quantum mantras

Upload: frank-munley

Post on 07-Apr-2018

318 views

Category:

Documents


3 download

TRANSCRIPT

  • 8/6/2019 Quantum Mantras

    1/24

    1

    QUANTUM MANTRAS

    Review of and supplement to P.C.W. Davies, Quantum Mechanics (1984)

    Frank Munley, June 2011

    CHAPTER 1: Preliminary concepts

    1. Quantum mechanics is weird in the sense that it describes properties of matter that dont seem to

    match and indeed contradict what we observe or are habituated to at the human level. One hallmark of this

    weirdness is wave-particle duality, i.e., depending on the experimental circumstances, it is sometimes

    convenient to treat a physical system as having wave properties, and at other times as having particle

    properties. How something can simultaneously behave like a wave and a particle is counterintuitive to the

    extreme. For example, just as a beam of light shows an interference pattern when it passes through two

    slits in an opaque screen, so an electron beam exhibits an interference pattern even though electrons behave

    like particles in many s ituations. The De Broglie relationship between the particle property of

    momentum and wave property wavelength is: p = h /"= hk, where h = h / 2" , h = Plancks constant,

    and k= 2" /#.

    2. Another manifestation of wave-particle duality is the relationship between energy and frequency:

    E= h" = h#. AlthoughE= h" holds for both a material particle which has a rest mass and a photon whose

    rest mass is zero, an important difference arises in the relationship between the wave numberkand the

    frequency " for these two particles. For a photon, "= kv , i.e., "= kc where c = speed of light, while for

    a material particle, "= (h / 2m)k2 . This important difference leads to the Schrdinger equation for material

    particles stated below.

    3. Uncertainty in frequency: Frequency is most directly measured by counting the number of wave peaks

    passing a fixed point. If the counting lasts a time " , andNpeaks are counted, then " = N /# . But the

    minimal uncertainty in the number is 1, so "# =1/$ . Measuring frequency by counting: = 1/, where

    is the time taken to count. " can also be a measure of the uncertainty in time " t.

    4. Items 2 and 3 lead to the Heisenberg uncertainty relation for energy and time is: "E"t # h , so it is

    impossible to carry out perfectly accurate measurements of both energy and time interval over which the

    measurement is made.

    5. " is the wave function which can be used to describe the energy, position, momentum, etc. of a

    material particle. All measurements give real numbers, but " is generally complex. So by itself, " is not

    observable; but "2

    links us to physical reality:

    In one dimension, "(x, t)2dx = probability that a particle can be found (when a position

    measurement is made) between x and x + dx. In three dimensions, "(r, t)2d# =

    probability that a particle can be found in the volume element d" about the position r.

    (N.B.:

    d" is a common notation for a 3-D volume element and has nothing to do with

    counting time.) The probability that a particle can be found in a limited region is given

    by the integral of "(r, t)2d# over that region. In particular, in one dimension, the

    probability that a particle can be found between x1 and x2 is "(x, t)2dx

    x1

    x2

    # . Since

    probabilities should sum to 1,

    "(x, t)2dx = 1

    x1

    x2

    # (or, in 3-D, "(r, t) 2d# = 1x1

    x2

    $ ) withthe integration carried out over all space. In this case, we say the wave function is

    normalized.

    6. Consider a particle localized on a one-dimensional line in a region of size

    " x. Since the particle also has

    a wave nature, the question arises: What range of wavelengths have to be added together to form a wave

    function, or in this context a localized wavepacket, of size " x? A rather straightforward application of

    Fourier analysis gives us the answer. To suggest how this works, suppose the particle is known to be in a

  • 8/6/2019 Quantum Mantras

    2/24

    2

    one-dimensional line box of size " x and we want to determine what range of wavelengths that are

    required to cancel outside the box but add together in the box. This can be done precisely using Fourier

    analysis, but we can guesstimate the range of wavelengths needed as follows.

    The diagram shows the one-dimensional box containing the particle. The wave function has the

    constant value "0

    inside the box and zero outside. Also shown are three sinusoidal waves, each wave

    representing a particular possible momentum of the particle if the momentum is measured. The figuresuggests that the waves add together inside the box but outside sometimes add but sometimes cancel.

    These waves help out, because they are all non-zero fromL/2 to +L/2 and tend to cancel outside. In other

    words, waves with "# L / 2 , if combined in the correct proportions as determined by Fourier analysis, canproduce a wave function which is equal to "

    0betweenL/2 and +L/2. Contrast these waves with shorter-

    wavelength waves:

    As the diagram suggests, these waves tend to cancel out almost everywhere inside the box, and are largely

    unnecessary to produce the desired ". But they cant be completely ignored, because combined in proper

    proportion, as determined by Fourier analysis, they help the larger waves to cancel outside the box and to

    produce the constant "0 in the box.

    A precise Fourier analysis yields the amplitude as a function of wavelength in order to achieve

    precisely the desired ". It is more convenient to do the analysis in terms of the amplitude as a function of

    wave numberkrather than wavelength. The following diagram shows the result, whereA(k) is the

    amplitude forL = 1 (and "0= 1):

    "

    x

    "0

    -L/2 L/2

    2" /L

    A(k)

    k

    "

    x

    "0

    The particle is equally likely to be found

    anywhere between L/2 and L/2. Theprobability of being anywhere in the boxis constant from point to point, so the

    probability of being betweenx andx +dx is constant: Prob(x, x+dx) =

    |0|2dx.

    -L/2 L/2

  • 8/6/2019 Quantum Mantras

    3/24

    3

    In general, for a 1-dimensional box between L/2 and +L/2, A(k) =2"0

    2#

    sin(kL /2)

    k, and

    "0 =1/ L . As the graph shows, the greatest amplitude is for waves between 2" /L for a spread of

    4" /L . Butthe amplitude is small as one goes to the limits of this range, so a reasonable estimate of the

    spread or uncertainty in kis about half of this, or 2" /L . Since p = hk, then "p = h"k# 2$h /L = h /L .

    But the particle can be anywhere inside the box with equal probability, soL is a measure of the uncertainty

    in the position of the box, i.e., "x = L . Therefore, "p"x # h .

    In our box example, the particle is free and just bounces inside the box between its ends. In a

    more complicated situation, the particle might be on a rubber band while bouncing back and forth, or

    subject to some other type of potential energy function. In any case, the uncertainty relation is an order-of-

    magnitude estimate, and the more general formulation of the uncertainty relation is framed in terms of h :

    The Heisenberg uncertainty relation for position and momentum: "x"p # h .

    The significance is that a more accurate measurement ofp creates a greater uncertainty inx, and vice versa.

    Most significantly, perfectly accurate measurements of position and momentum are impossible to carry out:

    the product of their uncertainties must always be h .

    7. " must be single-valued, well-behaved (cant blow up), and it and its first derivative (or gradient in 3-D)must be continuous. One exception to the latter rule: where the potential is infinite, just set " =0.

    8. Just as particles can flow from one place to another, so the probability of the particles being found at one

    place or another can flow. The probability current density in one dimension is:

    *)*(2

    ),( !!!! "#"#=m

    it

    hrj .

    9. The weighted average of something is just a sum of the values it can have times the probability of

    having the values. E.g., in one dimension,

    xn= "

    #x

    n"dx

    $%

    %

    & .

    10. " can be determined from the Schrdinger Equation (V(r) is the potential affecting the particle of

    mass m):

    tiV

    m !

    !=+"

    # $$$ h

    h)(

    2

    22

    r (A KEY EQUATION!)

    CHAPTER 2: Wave mechanics I

    1. The Schrdinger Equation is separable into time and space parts. The space part of(r,t) is u(r) or just

    u:

    "h2

    2m#

    2u +V(r)u = Eu (ANOTHER KEY EQUATION!)

    This equation is a recipe for grinding out allowed energy values (oreigenvalue) for a particle bound in a

    potential. The wave function associated with each energy eigenvalue is called an energy eigenfunction.

    2. When a particle is free, the allowed energy values in the Schrdinger equation are continuous; if the

    particle is bound (e.g., in an infinite or finite square well, by a harmonic force, etc.), the energy values are

    discreteonly special values, E1, E2, etc., are allowed.

    3. The time part of the Schrdinger equation tells us thath/)(),(

    iEtexutx!

    =" for a particle in a well-

    defined energy state. Therefore, ||2

    = |u|2

    and to normalize u is to normalize .

  • 8/6/2019 Quantum Mantras

    4/24

    4

    4. Most problems are very difficult to solve in quantum mechanics, but the infinite square well is an

    exception. In one dimension (1-D), this corresponds to a line with rigid walls at both ends. The form of

    the wave functions is completely similar to waves on a string, and that is the easiest way to generate them.

    However, you should also know how to obtain them by a systematic solution of the 1-D Schrdinger

    equation. The solutions are shown here (think of waves on a string with the endpoints fixed, so the wave is

    zero there), and the general form is: for n odd, and for n odd.

    un=

    1

    a

    cosn"x

    2a

    #

    $%&

    '(forn odd, and u

    n=

    1

    a

    sinn"x

    2a

    #

    $%&

    '(forn even.

    As the captions of the diagrams show, the de Broglie wavelengths are, in general, 4a/n.

    5. The energies of the particle in an infinite square well are purely kinetic. Since k = 2/ and p = k, we

    have the following equivalent expressions for En: En = pn2/2m = 2kn

    2/2m = 2(42)/(2mn2) =

    2

    (42

    )/[2m(4a/n)2

    ] = n2

    2

    2

    /(8ma2

    ), n = 1,2,3,

    6. Theorem: IfV(x) is symmetric, then the non-degenerate wave functions un have definite odd or even

    parity. (Example: the infinite square well potential. The uns are either sines or cosines with definite

    parity.)

    7. The energy eigenvalues of the finite square well (from a to +a, i.e., V(x) is symmetric) are determined

    by graphically solving the following equations:

    tan(a) = / (even parity), cot(a) = -/ (odd parity), with

    2/12 )/2( hmE=! , and .]/)(2[ 2/12hEVmo!="

    CHAPTER 3: Wave mechanics II

    1. The simple harmonic oscillator, for which V(x) = Kx2/2, has energy values !h)(21

    += nEn

    , and

    eigenfunctions )x(Hand)/m(with,e)x(H!n

    un

    //x

    n

    /

    n/n!"=!!#$

    %&'

    (

    )

    !=

    !* 212

    21

    21

    22

    2h is the

    nth Hermite polynomial. Note that the lowest energy state has non-zero energy of/2. This is a reflection

    of the uncertainty principle: zero energy would imply that momentum and position are both exactly equal

    to zero, which is prohibited by the uncertainty principle.

    V(x) oru(x)

    -a +a

    V(x) oru(x)

    -a +a

    V(x) oru(x)

    -a +a

    V(x) oru(x)

    -a +a

    u1; = 4a

    (i.e., 4a/1)

    u3; = 4a/3 u4; = a

    (i.e., 4a/4)

    u2; = 2a

    (i.e., 4a/2)

  • 8/6/2019 Quantum Mantras

    5/24

    5

    2. The hydrogen atom in its simplest treatment has the electron subjected to the Coulomb potential V(r) = -

    e2/(4or). By neglecting angular motions of the electron, we can rather simply determine the energy levels

    of the atom:

    222

    4

    )4(2 n

    meE

    o

    n

    h!"#= .

    Surprisingly, this is precisely what the simple Bohr theory of the H-atom predicts. The independence of the

    energy eigenvalues on angular motions is analogous to the result from classical mechanics that the energy

    of a planet or satellite under the influence of one other bodys gravitational force depends only on the major

    axis of the elliptical orbit, not on its eccentricity. (It is the eccentricity which involves angular motions,

    specifically angular momentum.)

    3. The energy levels in the previous item can be determined by assuming that2/

    e!"

    !=! )F()u( (where

    r!=" and2

    mE/h82

    !=" ). Then )F(! obeys the differential equation:

    0Fd

    dF

    d

    Fd2

    2

    =!!

    "

    #$$

    %

    &

    '

    ()+

    '!!

    "

    #$$

    %

    &(

    '+

    '

    11

    2,

    where2

    4

    2

    h!"#=$

    o

    2me

    . Assume F is a polynomial of degree k and substitute it into the d.e. The

    coefficient of the leading term in the differential equation is then1-k

    ! , and since this term is not canceled

    by anything else on the left side of the equation, it must must be 0, and this means that 1k +=! , k = 0,

    1, 2,, i.e., 3,..2,1,n =!" . Since ! (which is in ! ) has E in it, solving for E gives the energylevels.

    4. Free particles. Free particles are bothersome, because strictly speaking, their wave functions, of the

    form)/(),( hEtiet !"=# rr , are not normalizable. It is convenient, then, to imagine free particles to be

    in fact confined to a very large cubical box with sides of length L. In one dimension, we take this to be a

    very long line of length L. In these cases, the normalized wave functions for are:

    "(x, t) =1

    L1/ 2

    ei(kx#Et/h )

    (1# D), and for 3 - D, "(r, t) =1

    L3 / 2

    ei(k$r#Et/h )

    (3 # D),

    where the 1-D case is appropriate for a particle moving in the +x direction. For the 1-D case, the

    probability current (see Chapter 1, item 8 above) is: j=dProb

    dt=

    v

    L, where v is the velocity of the particle

    andProb is the probability that the particle leaves the box. Therefore, Prob =vt

    L

    . Thus, when t = 0, Prob

    = 0, since we assume the particle is initially in the box. After a time t = L/v, the particle must have left the

    box. (If it starts at the left end of the box, it takes it this time to reach the right side and leave the box.)Therefore, when t = L/v,Prob = 1. More often, we are interested in the flux of particles in a beam, e.g., the

    number of particles per second passing a given point or impacting a target. This can be represented in one

    dimension by a wave function "(x, t) = Aei(kx#Et/h )

    , where the flux is represented by the symbol A. It is

    typical to take A = 1 for the incident beam wave function, since all we are interested in are ratios of fluxes.

    (See next item.)

    5. Scattering from a potential step:Considera beam of energy E = 2k2/2m coming towards a potential

    step of height Vo

  • 8/6/2019 Quantum Mantras

    6/24

    6

    classical physics cannot explain such a phenomenon. The reflection and transmission fractions, R and T,

    are:

    .)(

    4,

    )(

    )(22

    2

    kk

    kkT

    kk

    kkR

    !+

    !=

    !+

    !"=

    6. Tunneling through a rectangular barrier. Consider a potential V(x) = Vo for 0 < x < a, and V(x) = 0everywhere else. Classically, a beam with energy E < Vo would be totally reflected, but in quantum

    mechanics, part of the beam tunnels through:

    .)(2

    ,)(4

    sinh1

    2/1

    2

    122

    !"

    #$%

    & '=

    !!"

    #

    $$%

    &

    '+=

    '

    h

    EVm

    EVE

    aVT

    o

    o

    o(

    (

    7. Consider a set of N non-interacting distinguishable particles, described by wave functions

    N!!! ,...,

    21. The joint probability that #1 is in volume element

    1!d , #2 is in

    2!d ,etc., is the product

    of the separate particle probabilities: Prob =NN

    d...dd !"!"!"2

    2

    2

    21

    2

    1. We will get this result if

    we take the combined wave function to be the product of the individual ones:N

    ...!!!="21

    . But if

    the particles are indistinguishable, we must have a combined wave function which reflects thisindistinguishablity. Two cases arise. For some particles, called bosons (e.g., photons), the wave function

    must be symmetric under an exchange of particle coordinates. For other particles, called fermions (e.g.,

    electrons, protons, and neutrons), the wave function must be antisymmetric under an exchange of particle

    coordinates. See Davies for simple examples of each for two particles.

    CHAPTER 4: The formal rules of quantum mechanics

    1. Ordinary functions as vectors in infinite-dimensional space. In ordinary geometrical space, we have

    three dimensions, which we can label 1, 2, and 3 (usually corresponding to x,y, and z), so a vector in 3-

    space can be written as A = (A1,A2,A3). The scalar or dot product between this A and another vectorB =

    (B1,B2,B3) is, as everyone knows, just A "B= A1B1 , A2B2, A3B3.

    A function ofxcan be thought of as a vector in an infinite-dimensional Hilbert space withthe variablex playing the indexing role of 1,2,3 except that it can take on an infinite number of values.

    To understand how this works, consider first a five-dimensional space defined atx = 1.5,x = 2.2,x = 2.4,

    x = 3.5, andx = 3.8, and a functionf(x) =x2, defined over these values. Then everything we want to knowabout this function is represented in the vectorf= (2.25, 4.84, 5.76, 12.25, 14.44). Similarly, a function

    g(x) =x1.5 isg= (1.837, 3.263, 3.718, 6.548, 7.408). And the dot product offandgis just (2.25*1.837 +

    + 14.44*7.408) = 228.5. Now imagine we define functions over not five values ofx between 1 and 4 but

    over a billion values, more or less equally spaced. Thenfandgwould be billion-component vectors, and

    the scalar product between them would be very close to the integral of (x2)(x1.5) from 1 to 4. Going to the

    limit of continuousx over the entire real line leads to a vector representation of the functions, which we can

    just represent as (fx) and (gx), or more conventionally as f(x) =x2 andg(x) =x1.5, with the scalar product

    betweenx = a andx = b being:

    f(x)g(x)dx =a

    b

    " x3.5

    dxa

    b

    " .

    2. Two functions, f(x) and g(x), are considered mutually orthogonal (or just orthogonal) on the interval

    [a,b] if!b

    a

    f(x)g(x)dx = 0, with an analogous equation for three dimensions.. Usually, a = - and b = +.

    3. Solutions to the Schrdinger equation corresponding to different energies are mutually orthogonal:

  • 8/6/2019 Quantum Mantras

    7/24

    7

    ! =b

    a

    mnnmdx ,* "## (A KEY EQUATION!)

    with an analogous equation for three dimensions.

    4. For a given physical system, the solutions (i.e., eigenfunctions) of the Schrdinger equation form a basis

    (analogous to z,y,x

    ), so that just as any three-dimensional vectorr can be written as a sum of z,y,x

    , socan any arbitrary function f(x) be written as a sum of the solutions to the Schrdinger equation. Thus, (x)

    = cnn, and the sum is in general, over an infinite number of eigenfunctions. Fourier analysis is a good

    example of this for a free particle. The cn numbers are called expansion coefficients.

    5. To find the expansion coefficients in (x) = cnn, we can use the orthogonal property of the n

    eigenfunctions. Thus, if we want cm, multiply the equation by m*, integrate, and we get:

    cm= ! m*dx, (A KEY EQUATION!)with an analogous equation for three dimensions.

    6. The probability that a measurement on a system in state will yield the energy Em is:

    Pmc = m2. (A KEY EQUATION!)

    7. Collapse of the wave function: If a measurement shows that the system has energy En, then the system

    is in the staten

    ! right after the measurementi.e., the wave function collapses from (x) = cnn ton

    ! .

    7. The time-independent Schrdinger Equation can be written in operator form H un = Enun, where

    H = -(h2/2m)

    2! + V(r), and

    2! =

    2

    2

    2

    2

    2

    2

    zyx !

    !+

    !

    !+

    !

    !.

    8. An eigenvalue equation has the form A = , where A is any operator and is a number which in

    general can be complex but in quantum mechanics must be real if it is to correspond to a measured

    quantity. The function is an eigenfunction of the operator A . So the eigenvalue equation shows a

    special relationship between A and : thinking of as a vector, when A operates on it, it doesnt change

    its direction; it only changes its length by the factor.

    9. Any legitimate operator in quantum mechanics, i.e., any operator which corresponds to an observable

    (like position or momentum) must be Hermetian, i.e., it must satisfy the following relationship for any two

    arbitrary functions and :

    ! !"" #$=$# %% dAdA )()( .

    In other words, if we always assume the operator operates to the right in the above expression, A can

    operate on , but it can also operate on *if we change A to A

    *.

    10. Another way to state the Hermetian relationship in the previous item is to think ofA

    being able tooperate right or left, but if it goes to the left and operates on * , we must complex-conjugate A first.

    This symmetry of going left or right, with everything on the left being understood to be complex-

    conjugated, leads to Diracs elegant bracket notation, whereby the integrals in the previous item can both

    be represented as:

    !"="!=!"# #$$

    AdAdA )()( %% .

    In this notation, ! is the bra part of the backet and ! is the ket part.

  • 8/6/2019 Quantum Mantras

    8/24

    8

    11. In the bra and ket notation, the expected value of an operator in a state ! is:

    !!= AA . (KEY EQUATION)

    12. The commutator, [ BA , ] , of two operators is defined as ABBA ! . Usually, we must operate onsome actual function to obtain the correct properties of the commutator.

    13. If two operators commute, i.e., if their commutator is zero, then the operators can have simultaneous

    eigenfunctions. If they dont commute, then they cannot have simultaneous eigenfunctions.

    14. Two physical observables can be simultaneously measured without any error if and only if their

    operators commute. In other words, a measurement of one will not affect the measurement of the other if

    their operators commute.

    15. The operators for position and momentum do not commute: [ ] hipx =, , i.e., hh ix

    i,x =!"

    #$%

    &

    '

    '( .

    Because they dont commute, they cant be simultaneously measured without disturbing each other.

    Therefore, they obey the Heisenberg uncertainty principle. (See Chapter 1 Quantum Mantras.)

    16. In general, for any two operators A and B , which have eigenvalues and respectively, we define

    the uncertainty of the measured values as =2

    2 AA ! , and likewise for.

    17. In general, for any two operators A and B , which have eigenvalues and respectively,

    >1

    2A,B[ ] . This again implies that they are simultaneously measurable, i.e., = 0, only

    if the operators commute. If the operators dont commute, they obey the uncertainty principle.

    18. The commutator of an operator with the Hamiltonian operator tells us what the average time rate of

    change of that operator and its observable are:

    d

    dt"A" =

    i

    h"[H,A]" . (A KEY EQUATION!)

    In particular, if the commutator of an operator with H is zero, the observable corresponding to the

    operator is conserved, i.e., it is constant in time.

    CHAPTER 5: Angular Momentum

    1. Classically, the angular momentum is L = r!p, and Lx = ypz zpy, etc. We obtain the quantum

    mechanical operator by substituting the operators for y, pz, etc., soy

    ziz

    iyLx

    !

    !+

    !

    !"= hh )( , etc.

    2. The operators for Lx, Ly, and Lz do not commute: zyx LiLL, h= , etc. Therefore, these components

    cannot be measured with perfect accuracy simultaneously. In a sense, a measurement of one interferes ordisturbs a measurement of the other.

    3.2L does commute with all three components ofL. Since

    zL is very simple in spherical coordinates

    (!"

    "#= hiL

    z ), we usually choose to work with simultaneous eigenfunctions of

    2L andz

    L .

  • 8/6/2019 Quantum Mantras

    9/24

    9

    un=

    1

    a

    cosn"x

    2a

    #

    $%&

    '(

    4. The eigenfunctions of2L and

    zL are designated by the notation m,l , where l and m are the

    quantum numbers of2L and

    zL respectively. For well-behaved eigenfunctions (i.e., single-valued &

    finite), we must have l a positive integer: l =0,1,2,3,. For each l , m = - l , - l +1, - l +2,-2, -1, 0, 1,

    2, l -2, l -1, l , for a total of 2 l +1 values ofm . The eigenvalues of2L and

    zL are, respectively,

    2)1( hll + and hm .

    5. In spherical coordinates, m,l =m

    imm

    mYePN

    lll=

    !")(cos , where

    mN

    lis the normalization

    constant:

    2/1

    )!(

    )!(

    4

    12)1( !

    "

    #$%

    &

    +

    '+'=

    m

    mN

    m

    m

    l

    lll

    (

    , and )(cos!m

    Pl

    is an associated Legendre polynomial.

    The entire wave function is the well-known spherical harmonic,m

    Yl

    .

    6. The mYl functions are orthonormal: ''',', mmmm !!llll = .

    7. Them

    Yl

    functions can be used to construct matrices for2L and

    zL . For each value ofl , the 2 l +1

    values ofm define a 2 l +1 by 2 l +1 matrix representation for2L and

    zL . Therefore, for l =0, we have a

    1x1 matrix (the sole element is 0); for l =1, a 3x3 matrix; for l =2, a 5x5 matrix; etc. (Where are the

    even matrices?? See below!) The matrices for2L and

    zL are:

    !!!!!!

    "

    #

    $$$$$$

    %

    &

    +=

    1......00

    0...::......:

    0......10

    0......01

    )1( 22 hllL

    , and zL

    =

    !!!!!!

    "

    #

    $$$$$$

    %

    &

    '

    +'

    '

    l

    l

    l

    l

    h

    ....00

    01......0

    0...0:

    0..010

    0....0

    .

    Note that the matrices for2L and

    zL are both diagonal, reflecting the fact that they commute and hence

    have common eigenfunctions when expressed in terms of the eigenvectors ofz

    L . Also, note that the

    diagonal elements are their eigenvalues. A similar process allows matrices forx

    L and yL to be obtained.

    For l =1, the complete set of matrices are as follows.

    !!!

    "

    #

    $$$

    %

    &

    '

    '

    =

    !!!

    "

    #

    $$$

    %

    &

    =

    !!!

    "

    #

    $$$

    %

    &

    '

    =

    !!!

    "

    #

    $$$

    %

    &

    =

    00

    0

    00

    2,

    010

    101

    010

    2,

    100

    000

    001

    ,

    100

    010

    001

    222

    i

    ii

    i

    LLLL yxzhh

    hh

    Note that the matrices matrices forx

    L and yL are not diagonal; the off-diagonal elements reflect the

    fact that we have chosen to express everything in simultaneous eigenstates of2L and

    zL , but

    xL and

  • 8/6/2019 Quantum Mantras

    10/24

    10

    yL do not commute with

    zL . Note also that all four matrices are Hermetian, which for a matrix

    means that the complex conjugate of the transpose of the matrix leaves the matrix unchanged.

    8. For a givern l , the most general eigenfunction is a 12 +l -component vector:

    !!!!!!

    "

    #

    $$$$$$

    %

    &

    =

    +12

    2

    1

    .

    .,

    l

    l

    a

    a

    a

    Cm ,*

    12

    *

    2

    *

    1 ....*, += ll aaaCm

    where C is a normalization factor. (N.B.: the bra form is complex-conjugated!) The expectation of any

    matrix operator A is just mAm ,, ll .

    9. The eigenvectors of2L and

    zL are just like m,l shown in #19, but with all zeros except for the

    particular value ofz

    L which is measured. For example, if l =5, there are 10 eigenvalues ofm , from +5

    to -5. The eigenvector which will yield an zL

    value of +3 will have a3 = 1 while all othera-values are

    zero; the eigenvector for 5 will have a10 = 1 while all others are 0.

    10. As we saw above, the r!p type of momentum yields only integral values of l and square matrices

    with an odd number of elements in a row or column. But not all angular momentum is of the r!p type.

    Intrinsic angular momentum, more commonly called spin, fills in the gaps with half-integral values of l .

    For example, if l =1/2, we get a two-by-two matrix, meaning that only two values ofz

    L are possible.

    Thats experimentally true for an electron. Any particle with a half-integral value ofl is called a

    fermion, while one with an integral value is called a boson. The crucial difference is that they obey

    different types of statistics, but this will not be explored here.

    11. For an electron, we designate its 2 by 2 matrices by the letter S to remind ourselves that were

    dealing with spin. Because of the great importance of these matrices, they are called the Pauli spin

    matrices after their formulator Wolfgang Pauli (1900-1958).

    ,10

    012

    4

    32

    !!"

    #$$%

    &= hS !!

    "

    #$$%

    &

    '=

    10

    01

    2

    1

    hz

    S , !!"

    #$$%

    &=

    01

    10

    2

    1

    hx

    S , !!"

    #$$%

    & '=

    0

    0

    2

    1

    i

    iSy h .

    12. MEASUREMENT RULE: If a spin is in a state !!"

    #$$%

    &

    b

    a, and a measurement of

    zL is made, the state

    immediately changes to either !!"

    #$$%

    &

    0

    1if the measurement yields +/2 or !!

    "

    #$$%

    &

    1

    0if the measurement yields

    /2. This is true in general. For example, if a system is in a state in which the probability of obtaining a

    momentum between xp

    and xp

    +d xp

    is some continuous function of xp

    , then if a measurement ofmomentum yields the magnitude px, the system immediately jumps into the wavefunction or eigenstate

    corresponding to this value, i.e., the wave function immediately changes toh/xipxe .

    13. If the axis z! makes an angle with the z axis, then the state corresponding to +/2 along z, i.e., the

    state ! , is : !"#$"=$ )2/sin()2/cos( %% . Therefore, if a measurement puts the system in the

    state ! , and then a measurement is made along z! , the probability of finding the spin up along z! is

  • 8/6/2019 Quantum Mantras

    11/24

    11

    )2/(cos2 ! . The probability of finding the spin down along z! is )2/(sin 2 ! . Similar results apply if

    we start with a spin along z or + z! : !"+#"=! )2/cos()2/sin( $$ ;

    !+"="# )2/sin()2/cos( $$ ; !+"#=!$ )2/cos()2/sin( %% .

    CHAPTER 7: Approximation methods

    1. If Eno is the unperturbed energy of the non-degenerate nth energy state u

    nofor the Hamiltonian H

    o,

    and En is the energy of this state when the Hamiltonian isH

    o+ !H ( !H is the perturbation), then for

    !H

  • 8/6/2019 Quantum Mantras

    12/24

    12

    For example, assume we have a one-dimensional system and ut has just two adjustable parameters:

    ut(,;x). Then the variational recipe is as follows:

    (1) Calculatett

    uHuE ),(0 =!" .

    (2) Minimize this with respect to and :

    .0),(

    ,0),( 00

    ==

    !"

    #"!

    !"

    #"! Eand

    E

    (3) The two equations in step (2) can be used to solve for the minimizing values of and , call

    them min and min.

    (4) Now the minimizing values from step (3) are plugged into the average for0

    E in step 1:

    ),( minmin0,0 !"EE estimate = .

    This estimate is always greater than or equal to the actual ground state energy. It cannot be less.

    CHAPTER 8: Transitions

    The following review includes a discussion of electromagnetic (EM) transitions, analogous to the simple

    Bohr model in which EM radiation is produced by an electron undergoing a transition from one state to

    another, or in which EM radiation induces a transition from one state to another. In the latter case, there are

    two possibilities. Either the radiation pumps an electron from a lower to a higher state, i.e., the radiation is

    absorbed, or the radiation induces a transition from a higher to a lower state, i.e., we have stimulated

    emission which results in additional EM radiation. In all of these cases, the EM radiation will be treated as

    a classical continuous electromagnetic field (which you are familiar with from Maxwell's equations),

    because a full quantum mechanical treatment of the radiation field is beyond the scope of this course. In

    other words, the treatment here is "semi-classical": classical for the EM radiation and quantum mechanical

    for the atomic transition itself. (A few brief remarks on the full quantum mechanical treatment can be

    found at the end of Section 11.) The discussion starts off with a general treatment of transitions due to a

    time-dependent perturbation. Electromagnetic radiation will not be specifically discussed until Section 8.

    So keep in mind that everything before that is wondrously general, particularly Fermi's Golden Rule, which

    has a multitude of applications throughout modern physics.

    1. TIME-DEPENDENT PERTURBATION THEORY. Consider a system with Hamiltoniano

    H and

    eigenvectors htiE

    nono

    no

    eut!

    = )(),( rr" . The system is in eigenstate kand at t= 0 it is suddenly

    subject to asmallperturbation )t,(H r! . This changes the wave functions and energy levels, and the new

    wave functions can be expanded in terms of the original wave functions:

    !="#

    h

    tiE

    nonn

    no

    e)(u)t(at,( rr .

  • 8/6/2019 Quantum Mantras

    13/24

    13

    In what follows, the subscript o will be dropped, since it will be understood that everything now refers to

    the original unperturbed wavefunctions and energies. To a first-order approximation, the time-dependent

    expansion coefficient, am(t), is:

    hh

    km

    mk

    tti

    mkm

    EEdteH

    ita mk

    !="= # $

    $

    where,1

    )(0

    .

    Therefore, the probability of a transition from state kto state m is am(t)

    2.

    2. HARMONIC PERTURBATION. The most common type of a time-dependent perturbation is a

    harmonic one which is turned on at t= 0: tcos)(H)t,(H !=" rr . This leads to:

    am= "

    Hmk

    2h

    ei(#

    mk"#)t"1

    #mk

    "#+

    ei(#

    mk+#)t"1

    #mk

    +#

    $

    %&

    '

    ().

    We expect transitions to be significant when " #"mk

    , in which case the first term in the brackets is much

    bigger than the second. The so-called rotating wave approximation consists in neglecting the second term

    for" #"mk

    . In this case, the probability )t,(Pm ! of making a transition in time t to a state m starting

    from a state k withEm > Ek is, to first order, for a small perturbation:

    ./)EE(with,t)(

    t)(sin

    tHa)t,(

    mP

    kmmk

    mk

    mk

    mk

    mh

    h

    !="

    ##$

    %

    &&'

    ( "!"

    ##$

    %

    &&'

    ( "!"

    =="2

    2

    2

    2

    2

    24

    2

    2

    Since the system goes from a lower energy,Ek, to a higher one,Em, this represents absorption.

    Please note th at the perturbation frequency is by definition positive, but mkcan be positive or

    negative. Specifically, mkis negative ifEk > Em, in which case !"!mk in the preceding

    equat ion is replaced by !+!mk

    . This representsstimulated emission (predicted by Einstein in 1917,

    before quantum mechanics was developed in the mid-1920s!) in whic h t he incident radiation induces

    a tr ansition from the sta te wit h energy Em down to the lower energy Ek. In this case,

    Pm

    (", t) = am

    2=

    Hmk

    2t

    2

    4h2

    sin2 ("mk+")t

    2

    #

    $%&

    '(

    ("mk

    +")t

    2

    #

    $%&

    '(

    2, with "

    mk= (E

    m) E

    k) /h < 0.

    3. PERTURBATION FREQUENCY !EQUAL TOmk

    ! . Suppose th at t he frequency of a

    perturbation precisely matches the energy difference between two levels: " ="mk

    . Then the

    probabi li ty of making a transition to the sta te kgrows wit h t ime. This probabil it y can be

    obtained from the equat ion for )t,(Pm ! by taking the limit asmk, which yields:

    (ABSORPTION)

    (STIMULATED

    EMISSION)

  • 8/6/2019 Quantum Mantras

    14/24

    14

    2

    22

    4h

    tH)t,(P

    mk

    mkm =! .

    Please remember that t his ho lds only for a smal l perturbat ion, i.e.,since ),( tPmkm

    ! increases

    with time, it holds only for t imes sufficiently small so that ),( tPmkm

    !

  • 8/6/2019 Quantum Mantras

    15/24

    15

    If the perturbation has a frequency "o, then the probability of a transition at time tis given by the

    magnitude of the )t,(Pm ! curve at that point. As shown in this diagram, o! is between the first and second

    zeros to the left of mk! , so the probability is rather small compared to the maximum 2

    22

    4h

    tHmk

    .

    Note from the equation for )t,(Pm ! that at t= 0, the sine function is identically zero, so the

    Pm(",0) = 0everywhere for any arbitrary ! . This is consistent with the graph of )t,(Pm ! as a function of

    tshown in Figure 8.5.1. This is quite reasonable, since we can't expect a perturbation to cause something to

    happen immediately after it is turned on. Note also the first zeros to the left and right of the central

    resonance peak, which are separated from that peak by t/2! . Therefore, as tincreases, these zeros get

    closer and closer tomk

    ! , and for a fixedo

    ! , these zeros sweep through this value of! , again consistent

    with the graph of )t,(Pm ! vs. t. Simultaneously, as tincreases, all of the peaks must grow in height,

    because as the )t,(Pm ! vs. tgraph shows, by the time the very low peaks initially far from it sweep through

    a given ! , they all have the same height! This is exactly consistent with Figure 8.4.1.

    Figure 8.5.1 illustrates the energy-time uncertainty principle. Imagine that the energy

    differenceEm Ek, or equivalently the frequency "mk , is unknown. To measure it, we subject the system

    to a perturbation of frequency ! . If! is sufficiently close tomk

    ! , then the absorption will be sizable. If

    it is far away, then the absorption will be small. But most of the probability for the transition is in the main

    central peak, which covers a frequency interval ]/2,/2[ ttmk

    !"!" +# . In other words, a perturbation

    which is within this frequency interval has a good chance of inducing a transition from state kto state m.

    We see that if our experiment lasts a time t, then a spread of input frequencies in that interval, i.e., an

    energy range of h("mk

    # 2$ / t) to h("mk

    + 2$ / t) , will be strongly absorbed. Therefore, the spread of

    energies strongly absorbed is tE /4 h!=" , which is also a measure of the uncertainty in the measured

    value of the energy difference between the two levels. Thinking of the measuring time tas the

    uncertainty in time, t! , we can cross-multiply by tt !" and we get hh !"=## 4tE , QED.

    Why is the sky blue? An interesting consequence of the uncertainty principle involves scattering

    of radiation from a molecule. If the radiation is close to a resonant frequency, i.e., if the energy of the

    radiation is close to the energy difference between two molecular levels, then E! is small, and

    "t= 4#h /"E is large. On the other hand, if E! is large, then "t is small. But what is"t? Assume

    that there is absorption of the radiation energy by the molecule. Then there are two interpretations of"t,

    depending on if one is looking at the wave or particle aspect of the interaction between radiation and

    molecule. From the wave aspect, i.e., considering the radiation to be wave-like, "t is the time the

    molecule is subject to the perturbation. From the particle aspect in which the radiation consists of photons

    and we are considering absorption of a photojn, the absorption is for all practical means instantaneous, but

  • 8/6/2019 Quantum Mantras

    16/24

    16

    now "t represents the time the molecule holds on to the radiation. For small E! , the molecule holds

    onto the radiation for a long time, long enough for collision with another molecule to occur. The collision

    will typically thermalize the energy, i.e., the energy absorbed by the molecule is transformed into

    molecular kinetic energy, so the colliding molecules rebound with greater kinetic energy than they started

    with. On the other hand, if E! is large, then the molecule holds onto the radiation for a shorter time, sothe radiation is can be quickly emitted. If emission happens before a collision occurs, then we have

    scattering of the radiation. The small- E! situation helps tell the story of absorption of solar uv radiation

    ozone in the stratosphere. Ozone, and also O2 and N2, have absorption peaks in the uv so the uv energy is

    resonantly absorbed, a collision occurs, and the energy is safely thermalized. Blue light, on the other hand,

    is farther removed from the uv resonant peak, so the large E! case holds, and it is scattered. Of course,

    green and red light is also scattered, but they are farther removed from the resonant peak so they arent

    absorbed and scattered as much as blue. In fact, the absorption is proportional to the inversefourth power

    of the wavelength, so blue is scattered about 16 times as much as red. Thats why we have a b lue sky!

    6. A FOURIER TRANSFORM INTERPRETATION OF )t,(Pm ! .The graph of )t,(Pm ! for fixed time

    is precisely a graph of the Fourier transform of a cosine segment as shown in Figure 8.6.1. This is not a

    coincidence. Assuming that a harmonic cosine perturbation in the form of an electromagnetic wave is

    turned on at t= 0, then at time tit looks like the cosine segment shown in the figure. (The question mark

    just indicates we dont know what the future holds past time tthe perturbation might continue or it might

    abruptly end.) If

    we assume this segment is of the form tcoso

    ! for a particular value ofo

    ! , then the Fourier transform of

    the segment will be centered ato

    ! .

    Figure 8.6.2 shows the Fourier transform of this segment. A comparison of this figure with Figure

    8.5.1 shows that the probability of a transition, now given by the s trength of the Fourier frequency atmk

    ! ,

    is exactly what Figure 8.5.1 shows it to be ato

    ! . This is because the equation forPm(", t) is symmetric

    in mko !"! , so it depends only the distance from the center, i.e., on mko !"! . And as before, for

    longer times, the longer the cosine segment will be and the more tightly its Fourier transform will be

    located near the central frequency (in this case,o

    ! ).

    t

    FIGURE 8.6.1

    ),( tH r!

    ?

    time

  • 8/6/2019 Quantum Mantras

    17/24

    17

    7. FERMIS GOLDEN RULE. Consider the absorption of light by atoms. Say they are all hydrogen

    atoms. We are prone to think that the energy levels of one hydrogen atom are the same as for another

    hydrogen atom, but this is true only at absolute zero, i.e., for motionless atoms, with lots of space betweenthe atoms. If the atoms are warm, i.e., at any temperature above absolute zero, the levels and the

    frequencies corresponding to differences between energy levels are Doppler-shifted because of the thermal

    motion. Broadening of energy levels can be caused by a number of other factors, such as crowding the

    atoms together (pressure broadening) or just the natural lifetime of the state which, by the uncertainty

    principle, causes a smearing out of the energy level. Fermis "Golden Rule" leads to a simple result for the

    absorption in cases where the broadening is sufficiently large.

    To derive the Golden Rule, assume to begin with that we have only a few energy levels, say 4,

    labeled m1, m2, m3, and m4, separated but not too much so, as shown in Figure 8.7.1 (a). We can think of

    these as really distinct levels or just as an approximation to the result of broadening a single level. Also

    shown is the initial state,Ek, and the energy of incident radiation, mk!h . This harmonic perturbation can

    induce transitions fromEk to one of the four higher states.

    Figure 8.7.1(b) shows the probability curves, )t,(Pim! , for the four levels as a function of! ,

    with their bases placed for comparison at the energy levels shown in 8.7.1(a). (Note that the curves have

    different heights, corresponding to different matrix elements for )(H kmi r ). The peak of )t,(P im ! is at

    "m

    i

    =

    Emi# E

    k

    h. The incoming radiation of frequency

    mk! has probabilities ,a,a,a

    2

    3

    2

    2

    2

    1and

    2

    4a , of

    inducing a transition to each of the four excited states. Just looking at the curves, it is easy to see that theprobability that a transition to state 3 has the greatest probability. This is to be expected, since 8.7.1(a)

    shows Em3 to be closest to mk!h . The probability that a transition will occur to any of the four excited

    states is just the sum of the separate probabilities, i.e.,2

    4

    2

    3

    2

    2

    2

    1aaaa +++ , with a1, a2, and a4 very

    small compared to a3.

    2

    22

    4h

    tHmk

    ),( tPm

    !

    != resonant

    absorption

    frequencymk

    !

    Area under entire

    curve =2

    2

    2h

    tHmk!

    Probability of absorption as a function of resonant absorption

    frequency for fixed time and fixed perturbing frequency

    o!

    "o# 2$ / t

    "o+ 2# / t

    FIGURE 8.6.2

  • 8/6/2019 Quantum Mantras

    18/24

    18

    Now suppose all of the levels have the same values of km iH between states kand mi (i = 1,2,3,4).

    Then all of the four probability curves are identical. As mentioned in the previous item, the probability of

    radiation of frequencymk! being absorbed depends only on the difference

    mimk!"! , so we can replace

    the four probability curves by one centered onkm! (and for convenience, placed right on the ! axis). This

    is shown in Figure 8.7.1(c).

    In practice, physical effects like Doppler broadening produce not just four or ten or twenty levels,

    but a virtual continuum of energy levels, which we initially assume for convenience to be very large in

    number but not infinite. This is depicted in Figure 8.7.2 (a), with the density, ! , of the smear of states

    represented by the darkness of the shading. If there are dNstates in the frequency interval [ ]!+!! d, ,

    then )(!" is defined by the relationship!

    =!"d

    dN)( , i.e., the number of states in the interval is just

    !!"= d)(dN . If we reasonably assume that the probability curves of all the states making up the

    continuum are equal, then as before we can determine the probability of a transition by looking at the

    overlap of the continuum with one single absorption curve centered onmk! . This is shown in Figure

    8.7.2(b) where, in a small interval aroundmk!h , )(!" is approximately constant. Both )t,(Pm ! and the

    density of the states, )(!" , are plotted on the vertical axis.

    The approximation which makes up Fermis Golden Rule consists in noting, first, that the

    separation between adjacent energy levels, shown much expanded in Figure 8.7.1 above, is very small

    compared to the width of the absorption curve, as suggested in Figure 8.7.2. And second, the density of

    states doesnt change very much across the range !" . So approximately, the number of states absorbing

    significant amounts of energy is approximately N)( mk !"#!#$ . The probability of a transition to any of

    the N! states is simply the sum of the separate probabilities:

    Em

    1

    Em2

    Em

    3

    Em

    4

    Ek

    E

    mk!h

    !

    E, Pm

    (", t)

    1m!

    2m!

    3m!

    4m! mk

    !

    FIGURE 8.7.1

    !

    1m!

    2m!

    3m!

    4m!

    E, Pm(",t)

    (a)

    (c)(b)

    mk!

  • 8/6/2019 Quantum Mantras

    19/24

    19

    Prob = ai

    2"

    i=1

    #N

    $ ( ai2(1)

    i=1

    #N

    $ )

    The reason for writing the number (1) in each term of the sum is as follows. Let the separation between

    adjacent states i! and 1+!i be i!" , which for a virtual continuum of states will be very small. Then the

    density of states ati! is

    i

    ii )(!"

    =!#1

    since there is just one state in this frequency interval, i.e.,

    1=!"!# ii )( . Now we can substitute ii )( !"!# for the number 1, and going to the limit of a continuum

    of states, and taking the different )( i!" all to be approximately )( mk!" , we get:

    ! """#$"% ! ""#&"'"#% ==(

    ()

    '

    =

    (

    ()

    '

    =

    d)t,(P)(d)(a)()(a)(a)t(obPr mkN

    imkiii

    N

    ii

    1

    22

    1

    2

    1 .

    But the last integral is just2

    2

    2h

    tHkm

    !

    , where2

    km

    H is the matrix element between state k and a typical state

    with an energy in the neighborhood ofkm!h . Therefore, we get:

    Prob(t) ="H

    km

    2

    t

    2h2

    #($mk).

    More usually, were interested in the transition rate," which is the transition probability per unit time:

    t

    )t(obPrw = . SinceProb(t) is proportional to t, w is constant:

    )(H

    wmk

    km!"

    #=

    2

    2

    2h

    . (FERMIS GOLDEN RULE)

    8. THE DIPOLE APPROXIMATION AND BROADBAND RADIATION. Consider an atom bathed

    in monochromatic electromagnetic radiation (i.e., radiation of single frequency, ! ) for which the

    wavelength is much greater than the linear size of the atom. (As mentioned at the beginning of this review,

    no attempt will be made to treat the radiation field quantum mechanically. A "semi-classical" treatment

    will be sufficient.) Then at any instant of time the electric field, which we assume to be varying

    Ek

    km!h

    E

    Em

    (a)

    FIGURE 8.7.2

    k!

    !" ),t,(mP

    !

    !" )(!"

    )( k!"

    (b)

    P(", t)

  • 8/6/2019 Quantum Mantras

    20/24

    20

    harmonically, is constant over the atom. Therefore, "H = #er $Eocos(%t) , so rEr !"= oe)(H where r is

    the coordinate of the electron (taking the origin conveniently to be at the nucleus of the atom, which is so

    massive we can neglect its motion). For most electromagnetic radiations of interest, e.g., anything longer

    than hard ultraviolet, the dipole approximation is excellent. It fails only when we get into the X-ray region.

    An electromagnetic wave, by its very name, includes a magnetic component,B(t), too. The

    perturbation )tcos(eH o !"#=$ Er ignores the magnetic part of the wave because in the dipole

    approximation, it is mainly the electric field that affects the motion of the electron around the nucleus. This

    can be appreciated by noting that, in this approximation, the field changes only slowly compared to the

    time it takes the electron to react to the field, but by Faraday's law, the magnetic effect, which depends on

    the (slow) rate of change ofB, is small. In essence, then, at any instant of time, the atom is subject to a

    quasi-static electric field which produces the change in potential energy given by H! .

    The equation for )t,(Pm ! involves the matrix element mkH . In the present case, this is equal to

    kme rEo!" . Thenthe equation for )t,(P

    m! is:

    2

    2

    oz

    22

    2

    2

    2

    24

    EEE

    !!"

    #

    $$%

    & '('

    !!"

    #

    $$%

    & '('

    ++

    ='t)(

    t)(sin

    kzyxmte)t,(P

    mk

    mk

    oyox

    m

    h

    .

    For plane-polarized light, it is convenient to take the direction of polarization along the z axis (Eo = Eo z )

    (and the dirction of propagation perpendicular toz), so we have only one integral to do in the matrix

    element for )t,(Pm ! .

    Suppose again that the atom is bathed in monochromatic radiation but with random (isotropic)

    polarizations and directions of propagation. Then Eox

    2= E

    oy

    2= E

    oz

    2= E

    o

    2/3 . Therefore, the matrix

    element is justE

    o

    2

    3m r k

    2

    , so:

    2

    22

    o

    22

    2

    2

    2

    212

    E

    !

    !

    "

    #

    $

    $

    %

    & '('

    !!"

    #

    $$%

    & '('

    ='t)(

    t)(sin

    kmte)t,(P

    mk

    mk

    m

    h

    r

    .

    Our result for )t,(Pm ! is not realized in practice because radiation, even from a laser, is never

    strictly monochromatic. Instead, we must integrate )t,(Pm ! over" to obtain a frequency-independent

    transition probability Pm

    (t) . As a first step towards obtaining Pm

    (t) , Eo2

    can be expressed in terms of the

    energy density of the wave, )(u ! . For a single-frequency plane wave, u is the sum of the electric and

  • 8/6/2019 Quantum Mantras

    21/24

    21

    magnetic energy densities:o

    oBE

    u

    +!

    =22

    22

    . But for a plane wave,c

    EB = , where c = speed of light.

    Also, the speed of light is given by

    oo

    c

    !

    =1

    . Therefore, )t(cosEEu ooo !"="=222

    . Averaged over a

    cycle of time, we get the average energy density2

    2

    ooEu

    !

    = . Solving for Eo2

    and substituting into the

    previous equation for )t,(Pm ! gives:

    2

    2

    2

    222

    2

    2

    6

    !!"

    #

    $$%

    & '('

    !!"

    #

    $$%

    & '('

    )='

    t)(

    t)(sin

    kmtue)t,(P

    mk

    mk

    o

    mh

    r

    .

    Next, turning to the problem at hand of radiation covering a continuous range of frequencies, we

    must suppose we have a differential of probability, )t,(dPm ! , proportional to the differential of energy in

    the frequency interval !+!! d, . In other words, )t,(dPm ! is the probability that radiation in that

    frequency interval induces a transition. The differential of energy can be written as du(") = u1(")d",

    which defines u1(") as the radiation density (per unit volume per unit frequency interval). We can get the

    transition probability from the previous equation by replacing u by u1(")d" and integrating over! . (In

    probability terms, we want the probability of a transition due to frequencies in the range [ ]!+!! d, , plus

    the probability due to frequencies in the range [ ]!+!!+! d,d 2 , etc.) We assume the radiation density

    covers a wide range of frequencies compared to the region over which )t,(Pm ! is appreciable, as shown in

    Figure 8.8.1. When the integration is carried, out, most of the contribution comes from the narrow region

    aroundmk! , as suggested

    by the figure. In this narrow region, u1(") is practically constant, of magnitude u

    1("

    mk) , which can be

    taken out of the integral. Taking the limits of the integral with respect to ! from 0 to +! is equivalent to

    taking the limits of the relevant variable in the integrand, tmk !"

    #$%

    & '('

    2, from !"+! to , since "

    mkis

    !

    mk!

    !"

    u1(") FIGURE 8.8.1

  • 8/6/2019 Quantum Mantras

    22/24

    22

    essentially infinite in the narrow region over which )t,(Pm ! is appreciable. The result of integrating

    sin2("

    mk

    #")t

    2

    $%&&

    '())

    ("mk

    #")t

    2

    $%&&

    '())2

    ist

    !2. Therefore, we end up with:

    Pm(t) =

    "e2u1(#mk) m r k

    2

    t

    3$oh

    2= B

    mku1(#mk)t,

    which defines the constant Bmk

    . This constant will be met again in section 11. As with Fermi's Golden

    Rule, this leads to a transition rate, w, which is constant;

    w =Pm(t)

    t= B

    mku1("mk) .

    Note that Davies chooses to express w in terms of the radiant intensity I(") # u1(")c , where c is the

    speed of light.

    The foregoing result is really another form of Fermi's Golden Rule, which we should now

    understand to apply if narrow-band radiation causes a transition to a continuum of states or if broad-band

    radiation, essentially a continuum of radiation states, causes a transition to a narrow s tate.

    9. SELECTION RULES FOR THE HYDROGEN ATOM. As the equations for )t,(Pm ! and

    )t(Pm show, they depend on the matrix element22

    k)(HmHmk r! . For the hydrogen atom, m and k

    each consist of four quantum numbers, n, l, m, ands, so we could write

    Hmk

    2" nml mmmsm H(r) nkl kmksk

    2, but with this understood, the simpler m H(r) k

    2notation can

    be used. Forgetting spin, this matrix element, which depends only on space variables, is very important

    because it is zero for the dominant dipole interaction rEr !"= oe)(H except for certain values ofm and k.

    For example, it can be shown that transitions for any system with a spherically symmetric potential (e.g.,

    the H atom or hydrogen-like atoms) must obey the following selection rules:

    "l = 1

    "m = 0 for z polarization

    "m = 1 for polarization in the xy plane

    As an application, consider the H-atom states 012013 ,,and,, . Can a transition occur between these

    states? No, because the selection rule on l! is violated (i.e., it didnt change at all), even though the

    selection rule on m is satisfied. What about a transition from 003 ,, to 112 !,, ? This can occur for

  • 8/6/2019 Quantum Mantras

    23/24

    23

    radiation polarized in thexy plane because the conditions on both l! and m are satisfied (m = -1 and

    l! = +1).

    10. WHERE IS THE LORENTZIAN? The absorption curves so far are similar to the famous,

    ubiquitous Lorentzian that appears in mechanics, optics, and nuclear radiation (in the Mssbauer effect),

    but they have the oscillating wings which the Lorentzian doesn't have. This is because all of our resultshave been for a truncated sinusoidal radiation, i.e., radiation that has a pure sinusoidal form for a finite

    period of time. A pure Lorentzian does appear for a perturbation of the form "H = #erEocos($t)e

    #%t,

    i.e., an exponentially dampedsinusoidal electric field. (You might remember, we got a Lorentzian shape in

    Davies, Problem 5, Chapter 1.) We expect such an exponentially damped classical field to result from an

    atomic transition. The Lorentzian is also appropriate for the inverse case of radiation being absorbed by a

    gas.

    11. SPONTANEOUS EMISSIONEINSTEIN'S A AND B COEFFICIENTS.

    In 1917, Einstein developed an argument based on statistical mechanics and thermal equilibrium

    which related the spontaneous transition rate to the absorption rate Bmk

    , which was calculated above in

    section 8. His great achievement was to do this before the development of quantum mechanics almost 10

    years later. Consider a collection of identical atoms in an excited state n inside a cavity in which there is

    black body radiation at temperature T. For simplicity, it is assumed that none of the levels are degenerate.

    As Planck showed in 1900, the energy density of the radiation as a function of frequency is:

    u1(") =

    h"3

    #2c

    3

    1

    eh" / kT

    $1,

    where kas usual is Boltzmann's constat. As in Section 8, u1(")d" = radiation energy per volume in the

    frequency range ",

    "+ d"[ ] . Let Amn be the probability per unit time that the atom decaysspontaneously from state n to a lower-energy state m, let B

    mku1(") be (as in Section 8) the absorption

    probability per unit time for a transition from state m to state n, and let Bkmu1(") be the stimulated

    emission probability per unit time for a transition from state n to state m. In thermal equilibrium, let Nm

    =

    the number of atoms in the lower state and Nn

    = the number atoms in the excited state. Then thermal

    equilibrium requires that the number of transitions from m to n equals the number of transitions from n to

    m:

    NmB

    mnu1(") = Nn Anm + Bnmu1(")[ ] .

    Now we can solve for u1("):

    u1(") =A

    nm

    (Nm/N

    n)B

    mn# B

    nm

    .

    But the Boltzmann equation gives:

    Nn

    Nm

    =

    e"En / kT

    e"Em / kT

    = e"h#

    mn/ kT

    .

  • 8/6/2019 Quantum Mantras

    24/24

    24

    Therefore,

    u1(") =A

    nm

    Bmneh" / kT

    # Bnm

    .

    Comparison with Planck's law shows that:

    Bnm = Bmn ,

    Anm=

    h"3

    #2c3Bmn

    Using the previous value of Bmn

    for a dipole transition, Anm

    is:

    Anm=

    e2"

    3

    3#c3$3h

    mr n2

    .

    Einstein provided a clever and accessible way to determine the spontaneous transition rate. A

    strictly quantum mechanical treatment explains spontaneous transition in terms of the quantized

    electromagnetic field. In this picture, the EM field is viewed as an infinite number of harmonic oscillators,

    each oscillator corresponding to a radiation frequency. The amplitude of a harmonic oscillator mode with a

    particular frequency is represented by the number of photons of that frequency. The elementary quantum

    mechanical theory of the harmonic oscillator shows that the ground state does not have zero energy but

    rather has an energy of h" /2 . This "zero point energy" is always there for every mode of the quantized

    EM field, and this energy interacts with an electron in an excited state to "rattle" it out of that state, with the

    consequence that it falls into a lower-energy state and in the process (a la Bohrs or iginal idea)

    spontaneously emits radiation.