comparison org/stat-mech.org @ 5:e7185b523c80 tip

Added Began Gibbs formalism.
author Dylan Holmes <ocsenave@gmail.com>
date Mon, 30 Apr 2012 19:10:15 -0500
parents 299a098a30da
children
comparison
equal deleted inserted replaced
4:299a098a30da 5:e7185b523c80
28 Kelvin. 28 Kelvin.
29 29
30 The /fact/ that this has proved possible, and the main technical 30 The /fact/ that this has proved possible, and the main technical
31 ideas involved, are assumed already known to the reader; 31 ideas involved, are assumed already known to the reader;
32 and we are not concerned here with repeating standard material 32 and we are not concerned here with repeating standard material
33 already available in a dozen other textbooks . However 33 already available in a dozen other textbooks. However
34 thermodynamics, in spite of its great successes, firmly established 34 thermodynamics, in spite of its great successes, firmly established
35 for over a century, has also produced a great deal of confusion 35 for over a century, has also produced a great deal of confusion
36 and a long list of \ldquo{}paradoxes\rdquo{} centering mostly 36 and a long list of \ldquo{}paradoxes\rdquo{} centering mostly
37 around the second law and the nature of irreversibility. 37 around the second law and the nature of irreversibility.
38 For this reason and others noted below, we want to dwell here at 38 For this reason and others noted below, we want to dwell here at
569 sample of nitrobenzene is now a thermodynamic system of $(n+1)$ 569 sample of nitrobenzene is now a thermodynamic system of $(n+1)$
570 degrees of freedom. This number may be as large as we please, limited 570 degrees of freedom. This number may be as large as we please, limited
571 only by our patience in constructing the apparatus needed to control 571 only by our patience in constructing the apparatus needed to control
572 or measure all these quantities. 572 or measure all these quantities.
573 573
574 We leave it as an exercise for the reader (Problem 1) to find the most 574 We leave it as an exercise for the reader (Problem 1.1) to find the most
575 general condition on the variables \(\{v_1, q_1, v_2, q_2, \ldots 575 general condition on the variables \(\{v_1, q_1, v_2, q_2, \ldots
576 v_n,q_n\}\) which will ensure that a definite equation of state 576 v_n,q_n\}\) which will ensure that a definite equation of state
577 $f(P,V,t)=0$ is observed in spite of all these new degrees of 577 $f(P,V,t)=0$ is observed in spite of all these new degrees of
578 freedom. The simplest special case of this relation is, evidently, to 578 freedom. The simplest special case of this relation is, evidently, to
579 ground all electrodes, thereby inposing the conditions $v_1 = v_2 = 579 ground all electrodes, thereby inposing the conditions $v_1 = v_2 =
1943 (C_v-C_\ell)/R$ are constants. 1943 (C_v-C_\ell)/R$ are constants.
1944 1944
1945 1945
1946 Equation (1-66) is recognized as an approximate form of the Vapor 1946 Equation (1-66) is recognized as an approximate form of the Vapor
1947 pressure formula 1947 pressure formula
1948 We note that AQ, AV, which appeared first as integration 1948 We note that $A_\ell$, $A_v$, which appeared first as integration
1949 constants for the entropy with no parti cular physical meaning, 1949 constants for the entropy with no particular physical meaning,
1950 now play a role in determining the vapor pressure. 1950 now play a role in determining the vapor pressure.
1951 1951
1952 ** The Second Law: Discussion 1952 ** The Second Law: Discussion
1953 1953
1954 We have emphasized the distinction between the weak and strong forms 1954 We have emphasized the distinction between the weak and strong forms
2047 persuade it to \ldquo{}organize\rdquo{} itself enough to perform useful work 2047 persuade it to \ldquo{}organize\rdquo{} itself enough to perform useful work
2048 against pistons, magnets, gravitational or electric fields, 2048 against pistons, magnets, gravitational or electric fields,
2049 chemical activation energy hills, etc. 2049 chemical activation energy hills, etc.
2050 2050
2051 2051
2052 It was Maxwell himself who first ([[../sources/Maxwell-Heat.pdf][1871]])[fn::See also, the [[http://openlibrary.org/books/OL7243600M/Theory_of_heat][Open Library 2052 It was Maxwell himself who first ([[../sources/Maxwell-Heat.pdf][1871]])[fn::Edit: See also, the [[http://openlibrary.org/books/OL7243600M/Theory_of_heat][Open Library
2053 page]], where you can read and download Maxwell's book in a variety of formats.] suggested such 2053 page]], where you can read and download Maxwell's book in a variety of formats.] suggested such
2054 possibilities, in his invention of the \ldquo{}Maxwell Demon\rdquo{}, 2054 possibilities, in his invention of the \ldquo{}Maxwell Demon\rdquo{},
2055 an imaginary being (or mechanism) which can regulate valves so as to allow 2055 an imaginary being (or mechanism) which can regulate valves so as to allow
2056 fast molecules to pass through a partition in one direction only, 2056 fast molecules to pass through a partition in one direction only,
2057 thus heating up one side at the expense of the other. We could 2057 thus heating up one side at the expense of the other. We could
2073 convinced. This is particularly so when we recall the lessons 2073 convinced. This is particularly so when we recall the lessons
2074 of history; clever experimenters have, over and over again, made 2074 of history; clever experimenters have, over and over again, made
2075 fools of theorists who were too quick to assert that something 2075 fools of theorists who were too quick to assert that something
2076 cannot be done. 2076 cannot be done.
2077 2077
2078
2079 A recent example worth recalling concerns the Overhauser 2078 A recent example worth recalling concerns the Overhauser
2080 effect in magnetic resonance (enhancement of the polarization 2079 effect in magnetic resonance (enhancement of the polarization
2081 of one set of spins by irradiation of another set coupled to them). 2080 of one set of spins by irradiation of another set coupled to them).
2082 When this effect was first proposed, several well-known 2081 When this effect was first proposed, several well-known
2083 authorities on thermodynamics and statistical mechanics ridiculed the 2082 authorities on thermodynamics and statistical mechanics ridiculed the
2085 because it violated the second law of thermodynamics. This 2084 because it violated the second law of thermodynamics. This
2086 incident is a valuable reminder of how little we really understand 2085 incident is a valuable reminder of how little we really understand
2087 the second law, or how to apply it in new situations. 2086 the second law, or how to apply it in new situations.
2088 2087
2089 In this connection, there is a fascinating little gadget 2088 In this connection, there is a fascinating little gadget
2090 known as the Hilsch tube or Vortex tube, in which a jet of 2089 known as the [[http://en.wikipedia.org/wiki/Vortex_tube][Hilsch tube]] or Vortex tube, in which a jet of
2091 compressed air is injected into a pipe at right angles to its 2090 compressed air is injected into a pipe at right angles to its
2092 axis, but off center so that it sets up a rapid rotational 2091 axis, but off center so that it sets up a rapid rotational
2093 motion of the gas. In some manner, this causes a separation of 2092 motion of the gas. In some manner, this causes a separation of
2094 the fast and slow molecules, cold air collecting along the axis 2093 the fast and slow molecules, cold air collecting along the axis
2095 of the tube, and hot air at the walls. On one side of the jet, 2094 of the tube, and hot air at the walls. On one side of the jet,
2096 a diaphragm with a small hole at the center allows only the cold 2095 a diaphragm with a small hole at the center allows only the cold
2097 air to escape, the other side is left open so that the hot air 2096 air to escape, the other side is left open so that the hot air
2098 can escape. The result is that when compressed air at room 2097 can escape. The result is that when compressed air at room
2099 temperature is injected, one can obtain air from the hot side 2098 temperature is injected, one can obtain air from the hot side
2100 at +100^\circ F from the cold side at -70^\circ F, in sufficient quantities 2099 at $+100^\circ$ F from the cold side at $-70^\circ$ F, in sufficient quantities
2101 to be used for quick-freezing small objects, or for cooling 2100 to be used for quick-freezing small objects, or for cooling
2102 photomultiplier tubes [for construction drawings and experi 2101 photomultiplier tubes [for construction drawings and
2103 mental data, see Stong (1960); for a partial thermodynamic 2102 experimental data, see [[http://books.google.com/books?id=yOUWAAAAIAAJ][Stong (1960)]]; for a partial thermodynamic
2104 analysis, see Hilsch (19-47)]. 2103 analysis, see Hilsch (1947)[fn::Edit: Hilsch's paper is entitled /The use of the expansion of gases in
2104 a centrifugal field as a cooling process./]].
2105 2105
2106 Of course, the air could also be cooled by adiabatic expansion 2106 Of course, the air could also be cooled by adiabatic expansion
2107 (i.e., by doing work against a piston); and it appears that 2107 (i.e., by doing work against a piston); and it appears that
2108 the amount of cooling achieved in vortex tubes is comparable to, 2108 the amount of cooling achieved in vortex tubes is comparable to,
2109 but somewhat less than, what could be obtained this way for the 2109 but somewhat less than, what could be obtained this way for the
2189 where confident, dogmatic statements on either side now seem 2189 where confident, dogmatic statements on either side now seem
2190 imprudent. For the present, therefore, we leave it as an open 2190 imprudent. For the present, therefore, we leave it as an open
2191 question whether such machines can or cannot be made. 2191 question whether such machines can or cannot be made.
2192 2192
2193 2193
2194
2195
2196
2197
2198
2199
2200
2201
2202
2203
2204
2205
2206
2207
2208
2209
2210
2211
2212
2213
2214
2215
2216
2217
2218
2219
2220 * COMMENT Use of Jacobians in Thermodynamics
2221
2222 Many students find that thermodynamics, although mathematically almost
2223 trivial, is nevertheless one of the most difficult subjects in their program.
2224 A large part of the blame for this lies in the extremely cumbersome partial
2225 derivative notation. In this chapter we develop a different mathematical
2226 scheme, with which thermodynamic derivations can be carried out more easily,
2227 and which gives a better physical insight into the meaning of thermodynamic
2228 relations.
2229
2230 *** COMMENT Editor's addendum
2231 #+begin_quote
2232 In order to help readers with the Jacobian material that follows, I
2233 have included this section of supplementary material. --- Dylan
2234 #+end_quote}
2235
2236 Suppose your experimental parameters consist of three variables
2237 $X,Y,Z$---say, volume, pressure, and temperature. Then the
2238 physically allowed combinations $\langle x,y,z\rangle$ of $X,Y,Z$
2239 comprise the /(equilibrium) state space/
2240 of your thermodynamic system; the set of these combinations forms a
2241 subset $\Omega$ of $\mathbb{R}^3$. (If there were four experimental
2242 parameters, the state space would be a subset of $\mathbb{R}^4$, and
2243 so on).
2244
2245 You can represent the flux of some physical quantities (such as
2246 heat, entropy, or number of moles) as a vector field spread throughout
2247 $\Omega$, i.e., a function $F:\Omega\rightarrow \mathbb{R}^n$ sending
2248 each state to the value of the vector at that state.
2249 When you trace out different paths through the state space
2250 $\gamma:[a,b]\rightarrow \Omega$, you can measure the net quantity
2251 exchanged by
2252
2253 \begin{equation}
2254 \text{net exchange} = \int_a^b (F\circ \gamma)\cdot \gamma^\prime.
2255 \end{equation}
2256
2257 Some quantities are conservative.
2258
2259 - If the vector field $F$ (representing the flux of a physical
2260 quantity) is in fact the gradient of some function
2261 $\varphi:\Omega\rightarrow \mathbb{R}$, then $F$ is conservative and
2262 $\varphi$ represents the value of the conserved quantity at each state.
2263 - In this case, the value of $\varphi$ is completely determined by
2264 specifying the values of the experimental parameters $X, Y, Z$. In
2265 particular, it doesn't matter by which path the state was reached.
2266
2267
2268 Some physical quantities (such as entropy or number of moles) are
2269 completely determined by your experimental parameters $X, Y, Z$. Others (such as
2270 heat) are not. For those quantities that are,
2271 you have functions $\phi:\Omega\rightarrow \mathbb{R}$ sending each state
2272 to the value of the quantity at that state.
2273
2274
2275
2276 and measure the change in physical
2277 quantities (like entropy or number of moles)
2278
2279
2280 Given your experimental parameters $X,Y,Z$, there may be other
2281 physical quantities (such as entropy or number of moles) which are uniquely
2282 defined by each combination of $\langle x,y,z\rangle$. Stated
2283 mathematically, there is a function $f:\Omega\rightarrow \mathbb{R}$
2284 sending each state to the value of the quantity at that state.
2285
2286
2287
2288 Now, sometimes you would like to use a different coordinate system to
2289 describe the same physical situation.
2290 A /change of variables/ is an
2291 invertible differentiable transformation $g:\mathbb{R}^n\rightarrow
2292 \mathbb{R}^n$---a function with $n$ input components (the $n$ old
2293 variables) and $n$ output components (the $n$ new variables), where
2294 each output component can depend on any number of the input components. For
2295 example, in two dimensions you can freely switch between Cartesian
2296 coordinates and polar coordinates; the familiar transformation is
2297
2298 \(g\langle x, y\rangle \mapsto \langle \sqrt{x^2+y^2}, \arctan{(y/x)}\rangle\)
2299
2300
2301
2302
2303
2304 ** Statement of the Problem
2305 In fields other than thermodynamics , one usually starts out by stating
2306 explicitly what variables shall be considered the independent ones, and then
2307 uses partial derivatives without subscripts, the understanding being that all
2308 independent variables other than the ones explicitly present are held constant
2309 in the differentiation. This convention is used in most of mathematics and
2310 physics without serious misunderstandings. But in thermodynamics, one never
2311 seems to be able to maintain a fixed set of independent variables throughout
2312 a derivation, and it becomes necessary to add one or more subscripts to every
2313 derivative to indicate what is being held constant. The often-needed
2314 transformation from one constant quantity to another involves the
2315 relation
2316
2317 \begin{equation}
2318 \left(\frac{\partial A}{\partial B}\right)_C = \left(\frac{\partial
2319 A}{\partial B}\right)_D + \left(\frac{\partial A}{\partial D}\right)_B \left(\frac{\partial D}{\partial B}\right)_C
2320 \end{equation}
2321
2322 which, although it expresses a fact that is mathematically trivial, assumes
2323 such a complicated form in the usual notation that few people can remember it
2324 long enough to write it down after the book is closed.
2325
2326 As a further comment on notation, we note that in thermodynamics as well
2327 as in mechanics and electrodynamics, our equations are made cumbersome if we
2328 are forced to refer at all times to some particular coordinate system (i.e.,
2329 set of independent variables). In the latter subjects this needless
2330 complication has long since been removed by the use of vector
2331 notation,
2332 which enables us to describe physical relationships without reference to any particular
2333 coordinate system. A similar house-cleaning can be effected for thermodynamics
2334 by use of jacobians, which enable us to express physical relationships without
2335 committing ourselves to any particular set of independent variables.
2336 We have here an interesting example of retrograde progress in science:
2337 for the historical fact is that use of jacobians was the original mathematical
2338 method of thermodynamics. They were used extensively by the founder of modern
2339 thermodynamics, Rudolph Clausius, in his work dating from about 1850. He used
2340 the notation
2341
2342 \begin{equation}
2343 D_{xy} \equiv \frac{\partial^2 Q}{\partial x\partial y} -
2344 \frac{\partial^2 Q}{\partial y \partial x}
2345 \end{equation}
2346
2347
2348 where $Q$ stands, as always, for heat, and $x$, $y$ are any
2349 two thermodynamic quantities. Since $dQ$ is not an exact differential,
2350 $D_{xy}$ is not identically zero. It is understandable that this notation, used in his published works, involved
2351 Clausius in many controversies, which in retrospect appear highly amusing. An
2352 account of some of them may be found in his book (Clausius, 1875). On the
2353 other hand, it is unfortunate that this occurred, because it is probably for
2354 this reason that the quantities $D_{xy}$ went out of general use for many years,
2355 with only few exceptions (See comments at the end of this chapter).
2356 In a footnote in Chapter II of Planck's famous treatise (Planck, 1897), he explains
2357 that he avoids using $dQ$ to represent an infinitesimal quantity of heat, because
2358 that would imply that it is the differential of some quantity $Q$. This in turn
2359 leads to the possibility of many fallacious arguments, all of which amount to
2360 setting $D_{xy}=0$. However, a reading of Clausius‘ works makes it clear that
2361 the quantities $D_{xy}$, when properly used, form the natural medium for discussion
2362 of thermodynamics. They enabled him to carry out certain derivations with a
2363 facility and directness which is conspicuously missing in most recent
2364 expositions. We leave it as an exercise for the reader to prove that $D_{xy}$ is a
2365 jacobian (Problem 2.1).
2366
2367 We now develop a condensed notation in which the algebra of jacobians
2368 may be surveyed as a whole, in a form easy to remember since the abstract
2369 relations are just the ones with which we are familiar in connection with the
2370 properties of commutators in quantum mechanics.
2371
2372 ** Formal Properties of Jacobians[fn::For any function $F:\mathbb{R}^n\rightarrow \mathbb{R}^n$, $F:\langle x_1,\ldots, x_n\rangle \mapsto \langle F_1(x), F_2(x),\ldots F_n(x)\rangle$ we can define the Jacobian matrix of $F$ to be \(JF = \begin{bmatrix}\partial_1{F_1}&\ldots& \partial_n{F_n}\\\vdots&\ddots&\vdots\\\partial_1 F_n & \ldots & \partial_n F_n\\\end{bmatrix}\), and the Jacobian (determinant) of $f$ to be the determinant of this matrix (provided all partial derivatives exist). ]
2373 Consider first a system with only two degrees of freedom. We define
2374
2375 \begin{equation}
2376 [A,B] \equiv \frac{\partial(A,B)}{\partial(x,y)} =
2377 \left|\begin{matrix}\frac{\partial A}{\partial x}& \frac{\partial
2378 A}{\partial y} \\
2379 \frac{\partial B}{\partial x} & \frac{\partial B}{\partial y} \end{matrix}\right|
2380 \end{equation}
2381 where $x$, $y$ are any variables adequate to determine the state of the system.
2382
2383 Since for any change of variables, $x,y \mapsto x^\prime, y^\prime$ we
2384 have
2385
2386 \begin{equation}
2387 \frac{\partial(A,B)}{\partial(x^\prime,y^\prime)} = \frac{\partial(A,B)}{\partial(x,y)}\frac{\partial(x,y)}{\partial(x^\prime,y^\prime)}
2388 \end{equation}
2389
2390 or, in an easily understandable condensed notation,
2391
2392 \begin{equation}
2393 [A,B]^\prime = [A,B][x,y]^\prime
2394 \end{equation}
2395
2396 It follows that any equations that are homogeneous in the jacobians are in
2397 variant in form under "coordinate transformations“, so that we can suppress
2398 the independent variables x, y and carry out derivations without committing
2399 ourselves to any particular set.
2400 The algebra of these symbols is characterized by the following identities
2401 (the comma may be omitted if A, B are single letters). The properties of
2402 antisymmetry, linearity, and composition have the familiar form
2403 In addition we have three cyclic identities, easily proved:
2404 These relations are not all independent; for example, (2—ll) follows from
2405 (2-9) and (2-13).
2406 Putting dC = O in (2-9) , we obtain the rule
2407 by means of which equations are translated from one language to the other.
2408
2409
2410 From it one sees that the transformation law (2-l) now appears as a special
2411 case of the identity (2-11) . Writing for the enthalpy, free energy, and Gibbs
2412 function respectively ,
2413 where U is the internal energy with the property dU = t :35 — P (N, we have as
2414 consequences of (2-13) the relations
2415 The advantages of this notation is shown particularly when we consider the
2416 four Maxwe ll equati ons
2417 Applying (2-14) , we see that each reduces to the single identity
2418
2419
2420 Thus, all of the Maxwell equations are expressions in different "coordinate
2421 systems" of the same basic fact (2-18) , which will receive a physical inter
2422 pretation in Sec. 2.4. In a derivation, such as that of Eq. (1-49) , every
2423 thing that can be gained by using any of the equations (2-17) is already
2424 accomplished by application of the single relation (2-18).
2425 Jacobians which involve the entropy in combinations other than are
2426 related to various specific heats. The heat capacity at constant X is
2427 and, using (2-14) we obtain the identity
2428 C
2429 In the simplest derivations, application of (2-18) or (2—20) is the essential
2430 step.
2431 In his well-known textbook, Zemansky (1943) shows that many of the ele
2432 mentary derivations in thermodynamics may be reduced to application of the
2433 In the above notation these equations are far from obvious and not easy to
2434 remember. Note, however, that the T :38 equations are special cases of the
2435 cyclic identity (2-9) for the sets of variables {TVS}, respectively,
2436 while the energy equation is a consequence of (2-13) and the Maxwell relation:
2437
2438
2439 From (2~l4) we see that this is the energy equation in jacobian notation.
2440 2 .3 Elementary Examples
2441 In a large class of problems, the objective is to express some quantity
2442 of interest, or some condition of interest, in terms of experimentally mea
2443 surable quantities. Therefore, the “sense of direction“ in derivations is
2444 provided by the principle that we want to get rid of any explicit appearance
2445 of the entropy and the various energies U, H, F, G. Thus, if the entropy
2446 appears in the combination [TS], we use the Maxwell relation to replace it
2447 with . If it appears in some other combination , we can use the
2448 identity (2-20) .
2449 Similarly, if combinations such as or [UX] appear, we can use (2-16)
2450 and replace them with
2451 it cannot be eliminated in this way. However, since in phenomenological
2452 thermodynamics the absolute value of the entropy has no meaning, this situa
2453 tion cannot arise in any expression representing a definite physical quantity.
2454 For problems of this simplest type, the jacobian formalism works like a
2455 well-oiled machine, as the following examples show. We denote the isothermal
2456 compressibility, thermal expansion coefficient, and ratio of specific heats
2457 bY K1 5: Y, réspectively:
2458
2459
2460 and note that from (2-27) and (2-28) we have
2461 (2-30)
2462 Several derivatives, chosen at random, are now evaluated in terms of these
2463 quantities:
2464 A more difficult type of problem is the following: We have given a num
2465 ber of quantities and wish to find the general relation, if any, connecting
2466 them. In one sense, the question whether relations exist can be answered
2467
2468
2469 immediately; for any two quantities A, B a necessary and sufficient condition
2470 for the existence of a functional relation A f(B) in a region R is:
2471 = O in R}. In a system of two degrees of freedom it is clear that between
2472 any three quantities A, B, C there is necessarily at least one functional
2473 relation f(A,B,C) = O, as is implied by the identity (2-9) [Problem 2.2] . An
2474 example is the equation of state f(PVT) = O. This , however, is not the type
2475 of relation one usually has in mind. For each choice of A, B, C and each
2476 particular system of two degrees of freedom, some functional relationship
2477 must exist, but in general it will depend on the physical nature of the system
2478 and can be obtained only when one has sufficient information, obtained from
2479 measurement or theory, about the system.
2480 The problem is rather to find those relations between various quantities
2481 which hold generally, regardless of the nature of the particular system.
2482 Mathematically, all such relations are trivial in the sense that they must be
2483 special cases of the basic identities already given. Their physical meaning
2484 may, however, be far from trivial and they may be difficult to find. Note,
2485 for example, that the derivative computed in (2-35) is just the Joule—Thomson
2486 coefficient 11. Suppose the problem had been stated as: "Given the five
2487 quantities V, Cp, 8, determine whether there is a general relation
2488 between them and if so find it." Now, although a repetition of the argument
2489 of (2-35) would be successful in this case, this success must be viewed as a
2490 lucky accident from ‘the standpoint of the problem just formulated. It is not
2491 a general rule for attacking this type of problem because there is no way of
2492 ensuring that the answer will come out in terms of the desired quantities.
2493 To illustrate a general rule of procedure, consider the problem of find
2494 ing a relationship, if any, between iCp, CV, V, T, B, K}. First we write
2495 these quantities in terms of jacobians.
2496
2497
2498 At this point we make a definite choice of some coordinate system. Since
2499 [TP] occurs more often than any other jacobian, we adopt x = T, y = P as the
2500 The variables in jacobians are P, V, T, S, for which (2-11) gives
2501 [PV][TS] + [VT] [PS] + = 0 (2-40)
2502 or, in this case
2503 Substituting the expressions (2-39) into this we obtain
2504 or, rearranging, we have the well—known law
2505 which is now seen as a special case of (2-11).
2506 There are several points to notice in this derivation: (1) no use has
2507 been made of the fact that the quantities T, V were given explicitly; the
2508
2509
2510
2511
2512
2513
2514
2515
2516
2517
2518
2519
2520
2521
2522
2523
2524
2525
2526
2527
2528
2529
2530
2531
2532
2533
2534
2535
2536
2537
2538
2539
2540
2541
2542
2543
2544
2545
2546
2547
2548
2549 * Gibbs Formalism \mdash{} Physical Derivation
2550
2551
2552 In this Chapter we present physical arguments by which the Gibbs
2553 formalism can be derived and justified, deliberately avoiding all use
2554 of probability theory. This will serve to convince us of the /validity/ of Gibbs’ formalism
2555 for the particular applications given by Gibbs, and will give us an intuitive
2556 physical understanding of the second law, as well as the physical meaning of
2557 the Kelvin temperature.
2558
2559 Later on (Chapter 9) we will present an entirely different derivation in
2560 terms of a general problem of statistical estimation, deliberately avoiding
2561 all use of physical ideas, and show that the identical mathematical formalism
2562 emerges. This will serve to convince us of the /generality/ of the
2563 Gibbs methods, and show that their applicability is in no way restricted to equilibrium
2564 problems; or indeed, to physics.
2565
2566
2567 It is interesting to note that most of Gibbs‘ important results were
2568 found independently and almost simultaneously by Einstein (1902); but it is
2569 to Gibbs that we owe the elegant mathematical formulation of the theory. In
2570 the following we show how, from mechanical considerations involving
2571 the microscopic state of a system, the Gibbs rules emerge as a
2572 description of equilibrium macroscopic properties. Having this, we can then reason
2573 backwards, and draw inferences about microscopic conditions from macroscopic experimental
2574 data. We will consider only classical mechanics here; however, none of this
2575 classical theory will have to be unlearned later, because the Gibbs formalism
2576 lost none of its validity through the development of quantum theory. Indeed,
2577 the full power of Gibbs‘ methods has been realized only through their
2578 successful application to quantum theory.
2579
2580 ** COMMENT Review of Classical Mechanics (SICM)
2581 In [[http://mitpress.mit.edu/sicm/][classical mechanics]] a complete description of the state of a system is
2582 given by specifying $n$ coordinates $q_1\ldots q_n$, and the corresponding velocities
2583 $D{q}_1\ldots Dq_n$. The equations of motion are then determined by a Lagrangian function
2584 which in simple mechanical problems is
2585
2586 \begin{equation}
2587 L(t,q(t),Dq(t)) = T - V
2588 \end{equation}
2589
2590
2591 where $T$ and $V$ are the kinetic and potential energies. In problems involving
2592 coupling of particles to an electromagnetic field, the Lagrangian function
2593 takes a more general form, as we will see later. In either case, the
2594 equations of motion are
2595
2596 \begin{equation}
2597 D(\partial_2 L \circ \Gamma[q]) - \partial_1 L \circ \Gamma[q] = 0
2598 \end{equation}
2599
2600 where $\Gamma[q]$ is the function $t\mapsto \langle
2601 t,q(t),Dq(t)\rangle$, and $\partial_i$ denotes the derivative with
2602 respect to the \(i\)th argument ($i=0,1,2,\ldots$).
2603
2604 The advantage of the Lagrangian form (5-2) over the original Newtonian form
2605 (to which it is completely equivalent in simple mechanical problems)
2606
2607 \begin{equation}
2608 D^2 (m\cdot x(t)) = -\partial_1 V \circ \Gamma[x]
2609 \end{equation}
2610
2611 is that (5-2) holds for arbitrary choices of the coordinates $q_i$;
2612 they can include angles, or any other parameters which serve to locate a particle in
2613 space. The Newtonian equations (5-3), on the other hand, hold only when the
2614 $x_i$ are rectangular (cartesian) coordinates of a particle.
2615 Still more convenient for our purposes is the Hamiltonian form of the
2616 equations of motion. Define the momentum \ldquo{}canonically
2617 conjugate\rdquo{} to the
2618 coordinate $q$ by
2619
2620 \begin{equation}
2621 p(t) \equiv \partial_1 L \circ \Gamma[q]
2622 \end{equation}
2623
2624 let $\mathscr{V}(t,q,p) = Dq$, and define a Hamiltonian function $H$ by
2625
2626 \begin{equation}
2627 H(t,q,p) = p \cdot V(t,q,p) - L(t,q, V(t,q,p)
2628 \end{equation}
2629
2630 the notation indicating that after forming the right-hand side of (5-5) the
2631 velocities $\dot{q}_i$ are eliminated mathematically, so that the
2632 Hamiltonian is
2633 expressed as a function of the coordinates and momenta only.
2634
2635 #+begin_quote
2636 ------
2637 *Problem (5.1).* A particle of mass $m$ is located by specifying
2638 $(q_1,q_2,q_3)=(r,\theta,z)$ respectively, where $r$, $\theta$, $z$
2639 are a cylindrical coordinate system
2640 related to the cartesian $x, y, z$ by $x + iy = re^{i\theta}$, $z=z$. The
2641 particle moves in a potential $V(q_1,q_2,q_3)$. Show that the
2642 Hamiltonian in this coordinate system is
2643
2644 \begin{equation}
2645 H = \frac{p_1^2}{2m} + \frac{p_2^2}{2m}+\frac{p_3^2}{2m} + V(q_1,q_2,q_3)
2646 \end{equation}
2647
2648 and discuss the physical meaning of $p_1$, $p_2$, $p_3$.
2649 ------
2650
2651
2652
2653 *Problem (5.2).* Find the Hamiltonian for the same particle, in the spherical
2654 coordinate system $(q_1,q_2,q_3) = (r,\theta,\phi)$ related to the
2655 Cartesian by
2656 $x + iy = r\,\sin{\theta}\,e^{i\phi}$, $z=r\,\cos{\theta}$., and again
2657 discuss the physical meaning of $p_1$, $p_2$, $p_3$ .
2658 ------
2659 #+end_quote
2660
2661
2662
2663
2664
2665
2666
2667
2668
2669
2670 ** Review of Classical Mechanics
2671 In [[http://mitpress.mit.edu/sicm/][classical mechanics]] a complete description of the state of a system is
2672 given by specifying $n$ coordinates $q_1\ldots q_n$, and the corresponding velocities
2673 $\dot{q}_1\ldots \dot{q}_n$. The equations of motion are then determined by a Lagrangian function
2674 which in simple mechanical problems is
2675
2676 \begin{equation}
2677 L(q_i,\dot{q}_i) = T - V
2678 \end{equation}
2679
2680
2681 where $T$ and $V$ are the kinetic and potential energies. In problems involving
2682 coupling of particles to an electromagnetic field, the Lagrangian function
2683 takes a more general form, as we will see later. In either case, the
2684 equations of motion are
2685
2686 \begin{equation}
2687 \frac{\partial L}{\partial q_i} - \frac{d}{dt}\frac{\partial
2688 L}{\partial \dot{q}_i} = 0.
2689 \end{equation}
2690
2691 The advantage of the Lagrangian form (5-2) over the original Newtonian form
2692 (to which it is completely equivalent in simple mechanical problems)
2693
2694 \begin{equation}
2695 m\ddot{x}_i = -\frac{\partial V}{\partial x_i}
2696 \end{equation}
2697
2698 is that (5-2) holds for arbitrary choices of the coordinates $q_i$;
2699 they can include angles, or any other parameters which serve to locate a particle in
2700 space. The Newtonian equations (5-3), on the other hand, hold only when the
2701 $x_i$ are rectangular (cartesian) coordinates of a particle.
2702 Still more convenient for our purposes is the Hamiltonian form of the
2703 equations of motion. Define the momentum \ldquo{}canonically
2704 conjugate\rdquo{} to the
2705 coordinate $q_i$ by
2706
2707 \begin{equation}
2708 p_i \equiv \frac{\partial L}{\partial q_i}
2709 \end{equation}
2710
2711 and a Hamiltonian function $H$ by
2712
2713 \begin{equation}
2714 H(q_1,p_1;\cdots ; q_n,p_n) \equiv \sum_{i=1}^n p\cdot \dot{q}_i -
2715 L(q_1,\ldots, q_n).
2716 \end{equation}
2717
2718 the notation indicating that after forming the right-hand side of (5-5) the
2719 velocities $\dot{q}_i$ are eliminated mathematically, so that the Hamiltonian is ex
2720 pressed as a function of the coordinates and momenta only.
2721
2722 #+begin_quote
2723 ------
2724 *Problem (5.1).* A particle of mass $m$ is located by specifying
2725 $(q_1,q_2,q_3)=(r,\theta,z)$ respectively, where $r$, $\theta$, $z$
2726 are a cylindrical coordinate system
2727 related to the cartesian $x, y, z$ by $x + iy = re^{i\theta}$, $z=z$. The
2728 particle moves in a potential $V(q_1,q_2,q_3)$. Show that the
2729 Hamiltonian in this coordinate system is
2730
2731 \begin{equation}
2732 H = \frac{p_1^2}{2m} + \frac{p_2^2}{2m}+\frac{p_3^2}{2m} + V(q_1,q_2,q_3)
2733 \end{equation}
2734
2735 and discuss the physical meaning of $p_1$, $p_2$, $p_3$.
2736 ------
2737
2738
2739
2740 *Problem (5.2).* Find the Hamiltonian for the same particle, in the spherical
2741 coordinate system $(q_1,q_2,q_3) = (r,\theta,\phi)$ related to the
2742 Cartesian by
2743 $x + iy = r\,\sin{\theta}\,e^{i\phi}$, $z=r\,\cos{\theta}$., and again
2744 discuss the physical meaning of $p_1$, $p_2$, $p_3$ .
2745 ------
2746 #+end_quote
2747
2748 In terms of the Hamiltonian, the equations of motion assume a more
2749 symmetrical form:
2750
2751 \begin{equation}
2752 \cdot{q}_i = \frac{\partial H}{\partial p_i}\qquad \dot{p}_i =
2753 -\frac{\partial H}{\partial q_i}
2754 \end{equation}
2755
2756 of which the first follows from the definition (5-5) , while the second is
2757 equivalent to (5-2).
2758
2759 The above formulation of mechanics holds only when all forces are
2760 conservative; i.e. derivable from a potential energy function
2761 $V(q_1,\ldots q_n)$ , and
2762 in this case the Hamiltonian is numerically equal to the total energy $(T + V)$.
2763 Often, in addition to the conservative forces we have non-conservative ones
2764 which depend on the velocities as well as the coordinates. The Lagrangian
2765 and Hamiltonian form of the equations of motion can be preserved if there
2766 exists a new potential function $M(q_i,\dot{q}_i)$ such that the non-conservative force
2767 acting on coordinate $q_i$ is
2768
2769 \begin{equation}
2770 F_i = \frac{d}{dt}\frac{\partial M}{\partial \dot{q}_i} -
2771 \frac{\partial M}{\partial q_i}
2772 \end{equation}
2773
2774 We then define the Lagrangian as $L \equiv T - V - M$.
2775
2776 #+begin_quote
2777 ------
2778 *Problem (5.3).* Show that the Lagrangian equations of motion (5-2)
2779 are correct with this modified Lagrangian. Find the new momenta and
2780 Hamiltonian. Carry this through explicitly for the case of a charged particle moving in a
2781 time-varying electromagnetic field $\vec{E}(x,y,z,t),
2782 \vec{H}(x,y,z,t)$, for which the
2783 non-conservative force is given by the Lorentz force law,
2784
2785 \(\vec{F} = e\left(\vec{E} + \frac{1}{c}\vec{v} \times \vec{B}\right)\)
2786
2787 # Jaynes wrote \dot{A}. typo?
2788 /Hint:/ Express the potential $M$ in terms of the vector and scalar
2789 potentials of the field \(\vec{A},\phi,\) defined by
2790 \(\vec{B}=\vec{\nabla}\times\vec{A},
2791 \vec{E}=-\vec{\nabla}{\phi}-\frac{1}{c}\vec{A}\).
2792 Notice that, since the potentials are not uniquely determined by $E$, $H$, there is no longer any
2793 unique connection between momentum and velocity; or between the Hamiltonian
2794 and the energy. Nevertheless, the Lagrangian and Hamiltonian equations of
2795 motion still describe the correct physical laws.
2796 -----
2797 #+end_quote
2798 ** Liouville's Theorem
2799 The Hamiltonian form (5-7) is of particular value because of the following
2800 property. Let the coordinates and momenta $(q_1,p_1;\ldots;q_n,p_n)$
2801 be regarded as coordinates of a single point in a $2n$-dimensional /phase space/. This point moves,
2802 by virtue of the equations of motion, with a velocity $v$ whose
2803 components are $\langle \dot{q}_1, \dot{p}_1; \ldots; \dot{q}_n,\dot{p}_n\rangle$.
2804 At each point of phase space there is specified in this way a
2805 particular velocity, and the equations of motion thus generate a continuous
2806 flow pattern in phase space, much like the flow pattern of a fluid in ordinary
2807 space. The divergence of the velocity of this flow pattern is
2808
2809 \begin{eqnarray}
2810 \vec{\nabla}\cdot {v}&=&\sum_{i=1}^n \left[\frac{\partial \dot{q}_i}{\partial q_i} +
2811 \frac{\partial \dot{p}_i}{\partial p_i}\right]\\
2812 &=& \sum_{i=1}^n \left[\frac{\partial^2 H}{\partial q_i \partial
2813 p_i}-\frac{\partial^2 H}{\partial p_i \partial q_i}\right]\\
2814 &=& 0
2815 \end{eqnarray}
2816
2817 # note: this is a sort of Jacobian determinant/commutator|((d_q q_p)(d_q d_p))|
2818
2819 so that the flow in phase space corresponds to that of an [[http://en.wikipedia.org/wiki/Incompressible_flow][incompressible fluid]].
2820 In an incompressible flow, the volume occupied by any given mass of the
2821 fluid remains constant as time goes on and the mass of fluid is carried into
2822 various regions. An exactly analogous property holds in phase space by virtue
2823 of (5-9). Consider at time $t = 0$ any $2n$-dimensional region
2824 $\Gamma_0$ consisting of some possible range of initial conditions
2825 $q_i(O), p_i(O)$ for a mechanical system, as shown in Fig. (5.1). This region has a total phase volume
2826
2827 \begin{equation}
2828 \Omega(0) = \int_{\Gamma_{0}} dq_1\ldots dp_n
2829 \end{equation}
2830
2831 In time t, each point $\langle q_1(O) \ldots p_n(O)\rangle$ of
2832 $\Gamma_0$ is carried, by the equations of
2833 motion, into a new point $\langle q_1(t),\ldots,p_n(t)\rangle$. The totality of all points which
2834 were originally in $\Gamma_0$ now defines a new region $\Gamma_t$ with phase volume
2835
2836 \(\Omega(t) = \int_{\Gamma_{t}} dq_1\ldots dp_n\)
2837
2838 and from (5-9) it can be shown that
2839
2840 \begin{equation}
2841 \Omega(t) = \Omega(0)
2842 \end{equation}
2843
2844 #+caption: Figure 5.1: Volume-conserving flow in phase space.
2845 [[../images/volume-conserved.jpg]]
2846
2847
2848 An equivalent statement is that the Jacobian determinant of the
2849 transformation \( \langle q_1(0), \ldots, p_n(0)\rangle \mapsto
2850 \langle q_1(t), \ldots , p_n(t)\rangle \) is identically equal to
2851 unity:
2852
2853 \begin{equation}
2854 \frac{\partial(q_{1t},\ldots p_{nt})}{\partial(q_{10}\ldots q_{n0})} =
2855 \left|
2856 \begin{matrix}
2857 \frac{\partial q_{1t}}{\partial q_{10}}&\cdots &
2858 \frac{\partial p_{nt}}{\partial q_{10}}\\
2859 \vdots&\ddots&\vdots\\
2860 \frac{\partial q_{1t}}{\partial p_{n0}}&\cdots &
2861 \frac{\partial p_{nt}}{\partial p_{n0}}\\
2862 \end{matrix}\right| = 1
2863 \end{equation}
2864
2865 #+begin_quote
2866 ------
2867 *Problem (5.4).* Prove that (5-9), (5-11), and (5-12) are equivalent statements.
2868 (/Hint:/ See A. I. Khinchin, /Mathematical Foundations of Statistical
2869 Mechanics/, Chapter II.)
2870 ------
2871 #+end_quote
2872
2873 This result was termed by Gibbs the \ldquo{}Principle of conservation
2874 of extension-in—phase\rdquo{}, and is usually referred to nowadays as /Liouville's theorem/.
2875 An important advantage of considering the motion of a system referred to phase
2876 space (coordinates and momenta) instead of the coordinate—velocity space of
2877 the Lagrangian is that in general no such conservation law holds in the latter
2878 space (although they amount to the same thing in the special case where all
2879 the $q_i$ are cartesian coordinates of particles and all forces are conservative
2880 in the sense of Problem 5.3).
2881
2882 #+begin_quote
2883 ------
2884 *Problem (5.5).* Liouville's theorem holds only because of the special form of
2885 the Hamiltonian equations of motion, which makes the divergence (5-9)
2886 identically zero. Generalize it to a mechanical system whose state is defined by a
2887 set of variables $\{x_1,x_2,\ldots,x_n\}$ with equations of motion for
2888 $x_i(t)$:
2889 \begin{equation}
2890 \dot{x}_i(t) = f_i(x_1,\ldots,x_n),\qquad i=1,2,\ldots,n
2891 \end{equation}
2892
2893 The jacobian (5-12) then corresponds to
2894
2895 \begin{equation}
2896 J(x_1(0),\ldots,x_n(0);t) \equiv \frac{\partial[x_1(t),\ldots, x_n(t)]}{\partial[x_1(0),\ldots,x_n(0)]}
2897 \end{equation}
2898
2899 Prove that in place of Liouville's theorem $J=1=\text{const.}$, we now
2900 have
2901
2902 \begin{equation}
2903 $J(t) = J(0)\,\exp\left[\int_0^t \sum_{i=1}^n \frac{\partial
2904 f[x_1(t),\ldots, x_n(t)]}{\partial x_i(t)}
2905 dt\right].
2906 \end{equation}
2907 ------
2908 #+end_quote
2909
2910 ** The Structure Function
2911
2912 One of the essential dynamical properties of a system, which controls its
2913 thermodynamic properties, is the total phase volume compatible with various
2914 experimentally observable conditions. In particular, for a system in which
2915 the Hamiltonian and the energy are the same, the total phase volume below a
2916 certain energy $E$ is
2917
2918 \begin{equation}
2919 \Omega(E) = \int \vartheta[E-H(q_i,p_i)] dq_i\ldots dp_n
2920 \end{equation}
2921 (When limits of integration are unspecified, we understand integration over
2922 all possible values of qi, pi.) In (5-16) , $\vartheta(x)$ is the unit
2923 step function
2924
2925 \begin{equation}
2926 \vartheta(x) \equiv \begin{cases}1,&x>0\\ 0,&x<0\end{cases}
2927 \end{equation}
2928
2929 The differential phase volume, called the /structure function/, is
2930 given by
2931 \begin{equation}
2932 \rho(E) = \frac{d\Omega}{dE} = \int \delta[E-H(q_i,p_i)] dq_1\ldots dp_n
2933 \end{equation}
2934
2935 and it will appear presently that essentially all thermodynamic properties of
2936 the system are known if $\rho(E)$ is known, in its dependence on such parameters
2937 as volume and mole numbers.
2938
2939
2940 Calculation of $\rho(E)$ directly from the definition (5-18) is generally
2941 very difficult. It is much easier to calculate first its [[http://en.wikipedia.org/wiki/Laplace_transform][Laplace transform]],
2942 known as the /partition function/:
2943
2944 \begin{equation}
2945 Z(\beta) = \int_0^\infty e^{-\beta E} \rho(E)\, dE
2946 \end{equation}
2947
2948 where we have assumed that all possible values of energy are positive; this
2949 can always be accomplished for the systems of interest by
2950 appropriately choosing the zero from which we measure energy. In addition, it will develop that
2951 full thermodynamic information is easily extracted directly from the partition
2952 function $Z(\beta)$ , so that calculation of the structure function
2953 $\rho(E)$ is
2954 unnecessary for some purposes.
2955
2956 * COMMENT
2957 Using (1-18) , the partition function can be written as
2958 which is the form most useful for calculation. If the structure function p (E)
2959 is needed, it is then found by the usual rule for inverting a Laplace trans
2960 form:
2961 the path of integration passing to the right of all singularities of Z(B) , as
2962 in Fig. (5.2) -
2963
2964
2965 Figure 5.2. Path of integration in Equation (5-21) .
2966 To illustrate the above relations, we now compute the partition function
2967 and structure function in two simple examples.
2968 Example 1. Perfect monatomic gas. We have N atoms, located by cartesian co
2969 ordinates ql...qN, and denote a particular component (direction in space) by
2970 an index oz, 0: = l, 2, 3. Thus, qia denotes the component of the position
2971 vector of the particle. Similarly, the vector momenta of the particles
2972 are denoted by pl.. .pN, and the individual components by pig. The Hamiltonian
2973 and the potential function u(q) defines the box of volume V containing the
2974
2975
2976 otherwise
2977 The arbitrary additive constant uo, representing the zero from which we
2978 measure our energies, will prove convenient later. The partition function is
2979 then
2980
2981 If N is an even number, the integrand is analytic everywhere in the com
2982 plex except for the pole of order 3N/2 at the origin. If E > Nuo,
2983 the integrand tends to zero very rapidly as GO in the left half—plane
2984 Re(;,%) 5 O. The path of integration may then be extended to a closed one by
2985 addition of an infinite semicircle to the left, as in Fig. (5.3), the integral
2986 over the semicircle vanishing. We can then apply the Cauchy residue theorem
2987 where the closed contour C, illustrated in Fig. (5.4) , encloses the point
2988 z = a once in a counter—clockwise direction, and f(z) is analytic everywhere
2989 on and within C.
2990
2991
2992
2993
2994
2995
2996
2997
2194 * COMMENT Appendix 2998 * COMMENT Appendix
2195 2999
2196 | Generalized Force | Generalized Displacement | 3000 | Generalized Force | Generalized Displacement |
2197 |--------------------+--------------------------| 3001 |--------------------+--------------------------|
2198 | force | displacement | 3002 | force | displacement |