set8.pdf

  • Upload
    cemnuy

  • View
    212

  • Download
    0

Embed Size (px)

Citation preview

  • 8/22/2019 set8.pdf

    1/16

    Lecture 22

    Higher-dimensional PDEs

    Relevant section of text: Chapter 7

    We now examine some PDEs in higher dimensions, i.e., R2 and R3. In general, the heat and wave

    equations in higher dimensions are given by

    u

    t= k2u (heat equation), (1)

    2u

    t2= c22u (wave equation). (2)

    The higher-dimensional form of the heat equation comes from the application of the conservation of

    thermal energy to regions in Rn, along with Fouriers law of heat conduction and the Divergence

    Theorem. A derivation of the wave equation in higher dimensions would require a detailed discussion

    of stress and strain in solids, which we forego.

    In what follows, we examine the two-dimensional wave equation, since it leads to some interesting

    and quite visualizable solutions.

    Vibrating rectangular, homogeneous membrane

    The vibration of a thin membrane is, for small displacements, modelled well by the two-dimensional

    wave equation 2u

    t2= c2

    2u

    x2+

    2u

    y2

    . (3)

    Here u(x,y,t) represents the displacement of the membrane from its equilibrium position, which is

    assumed to be the surface u(x, y) = 0.

    The simplest problem is that of a rectangular membrane, represented by the region

    0 x L, 0 y H, (4)

    which we shall simply denote as D R2. Sometimes, we shall also write the above region as [0, L] [0, H].

    151

  • 8/22/2019 set8.pdf

    2/16

    Well assume that the membrane is clamped, leading to the following boundary conditions on

    u(x,y,t):

    u(x, 0, t) = 0,

    u(L,y,t) = 0,

    u(x,H,t) = 0,

    u(0, y , t) = 0. (5)

    Well also need initial conditions:

    u(x,y, 0) = (x, y) (initial displacement)

    u

    t(x,y, 0) = (x, y) (initial velocity). (6)

    The above boundary conditions are homogeneous if two solutions u1(x,y,t) and u2(x,y,t) satisfy

    the BCs, then a linear combination also satisfies the BCs. This suggest that we can try the separation

    of variables method to produce an infinite set of solutions.

    Well first try a separation of spatial and time-dependent solutions:

    u(x,y,t) = (x, y)h(t). (7)

    Substitution into Eq. (3) yields

    h = c2xxh + c2yyh. (8)

    We may separate solutions as follows:

    h

    c2h=

    xx + yy

    = , > 0, (9)

    where we have introduced the negative separation constant , based upon past results. The time-dependent equation for h(t),

    h + c2h = 0, (10)

    has oscillatory solutions, which is desirable. (We have skipped a longer analysis of the problem which

    would have involved an examination of all cases, i.e., (i) = 0, (ii) < 0 and (iii) > 0.)

    The spatial function satisfies the equation

    2

    x2+

    2

    y2= , (11)

    152

  • 8/22/2019 set8.pdf

    3/16

    or, simply,

    2 + = 0, (12)

    which is a two-dimensional regular Sturm-Liouville eigenvalue equation, with boundary conditions,

    (0, y) = 0, (x, 0) = 0, (L, y) = 0, (x, H) = 0. (13)

    These are homogeneous BCs, suggesting that we can attempt a separation-of-variables-type solu-

    tion for (x, y), i.e.,

    (x, y) = f(x)g(y). (14)

    (Note that this implies that our original function u(x,y ,t) has been separated into three functions,

    u(x,y,t) = f(x)g(y)h(t). (15)

    We could have started with this form, but it probably would not have been clear why it should work.)

    Substitution of (14) into (11) yields

    fg + f g = fg. (16)

    In an attempt to separate variables, we now divide by the term f g to give

    f

    f+

    g

    g= . (17)

    We have separated variables: the first term involves x and the second term involves y, but this doesnt

    allow us to solve for either f or g. What we have to do is to perform another separation a separationof x and y terms as follows,

    f

    f= g

    g. (18)

    Now we have all x-dependent parts on the left and y-dependent parts on the right. We now introduce

    a second separation constant:f

    f= g

    g= , > 0, (19)

    which yields the following equations for f and g:

    d2fdx2

    + f = 0, f(0) = f(L) = 0,

    d2g

    dy2+ ( )g = 0, g(0) = g(H) = 0. (20)

    This is a system of two Sturm-Liouville eigenvalue problems:

    153

  • 8/22/2019 set8.pdf

    4/16

    1. An equation for f with eigenvalue ,

    2. An equation for g with eigenvalue .

    In fact, the system is coupled because of the appearance of in both eigenvalues. It might seem

    somewhat confusing that we have to deal with another separation constant. However, all will be fine

    if we proceed systematically.

    We can solve the eigenvalue equation for f very easily weve done it many times! The solutions

    are

    fn(x) = sinnx

    L

    , n =

    nL

    2. (21)

    For each value of n, n = 1, 2, , the equation for g becomes the eigenvalue problem

    d2g

    dy2+ ( n)g = 0, g(0) = g(H) = 0. (22)

    For simplicity, well define

    n = n, (23)

    so that the above equation becomes

    d2g

    dy2+ ng = 0, g(0) = g(H) = 0. (24)

    For each n, n = 1, 2, , the solutions are known:

    gnm(y) = sinmy

    H

    , (25)

    with eigenvalues

    nm =m

    H

    2, m = 1, 2, . (26)

    Note that we had to introduce another index, m, so that the eigenvalue is now doubly indexed.

    We now resubstitute for nm in terms of :

    n =m

    H

    2, (27)

    so that the eigenvalues of the 2D Sturm-Liouville equation (11) become

    nm =n

    L

    2+m

    H

    2. (28)

    154

  • 8/22/2019 set8.pdf

    5/16

    The associated eigenfunctions will be given by

    nm(x, y) = fn(x)gnm(y)

    = sinnx

    L

    sin

    myH

    . (29)

    We now go back to the time-dependent equation for h(t):

    h + c2h = 0. (30)

    The functions hnm(t) associated with the spatial functions nm(x, y) will satisfy the equation

    hnm + nmc2hnm(t) = 0. (31)

    The general solution is given by

    hnm(t) = anm cos(nmct) + bnm sin(nmct). (32)The product solutions, the normal modes of vibration of the membrane, are given by

    unm(x,y,t) = nm(x, y)hnm(t)

    = sinnx

    L

    sin

    myH

    anm cos(

    nmct) + bnm sin(

    nmct)

    , n, m = 1, 2, .(33)

    Each mode is composed of a spatial profile nm(x) that is modulated in time with frequency

    nm = cmn = c n

    L2

    + m

    H2

    1/2

    . (34)

    We now examine the first few modes nice pictures of these modes are to be found in the textbook.

    1. n = m = 1. This is the mode with the lowest frequency,

    11 = c

    L

    2+

    H

    21/2. (35)

    The spatial profile,

    11(x, y) = sinxL siny

    H , (36)

    has no zeros in the interior of the rectangle. As a result, this mode has no nodes the entire

    membrane executes a uniform up-and-down motion.

    155

  • 8/22/2019 set8.pdf

    6/16

    2. n = 2, m = 1. The spatial profile

    21(x, y) = sin

    2x

    L

    sin

    yH

    , (37)

    has a zero at x = L/2, which implies that the vibrational mode has a node along the line

    x = L/2, 0 y H. As a result, the membrane is divided into two regions, the displacements

    of which will lie on opposite sides of the xy-plane. The frequency of this mode is

    21 = c

    2

    L

    2+

    H

    21/2. (38)

    3. n = 1, m = 2. The spatial profile

    12(x, y) = sinx

    L

    sin

    2y

    H

    , (39)

    has a zero at y = H/2, which implies that the vibrational mode has a node along the line

    y = H/2, 0 x L. As a result, the membrane is divided into two regions, the displacementsof which will lie on opposite sides of the xy-plane. The frequency of this mode is

    12 = c

    L

    2+

    2

    H

    21/2. (40)

    Note that if L < H, then 21 > 12 and vice versa. If L = H, then 21 = 12 and the two

    modes have the same frequency of vibration. In this case, the membrane is a square, and the

    two modes are identical, up to a rotation of /2 around the origin.

    4. n = 2, m = 2. As in the previous case, the spatial profile

    22(x, y) = sin

    2x

    L

    sin

    2y

    H

    , (41)

    has a zero at y = H/2, which implies that the vibrational mode has a node along the line

    y = H/2, 0 x L. But it also has a zero at x = L/2, which implies that the mode has anode along this line. As a result, the membrane is divided into four regions, the displacements

    of which will lie on opposite sides of the xy-plane. The frequency of this mode is

    12 = c

    2

    L

    2+

    2

    H

    21/2. (42)

    Note that the frequency of this mode, 22, is greater than the frequencies 12 and 21.

    156

  • 8/22/2019 set8.pdf

    7/16

    Lecture 23

    Vibrating rectangular membrane (conclusion)

    In the previous lecture, using separation of variables, we obtained the normal modes of vibration for

    this problem,

    unm(x,y,t) = nm(x, y)hnm(t)

    = sinnx

    L

    sin

    myH

    anm cos(

    nmct) + bnm sin(

    nmct)

    , n, m = 1, 2, .(43)

    Each mode is composed of a spatial profile nm(x) that is modulated in time with frequency

    nm = c

    mn = c

    nL

    2+m

    H

    21/2. (44)

    Because of the homogeneous boundary conditions, the general solution to the vibrating membrane

    problem may be written as a superposition of these normal mode solutions,

    u(x,y,t) =n=1

    m=1

    nm(x, y)hnm(t)

    =

    n=1

    m=1

    sinnx

    L

    sin

    myH

    anm cos(

    nmct) + bnm sin(

    nmct)

    . (45)

    Imposition of the initial condition u(x,y, 0) = (x, y) implies that

    (x, y) =

    n=1

    m=1anm sin

    nx

    L sin

    my

    H =

    n=1

    m=1

    amnnm(x, y). (46)

    We now make use of the fact that the product basis nm(x, y) forms an orthogonal set over the

    region D = [0, L] [0, H], i.e.,

    Dn1,m1(x, y)n2,m2(x, y) dxdy =

    H0

    L0

    sinn1x

    L

    sin

    m1yH

    sin

    n2xL

    sin

    m2yH

    dxdy

    = L0

    sinn1xL

    sinn2xL

    dx H0

    sinm1yH

    sinm2yH

    dy=

    L2

    H2

    , (n1, m1) = (n2, m2)

    0, (n1, m1) = (n2, m2).(47)

    157

  • 8/22/2019 set8.pdf

    8/16

    We also state, without proof, that this product basis sometimes called a tensor product basis is

    complete in the function space L2[D], the set of functions f : D R (or C) that are square-integrableon D, i.e.,

    D

    |f(x, y)|2 dA < . (48)

    As such, we may multiply both sides of Eq. (46) by kl(x, y) for a k

    1 and l

    1, and integrate over

    D to give

    akl =4

    LH

    D

    (x, y)nm(x, y) dxdy. (49)

    Likewise, the initial velocity condition,

    u

    t(x,y, 0) = (x, y), (50)

    yields

    bkl =1

    c

    nm

    4

    LH D (x, y)nm(x, y) dxdy. (51)This concludes our discussion of the (clamped) rectangular vibrating membrane problem.

    The relevance of the vibrating membrane problem to quantum mechanics

    The 1D vibrating string and 2D rectangular vibrating membrane problems are quite relevant to quan-

    tum mechanics, as we now discuss briefly. In one-dimension, the time-dependent Schrodinger equation

    for a particle in a box is given by the eigenvalue/BVP,

    2

    2md2

    dx2 = E, (0) = (L) = 0. (52)

    The wavefunction (x) provides a description of a particle that is confined inside a potential We

    rewrite Eq. (52) asd2

    dx2 +

    2mE

    2 = 0, (0) = (L) = 0. (53)

    The solutions of this BVP are well known to us:

    n =2mE

    2=

    n22

    L2, n(x) = sin

    nx. (54)

    This implies the existence of a discrete set of energy eigenvalues,

    En =2

    2m

    nL

    2, n = 1, 2, . (55)

    158

  • 8/22/2019 set8.pdf

    9/16

    In two dimensions, we consider a particle confined in the two-dimensional region [0, L] [0, H]which is surrounded by potential walls of infinite height. The time-dependent Schrodinger equation

    for the particle is2

    2m2 = E, (56)

    with the same boundary conditions as the clamped rectangular membrane. It may be rewritten as

    2 + 2mE2

    = 0. (57)

    We have derived the solutions earlier:

    nm =2mEnm

    2=n

    L

    2+m

    H

    2, nm(x, y) = sin

    nxL

    sin

    mH

    . (58)

    The energy eigenvalues of these wavefunctions are

    Enm =2

    2mn

    L2 + m

    H2 . (59)

    159

  • 8/22/2019 set8.pdf

    10/16

    Vibrating circular membrane

    Relevant section of text: 7.7

    We now consider the problem of a clamped, vibrating circular membrane, i.e., a drum. The

    vertical displacement satisfies the 2D wave equation,

    2ut2

    = c22u. (60)

    Since the drum is circular, it is convenient to use polar coordinates, i.e., u = u(r, ). We assume that

    the drum has radius a > 0. The clamping of the drum along the outer boundary implies the following

    boundary condition

    u(a,,t) = 0, /2 /2, t 0. (61)

    As was the case for Laplaces equation on a circular region, we can impose only one boundary condition

    at this time.

    In order to determine a unique solution, well need initial conditions, i.e., initial displacement and

    velocity:

    u(r,, 0) = (r, ),u

    t(r,, 0) = (r, ). (62)

    Since the boundary condition is homogeneous, well try separation of variables, i.e.,

    u(r,,t) = (r, )h(t). (63)

    Substitution into (60) yields

    h = c2(2)h. (64)

    Separating variables:h

    c2h=2

    = . (65)

    This yields the following ODEs:

    h + c2h = 0, (66)

    and

    2 + = 0. (67)

    Note that these ODEs have the same form as those for the rectangular vibrating membrane. If > 0,

    the solutions for h(t) are oscillatory, i.e.,

    h(t) = c1 cos(c

    t) + c2 sin(c

    t). (68)

    160

  • 8/22/2019 set8.pdf

    11/16

    We now examine the eigenvalue equation (67) for (r, ). In polar coordinates, the Laplacian operator

    has the form

    2 = 1r

    r

    r

    r

    +

    1

    r22

    2, (69)

    so that the eigenvalue equation assumes the form

    1

    r

    r

    r

    r

    +

    1

    r22

    2 + = 0. (70)

    We now assume a separation-of-variables solution for , i.e.,

    (r, ) = f(r)g(), (a, ) = 0. (71)

    Substitution into (70) yieldsg

    r

    d

    dr

    r

    df

    dr

    +

    f

    r2d2g

    d2+ f g = 0. (72)

    Now divide by f g and multiply by r2:

    r

    f

    d

    dr

    r

    df

    dr

    +

    1

    g

    d2g

    d2+ r2 = 0. (73)

    We havent quite separated variables, so we do so now:

    1g

    d2g

    d2=

    r

    f

    d

    dr

    r

    df

    dr

    + r2 = , (74)

    where we have introduced another separation constant, . This yields the two equations,

    d2g

    d2

    + g = 0, (75)

    and

    rd

    dr

    r

    df

    dr

    + (r2 )f = 0, f(a) = 0. (76)

    The equation for g() will have the following periodicity conditions,

    g() = g() dgd

    () = dgd

    (). (77)

    Periodic solutions to (75) exist for 0:

    gm() = cos(m), m 0 and sin(m), m 1. (78)

    The eigenvalues are given by m = m2.

    161

  • 8/22/2019 set8.pdf

    12/16

    Lecture 24

    Vibrating circular membrane (contd)

    Relevant section of text: 7.7

    We now turn to the radial equation (76) for f(r). If we divide by r, the equation is cast into

    Sturm-Liouville eigenvalue form:

    d

    dr

    r

    df

    dr

    m

    2

    rf + rf = 0, 0 < r a. (79)

    Here, p(r) = r, q(r) = m2/r and (r) = r. Note that r = 0 is a singular point because q(r) isundefined there. As such this Sturm-Liouville eigenvalue problem is not regular. Nevertheless, we

    claim that solutions exist for this problem as is the case for regular S-L problems, that is,

    1. A set of eigenvalues nm, n = 1, 2, 3, , m = 0, 1, 2, ,

    2. A set of corresponding eigenfunctions fnm: For each fixed m, the functions fnm(r) are orthogonal

    with respect to the weight function (r) = r, i.e.,a0

    fnm(r)fnm(r) rdr = 0, for n = n. (80)

    To produce these solutions, we make the change of variable z =

    r in Eq. (79). The result is

    (Exercise):

    z2d2F

    dz2+ z

    dF

    dz+ (z2 m2)F = 0, m = 0, 1, 2, . (81)

    where F(z) = f(z/). (In the lecture, I wrote, erroneously, F(z) = f(r). Sorry about that!) Notethat this removes the parameter from the ODE of course, it is now hidden in the independent

    variable z.

    You may recognize Eq. (81) it is Bessels equation, which is studied in AMATH 351. Very briefly,

    z = 0 is a regular singular point. The standard practice is to assume a Frobenius series solution of

    the form

    zpn=0

    anzn =

    n=0

    an+pzn+p, a0 = 0, (82)

    where p is to b e determined. One substitutes the Frobenius series into Eq. (81), collects like termsin zk, and demands that the series begins with a term that starts with a nonzero term involving a0.

    The result is a quadratic equation in p. For Eq. (81), the indicial equation for p is

    p2 m2 = 0 p = m. (83)

    162

  • 8/22/2019 set8.pdf

    13/16

    For m = 0, the two distinct values of p, i.e., p1 = m and p2 = m, yield two linearly independentseries solutions,

    Jm(z) = zm[am,0 + am,1z + ],

    Ym(z) = Jm(z) = zm[bm,0 + bm,1z + ]. (84)

    The actual values of the coefficients am,i and bm,i are not important for our discussion.

    For m = 0, we have p1 = p2 = 0, so only one solution is produced by the series solution method:

    J0(z) = 1 + a0,1z + . (85)

    A second, linearly independent solution may be produced by the variation of parameters applied to

    J0(z). The result is the following function

    Y0(z) =2

    ln z +

    . (86)

    It is also important to mention that for all m 0, the functions Jm(z) are oscillatory for z > 0.(Another important result that is shown in AMATH 351 is that as z , the spacing of consecutivezeros of Jm(z) approaches .)

    For m = 0, 1, 2, , the general solution to the Bessel equation (81) will have the form,

    Fm(z) = c1Jm(z) + c2Ym(z). (87)

    We then transform back to the r variable to produce the general solution f(r) to Eq. (76),

    fm(r) = c1Jm(

    r) + c2Ym(

    r). (88)

    We now apply these solutions to the vibrating circular membrane problem. Recall that the

    membrane being clamped at r = a, i.e., f(a) = 0, gave us one boundary condition. Before considering

    that boundary condition, we consider another condition which deals with the singular nature of the

    point r = 0: It is the condition

    f must be finite at r = 0, i.e., |f(0)| < . (89)

    Recall that we imposed such a condition on solutions to Laplaces equation on the circular disk.

    The point r = 0 corresponds to the point z = 0. From a look at the behaviour solutions Jm(z)

    above, we note that, in all cases,

    |Ym(z)| as z 0. (90)

    163

  • 8/22/2019 set8.pdf

    14/16

    This implies that these functions must be excluded from the general solution fm(r) in Eq. (88), i.e.,

    c2 = 0 so that

    f(r) = c1Jm(

    r). (91)

    We now return to the boundary condition f(a) = 0. It becomes the condition,

    Jm(a) = 0. (92)

    Given that the functions Jm(z) are oscillatory for z > 0, it follows that the argument

    a is a zero of

    the function Jm(z). In other words, a = zmn > 0, (93)

    where zmn denotes the nth zero of the Bessel function Jm(z). The zeros of the Bessel functions are

    presented in standard mathematical tables. (Indeed, the results of this lecture give an idea of why

    these zeros are so important.) For example, approximate values of the first three zeros of J0(z) are

    z0,1 = 2.40482, z0,2 = 5.52007, z0,3 = 8.65372. (94)

    The graph of J0(x) for 0 x 20 is presented below. (Just for interests sake, the numerical valuesused to plot the graph were computed using the series expansion for J0(z).)

    -1

    -0.75

    -0.5

    -0.25

    0

    0.25

    0.5

    0.75

    1

    0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

    x

    Bessel function J_0(x)

    For each m 0, the above conditions produce an infinite sequence of eigenvalues mn given by

    mn =zmn

    a

    2, n = 1, 2, . (95)

    164

  • 8/22/2019 set8.pdf

    15/16

    The resulting radial functions,

    fmn(r) = Jm(

    mnr), (96)

    have n positive zeros, with the nth zero located at r = a.

    Note: Before going on to construct the full solutions to the clamped vibrating mem-

    brane problem, let us stop for a moment and look again at the clamping condition (93)

    that produces the eigenvalues mn in (95). This condition is analogous to the boundary

    condition

    sin(

    L) = 0, (97)

    encountered in the 1D wave equation (clamped string, zero-endpoint conditions), which

    implied the result,

    L = n, n = 1, 2, , (98)

    yielding the eigenvalue condition,

    n =n

    L

    2, n = 1, 2, . (99)

    This was a relatively straightforward result, since the zeros of the sin function are so well

    known. Comparing Eqs. (99) with (95),

    1. the zeros zmn of Jm(z) correspond to the zeros zn = n of sin(z) and

    2. a corresponds to L.

    The solutions to the clamped vibrating membrane will then have the form

    umn(r,,t) = fmn(r)gm()hmn(t), m = 0, 1, 2, , (100)

    where

    fmn(r) = Jm(

    mnr),

    gm() = am cos m + bm sin m,hmn(t) = cmn cos(

    mnct) + dmn sin(

    mnct). (101)

    These solutions correpond to the normal modes of vibration of the circular membrane. We shall return

    to them in the next lecture.

    165

  • 8/22/2019 set8.pdf

    16/16

    The general solution to the vibration problem will be given by

    u(r,,t) =

    m=0

    n=1

    umn(r,,t). (102)

    If we assume that the membrane is initially at rest, i.e., u/t = 0, then the sin(

    mnct) terms

    vanish, so that the general solution can be expressed in the following form,

    u(r,,t) =

    m=0

    n=1

    AmnJm(

    mnr) cos(m)cos(

    mnct)

    +

    m=0

    n=1

    BmnJm(

    mnr)sin(m)cos(

    mnct). (103)

    166