek independence over real- algebraic/analytic functions in n variables. As an example, note that for i 6= j there exists a relation 2 2 2 2 between the three commuting operators F1 = @ =@qi , F2 = @ =@qj , and F3 = @=@qi + @=@qj, since 2 2 for P (z1; z2; z3) = (z3 − z1 − z2) − 4z1z2 ; we have P (F1; F2; F3) = 0 : (1.15) There is a widespread use of a more convenient definition of quantum integrability which is motivated by the classical case, and it has the advantage of getting the independence of the operators \for free". This is based on the method of deformation/quantization. The quantization process consists in making the formal replacement @ pj −i~ ; (1.16) @qj 3 to go from a polynomial function on M to an operator belonging to A. There is an obvious issue of ordering here, since the function qjpj can be quantized in the many forms @ @ @ αqjpj + (1 − α)pjqj −αi~qj − (1 − α)i~ qj = −i~qj + (α − 1)i~ ; α 2 C : @qj @qj @qj Nevertheless, it is well-defined for the Hamiltonian operators of interest since n n 2 1 X 1 X @ H = p2 + V (q) H = −i + V (q) : (1.17) 2 j 2 ~@q j=1 j=1 j The \opposite" of the map (1.16) is given by the replacement @ − i~ pj ; then ~ ! 0 ; (1.18) @qj and it is called the reduction. For example, H is the reduction of H in (1.17). Note that performing quantization (1.16) then reduction (1.18) returns the original function, but the opposite may not be true. With this set up, we have the following alternative definition of quantum integrability. Definition 1.2 (Alternative to Definition 1.1). A set of operators H1;:::; Hn 2 A is a quan- tum (completely) integrable system if these operators pairwise commute ([Hi; Hj] = 0) and their reductions H1;:::;Hn are functionally independent. 2 Examples of quantum integrable systems Remark 2.1. To ease notations when dealing with examples, it is usual to fix ~ to some value in C n f0g. In our examples, we will work with ~ = −i so that the Hamiltonian operators are of the 1 Pn @2 @ form H = 2 j=1 2 + V , and quantization is just pj @q .