Добавил:
Upload Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:
intro_physics_1.pdf
Скачиваний:
52
Добавлен:
08.02.2016
Размер:
5.79 Mб
Скачать

414

Week 9: Oscillations

f) In this way one can (for weak damping) determine that:

Q =

1

=

0

= ω0τ

(870)

2ζ

b

 

 

 

 

where τ = m/b is the exponential damping time of the energy of the associated damped oscillator. This is one of the reasons defining dimensionless ζ is convenient. ζ is basically the inverse of Q, so that the larger Q (smaller ζ) is, the weaker the damping and the better the resonance will be!

It is worth mentioning in passing that one can easily reformulate the instantaneous or average power in terms of A, the driven amplitude, instead of F0, the driving force. The point then is that the power will be related to the amplitude of oscillation squared. This is intuitively reasonable, as the work removed per cycle is the damping force (proportional to A) integrated over the distance travelled (proportional to A). This is consistent with our developing rule of thumb that oscillator or wave energies or powers are (almost) always proportional to the oscillation or wave amplitude squared.

So what of this are you responsible for knowing? That depends on your interest and the level of your class. If you are a physics or math major or an engineer, you should probably work through the math in some detail. If you are a major in some other science or are premed, you probably don’t need to know all of the details, but you should still work to understand the general idea of resonance, as it is actually relevant to various aspects of biology and medicine and can occur in many other disciplines as well. At the very least, all students should be able to semi-quantitatively draw, and understand, Pavg(ω) for any given Q in the range from 3 to maybe 20 visibly to scale, specifically so that Pmax = F02/2b and Q ≈ ω0/ ω in your graph. Practice this on your homework! I nearly always ask this on one exam or quiz in addition to the homework.

The point of understanding Q pretty thoroughly is that oscillators with low Q quickly damp and don’t build up much amplitude even from a perfectly resonant ω = ω0 driving force. This is “good” when you are building bridges and skyscrapers. High Q means you get a large amplitude, eventually, even from a small but perfectly resonant driving force. This is “good” for jackhammers, musical instruments, understanding a good walking pace, and many other things.

9.5: Elastic Properties of Materials

It is now time to close a very important bootstrapping loop in your understanding of physics. From the beginning, we have used a number of force rules like “the normal force”, and “Hooke’s Law” because they were simple rules that we could directly observe and use to help us both understand ubiquitous phenomena and learn to use Newton’s Laws to quantitatively describe many of them.

We had to do this first, because until you understand force, work, energy, equations of motion and conservation principles – basic mechanics – you cannot start to understand the microscopic basis for the macroscopic “rules” that govern both everyday Newtonian physics and things like thermodynamics and chemistry and biology (all of which have rules at the macroscopic scale that follow from physics at the microscopic scale).

We’ve done a bit of this along the way – thought about microscopic causes of friction and drag forces, derived the ways in which a macroscopic object can be thought of as a pointlike microscopic object located at its center of mass (and ways it cannot, e.g. when describing its rotation). We’ve even thought a bit about things like compressibility of fluids, but we haven’t really thought about this enough.

Let’s fix that.

Week 9: Oscillations

415

To do so, we need a microscopic model for a solid, in particular for the molecular bonds within a solid.

9.5.1: Simple Models for Molecular Bonds

Consider, then a very crude microscopic model for a solid. We know that this solid is made up of many elementary particles, and that those elementary particles interact to form nucleons, which bind together to form nuclei, which bind elementary electrons to form atoms and that the atoms in turn are bound together by short range (nearest neighbor) interatomic forces – “chemical bond” if you like – to actually form the solid.

From both the text and some of the homework problems you should have learned that a “generic” potential energy associated with the interaction of a pair of atoms with a chemical bond between them is given by a short range repulsion followed by a long range attraction. We saw a potential energy form like this as the “e ective potential energy” in gravitation (the form that contained an angular momentum barrier with L2 in it), but the physical origin of the terms is very di erent.

The repulsion in the molecular potential energy comes from first the Pauli exclusion principle189 in quantum mechanics (that makes the interpenetration of the electron clouds surrounding atoms energetically “expensive” as the underlying quantum states rearrange to satisfy it) plus the penetration of a screened Coulomb interaction. The former is truly beyond the scope of this course – it is quantum magic associated with electrons190 as fermions191 , where I’m inserting wikipedia links to lots of these terms so that interested students can use them as the starting points of wikipedia romps, as a lot of this is all absolutely fascinating and is one of the reasons physicists love physics, it is all just so very amazing.

Next semester you will learn about Coulomb’s Law192 , that describes the forces between two charged particles, and (using Gauss’s Law193 ) you will be able to understand how the electron cloud normally “screens” the nuclei as eventually the rearrangement brings increasingly “bare” nuclei close enough so that the atoms have a very strong net repulsion.

One thing that we won’t cover then, however, is how this simple/naive model, which leads to two atoms not interacting at all as soon as they are not “touching” (electron clouds interpenetrating) is replaced by one where two neutral atoms have a residual long range interaction due to dipoleinduced dipole forces, leading to what is called a London dispersion force194 . This force has the generic form of an attractive −C/r6 for a rather complicated C that parameterizes various details of the interatomic interaction.

Physicists and quantum chemists or engineers often idealize the exact/quantum theory with an approximate (semi)classical potential energy function that models these important generic features. For example, two very common models (one of which we already briefly explored in week 4, homework problem 4) is the Lennard-Jones potential195 (energy):

ULJ (r) = Umin

½2

³ rb

´

 

³ rb ´

¾

(871)

 

 

 

r

 

6

 

 

r

12

 

189Wikipedia: http://www.wikipedia.org/wiki/Pauli Exclusion Principle.

190Wikipedia: http://www.wikipedia.org/wiki/Electron.

191Wikipedia: http://www.wikipedia.org/wiki/Fermion.

192Wikipedia: http://www.wikipedia.org/wiki/Coulomb’s Law.

193Wikipedia: http://www.wikipedia.org/wiki/Gauss’ Law.

194Wikipedia: http://www.wikipedia.org/wiki/London dispersion force. This force is due to Fritz London who was

a Duke physicist of great reknown, although he derived this force from second-order perturbation theory long before he fled the rise of the Nazi party in Germany and eventually moved to the United States and took a position at Duke. London is honored with an special invited “London Lecture” at Duke every year. Just an interesting True Fact for my Duke students.

195Wikipedia: http://www.wikipedia.org/wiki/Lennard-Jones potential.

416 Week 9: Oscillations

In this function, Umin is the minimum of the potential energy curve (evaluated with the usual convention that r → ∞ is ULJ () = 0), rb is the radius where the minimum occurs (and hence is the equilibrium bond length), and r is along the bond axis. This function is portrayed as the solid line in figure 126.

An alternative is the Morse potential196 (energy):

 

UM (r) = Umin ³1 − ea(rrb )´2

(872)

In this expression Umin and rb have the same meaning, but they are joined by a, a parameter than sets the width of the well. The only problem with this form of the potential is that it is di cult to compare it to the Lennard-Jones potential above because the LJ potential is zero at infinity and the Morse potential is Umin at in infinity. Of course, potential energy is always only defined within an additive constant, so I will actually subtract a constant Umin and display:

UM (r) = −Umin + Umin ³1 − ea(rrb )´2)

(873)

as the dashed line in figure 126 below. This potential now correctly vanishes at .

U(R)

2

1.5

1

0.5

0

-0.5

-1

0

0.5

1

1.5

2

R

Figure 126: Two “generic” classical potential energy functions associated with atomic bonds on a common scale Umin = 1, rb = 1.0, and a = 6.2, the latter a value that makes the two potential have roughly the same force constant for small displacements.The solid line is the LennardJones potential UL(r) and the dashed line is the Morse potential UM (r). Note that the two are very closely matched for short range repulsion, but the Morse potential dies o faster than the 1/r6 London form expected at longer range.

I should emphasize that neither of these potentials is in any sense a law of nature. They are e ective potentials, idealized model potentials that have close to the right shape and that can be used to study and understand molecular bonding in a semi-quantitative way – good enough to be compared to experimental results in an understandable if approximate way, but hardly exact.

The exact (relevant) laws of nature are those of electromagnetism and quantum mechanics, where the many electron problem must be solved, which is very di cult. The problem rapidly becomes too complex to really be solvable/computable as the number of electrons grows and more atoms are involved. So even in quantum mechanics people not infrequently work with Lennard-Jones or Morse

196Wikipedia: http://www.wikipedia.org/wiki/Morse potential.

Week 9: Oscillations

417

potentials (or any of a number of other related forms with more or less virtue for any particular problem) and sacrifice precision in the result for computability.

In our case, even these relatively simple e ective potentials are too complex. We therefore take advantage of something I have written about extensively up above. Nearly all potential energy functions that have a true minimum (so that there is a force equilibrium there, recalling that the force is the negative slope (gradient) of the potential energy function) vary quadratically for small displacements from equilibrium. A quadratic potential energy corresponds to a harmonic oscillator and a linear restoring force. Therefore all solids made up of atoms bound by interactions like the e ective molecular potential energies above will inherit certain properties from the tiny “springs” of the bonds between them!

9.5.2: The Force Constant

In both of the cases above, we can derive a force constant – e ective “spring constant” of the interatomic bonds that hold atoms in their place relative to a neighboring atom. The easiest way to do so is to do a Taylor Series Expansion197 of the potential energy function around rb. This is:

U (x0

+ x) = U (x0) + µ dx ¶¯x=x0

x + 2!

µ dx2

¶¯x=x0

x2 + ...

(874)

 

 

dU

¯

 

 

 

 

1

 

 

d2U

¯

 

 

 

 

 

 

 

¯

 

 

 

 

 

 

 

 

¯

 

 

 

 

 

 

 

¯

 

 

 

 

 

 

 

 

¯

 

 

At a stable equilibrium point x0 the force vanishes:

 

 

 

 

 

 

 

 

 

 

Fx(x0) =

µ dx ¶¯x=x0

= 0

 

 

 

(875)

 

 

 

 

 

 

dU

¯

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

¯

 

 

 

 

 

 

 

 

so that:

U (x0

+ x) = U (x0) + 2!

¯

¶¯x=x0

x2 + ...

 

(876)

 

µ dx2

 

 

 

 

 

 

1

 

d2U

 

¯

 

 

 

 

 

¯

¯

We can now identify this with the form of a potential energy equation for a mass on a spring in a coordinate system where the equilibrium position of the spring is at x0198:

U (x0

+ x) = U (x0) + 2!

µ dx2

¶¯x=x0

x2

+ ... = U0 +

2 ke x2

(877)

 

1

d2U

¯

 

 

 

1

 

 

 

 

 

 

 

 

¯

 

 

 

 

 

 

 

where

ke =

 

¯

¶¯x=x0

 

 

 

 

(878)

 

µ dx2

 

 

 

 

 

 

 

 

 

d2U

¯

 

 

 

 

 

 

¯

¯

What this means is that any mass at a stable equilibrium point will usually behave like a mass on a spring for motion “close to” the equilibrium point! If pulled a short distance away and released, it will oscillate nearly harmonically around the equilibrium. The terms “usually” and “nearly” can be made more precise by considering the neglected third and higher order derivatives in the Taylor series – as long as they are negligible compared to the second order derivative with its quadratic x2 dependence, the approximation will be a good one.

We have already seen this to be true and used it for the simple and physical pendulum problem, where we used a Taylor series expansion for the force or torque and for the energy and kept only the leading order term – the small angle approximation for sin(θ) and cos(θ). You will see it in homework problems next semester as well, if you continue using this textbook, where you will from time to time use the binomial expansion (the Taylor series for a particular form of polynomial) to

197Wikipedia: http://www.wikipedia.org/wiki/Taylor series expansion.

198Remember, any potential energy function is defined only to within an additive constant, so the constant term

U0 simply sets the scale for the potential energy without a ecting the actual force derived from the potential energy. The force is what makes things happen – it is, in a sense, all that matters.