R 3 ψ 2 d 3 x = 1. Observables are constructed by tripling everything we did before. The position operator ˆr is*

Size: px
Start display at page:

Download "R 3 ψ 2 d 3 x = 1. Observables are constructed by tripling everything we did before. The position operator ˆr is*"

Transcription

1 Quantum Mechanics in 3-d Relevant sections in text: Chapter 4 Quantum mechanics in three dimensions Our study of a particle moving in one dimension has been instructive, but it is now time to consider more realistic models of nature by working in three dimensions. It is reasonably straightforward to generalize our model of a particle moving in one dimension to a particle moving in three dimensions. In terms of our postulates, the generalization is as follows. The Hilbert space is the space of square-integrable, complex valued functions of three variables: ψ ψ(x, y, z) ψ(r). The scalar product is φ ψ = φ (r)ψ(r) d 3 x φ (x, y, z)ψ(x, y, z) dx dy dz. R 3 The meaning of the wave function ψ(r) is that ψ(r) 2 is the probability density for position measurements to find the particle at the position r = xi + yj + zk. The probability P (R) for finding the particle in some region R is P (R) = ψ(r) 2 d 3 x. R Normalization of the wave function now means R 3 ψ 2 d 3 x = 1. Observables are constructed by tripling everything we did before. The position operator ˆr is* (ˆrψ)(r) = rψ(r), that is, (ˆxψ)(r) = xψ, (ŷψ)(r) = yψ(r), (ẑψ)(r) = zψ(r). Momentum operators are (ˆpψ)(r) = h i ψ, * Note that we are using the hatˆto denote linear operators not unit vectors! 1

2 that is, (ˆp x ψ)(r) = h i ψ x, (ˆp yψ)(r) = h i ψ y, (ˆp zψ)(r) = h i ψ z. Let us note that the different components of the position commute, as do the different components of momentum (exercise). Moreover, each canonical pair of coordinates and momenta have the usual canonical commutation relation: [ˆx, ˆp x ] = i hˆ1, [ŷ, ˆp y ] = i hˆ1, [ẑ, ˆp z ] = i hˆ1. Consequently, we can determine all three components of the position of the particle to arbitrary statistical accuracy and likewise for the momentum. Corresponding components of the position and momentum are incompatible. But non-corresponding components are compatible. So, for example, while there is a non-trivial uncertainty relation for x and p x it is possible to determine both x and p y with arbitrary statistical accuracy. so that The energy operator, i.e., the Hamiltonian, is given by Ĥ = 1 2m ˆp2 + ˆV (ˆr, t), h2 (Ĥψ)(r) = 2m 2 ψ(r) + V (r, t)ψ(r), where V (r, t) V (x, y, z, t) is the potential energy function and 2 is the Laplacian. A significant new feature of particle motion three dimensions is that we can now build an operator representative of the angular momentum. Recall that classically the angular momentum is given by L = r p. The quantum operator representative of this triplet of observables is that is, (ˆLψ)(r) = r ˆL = ˆr ˆp, ) ( h i ψ(r) = h i r ψ. More on angular momentum later. A good exercise for you at this point is to verify that the ordering of the position and momentum operators does not matter in the angular momentum. So, for example, we could just as well define the operator as ˆL = ˆp ˆr. The Schrödinger equation in three dimensions The Schrödinger equation, ĤΨ = i h Ψ t 2

3 in 3-d is a partial differential equation for Ψ = Ψ(r, t): h2 2m 2 Ψ + V (r, t)ψ = i h Ψ t. For time independent potentials, V = V (r), separation of the time variable works as before. With Ψ(r, t) = ψ(r)e ī h Et, we get the time independent Schrödinger equation: Ĥψ = Eψ. As usual, the spectrum of Ĥ may be discrete, continuous, or both. Normalizable solutions of the TISE corresponding to eigenvectors of Ĥ will have discrete eigenvalues E n, and corresponding stationary states Ψ n (r, t). The corresponding stationary states Ψ n evolve in time via the (physically irrelevant) phase factor e ī h Ent. The general solution to the SE is given by a superposition Ψ(r, t) = n c n ψ n (r)e ī h E nt. Solving the TISE for central forces Let us turn to a simple but important class of interactions in which the potential energy function for the particle depends only upon the distance from a given point. Placing our origin at this point we have V = V ( r ) V (r). These potentials correspond to central forces, since the classical force field is given by (exercise) F(r) = V (r) = V (r) r r. Three good examples of central force potentials are the isotropic oscillator V (r) = 1 2 mω2 r 2, the Coulomb or Kepler potential and the Yukawa potential V (r) = k r, V (r) = αe µr. 3

4 The isotropic harmonic oscillator gets its name from the fact that, for any displacement in a radial direction, the restoring force is proportional to the displacement. In particular, isotropic indicates that the restoring force is independent of (radial) direction of displacement. The Coulomb potential is of course quite familiar to you and is the basis for the simplest model of an atom. The Yukawa potential models the strong, short-ranged attractive force between nucleons. Without this force nuclei would not exist. The nice thing about central potentials is that they allow a straightforward solution of the TISE using spherical coordinates (r, θ, φ) and separation of variables. Let us now investigate this. To begin, the Laplacian in spherical polar coordinates is 2 ψ = 1 ( r 2 r 2 ψ ) ( 1 + r r r 2 sin θ ψ ) 1 + sin θ θ θ r 2 sin 2 θ We use this in the TISE h2 2µ 2 ψ(r, θ, φ) + V (r)ψ(r, θ, φ) = Eψ(r, θ, φ). ( 2 ) ψ φ 2. Note that we use µ to denote the mass of the particle; this will avoid a clash of conventions later. It is not obvious, but this partial differential equation can be solved by separation of variables. So, let us try ψ(r, θ, φ) = R(r)Y (θ, φ). We substitute this into the TISE, and (i) divide both sides by ψ, (ii) multiply both sides by 2µr2 h 2 to find (exercise) 1 (r 2 R ) 2µr 2 ( 1 [V E] = sin θ Y ) R h 2 Y sin θ θ θ 1 Y sin 2 θ 2 Y φ 2. We see the familiar separation of variables result: the left-hand side is a function of r, the right hand side is independent of r. Therefore, each side must equal a constant (exercise), which we denote for later convenience by l(l + 1): 1 (r 2 R ) 2µr 2 ( 1 [V E] = l(l + 1) = sin θ Y ) + R h 2 Y sin θ θ θ 1 Y sin 2 θ 2 Y φ 2. The strategy now is to solve these two equations separately. Although I won t quite prove it, the definition of L as self-adjoint operators will require l to be a non-negative integer. For each choice of that integer there is a solution of the radial equation. Note that it is the radial equation that will be used to find the allowed energies (although the angular equation s influence will be felt through the integer values of l). Of course, to solve the radial equation we need to know what is the potential energy function. 4

5 Let us begin with the angular equation. Note that it is universal in the sense that it is the same for all choices of the central force potential energy function, and for all possible energies E. You probably have encountered this equation before. Note that if we consider the TISE with E = V = 0, then we are solving the Laplace equation. In this case the radial equation is different, but the angular equation is unchanged. Thus the angular part of the solution to the TISE for central forces is the same as the angular part of solutions to the Laplace equation for an electrostatic potential in a charge free region. When we solve the angular equation we are solving 1 sin θ ( sin θ Y ) 1 θ θ sin 2 θ 2 Y = l(l + 1)Y. φ2 This is the eigenvalue equation for the operator which is the angular part of Laplacian. We shall see that the angular part of the Laplacian represents the observable 1 h 2 L 2, so h l(l + 1) will be the possible values of the magnitude of angular momentum vector. The angular part of the Laplace eigenvalue problem can also be solved by separation of variables. The result is Y (θ, φ) = AP m l (cos θ)e imφ, where l is a non-negative integer, m is any integer l m l, and P m l is the associated Legendre function Pl m (x) = 1 ( ) 2 l l! (1 x2 ) m /2 d m +l (x 2 1) l. dx The associated Legendre functions look a little messy, but they are not so bad. For example (exercises) P 0 0 = 1, P 0 1 (x) = x = P 1(cos θ) = cos θ, P1 1 (x) = P 1 1 (x) = (1 x2 ) 1/2 = P1 1 and so on. See your text for more examples. (cos θ) = P 1 1 (cos θ) = sin θ, It can be shown that the associated Legendre functions Pl 0 (x) form a complete set of orthogonal polynomials (like the Hermite polynomials). This means in particular that these functions form a basis for the vector space of square-integrable functions on the interval [ 1, 1]. These polynomial functions Pl 0 (x) are usually called the Legendre functions. 5

6 The integer values of l and m are required so that the wave function is suitably nonsingular, single-valued, and define a domain for the angular momentum operators such that they are self-adjoint. The single-valued requirement is ψ(r, θ, 0) = ψ(r, θ, 2π). This single-valuedness requirement might not appear to be necessary a priori (like it would be for electrostatics) since we only require that ψ be square-integrable. However, in order for ψ to be in the domain of self-adjoint operators H and L this restriction is in fact needed. It is also for this reason that l should be an integer. We normalize the solutions of the TISE via 2π π 1 = dr R(r) 2 r 2 dφ dθ Y (θ, φ) 2 sin θ It is conventional to normalize the radial and angular parts separately: If we define Y m l = ɛ 2π π 1 = dr R(r) 2 r 2 dr, 1 = dφ dθ Y (θ, φ) 2 sin θ (2l + 1)(l m )! e imφ P m 4π(l + m )! l (cos θ), ɛ = ( 1) m, m 0, ɛ = 1, m 0. then the angular functions are normalized. In fact, they form the orthonormal set of functions on the unit sphere called spherical harmonics. See table 4.2 in your text for a list of several spherical harmonics. The orthonormality relation for spherical harmonics is 2π π (Yl m 0 0 ) Yl m sin θ dθ dφ = δ ll δ mm. Exercise: Show that (Yl m ) = ( 1) m Yl m. The radial equation We now have a look at radial equation, (r 2 R ) 2µr 2 h 2 (V E)R = l(l + 1)R. 6

7 The detailed nature of the solutions will depend upon the form of the potential energy and on the boundary conditions, but we can make some progress without committing to either of these. Make a change of variables u = rr, and you get (exercise) ( ) h2 2µ u + V + h2 l(l + 1) 2µr 2 u = Eu. This is mathematically the same as the 1-d TISE for a particle moving on the positive real line in a potential V eff = V + h2 l(l + 1) 2µr 2. Indeed, our normalization convention for R becomes u 2 dr = 1. 0 Without loss of generality you can assume that u (and hence R) is a real function (exercise). You can think of u as defining the radial position probability distribution for a particle moving in a central force field. The effective potential V eff governing the radial part of the wave function includes the central force potential and the centrifugal potential h2 l(l+1). You can check that the 2µr 2 centrifugal potential corresponds (classically) to a repulsive force. This kind of effective potential also arises in classical mechanics. It is worth spending a moment to recall this. In classical mechanics, a particle moving in a central force field always moves in a plane orthogonal to the conserved angular momentum vector L. If you introduce polar coordinates (r, φ) in the plane, the equations of motion are µ d2 r dt 2 = d ) (V (r) + L2 dr 2µr 2, and d dφ (µr2 dt dt ) = 0. The latter equation is conservation of the magnitude of angular momentum (exercise), where L = µr 2 dφ dt. The former equation governs the radial motion. You see that the effective potential energy governing the radial motion involves a centrifugal term representing the effect of the angular 7

8 motion on the radial motion. This classical picture suggests that we might identify h 2 l(l+1) with the squared-magnitude of the angular momentum in the quantum description. This is correct, and will be shown later. To proceed further, we need to specify the central force potential energy function. Let us briefly consider a couple of elementary examples, and then have a more detailed look at the Coulomb potential. Example: The free particle in 3-d A free particle certainly has a central force (rather trivially), since V (r) = 0. The radial equation is h2 2µ u + h2 l(l + 1) 2µr 2 u = Eu. This differential equation is a form of the spherical Bessel equation. The solutions, labeled by l, are r times the spherical Bessel functions and spherical Neumann functions of order l. Only the spherical Bessel functions are non-singular as r 0, so we only use them. They are denoted by j l (x). In terms of these special functions, the solution to the radial equation is u = rj l (kr), where can take any non-negative value. k = 1 h 2µE The spherical Bessel functions are oscillatory in functions with decreasing amplitude as the argument (x = kr) increases. They all vanish at r = 0 except the function defined by l = 0 (see below for more on this function). For details on the behavior of the spherical Bessel functions see the text. We will look at the simplest of them in a moment. Keeping in mind that u = rr, we have then the solution to the TISE of the form ψ(r, θ, φ) = Aj l (kr)yl m (θ, φ). Note that these solutions are determined by (i) a choice of E (through k), (ii) a choice of l (non-negative integer) and (iii) a choice of m (an integer such that l m l). Thus there is quite a bit of degeneracy, that is, there are many solutions (idealized stationary states) with the same energy. This corresponds to the fact in classical mechanics that a free particle can move in a variety of ways (same speed, different directions) and still have the same energy (exercise). We saw (in a homework problem) that for normalizable solutions of the TISE in 1-d there can be no degeneracy. Here this result doesn t apply for two reasons. First, the TISE was solved in 3-d. Still, one might argue, the radial equation was equivalent to a TISE in 1-d. We escape that objection by pointing out that (1) the degeneracy does not appear via the radial equation, for each fixed l and E there is only 8

9 one solution to that equation, (2) the solutions in this example cannot, in any case, be normalizable (since, e.g., the spectrum of H is continuous). We will content ourselves with examining the simplest of the solutions to the TISE. Suppose we choose l = 0. Then the radial TISE is (exercise) h2 2µ u = Eu. This is the familiar harmonic oscillator equation, with solutions u = A cos(kr) + B sin(kr). Since R = u r, we will only get a non-singular solution to the TISE if A = 0. The spherical Bessel function of order 0 is in fact j 0 (kr) = sin(kr), kr in agreement with our direct solution when l = 0. When l = 0 we must have m = 0 (exercise), in which case the angular functions are just a constant so the corresponding solution of the TISE with energy E is of the form ψ k,l=0=m = A sin(kr). kr Thus the probability distribution is spherically symmetric, and oscillatory with amplitude decreasing with increasing radius. The frequency of the oscillation increases with increasing energy. You can check that this solution is not normalizable, as expected (exercise). It is interesting to think about what the particle is actually doing in this family of stationary states, especially in comparison to the classical mechanics behavior. What do we know? Well, we have a free particle with some statistically definite energy and vanishing angular momentum (relative to the origin). In classical mechanics with that information, since L = r p, the particle is either at rest (in which case the energy vanishes) or its momentum is parallel to r. Just given the energy and vanishing of angular momentum, any radial motion at constant speed could occur. Of course, we know that in classical mechanics the particle traces out a definite path from one of the possibilities. We could in principle use additional measurements in position and momentum to prepare the particle in a state where it follows a single radial trajectory with the given energy. In the quantum state you can imagine similar possibilities for motion given the energy and angular momentum. But you should keep in mind that the state is already completely determined by the specification of energy and vanishing angular momentum, so the particle cannot really be said to take this or that radial path.* Indeed, you have to keep in mind the uncertainty * Notice that the wave function is spherically symmetric, so perhaps it is not surprising to learn that all directions of motion are equally likely. 9

10 relation between position and momentum which prohibits the classical path of motion to be determined with statistical certainty. Indeed, in this (idealized) stationary state the angular momentum is determined with certainty, which implies maximum uncertainty in angular position. The angular momentum and linear momentum are also incompatible, leading to maximum uncertainty in the direction of motion. So, while the particle is in some sense undergoing radial motion, the position and direction of motion are not determined! One can also prove that in this state (with proper care about its non-normalizability) the radial motion is equally likely to be toward the origin as away from it. One way to check this is to compute the integrand for, say, the expectation value for the x component of momentum. You can check that this integrand is an odd function of x. Consequently any symmetric integral over an arbitrarily large (but finite) region of this integrand will vanish. One can conlude that the expectation value of momentum is zero, even though the magnitude of the momentum need not be zero. This is consistent with equal probability for motion toward or away from the origin. Finally, I remind you that this idealized state is an eigenfunction of energy and hence determines an idealized stationary state. In contrast to the classical motion, no observable quantities for the particle will change in time in this state. The probability current vanishes for this wave function. Particle in a spherical box Let us now consider the stationary states of a particle confined to a region r < a. The potential energy function is that of an infinite spherical well. We can view the free particle of the last section as a limiting case a. We set ψ(r, θ, φ) = 0 when r > a. Inside the well, the radial equation is exactly as before. The only difference now is that we must take account of the boundary condition that R(a) = 0. In general, this leads to a transcendental equation involving the spherical Bessel functions: j l (ka) = 0. The zeros of the spherical Bessel functions are known; they form a discrete set. Thus k is determined by these discrete values. In this way we get a discrete set of allowed energies, as expected. Let us revisit the l = 0 case. We found that R(r) = A sin(kr) kr solves the radial equation (for r < a). To impose the boundary condition is now easy (exercise) k = nπ, n = 1, 2,.... a 10

11 (Exercise: why do we exclude n = 0, 1, 2,...?). The allowed energies are Quantum Mechanics in 3-d E n,l=0 = n2 π 2 h 2 2µa 2, which are identical to the 1-d particle in a box! This is really not so surprising. By ignoring angular motion effects (l = 0), we have essentially reduced the problem to a 1-d problem (exercise). The normalized stationary states that have this energy are (exercise) ψ n,l=0 = 1 1 2πa r sin(nπr a ). Keep in mind that these states and their energies correspond to choosing l = 0. For each l 0 there will also be a sequence of allowed energies; these energies will depend upon the choice of l and another integer labeling the zeros of the spherical Bessel function. This integer is usually called the principal quantum number, n. When l 0, it is also possible to have stationary states with m 0. Because m does not enter the radial equation, it will play no role in determining the energy. This is quite general: associated with each l there will be 2l+1 states, corresponding to m = l, l+1,..., l 1, l (exercise), all having the same energy. One says that a stationary state with a given value of l is (2l + 1)-fold degenerate. This degeneracy reflects the rotational invariance of the system. Hydrogen atom The simplest successful model of a hydrogen atom is the quantum mechanical system associated with the central force potential V (r) = e2 4πɛ 0 1 r. Here e is the charge on the electron, r is the value of the observable interpreted as distance from proton to electron, and ɛ 0 is the permittivity of the vacuum, familiar from eletrostatics. The Coulomb potential being used here is everywhere negative. It diverges as r 0, and it vanishes as r. The behavior as r 0 suggests, at least classically, that the system is not stable since the energy is unbounded from below. The behavior as r means that positive energy, unbound (scattering) states can exist. This is good, since our model should be able to describe ionized states and/or scattering states. We shall restrict our attention to the bound states, so as to understand the famous success of quantum mechanics in predicting the spectrum of hydrogen. The TISE for the hydrogen atom separates into the angular equations, which we solve using spherical harmonics, and the radial equation: [ ] h2 2µ u + e2 1 4πɛ 0 r + h2 l(l + 1) 2µ r 2 u = Eu. 11

12 The effective potential energy for the radial equation includes a centrifugal barrier when l 0. Recall that a similar term arises in classical mechanics and guarantees that bound states are kept away from r = 0. When l = 0 the classical particle may reach r = 0. We shall see how these features fare in the quantum domain. In the radial equation above, the mass µ can be interpreted as the reduced mass µ = m em p m e + m p, built from the electron mass m e and the proton mass m p. Because m p >> m e we have µ m e (exercise). We seek solutions to the radial equation that are well behaved everywhere and that vanish as r, so that the solutions define normalizable functions, i.e., bound states. As usual, we must also determine the allowed values of the energy E; these energy eigenvalues will form a discrete spectrum since we are looking for normalizable solutions. The text provides all the gory details; we shall be content with understanding the results. Energy spectrum To begin, the allowed energies are labeled by a positive definite integer n = 1, 2,.... This integer is called the principal quantum number. In terms of it the energy eigenvalues are { ( µ e 2 ) 2 } 1 E n = 2 h 2 4πɛ 0 n 2. We can write this as (exercise) where E n = e2 8πɛ 0 n 2 a, a = 4 h2 πɛ 0 µe m. The quantity a, which has dimensions of length, is the Bohr radius. It gets its name because, in the Bohr model of the atom, this is the radius a classical electron must maintain to be in the ground state. Indeed, you can see that when n = 1 the atom is in its lowest allowed energy state and the resulting energy is precisely that of a classical particle in a circular orbit at the Bohr radius (nice exercise). More generally, the energies for any n are the same you would get for a classical particle at the radius na. This is why the Bohr model was successful for the hydrogen atom. The Bohr model is now known to be physically incorrect as a model of nature since (a) it fails to describe any other atom, and (b) it does not provide a satisfactory means for describing any phenomenon besides the energy levels of hydrogen. The energy, expressed in terms of the principal quantum 12

13 number and the Bohr radius is, nevertheless, often called the Bohr formula for the hydrogen energies. I emphasize that this formula was obtained by Bohr by using a voodoo blend of classical mechanics and preliminary ideas from the not yet formed quantum theory. It is unfortunate that modern textbooks propagate this intermediate step (Bohr theory) on the road to quantum theory as if it were a viable model of nature*. The ground state, when n = 1, has energy E 1 = 13.6eV. The energy is negative, positive energies correspond to unbound (scattering states). This ground state energy, therefore, is the energy needed to ionize the atom, in good agreement with experiment. For this reason, E 1 is often called the binding energy of the atom in its ground state. Excited states have, of course, greater energy and hence less binding energy. Stationary states The stationary states are, of course, labeled by n. But they also depend upon l and m. One can anticipate a relation between l and n since l appears in the radial equation. What one finds when solving the radial equation and imposing boundary conditions is that, for each value of n, one can have any value of l between 0 and n 1, inclusive: l = 0, 1, 2,... n 1. Keep in mind also that for each l there are the 2l + 1 allowed values of m. Thus, the stationary states are labeled as follows. Pick any n > 0 this determines the energy level ; pick any integer l such that 0 l n 1; pick any integer m such that l m l. As we shall see, the choices of l and m correspond to specifying the angular momentum of the stationary state. With this algorithm in mind, the orthonormal basis of stationary states is, in all of its glory: ( ) ψ nlm (r, θ, φ) = N nlm e r 2r l na L 2l+1 na n l 1 (2r/na)Y l m (θ, φ). Here N nlm is a normalization constant (see the text for a formula) and L s r is the associated Laguerre polynomial, defined by ) L s r(x) = ( 1) (e s ds dx s x dr+s dx r+s (e x x r+s ). Note that this special function appears in ψ nlm with x = 2r na. * In my mind, teaching the Bohr model as fact is analogous to using the Ptolemaic model of the universe (geocentric, epicycles and all that) as a viable model of the universe just because it appeared on the road from the theory in which heavenly bodies were gods to the Copernican theory of the solar system. 13

14 The radial dependence of the stationary states is that of a polynomial in r times a decaying exponential e r na in r. Here are a few of the associated Laguerre polynomials: L 0 0 (x) = 1, L0 1 (x) = 1 x, L2 2 (x) = 12x2 96x + 144; See your text for some more. Such functions polynomials times decaying exponentials are normalizable because of the decaying exponential. Note that the exponential decay of the wave function is characterized by n times the Bohr radius, so we can say that measurements of the size of the atom (via observables that characterize the position probability distribution) will yield values on the order of magnitude of na. One amusing feature of the stationary states stems from the fact that the states with l = 0 have wave functions that do not vanish at the origin (exercise), since R(0) 0. For example, in its ground state we have ψ n=1,l=0,m=0 = 1 πa 3 e r/a. When l 0 it is not hard to see that R(0) = 0 (exercise). Does this means that the atom has (in states with l = 0) a non-zero probability for finding the electron occupying the same space as the proton? Let s see. The probability P (ɛ) for finding a ground state electron in, say, a spherical region of radius ɛ about the origin is 2π π ɛ [ 1 ( P (ɛ) = dφ dθ dr πa 3 e r/a 2 r 2 ɛ ) 2 ( ] ɛ sin(θ) = e a a) 2ɛ/a As ɛ 0 this probability vanishes. On the other hand, there is a real sense in which the nucleus (here, a proton) has a finite size (on the order of Fermis). From the preceding computation we can expect that the probability for finding the electron at the surface of the proton is non-zero, and by continuity we expect that probability to remain nonzero inside the proton. Explicit computations with a finite-sized charge density confirm this. Thus, figuratively speaking, it is possible that the electron is inside the proton! Again we see the tension between the quantum meaning of the word particle and the classical, macroscopic meaning. Because the nucleus has a finite size, the overlap between the electron wave function and the nuclear charge density indicates that we could improve our description of the atom by modifying the potential energy function (1/r outside the proton, r 2 inside exercise). This results in a small change in the l = 0 energy spectrum (thus partially lifting some degeneracy (see below)). For a hydrogen atom this effect is very small, on the order of 10 9 ev. For hydrogenic atoms with larger atomic number, the effect is more pronounced. This effect has been measured experimentally. More on the spectrum of hydrogen There is a good bit of degeneracy in the energies of hydrogen. Recall that an eigenvalue is called p-fold degenerate if there are p linearly independent eigenvectors with that 14

15 eigenvalue. Physically, this means that the outcome of a measurement corresponding to that eigenvalue can occur with probability one via a p-dimensional set of states. For the hydrogen atom, the energy is only determined by n. For a given n, there are n possible l values (exercise). For each l there are 2l+1 possible m values. All together, for each energy one has (exercise) n 2 different (i.e., linearly independent) energy eigenfunctions obtained by varying l and m in their allowed ranges. The freedom to vary m reflects the rotational symmetry of the problem: by rotating the atom we can obtain new states with the same energy. The freedom to vary l is a special feature of the Coulomb potential and reflects a hidden symmetry in the Coulomb (and Kepler) dynamical motion. Classically, this hidden symmetry is responsible for the fact that the motion of a particle in a r 1 central 2 force field has bounded orbits that are closed (in fact, ellipses).* One can probe the energy levels of hydrogen experimentally by adding energy to a hydrogen atom and looking at the electromagnetic radiation that comes out when the atom decays. For example, one can pass current through a container containing hydrogen gas. This has the effect of transferring energy to and from to the atom. The atom absorbs energy from the perturbation and moves to an excited state. The perturbation will then cause the atom to emit energy (via one or more photons) as it makes a transition to a state of lower energy. This process is called stimulated emission. This picture of stimulated emission is understood by looking at the time dependent Schrödinger equation. We will develop techniques for understanding stimulated emission via the SE next semester. (Why the atom decays at all in the absence of a perturbation (spontaneous emission) is another story... wait until next semester.) The following ideas will be given a full quantum mechanical treatment next semester. The bottom line is as follows. The atom will absorb any energy equal to the difference between any two allowed energies. The effect is to raise that atom into a higher energy state. After a certain amount of time, the atom will decay to a lower energy state by emission of a photon. The energy of this photon is the energy difference between the stationary states before and after the emission of the photon. Typically we measure the energy of the photons by using a spectrometer, which discerns the wavelength λ of the photon. The relation between the energy level difference 1 1 E = ( n 2 final n 2 )( 13.6 ev ), initial and the photon wavelength is (c is the speed of light) E = 2πc h λ. * You may have encountered the Laplace-Runge-Lenz vector in the study of the 2-body inverse-square central force problem in classical mechanics. This conserved vector, like all conservation laws, comes from a symmetry and it is this symmetry which is responsible for the degeneracy of energies in hydrogen associated to varying l for a given n. 15

16 Transitions from excited states to the ground state (n final = 1) give off the most energy; the emitted light is ultraviolet. Transitions to the first excited state yield visible light. Transitions down to the second excited state give off infrared radiation, etc. The wavelengths predicted by quantum mechanics are in very good agreement with experiment. There are differences, however, since there are other physical effects not accounted for in this simplest of models of the hydrogen atom. Angular momentum So far we have focused on the stationary states of a particle moving in a central force field. In particular, we have focused on the energies of the hydrogen atom. Another very important observable that one can study for a particle moving in three dimensions is its angular momentum. We shall spend a fair amount of time talking about this observable. Recall that we defined the operator representative of angular momentum via ˆL = ˆr ˆp. What this means is that, on a suitably well-behaved wave function ψ, the linear operation is ˆLψ = h i r ψ. More explicitly, L x = h i (y z z y ), L y = h i (z x x z ), L z = h i (x y y x ), Note that once you know the first of these formulas the other two follow by cyclic permutations (exercise). Probably the most important aspects of angular momentum are (1) it is conserved by central forces (more generally, we believe that the total angular momentum of a closed system is always conserved); (2) the 3 components of L are incompatible, that is, they do not commute. A direct computation (see the text) shows that [L x, L y ] = i hl z, [L y, L z ] = i hl x, [L z, L x ] = i hl y. 16

17 These are the very important angular momentum commutation relations, or angular momentum algebra. Note again that any two formulas follow from a third by cyclic permutations. If you have ever studied the Lie algebra of infinitesimal rotations, you will notice a similarity with the angular momentum algebra. This is no accident, the linear transformation defined by the angular momentum operators can be viewed as the transformation of a state under an infinitesimal rotation. Another way to compute the angular momentum algebra is outlined in one of your homework problems. The idea is to find out how the commutators of components of L with components of r and p look, and then use the fact about commutators (nice exercise) [A, BC] = [A, B]C + B[A, C]. (The order of the operators is crucial for the formula to work.) Because angular momentum is conserved for central potentials, the angular momentum is a useful observable for describing states and their dynamical evolution in this setting. The incompatibility of the components of L means that angular momentum is going to have some surprising properties relative to its classical behavior (recall that position and momentum are also incompatible, hence the uncertainty principle, etc.) In particular, we will find that the possible values of any component of angular momentum do not form a continuum but rather a discrete set. The spectra of the angular momentum operators (L x, L y, L z ) are discrete. One says that angular momentum is quantized. Because of the incompatibility of (L x, L y, L z ), it is generally impossible to have a particle in a state in which two or more components of L are known with certainty, unless the angular momentum is in fact zero. To see this, simply apply our general form of the uncertainty principle to, say, L x and L y to find (exercise) σ Lx σ Ly h 2 L z. As long as the state is such that L z 0, the variances σ Lx and σ Ly cannot vanish, as they should if one knows either of these observables with probability unity. Indeed, suppose that the state ψ is an eigenvector of L x and L y, then it is easy to see that (exercise) [L x, L y ] ψ = 0. But, by the angular momentum commutation relations, this implies L z ψ = 0. So, in fact, the state must be a state of definite value for all three components. By using the other two commutation relations you can show that (exercise) L x ψ = 0 = L y ψ, 17

18 so that the angular momentum is zero with probability unity in such a state. Quantum Mechanics in 3-d Each of the components (L x, L y, L z ) has its basis of eigenvectors, where the value of that observable is known with probability one. But because the three angular momentum operators do not commute, these basis are not the same. If the particle is in a state in which the angular momentum component, say, L x is known to be (non-zero) with certainty, then in general there will be statistical uncertainty in the values of L y and L z. When working in a basis of eigenvectors of angular momentum the component which is determined with certainty is conventionally chosen to be L z. There is nothing special about the z-axis, since we can call any direction in space the z direction. Classically we are used to thinking of this observable as a vector. Quantum mechanically, there are some vectorial aspects to angular momentum (e.g., it has three components), but it is impossible to say with certainty which way the vector points. As it turns out, the length of the angular momentum vector and any one component (e.g., L z ) can be determined with statistical certainty. This is because the operator representing the length-squared, L 2 = L 2 x + L 2 y + L 2 z, commutes with any component of L. For example (exercise), [L z, L 2 ] = [L z, L 2 x + L 2 y + L 2 z] = [L z, L x ]L x + L x [L z, L x ] + [L z, L y ]L y + L y [L z, L y ] = i h ( L y L x + L x L y L x L y L y L x ) = 0. Recall that commuting Hermitian operators always admit a simultaneous basis of eignvectors. Thus it is possible to find a basis of vectors that are eigenvectors of both one component (conventionally taken to be L z ) and L 2. The states defined by these angular momentum eigenvectors are states in which the value of L z and L 2 are known with probability unity. Our task now is to determine these eigenvectors. We will see that they correspond to the spherical harmonics. Angular momentum eigenvalues and eigenfunctions It is possible to systematically deduce the eigenvalues of L 2 and L z just using the angular momentum commutator algebra. This is logically the same game we played when finding the spectrum of the harmonic oscillator Hamiltonian using raising and lowering ( ladder ) operators. See the text for details. The result is that the operator L 2 has eigenvalues h 2 l(l + 1), where l = 0, 1, 2,.... Given the value l, any given component, say L z, has eigenvalues m h, where m = l, l + 1,..., l 1, l. This is true for L x and L y, too. The catch is that the eigenfunctions for each component are not, in general, the 18

19 same. We can pick any one component, find its eigenfunctions, and arrange it so that these eigenfunctions are also the eigenfunction of L 2. It turns out that the simultaneous eigenfunctions of L z and L 2 are given by any function of r times the spherical harmonics Y m l. Let us see a little bit more how this works. If we write the angular momentum operators L and L 2 in terms of spherical polar coordinates we find (via a nice, but long exercise) L x = h i L y = h i ( sin φ ) cos φ cot θ. θ φ ( cos φ ) sin φ cot θ. θ φ L z = h i You can see why one likes to use the component L z ; the spherical polar coordinates are adapted to the z and L z takes a very simple form. You really should try to at least verify the formula for L z. φ. It is easy to see that eigenfunctions of L z are any function of r and θ times e imφ : ψ m (r, θ, φ) = F (r, θ)e imφ, m = 0, ±1, ±2,..., Here m is any integer. The restriction to integer values is so that the eigenfunction is a well-defined function in space. You can see that there is a lot of degeneracy here. Since F is arbitrary, there are many, many functions that have the same eigenvalue m h for L z. We can narrow down the possibilities by demanding that ψ m also be an eigenfunction of L 2. The operator L 2 is a little more complicated to write out in spherical polar coordinates. Up to a factor of r 2, it turns out to be the angular part of the Laplacian: L 2 ψ = h 2 [ 1 sin θ ( sin θ ψ ) + 1 θ θ sin 2 θ ( 2 )] ψ φ 2. This differential operator is the restriction of the Laplace operator to a (unit) sphere (exercise). To find the eigenfunctions is a story we have already touched upon when we solved the TISE for central forces via separation of variables (exercise). Indeed, we recover our results from that analysis. In particular, we have that (i) the eigenvalues of L 2 are l(l + 1) h 2, where l = 0, 1,...; (ii) for a given choice of l, the eigenvalues of L z are restricted by m = l, l + 1,..., l 1, l; (iii) the eigenfunctions ψ l,m (r, θ, φ) for a given choice of l and m are built from the spherical harmonics Y m l : ψ l,m (r, θ, φ) = f(r)yl m (θ, φ). 19

20 We see that there is still some degeneracy owing to the appearance of the arbitrary function f(r). Thus, many different states correspond to the same values of L 2 and L z. These states all differ by the choice of f(r). To uniquely specify the states one must find another observable which commutes with L z and L 2 and demand that ψ l,m also be an eigenfunction of that operator with a specific eigenvalue. For example, in the central force problems it turns out (see homework) that [H, L] = [ p2 + V (r), L] = 0. 2m This fact is equivalent to the statement that the Hamiltonian operator is rotationally invariant. As you will see in a homework problem, the fact that H commutes with L means that angular momentum is conserved (probabilities for outcomes of momentum measurements will not change in time.) This is important in the present discussion because it means that it is possible to further specialize the eigenfunctions of angular momentum to also be energy eigenfunctions. How is this done? Well, you have already done it! The stationary states in a central force problem are radial functions times the spherical harmonics, as above. The radial functions are determined by the radial equation, which also determines the allowed energies. Thus f(r) can be determined by a choice of energy. Indeed, the stationary states of the hydrogen atom (more generally, any central force system) are simultaneously (i) energy eigenfunctions, (ii) eigenfunctions of a component (L z ) of angular momentum and (iii) eigenfunctions of the total angular momentum L 2. We now see the physical significance of l and m. Infinitesimal translations and rotations There is a nice (if a little sophisticated) way to think of momentum-like quantities in classical and quantum mechanics, namely, as generators of transformations. In fact, linear momentum generates translations while angular momentum generates rotations. Let us briefly explore this. We begin with translations. Let ψ(r) describe the state of a system. How does the value of ψ change as we translate from r to the point r + a? Simple: ψ(r) ψ(r + a). Now, we can built any translation by a sequence of many small translations. Suppose that the magnitude a is very small (infinitesimal), then a standard result from calculus is that ψ(r + a) ψ(r) + a ψ(r) = ψ(r) + ī h a (ˆpψ)(r). We say that the momentum is the infinitesimal generator of translations, since a finite translation can be built by many infinitesimal translations associated to p as above. In 20

21 general, the infinitesimal change in a wave function associated with a translation along the infinitesimal vector a is given by ī h a pψ(r). A similar result is available for angular momentum and infinitesimal rotations. Every rotation is a rotation by some angle about some axis. For a given rotation, we can choose our z axis to be along the axis of rotation. In these coordinates a rotation by an angle φ about z is (exercise) x x = x cos φ y sin φ, y y = y cos φ + x sin φ, z z. Again, we can think of any rotation as the net result of a sequence of many small rotations. Suppose φ << 1, then to first order in φ (exercise) x x φy, y y + φx. Consider a function ψ(r), how does it change under an infinitesimal rotation? We have (exercise) ψ(x, y, z ψ(x, y, z) ψ(x, y, z) ) ψ(x, y, z) φy + φx. x y We can write this as Better yet, we can write ψ(x, y, z ) ψ(x, y, z) + ī h φ(ˆl z ψ)(x, y, z). ψ(x, y, z ) ψ(x, y, z) + ī h φn (ˆLψ)(x, y, z), where n is a unit vector along the rotation axis. We say that angular momentum is the infinitesimal generator of rotations since n L gives the infinitesimal change in the wave function due to a rotation about the axis along n. A finite rotation can be built by many infinitesimal rotations, generated by L. Physically, you can think of the translations and rotations as transformations of the physical system (here a particle). Thus the operators that represent momentum and angular momentum give the infinitesimal change in the state of the system as one performs the indicated transformation (translation and/or rotation) of the system. Exercise: What operator is the infinitesimal generator of time translation of a given system? 21

22 Spin I am assuming you have had an introduction to the phenomenology of intrinsic spin in an earlier class in modern physics. Here I just want to introduce you to the quantum mechanical basic formalism and a couple of nice results. Spin is an intrinsic property of an elementary particle, like its mass and charge. Spin is a form of angular momentum conservation of angular momentum requires the spin to be taken into account. Charged particles with spin, e.g., the electron, have an intrinsic magnetic moment which is proportional to the spin vector, but neutral particles can have spin, e.g., the neutrino. Unlike mass and charge, which only take a single value for a given particle, the spin can be in various states. More precisely, the magnitude of the spin vector is fixed for a given type of particle unlike the orbital angular momentum whose magnitude can vary while any given component of that spin vector can take various discrete values. Here I will focus on spin-1/2, which pertains to the most significant matter in the universe, e.g., particles such as electrons, protons, neutrons, positrons, neutrinos, quarks, etc. The observable of interest is the spin vector, which represents angular momentum carried intrinsically by the particle. Like orbital angular momentum it can be represented by 3 Hermitian operators, S = (S x, S y, S z ) satisfying the angular momentum commutation relations: [S x, S y ] = i hs z, [S y, S z ] = i hs x, [S z, S x ] = i hs y. As mentioned when we studied the orbital angular momentum, these relationships can be used to deduce the possible eigenvalues and eigenvectors for any given component. It is not too hard to do, but we don t have time for it. Let me just state the main results. First of all, while each component will have a basis of eigenvectors, the bases will be different for different components. The magnitude squared of the spin, is compatible with any component: S 2 = S 2 x + S 2 y + S 2 z [S 2, S x ] = [S 2, S y ] = [S 2, S z ] = 0. Consequently, we can find a basis of eigenvectors of S 2 and any component conventionally we work with eigenvectors of S z. It can be shown from the commutator algebra that the eigenvalues of S 2 are given by s(s + 1) h 2, s = 0, 1 2, 1, 3 2, 2,... where it is understood that for a given type of particle the s value is fixed. Conventionally, it is the s-value which is used to characterize the spin. The spin 1/2 particles are those 22

23 with s = 1/2. It is pedagogically unfortunate that the particles with spin s actually have their total angular momentum taking the value s(s + 1) h. The eigenvalues of any one component, say S z, are given by m s h, m s = s, s + 1,... s 1, s. This is all very much like orbital angular momentum, of course. But there are two key differences to keep in mind. First, the quantum number s controlling the total spin value is not subject to change with the change of state of the particle; it is fixed by the type of particle (as is the mass and electric charge). The electron may have different l values in an atom, but its s value is always 1/2. This means that every state has the value of S 2 determined with certainty, and the operator representing S 2 must take the form Ŝ 2 = s(s + 1) h 2ˆ1, where s is fixed and 1 is the identity operator. Second, the orbital angular momentum operators require that only integer values of l (and m l ) can occur. This stems from the fact that L = r p with the momentum acting via the gradient on wave functions. The integer values of l are needed so the wave functions are sufficiently well-defined for L to act on them. The spin operators are not constructed as r p and do not have that restriction; half integral values of s (and m s ) may occur for elementary particles. Halfintegral spin particles of a given type (e.g., electrons) obey Fermi-Dirac statistics and are called fermions. Integral spin particles of a given type (e.g., photons) obey Bose-Einstein statistics and are called bosons. Even though the spin (s) is fixed for a given type of particle, if s 0 the value of m s can vary so that the particle can have more than one spin state. Of course the particle can also have a variety of states of motion, but we will ignore that for now and just consider the spin states as if that is all the particle can ever do. This is for pedagogical reasons, but there are physical situations where the spin state is pretty much all one is interested in. For example it is the spin state of the electron and proton in interstellar hydrogen which explains the famous 21 centimeter hydrogen emission line, and it is the spin state of electrons in a lattice which is used to explain ferromagnetic phase transitions in crystalline solids. Spin 1/2 Let s now specialize to the consideration of spin 1/2, i.e., particles with s = 1/2. This means we are studying the spin properties of particles like the electron. The magnitude of the spin for an electron is therefore 3 h/2. The m s values are ± 1 2, so that the component of the spin vector along any direction takes only two values, ± h/2. Therefore the basis of 23

24 eigenvectors of any of the spin components will have only 2 elements and the space of spin states is two-dimensional.* Because the space of states of a spin 1/2 particle is 2-dimensional, we can fruitfully use a matrix representation of vectors and operators. The vectors representing states are then column vectors with two components: ( ) ψ1 ψ =. ψ 2 The scalar product between two vectors is the usual one from matrix algebra, but keep in mind there is a complex conjugation: ( ) φ ψ = (φ 1 φ 2 ) ψ1 = φ ψ 1 ψ 1 + φ 2 ψ 2. 2 The normalization condition for ψ is then S x = h 2 1 = ψ ψ = ψ ψ 2 2. The spin operators can be constructed as h/2 times the Pauli spin matrices: ( ( ( ) i i ), S y = h 2 ), S z = h 2 It is straightforward, if a bit tedious (unless you use a computer) to check that these matrices do satisfy the angular momentum commutation relations. It is also straightfoward to check that S 2 = Sx 2 + Sy 2 + Sz 2 = 3 ( ) h2, 0 1 as it should. Notice that the basis is a basis of eigenvectors of S z : S z, + ( ) ( ) 1 0, S 0 z, 1 S z S z, ± = ± h 2 S z, ±. It is straightforward to show that the other spin components have the following column vectors representing their eigenvectors: S x, ± = 1 ( ) 1, S 2 ±1 y, ± = 1 ( ) 1. 2 ±i * Thus we have an example of a 2 state system. This does not mean there are only two states there are infinitely many vectors in a 2-d vector space. What it does mean is that any observable can have (at most) 2 eigenvalues and 2 eigenvectors, so there are 2 possible outcomes of a measurement and two corresponding states. 24

The 3 dimensional Schrödinger Equation

The 3 dimensional Schrödinger Equation Chapter 6 The 3 dimensional Schrödinger Equation 6.1 Angular Momentum To study how angular momentum is represented in quantum mechanics we start by reviewing the classical vector of orbital angular momentum

More information

Lecture 4 Quantum mechanics in more than one-dimension

Lecture 4 Quantum mechanics in more than one-dimension Lecture 4 Quantum mechanics in more than one-dimension Background Previously, we have addressed quantum mechanics of 1d systems and explored bound and unbound (scattering) states. Although general concepts

More information

Lecture 4 Quantum mechanics in more than one-dimension

Lecture 4 Quantum mechanics in more than one-dimension Lecture 4 Quantum mechanics in more than one-dimension Background Previously, we have addressed quantum mechanics of 1d systems and explored bound and unbound (scattering) states. Although general concepts

More information

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

Lecture 11 Spin, orbital, and total angular momentum Mechanics. 1 Very brief background. 2 General properties of angular momentum operators Lecture Spin, orbital, and total angular momentum 70.00 Mechanics Very brief background MATH-GA In 9, a famous experiment conducted by Otto Stern and Walther Gerlach, involving particles subject to a nonuniform

More information

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

20 The Hydrogen Atom. Ze2 r R (20.1) H( r, R) = h2 2m 2 r h2 2M 2 R 20 The Hydrogen Atom 1. We want to solve the time independent Schrödinger Equation for the hydrogen atom. 2. There are two particles in the system, an electron and a nucleus, and so we can write the Hamiltonian

More information

Physics 342 Lecture 23. Radial Separation. Lecture 23. Physics 342 Quantum Mechanics I

Physics 342 Lecture 23. Radial Separation. Lecture 23. Physics 342 Quantum Mechanics I Physics 342 Lecture 23 Radial Separation Lecture 23 Physics 342 Quantum Mechanics I Friday, March 26th, 2010 We begin our spherical solutions with the simplest possible case zero potential. Aside from

More information

1.6. Quantum mechanical description of the hydrogen atom

1.6. Quantum mechanical description of the hydrogen atom 29.6. Quantum mechanical description of the hydrogen atom.6.. Hamiltonian for the hydrogen atom Atomic units To avoid dealing with very small numbers, let us introduce the so called atomic units : Quantity

More information

1 Commutators (10 pts)

1 Commutators (10 pts) Final Exam Solutions 37A Fall 0 I. Siddiqi / E. Dodds Commutators 0 pts) ) Consider the operator  = Ĵx Ĵ y + ĴyĴx where J i represents the total angular momentum in the ith direction. a) Express both

More information

Schrödinger equation for central potentials

Schrödinger equation for central potentials Chapter 2 Schrödinger equation for central potentials In this chapter we will extend the concepts and methods introduced in the previous chapter ifor a one-dimenional problem to a specific and very important

More information

Phys 622 Problems Chapter 5

Phys 622 Problems Chapter 5 1 Phys 622 Problems Chapter 5 Problem 1 The correct basis set of perturbation theory Consider the relativistic correction to the electron-nucleus interaction H LS = α L S, also known as the spin-orbit

More information

Total Angular Momentum for Hydrogen

Total Angular Momentum for Hydrogen Physics 4 Lecture 7 Total Angular Momentum for Hydrogen Lecture 7 Physics 4 Quantum Mechanics I Friday, April th, 008 We have the Hydrogen Hamiltonian for central potential φ(r), we can write: H r = p

More information

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

Page 404. Lecture 22: Simple Harmonic Oscillator: Energy Basis Date Given: 2008/11/19 Date Revised: 2008/11/19 Page 404 Lecture : Simple Harmonic Oscillator: Energy Basis Date Given: 008/11/19 Date Revised: 008/11/19 Coordinate Basis Section 6. The One-Dimensional Simple Harmonic Oscillator: Coordinate Basis Page

More information

PHY4604 Introduction to Quantum Mechanics Fall 2004 Final Exam SOLUTIONS December 17, 2004, 7:30 a.m.- 9:30 a.m.

PHY4604 Introduction to Quantum Mechanics Fall 2004 Final Exam SOLUTIONS December 17, 2004, 7:30 a.m.- 9:30 a.m. PHY464 Introduction to Quantum Mechanics Fall 4 Final Eam SOLUTIONS December 7, 4, 7:3 a.m.- 9:3 a.m. No other materials allowed. If you can t do one part of a problem, solve subsequent parts in terms

More information

Quantum Physics II (8.05) Fall 2002 Assignment 11

Quantum Physics II (8.05) Fall 2002 Assignment 11 Quantum Physics II (8.05) Fall 00 Assignment 11 Readings Most of the reading needed for this problem set was already given on Problem Set 9. The new readings are: Phase shifts are discussed in Cohen-Tannoudji

More information

1 Mathematical preliminaries

1 Mathematical preliminaries 1 Mathematical preliminaries The mathematical language of quantum mechanics is that of vector spaces and linear algebra. In this preliminary section, we will collect the various definitions and mathematical

More information

The Hydrogen Atom. Dr. Sabry El-Taher 1. e 4. U U r

The Hydrogen Atom. Dr. Sabry El-Taher 1. e 4. U U r The Hydrogen Atom Atom is a 3D object, and the electron motion is three-dimensional. We ll start with the simplest case - The hydrogen atom. An electron and a proton (nucleus) are bound by the central-symmetric

More information

The Hydrogen atom. Chapter The Schrödinger Equation. 2.2 Angular momentum

The Hydrogen atom. Chapter The Schrödinger Equation. 2.2 Angular momentum Chapter 2 The Hydrogen atom In the previous chapter we gave a quick overview of the Bohr model, which is only really valid in the semiclassical limit. cf. section 1.7.) We now begin our task in earnest

More information

E = φ 1 A The dynamics of a particle with mass m and charge q is determined by the Hamiltonian

E = φ 1 A The dynamics of a particle with mass m and charge q is determined by the Hamiltonian Lecture 9 Relevant sections in text: 2.6 Charged particle in an electromagnetic field We now turn to another extremely important example of quantum dynamics. Let us describe a non-relativistic particle

More information

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

Physics 221A Fall 1996 Notes 12 Orbital Angular Momentum and Spherical Harmonics Physics 221A Fall 1996 Notes 12 Orbital Angular Momentum and Spherical Harmonics We now consider the spatial degrees of freedom of a particle moving in 3-dimensional space, which of course is an important

More information

2m r2 (~r )+V (~r ) (~r )=E (~r )

2m r2 (~r )+V (~r ) (~r )=E (~r ) Review of the Hydrogen Atom The Schrodinger equation (for 1D, 2D, or 3D) can be expressed as: ~ 2 2m r2 (~r, t )+V (~r ) (~r, t )=i~ @ @t The Laplacian is the divergence of the gradient: r 2 =r r The time-independent

More information

1 r 2 sin 2 θ. This must be the case as we can see by the following argument + L2

1 r 2 sin 2 θ. This must be the case as we can see by the following argument + L2 PHYS 4 3. The momentum operator in three dimensions is p = i Therefore the momentum-squared operator is [ p 2 = 2 2 = 2 r 2 ) + r 2 r r r 2 sin θ We notice that this can be written as sin θ ) + θ θ r 2

More information

Atomic Structure and Processes

Atomic Structure and Processes Chapter 5 Atomic Structure and Processes 5.1 Elementary atomic structure Bohr Orbits correspond to principal quantum number n. Hydrogen atom energy levels where the Rydberg energy is R y = m e ( e E n

More information

Angular momentum. Quantum mechanics. Orbital angular momentum

Angular momentum. Quantum mechanics. Orbital angular momentum Angular momentum 1 Orbital angular momentum Consider a particle described by the Cartesian coordinates (x, y, z r and their conjugate momenta (p x, p y, p z p. The classical definition of the orbital angular

More information

(relativistic effects kinetic energy & spin-orbit coupling) 3. Hyperfine structure: ) (spin-spin coupling of e & p + magnetic moments) 4.

(relativistic effects kinetic energy & spin-orbit coupling) 3. Hyperfine structure: ) (spin-spin coupling of e & p + magnetic moments) 4. 4 Time-ind. Perturbation Theory II We said we solved the Hydrogen atom exactly, but we lied. There are a number of physical effects our solution of the Hamiltonian H = p /m e /r left out. We already said

More information

Brief review of Quantum Mechanics (QM)

Brief review of Quantum Mechanics (QM) Brief review of Quantum Mechanics (QM) Note: This is a collection of several formulae and facts that we will use throughout the course. It is by no means a complete discussion of QM, nor will I attempt

More information

Physics 115C Homework 2

Physics 115C Homework 2 Physics 5C Homework Problem Our full Hamiltonian is H = p m + mω x +βx 4 = H +H where the unperturbed Hamiltonian is our usual and the perturbation is H = p m + mω x H = βx 4 Assuming β is small, the perturbation

More information

Quantum Mechanics Solutions

Quantum Mechanics Solutions Quantum Mechanics Solutions (a (i f A and B are Hermitian, since (AB = B A = BA, operator AB is Hermitian if and only if A and B commute So, we know that [A,B] = 0, which means that the Hilbert space H

More information

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

Outline Spherical symmetry Free particle Coulomb problem Keywords and References. Central potentials. Sourendu Gupta. TIFR, Mumbai, India Central potentials Sourendu Gupta TIFR, Mumbai, India Quantum Mechanics 1 2013 3 October, 2013 Outline 1 Outline 2 Rotationally invariant potentials 3 The free particle 4 The Coulomb problem 5 Keywords

More information

Mathematical Tripos Part IB Michaelmas Term Example Sheet 1. Values of some physical constants are given on the supplementary sheet

Mathematical Tripos Part IB Michaelmas Term Example Sheet 1. Values of some physical constants are given on the supplementary sheet Mathematical Tripos Part IB Michaelmas Term 2015 Quantum Mechanics Dr. J.M. Evans Example Sheet 1 Values of some physical constants are given on the supplementary sheet 1. Whenasampleofpotassiumisilluminatedwithlightofwavelength3

More information

1 Multiplicity of the ideal gas

1 Multiplicity of the ideal gas Reading assignment. Schroeder, section.6. 1 Multiplicity of the ideal gas Our evaluation of the numbers of microstates corresponding to each macrostate of the two-state paramagnet and the Einstein model

More information

Chem 467 Supplement to Lecture 19 Hydrogen Atom, Atomic Orbitals

Chem 467 Supplement to Lecture 19 Hydrogen Atom, Atomic Orbitals Chem 467 Supplement to Lecture 19 Hydrogen Atom, Atomic Orbitals Pre-Quantum Atomic Structure The existence of atoms and molecules had long been theorized, but never rigorously proven until the late 19

More information

The Postulates of Quantum Mechanics Common operators in QM: Potential Energy. Often depends on position operator: Kinetic Energy 1-D case: 3-D case

The Postulates of Quantum Mechanics Common operators in QM: Potential Energy. Often depends on position operator: Kinetic Energy 1-D case: 3-D case The Postulates of Quantum Mechanics Common operators in QM: Potential Energy Often depends on position operator: Kinetic Energy 1-D case: 3-D case Time Total energy = Hamiltonian To find out about the

More information

Quantum Theory of Angular Momentum and Atomic Structure

Quantum Theory of Angular Momentum and Atomic Structure Quantum Theory of Angular Momentum and Atomic Structure VBS/MRC Angular Momentum 0 Motivation...the questions Whence the periodic table? Concepts in Materials Science I VBS/MRC Angular Momentum 1 Motivation...the

More information

Introduction to Quantum Mechanics PVK - Solutions. Nicolas Lanzetti

Introduction to Quantum Mechanics PVK - Solutions. Nicolas Lanzetti Introduction to Quantum Mechanics PVK - Solutions Nicolas Lanzetti lnicolas@student.ethz.ch 1 Contents 1 The Wave Function and the Schrödinger Equation 3 1.1 Quick Checks......................................

More information

The Schrödinger Equation

The Schrödinger Equation Chapter 13 The Schrödinger Equation 13.1 Where we are so far We have focused primarily on electron spin so far because it s a simple quantum system (there are only two basis states!), and yet it still

More information

Approximation Methods in QM

Approximation Methods in QM Chapter 3 Approximation Methods in QM Contents 3.1 Time independent PT (nondegenerate)............... 5 3. Degenerate perturbation theory (PT)................. 59 3.3 Time dependent PT and Fermi s golden

More information

Quantum Mechanics: The Hydrogen Atom

Quantum Mechanics: The Hydrogen Atom Quantum Mechanics: The Hydrogen Atom 4th April 9 I. The Hydrogen Atom In this next section, we will tie together the elements of the last several sections to arrive at a complete description of the hydrogen

More information

Schrödinger equation for central potentials

Schrödinger equation for central potentials Chapter 2 Schrödinger equation for central potentials In this chapter we will extend the concepts and methods introduced in the previous chapter for a one-dimensional problem to a specific and very important

More information

Physics 221A Fall 1996 Notes 19 The Stark Effect in Hydrogen and Alkali Atoms

Physics 221A Fall 1996 Notes 19 The Stark Effect in Hydrogen and Alkali Atoms Physics 221A Fall 1996 Notes 19 The Stark Effect in Hydrogen and Alkali Atoms In these notes we will consider the Stark effect in hydrogen and alkali atoms as a physically interesting example of bound

More information

Basic Physical Chemistry Lecture 2. Keisuke Goda Summer Semester 2015

Basic Physical Chemistry Lecture 2. Keisuke Goda Summer Semester 2015 Basic Physical Chemistry Lecture 2 Keisuke Goda Summer Semester 2015 Lecture schedule Since we only have three lectures, let s focus on a few important topics of quantum chemistry and structural chemistry

More information

H atom solution. 1 Introduction 2. 2 Coordinate system 2. 3 Variable separation 4

H atom solution. 1 Introduction 2. 2 Coordinate system 2. 3 Variable separation 4 H atom solution Contents 1 Introduction 2 2 Coordinate system 2 3 Variable separation 4 4 Wavefunction solutions 6 4.1 Solution for Φ........................... 6 4.2 Solution for Θ...........................

More information

Rotations in Quantum Mechanics

Rotations in Quantum Mechanics Rotations in Quantum Mechanics We have seen that physical transformations are represented in quantum mechanics by unitary operators acting on the Hilbert space. In this section, we ll think about the specific

More information

Lectures 21 and 22: Hydrogen Atom. 1 The Hydrogen Atom 1. 2 Hydrogen atom spectrum 4

Lectures 21 and 22: Hydrogen Atom. 1 The Hydrogen Atom 1. 2 Hydrogen atom spectrum 4 Lectures and : Hydrogen Atom B. Zwiebach May 4, 06 Contents The Hydrogen Atom Hydrogen atom spectrum 4 The Hydrogen Atom Our goal here is to show that the two-body quantum mechanical problem of the hydrogen

More information

CHAPTER 8 The Quantum Theory of Motion

CHAPTER 8 The Quantum Theory of Motion I. Translational motion. CHAPTER 8 The Quantum Theory of Motion A. Single particle in free space, 1-D. 1. Schrodinger eqn H ψ = Eψ! 2 2m d 2 dx 2 ψ = Eψ ; no boundary conditions 2. General solution: ψ

More information

1 Reduced Mass Coordinates

1 Reduced Mass Coordinates Coulomb Potential Radial Wavefunctions R. M. Suter April 4, 205 Reduced Mass Coordinates In classical mechanics (and quantum) problems involving several particles, it is convenient to separate the motion

More information

Physics 228 Today: Ch 41: 1-3: 3D quantum mechanics, hydrogen atom

Physics 228 Today: Ch 41: 1-3: 3D quantum mechanics, hydrogen atom Physics 228 Today: Ch 41: 1-3: 3D quantum mechanics, hydrogen atom Website: Sakai 01:750:228 or www.physics.rutgers.edu/ugrad/228 Happy April Fools Day Example / Worked Problems What is the ratio of the

More information

PHYS 3313 Section 001 Lecture # 22

PHYS 3313 Section 001 Lecture # 22 PHYS 3313 Section 001 Lecture # 22 Dr. Barry Spurlock Simple Harmonic Oscillator Barriers and Tunneling Alpha Particle Decay Schrodinger Equation on Hydrogen Atom Solutions for Schrodinger Equation for

More information

Quantum Mechanics in Three Dimensions

Quantum Mechanics in Three Dimensions Physics 342 Lecture 21 Quantum Mechanics in Three Dimensions Lecture 21 Physics 342 Quantum Mechanics I Monday, March 22nd, 21 We are used to the temporal separation that gives, for example, the timeindependent

More information

Introduction to Electronic Structure Theory

Introduction to Electronic Structure Theory Introduction to Electronic Structure Theory C. David Sherrill School of Chemistry and Biochemistry Georgia Institute of Technology June 2002 Last Revised: June 2003 1 Introduction The purpose of these

More information

C/CS/Phys C191 Particle-in-a-box, Spin 10/02/08 Fall 2008 Lecture 11

C/CS/Phys C191 Particle-in-a-box, Spin 10/02/08 Fall 2008 Lecture 11 C/CS/Phys C191 Particle-in-a-box, Spin 10/0/08 Fall 008 Lecture 11 Last time we saw that the time dependent Schr. eqn. can be decomposed into two equations, one in time (t) and one in space (x): space

More information

14. Structure of Nuclei

14. Structure of Nuclei 14. Structure of Nuclei Particle and Nuclear Physics Dr. Tina Potter Dr. Tina Potter 14. Structure of Nuclei 1 In this section... Magic Numbers The Nuclear Shell Model Excited States Dr. Tina Potter 14.

More information

Quantum Physics II (8.05) Fall 2002 Outline

Quantum Physics II (8.05) Fall 2002 Outline Quantum Physics II (8.05) Fall 2002 Outline 1. General structure of quantum mechanics. 8.04 was based primarily on wave mechanics. We review that foundation with the intent to build a more formal basis

More information

Quantum Physics 2006/07

Quantum Physics 2006/07 Quantum Physics 6/7 Lecture 7: More on the Dirac Equation In the last lecture we showed that the Dirac equation for a free particle i h t ψr, t = i hc α + β mc ψr, t has plane wave solutions ψr, t = exp

More information

Physics 4617/5617: Quantum Physics Course Lecture Notes

Physics 4617/5617: Quantum Physics Course Lecture Notes Physics 467/567: Quantum Physics Course Lecture Notes Dr. Donald G. Luttermoser East Tennessee State University Edition 5. Abstract These class notes are designed for use of the instructor and students

More information

Time part of the equation can be separated by substituting independent equation

Time part of the equation can be separated by substituting independent equation Lecture 9 Schrödinger Equation in 3D and Angular Momentum Operator In this section we will construct 3D Schrödinger equation and we give some simple examples. In this course we will consider problems where

More information

Quantum Mechanics-I Prof. Dr. S. Lakshmi Bala Department of Physics Indian Institute of Technology, Madras. Lecture - 21 Square-Integrable Functions

Quantum Mechanics-I Prof. Dr. S. Lakshmi Bala Department of Physics Indian Institute of Technology, Madras. Lecture - 21 Square-Integrable Functions Quantum Mechanics-I Prof. Dr. S. Lakshmi Bala Department of Physics Indian Institute of Technology, Madras Lecture - 21 Square-Integrable Functions (Refer Slide Time: 00:06) (Refer Slide Time: 00:14) We

More information

The two body problem involves a pair of particles with masses m 1 and m 2 described by a Lagrangian of the form:

The two body problem involves a pair of particles with masses m 1 and m 2 described by a Lagrangian of the form: Physics 3550, Fall 2011 Two Body, Central-Force Problem Relevant Sections in Text: 8.1 8.7 Two Body, Central-Force Problem Introduction. I have already mentioned the two body central force problem several

More information

Physics 828 Problem Set 7 Due Wednesday 02/24/2010

Physics 828 Problem Set 7 Due Wednesday 02/24/2010 Physics 88 Problem Set 7 Due Wednesday /4/ 7)a)Consider the proton to be a uniformly charged sphere of radius f m Determine the correction to the s ground state energy 4 points) This is a standard problem

More information

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 20, March 8, 2006

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 20, March 8, 2006 Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer Lecture 20, March 8, 2006 Solved Homework We determined that the two coefficients in our two-gaussian

More information

Attempts at relativistic QM

Attempts at relativistic QM Attempts at relativistic QM based on S-1 A proper description of particle physics should incorporate both quantum mechanics and special relativity. However historically combining quantum mechanics and

More information

Lecture #13 1. Incorporating a vector potential into the Hamiltonian 2. Spin postulates 3. Description of spin states 4. Identical particles in

Lecture #13 1. Incorporating a vector potential into the Hamiltonian 2. Spin postulates 3. Description of spin states 4. Identical particles in Lecture #3. Incorporating a vector potential into the Hamiltonian. Spin postulates 3. Description of spin states 4. Identical particles in classical and QM 5. Exchange degeneracy - the fundamental problem

More information

G : Quantum Mechanics II

G : Quantum Mechanics II G5.666: Quantum Mechanics II Notes for Lecture 5 I. REPRESENTING STATES IN THE FULL HILBERT SPACE Given a representation of the states that span the spin Hilbert space, we now need to consider the problem

More information

Chemistry 532 Practice Final Exam Fall 2012 Solutions

Chemistry 532 Practice Final Exam Fall 2012 Solutions Chemistry 53 Practice Final Exam Fall Solutions x e ax dx π a 3/ ; π sin 3 xdx 4 3 π cos nx dx π; sin θ cos θ + K x n e ax dx n! a n+ ; r r r r ˆL h r ˆL z h i φ ˆL x i hsin φ + cot θ cos φ θ φ ) ˆLy i

More information

P3317 HW from Lecture and Recitation 7

P3317 HW from Lecture and Recitation 7 P3317 HW from Lecture 1+13 and Recitation 7 Due Oct 16, 018 Problem 1. Separation of variables Suppose we have two masses that can move in 1D. They are attached by a spring, yielding a Hamiltonian where

More information

Physics 342 Lecture 22. The Hydrogen Atom. Lecture 22. Physics 342 Quantum Mechanics I

Physics 342 Lecture 22. The Hydrogen Atom. Lecture 22. Physics 342 Quantum Mechanics I Physics 342 Lecture 22 The Hydrogen Atom Lecture 22 Physics 342 Quantum Mechanics I Friday, March 28th, 2008 We now begin our discussion of the Hydrogen atom. Operationally, this is just another choice

More information

The Wave Function. Chapter The Harmonic Wave Function

The Wave Function. Chapter The Harmonic Wave Function Chapter 3 The Wave Function On the basis of the assumption that the de Broglie relations give the frequency and wavelength of some kind of wave to be associated with a particle, plus the assumption that

More information

Quantum Orbits. Quantum Theory for the Computer Age Unit 9. Diving orbit. Caustic. for KE/PE =R=-3/8. for KE/PE =R=-3/8. p"...

Quantum Orbits. Quantum Theory for the Computer Age Unit 9. Diving orbit. Caustic. for KE/PE =R=-3/8. for KE/PE =R=-3/8. p... W.G. Harter Coulomb Obits 6-1 Quantum Theory for the Computer Age Unit 9 Caustic for KE/PE =R=-3/8 F p' p g r p"... P F' F P Diving orbit T" T T' Contact Pt. for KE/PE =R=-3/8 Quantum Orbits W.G. Harter

More information

IV. Electronic Spectroscopy, Angular Momentum, and Magnetic Resonance

IV. Electronic Spectroscopy, Angular Momentum, and Magnetic Resonance IV. Electronic Spectroscopy, Angular Momentum, and Magnetic Resonance The foundation of electronic spectroscopy is the exact solution of the time-independent Schrodinger equation for the hydrogen atom.

More information

Preliminary Quantum Questions

Preliminary Quantum Questions Preliminary Quantum Questions Thomas Ouldridge October 01 1. Certain quantities that appear in the theory of hydrogen have wider application in atomic physics: the Bohr radius a 0, the Rydberg constant

More information

QUANTUM MECHANICS. Franz Schwabl. Translated by Ronald Kates. ff Springer

QUANTUM MECHANICS. Franz Schwabl. Translated by Ronald Kates. ff Springer Franz Schwabl QUANTUM MECHANICS Translated by Ronald Kates Second Revised Edition With 122Figures, 16Tables, Numerous Worked Examples, and 126 Problems ff Springer Contents 1. Historical and Experimental

More information

Quantum Mechanics I Physics 5701

Quantum Mechanics I Physics 5701 Quantum Mechanics I Physics 5701 Z. E. Meziani 02/24//2017 Physics 5701 Lecture Commutation of Observables and First Consequences of the Postulates Outline 1 Commutation Relations 2 Uncertainty Relations

More information

1 The postulates of quantum mechanics

1 The postulates of quantum mechanics 1 The postulates of quantum mechanics The postulates of quantum mechanics were derived after a long process of trial and error. These postulates provide a connection between the physical world and the

More information

Atomic Structure and Atomic Spectra

Atomic Structure and Atomic Spectra Atomic Structure and Atomic Spectra Atomic Structure: Hydrogenic Atom Reading: Atkins, Ch. 10 (7 판 Ch. 13) The principles of quantum mechanics internal structure of atoms 1. Hydrogenic atom: one electron

More information

P3317 HW from Lecture 15 and Recitation 8

P3317 HW from Lecture 15 and Recitation 8 P3317 HW from Lecture 15 and Recitation 8 Due Oct 23, 218 Problem 1. Variational Energy of Helium Here we will estimate the ground state energy of Helium. Helium has two electrons circling around a nucleus

More information

Atomic Structure. Chapter 8

Atomic Structure. Chapter 8 Atomic Structure Chapter 8 Overview To understand atomic structure requires understanding a special aspect of the electron - spin and its related magnetism - and properties of a collection of identical

More information

d 3 r d 3 vf( r, v) = N (2) = CV C = n where n N/V is the total number of molecules per unit volume. Hence e βmv2 /2 d 3 rd 3 v (5)

d 3 r d 3 vf( r, v) = N (2) = CV C = n where n N/V is the total number of molecules per unit volume. Hence e βmv2 /2 d 3 rd 3 v (5) LECTURE 12 Maxwell Velocity Distribution Suppose we have a dilute gas of molecules, each with mass m. If the gas is dilute enough, we can ignore the interactions between the molecules and the energy will

More information

Opinions on quantum mechanics. CHAPTER 6 Quantum Mechanics II. 6.1: The Schrödinger Wave Equation. Normalization and Probability

Opinions on quantum mechanics. CHAPTER 6 Quantum Mechanics II. 6.1: The Schrödinger Wave Equation. Normalization and Probability CHAPTER 6 Quantum Mechanics II 6.1 The Schrödinger Wave Equation 6. Expectation Values 6.3 Infinite Square-Well Potential 6.4 Finite Square-Well Potential 6.5 Three-Dimensional Infinite- 6.6 Simple Harmonic

More information

ψ s a ˆn a s b ˆn b ψ Hint: Because the state is spherically symmetric the answer can depend only on the angle between the two directions.

ψ s a ˆn a s b ˆn b ψ Hint: Because the state is spherically symmetric the answer can depend only on the angle between the two directions. 1. Quantum Mechanics (Fall 2004) Two spin-half particles are in a state with total spin zero. Let ˆn a and ˆn b be unit vectors in two arbitrary directions. Calculate the expectation value of the product

More information

CONTENTS. vii. CHAPTER 2 Operators 15

CONTENTS. vii. CHAPTER 2 Operators 15 CHAPTER 1 Why Quantum Mechanics? 1 1.1 Newtonian Mechanics and Classical Electromagnetism 1 (a) Newtonian Mechanics 1 (b) Electromagnetism 2 1.2 Black Body Radiation 3 1.3 The Heat Capacity of Solids and

More information

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

Physics 342 Lecture 26. Angular Momentum. Lecture 26. Physics 342 Quantum Mechanics I Physics 342 Lecture 26 Angular Momentum Lecture 26 Physics 342 Quantum Mechanics I Friday, April 2nd, 2010 We know how to obtain the energy of Hydrogen using the Hamiltonian operator but given a particular

More information

Physics 342 Lecture 27. Spin. Lecture 27. Physics 342 Quantum Mechanics I

Physics 342 Lecture 27. Spin. Lecture 27. Physics 342 Quantum Mechanics I Physics 342 Lecture 27 Spin Lecture 27 Physics 342 Quantum Mechanics I Monday, April 5th, 2010 There is an intrinsic characteristic of point particles that has an analogue in but no direct derivation from

More information

Lecture #1. Review. Postulates of quantum mechanics (1-3) Postulate 1

Lecture #1. Review. Postulates of quantum mechanics (1-3) Postulate 1 L1.P1 Lecture #1 Review Postulates of quantum mechanics (1-3) Postulate 1 The state of a system at any instant of time may be represented by a wave function which is continuous and differentiable. Specifically,

More information

Angular momentum and spin

Angular momentum and spin Luleå tekniska universitet Avdelningen för Fysik, 007 Hans Weber Angular momentum and spin Angular momentum is a measure of how much rotation there is in particle or in a rigid body. In quantum mechanics

More information

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

which implies that we can take solutions which are simultaneous eigen functions of Module 1 : Quantum Mechanics Chapter 6 : Quantum mechanics in 3-D Quantum mechanics in 3-D For most physical systems, the dynamics is in 3-D. The solutions to the general 3-d problem are quite complicated,

More information

1 Measurement and expectation values

1 Measurement and expectation values C/CS/Phys 191 Measurement and expectation values, Intro to Spin 2/15/05 Spring 2005 Lecture 9 1 Measurement and expectation values Last time we discussed how useful it is to work in the basis of energy

More information

Quantum mechanics in one hour

Quantum mechanics in one hour Chapter 2 Quantum mechanics in one hour 2.1 Introduction The purpose of this chapter is to refresh your knowledge of quantum mechanics and to establish notation. Depending on your background you might

More information

13 Spherical Coordinates

13 Spherical Coordinates Utah State University DigitalCommons@USU Foundations of Wave Phenomena Library Digital Monographs 8-204 3 Spherical Coordinates Charles G. Torre Department of Physics, Utah State University, Charles.Torre@usu.edu

More information

Schrödinger equation for the nuclear potential

Schrödinger equation for the nuclear potential Schrödinger equation for the nuclear potential Introduction to Nuclear Science Simon Fraser University Spring 2011 NUCS 342 January 24, 2011 NUCS 342 (Lecture 4) January 24, 2011 1 / 32 Outline 1 One-dimensional

More information

Coupling of Angular Momenta Isospin Nucleon-Nucleon Interaction

Coupling of Angular Momenta Isospin Nucleon-Nucleon Interaction Lecture 5 Coupling of Angular Momenta Isospin Nucleon-Nucleon Interaction WS0/3: Introduction to Nuclear and Particle Physics,, Part I I. Angular Momentum Operator Rotation R(θ): in polar coordinates the

More information

Introduction to Quantum Physics and Models of Hydrogen Atom

Introduction to Quantum Physics and Models of Hydrogen Atom Introduction to Quantum Physics and Models of Hydrogen Atom Tien-Tsan Shieh Department of Applied Math National Chiao-Tung University November 7, 2012 Physics and Models of Hydrogen November Atom 7, 2012

More information

A few principles of classical and quantum mechanics

A few principles of classical and quantum mechanics A few principles of classical and quantum mechanics The classical approach: In classical mechanics, we usually (but not exclusively) solve Newton s nd law of motion relating the acceleration a of the system

More information

3. Quantum Mechanics in 3D

3. Quantum Mechanics in 3D 3. Quantum Mechanics in 3D 3.1 Introduction Last time, we derived the time dependent Schrödinger equation, starting from three basic postulates: 1) The time evolution of a state can be expressed as a unitary

More information

Representation theory and quantum mechanics tutorial Spin and the hydrogen atom

Representation theory and quantum mechanics tutorial Spin and the hydrogen atom 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

More information

Atomic Systems (PART I)

Atomic Systems (PART I) Atomic Systems (PART I) Lecturer: Location: Recommended Text: Dr. D.J. Miller Room 535, Kelvin Building d.miller@physics.gla.ac.uk Joseph Black C407 (except 15/1/10 which is in Kelvin 312) Physics of Atoms

More information

(Again, this quantity is the correlation function of the two spins.) With z chosen along ˆn 1, this quantity is easily computed (exercise):

(Again, this quantity is the correlation function of the two spins.) With z chosen along ˆn 1, this quantity is easily computed (exercise): Lecture 30 Relevant sections in text: 3.9, 5.1 Bell s theorem (cont.) Assuming suitable hidden variables coupled with an assumption of locality to determine the spin observables with certainty we found

More information

QM and Angular Momentum

QM and Angular Momentum Chapter 5 QM and Angular Momentum 5. Angular Momentum Operators In your Introductory Quantum Mechanics (QM) course you learned about the basic properties of low spin systems. Here we want to review that

More information

The Wave Function. Chapter The Harmonic Wave Function

The Wave Function. Chapter The Harmonic Wave Function Chapter 3 The Wave Function On the basis of the assumption that the de Broglie relations give the frequency and wavelength of some kind of wave to be associated with a particle, plus the assumption that

More information

Quantum Physics III (8.06) Spring 2007 FINAL EXAMINATION Monday May 21, 9:00 am You have 3 hours.

Quantum Physics III (8.06) Spring 2007 FINAL EXAMINATION Monday May 21, 9:00 am You have 3 hours. Quantum Physics III (8.06) Spring 2007 FINAL EXAMINATION Monday May 21, 9:00 am You have 3 hours. There are 10 problems, totalling 180 points. Do all problems. Answer all problems in the white books provided.

More information

One-electron Atom. (in spherical coordinates), where Y lm. are spherical harmonics, we arrive at the following Schrödinger equation:

One-electron Atom. (in spherical coordinates), where Y lm. are spherical harmonics, we arrive at the following Schrödinger equation: One-electron Atom The atomic orbitals of hydrogen-like atoms are solutions to the Schrödinger equation in a spherically symmetric potential. In this case, the potential term is the potential given by Coulomb's

More information

Physics 221A Fall 1996 Notes 16 Bloch s Theorem and Band Structure in One Dimension

Physics 221A Fall 1996 Notes 16 Bloch s Theorem and Band Structure in One Dimension Physics 221A Fall 1996 Notes 16 Bloch s Theorem and Band Structure in One Dimension In these notes we examine Bloch s theorem and band structure in problems with periodic potentials, as a part of our survey

More information