Class Notes
These notes are NOT intended to be a complete record of what has been covered in class. Students are responsible for all material discussed in class and in the assigned readings, not just what appears here. I will try (but not always succeed) to be brief here. I will focus on material not adequately discussed in the text.

Mon, Tues, Wed, Thurs, 12/9-12

- Sorry I have no time to write this up in any detail. We covered (from the course plan page):
two spatially separated souce interference, sound interferometer
Michelson interferometer, combined double/single slit interference patterns
 circular aperture, N slits, diffraction gratings, resolving power
diffraction grating, spectra, electron (matter wave) diffraction
This material is all covered fairly nicely in the textbook, except  for the electron diffraction.

Thurs, 12/5 no class: snow
Wed, 12/4

- interference from two spatially separated sources: the phase difference is then determined by the path length difference. If the sources lie symmetrically on the y-axis, we get a simple formula for the amplutude and intensity on an x=const line far from the sources. For the intensity, I = 4 cos^2(D phi/2) I_0, where the phase difference D phi = 2 pi d sin(theta)/lambda, theta is the angle from the x-axis, lambda is the wavelength, and I_0 is the intensity at theta=0.

Tues, 12/3

- more on polarization: We had an extensive discussion about how the quarter wave plate works to produce circular polarization from linear polarization. One key idea is that we can choose to write, say, a vertical polarization as a superposition of two linear polarizations at 45 degrees tilted to the left and right, for which the index of refraction of the birefringent material is different. Then, because the material acts linearly on the electromagnetic input, we can analyze separately what happens to the two 45 degree poalrizations, and add the results. A good question was raised about the time delay: if the incoming wave has finite extent, i.e. is a wavepacket rather than an infinite wave train, then if it passes through enough birefringent material the two polarization wavepackets will separate spatially so they no longer overlap. That's true. But for light coming from atomic transitions the wavepackets are very long: since it takes more than a nanosecond for the atomic transition, the emitted wavepacket is more than one light-nanosecond long, i.e. 30 cm. (This is huge compared to the wavelength which is, say, 0.5 microns, so it contains 600,000 wavelengths.) Thus it would take quite a large thickness of birefringent material to separate the two component wavepackets.

- interference: we considered thin film interference, in three examples: Newton's rings formed with a convex lens touching a flat glass plate, an oil slick on water, and a soap film. In each case two reflected waves combine, coming from reflection off the top or bottom interface, with some relative phase. If they differ by an integer times 2pi they add constructively and give a maximum. If they differ by an integer plus a half times 2pi they cancel (interfere destructively) and produce a minimum. The relative phase is governed by two factors:
1) the phase change of each upon reflection, and
2) the extra path length of the wave that reflects off the bottom interface.
1) It turns out that there is a pi phase change upon reflection if the second medium has a higher index of refraction than the first. If both reflections involve a phase change or both do not, then the reflection produces no relative phase shift. If only one reflection changes phase there is a relative phase shift of pi. In the example of oil (n=1.2) on water (1.33) both reflections produce a phase change, while in the soap film the top reflection changes phase while the bottom doesn't. For the lens and plate it is the reverse.

2) Aside from the possible phase change upon reflection, the total phase delay acquired by the wave that reflects off the bottom surface is the number of wavelengths times 2pi. If the reflection is perpendicular to the film or gap of width d, the numer of wavelengths is 2d/lambda_n, where lambda_n is the wavelength in the medium with index of refraction n (if any), lambda_n = lambda/n.

Thus, the conditions for a maximum are, with m=0,1,2,3,...:
2d = m lambda_n                      if both or neither change phase upon reflection
2d = (m+1/2) lambda_n              if one changes phase upon reflection
For a minimum, just interchange these.

Mon, 12/2

- Circular polarization: (See the supplement and/or textbook.) Superposition of two in-phase linear polarizations yields a new linear polarization, the vector sum of the two. However superposition of two equal amplitude orthogonal linear polarizations 90 degrees out of phase yields right or left circular polarization.   If the amplitudes are not equal one gets elliptical polarization. Conversely, linear polarization can be expressed as a superposition of right and left circular polarization. According to the book's definition, the electric field vector at a fixed position rotates clockwise when viewed head on for a right circular polarization. Most treatments, especially involving quantum processes, use the opposite definition, since the angular momentum of right circular polarized light is then positive using the right hand rule in the direction of propagation.

- Optical activity can be understood as a result of the fact that there is a different index of refraction for right and left circular polarization. As explained in class, and in this week's homework, that results in a rotation of the linear polarization direction.

- Circular polarization can be produced by a circular current acting as a source for the wave, either macroscopically with an antenna, or microscopically with atomic transitions. Circular polarization can also be produced from linear polarization using a "quarter wave plate". This is a material with a different index of refraction for, say, vertical and horizontal polarizations. (This is called a birefringent material.) If light with 45 degrees linear polarization enters the material, the passage results in a phase shift of the vertical polarization component relative to the horizontal one. If the thickness of the plate is adjusted to produce a 90 degree relative phase shift (for a particular wavelength), the light emerges circularly polarized.


Wed, 11/27

- Liquid crystal display: works on the principle of crossed polarizers, with a mirror below if not back-lit. Between the polarizers is a liquid crystal that has chain molecules arranged in a corkscrew that rotates the polarization direction so as to pass the light through. When a voltage is placed across the plates in one spot, the molecules (being electrically polarized) line up in the direction of the elctric field and so eliminate the corkscrew effect that rotates the polarization of the light, thus producing a dark spot.

- Polarization by reflection: Light reflecting off an interface tends to be transmitted if it's polarization direction is not parallel to the interface. In fact, at a critical angle, called Brewster's angle or the polarizing angle, the reflected wave is 100 % polarized in the plane of the interface. This angle of incidence theta_B is such that the refracted and reflected waves are at right angles to each other, which means that
tan(theta_B) = n_2/n_1.  (Brewster's angle)

- Polarization by scattering: light scattering at a right angle is polarized in the direction that is perpendicular to the plane formed by the incident and scattered waves.

- Optical activity: rotation of the polarization plane by passing though a "chiral" ("handed") medium, such as sugar syrup. The biologically produced sugar molecules have a preferred handedness (dextrose rather than levulose), which somehow produces this rotation of the polarization direction. The amount of rotation for a given sample depends on the wavelength of the light, hence if the sugar syrup is placed between two crossed polarizers, white light incident on the first polarizer emerges from the second polarizer colored, due to the unequal mix of intensities for the different wavelengths passed by the second polarizer.
For demonstrations, see  http://www.physics.umd.edu/deptinfo/facilities/lecdem/services/demos/demosm8/m8-01.htm
http://www.physics.umd.edu/deptinfo/facilities/lecdem/services/demos/demosm8/m8-03.htm

Tues, 11/26

- Polarization: definition, demonstration with radio wave generator. Amplitude of wave received by diploe antenna making an angle theta with respect to the tranmitting antenna is E_0 cos(theta), where E_0 is the incident amplitude. The intensity received goes as teh square of the amplitude, hence
I = I_0 cos^2(theta).   (Malus' law)
A polarizing filter absorbs (or reflects) one linear polarization and transmits the orthogonal one. We demonstrated this with microwaves. The transmitted intensity follows the same relation just discussed, which is known as Malus' Law. Optical polarizing filters absorb one polarization and transmit the other. They effectively pass current in one direction, along chain molecules, but with resistance that absorbs that part of the wave. Optical sources of light are usually unpolarized to begin with, since they contain many accelerating charges with random orientations. 
Mon, 11/25:

- Mach cone: when source moves faster thant he wave speed in a medium, it emits a shock wave in the form of a cone with vertex at the instantaneous position, with an opening angle from the axis given by sin(theta) = v_w/v_s. The ratio v_s/v_w is called the Mach number. In the case of charged particles in a medium this is called Cerenkov radiation. The cone orientation and angle indicate the direction and speed of the source, so the effect is used in high energy charged particle detectors, usually with the medium being water.

- Doppler effect for EM radiation: the frequency shift can only depend on the relative motion, since there is no meaning to "motion relative to the medium". How can this be right? Something must be missing in the derivation of the previous results (see Mon, 11/18). The case of moving source and moving observer differ by terms of order (v_s/v_w)^2 and (v_o/v_w)^2. In the case of EM waves, we should take into account another effect of this same order: the time dilation effect. The period of the source as measured in the rest frame of the source T_s,s is not the same as the period of the source as measured in the rest frame of the observer T_s,o. Rather, T_s,o = T_s,s/Sqrt[1 - (v/c)^2], hence f_s,o = f_s,s Sqrt[1 - (v/c)^2]. Taking this into account, we derived the relativistic Doppler shift formula:

f_o,o = f_s,s Sqrt[(c+v)/(c-v)] = f_s,s Sqrt[(1+v/c)/(1-v/c)] =~ f_s,s (1 + v/c) for v/c << 1.

For small v/c this agrees with both the moving source and moving obserber results derived before. By the way, rather than using the time dilation formula to derive the Doppler formula, one can turn it around and use the fact that the Doppler formula must depend only on the relative motion of the source and observer to derive the time dilation formula.

- examples of EM Doppler: police and baseball Doppler radar; mass of Saturn by measuring the orbital speed of Saturn's rings using the difference between the blueshift of spectral lines from the approaching side and the redshift from the receding side; mass of giant black hole in the center of a galaxy by a similar method using absorption lines of orbiting gas.


Thurs, 11/21: went over exam 2.
Wed, 11/20: exam 2
Tues, 11/19: review for exam 2

Mon, 11/18

- Doppler effect:  The effect of motion on frequency and wavelength. Motion relative to WHAT? One can consider motion of the source or observer relative to the medium in which the wave propagates,  as well as motion of the source and observer relative to each other. For this class we restrict to motion along the wave propagation direction. We consider two settings: I. (non-relativistic) waves in a medium, and II. waves at the speed of light in vacuum. Under setting I we consider three cases. Let f_s be the source frequency and f_o be the observed frequency, and let v_w,s,o be the wave speed, the source velocity, and the observer velocity respectively. Then by examining the time between the observer's reception of successive wavefronts we deduced (for cases A & B) the following relations:

A. source at rest, observer moving:  f_o = f_s (1 + v_o/v_w)   [v_o>0 toward source]

B. source moving, observer at rest:  
f_o = f_s /(1 - v_s/v_w)   [v_s>0 toward observer]

C. both moving:                              
f_o = f_s (1 + v_o/v_w) /(1 - v_s/v_w) = f_s (v_w + v_o)/(v_w - v_s)

Note that A and B differ, but for velocities small compared to the wave speed they are approximately equal for equal relative velocities. Example: Consider sound wave in air at speed 340 m/s. If v_o or v_s are 10 m/s, then since 10/340 is around 0.03 we have in either case f_0 =~ 1.03 f_s. Compare this to a musical interval of a half step, which is 2^(1/12) = 1.06. So the note change is a quarter step. For speeds of 20 m/s the shift would be about a half step. Another example: Police siren and speeder. This is on the homework this week. We discussed at length the different cases, whether and how the frequency and wavelength are changed.

Thur, 11/14

- refraction in inhomogeneous media: examples: atmosphere of the earth, mirage over hot sand, Einstein's original calculation of bending of starlight by the Sun (he thought at that time that the speed of light was position dependent and larger farther from the earth. These can all be understood using wavefront propagation and Snell's law, but another point of view is Fermat's principle of least time. A nice example is the lifeguard and swimmer: to miniize the time from his bench to the swimmer the lifeguard runs farther on the beach where he runs faster, so that he can swim a shorter distance in the water where he moves more slowly. Another example: seismic waves in the earth, generated by an earthquake or (in the past) nuclear detonation tests. Both compressional and shear waves are generated. The former have the higher wave speed. The waves refract and bend back up to the surface of the earth, where they reflect and go back down again. Shear waves cannot penetrate the liquid core. From arrays of seismic measurements the properties of the interior of the earth can be reconstructed, as in tomography. The earth also has free oscillations, with normal modes, which depend on the structure of the earth and hence provide information about that structure.

- spectrum of white light from a prism: Due to dispersion. refractive index slightly different for different colors (i.e. different frequencies). Blue refracted more than red since it is closer to the resonant frequencies (UV) of the dipoles in the matter.  Explained how a rainbow works: when parallel light impinges on a spherical water droplet, the light emerges preferentially at the maximum angle of refraction, called the rainbow angle. This is demonstrated and explained nicely in this rainbow applet.


Wed, 11/13

- Huygen's principle (1678) is what we implicitly used when propagating the wavefronts to derive Snell's law. Huygen's principle can be mathematically derived from the wave equation, including the direction dependence of the amplitude of the secondary wavelets. The theory of this was completed in the 19th century, by Helmholtz (1859) and Kirchoff (1882). Note Huygen's principle does not apply for the wave equation in two dimensions, or any even number of dimensions. This is related to the fact, mentioned earlier, that the shape of a circular wave pulse is not preserved in 2d.

- Total internal relflection. A nice example not in the text: a light source under water in a lake sends out rays in all directions, but only those within a certain cone refract out into the air. Beyond a critical angle they reflect back underwater. Conversely, if you are looking around underwater, you see above you in a cone the full hemisphereabove the water, and outside the cone angle you see reflections of things such as fish under the water.

Tues, 11/12

-  Speed of light: Maxwell's equations imply EM waves travelling at speed c = 1/Sqrt[mu_0 epsilon_0] = 3 x 10^8 m/s = 3 x 10^5 km/s...relative to whom?? How could this be the correct speed independent of the motion of the observer? If I chase a sound wave, by travelling at 340 m/s, it will appear as a stationary pattern of compression and rarefaction, not  propagating at all. In the sound case, my motion can be reckoned relative to the air in which the sound propagates. Therefore one might think---and Maxwell did think---that EM waves propagate in a medium, called the ether, and the equations can only be applied in the rest frame of that medium. Either Maxwell's equations only apply in the rest frame of the ether, or there is something wrong with the apparently obvious conclusion that by chasing a light wave you could make it stand still.

Ever since Galileo it was clear that the laws of physics do not distinguish different states of relative unaccelerated motion, so that all inertial reference frames are equivalent. It would therefore be surprising if Maxwell's equations applied only in a preferred frame.  Moreover, Einstein pointed out that the phenomena of electrodynamics only depend on the relative motion of things, supporting the idea that Maxwell's equations apply in any inertial frame. For example he pointed to Faraday's law: whether the magnetic flux changes because a magnet is moving relative to a fixed loop of wire, or because the loop of wire moves in the opposite way relative to a fixed magnet, one observes the same emf in the loop. That is, the emf depends only on the relative motion. In the first case the emf is due to the induced electric field produced by the changing magnetic field, while in the second case it is due to the magnetic Lorentz force on the charges that are made to move along with the moving wire.  If Maxwell's equations do apply in any inertial frame however, the only way to explain the fact that the speed of light is the same for all observers is if there is something profoundly wrong with how we relate the time and space measurements of one observer to those of another. Thinking along these lines led Einstein to the theory of relativity.

- How fast is c? Light travel time to the sun: 150 Mkm/(3 x 10^5 km/s) = 500 s ~ 8 minutes. To moon: (3.8 x 10^5 km)/(3 X 10^5 km/s) ~ 1.3 s. Orbital speed of earth 29.8 km/s ~ 10^-4 c (not so slow, eh?). Electron in H atom: ~ alpha c = (e^2/hbar c) c ~ (1/137) c (zipping right along, but not quite so fast that relativistic corrections are very important for electrons in H atoms).                       
   
- Speed of light in matter: use dielectric constant and magnetic relative permeability (~ 1 in non ferromagnetic matter) in Maxwell's eqns, find wave speed (kappa_e kappa_m)^-1/2 c.  What's going on? Electric field stretches dipoles in matter, which "slows down" the wave. Really, the wave keeps going, but the stretched dipoles generate a secondary wave that is superposed with the incoming wave to produce a net wave that has a lower phase velocity. See Ch. 31 of vol. 1 of The Feynman Lectures, "The Origin of the Refractive Index", for a great explanation of this.  The response of the medium depends on the relation between the natural frequency of the dipoles in the medium and the wave frequency, so the dielectric constant is frequency dependent and hence so is the wave speed. In a simple model of the diples as harmonic oscillators with natural frequency w0 Feynman derives a formula for the index of refraction: n = c/v: n = 1+ Ne^2/(2eps0m(w0^2-w^2)). So you see that the closer to the resonance the greater the index of refraction differs from 1 by a greater amount. Below resonance the wave is slowed down. Above resonance it is speeded up! Examples of refractive indices...(see textbook here and for much of what follows).

- Reflection and transmission at a nonzero angle of incidence: law of reflection and Snell's law. Example of sound and light at an air/water interface.
                                                                                                                       
- Wavefronts and rays, derived Snell's law from a picture of propagating wavefronts.  
                                                       

Mon, 11/11

- Radiation pressure: EM waves have momentum equal to 1/c times their energy. One way to understand this is to think about a particle first. Nonrelativistically, p=mv. The correct relation between momentum and velocity in relativity is obtained by replacing m by E/c^2, thus p = (E/c^2)v. This E is the total energy, rest energy mc^2 plus  the kinetic energy. For v<<c, E is approximately just mc^2, so the relativistic momentum agrees with the nonrelativistic one. More generally, E = mc^2/Sqrt[1-(v^2/c^2)] = mc^2 + 1/2 mv^2 + 3/8 m v^4/c^2 + ... . Setting v =c is only possible if m = 0, i.e. something can go at the speed of light only if it has zero rest mass. In this case, the momentum is given by p = E/c. This suggests that EM waves have a momentum density equal to 1/c times their energy density.

An EM wave of intensity I thus has a momentum flux per unit time per unit area of I/c. If the radiation is absorbed, this rate of change of momentum produces a force per unit area, i.e. a radiation pressure, equal to I/c. If the radiation is reflected, the pressure is 2I/c.  We looked at the example of a 60W light bulb viewed from a distance of one meter: I = 60W/(4pi m^2) =~ 5 W/m^2, and pressure = I/c =~ 1.7 10^-8 N/m^2. Small, but not zero. The same effect holds the sun up! And can be used to produce fusion using high powered lasers.

How does the EM wave push on charges? Answer: the magnetic field must do it, since the electric field is transverse to the wave direction. This can be used to understand in another way the relation p = E/c. Consider an EM wave propagating in the z direction with electric field E_x in the x direction and magnetic field B_y=E_x/c in the y direction. The rate of change of momentum of a charge q in the z-direciton is then
dp_z/dt = F_z = q(vxB)_z = q v_x B_y = q v_x E_x /c =  F.v/c = (dE/dt)/c = d(E/c)/dt.
(The step where F.v appears is justified by the fact that only the electric force contributes, since the magnetic force is perpendicular to v.) Thus, the charge absorbs z component of momentum at 1/c times the rate at which it absorbs energy from the wave. This makes it plausible that the wave contains momentum in that proportion to energy.

- Electromagnetic spectrum: there is no length scale in Maxwell's equations, so waves of all wavelengths are "equivalent". Moreover, what is a low frequency to one observer is a short wavelength to another observer running towards the direction the wave is coming from. Since the wave speed is c in any inertial reference frame, this means that the wavelength is changed inversely to the frequency by such relative motion. As far as we know so far, there is no violation of relativity, and all wavelengths travel at exactly the same speed. I discussed the different parts of the EM spectrum, much as the book does. Gave examples of gamma ray bursts and cosmic microwave background radiation, and lengthening of wavelengths due to the expansion of the universe.


Thurs, 11/7

- Displacement current: recall we introduced the displacement current density j_d so that  
div(j+j_d) = 0                  (1)
would hold identically, as required by consistency of the Ampere-Maxwell law: curl B = mu_0 (j+j_d). This gave us j_d = epsilon_0 E,t. Going back to the integral form, this corresponds to \oint B.dl = mu_0 (i + i_d), where i = \int j.dA  is the current through a surface spanning the loop and i_d = \int j_d.dA = epsilon_0 d/dt (\int E.dA) is the displacement current though a surface spanning the loop. Thus i_d is epsilon_0 times the rate of change of electric flux through the loop. Integrating (1) over any volume, and converting into a surface flux integral using the divergence theorem, we get that the sum of  i + i_d into the closed surface bounding the volume is zero. That is, any charge current flowing into a closed surface is balanced by a displacement current flowing out of the surface. We applied these ideas to understand the magnetic field at a distance r from a wire with a circular capacitor (see figure 38-24 in problem 1 on page 879 of HRK). No matter what surface spans the Amperian loop, we get the same value for the right hand side of the Maxwell-Ampere law. We can even take the surface to miss the wire althogether by going between the capacitor plates! In that case the whole contribution comes from the displacement current. The magnetic field at a point next to the capacitor at distance R from the wire is thus the same as at a point far from the capacitor, as long as R is outside the capacitor (neglecting the fringing fields).

- Generation of electromagnetic waves: See section 38-4 of HRK. A shaking charge will "shake the electric field lines" and generate waves, sort of like transverse waves on a string. The changing electric field induces a magnetic field via the Ampere-Maxwell law, which then induces a further electric field via Faraday's law, and the whole thing propagates away at the speed of light. Two examples: 1) X-ray generation: high speed electrons slam into a target, and as they decelerate shake off X-rays, 2) radio waves generated by a radio transmitting anetenna with oscillating currents. A nice visualization of the electric field lines from a wiggling charge is in this radiating charge applet .  I demonstrated dipole radiation in class with the demo K8-42: RADIOWAVES - ENERGY AND DIPOLE PATTERN. We observed the fact that the wave amplitude is zero along the axis of the antenna, and the electric field vector is in the plane spanned by the antenna and the line joining the observation point to the antenna. That is, the wave is polarized in that direction. We also saw that the intensity falls off as the distance from the transmitter grows. This pattern, and the time-development of the waves, is nicely represented in a visualization of dipole radiation made at MIT. One thing about this visualization that really bugs me is they don't tell us how the electric field lines diplayed at each instant of time are selected. Is the point on the axis just poapagated at the speed of light, or what?

- Radiation pressure: it turns out, as we'll discuss more on Monday, that light with energy density u has momentum density u/c, so if it is absorbed or refelcted by a surface it transfers momentum and therefore exerts a force, called radiation pressure. This is fairly important to us, since it is what holds the sun up!

Wed, 11/6

- Electromagnetic waves: With the displacement current term the Maxwell equations are more symmetric..., but
how could it have been missed before Maxwell? The displacement current density is (epsilon_0 mu_0) E,t. That dimensionful coefficient is very small: (epsilon_0 mu_0)= (3 10^8 m/s)^-2, so it will not be important compared to any charge current unless the electric field is changing rapidly enough. It does have a profound effect though: not only can a changing magnetic field produce an electric field (Faraday's law of  induction), but a changing electric field can produce a magnetic field. This raises the possibility of a self-sustaining electromagnetic field in vacuum (i.e. with no charges or currents). We showed in fact that the vacuum Maxwell equations imply that E and B satisfy the wave equation with a wave speed c = (epsilon_0 mu_0)^(-1/2) = 3 10^8 m/s = the speed of light! For plane symmetric waves depending only on z and t, and propagating only in the k (+z) direction, the electric and magnetic fields are tightly linked to each other and to k:
E.k = 0 = B.k      E.B = 0,      E = cB,      E x B parallel to k
The energy density in an electromagnetic field is u = (1/2) epsilon_0 E^2 + (1/2 mu_0) B^2. These two terms are equal in a plane, unidirectional wave, which is reminiscent of the equality of kinetic and potential energy in string or sounds waves. The intensity is I = uc = EB/mu_0 = c epsilon_0 E^2 = c B^2/mu_0. The energy flux vector is called the Poynting vector: S = (E x B)/mu_0. Since the electric and magnetic fields are perpendicular, the magnitude of  the Poynting vector is equal to the intensity, while the direction is the direction of propagation of the plane wave. In fact, the Poynting vector describes the field energy flux in all situtations, not just for plane waves. 

Demos:
K8-01: ELECTROMAGNETIC WAVE - MODEL

K8-05: ELECTROMAGNETIC PLANE WAVE MODEL
K8-03: LIGHT NANOSECOND (we also discussed picoseconds (10^-12 s) and femtoseconds (10^-15 s)---one light-femtosecond is about  3/4 the wavelength of blue light.)

Tues, 11/5

- Discussed aspects of the homework problems:
1) the integral for the time translated wavepacket is not just a  complex Gaussian (due to the finite range of integration); 2) dispersion of light in  vacuo:  the fractional difference of the speeds in part 4(ii) will be proportional to a; 3) the equation of motion F = ma for the nth mass on the linear chain is obtained by adding the forces from the two adjacent springs, -k(x_n - x_(n-1)) + k(x_(n+1) - x_n).

- Maxwell's equations: we recalled various facts about vector calculus, in particular Stoke's theorem and the divergence theorem, and used these to infer the differential form of Maxwell's equations from the integral form. The supplement Differential form of Maxwell's equations and electromagnetic waves covers this topic. We saw that the equations are inconsistent with charge conservation unless Maxwell's displacement current term is added to Ampere's law. Charge conservation is expressed locally by the continuity equation, rho_t + div j = 0.


Mon, 11/4

- Showed how to get the dispersion relation from the differential equation for waves: assume a solution of the form exp(ikx -w(k)t) and insert into the equation, which becomes an algebraic equation for w(k).

- Schrodinger's equation for a particle in one dimension: i hbar psi_t =  - (hbar^2/2m) psi_xx. A brief  (one page) synopsis of quantum mechanics is given in the supplements. We looked at the dispersion relation and the interpretation in terms of energy and momentum, and discussed the probability interpretation and the spreading of the wavepacket, as well as theHeisenberg uncertainty relation.

- Wavepackets: looked in more detail at wavepackets constructed with square and Gaussian window functions A(k) (see notes from 10/30).  Explicitly, consider the two normalized amplitude functions: A1(k) = (1/2Dk) for k_0-Dk<k<k_0+Dk and 0 otherwise; and the Gaussian A2(k) = (1/Sqrt[pi] Dk)exp[-((k-k0)/Dk)^2]. These give for Y(x) = \int dk A(k) exp(ikx) respectively Y1(x) = exp(ik_0 x) sin(x Dk)/(x Dk) and Y2(x) =
exp(ik_0 x) exp[-(x Dk/2)^2]. You can see graphs of these here. It is remarkable how much difference the smoothing of A(k) makes in localizing Y(x).

Discussed how the latter is more localized than the former, and explained the limitation Dx Dk >~ 1 in terms of the requirement of phase cancellation to localize the waves: consider exp(i k Dx)+exp(i(k+Dk)Dx). These will first cancel completely when Dx Dk = pi. So you see that there is going to be some inverse realtion between the localization in x and the localization in k. Multiplying by hbar, this  becomes Dx Dp  >~ hbar, Heisenberg's uncertainty relation.


We also discussed propagation and spreading of wavepackets, and interpreted the graphs in terms of the group velocity and its variation over different k's in the wavepackets.

Thurs, 10/31

-
group velocity: showed transparencies that illustrate group velocity: the two transparencies have black stripes, made with slightly different magnifications on a copy machine, so one set of stripes is slightly farther apart than the other. When they are placed one atop the other a pattern of beats is seen. When the shorter "wavelength" one is moved the beat pattern moves with it, much more quickly than the sheet itself. When the longer wavelength pattern is moves the beat pattern moves opposite. The beat pattern moves at what is called the group velocity. We analyzed this for a pair of harmonic waves:
exp(ik_1 x - iw_1 t) + exp(ik_1 x - iw_1 t) = exp(ik_0 x - iw_0 t) 2 cos(Dk x - Dw t)
where k_0 = (k_1 + k_2)/2, and Dk = (k_2 - k_1)/2, and similarly for w. The first factor is the carrier wave. It is rapidly varying. The cosine factor is a slowly varying modulation if k_1 and k_2 are close to each other. It defines an envelope, which moves at the speed Dw/Dk, which is the group velocity. In the limit Dk -> dk, this becomes dw/dk.

- Surface waves on water: We can get a lot of information from dimensional analysis. On shallow water of depth h the wave speed might depend on h, g, and rho, the mass density of the water. However the only combination with dimensions of speed is Sqrt[gh]. It turns out this is in fact the wave speed, with a numerical coefficient of unity.  On very deep water the depth h can't matter, so there is no way to make a speed independently of the wavenumber k. Waves on deep water are thus dispersive, even if they have linear wavefronts (i.e. not just for circular waves). Using k, one can form Sqrt[g/k]. It turns out that with a coefficient of unity this is the phase velocity. That is, w(k) = Sqrt[gk]. The group velocity dw/dk is thus 1/2 the phase velocity. (Check this.) The shallow water dispersion relation applies when the wavelength is much longer than the depth, and the deep water relation holds in the opposite case. The general case turns out to be governed by w = Sqrt[gk tanh(kh).] When kh << 1, tanh(kh) ~= kh, so this reduces to the shallow water case. When kh >> 1, tanh(kh) ~= 1, so it reduces to the deep water case.

- Watched a film loop that illustrated the meaning of phase and group velocity, and how to make a localized wavepacket by superposing component waves of different wavelengths.


Wed, 10/30

- Dispersion: w(k) not proportional to k, so phase velocity w(k)/k not same for different k. Thus wavepackets spread, or disperse. The relation w = w(k) is called the dispersion relation for the waves. The group velocity , i.e. the speed of the center of a wavepacket centered on the wavenumber k, is given by dw/dk. An example I gave was waves on a chain of masses connected by springs, with equilibrium spacing a, for which w(k) = (2v_0/a) sin(ka/2). When ka << 1, this is approximately v_0 k. But when ka = 2 pi it is zero! As you'll see in the homework, this is actually the same as k =0 in disguise. When ka = pi one gets the maximum possible w, which turns out to correspond to a standing wave with vanishing group velocity, corresponding to a normal mode with alternate masses vibrating in opposite directions.

- One can make a localized wavepacket by superposing an infinite number of waves of different wavenumbers, by integrating:
Y(x,t) = \int dk A(k) exp(ikx - iw(k)t).
 If A(k)  is peaked around a central value k_0 with width Dk, then Y(x,0) will be peaked around x=0 with width Dx ~ 1/ Dk. That is, the more sharply localized in k, the more weakly localized in x, and vice versa. The form of the wavepacket at time t will depend on the form of the dispersion relation w(k).

Tues, 10/29

- Matching conditions for waves at an interface between two media: the for both string waves and sound waves, the displacements must match, and the forces must be equal and opposite (Newton's third law). In the sound case, the two bulk moduli generally differ, as well as the mass density differing.
Treated the case of string in class, allowing for different tensions, which could be achieved e.g. with the help of a frictionless rod and massless ring. This is all treated carefully and clearly in the textbook, Chapter 8, pp. 256-264. In the first part the problem is treated assuming the tensions in the two strings are equal. The second part allows for different tensions, and shows that the relevant quantity for each medium is the impedance, Z = T/v.  If the impedances match, then there is no reflection at all, even if the density and tension both differ. I talked about the idea of avoiding reflection by putting in a gradual interpolation between the impedances of two media. We have a nice demo of   impedance interpolation. (One can also make an impedance match without even having another wave medium, but rather just a kind of damper. We also have a demo of  impedance-matched wave absorption.


Mon, 10/28  No class (fire in Physics building)

Thur, 10/24

- Discussed the difference between 2d and 3d circular/spherical waves: the latter preserve pulse shape, the former not. We could see this by a trick: make a 2d wave using a 3d wave with cylindrical symmetry. The same argument cannot be used to show that 1d waves are not shape preserving, since the integral over the plane diverges.

- Discussed CO_2 molecule model, and the reason for the disagreement between the model and the experimental results. The key thing (I think) is that the position of one oxygen affects the spring between the carbon and the other oxygen. See discussion in the homework 5.2d solution.

- mid-semester course evaluations

-
Partial reflection and transmission at an interface between two media: examples of string, sound from air to water, and light from air to glass. The boundary conditions are 1) equal displacements and 2) equal slopes in the case of equal string tensions, equal pressures in the case of sound waves. We had some trouble understanding the reason for the equal slopes and pressures. I explained it interms of an infinitesimal mass element at the interface: this has finite acceleration, but infinitesimal mass, hence Newton's second law says that it must have infinitesimal net force on it. It was clear that this line of argument was difficult for the students to follow, I'm not sure why. Next Tuesday I gave a simpler reason for this boundary condition, using Newton's third law.


Wed, 10/23

- Reflection from fixed and free ends. Treated with the method of the virtual pulse. See Ch. 8, pp. 253-256.


Tues, 10/22  

-  Energy density in a 3d plane wave is better measured per unit volume (= per unit area per unit length)  rather than per unit length as for waves on strings. The energy of sound waves in a slab of thickness dx and area A is dK = 1/2 rho A (s_t)^2 dx. For a purely unidirectional wave s_t = +/- v s_x, so dK = 1/2 rho v^2 A (s_x)^2. As you'll show in this week's homework, the potential energy is dU = 1/2 K A
(s_x)^2 = 1/2 rho v^2 A (s_x)^2. THis is equal to the kinetic energy for directional waves.

- Intensity of a unidirectional wave: let rho_E be the energy density. The energy in a slab is rho_E A dx, which passes a fixed plane in a time dx/v, hence the intensity is 
(rho_E A dx)/( dx/v) = rho_E v, i.e. (energy density) x (wave speed).

- Plane waves in other directions: introduce the wave vector
  k = k n, where n is a unit vector. The magnitude of the wave vector k is the wave number. A sinusoidal plane wave in the n direction can then be written A sin(k.x - wt), where x is the position vector.

- Plane waves are an idealization: far from the source, in a small enough region, all waves will look like plane waves. On a larger scale, another useful idealization is spherical waves, which have constant amplitude on spheres of constant radius about some origin. Far from the origin in a small enough solid angle these look like plane waves as well, except that their amplitude decreases as 1/r. Why? If the energy is conserved, and all flows at rate v, then the rate of energy flow of an outgoing wave through a sphere of radius r must be independent of r. Denoting the intensity at radius r by I(r), this says that I(r) 4pi r^2 = constant, so I(r) ~ 1/r^2. On the other hand, energy density is proportional to the square of the amplitude A(r), hence A(r) ~ 1/r.  In fact, the general spherical wave solution to the wave eqn has the form f(r-vt)/r + g(r+vt)/r, where the two terms correspond to outgoing and ingoing spherical waves respectively. We illustrated this with the example of using intensity to measure the distance to a star.The intrinsic luminosity L, i.e the energy per unit time emitted, must be equal to the flux of energy through a large sphere at the location of the earth, a distance d from the star: L =  I(d)(4pi d^2). This can be solved for d = Sqrt[4pi L/I(d)]. If we know L, we can measure the distance to the star just by measuring the intensity of its light as viewed at the location of earth.


- Wave eqn in 3d: s_tt = v^2 (s_xx + s_yy + s_zz) = v^2 div.grad s = v^2 Laplacian s. Although the first form refers explicitly to the x,y,z axes, the laplacian differential operator div.grad is independent of the orientation of the axes. This should be reasonably evident from the fact that it can be written as the dot  product of the gradient, a vector operator, with itself. It can be shown---in fact you show in the homework---that the spherical solutions to this equation really have the form indicated in the previous paragraph. You might think that the 2d case is just as straightforward, but this turns out not to be true! It is more complicated...

Mon, 10/21

- discussed the curving of the exam: what and why; please see me if you have concerns about the course, your grade, etc.

- Superposition of waves, contined: Similarly electromagnetic waves can be superposed, due to the linearity of Maxwell's equations. The electric field vector at each point in sapce and time is the vector sum of the ones produced from all different sources. A radio receiver works by tuning to a resonant curcuit at a certain frequency, to pluck out the signal from one transmitter...which thanks to linearity is completely undisturbed by the presence of the other signals.

- Energy and superposition: two facts derived from waves on strings and applicable also to planar sound waves:
1) purely left or right moving waves have equal kinetic and potential energy densities;
2) the total energy density of a general wave is the sum of the right moving and left moving energy densities.

We showed this by working out the kinetic and potential energy densities. I think the same thing is shown in the textbook.

- Sound waves
: these propagate in 3d, so the displacement field s is a function of four variables: s(x,y,z,t), but for a plane wave it is just a function of x and t: s(x,t). The energy density is then properly given as energy per unit area per unit length, i.e. energy per unit volume. We discussed the form this takes for kinetic and potential energy. The intensity is the energy per unit time per unit area carried by the wave. For a purely unidirectional wave this is given by Intensity = (energy density) x (wave speed).

- Bulk modulus vs. Young's modulus: the book says that the former is always larger than the latter, and gives a reason, however this seems not to be borne out in the data. I think the reason given is specious. Also the some of the numerical values given in the book disagree with other sources. In particular, for aluminum Y is 7 10^10 N/m^2, not 6.

Thurs, 10/17

speed of sound in gases:
air(79% N2, 21% O2)     0 C
331 m/s
air                                    20 C
343 m/s
He                                     0 C
965 m/s
H2                                     0 C
1284 m/s

Why the difference with type? Why the dependence on temperature? For sound in air, Newton figured he could assume the air is at constant temperature and found K_isothermal = p (see notes for 10/9/02). Thus he found v = Sqrt[p/rho], which yields 289 m/s at standard temperature (0 C) and pressure (1 atm). Not a bad start, but something is obviously wrong. On the other hand, something's right: using the ideal gas law pV = NkT, we get p/rho = kT/m, where m is the mass of the molecule, so v = Sqrt[kT/m]. The square root dependence on temperature explains the temperature dependence of the speed of sound in air: Sqrt[(273+20)/273] x 331 = 343! Also the square root dependence on the mass explains the ratio of the speed in air to the speed in hydrogen:
atomic weight is 28 for N2 and 32 for O2, compared with 2 for hydrogen, so the ratio should be slightly higher than Sqrt[28/2] x 331=1238, and lower than Sqrt[32/2] x 331=1324. Not bad. However the ratio to the speed in helium doesn't work out: the atomic weight is 4, and Sqrt[28/4] x 331 = 876, whereas the speed in helium is 965 m/s.

What's wrong is that the gas is not at a constant temperature in the wave. Instead, it is at constant energy, i.e. adiabatic: no heat flow in or out. The adiabatic bulk compressibility differs from p by a factor called the adiabatic index gamma, which depends on the type of gas (and also on the temperature in a stepwise fashion---see below):  K_adiabatic = 
gamma p, hence v = Sqrt[gamma kT/m], which agrees with observation when the correct adiabatic index is used. The fact that the wave is not at constant temperature is plausible when one looks at the rms velocity of the molecules: kinetic theory of gases shows that kT/m = <v_x^2>, the average of the squared x-component of velocity, so the isothermal wave speed formula of Newton would yield a wave speed equal to the rms molecule speed in the direction of propagation. But if the wave is going as fast as the average molecule, then it seems plausible that there isn't enough time for thermal equilibrium to be maintained with the surroundings.

Physics of the adiabatic index: If the volume of gas is compressed a given amount, the pressure goes up. The amount by which the pressure rises depends on how many degrees of freedom there are other than the translational ones that give rise to pressure. The adiabatic index is determined the the number of accessible degrees of freedom (d.o.f.) D of a molecule, hence is different for monatomic (3 d.o.f.) , diatomic (5 d.o.f.), and polyatomic (6 d.o.f.) molecules. It is equal to the ratio c_P/c_V of the specific heat at constant pressure to the specific heat at constant volume. It is also given by the formula gamma = 1 + 2/D. Thus gamma_air = 1 + 2/5 = 7/5, while
gamma_He = 1 + 2/3 = 5/3. For air then, the sound speed should be Sqrt[7/5] times Newton's value of 289, i.e. 342 m/s. Pretty good, but not dead on. Not sure why. For the ratio of the speeds in helium to air, we have the square root of the ratio of the adiabatic indices: Sqrt[(5/3)/(7/5)]=Sqrt[25/21]=1.09. Multiplying this by the 876 found above from the mass ratio alone gives 955 m/s, which is pretty darn close. The extra mass of the O2 in air probably explains the difference.

The number of accessible degrees of freedom D actually depends on temperature in a stepwise fashion, because of quantum mechanics!  For temperatures around 0 C it turns out that D is indeed given by 5 for N2, O2, and H2, while it is 3 for He. However, the general quantum story here is that, according to the equipartition of energy, each accessible d.o.f. has on average 1/2 kT of energy, and this holds in the sound wave, at the local temperature which varies with the local pressure. However, quantum mechanics asserts that the angular momentum of anything, and molecules in particular, is quantized in units of Planck's constant hbar. The energy of rotation can be written as L^2/2I, where I is the moment of inertia. The squared angular momentum of the molecule is quantized as N(N+1)hbar^2, N=0,1,2,3,...hence the lowest energy of rotation is E_0 = hbar^2/I. If this smallest energy is much larger  than 1/2 kT, then the rotation is not accessible, and doesn't carry any energy. It turns out that the rotational degrees of freedom of N2 and O2 are accessible way below 0 C, while those of  H2 are fully accessible just around 0 C, since the moment of inertia of H2 is much smaller due to the low mass of hydrogen compared to nitrogen and oxygen (the size of the molecules is comparable). (Molecules also have vibrational degrees of freedom, but these are unaccessible except for much higher temperatures, due to the quantization of the energy of the vibrational motion.)

superposition of traveling waves:  Because of linearity of the wave equation: the sum of two solut                 ions is a solution. This means that right and left moving wave pulses just travel through each other. Were it not for this we could not make out the sounds from different instruments in an orchestra, for example. An example shows that left and right moving pusles ona string can cancel exactly at a moment, so there is no potential energy. At that point, however, the string has transverse velocity, and all the energy is kinetic.


Wed, 10/16   traveling waves: So far we discussed the equation of motion of string, and the normal modes that result. A string can also carry travelling waves, which must also be described by the same equation of motion. To see this, we began with a normal mode, in the form

 y(x,t) = A sin kx cos wt.

Terminology:
time
w: angular frequency
f  = w/2pi: frequency
T = 1/f = 2pi/w: period
space
k: wave number
k_book = k/2pi
lambda = 1/k_book = 2pi/k: wavelength
space/time
v = w/k = lambda.f = lambda/T: wave speed




Using a trig identity, we can rewrite the above normal mode as a superposition of right and left moving traveling waves:


y(x,t) = (A/2){sin[k(x-vt)] + sin[k(x+vt)]}

The first term is right-moving at speed v, the second is left-moving. Why? Consider any function of the form y(x,t) = f(x-vt). If t is increased by Dt and x is increased by Dx =vDt then the argument of f is unchanged, so y(x+vDt, t+Dt) = y(x,t). That is, the shape of the function y(x,t) at time t+Dt is exactly the same as it was at time t, but just shifted over to the right by a distance vDt. Similarly for y(x,t) = g(x+vt), moving to the left.  Hence v really is the wave speed.

Fact (to be proved in the homework): the general solution to the wave equation is of the form
y(x,t) = f(x-vt) + g(x+vt)
Wave speeds: string: Sqrt[T/mu], rod: Sqrt[Y/rho], bulk(e.g. sound): Sqrt[K/rho], where T= tension, Y= Young's (stretch) modulus,
K=bulk modulus, mu= mass/length, and rho=mass/volume. I discussed sound in air, but let me put all of this in Wednesday's notes, since we finished it then...

Tues, 10/15  discussed solutions to exam 1

Mon, 10/14  Exam 1

Thur, 10/10  review for exam 1, went over last year's exam1 (see supplements for a copy)
Wed, 10/9

- microscopic energy of atomic springs underlying Young's modulus: see notes for Tuesday

- Bulk modulus: (
see notes for Tuesday). Bulk modulus of gas: Newton was the first to compute this, as far as I know. He used it to compute the speed of sound. He knew Boyle's law: pV = constant for a gas at constant temperature. Thus 0 = d(pV) = dp V + p dV,  so dp = - p (dV/V). Compare this to the definition of Bulk modulus: F/A = -K DV/V. The force here is due to the overpressure, that is the pressure over (or under) the ambient pressure. Hence it is the same as dp, so Boyle's law gives us K=p. It turns out that in a sound wave the temperature is NOT constant, so Boyle's law does not apply. Instead we need the "adiabatic" Bulk modulus, which is a constant times p. More on this later.

- Discussed some hw problems. For the piano wire, you don't know the cross sectional area, but both the tension and mass per unit length are proportional to it, so it cancels out in v. For the CO2 molecule, clarified the assumption about the nature of the other normal mode, and how to impose the condition that the center of mass remains at rest. For the energy in the vibrating string, the kinetic energy is an integral over the string: K = \int dK = \int 1/2 dm v^2 = \int 1/2 M (dx/L) (y_t)^2. When there are two (or more) modes simultaneously excited, you need to show that the cross terms between the two modes vanish upon integration from 0 to L.


Tues, 10/8

- normal modes of aluminum rod demo: first three normal modes excited, by stroking the rod while holding it at the different node locations.
- Young's modulus: intrinsic property of a material that measures its stiffness, independent of the particular size and shape of the chunk of material being considered. Stress = F/A is proportional to strain = DL/L (fractional change in length) . The proportionality constant is Y, the Young's modulus: F/A = - Y DL/L. (The minus sign is because the book likes to let F be the reaction force of the stretched stuff on whatever agent is stretching it.) The relation to the spring constant k of the rod is seen by writing this as F = - (AY/L) DL, so k = AY/L. The Young's modulus of aluminum is 6 x 10^10 N/m^2, and the propotionality holds for strains less than 0.1 % or so.

- microscopic picture of Young's modulus: material is like a bunch of microscopic springs Hooked together (pun intended). Spring constants add in parallel and their inverses add in series, so the total spring constant of the rod is k = (N_parallel/N_series) k_0, where k_0 is the
microscopic sping constant for each atomic spring. Say the longitudingal distance between the atoms is a_0, the transverse distance between them is b_0, and the cross sectional area and length of the rod are A and L. Then N_parallel = A/b_0^2 and N_series = L/a_0, so k = (A/L)(a_0 k_0/b_0^2). Thus Y = a_0 k_0/b_0^2 in this crude model. This is a formula for Y in terms of microscopic properties of the material. [It is fun to put in the numbers for aluminum just to see what microscopic spring constant k_0 = Y b_0^2/a_0 results. Aluminum is a cubic crystal, with a_0=b_0 = 4 x 10^-11 m. Hence k_0 = Y a_0 = (6 x 10^10 N/m^2)(4 x 10^-11 m)=2.4 N/m. Is this reasonable? Work out the potential energy when the spring is stretched by an amount a_0:
U = 1/2 k_0 a_0^2 = 1/2 Y a_0^3 = (0.5)
(6 x 10^10 N/m^2)(4 x 10^-11 m)^3 = 1.9 x 10^-19 J = 1.2 eV. This says that to pull apart one atom pair spring to a distance equal to the interatomic spacing would cost an energy equal to 1.2 eV, which is quite reasonable in order of magnitude, since a typical covalent bond energy is of the order of an electron volt...]

- wave equation for compressional waves
: we derived this by applying F = ma to the longitudinal motion of each little bit Dm of the rod. See textbook for details. The key step was to note that the local strain is equal to s_x, the partial derivative of the displacement wrt x, and to use this to write the force of the material on the right hand side of Dm as AYs_x. The net force is the difference between this and the force on the left, which gives approximately AYs_xx Dx. Equating this to (rho A Dx) s_tt  yields s_tt = (Y/rho) s_xx, where rho is the mass per unit volume. This is the equation of motion for the rod and it is mathematically identical to the equation for transverse vibrations of a stretched string. The same story goes for sound waves except there we speak of not the Young's modulus, but the bulk modulus K, defined by F/A = - K DV/V, where V is the volume. If the change of volume is only along one direction then DV/V=DL/L so the bulk and Young's moduli are the same. (For the rod this is not the case, since when the ron stretches it gets a bit skinnier in the transverse direction, and when it compresses it gets a bit fatter.)


Mon, 10/7:

- Wavelength: distance for sin function to complete one cycle. If mode function is f(x) = A sin((w/v) x + phi), then wavelength satisfies (w/v) lambda = 2pi, or w = 2pi v/lambda. Since ordinary frequency is f = w/2pi, this is the same as
 (wavelength)*(frequency)=(wavespeed),           or           (wavelength)/(period)=(wavespeed)
- Dependence of normal mode frequencies and shapes on boundary conditions: one fixed and one free end gives the spectrum of frequencies w_n = (n - 1/2) pi v/L. The mode functions in this case have an odd number of quarter wavelengths. If both ends are free, the frequencies are the same as if both are fixed, but the mode functions are shifted by pi/2 in phase, i.e. nodes and anti-nodes are interchanged. Sound waves in a tube is an example: an open end of a tube is a node for pressure (since the pressure must match the ambient air pressure in the room) but an anti-node for displacement. At a closed end of the tube it is the reverse. We did a demo with two identical length tubes, one with both ends open and the other with one open and one closed end. The fundamental of the closed/open tube has one quarter wavelength, while that of the open/open tube has one half wavelength, hence the latter has a frequency double the former, i.e. one octave above.


Thur, 10/3:

-  normal modes of string: The general solution to the equation for the normal mode amplitude function (see the previous lecture) is f(x) = A sin((w/v) x + phi).
There must be something missing, since this holds for ANY w,  whereas we know that only certain special frequencies are allowed for normal modes. What is missing is the boundary conditions, which enforce the fact that the ends of the string are not moving: y(0,t) = 0 = y(L,t). The condition at x = 0 implies that phi = 0 (or pi, but that can be absorbed in a change of the sign of A), and the condition at x = L then implies that the frequency must be such that wL/v is an integer multiple of pi. Therefore the frequencies of the normal modes are multiples of a lowest one w_n = n w_1, where w_1 = pi v/L. The lowest mode is called the fundamental, and the rest are called harmonics.

- The amplitude function in the n^th mode is thus f_n(x) = A_n sin(n pi x/L). The points where the amplitude vanishes are called nodes. The number of nodes is equal to n+1, including the two nodes at the endpoints. The maxima of the amplitude are called anti-nodes. The integer n is the number of half-cycles of the sin function that fit in the length L. The distance for one cycle is called the wavelength, and it is given in the n^th mode by lambda_n = 2L/n.

- The total number of normal modes is infinite, since the system has an infinite number of masses, continuously strung together. However, there are really only a finite number of atoms in the string, and we know that a system with N degrees of freedom has N normal modes, so there can only be a finite number. When the wavelength is shorter than the inter-atomic spacing the continuous treatment of the string fails.

- In the n^th mode, we have y_n(x,t) = A_n sin(
n pi x/L) cos(n pi v t/L). From this we can find the speed of a point on the string at position x and the slope of the string using partial derivatives with respect to t and x respectively. The maximum speed is thus A_n n pi v/L and the maximum slope is A_n n pi/L.

- Different boundary conditions: if an end of the string is free to move up and down (say if it is tied to a massless ring that slides up and down a frictionless post) then the slope of the string must vanish at that end, i.e. it must lie at an anti-node. Why? Think of the force on the ring: it is pulled vertically only on one side, so the vertical force would be finite while the mass is zero. This would cause an infinite acceleration that would flatten out the string. Put differently, since the acceleration of the ring is finite, and the mass is zero, the force must be zero, which is only the case if the slope is zero.

Wed, 10/2:

-  Equation of motion of string: A string under tension is like an infinite number of coupled oscillators. The restoring force is due to the tension T. The inertia is from the mass per unit length (linear mass density) mu. We consider only small transverse vibrations. Then each bit of string only moves perpendicular to the equilibrium line. The string configuration is described by a function y(x,t), the displacement of the bit of string at coordinate x and time t. We assume the slope always remains much less than unity, i.e. y_x <<1, where y_x denotes the partial derivative of y wrt x. Under these assumptions also the tension is constant to a good approximation. To find the equation of motion we consider a little section of string of mass Dm = mu Dx, and apply Newton's second law to it. The acceleration of this bit in the y direction is y_tt, the second partial derivative of y wrt t. Hence we have mu Dx y_tt = F_y. The force is the y-direction is due to the y-component of the tension, T_y.  Similar triangles give T_y/T_x = y_x, but T_x is approximately T, hence T_y = T y_x + terms of higher order in the small quantity y_x. Now the net force in the y-direction comes from the mis-match in the slopes at x and x+Dx:
F_y = T_y(x+Dx) - T_y(x) = T [ y_x(x+Dx) - y_x(x) ] = T y_xx(x)  Dx  + O(Dx^2).
Thuse we have Dx y_tt = F_y = T y_xx(x)  Dx  + O(Dx^2). Dividing through by Dx and taking the limit Dx->0 then gives
y_tt = v^2 y_xx,    (string wave equation)
where v^2 = T/mu is what turns out to be the square of the wave speed. (Check that it has the dimensions of speed squared.)
- Normal modes of string: Assume each point undergoes SHM with same frequency but different amplitude: y(x,t) = f(x) cos wt. Then y_tt = -w^2 y and y_xx = f_xx cos wt = (f_xx/f) y, so the amplitude must satisfy
f_xx = -(v/w)^2 f   (string normal mode equation)

Tues, 10/1:

- Examples of coupled oscillators and normal modes: 2,3,4 coupled pendula, Wilberforce pendulum, pair of hanging masses.
- superposition of normal modes.

- energy transfer in coupled oscillators viewed as one oscillator forcing the other; if they have the same or nearby frequencies there is resonance, and most or all of the energy can be transferred.
-molecules have resonant frequencies, and they absorb electromagnetic (infrared) radiation of the  resonant frequencies. This produces absorption lines, by which chemical species can be detected from far away. In optical fibers, one tries to use frequencies that do not resonate with any atomic structures in the fiber material.
- if you know linear algebra: the general problem of finding the normal mode frequencies and amplitudes amounts to finding the eigenvalues and eigenvectors of a matrix.


Mon, 9/30:

- coupled oscillators: demo of two pendula connected by a spring. Motion can start in one mass, and then transfer to the other and back again. So the motion of one mass exhibits beats. On the other hand there are two very special motions which are simple harmonic oscillations, in which all parts of the system oscillate with the same angular frequency and fixed amplitudes. These are called normal modes. We found the normal mode frequencies first by just writing down F=ma for these two motions.
(See textbook for the details). Then we went back and solved it more generally, not assuming any special relation between the motion of the two pendula. Adding and subtracting we found that the normal mode coordinates
q_1 =  x_A + x_B and q_2 = x_B - x_A each satisfy simple harmonic oscillator equations, with different natural frequencies. Using  the general solution for these normal mode coordinates  we can go back and find the solution for the coordinates of the masses x_A = (q_1 - q_2)/2 and x_B = (
q_1 + q_2)/2.  For example, if the initial conditions are x_A(0)  = A,  x_B(0)  = 0, and the time derivatives both vanish, then the corresponding initial conditions on q_1 and q_2 are q_1(0) = A,  q_2(0) = -A, and the time derivates vanish. Thus q_1 = A cos(w_1 t) and q_2 = - A cos(w_2 t), which corresponds to  x_A = (A/2)[ cos(w_1 t) + cos(w_2 t) ] and  x_B = (A/2)[cos(w_1 t) - cos(w_2 t) ]Each of these exhibits beats at the beat frequency equal to the difference of the normal mode frequencies (w1 - w2).

Thur, 9/19, Mon-Thur 9/22-26
(I may try to fill this in later...haven't had the time)
power in AC circuits

series RC circuit
complex impedance, parallel RC circuit
RLC circuit, complex impedance
power & resonance
transients, damped driven oscillator & resonance


Wed, 9/18:


- example of expressing sum of oscillations with same frequency in the form Re[A e^i(wt+a)]:
2 cos(wt) + 3 sin (wt) = Re[2 e^iwt] + Re[-3i e^iwt] = Re[(2-3i)e^iwt] = Re[Sqrt[13]e^i(wt - tan^-1(3/2))]
. Another method is to just set the left hand side equal to Acos(wt + a) = A cos wt cos a - A sin wt sin a, and equate the coefficients of the cos and sin terms: A cos a = 2 and A sin a = -3, which has the solution A = Sqrt[13] and a = tan^-1(3/2).
- Transients: when we demonstrated the driven oscillator on Tuesday using the torsional oscillator demo, it didn't behave quite like the steady state solution would indicate. In fact, the amplitude seemed to slowly oscillate from being large to being small and back again. The reason is that I did not turn on the damping force, so the transients were not going away. What is going on here is that in addition to the steady state solution there is another component to the motion. The general solution has the form x(t) = x_ss(t) + x_free(t), where x_free(t) is any solution to the free oscillator equation. (The fact that we can just add any solution x_free(t) is due to the fact that  x appears linearly in the equation of motion.) More explicitly, x(t) = A_ss cos(wt -d) + A_free cos(wt + a), so x(t) is the sum of two harmonic oscillations at the same frequency with different amplitude and different phase. This produces beats, as we have previously heard with a guitar string and now see with the torsional oscillator. What is different in the damped case is only that x_free(t) is a solution to the damped free oscillator equation. Those solutions all have the property that they die away exponentially in time, due to the damping. Hence at late times only the steady state solution remains! The greater the damping, the faster the total solution approaches the steady state solution.
- steady state solution in the driven case: we found this using the complex eponential method. The result:
x(t) = Re[C e^(iwt)], where C = (F_0/m)/(w0^2 - w^2 + i gamma w). Put differently, C = A e^(-id), where A = (F_0/m)/Sqrt[(w0^2 - w^2)^2 + (gamma w)^2] and tan d =  gamma w/(w0^2 - w^2). Note that the denominator of C has a positive imaginary part, hence its phase is between 0 and pi, so d runs from 0 to pi. Features:
(1) When w = w_0, the amplitude is finite: A = F_0/(m gamma w), and the phase is -pi/2. The position thus lags the force by 90 degrees. The smaller the damping, the larger the amplitude.
(2) For w < w_0 the position lags the force by less than 90 degrees, while for w > w_0 it lags by more than 90 degrees, approaching 180 degrees for w >> w_0.
(3) The amplitude is not maximized at exactly w = w_0 unless gamma = 0. Rather it is maximized below w_0.


Tues, 9/17:

- we found the combination of the two solutions in the overdamped case for the intial condition v(0)=0. both solutions are involved. for large damping, the more slowly damped solution dominates.
- driven, undamped oscillator: we found the steady state solution. it has the following features:
(1) the amplitude is determined, not free to be specified;
(2) the amplitude diverges as w approaches w_0, which is called resonance;
(3) x is in phase with F below resonance (w < w_0), and x is 180 degrees out of phase with F above resonance (w>w_0).
- if a damping term is added, then the larger the oscillation the larger the rate of energy loss by damping. Thus, instead of diverging, the amplitude will reach a maximum at (or near---see Wednesday) w = w_0.

Mon, 9/16:

- Please don't use calculators to do the complex number manipulations on the homework (like converting between polar and cartesian form). You need to really understand how to do this yourself! (You can of course use the caculator to evaluate trig functions.)  
- more on the damped oscillator (see textbook):
- damping force power drain: P = Fv = -bv^2, leads to decrease of the energy of the oscillator.
- meaning of gamma: dimensions are 1/T; 1/(time for energy to drop to 1/e of initial value); 2/(time for amplitude to drop to 1/e of initial value;gamma = |dE/dt|/E,
the fractional rate of change of energy loss. Note however that this only holds on the average. The instantaneous rate of energy loss is zero when the oscillator is at rest at the maxima.
- Quality factor Q = w_0/gamma, which is approximately (2pi)(# of cycles for energy to drop by 1/e)
- overdamped case, gamma/2 > w_0: There is no oscillation. The general solution is a combination of two different exponential decay rates!
- critically damped case: gamma/2 = w_0: the exponential method yields only one solution, x(t) = Cexp(-(gamma/2) t).  But there must be two independent parameters in the general solution, corresponding to the freedom to choose the initial position and velocity! What is the other solution?? To find it, one can take the limit of the overdamped case as gamma/2 approaches w_0. Alternatively, look at the very special case where there is no damping and no restoring force: d^2x/dt^2 = 0. This is just a free particle, and the general solution is x(t) = C + Dt. This might lead you to guess that the solutoin for the critically damped oscillator is simply x(t) = (C + Dt)exp(-(gamma/2) t). This guess is correct, as you can easily verify by plugging it into the equation of motion.
- I ended class today with a question: if you start out an over-damped harmonic oscillator at rest and displaced some distance from equilibrium, what will its subsequent motion be? Will it involve both of the damping rates, and if so with what relative weight?



Thu, 9/12:

- example of beats with equal amplitudes and zero phase shift: A1=A2=A and d1=d2=0 (see textbook).
- damped oscillator (see textbook).
- to be emphasized: the reason the complex method works is that (1) the equation of motion is linear, and (2) the coefficients are real. In detail: the equation is of the form x'' +  r x' + s x = 0, where the prime denotes derivative wrt time and r = gamma = b/m and s = w0^2 = k/m. We replace this by the same equation for a complex function z(t) = x(t) + i y(t),  z'' +  r z' + s z = 0. The real and imaginary parts of  z both separately satify the orginal equation:

0 = (x + iy)'' + r (x + iy)' + s (x + iy) = (x'' +  r x' + s x) + i(y'' +  r y' + s y).
If a complex number is zero, then both its real and imaginary parts must be zero, hence it follows that (x'' +  r x' + s x) = 0 and (y'' +  r y' + s y) = 0.
- To better understand why this worked, consider a case where it would not work: suppose the original equation were x'' + q x^2 = 0...
- We inserted z(t) = C e^pt and saw that the differential equation became an algebraic equation, since the operation of differentiation became the operation of multiplication by p! Solving the quadratic equation we obtained the solution for the under-damped case, i.e. when gamma/2 < w0.


Wed, 9/11:

- z = x + iy is the cartesian representation and z = r exp(i q) is the polar representation of the complex number z. r = |z| is the modulus of z and q is the argument of z, Arg[z] (defined up to an integer multiple of 2pi). It is also called the phase of z.
- cos q = [exp(iq) + exp(-iq)]/2, and sin q = [exp(iq) - exp(-iq)]/2i.
- did several examples of computing modulus, phase, inverse, polar and cartesian forms of complex numbers, both geometrically and algebraically.
- circular or rotating vector representation of harmonic oscillation: x(t) = A cos(w t + d) = Re[z (t)], where z(t) = A exp[i(w t + d)]. The complex number z(t) rotates counterclockwise with angular frequency w on a circle of radius A in the complex plane. In this way a harmonic oscillation in x is represented as the projection on the x-axis of a uniform circular motion in the plane.
- Beats: when two oscillating signals with different frequencies are combined the total signal has an intensity that oscillates with a beat frequency that is the difference of the two individual frequencies. This can be understood nicely using the rotating vector representation: Say x1(t) =  A1 cos(w1 t + d1) and x2(t) =  A2 cos(w2 t + d2) are the two signals to be added. Then x_tot(t) = Re[z1(t) + z2(t)], where z1(t) = A1 exp[i(w1 t + d1)] and similarly for z2(t). The resultant z1(t) + z2(t) has maximum modulus when the two angles are lined up, i.e. when w1 t + d1 = w2 t + d2 + 2pi n, i.e. when t = [2pi/(w1-w2)]n + (d2-d1)/(w1-w2). That means the interval T_beat between successive maxima of the modulus is T_beat = 2pi/(w1-w2). Hence the beat frequency is w_beat = w1-w2. (The exact line-up generally occurs when z1(t) + z2(t) is not along the real axis, however if w1 and w2 are close to each other, there will be many trips around the circle while the two are almost lined up, which will produce the observed beats.
(You may like to look at this  visual applet illustrating beats, and this audio one.)


Tues, 9/10:

- direction angle for a complex number is q = tan^-1(y/x) if x>0, i.e. in the right half of the complex plane. For x<0 it is pi + tan^-1(y/x), while on the imaginary axis it is + pi/2 for y>0 and - pi/2 for y<0.
- geometrical representation of complex numbers: a complex number is a vector on the complex plane. Addition of complex numbers corresponds to vector addition. Mutliplication by a real number is scalar multiplication of the vector. Multiplication by i is counter-clockwise rotation through 90 degrees, i.e. pi/2 radians. Thus i^2 = ii = counter-clockwise rotation through 180 degrees, which is the same as multiplication by -1. So you can understand the "number" i as an operation of  pi/2 rotation in the complex plane.
- complex exponential function, Euler's identity: exp(i q) = cos q + i sin q.
- multiplication by exp(iq) corresponds to counterclockwise rotation through the angle q.
- z^w = exp(w ln z). So i^i = exp(i ln i). But i = exp(i pi/2), so ln i = i pi/2, so i^i = exp(-pi/2). Note however that the ln function is multi-valued: you can add any inter multiple of i 2pi to the exponent and change nothing, since exp(i 2pi n) = 1 for all integers n. Thus to be more general, i^i = exp(-(pi/2 + 2pi n).
- complex harmonic oscillator solution to the equation  d^2z/dt^2 = -w^2 z

z(t) = (A + iB) exp(iwt) = (A cos wt - B sin wt) + i(A sin wt + B cos wt)
the real and imaginary parts are separately solutions, and in fact by choosing A and B one gets the most general solution this way.

Mon, 9/9:

- molecular potential example
- took photos of the class
- Complex numbers: oscillator equation d^2x/dt^2 = - w^2 x. Exponential solution: x(t) = A e^pt works if p^2 = - w^2, i.e. p = w Sqrt[-1]. So invent a new number, i = Sqrt[-1], called the "imaginary unit". Then i^2 = -1. A real oscillator position should be described by a real number, not an imaginary one. We shall use imaginary (more generally, complex) numbers to construct REAL solutions.
- Define general complex number as z = x + iy. x = Re[z] = real part of z; y = Im[z] = imaginary part of z (sometimes iy called "imaginary part of z"). Addition defined by z1 + z2 = (x1 + x2) + i (y1 + y2), multiplication defined to satisfy the usual rules: communtative, associative, distributive. Hence z1 z2 = (x1 x2 - y1 y2) + i (x1 y2 + y1 x2).
- Do we have to invent more numbers? No! All equations can be solved with complex numbers. E.g. sin z = 2, and e^z = -1 have solutions (see later).
- fundamental theorem of algebra: any polynomial z^n + c_(n-1) z^(n-1) + ... + c2 z^2 + c_1 z + c_0 can be factorized to the form (z - w1)(z - w2)...(z - wn), so any nth order polynomial equation has exactly n solution ("roots").
- complex plane, representation of complex numbers as two-dimensional vector with magnitude or "modulus"  |z| = Sqrt[x2 + y2] and direction angle q = tan^-1(y/x).


Thurs, 9/5:

- theme: everything is (approximately) a harmonic oscillator near equilibrium.
- Simple pendulum (point mass at the end of a massless string): NOT a harmonic oscillator...except for small enough amplitude. If q is the angle, we showed that d^2q/dt^2 = -(g/l) sin q. Since sinq = q - q^3/3! + q^5/5! - ...this is approximately the h.o. equation  when q is not too large, and the angular frequency is Sqrt[g/l]. Another way to look at it: the potential energy of the pendulum is U(q) = mgl(1-cos q) = mgl (q^2/2 - q^4/4! + ...), so the potential energy function is approximately a parabola for small enough q.
- Galileo noticed that the period of a pendulum is (nearly) independent of the amplitude, and used that to make a good clock for timing physics experiments.
- Dimensional analysis would tell us T(m,l,g,q_max) = f(q_max) Sqrt[g/l], without any other analysis! But it tells us nothing about the function f(q_max) of the dimensionless maximum angle q_max. As we showed above, for small q_max, f is approximately 1. The next correction is -q_max^2/16. (The honors students will show this in the first homework.) (In class I mistakenly said it was -q_max/16, without the square.) The exact result is not an elementary function, but can be expressed as a definite integral.
- Physical pendulum: extended rigid body suspended from a point. Also approximately a h.o.
- completely general situation: make Taylor expansion of potential about the point x_e:

U(x) = U(x_e) + U'(x_e) (x - x_e) + 1/2! U''(x_e) (x - x_e)^2 + 1/3! U'''(x_e) (x - x_e)^3 + ...
where the primes denote derivates. If x_e is a point of stable equilibrium, then U'(x_e) = 0 and (assuming it is nonzero) U''(x_e) > 0. Thus sufficiently near the bottom of the bowl we have the approximation U(x) = U(x_e) + 1/2 U''(x_e) (x - x_e)^2 which has the parabolic form of a harmonic oscillator potential with effective spring constant
k_eff = U''(x_e)
The frequency of small oscillations about x_e is thus Sqrt[U''(x_e)/m].


Wed, 9/4:

- more syllabus discussion
- dimensional analsis: how to do it and what it's good for:
(1) Catch errors, (2) infer answers without solving the whole problem, (3) check answers, (4) imbue formulae with meaning. For example, using just dimensional analysis we can find that the angular frequency of the harmonic oscillator must be of the form  # Sqrt[k/m], where # is a dimensionless number the same for all oscillators. So we find how the frequency depends on k and m without solving the differential equation of motion!
- important observation: argument of trig functions and exponential must be dimensionless!
- energy conservation for the harmonic oscillator
- derivation of equation of motion from energy conservation:

0= dE/dt = d/dt(1/2 m v^2 + U) = mv dv/dt + dU/dx dx/dt.
If v is not zero, this implies m dv/dt = - dU/dx, i.e. ma = F, Newton's 2nd law.
- example of the LC oscillator: Total energy: E = 1/2 L I^2 + 1/2 (1/C) Q^2. Perfect analogy to a mechanical oscillator:


x v = dx/dt m k k/m
Q I = dQ/dt L 1/C 1/LC
 The total energy in the LC circuit is conserved, and oscillates between electric field energy stored in the capacitor and magnetic field energy stored in the inductor. Q oscillates like a sin (or cosine) function of time with angular frequency
1/Sqrt[LC]. You can also find the equation of motion not from energy conservation but from Kirchoff's law that the sum of the voltage drops around a closed loop is zero. In this case that gives: L dI/dt + Q/C = 0, or dI/dt = -(1/LC) Q. Since I = dQ/dt this is just the harmonic oscillator equation.


Tue, 9/3:

- Syllabus discussion
- Course intoduction, everything is waves
- mathematical techniques used in course: complex number, ordinary & partial differential equations, Taylor expansions, vector calculus.
- Harmonic oscillator: prototype of all oscillations and waves:
    - restoring force and oscillatory motion
    - (second order, linear, ordinary) differential equation and general solution
    - two initial conditions (e.g. position and velocity at time t=0) required to determine solution, hence general solution has two adjustable parameters. The way we wrote it, these are the amplitude and phase angle.
    - amplitude, angular frequency, frequency, period, phase angle