Representation theory and quantum mechanics tutorial Spin and the hydrogen atom

Similar documents
Chapter 2 The Group U(1) and its Representations

Physics 221A Fall 1996 Notes 12 Orbital Angular Momentum and Spherical Harmonics

The 3 dimensional Schrödinger Equation

Introduction to Group Theory

MAT265 Mathematical Quantum Mechanics Brief Review of the Representations of SU(2)

REPRESENTATION THEORY WEEK 7

Quantum Theory and Group Representations

Total Angular Momentum for Hydrogen

Symmetries, Groups, and Conservation Laws

Lecture 11 Spin, orbital, and total angular momentum Mechanics. 1 Very brief background. 2 General properties of angular momentum operators

Algebra I Fall 2007

THE EULER CHARACTERISTIC OF A LIE GROUP

Math 594. Solutions 5

Spherical Harmonics on S 2

ALGEBRA QUALIFYING EXAM PROBLEMS LINEAR ALGEBRA

1 Classifying Unitary Representations: A 1

Angular momentum. Quantum mechanics. Orbital angular momentum

MAT 445/ INTRODUCTION TO REPRESENTATION THEORY

Recall that any inner product space V has an associated norm defined by

Physics 221A Fall 1996 Notes 14 Coupling of Angular Momenta

1 Revision to Section 17.5: Spin

MATH 423 Linear Algebra II Lecture 33: Diagonalization of normal operators.

Notation. For any Lie group G, we set G 0 to be the connected component of the identity.

INTRODUCTION TO LIE ALGEBRAS. LECTURE 2.

Lecture 10: A (Brief) Introduction to Group Theory (See Chapter 3.13 in Boas, 3rd Edition)

Quantum Mechanics for Mathematicians: Energy, Momentum, and the Quantum Free Particle

Quantum Physics and the Representation Theory of SU(2)

Isotropic harmonic oscillator

Second quantization: where quantization and particles come from?

Symmetries for fun and profit

L(C G (x) 0 ) c g (x). Proof. Recall C G (x) = {g G xgx 1 = g} and c g (x) = {X g Ad xx = X}. In general, it is obvious that

SPHERICAL UNITARY REPRESENTATIONS FOR REDUCTIVE GROUPS

BRST and Dirac Cohomology

Symmetries and particle physics Exercises

Chapter 2 Linear Transformations

Addition of Angular Momenta

Since G is a compact Lie group, we can apply Schur orthogonality to see that G χ π (g) 2 dg =

Quantum Mechanics Solutions

Problems in Linear Algebra and Representation Theory

Topics in Representation Theory: Fourier Analysis and the Peter Weyl Theorem

Linear Algebra: Matrix Eigenvalue Problems

LECTURE 4: REPRESENTATION THEORY OF SL 2 (F) AND sl 2 (F)

Quantum Mechanics Solutions. λ i λ j v j v j v i v i.

LINEAR ALGEBRA BOOT CAMP WEEK 4: THE SPECTRAL THEOREM

Lecture 4 Quantum mechanics in more than one-dimension

Lecture 45: The Eigenvalue Problem of L z and L 2 in Three Dimensions, ct d: Operator Method Date Revised: 2009/02/17 Date Given: 2009/02/11

G : Quantum Mechanics II

1 Mathematical preliminaries

CHAPTER 6. Representations of compact groups

Highest-weight Theory: Verma Modules

Lecture 4 Quantum mechanics in more than one-dimension

Notes on nilpotent orbits Computational Theory of Real Reductive Groups Workshop. Eric Sommers

QM and Angular Momentum

Rotations in Quantum Mechanics

Linear Algebra. Min Yan

Physics 342 Lecture 26. Angular Momentum. Lecture 26. Physics 342 Quantum Mechanics I

Consistent Histories. Chapter Chain Operators and Weights

Notes on SU(3) and the Quark Model

Summary: angular momentum derivation

Quantum Theory of Angular Momentum and Atomic Structure

The AKLT Model. Lecture 5. Amanda Young. Mathematics, UC Davis. MAT290-25, CRN 30216, Winter 2011, 01/31/11

Topics in Representation Theory: Roots and Complex Structures

Physics 221A Fall 2017 Notes 27 The Variational Method

LECTURE 25-26: CARTAN S THEOREM OF MAXIMAL TORI. 1. Maximal Tori

Generators for Continuous Coordinate Transformations

Time Independent Perturbation Theory Contd.

Differential Topology Final Exam With Solutions

Induced Representations and Frobenius Reciprocity. 1 Generalities about Induced Representations

The following definition is fundamental.

Page 404. Lecture 22: Simple Harmonic Oscillator: Energy Basis Date Given: 2008/11/19 Date Revised: 2008/11/19

-state problems and an application to the free particle

MATH 583A REVIEW SESSION #1

is an isomorphism, and V = U W. Proof. Let u 1,..., u m be a basis of U, and add linearly independent

Angular momentum and spin

MATHEMATICS 217 NOTES

20 The Hydrogen Atom. Ze2 r R (20.1) H( r, R) = h2 2m 2 r h2 2M 2 R

PHY 407 QUANTUM MECHANICS Fall 05 Problem set 1 Due Sep

REPRESENTATION THEORY NOTES FOR MATH 4108 SPRING 2012

1. General Vector Spaces

which implies that we can take solutions which are simultaneous eigen functions of

IRREDUCIBLE REPRESENTATIONS OF SEMISIMPLE LIE ALGEBRAS. Contents

Appendix: SU(2) spin angular momentum and single spin dynamics

Outline Spherical symmetry Free particle Coulomb problem Keywords and References. Central potentials. Sourendu Gupta. TIFR, Mumbai, India

Lecture If two operators A, B commute then they have same set of eigenkets.

Introduction to Modern Quantum Field Theory

LECTURE 3: REPRESENTATION THEORY OF SL 2 (C) AND sl 2 (C)

Part III Symmetries, Fields and Particles

Math 350 Fall 2011 Notes about inner product spaces. In this notes we state and prove some important properties of inner product spaces.

Angular Momentum in Quantum Mechanics

The Spinor Representation

12. Hilbert Polynomials and Bézout s Theorem

Symmetries, Fields and Particles 2013 Solutions

Geometry of the Special Unitary Group

Math 396. Quotient spaces

Systems of Identical Particles

Math 210B. Why study representation theory?

(1.1) In particular, ψ( q 1, m 1 ; ; q N, m N ) 2 is the probability to find the first particle

1. Matrix multiplication and Pauli Matrices: Pauli matrices are the 2 2 matrices. 1 0 i 0. 0 i

CLASSICAL GROUPS DAVID VOGAN

1 Commutators (10 pts)

Transcription:

Representation theory and quantum mechanics tutorial Spin and the hydrogen atom Justin Campbell August 3, 2017 1 Representations of SU 2 and SO 3 (R) 1.1 The following observation is long overdue. Proposition 1.1.1. The groups SU n and SO n (R) are compact. Proof. Recall that the Heine-Borel theorem says that a subset of Euclidean space is compact if and only if it is closed and bounded. First, notice that SU n and SO n (R) are closed in Mat n n (R) = R n2, being defined by the equations AA = I = A A and AA = I = A A respectively. They are bounded because all columns of a unitary or orthogonal matrix have unit length (for the Hermitian and ordinary dot products respectively). In particular, all representations of SU n and SO n (R) are completely decomposable. Thus the problem of classifying representations of these groups up to isomorphism reduces to the problem of classifying their irreducible representations up to isomorphism. Consider the case G = SU 2. We will now construct a family of irreducible SU 2 -representations, and then prove that our list is complete. Namely, let V n := {degree n homogeneous polynomials in two variables with C-coefficients}. More concretely, as a C-vector space V n has a basis consisting of the monomials x i y j for i + j = n, with i, j 0. In particular V n is (n + 1)-dimensional over C. Thinking of f(x, y) V n as a polynomial function on C 2, the action of A SU 2 is defined by the formula A f(x, y) = f(a 1 (x, y)). It is not hard to see that this defines a smooth (even R-polynomial) homomorphism SU 2 GL(V n ). For example, the representation V 0 = C 1 is trivial, and V 1 is the standard representation of SU 2 on C 2. Theorem 1.1.2. The SU 2 -representations V n for n 0 are all irreducible, and any (n + 1)-dimensional irreducible representation of SU 2 is isomorphic to V n. The following results demonstrate the utility of Lie algebras for studying group representations. Proposition 1.1.3. Suppose that G is connected and that V is a G-representation. Then a subspace of V is G-stable if and only if it is g-stable. Corollary 1.1.3.1. A representation V of a connected group G is irreducible as a G-representation if and only if it is irreducible as a g-representation. Moreover, it is not hard to show that we can replace g by g R C in these statements. 1

1.2 In this section we prove Theorem 1.1.2. By Corollary 1.1.3.1, in order to show that V n is irreducible as an SU 2 -representation it suffices to prove that it is irreducible as an su 2 -representation, or equivalently as an sl 2 (C)-representation because su 2 R C = sl 2 (C). Recall that sl 2 (C) has the standard C-basis e, f, h. The basic relations are [h, e] = 2e, [h, f] = 2f, and [e, f] = h. One computes that, as operators on polynomials in two variables x, y, we have e = x y, f = y x, and h = x x y y. In particular, on a basis vector x i y j for i + j = n, we have e x i y j = jx i+1 y j 1, f x i y j = ix i 1 y j+1, and h x i y j = (i j)x i y j. Thus h is diagonalizable with eigenvalues n, n 2,, n + 2, n. For this reason we call (C-multiples of) x n and y n highest weight and lowest weight vectors respectively, and e and f are called raising and lowering operators respectively. Irreducibility of V n is immediate from this description. If g V n and d is the y-degree of g, then f d g is a heighest weight vector, i.e. is proportional to x n. Since e j x n is proportional to x i y j for any 0 j n, it follows that g generates V n as an sl 2 (C)-representation. Suppose w W is an eigenvector of h with eigenvalue λ C. Then we have h (e w) = [h, e] w + e (h w) = (λ + 2)e w, so that e w is either zero or an eigenvector of h with eigenvalue λ + 2. Similarly f w is either zero or an eigenvector of h with eigenvalue λ 2. Let λ be the eigenvalue of h acting on W with the greatest real part, and fix an eigenvector w λ W of h with eigenvalue λ. Define w λ 2j := 1 j! f j w λ for j 0. By the computation just performed w λ 2j is either zero or an eigenvector with eigenvalue λ 2j. Moreover, by induction we have e w λ 2j = (λ j + 1)w λ 2j+2. Let k 1 be the least positive integer such that w λ 2k = 0. Then 0 = e w λ 2k = (λ k + 1)w λ 2k+2, and since w λ 2k+2 0, it follows that k = λ + 1 (in particular λ is a nonnegative integer). Moreover, our calculations show that w λ, w λ 2,, w λ span an sl 2 (C)-invariant subspace of W, which by irreducibility must be W itself. Since the w λ, w λ 2,, w λ have distinct eigenvalues they are independent and hence form a basis. In particular λ = n. Now it is easy to check, using the calculations already performed, that the linear map V n W defined by x i y j w n 2j is an sl 2 (C)-equivariant isomorphism. 1.3 Next, we classify representations of G = SO 3 (R). The groups SO 3 (R) and SU 2 are related as follows. Consider the adjoint action of SU 2 on its Lie algebra su 2, given by conjugation: A B = ABA 1 for A SU 2 and B su 2. We take as a basis of su 2 the skew-hermitian Pauli matrices iσ 1, iσ 2, iσ 3, which yields an isomorphism R 3 = su 2. Thus we obtain a three-dimensional real matrix representation of SU 2, i.e. a smooth group homomorphism π : SU 2 GL 3 (R). Proposition 1.3.1. The image of π : SU 2 GL 3 (R) is SO 3 (R), and its kernel is {±I}. 2

Proof. To see that π(su 2 ) O 3 (R), we first observe that the dot product in R 3 corresponds to the pairing, : su 2 su 2 R given by B, C = 1 2 tr(bc). But SU 2 preserves this inner product by conjugation-invariance of trace: AB 1 A 1, AB 2 A 1 = 1 2 tr(ab 1B 2 A 1 ) = 1 2 tr(b 1B 2 ) = B 1, B 2. Now we show that det π(a) = 1 for A SU 2. For this, recall that O 3 (R) is the disjoint union of two components, and SO 3 (R) is the component containing I. On the other hand SU 2 is connected and π is continuous, so π(i) = I implies π(su 2 ) SO 3 (R) as desired. For the surjectivity of π, recall that SO 3 (R) is generated by rotations Rθ x, Ry θ, Rz θ about the x-, y-, and z-axes respectively. We show that Rθ z is in the image of π, the other axes being similar. For this, put [ ] e iθ 0 A θ := 0 e iθ. We compute directly that A θ iσ 1 A 1 θ = cos(2θ)iσ 1 sin(2θ)iσ 2, A θ iσ 2 A 1 θ = sin(2θ)iσ 1 + cos(2θ)iσ 2, and A θ iσ 3 A 1 θ = iσ 3. This implies that π(a θ ) = R2θ z, whence the surjectivity. Finally, we show that ker π = {±I}. If π(a) = I, then Aσ 1 = σ 1 A, which immediately implies that A = A θ for some θ R. Now σ 1 = Aσ 1 A 1 = cos(2θ)σ 1 sin(2θ)σ 2 implies that sin(2θ) = 0, so A = ±I as desired. From now on we view π as a surjective homomorphism π : SU 2 SO 3 (R). In particular, any representation of SO 3 (R) gives rise to an SU 2 -representation by restriction along π. A representation of SU 2 comes from an SO 3 (R)-representation if and only if I SU 2 acts trivially on the representation. Proposition 1.3.2. The Lie algebra homomorphism π : su 2 so 3 (R) is an isomorphism which sends iσ 1 2r x, iσ 2 2r y, and iσ 3 2r z. Proof. The calculation π (iσ 3 ) = 2r z was essentially done in the proof of Proposition 1.3.1, and the others are similar. In particular, it follows that so 3 (R) R C is canonically isomorphic to sl 2 (C). Proposition 1.3.3. The SU 2 -representation V n comes from SO 3 (R) by restriction along π if and only if n is even, i.e. V n has odd dimension. Any irreducible representation of SO 3 (R) is isomorphic to one of this form, and in particular has odd dimension. Proof. The action of SU 2 on V n factors through π if and only if ker π = {±I} acts trivially on V n. Since I SU 2 acts by ( I f)(x, y) = f( x, y), this endomorphism of V n is trivial if and only if n is even (so that f(x, y) has even degree). Suppose W is an irreducible representation of SO 3 (R), hence of so 3 (R). Since so 3 (R) R C = sl 2 (C), we can view W as an irreducible sl 2 (C)-representation. But any such is isomorphic to V n for some n 0. 3

1.4 The irreducible representations V 0, V 2, V 4, of SO 3 (R) from the previous sections are simple to construct as vector spaces, but the action of SO 3 (R) is not manifest. Rather, the SU 2 -action on V n is obvious, and then one checks that it is induced by an SO 3 (R)-action for n even. We now construct the irreducible representations of SO 3 (R) in a way better-adapted to this group. Namely, we realize these representations using functions on a space with SO 3 (R)-action. To this end, consider the tautological action of SO 3 (R) on R 3 by rotations. This induces an SO 3 (R)-action on the space C (R 3, C) of smooth functions R 3 C via the formula We write (A f)(x, y, z) = f(a 1 (x, y, z)). ρ : SO 3 (R) GL(C (R 3, C)) for the action homomorphism. Differentiating the action of SO 3 (R) on R 3 gives rise to a Lie algebra homomorphism ϕ : so 3 (R) C (R 3, R 3 ) into smooth vector fields on R 3. Recall that we introduced basis elements r x, r y, r z so 3 (R), and one computes that ϕ(r x ) = z y y z, ϕ(ry ) = x z z x, and ϕ(rz ) = y x x y. Now we observe that so 3 (R)-representation obtained from ρ is the composition ρ : so 3 (R) ϕ C (R 3, R 3 ) gl(c (R 3, C)), where the second Lie algebra homomorphism is the action of vector fields on functions. That is, we can compute the action of so 3 (R) on a function R 3 C by applying the differential operators written above. In order to find an SO 3 (R)-subrepresentation of C (R 3, C) isomorphic to V 2l where l 0, it suffices to do the same for the action of so 3 (R), or equivalently so 3 (R) R C = sl 2 (C), on C (R 3, C). Based on our understanding of sl 2 (C)-representations, we should look for a highest weight vector, i.e. an eigenfunction Y l l : R3 C for h with eigenvalue 2l such that e Y l l = 0. It is convenient to work with the spherical coordinates r, θ, φ on R 3. In fact, the space of functions on R 3 is somewhat too large for our present purposes, because the radial coordinate r is SO 3 (R)-invariant. Therefore we consider only functions f(θ, φ) of the two angular coordinates, which can be viewed as smooth functions f : S 2 C on the 2-sphere The equations h Y l l = 2lY l l The first of these implies that S 2 := {(x, y, z) R 3 x 2 + y 2 + z 2 = 1}. and e Y l l = 0 become the differential equations i φ Y l l (θ, φ) = ly l l (θ, φ) and e iφ ( θ + i cot θ φ Y l l (θ, φ) = e ilφ F l (θ) for some smooth function F l (θ), and then the second gives The solutions of the latter equation have the form θ F l(θ) = l cot(θ)f l (θ). F l (θ) = C ll sin l θ for C ll C, which means our desired highest weight vector is Y l l (θ, φ) = C ll e ilφ sin l θ. ) Y l l (θ, φ) = 0. 4

We remark that the elements r ± := r x ± ir y of so 3 (R) are raising and lowering operators, i.e. up to scaling they correspond to e and f under the isomorphism so 3 (R) R C = sl 2 (C). Define W l C (S 2, C) to be the subspace spanned by the functions ( ( Yl m (θ, φ) := C lm (r ) l m Yl l (θ, φ) = C lm e iφ θ + i cot θ )) l m Yl l (θ, φ) φ for l m l. Note that r z Yl m = myl m. By our classification of representations it follows that W l is SO 3 (R)-stable, and that it is moreover isomorphic to V 2l. The functions Yl m are called spherical harmonics. One can show that the space of functions C (S 2, C) almost decomposes into the direct sum W l. Namely, an arbitrary smooth function f : S 2 C can be written uniquely as an infinite sum f(θ, φ) = C lm Yl m (θ, φ) l 0 l m l for some C ml C. Compare to the Fourier series decomposition of C (S 1, C). 2 Quantization and the hydrogen atom l 0 2.1 Before analyzing the hydrogen atom, we need to discuss the process of passing from classical mechanical systems to quantum systems, called quantization. Consider the classical mechanical situation of a particle of mass m > 0 moving in space R 3, so that the phase space is R 6 with the usual coordinates q 1, q 2 q 3, p 1, p 2, p 3. The Hamiltonian of this system generally takes the form H(q, p) = p 2 2m + V (q), where V : R 3 R is the potential function. How do we produce a quantum mechanical system analogous to this one? The first question is how to construct the Hilbert space space of states. We propose H = L 2 (R 3 ), the space of measurable (on a first pass, ignore this condition) functions ψ : R 3 C such that R 3 ψ(q) 2 dq <. Moreover, if ψ : R 3 C is such that R 3 ψ(q) 2 dq = 0 (equivalently, ψ vanishes outside a set of measure zero) we identify ψ with the zero function. We call the state vectors ψ L 2 (R 3 ) wavefunctions of the particle. For any ψ 1, ψ 2 L 2 (R 3 ), we can define ψ 1, ψ 2 := ψ 1 (q)ψ 2 (q) dq, R 3 which converges by virtue of the L 2 condition. This Hermitian inner product makes L 2 (R 3 ) into a Hilbert space, which we view as the state space of the particle. 5

2.2 The next step is to quantize observables, meaning pass from functions R 6 R to self-adjoint operators on L 2 (R 3 ). The most basic observables for our classical system are the position and momentum coordinates, so we quantize these first. The position operators Q i, corresponding to measurement of the position coordinate q i, are defined by Q i (ψ)(q) := q i ψ(q). Unfortunately, the resulting function R 3 R need not belong to L 2 (R 3 ), but we simply ignore this issue. The operators Q i are evidently self-adjoint to the extent that it makes sense. A related problem is that the eigenstates of Q i do not belong to L 2 (R): they are the so-called delta distributions δ q for q R 3. These objects are characterized by the formula R 3 ψδ q = ψ(q), and clearly there no functions with this property. They form an eigenbasis of L 2 (R) in the sense that any function ψ : R 3 R can be written as an integral ψ = ψδ q, q R 3 so that the coefficient of δ q in this expansion of ψ is ψ(q). The upshot of this discussion is that the probability of observing a particle with wavefunction ψ at the position q is ψ(q) 2. The momentum operators are given by the formula P i := i q i. Like the position operators, they are not well-defined on a general wavefunction, but we will not comment on this. Another similarity with position is that the eigenstates of P i do not belong to L 2 (R): they have the form e ikq for k R. The eigenvalue of Q i acting on e ikq is k, the so-called de Broglie momentum. Exercise 2.2.1. Use integration by parts to check the the momentum operators are self-adjoint, insofar as they are defined. The transition map between the position and momentum eigenbases is the Fourier transform. In particular, if ˆψ : R 3 C is the Fourier transform of ψ, suitably normalized, then the probability of measuring the momentum of a particle with wavefunction ψ to be p is ˆψ(p) 2. Another observable we must quantize is the Hamiltonian. Having done this for position and momentum coordinates, this is fairly straightforward: the Hamiltonian operator on L 2 (R 3 ) is H = 1 2m (P 2 1 + P 2 2 + P 2 3 ) + V (Q 1, Q 2, Q 3 ). Here V (Q 1, Q 2, Q 3 ) is the operator of multiplying by V (q), so there is not much harm in simply writing V (q) as the potential term in the Hamiltonian operator. It is convenient to use the notation := 2 q1 2 + 2 q2 2 + 2 q3 2 for the Laplace operator. The Hamiltonian can then be rewritten as H = 2 2m + V (q). 2.3 Now we introduce SO 3 (R) symmetry. The action of SO 3 (R) on R 3 defines a representation ρ : SO 3 (R) GL(L 2 (R 3 )) via the usual formula (A ψ)(q) = ψ(a 1 q). 6

This is a unitary representation because rotating a function in this manner evidently leaves its integral unchanged. One then computes that the induced Lie algebra representation ρ : so 3 (R) gl(l 2 (R 3 )) sends ( ρ (r x ) = i(q 2 P 3 Q 3 P 2 ) = ( ρ (r y ) = i(q 3 P 1 Q 1 P 3 ) = ( ρ (r z ) = i(q 1 P 2 Q 2 P 1 ) = q 2 q 3 q 1 ) q 3, q 3 q 2 ) q 1, q 1 q 3 ) q 2. q 2 q 1 It therefore seems reasonable to define the angular momentum operators by the formulas L 1 := Q 2 P 3 Q 3 P 2, L 2 := Q 3 P 1 Q 1 P 3, L 3 := Q 1 P 2 Q 2 P 1. These self-adjoint operators correspond to measurement of angular momentum with respect to the x-, y-, or z-axes respectively. Since ρ is a Lie algebra homomorphism, we automatically have the relations [ il 1, il 2 ] = il 3, [ il 2, il 3 ] = il 1, [ il 3, il 1 ] = il 2. (2.3.1) The Casimir operator for so 3 (R) acting on L 2 (R 3 ) is defined by L 2 := L 2 1 + L 2 2 + L 2 3. Exercise 2.3.1. Use the relations (2.3.1) to show that L 2 commutes with ρ (x) for any x so 3 (R). Thus L 2 also commutes with the action of SO 3 (R). One can show that in spherical coordinates, ( 1 L 2 = sin θ θ ( sin θ ) + 1 θ sin 2 θ 2 ) φ 2. Then we rewrite the Laplacian as = 2 r 2 + 2 r r 1 r 2 L2. Since the action of SO 3 (R) leaves the radial coordinate r = q invariant, Exercise 2.3.1 implies that commutes with the SO 3 (R)-action. In particular, if the potential V : R 3 R only depends on r, then H commutes with the SO 3 (R)-action, and consequently angular momentum is conserved. 2.4 Finally, we come to the hydrogen atom. More precisely, we consider the case of an electron moving in a Coulomb potential produced by a proton fixed at the origin. In this case the Hamiltonian is H = 2 2m e2 r, where e is the electron s charge. Since the potential is rotation-invariant, we know by the previous section that the SO 3 (R)-action commutes with H, and therefore that L 1, L 2, and L 3 are conserved quantities. The upshot of this discussion is that any eigenspace of H, consisting of states with a definite energy, will be SO 3 (R)-stable. In particular, we may ask how they decompose into irreducible SO 3 (R)-representations, which we previously classified. When searching for copies of the irreducible representation V 2l = Wl, the following will be helpful. 7

Proposition 2.4.1. If ϕ : so 3 (R) gl(w l ) is the action map, then the operator acts on W l by the scalar l(l + 1). ϕ(r x ) 2 + ϕ(r y ) 2 + ϕ(r z ) 2 Therefore wavefunctions satisfy the differential equation L 2 ψ = l(l + 1)ψ if and only if they generate a copy of W l in L 2 (R 3 ) as an SO 3 (R)-subrepresentation. In particular, to find wavefunctions with energy (i.e. H-eigenvalue) E which belong to W l, we can first look for their radial component by solving the equation Hg l,e (r) = ( ( 2 d 2 2m dr 2 + 2 r Then, for any l m l, we will have and the wavefunctions d l(l + 1) dr r 2 Hg l,e (r)y l m(θ, ψ) = Eg l,e (r)y l m(θ, ψ), ψ(r, θ, ψ) = g l,e (r)y l m(θ, ψ) ) ) e2 g l,e (r) = Eg l,e (r). r will form a basis for a copy of W l in the E-eigenspace of H. We assume that E < 0, which only allows bound states corresponding to electron orbitals, as opposed the scattering states which have E > 0. It turns out that for fixed l 0, there is precisely one solution g l,en (r) for each n > l, where E n := me4 2 2 n 2. In particular, at a fixed energy E n there are n solutions g l,en (r) corresponding to 0 l < n. Thus the E n -eigenspace of H decomposes as W 0 W 1 W n 1. Using this decomposition, we can explain three of the so-called quantum numbers which classify states of an electron in a hydrogen atom. Namely, n is the so-called principal quantum number, which determines the energy. The orbital quantum number l is responsible for the letter in the spectroscopic notation for orbitals: s means l = 0, p means l = 1, etc. The magnetic quantum number is m, the eigenvalue of L 3. Finally, we have until now ignored the spin of the electron. This doubles the number of states and introduces a new spin quantum number s = ± 1 2. 8