ocsenave@0: #+TITLE: Statistical Mechanics ocsenave@0: #+AUTHOR: E.T. Jaynes; edited by Dylan Holmes ocsenave@0: #+EMAIL: rlm@mit.edu ocsenave@0: #+KEYWORDS: statistical mechanics, thermostatics, thermodynamics, temperature, paradoxes, Jaynes ocsenave@0: #+SETUPFILE: ../../aurellem/org/setup.org ocsenave@0: #+INCLUDE: ../../aurellem/org/level-0.org ocsenave@0: #+MATHJAX: align:"left" mathml:t path:"http://www.aurellem.org/MathJax/MathJax.js" ocsenave@0: ocsenave@0: # "extensions/eqn-number.js" ocsenave@0: ocsenave@0: #+begin_quote ocsenave@0: *Note:* The following is a typeset version of ocsenave@0: [[../sources/stat.mech.1.pdf][this unpublished book draft]], written by [[http://en.wikipedia.org/wiki/Edwin_Thompson_Jaynes][E.T. Jaynes]]. I have only made ocsenave@0: minor changes, e.g. to correct typographical errors, add references, or format equations. The ocsenave@0: content itself is intact. --- Dylan ocsenave@0: #+end_quote ocsenave@0: ocsenave@0: * Development of Thermodynamics ocsenave@0: Our first intuitive, or \ldquo{}subjective\rdquo{} notions of temperature ocsenave@0: arise from the sensations of warmth and cold associated with our ocsenave@0: sense of touch . Yet science has been able to convert this qualitative ocsenave@0: sensation into an accurately defined quantitative notion, ocsenave@0: which can be applied far beyond the range of our direct experience. ocsenave@0: Today an experimentalist will report confidently that his ocsenave@0: spin system was at a temperature of 2.51 degrees Kelvin; and a ocsenave@0: theoretician will report with almost as much confidence that the ocsenave@0: temperature at the center of the sun is about \(2 \times 10^7\) degrees ocsenave@0: Kelvin. ocsenave@0: ocsenave@0: The /fact/ that this has proved possible, and the main technical ocsenave@0: ideas involved, are assumed already known to the reader; ocsenave@0: and we are not concerned here with repeating standard material ocsenave@0: already available in a dozen other textbooks . However ocsenave@0: thermodynamics, in spite of its great successes, firmly established ocsenave@0: for over a century, has also produced a great deal of confusion ocsenave@0: and a long list of \ldquo{}paradoxes\rdquo{} centering mostly ocsenave@0: around the second law and the nature of irreversibility. ocsenave@0: For this reason and others noted below, we want to dwell here at ocsenave@0: some length on the /logic/ underlying the development of ocsenave@0: thermodynamics . Our aim is to emphasize certain points which, ocsenave@0: in the writer's opinion, are essential for clearing up the ocsenave@0: confusion and resolving the paradoxes; but which are not ocsenave@0: sufficiently ernphasized---and indeed in many cases are ocsenave@0: totally ignored---in other textbooks. ocsenave@0: ocsenave@0: This attention to logic ocsenave@0: would not be particularly needed if we regarded classical ocsenave@0: thermodynamics (or, as it is becoming called increasingly, ocsenave@0: /thermostatics/) as a closed subject, in which the fundamentals ocsenave@0: are already completely established, and there is ocsenave@0: nothing more to be learned about them. A person who believes ocsenave@0: this will probably prefer a pure axiomatic approach, in which ocsenave@0: the basic laws are simply stated as arbitrary axioms, without ocsenave@0: any attempt to present the evidence for them; and one proceeds ocsenave@0: directly to working out their consequences. ocsenave@0: However, we take the attitude here that thermostatics, for ocsenave@0: all its venerable age, is very far from being a closed subject, ocsenave@0: we still have a great deal to learn about such matters as the ocsenave@0: most general definitions of equilibrium and reversibility, the ocsenave@0: exact range of validity of various statements of the second and ocsenave@0: third laws, the necessary and sufficient conditions for ocsenave@0: applicability of thermodynamics to special cases such as ocsenave@0: spin systems, and how thermodynamics can be applied to such ocsenave@0: systems as putty or polyethylene, which deform under force, ocsenave@0: but retain a \ldquo{}memory\rdquo{} of their past deformations. ocsenave@0: Is it possible to apply thermodynamics to a system such as a vibrating quartz crystal? We can by ocsenave@0: no means rule out the possibility that still more laws of ocsenave@0: thermodynamics exist, as yet undiscovered, which would be ocsenave@0: useful in such applications. ocsenave@0: ocsenave@0: ocsenave@0: It is only by careful examination of the logic by which ocsenave@0: present thermodynamics was created, asking exactly how much of ocsenave@0: it is mathematical theorems, how much is deducible from the laws ocsenave@0: of mechanics and electrodynamics, and how much rests only on ocsenave@0: empirical evidence, how compelling is present evidence for the ocsenave@0: accuracy and range of validity of its laws; in other words, ocsenave@0: exactly where are the boundaries of present knowledge, that we ocsenave@0: can hope to uncover new things. Clearly, much research is still ocsenave@0: needed in this field, and we shall be able to accomplish only a ocsenave@0: small part of this program in the present review. ocsenave@0: ocsenave@0: ocsenave@0: It will develop that there is an astonishingly close analogy ocsenave@0: with the logic underlying statistical theory in general, where ocsenave@0: again a qualitative feeling that we all have (for the degrees of ocsenave@0: plausibility of various unproved and undisproved assertions) must ocsenave@0: be convertefi into a precisely defined quantitative concept ocsenave@0: (probability). Our later development of probability theory in ocsenave@0: Chapter 6,7 will be, to a considerable degree, a paraphrase ocsenave@0: of our present review of the logic underlying classical ocsenave@0: thermodynamics. ocsenave@0: ocsenave@0: ** The Primitive Thermometer. ocsenave@0: ocsenave@0: The earliest stages of our ocsenave@0: story are necessarily speculative, since they took place long ocsenave@0: before the beginnings of recorded history. But we can hardly ocsenave@0: doubt that primitive man learned quickly that objects exposed ocsenave@0: to the sun‘s rays or placed near a fire felt different from ocsenave@0: those in the shade away from fires; and the same difference was ocsenave@0: noted between animal bodies and inanimate objects. ocsenave@0: ocsenave@0: ocsenave@0: As soon as it was noted that changes in this feeling of ocsenave@0: warmth were correlated with other observable changes in the ocsenave@0: behavior of objects, such as the boiling and freezing of water, ocsenave@0: cooking of meat, melting of fat and wax, etc., the notion of ocsenave@0: warmth took its first step away from the purely subjective ocsenave@0: toward an objective, physical notion capable of being studied ocsenave@0: scientifically. ocsenave@0: ocsenave@0: One of the most striking manifestations of warmth (but far ocsenave@0: from the earliest discovered) is the almost universal expansion ocsenave@0: of gases, liquids, and solids when heated . This property has ocsenave@0: proved to be a convenient one with which to reduce the notion ocsenave@0: of warmth to something entirely objective. The invention of the ocsenave@0: /thermometer/, in which expansion of a mercury column, or a gas, ocsenave@0: or the bending of a bimetallic strip, etc. is read off on a ocsenave@0: suitable scale, thereby giving us a /number/ with which to work, ocsenave@0: was a necessary prelude to even the crudest study of the physical ocsenave@0: nature of heat. To the best of our knowledge, although the ocsenave@0: necessary technology to do this had been available for at least ocsenave@0: 3,000 years, the first person to carry it out in practice was ocsenave@0: Galileo, in 1592. ocsenave@0: ocsenave@0: Later on we will give more precise definitions of the term ocsenave@0: \ldquo{}thermometer.\rdquo{} But at the present stage we ocsenave@0: are not in a position to do so (as Galileo was not), because ocsenave@0: the very concepts needed have not yet been developed; ocsenave@0: more precise definitions can be ocsenave@0: given only after our study has revealed the need for them. In ocsenave@0: deed, our final definition can be given only after the full ocsenave@0: mathematical formalism of statistical mechanics is at hand. ocsenave@0: ocsenave@0: Once a thermometer has been constructed, and the scale ocsenave@0: marked off in a quite arbitrary way (although we will suppose ocsenave@0: that the scale is at least monotonic: i.e., greater warmth always ocsenave@0: corresponds to a greater number), we are ready to begin scien ocsenave@0: tific experiments in thermodynamics. The number read eff from ocsenave@0: any such instrument is called the /empirical temperature/, and we ocsenave@0: denote it by \(t\). Since the exact calibration of the thermometer ocsenave@0: is not specified), any monotonic increasing function ocsenave@0: \(t‘ = f(t)\) provides an equally good temperature scale for the ocsenave@0: present. ocsenave@0: ocsenave@0: ocsenave@0: ** Thermodynamic Systems. ocsenave@0: ocsenave@0: The \ldquo{}thermodynamic systems\rdquo{} which ocsenave@0: are the objects of our study may be, physically, almost any ocsenave@0: collections of objects. The traditional simplest system with ocsenave@0: which to begin a study of thermodynamics is a volume of gas. ocsenave@0: We shall, however, be concerned from the start also with such ocsenave@0: things as a stretched wire or membrane, an electric cell, a ocsenave@0: polarized dielectric, a paramagnetic body in a magnetic field, etc. ocsenave@0: ocsenave@0: The /thermodynamic state/ of such a system is determined by ocsenave@0: specifying (i.e., measuring) certain macrcoscopic physical ocsenave@0: properties. Now, any real physical system has many millions of such ocsenave@0: preperties; in order to have a usable theory we cannot require ocsenave@0: that /all/ of them be specified. We see, therefore, that there ocsenave@0: must be a clear distinction between the notions of ocsenave@0: \ldquo{}thermodynamic system\rdquo{} and \ldquo{}physical ocsenave@0: system.\rdquo{} ocsenave@0: A given /physical/ system may correspond to many different ocsenave@0: /thermodynamic systems/, depending ocsenave@0: on which variables we choose to measure or control; and which ocsenave@0: we decide to leave unmeasured and/or uncontrolled. ocsenave@0: ocsenave@0: ocsenave@0: For example, our physical system might consist of a crystal ocsenave@0: of sodium chloride. For one set of experiments we work with ocsenave@0: temperature, volume, and pressure; and ignore its electrical ocsenave@0: properties. For another set of experiments we work with ocsenave@0: temperature, electric field, and electric polarization; and ocsenave@0: ignore the varying stress and strain. The /physical/ system, ocsenave@0: therefore, corresponds to two entirely different /thermodynamic/ ocsenave@0: systems. Exactly how much freedom, then, do we have in choosing ocsenave@0: the variables which shall define the thermodynamic state of our ocsenave@0: system? How many must we choose? What [criteria] determine when ocsenave@0: we have made an adequate choice? These questions cannot be ocsenave@0: answered until we say a little more about what we are trying to ocsenave@0: accomplish by a thermodynamic theory. A mere collection of ocsenave@0: recorded data about our system, as in the [[http://en.wikipedia.org/wiki/CRC_Handbook_of_Chemistry_and_Physics][/Handbook of Physics and ocsenave@0: Chemistry/]], is a very useful thing, but it hardly constitutes ocsenave@0: a theory. In order to construct anything deserving of such a ocsenave@0: name, the primary requirement is that we can recognize some kind ocsenave@0: of reproducible connection between the different properties con ocsenave@0: sidered, so that information about some of them will enable us ocsenave@0: to predict others. And of course, in order that our theory can ocsenave@0: be called thermodynamics (and not some other area of physics), ocsenave@0: it is necessary that the temperature be one of the quantities ocsenave@0: involved in a nontrivial way. ocsenave@0: ocsenave@0: The gist of these remarks is that the notion of ocsenave@0: \ldquo{}thermodynamic system\rdquo{} is in part ocsenave@0: an anthropomorphic one; it is for us to ocsenave@0: say which set of variables shall be used. If two different ocsenave@0: choices both lead to useful reproducible connections, it is quite ocsenave@0: meaningless to say that one choice is any more \ldquo{}correct\rdquo{} ocsenave@0: than the other. Recognition of this fact will prove crucial later in ocsenave@0: avoiding certain ancient paradoxes. ocsenave@0: ocsenave@0: At this stage we can determine only empirically which other ocsenave@0: physical properties need to be introduced before reproducible ocsenave@0: connections appear. Once any such connection is established, we ocsenave@0: can analyze it with the hope of being able to (1) reduce it to a ocsenave@0: /logical/ connection rather than an empirical one; and (2) extend ocsenave@0: it to an hypothesis applying beyond the original data, which ocsenave@0: enables us to predict further connections capable of being ocsenave@0: tested by experiment. Examples of this will be given presently. ocsenave@0: ocsenave@0: ocsenave@0: There will remain, however, a few reproducible relations ocsenave@0: which to the best of present knowledge, are not reducible to ocsenave@0: logical relations within the context of classical thermodynamics ocsenave@0: (and. whose demonstration in the wider context of mechanics, ocsenave@0: electrodynamics, and quantum theory remains one of probability ocsenave@0: rather than logical proof); from the standpoint of thermodynamics ocsenave@0: these remain simply statements of empirical fact which must be ocsenave@0: accepted as such without any deeper basis, but without which the ocsenave@0: development of thermodynamics cannot proceed. Because of this ocsenave@0: special status, these relations have become known as the ocsenave@0: \ldquo{}laws\rdquo{} ocsenave@0: of thermodynamics . The most fundamental one is a qualitative ocsenave@0: rather than quantitative relation, the \ldquo{}zero'th law.\rdquo{} ocsenave@0: ocsenave@0: ** Equilibrium; the \ldquo{}Zero‘th Law.\rdquo{} ocsenave@0: ocsenave@0: It is a common experience ocsenave@0: that when objects are placed in contact with each other but ocsenave@0: isolated from their surroundings, they may undergo observable ocsenave@0: changes for a time as a result; one body may become warmer, ocsenave@0: another cooler, the pressure of a gas or volume of a liquid may ocsenave@0: change; stress or magnetization in a solid may change, etc. But ocsenave@0: after a sufficient time, the observable macroscopic properties ocsenave@0: settle down to a steady condition, after which no further changes ocsenave@0: are seen unless there is a new intervention from the outside. ocsenave@0: When this steady condition is reached, the experimentalist says ocsenave@0: that the objects have reached a state of /equilibrium/ with each ocsenave@0: other. Once again, more precise definitions of this term will ocsenave@0: be needed eventually, but they require concepts not yet developed. ocsenave@0: In any event, the criterion just stated is almost the only one ocsenave@0: used in actual laboratory practice to decide when equilibrium ocsenave@0: has been reached. ocsenave@0: ocsenave@0: ocsenave@0: A particular case of equilibrium is encountered when we ocsenave@0: place a thermometer in contact with another body. The reading ocsenave@0: \(t\) of the thermometer may vary at first, but eventually it reach es ocsenave@0: a steady value. Now the number \(t\) read by a thermometer is always. ocsenave@0: by definition, the empirical temperature /of the thermometer/ (more ocsenave@0: precisely, of the sensitive element of the thermometer). When ocsenave@0: this number is constant in time, we say that the thermometer is ocsenave@0: in /thermal equilibrium/ with its surroundings; and we then extend ocsenave@0: the notion of temperature, calling the steady value \(t\) also the ocsenave@0: /temperature of the surroundings/. ocsenave@0: ocsenave@0: We have repeated these elementary facts, well known to every ocsenave@0: child, in order to emphasize this point: Thermodynamics can be ocsenave@0: a theory /only/ of states of equilibrium, because the very ocsenave@0: procedure by which the temperature of a system is defined by ocsenave@0: operational means, already presupposes the attainment of ocsenave@0: equilibrium. Strictly speaking, therefore, classical ocsenave@0: thermodynamics does not even contain the concept of a ocsenave@0: \ldquo{}time-varying temperature.\rdquo{} ocsenave@0: ocsenave@0: Of course, to recognize this limitation on conventional ocsenave@0: thermodynamics (best emphasized by calling it instead, ocsenave@0: thermostatics) in no way rules out the possibility of ocsenave@0: generalizing the notion of temperature to nonequilibrium states. ocsenave@0: Indeed, it is clear that one could define any number of ocsenave@0: time-dependent quantities all of which reduce, in the special ocsenave@0: case of equilibrium, to the temperature as defined above. ocsenave@0: Historically, attempts to do this even antedated the discovery ocsenave@0: of the laws of thermodynamics, as is demonstrated by ocsenave@0: \ldquo{}Newton's law of cooling.\rdquo{} Therefore, the ocsenave@0: question is not whether generalization is /possible/, but only ocsenave@0: whether it is in any way /useful/; i.e., does the temperature so ocsenave@0: generalized have any connection with other physical properties ocsenave@0: of our system, so that it could help us to predict other things? ocsenave@0: However, to raise such questions takes us far beyond the ocsenave@0: domain of thermostatics; and the general laws of nonequilibrium ocsenave@0: behavior are so much more complicated that it would be virtually ocsenave@0: hopeless to try to unravel them by empirical means alone. For ocsenave@0: example, even if two different kinds of thermometer are calibrated ocsenave@0: so that they agree with each other in equilibrium situations, ocsenave@0: they will not agree in general about the momentary value a ocsenave@0: \ldquo{}time-varying temperature.\rdquo{} To make any real ocsenave@0: progress in this area, we have to supplement empirical observation by the guidance ocsenave@0: of a rather hiqhly-developed theory. The notion of a ocsenave@0: time-dependent temperature is far from simple conceptually, and we ocsenave@0: will find that nothing very helpful can be said about this until ocsenave@0: the full mathematical apparatus of nonequilibrium statistical ocsenave@0: mechanics has been developed. ocsenave@0: ocsenave@0: Suppose now that two bodies have the same temperature; i.e., ocsenave@0: a given thermometer reads the same steady value when in contact ocsenave@0: with either. In order that the statement, \ldquo{}two bodies have the ocsenave@1: same temperature\rdquo{} shall describe a physical property of the bodies, ocsenave@0: and not merely an accidental circumstance due to our having used ocsenave@0: a particular kind of thermometer, it is necessary that /all/ ocsenave@0: thermometers agree in assigning equal temperatures to them if ocsenave@0: /any/ thermometer does . Only experiment is competent to determine ocsenave@0: whether this universality property is true. Unfortunately, the ocsenave@0: writer must confess that he is unable to cite any definite ocsenave@0: experiment in which this point was subjected to a careful test. ocsenave@0: That equality of temperatures has this absolute meaning, has ocsenave@0: evidently been taken for granted so much that (like absolute ocsenave@0: sirnultaneity in pre-relativity physics) most of us are not even ocsenave@0: consciously aware that we make such an assumption in ocsenave@0: thermodynamics. However, for the present we can only take it as a familiar ocsenave@0: empirical fact that this condition does hold, not because we can ocsenave@0: cite positive evidence for it, but because of the absence of ocsenave@0: negative evidence against it; i.e., we think that, if an ocsenave@0: exception had ever been found, this would have created a sensation in ocsenave@0: physics, and we should have heard of it. ocsenave@0: ocsenave@0: We now ask: when two bodies are at the same temperature, ocsenave@0: are they then in thermal equilibrium with each other? Again, ocsenave@0: only experiment is competent to answer this; the general ocsenave@0: conclusion, again supported more by absence of negative evidence ocsenave@0: than by specific positive evidence, is that the relation of ocsenave@0: equilibrium has this property: ocsenave@0: #+begin_quote ocsenave@0: /Two bodies in thermal equilibrium ocsenave@0: with a third body, are thermal equilibrium with each other./ ocsenave@0: #+end_quote ocsenave@0: ocsenave@0: This empirical fact is usually called the \ldquo{}zero'th law of ocsenave@0: thermodynamics.\rdquo{} Since nothing prevents us from regarding a ocsenave@0: thermometer as the \ldquo{}third body\rdquo{} in the above statement, ocsenave@0: it appears that we may also state the zero'th law as: ocsenave@0: #+begin_quote ocsenave@0: /Two bodies are in thermal equilibrium with each other when they are ocsenave@0: at the same temperature./ ocsenave@0: #+end_quote ocsenave@0: Although from the preceding discussion it might appear that ocsenave@0: these two statements of the zero'th law are entirely equivalent ocsenave@0: (and we certainly have no empirical evidence against either), it ocsenave@0: is interesting to note that there are theoretical reasons, arising ocsenave@0: from General Relativity, indicating that while the first ocsenave@0: statement may be universally valid, the second is not. When we ocsenave@0: consider equilibrium in a gravitational field, the verification ocsenave@0: that two bodies have equal temperatures may require transport ocsenave@0: of the thermometer through a gravitational potential difference; ocsenave@0: and this introduces a new element into the discussion. We will ocsenave@0: consider this in more detail in a later Chapter, and show that ocsenave@0: according to General Relativity, equilibrium in a large system ocsenave@0: requires, not that the temperature be uniform at all points, but ocsenave@0: rather that a particular function of temperature and gravitational ocsenave@0: potential be constant (the function is \(T\cdot \exp{(\Phi/c^2})\), where ocsenave@0: \(T\) is the Kelvin temperature to be defined later, and \(\Phi\) is the ocsenave@0: gravitational potential). ocsenave@0: ocsenave@0: Of course, this effect is so small that ordinary terrestrial ocsenave@0: experiments would need to have a precision many orders of ocsenave@0: magnitude beyond that presently possible, before one could hope even ocsenave@0: to detect it; and needless to say, it has played no role in the ocsenave@0: development of thermodynamics. For present purposes, therefore, ocsenave@0: we need not distinguish between the two above statements of the ocsenave@0: zero'th law, and we take it as a basic empirical fact that a ocsenave@0: uniform temperature at all points of a system is an essential ocsenave@0: condition for equilibrium. It is an important part of our ocsenave@0: ivestigation to determine whether there are other essential ocsenave@0: conditions as well. In fact, as we will find, there are many ocsenave@0: different kinds of equilibrium; and failure to distinguish between ocsenave@0: them can be a prolific source of paradoxes. ocsenave@0: ocsenave@0: ** Equation of State ocsenave@0: Another important reproducible connection is found when ocsenave@0: we consider a thermodynamic system defined by ocsenave@0: three parameters; in addition to the temperature we choose a ocsenave@0: \ldquo{}displacement\rdquo{} and a conjugate \ldquo{}force.\rdquo{} ocsenave@0: Subject to some qualifications given below, we find experimentally ocsenave@0: that these parameters are not independent, but are subject to a constraint. ocsenave@0: For example, we cannot vary the equilibrium pressure, volume, ocsenave@0: and temperature of a given mass of gas independently; it is found ocsenave@0: that a given pressure and volume can be realized only at one ocsenave@0: particular temperature, that the gas will assume a given tempera~ ocsenave@0: ture and volume only at one particular pressure, etc. Similarly, ocsenave@0: a stretched wire can be made to have arbitrarily assigned tension ocsenave@0: and elongation only if its temperature is suitably chosen, a ocsenave@0: dielectric will assume a state of given temperature and ocsenave@0: polarization at only one value of the electric field, etc. ocsenave@0: These simplest nontrivial thermodynamic systems (three ocsenave@0: parameters with one constraint) are said to possess two ocsenave@0: /degrees of freedom/; for the range of possible equilibrium states is defined ocsenave@0: by specifying any two of the variables arbitrarily, whereupon the ocsenave@0: third, and all others we may introduce, are determined. ocsenave@0: Mathematically, this is expressed by the existence of a functional ocsenave@1: relationship of the form[fn:: /Edit./: The set of solutions to an equation ocsenave@0: like /f(X,x,t)=/ const. is called a /level set/. Here, Jaynes is ocsenave@0: saying that the quantities /X/, /x/, and /t/ follow a \ldquo{}functional ocsenave@0: rule\rdquo{}, so the set of physically allowed combinations of /X/, ocsenave@0: /x/, and /t/ in equilibrium states can be ocsenave@0: expressed as the level set of a function. ocsenave@0: ocsenave@0: But not every function expresses a constraint relation; for some ocsenave@0: functions, you can specify two of the variables, and the third will ocsenave@0: still be undetermined. (For example, if f=X^2+x^2+t^2-3, ocsenave@0: the level set /f(X,x,t)=0/ is a sphere, and specifying /x=1/, /t=1/ ocsenave@0: leaves you with two potential possibilities for /X/ =\pm 1.) ocsenave@0: ocsenave@1: A function like /f/ has to possess one more propery in order for its ocsenave@1: level set to express a constraint relationship: it must be monotonic in ocsenave@0: each of its variables /X/, /x/, and /t/. ocsenave@0: #the partial derivatives of /f/ exist for every allowed combination of ocsenave@0: #inputs /x/, /X/, and /t/. ocsenave@0: In other words, the level set has to pass a sort of ocsenave@0: \ldquo{}vertical line test\rdquo{} for each of its variables.] ocsenave@0: ocsenave@0: #Edit Here, Jaynes ocsenave@0: #is saying that it is possible to express the collection of allowed combinations \(\langle X,x,t \rangle\) of force, quantity, and temperature as a ocsenave@0: #[[http://en.wikipedia.org/wiki/Level_set][level set]] of some function \(f\). However, not all level sets represent constraint relations; consider \(f(X,x,t)=X^2+x^2+t^2-1\)=0. ocsenave@0: #In order to specify ocsenave@0: ocsenave@0: \begin{equation} ocsenave@0: f(X,x,t) = O ocsenave@0: \end{equation} ocsenave@0: ocsenave@0: where $X$ is a generalized force (pressure, tension, electric or ocsenave@0: magnetic field, etc.), $x$ is the corresponding generalized ocsenave@0: displacement (volume, elongation, electric or magnetic polarization, ocsenave@1: etc.), and $t$ is the empirical temperature. Equation (1-1) is ocsenave@0: called /the equation of state/. ocsenave@0: ocsenave@0: At the risk of belaboring it, we emphasize once again that ocsenave@0: all of this applies only for a system in equilibrium; for ocsenave@0: otherwise not only.the temperature, but also some or all of the other ocsenave@0: variables may not be definable. For example, no unique pressure ocsenave@0: can be assigned to a gas which has just suffered a sudden change ocsenave@0: in volume, until the generated sound waves have died out. ocsenave@0: ocsenave@0: Independently of its functional form, the mere fact of the ocsenave@0: /existence/ of an equation of state has certain experimental ocsenave@0: consequences. For example, suppose that in experiments on oxygen ocsenave@0: gas, in which we control the temperature and pressure ocsenave@0: independently, we have found that the isothermal compressibility $K$ ocsenave@0: varies with temperature, and the thermal expansion coefficient ocsenave@0: \alpha varies with pressure $P$, so that within the accuracy of the data, ocsenave@0: ocsenave@0: \begin{equation} ocsenave@0: \frac{\partial K}{\partial t} = - \frac{\partial \alpha}{\partial P} ocsenave@0: \end{equation} ocsenave@0: ocsenave@0: Is this a particular property of oxygen; or is there reason to ocsenave@0: believe that it holds also for other substances? Does it depend ocsenave@0: on our particular choice of a temperature scale? ocsenave@0: ocsenave@0: In this case, the answer is found at once; for the definitions of $K$, ocsenave@0: \alpha are ocsenave@0: ocsenave@0: \begin{equation} ocsenave@0: K = -\frac{1}{V}\frac{\partial V}{\partial P},\qquad ocsenave@0: \alpha=\frac{1}{V}\frac{\partial V}{\partial t} ocsenave@0: \end{equation} ocsenave@0: ocsenave@0: which is simply a mathematical expression of the fact that the ocsenave@0: volume $V$ is a definite function of $P$ and $t$; i.e., it depends ocsenave@0: only ocsenave@0: on their present values, and not how those values were attained. ocsenave@0: In particular, $V$ does not depend on the direction in the \((P, t)\) ocsenave@0: plane through which the present values were approached; or, as we ocsenave@0: usually say it, \(dV\) is an /exact differential/. ocsenave@0: ocsenave@1: Therefore, although at first glance the relation (1-2) appears ocsenave@0: nontrivial and far from obvious, a trivial mathematical analysis ocsenave@0: convinces us that it must hold regardless of our particular ocsenave@0: temperature scale, and that it is true not only of oxygen; it must ocsenave@0: hold for any substance, or mixture of substances, which possesses a ocsenave@0: definite, reproducible equation of state \(f(P,V,t)=0\). ocsenave@0: ocsenave@0: But this understanding also enables us to predict situations in which ocsenave@1: (1-2) will /not/ hold. Equation (1-2), as we have just learned, expresses ocsenave@0: the fact that an equation of state exists involving only the three ocsenave@0: variables \((P,V,t)\). Now suppose we try to apply it to a liquid such ocsenave@0: as nitrobenzene. The nitrobenzene molecule has a large electric dipole ocsenave@0: moment; and so application of an electric field (as in the ocsenave@0: [[http://en.wikipedia.org/wiki/Kerr_effect][electro-optical Kerr cell]]) causes an alignment of molecules which, as ocsenave@0: accurate measurements will verify, changes the pressure at a given ocsenave@0: temperature and volume. Therefore, there can no longer exist any ocsenave@0: unique equation of state involving \((P, V, t)\) only; with ocsenave@0: sufficiently accurate measurements, nitrobenzene must be regarded as a ocsenave@0: thermodynamic system with at least three degrees of freedom, and the ocsenave@0: general equation of state must have at least a complicated a form as ocsenave@0: \(f(P,V,t,E) = 0\). ocsenave@0: ocsenave@0: But if we introduce a varying electric field $E$ into the discussion, ocsenave@0: the resulting varying electric polarization $M$ also becomes a new ocsenave@0: thermodynamic variable capable of being measured. Experimentally, it ocsenave@0: is easiest to control temperature, pressure, and electric field ocsenave@0: independently, and of course we find that both the volume and ocsenave@0: polarization are then determined; i.e., there must exist functional ocsenave@0: relations of the form \(V = V(P,t,E)\), \(M = M(P,t,E)\), or in more ocsenave@0: symmetrical form ocsenave@0: ocsenave@0: \begin{equation} ocsenave@0: f(V,P,t,E) = 0 \qquad g(M,P,t,E)=0. ocsenave@0: \end{equation} ocsenave@0: ocsenave@0: In other words, if we regard nitrobenzene as a thermodynamic system of ocsenave@0: three degrees of freedom (i.e., having specified three parameters ocsenave@0: arbitrarily, all others are then determined), it must possess two ocsenave@0: independent equations of state. ocsenave@0: ocsenave@0: Similarly, a thermodynamic system with four degrees of freedom, ocsenave@0: defined by the termperature and three pairs of conjugate forces and ocsenave@0: displacements, will have three independent equations of state, etc. ocsenave@0: ocsenave@0: Now, returning to our original question, if nitrobenzene possesses ocsenave@0: this extra electrical degree of freedom, under what circumstances do ocsenave@0: we exprect to find a reproducible equation of state involving ocsenave@0: \((p,V,t)\) only? Evidently, if $E$ is held constant, then the first ocsenave@0: of equations (1-5) becomes such an equation of state, involving $E$ as ocsenave@0: a fixed parameter; we would find many different equations of state of ocsenave@0: the form \(f(P,V,t) = 0\) with a different function $f$ for each ocsenave@0: different value of the electric field. Likewise, if \(M\) is held ocsenave@0: constant, we can eliminate \(E\) between equations (1-5) and find a ocsenave@0: relation \(h(P,V,t,M)=0\), which is an equation of state for ocsenave@0: \((P,V,t)\) containing \(M\) as a fixed parameter. ocsenave@0: ocsenave@0: More generally, if an electrical constraint is imposed on the system ocsenave@0: (for example, by connecting an external charged capacitor to the ocsenave@0: electrodes) so that \(M\) is determined by \(E\); i.e., there is a ocsenave@0: functional relation of the form ocsenave@0: ocsenave@0: \begin{equation} ocsenave@0: g(M,E) = \text{const.} ocsenave@0: \end{equation} ocsenave@0: ocsenave@0: then (1-5) and (1-6) constitute three simultaneous equations, from ocsenave@0: which both \(E\) and \(M\) may be eliminated mathematically, leading ocsenave@0: to a relation of the form \(h(P,V,t;q)=0\), which is an equation of ocsenave@0: state for \((P,V,t)\) involving the fixed parameter \(q\). ocsenave@0: ocsenave@0: We see, then, that as long as a fixed constraint of the form (1-6) is ocsenave@0: imposed on the electrical degree of freedom, we can still observe a ocsenave@0: reproducible equation of state for nitrobenzene, considered as a ocsenave@0: thermodynamic system of only two degrees of freedom. If, however, this ocsenave@0: electrical constraint is removed, so that as we vary $P$ and $t$, the ocsenave@0: values of $E$ and $M$ vary in an uncontrolled way over a ocsenave@0: /two-dimensional/ region of the \((E, M)\) plane, then we will find no ocsenave@0: definite equation of state involving only \((P,V,t)\). ocsenave@0: ocsenave@0: This may be stated more colloqually as follows: even though a system ocsenave@0: has three degrees of freedom, we can still consider only the variables ocsenave@0: belonging to two of them, and we will find a definite equation of ocsenave@0: state, /provided/ that in the course of the experiments, the unused ocsenave@0: degree of freedom is not \ldquo{}tampered with\rdquo{} in an ocsenave@0: uncontrolled way. ocsenave@0: ocsenave@0: We have already emphasized that any physical system corresponds to ocsenave@0: many different thermodynamic systems, depending on which variables we ocsenave@0: choose to control and measure. In fact, it is easy to see that any ocsenave@0: physical system has, for all practical purposes, an /arbitrarily ocsenave@0: large/ number of degrees of freedom. In the case of nitrobenzene, for ocsenave@0: example, we may impose any variety of nonuniform electric fields on ocsenave@1: our sample. Suppose we place $(n+1)$ different electrodes, labelled ocsenave@1: \(\{e_0,e_1, e_2 \ldots e_n\}\) in contact with the liquid in various ocsenave@1: positions. Regarding \(e_0\) as the \ldquo{}ground\rdquo{}, maintained ocsenave@1: at zero potential, we can then impose $n$ different potentials ocsenave@1: \(\{v_1, \ldots, v_n\}\) on the other electrodes independently, and we ocsenave@1: can also measure the $n$ different conjugate displacements, as the ocsenave@1: charges \(\{q_1,\ldots, q_n\}\) accumulated on electrodes ocsenave@1: \(\{e_1,\ldots e_n\}\). Together with the pressure (understood as the ocsenave@1: pressure measured at one given position), volume, and temperature, our ocsenave@1: sample of nitrobenzene is now a thermodynamic system of $(n+1)$ ocsenave@1: degrees of freedom. This number may be as large as we please, limited ocsenave@1: only by our patience in constructing the apparatus needed to control ocsenave@1: or measure all these quantities. ocsenave@1: ocsenave@1: We leave it as an exercise for the reader (Problem 1) to find the most ocsenave@1: general condition on the variables \(\{v_1, q_1, v_2, q_2, \ldots ocsenave@1: v_n,q_n\}\) which will ensure that a definite equation of state ocsenave@1: $f(P,V,t)=0$ is observed in spite of all these new degrees of ocsenave@1: freedom. The simplest special case of this relation is, evidently, to ocsenave@1: ground all electrodes, thereby inposing the conditions $v_1 = v_2 = ocsenave@1: \ldots = v_n = 0$. Equally well (if we regard nitrobenzene as having ocsenave@1: negligible electrical conductivity) we may open-circuit all ocsenave@1: electrodes, thereby imposing the conditions \(q_i = \text{const.}\) In ocsenave@1: the latter case, in addition to an equation of state of the form ocsenave@1: \(f(P,V,t)=0\), which contains these constants as fixed parameters, ocsenave@1: there are \(n\) additional equations of state of the form $v_i = ocsenave@1: v_i(P,t)$. But if we choose to ignore these voltages, there will be no ocsenave@1: contradiction in considering our nitrobenzene to be a thermodynamic ocsenave@1: system of two degrees of freedom, involving only the variables ocsenave@1: \(P,V,t\). ocsenave@1: ocsenave@1: Similarly, if our system of interest is a crystal, we may impose on it ocsenave@1: a wide variety of nonuniform stress fields; each component of the ocsenave@1: stress tensor $T_{ij}$ may bary with position. We might expand each of ocsenave@1: these functions in a complete orthonormal set of functions ocsenave@1: \(\phi_k(x,y,z)\): ocsenave@1: ocsenave@1: \begin{equation} ocsenave@1: T_{ij}(x,y,z) = \sum_k a_{ijk} \phi_k(x,y,z) ocsenave@1: \end{equation} ocsenave@1: ocsenave@1: and with a sufficiently complicated system of levers which in various ocsenave@1: ways squeeze and twist the crystal, we might vary each of the first ocsenave@1: 1,000 expansion coefficients $a_{ijk}$ independently, and measure the ocsenave@1: conjugate displacements $q_{ijk}$. Our crystal is then a thermodynamic ocsenave@1: system of over 1,000 degrees of freedom. ocsenave@1: ocsenave@1: The notion of \ldquo{}numbers of degrees of freedom\rdquo{} is ocsenave@1: therefore not a /physical property/ of any system; it is entirely ocsenave@1: anthropomorphic, since any physical system may be regarded as a ocsenave@1: thermodynamic system with any number of degrees of freedom we please. ocsenave@1: ocsenave@1: If new thermodynamic variables are always introduced in pairs, ocsenave@1: consisting of a \ldquo{}force\rdquo{} and conjugate ocsenave@1: \ldquo{}displacement\rdquo{}, then a thermodynamic system of $n$ ocsenave@1: degrees of freedom must possess $(n-1)$ independent equations of ocsenave@1: state, so that specifying $n$ quantities suffices to determine all ocsenave@1: others. ocsenave@1: ocsenave@1: This raises an interesting question; whether the scheme of classifying ocsenave@1: thermodynamic variables in conjugate pairs is the most general ocsenave@1: one. Why, for example, is it not natural to introduce three related ocsenave@1: variables at a time? To the best of the writer's knowledge, this is an ocsenave@1: open question; there seems to be no fundamental reason why variables ocsenave@1: /must/ always be introduced in conjugate pairs, but there seems to be ocsenave@1: no known case in which a different scheme suggests itself as more ocsenave@1: appropriate. ocsenave@1: ocsenave@1: ** Heat ocsenave@1: We are now in a position to consider the results and interpretation of ocsenave@1: a number of elementary experiments involving ocsenave@2: thermal interaction, which can be carried out as soon as a primitive ocsenave@2: thermometer is at hand. In fact these experiments, which we summarize ocsenave@2: so quickly, required a very long time for their first performance, and ocsenave@2: the essential conclusions of this Section were first arrived at only ocsenave@2: about 1760---more than 160 years after Galileo's invention of the ocsenave@2: thermometer---by Joseph Black, who was Professor of Chemistry at ocsenave@2: Glasgow University. Black's analysis of calorimetric experiments ocsenave@2: initiated by G. D. Fahrenheit before 1736 led to the first recognition ocsenave@2: of the distinction between temperature and heat, and prepared the way ocsenave@2: for the work of his better-known pupil, James Watt. ocsenave@1: ocsenave@2: We first observe that if two bodies at different temperatures are ocsenave@2: separated by walls of various materials, they sometimes maintain their ocsenave@2: temperature difference for a long time, and sometimes reach thermal ocsenave@2: equilibrium very quickly. The differences in behavior observed must be ocsenave@2: ascribed to the different properties of the separating walls, since ocsenave@2: nothing else is changed. Materials such as wood, asbestos, porous ocsenave@2: ceramics (and most of all, modern porous plastics like styrofoam), are ocsenave@2: able to sustain a temperature difference for a long time; a wall of an ocsenave@2: imaginary material with this property idealized to the point where a ocsenave@2: temperature difference is maintained indefinitely is called an ocsenave@2: /adiabatic wall/. A very close approximation to a perfect adiabatic ocsenave@2: wall is realized by the Dewar flask (thermos bottle), of which the ocsenave@2: walls consist of two layers of glass separated by a vacuum, with the ocsenave@2: surfaces silvered like a mirror. In such a container, as we all know, ocsenave@2: liquids may be maintained hot or cold for days. ocsenave@1: ocsenave@2: On the other hand, a thin wall of copper or silver is hardly able to ocsenave@2: sustain any temperature difference at all; two bodies separated by ocsenave@2: such a partition come to thermal equilibrium very quickly. Such a wall ocsenave@2: is called /diathermic/. It is found in general that the best ocsenave@2: diathermic materials are the metals and good electrical conductors, ocsenave@2: while electrical insulators make fairly good adiabatic walls. There ocsenave@2: are good theoretical reasons for this rule; a particular case of it is ocsenave@2: given by the [[http://en.wikipedia.org/wiki/Wiedemann_franz_law][Wiedemann-Franz law]] of solid-state theory. ocsenave@2: ocsenave@2: Since a body surrounded by an adiabatic wall is able to maintain its ocsenave@2: temperature independently of the temperature of its surroundings, an ocsenave@2: adiabatic wall provides a means of thermally /isolating/ a system from ocsenave@2: the rest of the universe; it is to be expected, therefore, that the ocsenave@2: laws of thermal interaction between two systems will assume the ocsenave@2: simplest form if they are enclosed in a common adiabatic container, ocsenave@2: and that the best way of carrying out experiments on thermal ocsenave@2: peroperties of substances is to so enclose them. Such an apparatus, in ocsenave@2: which systems are made to interact inside an adiabatic container ocsenave@2: supplied with a thermometer, is called a /calorimeter/. ocsenave@2: ocsenave@2: Let us imagine that we have a calorimeter in which there is initially ocsenave@2: a volume $V_W$ of water at a temperature $t_1$, and suspended above it ocsenave@2: a volume $V_I$ of some other substance (say, iron) at temperature ocsenave@2: $t_2$. When we drop the iron into the water, they interact thermally ocsenave@2: (and the exact nature of this interaction is one of the things we hope ocsenave@2: to learn now), the temperature of both changing until they are in ocsenave@2: thermal equilibrium at a final temperature $t_0$. ocsenave@2: ocsenave@2: Now we repeat the experiment with different initial temperatures ocsenave@2: $t_1^\prime$ and $t_2^\prime$, so that a new equilibrium is reached at ocsenave@2: temperature $t_0^\prime$. It is found that, if the temperature ocsenave@2: differences are sufficiently small (and in practice this is not a ocsenave@2: serious limitation if we use a mercury thermometer calibrated with ocsenave@2: uniformly spaced degree marks on a capillary of uniform bore), then ocsenave@2: whatever the values of $t_1^\prime$, $t_2^\prime$, $t_1$, $t_2$, the ocsenave@2: final temperatures $t_0^\prime$, $t_0$ will adjust themselves so that ocsenave@2: the following relation holds: ocsenave@2: ocsenave@2: \begin{equation} ocsenave@2: \frac{t_2 - t_0}{t_0 - t_1} = \frac{t_2^\prime - ocsenave@2: t_0^\prime}{t_0^\prime - t_1^\prime} ocsenave@2: \end{equation} ocsenave@2: ocsenave@2: in other words, the /ratio/ of the temperature changes of the iron and ocsenave@2: water is independent of the initial temperatures used. ocsenave@2: ocsenave@2: We now vary the amounts of iron and water used in the calorimeter. It ocsenave@2: is found that the ratio (1-8), although always independent of the ocsenave@2: starting temperatures, does depend on the relative amounts of iron and ocsenave@2: water. It is, in fact, proportional to the mass $M_W$ of water and ocsenave@2: inversely proportional to the mass $M_I$ of iron, so that ocsenave@2: ocsenave@2: \begin{equation} ocsenave@2: \frac{t_2-t_0}{t_0-t_1} = \frac{M_W}{K_I M_I} ocsenave@2: \end{equation} ocsenave@2: ocsenave@2: where $K_I$ is a constant. ocsenave@2: ocsenave@2: We next repeat the above experiments using a different material in ocsenave@2: place of the iron (say, copper). We find again a relation ocsenave@2: ocsenave@2: \begin{equation} ocsenave@2: \frac{t_2-t_0}{t_0-t_1} = \frac{M_W}{K_C \cdot M_C} ocsenave@2: \end{equation} ocsenave@2: ocsenave@2: where $M_C$ is the mass of copper; but the constant $K_C$ is different ocsenave@2: from the previous $K_I$. In fact, we see that the constant $K_I$ is a ocsenave@2: new physical property of the substance iron, while $K_C$ is a physical ocsenave@2: property of copper. The number $K$ is called the /specific heat/ of a ocsenave@2: substance, and it is seen that according to this definition, the ocsenave@2: specific heat of water is unity. ocsenave@2: ocsenave@2: We now have enough experimental facts to begin speculating about their ocsenave@2: interpretation, as was first done in the 18th century. First, note ocsenave@2: that equation (1-9) can be put into a neater form that is symmetrical ocsenave@2: between the two substances. We write $\Delta t_I = t_0 - t_2$, $\Delta ocsenave@2: t_W = t_0 - t_1$ for the temperature changes of iron and water ocsenave@2: respectively, and define $K_W \equiv 1$ for water. Equation (1-9) then ocsenave@2: becomes ocsenave@2: ocsenave@2: \begin{equation} ocsenave@2: K_W M_W \Delta t_W + K_I M_I \Delta t_I = 0 ocsenave@2: \end{equation} ocsenave@2: ocsenave@2: The form of this equation suggests a new experiment; we go back into ocsenave@2: the laboratory, and find $n$ substances for which the specific heats ocsenave@2: \(\{K_1,\ldots K_n\}\) have been measured previously. Taking masses ocsenave@2: \(\{M_1, \ldots, M_n\}\) of these substances, we heat them to $n$ ocsenave@2: different temperatures \(\{t_1,\ldots, t_n\}\) and throw them all into ocsenave@2: the calorimeter at once. After they have all come to thermal ocsenave@2: equilibrium at temperature $t_0$, we find the differences $\Delta t_j ocsenave@2: = t_0 - t_j$. Just as we suspected, it turns out that regardless of ocsenave@2: the $K$'s, $M$'s, and $t$'s chosen, the relation ocsenave@2: \begin{equation} ocsenave@2: \sum_{j=0}^n K_j M_j \Delta t_j = 0 ocsenave@2: \end{equation} ocsenave@2: is always satisfied. This sort of process is an old story in ocsenave@2: scientific investigations; although the great theoretician Boltzmann ocsenave@2: is said to have remarked: \ldquo{}Elegance is for tailors \rdquo{}, it ocsenave@2: remains true that the attempt to reduce equations to the most ocsenave@2: symmetrical form has often suggested important generalizations of ocsenave@2: physical laws, and is a great aid to memory. Witness Maxwell's ocsenave@2: \ldquo{}displacement current\rdquo{}, which was needed to fill in a ocsenave@2: gap and restore the symmetry of the electromagnetic equations; as soon ocsenave@2: as it was put in, the equations predicted the existence of ocsenave@2: electromagnetic waves. In the present case, the search for a rather ocsenave@2: rudimentary form of \ldquo{}elegance\rdquo{} has also been fruitful, ocsenave@2: for we recognize that (1-12) has the standard form of a /conservation ocsenave@2: law/; it defines a new quantity which is conserved in thermal ocsenave@2: interactions of the type just studied. ocsenave@2: ocsenave@2: The similarity of (1-12) to conservation laws in general may be seen ocsenave@2: as follows. Let $A$ be some quantity that is conserved; the $i$th ocsenave@2: system has an amount of it $A_i$. Now when the systems interact such ocsenave@2: that some $A$ is transferred between them, the amount of $A$ in the ocsenave@2: $i$th system is changed by a net amount \(\Delta A_i = (A_i)_{final} - ocsenave@2: (A_i)_{initial}\); and the fact that there is no net change in the ocsenave@2: total amount of $A$ is expressed by the equation \(\sum_i \Delta ocsenave@2: A_i = 0$. Thus, the law of conservation of matter in a chemical ocsenave@2: reaction is expressed by \(\sum_i \Delta M_i = 0\), where $M_i$ is the ocsenave@2: mass of the $i$th chemical component. ocsenave@2: ocsenave@2: what is this new conserved quantity? Mathematically, it can be defined ocsenave@2: as $Q_i = K_i\cdot M_i cdot t_i; whereupon (1-12) becomes ocsenave@2: ocsenave@2: \begin{equation} ocsenave@2: \sum_i \Delta Q_i = 0 ocsenave@2: \end{equation} ocsenave@2: ocsenave@2: and at this point we can correct a slight quantitative inaccuracy. As ocsenave@2: noted, the above relations hold accurately only when the temperature ocsenave@2: differences are sufficiently small; i.e., they are really only ocsenave@2: differential laws. On sufficiently accurate measurements one find that ocsenave@2: the specific heats $K_i$ depend on temperature; if we then adopt the ocsenave@2: integral definition of $\Delta Q_i$, ocsenave@2: \begin{equation} ocsenave@2: \Delta Q_i = \int_{t_{i}}^{t_0} K_i(t) M_i dt ocsenave@2: \end{equation} ocsenave@2: ocsenave@2: the conservation law (1-13) will be found to hold in calorimetric ocsenave@2: experiments with liquids and solids, to any accuracy now feasible. And ocsenave@2: of course, from the manner in which the $K_i(t)$ are defined, this ocsenave@2: relation will hold however our thermometers are calibrated. ocsenave@2: ocsenave@2: Evidently, the stage is now set for a \ldquo{}new\rdquo{} physical ocsenave@2: theory to account for these facts. In the 17th century, both Francis ocsenave@2: Bacon and Isaac Newton had expressed their opinions that heat was a ocsenave@2: form of motion; but they had no supporting factual evidence. By the ocsenave@2: latter part of the 18th century, one had definite factual evidence ocsenave@2: which seemed to make this view untenable; by the calorimetric ocsenave@2: \ldquo{}mixing\rdquo{} experiments just described, Joseph Black had ocsenave@2: recognized the distinction between temperature $t$ as a measure of ocsenave@2: \ldquo{}hotness\rdquo{}, and heat $Q$ as a measure of /quantity/ of ocsenave@2: something, and introduced the notion of heat capacity. He also ocsenave@2: recognized the latent heats of freezing and vaporization. To account ocsenave@2: for the conservation laws thus discovered, the theory then suggested ocsenave@2: itself, naturally and almost inevitably, that heat was /fluid/, ocsenave@2: indestructable and uncreatable, which had no appreciable weight and ocsenave@2: was attracted differently by different kinds of matter. In 1787, ocsenave@2: Lavoisier invented the name \ldquo{}caloric\rdquo{} for this fluid. ocsenave@2: ocsenave@2: Looking down today from our position of superior knowledge (i.e., ocsenave@2: hindsight) we perhaps need to be reminded that the caloric theory was ocsenave@2: a perfectly respectable scientific theory, fully deserving of serious ocsenave@2: consideration; for it accounted quantitatively for a large body of ocsenave@2: experimental fact, and made new predictions capable of being tested by ocsenave@2: experiment. ocsenave@2: ocsenave@2: One of these predictions was the possibility of accounting for the ocsenave@2: thermal expansion of bodies when heated; perhaps the increase in ocsenave@2: volume was just a measure of the volume of caloric fluid ocsenave@2: absorbed. This view met with some disappointment as a result of ocsenave@2: experiments which showed that different materials, on absorbing the ocsenave@2: same quantity of heat, expanded by different amounts. Of course, this ocsenave@2: in itself was not enough to overthrow the caloric theory, because one ocsenave@2: could suppose that the caloric fluid was compressible, and was held ocsenave@2: under different pressure in different media. ocsenave@2: ocsenave@2: Another difficulty that seemed increasingly serious by the end of the ocsenave@2: 18th century was the failure of all attempts to weigh this fluid. Many ocsenave@2: careful experiments were carried out, by Boyle, Fordyce, Rumford and ocsenave@2: others (and continued by Landolt almost into the 20th century), with ocsenave@2: balances capable of detecting a change of weight of one part in a ocsenave@2: million; and no change could be detected on the melting of ice, ocsenave@2: heating of substances, or carrying out of chemical reactions. But even ocsenave@2: this is not really a conclusive argument against the caloric theory, ocsenave@2: since there is no /a priori/ reason why the fluid should be dense ocsenave@2: enough to weigh with balances (of course, we know today from ocsenave@2: Einstein's $E=mc^2$ that small changes in weight should indeed exist ocsenave@2: in these experiments; but to measure them would require balances about ocsenave@2: 10^7 times more sensitive than were available). ocsenave@2: ocsenave@2: Since the caloric theory derives entirely from the empirical ocsenave@2: conservation law (1-33), it can be refuted conclusively only by ocsenave@2: exhibiting new experimental facts revealing situations in which (1-13) ocsenave@2: is /not/ valid. The first such case was [[http://www.chemteam.info/Chem-History/Rumford-1798.html][found by Count Rumford (1798)]], ocsenave@2: who was in charge of boring cannon in the Munich arsenal, and noted ocsenave@2: that the cannon and chips became hot as a result of the cutting. He ocsenave@2: found that heat could be produced indefinitely, as long as the boring ocsenave@2: was continued, without any compensating cooling of any other part of ocsenave@2: the system. Here, then, was a clear case in which caloric was /not/ ocsenave@2: conserved, as in (1-13); but could be created at will. Rumford wrote ocsenave@2: that he could not conceive of anything that could be produced ocsenave@2: indefinitely by the expenditure of work, \ldquo{}except it be /motion/\rdquo{}. ocsenave@2: ocsenave@2: But even this was not enough to cause abandonment of the caloric ocsenave@2: theory; for while Rumford's observations accomplished the negative ocsenave@2: purpose of showing that the conservation law (1-13) is not universally ocsenave@2: valid, they failed to accomplish the positive one of showing what ocsenave@2: specific law should replace it (although he produced a good hint, not ocsenave@2: sufficiently appreciated at the time, in his crude measurements of the ocsenave@2: rate of heat production due to the work of one horse). Within the ocsenave@2: range of the original calorimetric experiments, (1-13) was still ocsenave@2: valid, and a theory successful in a restricted domain is better than ocsenave@2: no theory at all; so Rumford's work had very little impact on the ocsenave@2: actual development of thermodynamics. ocsenave@2: ocsenave@2: (This situation is a recurrent one in science, and today physics offers ocsenave@2: another good example. It is recognized by all that our present quantum ocsenave@2: field theory is unsatisfactory on logical, conceptual, and ocsenave@2: mathematical grounds; yet it also contains some important truth, and ocsenave@2: no responsible person has suggested that it be abandoned. Once again, ocsenave@2: a semi-satisfactory theory is better than none at all, and we will ocsenave@2: continue to teach it and to use it until we have something better to ocsenave@2: put in its place.) ocsenave@2: ocsenave@2: # what is "the specific heat of a gas at constant pressure/volume"? ocsenave@2: # changed t for temperature below from capital T to lowercase t. ocsenave@2: Another failure of the conservation law (1-13) was noted in 1842 by ocsenave@2: R. Mayer, a German physician, who pointed out that the data already ocsenave@2: available showed that the specific heat of a gas at constant pressure, ocsenave@2: C_p, was greater than at constant volume $C_v$. He surmised that the ocsenave@2: difference was due to the work done in expansion of the gas against ocsenave@2: atmospheric pressure, when measuring $C_p$. Supposing that the ocsenave@2: difference $\Delta Q = (C_p - C_v)\Delta t$ calories, in the heat ocsenave@2: required to raise the temperature by $\Delta t$ was actually a ocsenave@2: measure of amount of energy, he could estimate from the amount ocsenave@2: $P\Delta V$ ergs of work done the amount of mechanical energy (number ocsenave@2: of ergs) corresponding to a calorie of heat; but again his work had ocsenave@2: very little impact on the development of thermodynamics, because he ocsenave@2: merely offered this notion as an interpretation of the data without ocsenave@2: performing or suggesting any new experiments to check his hypothesis ocsenave@2: further. ocsenave@2: ocsenave@2: Up to the point, then, one has the experimental fact that a ocsenave@2: conservation law (1-13) exists whenever purely thermal interactions ocsenave@2: were involved; but in processes involving mechanical work, the ocsenave@2: conservation law broke down. ocsenave@2: ocsenave@2: ** The First Law ocsenave@2: ocsenave@2: ocsenave@2: ocsenave@2: * COMMENT Appendix ocsenave@1: ocsenave@1: | Generalized Force | Generalized Displacement | ocsenave@1: |--------------------+--------------------------| ocsenave@1: | force | displacement | ocsenave@1: | pressure | volume | ocsenave@1: | electric potential | charge |