# HG changeset patch # User Dylan Holmes # Date 1335685102 18000 # Node ID 8f3b6dcb9addf1a1eff79540ea3c4a7b7b6d27d6 # Parent afbe1fe19b365863974150fd318b51e0b6f50a6c Transcribed up to section 1.9, Entropy of an Ideal Boltzmann Gas diff -r afbe1fe19b36 -r 8f3b6dcb9add org/stat-mech.org --- a/org/stat-mech.org Sat Apr 28 23:06:48 2012 -0500 +++ b/org/stat-mech.org Sun Apr 29 02:38:22 2012 -0500 @@ -91,7 +91,7 @@ of our present review of the logic underlying classical thermodynamics. -** The Primitive Thermometer. +** The Primitive Thermometer The earliest stages of our story are necessarily speculative, since they took place long @@ -145,7 +145,7 @@ present. -** Thermodynamic Systems. +** Thermodynamic Systems The \ldquo{}thermodynamic systems\rdquo{} which are the objects of our study may be, physically, almost any @@ -226,7 +226,7 @@ of thermodynamics . The most fundamental one is a qualitative rather than quantitative relation, the \ldquo{}zero'th law.\rdquo{} -** Equilibrium; the \ldquo{}Zero‘th Law.\rdquo{} +** Equilibrium; the Zeroth Law It is a common experience that when objects are placed in contact with each other but @@ -390,7 +390,7 @@ by specifying any two of the variables arbitrarily, whereupon the third, and all others we may introduce, are determined. Mathematically, this is expressed by the existence of a functional -relationship of the form[fn:: /Edit./: The set of solutions to an equation +relationship of the form[fn:: The set of solutions to an equation like /f(X,x,t)=/ const. is called a /level set/. Here, Jaynes is saying that the quantities /X/, /x/, and /t/ follow a \ldquo{}functional rule\rdquo{}, so the set of physically allowed combinations of /X/, @@ -753,7 +753,7 @@ \end{equation} is always satisfied. This sort of process is an old story in scientific investigations; although the great theoretician Boltzmann -is said to have remarked: \ldquo{}Elegance is for tailors \rdquo{}, it +is said to have remarked: \ldquo{}Elegance is for tailors\rdquo{}, it remains true that the attempt to reduce equations to the most symmetrical form has often suggested important generalizations of physical laws, and is a great aid to memory. Witness Maxwell's @@ -767,18 +767,18 @@ interactions of the type just studied. The similarity of (1-12) to conservation laws in general may be seen -as follows. Let $A$ be some quantity that is conserved; the $i$th +as follows. Let $A$ be some quantity that is conserved; the \(i\)th system has an amount of it $A_i$. Now when the systems interact such that some $A$ is transferred between them, the amount of $A$ in the -$i$th system is changed by a net amount \(\Delta A_i = (A_i)_{final} - +\(i\)th system is changed by a net amount \(\Delta A_i = (A_i)_{final} - (A_i)_{initial}\); and the fact that there is no net change in the total amount of $A$ is expressed by the equation \(\sum_i \Delta -A_i = 0$. Thus, the law of conservation of matter in a chemical +A_i = 0\). Thus, the law of conservation of matter in a chemical reaction is expressed by \(\sum_i \Delta M_i = 0\), where $M_i$ is the -mass of the $i$th chemical component. +mass of the \(i\)th chemical component. -what is this new conserved quantity? Mathematically, it can be defined -as $Q_i = K_i\cdot M_i cdot t_i; whereupon (1-12) becomes +What is this new conserved quantity? Mathematically, it can be defined +as $Q_i = K_i\cdot M_i \cdot t_i$; whereupon (1-12) becomes \begin{equation} \sum_i \Delta Q_i = 0 @@ -883,7 +883,7 @@ # what is "the specific heat of a gas at constant pressure/volume"? # changed t for temperature below from capital T to lowercase t. -Another failure of the conservation law (1-13) was noted in 1842 by +Another failure of the conservation law (1-13) was [[http://web.lemoyne.edu/~giunta/mayer.html][noted in 1842]] by R. Mayer, a German physician, who pointed out that the data already available showed that the specific heat of a gas at constant pressure, C_p, was greater than at constant volume $C_v$. He surmised that the @@ -905,6 +905,940 @@ conservation law broke down. ** The First Law +Corresponding to the partially valid law of \ldquo{}conservation of +heat\rdquo{}, there had long been known another partially valid +conservation law in mechanics. The principle of conservation of +mechanical energy had been given by Leibnitz in 1693 in noting that, +according to the laws of Newtonian mechanics, one could define +potential and kinetic energy so that in mechanical processes they were +interconverted into each other, the total energy remaining +constant. But this too was not universally valid---the mechanical +energy was conserved only in the absence of frictional forces. In +processes involving friction, the mechanical energy seemed to +disappear. + +So we had a law of conservation of heat, which broke down whenever +mechanical work was done; and a law of conservation of mechanical +energy, which broke down when frictional forces were present. If, as +Mayer had suggested, heat was itself a form of energy, then one had +the possibility of accounting for both of these failures in a new law +of conservation of /total/ (mechanical + heat) energy. On one hand, +the difference $C_p-C_v$ of heat capacities of gases would be +accounted for by the mechanical work done in expansion; on the other +hand, the disappearance of mechanical energy would be accounted for by +the heat produced by friction. + +But to establish this requires more than just suggesting the idea and +illustrating its application in one or two cases --- if this is really +a new conservation law adequate to replace the two old ones, it must +be shown to be valid for /all/ substances and /all/ kinds of +interaction. For example, if one calorie of heat corresponded to $E$ +ergs of mechanical energy in the gas experiments, but to a different +amoun $E^\prime$ in heat produced by friction, then there would be no +universal conservation law. This \ldquo{}first law\rdquo{} of +thermodynamics must therefore take the form: +#+begin_quote +There exists a /universal/ mechanical equivalent of heat, so that the +total (mechanical energy) + (heat energy) remeains constant in all +physical processes. +#+end_quote + +It was James Prescott Joule who provided the [[http://www.chemteam.info/Chem-History/Joule-Heat-1845.html][first experimental data]] +indicating this universality, and providing the first accurate +numerical value of this mechanical equivalent. The calorie had been +defined as the amount of heat required to raise the temperature of one +gram of water by one degree Centigrade (more precisely, to raise it +from 14.5 to 15.5$^\circ C$). Joule measured the heating of a number +of different liquids due to mechanical stirring and electrical +heating, and established that, within the experimental accuracy (about +one percent) a /calorie/ of heat always corresponded to the same +amount of energy. Modern measurements give this numerical value as: 1 +calorie = 4.184 \times 10^7 ergs = 4.184 joules. +# capitalize Joules? I think the convention is to spell them out in lowercase. + +The circumstances of this important work are worth noting. Joule was +in frail health as a child, and was educated by private tutors, +including the chemist, John Dalton, who had formulated the atomic +hypothesis in the early nineteenth century. In 1839, when Joule was +nineteen, his father (a wealthy brewer) built a private laboratory for +him in Manchester, England; and the good use he made of it is shown by +the fact that, within a few months of the opening of this laboratory +(1840), he had completed his first important piece of work, at the +age of twenty. This was his establishment of the law of \ldquo{}Joule +heating,\rdquo{} $P=I^2 R$, due to the electric current in a +resistor. He then used this effect to determine the universality and +numerical value of the mechanical equivalent of heat, reported +in 1843. His mechanical stirring experiments reported in 1849 yielded +the value 1 calorie = 4.154 \times 10^7 ergs, amount 0.7% too low; +this determination was not improved upon for several decades. + +The first law of thermodynamics may then be stated mathematically as +follows: + +#+begin_quote +There exists a state function (i.e., a definite function of the +thermodynamic state) $U$, representing the total energy of any system, +such that in any process in which we change from one equilibrium to +another, the net change in $U$ is given by the difference of the heat +$Q$ supplied to the system, and the mechanical work $W$ done by the +system. +#+end_quote +On an infinitesimal change of state, this becomes + +\begin{equation} +dU = dQ - dW. +\end{equation} + +For a system of two degrees of freedom, defined by pressure $P$, +volume $V$, and temperature $t$, we have $dW = PdV$. Then if we regard +$U$ as a function $U(V,t)$ of volume and temperature, the fact that +$U$ is a state function means that $dU$ must be an exact differential; +i.e., the integral + +\begin{equation} +\int_1^2 dU = U(V_2,t_2) - U(V_1,t_1) +\end{equation} +between any two thermodynamic states must be independent of the +path. Equivalently, the integral $\oint dU$ over any closed cyclic +path (for example, integrate from state 1 to state 2 along path A, +then back to state 1 by a different path B) must be zero. From (1-15), +this gives for any cyclic integral, + +\begin{equation} +\oint dQ = \oint P dV +\end{equation} + +another form of the first law, which states that in any process in +which the system ends in the same thermodynamic state as the initial +one, the total heat absorbed by the system must be equal to the total +work done. + +Although the equations (1-15)-(1-17) are rather trivial +mathematically, it is important to avoid later conclusions that we +understand their exact meaning. In the first place, we have to +understand that we are now measuring heat energy and mechanical energy +in the same units; i.e. if we measured $Q$ in calories and $W$ in +ergs, then (1-15) would of course not be correct. It does +not matter whether we apply Joule's mechanical equivalent of heat +to express $Q$ in ergs, or whether we apply it in the opposite way +to express $U$ and $W$ in calories; each procedure will be useful in +various problems. We can develop the general equations of +thermodynamics +without committing ourselves to any particular units, +but of course all terms in a given equation must be expressed +in the same units. + +Secondly, we have already stressed that the theory being +developed must, strictly speaking, be a theory only of +equilibrium states, since otherwise we have no operational definition +of temperature . When we integrate over any \ldquo{}path\rdquo{} in the $(V-t)$ +plane, therefore, it must be understood that the path of +integration is, strictly speaking, just a /locus of equilibrium +states/; nonequilibrium states cannot be represented by points +in the $(V-t)$ plane. + +But then, what is the relation between path of equilibrium +states appearing in our equations, and the sequence of conditions +produced experimentally when we change the state of a system in +the laboratory? With any change of state (heating, compression, +etc.) proceeding at a finite rate we do not have equilibrium in +termediate states; and so there is no corresponding \ldquo{}path\rdquo{} in +the $(V-t)$ plane ; only the initial and final equilibrium states +correspond to definite points. But if we carry out the change +of state more and more slowly, the physical states produced are +nearer and nearer to equilibrium state. Therefore, we interpret +a path of integration in the $(V-t)$ plane, not as representing +the intermediate states of any real experiment carried out at +a finite rate, but as the /limit/ of this sequence of states, in +the limit where the change of state takes place arbitrarily +slowly. + +An arbitrarily slow process, so that we remain arbitrarily +near to equilibrium at all times, has another important property. +If heat is flowing at an arbitrarily small rate, the temperature +difference producing it must be arbitrarily small, and therefore +an arbitrarily small temperature change would be able to reverse +the direction of heat flow. If the Volume is changing very +slowly, the pressure difference responsible for it must be very +small; so a small change in pressure would be able to reverse +the direction of motion. In other words, a process carried out +arbitrarily slowly is /reversible/; if a system is arbitrarily +close to equilibrium, then an arbitrarily small change in its +environment can reverse the direction of the process. +Recognizing this, we can then say that the paths of integra +tion in our equations are to be interpreted physically as +/reversible paths/ . In practice, some systems (such as gases) +come to equilibrium so rapidly that rather fast changes of +state (on the time scale of our own perceptions) may be quite +good approximations to reversible changes; thus the change of +state of water vapor in a steam engine may be considered +reversible to a useful engineering approximation. + + +** Intensive and Extensive Parameters + +The literature of thermodynamics has long recognized a distinction between two +kinds of quantities that may be used to define the thermodynamic +state. If we imagine a given system as composed of smaller +subsystems, we usually find that some of the thermodynamic variables +have the same values in each subsystem, while others are additive, +the total amount being the sum of the values of each subsystem. +These are called /intensive/ and /extensive/ variables, respectively. +According to this definition, evidently, the mass of a system is +always an extensive quantity, and at equilibrium the temperature +is an intensive ‘quantity. Likewise, the energy will be extensive +provided that the interaction energy between the subsystems can +be neglected. + +It is important to note, however, that in general the terms +\ldquo{}intensive\rdquo{} and \ldquo{}extensive\rdquo{} +so defined cannot be regarded as +establishing a real physical distinction between the variables. +This distinction is, like the notion of number of degrees of +freedom, in part an anthropomorphic one, because it may depend +on the particular kind of subdivision we choose to imagine. For +example, a volume of air may be imagined to consist of a number +of smaller contiguous volume elements. With this subdivision, +the pressure is the same in all subsystems, and is therefore in +tensive; while the volume is additive and therefore extensive . +But we may equally well regard the volume of air as composed of +its constituent nitrogen and oxygen subsystems (or we could re +gard pure hydrogen as composed of two subsystems, in which the +molecules have odd and even rotational quantum numbers +respectively, etc.) . With this kind of subdivision the volume is the +same in all subsystems, while the pressure is the sum of the +partial pressures of its constituents; and it appears that the +roles of \ldquo{}intensive\rdquo{} and \ldquo{}extensive\rdquo{} + have been interchanged. Note that this ambiguity cannot be removed by requiring +that we consider only spatial subdivisions, such that each sub +system has the same local composi tion . For, consider a s tressed +elastic solid, such as a stretched rubber band. If we imagine +the rubber band as divided, conceptually, into small subsystems +by passing planes through it normal to its axis, then the tension +is the same in all subsystems, while the elongation is additive. +But if the dividing planes are parallel to the axis, the elonga +tion is the same in all subsystems, while the tension is +additive; once again, the roles of \ldquo{}extensive\rdquo{} and +\ldquo{}intensive\rdquo{} are +interchanged merely by imagining a different kind of subdivision. +In spite of the fundamental ambiguity of the usual definitions, +the notions of extensive and intensive variables are useful, +and in practice we seem to have no difficulty in deciding +which quantities should be considered intensive. Perhaps the +distinction is better characterized, not by considering +subdivisions at all, but by adopting a different definition, in which +we recognize that some quantities have the nature of a \ldquo{}force\rdquo{} +or \ldquo{}potential\rdquo{}, or some other local physical property, and are +therefore called intensive, while others have the nature of a +\ldquo{}displacement\rdquo{} or a \ldquo{}quantity\rdquo{} of +something (i.e. are proportional to the size of the system), +and are therefore called extensive. Admittedly, this definition is somewhat vague, in a +way that can also lead to ambiguities ; in any event, let us agree +to class pressure, stress tensor, mass density, energy density, +particle density, temperature, chemical potential, angular +velocity, as intensive, while volume, mass, energy, particle +numbers, strain, entropy, angular momentum, will be considered +extensive. + +** The Kelvin Temperature Scale +The form of the first law, +$dU = dQ - dW$, expresses the net energy increment of a system as +the heat energy supplied to it, minus the work done by it. In +the simplest systems of two degrees of freedom, defined by +pressure and volume as the thermodynamic variables, the work done +in an infinitesimal reversible change of state can be separated +into a product $dW = PdV$ of an intensive and an extensive quantity. +Furthermore, we know that the pressure $P$ is not only the +intensive factor of the work; it is also the \ldquo{}potential\rdquo{} +which governs mechanical equilibrium (in this case, equilibrium with respect +to exchange of volume) between two systems; i .e., if they are +separated by a flexible but impermeable membrane, the two systems +will exchange volume $dV_1 = -dV_2$ in a direction determined by the +pressure difference, until the pressures are equalized. The +energy exchanged in this way between the systems is a product +of the form +#+begin_quote +(/intensity/ of something) \times (/quantity/ of something exchanged) +#+end_quote + +Now if heat is merely a particular form of energy that can +also be exchanged between systems, the question arises whether +the quantity of heat energy $dQ$ exchanged in an infinitesimal +reversible change of state can also be written as a product of one +factor which measures the \ldquo{}intensity\rdquo{} of the heat, +times another that represents the \ldquo{}quantity\rdquo{} + of something exchanged between +the systems, such that the intensity factor governs the +conditions of thermal equilibrium and the direction of heat exchange, +in the same way that pressure does for volume exchange. + + +But we already know that the /temperature/ is the quantity +that governs the heat flow (i.e., heat flows from the hotter to +the cooler body until the temperatures are equalized) . So the +intensive factor in $dQ$ must be essentially the temperature. But +our temperature scale is at present still arbitrary, and we can +hardly expect that such a factorization will be possible for all +calibrations of our thermometers. + +The same thing is evidently true of pressure; if instead of +the pressure $P$ as ordinarily defined, we worked with any mono +tonic increasing function $P_1 = P_1 (P)$ we would find that $P_1$ is +just as good as $P$ for determining the direction of volume +exchange and the condition of mechanical equilibrium; but the work +done would not be given by $PdV$; in general, it could not even +be expressed in the form $P_1 \cdot dF(V)$, where $F(V)$ is some function +of V. + + +Therefore we ask: out of all the monotonic functions $t_1(t)$ +corresponding to different empirical temperature scales, is +there one (which we denote as $T(t)$) which forms a \ldquo{}natural\rdquo{} +intensity factor for heat, such that in a reversible change +$dQ = TdS$, where $S(U,V)$ is a new function of the thermodynamic +state? If so, then the temperature scale $T$ will have a great +theoretical advantage, in that the laws of thermodynamics will +take an especially simple form in terms of this particular scale, +and the new quantity $S$, which we call the /entropy/, will be a +kind of \ldquo{}volume\rdquo{} factor for heat. + +We recall that $dQ = dU + PdV$ is not an exact differential; +i.e., on a change from one equilibrium state to another the +integral + +\[\int_1^2 dQ\] + +cannot be set equal to the difference $Q_2 - Q_1$ of values of any +state function $Q(U,V)$, since the integral has different values +for different paths connecting the same initial and final states. +Thus there is no \ldquo{}heat function\rdquo{} $Q(U,V)$, and the notion of +\ldquo{}amount of heat\rdquo{} $Q$ stored in a body has no meaning +(nor does the \ldquo{}amount of work\rdquo{} $W$; +only the total energy is a well-defined quantity). +But we want the entropy $S(U,V)$ to be a definite quantity, +like the energy or volume, and so $dS$ must be an exact differential. +On an infinitesimal reversible change from one equilibrium state +to another, the first law requires that it satisfy[fn:: The first +equality comes from our requirement that $dQ = T\,dS$. The second +equality comes from the fact that $dU = dQ - dW$ (the first law) and +that $dW = PdV$ in the case where the state has two degrees of +freedom, pressure and volume.] + +\begin{equation} +dS(U,V) = \frac{dQ}{T} = \frac{dU}{T} + \frac{P}{T}dV +\end{equation} + +Thus $(1/T)$ must be an /integrating factor/ which converts $dQ$ into +an exact differential [[fn::A differential $M(x,y)dx + +N(x,y)dy$ is called /exact/ if there is a scalar function +$\Phi(x,y)$ such that $M = \frac{\partial \Phi}{\partial x}$ and +$N=\frac{\partial \Phi}{\partial y}$. If there is, \Phi is called the +/potential function/ of the differential, Conceptually, this means +that M(x,y)dx + N(x,y) dy is the derivative of a scalar potential and +so consequently corresponds to a conservative field. + +Even if there is no such potential function +\Phi for the given differential, it is possible to coerce an +inexact differential into an exact one by multiplying by an unknown +function $\mu(x,y)$ (called an /integrating factor/) and requiring the +resulting differential $\mu M\, dx + \mu N\, dy$ to be exact. + +To complete the analogy, here we have the differential $dQ = +dU + PdV$ (by the first law) which is not exact---conceptually, there +is no scalar potential nor conserved quantity corresponding to +$dQ$. We have introduced a new differential $dS = \frac{1}{T}dQ$, and we +are searching for the temperature scale $T(U,V)$ which makes $dS$ +exact (i.e. which makes $S$ correspond to a conserved quantity). This means +that $\frac{1}{T}$ is playing the role of the integrating factor +\ldquo{}\mu\rdquo{} for the differential $dQ$.]] + +Now the question of the existence and properties of +integrating factors is a purely mathematical one, which can be +investigated independently of the properties of any particular +substance. Let us denote this integrating factor for the moment +by $w(U,V) = T^{-1}$; then the first law becomes + +\begin{equation} +dS(U,V) = w dU + w P dV +\end{equation} + +from which the derivatives are + +\begin{equation} +\left(\frac{\partial S}{\partial U}\right)_V = w, \qquad +\left(\frac{\partial S}{\partial V}\right)_U = wP. +\end{equation} + +The condition that $dS$ be exact is that the cross-derivatives be +equal, as in (1-4): + +\begin{equation} +\frac{\partial^2 S}{\partial U \partial V} = \frac{\partial^2 +S}{\partial V \partial U}, +\end{equation} + +or + +\begin{equation} +\left(\frac{\partial w}{\partial V}\right)_U = \left(\frac{\partial +P}{\partial U}\right)_V + P\cdot \left(\frac{\partial w}{\partial U}\right)_V. +\end{equation} + +Any function $w(U,V)$ satisfying this differential equation is an +integrating factor for $dQ$. + +But if $w(U,V)$ is one such integrating factor, which leads +to the new state function $S(U,V)$, it is evident that +$w_1(U,V) \equiv w \cdot f(S)$ is an equally good integrating factor, where +$f(S)$ is an arbitrary function. Use of $w_1$ will lead to a +different state function + +#what's with the variable collision? +\begin{equation} +S_1(U,V) = \int^S f(S) dS +\end{equation} + +The mere conversion of into an exact differential is, therefore, +not enough to determine any unique entropy function $S(U,V)$. +However, the derivative + +\begin{equation} +\left(\frac{dU}{dV}\right)_S = -P +\end{equation} + +is evidently uniquely determined; so also, therefore, is the +family of lines of constant entropy, called /adiabats/, in the +$(U-V)$ plane. But, as (1-24) shows, the numerical value of $S$ on +each adiabat is still completely undetermined. + +In order to fix the relative values of $S$ on different +adiabats we need to add the condition, not yet put into the equations, +that the integrating factor $w(U,V) = T^{-1}$ is to define a new +temperature scale . In other words, we now ask: out of the +infinite number of different integrating factors allowed by +the differential equation (1-23), is it possible to find one +which is a function only of the empirical temperature $t$? If +$w=w(t)$, we can write + +\begin{equation} +\left(\frac{\partial w}{\partial V}\right)_U = \frac{dw}{dt}\left(\frac{\partial +t}{\partial V}\right)_U +\end{equation} +\begin{equation} +\left(\frac{\partial w}{\partial U}\right)_V = \frac{dw}{dt}\left(\frac{\partial +t}{\partial U}\right)_V +\end{equation} + + +and (1-23) becomes +\begin{equation} +\frac{d}{dt}\log{w} = \frac{\left(\frac{\partial P}{\partial +U}\right)_V}{\left(\frac{\partial t}{\partial V}\right)_U-P\left(\frac{\partial t}{\partial U}\right)_V} +\end{equation} + + +which shows that $w$ will be determined to within a multiplicative +factor. + +Is the temperature scale thus defined independent of the +empirical scale from which we started? To answer this, let +$t_1 = t_1(t)$ be any monotonic function which defines a different +empirical temperature scale. In place of (1-28), we then have + +\begin{equation} +\frac{d}{dt_1}\log{w} \quad=\quad \frac{\left(\frac{\partial P}{\partial +U}\right)_V}{\left(\frac{\partial t_1}{\partial V}\right)_U-P\left(\frac{\partial t_1}{\partial U}\right)_V} +\quad = \quad + \frac{\left(\frac{\partial P}{\partial +U}\right)_V}{\frac{dt_1}{dt}\left[ \left(\frac{\partial t}{\partial +V}\right)_U-P\left(\frac{\partial t}{\partial U}\right)_V\right]}, +\end{equation} +or +\begin{equation} +\frac{d}{dt_1}\log{w_1} = \frac{dt}{dt_1}\frac{d}{dt}\log{w} +\end{equation} + +which reduces to $d \log{w_1} = d \log{w}$, or +\begin{equation} +w_1 = C\cdot w +\end{equation} + +Therefore, integrating factors derived from whatever empirical +temperature scale can differ among themselves only by a +multiplicative factor. For any given substance, therefore, except +for this factor (which corresponds just to our freedom to choose +the size of the units in which we measure temperature), there is +only /one/ temperature scale $T(t) = 1/w$ with the property that +$dS = dQ/T$ is an exact differential. + +To find a feasible way of realizing this temperature scale +experimentally, multiply numerator and denominator of the right +hand side of (1-28) by the heat capacity at constant volume, +$C_V^\prime = (\partial U/\partial t) V$, the prime denoting that +it is in terms of the empirical temperature scale $t$. +Integrating between any two states denoted 1 and 2, we have + +\begin{equation} +\frac{T_1}{T_2} = \exp\left\{\int_{t_1}^{t_2} +\frac{\left(\frac{\partial P}{\partial t}\right)_V dt}{P - C_V^\prime +\left(\frac{\partial t}{\partial V}\right)_U} \right\} +\end{equation} + +If the quantities on the right-hand side have been determined +experimentally, then a numerical integration yields the ratio +of Kelvin temperatures of the two states. + +This process is particularly simple if we choose for our +system a volume of gas with the property found in Joule's famous +expansion experiment; when the gas expands freely into a vacuum +(i.e., without doing work, or $U = \text{const.}$), there is no change in +temperature. Real gases when sufficiently far from their condensation +points are found to obey this rule very accurately. +But then + +\begin{equation} +\left(\frac{dt}{dV}\right)_U = 0 +\end{equation} + +and on a change of state in which we heat this gas at constant +volume, (1-31) collapses to + +\begin{equation} +\frac{T_1}{T_2} = \exp\left\{\int_{t_1}^{t_2} +\frac{1}{P}\left(\frac{\partial P}{\partial t}\right)_V dt\right\} = \frac{P_2}{P_1}. +\end{equation} + +Therefore, with a constant-volume ideal gas thermometer, (or more +generally, a thermometer using any substance obeying (1-32) and +held at constant volume), the measured pressure is directly +proportional to the Kelvin temperature. + +For an imperfect gas, if we have measured $(\partial t /\partial +V)_U$ and $C_V^\prime$, Eq. (1-31) determines the necessary +corrections to (1-33). However, an alternative form of (1-31), in +which the roles of pressure and volume are interchanged, proves to be +more convenient for experimental determinations. To derive it, introduce the +enthalpy function + +\begin{equation}H = U + PV\end{equation} + +with the property + +\begin{equation} +dH = dQ + VdP +\end{equation} + +Equation (1-19) then becomes + +\begin{equation} +dS = \frac{dH}{T} - \frac{V}{T}dP. +\end{equation} + +Repeating the steps (1-20) to (1-31) of the above derivation +starting from (1-36) instead of from (1-19), we arrive at + +\begin{equation} +\frac{T_2}{T_1} = \exp\left\{\int_{t_1}^{t_2} +\frac{\left(\frac{dV}{dt}\right)_P dt}{V + C_P^\prime +\left(\frac{\partial t}{\partial P}\right)_H}\right\} +\end{equation} + +or + +\begin{equation} +\frac{T_2}{T_1} = \exp\left\{\frac{\alpha^\prime +dt}{1+\left(C_P^\prime \cdot \mu^\prime / V\right)}\right\} +\end{equation} + +where +\begin{equation} +\alpha^\prime \equiv \frac{1}{V}\left(\frac{\partial V}{\partial t}\right)_P +\end{equation} +is the thermal expansion coefficient, +\begin{equation} +C_P^\prime \equiv \left(\frac{\partial H}{\partial t}\right)_P +\end{equation} +is the heat capacity at constant pressure, and +\begin{equation} +\mu^\prime \equiv \left(\frac{dt}{dP}\right)_H +\end{equation} + +is the coefficient measured in the Joule-Thompson porous plug +experiment, the primes denoting again that all are to be measured +in terms of the empirical temperature scale $t$. +Since $\alpha^\prime$, $C_P^\prime$, $\mu^\prime$ are all +easily measured in the laboratory, Eq. (1-38) provides a +feasible way of realizing the Kelvin temperature scale experimentally, +taking account of the imperfections of real gases. +For an account of the work of Roebuck and others based on this +relation, see [[http://books.google.com/books?id=KKJKAAAAMAAJ][Zemansky (1943)]]; pp. 252-255. + +Note that if $\mu^\prime = O$ and we heat the gas at constant +pressure, (1-38) reduces to + +\begin{equation} +\frac{T_2}{T_1} = \exp\left\{ \int_{t_1}^{t_2} +\frac{1}{V}\left(\frac{\partial V}{\partial t}\right)_P dt \right\} = \frac{V_2}{V_1} +\end{equation} + +so that, with a constant-pressure gas thermometer using a gas for +which the Joule-Thomson coefficient is zero, the Kelvin temperature is +proportional to the measured volume. + +Now consider another empirical fact, [[http://en.wikipedia.org/wiki/Boyle%27s_law][Boyle's law]]. For gases +sufficiently far from their condensation points---which is also +the condition under which (1-32) is satisfied---Boyle found that +the product $PV$ is a constant at any fixed temperature. This +product is, of course proportional to the number of moles $n$ +present, and so Boyle's equation of state takes the form + +\begin{equation}PV = n \cdot f(t)\end{equation} + +where f(t) is a function that depends on the particular empirical +temperature scale used. But from (1-33) we must then have +$f(t) = RT$, where $R$ is a constant, the universal gas constant whose +numerical value (1.986 calories per mole per degree K) , depends +on the size of the units in which we choose to measure the Kelvin +temperature $T$. In terms of the Kelvin temperature, the ideal gas +equation of state is therefore simply + +\begin{equation} +PV = nRT +\end{equation} + + +The relations (1-32) and (1-44) were found empirically, but +with the development of thermodynamics one could show that they +are not logically independent. In fact, all the material needed +for this demonstration is now at hand, and we leave it as an +exercise for the reader to prove that Joule‘s relation (1-32) is +a logical consequence of Boyle's equation of state (1-44) and the +first law. + + +Historically, the advantages of the gas thermometer were +discovered empirically before the Kelvin temperature scale was +defined; and the temperature scale \theta defined by + +\begin{equation} +\theta = \lim_{P\rightarrow 0}\left(\frac{PV}{nR}\right) +\end{equation} + +was found to be convenient, easily reproducible, and independent +of the properties of any particular gas. It was called the +/absolute/ temperature scale; and from the foregoing it is clear +that with the same choice of the numerical constant $R$, the +absolute and Kelvin scales are identical. + + +For many years the unit of our temperature scale was the +Centigrade degree, so defined that the difference $T_b - T_f$ of +boiling and freezing points of water was exactly 100 degrees. +However, improvements in experimental techniques have made another +method more reproducible; and the degree was redefined by the +Tenth General Conference of Weights and Measures in 1954, by +the condition that the triple point of water is at 273.l6^\circ K, +this number being exact by definition. The freezing point, 0^\circ C, +is then 273.15^\circ K. This new degree is called the Celsius degree. +For further details, see the U.S. National Bureau of Standards +Technical News Bulletin, October l963. + + +The appearance of such a strange and arbitrary-looking +number as 273.16 in the /definition/ of a unit is the result of +the historical development, and is the means by which much +greater confusion is avoided. Whenever improved techniques make +possible a new and more precise (i.e., more reproducible) +definition of a physical unit, its numerical value is of course chosen +so as to be well inside the limits of error with which the old +unit could be defined. Thus the old Centigrade and new Celsius +scales are the same, within the accuracy with which the +Centigrade scale could be realized; so the same notation, ^\circ C, is used +for both . Only in this way can old measurements retain their +value and accuracy, without need of corrections every time a +unit is redefined. + +#capitalize Joules? +Exactly the same thing has happened in the definition of +the calorie; for a century, beginning with the work of Joule, +more and more precise experiments were performed to determine +the mechanical equivalent of heat more and more accurately . But +eventually mechanical and electrical measurements of energy be +came far more reproducible than calorimetric measurements; so +recently the calorie was redefined to be 4.1840 Joules, this +number now being exact by definition. Further details are given +in the aforementioned Bureau of Standards Bulletin. + + +The derivations of this section have shown that, for any +particular substance, there is (except for choice of units) only +one temperature scale $T$ with the property that $dQ = TdS$ where +$dS$ is the exact differential of some state function $S$. But this +in itself provides no reason to suppose that the /same/ Kelvin +scale will result for all substances; i.e., if we determine a +\ldquo{}helium Kelvin temperature\rdquo{} and a +\ldquo{}carbon dioxide Kelvin temperature\rdquo{} by the measurements +indicated in (1-38), and choose the units so that they agree numerically at one point, will they then +agree at other points? Thus far we have given no reason to +expect that the Kelvin scale is /universal/, other than the empirical +fact that the limit (1-45) is found to be the same for all gases. +In section 2.0 we will see that this universality is a conse +quence of the second law of thermodynamics (i.e., if we ever +find two substances for which the Kelvin scale as defined above +is different, then we can take advantage of this to make a +perpetual motion machine of the second kind). + + +Usually, the second law is introduced before discussing +entropy or the Kelvin temperature scale. We have chosen this +unusual order so as to demonstrate that the concepts of entropy +and Kelvin temperature are logically independent of the second +law; they can be defined theoretically, and the experimental +procedures for their measurement can be developed, without any +appeal to the second law. From the standpoint of logic, there +fore, the second law serves /only/ to establish that the Kelvin +temperature scale is the same for all substances. + + +** COMMENT Entropy of an Ideal Boltzmann Gas + +At the present stage we are far from understanding the physical +meaning of the function $S$ defined by (1-19); but we can investigate its mathematical +form and numerical values. Let us do this for a system con +sisting cf n moles of a substance which obeys the ideal gas +equation of state +and for which the heat capacity at constant volume CV is a +constant. The difference in entropy between any two states (1) +and (2) is from (1-19), + + +where we integrate over any reversible path connecting the two +states. From the manner in which S was defined, this integral +must be the same whatever path we choose. Consider, then, a +path consisting of a reversible expansion at constant tempera +ture to a state 3 which has the initial temperature T, and the +.L ' "'1 final volume V2; followed by heating at constant volume to the final temperature T2. Then (1-47) becomes +3 2 I If r85 - on - db — = d — -4 S2 51 J V [aT]v M (1 8) +1 3 +To evaluate the integral over (1 +3) , note that since +dU = T :15 — P dV, the Helmholtz free energy function F E U — TS +has the property dF = --S - P 61V; and of course is an exact +differential since F is a definite state function. The condition +that dF be exact is, analogous to (1-22), +which is one of the Maxwell relations, discussed further in +where CV is the molar heat capacity at constant volume. Collec +ting these results, we have +3 +l 3 +1 nR log(V2/V1) + nCV log(T2/Tl) (1-52) +since CV was assumed independent of T. Thus the entropy function +must have the form +S(n,V,T) = nR log V + n CV log T + (const.) (l~53) + + +From the derivation, the additive constant must be independent +of V and T; but it can still depend on n. We indicate this by +writing +where f (n) is a function not determined by the definition (1-47). +The form of f (n) is , however, restricted by the condition that +the entropy be an extensive quantity; i .e . , two identical systems +placed together should have twice the entropy of a single system; +Substituting (l—-54) into (1-55), we find that f(n) must satisfy +To solve this, one can differentiate with respect to q and set +q = 1; we then obtain the differential equation +n f ' (n) — f (n) + Rn = 0 (1-57) +which is readily solved; alternatively, just set n = 1 in (1-56) +and replace q by n . By either procedure we find +f (n) = n f (1) — Rn log n . (1-58) +As a check, it is easily verified that this is the solution of +where A E f (l) is still an arbitrary constant, not determined +by the definition (l—l9) , or by the condition (l-55) that S be +extensive. However, A is not without physical meaning; we will +see in the next Section that the vapor pressure of this sub +stance (and more generally, its chemical potential) depends on +A. Later, it will appear that the numerical value of A involves +Planck's constant, and its theoretical determination therefore +requires quantum statistics . +We conclude from this that, in any region where experi +mentally CV const. , and the ideal gas equation of state is + + +obeyed, the entropy must have the form (1-59) . The fact that +classical statistical mechanics does not lead to this result, +the term nR log (l/n) being missing (Gibbs paradox) , was his +torically one of the earliest clues indicating the need for the +quantum theory. +In the case of a liquid, the volume does not change appre +ciably on heating, and so d5 = n CV dT/T, and if CV is indepen +dent of temperature, we would have in place of (1-59) , +where Ag is an integration constant, which also has physical +meaning in connection with conditions of equilibrium between +two different phases. +1.1.0 The Second Law: Definition. Probably no proposition in +physics has been the subject of more deep and sus tained confusion +than the second law of thermodynamics . It is not in the province +of macroscopic thermodynamics to explain the underlying reason +for the second law; but at this stage we should at least be able +to state this law in clear and experimentally meaningful terms. +However, examination of some current textbooks reveals that, +after more than a century, different authors still disagree as +to the proper statement of the second law, its physical meaning, +and its exact range of validity. +Later on in this book it will be one of our major objectives +to show, from several different viewpoints , how much clearer and +simpler these problems now appear in the light of recent develop +ments in statistical mechanics . For the present, however, our +aim is only to prepare the way for this by pointing out exactly +what it is that is to be proved later. As a start on this at +tempt, we note that the second law conveys a certain piece of +informations about the direction in which processes take place. +In application it enables us to predict such things as the final +equilibrium state of a system, in situations where the first law +alone is insufficient to do this. +A concrete example will be helpful. We have a vessel +equipped with a piston, containing N moles of carbon dioxide. + + +The system is initially at thermal equilibrium at temperature To, volume V0 and pressure PO; and under these conditions it contains +n moles of CO2 in the vapor phase and moles in the liquid +phase . The system is now thermally insulated from its surround +ings, and the piston is moved rapidly (i.e. , so that n does not +change appreciably during the motion) so that the system has a +new volume Vf; and immediately after the motion, a new pressure +PI . The piston is now held fixed in its new position , and the +system allowed to come once more to equilibrium. During this +process, will the CO2 tend to evaporate further, or condense further? What will be the final equilibrium temperature Teq, the final pressure PeCE , and final value of n eq? +It is clear that the firs t law alone is incapable of answering +these questions; for if the only requirement is conservation of +energy, then the CO2 might condense , giving up i ts heat of vapor +ization and raising the temperature of the system; or it might +evaporate further, lowering the temperature. Indeed, all values +of neq in O i neq i N would be possible without any violation of the first law. In practice, however, this process will be found +to go in only one direction and the sys term will reach a definite +final equilibrium state with a temperature, pressure, and vapor +density predictable from the second law. +Now there are dozens of possible verbal statements of the +second law; and from one standpoint, any statement which conveys +the same information has equal right to be called "the second +law." However, not all of them are equally direct statements of +experimental fact, or equally convenient for applications, or +equally general; and it is on these grounds that we ought to +choose among them . +Some of the mos t popular statements of the s econd law be +long to the class of the well-—known "impossibility" assertions ; +i.e. , it is impossible to transfer heat from a lower to a higher +temperature without leaving compensating changes in the rest of +the universe , it is imposs ible to convert heat into useful work +without leaving compensating changes, it is impossible to make +a perpetual motion machine of the second kind, etc. + + +Suoh formulations have one clear logical merit; they are +stated in such a way that, if the assertion should be false, a +single experiment would suffice to demonstrate that fact conclu +sively. It is good to have our principles stated in such a +clear, unequivocal way. +However, impossibility statements also have some disadvan +tages . In the first place, their_ are not, and their very +nature cannot be, statements of eiperimental fact. Indeed, we +can put it more strongly; we have no record of anyone having +seriously tried to do any of the various things which have been +asserted to be impossible, except for one case which actually +succeeded‘. In the experimental realization of negative spin +temperatures , one can transfer heat from a lower to a higher +temperature without external changes; and so one of the common +impossibility statements is now known to be false [for a clear +discussion of this, see the article of N. F . Ramsey (1956) ; +experimental details of calorimetry with negative temperature +spin systems are given by Abragam and Proctor (1958) ] . +Finally, impossibility statements are of very little use in +applications of thermodynamics; the assertion that a certain kind +of machine cannot be built, or that a -certain laboratory feat +cannot be performed, does not tell me very directly whether my +carbon dioxide will condense or evaporate. For applications, +such assertions must first be converted into a more explicit +mathematical form. +For these reasons, it appears that a different kind of +statement of the second law will be, not necessarily more +"correct,” but more useful in practice. Now both Clausius (3.875) +and Planck (1897) have laid great stress on their conclusion +that the most general statement, and also the most immediately +useful in applications, is simply the existence of a state +function, called the entropy, which tends to increase. More +precisely: in an adiabatic change of state, the entropy of +a system may increase or may remain constant, but does not +decrease. In a process involving heat flow to or from the +system, the total entropy of all bodies involved may increase + + +or may remain constant; but does not decrease; let us call this +the “weak form" of the second law. +The weak form of the second law is capable of answering the +first question posed above; thus the carbon dioxide will evapo +rate further if , and only if , this leads to an increase in the +total entropy of the system . This alone , however , is not enough +to answer the second question; to predict the exact final equili +brium state, we need one more fact. +The strong form of the second law is obtained by adding the +further assertion that the entropy not only “tends" to increase; +in fact it will increase, to the maximum value permitted E2 the +constraints imposed.* In the case of the carbon dioxide, these +constraints are: fixed total energy (first law) , fixed total +amount of carbon dioxide , and fixed position of the piston . The +final equilibrium state is the one which has the maximum entropy +compatible with these constraints , and it can be predicted quan +titatively from the strong form of the second law if we know, +from experiment or theory, the thermodynamic properties of carbon +dioxide (i .e . , heat capacity , equation of state , heat of vapor +ization) . +To illus trate this , we set up the problem in a crude ap +proximation which supposes that (l) in the range of conditions +of interest, the molar heat capacity CV of the vapor, and C2 of +the liquid, and the molar heat of vaporization L, are all con +stants, and the heat capacities of cylinder and piston are neg +ligible; (2) the liquid volume is always a small fraction of the +total V, so that changes in vapor volume may be neglected; (3) the +vapor obeys the ideal gas equation of state PV = nRT. The in +ternal energy functions of liquid and vapor then have the form +UPb = + A} (1-61) +T T U = n‘ C '1‘ A + L] (1-62) +v , v +where A is a constant which plays no role in the problem. The +appearance of L in (1-62) recognizes that the zero from which we +*Note , however , that the second law has nothing to say about how rapidly this approach to equilibrium takes place. + + +measure energy of the vapor is higher than that of the liquid by +the energy L necessary to form the vapor. On evaporation of dn +moles of liquid, the total energy increment is (ill = + dUV= O, +or +[n CV [(CV — CQ)T + = O (l—63) +which is the constraint imposed by the first law. As we found +previously (l~59) , (1-60) the entropies of vapor and liquid are +given by +S = n [C 1n T + R ln (V/n) + A ] (1-64) +v v v +where AV, ASL are the constants of integration discussed in the +Si +last Section. +We leave it as an exercise for the reader to complete the +derivation from this point , and show that the total entropy +S = 82 + SV is maximized subject to the constraint (1-6 3) , when +R +the values 11 , T are related by +eq eq +Equation (1-66) is recognized as an approximate form of the Vapor +pressure formula . +We note that AQ, AV, which appeared first as integration +constants for the entropy with no parti cular physical meaning , +now play a role in determining the vapor pressure. +l.ll The Second Law: Discussion. We have emphasized the dis +tinction between the weak and strong forms of the second law +because (with the exception of Boltzmann ' s original unsuccessful +argument based on the H—theorem) , most attempts to deduce the +second law from statis tical mechanics have considered only the +weak form; whereas it is evidently the strong form that leads +to definite quantitative predictions, and is therefore needed