% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % % % % The Project Gutenberg EBook of Treatise on Thermodynamics, by Max Planck% % % % This eBook is for the use of anyone anywhere in the United States and most % other parts of the world at no cost and with almost no restrictions % % whatsoever. You may copy it, give it away or re-use it under the terms of % the Project Gutenberg License included with this eBook or online at % % www.gutenberg.org. If you are not located in the United States, you'll have % to check the laws of the country where you are located before using this ebook. % % % Title: Treatise on Thermodynamics % % % % Author: Max Planck % % % % Translator: Alexander Ogg % % % % Release Date: January 9, 2016 [EBook #50880] % % Most recently updated: June 11, 2021 % % % % Language: English % % % % Character set encoding: UTF-8 % % % % *** START OF THIS PROJECT GUTENBERG EBOOK TREATISE ON THERMODYNAMICS ***% % % % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % \def\ebook{50880} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% %% %% Packages and substitutions: %% %% %% %% book: Required. %% %% inputenc: Latin-1 text encoding. Required. %% %% %% %% babel: German hyphenation. Required. %% %% %% %% ifthen: Logical conditionals. Required. %% %% %% %% amsmath: AMS mathematics enhancements. Required. %% %% amssymb: Additional mathematical symbols. Required. %% %% %% %% array: Enhancements to array/tabular environments. Required. %% %% %% %% alltt: Fixed-width font environment. Required. %% %% %% %% footmisc: Start footnote numbering on each page. Required. %% %% %% %% mhchem: Chemical formulas. Required. %% %% %% %% multicol: Multicolumn environment for index. Required. %% %% makeidx: Indexing. Required. %% %% %% %% graphicx: Standard interface for graphics inclusion. Required. %% %% %% %% calc: Length calculations. Required. %% %% %% %% fancyhdr: Enhanced running headers and footers. Required. %% %% %% %% geometry: Enhanced page layout package. Required. %% %% hyperref: Hypertext embellishments for pdf output. Required. %% %% %% %% %% %% Producer's Comments: %% %% %% %% OCR text for this ebook was obtained December 28, 2015, from %% %% http://www.archive.org/details/treatiseonthermo00planrich. %% %% %% %% Minor changes to the original are noted in this file in three %% %% ways: %% %% 1. \Typo{}{} for typographical corrections, showing original %% %% and replacement text side-by-side. %% %% 2. \Chg{}{} and \Add{}, for inconsistent/missing punctuation,%% %% italicization, and capitalization. %% %% 3. [** TN: Note]s for lengthier or stylistic comments. %% %% %% %% %% %% Compilation Flags: %% %% %% %% The following behavior may be controlled by boolean flags. %% %% %% %% ForPrinting (false by default): %% %% If false, compile a screen optimized file (one-sided layout, %% %% blue hyperlinks). If true, print-optimized PDF file: Larger %% %% text block, two-sided layout, black hyperlinks. %% %% %% %% %% %% PDF pages: 352 (if ForPrinting set to false) %% %% PDF page size: 4.75 x 6.75" (non-standard) %% %% %% %% Summary of log file: %% %% * Two overfull hboxes (neither visible). %% %% %% %% Compile History: %% %% %% %% January, 2016: (Andrew D. Hwang) %% %% texlive2012, GNU/Linux %% %% %% %% Command block: %% %% %% %% pdflatex x3 %% %% makeindex %% %% pdflatex x3 %% %% %% %% %% %% January 2016: pglatex. %% %% Compile this project with: %% %% pdflatex 50880-t.tex ..... THREE times %% %% makeindex 50880-t.idx %% %% pdflatex 50880-t.tex ..... THREE times %% %% %% %% pdfTeX, Version 3.1415926-2.5-1.40.14 (TeX Live 2013/Debian) %% %% %% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \documentclass[12pt]{book}[2005/09/16] \listfiles %%%%%%%%%%%%%%%%%%%%%%%%%%%%% PACKAGES %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \usepackage[utf8]{inputenc}[2006/05/05] \usepackage[german,UKenglish]{babel} \usepackage{ifthen}[2001/05/26] %% Logical conditionals \usepackage{amsmath}[2000/07/18] %% Displayed equations \usepackage{amssymb}[2002/01/22] %% and additional symbols \usepackage{array}[2008/09/09] \usepackage{alltt}[1997/06/16] %% boilerplate, credits, license \usepackage[perpage,symbol]{footmisc}[2005/03/17] \usepackage[version=3]{mhchem}[2006/12/17] \usepackage{multicol}[2006/05/18] \usepackage{makeidx}[2000/03/29] \usepackage{graphicx}[1999/02/16]%% For diagrams \usepackage{calc}[2005/08/06] \usepackage{fancyhdr} %% For running heads %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%% Interlude: Set up PRINTING (default) or SCREEN VIEWING %%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % ForPrinting=true false (default) % Asymmetric margins Symmetric margins % 1 : 1.6 text block aspect ratio 3 : 4 text block aspect ratio % Black hyperlinks Blue hyperlinks % Start major marker pages recto No blank verso pages % \newboolean{ForPrinting} %% UNCOMMENT the next line for a PRINT-OPTIMIZED VERSION of the text %% %\setboolean{ForPrinting}{true} %% Initialize values to ForPrinting=false \newcommand{\Margins}{hmarginratio=1:1} % Symmetric margins \newcommand{\HLinkColor}{blue} % Hyperlink color \newcommand{\PDFPageLayout}{SinglePage} \newcommand{\TransNote}{Transcriber's Note} \newcommand{\TransNoteCommon}{% The camera-quality files for this public-domain ebook may be downloaded \textit{gratis} at \begin{center} \texttt{www.gutenberg.org/ebooks/\ebook}. \end{center} This ebook was produced using scanned images and OCR text generously provided by the University of California, Berkeley, through the Internet Archive. \bigskip Minor typographical corrections and presentational changes have been made without comment. \bigskip } \newcommand{\TransNoteText}{% \TransNoteCommon This PDF file is optimized for screen viewing, but may be recompiled for printing. Please consult the preamble of the \LaTeX\ source file for instructions and other particulars. } %% Re-set if ForPrinting=true \ifthenelse{\boolean{ForPrinting}}{% \renewcommand{\Margins}{hmarginratio=2:3} % Asymmetric margins \renewcommand{\HLinkColor}{black} % Hyperlink color \renewcommand{\PDFPageLayout}{TwoPageRight} \renewcommand{\TransNote}{Transcriber's Note} \renewcommand{\TransNoteText}{% \TransNoteCommon This PDF file is optimized for printing, but may be recompiled for screen viewing. Please consult the preamble of the \LaTeX\ source file for instructions and other particulars. } }{% If ForPrinting=false, don't skip to recto \renewcommand{\cleardoublepage}{\clearpage} } %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%% End of PRINTING/SCREEN VIEWING code; back to packages %%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \ifthenelse{\boolean{ForPrinting}}{% \setlength{\paperwidth}{8.5in}% \setlength{\paperheight}{11in}% % 1:1.6 \usepackage[body={4.5in,8in},\Margins]{geometry}[2002/07/08] }{% \setlength{\paperwidth}{4.75in}% \setlength{\paperheight}{6.75in}% \raggedbottom % 3:4 \usepackage[body={4.5in,5.875in},\Margins,includeheadfoot]{geometry}[2002/07/08] } \providecommand{\ebook}{00000} % Overridden during white-washing \usepackage[pdftex, hyperfootnotes=false, pdfauthor={Max Planck}, pdftitle={The Project Gutenberg eBook \#\ebook: Treatise on Thermodynamics.}, pdfkeywords={University of California, The Internet Archive, Andrew D. Hwang}, pdfstartview=Fit, % default value pdfstartpage=1, % default value pdfpagemode=UseNone, % default value bookmarks=true, % default value linktocpage=false, % default value pdfpagelayout=\PDFPageLayout, pdfdisplaydoctitle, pdfpagelabels=true, bookmarksopen=true, bookmarksopenlevel=0, colorlinks=true, linkcolor=\HLinkColor]{hyperref}[2007/02/07] %% Fixed-width environment to format PG boilerplate %% \newenvironment{PGtext}{% \begin{alltt} %**** \fontsize{8.1}{10}\ttfamily\selectfont}% {\end{alltt}} % Errors found during digitization \newcommand{\Typo}[2]{#2} % Stylistic changes made for consistency \newcommand{\Chg}[2]{#2} %\newcommand{\Chg}[2]{#1} % Use this to revert inconsistencies in the original \newcommand{\Add}[1]{\Chg{}{#1}} % Book's list of errata \newcommand{\Erratum}[2]{#2} %% Miscellaneous global parameters %% % No hrule in page header \renewcommand{\headrulewidth}{0pt} % Loosen spacing \setlength{\emergencystretch}{1em} \newcommand{\Loosen}{\spaceskip 0.375em plus 0.75em minus 0.25em} % Scratch pad for length calculations \newlength{\TmpLen} %% Tables \newcommand{\TableFont}{\footnotesize}%{\scriptsize}% \newcommand{\ColHead}[1]{\multicolumn{1}{c}{#1}} \newcommand{\ColumnHeading}[2]{% \settowidth{\TmpLen}{#1}% \parbox[c]{\TmpLen}{\centering\medskip #2\medskip}% } \newcommand{\Strut}[1][12pt]{\rule{0pt}{#1}} %% Running heads %% \newcommand{\FlushRunningHeads}{\cleardoublepage\fancyhf{}} \newcommand{\InitRunningHeads}{% \setlength{\headheight}{15pt} \pagestyle{fancy} \thispagestyle{empty} \ifthenelse{\boolean{ForPrinting}} {\fancyhead[RO,LE]{\thepage}} {\fancyhead[R]{\thepage}} } \newcommand{\RHead}[1]{\textsc{\MakeLowercase{#1}}} % Asymmetric running heads in print-formatted version \ifthenelse{\boolean{ForPrinting}}{% \newcommand{\SetRunningHeads}[1]{% \fancyhead[CE]{\RHead{Thermodynamics.}}% \fancyhead[CO]{\RHead{#1}}}% }{% Uniform running head in screen mode \newcommand{\SetRunningHeads}[1]{% \fancyhead[C]{\RHead{#1}}}% } \newcommand{\BookMark}[2]{\phantomsection\pdfbookmark[#1]{#2}{#2}} %% Major document divisions %% \newcommand{\PGBoilerPlate}{% \pagenumbering{Alph} \pagestyle{empty} \BookMark{0}{PG Boilerplate} } \newcommand{\FrontMatter}{% \cleardoublepage \frontmatter \BookMark{-1}{Front Matter} } \newcommand{\MainMatter}{% \FlushRunningHeads \InitRunningHeads \mainmatter \BookMark{-1}{Main Matter} } \newcommand{\BackMatter}{% \FlushRunningHeads \InitRunningHeads \backmatter \BookMark{-1}{Back Matter} } \newcommand{\PGLicense}{% \FlushRunningHeads \pagenumbering{Roman} \InitRunningHeads \BookMark{-1}{PG License} \SetRunningHeads{License} } %% ToC formatting %% \AtBeginDocument{\renewcommand{\contentsname}% {\protect\thispagestyle{empty}% \protect\SectTitle{CONTENTS}\protect\vspace{-1.5\baselineskip}} } \newcommand{\TableofContents}{% \FlushRunningHeads \InitRunningHeads \SetRunningHeads{Contents} \BookMark{0}{Contents} \tableofcontents } % Set the section number in a fixed-width box \newcommand{\ToCBox}[1]{\settowidth{\TmpLen}{99.}% \makebox[\TmpLen][r]{#1}\hspace*{1em}% } % For internal use, to determine if we need the Sect./Page line \newcommand{\ToCAnchor}{} % \ToCLine{Counter}{SecNo}{Title} \newcommand{\ToCLine}[3]{% \label{toc:#1}% \ifthenelse{\not\equal{\pageref{toc:#1}}{\ToCAnchor}}{% \renewcommand{\ToCAnchor}{\pageref{toc:#1}}% \noindent\makebox[\textwidth][r]{\scriptsize CHAPTER\hfill PAGE}\\% }{}% \settowidth{\TmpLen}{9999}% \noindent\strut\parbox[b]{\textwidth-\TmpLen}{\small% \ToCBox{#2}\hangindent4em\raggedright\textsc{#3}\dotfill}% \makebox[\TmpLen][r]{\pageref{chap:#1}}% \smallskip } %% Index formatting \makeindex \makeatletter \renewcommand{\@idxitem}{\par\hangindent 30\p@\global\let\idxbrk\nobreak} \renewcommand\subitem{\idxbrk\@idxitem \hspace*{12\p@}\let\idxbrk\relax} \renewcommand{\indexspace}{\par\penalty-3000 \vskip 10pt plus5pt minus3pt\relax} \renewenvironment{theindex}{% \setlength\columnseprule{0.5pt}\setlength\columnsep{18pt}% \begin{multicols}{2}[{\Appendix{Index}\small}]% \setlength\parindent{0pt}\setlength\parskip{0pt plus 0.3pt}% \thispagestyle{empty}\let\item\@idxitem\raggedright% }{% \end{multicols}\FlushRunningHeads } \makeatother \newcommand{\indexnote}[1]{\hyperpage{#1}~\textit{n}.} \renewcommand{\see}[1]{See~#1.} %% Sectional units %% % Typographical abstraction \newcommand{\ChapHead}[2]{% \section*{\centering\normalfont\large\MakeUppercase{Chapter #1}} \subsection*{\centering\normalfont\normalsize\itshape\MakeUppercase{#2}} } \newcommand{\SectTitle}[2][\large]{% \section*{\centering#1\normalfont #2} } \newcommand{\SubsectTitle}[2][\normalsize]{% \subsection*{\centering#1\normalfont\textsc{#2}} } % Internal bookkeeping \newcounter{ChapNo}% \newboolean{ChapterClear} % Title for title page and start of Part I. \newcommand{\BookHead}{% \LARGE TREATISE \\[8pt] \normalsize ON \\[12pt] \TPage{0.9}{THERMODYNAMICS} } \newcommand{\Part}[2]{% \FlushRunningHeads \InitRunningHeads % Cross-referencing \BookMark{0}{Part #1} \addtocontents{toc}{\protect\SectTitle{PART #1}} \addtocontents{toc}{\protect\SubsectTitle[\protect\footnotesize]{\MakeUppercase{#2}}} % Page heading \setboolean{ChapterClear}{false}% Tell \Chapter not to start of a new page \ifthenelse{\equal{#1}{I.}}{% Special heading for Part I. \begin{center} \BookHead \end{center} }{}% Part II., etc. \SectTitle{PART #1} \SubsectTitle{#2} } % \Chapter{Title} \newcommand{\Chapter}[3][]{% % Start on a new page unless we've just started a \Part \ifthenelse{\boolean{ChapterClear}}{% \FlushRunningHeads \InitRunningHeads }{% Else we've just started a part; set flag \setboolean{ChapterClear}{true}% } % ToC entry \refstepcounter{ChapNo}% \phantomsection\label{chap:\theChapNo}% \addtocontents{toc}{\protect\ToCLine{\theChapNo}{#2}{#3}} % PDF bookmark and running heads \ifthenelse{\equal{#1}{}}{% No manual running head \BookMark{1}{#3} \SetRunningHeads{#3} \ChapHead{#2}{#3} }{% Manual running head \BookMark{1}{#1} \SetRunningHeads{#1} \ChapHead{#2}{#3} } } \newcommand{\Section}[1]{ \medskip\par\textbf{§\;#1} \label{section:#1} } % Unnumbered \Chapter-like units with no ToC entry \newcommand{\Preface}[1]{% \FlushRunningHeads \InitRunningHeads \BookMark{0}{#1} \SetRunningHeads{#1} \SectTitle{\MakeUppercase{#1}} } % Unnumbered \Chapter-like units with ToC entry \newcommand{\Appendix}[1]{% \FlushRunningHeads \InitRunningHeads % ToC entry \refstepcounter{ChapNo}% \phantomsection\label{chap:\theChapNo}% \ifthenelse{\equal{#1}{Catalogue}}{% \addtocontents{toc}{\protect\ToCLine{\theChapNo}{}{Catalogue of the Author's Papers on Thermodynamics}} }{% \addtocontents{toc}{\protect\ToCLine{\theChapNo}{}{#1}} } % PDF bookmark \BookMark{0}{#1} \SetRunningHeads{#1} \SectTitle{\MakeUppercase{#1}} } %% Cross-references %% \newcommand{\SecNum}[1]{\hyperref[section:#1.]{{\upshape #1}}} \newcommand{\SecRef}[2][§\;]{\hyperref[section:#2.]{{\upshape #1#2}}} \newcommand{\SSecRef}[1]{\SecRef[§§\;]{#1}} \newcommand{\SecRefs}[3][§§\;]{#1\SecNum{#2}--\SecNum{#3}} % Equation tags \newcommand{\Tag}[1]{% \phantomsection\label{eqn:#1}\tag*{\ensuremath{#1}}% } \newcommand{\Eq}[1]{% \hyperref[eqn:#1]{\ensuremath{#1}}% } % Figure labels \newcommand{\Fig}[1]{\hyperref[fig:#1]{Fig.~{\upshape #1}}} %% Diagrams %% \renewcommand{\floatpagefraction}{0.8} %\renewcommand{\topfraction}{0.95} %\renewcommand{\bottomfraction}{0.95} \newcommand{\Graphic}[2]{% \phantomsection\label{fig:#2}% \ifthenelse{\equal{#1}{}}{% \includegraphics{./images/#2.pdf}% }{% else width specified \includegraphics[width=#1]{./images/#2.pdf}% } } % \Figure[width]{figure number} \newcommand{\Figure}[2][]{% \begin{figure}[hbtp!] \centering \phantomsection\label{fig:#2} \Graphic{#1}{fig#2} \end{figure}\ignorespaces% } % Page separators \newcommand{\PageSep}[1]{\ignorespaces} % Small semantic structures and abbreviations % \TPage{fraction}{content}: scale content for title page \newcommand{\TPage}[2]{\resizebox{#1\textwidth}{!}{#2}} \newcommand{\Signature}[4]{% \medskip \null\hfill \large #1\hspace{\parindent} \bigskip \small% \textsc{#2} \\ \hspace*{2.5\parindent}\textit{#3} #4\normalsize\par } \newcommand{\Topic}[1]{\textbf{#1}} \newcommand{\Bibitem}{\medskip\par\noindent\hangindent2em} \newcommand{\const}{\text{const.}} \newcommand{\eg}{\emph{e.g.}} \newcommand{\ie}{\emph{i.e.}} \renewcommand{\cf}{\emph{cf.}} \newcommand{\Cf}{\emph{Cf.}} \DeclareMathOperator{\aq}{\, aq.} \newcommand{\Tr}{\textsc{Tr}} \newcommand{\First}[1]{\textsc{#1}} \newcommand{\Emph}[1]{\emph{\textbf{#1}}} %% Miscellaneous mathematical formatting %% \newcommand{\efrac}[2]{\tfrac{#1}{#2}} \newcommand{\dd}{\partial} \newcommand{\tsum}{\mathop{\textstyle\sum}\limits} \newcommand{\Z}{\phantom{0}} \DeclareInputMath{176}{{}^{\circ}} \DeclareInputMath{183}{\cdot} \newcommand{\Unit}[1]{\text{\upshape #1}} \newcommand{\eps}{\varepsilon} % Large integral sign \newcommand{\Bigint}[1][2]{% \scalebox{#1}{$\displaystyle\int$}{\kern-4pt}% } \newcommand{\EmRule}{\rule{12pt}{0pt}} \newcommand{\Bigintlimits}[3][2]{% \mathop{\scalebox{#1}{$\displaystyle\int$}}\limits_{\mbox{$#2$\EmRule}}^{\mbox{\EmRule$#3$}}\kern-2pt% } % \PadTo[alignment]{width text}{visible text} \newcommand{\PadTo}[3][c]{% \settowidth{\TmpLen}{$#2$}% \makebox[\TmpLen][#1]{$#3$}% } %%%%%%%%%%%%%%%%%%%%%%%% START OF DOCUMENT %%%%%%%%%%%%%%%%%%%%%%%%%% \begin{document} %% PG BOILERPLATE %% \PGBoilerPlate \begin{center} \begin{minipage}{\textwidth} \small \begin{PGtext} The Project Gutenberg EBook of Treatise on Thermodynamics, by Max Planck This eBook is for the use of anyone anywhere in the United States and most other parts of the world at no cost and with almost no restrictions whatsoever. You may copy it, give it away or re-use it under the terms of the Project Gutenberg License included with this eBook or online at www.gutenberg.org. If you are not located in the United States, you'll have to check the laws of the country where you are located before using this ebook. Title: Treatise on Thermodynamics Author: Max Planck Translator: Alexander Ogg Release Date: January 9, 2016 [EBook #50880] Most recently updated: June 11, 2021 Language: English Character set encoding: UTF-8 *** START OF THIS PROJECT GUTENBERG EBOOK TREATISE ON THERMODYNAMICS *** \end{PGtext} \end{minipage} \end{center} \newpage %% Credits and transcriber's note %% \begin{center} \begin{minipage}{\textwidth} \begin{PGtext} Produced by Andrew D. Hwang \end{PGtext} \end{minipage} \vfill \end{center} \begin{minipage}{0.85\textwidth} \small \BookMark{0}{Transcriber's Note.} \subsection*{\centering\normalfont\scshape% \normalsize\MakeLowercase{\TransNote}}% \raggedright \TransNoteText \end{minipage} %%%%%%%%%%%%%%%%%%%%%%%%%%% FRONT MATTER %%%%%%%%%%%%%%%%%%%%%%%%%% \PageSep{i} \FrontMatter \ifthenelse{\boolean{ForPrinting}}{% \null\vfill \TPage{0.9}{TREATISE ON THERMODYNAMICS} \vfill \cleardoublepage }{}% Omit half-title in screen version \PageSep{ii} %[Blank page] \PageSep{iii} {\centering % [** TN: Commend prints ``Treatise on Thermodynamics''] \BookHead \vfill \footnotesize BY \\ \Large \textsc{Dr.\ MAX PLANCK} \\[-4pt] \TPage{0.9}{\scriptsize PROFESSOR OF THEORETICAL PHYSICS IN THE UNIVERSITY OF BERLIN} \bigskip \vfill \normalsize\textit{TRANSLATED WITH THE AUTHOR'S SANCTION} \\[8pt] \footnotesize BY \\[8pt] \large \textsc{ALEXANDER OGG, M.A., B.Sc., Ph.D.} \TPage{0.9}{\scriptsize LATE 1851 EXHIBITION SCHOLAR AND UNIVERSITY ASSISTANT, ABERDEEN UNIVERSITY} \\[-4pt] \TPage{0.8}{\scriptsize ASSISTANT MASTER, ROYAL NAVAL ENGINEERING COLLEGE, DEVONPORT} \vfill %% Publisher's device \Graphic{1.5in}{device} \vfill \TPage{0.8}{LONGMANS, GREEN, AND CO.} \\ \TPage{0.6}{39 PATERNOSTER ROW, LONDON} \\ \TPage{0.4}{NEW YORK AND BOMBAY} \\ 1903 \\ \footnotesize \textit{All rights reserved} }% End \centering \newpage \PageSep{iv} % [** Blank page] \PageSep{v} \Preface{Translator's Notice} \First{The} modern developments of Thermodynamics, and the applications to physical and chemical problems, have become so important, that I have ventured to translate Professor Planck's book, which presents the whole subject from a uniform point of view. A few notes have been added to the present English edition by Professor Planck. He has not found it necessary to change the original text in any way. To bring the notation into conformity with the usual English notation, several symbols have been changed. This has been done with the author's sanction. Here I have followed J.~J. van~Laar and taken $\Psi$ to signify what he calls the \emph{Planck'sches Potential}, \Chg{i.e.}{\ie}\ the thermodynamic potential of Gibbs and Duhem divided by~$-\theta$. Professor Planck's recent paper, ``Über die Grundlage der Lösungstheorie'' (Ann.\ d.\ Phys.\ \textbf{10}, p.~436, 1903), ought to be read in connection with his thermodynamical theory of solution. I am indebted to Herren Veit \&~Co., Leipzig, for kindly supplying the blocks of the five figures in the text. \Signature{A. O.}{Devonport,}{June,}{1903.} \PageSep{vi} %[** Blank page] \PageSep{vii} \Preface{Preface} \First{The} oft-repeated requests either to publish my collected papers on Thermodynamics, or to work them up into a comprehensive treatise, first suggested the writing of this book. Although the first plan would have been the simpler, especially as I found no occasion to make any important changes in the line of thought of my original papers, yet I decided to rewrite the whole subject-matter, with the intention of giving at greater length, and with more detail, certain general considerations and demonstrations too concisely expressed in these papers. My chief reason, however, was that an opportunity was thus offered of presenting the entire field of Thermodynamics from a uniform point of view. This, to be sure, deprives the work of the character of an original contribution to science, and stamps it rather as an introductory text-book on Thermodynamics for students who have taken elementary courses in Physics and Chemistry, and are familiar with the elements of the Differential and Integral Calculus. Still, I do not think that this book will entirely supersede my former publications on the same subject. Apart from the fact that these contain, in a sense, a more original presentation, there may be found in them a number of details expanded at greater length than seemed advisable in the more comprehensive treatment here required. To \PageSep{viii} enable the reader to revert in particular cases to the original form for comparison, a list of my publications on Thermodynamics has been appended, with a reference in each case to the section of the book which deals with the same point. The numerical values in the examples, which have been worked, as applications of the theory, have, almost all of them, been taken from the original papers; only a few, that have been determined by frequent measurement, have been taken from the tables in Kohlrausch's ``Leitfaden der praktischen Physik.'' It should be emphasized, however, that the numbers used, notwithstanding the care taken, have not undergone the same amount of critical sifting as the more general propositions and deductions. \bigskip Three distinct methods of investigation may be clearly recognized in the previous development of Thermodynamics. The first penetrates deepest into the nature of the processes considered, and, were it possible to carry it out exactly, would be designated as the most perfect. Heat, according to it, is due to the definite motions of the chemical molecules and atoms considered as distinct masses, which in the case of gases possess comparatively simple properties, but in the case of solids and liquids can be only very roughly sketched. This kinetic theory, founded by Joule, Waterston, Krönig and Clausius, has been greatly extended mainly by Maxwell and Boltzmann. Obstacles, at present unsurmountable, however, seem to stand in the way of its further progress. These are due not only to the highly complicated mathematical treatment, but principally to essential difficulties, not to be discussed here, in the mechanical interpretation of the fundamental principles of Thermodynamics. \PageSep{ix} Such difficulties are avoided by the second method, developed by Helmholtz. It confines itself to the most important hypothesis of the mechanical theory of heat, that heat is due to motion, but refuses on principle to specialize as to the character of this motion. This is a safer point of view than the first, and philosophically quite as satisfactory as the mechanical interpretation of nature in general, but it does not as yet offer a foundation of sufficient breadth upon which to build a detailed theory. Starting from this point of view, all that can be obtained is the verification of some general laws which have already been deduced in other ways direct from experience. A third treatment of Thermodynamics has hitherto proved the most fruitful. This method is distinct from the other two, in that it does not advance the mechanical theory of heat, but, keeping aloof from definite assumptions as to its nature, starts direct from a few very general empirical facts, mainly the two fundamental principles of Thermodynamics. From these, by pure logical reasoning, a large number of new physical and chemical laws are deduced, which are capable of extensive application, and have hitherto stood the test without exception. This last, more inductive, treatment, which is used exclusively in this book, corresponds best to the present state of the science. It cannot be considered as final, however, but may have in time to yield to a mechanical, or perhaps an electro-magnetic theory. Although it may be of advantage for a time to consider the activities of nature---Heat, Motion, Electricity, etc.---as different in quality, and to suppress the question as to their common nature, still our aspiration after a uniform theory of nature, on a mechanical basis or otherwise, which has derived such powerful encouragement from the discovery of the principle of the \PageSep{x} conservation of energy, can never be permanently repressed. Even at the present day, a recession from the assumption that all physical phenomena are of a common nature would be tantamount to renouncing the comprehension of a number of recognized laws of interaction between different spheres of natural phenomena. Of course, even then, the results we have deduced from the two laws of Thermodynamics would not be invalidated, but these two laws would not be introduced as independent, but would be deduced from other more general propositions. At present, however, no probable limit can be set to the time which it will take to reach this goal. \Signature{THE AUTHOR.}{Berlin,}{April,}{1897.} \PageSep{xi} %[** Table of contents] \TableofContents \iffalse CONTENTS PART I. Fundamental Facts and Definitions CHAPTER PAGE I. Temperature 1 II. Molecular Weight 22 III. Quantity of Heat 32 PART II. The First Fundamental Principle of Thermodynamics I. General Exposition 38 II. Applications to Homogeneous Systems 46 III. Applications to Non-Homogeneous Systems 67 PART III. The Second Fundamental Principle of Thermodynamics I. Introduction 77 II. Proof 86 III. General Deductions . 105 \PageSep{xii} PART IV. Applications to Special States of Equilibrium CHAPTER PAGE I. Homogeneous Systems . 119 II. System in Different States of Aggregation . . . . 132 III. System of any Number of Independent Constituents . 173 IV. Gaseous System 207 V. Dilute Solutions 223 Catalogue of the Author's Papers on Thermodynamics 264 Index . 267 \PageSep{xiii} ERRATA The reader is requested kindly to make the following corrections in the pages where they occur:--- Page 26, line 8, for " occurs " read " occur." 30, 13 from bottom, for " Hydrobromamylene " read " Amy- lene hydrobromide." 40, 4 in 58, for "different "read "definite." 78, 5, for " restablishment " read " re-establishment." 78, 14, for "stakes" read "states." 79, 9 in 108, for "Occasionally," etc., read "That attempts are still made to represent this law as contained in the principle of energy may be seen from the fact that the too restricted term ' Energetics ' is sometimes applied to all investigations on these questions." 87, 2 in 118, for " heat " read " work." 104, 10 from end, for " metaphysicists " read " metaphysicians." 149, 3, /or "v" read "V' 152, lines 3 and 1 from end, read and throughout. \O v /2 \O V 176, line 8, for AM a " read " AM,'." 177, equation (149), for " M 2 " " read " 5M 2 "." 180, line 22, for " quintiple " read " quintuple." 185, 2, for " dM 2 " read " dM 2 '." 186, equation 153, for dM 2 5 read dM 2 'S-, G/iVl 2 O^'-z 191, line 3, left-hand side of equation, for " d log p " read " d log p " 202, lines 6 and 8, for " </> " read " <p." 214. line 15, for " n^c^ " read " n^." 229, 14, for " are " read " is." 230, 14, for second minus read plus. 232, 2 from end, for " equations " read * equation." 241, equation (225), for " 2 " read " 2 ." 241, line 5 from end, for " carbonic " read " carbon." 242, 8,/or"0 2 "read0 2 ." 243, 11 from end, for "molecule " read "gram molecule." % End of table of contents and errata text \fi \PageSep{xiv} % [** Blank page] \PageSep{1} \MainMatter % [** TN: Text written by the \Part macro] %TREATISE %ON %THERMODYNAMICS \Part{I.}{Fundamental Facts and Definitions.} \index{Laws of thermodynamics@Laws of thermodynamics|see{\textit{First} and \text{Second}}}% \Chapter{I.}{Temperature.} \Section{1.} \First{The} conception of ``heat'' arises from that particular \index{Free energy}% \index{Heat!conception of}% sensation of warmth or coldness which is immediately experienced on touching a body. This direct sensation, however, furnishes no quantitative scientific measure of a body's state with regard to heat; it yields only qualitative results, which vary according to external circumstances. For quantitative purposes we utilize the change of volume which takes place in all bodies when heated under constant pressure, for this admits of exact measurement. Heating produces in most substances an increase of volume, and thus we can tell whether a body gets hotter or colder, not merely by the sense of touch, but also by a purely mechanical observation affording a much greater degree of accuracy. We can also tell accurately when a body assumes a former state of heat. \Section{2.} If two bodies, one of which feels warmer than the other, be brought together (for example, a piece of heated metal and cold water), it is invariably found that the hotter body is cooled, and the colder one is heated up to a certain \PageSep{2} point, and then all change ceases. The two bodies are then said to be in \emph{thermal equilibrium}. Experience shows that \index{Equilibrium!thermal}% \index{Thermal equilibrium}% such a state of equilibrium finally sets in, not only when two, but also when any number of differently heated bodies are brought into mutual contact. From this follows the important proposition: \emph{If a body,~$A$, be in thermal equilibrium with two other bodies, $B$~and~$C$, then $B$~and $C$ are in thermal equilibrium with one another.}\footnote {\Loosen As is well known, there exists no corresponding proposition for electrical equilibrium. For if we join together the substances \ce{Cu $\mid$ CuSO4 $\aq$ $\mid$ ZnSO4 $\aq$ $\mid$ Zn} to form a conducting ring, no electrical equilibrium is possible.} For, if we bring $A$,~$B$, and~$C$ together so that each touches the other two, then, according to our supposition, there will be equilibrium at the points of contact $AB$ and~$AC$, and, therefore, also at the contact~$BC$. If it were not so, no general thermal equilibrium would be possible, which is contrary to experience. \Section{3.} These facts enable us to compare the degree of heat of two bodies, $B$~and~$C$, without bringing them into contact with one another; namely, by bringing each body into contact with an arbitrarily selected standard body,~$A$ (for example, a mass of mercury enclosed in a vessel terminating in a fine capillary tube). By observing the volume of~$A$ in each case, it is possible to tell whether $B$~and $C$ are in thermal equilibrium or not. If they are not in thermal equilibrium, we can tell which of the two is the hotter. The degree of heat of~$A$, or of any body in thermal equilibrium with~$A$, can thus be very simply defined by the volume of~$A$, or, as is usual, by the difference between the volume of~$A$ and its volume when in thermal equilibrium with melting ice under atmospheric pressure. This volumetric difference, which, by an appropriate choice of unit, is made to read $100$ when $A$~is in contact with steam under atmospheric pressure, is called the \emph{temperature} in degrees Centigrade with regard \index{Temperature!definition of}% to~$A$ as thermometric substance. Two bodies of equal temperature are, therefore, in thermal equilibrium, and \textit{vice versâ}. \PageSep{3} \Section{4.} The temperature readings of no two thermometric \index{Temperature!definition of}% substances agree, in general, except at~$0°$ and~$100°$. The definition of temperature is therefore somewhat arbitrary. This we may remedy to a certain extent by taking gases, in particular those hard to condense, such as hydrogen, oxygen, nitrogen, and carbon monoxide, as thermometric substances. They agree almost completely within a considerable range of temperature, and their readings are sufficiently in accordance for most purposes. Besides, the coefficient of expansion of these different gases is the same, inasmuch as equal volumes of them expand under constant pressure by the same amount---about $\frac{1}{273}$~of their volume---when heated from $0°$~C. to $1°$~C\@. Since, also, the influence of the external pressure on the volume of these gases can be represented by a very simple law, we are led to the conclusion that these regularities are based on a remarkable simplicity in their constitution, and that, therefore, it is reasonable to define the common temperature given by them simply as temperature. We must consequently reduce the readings of other thermometers to those of the gas thermometer, and preferably \index{Gas!thermometer}% \index{Thermometer, gas}% to those of the hydrogen thermometer. \Section{5.} The definition of temperature remains arbitrary in cases where the requirements of accuracy cannot be satisfied by the agreement between the readings of the different gas thermometers, for there is no sufficient reason for the preference of any one of these gases. A definition of temperature completely independent of the properties of any individual substance, and applicable to all stages of heat and cold, becomes first possible on the basis of the \emph{second law of thermodynamics} (\SecRef{160}, etc.). In the mean time, only such temperatures will be considered as are defined with sufficient accuracy by the gas thermometer. \Section{6.} In the following we shall deal chiefly with homogeneous, isotropic bodies of any form, possessing throughout \index{Isotropic bodies}% their substance the same temperature and density, and subject to a uniform pressure acting everywhere perpendicular to the surface. They, therefore, also exert the same \PageSep{4} pressure outwards. Surface phenomena are thereby disregarded. The condition of such a body is determined by its chemical nature; its mass,~$M$; its volume,~$V$; and its temperature,~$t$. On these must depend, in a definite manner, all other properties of the particular state of the body, especially the pressure, which is uniform throughout, internally and externally. The pressure,~$p$, is measured by the force acting on the unit of area---in the C.G.S. system, in dynes per square centimeter, a \emph{dyne} being the force which \index{Dyne}% imparts to a mass of one gramme in one second a velocity of one centimeter per second. \Section{7.} As the pressure is generally given in atmospheres, \index{Atmospheric pressure}% the value of an atmosphere in absolute C.G.S. units is here calculated. The pressure of an atmosphere is the weight of a column of mercury at~$0°$~C., $76~\Unit{cm.}$ high, and $1~\Unit{sq.}\ \Unit{cm.}$ in cross-section, when placed in mean geographical latitude. This latter condition must be added, because the weight, \ie\ the force of the earth's attraction, varies with the locality. The volume of the column of mercury is $76~\Chg{\Unit{c.c.}}{\Unit{cm.}^{3}}$; and since the density of mercury at~$0°$~C. is $13.596$, the mass is $76 × 13.596$. Multiplying the mass by the acceleration of gravity in mean latitude, we find the pressure of one atmosphere in absolute units to be \[ 76 × 13.596 × 981 = 1,013,650\ \frac{\Unit{dynes}}{\Unit{cm.}^{2}} \quad\text{or}\quad \frac{\Unit{gr.}}{\Unit{cm.-sec.}^{2}}. \] This, then, is the factor for converting atmospheres into absolute units. If, as was formerly the custom in mechanics, we use as the unit of force the weight of a gramme in mean geographical latitude instead of the dyne, the pressure of an atmosphere would be $76 × 13.596 = 1033.3~\Unit{grms.}$ per square centimeter. \Section{8.} Since the pressure in a given substance is evidently controlled by its internal physical condition only, and not by its form or mass, it follows that $p$~depends only on the temperature and the ratio of the mass~$M$ to the volume~$V$ \PageSep{5} (\ie\ the density), or on the reciprocal of the density, the volume of unit mass--- \[ \frac{V}{M} = v, \] which is called the specific volume of the substance. For every substance, then, there exists a characteristic relation--- \[ p = f(v, t), \] which is called the \emph{characteristic equation} of the substance. \index{Characteristic equation}% \index{Equation!characteristic}% For gases, the function~$f$ is invariably positive; for liquids and solids, however, it may have also negative values under certain circumstances. \Section{9.} \Topic{Perfect Gases.}---The characteristic equation assumes \index{Boyle's law}% \index{Gases, perfect}% \index{Laws:!Boyle's}% \index{Laws:!Mariotte's}% \index{Mariotte's law}% \index{Perfect gases}% its simplest form for the substances which we used in \SecRef{4} for the definition of temperature. If the temperature be kept constant, then, according to the Boyle-Mariotte law, the product of the pressure and the specific volume remains constant for gases--- \[ pv = T, \Tag{(1)} \] where $T$, for a given gas, depends only on the temperature. But if the pressure be kept constant, then, according to~\SecRef{3}, the temperature is proportional to the difference between the present volume~$v$ and the volume~$v_{0}$ at~$0°$; \ie--- \[ t = (v - v_{0})P, \Tag{(2)} \] where $P$~depends only on the pressure~$p$. Equation~\Eq{(1)} for~$v_{0}$ becomes \[ pv_{0} = T_{0}, \Tag{(3)} \] where $T_{0}$~is the value of the function~$T$, when $t = 0°$~C\@. Finally, as has already been mentioned in~\SecRef{4}, the expansion of all permanent gases on heating from~$0°$~C. to $1°$~C. is the same fraction~$\alpha$ (about~$\frac{1}{273}$) of their volume at~$0°$ \PageSep{6} (Gay-Lussac's law). Putting $t = 1$, we have $v - v_{0} = \alpha v_{0}$, \index{Laws:!Gay-Lussac's}% and equation~\Eq{(2)} becomes \index{Equation!characteristic}% \[ 1 = \alpha v_{0} P\Add{.} \Tag{(4)} \] By eliminating $P$,~$v_{0}$, and~$v$ from \Eq{(1)}, \Eq{(2)}, \Eq{(3)},~\Eq{(4)}, we obtain the temperature function of the gas--- \index{Temperature!absolute}% \[ T = T_{0} (1 + \alpha t), \] which is seen to be a linear function of~$t$. The characteristic equation~\Eq{(1)} becomes \[ p = \frac{T_{0}}{v} (1 + \alpha t). \] \Section{10.} The form of this equation is considerably simplified by shifting the zero of temperature, arbitrarily fixed in~\SecRef{3}, by $\dfrac{1}{\alpha}$ degrees, and calling the melting point of ice, not $0°$~C., but $\dfrac{1°}{\alpha}$~C. (\ie\ about $273°$~C.). For, putting $t + \dfrac{1}{\alpha} = \theta$ (absolute temperature), and the constant $\alpha T_{0} = C$, the \index{Absolute temperature}% characteristic equation becomes \index{Characteristic equation}% \[ p = \frac{C}{v} \theta = \frac{CM}{V} \theta\Add{.} \Tag{(5)} \] This introduction of \emph{absolute} temperature is evidently tantamount to measuring temperature no longer, as in~\SecRef{3}, by a change of volume, but by the volume itself. \Section{11.} The constant~$C$, which is characteristic for the perfect gas under consideration, can be calculated, if the specific volume~$v$ be known for any pair of values of~$\theta$ and~$p$ (\eg\ $0°$~and $1$~atmosphere). For different gases, taken at the same temperature and pressure, the constants~$C$ evidently vary directly as the specific volumes, or inversely as the densities~$\dfrac{1}{v}$. It may be affirmed, then, that, taken at the same temperature and pressure, the densities of all perfect gases bear a constant ratio to one another. A gas is, therefore, often characterized by the constant ratio which its \PageSep{7} density bears to that of a normal gas at the same temperature \index{Density!specific}% and pressure (\emph{specific density} relative to air or hydrogen). \index{Specific density}% At $0°$~C. ($\theta = 273°$) and $1$~atmosphere pressure, the densities of the following gases are: % [** TN: Removing dot leaders] \begin{center} \TableFont \begin{tabular}{ll} Hydrogen & $0.00008988~\dfrac{\Unit{gr.}}{\Unit{cm.}^{3}}$ \\ Oxygen & $0.0014291$ \\ Nitrogen & $0.0012507$ \\ Atmospheric nitrogen & $0.0012571$ \\ Air & $0.0012930$ \\ \end{tabular} \end{center} whence the corresponding values of~$C$ in absolute units can be readily calculated. All questions with regard to the behaviour of a substance when subjected to changes of temperature, volume, and pressure are completely answered by the characteristic equation of the substance. \Section{12.} \Topic{Behaviour under Constant Pressure} (Isopiestic or Isobaric Changes).---\emph{Coefficient of expansion} is the name \index{Coefficient!of elasticity}% \index{Coefficient!of expansion}% \index{Coefficient!of pressure}% \index{Elasticity, coefficient of}% \index{Expansion, coefficient of}% \index{Isobaric change}% \index{Isochoric change}% \index{Isopiestic change}% \index{Isopycnic change}% given to the ratio of the increase of volume for a rise of temperature of $1°$~C. to the volume at $0°$~C\@. This increase for a perfect gas is, according to~\Eq{(5)}, $\dfrac{CM}{p}$. The same equation~\Eq{(5)} gives the volume of the gas at $0°$~C. as $\dfrac{CM}{p} × 273$, hence the ratio of the two quantities, or the coefficient of expansion, is $\frac{1}{273} = \alpha$. \Section{13.} \Topic{Behaviour at Constant Volume} (Isochoric or Isopycnic Changes).---The \emph{pressure coefficient} is the ratio \index{Pressure coefficient}% of the increase of pressure for a rise of temperature of~$1°$ to the pressure at $0°$~C\@. For a perfect gas, this increase, according to equation~\Eq{(5)}, is~$\dfrac{CM}{V}$. The pressure at $0°$~C. is $\dfrac{CM}{V} × 273$, whence the required ratio, \ie\ the pressure coefficient, is~$\frac{1}{273}$, therefore equal to the coefficient of expansion~$\alpha$. \Section{14.} \Topic{Behaviour at Constant Temperature} (Isothermal Changes).---\emph{Coefficient of elasticity} is the ratio of an \PageSep{8} infinitely small increase of pressure to the resulting contraction of unit volume of the substance. In a perfect gas, according to equation~\Eq{(5)}, the contraction of volume~$V$, in consequence of an increase of pressure~$dp$, is \[ dV = \frac{CM\theta}{p^{2}}\, dp = \frac{V}{p}\, dp. \] The contraction of unit volume is therefore \[ -\frac{dV}{V} = \frac{dp}{p}, \] and the coefficient of elasticity of the gas is \index{Coefficient!of compressibility}% \[ \frac{\;\;dp\;\;}{\dfrac{dp}{p}} = p, \] that is, equal to the pressure. The reciprocal of the coefficient of elasticity, \ie\ the ratio of an infinitely small contraction of unit volume to the corresponding increase of pressure, is called the \emph{coefficient of compressibility}. \Section{15.} The three coefficients which characterize the behaviour of a substance subject to isopiestic, isopycnic, and isothermal changes are not independent of one another, but are in every case connected by a definite relation. The general characteristic equation, on being differentiated, gives \[ dp = \left(\frac{\dd p}{\dd \theta}\right)_{v} d\theta + \left(\frac{\dd p}{\dd v}\right)_{\theta} dv, \] where the suffixes indicate the variables to be kept constant while performing the differentiation. By putting $dp = 0$ we impose the condition of an isopiestic change, and obtain the relation between $dv$ and~$d\theta$ in isopiestic processes:--- \[ \left(\frac{\dd v}{\dd \theta}\right)_{p} = -\frac{\left(\dfrac{\dd p}{\dd \theta}\right)_{v}} {\left(\dfrac{\dd p}{\dd v}\right)_{\theta}}\Add{.} \Tag{(6)} \] \PageSep{9} For every state of a substance, one of the three coefficients, viz.\ of expansion, of pressure, or of compressibility, may therefore be calculated from the other two. Take, for example, mercury at $0°$~C. and under atmospheric pressure. Its coefficient of expansion is (\SecRef{12}) \[ \left(\frac{\dd v}{\dd \theta}\right)_{p} · \frac{1}{v_{0}} = 0.00018, \] its coefficient of compressibility in atmospheres (\SecRef{14}) is \[ -\left(\frac{\dd v}{\dd p}\right)_{\theta} · \frac{1}{v_{0}} = 0.000003, \] therefore its pressure coefficient in atmospheres (\SecRef{13}) is \index{Pressure coefficient!of mercury}% \[ \left(\frac{\dd p}{\dd \theta}\right)_{v} = -\left(\frac{\dd p}{\dd v}\right)_{\theta} · \left(\frac{\dd v}{\dd \theta}\right)_{p} = -\frac{\left(\dfrac{\dd v}{\dd \theta}\right)_{p}} {\left(\dfrac{\dd v}{\dd p}\right)_{\theta}} = \frac{0.00018}{0.000003} = 60. \] This means that an increase of pressure of $60$~atmospheres is required to keep the volume of mercury constant when heated from $0°$~C. to $1°$~C\@. \Section{16.} \Topic{Mixture of Perfect Gases.}---If any quantities of \index{Gas mixture}% \index{Mixture of gases}% the \emph{same} gas at the same temperatures and pressures be at first separated by partitions, and then allowed to come suddenly in contact with another by the removal of these partitions, it is evident that the volume of the entire system will remain the same and be equal to the sum-total of the partial volumes. Starting with quantities of \emph{different} gases, experience still shows that, when pressure and temperature are maintained uniform and constant, the total volume continues equal to the sum of the volumes of the constituents, notwithstanding the slow process of intermingling---diffusion---which \index{Diffusion}% takes place in this case. Diffusion goes on until the mixture has become at every point of precisely the same composition, \ie\ physically homogeneous. \Section{17.} Two views regarding the constitution of mixtures thus formed present themselves. Either we might assume \PageSep{10} that the individual gases, while mixing, split into a large number of small portions, all retaining their original volumes and pressures, and that these small portions of the different gases, without penetrating each other, distribute themselves evenly throughout the entire space. In the end each gas would still retain its original volume (partial volume), and all the gases would have the same common pressure. Or, we might suppose---and this view will be shown below (\SecRef{32}) to be the correct one---that the individual gases change and interpenetrate in every infinitesimal portion of the volume, and that after diffusion each individual gas, in so far as one may speak of such, fills the total volume, and is consequently under a lower pressure than before, diffusion. This so-called partial pressure of a constituent of a gas mixture can easily be calculated. \Section{18.} Denoting the quantities referring to the individual gases by suffixes---$\theta$~and $p$ requiring no special designation, as they are supposed to be the same for all the gases,---the characteristic equation~\Eq{(5)} gives for each gas before diffusion \[ p = \frac{C_{1}M_{1} \theta}{V_{1}};\quad p = \frac{C_{2}M_{2} \theta}{V_{2}};\quad \dots\Add{.} \] The total volume, \[ V = V_{1} + V_{2} + \dots, \] remains constant during diffusion. After diffusion we ascribe to each gas the total volume, and hence the partial pressures become \[ p_{1} = \frac{C_{1}M_{1} \theta}{V} = \frac{V_{1}}{V} p;\quad p_{2} = \frac{C_{2}M_{2} \theta}{V} = \frac{V_{2}}{V} p;\quad \dots\Add{,} \Tag{(7)} \] and by addition \[ p_{1} + p_{2} + \dots = \frac{V_{1} + V_{2} + \dots}{V} p = p\Add{.} \Tag{(8)} \] This is Dalton's law, that in a homogeneous mixture of \index{Dalton's law}% \index{Laws:!Dalton's}% \index{Partial pressures}% \PageSep{11} gases the pressure is equal to the sum of the partial pressures of the gases. It is also evident that \[ p_{1} : p_{2} : \dots = V_{1} : V_{2} : \dots = C_{1}M_{1} : C_{2}M_{2} : \dots\Add{,} \Tag{(9)} \] \ie\ the partial pressures are proportional to the volumes of the gases before diffusion, or to the partial volumes which the gases would have according to the first view of diffusion given above. \Section{19.} The characteristic equation of the mixture, according \index{Characteristic equation}% \index{Equation!characteristic}% to \Eq{(7)}~and \Eq{(8)}, is \begin{align*} p &= (C_{1}M_{1} + C_{2}M_{2} + \dots) \frac{\theta}{V} \\ &= \left(\frac{C_{1}M_{1} + C_{2}M_{2} + \dots}{M}\right) \frac{M}{V} \theta \Tag{(10)} \end{align*} which corresponds to the characteristic equation of a perfect gas with the following characteristic constant:--- \index{Characteristic constant}% \[ C = \frac{C_{1}M_{1} + C_{2}M_{2} + \dots}{M_{1} + M_{2} + \dots}\Add{.} \Tag{(11)} \] Hence the question as to whether a perfect gas is a chemically simple one, or a mixture of chemically different gases, cannot in any case be settled by the investigation of the characteristic equation. \Section{20.} The composition of a gas mixture is defined, either \index{Air, composition of}% by the ratios of the masses, $M_{1}$,~$M_{2}$,~\dots or by the ratios of the partial pressures $p_{1}$,~$p_{2}$,~\dots or the partial volumes $V_{1}$,~$V_{2}$,~\dots of the individual gases. Accordingly we speak of per cent.\ by weight or by volume. Let us take for example atmospheric air, which is a mixture of oxygen~(1) and ``atmospheric'' nitrogen~(2). The ratio of the densities of oxygen, ``atmospheric'' nitrogen and air is, according to~\SecRef{11}, \[ 0.0014291 : 0.0012571 : 0.0012930 = \frac{1}{C_{1}} : \frac{1}{C_{2}} : \frac{1}{C_{3}}\Add{.} \] \PageSep{12} Taking into consideration the relation~\Eq{(11)}--- \[ C = \frac{C_{1}M_{1} + C_{2}M_{2}}{M_{1} + M_{2}}, \] we find the ratio $M_{1} : M_{2} = 0.2998$, \ie\ $23.1$ per~cent.\ by weight of oxygen and $76.9$ per~cent.\ of nitrogen. Furthermore, \[ C_{1}M_{1} : C_{2}M_{2} = p_{1} : p_{2} = V_{1} : V_{2} = 0.2637 \] \ie\ $20.9$ per~cent.\ by volume of oxygen and $79.1$ per~cent.\ of nitrogen. \Section{21.} \Topic{Characteristic Equation of Other Substances.}---The characteristic equation of perfect gases, even in the case of the substances hitherto discussed, is only an approximation, though a close one, to the actual facts. A still further deviation from the behaviour of perfect gases is shown by the other gaseous bodies, especially by those easily condensed, which for this reason were formerly classed as \emph{vapours}. For these a modification in the form of the characteristic equation is necessary. It is worthy of notice, however, that the more rarefied the state in which we observe these gases, the less does their behaviour deviate from that of perfect gases, so that all gaseous substances, when sufficiently rarefied, may be said in general to act like perfect gases. The general characteristic equation of gases and vapours, for very large values of~$v$, will pass over, therefore, into the special form for perfect gases. \Section{22.} We may obtain by various graphical methods an idea of the character and magnitude of the deviations from the ideal gaseous state. An isothermal curve may, \eg, be drawn, taking $v$~and $p$ for some given temperature as the abscissa and ordinate, respectively, of a point in a plane. The entire system of isotherms gives us a complete representation of the characteristic equation. The more the behaviour of the vapour in question approaches that of a perfect gas, the closer do the isotherms approach those of equilateral hyperbolæ having the rectangular co-ordinate \PageSep{13} axes for asymptotes, for $pv = \const$ is the equation of an isotherm of a perfect gas. The deviation from the hyperbolic form yields at the same time a measure of the departure from the ideal state. \Section{23.} The deviations become still more apparent when \index{Deviation from perfect gases}% the isotherms are drawn taking the product~$pv$ (instead of~$p$) as the ordinate and say $p$~as the abscissa. Here a perfect gas has evidently for its isotherms straight lines parallel to the axis of abscissæ. In the case of actual gases, however, the isotherms slope gently towards a minimum value of~$pv$, the position of which depends on the temperature and the nature of the gas. For lower pressures (\ie\ to the left of the minimum), the volume decreases at a more rapid rate, with increasing pressure, than in the case of perfect gases; for higher pressures (to the right of the minimum), at a slower rate. At the minimum point the compressibility coincides with that of a perfect gas. In the case of hydrogen the minimum lies far to the left, and it has hitherto been possible to observe it only at very low temperatures. \Section{24.} To van~der Waals is due the first analytical formula \index{Equation!Van der Waals'}% \index{Van der Waals'!equation}% for the general characteristic equation, applicable also to the liquid state. He also explained physically, on the basis of the kinetic theory of gases, the deviations from the behaviour of perfect gases. As we do not wish to introduce here the hypothesis of the kinetic theory, we consider van~der Waals' equation merely as an approximate expression of the facts. His equation is \[ p = \frac{R\theta}{v - b} - \frac{a}{v^{2}}, \] where $R$,~$a$, and~$b$ are constants which depend on the nature of the substance. For large values of~$v$, the equation, as required, passes into that of a perfect gas; for small values of~$v$ and corresponding values of~$\theta$, it represents the characteristic equation of a liquid. Expressing $p$~in atmospheres and calling the specific \PageSep{14} volume~$v$ unity for $\theta = 273$ and $p = 1$, van~der Waals' constants for carbon dioxide are \index{Carbon dioxide!Van der Waals' constants for}% \[ R = 0.00369;\quad a = 0.00874;\quad b = 0.0023. \] As the volume of $1~\Unit{gr.}$ of carbon dioxide at $0°$~C. and atmospheric pressure is $505~\Chg{\Unit{c.c.}}{\Unit{cm.}^{3}}$, the values of~$v$ calculated from the formula must be multiplied by~$505$ to obtain the specific volumes in absolute units. \Section{25.} Van~der Waals' equation not being sufficiently \index{Equation!Clausius'}% \index{Van der Waals'!constants for \ce{CO2}}% accurate, Clausius supplemented it by the introduction of an additional constant. Clausius' equation is \index{Clausius' equation}% \[ p = \frac{R\theta}{v - a} - \frac{c}{\theta(v + b)^{2}}. \Tag{(12)} \] For large values of~$v$, this too approaches the ideal characteristic equation. In the same units as above, Clausius' constants for carbon dioxide are: \[ R = 0.003688;\quad a = 0.000843;\quad b = 0.000977;\quad c = 2.0935. \] Andrews' observations on the compressibility of gaseous \index{Andrews}% and liquid carbon dioxide are satisfactorily represented by Clausius' equation. \Section{26.} If we draw the system of isotherms with the aid of Clausius' equation, employing the graphical method described in~\SecRef{22}, the characteristic graphs for carbon dioxide---\Fig{1}---are \index{Apt, Dr.\ Richard}% obtained.\footnote {For the calculation and construction of the curves, I am indebted to Dr.\ Richard Apt.} For high temperatures the isotherms approach equilateral hyperbolæ, as may be seen from equation~\Eq{(12)}. In general, however, the isotherm is a curve of the third degree, three values of~$v$ corresponding to one of~$p$. Hence, in general, a straight line parallel to the axis of abscissæ intersects an isotherm in three points, of which two, as actually happens for large values of~$\theta$, may be imaginary. At high temperatures there is, consequently, \PageSep{15} % [** TN: ~10pt too tall in screen format] \ifthenelse{\boolean{ForPrinting}}{% \Figure[\textwidth]{1} }{% \Figure[3.5in]{1} } \index{Carbon dioxide!isotherms of}% \index{Isotherms of \ce{CO2}}% \PageSep{16} only one real volume corresponding to a given pressure, while at lower temperatures, there are three real values of the volume for a given pressure. Of these three values (indicated on the figure by $\alpha$,~$\beta$,~$\gamma$, for instance) only the smallest~($\alpha$) and the largest~($\gamma$) represent practically realizable states, for at the middle point~($\beta$) the pressure along the isotherm would increase with increasing volume, and the compressibility would accordingly be negative. Such a state has, therefore, only a theoretical signification. \Section{27.} The point~$\alpha$ corresponds to liquid carbon dioxide, and $\gamma$~to the gaseous condition at the temperature of the isotherm passing through the points and under the pressure measured by the ordinates of the line~$\alpha\beta\gamma$. In general only one of these states is stable (in the figure, the liquid state at~$\alpha$). For, if we compress gaseous carbon dioxide, enclosed in a cylinder with a movable piston, at constant temperature, \eg\ at $20°$~C., the gas assumes at first states corresponding to consecutive points on the $20°$~isotherm to the extreme right. The point representative of the physical state of the gas, then moves farther and farther to the left until it reaches a certain place~$C$. After this, further compression does not move the point beyond~$C$, but there now takes place a partial condensation of the substance---a splitting into a liquid and a gaseous portion. Both parts, of course, possess common pressure and temperature. The state of the gaseous portion continues to be characterized by the point~$C$, that of the liquid portion by the point~$A$ of the same isotherm. $C$~is called the saturation point of carbon dioxide \index{Saturation point}% gas for the particular temperature considered. Isothermal compression beyond~$C$ merely results in precipitating more of the vapour in liquid form. During this part of the isothermal compression no change takes place but the condensation of more and more vapour; the internal conditions (pressure, temperature, specific volume) of both parts of the substance are always represented by the two points $A$~and~$C$. At last, when all the vapour has been condensed, the whole substance is in the liquid condition~$A$, and again behaves \PageSep{17} as a homogeneous substance, so that further compression gives an increase of density and pressure along the isotherm. The substance will now pass through the point~$\alpha$ of the \index{Point!of inflection}% figure. On this side, as may be seen from the figure, the isotherm is much steeper than on the other, \ie\ the compressibility is much smaller. At times, it is possible to follow the isotherm beyond the point~$C$ towards the point~$\gamma$, and to prepare a so-called supersaturated vapour. Then only a more or less unstable condition of equilibrium is obtained, as may be seen from the fact that the smallest disturbance of the equilibrium is sufficient to cause an immediate condensation. The substance passes by a jump into the stable condition. Nevertheless, by the study of supersaturated vapours, the theoretical part of the curve also receives a direct meaning. \Section{28.} On any isotherm, which for certain values of~$p$ admits of three real values of~$v$, there are, therefore, two definite points, $A$~and~$C$, corresponding to the state of saturation. The position of these points is not immediately deducible from the graph of the isotherm. The propositions of thermodynamics, however, lead to a simple way of finding these points, as will be seen in~\SecRef{172}. The higher the temperature, \index{Temperature!critical}% the smaller becomes the region in which lines drawn parallel to the axis of abscissæ intersect the isotherm in three real points, and the closer will these three points approach one another. The transition to the hyperbola-like isotherms, which any parallel to the axis of abscissæ cuts in one point only, is formed by that particular isotherm on which the three points of intersection coalesce into one, giving a point of inflection. The tangent to the curve at \index{Inflection!point of}% this point is parallel to the axis of abscissæ. It is called the \emph{critical point} ($K$~of \Fig{1}) of the substance, and its \index{Critical point}% \index{Critical point!pressure}% \index{Critical point!specific volume}% \index{Critical point!temperature}% position indicates the critical temperature, the critical specific volume, and the critical pressure of the substance. Here there is no longer any difference between the saturated vapour and its liquid precipitate. Above the critical temperature and critical pressure, condensation does not exist, \PageSep{18} as the diagram plainly shows. Hence all attempts to condense hydrogen, oxygen, and nitrogen necessarily failed as long as the temperature had not been reduced below the critical temperature, which is very low for these gases. \Section{29.} It further appears from our figure that there is no definite boundary between the gaseous and liquid states, since from the region of purely gaseous states, as at~$C$, that of purely liquid ones, as at~$A$, may be reached on a circuitous path that nowhere passes through a state of saturation---on a curve, for instance, drawn around the critical point. Thus a vapour may be heated at constant volume above the critical temperature, then compressed at constant temperature below the critical volume, and finally cooled under constant pressure below the critical temperature. Condensation nowhere occurs in this process, which leads, nevertheless, to a region of purely liquid states. The earlier fundamental distinction between liquids, vapours, and gases should therefore be dropped as no longer tenable. A more modern proposal to denote as gaseous all states above the critical temperature, and as vaporous or liquid all others according as they lie to the right or left of the theoretical regions (\Fig{1}), has also this disadvantage, that thereby a boundary is drawn between liquid and gas on the one hand, and vapour and gas on the other hand, which has no physical meaning. The crossing of the critical temperature at a pressure other than the critical pressure differs in no way from the crossing of any other temperature. \Section{30.} The position of the critical point may be readily calculated from the general characteristic equation. According to~\SecRef{28} we have \[ \left(\frac{\dd p}{\dd v}\right)_{\theta} = 0,\quad\text{and}\quad \left(\frac{\dd^{2} p}{\dd v^{2}}\right)_{\theta} = 0. \] The first of these means that the tangent to the isotherm at~$K$ is parallel to the axis of abscissæ; and the second, that \PageSep{19} the isotherm has a point of inflection at~$K$. On the basis of Clausius' form of the characteristic equation~\Eq{(12)}, we obtain for the critical point \index{Critical point!of \ce{CO2}}% \[ \theta^{2} = \frac{8c}{27(a + b)R},\quad p^{2} = \frac{cR}{216(a + b)^{3}},\quad v = 3a + 2b.\footnotemark \] \footnotetext{Obtained as follows:--- \begin{align*} p &= \frac{R\theta}{v - a} - \frac{c}{\theta(v + b)^{2}}\Add{,} \tag*{(1)} \\ \left(\frac{\dd p}{\dd v}\right)_{\theta} &= -\frac{R\theta}{(v - a)^{2}} + \frac{2c}{\theta(v + b)^{3}} = 0\Add{,} \tag*{(2)} \\ \left(\frac{\dd^{2} p}{\dd v^{2}}\right)_{\theta} &= \frac{2R\theta}{(v - a)^{3}} - \frac{6c}{\theta(v + b)^{4}} = 0\Add{.} \tag*{(3)} \end{align*} From (2) and~(3), \begin{align*} v &= 3a + 2b\Add{.} \tag*{(4)} \intertext{Substituting (4) in (2) and reducing, we get} \theta^{2} &= \frac{8c}{27(a + b)R}\Add{.} \tag*{(5)} \intertext{And substituting (4) and (5) in (1) and reducing, we have} p^{2} &= \frac{cR}{216(a + b)^{3}}\Add{.} \tag*{(6)\ \Tr.} \end{align*}} These equations give for carbon dioxide from the above data \[ \theta = 304 = 273° + 31°,\quad p = 77~\Unit{atm.},\quad v = 2.27\, \frac{\Chg{\Unit{c.c.}}{\Unit{cm.}^{3}}}{\Unit{gr.}}. \] Qualitatively, all substances conform to these regularities, but the values of the constants differ widely. \Section{31.} Regarding the transition from the liquid to the \index{Theoretical regions}% solid state, the same considerations hold as for that from the gaseous to the liquid state. The system of isotherms might be drawn for this process, and it is probable that \emph{theoretical} regions and a critical point would be verified here also, if the means of experimental investigation were adequate. A continuous passage from the liquid to the solid state would then become possible along a path intersecting the critical isotherm on either side of the critical \PageSep{20} point. In fact, there are certain substances which under ordinary pressures pass without appreciable discontinuity from the solid to the liquid state (pitch, glass, etc.), while others possess for a definite temperature a definite pressure \index{Liquefaction pressure}% \index{Pressure!of liquefaction}% of liquefaction or pressure of solidification, at which the \index{Solidification!pressure}% substance splits into two portions of different densities. The pressure of liquefaction, however, varies with temperature at a much greater rate than the pressure of the saturated vapour. This view is physically justified, in particular by the experiments of Barus and Spring, in which the pressures \index{Barus}% \index{Spring}% were varied within wide limits. In its most complete form the characteristic equation would comprise the gaseous, liquid, and solid states simultaneously. No formula of such generality, however, has as yet been established for any substance. \Section{32.} \Topic{Mixtures.}---While, as shown in~\SecRef{19}, the characteristic \index{Dalton's law}% \index{Laws:!Dalton's}% \index{Mixtures}% equation of a mixture of perfect gases reduces in a simple manner to that of its components, no such simplification takes place, in general, when substances of any kind are mixed. Only for gases and vapours does Dalton's law hold, at least with great approximation, that the total pressure of a mixture is the sum of the partial pressures which each gas would exert if it alone filled the total volume at the given temperature. This law enables us to establish the characteristic equation of any gas mixture, provided that of the constituent gases be known. It also decides the question, unanswered in~\SecRef{17}, whether to the individual gases of a mixture common pressure and different volumes, or common volume and different pressures, should be ascribed. From the consideration of a vapour differing widely from an ideal gas, it follows that the latter of these views is the only one admissible. Take, for instance, atmospheric air and water vapour at $0°$~C. under atmospheric pressure. Here the water vapour cannot be supposed to be subject to a pressure of $1~\Unit{atm.}$, since at $0°$~C. no water vapour exists at this pressure. The only choice remaining is to assign to the air and water vapour a common volume \PageSep{21} (that of the mixture) and different pressures (partial pressures). For mixtures of solid and liquid substances no law of general validity has been found, that reduces the characteristic equation of the mixture to those of its constituents. \PageSep{22} \Chapter{II.}{Molecular Weight.} \index{Molecular weight}% \index{Weights molecular and equivalent}% \Section{33.} \First{In} the preceding chapter only such physical changes have been discussed as concern temperature, pressure, and density. The chemical constitution of the substance or mixture in question has been left untouched. Cases are frequent, however (much more so, in fact, than was formerly supposed) in which the chemical nature of a substance is altered by a change of temperature or pressure. The more recent development of thermodynamics has clearly brought out the necessity of establishing a fundamental difference between physical and chemical changes such as will exclude continuous transition from the one kind to the other (\cf\ \SecRef{42}, \textit{et seq.}, and~\SecRef{238}). It has, however, as yet not been possible to establish a practical criterion for distinguishing them, applicable to all cases. However strikingly most chemical processes differ from physical ones in their violence, suddenness, development of heat, changes of colour and other properties, yet there are, on the other hand, numerous changes of a chemical nature that take place with continuity and comparative slowness; for example, dissociation. One of the main tasks of physical chemistry in the near future will be the further elucidation of this essential difference.\footnote {In a word, we may, in a certain sense, say, that physical changes take place continuously, chemical ones, on the other hand, discontinuously. In consequence, the science of physics deals, primarily, with continuously varying numbers, the science of chemistry, on the contrary, with whole, or with simple rational numbers.} \Section{34.} Experience shows that all chemical reactions take place according to constant proportions by weight. A \PageSep{23} \index{Weights molecular and equivalent}% certain weight (strictly speaking, a mass) may therefore be used as a characteristic expression for the nature of a given chemically homogeneous substance, whether an element or a compound. Such a weight is called an \emph{equivalent weight}. It is arbitrarily fixed for one element---generally for hydrogen at $1~\Unit{gr.}$---and then the equivalent weight of any other element (\eg~oxygen) is that weight which will combine with $1~\Unit{gr.}$ of hydrogen. The weight of the compound thus formed is, at the same time, its equivalent weight. By proceeding in this way, the equivalent weights of all chemically homogeneous substances may be found. The equivalent weights of elements that do not combine directly with hydrogen can easily be determined, since in every case a number of elements can be found that combine directly with the element in question and also with hydrogen. The total weight of a body divided by its equivalent \index{Equivalents, number of}% weight is called the \emph{number of equivalents} contained in the body. Hence we may say that, in every chemical reaction, an equal number of equivalents of the different substances react with one another. \Section{35.} There is, however, some ambiguity in the above definition, since two elements frequently combine in more ways than one. For such cases there would exist several values of the equivalent weight. Experience shows, however, \index{Equivalent weight}% that the various possible values are always simple multiples or submultiples of any one of them. The ambiguity in the equivalent weight, therefore, reduces itself to multiplying or dividing that quantity by a simple integer. We must accordingly generalize the foregoing statement, that an equal number of equivalents react with one another, and say, that the number of equivalents that react with one another are in simple numerical proportions. Thus $16$~parts by weight of oxygen combine with $28$~parts by weight of nitrogen to form nitrous oxide, or with $14$~parts to form \index{Nitrogen!oxides}% \index{Oxides of nitrogen}% nitric oxide, or with $9\frac{1}{3}$~parts to form nitrous anhydride, or with $7$~parts to form nitrogen tetroxide, or with $5\frac{3}{5}$~parts to \PageSep{24} form nitric anhydride. Any one of these numbers may be assigned to nitrogen as its equivalent weight, if $16$~be taken as that of oxygen. They are in simple rational proportions, since \[ 28 : 14 : 9\tfrac{1}{3} : 7 : 5\tfrac{3}{5} = 60 : 30 : 20 : 15 : 12. \] \Section{36.} The ambiguity in the definition of the equivalent weight of nitrogen, exemplified by the above series of numbers, is removed by selecting a particular one of them to denote the \emph{molecular weight} of nitrogen. In the definition of the molecular weight as a quite definite quantity depending only on the particular state of a substance, and independent of possible chemical reactions with other substances, lies one of the most important and most fruitful achievements of theoretical chemistry. Its exact statement can at present be given only for special cases, viz.\ for perfect gases and dilute solutions. We need consider only the former of these, as we shall see from thermodynamics that the latter is also thereby determined. The definition of the molecular weight for a chemically homogeneous perfect gas is rendered possible by the further empirical law, that gases combine, not only in simple multiples of their equivalents, but also, at the same temperature and pressure, in simple volume proportions (Gay-Lussac). \index{Gay-Lussac}% \index{Laws:!Gay-Lussac's}% It immediately follows that the number of equivalents, contained in equal volumes of different gases, must bear simple ratios to one another. The values of these ratios, however, are subject to the above-mentioned ambiguity in the selection of the equivalent weight. The ambiguity is, however, removed by putting all these ratios $= 1$, \ie\ by establishing the condition that equal volumes of different gases shall contain an equal number of equivalents. Thus a definite choice is made from the different possible values, and a definite equivalent weight obtained for the gas, which is henceforth denoted as the \emph{molecular weight} of the gas. At the same time the number of equivalents in a quantity of the gas, which may be found by dividing the total weight by the molecular weight, is defined as the \emph{number of \PageSep{25} molecules} contained in that quantity. Hence, \emph{equal volumes of perfect gases at the same temperature and pressure contain an equal number of molecules} (Avogadro's law). The molecular \index{Avogadro's law}% \index{Laws:!Avogadro's}% \index{Molecules, number of}% weights of chemically homogeneous gases are, therefore, directly proportional to the masses contained in equal volumes, \ie\ to the densities. The ratio of the densities is equal to the ratio of the molecular weights. \Section{37.} Putting the molecular weight of hydrogen $= m_{0}$, that of any other chemically homogeneous gas must be equal to $m_{0}$~multiplied by its specific density relative to hydrogen (\SecRef{11}). The following table gives the specific densities relative to hydrogen, and the molecular weights of several gases:--- \begin{center} \TableFont \begin{tabular}{lr<{\qquad} r<{\qquad}} & \ColHead{Specific Density.} & \ColHead{Molecular Weight.} \\ Hydrogen & $1.0$ & $m_{0}$ \\ Oxygen & $16.0$ & $16.0\, m_{0}$ \\ Nitrogen & $14.0$ & $14.0\, m_{0}$ \\ Water vapour & $9.0$ & $9.0\, m_{0}$ \\ Ammonia & $8.5$ & $8.5\, m_{0}$ \\ \end{tabular} \end{center} Now, since water vapour consists of $1$~part by weight of hydrogen and $8$~parts by weight of oxygen, the molecule of water vapour, $9\, m_{0}$, must consist of $m_{0}$~parts by weight of hydrogen and $8\, m_{0}$~parts by weight of oxygen---\ie, according to the above table, of one molecule of hydrogen and half a molecule of oxygen. In the same manner ammonia, according to analysis, consisting of $1$~part by weight of hydrogen and $4\frac{2}{3}$~parts by weight of nitrogen, its molecule $8.5\, m_{0}$ must necessarily contain $1.5\, m_{0}$~parts by weight of hydrogen and $7\, m_{0}$~parts by weight of nitrogen---\ie, according to the table, $1\frac{1}{2}$~molecules of hydrogen and $\frac{1}{2}$~molecule of nitrogen. Thus Avogadro's law enables us to give in quite definite numbers the molecular quantities of each constituent present in the molecule of any chemically homogeneous gas, provided we know its density and its chemical composition. \Section{38.} The smallest weight of a chemical element entering \index{Atom, definition of}% into the molecules of its compounds is called an \emph{atom}. \PageSep{26} Hence half a molecule of hydrogen is called an atom of hydrogen,~\ce{H}; similarly, half a molecule of oxygen an atom of oxygen,~\ce{O}; and half a molecule of nitrogen an atom of nitrogen,~\ce{N}. The diatomic molecules of these substances are represented by \ce{H2}, \ce{O2},~\ce{N2}. An atom of mercury, on the contrary, is equal to a whole molecule, because in the molecules of its compounds no fractions of the molecular weight of mercury vapour \Erratum{occurs}{occur}. It is usual to put the atomic weight of hydrogen $\ce{H} = 1$. Then its molecular weight becomes $\ce{H2} = m_{0} = 2$, and the molecular weights of our table become:\Add{---} \begin{center} \TableFont \settowidth{\TmpLen}{Molecular} \begin{tabular}{l>{\qquad}l} & \ColHead{Molecular Weight.} \\ Hydrogen & $\Z2 = \ce{H2}$ \\ Oxygen & $32 = \ce{O2}$ \\ Nitrogen & $28 = \ce{N2}$ \\ Water vapour & $18 = \ce{H2O}$ \\ Ammonia & $17 = \ce{H3N}$ \\ \end{tabular} \end{center} \Section{39.} In general, then, the molecular weight of a chemically homogeneous gas is twice its density relative to hydrogen. Conversely, the molecular weight,~$m$, of a gas being known, its specific density, and consequently the constant~$C$ in the characteristic equation~\Eq{(5)}, can be calculated. Denoting all quantities referring to hydrogen by the suffix~$0$, we have, at any temperature and pressure, for hydrogen, \[ p = \frac{C_{0} \theta}{v_{0}}, \] for any other gas at the same temperature and pressure, \begin{align*} p &= \frac{C\theta}{v}, \\ \therefore C : C_{0} &= \frac{1}{v_{0}} : \frac{1}{v} = m_{0} : m, \intertext{or} C &= \frac{m_{0}C_{0}}{m}\Add{.} \Tag{(13)} \end{align*} \PageSep{27} Now $m_{0} = 2$, and $C_{0}$~is to be calculated from the density of hydrogen at $0°$~C. and atmospheric pressure (\SecRef{11}). Since \begin{gather*} \frac{1}{v_{0}} = 0.00008988,\quad p = 1013650,\quad \theta = 273, \\ \therefore C = \frac{m_{0}C_{0}}{m} = \frac{m_{0}}{m} · \frac{pv_{0}}{\theta} = \frac{2 · 1013650}{m · 273 · 0.00008988} = \frac{82600000}{m}\Add{.} \end{gather*} Putting, for shortness, $82600000 = R$, the characteristic equation of a chemically homogeneous perfect gas of molecular \index{Gas!constant}% \index{Gas!volume}% weight~$m$ becomes \[ p = \frac{R}{m} · \frac{\theta}{v}, \Tag{(14)} \] where $R$, being independent of the nature of the individual gas, is generally called the absolute gas constant. The molecular weight may be deduced directly from the characteristic equation by the aid of the constant~$R$, since \[ m = \frac{R}{C}\Add{.} \Tag{(15)} \] Since $v = \dfrac{V}{M}$, we have \[ V = \frac{R\theta}{p} · \frac{M}{m}. \] But $\dfrac{M}{m}$~is the quantity defined above as the number of molecules in the gas, and, therefore, if $\dfrac{M}{m} = n$, \[ V = \frac{R\theta}{p} · n, \] which means that at a given temperature and pressure the volume of a quantity of gas depends only on the number of the molecules present, and not at all on the nature of the gas. \Section{40.} In a mixture of chemically homogeneous gases of \index{Avogadro's law}% \PageSep{28} \index{Gas mixture!volume}% molecular weights $m_{1}$,~$m_{2}$,~\dots the relation between the partial pressures is, according to~\Eq{(9)}, \[ p_{1} : p_{2} : \dots = C_{1}M_{1} : C_{2}M_{2} \Add{:} \dots\Add{.} \] But in~\Eq{(15)} we have \begin{gather*} C_{1} = \frac{R}{m_{1}};\quad C_{2} = \frac{R}{m_{2}};\quad \dots\Add{,} \\ \therefore p_{1} : p_{2} : \dots = \frac{M_{1}}{m_{1}} : \frac{M_{2}}{m_{2}} : \dots = n_{1} : n_{2} : \dots\Add{,} \end{gather*} \ie\ the ratio of the partial pressures is also the ratio of the number of molecules of each gas present. Equation~\Eq{(10)} gives for the total volume \begin{align*} V &= \frac{(C_{1}M_{1} + C_{2}M_{2} + \dots)\theta}{p} \\ &= \frac{R\theta}{p} \left(\frac{M_{1}}{m_{1}} + \frac{M_{2}}{m_{2}} + \dots\right) \\ &= \frac{R\theta}{p} (n_{1} + n_{2} + \dots) \\ &= \frac{R\theta}{p}\, n\Add{.} \Tag{(16)} \end{align*} The volume of the mixture is therefore determined by the total number of the molecules present, just as in the case of a chemically homogeneous gas. \Section{41.} It is evident that we cannot speak of the molecular weight of a mixture. Its \emph{apparent} molecular weight, however, \index{Molecular weight!apparent}% may be defined as the molecular weight which a chemically homogeneous gas would have if it contained in the same mass the same number of molecules as the mixture. If we denote the apparent molecular weight by~$m$, we have \[ \frac{M_{1} + M_{2} + \dots}{m} = \frac{M_{1}}{m_{1}} + \frac{M_{2}}{m_{2}} + \dots \] and \[ m = \frac{M_{1} + M_{2} + \dots}{\dfrac{M_{1}}{m_{1}} + \dfrac{M_{2}}{m_{2}} + \dots}\Add{.} \] \PageSep{29} The apparent molecular weight of air may thus be calculated. Since \[ m_{1} = \ce{O2} = 32;\quad m_{2} = \ce{N2} = 28;\quad M_{1} : M_{2} = 0.3 \] we have \[ m = \frac{0.3 + 1}{\dfrac{0.3}{32} + \dfrac{1}{28}} = 28.8, \] which is somewhat larger than the molecular weight of nitrogen. \Section{42.} The characteristic equation of a perfect gas, whether \index{Membranes, semipermeable}% \index{Semipermeable membranes}% chemically homogeneous or not, gives, according to~\Eq{(16)}, the total number of molecules, but yields no means of deciding whether or not these molecules are all of the same kind. In order to answer this question, other methods must be resorted to, none of which, however, is practically applicable to all cases. A decision is often reached by an observation of the process of diffusion through a porous or, better, a semipermeable membrane. The individual gases of a mixture will separate from each other by virtue of the differences in their velocities of diffusion, which may even sink to zero in the case of \Chg{semi-permeable}{semipermeable} membranes, and thus disclose the inhomogeneity of the substance. The chemical constitution of a gas may often be inferred from the manner in which it originated. It is by means of the expression for the entropy (\SecRef{237}) that we first arrive at a fundamental definition for a chemically homogeneous gas. \Section{43.} Should a gas or vapour not obey the laws of perfect gases, or, in other words, should its specific density depend on the temperature or the pressure, Avogadro's definition of molecular weight is nevertheless applicable. The number of molecules in this case, instead of being a constant, will be dependent upon the momentary physical condition of the substance. We may, in such cases, either assume the number of molecules to be variable, or refrain from applying Avogadro's definition of the number of molecules. In other words, the cause for the deviation from the ideal state may be \PageSep{30} sought for either in the chemical or physical conditions. The latter view preserves the chemical nature of the gas. The molecules remain intact under changes of temperature and pressure, but the characteristic equation is more complicated than that of Boyle and Gay-Lussac like that, for example, of van~der Waals or of Clausius. The other view differs essentially from this, in that it represents any gas, not obeying the laws of perfect gases, as a mixture of various kinds of molecules (in nitrogen peroxide \ce{N2O4} and~\ce{NO2}, in phosphorus \index{Nitrogen!peroxide}% pentachloride \ce{PCl5}, \ce{PCl3}, and~\ce{Cl2}). The volume of \index{Phosphorus pentachloride}% these is supposed to have at every moment the exact value theoretically required for the total number of molecules of the mixture of these gases. The volume, however, does not vary with temperature and pressure in the same way as that of a perfect gas, because chemical reactions take place between the different kinds of molecules, continuously altering the number of each kind present, and thereby also the total number of molecules in the mixture. This hypothesis has proved fruitful in cases of great differences of density---so-called \index{Density!abnormal vapour}% abnormal vapour densities---especially where, beyond \index{Abnormal vapour densities}% \index{Vapour densities, abnormal}% a certain range of temperature or pressure, the specific density once more becomes constant. When this is the case, the chemical reaction has been completed, and for this reason the molecules henceforth remain unchanged. \Erratum{Hydrobromamylene}{Amylene hydrobromide}, \Erratum{\index{Hydrobromamylene}}{\index{Amylene hydrobromide}}% for instance, acts like a perfect gas below $160°$ and above $360°$, but shows only half its former density at the latter temperature. The doubling of the number of molecules corresponds to the equation \[ \ce{C5H11Br} = \ce{C5H10 + HBr}. \] Mere insignificant deviations from the laws of perfect gases are generally attributed to physical causes---as, \eg, in water vapour and carbon dioxide---and are regarded as the forerunners of condensation. The separation of chemical from physical actions by a principle which would lead to a more perfect definition of molecular weight for variable vapour densities, cannot be accomplished at the present time. The increase in the specific density which many \PageSep{31} vapours exhibit near their point of condensation might just as well be attributed to such chemical phenomena as the formation of double or multiple molecules. In fact, differences of opinion exist in a number of such cases. The molecular weight of sulphur vapour below $800°$, for instance, \index{Sulphur}% is generally assumed to be $\ce{S6} = 192$; but some assume a mixture of molecules $\ce{S8} = 256$ and $\ce{S2} = 64$, and others still different mixtures. In doubtful cases it is safest, in general, to leave this question open, and to admit both chemical and physical changes as causes for the deviations from the laws of perfect gases. This much, however, may be affirmed, that for small densities the physical influences will be of far less moment than the chemical ones, for, according to experience, all gases approach the ideal condition as their densities decrease (\SecRef{21}). This is an important point, which we will make use of later. \PageSep{32} \Chapter{III.}{Quantity of Heat.} \index{Heat!quantity}% \index{Heat!unit}% \index{Quantity of heat}% \index{Unit of heat}% \Section{44.} If we plunge a piece of iron and a piece of lead, both of equal weight and at the same temperature ($100°$~C.), into two precisely similar vessels containing equal quantities of water at $0°$~C., we find that, after thermal equilibrium has been established in each case, the vessel containing the iron has increased in temperature much more than that containing the lead. Conversely, a quantity of water at~$100°$ is cooled to a much lower temperature by a piece of iron at~$0°$, than by an equal weight of lead at the same temperature. This phenomenon leads to a distinction between \emph{temperature} and \emph{quantity of heat}. As a measure of the heat given out or received by a body, we take the increase or decrease of temperature which some \emph{normal} substance (\eg\ water) undergoes when it alone is in contact with the body, provided all other causes of change of temperature (as compression, etc.)\ are excluded. The quantity of heat given out by the body is assumed to be equal to that received by the normal substance, and \textit{vice versâ}. The experiment described above proves, then, that a piece of iron in cooling through a given interval of temperature gives out more heat than an equal weight of lead (about four times as much), and conversely, that, in order to bring about a certain increase of temperature, iron requires a correspondingly larger supply of heat than lead. \Section{45.} It was, in general, customary to take as the unit of heat that quantity which must be added to $1~\Unit{gr.}$ of water to raise its temperature from $0°$~C. to $1°$~C. (zero \PageSep{33} calorie). This is almost equal to the quantity of heat which \index{Calorie!laboratory}% \index{Calorie!large}% \index{Calorie!mean}% \index{Calorie!small}% \index{Calorie!zero}% \index{Heat!capacity}% \index{Heat!specific (definition)}% will raise $1\Unit{gr.}$ of water $1°$~C. at any temperature. The refinement of calorimetric measurements has since made it necessary to take account of the initial temperature of the water, and it is often found convenient to define the calorie as that quantity of heat which will raise $1~\Unit{gr.}$ of water of mean laboratory temperature ($15°$~to~$20°$) $1$~degree of the Centigrade scale. This laboratory calorie is about $\dfrac{1}{1.006}$~of a zero calorie. Finally, a \emph{mean calorie} has been \index{Zero!calorie}% introduced, namely, the hundredth part of the heat required to raise $1~\Unit{gr.}$ of water from $0°$~C. to $100°$~C\@. The mean calorie is about equal to the zero calorie. Besides these so-called \emph{small} calories, there are a corresponding number of \emph{large} or kilogram calories, which contain $1000$~small calories. \Section{46.} The ratio of~$Q$, the quantity of heat each gram of a substance receives, to~$\Delta\theta$, the corresponding increase of temperature, is called the \emph{mean specific heat}, or \emph{mean heat \index{Specific heat}% capacity} of $1~\Unit{gr.}$ of the substance between the initial and final temperatures of the process--- \[ \frac{Q}{\Delta\theta} = c_{m}. \] Hence, the mean heat capacity of water between $0°$~and $1°$ is equal to one zero calorie. Passing to infinitely small differences of temperature, the specific heat of a substance, at the temperature~$\theta$, becomes \[ \frac{Q}{d\theta} = c. \] This, in general, varies with temperature, but very slowly for most substances. It is usually permissible to put the specific heat at a certain temperature equal to the mean specific heat of an adjoining interval of moderate size. \Section{47.} The heat capacity of solids and liquids is very \PageSep{34} nearly independent of any variations of external pressure that may take place during the process of heating. Hence the definition of the heat capacity is not, usually, encumbered \index{Atomic heat}% \index{Heat!atomic}% \index{Heat!molecular}% \index{Molecular heat}% with a condition regarding pressure. The specific heat of gases, however, is influenced considerably by the conditions of the heating process. In this case the definition of specific heat would, therefore, be incomplete without some statement as to the accompanying conditions. Nevertheless, we speak of the specific heat of a gas, without further specification, when we mean its specific heat at constant (atmospheric) pressure, as this is the value most readily determined. \Section{48.} That the heat capacities of different substances should be referred to unit mass is quite arbitrary. It arises from the fact that quantities of matter can be most easily compared by weighing them. Heat capacity might, quite as well, be referred to unit volume. It is more rational to compare masses which are proportional to the molecular and atomic weights of substances, for then certain regularities at once become manifest. The corresponding heat capacities are obtained by multiplying the specific heats (per unit mass) by the molecular or atomic weights. The values thus obtained are known as the \emph{molecular} or \emph{atomic heats}. \Section{49.} The chemical elements, especially those of high \index{Laws:!Dulong and Petit's}% atomic weight, are found to have nearly the constant atomic heat of~$6.4$ (Dulong and Petit). It cannot be claimed that this law is rigorously true, since the heat capacity depends on the molecular constitution, as in the case of carbon, and on the state of aggregation, as in the case of mercury, as well as on the temperature. The effect of temperature is especially marked in the elements, carbon, boron, and silicon, which show the largest deviations from Dulong and Petit's law. The conclusion is, however, justified, that Dulong and Petit's law is founded on some more general \index{Dulong and Petit's law}% law of nature, which has not yet been formulated. \PageSep{35} \Section{50.} Similar regularities, as appear in the atomic heats of \index{Laws:!Neumann's (Regnault)}% \index{Neumann, F.}% elements, are also found in the molecular heats of compounds, especially with compounds of similar chemical constitution. According to F.~Neumann's law, subsequently confirmed by Regnault, compounds of similar constitution, when solid, have equal molecular heats. Joule and Woestyn further extended this law by showing that the molecular heat is merely the sum of the atomic heats, or that in any combination every element preserves its atomic heat, whether or not the latter be~$6.4$, according to Dulong and Petit's law. This relation also is only approximately true. \Section{51.} Since all calorimetric measurements, according to \SecRef{44}, extend only to quantities of heat imparted to bodies or given out by them, they do not lead to any conclusion as to the total amount of heat \emph{contained in} a body of given temperature. It would be absurd to define the heat contained in a body of given temperature, density, etc., as the number of calories absorbed by the body in its passage from some normal state into its present state, for the quantity thus defined would assume different values according to the way in which the change was effected. A gas at~$0°$ and atmospheric pressure can be brought to a state where its temperature is~$100°$ and its pressure $10$~atmospheres, either by heating to~$100°$ under constant pressure, and then compressing at constant temperature; or by compressing isothermally to $10$~atmospheres, and then heating isopiestically to~$100°$; or, finally, by compressing and heating simultaneously or alternately in a variety of ways. The total number of calories absorbed would in each case be different (\SecRef{77}). It is seen, then, that it is useless to speak of a certain quantity of heat which must be applied to a body in a given state to bring it to some other state. If the ``total heat contained in a body'' is to be expressed numerically, as is done in the kinetic theory of heat, where the heat of a body is defined as the total energy of its internal motions, it must not be interpreted as the sum-total of the quantities of heat applied to the body. As we \PageSep{36} shall make no use of this quantity in our present work, no definition of it need be attempted. \Section{52.} In contrast to the above representation of the facts, the older (Carnot's) theory of heat, which started from the \index{Carnot's!theory}% \index{Heat!total}% hypothesis that heat is an indestructible substance, necessarily reached the conclusion that the ``heat contained in a body'' depends solely on the number of calories absorbed or given out by it. The heating of a body by other means than direct application of heat, by compression or by friction for instance, according to that theory produces no change in the ``total heat.'' To explain the rise of temperature which takes place notwithstanding, it was necessary to make the assumption that compression and friction so diminish the body's heat capacity, that the same amount of heat now produces a higher temperature, just as, for example, a moist sponge appears more moist if compressed, although the quantity of liquid in the sponge remains the same. In the meantime, Rumford and Davy proved by direct experiment \index{Davy}% \index{Rumford}% that bodies, in which any amount of heat can be generated by an adequate expenditure of work, do not in the least alter their heat capacities with friction. Regnault, likewise, showed, by accurate measurements, that the heat capacity of gases is independent of or only very slightly dependent on volume; that it cannot, therefore, diminish, in consequence of compression, as much as Carnot's theory would require. Finally, W.~Thomson and Joule have \index{Joule}% \index{Thomson}% demonstrated by careful experiments that a gas, when expanding without overcoming external pressure, undergoes no change of temperature, or an exceedingly small one (\cf~\SecRef{70}), so that the cooling of gases generally observed when they expand is not due to the increase of volume \textit{per~se}, but to the work done in the expansion. Each one of these experimental results would by itself be sufficient to disprove the hypothesis of the indestructibility of heat, and to overthrow the older theory. \Section{53.} While, in general, the heat capacity varies continuously with temperature, every substance possesses, \PageSep{37} \index{Heat effect}% \index{Singular values}% under certain external pressures, so-called \emph{singular} values of temperature, for which the heat capacity, together \index{Heat!of fusion}% \index{Heat!of sublimation}% \index{Heat!of vaporization}% with other properties, is discontinuous. At such temperatures the heat absorbed no longer affects the entire body, but only one of the parts into which it has split; and it no longer serves to increase the temperature, but simply to alter the state of aggregation, \ie\ to melt, evaporate, or sublime. Only when the entire substance has again become homogeneous will the heat imparted produce a rise in temperature, and then the heat capacity becomes once more capable of definition. The quantity of heat necessary to change $1$~gram of a substance from one state of aggregation to another is called the \emph{latent heat}, in particular, the \emph{heat of \index{Latent heat}% fusion}, \emph{of vaporization}, or \emph{of sublimation}. The same amount of heat is set free when the substance returns to its former state of aggregation. Latent heat, as in the case of specific heat, is best referred, not to unit mass, but to molecular or atomic weight. Its amount largely depends on the external conditions under which the process is carried out (\SecRef{47}), \index{Endothermal process}% \index{Exothermal process}% \index{Process!endothermal}% \index{Process!exothermal}% constant pressure being the most important condition. \Section{54.} Like the changes of the state of aggregation, all processes involving mixture, or solution, and all chemical reactions are accompanied by an evolution of heat of greater or less amount, which varies according to the external conditions. This we shall henceforth designate as the \emph{heat effect} (Wärmetonung) of the process under consideration, in particular as the heat of mixture, of solution, of combination, of dissociation, etc. It is reckoned \emph{positive} when heat is set free or developed, \ie\ given out by the body (exothermal processes); \emph{negative}, when heat is absorbed, or rendered latent, \ie\ taken up by the body (endothermal processes). \PageSep{38} \Part{II.}{The First Fundamental Principle of Thermodynamics.} \Chapter{I.}{General Exposition.} \index{First law of thermodynamics}% \Section{55.} \First{The} \emph{first law} of thermodynamics is nothing more than the principle of the conservation of energy applied \index{Conservation of energy}% \index{Energy!conservation of}% to phenomena involving the production or absorption of heat. Two ways lead to a deductive proof of this principle. We may take for granted the correctness of the mechanical view of nature, and assume that all changes in nature can be reduced to motions of material points between which there act forces which have a potential. Then the principle of energy is simply the well-known mechanical theorem of kinetic energy, generalized to include all natural processes. Or we may, as is done in this work, leave open the question concerning the possibility of reducing all natural processes to those of motion, and start from the fact which has been tested by centuries of human experience, and repeatedly verified, viz.\ that \emph{it is in no way possible, either by mechanical, thermal, chemical, or other devices, to obtain perpetual motion}, \ie\ it is impossible to construct an engine which will work in a cycle and produce continuous work, or kinetic energy, from nothing. We shall not attempt to show how this single fact of experience, quite independent of the mechanical view of nature, serves to prove the principle of energy in its generality, mainly for the reason that the validity of the energy principle is nowadays no longer disputed. It \PageSep{39} will be different, however, in the case of the \emph{second law} of thermodynamics, the proof of which, at the present stage of the development of our subject, cannot be too carefully presented. The general validity of this law is still contested from time to time, and its significance variously interpreted, even by the adherents of the principle. \Section{56.} \Topic{The energy} of a body, or system of bodies, is \index{Energy!definition of}% \index{External conditions of equilibrium!effect}% a magnitude depending on the momentary condition of the system. In order to arrive at a definite numerical expression for the energy of the system in a given state, it is necessary to fix upon a certain \emph{normal} arbitrarily selected state (\eg\ $0°$~C. and atmospheric pressure). The energy of the system in a given state, referred to the arbitrarily selected normal state, is then equal to \emph{the algebraic sum of the mechanical equivalents of all the effects produced outside the system when it passes in any way from the given to the normal state}. The energy of a system is, therefore, sometimes briefly denoted as the faculty to produce external effects. Whether or not the energy of a system assumes different values according as the transition from the given to the normal state is accomplished in different ways is not implied in the above definition. It will be necessary, however, for the sake of completeness, to explain the term ``mechanical equivalent of an external effect.'' \Section{57.} Should the external effect be mechanical in nature---should it consist, \eg, in lifting a weight, overcoming atmospheric pressure, or producing kinetic energy---then its mechanical equivalent is simply equal to the mechanical work done by the system on the external body (weight, atmosphere, projectile). It is positive if the displacement take place in the direction of the force exercised by the system---when the weight is lifted, the atmosphere pushed back, the projectile discharged,---negative in the opposite sense. But if the external effect be thermal in nature---if it consist, \eg, in heating surrounding bodies (the atmosphere, \PageSep{40} a calorimetric liquid, etc.)---then its mechanical equivalent is \index{Mechanical equivalent!of heat}% equal to the number of calories which will produce the same rise of temperature in the surrounding bodies multiplied by an absolute constant, which depends only on the units of heat and mechanical work, the so-called \emph{mechanical equivalent of heat}. This proposition, which appears here only as a definition, receives through the principle of the conservation of energy a physical meaning, which may be \index{Conservation of energy}% \index{Energy!conservation of}% put to experimental test. \Section{58.} \Topic{The Principle of the Conservation of Energy} asserts, generally and exclusively, that the energy of a system in a given state, referred to a fixed normal state, has a quite \Erratum{different}{definite} value; in other words---substituting the definition given in~\SecRef{56}---that the algebraic sum of the mechanical equivalents of the external effects produced outside the system, when it passes from the given to the normal state, is independent of the manner of the transformation. On passing into the normal state the system thus produces a definite total of effects, as measured in mechanical units, and it is this sum---the ``work-value'' of the external effects---that represents the energy of the system in the given state. \Section{59.} The validity of the principle of the conservation of energy may be experimentally verified by transferring a system in various ways from a given state to a certain other state, which may here be designated as the normal state, and measuring the mechanical equivalents of all external effects in each case. Special care must be taken, however, that the initial state of the system is the same each time, and that none of the external effects is overlooked or taken into account more than once. \Section{60.} As a first application we shall discuss Joule's famous \index{Joule's experiments|(}% experiments, in which the external effects produced by weights falling from a certain height were compared, first, when performing only mechanical work (\eg\ lifting a load), and second, when by suitable contrivances generating heat \PageSep{41} by friction. The initial and final position of the weights may be taken as the two states of the system, the work or heat produced, as the external effects. The first case, where the weights produce only mechanical work, is simple, and requires no experiment. Its mechanical equivalent is the product of the sum of the weights, and the height through which they fall. The second case requires accurate measurement of the increase of temperature, which the surrounding substances (water, mercury) undergo in consequence of the friction, as well as of their heat capacities, for the determination of the number of calories which will produce in them the same rise of temperature. It is, of course, entirely immaterial what our views may be with regard to the details of the frictional generation of heat, or with regard to the ultimate form of the heat thus generated. The only point of importance is that the state produced in the liquid by friction is identical with a state produced by the absorption of a definite number of calories. Joule, by equating the mechanical work, corresponding \index{Gram-calorie, mechanical equivalent of}% \index{Mechanical equivalent!of a gram-calorie}% to the fall of the weights, to the mechanical equivalent of the heat produced by friction, showed that the mechanical equivalent of a gram-calorie is, under all circumstances, equal to the work done in lifting a weight of a gram through a height of $423.55$~meters. That all his experiments with different weights, different calorimetric substances, and different temperatures, led to the same value, goes to prove the correctness of the principle of the conservation of energy. \Section{61.} In order to determine the mechanical equivalent of heat in absolute units, we must bear in mind that Joule's result refers to laboratory calories (\SecRef{45}), and the readings of a mercury thermometer. At the temperature of the laboratory, $1°$~of the mercury thermometer represents about $\dfrac{1}{1.007}$~of $1°$~of the gas thermometer. A calorie referred to the gas thermometer has, therefore, a mechanical equivalent of $423.55 × 1.007 = 427$. \PageSep{42} The acceleration of gravity must also be considered, since raising a gram to a certain height represents, in general, different amounts of work in different latitudes. The absolute value of the work done is obtained by multiplying the weight, \ie\ the product of the mass and the acceleration of gravity, by the height of fall. The following table gives the mechanical equivalent of heat in the \index{Mechanical equivalent!of heat!in absolute units}% different calories:--- \begin{center} \TableFont \begin{tabular}{l|c|c} \hline %[** TN: Re-breaking column headings] \ColumnHeading{Unit of heat referred}{Unit of heat referred \\ to gas thermometer.} & \ColumnHeading{Corresponding height in}{Corresponding height in \\ meters to which $1~\Unit{gr.}$ \\ must be raised in places of mean latitude.} & \ColumnHeading{equivalent (C.G.S.)}{Absolute value of \\ the mechanical \\ equivalent (C.G.S. \\ system, erg).} \\ \hline \Strut Laboratory caloric & $427$ & $419 × 10^{5}$ \\ Zero calorie & $430$ & $422 × 10^{5}$ \\ \hline \end{tabular} \end{center} The numbers of the last column are derived from those of the preceding one by multiplying by $98,100$, to reduce grams to dynes, and meters to centimeters. Joule's results have been substantially confirmed by recent careful measurements \index{Joule's experiments|)}% by Rowland and others. \Section{62.} The determination of the mechanical equivalent of heat enables us to express quantities of heat in ergs directly, instead of calories. The advantage of this is, that a quantity of heat is not only proportional to, but directly equal to its mechanical equivalent, whereby the mathematical expression for the energy is greatly simplified. This unit of heat will be used in all subsequent equations. The return to calories is, at any time, readily accomplished by dividing by $419 × 10^{5}$. \Section{63.} Some further propositions immediately follow from \index{First law of thermodynamics}% the above exposition of the principle of energy. The energy, as stated, depends on the momentary condition of the system. To find the change of energy, $U_{1} - U_{2}$, accompanying the transition of the system from a state~$1$ \PageSep{43} to a state~$2$, we should, according to the definition of the energy in~\SecRef{58}, have to measure~$U_{1}$ as well as~$U_{2}$ by the \index{Energy!change of}% mechanical equivalent of the external effects produced in passing from the given states to the normal state. But, supposing we so arrange matters that the system passes from state~$1$, through state~$2$, into the normal state, it is evident then that $U_{1} - U_{2}$ is simply the mechanical equivalent of the external effects produced in passing from~$1$ to~$2$. The decrease of the energy of a system subjected to any change is, then, the mechanical equivalent of the external effects resulting from that change; or, in other words, the \emph{increase of the energy} of a system which undergoes any change, is equal to the mechanical equivalent of the heat absorbed and the work expended in producing the change:\Add{---} \[ U_{2} - U_{1} = Q + W\Add{.} \Tag{(17)} \] $Q$~is the mechanical equivalent of the heat absorbed by the system, \eg\ by conduction, and $W$~is the amount of work expended on the system. $W$~is positive if the change takes place in the direction of the external forces. The sum $Q + W$ represents the mechanical equivalent of all the thermal and mechanical operations of the surrounding bodies on the system. We shall use $Q$~and $W$ always in this sense. The value of $Q + W$ is independent of the manner of the transition from~$1$ to~$2$, and evidently also of the selection of the normal state. When differences of energy of one and the same system are considered, it is, therefore, not even necessary to fix upon a normal state. In the expression for the energy of the system there remains then an arbitrary additive constant undetermined. \Section{64.} The difference $U_{2} - U_{1}$ may also be regarded as the energy of the system in state~$2$, referred to state~$1$ as the normal state. For, if the latter be thus selected, then $U_{1} = 0$, since it takes no energy to change the system from~$1$ to the normal state, and $U_{2} - U_{1} = U_{2}$. The normal \PageSep{44} state is, therefore, sometimes called the state of zero \index{Zero!energy}% energy. \index{Energy!zero}% \Section{65.} States $1$~and $2$ may be identical, in which case the system changing from~$1$ to~$2$ passes through a so-called \index{Perfect gases!system}% \index{System!perfect}% \emph{cycle of operations}. In this case, \index{Cycle of operations}% \[ U_{2} = U_{1}\quad\text{and}\quad Q + W = 0\Add{.} \Tag{(18)} \] The mechanical equivalent of the external effects is zero, or the external heat effect is equal in magnitude and opposite in sign to the external work. This proposition shows the impracticability of perpetual motion, which necessarily presupposes engines working in complete cycles. \Section{66.} If no external effects ($Q = 0$, $W = 0$) be produced by a change of state of the system, its energy remains constant (conservation of the energy). The quantities, on which the state of the system depends, may undergo considerable changes in this case, but they must obey the condition $U = \const$. A system which changes without being acted on by external agents is called a \emph{perfect system}. Strictly speaking, no perfect system can be found in nature, since there is constant interaction between all material bodies of the universe. It is, however, of importance to observe that by an adequate choice of the system which is to undergo the contemplated change, we have it in our power to make the external effect as small as we please, in comparison with the changes of energy of portions of the system itself. Any particular external effect may be eliminated by making the body which produces this effect, as well as the recipient, a part of the system under consideration. In the case of a gas which is being compressed by a weight sinking to a lower level, if the gas by itself be the system considered, the external effect on it is equal to the work done by the weight. The energy of the system accordingly increases. If, however, the weight and the earth be considered parts of the system, all external effects are eliminated, and the \PageSep{45} energy of this system remains constant. The expression \index{Energy!potential}% \index{Potential, energy}% for the energy now contains a new term representing the potential energy of the weight. The loss of the potential energy of the weight is exactly compensated by the gain of the internal energy of the gas. All other cases admit of similar treatment. \PageSep{46} \Chapter{II.}{Applications to Homogeneous Systems.} \Section{67.} \First{We} shall now apply the first law of thermodynamics \index{First law of thermodynamics}% as expressed in equation~\Eq{(17)}, \[ U_{2} - U_{1} = Q + W, \] to a homogeneous substance, whose state is determined, besides by its chemical nature and mass~$M$, by two variables, the temperature~$\theta$ and the volume~$V$, for instance. The term \emph{homogeneous} is used here in the sense of \emph{physically homogeneous}, and is applied to any system which appears of completely uniform structure throughout. The substance may be chemically homogeneous, \ie\ it may consist entirely of the same kind of molecules, or chemical transformations may take place at some stage of the process, as, for example, in the case of a vapour, which partially dissociates on being heated. The homogeneous state must, however, be a single valued function of the temperature and the volume. As long as the system is at rest, the total energy consists of the so-called \emph{internal} energy~$U$, which depends only on the internal state of the substance as determined by its density and temperature, and on its mass, to which it is evidently proportional. In other cases the total energy contains, besides the internal energy~$U$, another term, namely, the kinetic energy, which is known from the principles of mechanics. In order to determine the functional relation between $U$,~$\theta$, and~$V$, the state of the system must be changed, and the external effects of this change calculated. Equation~\Eq{(17)} then gives the corresponding change of energy. \PageSep{47} \Section{68.} If a gas, initially at rest and at uniform temperature, be allowed to suddenly expand by the opening of a \Chg{stopcock}{stop-cock}, which makes communication with a previously exhausted vessel, a number of intricate mechanical and thermal changes will at first take place. The portion of the gas flowing into the vacuum is thrown into violent motion, then heated by impact against the sides of the vessel and by compression of the particles crowding behind, while the portion remaining in the first vessel is cooled down by expansion, etc. Assuming the walls of the vessels to be absolutely rigid and non-conducting, and denoting by~$2$ any particular state after communication between the vessels has been established, then, according to equation~\Eq{(17)}, the total energy of the gas in state~$2$ is precisely equal to that \index{Energy!internal}% in state~$1$, for neither thermal nor mechanical forces have acted on the gas from without. The reaction of the walls does not perform any work. The energy in state~$2$ is, in general, composed of many parts, viz.\ the kinetic and internal energies of the gas particles, each one of which, if taken sufficiently small, may be considered as homogeneous and uniform in temperature and density. If we wait until complete rest and thermal equilibrium have been re-established, and denote this state by~$2$, then in~$2$, as in~$1$, the total energy consists only of the internal energy~$U$, and we have \index{Internal energy}% $U_{2} = U_{1}$. But the variables $\theta$~and~$V$, on which $U$~depends, have passed from $\theta_{1}$,~$V_{1}$ to $\theta_{2}$,~$V_{2}$, where $V_{2} > V_{1}$. By measuring the temperatures and the volumes, the relation between the temperature and the volume in processes where the internal energy remains constant may be established. \Section{69.} Joule performed such an experiment as described, \index{Joule's experiments}% and found that for perfect gases $\theta_{2} = \theta_{1}$. He put the two communicating vessels, one filled with air at high pressure, the other exhausted, into a common water-bath at the same temperature, and found that, after the air had expanded and equilibrium had been established, the change of temperature of the water-bath was inappreciable. It immediately follows that, if the walls of the vessels were \PageSep{48} non-conducting, the final temperature of the total mass of the gas would be equal to the initial temperature; for otherwise the change in temperature would have communicated itself to the water-bath in the above experiment. Hence, if the internal energy of a nearly perfect gas \index{Internal energy!of perfect gas}% remains unchanged after a considerable change of volume, then its temperature also remains almost constant. In other words, \emph{the internal energy of a perfect gas depends only on the temperature, and not on the volume}. \Section{70.} For a conclusive proof of this important deduction, \index{Joule and Thomson's absolute temperature!experiments}% \index{Porous plug experiments}% much more accurate measurements are required. In Joule's experiment described above, the heat capacity of the gas is so small compared with that of the vessel and the water-bath, that a considerable change of temperature in the gas would have been necessary to produce an appreciable change of temperature in the water. More reliable results are obtained by a modification of the above method devised by Sir~William Thomson (Lord Kelvin), and used by him, along with Joule, for accurate measurements. Here the outflow of the gas is artificially retarded, so that the gas passes immediately into its second state of equilibrium. The temperature~$\theta_{2}$ is then directly measured in the stream of outflowing gas. No limited quantity of gas rushes tumultuously into a vacuum, but a gas is slowly transferred in a steady flow from a place of high pressure,~$p_{1}$, to one of low pressure,~$p_{2}$ (the atmosphere), by forcing it through a boxwood tube stopped at one part of its length by a porous plug of cotton wool or filaments of silk. The results of the experiment show that when the flow has become steady there is, for air, a very small change of temperature, and, for hydrogen, a still smaller, hardly appreciable change. Hence the conclusion appears justified, that, for a perfect gas, the change of temperature vanishes entirely. This leads to an inference with regard to the internal energy of a perfect gas. When, after the steady state of the process has been established, a certain mass of the gas has been completely pushed through the plug, it has been \PageSep{49} operated upon by external agents during its change from the volume,~$V_{1}$, at high pressure, to the larger volume,~$V_{2}$, at atmospheric pressure. The mechanical equivalent of these operations, $Q + W$, is to be calculated from the external changes. The state of the porous plug remains the same throughout; hence the processes that take place in it may be neglected. No change of temperature occurs outside the tube, as the material of which it is made is practically non-conducting; hence $Q = 0$. The mechanical work done by a piston in pressing the gas through the plug at the constant pressure~$p_{1}$ is evidently~$p_{1}V_{1}$, and this for a perfect gas at constant temperature is, according to Boyle's law, equal to the work~$p_{2}V_{2}$, which is gained by the escaping gas pushing a second piston at pressure~$p_{2}$ through a volume~$V_{2}$. Hence the sum of the external work~$W$ is also zero, and therefore, according to equation~\Eq{(17)}, $U_{2} = U_{1}$. As the experimental results showed the temperature to be practically unchanged while the volume increased very considerably, the internal energy of a perfect gas can depend only on the temperature and not on the volume, \ie, \[ \left(\frac{\dd U}{\dd V}\right)_{\theta} = 0\Add{.} \Tag{(19)} \] For nearly perfect gases, as hydrogen, air, etc., the actual small change of temperature observed shows how far the internal energy depends on the volume. It must, however, be borne in mind that for such gases the external work, \[ W = p_{1}V_{1} - p_{2}V_{2}, \] does not vanish; hence the internal energy does not remain constant. For further discussion, see~\SecRef{158}. \Section{71.} Special theoretical importance must be attached to \index{Infinitely slow!process|(}% those thermodynamical processes which progress infinitely slowly, and which, therefore, consist of a succession of states of equilibrium. Strictly speaking, this expression is vague, since a process presupposes changes, and, therefore, disturbances of equilibrium. But where the time taken is \PageSep{50} immaterial, and the result of the process alone of consequence, these disturbances may be made as small as we please, certainly very small in comparison with the other quantities which characterize the state of the system under observation. Thus, a gas may be compressed very slowly to any fraction of its original volume, by making the external pressure, at each moment, just a trifle greater than the internal pressure of the gas. Wherever external pressure enters---as, for instance, in the calculation of the work of compression---a very small error will then be committed, if the pressure of the gas be substituted for the external pressure. On passing to the limit, even that error vanishes. In other words, the result obtained becomes rigorously exact for \emph{infinitely slow} compression. \index{Infinitely slow!compression}% This holds for compression at constant as well as at variable pressure. The latter may be given the required value at each moment by the addition or removal of small weights. This may be done either by hand (by pushing weights to one side), or by means of some automatic device which acts merely as a release, and therefore does not contribute towards the work done. \Section{72.} The conduction of heat to and from the system may be treated in the same way. When it is not a question of time, but only of the amount of heat received or given out by the system, it is sufficient, according as heat is to be added to or taken from the system, to connect it with a heat-reservoir of slightly higher or lower temperature than that of the system. This small difference serves, merely, to determine the direction of the flow of the heat, while its magnitude is negligible compared with the changes of the system, which result from the process. We, therefore, speak of the conduction of heat between bodies of equal temperature, just as we speak of the compression of a gas by an external pressure equal to that of the gas. This is merely anticipating the result of passing to the limit from a small finite difference to an infinitesimal difference of temperature between the two bodies. \PageSep{51} \index{External conditions of equilibrium!work in reversible process}% \index{Work!external, in reversible process}% This applies not only to strictly isothermal processes, but also to those of varying temperature. One heat-reservoir of constant temperature will not suffice for carrying out the latter processes. These will require either an auxiliary body, the temperature of which may be arbitrarily changed, \eg\ a gas that can be heated or cooled at pleasure by compression or expansion; or a set of constant-temperature reservoirs, each of different temperature. In the latter case, at each stage of the process we apply that particular heat-reservoir whose temperature lies nearest to that of the system at that moment. \Section{73.} The value of this method of viewing the process lies in the fact that we may imagine each \emph{infinitely slow} process to be carried out also in the opposite direction. If a process consist of a succession of states of equilibrium with the exception of very small changes, then evidently a suitable change, quite as small, is sufficient to reverse the process. This small change will vanish when we pass over to the limiting case of the infinitely slow process, for a \index{Infinitely slow!process|)}% definite result always contains a quite definite error, and if this error be smaller than any quantity, however small, it must be zero. \Section{74.} We pass now to the application of the first law to a process of the kind indicated, and, therefore, reversible in its various parts. Taking the volume~$V$ (abscissa) and the pressure~$p$ (ordinate) as the independent variables, we may graphically illustrate our process by plotting its successive states of equilibrium in the form of a curve in the plane of the co-ordinates. Each point in this plane corresponds to a certain state of our system, the chemical nature and mass of which are supposed to be given, and each curve corresponds to a series of continuous changes of state. Let the curve a from~$1$ to~$2$ represent a reversible process which takes the substance from a state~$1$ to a state~$2$ (\Fig{2}). Along~$\alpha$, according to equation~\Eq{(17)}, the increase of the energy is \[ U_{2} - U_{1} = W + Q, \] \PageSep{52} where $W$~is the mechanical work expended on the substance, and $Q$~the total heat absorbed by it. \Section{75.} The value of~$W$ can be readily determined. $W$~is \index{External conditions of equilibrium!work in reversible process}% made up of the elementary quantities of work done on the system during the infinitesimal changes corresponding to the elements of arc of the curve~$\alpha$. The external pressure is at any moment equal to that of the substance, since the process is supposed to be reversible. Consequently, by the \Figure{2} laws of hydrodynamics, the work done by the external forces in the infinitely small change is equal to the product of the pressure~$p$, and the decrease of the volume,~$-dV$, no matter what the geometrical form of the surface of the body may be. Hence the external work done during the whole process is \[ W = -\int_{1}^{2} p\, dV, \Tag{(20)} \] in which the integration extends from~$1$ to~$2$ along the curve~$\alpha$. If $p$~be positive, as in the case of gases, and $V_{2} > V_{1}$ as in \Fig{2}, $W$~is negative. \PageSep{53} In order to perform the integration, the curve~$\alpha$, \ie\ the relation between $p$ and~$V$, must be known. As long as only the points $1$~and $2$ are given, the integral has no definite value. In fact, it assumes an entirely different value along a different curve,~$\beta$, joining $1$~and~$2$. Therefore $p\, dV$~is not a perfect differential. Mathematically this depends on the fact that $p$~is in general not only a function of~$V$, but also of another variable, the temperature~$\theta$, which also changes along the path of integration. As long as $\alpha$~is not given, no statement can be made with regard to the relation between $\theta$~and~$V$, and the integration cannot be performed. The external work,~$W$, is evidently represented by the area (taken negative) of the plane figure bounded by the curve~$\alpha$, the ordinates at $1$~and~$2$, and the axis of abscissæ. This, too, shows that $W$~depends on the path of the curve~$\alpha$. Only for infinitesimal changes, \ie\ when $1$~and $2$ are infinitely near one another and $\alpha$~shrinks to a curve element, is $W$~determined by the initial and final points of the curve alone. \Section{76.} The second measurable quantity is~$Q$, the heat \index{Heat!absorbed}% \index{Heat and work, analogy between}% \index{Work!and heat, analogy between}% absorbed. It may be determined by calorimetric methods in calories, and then expressed in mechanical units by multiplying by the mechanical equivalent of heat. We shall now consider the theoretical determination of~$Q$. It is, like~$W$, the algebraical sum of the infinitely small quantities of heat added to the body during the elementary processes corresponding to the elements of the curve~$\alpha$. Such an increment of heat cannot, however, be immediately calculated, from the position of the curve element in the co-ordinate plane, in a manner similar to that of the increment of work. To establish an analogy between the two, one might, in imitation of the expression~$-p\, dV$, put the increment of heat $= C\, d\theta$, where $d\theta$~is the increment of temperature, and $C$~the heat capacity, which is usually a finite quantity. But $C$~has not, in general, a definite value. It does not depend, as the factor~$p$ in the expression for the increment of work, alone on the momentary state of the substance, \ie\ \PageSep{54} \index{External conditions of equilibrium!work in complete cycle}% on the position of the point of the curve considered, but also on the direction of the curve element. In isothermal changes $C$~is evidently $= \pm\infty$, because $d\theta = 0$, and the heat added or withdrawn is a finite quantity. In adiabatic changes $C = 0$, for here the temperature may change in any way, while no heat is added or withdrawn. For a given point, $C$~may, therefore, in contradistinction to~$p$, assume all values between~$+\infty$ and~$-\infty$. (\Cf~\SecRef{47}.) Hence the analogy is incomplete in one essential, and does not, in the general case, simplify the problem in hand. We shall also find that the breaking up of the heat absorbed into the two factors $\theta$~and~$d\Phi$ (\SecRef{120}), is permissible only in some very special cases. \Section{77.} Although the value of~$Q$ cannot, in general, be directly determined, equation~\Eq{(17)} enables us to draw some important inferences regarding it. Substituting the value of~$W$ from equation~\Eq{(20)} in equation~\Eq{(17)}, we obtain \[ Q = U_{2} - U_{1} + \int_{1}^{2} p\, dV, \Tag{(21)} \] which shows that the value of~$Q$ depends not only on the position of the points $1$~and~$2$, but also on the connecting path ($\alpha$~or~$\beta$). Carnot's theory of heat cannot be reconciled with this proposition, as we have shown at length in % [** Two section signs in the original] \SecRef{51}~and~\SecNum{52}. \Section{78.} The complete evaluation of~$Q$ is possible in the case where the substance returns to its initial state, having gone through a cycle of operations. This might be done by first bringing the system from~$1$ to~$2$ along~$\alpha$, then back from~$2$ to~$1$ along~$\beta$. Then, as in all cycles (\SecRef{65}), \[ Q = -W. \] The external work is \[ W = -\int_{1}^{1} p\, dV, \] the integral to be taken along the closed curve $1\alpha 2\beta 1$. $W$~evidently represents the area bounded by the curve, \PageSep{55} and is positive if the process follows the direction of the arrow in \Fig{2}. \Section{79.} We shall now consider the special case where the \index{Clausius' equation!notation}% curve~$\alpha$, which characterizes the change of state, shrinks into an element, so that the points $1$~and~$2$ lie infinitely near one another. $W$~here becomes the increment of work,~$-p\, dV$, and the change of the internal energy is~$dU$. Hence, according to~\Eq{(21)}, the heat absorbed assumes the value:\footnote {It is usual to follow the example of Clausius, and denote this quantity by~$dQ$, to indicate that it is infinitely small. This notation, however, has frequently given rise to misunderstanding, for $dQ$~has been repeatedly regarded as the differential of a known finite quantity~$Q$. We therefore adhere to the notation given above. Other authors use~$d'Q$, in order to obviate the aforesaid misunderstanding.} \[ Q = dU + p\, dV. \] Per unit mass, this equation becomes \[ q = du + p\, dv, \Tag{(22)} \] where the small letters denote the corresponding capitals divided by~$M$. In subsequent calculations it will often be advisable to use $\theta$~as an independent variable, either in conjunction with~$p$, or~$v$. We shall, in each case, select as independent variables those which are most conducive to a simplification of the problem in hand. The meaning of the differentiation will be indicated whenever a misunderstanding is possible. We shall now apply our last equation~\Eq{(22)} to the most important reversible processes. \Section{80.} It has been repeatedly mentioned that the specific heat of a substance may be defined in very different ways according to the manner in which the heating is carried out. But, according to \SecRef{46} and equation~\Eq{(22)}, we have, for any heating process, \[ c = \frac{q}{d\theta} = \frac{du}{d\theta} + p\, \frac{dv}{d\theta}. \Tag{(23)} \] \PageSep{56} In order to give a definite meaning to the differential coefficients, some arbitrary condition is required, which will prescribe the direction of the change. A single condition is sufficient, since the state of the substance depends on two variables only. \Section{81.} \Topic{Heating at Constant Volume.}---Here $dv = 0$, \index{Heating!at constant pressure}% \index{Heating!at constant volume}% $c = c_{v}$, the specific heat at constant volume. Hence, according \index{Specific heat!at constant pressure}% \index{Specific heat!at constant volume}% to equation~\Eq{(23)}, \begin{align*} c_{v} &= \left(\frac{\dd u}{\dd \theta}\right)_{v}\Add{,} \Tag{(24)} \\ \intertext{or} c_{v} &= \left(\frac{\dd u}{\dd p}\right)_{v} \left(\frac{\dd p}{\dd \theta}\right)_{v}\Add{.} \Tag{(25)} \end{align*} \Section{82.} \Topic{Heating under Constant Pressure.}---Here $dp = 0$, $c = c_{p}$, the specific heat at constant pressure. According to equation~\Eq{(23)}, \begin{align*} c_{p} &= \left(\frac{\dd u}{\dd \theta}\right)_{p} + p\left(\frac{\dd v}{\dd \theta}\right)_{p}\Add{,} \Tag{(26)} \\ \intertext{or} c_{p} &= \left[\left(\frac{\dd u}{\dd v}\right)_{p} + p\right]\left(\frac{\dd v}{\dd \theta}\right)_{p}\Add{.} \Tag{(27)} \end{align*} By the substitution of \[ \left(\frac{\dd u}{\dd \theta}\right)_{p} = \left(\frac{\dd u}{\dd \theta}\right)_{v} + \left(\frac{\dd u}{\dd v}\right)_{\theta} \left(\frac{\dd v}{\dd \theta}\right)_{p} \] in~\Eq{(26)}, $c_{p}$~may be written in the form \[ c_{p} = \left(\frac{\dd u}{\dd \theta}\right)_{v} + \left[\left(\frac{\dd u}{\dd v}\right)_{\theta} + p\right]\left(\frac{\dd v}{\dd \theta}\right)_{p}, \] or, by~\Eq{(24)}, \[ c_{p} = c_{v} + \left[\left(\frac{\dd u}{\dd v}\right)_{\theta} + p\right]\left(\frac{\dd v}{\dd \theta}\right)_{p}. \Tag{(28)} \] \Section{83.} By comparing \Eq{(25)} and~\Eq{(27)} and eliminating~$u$, we are led to a direct experimental test of the theory. By~\Eq{(25)}, \[ \left(\frac{\dd u}{\dd p}\right)_{v} = c_{v} \left(\frac{\dd \theta}{\dd p}\right)_{v}, \] and by~\Eq{(27)}, \[ \left(\frac{\dd u}{\dd v}\right)_{p} = c_{p} \left(\frac{\dd \theta}{\dd v}\right)_{p} - p; \] \PageSep{57} whence, differentiating the former equation with respect to~$v$, keeping $p$~constant, and the latter with respect to~$p$, keeping $v$~constant, and equating, we have \[ \frac{\dd}{\dd v}\left(c_{v}\, \frac{\dd \theta}{\dd p}\right) = \frac{\dd}{\dd p}\left(c_{p}\, \frac{\dd \theta}{\dd v} - p\right)\Add{,} \] or \[ (c_{p} - c_{v})\, \frac{\dd^{2} \theta}{\dd p\, \dd v} + \frac{\dd c_{p}}{\dd p} · \frac{\dd \theta}{\dd v} - \frac{\dd c_{v}}{\dd v} · \frac{\dd \theta}{\dd p} = 1\Add{.} \Tag{(29)} \] This equation contains only quantities which may be experimentally determined, and therefore furnishes a means for testing the first law of thermodynamics by observations on any homogeneous substance. \Section{84.} \Topic{Perfect Gases.}---The above equations undergo \index{Boyle's law}% \index{Gases, perfect}% \index{Joule and Thomson's absolute temperature!experiments}% \index{Laws:!Avogadro's}% \index{Laws:!Boyle's}% \index{Laws:!Gay-Lussac's}% \index{Perfect gases}% considerable simplifications for perfect gases. We have, from~\Eq{(14)}, \[ p = \frac{R}{m} · \frac{\theta}{v}, \Tag{(30)} \] where $R = 826 × 10^{5}$ and $m$~is the (real or apparent) molecular weight. Hence \[ \theta = \frac{m}{R}\, pv, \] and equation~\Eq{(29)} becomes \[ c_{p} - c_{v} + p\, \frac{\dd c_{p}}{\dd p} - v\, \frac{\dd c_{v}}{\dd v} = \frac{R}{m}. \] Assuming that only the laws of Boyle, Gay-Lussac, and \index{Gay-Lussac}% Avogadro hold, no further conclusions can be drawn from the first law of thermodynamics with regard to perfect gases. \Section{85.} We shall now make use of the additional property of perfect gases, established by Thomson and Joule (\SecRef{70}), that the internal energy of a perfect gas depends only on \index{Energy!internal!of perfect gas}% the temperature, and not on the volume, and that hence per unit mass, according to~\Eq{(19)}, \[ \left(\frac{\dd u}{\dd v}\right)_{\theta} = 0. \Tag{(31)} \] \PageSep{58} The general equation, \[ du = \left(\frac{\dd u}{\dd \theta}\right)_{v} d\theta + \left(\frac{\dd u}{\dd v}\right)_{\theta} dv, \] then becomes, for perfect gases, \[ du = \left(\frac{\dd u}{\dd \theta}\right)_{v} d\theta, \] and, according to~\Eq{(24)}, \[ du = c_{v} · d\theta. \Tag{(32)} \] It follows from~\Eq{(28)} that \[ c_{p} = c_{v} + p \left(\frac{\dd v}{\dd \theta}\right)_{p}, \] or, considering the relation~\Eq{(30)}, \[ c_{p} = c_{v} + \frac{R}{m}; \] \ie\ there is a constant difference between the specific heat \index{Heat!molecular!of perfect gases}% \index{Molecular heat!of perfect gases}% at constant pressure and the specific heat at constant volume. Referring the heat capacity to the molecular weight~$m$, instead of to unit mass, we have \[ mc_{p} - mc_{v} = R. \Tag{(33)} \] The difference is, therefore, independent even of the nature of the gas. \Section{86.} Only the specific heat at constant pressure,~$c_{p}$, is capable of direct experimental determination, because a quantity of gas enclosed in a vessel of constant volume has far too small a heat capacity to produce sufficient thermal effects on the surrounding bodies. Since $c_{v}$, according to~\Eq{(24)}, like~$u$, depends on the temperature only, and not on the volume, the same follows for~$c_{p}$, according to~\Eq{(33)}. This conclusion was first confirmed by Regnault's experiments. \index{Regnault}% He found $c_{p}$~constant within a considerable range of temperature. By~\Eq{(33)}, $c_{v}$~is constant within the same range. \PageSep{59} If the molecular heats be expressed in calories, $R$~must be divided by Joule's equivalent~$J$. The difference between the molecular heats at constant pressure and at constant volume is then \[ mc_{p} - mc_{v} = \frac{R}{J} = \frac{826 · 10^{5}}{419 · 10^{5}} = 1.971\Add{.} \Tag{(34)} \] \Section{87.} The following table contains the specific heats \index{Ratio of specific heats}% \index{Specific heats!ratio of}% \index{Specific heat!at constant pressure}% \index{Specific heat!at constant volume}% and molecular heats of several gases at constant pressure, measured by direct experiment; also the molecular heats at constant volume found by subtracting~$1.97$, and also the ratio $\dfrac{c_{p}}{c_{v}} = \gamma$:--- \begin{center} \TableFont \begin{tabular}{l*{5}{|@{\;}c@{\;}}} \hline \Strut[16pt] & $c_{p}$ & $m$ & $mc_{p}$ & $mc_{v}$ & $\dfrac{c_{p}}{c_{v}} = \gamma$ \\ %[** TN: Re-breaking column headings] & \ColumnHeading{Specific heat}{Specific heat \\ at const. \\ pressure.} & \ColumnHeading{Molecular}{Molecular \\ weight.} & \ColumnHeading{Molecular}{Molecular \\ heat at \\ const. \\ pressure.} & \ColumnHeading{Molecular}{Molecular \\ heat at \\ const. \\ volume.} & \\ \cline{2-6} \Strut Hydrogen & $3.410\Z$ & $\Z2.0$ & $6.82$ & $4.85$ & $1.41$ \\ Oxygen & $0.2175$ & $31.9$ & $6.94$ & $4.97$ & $1.40$ \\ Nitrogen & $0.2438$ & $28.0$ & $6.83$ & $4.86$ & $1.41$ \\ Air & $0.2375$ & $28.8$ & $6.84$ & $4.87$ & $1.41$ \\ \hline \end{tabular} \end{center} The specific heat generally increases slowly on considerable increase of temperature. Within the range of temperature in which the specific heat is constant, equation~\Eq{(32)} can be integrated, giving \[ u = c_{v} \theta + \const \Tag{(35)} \] The constant of integration depends on the selection of the zero point of energy. For perfect gases, we consider $c_{p}$~and $c_{v}$ as constants throughout, hence the last equation holds good in general. \Section{88.} \Topic{Adiabatic Process.}---The characteristic feature \index{Adiabatic process}% of the adiabatic process is that $q = 0$, and, according to equation~\Eq{(22)}, \[ 0 = du + p\, dv. \] \PageSep{60} Assuming, again, a perfect gas, and substituting the values of~$du$ from~\Eq{(32)} and of~$p$ from~\Eq{(30)}, we have \[ 0 = c_{v}\, d\theta + \frac{R}{m} · \frac{\theta}{v}\, dv, \Tag{(36)} \] or, on integrating, \[ c_{v} \log\theta + \frac{R}{m} \log v = \const \] Replacing $\dfrac{R}{m}$ according to~\Eq{(33)} by $c_{p} - c_{v}$, and dividing by~$c_{v}$, we get \[ \log \theta + (\gamma - 1) \log v = \const \Tag{(37)} \] (\ie\ during adiabatic expansion the temperature decreases)\Add{.} Remembering that according to the characteristic equation\Eq{(30)} \[ \log p + \log v - \log \theta = \const, \] we have, on eliminating~$v$, \[ -\gamma \log \theta + (\gamma - 1) \log p = \const \] (\ie\ during adiabatic compression the temperature rises); or, on eliminating~$\theta$, \[ \log p + \gamma \log v = \const \] The values of the constants of integration are given by the initial state of the process. If we compare our last equation in the form \[ pv^{\gamma} = \const \Tag{(38)} \] with Boyle's law $pv = \const$, it is seen that during adiabatic compression the volume decreases more slowly for an increase of pressure than during isothermal compression, because during adiabatic compression the temperature rises. The adiabatic curves in the $\Chg{pv}{(p, v)}$-plane (\SecRef{22}) are, therefore, steeper than the hyperbolic isotherms. \Section{89.} Adiabatic processes may be used in various ways \PageSep{61} for the determination of~$\gamma$, the ratio of the specific heats. The agreement of the results with the value calculated from the mechanical equivalent of heat forms an important confirmation of the theory. Thus, the measurement of the velocity of sound in a gas \index{Sound, velocity of}% may be used for determining the value of~$\gamma$. It is proved in hydrodynamics that the velocity of sound in a fluid is $\sqrt{\dfrac{dp}{d\rho}}$, where $\rho = \dfrac{1}{v}$, the density of the fluid. Since gases are bad conductors of heat, the compressions and expansions which accompany sound-vibrations must be considered as adiabatic, and not isothermal, processes. The relation between the pressure and the density is, therefore, in the case of perfect gases, not that expressed by Boyle's law $\dfrac{p}{\rho} = pv = \const$, but that given by equation~\Eq{(38)}, viz.--- \[ \frac{p}{\rho^{\gamma}} = \const \] Hence, by differentiation \[ \frac{dp}{d\rho} = \frac{\gamma p}{\rho} = \gamma pv, \] or, according to~\Eq{(30)}, \begin{align*} \frac{dp}{d\rho} &= \gamma \frac{R}{m} \theta, \\ \gamma &= \frac{m}{R\theta} · \frac{dp}{d\rho}. \end{align*} In air at~$0°$, the velocity of sound is $\sqrt{\dfrac{dp}{d\rho}} = 33280\, \dfrac{\Unit{cm.}}{\Unit{sec.}}$; hence, according to our last equation, taking the values of~$m$ from \SecRef{41}, and of~$R$ from \SecRef{84}, and $\theta = 273$, \[ \gamma = \frac{28.8}{826 · 10^{5}} · \frac{33280^{2}}{273} = 1.41. \] This agrees with the value calculated in~\SecRef{87}. Conversely, the value of~$\gamma$, calculated from the velocity \PageSep{62} \index{Meyer, Robert}% of sound, may be used in the calculation of~$c_{v}$ in calories, for the determination of the mechanical equivalent of heat from~\Eq{(33)}. This method of evaluating the mechanical equivalent of heat was first proposed by Robert Meyer in 1842. It is true that the assumption expressed in equation~\Eq{(31)}, that the internal energy of air depends only on the temperature, is essential to this method. In other words, this means that the difference of the specific heats at constant pressure and constant volume depends only on the external work. The direct proof of this fact, however, must be considered as first given by the experiments of Thomson and Joule, described in~\SecRef{70}. \Section{90.} We shall now consider a more complex process, a reversible cycle of a special kind, which has played an important part in the development of thermodynamics, known as Carnot's cycle, and shall apply the first law to it \index{Carnot's!cycle}% in detail. Let a substance of unit mass, starting from an initial state characterized by the values $\theta_{1}$,~$v_{1}$, first be compressed \emph{adiabatically} until its temperature rises to~$\theta_{2}$ ($\theta_{2} > \theta_{1}$) and its volume reduced to~$v_{2}$ ($v_{2} < v_{1}$) (\Fig{3}). Second, suppose it be now allowed to expand \emph{isothermally} to volume~$v_{2}'$ ($v_{2}' > v_{2}$), in constant connection with a heat-reservoir of constant temperature,~$\theta_{2}$, which gives out the heat of expansion~$Q_{2}$. Third, let it be further expanded \emph{adiabatically} until its temperature falls to~$\theta_{1}$, and the volume thereby increased to~$v_{1}'$. Fourth, let it be compressed \emph{isothermally} to the original volume~$v_{1}$, while a heat-reservoir maintains the temperature at~$\theta_{1}$, by absorbing the heat of compression. All these operations are to be carried out in the reversible manner described in~\SecRef{71}. The sum of the heat absorbed by the system, and the work done on the system during this cycle is, by the first law, \[ Q + W = 0. \Tag{(39)} \] The heat~$Q$, that has been absorbed by the substance, is \[ Q = Q_{1} + Q_{2} \Tag{(40)} \] \PageSep{63} ($Q_{1}$~is here negative). The external work~$W$ may be calculated from the adiabatic and the isothermal compressibility of the substance. According to~\Eq{(20)}, \[ W = -\int_{v_{1}, \theta_{1}}^{v_{2}, \theta_{2}} p\, dv -\int_{v_{2}, \theta_{2}}^{v_{2}', \theta_{2}} p\, dv -\int_{v_{2}', \theta_{2}}^{v_{1}', \theta_{1}} p\, dv -\int_{v_{1}', \theta_{1}}^{v_{1}, \theta_{1}} p\, dv. \] These integrals are to be taken along the curves $1$,~$2$, $3$,~$4$ respectively; $1$~and $3$ being adiabatic, $2$~and $4$ isothermal. \Figure{3} Assuming the substance to be a perfect gas, the above integrals can readily be found. If we bear in mind the relations \Eq{(30)} and~\Eq{(36)}, we have \[ W = \int_{\theta_{1}}^{\theta_{2}} c_{v}\, d\theta - \frac{R}{m} \int_{v_{2}}^{v_{2}'} \frac{\theta_{2}}{v}\, dv + \int_{\theta_{2}}^{\theta_{1}} c_{v}\, d\theta - \frac{R}{m} \int_{v_{1}'}^{v_{1}} \frac{\theta_{1}}{v}\, dv\Add{.} \Tag{(41)} \] The work of the adiabatic compression in the first part of the process is equal in value and opposite in sign to that of the adiabatic expansion in the third part of the process. \PageSep{64} There remains, therefore, the sum of the work in the isothermal portions\Add{:}--- \[ W = -\frac{R}{m} \left(\theta_{2} \log \frac{v_{2}'}{v_{2}} + \theta_{1} \log \frac{v_{1}'}{v_{1}}\right). \] Now, the state $(v_{2}, \theta_{2})$ was developed from $(v_{1}, \theta_{1})$ by an adiabatic process; therefore, by~\Eq{(37)}, \[ \log \theta_{2} + (\gamma - 1) \log v_{2} = \log \theta_{1} + (\gamma - 1) \log v_{1}. \] Similarly, for the adiabatic process, which leads from $(v_{2}', \theta_{2})$ to~$(v_{1}', \theta_{1})$, \[ \log \theta_{2} + (\gamma - 1) \log v_{2}' = \log \theta_{1} + (\gamma - 1) \log v_{1}'. \] From these equations, it follows that \[ \frac{v_{2}'}{v_{2}} = \frac{v_{1}'}{v_{1}}, \] and \[ \therefore W = -\frac{R}{m} (\theta_{2} - \theta_{1}) \log \frac{v_{1}'}{v_{1}}. \] Since, in the case considered, $\theta_{2} > \theta_{1}$, and $\dfrac{v_{1}'}{v_{1}} = \dfrac{v_{2}'}{v_{2}} > 1$, the total external work~$W$ is negative, \ie\ mechanical work has been gained by the process. But, from \Eq{(39)}~and~\Eq{(40)}, \[ Q = Q_{1} + Q_{2}= -W; \Tag{(42)} \] therefore $Q$~is positive, \ie\ the heat-reservoir at temperature~$\theta_{2}$ has lost more heat than the heat-reservoir at temperature~$\theta_{1}$ has gained. The value of~$W$, substituted in the last equation, gives \[ Q = Q_{1} + Q_{2} = \frac{R}{m} (\theta_{2} - \theta_{1}) \log \frac{v_{1}'}{v_{1}}. \Tag{(43)} \] The correctness of this equation is evident from the direct calculation of the values of $Q_{1}$ and~$Q_{2}$. The gas expands isothermally while the heat-reservoir at temperature~$\theta_{2}$ is in action. The internal energy of the gas therefore remains constant, and the heat absorbed is equal in magnitude and \PageSep{65} opposite in sign to the external work. Hence, by equating $Q_{2}$ to the second integral in~\Eq{(41)}, \[ Q_{2} = \frac{R}{m} \theta_{2} \log \frac{v_{2}'}{v_{2}} = \frac{R}{m} \theta_{2} \log \frac{v_{1}'}{v_{1}}, \] and, similarly, by equating $Q_{1}$ to the fourth integral in~\Eq{(41)}, \[ Q_{1} = \frac{R}{m} \theta_{1} \log \frac{v_{1}}{v_{1}'} = -\frac{R}{m} \theta_{1} \log \frac{v_{1}'}{v_{1}}, \] which agrees with equation~\Eq{(43)}. There exists, then, between the quantities $Q_{1}$,~$Q_{2}$,~$W$, besides the relation given in~\Eq{(42)}, this new relation--- \[ Q_{1} : Q_{2}: W = (-\theta_{1}): \theta_{2} : (\theta_{1} - \theta_{2})\Add{.} \Tag{(44)} \] \Section{91.} In order, now, to survey all the effects of the above Carnot cycle, we shall compare the initial and final states of all the bodies concerned. The gas operated upon has not been changed in any way by the process, and may be left out of account. It has done service only as a transmitting agent, in order to bring about changes in the surroundings. The two reservoirs, however, have undergone a change, and, besides, a positive amount of external work, $W' = -W$, has been gained; \ie\ at the close of the process certain weights, which were in action during the compression and the expansion, are found to be at a higher level than at the beginning, or a spring, serving similar purposes, is at a greater tension, etc. On the other hand, the heat-reservoir at~$\theta_{2}$ has given out heat to the amount~$Q_{2}$, and the cooler reservoir at~$\theta_{1}$ has received the smaller amount $Q_{1}' = -Q_{1}$. The heat that has vanished is equivalent to the work gained. This result may be briefly expressed as follows: The quantity of heat~$Q_{2}$, at temperature~$\theta_{2}$, has passed in part~($Q_{1}'$) to a lower temperature~($\theta_{1}$), and has in part ($Q_{2} - Q_{1}' = Q_{1} + Q_{2}$) been transformed into mechanical work. Carnot's cycle, performed with a perfect gas, thus affords a means of drawing heat from a body and of gaining work in its stead, without introducing any changes in nature except the \PageSep{66} transference of a certain quantity of heat from a body of higher temperature to one of lower temperature. But, since the process described is reversible in all its parts, it may be put into effect in such a way that all the quantities, $Q_{1}$,~$Q_{2}$,~$W$, change sign, $Q_{1}$~and $W$ becoming positive, $Q_{2} = -Q_{2}'$ negative. In this case the hotter reservoir at~$\theta_{2}$ receives heat to the amount~$Q_{2}'$, partly from the colder reservoir (at~$\theta_{1}$), and partly from the mechanical work expended~($W$). By reversing Carnot's cycle, we have, then, a means of transferring heat from a colder to a hotter body without introducing any other changes in nature than the transformation of a certain amount of mechanical work into heat. We shall see, later, that, for the success of Carnot's reversible cycle, the nature of the transmitting agent or working substance is immaterial, and that perfect gases are, in this respect, neither superior nor inferior to other substances (\cf~\SecRef{137}). \PageSep{67} \Chapter{III.}{Applications to Non-Homogeneous Systems.} \Section{92.} \First{The} propositions discussed in the preceding chapter are, in a large part, also applicable to substances which are not perfectly homogeneous in structure. We shall, therefore, in this chapter consider mainly such phenomena as characterize the inhomogeneity of a system. Let us consider a system composed of a number of homogeneous bodies in juxtaposition, separated by given bounding surfaces. Such a system may, or may not, be chemically homogeneous. A liquid in contact with its vapour is an example of the first case, if the molecules of the latter be identical with those of the former. The beginning of a chemical reaction, inasmuch as a substance is in contact with another of different chemical constitution, is an example of the second. Whether a system is physically homogeneous or not, can, in most cases, be ascertained beyond doubt, by finding surfaces of contact within the system, or, by other means---in the case of emulsions, for example, by determining the vapour pressure or the freezing point. The question as to the chemical homogeneity, \ie\ the presence of one kind of molecule only, is much more difficult, and has hitherto been answered only in special cases. For this reason we classify substances according to their physical and not according to their chemical homogeneity. \Section{93.} One characteristic of processes in non-homogeneous systems consists in their being generally accompanied by considerable changes of temperature, \eg\ in evaporation or \PageSep{68} in oxidation. To maintain the initial temperature and pressure consequently requires considerable exchange of heat with the surroundings and corresponding external work. The latter, however, is generally small compared with the external heat, and may be neglected in most chemical processes. In thermochemistry, therefore, the external effects, \[ Q + W = U_{2} - U_{1}, \Tag{(45)} \] are generally measured in calories (the heat equivalent of the external effects). The external work,~$W$, is small compared with~$Q$. Furthermore, most chemical processes are accompanied by a rise in temperature, or, if the initial temperature be re-established, by an external yield of heat (exothermal processes). Therefore, in thermochemistry, the heat given out to the surroundings in order to restore the initial temperature is denoted as the ``positive heat effect'' \index{Heat effect!in thermochemistry}% of the process. In our equations we shall therefore use~$Q$ (the heat absorbed) with the negative sign, in processes with positive heat effect (\eg\ combustion); with the positive sign, in those with negative heat effect (\eg\ evaporation, fusion, dissociation). \Section{94.} To make equation~\Eq{(45)} suitable for thermochemistry \index{Thermochemical symbols}% \index{Thomsen, J.}% it is expedient to denote the internal energy~$U$ of a system in a given state, by a symbol denoting its chemical constitution. J.~Thomsen introduced a symbol of this kind. He denoted by the formulæ for the atomic or molecular weight of the substances enclosed in brackets, the internal energy of a corresponding weight referred to an arbitrary zero of energy. Thus \ce{[Pb]}, \ce{[S]}, \ce{[PbS]} denote the energies of an atom of lead, an atom of sulphur, and a molecule of lead sulphide respectively. In order to express the fact that \index{Lead sulphide}% the formation of a molecule of lead sulphide from its atoms is accompanied by a heat effect of $18,400~\Unit{cal.}$, the external work of the process being negligible, we put \begin{gather*} U_{1} = \ce{[Pb] + [S]};\quad U_{2} = \ce{[PbS]}; \\ W = 0;\quad Q = -18,400~\Unit{cal.}, \end{gather*} \PageSep{69} and equation~\Eq{(45)} becomes \[ -18,400~\Unit{cal.} = \ce{[PbS] - [Pb] - [S]}, \] or, as usually written, \[ \ce{[Pb] + [S] - [PbS]} = 18,400~\Unit{cal.} \] This means that the internal energy of lead and sulphur, when separate, is $18,400$ calories greater than that of their combination at the same temperature. That the internal energies compared actually refer to the same material system, can be checked by the use of the molecular formulæ. The equation could be simplified by selecting the uncombined state of the elements \ce{Pb} and \ce{S} as the zero of energy. Then (\SecRef{64}), $\ce{[Pb] + [S]} = 0$, and \[ \ce{[PbS]} = -18,400~\Unit{cal.} \] \Section{95.} To define accurately the state of a substance, and \index{Aggregation, states of}% \index{States of aggregation}% thereby its energy, besides its chemical nature and mass, its temperature and pressure must be given. If no special statement is made, as in the above example, mean laboratory temperature, \ie\ about $18°$~C., is generally assumed, and the pressure is supposed to be atmospheric pressure. The pressure has, however, very little influence on the internal energy; in fact, none at all in the case of perfect gases [equation~\Eq{(35)}]. The state of aggregation should also be indicated. This may be done, where necessary, by using brackets for the solids, parentheses for liquids, and braces for gases. Thus \ce{[H2O]}, \ce{(H2O)}, \ce{\{H2O\}} denote the energies of a molecule of ice, water, and water vapour respectively. Hence, for the fusion of ice at~$0°$~C., \[ \ce{(H2O) - [H2O]} = 80 × 18 = 1440~\Unit{cal.} \] It is often desirable, as in the case of solid carbon, sulphur, arsenic, or isomeric compounds, to denote by some means the special modification of the substance. These symbols may be treated like algebraic quantities, whereby considerations, which would otherwise present \PageSep{70} considerable complications, may be materially shortened. Examples of this are given below. \Section{96.} To denote the energy of a solution or mixture of \index{Energy!of a solution}% several compounds, we may write the formulæ for the molecular weights with the requisite number of molecules. Thus, \[ \ce{(H2SO4) + 5(H2O) - (H2SO4 . 5H2O)} = 13,100~\Unit{cal.} \] means that the solution of $1$~molecule of sulphuric acid in $5$~molecules of water gives out $13,100$ calories of heat. Similarly, the equation \[ \ce{(H2SO4) + 10(H2O) - (H2SO4 . 10H2O)} = 15,100~\Unit{cal.} \] gives the heat effect on dissolving the same in ten molecules \index{Heat effect!of dilution of \ce{H2SO4}}% of water. By subtracting the first equation from the second, we get \[ \ce{(H2SO4 . 5H2O) + 5(H2O) - (H2SO4 . 10H2O)} = 2000~\Unit{cal.}, \] \ie\ on diluting a solution of $1$~molecule of sulphuric acid dissolved in $5$~molecules of water, by the addition of another $5$~molecules of water, $2000$ calories are given out. \Section{97.} As a matter of experience, in very dilute solutions further dilution no longer yields any appreciable amount of \index{Dilution!infinite}% heat. Thus, in indicating the internal energy of a dilute solution it is often unnecessary to give the number of molecules of the solvent. We write briefly \[ \ce{(H2SO4) + ($\aq$) - (H2SO4 $\aq$)} = 17,900~\Unit{cal.} \] to express the heat effect of infinite dilution of a molecule \index{Infinite dilution}% of sulphuric acid. Here ($\aq$)\ denotes any amount of water sufficient for the practical production of an infinitely dilute solution. \Section{98.} Volumetric changes being very slight in chemical processes which involve only solids and liquids, the heat equivalent of the external work~$W$ (\SecRef{93}) is a negligible \PageSep{71} quantity compared with the heat effect. The latter alone, \index{Heat effect!at constant pressure}% then, represents the change of energy of the system\Add{:}--- \[ U_{2} - U_{1} = Q. \] It, therefore, depends on the initial and final states only, and not on the intermediate steps of the process. These considerations do not apply, in general, when gaseous substances enter into the reaction. It is only in the combustions in the ``calorimetric bomb,'' extensively used by Berthelot \index{Berthelot|(}% \index{Calorimetric bomb}% and Stohmann in their investigations, that the volume \index{Stohmann}% remains constant and the external work is zero. In these reactions the heat effect observed represents the total change of energy. In other cases, however, the amount of external work~$W$ may assume a considerable value, and it is materially influenced by the process itself. Thus, a gas may be allowed to expand, at the same time performing work, which may have any value within certain limits, from zero upwards. But since its change of energy $U_{2} - U_{1}$ depends on the initial and final states only, a greater amount of work done against the external forces necessitates a smaller heat effect for the process, and \textit{vice versâ}. To find the latter, not only the change of the internal energy, but also the amount of the external work must be known. This renders necessary an account of the external conditions under which the process takes place. \Section{99.} Of all the external conditions that may accompany a chemical process, constant (atmospheric) pressure is the one which is of the most practical importance: $p = p_{0}$. The external work is then, according to equation~\Eq{(20)}, \[ W = -\int_{1}^{2} p_{0}\, dV = p_{0} (V_{1} - V_{2}); \Tag{(46)} \] that is, equal to the product of the pressure and the decrease of volume. This, according to~\Eq{(45)}, gives \[ U_{2} - U_{1} = Q + p_{0} (V_{1} - V_{2}). \Tag{(47)} \] Now, the total decrease of volume, $V_{1} - V_{2}$, may generally \PageSep{72} be put equal to the decrease of volume of the gaseous portions of the system, neglecting that of the solids and liquids. Since, by~\Eq{(16)}, \[ V_{1} - V_{2} = R \frac{\theta}{p_{0}} (n_{1} - n_{2}), \] where $n_{1}$,~$n_{2}$ are the number of gas molecules present before and after the reaction, the heat equivalent of the external work at constant pressure is, by \Eq{(46)}~and~\Eq{(34)}, \[ \frac{W}{J} = \frac{p_{0} (V_{1} - V_{2})}{J} = \frac{R}{J} \theta (n_{1} - n_{2}) = 1.97 \theta (n_{1} - n_{2})~\Unit{cal.} \] The heat effect of a process at constant pressure is therefore \[ -Q = U_{1} - U_{2} + 1.97 \theta (n_{1} - n_{2})~\Unit{cal.} \Tag{(48)} \] If, for instance, one gram molecule of hydrogen and half a gram molecule of oxygen, both at $18°$~C., combine at constant pressure to form water at $18°$~C., we put \[ U_{1} = \ce{\{H2\} + $\tfrac{1}{2}$ \{O2\}};\quad U_{2} = \ce{(H2O)};\quad n_{1} = \tfrac{3}{2};\ n_{2} = 0;\ \theta = 291. \] The heat of combustion is, therefore, by~\Eq{(48)}, \[ -Q = \ce{\{H2\} + $\tfrac{1}{2}$ \{O2\}} - \ce{(H2O)} + 860~\Unit{cal.}, \] \ie\ $860~\Unit{cal.}$ more than would correspond to the decrease of the internal energy, or to the combustion without the simultaneous performance of external work. \Section{100.} If we write equation~\Eq{(47)} in the form \[ (U + p_{0} V)_{2} - (U + p_{0} V)_{1} = Q, \Tag{(49)} \] it will be seen that, in processes under constant pressure~$p_{0}$, the heat effect depends only on the initial and final states, just as in the case when there is no external work. The heat effect, however, is not equal to the difference of the internal energies~$U$, but to the difference of the values of the quantity $(U + p_{0} V)$ at the beginning and end of the process. This quantity is Gibbs's ``\emph{heat function at constant \PageSep{73} \index{Gibbs}% \index{Heat function at constant pressure}% pressure}.'' If, then, only processes at constant pressure be considered, it will be expedient to regard the symbols \ce{\{H2\}}, \ce{\{H2O\}}, etc., as representing the above function $(U + p_{0} V)$, instead of simply the energy~$U$. Thus the difference in the two values of the function will, in all cases, directly represent the heat effect. This notation is therefore adopted \index{Heat!of neutralization}% \index{Neutralization, heat of}% in the following. \Section{101.} To determine the heat effect of a chemical reaction \index{Thomsen, J.}% at constant pressure, the initial and final values of the heat function, $U + p_{0} V$, of the system suffice. The general solution of this problem, therefore, amounts to finding the heat functions of all imaginable material systems in all possible states. Frequently, different ways of transition from one state of a system to another may be devised, which may serve either as a test of the theory, or as a check upon the accuracy of the observations. Thus J.~Thomsen found the heat of neutralization of a solution of sodium \index{Sodium!carbonate}% \index{Sodium!hydrate}% bicarbonate with caustic soda to be: \[ \ce{(NaHCO3 $\aq$) + (NaHO $\aq$) - (Na2CO3 $\aq$)} = 9200~\Unit{cal.} \] He also found the heat of neutralization of carbon dioxide to be: \[ \ce{(CO2 $\aq$) + 2(NaHO $\aq$) - (Na2CO3 $\aq$)} = 20,200~\Unit{cal.} \] By subtraction \[ \ce{(CO2 $\aq$) + (NaHO $\aq$) - (NaHCO3 $\aq$)} = 11,000~\Unit{cal.} \] This is the heat effect corresponding to the direct combination of carbon dioxide and caustic soda to form sodium bicarbonate. Berthelot verified this by direct \index{Berthelot|)}% measurement. \Section{102.} Frequently, of two ways of transition, one is better adapted for calorimetric measurements than the other. Thus, the heat effect of the decomposition of hydrogen peroxide into water and oxygen cannot readily be measured directly. Thomsen therefore oxidized a solution \PageSep{74} of stannous chloride in hydrochloric acid first by means of hydrogen peroxide: \index{Hydrogen peroxide}% \[ \ce{(SnCl2 . 2HCl $\aq$) + (H2O2 $\aq$) - (SnCl4 $\aq$)} = 88,800~\Unit{cal.}, \] then by means of oxygen gas: \[ \ce{(SnCl2 . 2HCl $\aq$) + $\tfrac{1}{2}$ \{O2\} - (SnCl4 $\aq$)} = 65,700~\Unit{cal.} \] Subtraction gives \[ \ce{(H2O2 $\aq$) - $\tfrac{1}{2}$ \{O2\} - ($\aq$)} = 23,100~\Unit{cal.} \] for the heat effect of the decomposition of dissolved hydrogen peroxide into oxygen and water. \Section{103.} The \emph{heat of formation} of carbon monoxide from \index{Carbon, combustion of}% \index{Combustion, of carbon}% solid carbon and oxygen cannot be directly determined, because carbon never burns completely to carbon monoxide, but always, in part, to carbon dioxide as well. Therefore Favre and Silbermann determined the heat effect of the \index{Favre}% \index{Silbermann}% complete combustion of carbon to carbon dioxide: \[ \ce{[C] + \{O2\} - \{CO2\}} = 97,000~\Unit{cal.}, \] and then determined the heat effect of the combustion of carbon monoxide to carbon dioxide: \[ \ce{\{CO\} + $\tfrac{1}{2}$ \{O2\} - \{CO2\}} = 68,000~\Unit{cal.} \] By subtraction we get \[ \ce{[C] + $\tfrac{1}{2}$ \{O2\} - \{CO\}} = 29,000~\Unit{cal.}, \] the required heat of formation of carbon monoxide. \Section{104.} According to the above, theory enables us to calculate the heat effect of processes which cannot be directly realized, for as soon as the heat function of a system has been found in any way, it may be compared with other heat functions. Let the problem be, \eg, to find the heat of formation of liquid carbon bisulphide from solid carbon and solid sulphur, which do not combine directly. The following represent the reactions:--- \PageSep{75} The combustion of solid sulphur to sulphur dioxide gas: \[ \ce{[S] + \{O2\} - \{SO2\}} = 71,100~\Unit{cal.} \] The combustion of solid carbon to carbon dioxide: \[ \ce{[C] + \{O2\} - \{CO2\}} = 97,000~\Unit{cal.} \] The combustion of carbon bisulphide vapour to carbon dioxide and sulphur dioxide: \[ \ce{\{CS2\} + 3\{O2\} - \{CO2\} - 2\{SO2\}} = 265,100~\Unit{cal.} \] The condensation of carbon bisulphide vapour: \[ \ce{\{CS2\} - (CS2)} = 6400~\Unit{cal.} \] Elimination by purely mathematical processes furnishes the required heat of formation: \index{Heat!of combustion}% \index{Heat!of formation of \ce{CO2}, of \ce{CS2}, of \ce{CH4}}% \[ \ce{[C] + 2[S] - (CS2)} = - 19,500~\Unit{cal.}, \] hence negative. In organic thermochemistry the most important method of determining the heat of formation of a compound consists in determining the heat of combustion, first of the compound, and then of its constituents. Methane (marsh gas) gives by the complete combustion to carbon dioxide and water (liquid): \begin{alignat*}{2} &&\ce{\{CH4\} + 2\{O2\} - \{CO2\} - 2(H2O)} &= 211,900~\Unit{cal.}, \\ \text{but}&& \ce{\{H2\} + $\tfrac{1}{2}$ \{O2\} - (H2O)} &= \Z68,400~\Unit{cal.}, \Tag{(50)} \\ \text{and}&& \ce{[C] + \{O2\} - \{CO2\}} &= \Z97,000~\Unit{cal.}; \end{alignat*} therefore, by elimination, we obtain the heat of formation of methane from solid carbon and hydrogen gas: \[ \ce{[C] + 2\{H2\} - \{CH4\}} = 21,900~\Unit{cal.} \] \Section{105.} The external heat,~$Q$, of a given change at constant pressure will depend on the temperature at which the process is carried out. In this respect the first law of thermodynamics leads to the following relation:--- \PageSep{76} From equation~\Eq{(49)} it follows that, for any two given temperatures, $\theta$~and~$\theta'$, \[ (U_{2} + p_{0} V_{2})_{\theta} - (U_{1} + p_{0} V_{1})_{\theta} = Q_{\theta} \] and \[ (U_{2} + p_{0} V_{2})_{\theta'} - (U_{1} + p_{0} V_{1})_{\theta'} = Q_{\theta'}. \] Hence, by subtraction, \begin{multline*} Q_{\theta'} - Q_{\theta} = \bigl[(U_{2} + p_{0} V_{2})_{\theta'} - (U_{2} + p_{0} V_{2})_{\theta}\bigr] \\ - \bigl[(U_{1} + p_{0} V_{1})_{\theta'} - (U_{1} + p_{0} V_{1})_{\theta}\bigr]; \end{multline*} \ie\ the difference in the heat effects ($Q_{\theta} - Q_{\theta'}$) resulting \index{Heat!of combustion}% from performing the process at different temperatures, is equal to the difference in the quantities of heat which, before and after the reaction, would be required to raise the temperature of the system from~$\theta$ to~$\theta'$. Thus the influence of the temperature on the combustion \index{Combustion, of carbon!influence of temperature on}% \index{Influence!of temperature on combustion}% of hydrogen to water (liquid) may be found by comparing the heat capacity of the mixture (\ce{H2 + O2}) with that of the water (\ce{H2O}). The former is equal to the molecular heat of hydrogen plus half the molecular heat of oxygen. According to the table in~\SecRef{87}, this is \[ 6.82 + 3.47 = 10.29. \] The latter is \[ 1 × 18 = 18. \] The difference between these values is~$-7.71$, and, therefore, the heat of combustion of a gram molecule of hydrogen decreases with rising temperature by $7.7~\Unit{cal.}$ per degree Centigrade. \PageSep{77} \Part{III.}{The Second Fundamental Principle of~Thermodynamics.} \Chapter{I.}{Introduction.} \Section{106.} \First{The} second law of thermodynamics is essentially \index{Second law of thermodynamics!introduction}% different from the first law, since it deals with a question in no way touched upon by the first law, viz.\ the direction in which a process takes place in nature. Not every change which is consistent with the principle of the conservation of energy satisfies also the additional conditions which the second law imposes upon the processes, which actually take place in nature. In other words, the principle of the conservation of energy does not suffice for a unique determination of natural processes. If, for instance, an exchange of heat by conduction takes place between two bodies of different temperature, the first law, or the principle of the conservation of energy, merely demands that the quantity of heat given out by the one body shall be equal to that taken up by the other. Whether the flow of heat, however, takes place from the colder to the hotter body, or \textit{vice versâ}, cannot be answered by the energy principle alone. The very notion of temperature is alien to that principle, as can be seen from the fact that it yields no exact definition of temperature. Neither does the general equation~\Eq{(17)} of the first law contain any statement with \PageSep{78} regard to the direction of the particular process. The special equation~\Eq{(50)}, for instance, \[ \ce{\{H2\} + $\tfrac{1}{2}$ \{O2\} - (H2O)} = 68,400~\Unit{cal.}, \] means only that, if hydrogen and oxygen combine under constant pressure to form water, the \Erratum{restablishment}{re-establishment} of the initial temperature requires a certain amount of heat to be given up to surrounding bodies; and \textit{vice versâ}, that this amount of heat is absorbed when water is decomposed into hydrogen and oxygen. It offers no information, however, as to whether hydrogen and oxygen actually combine to form water, or water decomposes into hydrogen and oxygen, or whether such a process can take place at all in either direction. From the point of view of the first law, the initial and final \Erratum{stakes}{states} of any process are completely equivalent. \Section{107.} In one particular case, however, does the principle of the conservation of energy prescribe a certain direction to a process. This occurs when, in a system, one of the various forms of energy is at an absolute maximum (or minimum). It is evident that, in this case, the direction of the change must be such that the particular form of energy will decrease (or increase). This particular case is realized in mechanics by a system of particles at rest. Here the kinetic energy is at an absolute minimum, and, therefore, any change of the system is accompanied by an increase of the kinetic energy, and, if it be an isolated system, by a decrease of the potential energy. This gives rise to an important proposition in mechanics, which characterizes the direction of possible motion, and lays down, in consequence, the general condition of mechanical equilibrium. It is evident that, if both the kinetic and potential energies be at a minimum, no change can possibly take place, since none of these can increase at the expense of the other. The system must, therefore, remain at rest. If a heavy liquid be initially at rest at different levels in two communicating tubes, then motion will set in, so as to \PageSep{79} equalize the levels, for the centre of gravity of the system is thereby lowered, and the potential energy diminished. Equilibrium exists when the centre of gravity is at its lowest, and therefore the potential energy at a minimum, \ie\ when the liquid stands at the same level in both tubes. If no special assumption be made with regard to the initial velocity of the liquid, the above proposition no longer holds. The potential energy need not decrease, and the higher level might rise or sink according to circumstances. If our knowledge of thermal phenomena led us to recognize a state of minimum energy, a similar proposition would hold for this, but only for this, particular state. In reality no such minimum has been detected. It is, therefore, hopeless to seek to reduce the general laws regarding the direction of thermodynamical changes, as well as those of thermodynamical equilibrium, to the corresponding propositions in mechanics which hold good only for systems at rest. \Section{108.} Although these considerations make it evident that the principle of the conservation of energy cannot serve to determine the direction of a thermodynamical process, and therewith the conditions of thermodynamical equilibrium, unceasing attempts have been made to make the principle of the conservation of energy in some way or other serve this purpose. These attempts have, in many encases, stood in the way of a clear presentation of the second law. \Erratum{Occasionally we still find the endeavour made to represent this law as contained in the energy principle, in that the doubtless too restricted term of ``energetics'' is \index{Energetics}% applied to all investigations on these questions.}{That attempts are still made to represent this law as contained in the principle of energy may be seen from the fact that the too restricted term \index{Energetics}% ``energetics'' is sometimes applied to all investigations on these questions.} The conception of energy is not sufficient for the second law. It cannot be exhaustively treated by breaking up a natural process into a series of changes of energy, and then investigating the direction of each change. We can always tell, it is true, what are the different kinds of energy exchanged for one another; for there is no doubt that the principle of energy must be fulfilled, but the expression of \PageSep{80} the conditions of these changes remains arbitrary, and this ambiguity cannot be completely removed by any general assumption. We often find the second law stated as follows: The change of mechanical work into heat may be complete, but, on the contrary, that of heat into work must needs be incomplete, since, whenever a certain quantity of heat is transformed into work, another quantity of heat must undergo a corresponding and compensating change; \eg\ transference from higher to lower temperature. This is quite correct in certain very special cases, but it by no means expresses the essential feature of the process, as a simple example will show. An achievement which is closely associated with the discovery of the principle of energy, and which is one of the most important for the theory of heat, is the proposition expressed in equation~\Eq{(19)}, \SecRef{70}, that the total internal energy of a gas depends only on the temperature, and not on the volume. If a perfect gas be allowed to expand, doing external work, and be prevented from cooling by connecting it with a heat-reservoir of higher temperature, the temperature of the gas, and at the same time its internal energy, remains unchanged, and it may be said that the amount of heat given out by the reservoir is completely changed into work without an exchange of energy taking place anywhere. Not the least objection can be made to this. The proposition of the ``incomplete transformability of heat into work'' cannot \index{Transformability of heat into work}% be applied to this case, except by a different way of viewing the process, which, however, changes nothing in the physical facts, and cannot, therefore, be confirmed or refuted by them, namely, by the introduction of new kinds of energy, only invented \textit{ad~hoc}. This consists in dividing the energy of the gas into several parts, which may then individually depend also on the volume. This division has, however, to be carried out differently for different cases (\eg, in one way for isothermal, in another for adiabatic processes), and necessitates complicated considerations even in cases of physical simplicity. But when we pass from the \PageSep{81} consideration of the first law of thermodynamics to that of the second, we have to deal with a new fact, and it is evident that no definition, however ingenious, although it contain no contradiction in itself, will ever permit of the deduction of a new fact. \Section{109.} There is but one way of clearly showing the significance of the second law, and that is to base it on facts by formulating propositions which may be proved or disproved by experiment. The following proposition is of this character: It is in no way possible to completely reverse any process in which heat has been produced by friction. For the sake of example we shall refer to Joule's experiments on friction, described in~\SecRef{60}, for the determination of the mechanical equivalent of heat. Applied to these, our proposition says that, when the falling weights have generated heat in water or mercury by the friction of the paddles, no process can be invented which will completely restore everywhere the initial state of that experiment, \ie\ which will raise the weights to their original height, cool the liquid, and otherwise leave no change. The appliances used may be of any kind whatsoever, mechanical, thermal, chemical, electrical, etc., but the condition of \emph{complete} restoration of the initial state renders it necessary that all materials and machines used must ultimately be left exactly in the condition in which they were before their application. Such a proposition cannot be proved \textit{a~priori}, neither does it amount to a definition, but it contains a definite assertion, to be stated precisely in each case, which may be verified by actual experiment. The proposition is therefore correct or incorrect. \Section{110.} Another proposition of this kind, and closely connected with the former, is the following: It is in no way possible to completely reverse any process in which a gas expands without performing work or absorbing heat, \ie\ with constant total energy (as described in~\SecRef{68}). The word ``completely'' again refers to the accurate reproduction of the initial conditions. To test this, the gas, after \PageSep{82} it had assumed its new state of equilibrium, might first be compressed to its former volume by a weight falling to a lower level. External work is done on the gas, and it is thereby heated. The problem is now to bring the gas to its initial condition, and to raise the weight. The gas might be reduced to its original temperature by conducting the heat of compression into a colder heat-reservoir. In order that the process may be completely reversed, the reservoir must be deprived of the heat gained thereby, and the weight raised to its original position. This is, however, exactly what was asserted in the preceding paragraph to be impracticable. \Section{111.} A third proposition in point refers to the conduction of heat. Supposing that a body receives a certain quantity of heat from another of higher temperature, the problem is to completely reverse this process, \ie\ to convey back the heat without leaving any change whatsoever. In the description of Carnot's reversible cycle it has been pointed out, that heat can at any time be drawn from a heat-reservoir and transferred to a hotter reservoir without leaving any change except the expenditure of a certain amount of work, and the transference of an equivalent amount of heat from one reservoir to the other. If this heat could be removed, and the corresponding work recovered without other changes, the process of heat-conduction would be completely reversed. Here, again, we have the problem which was declared in~\SecRef{109} to be impracticable. Further examples of processes to which the same considerations \index{Processes!reversible and irreversible}% apply are, diffusion, the freezing of an overcooled liquid, the condensation of a supersaturated vapour, all explosive reactions, and, in fact, every transformation of a system into a state of greater stability. \Section{112.} A process which can in no way be completely reversed is termed \emph{irreversible}, all other processes \emph{reversible}. That a process may be irreversible, it is not sufficient that it cannot be directly reversed. This is the \PageSep{83} case with many mechanical processes which are not irreversible \index{Processes!periodic}% (\cf~\SecRef{113}). The full requirement is, that it be impossible, even with the assistance of all agents in nature, to restore everywhere the exact initial state when the process has once taken place. The propositions of the three preceding paragraphs, therefore, declare, that the generation of heat by friction, the expansion of a gas without the performance of external work and the absorption of external heat, the conduction of heat, etc., are irreversible processes. \Section{113.} We now turn to the question of the actual existence of reversible and irreversible processes. Numerous reversible processes can at least be imagined, as, for instance, those consisting of a succession of states of equilibrium, as fully explained in~\SecRef{71}, and, therefore, directly reversible in all their parts. Further, all perfectly periodic processes, \eg\ an ideal pendulum or planetary motion, are reversible, for, at the end of every period, the initial state is completely restored. Also, all mechanical processes with absolutely rigid bodies and absolutely incompressible liquids, as far as friction can be avoided, are reversible. By the introduction of suitable machines with absolutely unyielding connecting rods, frictionless joints and bearings, inextensible belts, etc., it is always possible to work the machines in such a way as to bring the system completely into its initial state without leaving any change in the machines, for the machines of themselves do not perform work. If, for instance, a heavy liquid, originally at rest at different levels in two communicating tubes (\SecRef{107}), be set in motion by gravity, it will, in consequence of its kinetic energy, go beyond its position of equilibrium, and, since the tubes are supposed frictionless, again swing back to its exact original position. The process at this point has been completely reversed, and therefore belongs to the class of reversible processes. As soon as friction is admitted, however, its reversibility is at least questionable. Whether reversible processes exist in nature or not, is not \textit{a~priori} evident or demonstrable. There is, however, no purely \PageSep{84} logical objection to imagining that a means may some day be found of completely reversing some process hitherto considered irreversible: one, for example, in which friction or heat-conduction plays a part. But it can be demonstrated---and this will be done in the following chapter---that if, in a single instance, one of the processes declared to be irreversible in \SSecRef{109},~etc., should be found to be reversible, then all of these processes must be reversible in all cases. Consequently, either all or none of these processes are irreversible. There is no third possibility. If those processes are not irreversible, the entire edifice of the second law will crumble. None of the numerous relations deduced from it, however many may have been verified by experience, could then be considered as universally proved, and theoretical work would have to start from the beginning. (The so-called proofs of ``energetics'' are not a substitute, for a \index{Energetics}% closer test shows all of them to be more or less imperfect paraphrases of the propositions to be proved. This is not the place, however, to demonstrate this point.) It is this foundation on the physical fact of irreversibility which forms the strength of the second law. If, therefore, it must be admitted that a single experience contradicting that fact would render the law untenable, on the other hand, any confirmation of part supports the whole structure, and gives to deductions, even in seemingly remote regions, the full significance possessed by the law itself. \Section{114.} Since the decision as to whether a particular process is irreversible or reversible depends only on whether the process can in any manner whatsoever be completely reversed or not, the nature of the initial and final states, and not the intermediate steps of the process, entirely settle it. The question is, whether or not it is possible, starting from the final state, to reach the initial one in any way without any other change. The second law, therefore, furnishes a relation between the quantities connected with the initial and final states of any natural process. The final state of an irreversible process is evidently in some way discriminate \PageSep{85} from the initial state, while in reversible processes the two states are in certain respects equivalent. The second law points out this characteristic property of both states, and also shows, when the two states are given, whether a transformation is possible in nature from the first to the second, or from the second to the first, without leaving changes in other bodies. For this purpose, of course, the two states must be fully characterized. Besides the chemical constitution of the systems in question, the physical conditions---viz.\ the state of aggregation, temperature, and pressure in both states---must be known, as is necessary for the application of the first law. The relation furnished by the second law will evidently be simpler the nearer the two states are to one another. On this depends the great fertility of the second law in its treatment of cyclic processes, which, however complicated they may be, give rise to a final state only slightly different from the initial state~(\SecRef{91}). \Section{115.} Since there exists in nature no process entirely free from friction or heat-conduction, all processes which actually take place in nature, if the second law be correct, are in reality irreversible; reversible processes form only an ideal limiting case. They are, however, of considerable importance for theoretical demonstration and for application to states of equilibrium. \PageSep{86} \Chapter{II.}{Proof.} \index{Second law of thermodynamics!proof}% \Section{116.} \First{The} second fundamental principle of thermodynamics being, like the first, an empirical law, we can speak of its proof only in so far as its total purport may be deduced from a single self-evident proposition. We, therefore, put forward the following proposition as being given directly by experience: \emph{It is impossible to construct an engine which will work in a complete cycle, and produce no effect except the raising of a weight and the cooling of a heat-reservoir.} Such an engine could be used simultaneously as a motor and a refrigerator without any waste of energy or material, and would in any case be the most profitable engine ever made. It would, it is true, not be equivalent to perpetual motion, for it does not produce work from nothing, but from the heat, which it draws from the reservoir. It would not, therefore, like perpetual motion, contradict the principle of energy, but would, nevertheless, possess for man the essential advantage of perpetual motion, the supply of work without cost; for the inexhaustible supply of heat in the earth, in the atmosphere, and in the sea, would, like the oxygen of the atmosphere, be at everybody's immediate disposal. For this reason we take the above proposition as our starting point. Since we are to deduce the second law from it, we expect, at the same time, to make a most serviceable application of any natural phenomenon which may be discovered to deviate from the second law. As soon as a phenomenon is found to contradict any legitimate conclusions from the second law, this contradiction must arise from an inaccuracy in our first assumption, and the phenomenon could be used \PageSep{87} for the construction of the above-described engine. We shall in the following, according to the proposal of Ostwald, speak of perpetual motion of the second kind, since it stands in the same relation to the second law as perpetual motion of the first kind does to the first law. In connection with all objections to the second law, it must be borne in mind that, if no errors are to be found in the line of proof, they are ultimately directed against the impossibility of perpetual motion of the second kind (\SecRef{136}).\footnote {I desire to emphasize here, that the starting point selected by me for the proof of the second law coincides fundamentally with that which R.~Clausius, \index{Clausius}% or which Sir W.~Thomson, or which J.~Clerk Maxwell used for the \index{Maxwell}% \index{Thomson}% same purpose. The fundamental proposition which each of these investigators placed at the beginning of his deductions asserts each time, only in different form, the impossibility of the realization of perpetual motion of the second kind. I have selected the above form of expression, because of its apparent technical significance. Not a single really rational proof of the second law has thus far been advanced which does not require this fundamental principle, however numerous the attempts in this direction may have been in recent times, nor do I believe that such an attempt will ever meet with success.} \Section{117.} From the impossibility of perpetual motion of the second kind, it follows, in the first place, that the generation of heat by friction is \emph{irreversible} (\cf\ def.~\SecRef{112}). For supposing it were not so, \ie\ supposing a method could be found by which a process involving generation of heat by friction could be completely reversed, this very method would produce what is identically perpetual motion of the second kind: viz.\ a change which consists of nothing but the production of work, and the absorption of an equivalent amount of heat. \Section{118.} It follows, further, that the expansion of a gas without the performance of external \Erratum{heat}{work}, or the absorption of heat, is irreversible. For, suppose a method were known of completely reversing this process, \ie\ of reducing the volume of a gas, without leaving any other change whatsoever, this method could be utilized for the production of perpetual motion of the second kind in the following manner. Allow the gas to do work by expansion, supplying the energy \PageSep{88} lost thereby by the conduction of heat from a reservoir at the same or higher temperature, and then, by the assumed method, reduce the volume of the gas to its initial value without leaving any other change. This process might be repeated as often as we please, and would therefore represent an engine working in a complete cycle, and producing no effect except the performance of work, and the withdrawal of heat from a reservoir, \ie\ perpetual motion of the second kind. On the basis of the proposition we have just proved, that the expansion of a gas without the performance of work and the absorption of heat is irreversible, we shall now carry through the proof of the second law for those bodies whose thermodynamical properties are most completely known, viz.\ for perfect gases. \Section{119.} If a perfect gas be subjected to infinitely slow compression or expansion, and if, at the same time, heat be applied or withdrawn, we have, by equation~\Eq{(22)}, in each infinitely small portion of the process, per unit mass, \[ q = du + p\, dv \] or, since for a perfect gas, \[ du = c_{v}\, d\theta, \] and \begin{align*} p &= \frac{R}{m} · \frac{\theta}{v}, \\ q &= c_{v}\, d\theta + \frac{R}{m} · \frac{\theta}{v}\, dv. \end{align*} If the process be adiabatic, then $q = 0$, and the integration of the above equation gives (as in~\SecRef{88}) the function \[ c_{v} \log \theta + \frac{R}{m} \log v \] equal to a constant. We shall now put \[ \phi = c_{v} \log \theta + \frac{R}{m} \log v + \const, \Tag{(51)} \] \PageSep{89} and call this function, after Clausius, the \emph{entropy} of unit mass \index{Entropy!of a gas}% of the gas. The constant, which has to be added, can be determined by arbitrarily fixing the zero state. Accordingly \[ \Phi = M\phi = M \left(c_{v} \log \theta + \frac{R}{m} \log v + \const\right) \Tag{(52)} \] is the entropy of mass~$M$ of the gas. The entropy of the gas, therefore, remains constant during the described adiabatic change of state. \Section{120.} On the application of heat, the entropy of the gas changes, in the case considered, by \[ d\Phi = M \left(c_{v}\, \frac{d\theta}{\theta} + \frac{R}{m}\, \frac{dv}{v}\right) = \frac{M · q}{\theta} = \frac{Q}{\theta}. \Tag{(53)} \] It increases or decreases according as heat is absorbed or evolved. The absorbed heat~$Q$ has here been broken up into two factors, $\theta$~and~$d\Phi$. According to a view which has recently been brought forward, this breaking up of heat into factors is regarded as a general property of heat. It should, however, be emphasized that equation~\Eq{(53)} is by no means generally true. It holds only in the particular case where the external work performed by the gas is expressed by~$p\, dV$. The relation \[ d\Phi = M \left(c_{v}\, \frac{d\theta}{\theta} + \frac{R}{m}\, \frac{dv}{v}\right) = \frac{dU + p\, dV}{\theta} \] holds, quite generally, for any process in which the temperature of the gas is increased by~$d\theta$, and the volume by~$dV$. It is, in fact, only a different mathematical form for the definition of the entropy given in~\Eq{(52)}. On the other hand, the equation \[ Q = dU + p\, dV \] holds by no means in all cases, but should, in general, be replaced by \[ Q + W = dU, \] \PageSep{90} where $W$, the work done on the substance, may have any value within certain limits. For instance, $W = 0$, if the gas be conveyed into its new state of equilibrium without performing external work (as described in~\SecRef{68}). In this case, $Q = dU$, and the equation $Q = \theta\, d\Phi$ no longer holds. \Section{121.} We shall now consider two gases which can communicate heat to one another by conduction, but may, in general, be under different pressures. If the volume of one, or both, of the gases be changed by some reversible process, care being taken that the temperatures of the gases equalize at each moment, and that no exchange of heat takes place with surrounding bodies, we have, according to equation~\Eq{(53)}, during any element of time, for the first gas, \[ d\Phi_{1} = \frac{Q_{1}}{\theta_{1}}, \] and, for the second gas, \[ d\Phi_{2} = \frac{Q_{2}}{\theta_{2}}. \] According to the conditions of the process, \[ \theta_{1} = \theta_{2}\quad\text{and}\quad Q_{1} + Q_{2} = 0, \] whence, \[ d\Phi_{1} + d\Phi_{2} = 0 \] or, for a finite change, \[ \Phi_{1} + \Phi_{2} = \const \Tag{(54)} \] The sum of the entropies of the two gases remains constant during the described process. \Section{122.} Any such process with two gases is evidently reversible in all its parts, for it may be directly reversed without leaving changes in the surroundings. From this follows the proposition that it is always possible to bring two gases, by a reversible process, without leaving changes \PageSep{91} \index{Condition of complete reversibility!of a process}% in\footnote {The emphasis is to be put on the word ``in.'' Changes of position of ponderable bodies (for example, the raising or lowering of weights) are not internal changes; but, of course, temperature and density changes are.} other bodies, from any given state to any other given state, if the sum of the entropies in the two states be equal. Let an initial state of the gases be given by the temperatures $\theta_{1}$,~$\theta_{2}$, and the specific volumes $v_{1}$,~$v_{2}$; a second state by the corresponding values $\theta_{1}'$,~$\theta_{2}'$; $v_{1}'$,~$v_{2}'$. We now suppose that \[ \Phi_{1} + \Phi_{2} = \Phi_{1}' + \Phi_{2}'\Add{.} \Tag{(55)} \] Bring the first gas to the temperature~$\theta_{2}$ by a reversible adiabatic compression or expansion; then place the two gases in thermal contact with one another, and continue to compress or expand the first infinitely slowly. Heat will now pass between the two gases, and the entropy of the first one will change, and it will be possible to make this entropy assume the value~$\Phi_{1}'$. But, according to~\Eq{(54)}, during the above process the sum of the two entropies remains constant, and $= \Phi_{1} + \Phi_{2}$; therefore the entropy of the second gas is $(\Phi_{1} + \Phi_{2}) - \Phi_{1}'$, which is, according to~\Eq{(55)}, equal to~$\Phi_{2}'$. If we now separate the two gases, and compress or expand each one adiabatically and reversibly until they have the required temperatures $\theta_{1}'$ and~$\theta_{2}'$, the specific volumes must then be $v_{1}'$ and~$v_{2}'$, and the required final state has been reached. This process is reversible in all its parts, and no changes remain in other bodies; in particular, the surroundings have neither gained nor lost heat. The conditions of the problem have therefore been fulfilled, and the proposition proved. \Section{123.} A similar proposition can readily be proved for any number of gases. It is always possible to bring a system of $n$~gases from any one state to any other by a reversible process without leaving changes in other bodies, if the sum of the entropies of all the gases is the same in both states, \ie\ if \[ \Phi_{1} + \Phi_{2} + \dots + \Phi_{n} = \Phi_{1}' + \Phi_{2}' + \dots + \Phi_{n}'. \Tag{(56)} \] \PageSep{92} \index{Zero!state}% By the process described in the preceding paragraph we may, by the successive combination of pairs of gases of the system, bring the first, then the second, then the third, and so on to the $(n - 1)$th~gas, to the required entropy. Now, \index{Entropy!of a system of gases}% in each of the successive processes the sum of the entropies of all the gases remains constant, and, since the entropies of the first $(n - 1)$ gases are $\Phi_{1}'$, $\Phi_{2}'$\Add{,}~\dots\Add{,} $\Phi_{n-1}'$, the entropy of the $n$th~gas is necessarily \[ (\Phi_{1} + \Phi_{2} + \dots + \Phi_{n}) - (\Phi_{1}' + \Phi_{2}' + \dots + \Phi_{n-1}'). \] This is, according to~\Eq{(56)}, the required value~$\Phi_{n}'$. Each gas can now be brought by an adiabatic reversible process into the required state, and the problem is solved. If we call the sum of the entropies of all the gases the entropy of the whole system, we may then say: \emph{If a system of gases has the same entropy in two different states it may be transformed from the one to the other by a reversible process, without leaving changes in other bodies.} \Section{124.} We now introduce the proposition proved in~\SecRef{118}, that the expansion of a perfect gas, without performing external work or absorbing heat, is irreversible; or, what is the same thing, that the transition of a perfect gas to a state of greater volume and equal temperature, without external effects, as described in~\SecRef{68}, is irreversible. Such a process corresponds to an increase of the entropy, according to the definition~\Eq{(52)}. It immediately follows that it is altogether impossible to decrease the entropy of a gas without producing a change in surrounding objects. If this were possible, the irreversible expansion of a gas could be completely reversed. After the gas had expanded without external effects, and had assumed its new state of equilibrium, the entropy of the gas could be reduced to its initial value, without leaving changes in other bodies, by the supposed method, and then, by an adiabatic reversible process, brought to its initial temperature, and thereby also to its original volume. This would completely reverse the \PageSep{93} first expansion, and furnish, according to~\SecRef{118}, perpetual motion of the second kind. \Section{125.} A system of two or more gases behaves in the same way. There exists, in nature, no means of diminishing the entropy of a system of perfect gases, without leaving \index{Entropy!diminution of}% changes in bodies outside the system. A contrivance which would accomplish this, be it mechanical, thermal, chemical, or electrical in nature, might be used to reduce the entropy of a single gas without leaving changes in other bodies. Suppose a system of gases to have passed in any manner from one state in which their entropies are $\Phi_{1}$, $\Phi_{2}$\Add{,}~\dots\Add{,} $\Phi_{n}$, to a state where they are $\Phi_{1}'$, $\Phi_{2}'$\Add{,}~\dots\Add{,} $\Phi_{n}'$, and that no change has been produced in any body outside the system, and let \[ \Phi_{1}' + \Phi_{2}' + \dots + \Phi_{n}' < \Phi_{1} + \Phi_{2} + \dots + \Phi_{n}, \Tag{(57)} \] then it is possible, according to the proposition proved in~\SecRef{123}, to bring the system by a reversible process, without leaving changes in other bodies, into any other state in which the sum of the entropies is \[ \Phi_{1}' + \Phi_{2}' + \dots + \Phi_{n}', \] and accordingly into a state in which the first gas has the entropy~$\Phi_{1}$, the second the entropy~$\Phi_{2}$\Add{,}~\dots, the $(n - 1)$th the entropy~$\Phi_{n-1}$, and the $n$th in consequence the entropy \[ (\Phi_{1}' + \Phi_{2}' + \dots + \Phi_{n}') - \Phi_{1} - \Phi_{2} - \dots - \Phi_{n-1}\Add{.} \Tag{(58)} \] The first $(n - 1)$ gases may now be reduced to their original state by reversible adiabatic processes. The $n$th~gas possesses the entropy~\Eq{(58)}, which is, according to the supposition~\Eq{(57)}, smaller than the original entropy~$\Phi_{n}$. The entropy of the $n$th~gas has, therefore, been diminished without leaving changes in other bodies. This we have already proved in the preceding paragraph to be impossible. \PageSep{94} The general proposition has, therefore, been proved, and we may immediately add the following. \Section{126.} \emph{If a system of perfect gases pass in any way from \index{Reversibility of a process!complete!condition of}% one state to another, and no changes remain in surrounding bodies, the entropy of the system is certainly not smaller, but either greater than, or, in the limit, equal to that of the initial state; in other words, the total change of the entropy $\geq 0$.} The sign of inequality corresponds to an irreversible process, the sign of equality to a reversible one. The equality of the entropies in both states is, therefore, not only a sufficient, as described in~\SecRef{123}, but also a necessary condition of the complete reversibility of the transformation from the one state to the other, provided no changes are to remain in other bodies. \Section{127.} The scope of this proposition is considerable, since there have designedly been imposed no restrictions regarding the way in which the system passes from its initial to its final state. The proposition, therefore, holds not only for slow and simple processes, but also for physical and chemical ones of any degree of complication, provided that at the end of the process no changes remain in any body outside the system. It must not be supposed that the entropy of a gas has a meaning only for states of equilibrium. We may assume each sufficiently small particle, even of a gas in turmoil, to be homogeneous and at a definite temperature, and must, therefore, according to~\Eq{(52)}, assign to it a definite value of the entropy. $M$,~$v$, and~$\theta$ are then the mass, specific volume, and temperature of the particle under consideration. A summation extending over all the particles of the mass---within which the values of $v$~and~$\theta$ may vary from particle to particle---gives the entropy of the whole mass of the gas in the particular state. The proposition still holds, that the entropy of the whole gas must continually increase during any process which does not give rise to changes in other bodies, \eg\ when a gas flows from a vessel into a vacuum (\SecRef{68}). It will be seen that the velocity of the gas particles does not influence \PageSep{95} the value of the entropy; neither does their height above a certain horizontal plane, although they are considered to have weight. \Section{128.} The laws which we have deduced for perfect gases may be transferred to any substance in exactly the same way. The main difference is, that the expression for the entropy of any body cannot, in general, be written down in finite quantities, since the characteristic equation is not generally known. But it can be demonstrated---and this is the deciding point---that, for any other body, there exists a function with the characteristic properties of the entropy. Imagine any homogeneous body to pass through a certain reversible or irreversible cycle and to be brought back to its exact original state, and let the external effects of this process consist in the performance of work and in the addition or withdrawal of heat. The latter may be brought about by means of any required number of suitable heat-reservoirs. After the process, no changes remain in the substance itself; the heat-reservoirs alone have suffered change. We shall now assume all the heat-reservoirs to be perfect gases, kept either at constant volume or at constant pressure, but, at any rate, subject only to reversible changes of volume. According to our last proposition, the sum of the entropies of all these gases cannot have decreased, since after the process no change remains in any other body. If $Q$~denote the amount of heat given to the substance during an infinitely small element of time by one of the reservoirs; $\theta$,~the temperature of the reservoir at that moment; then, according to equation~\Eq{(53)}, the reservoir's change of entropy during that element of time is \[ -\frac{Q}{\theta}. \] The change of the entropy of all the reservoirs, during all the elements of time considered, is \[ -\tsum \frac{Q}{\theta}. \] \PageSep{96} \index{Clausius' equation!form of second law}% Now, according to~\SecRef{126}, we have the following condition:--- \[ -\tsum \frac{Q}{\theta} \geq 0 \] or \[ \tsum \frac{Q}{\theta} \leq 0. \] This is the form in which the second law was first enunciated by Clausius. A further condition is given by the first law; for, according to~\Eq{(17)} in~\SecRef{63}, we have, during every element of time of the process, \[ Q + W = dU, \] $U$~being the initial energy of the body, and $W$~the work done on the body during the element of time. \Section{129.} If we now make the special assumption that the external pressure is, at any moment, equal to the pressure~$p$ of the substance, the work of compression becomes, according to~\Eq{(20)}, \[ W = -p\, dV, \] whence \[ Q = dU + p\, dV. \] If, further, each heat-reservoir be exactly at the temperature of the substance at the moment when brought into operation, the cyclic process is reversible, and the inequality of the second law becomes an equality\Add{:}--- \[ \tsum \frac{Q}{\theta} = 0, \] or, on substituting the value of~$Q$, \[ \tsum \frac{dU + p\, dV}{\theta} = 0. \] All the quantities in this equation refer to the state of the substance itself. It admits of interpretation without reference to the heat-reservoirs, and amounts to the following proposition. \PageSep{97} \Section{130.} \emph{If a homogeneous body be taken through a series of states of equilibrium {\upshape(\SecRef{71})}, that follow continuously from one another, back to its initial state, then the summation of the differential \[ \frac{dU + p\, dV}{\theta} \] extending over all the states of that process gives the value zero.} It follows that, if the process be not continued until the initial state,~$1$, is again reached, but be stopped at a certain state,~$2$, the value of the summation \[ \int_{1}^{2} \frac{dU + p\, dV}{\theta} \Tag{(59)} \] depends only on the states $1$~and~$2$, not on the manner of the transformation from state~$1$ to state~$2$. If two series of changes leading from~$1$ to~$2$ be considered (\eg\ curves $\alpha$~and~$\beta$ in \Fig{2}, \SecRef{75}), these can be combined into an infinitely slow cyclic process. We may, for example, go from~$1$ to~$2$ along~$\alpha$, and return to~$1$ along~$\beta$. It has been demonstrated that over the entire cycle\Add{:}--- \[ \int_{1\; (\alpha)}^{2} \frac{dU + p\, dV}{\theta} + \int_{2\; (\beta)}^{1} \frac{dU + p\, dV}{\theta} = 0, \] whence \[ \int_{1\; (\alpha)}^{2} \frac{dU + p\, dV}{\theta} = \int_{1\; (\beta)}^{2} \frac{dU + p\, dV}{\theta}. \] The integral~\Eq{(59)} with the above-proved properties has been called by Clausius the \emph{entropy} of the body in state~$2$, \index{Entropy!definition}% referred to state~$1$ as the zero state. The entropy of a body in a given state, like the internal energy, is completely determined up to an additive constant, whose value depends on the zero state. \PageSep{98} Denoting the entropy, as formerly, by~$\Phi$, we have: \[ \Phi = \int \frac{dU + p\, dV}{\theta} \] and \[ d\Phi = \frac{dU + p\, dV}{\theta} \Tag{(60)} \] or per unit mass: \[ d\phi = \frac{du + p\, dv}{\theta}\Add{.} \Tag{(61)} \] This, again, leads to the value~\Eq{(51)} for a perfect gas. The expression for the entropy of any body may be found by immediate integration (\SecRef{254}), provided its energy, $U = Mu$, and its volume, $V = Mv$, are known as functions, say, of $\theta$~and~$p$. Since, however, these are not completely known except for perfect gases, we have to content ourselves in general with the differential equation. For the proof, and for many applications of the second law, it is, however, sufficient to know that this differential equation contains in reality a unique definition of the entropy. \Section{131.} We may, therefore, just as in the case of perfect gases, speak of the entropy of any substance as of a finite quantity determined by the momentary values of temperature and volume, even when the substance undergoes reversible or irreversible changes. The differential equation~\Eq{(61)} holds, as was stated in~\SecRef{120} in the case of perfect gases, for any change of state, including irreversible changes. This more general application of the conception of the entropy in no wise contradicts the manner of its deduction. The entropy of a body in any given state is measured by means of a reversible process which brings the body from that state to the zero state. This ideal process, however, has nothing to do with any actual reversible or irreversible changes which the body may have undergone or be about to undergo. On the other hand, it should be stated that the differential equation~\Eq{(60)}, while it holds for changes of volume and temperature, does not apply to changes of mass, for \PageSep{99} this kind of change was in no way referred to in the definition of the entropy. Finally, we shall call the sum of the entropies of a number of bodies briefly the entropy of the system composed of those bodies. Thus the entropy of a body whose particles are not at uniform temperature, and have different velocities, may be found, as in the case of gases (\SecRef{127}), by a summation extending over all its elements of mass, provided the temperature and density within each infinitely small element of mass may be considered uniform. Neither the velocity nor the weight of the particles enter into the expression for the entropy. \Section{132.} The existence and the value of the entropy having been established for all states of a body, there is no difficulty in transferring the proof, which was given for perfect gases (beginning in~\SecRef{119}), to any system of bodies. Just as in~\SecRef{119} we find that, during reversible adiabatic expansion or compression of a body, its entropy remains constant, while by the absorption of heat the change of the entropy is \[ d\Phi = \frac{Q}{\theta}\Add{.} \Tag{(62)} \] This relation holds only for reversible changes of volume, as was shown for perfect gases in~\SecRef{120}. Besides, it is found, as in~\SecRef{121}, that during reversible expansion or compression of two bodies at a common temperature, if they be allowed to exchange heat by conduction with one another, but not with surrounding bodies, the sum of their entropies remains constant. A line of argument corresponding fully to that advanced for perfect gases then leads to the following general result:\footnote {With regard to the generalization of the theorem which was proved for a perfect gas in~\SecRef{124}, it may be stated that a certain difficulty arises in the special case of an incompressible body. In this case the body cannot be expanded. Professor Krigar-Menzel, who drew my attention to this, sent \index{Krigar@Krigar-Manzel|indexnote}% me at the same time the following proof. \emph{Proposition}: It is impossible to diminish the entropy of an incompressible body without leaving changes in other bodies. \emph{Proof}: Bring the body into thermal contact with a perfect gas, isolate the system adiabatically, and diminish the volume of the gas by reversible compression. Heat thereby passes from the gas into the body, and the entropy of the gas diminishes in consequence, while that of the body increases by an equal amount. Now separate the body from the gas. If the proposition were false, and there existed an uncompensated entropy diminishing process, we could by means of it bring the body back to its original smaller entropy, and therewith to its initial state. The only outstanding change of the whole process would be the diminution of the entropy of the perfect gas. But this contradicts \SecRef{118}. The proposition is therefore not false, but true.} \emph{It is impossible in any way to diminish the \PageSep{100} entropy of a system of bodies without thereby leaving behind \index{Entropy!principle of increase of}% changes in other bodies.} If, therefore, a system of bodies has changed its state in a physical or chemical way, without leaving any change in bodies not belonging to the system, then the entropy in the final state is greater than, or, in the limit, equal to the entropy in the initial state. The limiting case corresponds to reversible, all others to irreversible, processes. \Section{133.} The restriction, hitherto indispensable, that no \index{Carnot's!cycle}% changes must remain in bodies outside the system is easily dispensed with by including in the system all bodies that may be affected in any way by the process considered. The proposition then becomes: \emph{Every physical or chemical process in nature takes place in such a way as to increase the sum of the entropies of all the bodies taking any part in the process. In the limit, \ie\ for reversible processes, the sum of the entropies remains unchanged.} This is the most general statement of the second law of Thermodynamics. \Section{134.} As the impossibility of perpetual motion of the first kind leads to the first law of Thermodynamics, or the principle of the conservation of energy; so the impossibility of perpetual motion of the second kind has led to the second law, properly designated as the \emph{principle of the increase of the entropy}. This principle may be presented under other forms, which possess certain practical advantages, especially for isothermal or isopiestic processes. They will be mentioned in our next chapter. It should be emphasized, however, that the form here given is the only one of unrestricted \PageSep{101} applicability to any finite process, and that no other universal measure of the irreversibility of processes exists than the amount of the increase of the entropy to which they lead. All other forms of the second law are either applicable to infinitesimal changes only, or presuppose, when extended to finite changes, the existence of some special condition imposed upon the process (\SecRef{140}, etc.). The real meaning of the second law has frequently been looked for in a ``dissipation of energy.'' This view, proceeding, as it does, from \index{Dissipation of energy.}% \index{Energy!dissipation of}% the irreversible phenomena of conduction and radiation of heat, presents only one side of the question. There are irreversible processes in which the final and initial states show exactly the same form of energy, \eg\ the diffusion of two perfect gases (\SecRef{238}), or further dilution of a dilute solution. Such processes are accompanied by no perceptible transference of heat, nor by external work, nor by any noticeable transformation of energy.\footnote {In reply to a criticism of this statement, I have simply to refer to \SecRef{108}, wherein the remark is made, that, to be sure, by the introduction of new kinds of energy, conceived \textit{ad~hoc}, it is possible to speak of an energy transformation even for the cases now under discussion. There is, however, nothing arbitrary in the statement made in the text, where the energy appears as completely defined by~\SecRef{56}, but rather in the introduction of the new kinds of energy.} They occur only for the reason that they lead to an appreciable increase of the entropy. The amount of ``lost work'' yields a no more definite general measure of irreversibility than does that of ``dissipated energy.'' This is possible only in the case of isothermal processes (\SecRef{143}). An exhaustive general statement of the second law can be made only by means of the conception of the entropy. \Section{135.} Clausius summed up the first law by saying that \index{Clausius' equation!statement of first and second laws}% the energy of the world remains constant; the second by saying that the entropy of the world tends towards a maximum. Objection has justly been raised to this form of expression. The energy and the entropy of the world have no meaning, because such quantities admit of no accurate definition. Nevertheless, it is not difficult to \PageSep{102} express the characteristic feature of those propositions of Clausius in such a way as to give them a meaning, and to bring out more clearly what Clausius evidently wished to express by them. The energy of any system of bodies changes according to the measure of the effects produced by external agents. It remains constant, only, if the system be isolated. Since, strictly speaking, every system is acted on by external agents---for complete isolation cannot be realized in nature---the energy of a finite system may be approximately, but never absolutely, constant. Nevertheless, the more extended the system, the more negligible, in general, will the external effects become, in comparison with the magnitude of the energy of the system, and the changes of energy of its parts (\SecRef{55}); for, while the external effects are of the order of magnitude of the surface of the system, the internal energy is of the order of magnitude of the volume. In very small systems (elements of volume) the opposite is the case for the same reason, since here the energy of the system may be neglected in comparison with any one of the external effects. Frequent use is made of this proposition, \eg\ in establishing the limiting conditions in the theory of the conduction of heat. In the case here considered, it may, therefore, be said that the more widely extended a system we assume, the more approximately, in general, will its energy remain constant. A comparatively small error will be committed in assuming the energy of our solar system to be constant, a proportionately smaller one if the system of all known fixed stars be included. In this sense an actual significance belongs to the proposition, that the energy of an infinite system, or the energy of the world, remains constant. The proposition regarding the increase of the entropy should be similarly understood. If we say that the entropy of a system increases quite regardless of all outside changes, an error will, in general, be committed, but the more comprehensive the system, the smaller does the proportional error become. \PageSep{103} \Section{136.} In conclusion, we shall briefly discuss the question \index{Second law of thermodynamics!possible limitations}% of the possible limitations to the second law. If there exist any such limitations---a view still held by many scientists and philosophers---this much may be asserted, that their existence presupposes an error in our starting-point, viz.\ the impossibility of perpetual motion of the second kind, or a fault in our method of proof. From the beginning we have recognized the legitimacy of the first of these objections, and it cannot be removed by any line of argument. The second objection generally amounts to the following. The impracticability of perpetual motion of the second kind is granted, yet its absolute impossibility is contested, since our limited experimental appliances, supposing it were possible, would be insufficient for the realization of the ideal processes which the line of proof presupposes. This position, however, proves untenable. It would be absurd to assume that the validity of the second law depends in any way on the skill of the physicist or chemist in observing or experimenting. The gist of the second law has nothing to do with experiment; the law asserts briefly that \emph{there exists in nature a quantity which changes always in the same sense in all natural processes}. The proposition stated in this general form may be correct or incorrect; but whichever it may be, it will remain so, irrespective of whether thinking and measuring beings exist on the earth or not, and whether or not, assuming they do exist, they are able to measure the details of physical or chemical processes more accurately by one, two, or a hundred decimal places than we can.\footnote {We do not say that the second law is applicable to every single detail of a process. Upon closer examination the matter appears to be thus. The entropy, like temperature, pressure, and density, cannot be defined as an absolute, continuous quantity, but as a certain average value of a large number of single values. As long, therefore, as we regard simply one or more single values, the entropy cannot be defined any more than the temperature or pressure, and the second law neither applied nor proved.} The limitations to the law, if any, must lie in the same province as its essential idea, in the observed Nature, and not in the Observer. That man's experience is called upon in the deduction of the law is of no consequence; for that is, in fact, \PageSep{104} our only way of arriving at a knowledge of natural law. But the law once discovered must receive recognition of its independence, at least in so far as Natural Law can be said to exist independent of Mind. Should any one deny this, he would have to deny the possibility of natural science. The case of the first law is quite similar. To most unprejudiced scientists the impossibility of perpetual motion of the first kind is certainly the most direct of the general proofs of the principle of energy. Nevertheless, hardly any one would now think of making the validity of that principle depend on the degree of accuracy of the experimental proof of that general empirical proposition. Presumably the time will come when the principle of the increase of the entropy will be presented without any connection with experiment. Some \Erratum{metaphysicists}{metaphysicians} may even put it forward as being \textit{a~priori} valid. In the mean time, no more effective weapon can be used by both champions and opponents of the second law than indefatigable endeavour to follow the real purport of this law to the utmost consequences, taking the latter one by one to the highest court of appeal---experience. Whatever the decision may be, lasting gain will accrue to us from such a proceeding, since thereby we serve the chief end of natural science---the enlargement of our stock of knowledge. \PageSep{105} \Chapter{III.}{General Deductions.} \index{Deductions from second law of thermodynamics}% \index{Second law of thermodynamics!deductions}% \Section{137.} \First{Our} first application of the principle of the entropy which was expressed in its most general form in the preceding chapter, will be to Carnot's cycle, described in detail for perfect gases in~\SecRef{90}. This time, the system operated upon may be of any character whatsoever, and chemical reactions, too, may take place, provided they are reversible. Resuming the notation used in~\SecRef{90}, we may at once state the result. In a cyclic process, according to the first law, the heat,~$Q_{2}$, given out by the hotter reservoir is equivalent to the sum of the work done by the system, $W' = -W$, and the heat received by the colder reservoir, $Q_{1}' = -Q_{1}$: \[ Q_{2} = W' + Q_{1}' \] or \[ Q_{1} + Q_{2} + W = 0\Add{.} \Tag{(63)} \] According to the second law, since the process is reversible, all bodies which show any change of state after the process, \ie\ the two heat-reservoirs only, possess the same total entropy as before the process. The change of the entropy of the two reservoirs is, according to~\Eq{(62)}: \[ \frac{Q_{1}'}{\theta_{1}} = -\frac{Q_{1}}{\theta_{1}} \quad\text{for the first, and}\quad -\frac{Q_{2}}{\theta_{2}}\quad\text{for the second,} \Tag{(64)} \] their sum: \[ \frac{Q_{1}}{\theta_{1}} + \frac{Q_{2}}{\theta_{2}} = 0 \Tag{(65)} \] whence, by~\Eq{(63)}, \[ Q_{1} : Q_{2} : W = (-\theta_{1}) : \theta_{2} : (\theta_{1} - \theta_{2}) \] \PageSep{106} as in~\Eq{(44)}, but without any assumption as to the nature of the substance passing through the cycle of operations. In order, therefore, to gain the mechanical work,~$W'$, by means of a reversible Carnot cycle of operations with any substance between two heat-reservoirs at the temperatures $\theta_{1}$~and~$\theta_{2}$ ($\theta_{2} > \theta_{1}$), the quantity of heat \[ Q_{1}' = \frac{\theta_{1}}{\theta_{2} - \theta_{1}} W' \] must pass from the hotter to the colder reservoir. In other words, the passage of the quantity of heat~$Q_{1}'$ from~$\theta_{2}$ to~$\theta_{1}$ may be taken advantage of to gain the mechanical work \[ W' = \frac{\theta_{2} - \theta_{1}}{\theta_{1}} Q_{1}'\Add{.} \Tag{(66)} \] \Section{138.} For an irreversible cycle, \ie\ one involving any irreversible physical or chemical changes of the substance operated upon, the equation of energy~\Eq{(63)} still holds, but the equation for the change of the entropy~\Eq{(65)} is replaced by the inequality: \[ - \frac{Q_{1}}{\theta_{1}} - \frac{Q_{2}}{\theta_{2}} > 0. \] Observe, however, that the expressions~\Eq{(64)} for the change of the entropy of the reservoirs are still correct, provided we assume that any changes of volume of the substances used as reservoirs are reversible. Thus, \[ \frac{Q_{1}}{\theta_{1}} + \frac{Q_{2}}{\theta_{2}} < 0 \Tag{(67)} \] hence, from \Eq{(67)}~and \Eq{(63)}, \[ W' = -W = Q_{1} + Q_{2} < Q_{1} + \frac{\theta_{2}}{\theta_{1}} Q_{1}'\Add{,} \] or \[ W' < \frac{\theta_{2} - \theta_{1}}{\theta_{1}} Q_{1}'. \] \PageSep{107} This means that the amount of work,~$W'$, to be gained by means of a cyclic process from the transference of the heat,~$Q_{1}'$, from a hotter to a colder reservoir, is always smaller for an irreversible process than for a reversible one. Consequently the equation~\Eq{(66)} represents the maximum amount of work to be gained from any cyclic process between heat-reservoirs at the temperatures $\theta_{2}$ and~$\theta_{1}$. In particular, if $W' = 0$, it follows from the equation of energy~\Eq{(63)} that \[ Q_{2} = -Q_{1} = Q_{1}' \] and the inequality~\Eq{(67)} becomes \[ Q_{2} \left(\frac{1}{\theta_{2}} - \frac{1}{\theta_{1}}\right) < 0. \] In this case the cyclic process results in the transference of heat~($Q_{2}$) from the reservoir of temperature~$\theta_{2}$ to that of temperature~$\theta_{1}$, and the inequality means that this flow of heat is always directed from the hotter to the colder reservoir. Again, a special case of this type of process is the direct passage of heat by conduction between heat-reservoirs, without any actual participation of the system supposed to pass through the cycle of operations. It is seen to be an irreversible change, since it brings about an increase of the sum of the entropies of the two heat-reservoirs. \Section{139.} We shall now apply the principle of the entropy to any reversible or irreversible cycle with any system of bodies, in the course of which only one heat-reservoir of constant temperature~$\theta$ is used. Whatever may be the nature of the process in detail, there remains at its close no change of the entropy except that undergone by the heat-reservoir. According to the first law, we have \[ W + Q = 0. \] $W$~is the work done on the system, and $Q$~the heat absorbed by the system from the reservoir. According to the second law, the change of the entropy \PageSep{108} of the reservoir, within which only reversible changes of volume are supposed to take place, is \[ -\frac{Q}{\theta} \geq 0 \] or \[ Q \leq 0, \] whence \[ W \geq 0. \] Work has been expended on the system, and heat added to the reservoir. If, in the limit, the process be reversible, the signs of inequality disappear, and both the work~$W$ and the heat~$Q$ are zero. On this proposition rests the great fertility of the second law in its application to isothermal reversible cycles. \Section{140.} We shall no longer deal with cycles, but shall \index{Direction of natural process}% \index{Natural process, direction of}% consider the general question of the direction in which a change will set in, when any system in nature is given. For chemical reactions in particular is this question of importance. It is completely answered by the second law in conjunction with the first, for the second law contains a condition necessary for all natural processes. Let us imagine any homogeneous or heterogeneous system of bodies at the common temperature~$\theta$, and investigate the conditions for the starting of any physical or chemical change. According to the first law, we have for any infinitesimal change: \[ dU = Q + W, \Tag{(68)} \] where $U$~is the total internal energy of the system, $Q$~the heat absorbed by the system during the process, and $W$~the work done on the system. According to the second law, the change of the total entropy of all the bodies taking part in the process is \[ d\Phi + d\Phi_{0} \geq 0\Add{,} \] where $\Phi$~is the entropy of the system, $\Phi_{0}$~the entropy of the surrounding medium (air, calorimetric liquid, walls of \PageSep{109} vessels, etc.). Here the sign of equality holds for reversible cases, which, it is true, should be considered as an ideal limiting case of actual processes (\SecRef{115}). If we assume that all changes of volume in the surrounding medium are reversible, we have, according to~\Eq{(62)}, \[ d\Phi_{0} = -\frac{Q}{\theta}, \] or, by~\Eq{(68)}, \[ d\Phi_{0} = -\frac{dU - W}{\theta}. \] On substituting the value of~$d\Phi_{0}$, we have \[ d\Phi - \frac{dU - W}{\theta} \geq 0\Add{,} \Tag{(69)} \] or \[ dU - \theta\, d\Phi \leq W\Add{.} \Tag{(70)} \] All conclusions with regard to thermodynamic chemical changes, hitherto drawn by different authors in different ways, culminate in this relation~\Eq{(70)}. It cannot in general be integrated, since the left-hand side is not, in general, a perfect differential. The second law, then, does not lead to a general statement with regard to finite changes of a system taken by itself unless something be known of the external conditions to which it is subject. This was to be expected, and holds for the first law as well. To arrive at a law governing finite changes of the system, the knowledge of such external conditions as will permit the integration of the differential is indispensable. Among these the following are singled out as worthy of note. \Section{141.} \Topic{Case I\@. Adiabatic Process.}---No exchange of \index{Adiabatic process}% \index{Process!adiabatic}% heat with the surroundings being permitted, we have $Q = 0$, and, by~\Eq{(68)}, \[ dU = W. \] Consequently, by~\Eq{(70)}, \[ d\Phi \geq 0. \] The entropy of the system increases or remains constant, a case which has already been sufficiently discussed. \PageSep{110} \Section{142.} \Topic{Case II\@. Isothermal Process.}---The temperature \index{Process!isothermal}% $\theta$ being kept constant, \Eq{(70)}~passes into \[ d(U - \theta\Phi) \leq W\Add{,} \] \ie\ the increment of the quantity $(U - \theta\Phi)$ is smaller than, or, in the limit, equal to, the work done on the system. This theorem is well adapted for application to chemical processes, since isothermal changes play an important part in nature. Putting \[ U - \theta\Phi = F, \Tag{(71)} \] we have, for reversible isothermal changes: \[ dF = W \] and, on integrating, \[ F_{2} - F_{1} = \tsum W\Add{.} \Tag{(72)} \] For finite reversible isothermal changes the total work done on the system is equal to the increase of~$F$; or, the entire work performed by the system is equal to the decrease of~$F$, and, therefore, depends only on the initial and final states of the system. Where $F_{1} = F_{2}$, as in cyclic processes, the external work is zero. The function~$F$, thus bearing the same relation to the external work that the energy~$U$ does to the sum of the \index{Energy!free}% \index{Energy!latent}% \index{Energy!total}% \index{Heat, latent!theory}% external heat and work, has been called by H.~v.~Helmholtz the \emph{free energy} (freie Energie) of the system. (It should \index{Free energy}% rather be called ``free energy for isothermal processes.'') \index{Isothermal processes}% Corresponding to this, he calls~$U$ the \emph{total energy} (Gesammtenergie), and the difference $U - F = \theta\Phi$, the \emph{latent energy} \index{Latent energy}% (gebundene Energie) of the system. The change of the latter in reversible isothermal processes gives the amount of the external heat absorbed. This splitting up of total energy into free and latent energy is applicable to isothermal processes only. In irreversible processes, on the other hand, $dF < W$, and on integrating we have \[ F_{2} - F_{1} < \tsum W\Add{.} \Tag{(73)} \] \PageSep{111} The free energy increases by a less amount than that which corresponds to the work done on the system. The results \index{Work!maximum}% for reversible and irreversible processes may be stated thus. In irreversible isothermal processes the work done on the system is more, or the work done by the system is less, than it would be if the same change were brought about by a reversible process, for in that case it would be the difference of the free energies at the beginning and end of the process~\Eq{(72)}. Hence, any reversible transformation of the system from one state to another yields the maximum amount of work that can be gained by any isothermal process between those two states. In all irreversible processes a certain amount of work is lost, viz.\ the difference between the maximum work \index{Maximum work}% to be gained (the decrease of the free energy) and the work actually gained. The fact that, in the above, irreversible as well as reversible processes between the same initial and final states were considered, does not contradict the proposition that between two states of a system either only reversible or only irreversible processes are possible, if no external changes are to remain in other bodies. In fact, the process here discussed involves such changes in the surrounding medium; for, in order to keep the system at constant temperature, an exchange of heat between it and the surrounding medium must take place in one direction or the other. \Section{143.} If the work done during an isothermal process vanish, as is practically the case in most chemical reactions, we have \[ \tsum W = 0, \] and, by~\Eq{(73)}, \[ F_{2} - F_{1} < 0, \] \ie\ the free energy decreases. The amount of this decrease may be used as a measure of the work done by the forces (chemical affinity) causing the process, for the same is not available for external work. For instance, let an aqueous solution of some non-volatile \PageSep{112} \index{Affinity of hydrogen for oxygen}% \index{Decrease of free energy by dilution}% \index{Hydrogen, affinity of, for oxygen}% salt be diluted isothermally, the heat of dilution being \index{Dilution!decrease of free energy by}% furnished or received by a heat-reservoir according as the energy,~$U_{2}$, of the diluted solution (final state) is greater or less than the sum,~$U_{1}$, of the energies of the undiluted solution and the water added (initial state). The free energy,~$F_{2}$, of the diluted solution, on the other hand, is necessarily smaller than the sum,~$F_{1}$, of the free energies of the undiluted solution and the water added. The amount of the decrease of the free energy, or the work done by the \index{Free energy!decrease of, by dilution}% ``affinity of the solution for water'' during the process of dilution may be measured. For this purpose, the dilution should be performed in some reversible isothermal manner, when, according to~\Eq{(72)}, the quantity to be measured is actually gained in the form of external work. For instance, evaporate the water, which is to be added, infinitely slowly under the pressure of its saturated vapour. When it has all been changed to water vapour, allow the latter to expand isothermally and reversibly until its density equals that which saturated water vapour would possess at that temperature when in contact with the solution. Now establish lasting contact between the water vapour and the solution, whereby the equilibrium will not be disturbed. Finally, by isothermal compression, condense the water vapour infinitely slowly when in direct contact with the solution. It will then be uniformly distributed throughout the latter. Such a process, as here described, is composed only of states of equilibrium. Hence it is reversible, and the external work thereby gained represents at the same time the decrease of the free energy, $F_{2} - F_{1}$, which takes place on directly mixing the solution and the water. As a further example, we shall take a mixture of hydrogen and oxygen which has been exploded by means of an electric spark. The spark acts only the secondary part of a release, its energy being negligible in comparison with the energies obtained by the reaction. The work of the chemical affinities in this process is equal to the mechanical work that might be gained by chemically combining the oxygen and hydrogen in some reversible and \PageSep{113} isothermal way. Dividing this quantity by the number of oxidized molecules of hydrogen, we obtain a measure of the force with which a molecule of hydrogen tends to become oxidized. This definition of chemical force, however, has only a meaning in so far as it is connected with that work. \Section{144.} In chemical processes the changes of the first \index{Berthelot's principle}% \index{Principle of Berthelot}% term,~$U$, of the expression for the free energy~\Eq{(71)}, frequently \index{Energy!free!of perfect gas}% \index{Free energy!change of, with temperature.}% \index{Free energy!of a perfect gas}% far surpass those of the second,~$\theta\Phi$. Under such circumstances, instead of the decrease of~$F$, that of~$U$, \ie\ the heat effect, may be considered as a measure of the chemical work. This leads to the proposition that chemical reactions, in which there is no external work, take place in such a manner as to give the greatest heat effects (Berthelot's principle). For high temperatures, where~$\theta$, and for gases and dilute solutions, where $\Phi$~is large, the term~$\theta\Phi$ can no longer be neglected without considerable error. In these cases, therefore, chemical changes often do take place in such a way as to increase the total energy, \ie\ with the absorption of heat. \Section{145.} It should be borne in mind that all these propositions refer only to isothermal processes. To answer the question as to how the free energy acts in other processes, it is only necessary to form the differential of~\Eq{(71)} viz.: \[ dF = dU - \theta\, d\Phi - \Phi\, d\theta\Add{,} \] and to substitute in the general relation~\Eq{(70)}. We have then \[ dF \leq W - \Phi\, d\theta \] for any physical or chemical process. This shows that, with change of temperature, the relation between the external work and the free energy is far more complicated. This relation cannot, in general, be used with advantage. \Section{146.} We shall now compute the value of the free energy of a perfect gas. Here, according to~\Eq{(35)}, \[ U = Mu = M(c_{v} \theta + \const), \] \PageSep{114} and, by~\Eq{(52)}, \[ \Phi = M\phi = M(c_{v} \log \theta + \frac{R}{m} \log v + \const). \] Substituting in~\Eq{(71)}, we obtain \[ F = M\{c_{v} \theta (\const - \log \theta) - \frac{R\theta}{m} \log v + \const\} \Tag{(74)} \] which contains an arbitrary linear function of~$\theta$. For isothermal changes of the gas, we have, by~\SecRef{142}, \[ dF \leq W, \] or, by~\Eq{(74)}, since $\theta$~is $\const$, \[ dF = -\frac{M\theta R}{m} · \frac{dv}{v} = -p\, dV \leq W. \] If the change be reversible, the external work on the gas is $W = -p\, dV$, but if it be irreversible, then the sign of inequality shows that the work of compression is greater, or that of expansion smaller, than in a reversible process. \index{Process!isothermal-isopiestic}% \Section{147.} \Topic{Case III\@. Isothermal-isopiestic Process.}---If, \index{Isothermal-isopiestic process}% besides the temperature~$\theta$, the external pressure~$p$ be also kept constant, then the external work is given by the formula, \[ W = -p\, dV, \] and the left-hand side of~\Eq{(69)} becomes a complete differential: \[ d\left(\Phi - \frac{U + pV}{\theta}\right) \geq 0\Add{.} \] In this case, it may be stated that for finite changes the \index{Function $\Psi$}% function, \[ \Phi - \frac{U + pV}{\theta} = \Psi, \Tag{(75)} \] \PageSep{115} must increase, and will remain constant only in the limit \index{Duhem@Duhem|indexnote}% \index{Potential, energy!thermodynamic@thermodynamic|indexnote}% \index{Thermodynamic potential}% when the change is reversible.\footnote {Multiplying \Eq{(75)} by~$-\theta$, we get P.~Duhem's \emph{thermodynamic potential at constant pressure}, \[ U + pV - \theta\Phi, \] for which, so long as $\theta$~remains constant, the same propositions hold as for the function~$\Psi$. However, the equation~\Eq{(153)} in \SecRef{211}, which is important for the dependence of the equilibrium on temperature and pressure, can be \index{Equilibrium!conditions of}% more conveniently deduced from the function~$\Psi$ than from the thermodynamic potential.} \Section{148.} \Topic{Conditions of Equilibrium.}---The most general \index{Condition of complete reversibility!of equilibrium}% condition of equilibrium for any system of bodies is derived from the proposition that no change can take place in the system if it be impossible to satisfy the condition necessary for a change. Now, by~\Eq{(69)}, for any actual change of the system, \[ d\Phi - \frac{dU - W}{\theta} > 0. \] The sign of equality is omitted, because it refers to ideal changes which do not actually occur in nature. Equilibrium is, therefore, maintained if the fixed conditions imposed on the system be such that they will permit only changes in which \[ \delta\Phi - \frac{\delta U - W}{\theta} \leq 0. \] Here $\delta$~is used to signify a virtual infinitely small change, in contrast to~$d$, which corresponds to an actual change. \Section{149.} In most of the cases subsequently discussed, if any given virtual change be compatible with the fixed conditions of the system, its exact opposite is also, and is represented by changing the sign of all variations involved. This is true if the fixed conditions be expressed by equations, not by inequalities. Assuming this to be the case, if we should have, for any particular virtual change, \[ \delta\Phi - \frac{\delta U - W}{\theta} < 0, \] \PageSep{116} which, by~\Eq{(69)}, would make its occurrence in nature impossible, its opposite would conform to the condition for actual changes~\Eq{(69)}, and could therefore take place in nature. To ensure equilibrium in such cases, it is necessary, therefore, that, for any virtual change compatible with the fixed conditions, \[ \delta\Phi - \frac{\delta U - W}{\theta} = 0\Add{.} \Tag{(76)} \] This equation contains a condition always sufficient, but, as we have seen, not always necessary to its full extent, for the maintenance of equilibrium. As a matter of experience, equilibrium will occasionally subsist when equation~\Eq{(76)} is not fulfilled, even though the fixed conditions permit of a change of sign of all variations. This is to say, that occasionally a certain change will not take place in nature, though it satisfy the fixed conditions as well as the demands of the second law. Such cases lead to the conclusion that in some way the setting in of a change meets with a certain resistance, which, on account of the direction in which it acts, has been termed inertia resistance, or passive resistance. \index{Inertia resistance}% \index{Resistance inertia}% States of equilibrium of this description are always unstable. Often a very small disturbance, not comparable in size with the quantities within the system, suffices to produce the change, which under these conditions often occurs with great violence. We have examples of this in overcooled liquids, supersaturated vapour, supersaturated solutions, explosive substances, etc. We shall henceforth discuss mainly the conditions of stable equilibrium deducible from~\Eq{(76)}. This equation may, under certain circumstances, be expressed in the form of a condition for a maximum or minimum. This can be done when, and only when, the conditions imposed upon the system are such that the left-hand side of~\Eq{(76)} represents the variation of some one function. The most important of these cases are dealt with separately in the following paragraphs. They correspond exactly to the propositions which we have already deduced \PageSep{117} for special cases. From these propositions it may at once be seen whether it is a case of a maximum or a minimum. \Section{150.} \Topic{First Case} (\SecRef{141}).---If no exchange of heat take place with the surrounding medium, the first law gives \[ \delta U = W, \] hence, by~\Eq{(76)}, \[ \delta \Phi = 0\Add{.} \Tag{(77)} \] Among all the states of the system which can proceed from one another by adiabatic processes, the state of equilibrium is distinguished by a maximum of the entropy. Should \index{Entropy!maximum value}% \index{Maximum value!of entropy}% \index{Maximum value!of free energy}% there be several states in which the entropy has a maximum value, each one of them is a state of equilibrium; but if the entropy be greater in one than in all the others, then that state represents absolutely stable equilibrium, for it could no longer be the starting-point of any change whatsoever. \Section{151.} \Topic{Second Case} (\SecRef{142}).---If the temperature be kept constant, equation~\Eq{(76)} passes into \[ \delta\left(\Phi - \frac{U}{\theta}\right) + \frac{W}{\theta} = 0, \] and, by~\Eq{(71)}, \[ - \delta F = -W. \] Among all the states which the system may assume at a given temperature, a state of equilibrium is characterized by the fact that the free energy of the system cannot \index{Free energy!minimum value of}% decrease without performing an equivalent amount of work. If the external work be a negligible quantity, as it is when the volume is kept constant or in numerous chemical processes, then $W = 0$, and the condition of equilibrium becomes \[ \delta F = 0, \] \ie\ among the states which can proceed from one another by isothermal processes, without the performance of external work, the state of most stable equilibrium is distinguished by an absolute minimum of the free energy. \PageSep{118} \Section{152.} \Topic{Third Case} (\SecRef{147}).---Keeping the temperature~$\theta$ \index{Maximum value!of $\Psi$}% and the pressure~$p$ constant and uniform, we have \[ W = -p\, \delta V, \Tag{(78)} \] and the condition of equilibrium~\Eq{(76)} becomes \[ \delta\left(\Phi - \frac{U + pV}{\theta}\right) = 0, \] or, by~\Eq{(75)}, \[ \delta\Psi = 0\Add{,} \Tag{(79)} \] \ie\ at constant temperature and constant pressure, the state of most stable equilibrium is characterized by an absolute maximum of the function~$\Psi$. We shall now proceed to consider, in succession, states of equilibrium of various systems by means of the theorems we have just deduced, going from simpler to more complicated cases. \PageSep{119} \Part{IV.}{Applications to Special States of Equilibrium.} \Chapter{I.}{Homogeneous Systems.} \index{Homogeneous!system|(}% \Section{153.} \First{Let} the state of a homogeneous system be determined, \index{System!homogeneous|(}% as hitherto, by its mass,~$M$; its temperature,~$\theta$; and either its pressure,~$p$, or its specific volume, $v = \dfrac{V}{M}$. For the present, besides~$M$, let $\theta$~and $v$ be the independent variables. Then the pressure~$p$, the specific energy $u = \dfrac{U}{M}$, and the specific entropy $\phi = \dfrac{\Phi}{M}$ are functions of $\theta$~and~$v$, \index{Entropy!specific}% \index{Specific entropy}% the definition of the specific entropy~\Eq{(61)} being \[ d\phi = \frac{du + p\, dv}{\theta} = \frac{1}{\theta} \left(\frac{\dd u}{\dd \theta}\right)_{v} d\theta + \frac{\left(\dfrac{\dd u}{\dd v}\right)_{\theta} + p}{\theta}\, dv. \] On the other hand, \[ d\phi = \left(\frac{\dd \phi}{\dd \theta}\right)_{v} d\theta + \left(\frac{\dd \phi}{\dd v}\right)_{\theta} dv. \] Therefore, since $d\theta$~and $dv$ are independent of each other, \begin{align*} \left(\frac{\dd \phi}{\dd \theta}\right)_{v} &= \frac{1}{\theta} \left(\frac{\dd u}{\dd \theta}\right)_{v} \Tag{(79a)} \\ \intertext{and} \left(\frac{\dd \phi}{\dd v}\right)_{\theta} &= \frac{\left(\dfrac{\dd u}{\dd v}\right)_{\theta} + p}{\theta}\Add{.} \end{align*} \PageSep{120} These two equations lead to an experimental test of the second law; for, differentiating the first with respect to~$v$, the second with respect to~$\theta$, we have \[ \frac{\dd^{2} \phi}{\dd \theta\, \dd v} = \frac{1}{\theta} · \frac{\dd^{2} u}{\dd \theta\, \dd v} = \frac{\dfrac{\dd^{2} u}{\dd \theta\, \dd v} + \left(\dfrac{\dd p}{\dd \theta}\right)_{v}}{\theta} - \frac{\left(\dfrac{\dd u}{\dd v}\right)_{\theta} + p}{\theta^{2}}\Add{,} \] or \[ \left(\frac{\dd u}{\dd v}\right)_{\theta} = \theta \left(\frac{\dd p}{\dd \theta}\right)_{v} - p\Add{.} \Tag{(80)} \] By this and equation~\Eq{(24)}, the above expressions for the differential coefficients of~$\phi$ become: \[ \left(\frac{\dd \phi}{\dd \theta}\right)_{v} = \frac{c_{v}}{\theta}; \quad\text{and}\quad \left(\frac{\dd \phi}{\dd v}\right)_{\theta} = \left(\frac{\dd p}{\dd \theta}\right)_{v}\Add{.} \Tag{(81)} \] \Section{154.} Equation~\Eq{(80)}, together with~\Eq{(28)} of the first law, gives the relation: \[ c_{p} - c_{v} = \theta \left(\frac{\dd p}{\dd \theta}\right)_{v} · \left(\frac{\dd v}{\dd \theta}\right)_{p}\Add{,} \Tag{(82)} \] which is useful either as a test of the second law or for the calculation of~$c_{v}$ when $c_{p}$~is given. But since in many cases $\left(\dfrac{\dd p}{\dd \theta}\right)_{v}$ cannot be directly measured, it is better to introduce the relation~\Eq{(6)}, and then \[ c_{p} - c_{v} = -\theta \left(\frac{\dd p}{\dd v}\right)_{\theta} · \left(\frac{\dd v}{\dd \theta}\right)_{p}^{2}\Add{.} \Tag{(83)} \] As $\left(\dfrac{\dd p}{\dd v}\right)_{\theta}$ is necessarily negative, $c_{p}$~is always greater than~$c_{v}$, except in the limiting case, when the coefficient of expansion is $= 0$, as in the case of water at $4°$~C.\Chg{,}{}; then $c_{p} - c_{v} = 0$. As an example, we may calculate the specific heat at \index{Specific heat!at constant pressure}% \index{Specific heat!at constant volume}% constant volume,~$c_{v}$, of mercury at $0°$~C. from the following data: \begin{gather*} c_{p} = 0.0333;\quad \theta = 273°; \\ \left(\frac{\dd p}{\dd v}\right)_{\theta} = -\frac{1014000}{0.00000295 · v}, \end{gather*} \PageSep{121} where the denominator is the coefficient of compressibility in atmospheres (\SecRef{15}); the numerator, the pressure of an atmosphere in absolute units (\SecRef{17}); $v = \dfrac{1}{13.6}$, the volume of $1~\Unit{gr.}$ of mercury at $0°$~C.; $\left(\dfrac{\dd v}{\dd p}\right)_{p} = 0.0001812 · v$, the coefficient of thermal expansion~(\SecRef{15}). To obtain $c_{v}$~in calories, it is necessary to divide by the mechanical equivalent of heat, $419 × 10^{5}$ (\SecRef{61}). Thus we obtain from~\Eq{(83)} \[ c_{p} - c_{v} = \frac{273 × 1014000 × 0.0001812^{2}} {0.00000295 × 13.6 × 419 × 10^{5}} = 0.0054, \] whence, from the above value for~$c_{p}$, \[ c_{v} = 0.0279. \] \Section{155.} This method of calculating the difference of the \index{Difference of specific heats}% \index{Specific heats!difference of}% specific heats $c_{p} - c_{v}$ applicable to any substance, discloses at the same time the order of magnitude of the different influences to which this quantity is subject. According to equation~\Eq{(28)} of the first law, the difference of the specific heats is \[ c_{p} - c_{v} = \left\{\left(\frac{\dd u}{\dd v}\right)_{\theta} + p\right\} \left(\frac{\dd v}{\dd \theta}\right)_{p}\Add{.} \] The two terms of this expression, $\left(\dfrac{\dd u}{\dd v}\right)_{\theta} \left(\dfrac{\dd v}{\dd \theta}\right)_{p}$ and $p\left(\dfrac{\dd v}{\dd \theta}\right)_{p}$, depend on the rate of change of the energy with the volume, and on the external work performed by the expansion respectively. In order to find which of these two terms has the greater influence on the quantity $c_{p} - c_{v}$, we shall find the ratio of the first to the second: \[ \frac{1}{p} · \left(\frac{\dd u}{\dd v}\right)_{\theta}, \] or, by~\Eq{(80)}, \[ \frac{\theta}{p} · \left(\frac{\dd p}{\dd \theta}\right)_{v} - 1\Add{,} \Tag{(84)} \] or by~\Eq{(6)}, \[ -\frac{\theta}{p} · \left(\frac{\dd v}{\dd \theta}\right)_{p} · \left(\frac{\dd p}{\dd v}\right)_{\theta} - 1\Add{.} \] \PageSep{122} A glance at the tables of the coefficients of thermal expansion and of the compressibility of solids and liquids shows that, in general, the first term of this expression is a large number, making the second,~$-1$, a negligible quantity. For mercury at~$0°$, \eg, the above data give the first term to be \[ 273 × \frac{0.0001812}{0.00000295} = 16800. \] Water at $4°$~C. is an exception. It follows that, for solids and liquids, the difference $c_{p} - c_{v}$ depends rather on the relation between the energy and the volume than on the external work of expansion. For perfect gases the reverse is the case, since the internal energy is independent of the volume, \ie--- \[ \left(\frac{\dd u}{\dd v}\right)_{\theta} = 0. \] During expansion, therefore, the influence of the internal energy vanishes in comparison with that of the external work; in fact, the expression~\Eq{(84)} vanishes for the characteristic equation of a perfect gas. With ordinary gases, however, both the internal energy and the external work must be considered. \Section{156.} The sum of both these influences, \ie\ the whole \index{Ratio of specific heats}% \index{Specific heats!ratio of}% expression $c_{p} - c_{v}$, may be said to have a small value for most solids and liquids; thus the ratio $\dfrac{c_{p}}{c_{v}} = \gamma$ is but slightly greater than unity. This means that in solids and liquids the energy depends far more on the temperature than on the volume. For gases, $\gamma$~is large; and, in fact, the fewer the number of atoms in a molecule of the gas, the larger does it become. Hydrogen, oxygen, and most gases with diatomic molecules have $\gamma = 1.41$ (\SecRef{87}). The largest value of~$\gamma$ ever observed is that found by Kundt and Warburg for the \index{Kundt}% \index{Warburg}% monatomic vapour of mercury, viz.~$1.666$. \Section{157.} For many applications of the second law it is \PageSep{123} \index{Deviation from perfect gases}% \index{Gay-Lussac's law, deviations from}% \index{Influence!of pressure on specific heat}% \index{Specific heat!influence of temperature on, at constant pressure}% convenient to introduce~$p$ instead of~$v$ as an independent variable. We have, by~\Eq{(61)}, %[** TN: Re-breaking] \begin{align*} d\phi &= \frac{du + p\, dv}{\theta} \\ &= \biggl\{\left(\frac{\dd u}{\dd \theta}\right)_{p} + p\left(\frac{\dd v}{\dd \theta}\right)_{p}\biggr\} \frac{d\theta}{\theta} + \biggl\{\left(\frac{\dd u}{\dd p}\right)_{\theta} + p\left(\frac{\dd v}{\dd p}\right)_{\theta}\biggr\} \frac{dp}{\theta}\Add{.} \end{align*} On the other hand, \[ d\phi = \left(\frac{\dd \phi}{\dd \theta}\right)_{p} d\theta + \left(\frac{\dd \phi}{\dd p}\right)_{\theta} dp\Add{,} \] % [** TN: Next two displayed equations brace-grouped in the original] whence, \[ \left(\frac{\dd \phi}{\dd \theta}\right)_{p} = \frac{\left(\dfrac{\dd u}{\dd \theta}\right)_{p} + p\left(\dfrac{\dd v}{\dd \theta}\right)_{p}}{\theta} \] and \[ \left(\frac{\dd \phi}{\dd p}\right)_{\theta} = \frac{\left(\dfrac{\dd u}{\dd p}\right)_{\theta} + p\left(\dfrac{\dd v}{\dd p}\right)_{\theta}}{\theta}\Add{.} \] Differentiating the first of these with respect to~$p$, the second with respect to~$\theta$, we get %[** TN: Re-breaking] \begin{align*} \frac{\dd^{2} \phi}{\dd \theta\, \dd p} &= \frac{\dfrac{\dd^{2} u}{\dd \theta\, \dd p} + p \dfrac{\dd^{2} v}{\dd \theta\, \dd p} + \left(\dfrac{\dd v}{\dd \theta}\right)_{p}}{\theta} \\ &= \frac{\dfrac{\dd^{2} u}{\dd \theta\, \dd p} + p \dfrac{\dd^{2} v}{\dd \theta\, \dd p}}{\theta} - \frac{\left(\dfrac{\dd u}{\dd p}\right)_{\theta} + p\left(\dfrac{\dd v}{\dd p}\right)_{\theta}}{\theta^{2}}\Add{,} \end{align*} whence \[ \left(\frac{\dd u}{\dd p}\right)_{\theta} = -\theta \left(\frac{\dd v}{\dd \theta}\right)_{p} - p\left(\frac{\dd v}{\dd p}\right)_{\theta}. \] The differential coefficients of~$\phi$ become, then, by~\Eq{(26)}, \[ \left(\frac{\dd \phi}{\dd \theta}\right)_{p} = \frac{c_{p}}{\theta}; \quad\text{and}\quad \left(\frac{\dd \phi}{\dd p}\right)_{\theta} = -\left(\frac{\dd v}{\dd \theta}\right)_{p}. \] Finally, differentiating the first of these with respect to~$p$, the second with respect to~$\theta$, and equating, we have \[ \left(\frac{\dd c_{p}}{\dd p}\right)_{\theta} = -\theta \left(\frac{\dd^{2} v}{\dd \theta^{2}}\right)_{p}\Add{.} \Tag{(85)} \] This equation contains only quantities that can be directly measured, and establishes a relation between the rate of change of the coefficient of thermal expansion of the substance with temperature (\ie\ the deviation from Gay-Lussac's \PageSep{124} law), and the rate of change of the specific heat with pressure. \Section{158.} By means of the relations furnished by the second \index{Joule and Thomson's absolute temperature!(theory)}% law we may also draw a further conclusion from Thomson and Joule's experiments (\SecRef{70}), in which a gas was slowly pressed through a tube plugged with cotton wool. The interpretation in~\SecRef{70} was confined to their bearing on the properties of perfect gases. It has been mentioned that the characteristic feature of these experiments consists in giving to a gas---without adding or withdrawing heat\footnotemark---an increase \footnotetext{Whether this condition is actually fulfilled may be ascertained by measurements in the medium surrounding the tube through which the gas flows.}% of volume, $V_{2} - V_{1}$, or $v_{2} - v_{1}$ per unit mass, while the external work done per unit mass is represented by \[ p_{1}v_{1} - p_{2}v_{2} = W. \] This expression vanishes in the case of perfect gases, since then the temperature remains constant. In the case of actual gases we may put \begin{alignat*}{3} p_{1} &= p,\quad & p_{2} &= p + \Delta p\quad && (\Delta p < 0) \\ v_{1} &= v, & v_{2} &= v + \Delta v && (\Delta v > 0) \end{alignat*} whence \[ W = -\Delta(pv), \] and by the first law, since $Q = 0$, \[ \Delta u = W + Q = -\Delta(pv). \] For the sake of simplicity we shall assume $\Delta p$~and $\Delta v$ to be small, and we may then write the above equation: \[ \left(\frac{\dd u}{\dd \theta}\right)_{v} \Delta \theta + \left(\frac{\dd u}{\dd v}\right)_{\theta} \Delta v = -v\, \Delta p - p\, \Delta v \] or, by \Eq{(24)},~\Eq{(82)}, and~\Eq{(80)}, \[ \biggl\{c_{p} - \theta \left(\frac{\dd p}{\dd \theta}\right)_{v} \left(\frac{\dd v}{\dd \theta}\right)_{p}\biggr\} \Delta \theta + \theta \left(\frac{\dd p}{\dd \theta}\right)_{v} \Delta v = -v\, \Delta p, \] \PageSep{125} and, by~\Eq{(6)} \begin{align*} c_{p}\, \Delta \theta &= -v\, \Delta p + \theta \left(\frac{\dd v}{\dd \theta}\right)_{p} · \biggl\{\left(\frac{\dd p}{\dd \theta}\right)_{v} \Delta \theta + \left(\frac{\dd p}{\dd v}\right)_{\theta} \Delta v\biggr\} \\ &= -v\, \Delta p + \theta \left(\frac{\dd v}{\dd \theta}\right)_{p} \Delta p\Add{,} \displaybreak[0] \\ \therefore \Delta \theta &= \frac{\theta \left(\dfrac{\dd v}{\dd \theta}\right)_{p} - v}{c_{p}}\, \Delta p\Add{.} \Tag{(86)} \end{align*} By means of this simple equation, the change of temperature ($\Delta \theta$) of the gas in Thomson and Joule's experiments, for a difference of pressure~$\Delta p$, may be found from its specific heat,~$c_{p}$, and its deviation from Gay-Lussac's law. If, under constant pressure, $v$~were proportional to~$\theta$, as in Gay-Lussac's law, then, by equation~\Eq{(86)}, $\Delta \theta = 0$, as is really the case for perfect gases. \Section{159.} Thomson and Joule embraced the results of their observations in the formula \[ \Delta \theta = \frac{\alpha}{\theta^{2}}\, \Delta p, \] where $\alpha$~is a constant. If we express~$p$ in atmospheres, we have, for air, \[ \alpha = 0.276 × (273)^{2}. \] No doubt the formula is only approximate. Within the region of its validity we get, from~\Typo{86}{\Eq{(86)}}, \[ \theta \left(\frac{\dd v}{\dd \theta}\right)_{p} - v = c_{p} \frac{\alpha}{\theta^{2}} \Tag{(87)} \] and, differentiating with respect to~$\theta$, \[ \theta \left(\frac{\dd^{2} v}{\dd \theta^{2}}\right)_{p} = \frac{\alpha}{\theta^{2}} \left(\frac{\dd c_{p}}{\dd \theta}\right)_{p} - \frac{2\alpha c_{p}}{\theta^{3}} \] whence, by the relation~\Eq{(85)}, \[ \left(\frac{\dd c_{p}}{\dd p}\right)_{\theta} + \frac{\alpha}{\theta^{2}} \left(\frac{\dd c_{p}}{\dd \theta}\right)_{p} - \frac{2\alpha c_{p}}{\theta^{3}} = 0. \] \PageSep{126} The general solution of this differential equation is \index{Equation!deduced from Thomson and Joule's experiments}% \[ c_{p} = \theta^{2} · f(\theta^{3} - 3\alpha p), \] where $f$~denotes an arbitrary function of its argument, $\theta^{3} - 3\alpha p$. If we now assume that, for small values of~$p$, the gas, at any temperature, approaches indefinitely near the ideal state, then, when $p = 0$, $c_{p}$~becomes a constant $= c_{p}^{(0)}$ (for air, $c_{p}^{(0)} = 0.238~\Unit{ calorie}$). Hence, generally, \[ c_{p} = c_{p}^{(0)} \theta^{2} (\theta^{3} - 3\alpha p)^{-\efrac{2}{3}}\Add{,} \] or \[ c_{p} = \frac{c_{p}^{(0)}}{\left(1 - \dfrac{3\alpha p}{\theta^{3}}\right)^{\efrac{2}{3}}}\Add{.} \Tag{(88)} \] This expression for~$c_{p}$ will serve further to determine~$v$ in terms of $\theta$~and~$p$. It follows from~\Eq{(87)} that \[ \theta^{2}\, \frac{\dd}{\dd \theta} \left(\frac{v}{\theta}\right)_{p} = \frac{\alpha c_{p}}{\theta^{2}} = \frac{\alpha c_{p}^{(0)}}{(\theta^{3} - 3\alpha p)^{\efrac{2}{3}}}, \] whence \[ \frac{v}{\theta} = \alpha c_{p}^{(0)} \Bigint \frac{d\theta}{\theta^{4} \left(1 - \dfrac{3\alpha p}{\theta^{3}}\right)^{\efrac{2}{3}}}, \] or \[ v = \frac{c_{p}^{(0)} \theta}{3p} \left(\sqrt[3]{1 - \frac{3\alpha p}{\theta^{3}}} + \beta\right). \Tag{(89)} \] This is the characteristic equation of the gas, and $\beta$, the \index{Characteristic equation!deduced from Thomson and Joule's experiments}% constant of integration, may be determined from its density at $0°$~C. and atmospheric pressure. Equations \Eq{(88)}~and \Eq{(89)}, like Thomson and Joule's formula, are valid only within certain limits. It is, however, of theoretical interest to see how the different relations necessarily follow from one another. \Section{160.} A further, theoretically important application of the second law is the determination of the absolute temperature of a substance by a method independent of the \PageSep{127} \index{Joule and Thomson's absolute temperature|(}% deviations of actual gases from the ideal state. In \SecRef{4} we defined temperature by means of the gas thermometer, but had to confine that definition to the cases in which the readings of the different gas thermometers (hydrogen, air, etc.)\ agree as nearly as the desired accuracy of the result requires. For all other cases (including mean temperatures, when a high degree of accuracy is desired) we postponed the definition of absolute temperature. Equation~\Eq{(80)} \index{Absolute temperature!deduced from Thomson and Joule's experiments|(}% enables us to give an exact definition of absolute temperature, entirely independent of the behaviour of special substances. Given the temperature readings,~$t$, of any arbitrary thermometer (mercury-thermometer, or the scale deflection of a thermo-element, or of a bolometer), our problem is to reduce the thermometer to an absolute one, or to express the absolute temperature~$\theta$ as a function of~$t$. We may by direct measurement find how the behaviour of some appropriate substance, \eg\ a gas, depends on~$t$ and either $v$~or~$p$. Introducing, then, $t$~and $v$ as the independent variables in~\Eq{(80)} instead of $\theta$~and~$v$, we obtain \[ \left(\frac{\dd u}{\dd v}\right)_{t} = \theta \left(\frac{\dd p}{\dd t}\right)_{v} · \frac{dt}{d\theta} - p, \] where $\left(\dfrac{\dd u}{\dd v}\right)_{t}$,~$p$, and $\left(\dfrac{\dd p}{\dd t}\right)_{v}$ represent functions of $t$~and~$v$, which can be experimentally determined. The equation can then be integrated thus: \[ \int \frac{d\theta}{\theta} = \Bigint \frac{\left(\dfrac{\dd p}{\dd t}\right)_{v} dt}{\left(\dfrac{\dd u}{\dd v}\right)_{t} + p}. \] If we further stipulate that at the freezing-point of water, where $t = t_{0}$, $\theta = \theta_{0} = 273$, then, \[ \log \frac{\theta}{\theta_{0}} = \Bigintlimits{t_{0}}{t} \frac{\left(\dfrac{\dd p}{\dd t}\right)_{v} dt}{\left(\dfrac{\dd u}{\dd v}\right)_{t} + p}. \] \PageSep{128} This completely determines $\theta$~as a function of~$t$. It is evident that the volume,~$v$, no longer enters into the expression under the sign of integration. \Section{161.} The numerator of this expression may be found directly from the characteristic equation of the substance. The denominator, however, depends on the amount of heat which the substance absorbs during isothermal reversible expansion. For, by~\Eq{(22)} of the first law, the ratio of the heat absorbed during isothermal reversible expansion to the change of volume is \[ \left(\frac{q}{dv}\right)_{t} = \left(\frac{\dd u}{dv}\right)_{t} + p. \] \Section{162.} Instead of measuring the quantity of heat absorbed during isothermal expansion, it may be more convenient, for the determination of the absolute temperature, to experiment on the changes of temperature of a slowly escaping gas, according to the method of Thomson and Joule. If we introduce~$t$ (of~\SecRef{160}) instead of~$\theta$ into equation~\Eq{(86)}, which represents the theory of those experiments on the absolute temperature scale, we have \begin{align*} \Delta \theta &= \frac{d\theta}{dt}\, \Delta t\Add{,} \\ \left(\frac{\dd v}{\dd \theta}\right)_{p} &= \left(\frac{\dd v}{\dd t}\right)_{p} · \frac{dt}{d\theta}, \\ c_{p} &= \left(\frac{q}{d\theta}\right)_{p} = \left(\frac{q}{dt}\right)_{p} · \frac{dt}{d\theta} = c_{p}'\, \frac{dt}{d\theta}, \end{align*} where $c_{p}'$~is the specific heat at constant pressure, determined by a $t$~thermometer. Consequently, by~\Eq{(86)}, \[ \Delta t = \frac{\theta \left(\dfrac{\dd v}{\dd t}\right)_{p} · \dfrac{dt}{d\theta} - v}{c_{p}'}\, \Delta p \] and again, by integration, \[ \log \frac{\theta}{\theta_{0}} = \Bigintlimits{t_{0}}{t} \frac{\left(\dfrac{\delta v}{\delta t}\right)_{p} dt}{v + c_{p}'\, \dfrac{\Delta t}{\Delta p}} = J. \Tag{(90)} \] \PageSep{129} The expression to be integrated again contains quantities which may be measured directly with comparative ease. \Section{163.} The stipulation of \SecRef{160}, that, at the freezing-point of water, $\theta = \theta_{0}= 273$, implies the knowledge of the coefficient of expansion,~$\alpha$, of perfect gases. Strictly speaking, however, all gases show at all temperatures deviations from the behaviour of perfect gases, and disagree with one another. To rid ourselves of any definite assumption about~$\alpha$, we return to our original definition of temperature, viz.\ that the difference between the absolute temperature of water boiling under atmospheric pressure~($\theta_{1}$), and that of water freezing under the same pressure~($\theta_{0}$), shall be \[ \theta_{1} - \theta_{0} = 100\Add{.} \Tag{(91)} \] Now, if $t_{1}$~be the boiling-point of water, measured by means of a $t$~thermometer, then, by~\Eq{(90)}, \[ \log \frac{\theta_{1}}{\theta_{0}} = \Bigintlimits{t_{0}}{t_{1}} \frac{\left(\dfrac{\dd v}{\dd t}\right)_{p} dt}{v + c_{p}'\, \dfrac{\Delta t}{\Delta p}} = J_{1}\Add{,} \Tag{(92)} \] and, eliminating $\theta_{0}$~and~$\theta_{1}$ from \Eq{(90)},~\Eq{(91)}, and~\Eq{(92)}, we find the absolute temperature: \[ \theta = \frac{100 e^{J}}{e^{J_{1}} - 1}. \Tag{(93)} \] From this we obtain the coefficient of thermal expansion of a perfect gas, independently of any gas thermometer, \[ \alpha = \frac{1}{\theta_{0}} = \frac{e^{J_{1}} - 1}{100}. \Tag{(94)} \] Since, in both $J$~and $J_{1}$, the expression to be integrated depends necessarily on $t$~only, it is sufficient for the calculation of the value of the integral to experiment at different temperatures under some simplifying condition, as, for instance, always at the same pressure (atmospheric pressure). \Section{164.} The formula may be still further simplified by \PageSep{130} using as thermometric substance in the $t$~thermometer the same gas as that on which Thomson and Joule's experiments are being performed. The coefficient of expansion,~$\alpha'$, referred to temperature~$t$, is then a constant, and if, as is usual, we put $t_{0} = 0$, and $t_{1} = 100$, \[ v = v_{0} (1 + \alpha' t), \] $v_{0}$~being the specific volume at the melting-point of ice under atmospheric pressure. Also \[ \left(\frac{\dd v}{\dd t}\right)_{p} = \alpha' v_{0}\Typo{,}{.} \] Hence, by~\Eq{(90)}, \[ J = \Bigintlimits{0}{t} \frac{\alpha'\, dt}{1 + \alpha' t + \dfrac{c_{p}'}{v_{0}} · \dfrac{\Delta t}{\Delta p}}, \] and, by~\Eq{(92)}, \[ J_{1} = \Bigintlimits{0}{\PadTo[l]{t}{100}} \frac{\alpha'\, dt}{1 + \alpha' t + \dfrac{c_{p}'}{v_{0}} · \dfrac{\Delta t}{\Delta p}}. \] In the case of an almost perfect gas (\eg\ air), $\Delta t$~is small, and the term $\dfrac{c_{p}'}{v_{0}} · \dfrac{\Delta t}{\Delta p}$ acts merely as a correction term, and, therefore, no great degree of accuracy is required in the determination of $c_{p}'$~and~$v_{0}$. For a perfect gas we should have $\Delta t = 0$, and, from the last two equations, \[ J = \log (1 + \alpha' t),\quad J_{1} = \log (1 + 100 \alpha'); \] therefore, by~\Eq{(93)}, \[ \theta = t + \frac{1}{\alpha'}, \] and, by~\Eq{(94)}, \[ \alpha = \frac{1}{\theta_{0}} = \alpha', \] as it should be. \PageSep{131} As soon as accurate measurement of even a single substance has determined $\theta$~as a function of~$t$, the question regarding the value of the absolute temperature may be considered as solved for all cases. The absolute temperature may be determined not only \index{Absolute temperature!deduced from Thomson and Joule's experiments|)}% by experiments on homogeneous substances, but also from \index{Homogeneous!system|)}% the theory of heterogeneous substances (\Typo{cf.}{\cf}~\SecRef{177}). \index{Joule and Thomson's absolute temperature|)}% \index{System!homogeneous|)}% \PageSep{132} \Chapter{II.}{System in Different States of Aggregation.} \index{Aggregation, states of}% \index{States of aggregation}% \Section{165.} \First{We} shall discuss in this chapter the equilibrium of a system which may consist of solid, liquid, and gaseous portions. We assume that the state of each of these portions is fully determined by mass, temperature, and volume; or, in other words, that the system is formed of but one independent constituent (\SecRef{198}). For this it is not necessary that any portion of the system should be chemically homogeneous. Indeed, the question with regard to the chemical homogeneity cannot, in general, be completely answered (\SecRef{92}). It is still very uncertain whether the molecules of liquid water are the same as those of ice. In fact, the anomalous properties of water in the neighbourhood of its freezing-point make it probable that even in the liquid state its molecules are of different kinds. The decision of such questions has no bearing on the investigations of this chapter. The system may even consist of a mixture of substances in any proportion; that is, it may be a solution or an alloy. What we assume is only this: that the state of each of its homogeneous portions is quite definite when the temperature~$\theta$ and the specific volume~$v$ are definitely given, and that, if the system consists of different substances, their proportion is the same in all portions of the system. We may now enunciate our problem in the following manner:--- Let us imagine a substance of given total mass,~$M$, enclosed in a receptacle of volume,~$V$, and the energy,~$U$, added to it by heat-conduction. If the system be now isolated and left to itself, $M$,~$V$, and~$U$ will remain constant, while the entropy,~$\Phi$, will increase. We shall now \PageSep{133} investigate the state or states of equilibrium which the system may assume, finding at the same time the conditions of its stability or instability. This investigation may be completely carried through by means of the proposition expressed in equation~\Eq{(77)}, that of all the states that may adiabatically arise from one another, the most stable state of equilibrium is characterized by an absolute maximum of the entropy. The entropy may in general, however, as we shall see, assume several relative maxima, under the given external conditions. Each maximum, which is not the absolute one, will correspond to a more or less unstable equilibrium. The system in a state of this kind (\eg\ as supersaturated vapour) may occasionally, upon appropriate, very slight disturbances, undergo a finite change, and pass into another state of equilibrium, which necessarily corresponds to a greater value of the entropy. \Section{166.} We have now to find, first of all, the states in which the entropy~$\Phi$ becomes a maximum. The most general assumption regarding the state of the system is that it consists of a solid, a liquid, and a gaseous portion. Denoting the masses of these portions by $M_{1}$,~$M_{2}$,~$M_{3}$, but leaving open, for the present, the question as to which particular portion each suffix refers, we have for the entire mass of the system $M_{1} + M_{2} + M_{3} = M$. All the quantities are positive, but some may be zero. Further, since the state under discussion is to be one of equilibrium, each portion of the system, also when taken alone, must be in equilibrium, and therefore of uniform temperature and density. To each of them, therefore, we may apply the propositions which were deduced in the preceding chapter for homogeneous substances. If $v_{1}$,~$v_{2}$,~$v_{3}$, denote the specific volumes, the given volume of the system is \[ M_{1} v_{1} + M_{2} v_{2} + M_{3} v_{3} = V. \] Similarly, the given energy is \[ M_{1} u_{1} + M_{2} u_{2} + M_{3} u_{3} = U, \] where $u_{1}$,~$u_{2}$,~$u_{3}$ denote the specific energies of the portions. \PageSep{134} These three equations represent the given \emph{external} conditions. \Section{167.} For the entropy of the system we have \[ \Phi = M_{1} \phi_{1} + M_{2} \phi_{2} + M_{3} \phi_{3}, \] $\phi_{1}$,~$\phi_{2}$,~$\phi_{3}$ being the specific entropies. For an infinitesimal change of state this equation gives \[ \delta \Phi = \tsum M_{1}\, \delta \phi_{1} + \tsum \phi_{1}\, \delta M_{1}. \] Since, by~\Eq{(61)}, we have, in general, \[ \delta \phi = \frac{\delta u + p\, \delta v}{\theta}. \] we obtain \[ \delta \Phi = \tsum \frac{M_{1}\, \delta u_{1}}{\theta_{1}} + \tsum \frac{M_{1}p_{1}\, \delta v_{1}}{\theta_{1}} + \tsum \phi_{1}\, \delta M_{1}\Add{.} \Tag{(95)} \] These variations are not all independent of one another. In fact, from the equations of the imposed (external) conditions, it follows that \[ \left. \begin{aligned} \tsum \delta M_{1} &= 0\Add{,} \\ \tsum \delta M_{1}\, \delta v_{1} + \tsum v_{1}\, \delta M_{1} &= 0\Add{,} \\ \tsum \delta M_{1}\, \delta u_{1} + \tsum u_{1}\, \delta M_{1} &= 0\Add{.} \\ \end{aligned} \right\} \Tag{(96)} \] With the help of these equations we must eliminate from~\Eq{(95)} any three variations, in order that it may contain only independent variations. If we substitute in~\Eq{(95)}, for instance, the values for $\delta M_{2}$,~$\delta v_{2}$, and~$\delta u_{2}$ taken from~\Eq{(96)}, the equation for~$\delta \Phi$ becomes \[ \left. \begin{aligned} \delta \Phi &= \left(\frac{1}{\theta_{1}} - \frac{1}{\theta_{2}}\right) M_{1}\, \delta u_{1} - \left(\frac{1}{\theta_{2}} - \frac{1}{\theta_{3}}\right) M_{3}\, \delta u_{3} \\ &+ \left(\frac{p_{1}}{\theta_{1}} - \frac{p_{2}}{\theta_{2}}\right) M_{1}\, \delta v_{1} - \left(\frac{p_{2}}{\theta_{2}} - \frac{p_{3}}{\theta_{3}}\right) M_{3}\, \delta v_{3} \\ &+ \left(\phi_{1} - \phi_{2} - \frac{u_{1} - u_{2}}{\theta_{2}} - \frac{p_{2}(v_{1} - v_{2})}{\theta_{2}}\right) \delta M_{1} \\ &- \left(\phi_{2} - \phi_{3} - \frac{u_{2} - u_{3}}{\theta_{2}} - \frac{p_{2}(v_{2} - v_{3})}{\theta_{2}}\right) \delta M_{3}\Add{.} \\ \end{aligned} \right\} \Tag{(97)} \] \PageSep{135} Since the six variations occurring in this expression are now independent of one another, it is necessary that each of their six coefficients should vanish, in order that $\delta \Phi$~may be zero for all changes of state. Therefore \[ \left. \begin{aligned} \theta_{1} &= \theta_{2} = \theta_{3} (= \theta)\Add{,} \\ p_{1} &= p_{2} = p_{3}\Add{,} \\ \phi_{1} - \phi_{2} &= \frac{(u_{1} - u_{2}) + p_{1}(v_{1} - v_{2})}{\theta}\Add{,} \\ \phi_{2} - \phi_{3} &= \frac{(u_{2} - u_{3}) + p_{2}(v_{2} - v_{3})}{\theta}\Add{.} \\ \end{aligned} \right\} \Tag{(98)} \] These six equations represent necessary properties of any state, which corresponds to a maximum value of the entropy, \ie\ of any state of equilibrium. As the first four refer to equality of temperature and pressure, the main interest centres in the last two, which contain the thermodynamical theory of fusion, evaporation, and sublimation. \index{Evaporation!theory of}% \Section{168.} These two equations may be considerably simplified \index{Fusion, curve!theory of}% \index{Sublimation, curve!theory of}% \index{Thermodynamical theory!of fusion, vaporization, and sublimation}% \index{Vaporization curve!theory of}% by substituting the value of the specific entropy~$\phi$, which, as well as $u$~and~$p$, is here considered as a function of $\theta$~and~$v$. For, since \Eq{(61)}~gives, in general, \[ d\phi = \frac{du + p\, dv}{\theta}, \] we get, by integration, \[ \phi_{1} - \phi_{2} = \int_{2}^{1} \frac{du + p\, dv}{\theta}, \] where the upper limit of the integral is characterized by the values $\theta_{1}$,~$v_{1}$, the lower by $\theta_{2}$,~$v_{2}$. The path of integration is arbitrary, and does not influence the value of $\phi_{1} - \phi_{2}$. Since, now, $\theta_{1} = \theta_{2} = \theta$ (by~\Typo{98}{\Eq{(98)}}), we may select an isothermal path of integration ($\theta = \const$). This gives \[ \phi_{1} - \phi_{2} = \frac{u_{1} - u_{2}}{\theta} + \frac{1}{\theta} \int_{v_{2}}^{v_{1}} p\, dv. \] \PageSep{136} The integration is to be taken along an isotherm, since $p$~is a known function of $\theta$~and $v$ determined by the characteristic equation of the substance. Substituting the value of $\phi_{1} - \phi_{2}$ in the equations~\Eq{(98)}, we have the relations: \[ \left. \begin{aligned} \int_{v_{2}}^{v_{1}} p\, dv &= p_{1}(v_{1} - v_{2})\Add{,} \\ \int_{v_{3}}^{v_{2}} p\, dv &= p_{2}(v_{2} - v_{3})\Add{,} \\ \llap{\text{to which we add}\qquad\qquad} p_{1} &= p_{2} = p_{3}\Add{.} \end{aligned} \right\} \Tag{(99)} \] With the four unknowns $\theta$, $v_{1}$, $v_{2}$,~$v_{3}$, we have four equations which the state of equilibrium must satisfy. The constants \index{Equilibrium!conditions of}% which occur in these equations depend obviously only on the chemical nature of the substance, and in no way on the given values of the mass,~$M$, the volume,~$V$, and the energy,~$U$, of the system. The equations~\Eq{(99)} might therefore be called the system's \emph{internal} or \emph{intrinsic} conditions of equilibrium, while those of~\SecRef{166} represent the \emph{external} conditions imposed on the system. \Section{169.} Before discussing the values which the equations~\Eq{(99)} \index{Condition of complete reversibility!of equilibrium}% \index{External conditions of equilibrium}% \index{Internal conditions of equilibrium}% give to the unknowns, we shall investigate generally whether, and under what condition, they lead to a maximum value of the entropy and not to a minimum value. It is necessary, for this purpose, to find the value of~$\delta^{2} \Phi$. If this be negative for all virtual changes, then the state considered is certainly one of maximum entropy. From the expression for~$\delta \Phi$ \Eq{(97)} we obtain~$\delta^{2} \Phi$, which may be greatly simplified with the help of the equations~\Eq{(98)}. The equations of the imposed external conditions, and the equations~\Eq{(96)} further simplify the result, and we obtain, finally, \[ \delta^{2} \Phi = -\tsum \frac{M_{1}\, \delta \phi_{1}\, \delta \theta_{1}}{\theta_{1}} + \tsum \frac{M_{1}\, \delta p_{1}\, \delta v_{1}}{\theta_{1}}. \] This may be written \[ \theta\, \delta^{2} \Phi = -\tsum M_{1} (\delta \phi_{1}\, \delta \theta_{1} - \delta p_{1}\, \delta v_{1})\Add{.} \] \PageSep{137} To reduce all variations to those of the independent variables, $\theta$~and~$v$, we may write, according to~\Eq{(81)}, \begin{align*} \delta \phi &= \left(\frac{\dd \phi}{\dd \theta}\right)_{v} \delta \theta + \left(\frac{\dd \phi}{\dd v}\right)_{\theta} \delta v = \frac{c_{v}}{\theta}\, \delta \theta + \left(\frac{\dd p}{\dd \theta}\right)_{v} \delta v \\ \intertext{and} \delta p &= \left(\frac{\dd p}{\dd \theta}\right)_{v} \delta \theta + \left(\frac{\dd p}{\dd v}\right)_{\theta} \delta v, \\ \therefore \theta\, \delta^{2} \Phi &= -\tsum M_{1} \left(\frac{(c_{v})_{1}}{\theta}\, \delta \theta_{1}^{2} - \left(\frac{\dd p_{1}}{\dd v}\right)_{\theta} \delta v_{1}^{2}\right). \Tag{(100)} \end{align*} Obviously, if the quantities $(c_{v})_{1}$, $(c_{v})_{2}$, $(c_{v})_{3}$ be all positive, and the quantities $\left(\dfrac{\dd p_{1}}{\dd v}\right)_{\theta}$\Add{,}~\dots all negative, $\delta^{2} \Phi$~is negative in all cases, and $\Phi$~is really a maximum, and the corresponding state is a state of equilibrium. Since $c_{v}$~is the specific heat at constant volume, and therefore always positive, the condition of equilibrium depends on whether $\left(\dfrac{\dd p}{\dd v}\right)_{\theta}$~is negative for all three portions of the system or not. In the latter case there is no equilibrium. Experience immediately shows, however, that in any state of equilibrium $\dfrac{\dd p}{\dd v}$~is negative, since the pressure, whether positive or negative, and the volume always change in opposite directions. A glance at the graphical representation of~$p$, as an isothermal function of~$v$ (\Fig{1}, \SecRef{26}), shows that there are certain states of the system in which $\dfrac{\dd p}{\dd v}$~is positive. These, however, can never be states of equilibrium, and are, therefore, not accessible to direct observation. If, on the other hand, $\dfrac{\dd p}{\dd v}$~be negative, it is a state of equilibrium, yet it need not be stable; for another state of equilibrium may be found to exist which corresponds to a greater value of the entropy. We shall now discuss the values of the unknowns, $\theta$,~$v_{1}$, $v_{2}$,~$v_{3}$, which represent solutions of the conditions of equilibrium~\Eq{(98)}. Several such systems may be found. Thereafter, we shall deal (beginning at \SecRef{189}) with the further \PageSep{138} question as to which of the different solutions in each case represents the most stable equilibrium under the given external conditions; \ie\ which one leads to the largest value of the entropy of the system. \Section{170.} \Topic{First Solution.}---If we put, in the first place, \[ v_{1} = v_{2} = v_{3} (= v) \] all the equations~\Eq{(98)} are satisfied, for, since the temperature is common to all three portions of the system, their states become absolutely identical. The entire system is, therefore, homogeneous. The state of the system is determined \index{Homogeneous!substance}% by the equations of~\SecRef{166}, which give the imposed conditions. In this case they are \begin{align*} M_{1} + M_{2} + M_{3} &= M\Add{,} \\ v(M_{1} + M_{2} + M_{3}) &= V\Add{,} \\ u(M_{1} + M_{2} + M_{3}) &= U\Add{,} \\ \therefore v = \frac{V}{M}\quad\text{and}\quad u &= \frac{U}{M}. \end{align*} From $v$~and~$u$, $\theta$~may be found, since $u$~was assumed to be a known function of $\theta$~and~$v$. This solution has always a definite meaning; but, as we saw in equation~\Eq{(100)}, it represents a state of equilibrium only when $\dfrac{\dd p}{\dd v}$~is negative. If this be the case, then, the equilibrium is stable or unstable, according as under the external conditions there exists a state of greater entropy or not. This will be discussed later. \Section{171.} \Topic{Second Solution.}---If, in the second case, we put \[ v_{1} \gtrless v_{2},\quad v_{2} = v_{3}, \] the states $2$~and $3$ coincide, and the equations~\Eq{(98)} reduce to \[ \left. \begin{aligned} p_{1} &= p_{2}\Add{,} \\ \phi_{1} - \phi_{2} &= \frac{u_{1} - u_{2} + p_{1}(v_{1} - v_{2})}{\theta}\Add{,} \end{aligned} \right\} \Tag{(101)} \] \PageSep{139} or, instead of the second of these equations, \[ \int_{v_{2}}^{v_{1}} p\, dv = p_{1} (v_{1} - v_{2})\Add{.} \Tag{(102)} \] In this case two states of the system coexist; for instance, the vapour and the liquid. The equations~\Eq{(101)} contain three unknowns, $\theta$,~$v_{1}$,~$v_{2}$; and hence may serve to express $v_{1}$~and $v_{2}$, consequently also the pressure $p_{1} = p_{2}$, and the specific energies $u_{1}$~and~$u_{2}$, as definite functions of the temperature~$\theta$. The internal state of two heterogeneous portions of the same substance in contact with one another is, therefore, completely determined by the temperature. The temperature, as well as the masses of the two portions, may be found from the imposed conditions (\SecRef{166}), which are, in this case, \[ \left. \begin{aligned} M_{1} + (M_{2} + M_{3}) &= M\Add{,} \\ M_{1} v_{1} + (M_{2} + M_{3}) v_{2} &= V\Add{,} \\ M_{1} u_{1} + (M_{2} + M_{3}) u_{2} &= U\Add{.} \\ \end{aligned} \right\} \Tag{(103)} \] These equations serve for the determination of the three last unknowns, $\theta$,~$M_{1}$, and $M_{2} + M_{3}$. This completely determines the physical state, for, in the case of the masses $M_{2}$~and~$M_{3}$, it is obviously sufficient to know their sum. Of course, the result can only bear a physical interpretation if both $M_{1}$~and $M_{2} + M_{3}$ have positive values. \Section{172.} An examination of equation~\Eq{(102)} shows that it can be satisfied only if the pressure,~$p$, which is known to have the same value ($p_{1} = p_{2}$) for both limits of the integral, assume between the limits values which are partly larger and partly smaller than~$p_{1}$. Some of these, then, must correspond to unstable states (\SecRef{169}), since in certain places $p$~and $v$ increase simultaneously $\left(\dfrac{\dd p}{\dd v} > 0\right)$. The equation admits of a simple geometrical interpretation with the help of the above-mentioned graphical representation of the characteristic equation by isotherms (\Fig{1}, \SecRef{26}). For the integral $\displaystyle\int_{2}^{1} p\, dv$ is represented by the area bounded by \PageSep{140} \index{Latent heat}% the isotherm, the axis of abscissæ, and the ordinates at $v_{1}$~and $v_{2}$, while the product $p_{1} (v_{1} - v_{2})$ is the rectangle formed by the same ordinates ($p_{1} = p_{2}$), and the length ($v_{1} - v_{2}$). We learn, therefore, from equation~\Eq{(102)} that in every isotherm the pressure, under which two states of aggregation of the substance may be kept in lasting contact, is represented by the ordinate of the straight line parallel to the axis of abscissæ, which intercepts equal areas on both sides of the isotherm. Such a line is represented by~$ABC$ in \Fig{1}. We are thus enabled to deduce directly from the characteristic equation for homogeneous, stable and unstable, states the functional relation between the pressure, the density of the saturated vapour and of the liquid in contact with it, and the temperature. Taking Clausius' equation~\Eq{(12)} as an empirical expression \index{Clausius' equation}% \index{Equation!Clausius'}% of the facts, we have, for the specific volume~$v_{1}$ of the saturated vapour, and $v_{2}$~of the liquid in contact with it, the two conditions \[ \frac{R\theta}{v_{1} - a} - \frac{c}{\theta (v_{1} + b)^{2}} = \frac{R\theta}{v_{2} - a} - \frac{c}{\theta (v_{2} + b)^{2}}, \] and, from~\Eq{(102)}, % [** TN: Re-breaking] \begin{multline*} R\theta \log \frac{v_{1} - a}{v_{2} - a} - \frac{c}{\theta} \left(\frac{1}{v_{2} + b} - \frac{1}{v_{1} + b}\right) \\ = (v_{1} - v_{2}) \left(\frac{R\theta}{v_{1} - a} - \frac{c}{\theta (v_{1} + b)^{2}}\right). \end{multline*} By means of these $v_{1}$,~$v_{2}$ and $p_{1} = p_{2}$ may be expressed as functions of~$\theta$, or, still more conveniently, $v_{1}$,~$v_{2}$, $p_{1}$, and~$\theta$ as functions of some appropriately selected independent variable. With Clausius' values of the constants for carbon dioxide (\SecRef{25}), this calculation furnishes results which show a satisfactory agreement with Andrews' observations. According \index{Andrews}% to Thiesen, however, Clausius' equation is by no means the \index{Thiesen}% general form of the characteristic equation. \Section{173.} We shall now follow the interpretation of the equation~\Eq{(101)} in other directions. If we put, for shortness, \[ u - \theta \phi = f \Tag{(104)} \] %[** TN: Next line set is small type in the original, as if squeezed in] (free energy per unit mass, by equation~\Eq{(71)}), \PageSep{141} the equations~\Eq{(101)} become, simply, \begin{align*} p_{1} &= p_{2}\Add{,} \Tag{(105)} \\ f_{2} - f_{1} &= p_{1} (v_{1} - v_{2})\Add{.} \Tag{(106)} \end{align*} The function~$f$ satisfies the following simple conditions. By~\Eq{(104)}, \begin{align*} \left(\frac{\dd f}{\dd \theta}\right)_{v} &= \left(\frac{\dd u}{\dd \theta}\right)_{v} - \theta \left(\frac{\dd \phi}{\dd \theta}\right)_{v} - \phi. \intertext{By~\Eq{(79a)},} \left(\frac{\dd f}{\dd \theta}\right)_{v} &= -\phi. \Tag{(107)} \intertext{Also, by~\Eq{(104)},} \left(\frac{\dd f}{\dd v}\right)_{\theta} &= \left(\frac{\dd u}{\dd v}\right)_{\theta} - \theta \left(\frac{\dd \phi}{\dd v}\right)_{\theta} \intertext{and, by \Eq{(80)} and \Eq{(81)},} \left(\frac{\dd f}{\dd v}\right)_{\theta} &= -p. \Tag{(108)} \end{align*} The conditions of equilibrium for two states of aggregation in mutual contact hold for the three possible combinations of the solid and liquid, liquid and gaseous, gaseous and solid states. In order to fix our ideas, however, we shall discuss that solution of those equations which corresponds to the contact of a liquid with its vapour. Denoting the vapour by the subscript~$1$, the liquid by~$2$, $v_{1}$~is then the specific volume of the saturated vapour at the temperature~$\theta$; $p_{1} = p_{2}$, its pressure; $v_{2}$~the specific volume of the liquid with which it is in contact. All these quantities, then, are functions of the temperature only, which agrees with experience. \Section{174.} Further theorems may be arrived at by the differentiation of the conditions of equilibrium with respect to~$\theta$. Since all variables now depend only on~$\theta$, we shall use~$\dfrac{d}{d\theta}$ to indicate this total differentiation, while partial differentiation with regard to~$\theta$ at constant~$v$ will be expressed, as hitherto, by~$\dfrac{\dd}{\dd \theta}$. \PageSep{142} Equations \Eq{(105)} and \Eq{(106)}, thus differentiated, give \[ \frac{dp_{1}}{d\theta} = \frac{dp_{2}}{d\theta} \] and \[ \frac{df_{2}}{d\theta} - \frac{df_{1}}{d\theta} = (v_{1} - v_{2})\, \frac{dp_{1}}{d\theta} + p_{1}\left(\frac{dv_{1}}{d\theta} - \frac{dv_{2}}{d\theta}\right). \] But, by \Eq{(107)} and \Eq{(108)}, we have \iffalse \begin{align*} \frac{df_{2}}{d\theta} - \frac{df_{1}}{d\theta} &\begin{aligned}[t] &= \left(\frac{\dd f_{2}}{\dd \theta}\right)_{v} + \left(\frac{\dd f_{2}}{\dd v}\right)_{\theta} · \frac{dv_{2}}{d\theta} \\ &- \left(\frac{\dd f_{1}}{\dd \theta}\right)_{v} - \left(\frac{\dd f_{1}}{\dd v}\right)_{\theta} · \frac{dv_{1}}{d\theta} \end{aligned} \\ &= -\phi_{2} - p_{2}\, \frac{dv_{2}}{d\theta} + \phi_{1} + p_{1}\, \frac{dv_{1}}{d\theta}, \end{align*} \fi \begin{align*} \frac{df_{2}}{d\theta} - \frac{df_{1}}{d\theta} &= \left(\frac{\dd f_{2}}{\dd \theta}\right)_{v} \!\!+ \left(\frac{\dd f_{2}}{\dd v}\right)_{\theta} · \frac{dv_{2}}{d\theta} - \left(\frac{\dd f_{1}}{\dd \theta}\right)_{v} \!\!- \left(\frac{\dd f_{1}}{\dd v}\right)_{\theta} · \frac{dv_{1}}{d\theta} \\ &= -\phi_{2} - p_{2}\, \frac{dv_{2}}{d\theta} + \phi_{1} + p_{1}\, \frac{dv_{1}}{d\theta}, \end{align*} whence, by substitution, \[ \phi_{1} - \phi_{2} = (v_{1} - v_{2})\, \frac{dp_{1}}{d\theta}, \] or, finally, by~\Eq{(101)}, \[ (u_{1} - u_{2}) + p_{1} (v_{1} - v_{2}) = \theta (v_{1} - v_{2})\, \frac{dp_{1}}{d\theta}. \Tag{(109)} \] Here the left-hand side of the equation, according to equation~\Eq{(17)} of the first law, represents the heat of vaporization,~$L$, of the liquid. It is the heat which must be added to unit mass of the liquid, in order to completely change it to vapour under the constant pressure of its saturated vapour. For the corresponding change of energy is $u_{1} - u_{2}$, and the external work performed, here negative, amounts to \begin{align*} W &= -p_{1}(v_{1} - v_{2}) \\ \therefore L &= u_{1} - u_{2} + p_{1}(v_{1} - v_{2}), \Tag{(110)} \\ \intertext{whence} L &= \theta (v_{1} - v_{2}). \Tag{(111)} \end{align*} This equation, deduced by Clapeyron from Carnot's theory, \index{Clapeyron}% but first rigorously proved by Clausius, may be used for the determination of the heat of vaporization at any temperature, if we know the specific volumes of the saturated vapour and the liquid, as well as the relation between the pressure of the saturated vapour and the \PageSep{143} temperature. This formula has been verified by experiment in a large number of cases. \Section{175.} As an example, we shall calculate the heat of \index{Heat, latent!approximation formula}% vaporization of water at $100°$~C., \ie\ under atmospheric pressure, from the following data:--- \settowidth{\TmpLen}{vapour at $100°$ in $\Unit{cm.}^{3}$, according to Wüllner.}% \begin{align*} \theta &= 273 + 100 = 373. \\ v_{1} &= 1658\quad\parbox[t]{\TmpLen}{(volume of $1~\Unit{gr.}$ of saturated water \\ vapour at $100°$ in $\Chg{\Unit{c.c.}}{\Unit{cm.}^{3}}$, according to Wüllner).} \\ v_{2} &= 1\quad\text{(volume of $1~\Unit{gr.}$ of water at $100°$ in $\Chg{\Unit{c.c.}}{\Unit{cm.}^{3}}$).} \end{align*} $\dfrac{dp_{1}}{d\theta}$~is found from the experiments of Regnault. Saturated \index{Regnault}% water vapour at $100°$~C. gave an increase of pressure of $27.2~\Unit{mm.}$ of mercury for a rise of $1°$~C\@. In absolute units, by~\SecRef{7}, \[ \frac{dp_{1}}{d\theta} = \frac{ 27.2}{760} × 1013650. \] Thus, the required latent heat of vaporization is \index{Latent heat}% \[ L = \frac{373 × 1657 × 27.2 × 1013650}{760 × 419 × 10^{5}} = 535~\Unit{cal.} \] By direct observation Regnault found the heat of vaporization of water at $100°$~C. to be~$536$. \Section{176.} As equation~\Eq{(110)} shows, part of the heat of vaporization,~$L$, corresponds to an increase of energy, and part to external work. To find the relation between these two it is most convenient to find the ratio of the external work to the latent heat of vaporization, viz.\ \[ \frac{p_{1} (v_{1} - v_{2})}{L} = \frac{p_{1}}{\theta\, \dfrac{dp_{1}}{d\theta}}. \] In the above case $p = 760~\Unit{mm.}$, $\theta = 373$, $\dfrac{dp}{d\theta} = 27.2~\Unit{mm.}$, and therefore, \[ \frac{p_{1} (v_{1} - v_{2})}{L} = \frac{760}{373 × 27.2} = 0.075. \] \PageSep{144} This shows that the external work forms only a small part of the value of the latent heat of vaporization. \Section{177.} Equation~\Eq{(111)} also leads to a method of calculating the absolute temperature~$\theta$, when the latent heat of vaporization,~$L$, as well as the pressure and the density of the saturated vapour and the liquid, have been determined by experiment in terms of any scale of temperature~$t$ (\SecRef{160}). We have \begin{align*} L &= \theta (v_{1} - v_{2})\, \frac{dp_{1}}{dt} · \frac{dt}{d\theta}, \\ \therefore \log \theta &= \int \frac{v_{1} - v_{2}}{L} · \frac{dp_{1}}{dt}\, dt, \end{align*} and therefore $\theta$~may be determined as a function of~$t$. It is obvious that any equation between measurable quantities, deduced from the second law, may be utilized for a determination of the absolute temperature. The question as to which of those methods deserves preference is to be decided by the degree of accuracy to be obtained in the actual measurements. \Section{178.} A simple approximation formula, which in many cases gives good, though in some, only fair results, may be obtained by neglecting in the equation~\Eq{(111)} the specific volume of the liquid,~$v_{2}$, in comparison with that of the vapour,~$v_{1}$, and assuming for the vapour the characteristic equation of a perfect gas. Then, by~\Eq{(14)}, \[ v_{1} = \frac{R\theta}{mp_{1}}, \] where $R$~is the absolute gas constant, and $m$~the molecular weight of the vapour. Equation~\Eq{(111)} then becomes \[ L = \frac{R}{m} · \frac{\theta^{2}}{p_{1}} · \frac{dp_{1}}{d\theta}. \Tag{(112)} \] For water at $100°$~C. we have $R = 1.971$; $m = \ce{H2O} = 18$; \PageSep{145} $\theta = 373$; $p_{1} = 760~\Unit{mm.}$; $\dfrac{dp_{1}}{d\theta} = 27.2~\Unit{mm}$. Hence the latent heat of vaporization is \[ L = \frac{1.971 × 373^{2} × 27.2}{18 × 760} = 545~\Unit{cal.} \] This value is somewhat large (\SecRef{175}). The cause of this lies in the fact that the volume of saturated water vapour at~$100°$ is in reality smaller than that calculated from the characteristic equation of a perfect gas of molecular weight~$18$. But, for this very reason, accurate measurement of the heat of vaporization may serve as a means of estimating from the second law the amount by which the density of a vapour deviates from the ideal value. Another kind of approximation formula, valid within the same limits, is found by substituting in~\Eq{(109)} the value of the specific energy $u_{1} = c_{v} \theta + \const$, which, by~\Eq{(39)}, holds for perfect gases. We may put the specific energy of the liquid $u_{2} = c_{2} \theta + \const$, if we assume its specific heat,~$c_{2}$, to be constant, and neglect the external work. It then follows from~\Eq{(109)} that \[ (c_{v} - c_{2}) \theta + \const + \frac{R \theta}{m} = \frac{R}{m} · \frac{\theta^{2}}{p_{1}} · \frac{dp_{1}}{d\theta}. \] If we multiply both sides by~$\dfrac{d\theta}{\theta^{2}}$, this equation may be integrated, term by term, and we finally obtain with the help of~\Eq{(33)} \[ p_{1} = ae^{-\efrac{b}{\theta}} · \theta^{\efrac{m}{R} (c_{p} - c_{2})} \] where $a$~and $b$ are positive constants; $c_{p}$~and $c_{2}$ the specific heats of the vapour and the liquid, at constant pressure. This relation between the pressure of the saturated vapour and the temperature is the more approximately true, the further the temperature lies below the critical temperature of the vapour. \PageSep{146} For mercury vapour, for example, according to a calculation \index{Hertz, H.}% by H.~Hertz, if $p_{1}$~be given in~$\Unit{mm.}$ of mercury, \[ a = 3.915 × 10^{10};\quad b = 7695;\quad \frac{m}{R} (c_{p} - c_{2}) = - 0.847. \] \Section{179.} Equation~\Eq{(111)} is applicable to the processes of fusion and sublimation in the same manner as to that of evaporation. In the first case $L$~denotes the latent heat of fusion of the substance, if the subscript~$1$ correspond to the liquid state and $2$~to the solid state, and $p_{1}$~the melting pressure, \ie\ the pressure under which the solid and the liquid substance may be in contact and in equilibrium. The melting pressure, therefore, just as the pressure of evaporation, depends on the temperature only. Conversely, a change of pressure produces a change in the melting point: \[ \frac{d\theta}{dp_{1}} = \frac{\theta (v_{1} - v_{2})}{L}. \Tag{(113)} \] For ice at $0°$~C. and under atmospheric pressure, we have \settowidth{\TmpLen}{(heat of fusion of $1~\Unit{gr.}$}% \begin{align*} %[** TN: Re-breaking text on following line] L &= 80 × 419 × 10^{5}\quad\parbox[t]{\TmpLen}{(heat of fusion of $1~\Unit{gr.}$ \\ of ice in C.G.S. units);} \\ \theta &= 273; \\ v_{1} &= 1.0\quad\text{(vol.\ of $1~\Unit{gr.}$ of water at $0°$~C. in~$\Chg{\Unit{c.c.}}{\Unit{cm.}^{3}}$);} \\ v_{2} &= 1.09\quad\text{(vol.\ of $1~\Unit{gr.}$ of ice at $0°$~C. in~$\Chg{\Unit{c.c.}}{\Unit{cm.}^{3}}$).} \end{align*} To obtain~$\dfrac{d\theta}{dp_{1}}$ in atmospheres we must multiply by $1,013,650$: \[ \frac{d\theta}{dp_{1}} = -\frac{273 × 0.09 × 1013650} {80 × 419 × 10^{5}} = - 0.0074. \Tag{(114)} \] On increasing the external pressure by $1$~atmosphere, the melting point of ice will, therefore, be lowered by $0.0074°$~C.; \index{Melting point of ice}% \index{Melting point of ice!lowering of, by pressure}% or, to lower the melting point of ice by $1°$~C., the pressure must be increased by about $130$~atmospheres. This was first verified by the measurements of W.~Thomson (Lord Kelvin). \index{Thomson}% Equation~\Eq{(113)} shows that, conversely, the melting point of substances, which expand on melting, is raised by an increase \PageSep{147} of pressure. This has been qualitatively and quantitatively verified by experiment. \Section{180.} By means of the equations~\Eq{(101)} still further \index{Second law of thermodynamics!test of}% important properties of substances in different states may be shown to depend on one another. From these, along with~\Eq{(110)}, we obtain \[ \frac{L}{\theta} = \phi_{1} - \phi_{2}. \] Differentiating this with respect to~$\theta$, we have \begin{align*} %[** TN: Re-breaking] \frac{1}{\theta} · \frac{dL}{d\theta} - \frac{L}{\theta^{4}} &= \left(\frac{\dd \phi_{1}}{\dd \theta}\right)_{v} + \left(\frac{\dd \phi_{1}}{\dd v}\right)_{\theta} · \frac{dv_{1}}{d\theta} \\ &- \left(\frac{\dd \phi_{2}}{\dd \theta}\right)_{v} - \left(\frac{\dd \phi_{2}}{\dd v}\right)_{\theta} · \frac{dv_{2}}{d\theta}, \intertext{or, by~\Eq{(81)},} &= \frac{(c_{v})_{1}}{\theta} + \left(\frac{\dd p_{1}}{\dd \theta}\right)_{v} · \frac{dv_{1}}{d\theta} - \frac{(c_{v})_{2}}{\theta} - \left(\frac{\dd p_{2}}{\dd \theta}\right)_{v} · \frac{dv_{2}}{d\theta}. \end{align*} We now introduce~$c_{p}$, the specific heat at constant pressure, for~$c_{v}$, that at constant volume, and obtain by~\Eq{(82)}, on multiplying by~$\theta$, \begin{align*} \frac{dL}{d\theta} - \frac{L}{\theta} &= (c_{p})_{1} - \theta \left(\frac{\dd p_{1}}{\dd \theta}\right)_{v} · \left(\frac{\dd v_{1}}{\dd \theta}\right)_{p} + \theta \left(\frac{\dd p_{1}}{\dd \theta}\right)_{v} · \frac{dv_{1}}{d\theta} \\ &- (c_{p})_{2} + \theta \left(\frac{\dd p_{2}}{\dd \theta}\right)_{v} · \left(\frac{\dd v_{2}}{\dd \theta}\right)_{p} - \theta \left(\frac{\dd p_{2}}{\dd \theta}\right)_{v} · \frac{dv_{2}}{d\theta}, \end{align*} or, since for both states, according to~\Eq{(6)}, \[ \left(\frac{\dd p}{\dd \theta}\right)_{v} = -\left(\frac{\dd v}{\dd \theta}\right)_{p} · \left(\frac{\dd p}{\dd v}\right)_{\theta}, \] \begin{align*} \frac{dL}{d\theta} - \frac{L}{\theta} &= (c_{p})_{1} - \theta \left(\frac{\dd v_{1}}{\dd \theta}\right)_{p} · \left[\left(\frac{\dd p_{1}}{\dd \theta}\right)_{v} + \left(\frac{\dd p_{1}}{\dd v}\right)_{\theta} · \frac{dv_{1}}{d\theta}\right] \\ &- (c_{p})_{2} + \theta \left(\frac{\dd v_{2}}{\dd \theta}\right)_{p} · \left[\left(\frac{\dd p_{2}}{\dd \theta}\right)_{v} + \left(\frac{\dd p_{2}}{\dd v}\right)_{\theta} · \frac{dv_{2}}{d\theta}\right]. \end{align*} Now, the expressions within the brackets are identical with \[ \frac{dp_{1}}{d\theta} = \frac{dp_{2}}{d\theta} = \frac{L}{\theta (v_{1} - v_{2})}. \] \PageSep{148} \index{Second law of thermodynamics!test of}% This gives, finally, \[ (c_{p})_{1} - (c_{p})_{2} = \frac{dL}{d\theta} - \frac{L}{\theta} + \frac{L}{v_{1} - v_{2}} · \left[\left(\frac{\dd v_{1}}{\dd \theta}\right)_{p} - \left(\frac{\dd v_{2}}{\dd \theta}\right)_{p}\right]. \Tag{(115)} \] This equation, which is rigorously true, again leads to a test of the second law, since all the quantities in it may be observed independently of one another. \Section{181.} We shall again take as example saturated water vapour at $100°$~C. (atmospheric pressure), and calculate its specific heat at constant pressure. We have the following \index{Specific heat!of steam}% data: \settowidth{\TmpLen}{decrease of the heat of vaporization with}% \begin{align*} (c_{p})_{2} &= 1.03\quad\text{($=$ spec.\ heat of liquid water at $100°$~C.);} \\ L &= 536; \\ \theta &= 373; \\ \frac{dL}{d\theta} &= -0.708\quad\parbox[t]{\TmpLen}{(decrease of the heat of vaporization with \\ increase of temperature, according to \\ Regnault's observations).} \index{Regnault}% \end{align*} We determine~$v$, and $\left(\dfrac{\dd v_{1}}{\dd \theta}\right)_{p}$ from the observations of Hirn, who found the volume of $1~\Unit{gr.}$ of steam, at $100°$~under \index{Hirn}% atmospheric pressure, to be $1650.4~\Chg{\Unit{c.c.}}{\Unit{cm.}^{3}}$; and at~$118.5°$, $1740~\Chg{\Unit{c.c.}}{\Unit{cm.}^{3}}$, whence \[ v_{1} = 1650.4;\quad \left(\frac{\dd v_{1}}{\dd \theta}\right)_{p} = \frac{1740 - 1650.4}{18.5} = 4.843. \] Also \[ v_{2} = 1.0;\quad \left(\frac{\dd v_{2}}{\dd \theta}\right)_{p} = 0.001. \] These values, substituted in~\Eq{(115)}, give \[ (c_{p})_{1} - (c_{p})_{2} = -0.56, \] or, \[ (c_{p})_{1} = (c_{p})_{2} - 0.56 = 1.03 - 0.56 = 0.47. \] By direct measurement, Regnault found the mean specific heat of steam under atmospheric pressure for temperatures somewhat higher than $100°$~C. to be~$0.48$. \PageSep{149} \Section{182.} The relation~\Eq{(115)} may be simplified, but is inaccurate, if we neglect the volume~$v_{2}$ of the liquid in comparison with~$\Erratum{v}{v_{1}}$, that of the vapour, and apply to the vapour the characteristic equation of a perfect gas. Then \begin{gather*} v_{1} = \frac{R\theta}{mp_{1}},\displaybreak[0] \\ \left(\frac{\dd v_{1}}{\dd \theta}\right)_{p} = \frac{R}{mp_{1}}, \end{gather*} and equation~\Eq{(115)} becomes, simply, \[ (c_{p})_{1} - (c_{p})_{2} = \frac{dL}{d\theta}. \] In our example, \begin{gather*} (c_{p})_{1} - (c_{p})_{2} = -0.71 \\ (c_{p})_{1} = 1.03 - 0.71 = 0.32, \end{gather*} a value considerably too small. \Section{183.} We shall now apply the relation~\Eq{(115)} to the melting of ice at $0°$~C. and under atmospheric pressure. The subscript~$1$ now refers to the liquid state, and $2$~to the solid state. The relation between the latent heat of fusion of ice and the temperature has probably never been measured. It may, however, be calculated from~\Eq{(115)}, which gives \[ \frac{dL}{d\theta} = (c_{p})_{1} - (c_{p})_{2} + \frac{L}{\theta} - \frac{L}{v_{1} - v_{2}} \left[\left(\frac{\dd v_{1}}{\dd \theta}\right)_{p} - \left(\frac{\dd v_{2}}{\dd \theta}\right)_{p}\right] \] in which \settowidth{\TmpLen}{thermal coeff.\ of expansion\quad}% % [** TN: Re-breaking] \begin{align*} (c_{p})_{1} &= 1\quad\text{(spec.\ heat of water at $0°$~C.);} \\ (c_{p})_{2} &= 0.505\quad\text{(spec.\ heat of ice at $0°$~C.);} \displaybreak[0] \\ L &= 80;\quad \theta = 273;\quad v_{1} = 1;\quad v_{2} = 1.09; \displaybreak[0] \\ \left(\frac{\dd v_{1}}{\dd \theta}\right)_{p} &= -0.00006\quad\parbox[t]{\TmpLen}{(therm.\ coeff.\ of expansion \\ of water at $0°$~C.);} \\ \left(\frac{\dd v_{1}}{\dd \theta}\right)_{p} &= 0.00011\quad\parbox[t]{\TmpLen}{(thermal coeff.\ of expansion \\ of ice at $0°$~C.).} \end{align*} \PageSep{150} Hence, by the above equation, \[ \frac{dL}{d\theta} = 0.64, \] \ie, if the melting point of ice be lowered $1°$~C. by an appropriate increase of the external pressure, its heat of fusion decreases by $0.64~\Unit{cal}$. \Section{184.} It has been repeatedly mentioned in the early chapters, that, besides the specific heat at constant pressure, \index{Specific heat!of saturated vapour}% or constant volume, any number of specific heats may be defined according to the conditions under which the heating takes place. Equation~\Eq{(23)} of the first law holds in each case: \[ c = \frac{du}{d\theta} + p\, \frac{dv}{d\theta}. \] In the case of saturated vapours special interest attaches to the process of heating, which keeps them permanently in a state of saturation. Denoting by~$h_{1}$ the specific heat of the vapour corresponding to this process (Clausius called it the specific heat of ``the saturated vapour''), we have \[ h_{1} = \frac{du_{1}}{d\theta} + p_{1}\, \frac{dv_{1}}{d\theta}. \Tag{(116)} \] No off-hand statement can be made with regard to the value of~$h_{1}$; even its sign must in the mean time remain uncertain. For, if during a rise of temperature of~$1°$ the vapour is to remain just saturated, it must evidently be compressed while being heated, since the specific volume of the saturated vapour decreases as the temperature rises. This compression, however, generates heat, and the question is, whether the latter is so considerable that it must be in part withdrawn by conduction, so as not to superheat the vapour. Two cases may, therefore, arise: (1)~The heat of compression may be considerable, and the withdrawal of heat is necessary to maintain saturation at the higher temperature, \ie\ $h_{1}$~is negative. (2)~The heat of compression may be too slight to prevent the compressed vapour, \PageSep{151} without the addition of heat, from becoming supersaturated. Then, $h_{1}$~has a positive value. Between the two there is a limiting case ($h_{1} = 0$), where the heat of compression is exactly sufficient to maintain saturation. In this case the curve of the saturated vapour coincides with that of adiabatic compression. Watt assumed this to be the case for steam. It is now easy to calculate~$h_{1}$ from the above formulæ. Calling $h_{2}$~the corresponding specific heat of the liquid, we have \[ h_{2} = \frac{du_{2}}{d\theta} + p_{2}\, \frac{dv_{2}}{d\theta}. \Tag{(117)} \] During heating, the liquid is kept constantly under the pressure of its saturated vapour. Since the external pressure, unless it amount to many atmospheres, has no appreciable influence on the state of a liquid, the value of~$h_{2}$ practically coincides with that of the specific heat at constant pressure, \[ h_{2} = (c_{p})_{2}. \Tag{(118)} \] Subtracting \Eq{(117)} from~\Eq{(116)}, we get \[ h_{1} - h_{2} = \frac{d(u_{1} - u_{2})}{d\theta} + p_{1} \frac{d(v_{1} - v_{2})}{d\theta}. \] But \Eq{(110)}, differentiated with respect to~$\theta$, gives \begin{align*} \frac{dL}{d\theta} &= \frac{d(u_{1} - u_{2})}{d\theta} + p_{1} \frac{d(v_{1} - v_{2})}{d\theta} + (v_{1} - v_{2})\, \frac{dp_{1}}{d\theta}, \\ \therefore h_{1} - h_{2} &= \frac{dL}{d\theta} - (v_{1} - v_{2})\, \frac{dp_{1}}{d\theta}, \end{align*} or, by \Eq{(118)} and~\Eq{(111)}, \[ h_{1} = (c_{p})_{2} + \frac{dL}{d\theta} - \frac{L}{\theta}. \] For saturated water vapour at~$100°$, we have, as above, \[ (c_{p})_{2} = 1.03;\quad \frac{dL}{d\theta} = -0.71;\quad L = 536;\quad \theta = 373; \] \PageSep{152} whence \[ h_{1} = 1.03 - 0.71 - \tfrac{536}{373} = -1.12. \] Water vapour at $100°$~C. represents the first of the cases described above, \ie\ saturated water vapour at~$100°$ is superheated by adiabatic compression. Conversely, saturated water vapour at~$100°$ becomes supersaturated by adiabatic expansion. The influence of the heat of compression (or expansion) is greater than the influence of the increase (or decrease) of the density. Some other vapours behave in the opposite way. \Section{185.} It may happen that, for a given value of~$\theta$, the \index{Critical point!temperature}% values of $v_{1}$~and~$v_{2}$, which are fully determined by the equation~\Eq{(101)}, become equal. Then the two states which are in contact with one another are identical. Such a value of~$\theta$ is called a \emph{critical temperature} of the substance. From a \index{Temperature!critical}% purely mathematical point of view, every substance must be supposed to have a critical temperature for each of the three combinations, solid-liquid, liquid-gas, gas-solid. This critical temperature, however, will not always be real. The critical temperature~$\theta$ and the critical volume $v_{1} = v_{2}$, fully determine the critical state. We may calculate it from the equations~\Eq{(101)} by finding the condition that $v_{1} - v_{2}$ should vanish. If we first assume $v_{1} - v_{2}$ to be very small, Taylor's theorem then gives for any volume~$v$, lying between $v_{1}$ and~$v_{2}$, \[ p = p_{2} + \left(\frac{\dd p}{\dd v}\right)_{2} (v - v_{2}) + \tfrac{1}{2} \left(\frac{\dd^{2} p}{\dd v^{2}}\right)_{2} (v - v_{2})^{2}\Add{,} \Tag{(119)} \] and therefore the first equation~\Eq{(101)} becomes % [** TN: Errata (three missing subscripts) applied to displays] \[ p_{2} + \left(\frac{\dd p}{\dd v}\right)_{2} (v_{1} - v_{2}) + \tfrac{1}{2} \left(\frac{\dd^{2} p}{\dd v^{2}}\right)_{2} (v_{1} - v_{2})^{2} = p_{2}, \] and equation~\Eq{(102)}, by the integration of~\Eq{(119)} with respect to~$v$, gives \[ p_{2} (v_{1} - v_{2}) + \tfrac{1}{2} \left(\frac{\dd p}{\dd v}\right)_{2}\!\!\! (v_{1} - v_{2})^{2} + \tfrac{1}{2 · 3} \left(\frac{\dd^{2} p}{\dd v^{2}}\right)_{2}\!\!\! (v_{1} - v_{2})^{3} = p_{2} (v_{1} - v_{2}). \] \PageSep{153} The last two equations give, as the conditions of the critical state, \[ \left(\frac{\dd p}{\dd v}\right)_{2} = 0,\quad\text{and}\quad \left(\frac{\dd^{2} p}{\dd v^{2}}\right)_{2} = 0. \] These conditions agree with those found in~\SecRef{30}. They are there geometrically illustrated by the curve of the critical isotherm. In the critical state the compressibility is infinite; so are also the thermal coefficient of expansion and the specific heat at constant pressure; the heat of vaporization is zero. For all temperatures other than the critical one, the values of $v_{1}$~and~$v_{2}$ are different. On one side of the critical isotherm they have real, on the other imaginary values. In this latter case our solution of the problem of equilibrium no longer admits of a physical interpretation. Several reasons may be given for assuming that not only for evaporation but also for fusion in the case of many substances there exists a real critical temperature at which the solid and liquid states are identical (\SecRef{31} and~\SecRef{191}). \Section{186.} \Topic{Third Solution.}---In the third place, we shall assume that in the conditions of equilibrium~\Eq{(98)} \[ v_{1} \gtrless v_{2} \gtrless v_{3}. \] We have then, without further simplification, \[ \left. \begin{aligned} p_{1} &= p_{2} = p_{3}\Add{,} \\ \phi_{1} - \phi_{2} &= \frac{u_{1} - u_{2} + p_{1}(v_{1} - v_{2})}{\theta}\Add{,} \\ \phi_{2} - \phi_{3} &= \frac{u_{2} - u_{3} + p_{1}(v_{2} - v_{3})}{\theta}\Add{.} \end{aligned} \right\} \Tag{(120)} \] These refer to a state in which the three states of aggregation \index{Aggregation, states of!co-existence of states of}% \index{Coexistence of states of aggregation}% \index{States of aggregation!co-existence of}% are simultaneously present. There are four equations, and these assign definite values to the four unknowns $\theta$,~$v_{1}$, $v_{2}$,~$v_{3}$. The coexistence of the three states of aggregation in equilibrium is, therefore, possible only at a definite temperature, and with definite densities; therefore, also, at a \PageSep{154} \index{Fundamental point (triple)!pressure}% \index{Fundamental point (triple)!temperature}% \index{Fundamental point (triple)!temperature of ice}% definite pressure. We shall call this temperature the \index{Pressure!fundamental}% \index{Temperature!fundamental}% \index{Temperature!fundamental!of ice}% \emph{fundamental temperature}, and the corresponding pressure the \emph{fundamental pressure} of the substance. According to equations~\Eq{(120)}, the fundamental temperature is characterized by the condition that at it the pressure of the saturated vapour is equal to the pressure of fusion. It necessarily follows, by addition of the last two equations, that this pressure is also equal to the pressure of sublimation. After the fundamental temperature and pressure have been found, the external conditions of~\SecRef{166}--- \[ \left. \begin{aligned} M_{1} + (M_{2} + M_{3}) &= M\Add{,} \\ M_{1} v_{1} + M_{2} v_{2} + M_{3} v_{3} &= V\Add{,} \\ M_{1} u_{1} + M_{2} u_{2} + M_{3} u_{3} &= U\Add{,} \\ \end{aligned} \right\} \Tag{(121)} \] uniquely determine the masses of the three portions of the substance. The solution, however, can be interpreted physically only if $M_{1}$,~$M_{2}$, and~$M_{3}$ are positive. \Section{187.} Let us determine, \eg, the fundamental state of water. $0°$~C. is not its fundamental temperature, for at $0°$~C. the maximum vapour pressure of water is $4.62~\Unit{mm.}$, but the melting pressure of ice is $760~\Unit{mm}$. Now, the latter decreases with rise of temperature, while the maximum vapour pressure increases. A coincidence of the two is, therefore, to be expected at a temperature somewhat higher than~$0°$~C\@. According to equation~\Eq{(114)}, the melting point of ice rises by $0.0074°$~C. approximately, when the pressure is lowered from $760~\Unit{mm.}$ to $4.62~\Unit{mm}$. The fundamental temperature of water is, then, approximately, $0.0074°$~C\@. At this temperature the maximum vapour pressure of water nearly coincides with the melting pressure of ice, and, therefore, also with the maximum vapour pressure of ice. The specific volumes of water in the three states are, therefore, \[ v_{1} = 205,000;\quad v_{2} = 1;\quad v_{3} = 1.09. \] For all temperatures other than the fundamental temperature, the pressure of vaporization, of fusion, and of sublimation differ from one another. \PageSep{155} \Section{188.} We return once more to the intrinsic conditions of equilibrium~\Eq{(101)} which hold for each of the three combinations of two states of aggregation. The pressure~$p_{1}$, and the specific volumes of the two portions of the substance, in each case depend only on the temperature, and are determined by~\Eq{(101)}. It is necessary, however, to distinguish whether the saturated vapour is in contact with the liquid or the solid, since in these two cases the functions which express the pressure and specific volume in terms of the temperature are quite different. The state of the saturated vapour is determined only when there is given, besides the temperature, the state of aggregation with which it is in contact, whether it is in contact with the liquid or solid. The same applies to the other two states of aggregation. If we henceforth use the suffixes $1$,~$2$,~$3$, in this order, to refer to the gaseous, liquid, and solid states, we shall be obliged to use two of them when we refer to a portion of the substance in a state of saturation. The first of these will refer to the state of the portion considered, the second to that of the portion with which it is in contact. Both the symbols $v_{12}$~and $v_{13}$ thus denote the specific volume of the saturated vapour,~$v_{12}$, in contact with the liquid, and $v_{13}$~in contact with the solid. Similarly $v_{23}$~and $v_{21}$,~$v_{31}$\Add{,} and~$v_{32}$ represent the specific volumes of the liquid, and of the solid in a state of saturation. Each of these six quantities is a definite function of the temperature alone. The corresponding pressures are \begin{center} \TableFont \begin{tabular}{ccc} Of vaporization. & Of fusion. & Of sublimation. \\ $p_{12} = p_{21}$ & $p_{23} = p_{32}$ & $p_{31} = p_{13}$ \\ \end{tabular} \end{center} These are also functions of the temperature alone. Only at the fundamental temperature do two of these pressures become equal, and therefore equal to the third. If we represent the relation between these three pressures and the temperature by three curves, the temperatures as abscissæ and the pressures as ordinates, these curves will meet in one point, the \emph{fundamental point}, also called the \emph{triple point}. \index{Fundamental point (triple)}% \index{Point!triple}% \index{Triple point}% \PageSep{156} The inclination of the curves to the abscissa is given by the differential coefficients \[ \frac{dp_{12}}{d\theta};\quad \frac{dp_{23}}{d\theta};\quad \frac{dp_{31}}{d\theta}. \] We have, therefore, according to equations~\Eq{(111)}, \begin{align*} \frac{dp_{12}}{d\theta} &= \frac{L_{12}}{\theta (v_{1} - v_{2})}, \\ \frac{dp_{23}}{d\theta} &= \frac{L_{23}}{\theta (v_{2} - v_{3})}, \\ \frac{dp_{31}}{d\theta} &= \frac{L_{31}}{\theta (v_{3} - v_{1})}, \\ \end{align*} where $v$~refers to the fundamental state, and, therefore, requires only one suffix. We can thus find the direction of each curve at the fundamental point if we know the heat of vaporization, of fusion, and of sublimation. Let us compare, for example, the curve of vaporization,~$p_{12}$, of water, with its curve of sublimation,~$p_{13}$, near the fundamental point, $0.0074°$~C\@. We have, in absolute units, % [** TN: Re-breaking] \settowidth{\TmpLen}{heat of vaporization of\quad}% \begin{align*} L_{12} &= 604 × 419 × 10^{5}\quad\parbox[t]{\TmpLen}{(heat of vaporization of \\ water at $0.0074°$~C.);} \\ L_{13} &= -L_{31} = (80 + 604) × 419 × 10^{5}\quad\parbox[t]{\TmpLen}{(heat of sublimation \\ of ice at $0.0074°$~C.);} \\ v_{1} &= 205000;\quad v_{2} = 1;\quad v_{3} = 1.09 \text{ (\SecRef{187})};\quad \theta = 273. \end{align*} Hence \begin{align*} \frac{dp_{12}}{d\theta} &= \frac{604 × 419 × 10^{5} × 760} {273 × 205000 × 1013650} = 0.339, \\ \frac{dp_{31}}{d\theta} &= \frac{684 × 419 × 10^{5} × 760} {273 × 205000 × 1013650} = 0.384, \end{align*} in millimeters of mercury. The curve of the sublimation pressure~$p_{13}$, is steeper at the fundamental point than the curve of the vaporization pressure~$p_{12}$. For temperatures above the fundamental one, therefore, $p_{13} > p_{12}$; for those below it, $p_{12} > p_{13}$. Their difference is \[ \frac{dp_{13}}{d\theta} - \frac{dp_{12}}{d\theta} = \frac{d(p_{13} - p_{12})}{d\theta} = 0.045\Add{.} \] \PageSep{157} If, therefore, the maximum vapour pressure of water be measured above the fundamental point, and of ice below it, the curve of pressure will show an abrupt bend at the fundamental point. This change of direction is measured by the discontinuity of the differential coefficient. At $-1°$~C., ($d\theta = -1$), we have, approximately, \[ p_{13} - p_{12} = -0.045; \] \ie\ at $1°$~C. the maximum vapour pressure of ice is $0.045~\Unit{mm.}$ less than that of water. This has been verified by experiment. The existence of a sharp bend in the curve, however, can only be inferred from theory. \Section{189.} We have hitherto extended our investigation only to the different admissible solutions of the equations which express the intrinsic conditions of equilibrium, and have deduced from them the properties of the states of equilibrium to which they lead. We shall now consider the relative merit of these solutions, \ie\ which of them represents the state of greatest stability. For this purpose we resume our original statement of the problem (\SecRef{165}), which is briefly as follows: Given the total mass~$M$, the total volume~$V$, and the total energy~$U$, it is required to find the state of most stable equilibrium, \ie\ the state in which the total entropy of the system is an absolute maximum. Instead of $V$~and~$U$, however, it is often more convenient to introduce $v = \dfrac{V}{M}$, the mean specific volume of the system, and $u = \dfrac{U}{M}$, the mean specific energy of the system. We have found that the conditions of equilibrium admit, in general, of three kinds of solution, according as the system is split into $1$,~$2$, or $3$~states of aggregation. When we come to consider which of these three solutions deserves the preference in a given case, we must remember that the second and third can be interpreted physically only if the values of the masses, as given by the equations \Eq{(103)} and~\Eq{(121)}, \PageSep{158} are positive. This restricts the region of validity of these two solutions. We shall first establish this region of validity, and then prove that within its region the third solution is always preferable to the other two, and, similarly, the second is preferable to the first. A geometrical representation may facilitate a general survey of the problem. We shall take the mean specific volume, $v = \dfrac{V}{M}$, and the mean specific energy, $u = \dfrac{U}{M}$, of the system as the rectangular co-ordinate axes. The value of~$M$ is here immaterial. Each point of this plane will, \ifthenelse{\boolean{ForPrinting}}{% \Figure[\textwidth]{4} }{% \Figure[4.25in]{4} } \index{Curves!of evaporation}% \index{Curves!of fusion}% \index{Curves!of sublimation}% \index{Fusion, curve}% \index{Sublimation, curve}% \index{Vaporization curve}% then, represent definite values of $u$~and~$v$. Our problem is, therefore, to find the kind of stable equilibrium which will correspond to any given point in this plane. \Section{190.} Let us consider the region of validity of the third \PageSep{159} solution. The values of the masses given by the equations~\Eq{(121)} are \begin{multline*} M_{1} : M_{2} : M_{3} : M \\ = \left|\begin{array}{@{}ccc@{}} 1 & 1 & 1 \\ v & v_{2} & v_{3} \\ u & u_{2} & u_{3} \\ \end{array}\right| : \left|\begin{array}{@{}ccc@{}} 1 & 1 & 1 \\ v & v_{3} & v_{1} \\ u & u_{3} & u_{1} \\ \end{array}\right| : \left|\begin{array}{@{}ccc@{}} 1 & 1 & 1 \\ v & v_{1} & v_{2} \\ u & u_{1} & u_{2} \\ \end{array}\right| : \left|\begin{array}{@{}ccc@{}} 1 & 1 & 1 \\ v_{1} & v_{2} & v_{3} \\ u_{1} & u_{2} & u_{3} \\ \end{array}\right| \Tag{(121a)} \end{multline*} where $v_{1}$,~$v_{2}$,~$v_{3}$, $u_{1}$,~$u_{2}$,~$u_{3}$, refer, as hereafter, to the special values which these quantities assume in the fundamental state. It is obvious from this that the values of $M_{1}$,~$M_{2}$,~$M_{3}$ can be simultaneously positive only when the point $(v, u)$ lies within the triangle formed by the points $(v_{1}, u_{1})$\Add{,} $(v_{2}, u_{2})$\Add{,} and $(v_{3}, u_{3})$. The area of this triangle then represents the region of validity of the third solution, and may be called the \emph{fundamental triangle} of the substance. In \Fig{4} this \index{Fundamental point (triple)!triangle}% \index{Triangle, fundamental}% triangle is represented by~\Eq{(123)}. The diagram is based on a substance for which, as for water, \[ v_{1} > v_{3} > v_{2}\quad\text{and}\quad u_{1} > u_{2} > u_{3}. \] \Section{191.} We shall now consider the region of validity of the second solution contained in equations \Eq{(101)} and~\Eq{(103)}. These equations furnish three sets of values for the three possible combinations, and no preference can be given to any one of these. If we consider first the combination of liquid and vapour, the equations referred to become, under our present notation, \begin{gather*} \left. \begin{aligned} \theta_{12} &= \theta_{21}\Add{,} \\ p_{12} &= p_{21}\Add{,} \\ \phi_{12} - \phi_{21} &= \frac{u_{12} - u_{21} + p_{12} (v_{12} - v_{21})}{\theta_{12}}\Add{;} \\ \end{aligned} \right\} \Tag{(122)}\displaybreak[0] \\ % \left. \begin{aligned} M_{12} + M_{21} &= M\Add{,} \\ M_{12} v_{12} + M_{21} v_{21} &= V = Mv\Add{,} \\ M_{12} u_{12} + M_{21} u_{21} &= U = Mu\Add{.} \\ \end{aligned} \right\} \Tag{(123)} \end{gather*} In order to determine the area within which the point $(v, u)$ must lie so that $M_{12}$~and $M_{21}$ may both be positive, we shall find the limits of that area, \ie\ the curves represented \PageSep{160} by the conditions $M_{12} = 0$, and $M_{21} = 0$. The introduction of the latter (no liquid mass) gives $M_{12} = M$ and \[ v = v_{12},\quad u = u_{12}. \Tag{(124)} \] Since $v_{12}$~and $u_{12}$ are functions of a single variable, the conditions~\Eq{(124)} restrict the point $(v, u)$ to a curve, one of the limits of the region of validity. The curve passes through the vertex~$1$ of the fundamental triangle, because, at the fundamental temperature, $v_{12} = v_{1}$, and $u_{12} = u_{1}$. To follow the path of the curve it is necessary to find the differential coefficient~$\dfrac{du_{12}}{dv_{12}}$. We, have \[ \frac{du_{12}}{dv_{12}} = \left(\frac{\dd u}{\dd v}\right)_{12} + \left(\frac{\dd u}{\dd \theta}\right)_{12} \frac{d\theta_{12}}{dv_{12}}. \] The partial differential coefficients here refer to the independent variables $\theta$~and~$v$. It follows from \Eq{(80)}~and \Eq{(24)} that \[ \frac{du_{12}}{dv_{12}} = \theta_{12} \left(\frac{\dd p}{\dd \theta}\right)_{12} - p_{12} + (c_{v})_{12}\, \frac{d\theta_{12}}{dv_{12}}. \] By means of this equation the path of the curve~\Eq{(124)} may be experimentally plotted by taking~$\theta_{12}$, or~$v_{12}$, or some other appropriate quantity as independent parameter. Similarly, the condition $M_{12} = 0$ (no vapour) gives another boundary of the region of validity, viz.\ the curve, \[ v = v_{21},\quad u = u_{21}, \] which passes through the vertex~$2$ of the fundamental triangle, and satisfies the differential equation \[ \frac{du_{21}}{dv_{21}} = \theta_{21} \left(\frac{\dd p}{\dd \theta}\right)_{21} - p_{12} + (c_{v})_{21}\, \frac{d\theta_{12}}{dv_{21}}, \] since $\theta_{21} = \theta_{12}$ and $p_{21} = p_{12}$. The two limiting curves, however, are merely branches of one curve, since they pass into one another at the critical point ($v_{12} = v_{21}$) without forming an angle or cusp at that \PageSep{161} \index{Corresponding point.}% \index{Curves!of evaporation}% \index{Curves!of fusion}% \index{Curves!of sublimation}% \index{Fusion, curve}% \index{Sublimation, curve}% point, as a further discussion of the values of~$\dfrac{du_{12}}{dv_{12}}$ and $\dfrac{du_{21}}{dv_{21}}$ will show. We may, therefore, include the two branches under the name of the \emph{vaporization curve}. Every point \index{Vaporization curve}% $(v_{12}, u_{12})$ of one branch has a \emph{corresponding point} $(v_{21}, u_{21})$ on the other, since these two represent the same temperature $\theta_{12} = \theta_{21}$, and the same pressure $p_{12} = p_{21}$. This co-ordination of points on the two branches is given by the equations~\Eq{(122)}, and has been indicated on our diagram (\Fig{4}) by drawing some dotted lines joining corresponding points. In this sense the vertices $1$~and~$2$ of the fundamental triangle are corresponding points, and the critical point is self-corresponding. This vaporization curve bounds the region of validity of that part of the second solution which refers to liquid in contact with its vapour. Equation~\Eq{(123)} makes it obvious that the region of validity lies within the concave side of the curve. The curve has not been produced beyond the vertices $1$~and~$2$ of the fundamental triangle, because we shall see later, that the side~$12$ of that triangle bounds the area within which this solution gives stable equilibrium. There may be found, quite analogous to the vaporization curve, also a \emph{fusion curve} the two branches of which are represented by \[ v = v_{23},\quad u = u_{23}, \] and \[ v = v_{32},\quad u = u_{32}, \] and a \emph{sublimation curve} represented by \[ v = v_{31},\quad u = u_{31}, \] and \[ v = v_{13},\quad u = u_{13}, \] The former passes through the vertices $2$~and~$3$, the latter through $3$~and~$1$, of the fundamental triangle. The region of validity of the three parts of the second solution have been marked $(12)$,~$(23)$, and $(31)$, respectively, in \Fig{4}. The relations which have been specially deduced for the area~$(12)$ apply to $(23)$ and~$(31)$ as well only with a corresponding \PageSep{162} interchange of the suffixes. Some pairs of corresponding points have again been joined by dotted lines. On the fusion curve a critical point has been marked, on the assumption that, with falling temperature, the latent heat of ice decreases by $0.64$~calorie per degree Centigrade (\SecRef{183}). If, as a rough approximation, we assume this same ratio to hold for much lower temperatures, the latent heat of fusion would be zero at about $-120°$~C., and this would be the critical point of the fusion curve. The pressure here would be about $17,000$~atmospheres, and water and ice would become identical. We might imagine this to be the result of a considerable increase in the viscosity of water and in the plasticity of ice, as they both approach this state. \Section{192.} Having thus fixed the region of validity for the second solution, we find that for all points $(v, u)$ outside this region only the first solution admits of physical interpretation. It follows that for such points the first solution represents the stable equilibrium. The areas where such is the case have been marked $(1)$,~$(2)$, and~$(3)$ in our figure, to signify the gaseous, liquid, and solid states respectively. \Section{193.} We have now to consider the following question: Which of the different states of equilibrium, that may correspond to given values $M$,~$v$,~$u$ (or to a given point of the figure), gives to the system the greatest value of the entropy? Since each of the three solutions discussed leads to a definite state of the system, we have for each given system $(M, v, u)$ as many values of the entropy as there are solutions applying to it. Denoting these by $\Phi$,~$\Phi'$, and~$\Phi''$, we get for the first solution \[ \Phi = M\phi\Add{;} \Tag{(125)} \] for the second \[ \Phi' = M\phi' = M_{12} \phi_{12} + M_{21} \phi_{21} \Tag{(126)} \] (or a cyclic interchange of the suffixes $1$,~$2$,~$3$); for the third \[ \Phi'' = M\phi'' = M_{1} \phi_{1} + M_{2} \phi_{2} + M_{3} \phi_{3}\Add{.} \Tag{(127)} \] \PageSep{163} All these quantities are fully determined by the given values of $M$,~$v$, and~$u$. Now, we can show that for any system $(M, v, u)$ we have $\Phi'' > \Phi' > \Phi$, or $\phi'' > \phi' > \phi$, provided all the partial masses are positive. It is more convenient to deal with the mean specific entropies than with the entropies themselves, because the former, being functions of $v$~and $u$~alone, are quite independent of~$M$. As a geometrical representation, we may imagine, on the plane of our figure (\Fig{4}), perpendiculars erected at each point $(v, u)$, proportional in length to the values of $\phi$,~$\phi'$, and $\phi''$~respectively, at that point. The upper ends of these perpendiculars will generate the three surfaces of entropy, $\phi$,~$\phi'$, and~$\phi''$. \Section{194.} We shall show that $\phi' - \phi$ is always positive, \ie\ that the surface of entropy,~$\phi'$, lies everywhere above the surface~$\phi$. While the value of~$\phi$ may be taken directly from~\Eq{(61)}, which contains the definition of the entropy for homogeneous substances, $\phi'$~may be found from \Eq{(126)}, \Eq{(122)}, and~\Eq{(123)}, in terms of $v$~and~$u$. The surface~$\phi'$ forms three sheets corresponding to the three combinations of two states of aggregation. We shall in the following refer to the combination of vapour and liquid. With regard to the relative position of the surfaces $\phi$ and~$\phi'$, it is obvious that they have one curve in common, the projection of which is the vaporization curve. At any point on the vaporization curve we have $v = v_{12}$, $u = u_{12}$, and for the first entropy surface, $\phi = \phi_{12}$; for the second we have, from~\Eq{(123)}, \[ M_{21} = 0,\quad M_{12} = M \Tag{(128)} \] and, from~\Eq{(126)}, \[ \phi' = \phi_{12}. \] In fact, for all points of the vaporization curve, both solutions coincide. The curve of intersection of the surfaces $\phi$~and~$\phi'$ is represented by \[ v = v_{12},\quad u = u_{12},\quad \phi = \phi_{12}, \] \PageSep{164} where $v$,~$u$, and~$\phi$ are the three rectangular co-ordinates of a point in space. $v_{12}$,~$u_{12}$,~$\phi_{12}$ depend on a single variable parameter, for example, the temperature, $\theta_{12} = \theta_{21}$. This curve passes through the point $(v_{1}, u_{1}, \phi_{1})$, which has the vertex~$1$ for its projection. A second branch of the same curve is given by the equations \[ v = v_{21},\quad u = u_{21},\quad \phi = \phi_{21}, \] and these branches meet in a point whose projection is the critical point. Each point of one branch has a corresponding point on the other, since both correspond to the same temperature, $\theta_{12} = \theta_{21}$, and the same pressure, $p_{12} = p_{21}$. Thus, $(v_{1}, u_{1}, \phi_{1})$ and $(v_{2}, u_{2}, \phi_{2})$ are corresponding points. It is further obvious that the surface~$\phi'$ is a ruled surface \index{Developable surface}% \index{Surface, developable}% and is developable. The first may be shown by considering any point with the co-ordinates \[ v = \frac{\lambda v_{12} + \mu v_{21}}{\lambda + \mu};\quad u = \frac{\lambda u_{12} + \mu u_{21}}{\lambda + \mu};\quad \phi = \frac{\lambda \phi_{12} + \mu \phi_{21}}{\lambda + \mu}; \] where $\lambda$~and~$\mu$ are arbitrary positive quantities. By giving $\lambda$~and $\mu$ all positive values, we obtain all points of the straight line joining the corresponding points $(v_{12}, u_{12}, \phi_{12})$ and $(v_{21}, u_{21}, \phi_{21})$. But this line lies on the surface~$\phi'$, since all the above values of $(v, u, \phi)$ satisfy the equations \Eq{(123)} and~\Eq{(126)} if we put $M_{12} = \lambda$ and $M_{21} = \mu$. The surface~$\phi'$, then, is formed of the lines joining the corresponding points on the curve in which the surfaces $\phi'$~and $\phi$ meet. One of these is the line joining the points $(v_{1}, u_{1}, \phi_{1})$ and $(v_{2}, u_{2}, \phi_{2})$, the projection of which is the side~$12$ of the fundamental triangle. At the critical point, the line shrinks to a point, and here the surface~$\phi$ ends. The other two sheets of the surface are quite similar. One begins at the line joining $(v_{2}, u_{2}, \phi_{2})$ and $(v_{3}, u_{3}, \phi_{3})$, the other at the line joining $(v_{3}, u_{3}, \phi_{3})$ and $(v_{1}, u_{1}, \phi_{1})$. The developability of the surface~$\phi'$ may best be inferred from the following equation of a plane:--- \[ p_{12} (v - v_{12}) + (u - u_{12}) - \theta_{12} (\phi - \phi_{12}) = 0, \] \PageSep{165} where $v$,~$u$,~$\phi$ are variable co-ordinates, while $p_{12}$,~$v_{12}$, $u_{12}$, $\theta_{12}$,~$\phi_{12}$, depend, by~\Eq{(122)}, on one parameter, \eg~$\theta_{12}$. This plane contains the point $(v_{12}, u_{12}, \phi_{12})$, and by the equations~\Eq{(122)} the point $(v_{21}, u_{21}, \phi_{21})$, which are corresponding points, and hence also the line joining them. But it also, by~\Eq{(61)}, contains the neighbouring corresponding points \[ (v_{12} + dv_{12}, u_{12} + du_{12}, \phi_{12} + d\phi_{12}) \] and \[ (v_{21} + dv_{21}, u_{21} + du_{21}, \phi_{21} + d\phi_{21})\Add{,} \] hence also the line joining them. Therefore, two consecutive generating lines are coplanar, which is the condition of developability of a surface. In order to determine the value of $\phi' - \phi$, we shall find the change which this difference undergoes on passing from a point $(v, u)$ to a neighbouring one $(v + \delta v, u + \delta u)$. During this passage we shall keep $M = M_{12} + M_{21}$ constant. This does not affect the generality of the result, since $\phi$~and $\phi'$ are functions of $v$~and $u$ only. From~\Eq{(126)} we have \[ M\, \delta \phi' = M_{12}\, \delta \phi_{12} + M_{21}\, \delta \phi_{21} + \phi_{12}\, \delta M_{12} + \phi_{21}\, \delta M_{21}, \] and, by~\Eq{(61)}, \[ \delta \phi = \frac{\delta u + p\, \delta v}{\theta}. \] But, by~\Eq{(123)}, \[ \left. \begin{aligned} \delta M_{12} + \delta M_{21} &= 0\Add{,} \\ M_{12}\, \delta v_{12} + M_{21}\, \delta v_{21} + v_{12}\, \delta M_{12} + v_{21}\, \delta M_{21} &= M\, \delta v\Add{,} \\ M_{12}\, \delta u_{12} + M_{21}\, \delta u_{21} + u_{12}\, \delta M_{12} + u_{21}\, \delta M_{21} &= M\, \delta u\Add{.} \\ \end{aligned} \right\} \Tag{(129)} \] Whence, by~\Eq{(122)}, \begin{align*} \delta \phi' &= \frac{\delta u + p_{12}\, \delta v}{\theta_{12}} \Tag{(130)} \\ \intertext{and} \delta (\phi' - \phi) &= \left(\frac{1}{\theta_{12}} - \frac{1}{\theta}\right) \delta u + \left(\frac{p_{12}}{\theta_{12}} - \frac{p}{\theta}\right) \delta v. \Tag{(131)} \end{align*} \PageSep{166} If we now examine the surfaces $\phi$~and $\phi'$ in the neighbourhood of their curve of contact, it is evident from the last equation that they touch one another along the whole of this curve. For, at any point of the vaporization curve, we have $v = v_{12}$ and $u = u_{12}$; therefore also \[ \theta = \theta_{12},\quad\text{and}\quad p = p_{12} \Tag{(132)} \] and hence, for the entire curve, $\delta (\phi' - \phi) = 0$. To find the kind of contact between the two surfaces, we form $\delta^{2} (\phi' - \phi)$ from~\Eq{(131)}, and apply it to the same points of the curve of contact. In general, \begin{align*} \delta^{2} (\phi' - \phi) &= \delta u \left(\frac{\delta \theta}{\theta^{2}} - \frac{\delta \theta_{12}}{\theta_{12}^{2}}\right) \\ &+ \delta v \left(\frac{\delta p_{12}}{\theta_{12}} - \frac{\delta p}{\theta} - \frac{p_{12}\, \delta \theta_{12}}{\theta_{12}^{2}} + \frac{p\, \delta \theta}{\theta^{2}}\right) \\ &+ \delta^{2} u \left(\frac{1}{\theta_{12}} - \frac{1}{\theta}\right) + \delta^{2} v \left(\frac{p_{12}}{\theta_{12}} - \frac{p}{\theta}\right). \end{align*} According to~\Eq{(132)} we have, at the points of contact of the surfaces, \[ \theta^{2}\, \delta^{2} (\phi' - \phi) = \delta u\, (\delta \theta - \delta \theta_{12}) + \delta v (\theta\, \delta p_{12} - \theta\, \delta p - p\, \delta \theta_{12} + p\,\delta \theta), \] or, by~\Eq{(61)}, \[ \theta\, \delta^{2} (\phi' - \phi) = (\delta \theta - \delta \theta_{12})\, \delta \phi + (\delta p_{12} - \delta p)\, \delta v. \Tag{(133)} \] All these variations may be expressed in terms of $\delta \theta$~and~$\delta v$, by putting \begin{align*} \delta \phi &= \frac{c_{v}}{\theta}\, \delta \theta + \frac{\dd p}{\dd \theta}\, \delta v\quad\text{\Chg{(by 81)}{by~\Eq{(81)}}}, \\ \delta p &= \frac{\dd p}{\dd \theta}\, \delta \theta + \frac{\dd p}{\dd v}\, \delta v, \\ \delta p_{12} &= \frac{dp_{12}}{d\theta_{12}}\, \delta \theta_{12}. \end{align*} We have now to express $\delta \theta_{12}$ in terms of $\delta \theta$~and~$\delta v$. Equations~\Eq{(129)}, here simplified by~\Eq{(128)}, give \[ \frac{\delta u_{12} - \delta u}{u_{12} - u_{21}} = \frac{\delta v_{12} - \delta v}{v_{12} - v_{21}}. \] \PageSep{167} In these we put \begin{align*} \delta u_{12} &= \frac{du_{12}}{d\theta_{12}}\, \delta \theta_{12},\quad \delta v_{12} = \frac{dv_{12}}{d\theta_{12}}\, \delta \theta_{12}\Add{,} \Tag{(134)} \\ \delta u &= c_{v}\, \delta \theta + \frac{\dd u}{\dd v}\, \delta v, \end{align*} and obtain \[ \delta \theta_{12} = \frac{c_{v}\, \delta \theta + \left(\dfrac{\dd u}{\dd v} - \dfrac{u_{12} - u_{21}}{v_{12} - v_{21}}\right) \delta v} {\dfrac{du_{12}}{d\theta_{12}} - \dfrac{u_{12} - u_{21}}{v_{12} - v_{21}} \Add{·} \dfrac{dv_{12}}{d\theta_{12}}}. \] If we consider that, by~\Eq{(109)}, \begin{align*} \frac{u_{12} - u_{21}}{v_{12} - v_{21}} &= \theta_{12}\, \frac{dp_{12}}{d\theta_{12}} - p_{12} \Tag{(135)} \\ &= \theta\, \frac{dp_{12}}{d\theta_{12}} - p, \end{align*} that, by~\Eq{(80)}, \[ \frac{\dd u}{\dd v} = \theta\, \frac{dp}{d\theta} - p \] and that \begin{align*} \frac{du_{12}}{d\theta_{12}} &= \left(\frac{\dd u}{\dd \theta}\right)_{12} + \left(\frac{\dd u}{\dd v}\right)_{12} · \frac{dv_{12}}{d\theta_{12}} \\ &= c_{v} + \left(\theta\, \frac{\dd p}{\dd \theta} - p\right) \frac{dv_{12}}{d\theta_{12}}\Add{;} \Tag{(136)} \end{align*} also that \[ \frac{dp_{12}}{d\theta_{12}} = \frac{\dd p}{\dd \theta} + \frac{\dd p}{\dd v} · \frac{dv_{12}}{d\theta_{12}}, \] we obtain \[ \delta \theta_{12} = \frac{c_{v}\, \delta \theta - \theta\, \dfrac{\dd p}{\dd v} · \dfrac{dv_{12}}{d\theta_{12}}\, \delta v}{c_{v} - \theta\, \dfrac{\dd p}{\dd v} \left(\dfrac{dv_{12}}{d\theta_{12}}\right)^{2}}. \] Equation~\Eq{(133)}, with all variations expressed in terms of $\delta \theta$ and~$\delta v$, finally becomes \[ \delta^{2} (\phi' - \phi) = -\frac{\dd p}{\dd v} · \frac{c_{v}}{\theta} · \frac{\left(\dfrac{dv_{12}}{d\theta_{12}}\, \delta \theta - \delta v\right)^{2}}{c_{v} - \theta\, \dfrac{\dd p}{\dd v} \left(\dfrac{dv_{12}}{d\theta_{12}}\right)^{2}}. \] This expression is essentially positive, since $c_{v}$~is positive \PageSep{168} on account of its physical meaning, and $\dfrac{\dd p}{\dd v}$~is negative for any state of equilibrium (\SecRef{169}). There is a limiting case, when \[ \frac{dv_{12}}{d\theta_{12}}\, \delta \theta - \delta v = 0, \] for, then, \[ \delta^{2} (\phi' - \phi) = 0. \] In this case the variation $(\delta \theta, \delta v)$ obviously takes place along the curve of contact $(\theta_{12}, v_{12})$ of the surfaces. Here we know that $\phi' = \phi$. It follows that the surface~$\phi'$, in the vicinity of all points of contact with~$\phi$, rises above the latter throughout, or that $\phi' - \phi$ is everywhere~$> 0$. This proves that the second solution of the conditions of equilibrium, within its region of validity, \ie\ in the areas $(12)$,~$(23)$, and~$(31)$, always represents the stable equilibrium. \Section{195.} Similarly, it may be shown that the third solution, within its region of validity, is preferable to the second one. The quantities $v$~and $u$ being given, the value of the mean specific entropy,~$\phi''$, corresponding to this solution is uniquely determined by the equations \Eq{(127)} and~\Eq{(121)}. The quantities $v_{1}$,~$v_{2}$,~$v_{3}$, $u_{1}$,~$u_{2}$,~$u_{3}$, and therefore also $\phi_{1}$,~$\phi_{2}$,~$\phi_{3}$, have definite numerical values, given by equations~\Eq{(120)}. In the first place, it is obvious that the surface~$\phi''$ is the plane triangle formed by the points $(v_{1}, u_{1}, \phi_{1})$, $(v_{2}, u_{2}, \phi_{2})$, and $(v_{3}, u_{3}, \phi_{3})$, the projection of which on the plane of the figure is the fundamental triangle, since any point with the co-ordinates \begin{align*} v &= \frac{\lambda v_{1} + \mu v_{2} + \nu v_{3}}{\lambda + \mu + \nu}, \\ u &= \frac{\lambda u_{1} + \mu u_{2} + \nu u_{3}}{\lambda + \mu + \nu}, \\ \phi &= \frac{\lambda \phi_{1} + \mu \phi_{2} + \nu \phi_{3}}{\lambda + \mu + \nu}, \end{align*} ($\lambda$,~$\mu$,~$\nu$ may have any positive values) satisfies the equations \Eq{(121)} and~\Eq{(127)}. To show this, we need only put $M_{1} = \lambda$, \PageSep{169} $M_{2} = \mu$, $M_{3} = \nu$. This plane meets the three sheets of the developable surface~$\phi'$ in the three lines joining the points $(v_{1}, u_{1}, \phi_{1})$, $(v_{2}, u_{2}, \phi_{2})$, $(v_{3}, u_{3}, \phi_{3})$. In fact, by making $\nu = 0$, \ie, by~\Eq{(121)}, $M_{3} = 0$, the third solution coincides with the second; for, then, \[ \left. \begin{gathered} M_{1} = M_{12};\quad M_{2} = M_{21};\quad v_{1} = v_{12};\quad u_{1} = u_{12}; \\ v_{2} = v_{2l};\quad \theta_{1} = \theta_{12};\quad\text{etc.} \end{gathered} \right\} \Tag{(137)} \] If we also put $\mu = 0$, then we have $M_{2} = 0$, $v_{1} = v$, $u_{1} = u$, which means the coincidence of all three surfaces, $\phi''$,~$\phi'$, and~$\phi$. In order to find the sign of $\phi'' - \phi'$, we again find $\delta (\phi'' - \phi')$ in terms of $\delta u$~and~$\delta v$. Equation~\Eq{(127)} gives \[ M\, \delta\phi'' = \phi_{1}\, \delta M_{1} + \phi_{2}\, \delta M_{2} + \phi_{3}\, \delta M_{3} \Tag{(138)} \] where, by~\Eq{(121)}, \begin{align*} \delta M_{1} + \delta M_{2} + \delta M_{3} &= 0\Add{,} \\ v_{1}\, \delta M_{1} + v_{2}\, \delta M_{2} + v_{3}\, \delta M_{3} &= M\, \delta v\Add{,} \\ u_{1}\, \delta M_{1} + u_{2}\, \delta M_{2} + u_{3}\, \delta M_{3} &= M\, \delta u\Add{.} \end{align*} Multiplying the last of these by~$\dfrac{1}{\theta_{1}}$, the second by~$\dfrac{p_{1}}{\theta_{1}}$, and adding to~\Eq{(138)}, we obtain, with the help of~\Eq{(120)}, \[ \delta \phi'' = \frac{\delta u + p_{1}\, \delta v}{\theta_{1}}. \] This, in combination with~\Eq{(130)}, gives \[ \delta (\phi'' - \phi') = \left(\frac{1}{\theta_{1}} - \frac{1}{\theta_{12}}\right) \delta u + \left(\frac{p_{1}}{\theta_{1}} - \frac{p_{12}}{\theta_{12}}\right) \delta v, \Tag{(139)} \] if the surface~$\phi'$ is represented by the sheet~$(12)$. This equation shows that the surface~$\phi''$ is a tangent to the sheet~$(12)$ along the line joining $(v_{1}, u_{1}, \phi_{1})$ and $(v_{2}, u_{2}, \phi_{2})$, for all points of this line have $\theta_{1} = \theta_{12}$, $p_{1} = p_{12}$, so that $\delta (\phi'' - \phi')$ vanishes. Thus, we find that the plane~$\phi''$ is a tangent plane to the three sheets of the surface~$\phi'$. The curves of contact are the three straight lines which form \PageSep{170} the sides of the plane triangle~$\phi''$. We have, from~\Eq{(139)}, for any point of contact \[ \delta^{2} (\phi'' - \phi') = \frac{\delta \theta_{12}}{\theta_{1}^{2}}\, \delta u + \left(\frac{p_{1}\, \delta \theta_{12}}{\theta_{1}^{2}} - \frac{\delta p_{12}}{\theta_{1}}\right) \delta v, \] since $\theta_{1}$~and $p_{1}$ are absolute constants; or \[ \theta_{1}^{2}\, \delta^{2} (\phi'' - \phi') = \left[\delta u - \left(\theta_{1}\, \frac{dp_{12}}{d\theta_{12}} - p_{1}\right) \delta v\right] \delta \theta_{12}. \Tag{(140)} \] Now, by the elimination of $\delta M_{12}$ and~$\delta M_{21}$, it follows, from~\Eq{(129)}, that \[ \frac{M_{12}\, \delta v_{12} + M_{21}\, \delta v_{21} - M\, \delta v}{v_{12} - v_{21}} = \frac{M_{12}\, \delta u_{12} + M_{21}\, \delta u_{21} - M\, \delta u}{u_{12} - u_{21}} \] or, by \Eq{(135)} and~\Eq{(134)}, % [** TN: Re-breaking] \begin{multline*} M \left[\delta u - \left(\theta_{1}\, \frac{dp_{12}}{d\theta_{12}} - p_{1}\right) \delta v\right] \\ = \delta \theta_{12} \left[M_{12}\, \frac{du_{12}}{d\theta_{12}} + M_{21}\, \frac{du_{21}}{d\theta_{12}}\right. \\ - \left.\left(\theta_{1}\, \frac{dp_{12}}{d\theta_{12}} - p_{1}\right) · \left(M_{12}\, \frac{dv_{12}}{d\theta_{12}} + M_{21}\, \frac{dv_{21}}{d\theta_{12}}\right)\right]. \end{multline*} Substituting this expression in~\Eq{(140)}, and replacing $\dfrac{du_{12}}{d\theta_{12}}$ and $\dfrac{du_{21}}{d\theta_{21}}$ by their values~\Eq{(136)}, we obtain \begin{multline*} \delta^{2} (\phi'' - \phi') = \frac{\delta \theta_{12}^{2}}{M \theta_{1}^{2}} \left[M_{12} \left((c_{v})_{12} - \theta_{1} \left(\frac{\dd p}{\dd v}\right)_{12} · \left(\frac{dv_{12}}{d\theta_{12}}\right)^{2}\right)\right. \\ + \left.M_{21} \left((c_{v})_{21} - \theta_{1} \left(\frac{\dd p}{\dd v}\right)_{21} · \left(\frac{dv_{21}}{d\theta_{12}}\right)^{2}\right)\right]. \end{multline*} This quantity is essentially positive, since $M_{12}$,~$M_{21}$, as well as~$c_{v}$, are always positive, and $\dfrac{\dd p}{\dd v}$~always negative for states of equilibrium. There is a limiting case, when $\delta \theta_{12} = 0$, \ie\ for a variation along the line of contact of the surfaces $\phi''$~and~$\phi'$, as is obvious. It follows that the plane area~$\phi''$ rises everywhere above the surface~$\phi'$, and that $\phi'' - \phi'$ is never negative. This proves that the third solution within \PageSep{171} its region of validity (the fundamental triangle of the substance) represents stable equilibrium. \Section{196.} We are now in a position to answer generally the question proposed in \SecRef{165} regarding the stability of the equilibrium. The total mass~$M$, the volume~$V$, and the energy~$U$ of a system being given, its corresponding state of stable equilibrium is determined by the position of the point $v = \dfrac{V}{M}$, $u = \dfrac{U}{M}$, in the plane of \Fig{4}. If this point lie within one of the regions $(1)$,~$(2)$, or~$(3)$, the system behaves as a homogeneous gas, liquid, or solid. If it lie within $(12)$, $(23)$, or~$(31)$, the system splits into two different states of aggregation, indicated by the numbers used in the notation of the region. In this case, the common temperature and the ratio of the two heterogeneous portions are completely determined. According to the equation~\Eq{(123)}, the point $(v, u)$ lies on the straight line joining two corresponding points of the limiting curve. If a straight line be drawn through the given point $(v, u)$, cutting the two branches of that curve in corresponding points, these points give the properties of the two states of aggregation into which the system splits. They have, of course, the same temperature and pressure. The proportion of the two masses, according to the equation~\Eq{(123)}, is given by the ratio in which the point $(v, u)$ divides the line joining the corresponding points. If, finally, the point $(v, u)$ lie within the region of the fundamental triangle~\Eq{(123)}, stable equilibrium is characterized by a division of the system into a solid, a liquid, and a gaseous portion at the fundamental temperature and pressure. The masses of these three portions may then be determined by the equations~\Eq{(121a)}. It will be seen that their ratio is that of the three triangles, which the point~$(v, u)$ makes with the three sides of the fundamental triangle. The conditions of stable equilibrium of any substance \PageSep{172} can thus be found, provided its fundamental triangle, its vaporization, fusion, and sublimation curves have been drawn once for all. To obtain a better view of the different relations, isothermal and isopiestic curves may be added to the figure. These curves coincide in the regions $(12)$, $(23)$, $(31)$, and form the straight lines joining corresponding points on the limiting curves. On the other hand, the area~$(123)$ represents one singular isothermal and isopiestic (the triple point). In this way we may find that ice cannot exist in stable equilibrium at a higher temperature than the fundamental temperature ($0.0074°$~C.), no matter how the pressure may be reduced. Liquid water, on the other hand, may, under suitable pressure, be brought to any temperature without freezing or evaporating. A question which may also be answered directly is the following. Through what stages will a body pass if subjected to a series of definite external changes? For instance, the behaviour of a body of mass~$M$, when cooled or heated at constant volume~$V$, may be known by observing the line $v = \dfrac{V}{M}$ parallel to the axis of ordinates. The regions which this line traverses show the states through which the body passes, \eg\ whether the substance melts during the process, or whether it sublimes, etc. \PageSep{173} \Chapter[Any Number of Independent Constituents.]{III.}{System of any Number of Independent Constituents.} \index{Constituents, independent}% \index{Independent constituents}% \Section{197.} \First{We} proceed to investigate quite generally the equilibrium of a system made up of distinct portions in contact with one another. The system, contrary to that treated of in the preceding chapter, may consist of any number of independent constituents. Following Gibbs, we shall call each \index{Gibbs}% one of these portions, inasmuch as it may be considered physically homogeneous (\SecRef{67}), a \emph{phase}. Thus, a quantity \index{Phase!defined}% of water partly gaseous, partly liquid, and partly solid, forms a system of three phases. The number of phases as well as the states of aggregation is quite arbitrary, although we at once recognize the fact that a system in equilibrium may consist of any number of solid and liquid phases, but only one single \emph{gaseous} phase, for two different gases in contact are never in equilibrium with one another. \Section{198.} A system is characterized by the number of its \emph{independent constituents},\footnote {Frequently termed \emph{components}.} in addition to the number of its phases. The main properties of the state of equilibrium depend upon these. We define the number of independent constituents as follows. First find the number of elements contained in the system, and from these discard, as dependent constituents, all those whose quantity is determined in each phase by the remaining ones. The number of the remaining elements will be the number of independent constituents of the system. It is immaterial which of the \PageSep{174} constituents we regard as independent and which as dependent, since we are here concerned with the number, and not with the kind, of the independent constituents. The question as to the number of the independent constituents has nothing at all to do with the chemical constitution of the substances in the different phases, in particular, with the number of different kinds of molecules. Thus, a quantity of water in any number of states forms but one independent constituent, however many associations and dissociations of \ce{H2O}~molecules may occur (it may be a mixture of hydrogen and oxygen or ions), for the mass of the oxygen in each phase is completely determined by that of the hydrogen, and \textit{vice versâ}. Should, however, an excess of oxygen or hydrogen be present in the vapour, we have then two independent constituents. An aqueous solution of sulphuric acid forms a system of three chemical elements, \ce{S},~\ce{H}, and~\ce{O}, but contains only two independent constituents, for, in each phase (\eg\ liquid, vapour, solid) the mass of~\ce{O} depends on that of \ce{S}~and~\ce{H}, while the masses of \ce{S}~and~\ce{H} are not in each phase interdependent. Whether the molecule \ce{H2SO4} dissociates in any way, or whether hydrates are formed or not, does not change the number of independent constituents of the system. \Section{199.} We denote the number of independent constituents of a system by~$\alpha$. By our definition of this number we see, at once, that each phase of a given system in equilibrium is determined by the masses of each one of its $\alpha$~constituents, the temperature~$\theta$, and the pressure~$p$. For the sake of uniformity, we assume that each of the $\alpha$~independent constituents actually occurs in each phase of the system in a certain quantity, which, in special cases, may become infinitely small. The selection of the temperature and the pressure as independent variables, produces a change in the form of the equations of the last chapter, where the temperature and the specific volume were considered as the independent variables. The substitution of the pressure \PageSep{175} for the volume is more convenient here, because the pressure is the same for all phases in free contact, and it can in most cases be more readily measured. \Section{200.} We shall now consider the thermodynamical equilibrium of a system, in which the total masses of the a independent constituents $M_{1}$,~$M_{2}$,~\dots\Add{,} $M_{\alpha}$ are given. Of the different forms of the condition of equilibrium it is best to use that expressed by equation~\Eq{(79)} \[ \delta \Psi = 0 \Tag{(141)} \] which holds, if $\theta$~and~$p$ remain constant, for any change compatible with the given conditions. The function~$\Psi$ is given in terms of the entropy~$\Phi$, the energy~$U$, and the volume~$V$, by the equation \[ \Psi = \Phi - \frac{U + pV}{\theta}. \] \Section{201.} Now, let $\beta$~be the number of phases in the system, then $\Phi$,~$U$, and~$V$, and therefore also~$\Psi$, are sums of $\beta$~terms, each of which refers to a single phase, \ie\ to a physically homogeneous body: \[ \Psi = \Psi' + \Psi'' + \dots +\Psi^{\beta}\Add{,} \Tag{(142)} \] where the different phases are distinguished from one another by dashes. For the first phase, \[ \Psi' = \Phi' - \frac{U' + pV'}{\theta}. \Tag{(143)} \] $\Phi'$,~$U'$,~$V'$\Add{,} and~$\Psi'$ are completely determined by $\theta$,~$p$, and the masses $M_{1}$,~$M_{2}$,~\dots\Add{,} $M_{\alpha}$ of the independent constituents in the phases. As to how they depend on the masses, all we can at present say is, that, if all the masses were increased in the same proportion (say doubled), each of these functions would be increased in the same proportion. Since the nature of the phase remains unchanged, the entropy, the energy, and the volume change in the same proportion as \PageSep{176} the mass; hence, also, the function~$\Psi'$. In other words, $\Psi'$~is a homogeneous function of the masses $M_{1}'$,~$M_{2}'$,~\dots\Add{,} $M_{\alpha}'$ of the first degree, but not necessarily linear. To express this analytically, let us increase all the masses in the same ratio $1 + \eps: 1$, where $\eps$~is very small. All changes are then small; and for the corresponding change of~$\Psi'$ we obtain \begin{align*} \Delta \Psi' &= \frac{\dd \Psi'}{\dd M_{1}}\, \Erratum{\Delta M_{1}}{\Delta M_{1}'} + \frac{\dd \Psi'}{\dd M_{2}}\, \Delta M_{2}' + \dots \\ &= \frac{\dd \Psi'}{\dd M_{1}'}\, \eps M_{1}' + \frac{\dd \Psi'}{\dd M_{2}'}\, \eps M_{2}' + \dots\Add{.} \end{align*} But, by supposition, \[ \Delta \Psi = \eps\Psi', \] and, therefore, \index{Euler}% \[ \Psi' = \frac{\dd \Psi'}{\dd M_{1}'}\, M_{1}' + \frac{\dd \Psi'}{\dd M_{2}'}\, M_{2}' + \dots + \frac{\dd \Psi'}{\dd M_{\alpha}'}\, M_{\alpha}'. \Tag{(144)} \] {\Loosen Various forms may be given to this Eulerian equation by further differentiation. The differential coefficients $\dfrac{\dd \Psi'}{\dd M_{1}''}$, $\dfrac{\dd \Psi'}{\dd M_{2}''}$\Add{,}~\dots\ evidently depend on the constitution of the phase, and not on its total mass, since a change of mass changes both numerator and denominator in the same proportion.} \Section{202.} By~\Eq{(142)}, the condition of equilibrium becomes \index{Condition of complete reversibility!of equilibrium}% \index{Equilibrium!conditions of}% \[ \delta \Psi' + \delta \Psi'' + \dots \Add{+} \delta \Psi^{\beta} = 0\Add{,} \Tag{(145)} \] or, since the temperature and pressure remain constant, \begin{align*} &\frac{\dd \Psi'}{\dd M_{1}'}\, \delta M_{1}' + \frac{\dd \Psi'}{\dd M_{2}'}\, \delta M_{2}' + \dots + \frac{\dd \Psi'}{\dd M_{\alpha}'}\, \delta M_{\alpha}' \\ % + & \frac{\dd \Psi''}{\dd M_{1}''}\, \delta M_{1}'' + \frac{\dd \Psi''}{\dd M_{2}''}\, \delta M_{2}'' + \dots + \frac{\dd \Psi''}{\dd M_{\alpha}''}\, \delta M_{\alpha}'' \\ + & \dots \\ % + & \frac{\dd \Psi^{\beta}}{\dd M_{1}^{\beta}}\, \delta M_{1}^{\beta} + \frac{\dd \Psi^{\beta}}{\dd M_{2}^{\beta}}\, \delta M_{2}^{\beta} + \dots + \frac{\dd \Psi^{\beta}}{\dd M_{\alpha}^{\beta}}\, \delta M_{\alpha}^{\beta} = 0\Add{.} \Tag{(146)} \end{align*} \PageSep{177} If the variation of the masses were quite arbitrary, then the equation could only be satisfied, if all the coefficients of the variations were equal to~$0$. According to \SecRef{200}, however, the following conditions exist between them, \[ \left. \begin{aligned} M_{1} &= M_{1}' + M_{1}'' + \dots + M_{1}^{\beta}\Add{,} \\ M_{2} &= M_{2}' + M_{2}'' + \dots + M_{2}^{\beta}\Add{,} \\ & \dots \\ M_{\alpha} &= M_{\alpha}' + M_{\alpha}'' + \dots + M_{\alpha}^{\beta}\Add{,} \end{aligned}\right\} \Tag{(147)} \] and, therefore, for any possible change of the system, \[ \left. \begin{aligned} 0 &= \delta M_{1}' + \delta M_{1}'' + \dots + \delta M_{1}^{\beta}\Add{,} \\ 0 &= \delta M_{2}' + \delta M_{2}'' + \dots + \delta M_{2}^{\beta}\Add{,} \\ & \dots \\ 0 &= \delta M_{\alpha}' + \delta M_{\alpha}'' + \dots + \delta M_{\alpha}^{\beta}\Add{.} \end{aligned}\right\} \Tag{(148)} \] For the expression~\Eq{(146)} to vanish, the necessary and sufficient condition is \[ \left. \begin{aligned} \frac{\dd \Psi'}{\dd M_{1}'} &= \frac{\dd \Psi''}{\dd M_{1}''} = \dots = \frac{\dd \Psi^{\beta}}{\dd M_{1}^{\beta}}\Add{,} \\ \frac{\dd \Psi'}{\dd M_{2}'} &= \frac{\dd \Psi''}{\Erratum{}{\dd} M_{2}''} = \dots = \frac{\dd \Psi^{\beta}}{\dd M_{2}^{\beta}}\Add{,} \\ & \dots \\ \frac{\dd \Psi'}{\dd M_{\alpha}'} &= \frac{\dd \Psi''}{\dd M_{\alpha}''} = \dots = \frac{\dd \Psi^{\beta}}{\dd M_{\alpha}^{\beta}}\Add{.} \end{aligned}\right\} \Tag{(149)} \] There are for each independent constituent $(\beta - 1)$ equations, which must be satisfied, and therefore for all the $\alpha$~independent constituents $\alpha (\beta - 1)$ conditions. Each of these equations refers to the transition from one phase into another, and asserts that this particular transition does not take place in nature. This condition depends, as it must, on the internal constitution of the phase, and not on its total mass. Since the equations in a single row with regard to a particular constituent may be arranged in any order, it follows that, if a phase be in equilibrium as regards a \PageSep{178} given constituent with two others, these two other phases are in equilibrium with one another with regard to that constituent (they coexist). This shows that, since any system in equilibrium can have only one gaseous phase, two coexisting phases must emit the same vapour. For, since each phase is in equilibrium with the other, and also with its own vapour with respect to all constituents, it must also coexist with the vapour of the second phase. The coexistence of solid and liquid phases may, therefore, be settled by comparing their vapours. \Section{203.} It is now easy to see how the state of equilibrium \index{External conditions of equilibrium!variable}% \index{Internal conditions of equilibrium!variable}% \index{Variable, internal and external}% of the system is determined, in general, by the given external conditions~\Eq{(147)}, and the conditions of equilibrium~\Eq{(149)}. There are a of the former and $\alpha(\beta - 1)$ of the latter, a total of $\alpha\beta$~equations. On the other hand, the state of the $\beta$~phases depends on $(\alpha\beta + 2)$ variables, viz.\ on the $\alpha\beta$~masses, $M_{1}'$,~\dots\Add{,} $M_{\alpha}^{\beta}$, the temperature~$\theta$, and the pressure~$p$. After all conditions have been satisfied, two variables still remain undetermined. In general, the temperature and the pressure may be arbitrarily chosen, but in special cases, as will be shown presently, these are no longer arbitrary, and in such cases two other variables, as the total energy and the total volume of the system, are undetermined. By disposing of the values of the arbitrary variables we completely determine the state of the equilibrium. \Section{204.} The $\alpha\beta + 2$ variables, which control the state of the system, may be separated into those which merely govern the composition of the phases (\emph{internal} variables), and those which determine only the total masses of the phases (\emph{external} variables). The number of the former is $(\alpha - 1)\beta + 2$, for in each of the $\beta$~phases there are $\alpha - 1$ ratios between its $\alpha$~independent constituents, to which must be added temperature and pressure. The number of the external variables is~$\beta$, viz.\ the total masses of all the phases. We found that the $\alpha(\beta - 1)$ equations~\Eq{(149)} contain \PageSep{179} only internal variables, and, therefore, after these have been satisfied, there remain \[ \bigl[(\alpha - 1)\beta + 2\bigr] - \bigl[\alpha (\beta - 1)\bigr] = \alpha - \beta + 2 \] of the internal variables, undetermined. This number cannot be negative, for otherwise the number of the internal variables of the system would not be sufficient for \index{Non-variant system}% \index{System!non-variant}% the solution of the equations~\Eq{(149)}. It, therefore, follows that \[ \beta \leq \alpha + 2\Add{.} \] The number of the phases, therefore, cannot exceed the number of the independent constituents by more than two; or, a system of $\alpha$~independent constituents will contain at most $(\alpha + 2)$ phases. In the limiting case, where $\beta = \alpha + 2$, the number of the internal variables are just sufficient to satisfy the internal conditions of equilibrium~\Eq{(149)}. Their values in the state of equilibrium are completely determined quite independently of the given external conditions. Decreasing the number of phases by one increases the number of the indeterminate internal variables by one. This proposition, first propounded by Gibbs and universally \index{Gibbs's phase rule}% known as the \emph{phase rule}, has been amply verified, \index{Phase!rule}% especially by the experiments of Bakhuis Roozeboom. \index{Roozeboom, Bakhuis}% \Section{205.} We shall consider, first, the limiting case: \[ \beta = \alpha + 2. \] (Non-variant systems.) Since all the internal variables are completely determined, they form an \emph{$(n + 2)$-ple point}. \index{Point!$(n + 2)$-ple}% Change of the external conditions, as heating, compression, further additions of the substances, alter the total masses of the phases, but not their internal nature, including temperature and pressure. This holds until the mass of some one phase becomes zero, and therewith completely vanishes from the system. If $\alpha = 1$, then $\beta = 3$. A single constituent may split \PageSep{180} \index{System!univariant}% \index{Univariant system}% into three phases at most, forming a triple point. An \index{Point!triple}% \index{Point!quadruple}% \index{Triple point}% example of this is a substance existing in the three states of aggregation, all in contact with one another. For water it was shown in~\SecRef{187}, that at the triple point the temperature is $0.0074°$~C., and the pressure $4.62~\Unit{mm.}$ of mercury. The three phases need not, however, be different states of aggregation. Sulphur, for instance, forms several modifications \index{Sulphur!dioxide and water equilibrium}% in the solid state. Each modification constitutes a separate phase, and the proposition holds that two modifications of a substance can coexist with a third phase of the same substance, for example, its vapour, only at a definite temperature and pressure. A quadruple point is obtained when $\alpha = 2$. Thus, the \index{Quadruple point}% two independent constituents, \ce{SO2} (sulphur dioxide) and \ce{H2O}, form the four coexisting phases: \ce{SO2}, \ce{7H2O} (solid), \ce{SO2} dissolved in~\ce{H2O} (liquid), \ce{SO2} (liquid), \ce{SO2} (gaseous), at a temperature of $12.1°$~C. and a pressure of $1770~\Unit{mm.}$ of mercury. The question as to the formation of hydrates by \ce{SO2} in aqueous solution does not influence the application of the phase rule (see~\SecRef{198}). Three independent constituents ($\alpha = 3$) lead to a \Erratum{quintiple}{quintuple} \Erratum{\index{Point!quintiple}\index{Quintiple point}}{\index{Point!quintuple}\index{Quintuple point}}% point. Thus \ce{Na2SO4}, \ce{MgSO4}, and \ce{H2O} give the double salt \ce{Na2Mg(SO4)2 4H2O} (astrakanite), the crystals of the two simple salts, aqueous solution, and water vapour, at a temperature of $22°$~C. and a pressure of $19.6~\Unit{mm.}$ of mercury. \Section{206.} We shall now take the case \index{Heterogeneous system}% \index{System!heterogeneous}% \[ \beta = \alpha + 1, \] that is, $\alpha$~independent constituents form $\alpha + 1$ phases (Univariant systems). The composition of all the phases is then completely determined by a single variable, \eg\ the temperature or the pressure. This case is generally called \emph{perfect heterogeneous} equilibrium. If $\alpha = 1$, then $\beta = 2$: one independent constituent in two phases, \eg\ a liquid and its vapour. The pressure and the density of the liquid and the vapour depend on the temperature alone, as was pointed out in the last chapter. \PageSep{181} Evaporation involving chemical decomposition also belongs to this class, since the system contains only one independent constituent. The evaporation of solid \ce{NH4Cl}\Typo{.}{} is a case in point. Unless there be present an excess of hydrochloric acid or ammonia gas, there will be for each temperature a quite definite dissociation pressure. If $\alpha = 2$, then $\beta = 3$, for instance when the solution of a salt is in contact with its vapour and with the solid salt, or when two liquids that cannot be mixed in all proportions (ether and water) are in contact with their common vapour. Vapour pressure, density and concentration in each phase, are here functions of the temperature alone. \Section{207.} We often take the pressure instead of the temperature as the variable which controls the phases in perfect heterogeneous equilibrium; namely, in systems which do not possess a gaseous phase, so-called \emph{condensed} systems. Upon these the influence of the pressure is so slight that, under ordinary circumstances, it may be considered as given, and equal to that of the atmosphere. The phase rule, therefore, gives rise to the following proposition: \emph{A condensed system of a independent constituents forms $\alpha + 1$ \index{Condensed system}% \index{System!condensed}% phases at most, and is then completely determined, temperature included.} The melting point of a substance, and the point of transition from one allotropic modification to another, are examples of $\alpha = 1$, $\beta = 2$. The point at which the cryohydrate (ice and solid salt) separates out from the solution of a salt, and also the point at which two liquid layers in contact begin to precipitate a solid (\eg\ \ce{AsBr3}, and \ce{H2O}) are examples of $\alpha = 2$, $\beta = 3$. We have an example of $\alpha = 3$, $\beta = 4$ when two salts, capable of forming a double salt, are in contact with the solid simple salts, and also with the double salt. \Section{208.} If \index{Divariant system}% \index{System!divariant}% \[ \beta = \alpha, \] then $\alpha$~independent constituents form $\alpha$~phases (Divariant systems). The internal nature of all the phases depends \PageSep{182} on two variables, \eg\ on temperature and pressure. Any \index{Temperature!critical solution}% homogeneous substance furnishes an example of $\alpha = 1$. A liquid solution of a salt in contact with its vapour is an example of $\alpha = 2$. The temperature and the pressure determine the concentration in the vapour as well as in the liquid. The concentration of the liquid and either the temperature or the pressure are frequently chosen as the independent variables. In the first case, we say that a solution of given concentration and given temperature emits a vapour of definite composition and definite pressure; and in the second case, that a solution of given concentration and given pressure has a definite boiling point, and at this temperature a vapour of definite composition may be distilled off. Corresponding regularities hold when the second phase is solid or liquid, as in the case of two liquids which do not mix in all proportions. The internal nature of the two phases, in our example the concentrations in the two layers of the liquids, depends on two variables---pressure and temperature. If, under special circumstances, the concentrations become equal, a phenomenon is obtained which is quite analogous to that of the critical point of a homogeneous \index{Critical point!solution temperature}% substance (critical solution temperature of two liquids). \Section{209.} Let us now consider briefly the case \index{Isomorphous substance}% \index{Substance, isomorphous}% \[ \beta = \alpha - 1, \] where the number of phases is one less than the number of the independent constituents, and the internal nature of all phases depends on a third arbitrary variable, besides temperature and pressure. Thus, $\alpha = 3$, $\beta = 2$ for an aqueous solution of two isomorphous substances (potassium chlorate \index{Potassium chlorate}% and thallium chlorate) in contact with a mixed crystal. The \index{Thallium chlorate}% concentration of the solution under atmospheric pressure and at a given temperature will vary according to the composition of the mixed crystal. We cannot, therefore, speak of a saturated solution of the two substances of definite composition. However, should a second solid phase---for \PageSep{183} instance, a mixed crystal of different composition---separate out, the internal nature of the system will be determined by temperature and pressure alone. The experimental investigation of the equilibrium of such systems may enable us to decide whether a precipitate from a solution of two salts forms one phase---for example, a mixed crystal of changing concentration---or whether the two substances are to be considered as two distinct phases in contact. If, at a given temperature and pressure, the concentration of the liquid in contact were quite definite, it would represent the former case, and, if not, the latter. \Section{210.} If the expressions for the functions $\Psi'$,~$\Psi''$,~\dots for each phase were known, the equations~\Eq{(149)} would give every detail regarding the state of the equilibrium. This, however, is by no means the case, for, regarding the relations between these functions and the masses of the constituents in the individual phases, all we can, in general, assert is that they are homogeneous functions of the first degree (\SecRef{201}). We can, however, tell how they depend upon temperature and pressure, since their differential coefficients with respect to $\theta$~and~$p$ can be given. This point leads to far-reaching conclusions concerning the variation of the equilibrium with temperature and pressure. Since, for the first phase, according to~\Eq{(143)}, \[ \Psi' = \Phi' - \frac{U' + pV'}{\theta}, \] we have, for an infinitely small change, \[ d\Psi' = d\Phi' - \frac{dU' + p\, dV' + V'\, dp}{\theta} + \frac{U' + pV'}{\theta^{2}}\, d\theta. \] Under the assumption that the change is produced only by variations of $p$~and~$theta$, and not by that of the masses $M_{1}'$, $M_{2}'$,~\dots\Add{,} $M_{\alpha}'$, equation~\Eq{(60)} gives \[ d\Phi' = \frac{dU' + p\, dV'}{\theta}, \] \PageSep{184} and, therefore, \[ d\Psi' = \frac{U' + pV'}{\theta^{2}}\, d\theta - \frac{V'\, dp}{\theta}, \] whence \[ \frac{\dd \Psi'}{\dd \theta} = \frac{U' + pV'}{\theta^{2}}, \quad\text{and}\quad \frac{\dd \Psi'}{\dd p} = -\frac{V'}{\theta}, \] and for the system, as the sum of all the phases, \[ \frac{\dd \Psi}{\dd \theta} = \frac{U + pV}{\theta^{2}}, \quad\text{and}\quad \frac{\dd \Psi}{\dd p} = -\frac{V}{\theta}. \Tag{(150)} \] \Section{211.} These relations may be used to determine how the equilibrium depends on the temperature and pressure. For this purpose we shall distinguish between two different kinds of infinitely small changes. The notation~$\delta$ will refer, as hitherto, to a change of the masses $M_{1}'$, $M_{2}'$,~\dots\Add{,} $M_{\alpha}^{\beta}$, consistent with the given external conditions, and, therefore, consistent with the equations~\Eq{(148)}, temperature and pressure being kept constant, \ie\ $\delta \theta = 0$ and $\delta p = 0$. The state, to which this variation leads, need not be one of equilibrium, and the equations~\Eq{(149)} need not, therefore, apply to it. The notation~$d$, on the other hand, will refer to a change from one state of equilibrium to another, only slightly different from it. All external conditions, including temperature and pressure, may be changed in any arbitrary manner. The problem is now to find the conditions of equilibrium of this second state, and to compare them with those of the original state. Since the condition of equilibrium of the first state is \[ \delta \Psi = 0, \] the condition for the second state is \[ \delta (\Psi + d\Psi) = 0, \] hence \[ \delta\, d\Psi = 0\Add{.} \Tag{(151)} \] \PageSep{185} But \[ d\Psi = \frac{\dd \Psi}{\dd \theta}\, d\theta + \frac{\dd \Psi}{\dd p}\, dp + \tsum^{\beta} \frac{\dd \Psi'}{\dd M_{1}'}\, dM_{1}' + \frac{\dd \Psi'}{\dd M_{2}'}\, \Erratum{dM_{2}}{dM_{2}'} + \dots \] where $\tsum$~denotes the summation over all the $\beta$~phases of the system, while the summation over the $\alpha$~constituents of a single phase is written out at length. This becomes, by~\Eq{(150)}, \[ d\Psi = \frac{U + pV}{\theta^{2}} - \frac{V}{\theta}\, dp + \tsum^{\beta} \frac{\dd \Psi'}{\dd M_{1}'}\, dM_{1}' + \frac{\dd \Psi'}{\dd M_{2}'}\, dM_{2}' + \dots\Add{.} \] The condition of equilibrium~\Eq{(151)} therefore becomes \begin{multline*} \frac{\delta U + p\, \delta V}{\theta^{2}}\, d\theta - \frac{\delta V}{\theta}\, dp + \tsum^{\beta} dM_{1}' \, \delta\frac{\dd \Psi'}{\dd M_{1}'} \\ + dM_{2}' \, \delta\frac{\dd \Psi'}{\dd M_{2}'} + \dots = 0\Add{.} \Tag{(152)} \end{multline*} All variations of $d\theta$, $dp$, $dM_{1}$, $dM_{2}'$,~\dots disappear because $\delta \theta = 0$ and $\delta p = 0$, and because in the sum \begin{align*} & \frac{\dd \Psi'}{\dd M_{1}'}\, \delta dM_{1}' + \frac{\dd \Psi'}{\dd M_{2}'}\, \delta dM_{2}' + \dots \\ + & \frac{\dd \Psi''}{\dd M_{1}''}\, \delta dM_{1}'' + \frac{\dd \Psi''}{\dd M_{2}''}\, \delta dM_{2}'' + \dots \\ + & \dots \\ +& \frac{\dd \Psi^{\beta}}{\dd M_{1}^{\beta}}\, \delta dM_{1}^{\beta} + \frac{\dd \Psi^{\beta}}{\dd M_{2}^{\beta}}\, \delta dM_{2}^{\beta} + \dots \end{align*} each vertical column vanishes. Taking the first column for example, we have, by~\Eq{(149)}, \[ \frac{\dd \Psi'}{\dd M_{1}'} = \frac{\dd \Psi''}{\dd M_{1}''} = \dots = \frac{\dd \Psi^{\beta}}{\dd M_{1}^{\beta}}, \] \PageSep{186} and also, by~\Eq{(148)}, \begin{multline*} \delta dM_{1}' + \delta dM_{1}'' + \dots + \delta dM_{1}^{\beta} \\ = d(\delta M_{1}' + \delta M_{1}'' + \dots + \delta M_{1}^{\beta}) = 0. \end{multline*} Furthermore, since, by the first law, $\delta U + p\, \delta V$ represents~$Q$, the heat absorbed by the system during the virtual change, the equation~\Eq{(152)} may also be written \[ \frac{Q}{\Typo{\theta_{2}}{\theta^{2}}}\, d\theta - \frac{\delta V}{\theta}\, dp + \tsum^{\beta} dM_{1}'\, \delta \frac{\dd \Psi'}{\dd M_{1}'} + \Erratum{dM_{2}\, \delta \frac{\dd \Psi''}{\dd M_{2}'}}{dM_{2}'\, \delta \frac{\dd \Psi'}{\dd M_{2}'}} + \dots = 0. \Tag{(153)} \] This equation shows how the equilibrium depends on the temperature, and the pressure, and on the masses of the independent constituents of the system. It shows, in the first place, that the influence of the temperature depends essentially on the heat effect which accompanies a virtual change of state. If this be zero, the first term vanishes, and a change of temperature does not disturb the equilibrium. If $Q$~change sign, the influence of the temperature is also reversed. It is quite similar with regard to the influence of the pressure, which, in its turn, depends essentially on the change of volume,~$\delta V$, produced by a virtual isothermal and isopiestic change of state. \Section{212.} We shall now apply the equation~\Eq{(153)} to several special cases; first, to those of perfect heterogeneous equilibrium, which are characterized (\SecRef{206}) by the relation \[ \beta = \alpha + 1. \] The internal nature of all the phases, including the pressure, is determined by the temperature alone. An isothermal, infinitely slow compression, therefore, changes only the total masses of the phases, but does not change either the composition or the pressure. We shall choose a change of this kind as the virtual change of state. In this special case it leads to a new state of equilibrium. The internal nature of all the phases, as well as the temperature and pressure, \PageSep{187} remain constant, and therefore the variations of the functions $\dfrac{\dd \Psi'}{\dd M_{1}'}$, $\dfrac{\dd \Psi'}{\dd M_{2}'}$,~\dots are all equal to zero, since these quantities depend only on the nature of the phases. The equation~\Eq{(153)} therefore becomes \[ \frac{dp}{d\theta} = \frac{Q}{\theta\, \delta V}. \Tag{(154)} \] This means that the heat effect in a variation that leaves the composition of all phases unchanged, divided by the change of volume of the system and by the absolute temperature, gives the rate of change of the equilibrium pressure with the temperature. Where application of heat increases the volume, as in the case of evaporation, the equilibrium pressure increases with temperature; in the opposite case, as in the melting of ice, it decreases with increase of temperature. \Section{213.} In the case of one independent constituent ($\alpha = 1$, \index{Latent heat!from phase rule|(}% and $\beta = 2$), equation~\Eq{(154)} leads immediately to the laws discussed at length in the preceding chapter; namely, those concerning the heat of vaporization, of fusion, and of sublimation. If, for instance, the liquid form the first phase, the vapour the second phase, and~$L$ denote the heat of vaporization per unit mass, we have \begin{align*} Q &= L\, \delta M''\Add{,} \\ \delta V &= (v'' - v')\, \delta M''\Add{,} \end{align*} where $v'$~and $v''$ are the specific volumes of liquid and vapour, and $\delta M''$~the mass of vapour formed during the isothermal and isopiestic change of state. Hence, by~\Eq{(154)}, \[ L = \theta\, \frac{dp}{d\theta} (v'' - v'), \] which is identical with the equation~\Eq{(111)}. This, of course, applies to chemical changes as well, \PageSep{188} \index{Ammonium carbamate, evaporation of}% \index{Ammonium chloride, evaporation of}% whenever the system under consideration contains one constituent in two distinct phases; for example, to the vaporization of ammonium chloride (first investigated with regard to this law by Horstmann), which decomposes into hydrochloric \index{Horstmann}% acid and ammonia; or to the vaporization of ammonium carbamate, which decomposes into ammonia and carbon dioxide. Here $L$~of our last equation denotes the heat of dissociation, and $p$~the dissociation pressure, which depends only on the temperature. \Section{214.} We shall also consider the perfect heterogeneous equilibrium of two independent constituents ($\alpha = 2$, $\beta = 3$); for example, water (suffix~$1$) and a salt (suffix~$2$) in three phases; the first, an aqueous solution ($M_{1}'$~the mass of the water, $M_{2}'$~that of the salt); the second, water vapour (mass~$M_{1}''$); the third, solid salt (mass~$M_{2}'''$). For a virtual change, therefore, \[ \delta M_{1}' + \delta M_{1}'' = 0,\quad\text{and}\quad \delta M_{2}' + \delta M_{2}''' = 0. \] According to the phase rule, the concentration of the solution $\left(\dfrac{M_{2}'}{M_{1}'} = c\right)$, as well as the vapour pressure~($p$) is a function of the temperature alone. By~\Eq{(154)}, the heat absorbed ($\theta$,~$p$,~$c$ remaining constant) is \[ Q = \theta · \frac{dp}{d\theta} · \delta V. \Tag{(155)} \] Let the virtual change consist in the evaporation of a \index{Evaporation!of ammonium carbamate}% \index{Evaporation!of ammonium chloride}% small quantity of water, \[ \delta M_{1}'' = - \delta M_{1}'. \] Then, since the concentration also remains constant, the quantity of salt \[ \delta M_{2}''' = -\delta M_{2}' = -c\, \delta M_{1}' = c\, \delta M_{1}'' \] is precipitated from the solution. All variations of mass have here been expressed in terms of~$\delta M_{1}''$. \PageSep{189} The total volume of the system \[ V = v' (M_{1}' + M_{2}') + v'' M_{1}'' + v''' M_{2}''', \] where $v'$,~$v''$, and~$v'''$ are the specific volumes of the phases, is increased by \begin{align*} \delta V &= v' (\delta M_{1}' + \delta M_{2}') + v''\, \delta M_{1}'' + v'''\, \delta M_{2}'''\Add{,} \\ \delta V &= \bigl[(v'' + cv''') - (1 + c) v'\bigr]\, \delta M_{1}''. \Tag{(156)} \end{align*} If $L$~be the quantity of heat that must be applied to evaporate unit mass of water from the solution, and to precipitate the corresponding quantity of salt, under constant pressure, temperature, and concentration, then the equation~\Eq{(155)}, since \[ Q = L\, \delta M_{1}'', \] becomes \[ L = \theta\, \frac{dp}{d\theta} \bigl(v'' + cv''' - (1 + c)v'\bigr). \] A useful approximation is obtained by neglecting $v'$~and $v'''$, the specific volumes of the liquid and solid, in comparison with~$v''$, that of the vapour, and considering the latter as a perfect gas. By~\Eq{(14)}, \[ v'' = \frac{R}{m} · \frac{\theta}{p} \] ($R = \text{gas constant}$, $m = \text{the molecular weight of the vapour}$) and we obtain \[ L = \frac{R}{m} \theta^{2} · \frac{d \log p}{d\theta}. \Tag{(157)} \] \Section{215.} Conversely, $L$~is at the same time the quantity of heat given out when unit mass of water vapour combines, at constant temperature and pressure, with the quantity of salt necessary to form a saturated solution. This process may be accomplished directly, or in two steps, viz.\ by condensing unit mass of water vapour into pure water, and then dissolving the salt in the water. According to the first law of thermodynamics, since the initial and final states are the \index{Latent heat!from phase rule|)}% \PageSep{190} same in both cases, the sum of the heat given out and the \index{Heat!of solution}% work done is the same. In the first case the heat given out is~$L$, the work done, $-p\, \dfrac{\delta V}{\delta M_{1}''}$; and the sum of these, by the approximation used above, is \[ \frac{R}{m} \theta^{2} · \frac{d \log p}{d\theta} - pv''. \Tag{(158)} \] To calculate the same sum for the second case, we must in the first place note that the vapour pressure of a solution \index{Solution!heat of}% is different from the vapour pressure of pure water at the same temperature. It will, in fact, in no case be greater, but smaller, otherwise the vapour would be supersaturated. Denoting the vapour pressure of pure water at the temperature~$\theta$ by~$p_{0}$, then $p < p_{0}$. We shall now bring, by isothermic compression, unit mass of water vapour from pressure~$p$ and specific volume~$v''$ to pressure~$p_{0}$ and specific volume~$v_{0}''$, \ie\ to a state of saturation. Work is thereby done on the substance, and heat is given out. The sum of both, which gives the decrease of the energy of the vapour, is zero, if we again assume that the vapour behaves as a perfect gas, \ie\ that its energy remains constant at constant temperature. If we then condense the water vapour of volume~$v_{0}''$, at constant temperature~$\theta$ and constant pressure~$p_{0}$, into pure water, the sum of the heat given out and work spent at this step is, by equation~\Eq{(112)}, \[ \frac{R}{m} \theta^{2} · \frac{d \log p_{0}}{d\theta} - p_{0} v_{0}''. \Tag{(159)} \] No appreciable external effects accompany the further change of the liquid water from pressure~$p_{0}$ to pressure~$p$. If, finally, we dissolve salt sufficient for saturation in the newly formed unit of water, at constant temperature~$\theta$ and constant pressure~$p$, the sum of the heat and work is simply the heat of solution \[ \lambda\Add{.} \Tag{(160)} \] \PageSep{191} By the first law, the sum of \Eq{(159)}~and \Eq{(160)} must be equal to~\Eq{(158)}, \[ \frac{R}{m} \theta^{2} · \frac{\Erratum{d \log p}{d \log p_{0}}}{d\theta} - p_{0} v_{0}'' + \lambda = \frac{R}{m} \theta^{2} · \frac{d \log p}{d\theta} - pv''; \] or, since, by Boyle's law $p_{0} v_{0}'' = pv''$, \[ \lambda = \frac{R}{m} \theta^{2} · \frac{d \log \dfrac{p}{p_{0}}}{d\theta}. \Tag{(161)} \] This formula, first established by Kirchhoff, gives the heat \index{Kirchhoff}% evolved when salt sufficient for saturation is dissolved in $1~\Unit{gr.}$ of pure water. To express~$\lambda$ in calories, $R$~must be divided by the mechanical equivalent of heat,~$J$\@. By~\Eq{(34)}, $\dfrac{R}{J} = 1.97$, and since $m = 18$, we have \[ \lambda = 0.11 \theta^{2} · \frac{d \log \dfrac{p}{p_{0}}}{d\theta}~\Unit{cal}. \] It is further worthy of notice that~$p$, the vapour pressure of a saturated solution, is a function of the temperature alone, since $c$,~the concentration of a saturated solution, changes in a definite manner with the temperature. The quantities neglected in this approximation may, if necessary, be put in without any difficulty. \Section{216.} We proceed now to the important case of two independent constituents in two phases ($\alpha = 2$, $\beta = 2$). We assume, for the present, that both constituents are contained in both phases in appreciable quantity, having the masses $M_{1}'$,~$M_{2}'$ in the first; $M_{1}''$,~$M_{2}''$, in the second phase. The internal variables are the temperature, the pressure, \PageSep{192} and the concentrations of the second constituent in both phases; \[ c' = \frac{M_{2}'}{M_{1}'} \quad\text{and}\quad c'' = \frac{M_{2}''}{M_{1}''}. \Tag{(162)} \] According to the phase rule, two of the variables, $\theta$,~$p$, $c'$,~$c''$, are arbitrary. Equation~\Eq{(153)} leads to the following law regarding the shift of the equilibrium corresponding to any change of the external conditions: \begin{multline*} \frac{Q}{\theta_{2}}\, d\theta - \frac{\delta V}{\theta}\, dp + dM_{1}'\, \delta \frac{\dd \Psi'}{\dd M_{1}'} + dM_{2}'\, \delta \frac{\dd \Psi'}{\dd M_{2}'} \\ + dM_{1}''\, \delta \frac{\dd \Psi''}{\dd M_{1}''} + dM_{2}''\, \delta \frac{\dd \Psi''}{\dd M_{2}''} = 0. \Tag{(163)} \end{multline*} Here, for the first phase, \[ \left. \begin{aligned} \delta \frac{\dd \Psi'}{\dd M_{1}'} &= \frac{\dd^{2} \Psi'}{\dd M_{1}'^{2}}\, \delta M_{1}' + \frac{\dd^{2} \Psi'}{\dd M_{1}'\, \dd M_{2}'}\, \delta M_{2}'\Add{,} \\ \delta \frac{\dd \Psi'}{\dd M_{1}'} &= \frac{\dd^{2} \Psi'}{\dd M_{1}'\, \dd M_{2}'}\, \delta M_{1}' + \frac{\dd^{2} \Psi'}{\dd M_{2}'^{2}}\, \delta M_{2}'\Add{.} \end{aligned} \right\} \Tag{(164)} \] Certain simple relations hold between the derived functions of~$\Psi'$ with respect to $M_{1}'$~and~$M_{2}'$. For, since, by~\Eq{(144)}, \[ \Psi' = M_{1}'\, \frac{\dd \Psi'}{\dd M_{1}'} + M_{2}'\, \frac{\dd \Psi'}{\dd M_{2}'}, \] partial differentiation with respect to $M_{1}'$~and $M_{2}'$ gives \begin{align*} 0 &= M_{1}'\, \frac{\dd^{2} \Psi'}{\dd M_{1}'^{2}} + M_{2}'\, \frac{\dd^{2} \Psi'}{\dd M_{1}'\, \dd M_{2}'}, \\ 0 &= M_{1}'\, \frac{\dd^{2} \Psi'}{\dd M_{1}'\, \dd M_{2}'} + M_{2}'\, \frac{\dd^{2} \Psi'}{\dd M_{2}'^{2}}. \end{align*} If we put, for shortness, \[ M_{1}'\, \frac{\dd^{2} \Psi'}{\dd M_{1}'\, \dd M_{2}'} = \varphi', \Tag{(165)} \] \PageSep{193} a quantity depending only on the nature of the first phase, % [** TN: Manually broke footnote] on $\theta$,~$p$, and~$c'$, and not on the masses $M_{1}'$~and~$M_{2}'$ individually,\footnote {The general integral of $\Psi' = M_{1}'\, \dfrac{\dd \Psi'}{\dd M_{1}'} + M_{2}'\, \dfrac{\dd \Psi'}{\dd M_{2}'}$ \\ is $\Psi' = M_{2}' f\left(\dfrac{M_{1}'}{M_{2}'}\right)$.---\Tr.} we have \[ \left. \begin{aligned} \frac{\dd^{2} \Psi'}{\dd M_{1}'\, \dd M_{2}'} &= \frac{\varphi'}{M_{1}'}\Add{,} \\ \frac{\dd^{2} \Psi'}{\dd M_{1}'^{2}} &= -\frac{M_{2}'}{M_{1}'^{2}} · \varphi'\Add{,} \\ \frac{\dd^{2} \Psi'}{\dd M_{2}'^{2}} &= -\frac{\varphi'}{M_{2}'}\Add{.} \end{aligned} \right\} \Tag{(166)} \] Analogous equations hold for the second phase if we put \[ \varphi'' = M_{1}'' · \frac{\dd \Psi''}{\dd M_{1}''\, \dd M_{2}''}. \] \Section{217.} With respect to the quantities $\varphi'$~and~$\varphi''$ all we can immediately settle is their sign. According to \SecRef{147}, $\Psi$~is a maximum in stable equilibrium if only processes at constant temperature and constant pressure be considered. Hence \[ \delta^{2} \Psi < 0. \Tag{(167)} \] But \[ \Psi = \Psi' + \Psi'', \] whence \[ \delta \Psi = \frac{\dd \Psi'}{\dd M_{1}'}\, \delta M_{1}' + \frac{\dd \Psi'}{\dd M_{2}'}\, \delta M_{2}' + \frac{\dd \Psi''}{\dd M_{1}''}\, \delta M_{1}'' + \frac{\dd \Psi''}{\dd M_{2}''}\, \delta M_{2}'' \] and \begin{multline*} \delta^{2} \Psi = \frac{\dd^{2} \Psi'}{\dd M_{1}'^{2}}\, \delta M_{1}'^{2} + 2 \frac{\dd^{2} \Psi'}{\dd M_{1}'\, \dd M_{2}'}\, \delta M_{1}'\, \delta M_{2}' + \frac{\dd^{2} \Psi'}{\dd M_{2}'^{2}}\, \delta M_{2}'^{2} \\ + \frac{\dd^{2} \Psi''}{\dd M_{1}''^{2}}\, \delta M_{1}''^{2} + 2 \frac{\dd^{2} \Psi''}{\dd M_{1}''\, \dd M_{2}''}\, \delta M_{1}''\, \delta M_{2}'' + \frac{\dd^{2} \Psi''}{\dd M_{2}''^{2}}\, \delta M_{2}''^{2}. \end{multline*} \PageSep{194} If we introduce the quantities $\varphi'$~and~$\varphi''$, then \[ \delta^{2} \Psi = -M_{2}' \varphi' \left(\frac{\delta M_{1}'}{M_{1}'} - \frac{\delta M_{2}'}{M_{2}'}\right)^{2} -M_{2}'' \varphi'' \left(\frac{\delta M_{1}''}{M_{1}''} - \frac{\delta M_{2}''}{M_{2}''}\right)^{2}. \] This relation shows that the inequality~\Eq{(167)} is satisfied, and only then, if both $\varphi'$~and $\varphi''$ are positive. \Section{218.} There are on the whole two kinds of changes possible, according as the first or the second constituent passes from the first to the second phase. We have, for the first, \[ \delta M_{1}' = -\delta M_{1}'';\quad \delta M_{2}' = \delta M_{2}'' = 0; \Tag{(168)} \] and for the second, \[ \delta M_{1}' = \delta M_{1}'' = 0;\quad \delta M_{2}' = -\delta M_{2}''. \] We shall distinguish~$Q$, the heat absorbed, and $\delta V$, the change of volume, in these two cases by the suffixes $1$ and~$2$. In the first case, the law for the displacement of the equilibrium, by \Eq{(163)}, \Eq{(164)}, \Eq{(168)}, \Eq{(166)}, and~\Eq{(162)}, reduces to \[ \frac{Q_{1}}{\theta_{2}}\, d\theta - \frac{\delta_{1}V}{\theta}\, dp - \delta M_{1}'' (\varphi'\, dc' - \varphi''\, dc'') = 0 \] and, introducing for shortness the finite quantities \[ L_{1} = \frac{Q_{1}}{\delta M_{1}''},\quad s_{1} = \frac{\delta_{1}V}{\delta M_{1}''}, \Tag{(169)} \] \ie\ the ratios of the heat absorbed and of the change of volume to the mass of the first constituent, which passes from the first to the second phase, we have \[ \frac{L_{1}}{\theta^{2}}\, d\theta - \frac{s_{1}}{\theta}\, dp - \varphi'\, dc' + \varphi''\, dc'' = 0. \Tag{(170)} \] \PageSep{195} Similarly, for the second constituent passing into the second phase, we get \[ \frac{L_{2}}{\theta^{2}}\, d\theta - \frac{s_{2}}{\theta}\, dp - \varphi'\, \frac{dc'}{c'} + \varphi''\, \frac{dc''}{c''} = 0. \Tag{(171)} \] These are the two relations connecting the four differentials $d\theta$, $dp$, $dc'$, $dc''$ in any displacement of the equilibrium. \Section{219.} To show the application of these laws, let us consider a mixture of two liquids (water and alcohol) in two phases, the first a liquid, the second a vapour. The phase rule leaves two of the variables $\theta$, $p$, $c'$, $c''$ arbitrary. The pressure~$p$, and the concentration~$c''$ of the vapour, for instance, are determined by the temperature and the concentration~$c'$ of the liquid mixture. Accordingly, for any changes $d\theta$~and $dc'$ we have, from \Eq{(170)} and~\Eq{(171)}, \begin{align*} dp &= \frac{\left(\dfrac{c''}{c'} - 1\right) \theta^{2} \varphi'\, dc' + (L_{1} + c'' L_{2})\, d\theta}{(s_{1} + c'' s_{2}) \theta},\displaybreak[0] \\ dc'' &= \frac{\left(\dfrac{1}{s_{1}} + \dfrac{1}{c's_{2}}\right) \varphi'\, dc' - \left(\dfrac{L_{1}}{s_{1}} - \dfrac{L_{2}}{s_{2}}\right) · \dfrac{d\theta}{\theta^{2}}}{\left(\dfrac{1}{s_{1}} + \dfrac{1}{c'' s_{2}}\right) \varphi''}. \end{align*} Of the many conclusions to be drawn from these equations, we mention only the following: Along an isotherm ($d\theta = 0$) the equations become \begin{align*} dp &= \frac{\left(\dfrac{c''}{c'} - 1\right) \theta \varphi'\, dc'}{s_{1} + c'' s_{2}}, \Tag{(172)}\displaybreak[0] \\ dc'' &= \frac{\left(\dfrac{1}{s_{1}} + \dfrac{1}{c's_{2}}\right)}{\left(\dfrac{1}{s_{1}} + \dfrac{1}{c'' s_{2}}\right)} · \frac{\varphi'}{\varphi''}\, dc'. \Tag{(173)} \end{align*} The vapour pressure~$p$ may rise or fall with increasing \PageSep{196} concentration. When $p$~shows a maximum or minimum value, as it does according to Konowalow for a $77 : 23$~mixture \index{Konowalow}% of propyl alcohol and water, then $\dfrac{dp}{dc'}$~vanishes, and, from equation~\Eq{(172)}, $c' = c''$, \ie\ the percentage composition of the liquid and the vapour is the same, or the liquid boils at constant concentration. But if, along an isotherm, $p$~varies with~$c'$, the concentration of the vapour will differ from that of the liquid; in fact, the concentration of the second constituent in the vapour will be more or less than in the liquid ($c'' >$ or $< c'$), according as the vapour pressure~$p$ rises or falls with increasing concentration. This is an immediate deduction from equation~\Eq{(172)} if we bear in mind that $\varphi'$, $s_{1}$,~$s_{2}$, and~$c''$ are always positive. The equation~\Eq{(173)} shows that along an isotherm the concentration of both phases always changes in the same sense. \Section{220.} In the following applications we shall restrict ourselves to the case in which the second constituent occurs only in the first phase, \[ c'' = 0, \] and \[ \therefore dc'' = 0\Add{.} \Tag{(174)} \] The first constituent which occurs along with the second in the first phase, and pure in the second, will be called the \emph{solvent}; the second, the \emph{dissolved substance}. By~\Eq{(174)}, \index{Solvent}% the equation~\Eq{(171)} is identically satisfied, and from~\Eq{(170)} there remains \[ \frac{L}{\theta^{2}}\, d\theta - \frac{s}{\theta}\, dp - \varphi\, dc = 0, \Tag{(175)} \] if we omit suffixes and dashes for simplicity. We shall take, first, a solution of a \Chg{nonvolatile}{non-volatile} salt in contact with the vapour of the solvent, and investigate the equation~\Eq{(175)} in three directions by keeping in turn the concentration~$c$, the temperature~$\theta$, and the pressure~$p$ constant. \PageSep{197} \Section{221.} \Topic{Concentration Constant: $dc = 0$.}---The relation \index{Boyle and Gay-Lussac}% between the vapour pressure and the temperature is, by~\Eq{(175)}, \[ \left(\frac{\dd p}{\dd \theta}\right)_{c} = \frac{L}{\theta · s}. \Tag{(176)} \] Here $L$~may be called briefly the heat of vaporization of the solution. If, instead of regarding $L$~as the ratio of two infinitely small quantities, we take it to be the heat of vaporization per unit mass of the solvent, then the mass of the solvent must be assumed so large that the concentration is not appreciably altered by the evaporation of unit mass. The quantity~$s$ may generally be put $= v$, the specific volume of the vapour. Assuming, further, that the laws of Boyle and \Typo{Gay Lussac}{Gay-Lussac} hold for the vapour, we get \[ s = v = \frac{R}{m} · \frac{\theta}{p} \Tag{(177)} \] and, by~\Eq{(176)}, \[ L = \frac{R}{m} \theta^{2} \left(\frac{\dd \log p}{\dd \theta}\right)_{c}. \] On the other hand, $L$~is also the quantity of heat given out when unit mass of the vapour of the solvent combines at constant temperature and pressure with a large quantity of a solution of concentration~$c$. This process may be performed directly, or unit mass of the vapour may be first condensed to the pure solvent and then the solution diluted with it. In the first case the sum of the heat given out and the work spent is \[ L - pv = \frac{R}{m} \theta^{2} \left(\frac{\dd \log p}{\dd \theta}\right)_{c} - pv. \] In the second case, by the method used in \SecRef{215} we obtain, as the sum of the heat given out and the work spent during condensation and dilution, \[ \frac{R}{m} \theta^{2} \frac{d \log p_{0}}{d\theta} - p_{0} v_{0} + \Delta, \] \PageSep{198} where $p_{0}$~is the pressure, $v$~the specific volume of the vapour of the solvent in contact with the pure liquid solvent, $\Delta$~the heat of dilution of the solution, \ie\ the heat given out on \index{Dilution!heat of}% \index{Heat!of dilution}% adding unit mass of the solvent to a large quantity of the solution of concentration~$c$. Both the above expressions being equal according to the first law, we obtain, on applying Boyle's law, \[ \Delta = \frac{R}{m} \theta^{2} \left(\frac{\dd \log \dfrac{p}{p_{0}}}{\dd \theta}\right)_{c}, \Tag{(178)} \] which is Kirchhoff's formula for the heat of dilution. \index{Kirchhoff's formula}% The quantities here neglected, by considering the vapour a perfect gas, and its specific volume large in comparison with that of the liquid, may readily be taken into account when necessary. The similarity of the expressions for~$\Delta$, the heat of dilution, and for~$\lambda$, the heat of saturation~\Eq{(161)}, is only external, since in this case the solution may be of any concentration, and therefore may be differentiated with respect to the temperature, $c$~being kept constant, while in~\Eq{(161)} the concentration of a saturated solution changes with temperature in a definite manner. \Section{222.} Since $\Delta$~is \emph{small} for small values of~$c$ (dilute solutions, \SecRef{97}), then, according to~\Eq{(178)}, the ratio of the vapour pressure of a dilute solution of fixed concentration to the vapour pressure of the pure solvent is practically independent of the temperature (Babo's law). \index{Babo's law}% \index{Laws:!Babo's}% \Section{223.} \Topic{Temperature Constant: $d\theta = 0$.}---The relation between the vapour pressure~($p$) and the concentration~($c$) of the solution is, according to~\Eq{(175)}, \[ \left(\frac{\dd p}{\dd c}\right)_{\theta} = -\frac{\theta\varphi}{s}. \Tag{(179)} \] Neglecting the specific volume of the liquid in comparison \PageSep{199} \index{Lowering!of vapour pressure}% \index{Vapour pressure, lowering of}% with that of the vapour, and considering the latter a perfect gas of molecular weight~$m$, equation~\Eq{(177)} gives \[ \left(\frac{\dd p}{\dd c}\right)_{\theta} = -\frac{mp\theta}{R}, \] or \[ \left(\frac{\dd \log p}{\dd c}\right)_{\theta} = -\frac{m}{R}\, \varphi. \] Since $\varphi$~is always positive (\SecRef{217}), the vapour pressure must decrease with increasing concentration. This proposition furnishes a means of distinguishing between a solution and an emulsion. In an emulsion the number of particles suspended in the solution has no influence on the vapour pressure. So long as the quantity~$\varphi$ is undetermined, nothing further can be stated with regard to the general relation between the vapour pressure and the concentration. \Section{224.} As we have $p = p_{0}$ when $c = 0$ (pure solvent), $p - p_{0}$ \index{Laws:!Wüllner's}% \index{Wüllner's law}% is small for small values of~$c$. We may, therefore, put \[ \frac{\dd p}{\dd c} = \frac{p - p_{0}}{c - 0} = \frac{p - p_{0}}{c}. \] Hence, by~\Eq{(179)}, \[ p - p_{0} = \frac{c\theta\varphi}{s} \Tag{(180)} \] and substituting for~$s$, as in~\Eq{(177)}, the specific volume of the vapour, considered a perfect gas, we get \[ \frac{p - p_{0}}{p} = \frac{cm\varphi}{R}. \Tag{(181)} \] This means that the relative decrease of the vapour pressure is proportional to the concentration of the solution (Wüllner's law). For further particulars, see~\SecRef{270}. \Section{225.} \Topic{Pressure Constant: $dp = 0$.}---The relation \PageSep{200} between the temperature (boiling point) and the concentration is, by~\Eq{(175)}, \[ \left(\frac{\dd \theta}{\dd c}\right)_{p} = \frac{\theta^{2} \varphi}{L}. \Tag{(182)} \] Since $\varphi$~is positive, the boiling point rises with increasing concentration. By comparing this with the formula~\Eq{(179)} for the decrease of the vapour pressure, we find that any solution gives \[ \left(\frac{\dd \theta}{\dd c}\right)_{p} : \left(\frac{\dd p}{\dd c}\right)_{\theta} = -\frac{\theta s}{L}, \] \ie\ for an infinitely small increase of the concentration the rise in the boiling point (at constant pressure) is to the decrease of the vapour pressure (at constant temperature) as the product of the absolute temperature and the specific volume of the vapour is to the heat of vaporization of the solution. Remembering that this relation satisfies the identity \[ \left(\frac{\dd \theta}{\dd c}\right)_{p} : \left(\frac{\dd p}{\dd c}\right)_{\theta} = -\left(\frac{\dd \theta}{\dd p}\right)_{c}, \] we come immediately to the equation~\Eq{(176)}. \Section{226.} Let $\theta_{0}$~be the boiling point of the pure solvent \index{Boiling point, elevation of}% \index{Elevation of boiling point}% ($c = 0$), then, for some values of~$c$, the difference between $\theta$~and~$\theta_{0}$ will be small, and we may put \[ \frac{\dd \theta}{\dd c} = \frac{\theta - \theta_{0}}{c - 0} = \frac{\theta - \theta_{0}}{c}, \] whereby the equation becomes \[ \theta - \theta_{0} = \frac{c\theta^{2} \varphi}{L}. \Tag{(183)} \] This means that the elevation of the boiling point is proportional to the concentration of the solution. For further details, see~\SecRef{269}. \PageSep{201} \Section{227.} Let the second phase consist of the pure solvent in the solid state instead of the gaseous state, as happens in the freezing of an aqueous salt solution or in the precipitation of salt from a saturated solution. In the latter case, in conformity with the stipulations of~\SecRef{220}, the salt will be regarded as the first constituent (the solvent), and water as the second constituent (the dissolved substance). The equation~\Eq{(175)} is then directly applicable, and may be discussed in three different ways. We may ask how the freezing point or the saturation point of a solution of definite concentration changes with the pressure ($dc = 0$); or, how the pressure must be changed, in order that a solution of changing concentration may freeze or become saturated at constant temperature ($d\theta = 0$); or, finally, how the freezing point or the saturation point of a solution under given pressure changes with the concentration ($dp = 0$). In the last and most important case, if we denote the freezing point or the saturation point as a function of the concentration by~$\theta'$, to distinguish it from the boiling point, equation~\Eq{(175)} gives \[ \left(\frac{\dd \theta'}{\dd c}\right)_{p} = \frac{\theta^{2} \varphi}{L}, \] $L$~being the heat absorbed when unit mass of the solvent \index{Heat!of precipitation}% \index{Heat!of solidification}% \index{Precipitation, heat of}% separates as a solid (ice, salt) from a large quantity of the solution of concentration~$c$. Since $L$~is often negative, we may put $L = -L'$ and call $L'$ the \emph{heat of solidification} of \index{Solidification!heat of}% the solution or the \emph{heat of precipitation} of the salt. We have, then, \[ \left(\frac{\dd \theta'}{\dd c}\right)_{p} = -\frac{\theta^{2} \varphi}{L'}. \Tag{(184)} \] The heat of solidification~($L'$) of a salt solution is always positive, hence the freezing point is lowered by an increase of concentration~$c$. On the other hand, if the heat of precipitation~($L'$) of a salt from a solution be positive, the saturation point~$theta'$ is lowered by an increase of the mass of water, or rises with an increase of the concentration of \PageSep{202} the salt. If $L'$~be negative, the saturation point is lowered by an increase of the concentration of the salt. Should we prefer to designate by~$c$, not the amount of water, but the amount of salt in a saturated solution, then, according to the definition of~$c$ in~\Eq{(162)} and of~$\varphi$ in~\Eq{(165)}, we should have $\dfrac{1}{c}$ replacing~$c$ in~\Eq{(184)} and $c\Erratum{\phi}{\varphi}$~replacing~$\varphi$, and therefore \[ \left(\frac{\dd \theta'}{\dd c}\right)_{p} = \frac{\theta^{2} \varphi}{cL'}. \Tag{(185)} \] Here $c$~and $\Erratum{\phi}{\varphi}$ have the same meaning as in equation~\Eq{(184)}, which refers to the freezing point of a solution. \Section{228.} {\Loosen Let $\theta_{0}'$~be the freezing point of the pure solvent ($c = 0$), then, for small values of~$c$, $\theta'$~will be nearly $= \theta_{0}'$, and we may put} \[ \frac{\dd \theta'}{\dd c} = \frac{\theta' - \theta_{0}'}{c - 0} = \frac{\theta' - \theta_{0}'}{c}. \] Equation~\Eq{(184)} then becomes \[ \theta_{0}' - \theta' = \frac{c\theta^{2} \varphi}{L'}\Add{,} \Tag{(186)} \] which means that the lowering of the freezing point is \index{Lowering!of freezing point}% proportional to the concentration. For further particulars, see~\SecRef{271}. \Section{229.} The positive quantity,~$\varphi$, which occurs in all these formulæ, has a definite value for a solution of given $c$,~$\theta$, and~$p$, and is independent of the nature of the second phase. Our last equations, therefore, connect in a perfectly general way the laws regarding the lowering of the vapour pressure, the elevation of the boiling temperature, the depression of the freezing point, and the change of the saturation point. Only one of these phenomena need be experimentally investigated in order to calculate~$\varphi$, and by means of the value thus determined the others may be deduced for the same solution. \PageSep{203} We shall now consider a further case for which $\varphi$~is of \index{Membranes, semipermeable}% \index{Semipermeable membranes}% fundamental importance, viz.\ the state of equilibrium which ensues when the pure liquid solvent forms the second phase, not in contact with a solution, for no equilibrium would thus be possible, but separated from it by a membrane, permeable to the solvent only. It is true that for no solution can perfectly \emph{semipermeable} membranes of this character be manufactured. In fact, the further development of this theory (\SecRef{259}) will exclude them as a matter of principle, for in every case the dissolved substance will also diffuse through the membrane, though possibly at an extremely slow rate. For the present it is sufficient that we may, without violating a law of thermodynamics, assume the velocity of diffusion of the dissolved substance as small as we please in comparison with that of the solvent. This assumption is justified by the fact that semipermeability may be very closely approximated in the case of many substances. The error committed in putting the rate of diffusion of a salt through such a membrane equal to zero, falls below all measurable limits. An exactly similar error is made in assuming that a salt does not evaporate or freeze from a solution, for, strictly speaking, this assumption is not admissible (\SecRef{259}). The condition of equilibrium of two phases separated by a semipermeable membrane is contained in the general thermodynamical condition of equilibrium~\Eq{(145)}, \[ \delta \Psi' + \delta \Psi'' = 0, \Tag{(187)} \] which holds for virtual changes at constant temperature and pressure in each phase. The only difference between this case and free contact is, that the pressures in the two phases may be different. Pressure always means hydrostatic pressure as measured by a manometer. If, in the general equation~\Eq{(76)}, we put \[ W = -p'\, \delta V' - p''\, \delta V'', \] it immediately follows that \Eq{(187)}~is the condition of equilibrium. The further conclusions from~\Eq{(187)} are completely \PageSep{204} analogous to those which are derived, when there is a free surface of contact. Corresponding to~\Eq{(163)} we have for any displacement of the equilibrium \[ \frac{Q}{\theta^{2}}\, d\theta - \frac{\delta V'}{\theta}\, dp' - \frac{\delta V''}{\theta}\, dp'' + dM_{1}'\, \delta \frac{\dd \Psi'}{\dd M_{1}'} + dM_{2}'\, \delta \frac{\dd \Psi'}{\dd M_{2}'} + \dots = 0. \] Since the constituent~$2$ occurs only in the first phase, we get, instead of~\Eq{(175)}, \[ \frac{L}{\theta^{2}}\, d\theta - \frac{s'}{\theta}\, dp' - \frac{s''}{\theta}\, dp'' - \varphi\, dc = 0. \Tag{(188)} \] Here, as in \SecRef{221}, $L$~is the ``\emph{heat of removal}'' of the solvent from the solution, \ie\ the heat absorbed when, at constant temperature and constant pressures $p'$~and~$p''$, unit mass of the solvent passes through the semipermeable membrane from a large quantity of the solution to the pure solvent. The change of volume of the solution during this process is~$s'$ (negative), that of the pure solvent~$s''$ (positive). In the condition of equilibrium~\Eq{(188)}, three of the four variables $\theta$, $p'$, $p''$, $c$ remain arbitrary, and the fourth is determined by their values. Consider the pressure~$p''$ in the pure solvent as given and \index{Pressure!osmotic}% constant, say one atmosphere, then $dp'' = 0$. If, further, we put $d\theta = 0$ and $dc$~not equal to zero, we are then considering solutions in which the concentration varies, but the temperature and the pressure in the pure solvent remains the same. Then, by~\Eq{(188)}, \[ \left(\frac{\dd p'}{\dd c}\right)_{\theta} = -\frac{\theta\varphi}{s'}. \] Since $\varphi > 0$, and $s' < 0$, $p'$~the pressure in the solution increases with the concentration. The difference of the pressures in the two phases, $p' - p'' = P$, has been called the \emph{osmotic pressure} of the \index{Osmotic pressure}% solution. Since $p''$~has been assumed constant, we may write \[ \left(\frac{\dd P}{\dd c}\right)_{\theta} = -\frac{\theta\varphi}{s'}. \Tag{(189)} \] \PageSep{205} Thus the laws of the osmotic pressure have also been expressed in terms of~$\varphi$, which controls those of the depression of the freezing point, the elevation of the boiling point, etc. Since $\varphi$~is positive, the osmotic pressure increases with increasing concentration, and also, since $p' - p''$ vanishes when $c = 0$, the osmotic pressure is necessarily positive. For small values of~$c$, \[ \frac{\dd P}{\dd c} = \frac{P - 0}{c - 0} = \frac{P}{c}, \] and $-s'$~is nearly equal to~$v$ the specific volume of the solution. It therefore follows from~\Eq{(189)} that \[ P = \frac{c\theta\varphi}{v}. \Tag{(190)} \] A further discussion of this question will be found in~\SecRef{272}. \Section{230.} In the preceding paragraphs we have expressed the laws of equilibrium of several systems, that fulfil the conditions of \SecRef{220}, in terms of a quantity~$\varphi$ which is characteristic for the thermodynamical behaviour of a solution. Starting from the two equations \Eq{(170)} and \Eq{(171)}, we find that all the relations in question depend on $\varphi'$~and~$\varphi''$. A better insight into the nature of these quantities is gained by extending to the liquid state the idea of the molecule, hitherto applied only to gases. This step is taken in the next two chapters, and it appears that the manner in which the idea applies is uniquely determined by the propositions of thermodynamics, which have been given. \Section{231.} Just as the conditions of equilibrium~\Eq{(170)} and \Eq{(171)} were deduced for two independent constituents in two phases from the general relation~\Eq{(153)}, so in the same way an entirely analogous deduction may be made in the general case. We shall conclude this chapter by giving, briefly, the \PageSep{206} results for a system of $\alpha$~independent constituents in $\beta$~phases. Denoting the concentrations of the independent constituents, relative to one fixed constituent~$1$, by \begin{align*} \frac{M_{2}'}{M_{1}'} &= c_{2}'; & \frac{M_{3}'}{M_{1}'} &= c_{3}'; & \frac{M_{4}'}{M_{1}'} &= c_{4}'; \dots \\ % \frac{M_{2}''}{M_{1}''} &= c_{2}''; & \frac{M_{3}''}{M_{1}''} &= c_{3}''; & \frac{M_{4}''}{M_{1}''} &= c_{4}''; \dots \\ \dots && \dots && \dots \end{align*} the condition that, by any infinitely small change of the system: $d\theta$, $dp$, $dc_{2}'$, $dc_{3}'$, $dc_{4}'$,~\dots\Add{,} $dc_{2}''$, $dc_{3}''$, $dc_{4}''$,~\dots, the equilibrium may remain stable with regard to the passage of the constituent~$1$ from the phase denoted by one dash to the phase denoted by two dashes is \[ \frac{L_{1}}{\theta_{2}}\, d\theta - \frac{s_{1}}{\theta}\, dp + (\varphi_{2}''\, dc_{2}'' - \varphi_{2}'\, dc_{2}') + (\varphi_{3}''\, dc_{3}'' - \varphi_{3}'\, dc_{3}') + \dots = 0, \] where, analogous to~\Eq{(165)}, \begin{align*} \varphi_{2}' &= M_{1}'\, \delta \frac{\dd^{2} \Psi'}{\dd M_{1}'\, \dd M_{2}'}; & \varphi_{3}' &= M_{1}'\, \delta \frac{\dd^{2} \Psi'}{\dd M_{1}'\, \dd M_{3}'}; & \dots \\ % \varphi_{2}'' &= M_{1}''\, \delta \frac{\dd^{2} \Psi''}{\dd M_{1}''\, \dd M_{2}''}; & \varphi_{3}'' &= M_{1}''\, \delta \frac{\dd^{2} \Psi''}{\dd M_{1}''\, \dd M_{3}''}; & \dots \\ \end{align*} and $L_{1}$,~$s_{1}$ denote the heat absorbed, and the increase of volume of the system during the isothermal and isopiestic transference of unit mass of constituent~$1$ from a large quantity of the phase denoted by one dash to a large quantity of the phase denoted by two dashes. The corresponding conditions of equilibrium for any possible passage of any constituent from any one phase to any other phase may be established in the same way. \PageSep{207} \Chapter{IV.}{Gaseous System.} \index{Gaseous system|(}% \index{System!gaseous|(}% \Section{232.} \First{The} relations, which have been deduced from the general condition of equilibrium~\Eq{(79)} for the different properties of thermodynamical equilibria, rest mainly on the relations between the characteristic function~$\Psi$, the temperature, and the pressure as given in the equations~\Eq{(150)}. It will be impossible to completely answer all questions regarding equilibrium until can be expressed in its functional relation to the masses of the constituents in the different phases. The introduction of the molecular weight serves this purpose. Having already defined the molecular weight of a chemically homogeneous gas as well as the number of molecules of a mixture of gases by Avogadro's law, we shall turn first to the investigation of a system consisting of one gaseous phase. The complete solution of the problem consists in expressing $\Psi$~in terms of $\theta$,~$p$, and $n_{1}$,~$n_{2}$, $n_{3}$,~\dots, the number of all the different kinds of molecules in the mixture. Since we have, in general, by~\Eq{(75)}, \[ \Psi = \Phi - \frac{U + pV}{\theta}, \] we are required to express the entropy~$\Phi$, the energy~$U$, and the volume~$V$ as functions of the above independent variables. This can be done, in general, on the assumption that the mixture obeys the laws of perfect gases. Such a restriction will not, in most cases, lead to appreciable errors. Even this assumption may be set aside by special measurement of the \PageSep{208} quantities $\Phi$,~$U$, and~$V$, as is given later. For the present, however, perfect gases will be assumed. \Section{233.} The laws of Boyle, \Typo{Gay Lussac}{Gay-Lussac}, and Dalton determine the volume of the mixture, for equation~\Eq{(16)} gives \[ V = \frac{R\theta}{p} (n_{1} + n_{2} + \dots) = \frac{R\theta}{p} \tsum n_{1}. \Tag{(191)} \] By the first law of thermodynamics, the energy~$U$ of a mixture of gases is given by the energies of its constituents, for, according to this law, the energy of the system remains unchanged, no matter what internal changes take place, provided there are no external effects. Experience shows that when diffusion takes place between a number of gases at constant temperature and pressure, neither does the volume change, nor is heat absorbed or evolved. The energy of the system, therefore, remains constant during the process. Hence, the energy of a mixture of perfect gases is the sum of the energies of the gases at the same temperature and pressure. Now the energy~$U_{1}$ of $n_{1}$~molecules of a perfect gas depends only on the temperature; it is, by~\Eq{(35)}, \[ U_{1} = \tsum n_{1} (c_{v_{1}}\theta + h_{1}), \Tag{(192)} \] where $c_{v_{1}}$~is the molecular heat of the gas at constant volume, and $h_{1}$~is a constant. Hence the total energy of the mixture is \[ U = \tsum n_{1} (c_{v_{1}}\theta + h_{1}). \Tag{(193)} \] \Section{234.} We have now to determine the entropy~$\Phi$ as a function of $\theta$,~$p$, and $n_{1}$,~$n_{2}$,~\dots the number of molecules. $\Phi$,~in so far as it depends on $\theta$~and~$p$, may be calculated from the equation~\Eq{(60)}, \[ d\Phi = \frac{dU + p\, dV}{\theta}, \] \PageSep{209} where the differentials correspond to variations of $\theta$~and~$p$, but not of the number of molecules. Now, by~\Eq{(193)}, \index{Energy!of gas mixture}% \index{Entropy!of gas mixture|(}% \index{Gas mixture!energy of}% \index{Gas mixture!entropy of|(}% \begin{align*} dU &= \tsum n_{1} c_{v_{1}}\, d\theta, \\ \intertext{and, by~\Eq{(191)},} dV &= R \tsum n_{1}\, d\left(\frac{\theta}{p}\right), \\ \therefore d\Phi &= \tsum n_{1}\, \left(c_{v_{1}}\, \frac{d\theta}{\theta} + \frac{R\, d\theta}{\theta} - \frac{R\, dp}{p}\right), \end{align*} and, by integration, \[ \Phi = \tsum n_{1} \left(c_{v_{1}} \log \theta + R \log \frac{\theta}{p}\right) + C. \Tag{(194)} \] The constant of integration~$C$ is independent of $\theta$~and~$p$, but may depend on the composition of the mixture, \ie\ on the numbers $n_{1}$,~$n_{2}$, $n_{3}$,~\dots. The investigation of this relation forms the most important part of our problem. The determination of the constant is not, in this case, a matter of definition. It can only be determined by applying the second law of thermodynamics to a reversible process which brings about a change in the composition of the mixture. A reversible process produces a definite change of the entropy which may be compared with the simultaneous changes of the number of molecules, and thus the relation between the entropy and the composition determined. If we select a process devoid of external effects either in work or heat, then the entropy remains constant during the whole process. We cannot, however, use the process of diffusion, which leads to the value of~$U$; for diffusion, as might be expected, and as will be shown in~\SecRef{238}, is irreversible, and therefore leads only to the conclusion that the entropy of the system is thereby increased. There is, however, a reversible process at our disposal, which will change the composition of the \PageSep{210} mixture, viz.\ the separation by a semipermeable membrane, as introduced and established in~\SecRef{229}. \Section{235.} Before we can apply a semipermeable membrane to the purpose in hand, we must acquaint ourselves with the nature of the thermodynamical equilibrium of a gas in contact with both sides of a membrane permeable to it. The membrane will act like a bounding wall to those gases to which it is impermeable, and will, therefore, not introduce any special conditions. Experience shows that a gas on both sides of a membrane permeable to it is in equilibrium when its partial pressures (\SecRef{18}) are the same on both sides, quite independent of the other gases present. This proposition is neither axiomatic nor a necessary consequence of the preceding considerations, but it commends itself by its simplicity, and has been confirmed without exception in the few cases accessible to direct experiment. A test of this kind may be established as follows: Platinum foil at a white heat is permeable to hydrogen, but impermeable to air. If a vessel having a platinum wall be filled with pure hydrogen, and hermetically sealed, and the platinum be then heated, the hydrogen must completely diffuse out against atmospheric pressure. As the air cannot enter, the vessel must finally become completely exhausted.\footnote {This inference was tested by me in the Physical Institute of the University of Munich in 1883, and was confirmed within the limits of experimental error as far as the actual deviation from ideal conditions might lead one to expect. As this experiment has not been published anywhere, I shall briefly describe it here. A glass tube of about $5~\Unit{mm.}$ internal diameter, blown out to a bulb at the middle, was provided with a stop-cock at one end. To the other end a platinum tube $10~\Unit{cm.}$ long was fastened, and closed at the end. The whole tube was exhausted by the mercury pump, filled with hydrogen at ordinary atmospheric pressure, and then closed. The closed end of the platinum portion was then heated in a horizontal position by a Bunsen burner. The connection between the glass and platinum tubes having been made by means of sealing-wax, had to be kept cool by a continuous current of water to prevent the softening of the wax. After four hours the tube was taken from the flame, cooled to the temperature of the room, and the stop-cock opened under mercury. The mercury rose rapidly, almost completely filling the tube, proving that the tube had been very nearly exhausted.} \PageSep{211} \Section{236.} We shall make use of the properties of semipermeable membranes to separate in a reversible and simple manner the constituents of a gas mixture. Let us consider the following example:--- Let there be four pistons in a hollow cylinder, two of them, $A$~and~$A'$, in fixed positions; two, $B$~and~$B'$, movable in such a way that the distance~$BB'$ remains constant, and equal to~$AA'$. This is indicated by the brackets in \Fig{5}. \Figure{5}% Further, let $A'$~(the bottom), and $B$~(the cover) be impermeable to any gas, while $A$~is permeable only to one gas~$(1)$, and $B'$~only to another one~$(2)$. The space above~$B$ is a vacuum. At the beginning of the process the piston~$B$ is close to~$A$, therefore $B'$~close to~$A'$, and the space between them contains a mixture of the two gases ($1$~and~$2$). The connected pistons $B$~and $B'$ are now very slowly raised. The gas~$1$ will pass into the space opening up between $A$ and~$B$, and the gas~$2$ into that between $A'$ and~$B'$. Complete separation will have been effected when $B'$~is in contact with~$A$. \PageSep{212} We shall now calculate the external work of this process. The pressure on the movable piston~$B$ consists only of the pressure of the gas~$1$, upwards, since there is a vacuum above~$B$; and on the other movable piston,~$B'$, there is only the partial pressure of the same gas, which acts downwards. According to the preceding paragraph both these pressures are equal, and since the paths of $B$ and~$B'$ are also equal, the total work done on the pistons is zero. If no heat be absorbed or given out, as we shall further assume, the energy of the system remains constant. But, by~\Eq{(193)}, the energy of a mixture of gases depends, like that of pure gases, on the temperature alone, so the temperature of the system remains constant throughout. Since this infinitely slow process is reversible, the entropy in the initial and final states is the same, if there are no external effects. Hence, the entropy of the mixture is equal to the sum of the entropies which the two gases would have, if at the same temperature each by itself occupied the whole volume of the mixture. This proposition may be easily extended to a mixture of any number of gases. \emph{The entropy of a mixture of gases is the sum of the entropies which the individual gases would have, if each at the same temperature occupied a volume equal to the total volume of the mixture.} This proposition was first established by Gibbs. \index{Gibbs}% \Section{237.} The entropy of a perfect gas of mass~$M$ and molecular weight~$m$ was found to be~\Eq{(52)} \[ M \left(\frac{c_{v}}{m} \log \theta + \frac{R}{m} \log v + \const\right), \] where $c_{v}$~is the molecular heat at constant volume, as in~\Eq{(192)}. By the gas laws~\Eq{(14)}, the volume of unit mass is \[ v = \frac{R}{m} · \frac{\theta}{p}, \] whence the entropy is \[ n (c_{v} \log \theta + R \log \frac{\theta}{p} + k), \Tag{(195)} \] \PageSep{213} where $n = \dfrac{M}{m}$, the number of molecules, and the constant~$k$ includes the term $\log \dfrac{R}{m}$. Hence, according to Gibbs's proposition, the entropy of the mixture is \[ \Phi = \tsum n_{1} (c_{v_{1}} \log \theta + R \log \frac{\theta}{p_{1}} + k_{1}), \] $p_{1}$~being the partial pressure of the first gas in the mixture. Now, by~\Eq{(8)}, the pressure of the mixture is the sum of the partial pressures, $\tsum p_{1} = p$, and, by \SecRef{40}, the partial pressures are proportional to the number of molecules of each gas, \[ p_{1} : p_{2} : \dots = n_{1} : n_{2} : \dots\Add{.} \] Hence \begin{align*} p_{1} &= \frac{n_{1}}{n_{1} + n_{2} + \dots}\, p\Add{,} \\ p_{2} &= \frac{n_{2}}{n_{1} + n_{2} + \dots}\, p\Add{,} \\ & \dots\Add{,} \end{align*} or, if we introduce the \emph{concentrations} of the different gases in the mixture, \begin{align*} c_{1} &= \frac{n_{1}}{n_{1} + n_{2} + \dots}; & c_{2} &= \frac{n_{2}}{n_{1} + n_{2} + \dots}; \\ p_{1} &= c_{1} p; & p_{2} &= c_{2} p. \Tag{(196)} \end{align*} Thus the expression for the entropy of a mixture as a function of $\theta$,~$p$, and $n$~the number of molecules, finally becomes \[ \Phi = \tsum n_{1} (c_{v_{1}} \log \theta + R \log \frac{\theta}{pc_{1}} + k_{1}). \Tag{(197)} \] Comparing this expression with the value of the entropy of \PageSep{214} a gas mixture given by~\Eq{(194)}, we see that the constant of \index{Gas mixture!entropy of|)}% integration which was left undetermined is \[ C = \tsum n_{1} (k_{1} - R \log c_{1})\Add{.} \Tag{(198)} \] \Section{238.} Knowing the value of the entropy of a gas \index{Diffusion!increase of entropy by}% \index{Diffusion!irreversible}% \index{Entropy!increase of, by diffusion}% \index{Irreversible diffusion}% mixture, we may answer the question which we discussed in~\SecRef{234}, whether and to what extent the entropy of a system of gases is increased by diffusion. Let us take the simplest case, that of two gases, the number of molecules being $n_{1}$~and $n_{2}$, diffusing into one another under common and constant pressure and temperature. Before diffusion begins, the entropy of the system is the sum of the entropies of the gases, by~\Eq{(195)}, \[ n_{1} (c_{v_{1}} \log \theta + R \log \frac{\theta}{\Typo{p}{p_{1}}} + k_{1}) + n_{2} (c_{v_{2}} \log \theta + R \log \frac{\theta}{\Typo{p}{p_{2}}} + k_{2}). \] After diffusion it is, by~\Eq{(197)}, \[ n_{1} (\Erratum{c_{v_{2}}}{c_{v_{1}}} \log \theta + R \log \frac{\theta}{pc_{1}} + k_{1}) + n_{2} (c_{v_{2}} \log \theta + R \log \frac{\theta}{pc_{2}} + k_{2}). \] Therefore, the change of the entropy of the system is, by~\Eq{(196)}, \[ -n_{1} R \log c_{1} - n_{2} R \log c_{2}\Add{,} \] an essentially positive quantity. This shows that diffusion is always irreversible. It also appears that the increase of the entropy depends solely on the number of the molecules $n_{1}$,~$n_{2}$, and not on the nature---\eg\ the molecular weight, of the diffusing gases. The increase of the entropy does not depend on whether the gases are chemically alike or not. By making the two gases the same, there is evidently no increase of the entropy, \index{Entropy!of gas mixture|)}% since no change of state ensues. It follows that the chemical difference of two gases, or, in general, of two substances, cannot be represented by a continuous variable; but that \PageSep{215} here we can speak only of a discontinuous relation, either of equality or inequality. This fact involves a fundamental distinction between chemical and physical properties, since the latter may always be regarded as continuous. \Section{239.} The values of the entropy~\Eq{(197)}, the energy~\Eq{(193)}, and the volume~\Eq{(191)}, substituted in~\Eq{(75)}, give the function~$\Psi$, \[ \Psi = \tsum n_{1} (c_{v_{1}} \log \theta + R \log \frac{\theta}{pc_{1}} + k_{1} - c_{v_{1}} - \frac{h_{1}}{\theta} - R); \] or, putting the quantity, which depends on $p$~and~$\theta$, and not on the number of molecules, \begin{gather*} c_{v_{1}} \log \theta - \frac{h_{1}}{\theta} + R \log \frac{\theta}{p} + k_{1} - c_{v_{1}} - R = \varphi_{1}, \Tag{(199)} \\ \Psi = \tsum n_{1} (\varphi_{1} - R \log c_{1})\Add{.} \end{gather*} \Section{240.} This enables us to establish the condition of \index{Condition of complete reversibility!of a gas mixture|(}% equilibrium. If in a gas mixture a chemical change, which \index{Equilibrium!of gas mixture}% changes the number of molecules $n_{1}$,~$n_{2}$,~\dots by $\delta n_{1}$,~$\delta n_{2}$,~\dots be possible, then such a change will not take place if the condition of equilibrium~\Eq{(79)} be fulfilled, \ie\ if, when $\delta \theta = 0$ and $\delta p = 0$, \[ \delta \Psi = 0, \] or \[ \tsum (\varphi_{1} - R \log c_{1})\, \delta n_{1} + \tsum n_{1}\, \delta(\varphi_{1} - R \log c_{1}) = 0. \Tag{(200)} \] The quantities $\varphi_{1}$, $\varphi_{2}$\Add{,}~\dots depend on $\theta$~and $p$~only, therefore \[ \delta \varphi_{1} = \delta \varphi_{2} = \dots = 0. \] Further, \[ n_{1}\, \delta \log c_{1} + n_{2}\, \delta \log c_{2} + \dots = \frac{n_{1}}{c_{1}}\, \delta c_{1} + \frac{n_{2}}{c_{2}}\, \delta c_{2} + \dots \] and, by~\Eq{(196)}, \[ = (n_{1} + n_{2} + \dots)(\delta c_{1} + \delta c_{2} + \dots) = 0, \] since \[ c_{1} + c_{2} + \dots = 1. \] \PageSep{216} The condition of equilibrium, therefore, reduces to \[ \tsum (\varphi_{1} - R \log c_{1})\, \delta n_{1} = 0. \] Since this equation does not involve the absolute values of the variations~$\delta n_{1}$, but only their ratios, we may put \[ \delta n_{1} : \delta n_{2} : \dots = \nu_{1} : \nu_{2} : \dots \Tag{(201)} \] and take $\nu_{1}$,~$\nu_{2}$\Add{,}~\dots to denote the number of molecules simultaneously passing into the mass of each constituent. They are simple integers, positive or negative, according as the gas in question is forming, or is being used up in the formation of others. The condition of equilibrium now becomes \[ \tsum (\varphi_{1} - R \log c_{1}) \nu_{1} = 0, \] or \[ \nu_{1} \log c_{1} + \nu_{2} \log c_{2} + \dots = \frac{\nu_{1} \varphi_{1} + \nu_{2} \varphi_{2} + \dots}{R}. \] The right-hand side of the equation depends only on temperature and pressure~\Eq{(199)}. The equation gives a definite relation between the concentrations of the different kinds of molecules for given temperature and pressure. \Section{241.} We shall now substitute the values of $\varphi_{1}$,~$\varphi_{2}$,~\dots\Add{.} If, for shortness, we put the constants \begin{align*} \frac{\tsum \nu_{1} (k_{1} - c_{v_{1}} - R)}{R} &= \log a\Add{,} \\ \frac{\tsum \nu_{1} h_{1}}{R} = b\Add{,} \Tag{(202)} \\ \frac{\tsum \nu_{1} c_{v_{1}}}{R} = c\Add{,} \Tag{(203)} \\ \end{align*} \PageSep{217} then \[ \nu_{1} \log c_{1} + \nu_{2} \log c_{2} + \dots = \log a + (\nu_{1} + \nu_{2} + \dots) \log \frac{\theta}{p} - \frac{b}{\theta} + c \log \theta, \] or \begin{align*} c_{1}^{\nu_{1}} c_{2}^{\nu_{2}} \dots &= a\left(\frac{\theta}{p}\right)^{\nu_{1} + \nu_{2} + \dots} e^{-\efrac{b}{\theta}} \theta^{c}\Add{,} \\ \prod c_{1}^{\nu_{1}} &= a\left(\frac{\theta}{p}\right)^{\tsum \nu_{1}} e^{-\efrac{b}{\theta}} \theta^{c}\Add{.} \end{align*} \Section{242.} This condition may be further simplified by making use of the experimental fact (\SecRef{50}) that the atomic heat of an element remains unchanged in its combinations. By equation~\Eq{(203)} $Rc$~is the change of the sum of the molecular heats of all the molecules of the system during the reaction. The sum of the molecular heats, however, being the sum of the atomic heats, remains unchanged, hence $c = 0$, and the equation becomes \[ \prod c_{1}^{\nu_{1}} = a e^{-\efrac{b}{\theta}} \left(\frac{\theta}{p}\right)^{\tsum \nu_{1}}. \] \Section{243.} According to this equation the influence of the pressure on the equilibrium depends entirely on the number~$\tsum \nu_{1}$, \index{Equilibrium!of gas mixture}% which gives the degree to which the total number of molecules, therefore also the volume of the mixture, is increased by the reaction considered. Where the volume remains unchanged, as, \eg, in the dissociation of hydriodic acid, considered below, the equilibrium is independent of the pressure. The influence of the temperature depends further on the constant~$b$, which is closely connected with the heat effect of the reaction. For, by the first law, \[ Q = \delta U + p\, \delta V, \] which, by \Eq{(193)} and~\Eq{(191)}, $\theta$~and $p$ being constant, becomes \[ Q = \tsum (c_{v_{1}} \theta + h_{1} + R\theta)\, \delta n_{1}. \] \index{Condition of complete reversibility!of a gas mixture|)}% \PageSep{218} If we refer the heat effect to the finite numbers~$\nu$, instead of the infinitely small numbers~$\delta n$, then the heat absorbed is: \[ L = \tsum (c_{v_{1}} \theta + h_{1} + R\theta) \nu_{1}, \] and by \Eq{(202)} and~\Eq{(203)}, again putting $c = 0$, \[ L = Rb + R\theta \tsum \nu_{1}, \] or, \[ L = 1.97 (b + \theta \tsum \nu_{1})~\Unit{cal.} \] The term containing~$b$ refers to the heat spent in the increase of the internal energy; the term containing~$\theta$ to that spent in external work. \Section{244.} Before proceeding to numerical applications, we shall enumerate the principal equations. Suppose that in a gaseous system \[ n_{1}, m_{1};\quad n_{2}, m_{2};\quad n_{3}, m_{3};\ \dots \] ($n$~the number of molecules, $m$~the molecular weight) any chemical change be possible, in which the simultaneous changes of the number of molecules are \[ \delta n_{1} : \delta n_{2} : \delta n_{3} \dots = \nu_{1} : \nu_{2} : \nu_{3} : \dots \] ($\nu$~simple, positive or negative integers) then there will be equilibrium, if the concentrations \[ c_{1} = \frac{n_{1}}{n_{1} + n_{2} + \Add{\dots}};\quad c_{2} = \frac{n_{2}}{n_{1} + n_{2} + \Add{\dots}};\ \dots \] satisfy the condition \begin{align*} c_{1}^{\nu_{1}} c_{2}^{\nu_{2}} c_{3}^{\nu_{3}} \dots &= ae^{-\efrac{b}{\theta}} \left(\frac{\theta}{p}\right)^{\nu_{1} + \nu_{2} + \nu_{3} \dots}\Add{,} \\ \prod c_{1}^{\nu_{1}} &= ae^{-\efrac{b}{\theta}} \left(\frac{\theta}{p}\right)^{\tsum \nu_{1}}\Add{.} \Tag{(204)} \end{align*} \PageSep{219} The heat absorbed during the change at constant temperature and pressure is \[ L = 1.97 \{b + (\nu_{1} + \nu_{2} + \dots) \theta\} = 1.97 (b + \theta \tsum \nu_{1})~\Unit{cal.}\Add{,} \Tag{(205)} \] while the change of volume is \[ s = R(\nu_{1} + \nu_{2} + \dots) \frac{\theta}{p} = R\, \frac{\theta}{p} \tsum \nu_{1}. \Tag{(206)} \] \Section{245.} \Topic{Dissociation of Hydriodic Acid.}---Since hydriodic \index{Dissociation!of hydriodic acid}% \index{Hydriodic acid, dissociation of}% acid gas splits partly into hydrogen and iodine vapour, the system is represented by three kinds of molecules: \[ \ce{$n_{1}$ HI};\quad \ce{$n_{2}$ H2};\quad \ce{$n_{3}$ I2}; \] The concentrations are: \[ c_{1} = \frac{n_{1}}{n_{1} + n_{2} + n_{3}};\quad c_{2} = \frac{n_{2}}{n_{1} + n_{2} + n_{3}};\quad c_{3} = \frac{n_{3}}{n_{1} + n_{2} + n_{3}}. \] The reaction consists in the transformation of two molecules of~\ce{HI} into one of~\ce{H2} and one of~\ce{I2}: \[ \nu_{1} = -2;\quad \nu_{2} = 1;\quad \nu_{3} = 1. \] By~\Eq{(204)}, therefore, in the state of equilibrium, \[ c_{1}^{-2} c_{2}^{1} c_{3}^{1} = ae^{-\efrac{b}{\theta}}\Add{,} \] or \[ \frac{c_{2}c_{3}}{c_{1}^{2}} = \frac{n_{2}n_{3}}{n_{1}^{2}} = ae^{-\efrac{b}{\theta}}\Add{.} \Tag{(207)} \] Since the total number of atoms of hydrogen $(n_{1} + 2n_{2})$ and of iodine $(n_{1} + 2n_{3})$ in the system are supposed to be known, equation~\Eq{(207)} is sufficient for the determination of the three quantities, $n_{1}$,~$n_{2}$, and~$n_{3}$, at any given temperature. The pressure has in this case no influence on the equilibrium. Any two measurements of the degree of dissociation are sufficient for the calculation of $a$~and~$b$. From Bodenstein's \index{Bodenstein}% measurements we have for \[ \theta = 273 + 448 = 712;\quad \frac{c_{2}c_{3}}{c_{1}^{2}} = 0.01984; \] \PageSep{220} and for \[ \theta = 273 + 350 = 623;\quad \frac{c_{2}c_{3}}{c_{1}^{2}} = 0.01494. \] Hence, by~\Eq{(207)}, \[ a = 0.120;\quad b = 1300. \] Thus the equilibrium of any mixture of hydriodic acid, hydrogen, and iodine vapour at any temperature, even when the hydrogen and the iodine are not present in equivalent quantities, is determined by~\Eq{(207)}. Equation~\Eq{(205)} gives the heat of dissociation of two molecules of \index{Dissociation!of iodine vapour}% \index{Iodine vapour, dissociation of}% hydriodic acid into a molecule of hydrogen and a molecule of iodine vapour: \[ L = 1.971 × 1300 = 2560~\Unit{cal.} \] \Section{246.} \Topic{Dissociation of Iodine Vapour.}---At high temperatures iodine vapour appreciably decomposes, leading to a system of two kinds of molecules: \[ \ce{$n_{1}$ I2};\quad \ce{$n_{2}$ I}. \] The concentrations are \[ c_{1} = \frac{n_{1}}{n_{1} + n_{2}};\quad c_{2} = \frac{n_{2}}{n_{1} + n_{2}}. \] The reaction consists in a splitting of the molecule~\ce{I2} into two molecules~\ce{I}, \[ \therefore \nu_{1} = - 1;\quad \nu_{2} = 2; \] and in equilibrium, by~\Eq{(204)}, \[ c_{1}^{-1} c_{2}^{2} = \frac{n_{2}^{2}}{n_{1} (n_{1} + n_{2})} = a' e^{-\efrac{b'}{\theta}} · \frac{\theta}{p}\Add{.} \Tag{(208)} \] $a'$~and $b'$ may be calculated from data given by Fr.~Meier and Crafts. When $p = 728~\Unit{mm.}$ of mercury, \index{Crafts}% \index{Meier, Fr.}% \begin{align*} \frac{n_{2}}{2n_{1} + n_{2}} &= 0.145\quad\text{when}\quad \theta = 273 + 940 = 1213, \\ \llap{\text{and\qquad\qquad}} &= 0.662\quad\text{when}\quad \theta = 273 + 1390 = 1663. \end{align*} \PageSep{221} This gives, if $p$~be measured in millimeters of mercury, \[ a' = 9375;\quad b' = 14690; \] from which the equilibrium of dissociation may be determined \index{Dissociation!graded}% for any temperature and pressure. The heat of dissociation of a molecule of iodine is, by~\Eq{(205)}, \[ L = 1.97 (14690 + \theta) = 28900 + 1.97\theta~\Unit{cal.} \] It will be seen that at such temperatures the external work, on which the second term depends, has an appreciable influence. At $1500°$~C. ($\theta = 1773$) it amounts to $3500~\Unit{cal.}$, making the heat of dissociation \[ L = 32400~\Unit{cal.} \] \Section{247.} \Topic{Graded Dissociation.}---Since, by equation~\Eq{(208)}, \index{Graded dissociation}% the concentration~$c_{2}$ of the monatomic iodine molecules does not vanish even at low temperatures, the decomposition of the iodine vapour should be taken into account in determining the dissociation of hydriodic acid. This will have practically no influence on the results of \SecRef{245}, but nevertheless we shall give the more rigorous solution on account of the theoretical interest which attaches to it. There are now four kinds of molecules in the system: \[ \ce{$n_{1}$ HI};\quad \ce{$n_{2}$ H2};\quad \ce{$n_{3}$ I2};\quad \ce{$n_{4}$ I}. \] Two kinds of chemical changes are possible: \begin{align*} \text{(1)}\quad \nu_{1} &= -2;\quad \nu_{2} = 1;\quad \nu_{3} = 1;\quad \nu_{4} = 0;\quad\text{and} \\ \text{(2)}\quad \nu_{1}' &= 0;\quad \nu_{2}' = 0;\quad \nu_{3}' = -1;\quad \nu_{4}' = 2. \end{align*} There will be equilibrium for each of these, if, by~\Eq{(204)}, \[ \text{(1)}\quad c_{1}^{\nu_{1}} c_{2}^{\nu_{2}} c_{3}^{\nu_{3}} c_{4}^{\nu_{4}} = \frac{c_{2}c_{3}}{c_{1}^{2}} = \frac{n_{2}n_{3}}{n_{1}^{2}} = ae^{-\efrac{b}{\theta}}, \] \PageSep{222} and \[ \text{(2)}\quad c_{1}^{\nu_{1}'} c_{2}^{\nu_{2}'} c_{3}^{\nu_{3}'} c_{4}^{\nu_{4}'} = \frac{c_{4}^{2}}{c_{3}} = \frac{n_{4}^{2}}{n_{3} (n_{1} + n_{2} + n_{3} + n_{4})} = a' e^{-\efrac{b'}{\theta}} · \frac{\theta}{p}. \] The constants $a$,~$b$, $a'$,~$b'$ have the values calculated above. The total number of hydrogen atoms $(n_{1} + 2n_{2})$ and of iodine atoms $(n_{1} + 2n_{3} + n_{4})$ being known, we have four equations for the complete determination of the four quantities $n_{1}$, $n_{2}$, $n_{3}$,~$n_{4}$. \Section{248.} The general equation of equilibrium~\Eq{(204)} also \index{Dilute solutions!thermodynamical theory of}% \index{Thermodynamical theory!of dilute solutions}% shows that at finite temperatures and pressures none of the concentrations,~$c$, can ever vanish; in other words, that the dissociation can never be complete, nor can it completely vanish. There is always present a finite, though perhaps a very small number of all possible kinds of molecules. Thus, in water vapour at any temperature at least a trace of oxygen and hydrogen must be present (see also \SecRef{259}). In a great number of phenomena, however, these quantities are too small to be of any importance. \index{Gaseous system|)}% \index{System!gaseous|)}% \PageSep{223} \Chapter{V.}{Dilute Solutions.} \index{Dilute solutions|(}% \index{Solutions, dilute|(}% \Section{249.} To determine~$\Psi$ as a function of the temperature~$\theta$, the pressure~$p$, and the number~$n$ of the different kinds of molecules in a system of any number of constituents and any number of phases, we may use the method of the preceding chapter. It is necessary first to find by suitable measurements the volume~$V$, and the internal energy~$U$ of each single phase, and then calculate the entropy~$\Phi$ from the definition~\Eq{(60)}. A simple summation extending over all the phases gives, by~\Eq{(75)}, the function~$\Psi$ for the whole system. On account of incomplete experimental data, however, the calculation of~$\Psi$ can be performed, besides for a gaseous phase, only for a \emph{dilute solution}, \ie\ for a phase in which one kind of molecule far outnumbers all the others in the phase. We shall in future call this kind of molecule the \emph{solvent}, the other kinds the \emph{dissolved substances}. This differs from the definition of~\SecRef{220}. If $n_{0}$~be the number of molecules of the solvent, $n_{1}$,~$n_{2}$, $n_{3}$,~\dots the number of molecules of the dissolved substances, then the solution may be considered dilute if $n_{0}$~be large in comparison with each of the numbers $n_{1}$,~$n_{2}$, $n_{3}$\Add{,}~\dots. The state of aggregation of the substance is of no importance, it may be solid, liquid, or gaseous. \Section{250.} We shall now determine by the above method the energy~$U$ and the volume~$V$ of a dilute solution. The important simplification, to which this definition of a dilute solution leads, rests on the mathematical theorem, that a finite, continuous, and differentiable function of several \PageSep{224} variables, which have very small values, is necessarily a \emph{linear} function of these variables. This determines $U$~and $V$ as functions of $n_{0}$,~$n_{1}$, $n_{2}$,~\dots\Add{.} Physically speaking, this means that the properties of a dilute solution, besides depending on the interactions between the molecules of the solvent, necessarily depend only on the interactions between the molecules of the solvent and the molecules of the dissolved substances, but not on the interactions of the dissolved substances among themselves, for these are small quantities of a higher order. \Section{251.} The quotient~$\dfrac{U}{n_{0}}$, \ie\ the internal energy divided \index{Energy!of dilute solution}% by the number of molecules of the solvent, remains unchanged if the numbers, $n_{0}$,~$n_{1}$, $n_{2}$\Add{,}~\dots be varied in the same proportion; for, by \SecRef{201}, $U$~is a homogeneous function of the number of molecules $n_{0}$,~$n_{1}$, $n_{2}$,~\dots, of the first degree. $\dfrac{U}{n_{0}}$~is, therefore, a function of the ratios $\dfrac{n_{1}}{n_{0}}$, $\dfrac{n_{2}}{n_{0}}$,~\dots, and also a linear function, since these ratios are small, and the function is supposed to be differentiable. The function is, therefore, of the form \[ \frac{U}{n_{0}} = u_{0} + u_{1}\, \frac{n_{1}}{n_{0}} + u_{2}\, \frac{n_{2}}{n_{0}} + \dots\Add{,} \] where, $u_{0}$, $u_{1}$, $u_{2}$ are quantities depending, not on the number of molecules, but only on the temperature~$\theta$, the pressure~$p$, and the nature of the molecules. In fact, $u_{0}$~depends only on the nature of the solvent, since the energy reduces to $n_{0} u_{0}$, when $n_{1} = 0 = n_{2} = \dots$, and $u_{1}$~only on the nature of the first dissolved substance and the solvent, and so on. $u_{0}$,~therefore, corresponds to the interactions between the molecules of the solvent, $u_{1}$~to those between the solvent and the first dissolved substance, and so on. This contains a refutation of an objection, which is often raised against the modern theory of dilute solutions, that it treats dilute solutions \index{Dilute solutions!energy of}% simply as gases, and takes no account of the influence of the solvent. \PageSep{225} \Section{252.} If the dilution is not sufficient to warrant the use \index{Dilute solutions!volume of}% \index{Entropy!of dilute solution}% of this very simple form of the function~$U$, a more accurate relation may be obtained by expanding Taylor's series still further, \[ \frac{U}{n_{0}} = u_{0} + u_{1}\, \frac{n_{1}}{n_{0}} + \dots + u_{11} \left(\frac{n_{1}}{n_{0}}\right)^{2} + 2u_{12}\, \frac{n_{1}}{n_{0}} · \frac{n_{2}}{n_{0}} + u_{22} \left(\frac{n_{2}}{n_{0}}\right)^{2} + \dots\Add{.} \] The coefficients $u_{11}$,~$u_{12}$, $u_{22}$,~\dots refer to the influence of the interactions of the dissolved molecules with one another. This, in fact, is the only practicable way of obtaining a rational thermodynamical theory of solutions of any concentration. \Section{253.} We shall here keep to the simple form, and write \[ \left. \begin{aligned} U &= n_{0} u_{0} + n_{1} u_{1} + n_{2} u_{2} + \dots\Add{,} \\ \llap{\text{and\qquad\qquad}} V &= n_{0} v_{0} + n_{1} v_{1} + n_{2} v_{2} + \dots\Add{.} \end{aligned} \right\} \Tag{(209)} \] How far these equations correspond to the facts may be determined by the inferences to which they lead. If we dilute the solution still further by adding one molecule of the solvent in the same state of aggregation as the solution, keeping meanwhile the temperature~$\theta$ and the pressure~$p$ constant, the corresponding change of volume and the heat effect may be calculated from the above equations. One molecule of the pure solvent, at the same temperature and pressure, has the volume~$v_{0}$ and the energy~$u_{0}$. After dilution, the volume of the solution becomes \[ V' = (n_{0} + 1) v_{0} + n_{1} v_{1} + n_{2} v_{2} + \dots \] and the energy \[ U' = (n_{0} + 1) u_{0} + n_{1} u_{1} + n_{2} u_{2} + \dots\Add{.} \] The increase of volume brought about by the dilution is therefore \[ V' - (V + v_{0}), \] \ie\ the increase of volume is zero. The heat absorbed is, by the first law~\Eq{(47)}, \[ U' - (U + u_{0}) + p \bigl\{V' - (V + v_{0})\bigr\}. \] \PageSep{226} This also vanishes. These inferences presuppose that the number of molecules of the dissolved substances remain unchanged, \ie\ that no chemical changes (\eg\ changes of the degree of dissociation) are produced by the dilution. If such were the case, the number of molecules of the dissolved substances would have values in the equations for $U'$~and~$V'$ different from those in the equations for $U$~and~$V$, and therefore would not disappear on subtraction. We may therefore enunciate the following proposition: \emph{Further dilution of a dilute solution, if no chemical changes accompany the process, produces neither an appreciable change of volume nor an appreciable heat effect}; or, in other words, \emph{any change of volume or any heat effect produced by further dilution of a dilute solution is due to chemical transformations among the molecules of the dissolved substances}. \Section{254.} We now turn to the calculation of the entropy~$\Phi$ \index{Dilute solutions!entropy of}% of a dilute solution. If the number of molecules $n_{0}$,~$n_{1}$, $n_{2}$,~\dots be constant, we have, by~\Eq{(60)}, \[ d\Phi = \frac{dU + p\, dV}{\theta}, \] and, by~\Eq{(209)}, \[ d\Phi = n_{0}\, \frac{du_{0} + p\, dv_{0}}{\theta} + n_{1}\, \frac{du_{1} + p\, dv_{1}}{\theta} + n_{2}\, \frac{du_{2} + p\, dv_{2}}{\theta} + \dots\Add{.} \] Since $u$~and $v$ are functions of $\theta$~and $p$~only, and not of~$n$, each of the coefficients of $n_{0}$,~$n_{1}$, $n_{2}$,~\dots, must be a perfect differential, \ie\ there must be certain functions~$\phi$, depending only on $\theta$~and~$p$, such that \[ \left. \begin{aligned} d\phi_{0} &= \frac{du_{0} + p\, dv_{0}}{\theta}\Add{,} \\ d\phi_{1} &= \frac{du_{1} + p\, dv_{1}}{\theta}\Add{,} \\ d\phi_{2} &= \frac{du_{2} + p\, dv_{2}}{\theta}\Add{.} \end{aligned} \right\} \Tag{(210)} \] \PageSep{227} We have, then, \[ \Phi = n_{0} \phi_{0} + n_{1} \phi_{1} + n_{2} \phi_{2} + \dots + C, \Tag{(211)} \] where the integration constant~$C$ cannot depend on $\theta$~and~$p$, but may be a function of the number of molecules. $C$~may be determined as a function of $n_{0}$,~$n_{1}$, $n_{2}$,~\dots for a particular temperature and pressure, and this will be the general expression for~$C$ at any temperature and pressure. We shall now determine~$C$ as a function of~$n$ taking the particular case of high temperature and small pressure. By increasing the temperature and diminishing the pressure, the solution, whatever may have been its original state of aggregation, will pass completely into the gaseous state. Chemical changes will certainly take place at the same time, \ie\ the number of molecules~$n$ will change. But we shall assume that the process takes place in such a way as to leave the number of the different kinds of molecules unaltered, because $C$~remains constant only in this case. Only an ideal process can accomplish this, since it passes through unstable states. There is, however, no objection to its use for our present purpose, since the above expression for~$\Phi$ holds not only for stable states of equilibrium, but for all states characterized by quite arbitrary values of $\theta$,~$p$, $n_{0}$,~$n_{1}$, $n_{2}$,~\dots\Add{.} Stable equilibrium is a special case, satisfying \index{Cantor@Cantor, Hr.|indexnote}% \Typo{\index{Planck}}{\index{Planck@Planck|indexnote}}% a further condition to be established below.\footnote {Hr.~Cantor maintains (Ann.\ d.\ Phys., \textbf{10}, p.~205, 1903) that it is not permissible to suppose that the gaseous state may be reached in this way. ``It must be proved that this represents a possible state of the substance, that it may be at least a momentary state. But no theoretical proof of this has been advanced, and direct experience does not at all justify such an assumption.'' In reply, it has first of all to be pointed out that the possibility of varying the temperature and the pressure, keeping the number of molecules constant, merely depends on the fact that the number of molecules together with the temperature and the pressure form the independent variables which are necessary for the unique determination of the state of solution under consideration. The variables are not subject to any limitations, except that the number of atoms must remain on the whole unchanged. This does not concern us here, and is chemically self-evident. Therefore, from a general thermodynamical point of view, nothing stands in the way of letting the pressure diminish and the temperature rise in any way, keeping the number of molecules constant, if the formation of a new phase is prevented. When this is recognized, it requires only the hypothesis that by continuing this process the ideal gaseous state is finally reached---a supposition which scarcely any one can object to, and which Hr.~Cantor does not, at least directly, contradict (Planck, Ann.\ d.\ Phys.\ \textbf{10}, p.~436, 1903).---\Tr.} \PageSep{228} At a sufficiently high temperature, and a sufficiently low pressure, any gaseous system possesses so small a density, that it may be regarded as a mixture of perfect gases (\SecRef{21}, and \SecRef{43}). We have, therefore, by~\Eq{(194)}, bearing in mind that here the first kind of molecule is denoted by the suffix~$0$, \begin{align*} \Phi &= n_{0} \left(c_{v_{0}} \log \theta + R \log \frac{\theta}{p}\right) \\ &+ n_{1} \left(c_{v_{1}} \log \theta + R \log \frac{\theta}{p}\right) + \dots + C. \Tag{(212)} \end{align*} The constant~$C$ is independent of $\theta$~and~$p$, and has the value given in~\Eq{(198)}. On comparing this with~\Eq{(211)}, it is seen that the expression for~$\Phi$ can pass from~\Eq{(211)} into~\Eq{(212)} by mere change of temperature and pressure, only if the constant~$C$ is the same in both expressions, \ie\ if, by~\Eq{(198)}, \[ C = n_{0} (k_{0} - R \log c_{0}) + n_{1} (k_{1} - R \log c_{1}) + \dots\Add{.} \] Here $k_{0}$,~$k_{1}$, $k_{2}$,~\dots are constants, and the concentrations are \[ c_{0} = \frac{n_{0}}{n_{0} + n_{1} + n_{2} + \dots};\quad c_{1} = \frac{n_{1}}{n_{0} + n_{1} + n_{2} + \dots}. \] By~\Eq{(211)}, the entropy of a dilute solution becomes \[ \Phi = n_{0} (\phi_{0} + k_{0} - R \log c_{0}) + n_{1} (\phi_{1} + k_{1} - R \log c_{1}) + \dots\Add{.} \Tag{(213)} \] If we put, for shortness, the quantities which depend only on $\theta$~and~$p$, \[ \left. \begin{aligned} \phi_{0} + k_{0} - \frac{u_{0} + pv_{0}}{\theta} &= \varphi_{0}\Add{,} \\ \phi_{1} + k_{1} - \frac{u_{1} + pv_{1}}{\theta} &= \varphi_{1}\Add{,} \\ \phi_{2} + k_{2} - \frac{u_{2} + pv_{2}}{\theta} &= \varphi_{2}\Add{,} \\ \end{aligned} \right\} \Tag{(214)} \] \PageSep{229} we have, finally, from \Eq{(75)}, \Eq{(213)} and~\Eq{(209)}, \begin{align*} \Psi = n_{0} (\varphi_{0} - R \log c_{0}) &+ n_{1} (\varphi_{1} - R \log c_{1}) \\ &+ n_{2} (\varphi_{2} - R \log c_{2}) + \dots\Add{.} \Tag{(215)} \end{align*} This equation determines the thermodynamical properties of a dilute solution. \Section{255.} We may now proceed to establish the conditions of equilibrium of a system consisting of several phases. As hitherto, the different kinds of molecules in the phase will be denoted by suffixes, and the different phases by dashes. For the sake of simplicity the first phase will be left without a dash. The entire system is then represented by \begin{gather*} n_{0}\ m_{0}, n_{1}\ m_{1}, n_{2}\ m_{2}, \dots \mid n_{0}'\ m_{0}', n_{1}'\ m_{1}', n_{2}'\ m_{2}', \dots \\ \mid n_{0}''\ m_{0}'', n_{1}''\ m_{1}'', n_{2}''\ m_{2}'', \dots \mid \Tag{(216)} \end{gather*} The number of molecules \Erratum{are}{is} denoted by~$n$, and the molecular weights by~$m$, and the individual phases are separated by vertical lines. In the general formula we signify the summation over the different kinds of molecules of one and the same phase by writing the individual terms of the summation; the summation over the different phases, on the other hand, by the symbol~$\tsum$. In order to enable us to apply the derived formulæ, we shall assume that each phase is either a mixture of perfect gases or a dilute solution. The latter designation will be applied to phases containing only one kind of molecule, \eg\ a chemically homogeneous solid precipitate from an aqueous solution. One kind of molecule represents the special case of a dilute solution in which the concentrations of all the dissolved substances are zero. \Section{256.} Suppose now that an isothermal isopiestic change be possible, corresponding to a simultaneous variation $\delta n_{0}$,~$\delta n_{1}$, $\delta n_{2}$,~\dots\Add{,} $\delta n_{0}'$,~$\delta n_{1}'$, $\delta n_{2}'$,~\dots... of the number of molecules $n_{0}$,~$n_{1}$, $n_{2}$,~\dots\Add{,} $n_{0}'$,~$n_{1}'$, $n_{2}'$,~\dots; then, by~\Eq{(79)}, this change \PageSep{230} will not take place, if at constant temperature and pressure \[ \delta \Psi = 0 \] or, by~\Eq{(215)}, if \begin{multline*} \tsum (\varphi_{0} - R \log c_{0})\, \delta n_{0} + (\varphi_{1} - R \log c_{1})\, \delta n_{1} + (\varphi_{2} - R \log c_{2})\, \delta n_{2} \\ \begin{aligned} &+ \dots + \tsum n_{0}\, \delta(\varphi_{0} - R \log c_{0}) + n_{1}\, \delta(\varphi_{1} - R \log c_{1}) \\ &+ n_{2}\, \delta(\varphi_{2} - R \log c_{2}) + \dots = 0\Add{.} \end{aligned} \end{multline*} The summation~$\tsum$ extends over all the phases of the system. The second series is identically equal to zero for the same reason as was given in connection with equation~\Eq{(200)}. If we again introduce the simple integral ratio \begin{multline*} \delta n_{0} : \delta n_{1} : \delta n_{2} : \dots : \delta n_{0}' : \delta n_{1}' : \delta n_{2}' : \dots \\ = \nu_{0} : \nu_{1} : \nu_{2} : \dots : \nu_{0}' : \nu_{1}' : \nu_{2}' : \dots \Tag{(217)} \end{multline*} then the equation of equilibrium becomes % [** TN: Re-breaking] %\begin{multline*} \[ \tsum (\varphi_{0} - R \log c_{0})\, \nu_{0} \Erratum{-}{+} (\varphi_{1} - R \log c_{1})\, \nu_{1} \\ + (\varphi_{2} - R \log c_{2})\, \nu_{2} + \dots = 0 \] %\end{multline*} or \begin{align*} \tsum \nu_{0} \log c_{0} + \nu_{1} \log c_{1} + \nu_{2} \log c_{2} + \dots &= \frac{1}{R} \tsum \nu_{0} \varphi_{0} + \nu_{1} \varphi_{1} + \dots \\ &= \log K. \Tag{(218)} \end{align*} $K$~like $\varphi_{0}$, $\varphi_{1}$, $\varphi_{2}$, is independent of the number of molecules~$n$. \Section{257.} The definition of~$K$ gives its functional relation to $\theta$~and~$p$. \begin{align*} \frac{\dd \log K}{\dd \theta} &= \frac{1}{R} \tsum \nu_{0}\, \frac{\dd \varphi_{0}}{\dd \theta} + \nu_{1}\, \frac{\dd \varphi_{1}}{\dd \theta} +\nu_{2}\, \frac{\dd \varphi_{2}}{\dd \theta} + \dots\Add{,} \\ \frac{\dd \log K}{\dd p} &= \frac{1}{R} \tsum \nu_{0}\, \frac{\dd \varphi_{0}}{\dd p} + \nu_{1}\, \frac{\dd \varphi_{1}}{\dd p} +\nu_{2}\, \frac{\dd \varphi_{2}}{\dd p} + \dots\Add{.} \end{align*} \PageSep{231} Now, by~\Eq{(214)}, we have for an infinitely small change of $\theta$~and~$p$ \[ d\varphi_{0} = d\phi_{0} - \frac{du_{0} + p\, dv_{0} + v_{0}\, dp}{\theta} + \frac{u_{0} + pv_{0}}{\theta^{2}}\, d\theta \] and therefore, by~\Eq{(210)}, \[ d\varphi_{0} = \frac{u_{0} + pv_{0}}{\theta^{2}}\, d\theta - \frac{v_{0}\, dp}{\theta}. \] From this it follows that \[ \frac{\dd \varphi_{0}}{\dd \theta} = \frac{u_{0} + pv_{0}}{\theta^{2}};\quad \frac{\dd \varphi_{0}}{\dd p} = -\frac{v_{0}}{\theta}. \] Similarly \[ \frac{\dd \varphi_{1}}{\dd \theta} = \frac{u_{1} + pv_{1}}{\theta^{2}};\quad \frac{\dd \varphi_{1}}{\dd p} = -\frac{v_{1}}{\theta}. \] Hence \begin{align*} \frac{\dd \log K}{\dd \theta} &= \frac{1}{R\theta^{2}} \tsum (\nu_{0} u_{0} + \nu_{1} u_{1} + \dots) + p(\nu_{0} v_{0} + \nu_{1} v_{1} + \dots), \\ \frac{\dd \log K}{\dd p} &= -\frac{1}{R\theta} \tsum \nu_{0} v_{0} + \nu_{1} v_{1} + \dots\Add{.} \end{align*} Denoting by~$s$ the increase of volume of the system, and by~$L$ the heat absorbed, when the change corresponding to~\Eq{(217)} takes place at constant temperature and pressure, then, by~\Eq{(209)}, \[ s = \tsum \nu_{0} v_{0} + \nu_{1} v_{1} + \nu_{2} v_{2} + \dots \] and, by the first law of thermodynamics, \[ L = \tsum (\nu_{0} u_{0} + \nu_{1} u_{1} + \dots) + p(\nu_{0} v_{0} + \nu_{1} v_{1} + \dots); \] therefore \[ \frac{\dd \log K}{\dd \theta} = \frac{L}{R\theta^{2}} \Tag{(219)} \] \PageSep{232} and \[ \frac{\dd \log K}{\dd p} = -\frac{s}{R\theta}\Add{.} \Tag{(220)} \] The influence of the temperature on~$K$, and therewith on the condition of equilibrium towards a certain chemical reaction, is controlled by the heat effect of that reaction, and the influence of the pressure is controlled by the corresponding change of volume of the system. If the reaction take place without the absorption or evolution of heat, the temperature has no influence on the equilibrium. If it produce no change of volume the pressure has no influence. The former equations \Eq{(205)} and \Eq{(206)} are particular cases of \Eq{(219)} and \Eq{(220)}, as may be seen by substituting for $\log K$ the special value obtained from \Eq{(218)} and~\Eq{(204)}: \[ \log K = \log a - \frac{b}{\theta} + (\nu_{1} + \nu_{2} + \dots) \log \frac{\theta}{p}. \] \Section{258.} By means of equation~\Eq{(218)} a condition of equilibrium may be established for each possible change in a given system subject to chemical change. Of course, $K$~will have a different value in each case. This corresponds to the requirements of Gibbs's phase rule, which is general \index{Gibbs}% \index{Gibbs's phase rule}% in its application (\SecRef{204}). The number of the different kinds of molecules in the system must be distinguished from the number of the independent constituents (\SecRef{198}). Only the latter determines the number and nature of the phases; while the number of the different kinds of molecules plays no part whatever in the application of the phase rule. If another kind of molecule be introduced the number of the variables increases, to be sure, but so does the number of the possible reactions, and therewith, the number of the conditions of equilibrium by the same amount, so that the number of independent variables is quite independent thereof. \Section{259.} \Erratum{Equations}{Equation}~\Eq{(218)} shows further that, generally speaking, all kinds of molecules possible in the system \PageSep{233} will be present in finite numbers in every phase; for instance, molecules of~\ce{H2O} must occur in any precipitate from an aqueous solution. Even solid bodies in contact must partially dissolve in one another, if sufficient time be given. The quantity~$K$, which determines the equilibrium, possesses, according to the definition~\Eq{(218)}, a definite, in general, a finite value for each possible chemical change, and none of the concentrations~$c$ can, therefore, vanish so long as the temperature and the pressure remain finite. This principle, based entirely on thermodynamical considerations, has already served to explain certain facts, \eg\ the impossibility of removing the last traces of impurity from gases, liquids, and even solids. It also follows from it that absolutely semipermeable membranes are non-existent, for the substance of any membrane would, in time, become saturated with the molecules of all the various kinds of substances in contact with one side of it, and thus give up each kind of substance to the other side. On the other hand, this view greatly complicates the calculation of the thermodynamical properties of a solution, since, in order to make no mistake, it is necessary to assume from the start the existence in every phase of all kinds of molecules possible from the given constituents. We must not neglect any kind of molecule until we have ascertained by a particular experiment that its quantity is inappreciable. Many cases of apparent discrepancy between theory and experiment may probably be explained in this way. We shall now discuss some of the most important particular cases. They have been arranged, in the first place, according to the number of the independent constituents of the system; in the second, according to the number of the phases. \Section{260.} \Topic{One Independent Constituent in One Phase.}---According to the phase rule, the nature of the phase depends on two variables, \eg\ on the temperature and the pressure. The phase may contain any number of different kinds of molecules. Water, for instance, will contain simple, double, \PageSep{234} and multiple \ce{H2O}-molecules; molecules of hydrogen and oxygen, \ce{H2} and~\ce{O2}; electrically charged ions \ce{H+},~\ce{HO-} and~\ce{O^{-\,-}}, etc., in finite quantities. The electrical charges of the ions do not play any important part in thermodynamics, so long as there is no direct conflict between the electrical and the thermodynamical forces. This happens when and only when the thermodynamical conditions of equilibrium call for such a distribution of the ions in the different phases of the system as would lead, on account of the constant charges of the ions, to free electricity in any phase. The electrical forces strongly oppose such a distribution, and the resulting deviation from the pure thermodynamical equilibrium is, however, compensated by differences of potential between the phases. A general view of these electromolecular phenomena may be got by generalizing the expressions for the entropy and the energy of the system by the addition of electrical terms. We shall restrict our discussion to states which do not involve electrical phenomena, and need not consider the charges of the ions, which we may treat like other molecules. In the case mentioned above, then, the concentrations of all kinds of molecules are determined by $\theta$~and~$p$. The calculation of the concentrations has succeeded so far only in the case of the \ce{H+}~and \ce{OH-} ions (the number of the \ce{O^{-\,-}}~ions is negligible), in fact, among other methods, by the measurement of the electrical conductivity of the solution, which depends only on the ions. Kohlrausch and Heydweiller \index{Heydweiller}% \index{Kohlrausch}% found the degree of dissociation of water, \ie\ the ratio of \index{Dissociation!of water}% \index{Water, dissociation of}% the mass of water split into \ce{H+}~and \ce{OH-} ions to the total mass of water to be, at $18°$~C., \[ 14.3 × 10^{-10}. \] This number represents the ratio of the number of dissociated molecules to the total number of molecules. We may determine by thermodynamics the change of the dissociation with temperature. \PageSep{235} The condition of equilibrium will now be established. The system is, by~\Eq{(216)}, \[ \ce{$n_{2}$ H2O};\quad \ce{$n_{1}$ H+};\quad \ce{$n_{2}$ OH-}. \] Let the total number of molecules be \[ n = n_{0} + n_{1} + n_{2}, \] the concentrations are, therefore, \[ c_{0} = \frac{n_{0}}{n};\quad c_{1} = \frac{n_{1}}{n};\quad c_{2} = \frac{n_{2}}{n}. \] The chemical reaction in question, \[ \nu_{0} : \nu_{1} : \nu_{2} = \delta n_{0} : \delta n_{1} : \delta n_{2}, \] consists in the dissociation of one \ce{H2O}~molecule into \ce{H+} and \ce{OH-}. \[ \nu_{0} = 1;\quad \nu_{1} = 1;\quad \nu_{2} = 1; \] and therefore, by~\Eq{(218)}, in the state of equilibrium \[ -\log c_{0} + \log c_{1} + \log c_{2} = K, \] or, since $c_{1} = c_{2}$, and $c_{0} = 1$ nearly, \[ 2 \log c_{1} = \log K. \] This gives, by~\Eq{(219)}, the relation between the concentration and the temperature: \[ 2\, \frac{\dd \log c_{1}}{\dd \theta} = \frac{1}{R} · \frac{L}{\theta^{2}}. \Tag{(221)} \] According to Arrhenius, $L$,~the heat necessary for the dissociation \index{Arrhenius}% of one molecule of~\ce{H2O} into \ce{H+}~and \ce{OH-}, is equal to the heat of neutralization of a strong monobasic acid and base in dilute aqueous solution. J.~Thomsen's experiments \index{Thomsen, J.}% give for mean temperatures: \[ L = \frac{4045000}{\theta}~\Unit{cal.} \] \PageSep{236} On reducing calories to C.G.S. units, we get \[ \frac{\dd \log c_{1}}{\dd \theta} = \frac{1}{2 × 1.971} × \frac{4045000}{\theta^{3}}. \] On integrating, we have \begin{align*} \log c_{1} &= -\frac{4045000}{7.884} · \frac{1}{\theta^{2}} = -\frac{513000}{\theta^{2}} + \const \\ c_{1} &= C e^{-\efrac{513000}{\theta^{2}}}. \end{align*} The value of the constant~$C$ is found from the degree of dissociation at $18°$~C. ($\theta = 291$); \begin{align*} c_{l} = c_{2} &= 14.3 × 10^{-10}\Add{,} \\ \therefore C &= 6.1 × 10^{-7}\Add{.} \end{align*} Hence the degree of dissociation for any temperature is, \[ c_{1} = 6.1 e^{-\efrac{513000}{\theta^{2}}} × 10^{-7}. \] This agrees well with the electrical conductivity of pure \index{Conductivity of water, electrical}% \index{Electrical conductivity of water}% water when measured at different temperatures. Only at the absolute zero of temperature does the dissociation, and with it the conductivity, vanish. On the other hand, it does not increase indefinitely with temperature, but reaches a maximum value~$C$. \Section{261.} \Topic{One Independent Constituent in Two or Three Phases.}---The main features of these cases have already been discussed in Chapter~II., \SSecRef{205} to~\SecNum{207}, and~\SecRef{213}. \Section{262.} \Topic{Two Independent Constituents in One Phase\Add{.}}\\ ---(A substance dissolved in a homogeneous solvent). According to the phase rule, one other variable besides the pressure and the temperature is arbitrary, \eg\ the number of the molecules dissolved in $1$~litre of the solution, a quantity which may be directly measured. The values of \PageSep{237} these three variables determine the concentrations of all kinds of molecules, whether they have their origin in dissociation, association, formation of hydrates, or hydrolysis of the dissolved molecules. Let us consider the simple case of a binary electrolyte, \eg\ acetic acid in water. The \index{Acetic acid}% \index{Binary electrolyte}% \index{Electrolyte, binary}% system is represented by \[ \ce{$n_{0}$ H2O},\quad \ce{$n_{1}$ CH3 . COOH},\quad \ce{$n_{2}$ H+},\quad \ce{$n_{3}$ CH3- . COO}. \] The total number of molecules, \[ n = n_{0} + n_{1} + n_{2} + n_{3}, \] is only slightly greater than~$n_{0}$. The concentrations are \[ c_{0} = \frac{n_{0}}{n};\quad c_{1} = \frac{n_{1}}{n};\quad c_{2} = \frac{n_{2}}{n};\quad c_{3} = \frac{n_{3}}{n}. \] The reaction to be considered is represented by \[ \nu_{0} : \nu_{1} : \nu_{2} : \nu_{3} = \delta n_{0} : \delta n_{1} : \delta n_{2} : \delta n_{3}, \] and consists in the dissociation of one molecule of \ce{CH3 . COOH} into its two ions. \[ \nu_{0} = 0;\quad \nu_{1} = -1;\quad \nu_{2} = 1;\quad \nu_{3} = 1. \] Therefore, in equilibrium, \[ -\log c_{1} + \log c_{2} + \log c_{3} = \log K; \] or, since $c_{2} = c_{3}$, \[ \frac{c_{2}^{2}}{c_{1}} = K\Add{.} \Tag{(222)} \] Now, we may regard the sum \[ c_{1} + c_{2} = c \] as known, since the total number ($n_{1} + n_{2}$) of the undissociated and the dissociated molecules of the acid, and the total number of water molecules, which may be put $= n$, are \PageSep{238} measured directly. Hence $c_{1}$~and $c_{2}$ may be calculated from the last two equations. \begin{align*} \frac{c_{1}}{c} &= \frac{n_{1}}{n_{1} + n_{2}} = 1 - \frac{K}{2c} \left(\sqrt{1 + \frac{4c}{K}} - 1\right); \\ \frac{c_{2}}{c} &= \frac{n_{2}}{n_{1} + n_{2}} = \frac{K}{2c} \left(\sqrt{1 + \frac{4c}{K}} - 1\right). \end{align*} With increasing dilution (decreasing~$c$), the ratio~$\dfrac{c_{2}}{c}$ increases \index{Dilution!law of, of binary electrolytes}% in a definite manner approaching the value~$1$, \ie\ complete dissociation. This also gives for the electrical conductivity \index{Dissociation!of \ce{H2SO4}}% \index{Laws:!Ostwald's}% \index{Ostwald's law}% \index{Sulphuric acid, dissociation of}% of a solution of given concentration Ostwald's so-called \emph{law of dilution of binary electrolytes},\footnote {$K = \dfrac{\lambda_{v}}{\lambda_{\infty} (\lambda_{\infty} - \lambda_{v})^{v}}$, where $\lambda_{v}$~is the molecular conductivity at dilution~$v$; $\lambda_{\infty}$~the molecular conductivity at infinite dilution; and $v$~the molecular volume of the electrolyte.---\Tr.} which has been experimentally verified in numerous cases. In a manner quite similar to that of~\SecRef{260}, the heat effect of the dissociation shows how the degree of dissociation depends on the temperature. Conversely, as was first shown by Arrhenius, the heat of dissociation may be calculated from \index{Arrhenius}% the rate of change of the dissociation with temperature. \Section{263.} Usually, however, in a solution, not one, but a large number of reactions will be possible. Accordingly, the complete system contains many kinds of molecules. As another example, we shall discuss the case of an electrolyte capable of splitting into ions in several ways, viz.\ an aqueous solution of sulphuric acid. The system is represented by \[ \ce{$n_{0}$ H2O},\quad \ce{$n_{1}$ H2SO4},\quad \ce{$n_{2}$ H+},\quad \ce{$n_{3}$ HSO4-},\quad \ce{$n_{4}$ SO4^{-\,-}}. \] The total number of molecules is \[ n = n_{0} + n_{1} + n_{2} + n_{3} + n_{4}\quad\text{(nearly equal to~$n_{0}$).} \] \PageSep{239} The concentrations are \[ c_{0} = \frac{n_{0}}{n};\quad c_{1} = \frac{n_{1}}{n};\quad c_{2} = \frac{n_{2}}{n};\quad c_{3} = \frac{n_{2}}{n};\quad c_{4} = \frac{n_{3}}{n}. \] Here two different kinds of reactions \[ \nu_{0} : \nu_{1} : \nu_{2} : \nu_{3} : \nu_{4} = \delta n_{0} : \delta n_{1} : \delta n_{2} : \delta n_{3} : \delta n_{4} \] must be considered; first, the dissociation of one molecule of \ce{H2SO4} into \ce{H+} and \ce{HSO4-}: \[ \nu_{0} = 0;\quad \nu_{1} = -1;\quad \nu_{2} = 1;\quad \nu_{3} = 1;\quad \nu_{4} = 0; \] second, the dissociation of the ion \ce{HSO4-} into \ce{H+} and \ce{SO4^{-\,-}}\Add{:} \[ \nu_{0} = 0;\quad \nu_{1} = 0;\quad \nu_{2} = 1;\quad \nu_{3} = -1;\quad \nu_{4} = 1. \] Hence, by~\Eq{(218)}, there are two conditions of equilibrium: \begin{align*} -\log c_{1} + \log c_{2} + \log c_{3} &= \log K \\ \intertext{and} \log c_{2} - \log c_{3} + \log c_{4} &= \log K'; \end{align*} or \[ \frac{c_{2} c_{3}}{c_{1}} = K \] and \[ \frac{c_{2} c_{4}}{c_{3}} = K'. \] This further condition must be added, viz.\ that the total number of \ce{SO4} radicals ($n_{1} + n_{2} + n_{3}$) must be equal to half the number of \ce{H}~atoms ($2n_{1} + n_{2} + n_{3}$); otherwise the system would contain more than two independent constituents. This condition is \[ 2c_{4} + c_{3} = c_{2}. \] Finally, the quantity of sulphuric acid in the solution is supposed to be given: \[ c_{1} + c_{3} + c_{4} = c. \] \PageSep{240} The last four equations determine $c_{1}$,~$c_{2}$, $c_{3}$,~$c_{4}$, and hence the state of equilibrium is found. For a more accurate determination it would be necessary to consider still other kinds of molecules. Every one of these introduces a new variable, but also a new possible reaction, and therefore a new condition of equilibrium, so that the state of equilibrium remains uniquely determined. \Section{264.} \Topic{Two Independent Constituents in Two Phases.}\\ ---The state of equilibrium, by the phase rule, depends on two variables, \eg\ temperature and pressure. The wide range of cases in point makes a subdivision desirable, according as only one phase contains both constituents in appreciable quantity, or both phases contains both constituents. Let us first take the simpler case, where one (first) phase contains both constituents, and the other (second) phase contains only one single constituent. Strictly speaking this never occurs (by~\Chg{259}{\SecRef{259}}), but in many cases it is a sufficient approximation to the actual facts. The application of the general condition of equilibrium~\Eq{(218)} to this case leads to different laws, according as the constituent in the second phase plays the part of dissolved substance or solvent (\SecRef{249}) in the first phase. We shall therefore divide this case into two further subdivisions. \Section{265.} \Topic{The Pure Substance in the Second Phase forms the Dissolved Body in the First.}---An example of this is the absorption of a gas, \eg\ carbon dioxide in a liquid of comparatively small vapour pressure. The system is represented by \[ \ce{$n$ H2O},\quad \ce{$n_{1}$ CO2} \mid \ce{$n_{0}'$ CO2}. \] The concentrations of the different kinds of molecule of the system in the two phases are \[ c_{0} = \frac{n_{0}}{n_{0} + n_{1}};\quad c_{1} = \frac{n_{1}}{n_{0} + n_{1}};\quad c_{0}' = \frac{n_{0}'}{n_{0}'} = 1. \] \PageSep{241} The reaction \[ \nu_{0} : \nu_{1} : \nu_{0}' = \delta n_{0} : \delta n_{1} : \delta n_{0}' \] consists in the evaporation of one molecule of carbon dioxide from the solution, therefore, \[ \nu_{0} = 0,\quad \nu_{1} = -1,\quad \nu_{0}' = 1. \] The condition of equilibrium \[ \nu_{0} \log c_{0} + \nu_{1} \log c_{1} + \nu_{0}' \log c_{0}' = \log K, \] is, therefore, \[ -\log c_{1} = \log K, \Tag{(223)} \] or, at a given temperature and pressure (for these determine~$K$), $c_{1}$~the concentration of the gas in the solution is determined. The change of concentration with pressure and temperature is found by substituting \Eq{(223)} in \Eq{(219)} and~\Eq{(220)}: \begin{align*} \frac{\dd \log c_{1}}{\dd p} &= \frac{1}{R} · \frac{s}{\theta}\Add{,} \Tag{(224)} \\ \frac{\dd \log c_{1}}{\dd \theta} &= -\frac{1}{R} · \frac{L}{\Erratum{\theta_{2}}{\theta^{2}}}\Add{.} \Tag{(225)} \end{align*} $s$~is the increase of volume of the system, $L$~the heat absorbed during isothermal-isopiestic evaporation of one gram molecule of~\ce{CO2}. Since $s$~represents nearly the volume of one gram molecule of \Erratum{carbonic}{carbon} dioxide gas, we may, by~\Eq{(16)}, put \[ s = \frac{R\theta}{p}, \] and equation~\Eq{(224)} gives \[ \frac{\dd \log c_{1}}{\dd p} = \frac{1}{p}. \] \PageSep{242} On integrating, we have \[ \log c_{1} = \log p + \const \] or \[ c_{1} = Cp \Tag{(226)} \] \ie\ \emph{the concentration of the dissolved gas is proportional to the pressure of the free gas on the solution} (Henry's law). The \index{Henry's law}% \index{Laws:!Henry's}% factor~$C$, which is a measure of the solubility of the gas, still depends on the temperature, since \Eq{(225)} and \Eq{(226)} give \[ \frac{\dd \log C}{\dd \theta} = -\frac{1}{R} · \frac{L}{\Erratum{\theta_{2}}{\theta^{2}}}\Add{.} \] If, therefore, heat is absorbed during the evaporation of the gas from the solution, $L$~is positive, and the solubility decreases with increase of temperature. Conversely, from the variation of~$C$ with temperature, the heat effect produced by the absorption may be calculated; \[ L = -\frac{R \theta^{2}}{C} · \frac{\dd C}{\dd \theta}. \] According to the experiments of Naccari and Pagliani, \index{Naccari}% \index{Pagliani}% the solubility of carbon dioxide in water at~$20°$ ($\theta = 293$), (expressed in a unit which need not be discussed here), is~$0.8928$, its temperature coefficient $-0.02483$; therefore, by~\Eq{(34)}, \[ L = \frac{1.971 × 293^{2} × 0.02483}{0.8928} = 4700~\Unit{cal.} \] Thomsen found the heat effect of the absorption of one gram molecule of carbon dioxide to be $5880~\Unit{cal}$. The error (according to Nernst) lies mainly in the determination of \index{Nernst}% the coefficient of solubility. Of the heat effect, the amount \[ R\theta = 1.97 × 293 = 586~\Unit{cal.} \] corresponds, by~\Eq{(48)}, to external work. \Section{266.} A further example is the saturation of a liquid \PageSep{243} with an almost insoluble salt; \eg\ succinic acid in water. \index{Succinic acid}% The system is represented by \[ \ce{$n_{0}$ H2O},\quad n_{1}\ \begin{array}{@{}l@{}} \ce{CH2} - \ce{COOH} \\ | \\ \ce{CH2} - \ce{COOH}, \\ \end{array} \quad\left|\quad n_{0}'\ \begin{array}{@{}l@{}} \ce{CH2} - \ce{COOH} \\ | \\ \ce{CH2} - \ce{COOH}, \\ \end{array} \right. \] if the slight dissociation of the acid in water be neglected. The calculation of the condition of equilibrium gives, as in~\SecRef{223}, \[ -\log c_{1} = \log K, \] $c_{1}$~is determined by temperature and pressure. Further, by~\Eq{(219)}, \[ L = -R\theta^{2} \frac{\dd \log c_{1}}{\dd \theta}\Add{.} \Tag{(227)} \] Van't Hoff was the first to calculate~$L$ by means of this equation from the solubility of succinic acid at $0°$~C. ($2.88$) and at $8.5°$~C. ($4.22$) \[ \frac{\dd \log c_{1}}{\dd \theta} = \frac{\log_{e} 4.22 - \log_{e} 2.88}{8.5} = 0.04494. \] This gives, for $\theta = 273$, $L = -1.971 × 273^{2} × 0.4494 = 6600~\Unit{cals.}$; \ie\ on the precipitation of one \Erratum{molecule}{gram molecule} of the solid from the solution, $6600~\Unit{cals.}$ are given out. Berthelot found the heat of solution to be $6700~\Unit{cals}$. \index{Berthelot}% If $L$~be regarded as independent of the temperature, which is permissible in many cases as a first approximation, the equation~\Eq{(227)} may be integrated with respect to~$\theta$, giving \[ \log c_{1} = \frac{L}{R\theta} + \const \] \Section{267.} The relation~\Eq{(227)} becomes inapplicable if the salt in solution undergoes an appreciable chemical transformation, \eg\ dissociation. For then, besides the ordinary \PageSep{244} molecules of the salt, the products of the dissociation are present in the solution; for example, in the system of water and silver acetate, \index{Acetate!silver}% \index{Silver!acetate}% \index{Silver!nitrate}% \[ \ce{$n_{0}$ H2O},\ \ce{$n_{1}$ CH3COOAg},\ \ce{$n_{2}$ Ag+},\ \ce{$n_{3}$ CH3- . COO} \mid \ce{$n_{0}'$ CH3COOAg}. \] The total number of molecules in the solution: \[ n = n_{0} + n_{1} + n_{2} + n_{3}\quad\text{(nearly $ = n_{0}$).} \] The concentrations of the different molecules in both phases are \[ c_{0} = \frac{n_{0}}{n};\quad c_{1} = \frac{n_{1}}{n};\quad c_{2} = \frac{n_{2}}{n};\quad c_{3} = \frac{n_{3}}{n};\quad c_{0}' = \frac{n_{0}'}{n_{0}'} = 1. \] The reactions, \[ \nu_{0} : \nu_{1} : \nu_{2} : \nu_{3} : \nu_{0}' = \delta n_{0} : \delta n_{1} : \delta n_{2} : \delta n_{3} : \delta n_{0}', \] are: (1) The precipitation of a molecule of the salt from the solution: \[ \nu_{0} = 0,\quad \nu_{1} = -1,\quad \nu_{2} = 0,\quad \nu_{3} = 0,\quad \nu_{0}' = 1. \] (2) The dissociation of a molecule of silver acetate: \[ \nu_{0} = 0,\quad \nu_{1} = -1,\quad \nu_{2} = 1,\quad \nu_{3} = 1,\quad \nu_{0}' = 0. \] Accordingly, the two conditions of equilibrium are: (1)\qquad $-\log c_{1} = \log K$\Add{,} (2)\qquad $-\log c_{1} + \log c_{2} + \log c_{3} = \log K'$; \\ or, since $c_{2} = c_{3}$, \[ \frac{c_{2}^{2}}{c_{1}} = K'. \] At given temperature and pressure, therefore, there is in the saturated solution of a salt a definite number of undissociated molecules; and the concentration~($c_{2}$) of the dissociated molecules may be derived from that of the \PageSep{245} undissociated~($c_{1}$) by the law of dissociation of an electrolyte, as given in~\Eq{(222)}. Now, since by measuring the solubility the value of $c_{1} + c_{2}$, and by measuring the electrical conductivity the value of~$c_{2}$, may be found, the quantities $K$~and $K'$ can be calculated for any temperature. Their dependence on temperature, by~\Eq{(219)}, serves as a measure of the heat effect of the precipitation of an undissociated molecule from the solution, and of the dissociation of a dissolved molecule. Jahn has thus \index{Jahn}% given a method of calculating the actual heat of solution of a salt, from measurements of the solubility of the salt and of the conductivity of saturated solutions at different temperatures; \ie\ the heat effect which takes place when one gram molecule of the solid salt is dissolved, and the fraction $\dfrac{c_{2}}{c_{1} + c_{2}}$ is dissociated into its ions, as is actually the case in the process of solution. \Section{268.} \Topic{The Pure Substance occurring in the Second Phase forms the Solvent in the First Phase.}---This case is realized when the pure solvent in any state of aggregation is separated out from a solution of another state of aggregation, \eg\ by freezing, evaporation, fusion, and sublimation. The type of such a system is \[ n_{0}\ m_{0},\quad n_{1}\ m_{1},\quad n_{2}\ m_{2},\quad n_{3}\ m_{3},\ \dots \mid n_{0}'\ m_{0}'. \] The question whether the solvent has the same molecular weight in both phases, or not, is left open. The total number of molecules in the solution is \[ n = n_{0} + n_{1} + n_{2} + n_{3} + \dots\quad\text{(nearly $= n_{0}$).} \] The concentrations are \[ c_{0} = \frac{n_{0}}{n};\quad c_{1} = \frac{n_{1}}{n};\quad c_{2} = \frac{n_{2}}{n};\ \dots c_{0}' = \frac{n_{0}'}{n_{0}'} = 1. \] A possible transformation, \[ \nu_{0} : \nu_{1} \Add{:} \dots : \nu_{0}' = \delta n_{0} : \delta n_{1} \Add{:} \dots : \delta n_{0}', \] \PageSep{246} is the passage of a molecule of the solvent from the first phase to the second phase, \ie\ \[ \nu_{0} = -1;\quad \nu_{1} = 0;\quad \nu_{2} = 0;\ \dots \nu_{0}' = \frac{m_{0}}{m_{0}'}. \Tag{(228)} \] Equilibrium demands, by~\Eq{(218)}, that \[ -\log c_{0} + \frac{m_{0}}{m_{0}'} \log c_{0}' = \log K, \] and, therefore, on substituting the above values of $c_{0}$~and~$c_{0}'$, \[ \log \frac{n}{n_{0}} = \log K. \] But \[ \frac{n}{n_{0}} = 1 + \frac{n_{1} + n_{2} + n_{3} + \dots}{n_{0}}, \] and, therefore, since the fraction on the right is very small, \[ \frac{n_{1} + n_{2} + n_{3} + \dots}{n_{0}} = \log K. \Tag{(229)} \] By the general definition~\Eq{(218)}, we have \[ \log K = \frac{1}{R} (\nu_{0} \varphi_{0} + \nu_{1} \varphi_{1} + \nu_{2} \varphi_{2} + \dots + \nu_{0}' \varphi_{0}'), \] and, therefore, on substituting the values of~$\nu$ from~\Eq{(228)}, \[ \frac{n_{1} + n_{2} + n_{3} + \dots}{n_{0}} = \frac{1}{R} \left(\frac{m_{0}}{m_{0}'}\, \varphi_{0}' - \varphi_{0}\right)\Add{.} \Tag{(230)} \] This expression shows that $\log K$ also has a small value. Suppose for the moment that $\log K = 0$, \ie\ that the pure solvent takes the place of the solution \[ n_{1} + n_{2} + \dots = 0, \] then, by~\Eq{(230)}, \[ \frac{\varphi_{0}}{m_{0}} = \frac{\varphi_{0}'}{m_{0}'}. \] \PageSep{247} Since $\varphi_{0}$~and $\varphi_{0}'$ depend only on $\theta$,~$p$ and the nature of the solvent, and not on the dissolved substances, the above equation asserts a definite relation between temperature and pressure, which is, in fact, the condition which $\theta$~and~$p$ must fulfil, in order that the two states of aggregation of the pure solvent may exist in contact. On substituting the values of $\varphi_{0}$~and $\varphi_{0}'$ from~\Eq{(214)}, we return immediately to the condition of equilibrium~\Eq{(101)} which we deduced in the second chapter. The pressure (vapour pressure) may be taken as depending on the temperature, or the temperature (boiling point, melting point) as depending on the pressure. Returning now to the general case expressed in equation~\Eq{(230)}, we find that the solution of foreign molecules, $n_{1}$,~$n_{2}$, $n_{3}$,~\dots affects the functional relation between $\theta$~and~$p$, which holds for the pure solvent. The deviation, in fact, depends only on the total number of dissolved molecules, and not on their nature. To find its amount in measurable quantities, we may introduce either~$p_{0}$, the pressure which would exist in the system at the given temperature~$\theta$, if there were no dissolved molecules (lowering of the vapour pressure), or the temperature~$\theta_{0}$ which would exist at the given pressure~$p$, if there were no dissolved molecules (elevation of the boiling point, depression of the freezing point). If we take the second alternative, $\theta - \theta_{0}$ will be very small, and we may, therefore, put \[ \log K = \frac{\dd \log K}{\dd \theta}\; (\theta - \theta_{0}), \] or, by~\Eq{(219)}, \[ \log K = \frac{1}{R} · \frac{L}{\theta^{2}} (\theta - \theta_{0}), \] and \[ \therefore \frac{n_{1} + n_{2} + n_{3} + \dots}{n_{0}} = \frac{L}{R\theta^{2}} (\theta - \theta_{0}), \] or \[ \theta - \theta_{0} = \frac{R\theta^{2}}{n_{0} L} (n_{1} + n_{2} + n_{3} + \dots)\Add{.} \Tag{(231)} \] By this formula the elevation of the boiling point may be calculated directly from the number of the dissolved molecules, the temperature, and the heat of vaporization. \PageSep{248} Since $L$~refers to the evaporation of one gram molecule of the liquid, the product~$n_{0} L$ depends only on the mass, and not on the molecular weight~($m_{0}$) of the liquid solvent. If $L$~is to be expressed in calories, we must put $R = 1.97$ (by~\Chg{34}{\Eq{(34)}}). For instance, for one litre of water under atmospheric pressure, \[ n_{0} L = 1000 × 536~\Unit{cal.}\quad\text{(approximately)},\quad \theta = 373, \] and, therefore, the elevation of the boiling point is \begin{align*} \theta - \theta_{0} &= \frac{1.97 × 373^{2}}{1000 × 536} (n_{1} + n_{2} + \dots) \\ &= 0.51 (n_{1} + n_{2} + \dots)°~\Unit{C.} \end{align*} \Section{269.} Let us now compare equation~\Eq{(231)} with the relation~\Eq{(183)}, also referring to the elevation of the boiling point, but deduced from more general principles independent of any molecular theory. The equation is \[ \theta - \theta_{0} = \frac{c \theta^{2} \varphi}{L}\Add{.} \Tag{(232)} \] Here $c$~denotes the ratio of the mass~$M_{2}$ of the dissolved non-volatile substance to the mass~$M_{1}$ of the solvent. In the present notation, \[ c = \frac{n_{1} m_{1} + n_{2} m_{2} + \dots}{n_{0} m_{0}}. \Tag{(233)} \] $L$,~in~\Eq{(232)}, is the heat of vaporization per unit mass of the solvent; therefore, in the present notation, \[ \frac{L}{m_{0}}\Add{.} \Tag{(234)} \] The equation~\Eq{(232)}, therefore, becomes \[ \theta - \theta_{0} = \frac{(n_{1} m_{1} + n_{2} m_{2} + \dots) \theta^{2} \varphi}{n_{0} L}. \] \PageSep{249} Comparison with~\Eq{(231)} shows that the two theories will agree perfectly only if \[ \varphi = \frac{R(n_{1} + n_{2} + \dots)}{n_{1} m_{1} + n_{2} m_{2} + \dots}. \Tag{(235)} \] The molecular theory here set forth specializes the previous more general theory in such a way as to assign the particular value~\Eq{(235)} to the quantity~$\varphi$, formerly defined by~\Eq{(165)}. \Section{270.} The quantity~$\varphi$ was found to be of importance for a whole series of other properties of solutions besides the elevation of the boiling point. These relations may at once be specialized for dilute solutions by substituting the value of~$c\varphi$ from \Eq{(233)} and~\Eq{(235)}, \[ c\varphi = \frac{R(n_{1} + n_{2} + n_{3} + \dots)}{n_{0} m_{0}}, \Tag{(236)} \] and for $L$~and~$s$, by~\Eq{(234)}, the values \[ \frac{L}{m_{0}}\quad\text{and}\quad \frac{s}{m_{0}}\Add{.} \Tag{(237)} \] In this way, for the lowering of the vapour pressure of dilute solutions, we deduce, from~\Eq{(180)}, \[ p_{0} - p = \frac{R\theta}{n_{0}s} (n_{1} + n_{2} + n_{3} + \dots). \Tag{(238)} \] If the vapour of the solvent form a perfect gas, and the specific volume of the solution be negligible in comparison with that of the vapour, then $s$~(the change of volume of the system produced by the evaporation of a gram molecule of the liquid) is equal to the volume of the vapour formed. By~\Eq{(228)}, \[ s = R\, \frac{m_{0}}{m_{0}'} · \frac{\theta}{p}, \] \PageSep{250} therefore, by~\Eq{(238)}, \[ p_{0} - p = \frac{m_{0}'p (n_{1} + n_{2} + \dots)}{n_{0} m_{0}}, \] or, the relative lowering of the vapour pressure, \index{Lowering!of vapour pressure}% \[ \frac{p_{0} - p}{p} = (n_{1} + n_{2} + n_{3} + \dots)\, \frac{m_{0}'}{n_{0} m_{0}}. \] This relation is frequently stated thus:---\emph{The relative lowering of the vapour pressure of a solution is equal to the ratio of the number of the dissolved molecules $(n_{1} + n_{2} + n_{3} + \dots)$ to the number of the molecules of the solvent~$(n_{0})$, or, what is the same thing in dilute solutions, to the total number of the molecules of the solution.} This proposition holds only, as is evident, if $m_{0} = m_{0}'$, \ie\ if the molecules of the solvent possess the same molecular weight in the vapour as in the liquid. This, however, is not generally true, as, for example, in the case of water. It may be well therefore to emphasize this fact, that nothing concerning the molecular weight of the solvent can be inferred from the relative lowering of the vapour pressure, any more than from its boiling point, freezing point, or osmotic pressure. Measurements of this kind will not, under any circumstances, lead to anything but the total number ($n_{1} + n_{2} + \dots$) of the dissolved molecules. Thus, in the last equation the product~$n_{0} m_{0}$ is immediately determined by the mass of the liquid solvent, and the mole-weight,~$m_{0}'$, of the vapour by its density. \Section{271.} For the depression of the freezing point of a \index{Depression of freezing point}% \index{Freezing point!depression of}% dilute solution, it follows from \Eq{(186)}, \Eq{(236)}, and~\Eq{(237)}, that \[ \theta_{0}' - \theta' = \frac{R \theta^{2}}{n_{0} L'} (n_{1} + n_{2} + n_{3} + \dots), \] $L'$~being the heat of solidification of a gram molecule of the solvent. The product,~$n_{0} L'$, is given by the mass of the solvent; it is independent of its molecular weight. To express~$L'$ in calories we must put $R = 1.97$ (by~\Chg{34}{\Eq{(34)}}). \PageSep{251} Take water as an example: For $1$~litre of water under atmospheric pressure, $n_{0} L' = 1000 × 80~\Unit{cal.}$ approximately. \index{Pressure!osmotic}% $\theta_{0}' = 273$, and therefore the depression of the freezing point is \[ \theta_{0}' - \theta' = \frac{1.97 × 273^{2}}{1000 × 80} (n_{1} + n_{2} + \dots) = 1.84 (n_{1} + n_{2} + \dots)°~\Unit{C.} \] \Section{272.} Finally, for the osmotic pressure~$P$ we have, from \index{Osmotic pressure}% \Eq{(190)}, \[ P = \frac{R\theta}{n_{0} m_{0} v} (n_{1} + n_{2} + n_{3} + \dots), \] $v$~is the specific volume of the solution, and therefore the product~$n_{0} m_{0} v$ is approximately its whole volume~$V$. Hence \[ P = \frac{R\theta}{V} (n_{1} + n_{2} + n_{3} + \dots), \] an expression identical with the characteristic equation of a mixture of perfect gases with the number of molecules, $n_{1}$,~$n_{2}$, $n_{3}$,~\dots\Add{.} \Section{273.} Each of the theorems deduced in the preceding paragraphs contains a method of determining the total number of the dissolved molecules in a dilute solution. Should the number calculated from such a measurement disagree with the number calculated from the percentage composition of the solution on the assumption of normal molecules, some chemical change of the dissolved molecules must have taken place by dissociation, association, hydrolysis, or the like. This inference is of great importance in the determination of the chemical nature of dilute solutions. The number and nature of the different kinds of molecules are uniquely determined by the total number of molecules only in quite special cases, viz.\ when the dissolved substance undergoes a chemical change only in one way. In this case the total mass of the dissolved substance and the total number of molecules formed by it in the \PageSep{252} solution are sufficient for the calculation of the number of all the different kinds of molecules present. This case is exceptional, however, for we have seen (\SecRef{259}) that all the molecules a substance is capable of forming necessarily occur in the solution in finite quantities. As soon as two reactions (\eg\ \ce{H2SO4 $=$ 2H+ $+$ SO4^{-\,-}} and \ce{H2SO4 $=$ H+ $+$ HSO4-}) must be considered, the analysis of the equilibrium remains indeterminate, since there are more unknown quantities than determining equations. For this reason there is no direct connection between the depression of freezing point, the elevation of the boiling point, etc., on the one hand, and electrical conductivity on the other. For the one set of quantities depends on the total number of the dissolved molecules, charged or uncharged, while the other depends on the number and nature of molecules charged with electricity (ions), which cannot, in general, be calculated from the former. Conversely, a disagreement between the depression of the freezing point as calculated from the conductivity, and as observed, is not in itself an objection to the theory, but rather to the assumptions made in the calculation concerning the kinds of molecules present. Raoult was the first to establish rigorously by experiment the relation between the depression of the freezing point and the number of the molecules of the dissolved substance; and van't Hoff gave a thermodynamical explanation and generalization of it by means of his theory of osmotic pressure. Application to electrolytes was rendered possible by Arrhenius' theory of electrolytic dissociation. \index{Arrhenius'!theory of electrolytic dissociation}% \index{Dissociation!Arrhenius' theory of electrolytic}% \index{Electrolytic dissociation!Arrhenius' theory of}% Thermodynamics has led quite independently, by the method here described, to the necessity of postulating chemical changes of the dissolved substances in dilute solutions. \Section{274.} \Topic{Each Phase contains both Constituents in Appreciable Quantity.}---The most important case is the evaporation of a liquid solution, in which not only the solvent, but also the dissolved substance is volatile. The general equation of equilibrium~\Eq{(218)}, being applicable to \PageSep{253} mixtures of perfect gases whether the mixture may be supposed dilute or not, holds with corresponding approximation for a vapour of any composition. The liquid, on the other hand, must be assumed to be a dilute solution. In general, all kinds of molecules will be present in \index{Van't Hoff's laws}% both phases, and therefore the system is represented by \[ n_{0}\ m_{0},\quad n_{1}\ m_{1},\quad n_{2}\ m_{2},\ \dots \mid n_{0}'\ m_{0},\quad n_{1}'\ m_{1},\quad n_{2}'\ m_{2},\ \dots\Add{.} \] The molecules have the same molecular weight in both phases. The total number of molecules in the liquid is \[ n = n_{0} + n_{1} + n_{2} + \dots\quad\text{(nearly $= n_{0}$)}, \] in the vapour \[ n' = n_{0}' + n_{1}' + n_{2}' + \dots\Add{.} \] The concentrations of the different kinds of molecules are, in the liquid, \[ c_{0} = \frac{n_{0}}{n};\quad c_{1} = \frac{n_{1}}{n};\quad c_{2} = \frac{n_{2}}{n};\ \dots \] in the vapour, \[ c_{0}' = \frac{n_{0}'}{n'};\quad c_{1}' = \frac{n_{1}'}{n'};\quad c_{2}' = \frac{n_{2}'}{n'};\ \dots\Add{.} \] The reaction \begin{multline*} \nu_{0} : \nu_{1} : \nu_{2} : \dots : \nu_{0}' : \nu_{1}' : \nu_{2}' : \dots \\ = \delta n_{0} : \delta n_{1} : \delta n_{2} : \dots : \delta n_{0}' : \delta n_{1}' : \delta n_{2}' : \dots \end{multline*} consists in the evaporation of a molecule of the first kind, and therefore \[ \nu_{0} = 0,\ \nu_{1} = -1,\ \nu_{2} = 0,\ \dots\quad \nu_{0}' = 0,\ \nu_{1}' = 1,\ \nu_{2}' = 0,\ \dots\Add{.} \] The equation of equilibrium becomes \[ -\log c_{1} + \log c_{1}' = \log K, \] or \[ \frac{c_{1}'}{c_{1}} = K. \] \PageSep{254} \emph{For every kind of molecule, which possesses the saute molecular weight in both phases, there is a constant ratio of distribution, which is independent of the presence of other molecules} (Nernst's law of distribution). \index{Distribution law (Nernst's)}% \index{Laws:!Nernst's}% \index{Nernst}% \index{Nernst's law of distribution}% If, on the other hand, a molecule of the solvent evaporate, we have, \[ \nu_{0} = -1,\ \nu_{1} = 0,\ \nu_{2} = 0,\ \dots\quad \nu_{0}' = 1,\ \nu_{1}' = 0,\ \nu_{2}' = 0\Add{,}\ \dots; \] and the equation of equilibrium becomes \[ -\log c_{0} + \log c_{0}' = \log K, \] where \begin{align*} -\log c_{0} &= \log \frac{n}{n_{0}} = \log \left(1 + \frac{n_{1} + n_{2} + \dots}{n_{0}}\right) = \frac{n_{1} + n_{2} + \dots}{n_{0}} \\ &= c_{1} + c_{2} + \dots \Tag{(239)} \end{align*} \[ \therefore c_{1} + c_{2} + \dots + \log c_{0}' = \log K, \Tag{(240)} \] where $c_{1}$,~$c_{2}$,~\dots, the concentrations of the molecules dissolved in the liquid, have small values. Two cases must be considered. Either, the molecules in~$m_{0}$ in the vapour form only a small or at most a moderate portion of the number of the vapour molecules. Then the small numbers $c_{1}$,~$c_{2}$,~\dots, may be neglected in comparison with the logarithm, and therefore \[ \log c_{0}' = \log K. \] This asserts that the concentration of the molecules of the solvent in the vapour does not depend on the composition of the solution. An example of this is the evaporation of a dilute solution, when the solvent is not very volatile, \eg\ alcohol in water. The partial pressure of the solvent (water) in the vapour is not at all dependent on the concentration of the solution, but is equal to that of the pure solvent. Or, the molecules~$m_{0}$ in the vapour far outnumber all the other molecules, as, \eg, when alcohol is the solvent in \PageSep{255} the liquid phase, water the dissolved substance. The concentrations $c_{1}$,~$c_{2}$,~\dots must not be neglected, and, as in~\Eq{(239)}, \[ \log c_{0}' = -(c_{1}' + c_{2}' + \dots); \] equation~\Eq{(240)} therefore becomes \[ (c_{1} + c_{2} + \dots) - (c_{1}' + c_{2}' + \dots) = \log K. \] This relation contains an extension of van't Hoff's laws \index{Laws:!Van't Hoff's}% concerning the elevation of the boiling point, the diminution of the vapour pressure, etc., and asserts that \emph{when the substance dissolved in the liquid also passes in part into the vapour, the elevation of boiling point or the diminution of the vapour pressure depends no longer on the concentrations of the molecules dissolved in the liquid, but on the \Emph{difference} of their concentrations in the liquid and in the vapour}. If this difference be zero, the distillate being of the same composition as the liquid, the elevation of the boiling point and the diminution of the vapour pressure vanish. This conclusion has already been reached from a more general point of view (\SecRef{219}). If the concentration of the dissolved substance in the vapour be larger than that in the liquid, as may happen in the evaporation of an aqueous solution of alcohol, the boiling point falls, while the vapour pressure rises. Exactly analogous theorems may, of course, be deduced for other states of aggregation. Thus, the more general statement of the law concerning the freezing point would be: \emph{If both the solvent and the dissolved substance of a dilute solution solidify in such a way as to form another dilute solution, the depression of the freezing point is not proportional to the concentrations of the dissolved substances in the liquid, but to the \Emph{difference} of the concentrations of the dissolved substances in the liquid and solid phases, and changes sign with this difference.} The solidification of some alloys is an example. While these laws govern the distribution of the molecules in both phases, the equilibrium within each phase \PageSep{256} obeys the laws, which were deduced in \SecRef{262}, etc. We again meet with the laws of dissociation, association, etc. (Nernst). \Section{275.} \Topic{Three Independent Constituents in one Phase.} \\ ---Two dissolved substances in a dilute solution will not affect one another unless they have certain kinds of molecules in common, for there is no transformation possible, and therefore no special condition of equilibrium to fulfil. If two dilute solutions of totally different electrolytes in the same solvent be mixed, each solution will behave as if it had been diluted with a corresponding quantity of the pure solvent. The degree of the dissociation will rise to correspond to the greater dilution. It is different when both electrolytes have an ion in common, as, for example, acetic acid and sodium acetate. \index{Acetate!sodium}% In this case, before mixing there are two systems: \[ \ce{$n_{0}$ H2O},\quad \ce{$n_{1}$ CH3 . COOH},\quad \ce{$n_{2}$ H+},\quad \ce{$n_{3}$ CH3- . COO}, \] and \[ \ce{$n_{0}'$ H2O},\quad \ce{$n_{1}'$ CH3 . COONa},\quad \ce{$n_{2}'$ Na+},\quad \ce{$n_{3}'$ CH3- . COO}. \] As in~\Eq{(222)}, for the first solution, \[ \frac{c_{2}^{2}}{c_{1}} = K,\quad\text{or}\quad \frac{n_{2}^{2}}{n_{1} n_{0}} = K, \Tag{(241)} \] for the second, \[ \frac{c_{2}'^{2}}{c_{1}'} = K',\quad\text{or}\quad \frac{n_{2}'^{2}}{n_{1}' n_{0}'} = K'. \Tag{(242)} \] After mixing the two, we have the system \begin{multline*} \ce{$\bar{n}_{0}$ H2O},\quad \ce{$\bar{n}_{1}$ CH3 . COOH},\quad \ce{$\bar{n}_{2}$ CH3 . COONa}, \\ \ce{$\bar{n}_{3}$ H+},\quad \ce{$\bar{n}_{4}$ Na+},\quad \ce{$\bar{n}_{5}$ CH3- . COO}, \end{multline*} where, necessarily, \settowidth{\TmpLen}{\null\qquad$=$ number of $-$~ions)}% \[ \left. \begin{aligned} \bar{n}_{0} &= n_{0} + n_{0}'\quad\text{(number of \ce{H2O} molecules)\Add{,}} \\ \bar{n}_{2} + \bar{n}_{4} &= n_{1}' + n_{2}'\quad\text{(number of \ce{Na} atoms)\Add{,}} \\ \bar{n}_{1} + \bar{n}_{3} &= n_{1} + n_{2}\quad\text{(number of \ce{H} atoms)\Add{,}} \\ \bar{n}_{3} + \bar{n}_{4} &= \bar{n}_{5}\quad\parbox[t]{\TmpLen}{(number of $+$~ions \\ \null\qquad$=$ number of $-$~ions)\Add{.}} \end{aligned} \right\} \Tag{(243)} \] \PageSep{257} The total number of molecules in the system is \[ \bar{n} = \bar{n}_{0} + \bar{n}_{1} + \bar{n}_{2} + \bar{n}_{3} + \bar{n}_{4} + \bar{n}_{5}\quad\text{(nearly $= \bar{n}_{0}$).} \] The concentrations are \[ \bar{c}_{0} = \frac{\bar{n}_{0}}{\bar{n}};\ \bar{c}_{1} = \frac{\bar{n}_{1}}{\bar{n}};\ \bar{c}_{2} = \frac{\bar{n}_{2}}{\bar{n}};\ \bar{c}_{3} = \frac{\bar{n}_{3}}{\bar{n}};\ \bar{c}_{4} = \frac{\bar{n}_{4}}{\bar{n}};\ \bar{c}_{5} = \frac{\bar{n}_{5}}{\bar{n}}. \] In the system there are two different reactions, \[ \nu_{0} : \nu_{1} : \nu_{2} : \nu_{3} : \nu_{4} : \nu_{5} = \delta \bar{n}_{0} : \delta \bar{n}_{1} : \delta \bar{n}_{2} : \delta \bar{n}_{3} : \delta \bar{n}_{4} : \delta \bar{n}_{5}, \] possible; first, the dissociation of one molecule of acetic acid, \[ \nu_{0} = 0,\quad \nu_{1} = -1,\quad \nu_{2} = 0,\quad \nu_{3} = 1,\quad \nu_{4} = 0,\quad \nu_{5} = 1, \] and therefore the condition of equilibrium is, by~\Eq{(218)}, \[ -\log \bar{c}_{1} + \log \bar{c}_{3} + \log \bar{c}_{5} = \log K, \] or \[ \frac{\bar{c}_{3} \bar{c}_{5}}{\bar{c}_{1}} = K,\quad\text{or}\quad \frac{\bar{n}_{3} · \bar{n}_{5}}{\bar{n}_{1} · \bar{n}_{0}} = \frac{\bar{n}_{3} · \bar{n}_{5}}{\bar{n}_{1} (n_{0} + n_{0}')} = K; \Tag{(244)} \] second, the dissociation of a molecule of sodium acetate, \[ \nu_{0} = 0,\quad \nu_{1} = 0,\quad \nu_{2} = -1,\quad \nu_{3} = 0,\quad \nu_{4} = 1,\quad \nu_{5} = 1, \] whence, for equilibrium, \[ -\log \bar{c}_{2} + \log \bar{c}_{4} + \log \bar{c}_{5} = \log K', \quad\text{or}\quad \frac{\bar{c}_{4} \bar{c}_{5}}{\bar{c}_{2}} = K', \] or \[ \frac{\bar{n}_{4} · \bar{n}_{5}}{\bar{n}_{2} · \bar{n}_{0}} = \frac{\bar{n}_{4} · \bar{n}_{5}}{\bar{n}_{2} (n_{0} + n_{0}')} = K'. \Tag{(245)} \] The quantities $K$~and $K'$ are the same as those in \Eq{(241)} and~\Eq{(242)}. They depend, besides on $\theta$~and~$p$, only on the nature of the reaction, and not on the concentrations, nor on other possible reactions. By the conditions of equilibrium \Eq{(244)} and~\Eq{(245)}, together with the four equations~\Eq{(243)}, the values of the six quantities $\bar{n}_{0}$, $\bar{n}_{1}$,~\dots\Add{,} $\bar{n}_{5}$ are uniquely \PageSep{258} determined, if the original solutions and also the number of molecules $n_{0}$, $n_{1}$,~\dots and $n_{0}'$, $n_{1}'$\Add{,}~\dots be given. \Section{276.} The condition that the two solutions should be \index{Solution!isohydric}% \emph{isohydric}, \ie\ that their degree of dissociation should remain unchanged on mixing them, is evidently expressed by the two equations \[ \bar{n}_{1} = n_{1},\quad\text{and}\quad \bar{n}_{2} = n_{1}', \] \ie\ the number of undissociated molecules of both acetic acid and sodium acetate must be the same in the original solutions as in the mixture. It immediately follows, by~\Eq{(243)}, that \[ \bar{n}_{3} = n_{2},\quad \bar{n}_{4} = n_{2}',\quad \bar{n}_{5} = n_{2} + n_{2}'. \] These values, substituted in \Eq{(244)} and~\Eq{(245)}, and combined with \Eq{(241)} and~\Eq{(242)}, give \begin{align*} \frac{n_{2} (n_{2} + n_{2}')}{n_{1} (n_{0} + n_{0}')} &= K = \frac{n_{2}^{2}}{n_{1} n_{0}}, \\ \frac{n_{2}' (n_{2} + n_{2}')}{n_{1}' (n_{0} + n_{0}')} &= K' = \frac{n_{2}'^{2}}{n_{1}' n_{0}'}, \end{align*} whence the single condition of isohydric solutions is \index{Isohydric solutions}% \[ \frac{n_{2}}{n_{0}} = \frac{n_{2}'}{n_{0}'},\quad\text{or}\quad c_{2} = c_{2}'\ (= c_{3} = c_{3}'), \] or, \emph{the two solutions are isohydric if the concentration of the common ion \emph{\ce{CH3- COO}} is the same in both}. This proposition was enunciated by Arrhenius, who verified it by \index{Arrhenius}% numerous experiments. In all cases where this condition is not realized, chemical changes must take place on mixing the solutions, either dissociation or association. The direction and amount of these changes may be estimated by imagining the dissolved substances separate, and the entire solvent distributed over the two so as to form isohydric solutions. \PageSep{259} If, for instance, both solutions are originally normal ($1$~gram molecule in $1$~litre of solution), they will not be isohydric, since sodium acetate in normal solution is more strongly dissociated, and has, therefore, a greater concentration of \ce{CH3- . COO}-ions, than acetic acid. In order to distribute the solvent so that the concentration of the common ion \ce{CH3- . COO} may be the same in both solutions, some water must be withdrawn from the less dissociated electrolyte (acetic acid), and added to the more strongly dissociated (\ce{Na}-acetate). For, though it is true that with decreasing dilution the dissociation of the acid becomes less, the concentration of free ions increases, as \Typo{(262)}{\SecRef{262}}~shows, because the ions are now compressed into a smaller quantity of water. Conversely, the dissociation of the sodium acetate increases on the addition of water, but the concentration of the free ions decreases, because they are distributed over a larger quantity of water. In this way the concentration of the common ion \ce{CH3- . COO} may be made the same in both solutions, and then their degree of dissociation will not be changed by mixing. This is also the state ultimately reached by the two normal solutions, when mixed. It follows, then, that when two equally diluted solutions of binary electrolytes are mixed, the dissociation of the more weakly dissociated recedes, while that of the more strongly dissociated increases still further. \Section{277.} \Topic{Three Independent Constituents in Two Phases.}---We shall first discuss the simple case, where the second phase contains only one constituent in appreciable quantity. A solution of an almost insoluble salt in a liquid, to which a small quantity of a third substance has been added, forms an example of this case. Let us consider an aqueous solution of silver bromate and silver nitrate. This \index{Silver!bromate}% \index{Silver!nitrate}% two-phase system is represented by \begin{multline*} \ce{$n_{0}$ H2O},\quad \ce{$n_{1}$ AgBrO3},\quad \ce{$n_{2}$ AgNO3}, \\ \ce{$n_{3}$ Ag+},\quad \ce{$n_{4}$ BrO3-},\quad \ce{$n_{5}$ NO3-} \mid \ce{$n_{0}'$ AgBrO3}. \end{multline*} \PageSep{260} The concentrations are \[ c_{0} = \frac{n_{0}}{n};\quad c_{1} = \frac{n_{1}}{n};\quad c_{2} = \frac{n_{2}}{n};\ \dots;\quad c_{0}' = \frac{n_{0}'}{n_{0}'}, \] where \[ n = n_{0} + n_{1} + n_{2} + n_{3} + n_{4} + n_{5}\quad\text{(nearly $= n_{0}$).} \] Of the possible reactions, \begin{multline*} \nu_{0} : \nu_{1} : \nu_{2} : \nu_{3} : \nu_{4} : \nu_{5} : \nu_{0}' \\ = \delta n_{0} : \delta n_{1} : \delta n_{2} : \delta n_{3} : \delta n_{4} : \delta n_{5} : \delta n_{0}', \end{multline*} we shall first consider the passage of one molecule of \ce{AgBrO3} from the solution, viz.\ \[ \nu_{0} = 0,\quad \nu_{1} = -1,\quad \nu_{2} = 0,\ \dots\Add{,}\quad \nu_{0}' = 1. \] The condition of equilibrium is, therefore, \[ -\log c_{1} + \log c_{0}' = \log K \] or \[ c_{1} = \frac{1}{K}. \Tag{(246)} \] The concentration of the undissociated molecules of silver bromate in the saturated solution depends entirely on the temperature and the pressure. We may now consider the dissociation of a molecule of \ce{AgBrO3} into its two ions. \[ \nu_{0} = 0,\ \nu_{1} = -1,\ \nu_{2} = 0,\ \nu_{3} = 1,\ \nu_{4} = 1,\ \nu_{5} = 0,\ \nu_{0}' = 0, \] and, therefore, \begin{gather*} -\log c_{1} + \log c_{3} + \log c_{4} = \log K', \\ \frac{c_{3} c_{4}}{c_{1}} = K', \end{gather*} or, by~\Eq{(246)}, \[ c_{3} c_{4} = \frac{K'}{K}, \Tag{(247)} \] \ie\ the product of the concentrations of the \ce{Ag+}~and \ce{BrO3-} ions depends only on temperature and pressure. The concentration of the \ce{Ag+}-ions is inversely proportional to the \PageSep{261} concentration of the \ce{BrO3-}-ions. Since the addition of silver nitrate increases the number of the \ce{Ag+}-ions, it diminishes the number of the \ce{BrO3-}-ions, and thereby the solubility of the bromate, which is evidently measured by the sum $c_{1} + c_{4}$. We shall, finally, consider the dissociation of a molecule of~\ce{AgNO3} into its ions. \[ \nu_{0} = 0,\ \nu_{1} = 0,\ \nu_{2} = -1,\ \nu_{3} = 1,\ \nu_{4} = 0,\ \nu_{5} = 1,\ \nu_{0}' = 0, \] whence, by~\Eq{(218)}, \[ \frac{c_{3} c_{5}}{c_{2}} = K''. \Tag{(248)} \] To equations \Eq{(246)}, \Eq{(247)}, and~\Eq{(248)}, must be added, as a fourth, the condition \[ c_{3} = c_{4} + c_{5}, \] and, as a fifth, the value of $c_{2} + c_{5}$, given by the quantity of the nitrate added, so that the five unknown quantities, $c_{1}$,~$c_{2}$, $c_{3}$, $c_{4}$,~$c_{5}$, are uniquely determined. The theory of such influences on solubility was first established by Nernst, and has been experimentally verified \index{Nernst}% by him, and more recently by Noyes. \index{Noyes}% \Section{278.} The more general case, where each of the two phases contains all three constituents, is realized in the distribution of a salt between two solvents, which are themselves soluble to a small extent in one another (\eg\ water and ether). The equilibrium is completely determined by a combination of the conditions holding for the transition of molecules from one phase to another with those holding for the chemical reactions of the molecules within one and the same phase. The former set of conditions may be summed up in Nernst's law of distribution (\SecRef{274}). It assigns to each kind of molecule in the two phases a constant ratio of distribution, which is independent of the presence of other dissolved molecules. The second set is the conditions of \PageSep{262} the coexistence of three independent constituents in one phase (\SecRef{275}), to which must be added Arrhenius' theory of \index{Arrhenius'!theory of isohydric solutions}% isohydric solutions. \index{Isohydric solutions!Arrhenius' theory of}% \Section{279.} The same method applies to four or more independent constituents combined into one or several phases. The notation of the system is given in each case by~\Eq{(216)}, and any possible reaction of the system may be reduced to the form~\Eq{(217)}, which corresponds to the condition of equilibrium~\Eq{(218)}. All the conditions of equilibrium, together with the given conditions of the system, give the number of equations which the phase rule prescribes for the determination of the state of equilibrium. When chemical interchanges between the different substances in solution are possible, as, \eg, in a solution of dissociating salts and acids with common ions, the term \emph{degree of dissociation} has no meaning, for the ions may be combined arbitrarily into dissociated molecules. For instance, in the solution \begin{multline*} \ce{$n_{0}$ H2O},\quad \ce{$n_{1}$ NaCl},\quad \ce{$n_{2}$ KCl},\quad \ce{$n_{3}$ NaNO3}, \\ \ce{$n_{4}$ KNO3},\quad \ce{$n_{5}$ Na+},\quad \ce{$n_{6}$ K+},\quad \ce{$n_{7}$ Cl-},\quad \ce{$n_{8}$ NO3-} \end{multline*} we cannot tell which of the \ce{Na+}-ions should be regarded as belonging to~\ce{NaCl}, and which to~\ce{NaNO3}. In such cases the only course is to characterize the state by the concentrations of the dissolved molecules. The above system consists of water and four salts, but, besides the solvent, only three are independent constituents, for the quantities of the \ce{Na}, the~\ce{K}, and the~\ce{Cl} determine that of the~\ce{NO3}. Accordingly, by \SecRef{204} ($\alpha = 4$, $\beta = 1$) all the concentrations are completely determined at given temperature and pressure by three of them. This is independent of other kinds of molecules, and other reactions, which, as is likely, may have to be considered in establishing the conditions of equilibrium. \Section{280.} If in a system of any number of independent \PageSep{263} constituents in any number of phases, the condition of equilibrium~\Eq{(218)} is not satisfied, \ie\ if for any virtual isothermal-isopiestic change \[ \tsum \nu_{0} \log c_{0} + \nu_{1} \log c_{1} + \nu_{2} \log c_{2} + \dots \gtrless \log K, \] then the direction of the change which will actually take place in nature is given by the condition $d\Psi > 0$ (\SecRef{147}). If we now denote by $\nu_{0}$,~$\nu_{1}$, $\nu_{2}$,~\dots, simple whole numbers, which are not only proportional to, but also of the same sign as the actual changes which take place, then we have, by~\Eq{(215)}, \[ \tsum \nu_{0} \log c_{0} + \nu_{1} \log c_{1} + \nu_{2} \log c_{2} + \dots < \log K, \] for the direction of any actual isothermal isopiestic change, whether it be a chemical change inside any single phase, or the passage of molecules between the different phases. The constant~$K$ is defined by~\Eq{(218)}. To find the connection between the difference of the expressions on the right and left and the time of the reaction is immediately suggested, and, in fact, a general law for the velocity of an irreversible isothermal isopiestic process may be thus deduced. We shall not, however, enter further into these considerations in this book. \index{Dilute solutions|)}% \index{Solutions, dilute|)}% \PageSep{264} \BackMatter \Appendix{Catalogue} \index{Catalogue}% Of the Author's Publications on Thermodynamics, excluding the applications to Electricity, with a reference to the paragraphs of this book, which deal with the same point. \selectlanguage{german} \Bibitem ``\Chg{Ueber}{Über} den zweiten Hauptsatz der mechanischen Wärmetheorïe. Inaugural-dissertation.'' München. Th.\ Ackermann. S.~1--61. 1879. (\SecRefs{106}{136}.) \Bibitem ``Gleichgewichtszustände isotroper Körper in verschieden Temperaturen, Habitationsschrift.'' München. Th.\ Ackermann. S.~1--68. 1880. (\SecRefs{153}{187}.) \Bibitem ``Die Theorie des Sättigungsgesetzes.'' Wied.\ Ann.~13. S.~535--543. 1881. (\SecRef{172}.) \Bibitem ``Verdampfen, Schmelzen, und Sublimiren.'' Wied.\ Ann.~15. S.~446--475. 1882. (\SecRefs{188}{196}.) \Bibitem ``\Chg{Ueber}{Über} das thermodynamische Gleichgewicht von Gasgemengen.'' Wied.\ Ann.~19. S.~358--378. 1883. (\SecRefs{232}{248}.) \Bibitem ``Das Princip der Erhaltung der Energie.'' Leipzig. B.~G. Teubner. S.~1--247. 1887. (\SecRefs{55}{105}.) \Bibitem ``\Chg{Ueber}{Über} das Princip der Vermehrung der Entropie.'' Erste Abhandlung. Gesetze des Verlaufs von Reaktionen, die nach constanten Gewichts-verhältnissen vor sich gehen. Wied.\ Ann.~30. S.~562--582. 1887. (\SecRefs{206}{212}.) \Bibitem ``\Chg{Ueber}{Über} das Princip der Vermehrung der Entropie.'' Zweite Abhandlung. Gesetze der Dissociation gasförmiger Verbindungen. Wied.\ Ann.~31. S.~189--203. 1887. (\SecRefs{232}{248}.) \Bibitem ``\Chg{Ueber}{Über} das Princip der Vermehrung der Entropie.'' Dritte Abhandlung. Gesetze des Eintritts beliebiger thermodynamischer und chemischer Reaktionen. Wied.\ Ann.~32. S.~462--503. 1887. (\SecRefs{232}{279}.) \Bibitem ``\Chg{Ueber}{Über} die Molekulare Constitution verdünnter Lösungen.'' \PageSep{265} Zeitschr.\ f.~phys.\ Chem.~1. S.~577--582. 1887. (\SSecRef{271}, \SecNum{273}.) \Bibitem ``Das chemische Gleichgewicht in verdünnten Lösungen.'' Wied.\ Ann.~34. S.~139--154. 1888. (\SecRef{262}~f., \SecRefs{268}{273}.) \Bibitem {\Loosen ``\Chg{Ueber}{Über} die Hypothese der Dissociation der Salze in sehr verdünnten Lösungen.'' Zeitschr.\ f.~phys.\ Chem.~2. S.~343. 1888. (\SecRef{271}.)} \Bibitem ``\Chg{Ueber}{Über} die Dampfspannung von verdünnten Lösungen flüchtiger Stoffe.'' Zeitchr.\ f.~phys.\ Chem.~2. S.~405--414. 1888. (\SecRef{274}). \Bibitem ``\Chg{Ueber}{Über} den osmotischen Druck.'' Zeitschr.\ f.~phys.\ Chem.~6. S.~187--189. 1890. (\SSecRef{229}, \SecNum{272}.) \Bibitem ``Allgemeines zur neuren Entwicklung der Wärmetheorie.'' Zeitschr.\ f.~phys.\ Chem.~8. S.~647--656. 1891. (\SecRef{136}.) \Bibitem ``Bemerkung über das Carnot-Clausius'sche Princip.'' Wied.\ Ann.~46. S.~162--166. 1892. (\SecRef{134}.) \Bibitem ``Erwiderung auf einen von Herrn Arrhenius erhobenen Einwand.'' Zeitschr.\ f.~phys.\ Chem.~9. S.~636~f. 1892. (\SecRef{253}.) \Bibitem ``Der Kern des zweiten Hauptsatzes der Wärmetheorie.'' Zeitschr.\ f.~d.\ phys.\ und chem.\ Untemcht~6. S.~217--221. 1893. (\SecRefs{106}{115}.) \Bibitem ``Grundriss der allgemeinen Thermochemie.'' Breslau. E.\ Trewendt. S.~1--140. 1893. (\SecRefs{1}{66}, \SecRefs[]{92}{152}, \SecRefs[]{197}{279}.) \Bibitem ``Gegen die neuere Energetik.'' Wied.\ Ann.~57. S.~72--78. 1896. (\SecRefs{108}{113}.) \PageSep{266} % [** TN: Blank Page] \PageSep{267} \selectlanguage{UKenglish} \printindex % [** TN: Index text] \iffalse INDEX The numbers refer to pages. % A Abnormal vapour densities 30 Absolute temperature 6 deduced from Thomson and Joule's experiments 127-131 Acetate silver 244 sodium 256 Acetic acid 237 Adiabatic process 59, 109 Affinity of hydrogen for oxygen 112 Aggregation, states of 69, 132 co-existence of states of 153 Air, composition of 11 Ammonium carbamate, evaporation of 188 Ammonium chloride, evaporation of 188 Andrews 14, 140 Apt, Dr.\ Richard 14 Arrhenius 235, 238, 258 Arrhenius' theory of electrolytic dissociation 252 theory of isohydric solutions 262 Atmospheric pressure 4 Atom, definition of 25 Atomic heat 34 Avogadro's law 25, 27 % B Babo's law 198 Barus 20 Berthelot 71-73, 243 Berthelot's principle 113 Binary electrolyte 237 Bodenstein 219 Boiling point, elevation of 200 Boyle and Gay-Lussac 197 Boyle's law 5, 57 % C Calorie laboratory 33 large 33 mean 33 small 33 zero 33 Calorimetric bomb 71 Cantor@Cantor, Hr.|indexnote#Cantor 227 Carbon, combustion of 74 Carbon dioxide Van der Waals' constants for 14 isotherms of 15 Carnot's theory 36 cycle 62, 100 Catalogue 264 Characteristic constant 11 Characteristic equation 5, 6, 11 deduced from Thomson and Joule's experiments 126 Clapeyron 142 Clausius 87 Clausius' equation 14, 140 form of second law 96 notation 55 statement of first and second laws 101 Coefficient of compressibility 8 of elasticity 7 of expansion 7 of pressure 7 Coexistence of states of aggregation 153 Combustion, of carbon 74 influence of temperature on 76 Condensed system 181 Condition of complete reversibility %\PageSep{268} of a process 91 of equilibrium 115, 136, 176 of a gas mixture 215-217 Conductivity of water, electrical 236 Conservation of energy 38, 40 Constituents, independent 173 Corresponding point. 161 Crafts 220 Critical point 17 of \ce{CO2}#CO 19 pressure 17 solution temperature 182 specific volume 17 temperature 17, 152 Curves of evaporation 158, 161 of fusion 158, 161 of sublimation 158, 161 Cycle of operations 44 % D Dalton's law 10, 20 Davy 36 Decrease of free energy by dilution 112 Deductions from second law of thermodynamics 105 Density specific 7 abnormal vapour 30 Depression of freezing point 250 Developable surface 164 Deviation from perfect gases 13, 123 Difference of specific heats 121 Diffusion 9 increase of entropy by 214 irreversible 214 Dilute solutions 223-263 energy of 224 entropy of 226 thermodynamical theory of 222 volume of 225 Dilution decrease of free energy by 112 heat of 198 infinite 70 law of, of binary electrolytes 238 Direction of natural process 108 Dissipation of energy. 101 Dissociation graded 221 of \ce{H2SO4}#SO 238 of hydriodic acid 219 of iodine vapour 220 of water 234 Arrhenius' theory of electrolytic 252 Distribution law (Nernst's) 254 Divariant system 181 Duhem@Duhem|indexnote#Duhem 115 Dulong and Petit's law 34 Dyne 4 % E Elasticity, coefficient of 7 Electrical conductivity of water 236 Electrolyte, binary 237 Electrolytic dissociation Arrhenius' theory of 252 Elevation of boiling point 200 Endothermal process 37 Energetics 79, 84 Energy change of 43 conservation of 38, 40 definition of 39 dissipation of 101 free 110 of perfect gas 113 internal 47 of perfect gas 57 latent 110 of a solution 70 of dilute solution 224 of gas mixture 209 potential 45 total 110 zero 44 Entropy definition 97 diminution of 93 increase of, by diffusion 214 maximum value 117 of a gas 89 of a system of gases 92 of dilute solution 225 of gas mixture 209-214 principle of increase of 100 specific 119 Equation characteristic 5, 6, 11 deduced from Thomson and Joule's experiments 126 Clausius' 14, 140 Van der Waals' 13 Equilibrium thermal 2 conditions of 115, 136, 176 of gas mixture 215, 217 Equivalent weight 23 Equivalents, number of 23 Euler 176 Evaporation of ammonium carbamate 188 of ammonium chloride 188 theory of 135 Exothermal process 37 Expansion, coefficient of 7 External conditions of equilibrium 136 effect 39 variable 178 work in complete cycle 54 work in reversible process 51, 52 %\PageSep{269} % F Favre 74 First law of thermodynamics 38, 42, 46 Free energy 110 change of, with temperature. 113 decrease of, by dilution 112 minimum value of 117 of a perfect gas 113 Freezing point depression of 250 Function $\Psi$ 114 Fundamental point (triple) 155 pressure 154 temperature 154 temperature of ice 154 triangle 159 Fusion, curve 158, 161 theory of 135 % G Gas constant 27 thermometer 3 volume 27 Gas mixture 9 energy of 209 entropy of 209-214 volume 28 Gases, perfect 5, 57 Gaseous system 207-222 Gay-Lussac 24, 57 Gay-Lussac's law, deviations from 123 Gibbs 73, 173, 212, 232 Gibbs's phase rule 179, 232 Graded dissociation 221 Gram-calorie, mechanical equivalent of 41 % H Heat absorbed 53 atomic 34 capacity 33 conception of 1 molecular 34 of perfect gases 58 of combustion 75, 76 of dilution 198 of formation of \ce{CO2}, of \ce{CS2}, of \ce{CH4}#formation 75 of fusion 37 of neutralization 73 of precipitation 201 of solidification 201 of solution 190 of sublimation 37 of vaporization 37 quantity 32 specific (definition) 33 total 36 unit 32 Heat and work, analogy between 53 Heat effect 37 at constant pressure 71 in thermochemistry 68 of dilution of \ce{H2SO4}#SO 70 Heat function at constant pressure 73 Heat, latent theory 110 approximation formula 143 Heating at constant pressure 56 at constant volume 56 Henry's law 242 Hertz, H. 146 Heterogeneous system 180 Heydweiller 234 Hirn 148 Homogeneous substance 138 system 119-131 Horstmann 188 Hydriodic acid, dissociation of 219 Hydrobromamylene 30 Hydrogen, affinity of, for oxygen 112 Hydrogen peroxide 74 % I Independent constituents 173 Inertia resistance 116 Infinite dilution 70 Infinitely slow compression 50 process 49-51 Inflection point of 17 Influence of pressure on specific heat 123 of temperature on combustion 76 Internal conditions of equilibrium 136 variable 178 Internal energy 47 of perfect gas 48 Iodine vapour, dissociation of 220 Irreversible diffusion 214 Isobaric change 7 Isochoric change 7 Isohydric solutions 258 Arrhenius' theory of 262 Isomorphous substance 182 Isopiestic change 7 Isopycnic change 7 Isothermal processes 110 %\PageSep{270} Isothermal-isopiestic process 114 Isotherms of \ce{CO2}#CO 15 Isotropic bodies 3 % J Jahn 245 Joule 36 Joule's experiments 40-42, 47 Joule and Thomson's absolute temperature 127-131 experiments 48, 57 (theory) 124 % K Kirchhoff 191 Kirchhoff's formula 198 Kohlrausch 234 Konowalow 196 Krigar@Krigar-Manzel|indexnote#Krigar 99 Kundt 122 % L Latent energy 110 Latent heat 37, 140, 143 from phase rule 187-189 Laws: Avogadro's 25, 57 Babo's 198 Boyle's 5, 57 Dalton's 10, 20 Dulong and Petit's 34 Gay-Lussac's 6, 24, 57 Henry's 242 Mariotte's 5 Nernst's 254 Neumann's (Regnault) 35 Ostwald's 238 Van't Hoff's 255 Wüllner's 199 Laws of thermodynamics@Laws of thermodynamics|see{\textit{First} and \text{Second}}#Thermodynamics 0 Lead sulphide 68 Liquefaction pressure 20 Lowering of freezing point 202 of vapour pressure 199, 250 % M Mariotte's law 5 Maximum value of entropy 117 of free energy 117 of $\Psi$ 118 Maximum work 111 Maxwell 87 Mechanical equivalent of a gram-calorie 41 of heat 40 in absolute units 42 Meier, Fr. 220 Melting point of ice 146 lowering of, by pressure 146 Membranes, semipermeable 29, 203 Meyer, Robert 62 Mixture of gases 9 Mixtures 20 Molecular heat 34 of perfect gases 58 Molecular weight 22 apparent 28 Molecules, number of 25 % N Naccari 242 Natural process, direction of 108 Nernst 242, 254, 261 Nernst's law of distribution 254 Neumann, F. 35 Neutralization, heat of 73 Nitrogen oxides 23 peroxide 30 Non-variant system 179 Noyes 261 % O Osmotic pressure 204, 251 Ostwald's law 238 Oxides of nitrogen 23 % P Pagliani 242 Partial pressures 10 Perfect gases 5, 57 system 44 Phase defined 173 rule 179 Phosphorus pentachloride 30 Planck 228 Point $(n + 2)$-ple#-ple 179 of inflection 17 triple 155, 180 quadruple 180 quintuple 180 Porous plug experiments 48 Potassium chlorate 182 Potential, energy 45 thermodynamic@thermodynamic|indexnote#thermodynamic 115 %\PageSep{271} Precipitation, heat of 201 Pressure coefficient 7 of mercury 9 Pressure fundamental 154 osmotic 204, 251 of liquefaction 20 Principle of Berthelot 113 Process adiabatic 109 endothermal 37 isothermal 110 isothermal-isopiestic 114 exothermal 37 Processes periodic 83 reversible and irreversible 82 % Q Quadruple point 180 Quantity of heat 32 Quintuple point 180 % R Ratio of specific heats 59, 122 Regnault 58, 143, 148 Resistance inertia 116 Reversibility of a process complete condition of 94 Roozeboom, Bakhuis 179 Rumford 36 % S Saturation point 16 Second law of thermodynamics introduction 77 proof 86 possible limitations 103 deductions 105 test of 147, 148 Semipermeable membranes 29, 203 Silbermann 74 Silver acetate 244 bromate 259 nitrate 244, 259 Singular values 37 Sodium carbonate 73 hydrate 73 Solidification heat of 201 pressure 20 Solution heat of 190 isohydric 258 Solutions, dilute 223-263 Solvent 196 Sound, velocity of 61 Specific density 7 Specific entropy 119 Specific heat 33 at constant pressure 56, 59, 120 at constant volume 56, 59, 120 influence of temperature on, at constant pressure 123 of saturated vapour 150 of steam 148 Specific heats difference of 121 ratio of 59, 122 Spring 20 States of aggregation 69, 132 co-existence of 153 Stohmann 71 Sublimation, curve 158, 161 theory of 135 Substance, isomorphous 182 Succinic acid 243 Sulphur 31 dioxide and water equilibrium 180 Sulphuric acid, dissociation of 238 Surface, developable 164 System condensed 181 divariant 181 gaseous 207-222 heterogeneous 180 homogeneous 119-131 non-variant 179 perfect 44 univariant 180 % T Temperature absolute 6 critical 17, 152 critical solution 182 definition of 2, 3 fundamental 154 of ice 154 Thallium chlorate 182 Theoretical regions 19 Thermal equilibrium 2 Thermochemical symbols 68 Thermodynamic potential 115 Thermodynamical theory of dilute solutions 222 of fusion, vaporization, and sublimation 135 Thermometer, gas 3 Thiesen 140 Thomsen, J. 68, 73, 235 Thomson 36, 87, 146 Thomson@Thomson|see{Joule and Thomson}#Thomson Transformability of heat into work 80 Triangle, fundamental 159 Triple point 155, 180 %\PageSep{272} % U Unit of heat 32 Univariant system 180 % V Van der Waals' constants for \ce{CO2}#CO 14 equation 13 Van't Hoff's laws 253 Vaporization curve 158, 161 theory of 135 Vapour densities, abnormal 30 Vapour pressure, lowering of 199 Variable, internal and external 178 % W Warburg 122 Water, dissociation of 234 Weights molecular and equivalent 22, 23 Work and heat, analogy between 53 external, in reversible process 51 maximum 111 Wüllner's law 199 % Z Zero calorie 33 energy 44 state 92 %[** TN End of index] \fi % PRINTED BY WILLIAM CLOWES AND SONS, LIMITED, LONDON AND BECCLES. %%%%%%%%%%%%%%%%%%%%%%%%% GUTENBERG LICENSE %%%%%%%%%%%%%%%%%%%%%%%%%% \PGLicense \begin{PGtext} End of Project Gutenberg's Treatise on Thermodynamics, by Max Planck *** END OF THIS PROJECT GUTENBERG EBOOK TREATISE ON THERMODYNAMICS *** ***** This file should be named 50880-t.tex or 50880-pdf.zip ***** This and all associated files of various formats will be found in: http://www.gutenberg.org/5/0/8/8/50880/ Produced by Andrew D. Hwang Updated editions will replace the previous one--the old editions will be renamed. Creating the works from print editions not protected by U.S. copyright law means that no one owns a United States copyright in these works, so the Foundation (and you!) can copy and distribute it in the United States without permission and without paying copyright royalties. Special rules, set forth in the General Terms of Use part of this license, apply to copying and distributing Project Gutenberg-tm electronic works to protect the PROJECT GUTENBERG-tm concept and trademark. Project Gutenberg is a registered trademark, and may not be used if you charge for the eBooks, unless you receive specific permission. If you do not charge anything for copies of this eBook, complying with the rules is very easy. You may use this eBook for nearly any purpose such as creation of derivative works, reports, performances and research. They may be modified and printed and given away--you may do practically ANYTHING in the United States with eBooks not protected by U.S. copyright law. Redistribution is subject to the trademark license, especially commercial redistribution. START: FULL LICENSE THE FULL PROJECT GUTENBERG LICENSE PLEASE READ THIS BEFORE YOU DISTRIBUTE OR USE THIS WORK To protect the Project Gutenberg-tm mission of promoting the free distribution of electronic works, by using or distributing this work (or any other work associated in any way with the phrase "Project Gutenberg"), you agree to comply with all the terms of the Full Project Gutenberg-tm License available with this file or online at www.gutenberg.org/license. Section 1. General Terms of Use and Redistributing Project Gutenberg-tm electronic works 1.A. By reading or using any part of this Project Gutenberg-tm electronic work, you indicate that you have read, understand, agree to and accept all the terms of this license and intellectual property (trademark/copyright) agreement. If you do not agree to abide by all the terms of this agreement, you must cease using and return or destroy all copies of Project Gutenberg-tm electronic works in your possession. If you paid a fee for obtaining a copy of or access to a Project Gutenberg-tm electronic work and you do not agree to be bound by the terms of this agreement, you may obtain a refund from the person or entity to whom you paid the fee as set forth in paragraph 1.E.8. 1.B. "Project Gutenberg" is a registered trademark. It may only be used on or associated in any way with an electronic work by people who agree to be bound by the terms of this agreement. There are a few things that you can do with most Project Gutenberg-tm electronic works even without complying with the full terms of this agreement. See paragraph 1.C below. There are a lot of things you can do with Project Gutenberg-tm electronic works if you follow the terms of this agreement and help preserve free future access to Project Gutenberg-tm electronic works. See paragraph 1.E below. 1.C. The Project Gutenberg Literary Archive Foundation ("the Foundation" or PGLAF), owns a compilation copyright in the collection of Project Gutenberg-tm electronic works. Nearly all the individual works in the collection are in the public domain in the United States. If an individual work is unprotected by copyright law in the United States and you are located in the United States, we do not claim a right to prevent you from copying, distributing, performing, displaying or creating derivative works based on the work as long as all references to Project Gutenberg are removed. Of course, we hope that you will support the Project Gutenberg-tm mission of promoting free access to electronic works by freely sharing Project Gutenberg-tm works in compliance with the terms of this agreement for keeping the Project Gutenberg-tm name associated with the work. You can easily comply with the terms of this agreement by keeping this work in the same format with its attached full Project Gutenberg-tm License when you share it without charge with others. 1.D. The copyright laws of the place where you are located also govern what you can do with this work. Copyright laws in most countries are in a constant state of change. If you are outside the United States, check the laws of your country in addition to the terms of this agreement before downloading, copying, displaying, performing, distributing or creating derivative works based on this work or any other Project Gutenberg-tm work. The Foundation makes no representations concerning the copyright status of any work in any country outside the United States. 1.E. Unless you have removed all references to Project Gutenberg: 1.E.1. The following sentence, with active links to, or other immediate access to, the full Project Gutenberg-tm License must appear prominently whenever any copy of a Project Gutenberg-tm work (any work on which the phrase "Project Gutenberg" appears, or with which the phrase "Project Gutenberg" is associated) is accessed, displayed, performed, viewed, copied or distributed: This eBook is for the use of anyone anywhere in the United States and most other parts of the world at no cost and with almost no restrictions whatsoever. You may copy it, give it away or re-use it under the terms of the Project Gutenberg License included with this eBook or online at www.gutenberg.org. If you are not located in the United States, you'll have to check the laws of the country where you are located before using this ebook. 1.E.2. If an individual Project Gutenberg-tm electronic work is derived from texts not protected by U.S. copyright law (does not contain a notice indicating that it is posted with permission of the copyright holder), the work can be copied and distributed to anyone in the United States without paying any fees or charges. If you are redistributing or providing access to a work with the phrase "Project Gutenberg" associated with or appearing on the work, you must comply either with the requirements of paragraphs 1.E.1 through 1.E.7 or obtain permission for the use of the work and the Project Gutenberg-tm trademark as set forth in paragraphs 1.E.8 or 1.E.9. 1.E.3. If an individual Project Gutenberg-tm electronic work is posted with the permission of the copyright holder, your use and distribution must comply with both paragraphs 1.E.1 through 1.E.7 and any additional terms imposed by the copyright holder. Additional terms will be linked to the Project Gutenberg-tm License for all works posted with the permission of the copyright holder found at the beginning of this work. 1.E.4. Do not unlink or detach or remove the full Project Gutenberg-tm License terms from this work, or any files containing a part of this work or any other work associated with Project Gutenberg-tm. 1.E.5. Do not copy, display, perform, distribute or redistribute this electronic work, or any part of this electronic work, without prominently displaying the sentence set forth in paragraph 1.E.1 with active links or immediate access to the full terms of the Project Gutenberg-tm License. 1.E.6. You may convert to and distribute this work in any binary, compressed, marked up, nonproprietary or proprietary form, including any word processing or hypertext form. However, if you provide access to or distribute copies of a Project Gutenberg-tm work in a format other than "Plain Vanilla ASCII" or other format used in the official version posted on the official Project Gutenberg-tm web site (www.gutenberg.org), you must, at no additional cost, fee or expense to the user, provide a copy, a means of exporting a copy, or a means of obtaining a copy upon request, of the work in its original "Plain Vanilla ASCII" or other form. Any alternate format must include the full Project Gutenberg-tm License as specified in paragraph 1.E.1. 1.E.7. Do not charge a fee for access to, viewing, displaying, performing, copying or distributing any Project Gutenberg-tm works unless you comply with paragraph 1.E.8 or 1.E.9. 1.E.8. You may charge a reasonable fee for copies of or providing access to or distributing Project Gutenberg-tm electronic works provided that * You pay a royalty fee of 20% of the gross profits you derive from the use of Project Gutenberg-tm works calculated using the method you already use to calculate your applicable taxes. The fee is owed to the owner of the Project Gutenberg-tm trademark, but he has agreed to donate royalties under this paragraph to the Project Gutenberg Literary Archive Foundation. Royalty payments must be paid within 60 days following each date on which you prepare (or are legally required to prepare) your periodic tax returns. Royalty payments should be clearly marked as such and sent to the Project Gutenberg Literary Archive Foundation at the address specified in Section 4, "Information about donations to the Project Gutenberg Literary Archive Foundation." * You provide a full refund of any money paid by a user who notifies you in writing (or by e-mail) within 30 days of receipt that s/he does not agree to the terms of the full Project Gutenberg-tm License. You must require such a user to return or destroy all copies of the works possessed in a physical medium and discontinue all use of and all access to other copies of Project Gutenberg-tm works. * You provide, in accordance with paragraph 1.F.3, a full refund of any money paid for a work or a replacement copy, if a defect in the electronic work is discovered and reported to you within 90 days of receipt of the work. * You comply with all other terms of this agreement for free distribution of Project Gutenberg-tm works. 1.E.9. If you wish to charge a fee or distribute a Project Gutenberg-tm electronic work or group of works on different terms than are set forth in this agreement, you must obtain permission in writing from both the Project Gutenberg Literary Archive Foundation and The Project Gutenberg Trademark LLC, the owner of the Project Gutenberg-tm trademark. Contact the Foundation as set forth in Section 3 below. 1.F. 1.F.1. Project Gutenberg volunteers and employees expend considerable effort to identify, do copyright research on, transcribe and proofread works not protected by U.S. copyright law in creating the Project Gutenberg-tm collection. Despite these efforts, Project Gutenberg-tm electronic works, and the medium on which they may be stored, may contain "Defects," such as, but not limited to, incomplete, inaccurate or corrupt data, transcription errors, a copyright or other intellectual property infringement, a defective or damaged disk or other medium, a computer virus, or computer codes that damage or cannot be read by your equipment. 1.F.2. LIMITED WARRANTY, DISCLAIMER OF DAMAGES - Except for the "Right of Replacement or Refund" described in paragraph 1.F.3, the Project Gutenberg Literary Archive Foundation, the owner of the Project Gutenberg-tm trademark, and any other party distributing a Project Gutenberg-tm electronic work under this agreement, disclaim all liability to you for damages, costs and expenses, including legal fees. YOU AGREE THAT YOU HAVE NO REMEDIES FOR NEGLIGENCE, STRICT LIABILITY, BREACH OF WARRANTY OR BREACH OF CONTRACT EXCEPT THOSE PROVIDED IN PARAGRAPH 1.F.3. YOU AGREE THAT THE FOUNDATION, THE TRADEMARK OWNER, AND ANY DISTRIBUTOR UNDER THIS AGREEMENT WILL NOT BE LIABLE TO YOU FOR ACTUAL, DIRECT, INDIRECT, CONSEQUENTIAL, PUNITIVE OR INCIDENTAL DAMAGES EVEN IF YOU GIVE NOTICE OF THE POSSIBILITY OF SUCH DAMAGE. 1.F.3. LIMITED RIGHT OF REPLACEMENT OR REFUND - If you discover a defect in this electronic work within 90 days of receiving it, you can receive a refund of the money (if any) you paid for it by sending a written explanation to the person you received the work from. If you received the work on a physical medium, you must return the medium with your written explanation. The person or entity that provided you with the defective work may elect to provide a replacement copy in lieu of a refund. If you received the work electronically, the person or entity providing it to you may choose to give you a second opportunity to receive the work electronically in lieu of a refund. If the second copy is also defective, you may demand a refund in writing without further opportunities to fix the problem. 1.F.4. Except for the limited right of replacement or refund set forth in paragraph 1.F.3, this work is provided to you 'AS-IS', WITH NO OTHER WARRANTIES OF ANY KIND, EXPRESS OR IMPLIED, INCLUDING BUT NOT LIMITED TO WARRANTIES OF MERCHANTABILITY OR FITNESS FOR ANY PURPOSE. 1.F.5. Some states do not allow disclaimers of certain implied warranties or the exclusion or limitation of certain types of damages. If any disclaimer or limitation set forth in this agreement violates the law of the state applicable to this agreement, the agreement shall be interpreted to make the maximum disclaimer or limitation permitted by the applicable state law. The invalidity or unenforceability of any provision of this agreement shall not void the remaining provisions. 1.F.6. INDEMNITY - You agree to indemnify and hold the Foundation, the trademark owner, any agent or employee of the Foundation, anyone providing copies of Project Gutenberg-tm electronic works in accordance with this agreement, and any volunteers associated with the production, promotion and distribution of Project Gutenberg-tm electronic works, harmless from all liability, costs and expenses, including legal fees, that arise directly or indirectly from any of the following which you do or cause to occur: (a) distribution of this or any Project Gutenberg-tm work, (b) alteration, modification, or additions or deletions to any Project Gutenberg-tm work, and (c) any Defect you cause. Section 2. Information about the Mission of Project Gutenberg-tm Project Gutenberg-tm is synonymous with the free distribution of electronic works in formats readable by the widest variety of computers including obsolete, old, middle-aged and new computers. It exists because of the efforts of hundreds of volunteers and donations from people in all walks of life. Volunteers and financial support to provide volunteers with the assistance they need are critical to reaching Project Gutenberg-tm's goals and ensuring that the Project Gutenberg-tm collection will remain freely available for generations to come. In 2001, the Project Gutenberg Literary Archive Foundation was created to provide a secure and permanent future for Project Gutenberg-tm and future generations. To learn more about the Project Gutenberg Literary Archive Foundation and how your efforts and donations can help, see Sections 3 and 4 and the Foundation information page at www.gutenberg.org Section 3. Information about the Project Gutenberg Literary Archive Foundation The Project Gutenberg Literary Archive Foundation is a non profit 501(c)(3) educational corporation organized under the laws of the state of Mississippi and granted tax exempt status by the Internal Revenue Service. The Foundation's EIN or federal tax identification number is 64-6221541. Contributions to the Project Gutenberg Literary Archive Foundation are tax deductible to the full extent permitted by U.S. federal laws and your state's laws. The Foundation's principal office is in Fairbanks, Alaska, with the mailing address: PO Box 750175, Fairbanks, AK 99775, but its volunteers and employees are scattered throughout numerous locations. Its business office is located at 809 North 1500 West, Salt Lake City, UT 84116, (801) 596-1887. Email contact links and up to date contact information can be found at the Foundation's web site and official page at www.gutenberg.org/contact For additional contact information: Dr. Gregory B. Newby Chief Executive and Director gbnewby@pglaf.org Section 4. Information about Donations to the Project Gutenberg Literary Archive Foundation Project Gutenberg-tm depends upon and cannot survive without wide spread public support and donations to carry out its mission of increasing the number of public domain and licensed works that can be freely distributed in machine readable form accessible by the widest array of equipment including outdated equipment. Many small donations ($1 to $5,000) are particularly important to maintaining tax exempt status with the IRS. The Foundation is committed to complying with the laws regulating charities and charitable donations in all 50 states of the United States. Compliance requirements are not uniform and it takes a considerable effort, much paperwork and many fees to meet and keep up with these requirements. We do not solicit donations in locations where we have not received written confirmation of compliance. To SEND DONATIONS or determine the status of compliance for any particular state visit www.gutenberg.org/donate While we cannot and do not solicit contributions from states where we have not met the solicitation requirements, we know of no prohibition against accepting unsolicited donations from donors in such states who approach us with offers to donate. International donations are gratefully accepted, but we cannot make any statements concerning tax treatment of donations received from outside the United States. U.S. laws alone swamp our small staff. Please check the Project Gutenberg Web pages for current donation methods and addresses. Donations are accepted in a number of other ways including checks, online payments and credit card donations. To donate, please visit: www.gutenberg.org/donate Section 5. General Information About Project Gutenberg-tm electronic works. Professor Michael S. Hart was the originator of the Project Gutenberg-tm concept of a library of electronic works that could be freely shared with anyone. For forty years, he produced and distributed Project Gutenberg-tm eBooks with only a loose network of volunteer support. Project Gutenberg-tm eBooks are often created from several printed editions, all of which are confirmed as not protected by copyright in the U.S. unless a copyright notice is included. Thus, we do not necessarily keep eBooks in compliance with any particular paper edition. Most people start at our Web site which has the main PG search facility: www.gutenberg.org This Web site includes information about Project Gutenberg-tm, including how to make donations to the Project Gutenberg Literary Archive Foundation, how to help produce our new eBooks, and how to subscribe to our email newsletter to hear about new eBooks. \end{PGtext} % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % % % % End of Project Gutenberg's Treatise on Thermodynamics, by Max Planck % % % % *** END OF THIS PROJECT GUTENBERG EBOOK TREATISE ON THERMODYNAMICS *** % % % % ***** This file should be named 50880-t.tex or 50880-pdf.zip ***** % % This and all associated files of various formats will be found in: % % http://www.gutenberg.org/5/0/8/8/50880/ % % % % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % \end{document} ### @ControlwordReplace = ( ['\\BookHead', 'Treatise on Thermodynamics'], ['\\cf', 'cf.'], ['\\Cf', 'Cf.'], ['\\ie', 'i.e.'], ['\\ie', 'i.e.'], ['\\eg', 'e.g.'] ); @ControlwordArguments = ( ['\\TPage', 1, 0, '', '', 1, 1, '', ''], ['\\BookMark', 1, 0, '', '', 1, 0, '', ''], ['\\Graphic', 1, 0, '', '', 1, 0, '', ''], ['\\Signature', 1, 1, '', ' ', 1, 1, '', ' ', 1, 1, '', ' ', 1, 1, '', ''], ['\\Figure', 0, 0, '', '', 1, 0, '
', ''], ['\\ce', 1, 0, '', ''], ['\\First', 1, 1, '', ''], ['\\Part', 1, 1, 'Part ', ' ', 1, 1, '', ''], ['\\Chapter', 0, 0, '', '', 1, 1, 'Chapter ', ' ', 1, 1, '', ''], ['\\Section', 1, 1, '§ ', ''], ['\\SecRef', 1, 1, '§', ''], ['\\SSecRef', 1, 1, '§§', ''], ['\\SecRefs', 0, 0, '', '', 1, 1, '§§', '--', 1, 1, '', ''], ['\\SecNum', 1, 1, '', ''], ['\\Fig', 1, 1, 'Fig. ', ''], ['\\Eq', 1, 1, '', ''], ['\\Erratum', 1, 0, '', '', 1, 1, '', ''], ['\\Typo', 1, 0, '', '', 1, 1, '', ''], ['\\Add', 1, 1, '', ''], ['\\Chg', 1, 0, '', '', 1, 1, '', ''] ); $PageSeparator = qr/^\\PageSep/; $CustomClean = 'print "\\nCustom cleaning in progress..."; my $cline = 0; while ($cline <= $#file) { $file[$cline] =~ s/--------[^\n]*\n//; # strip page separators $cline++ } print "done\\n";'; ### This is pdfTeX, Version 3.1415926-2.5-1.40.14 (TeX Live 2013/Debian) (format=pdflatex 2015.9.16) 8 JAN 2016 19:13 entering extended mode %&-line parsing enabled. **50880-t.tex (./50880-t.tex LaTeX2e <2011/06/27> Babel <3.9h> and hyphenation patterns for 78 languages loaded. (/usr/share/texlive/texmf-dist/tex/latex/base/book.cls Document Class: book 2007/10/19 v1.4h Standard LaTeX document class (/usr/share/texlive/texmf-dist/tex/latex/base/bk12.clo File: bk12.clo 2007/10/19 v1.4h Standard LaTeX file (size option) ) \c@part=\count79 \c@chapter=\count80 \c@section=\count81 \c@subsection=\count82 \c@subsubsection=\count83 \c@paragraph=\count84 \c@subparagraph=\count85 \c@figure=\count86 \c@table=\count87 \abovecaptionskip=\skip41 \belowcaptionskip=\skip42 \bibindent=\dimen102 ) (/usr/share/texlive/texmf-dist/tex/latex/base/inputenc.sty Package: inputenc 2008/03/30 v1.1d Input encoding file \inpenc@prehook=\toks14 \inpenc@posthook=\toks15 (/usr/share/texlive/texmf-dist/tex/latex/base/latin1.def File: latin1.def 2008/03/30 v1.1d Input encoding file )) (/usr/share/texlive/texmf-dist/tex/generic/babel/babel.sty Package: babel 2013/12/03 3.9h The Babel package (/usr/share/texlive/texmf-dist/tex/generic/babel-german/germanb.ldf Language: germanb 2013/12/13 v2.7 German support for babel (traditional orthogr aphy) (/usr/share/texlive/texmf-dist/tex/generic/babel/babel.def File: babel.def 2013/12/03 3.9h Babel common definitions \babel@savecnt=\count88 \U@D=\dimen103 ) \l@austrian = a dialect from \language\l@german Package babel Info: Making " an active character on input line 107. ) (/usr/share/texlive/texmf-dist/tex/generic/babel-english/english.ldf Language: english 2012/08/20 v3.3p English support from the babel system \l@canadian = a dialect from \language\l@american \l@australian = a dialect from \language\l@british \l@newzealand = a dialect from \language\l@british )) (/usr/share/texlive/texmf-dist/tex/latex/base/ifthen.sty Package: ifthen 2001/05/26 v1.1c Standard LaTeX ifthen package (DPC) ) (/usr/share/texlive/texmf-dist/tex/latex/amsmath/amsmath.sty Package: amsmath 2013/01/14 v2.14 AMS math features \@mathmargin=\skip43 For additional information on amsmath, use the `?' option. (/usr/share/texlive/texmf-dist/tex/latex/amsmath/amstext.sty Package: amstext 2000/06/29 v2.01 (/usr/share/texlive/texmf-dist/tex/latex/amsmath/amsgen.sty File: amsgen.sty 1999/11/30 v2.0 \@emptytoks=\toks16 \ex@=\dimen104 )) (/usr/share/texlive/texmf-dist/tex/latex/amsmath/amsbsy.sty Package: amsbsy 1999/11/29 v1.2d \pmbraise@=\dimen105 ) (/usr/share/texlive/texmf-dist/tex/latex/amsmath/amsopn.sty Package: amsopn 1999/12/14 v2.01 operator names ) \inf@bad=\count89 LaTeX Info: Redefining \frac on input line 210. \uproot@=\count90 \leftroot@=\count91 LaTeX Info: Redefining \overline on input line 306. \classnum@=\count92 \DOTSCASE@=\count93 LaTeX Info: Redefining \ldots on input line 378. LaTeX Info: Redefining \dots on input line 381. LaTeX Info: Redefining \cdots on input line 466. \Mathstrutbox@=\box26 \strutbox@=\box27 \big@size=\dimen106 LaTeX Font Info: Redeclaring font encoding OML on input line 566. LaTeX Font Info: Redeclaring font encoding OMS on input line 567. \macc@depth=\count94 \c@MaxMatrixCols=\count95 \dotsspace@=\muskip10 \c@parentequation=\count96 \dspbrk@lvl=\count97 \tag@help=\toks17 \row@=\count98 \column@=\count99 \maxfields@=\count100 \andhelp@=\toks18 \eqnshift@=\dimen107 \alignsep@=\dimen108 \tagshift@=\dimen109 \tagwidth@=\dimen110 \totwidth@=\dimen111 \lineht@=\dimen112 \@envbody=\toks19 \multlinegap=\skip44 \multlinetaggap=\skip45 \mathdisplay@stack=\toks20 LaTeX Info: Redefining \[ on input line 2665. LaTeX Info: Redefining \] on input line 2666. ) (/usr/share/texlive/texmf-dist/tex/latex/amsfonts/amssymb.sty Package: amssymb 2013/01/14 v3.01 AMS font symbols (/usr/share/texlive/texmf-dist/tex/latex/amsfonts/amsfonts.sty Package: amsfonts 2013/01/14 v3.01 Basic AMSFonts support \symAMSa=\mathgroup4 \symAMSb=\mathgroup5 LaTeX Font Info: Overwriting math alphabet `\mathfrak' in version `bold' (Font) U/euf/m/n --> U/euf/b/n on input line 106. )) (/usr/share/texlive/texmf-dist/tex/latex/tools/array.sty Package: array 2008/09/09 v2.4c Tabular extension package (FMi) \col@sep=\dimen113 \extrarowheight=\dimen114 \NC@list=\toks21 \extratabsurround=\skip46 \backup@length=\skip47 ) (/usr/share/texlive/texmf-dist/tex/latex/base/alltt.sty Package: alltt 1997/06/16 v2.0g defines alltt environment ) (/usr/share/texlive/texmf-dist/tex/latex/footmisc/footmisc.sty Package: footmisc 2011/06/06 v5.5b a miscellany of footnote facilities \FN@temptoken=\toks22 \footnotemargin=\dimen115 \c@pp@next@reset=\count101 \c@@fnserial=\count102 Package footmisc Info: Declaring symbol style bringhurst on input line 855. Package footmisc Info: Declaring symbol style chicago on input line 863. Package footmisc Info: Declaring symbol style wiley on input line 872. Package footmisc Info: Declaring symbol style lamport-robust on input line 883. Package footmisc Info: Declaring symbol style lamport* on input line 903. Package footmisc Info: Declaring symbol style lamport*-robust on input line 924 . ) (/usr/share/texlive/texmf-dist/tex/latex/mhchem/mhchem.sty Package: mhchem 2014/02/01 v3.15 for typesetting chemical formulae (/usr/share/texlive/texmf-dist/tex/latex/tools/calc.sty Package: calc 2007/08/22 v4.3 Infix arithmetic (KKT,FJ) \calc@Acount=\count103 \calc@Bcount=\count104 \calc@Adimen=\dimen116 \calc@Bdimen=\dimen117 \calc@Askip=\skip48 \calc@Bskip=\skip49 LaTeX Info: Redefining \setlength on input line 76. LaTeX Info: Redefining \addtolength on input line 77. \calc@Ccount=\count105 \calc@Cskip=\skip50 ) (/usr/share/texlive/texmf-dist/tex/latex/oberdiek/twoopt.sty Package: twoopt 2008/08/11 v1.5 Definitions with two optional arguments (HO) ) (/usr/share/texlive/texmf-dist/tex/latex/graphics/keyval.sty Package: keyval 1999/03/16 v1.13 key=value parser (DPC) \KV@toks@=\toks23 ) (/usr/share/texlive/texmf-dist/tex/latex/graphics/graphics.sty Package: graphics 2009/02/05 v1.0o Standard LaTeX Graphics (DPC,SPQR) (/usr/share/texlive/texmf-dist/tex/latex/graphics/trig.sty Package: trig 1999/03/16 v1.09 sin cos tan (DPC) ) (/usr/share/texlive/texmf-dist/tex/latex/latexconfig/graphics.cfg File: graphics.cfg 2010/04/23 v1.9 graphics configuration of TeX Live ) Package graphics Info: Driver file: pdftex.def on input line 91. (/usr/share/texlive/texmf-dist/tex/latex/pdftex-def/pdftex.def File: pdftex.def 2011/05/27 v0.06d Graphics/color for pdfTeX (/usr/share/texlive/texmf-dist/tex/generic/oberdiek/infwarerr.sty Package: infwarerr 2010/04/08 v1.3 Providing info/warning/error messages (HO) ) (/usr/share/texlive/texmf-dist/tex/generic/oberdiek/ltxcmds.sty Package: ltxcmds 2011/11/09 v1.22 LaTeX kernel commands for general use (HO) ) \Gread@gobject=\count106 )) (/usr/share/texlive/texmf-dist/tex/generic/oberdiek/pdftexcmds.sty Package: pdftexcmds 2011/11/29 v0.20 Utility functions of pdfTeX for LuaTeX (HO ) (/usr/share/texlive/texmf-dist/tex/generic/oberdiek/ifluatex.sty Package: ifluatex 2010/03/01 v1.3 Provides the ifluatex switch (HO) Package ifluatex Info: LuaTeX not detected. ) (/usr/share/texlive/texmf-dist/tex/generic/oberdiek/ifpdf.sty Package: ifpdf 2011/01/30 v2.3 Provides the ifpdf switch (HO) Package ifpdf Info: pdfTeX in PDF mode is detected. ) Package pdftexcmds Info: LuaTeX not detected. Package pdftexcmds Info: \pdf@primitive is available. Package pdftexcmds Info: \pdf@ifprimitive is available. Package pdftexcmds Info: \pdfdraftmode found. ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/expl3.sty (/usr/share/texli ve/texmf-dist/tex/latex/l3kernel/l3names.sty (/usr/share/texlive/texmf-dist/tex /latex/l3kernel/l3bootstrap.sty Package: l3bootstrap 2014/01/04 v4640 L3 Experimental bootstrap code ) Package: l3names 2014/01/04 v4640 L3 Namespace for primitives ) (/usr/share/texlive/texmf-dist/tex/latex/etex-pkg/etex.sty Package: etex 1998/03/26 v2.0 eTeX basic definition package (PEB) \et@xins=\count107 ) Package: expl3 2014/01/07 v4646 L3 Experimental code bundle wrapper (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3basics.sty Package: l3basics 2014/01/04 v4642 L3 Basic definitions ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3expan.sty Package: l3expan 2014/01/04 v4642 L3 Argument expansion ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3tl.sty Package: l3tl 2013/12/27 v4625 L3 Token lists ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3seq.sty Package: l3seq 2013/12/14 v4623 L3 Sequences and stacks ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3int.sty Package: l3int 2013/08/02 v4583 L3 Integers \c_max_int=\count108 \l_tmpa_int=\count109 \l_tmpb_int=\count110 \g_tmpa_int=\count111 \g_tmpb_int=\count112 ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3quark.sty Package: l3quark 2013/12/14 v4623 L3 Quarks ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3prg.sty Package: l3prg 2014/01/04 v4642 L3 Control structures \g__prg_map_int=\count113 ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3clist.sty Package: l3clist 2013/07/28 v4581 L3 Comma separated lists ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3token.sty Package: l3token 2013/08/25 v4587 L3 Experimental token manipulation ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3prop.sty Package: l3prop 2013/12/14 v4623 L3 Property lists ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3msg.sty Package: l3msg 2013/07/28 v4581 L3 Messages ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3file.sty Package: l3file 2013/10/13 v4596 L3 File and I/O operations \l_iow_line_count_int=\count114 \l__iow_target_count_int=\count115 \l__iow_current_line_int=\count116 \l__iow_current_word_int=\count117 \l__iow_current_indentation_int=\count118 ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3skip.sty Package: l3skip 2013/07/28 v4581 L3 Dimensions and skips \c_zero_dim=\dimen118 \c_max_dim=\dimen119 \l_tmpa_dim=\dimen120 \l_tmpb_dim=\dimen121 \g_tmpa_dim=\dimen122 \g_tmpb_dim=\dimen123 \c_zero_skip=\skip51 \c_max_skip=\skip52 \l_tmpa_skip=\skip53 \l_tmpb_skip=\skip54 \g_tmpa_skip=\skip55 \g_tmpb_skip=\skip56 \c_zero_muskip=\muskip11 \c_max_muskip=\muskip12 \l_tmpa_muskip=\muskip13 \l_tmpb_muskip=\muskip14 \g_tmpa_muskip=\muskip15 \g_tmpb_muskip=\muskip16 ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3keys.sty Package: l3keys 2013/12/08 v4614 L3 Experimental key-value interfaces \g__keyval_level_int=\count119 \l_keys_choice_int=\count120 ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3fp.sty Package: l3fp 2014/01/04 v4642 L3 Floating points \c__fp_leading_shift_int=\count121 \c__fp_middle_shift_int=\count122 \c__fp_trailing_shift_int=\count123 \c__fp_big_leading_shift_int=\count124 \c__fp_big_middle_shift_int=\count125 \c__fp_big_trailing_shift_int=\count126 \c__fp_Bigg_leading_shift_int=\count127 \c__fp_Bigg_middle_shift_int=\count128 \c__fp_Bigg_trailing_shift_int=\count129 ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3box.sty Package: l3box 2013/07/28 v4581 L3 Experimental boxes \c_empty_box=\box28 \l_tmpa_box=\box29 \l_tmpb_box=\box30 \g_tmpa_box=\box31 \g_tmpb_box=\box32 ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3coffins.sty Package: l3coffins 2013/12/14 v4624 L3 Coffin code layer \l__coffin_internal_box=\box33 \l__coffin_internal_dim=\dimen124 \l__coffin_offset_x_dim=\dimen125 \l__coffin_offset_y_dim=\dimen126 \l__coffin_x_dim=\dimen127 \l__coffin_y_dim=\dimen128 \l__coffin_x_prime_dim=\dimen129 \l__coffin_y_prime_dim=\dimen130 \c_empty_coffin=\box34 \l__coffin_aligned_coffin=\box35 \l__coffin_aligned_internal_coffin=\box36 \l_tmpa_coffin=\box37 \l_tmpb_coffin=\box38 \l__coffin_display_coffin=\box39 \l__coffin_display_coord_coffin=\box40 \l__coffin_display_pole_coffin=\box41 \l__coffin_display_offset_dim=\dimen131 \l__coffin_display_x_dim=\dimen132 \l__coffin_display_y_dim=\dimen133 ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3color.sty Package: l3color 2012/08/29 v4156 L3 Experimental color support ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3luatex.sty Package: l3luatex 2013/07/28 v4581 L3 Experimental LuaTeX-specific functions \g__cctab_allocate_int=\count130 \g__cctab_stack_int=\count131 ) (/usr/share/texlive/texmf-dist/tex/latex/l3kernel/l3candidates.sty Package: l3candidates 2014/01/06 v4643 L3 Experimental additions to l3kernel \l__box_top_dim=\dimen134 \l__box_bottom_dim=\dimen135 \l__box_left_dim=\dimen136 \l__box_right_dim=\dimen137 \l__box_top_new_dim=\dimen138 \l__box_bottom_new_dim=\dimen139 \l__box_left_new_dim=\dimen140 \l__box_right_new_dim=\dimen141 \l__box_internal_box=\box42 \l__coffin_bounding_shift_dim=\dimen142 \l__coffin_left_corner_dim=\dimen143 \l__coffin_right_corner_dim=\dimen144 \l__coffin_bottom_corner_dim=\dimen145 \l__coffin_top_corner_dim=\dimen146 \l__coffin_scaled_total_height_dim=\dimen147 \l__coffin_scaled_width_dim=\dimen148 )) \mhchem@ce@result=\toks24 \mhchem@ce@part=\toks25 \mhchem@arrow@deployType=\toks26 \mhchem@arrow@params=\toks27 \mhchem@arrowlength@pgf=\skip57 \mhchem@bondwidth=\skip58 \mhchem@bondheight=\skip59 \mhchem@smallbondwidth@tmpA=\skip60 \mhchem@smallbondwidth@tmpB=\skip61 \mhchem@smallbondwidth=\skip62 \__mhchem_option_version_int=\count132 \mhchem@prepostscript@tmp@i=\skip63 \mhchem@prepostscript@tmp@ii=\skip64 \mhchem@mathbox@tmp@i=\skip65 \mhchem@mathbox@tmp@ii=\skip66 \mhchem@minispace@tmp=\skip67 \mhchem@option@minussidebearingleft=\skip68 \mhchem@option@minussidebearingright=\skip69 ) (/usr/share/texlive/texmf-dist/tex/latex/tools/multicol.sty Package: multicol 2011/06/27 v1.7a multicolumn formatting (FMi) \c@tracingmulticols=\count133 \mult@box=\box43 \multicol@leftmargin=\dimen149 \c@unbalance=\count134 \c@collectmore=\count135 \doublecol@number=\count136 \multicoltolerance=\count137 \multicolpretolerance=\count138 \full@width=\dimen150 \page@free=\dimen151 \premulticols=\dimen152 \postmulticols=\dimen153 \multicolsep=\skip70 \multicolbaselineskip=\skip71 \partial@page=\box44 \last@line=\box45 \mult@rightbox=\box46 \mult@grightbox=\box47 \mult@gfirstbox=\box48 \mult@firstbox=\box49 \@tempa=\box50 \@tempa=\box51 \@tempa=\box52 \@tempa=\box53 \@tempa=\box54 \@tempa=\box55 \@tempa=\box56 \@tempa=\box57 \@tempa=\box58 \@tempa=\box59 \@tempa=\box60 \@tempa=\box61 \@tempa=\box62 \@tempa=\box63 \@tempa=\box64 \@tempa=\box65 \@tempa=\box66 \c@columnbadness=\count139 \c@finalcolumnbadness=\count140 \last@try=\dimen154 \multicolovershoot=\dimen155 \multicolundershoot=\dimen156 \mult@nat@firstbox=\box67 \colbreak@box=\box68 \multicol@sort@counter=\count141 ) (/usr/share/texlive/texmf-dist/tex/latex/base/makeidx.sty Package: makeidx 2000/03/29 v1.0m Standard LaTeX package ) (/usr/share/texlive/texmf-dist/tex/latex/graphics/graphicx.sty Package: graphicx 1999/02/16 v1.0f Enhanced LaTeX Graphics (DPC,SPQR) \Gin@req@height=\dimen157 \Gin@req@width=\dimen158 ) (/usr/share/texlive/texmf-dist/tex/latex/fancyhdr/fancyhdr.sty \fancy@headwidth=\skip72 \f@ncyO@elh=\skip73 \f@ncyO@erh=\skip74 \f@ncyO@olh=\skip75 \f@ncyO@orh=\skip76 \f@ncyO@elf=\skip77 \f@ncyO@erf=\skip78 \f@ncyO@olf=\skip79 \f@ncyO@orf=\skip80 ) (/usr/share/texlive/texmf-dist/tex/latex/geometry/geometry.sty Package: geometry 2010/09/12 v5.6 Page Geometry (/usr/share/texlive/texmf-dist/tex/generic/oberdiek/ifvtex.sty Package: ifvtex 2010/03/01 v1.5 Detect VTeX and its facilities (HO) Package ifvtex Info: VTeX not detected. ) (/usr/share/texlive/texmf-dist/tex/generic/ifxetex/ifxetex.sty Package: ifxetex 2010/09/12 v0.6 Provides ifxetex conditional ) \Gm@cnth=\count142 \Gm@cntv=\count143 \c@Gm@tempcnt=\count144 \Gm@bindingoffset=\dimen159 \Gm@wd@mp=\dimen160 \Gm@odd@mp=\dimen161 \Gm@even@mp=\dimen162 \Gm@layoutwidth=\dimen163 \Gm@layoutheight=\dimen164 \Gm@layouthoffset=\dimen165 \Gm@layoutvoffset=\dimen166 \Gm@dimlist=\toks28 ) (/usr/share/texlive/texmf-dist/tex/latex/hyperref/hyperref.sty Package: hyperref 2012/11/06 v6.83m Hypertext links for LaTeX (/usr/share/texlive/texmf-dist/tex/generic/oberdiek/hobsub-hyperref.sty Package: hobsub-hyperref 2012/05/28 v1.13 Bundle oberdiek, subset hyperref (HO) (/usr/share/texlive/texmf-dist/tex/generic/oberdiek/hobsub-generic.sty Package: hobsub-generic 2012/05/28 v1.13 Bundle oberdiek, subset generic (HO) Package: hobsub 2012/05/28 v1.13 Construct package bundles (HO) Package hobsub Info: Skipping package `infwarerr' (already loaded). Package hobsub Info: Skipping package `ltxcmds' (already loaded). Package hobsub Info: Skipping package `ifluatex' (already loaded). Package hobsub Info: Skipping package `ifvtex' (already loaded). Package: intcalc 2007/09/27 v1.1 Expandable calculations with integers (HO) Package hobsub Info: Skipping package `ifpdf' (already loaded). Package: etexcmds 2011/02/16 v1.5 Avoid name clashes with e-TeX commands (HO) Package etexcmds Info: Could not find \expanded. (etexcmds) That can mean that you are not using pdfTeX 1.50 or (etexcmds) that some package has redefined \expanded. (etexcmds) In the latter case, load this package earlier. Package: kvsetkeys 2012/04/25 v1.16 Key value parser (HO) Package: kvdefinekeys 2011/04/07 v1.3 Define keys (HO) Package hobsub Info: Skipping package `pdftexcmds' (already loaded). Package: pdfescape 2011/11/25 v1.13 Implements pdfTeX's escape features (HO) Package: bigintcalc 2012/04/08 v1.3 Expandable calculations on big integers (HO ) Package: bitset 2011/01/30 v1.1 Handle bit-vector datatype (HO) Package: uniquecounter 2011/01/30 v1.2 Provide unlimited unique counter (HO) ) Package hobsub Info: Skipping package `hobsub' (already loaded). Package: letltxmacro 2010/09/02 v1.4 Let assignment for LaTeX macros (HO) Package: hopatch 2012/05/28 v1.2 Wrapper for package hooks (HO) Package: xcolor-patch 2011/01/30 xcolor patch Package: atveryend 2011/06/30 v1.8 Hooks at the very end of document (HO) Package atveryend Info: \enddocument detected (standard20110627). Package: atbegshi 2011/10/05 v1.16 At begin shipout hook (HO) Package: refcount 2011/10/16 v3.4 Data extraction from label references (HO) Package: hycolor 2011/01/30 v1.7 Color options for hyperref/bookmark (HO) ) (/usr/share/texlive/texmf-dist/tex/latex/oberdiek/auxhook.sty Package: auxhook 2011/03/04 v1.3 Hooks for auxiliary files (HO) ) (/usr/share/texlive/texmf-dist/tex/latex/oberdiek/kvoptions.sty Package: kvoptions 2011/06/30 v3.11 Key value format for package options (HO) ) \@linkdim=\dimen167 \Hy@linkcounter=\count145 \Hy@pagecounter=\count146 (/usr/share/texlive/texmf-dist/tex/latex/hyperref/pd1enc.def File: pd1enc.def 2012/11/06 v6.83m Hyperref: PDFDocEncoding definition (HO) ) \Hy@SavedSpaceFactor=\count147 (/usr/share/texlive/texmf-dist/tex/latex/latexconfig/hyperref.cfg File: hyperref.cfg 2002/06/06 v1.2 hyperref configuration of TeXLive ) Package hyperref Info: Option `hyperfootnotes' set `false' on input line 4319. Package hyperref Info: Option `bookmarks' set `true' on input line 4319. Package hyperref Info: Option `linktocpage' set `false' on input line 4319. Package hyperref Info: Option `pdfdisplaydoctitle' set `true' on input line 431 9. Package hyperref Info: Option `pdfpagelabels' set `true' on input line 4319. Package hyperref Info: Option `bookmarksopen' set `true' on input line 4319. Package hyperref Info: Option `colorlinks' set `true' on input line 4319. Package hyperref Info: Hyper figures OFF on input line 4443. Package hyperref Info: Link nesting OFF on input line 4448. Package hyperref Info: Hyper index ON on input line 4451. Package hyperref Info: Plain pages OFF on input line 4458. Package hyperref Info: Backreferencing OFF on input line 4463. Package hyperref Info: Implicit mode ON; LaTeX internals redefined. Package hyperref Info: Bookmarks ON on input line 4688. \c@Hy@tempcnt=\count148 (/usr/share/texlive/texmf-dist/tex/latex/url/url.sty \Urlmuskip=\muskip17 Package: url 2013/09/16 ver 3.4 Verb mode for urls, etc. ) LaTeX Info: Redefining \url on input line 5041. \XeTeXLinkMargin=\dimen168 \Fld@menulength=\count149 \Field@Width=\dimen169 \Fld@charsize=\dimen170 Package hyperref Info: Hyper figures OFF on input line 6295. Package hyperref Info: Link nesting OFF on input line 6300. Package hyperref Info: Hyper index ON on input line 6303. Package hyperref Info: backreferencing OFF on input line 6310. Package hyperref Info: Link coloring ON on input line 6313. Package hyperref Info: Link coloring with OCG OFF on input line 6320. Package hyperref Info: PDF/A mode OFF on input line 6325. LaTeX Info: Redefining \ref on input line 6365. LaTeX Info: Redefining \pageref on input line 6369. \Hy@abspage=\count150 \c@Item=\count151 ) Package hyperref Message: Driver: hpdftex. (/usr/share/texlive/texmf-dist/tex/latex/hyperref/hpdftex.def File: hpdftex.def 2012/11/06 v6.83m Hyperref driver for pdfTeX \Fld@listcount=\count152 \c@bookmark@seq@number=\count153 (/usr/share/texlive/texmf-dist/tex/latex/oberdiek/rerunfilecheck.sty Package: rerunfilecheck 2011/04/15 v1.7 Rerun checks for auxiliary files (HO) Package uniquecounter Info: New unique counter `rerunfilecheck' on input line 2 82. ) \Hy@SectionHShift=\skip81 ) \TmpLen=\skip82 \@indexfile=\write3 \openout3 = `50880-t.idx'. Writing index file 50880-t.idx \c@ChapNo=\count154 (./50880-t.aux) \openout1 = `50880-t.aux'. LaTeX Font Info: Checking defaults for OML/cmm/m/it on input line 590. LaTeX Font Info: ... okay on input line 590. LaTeX Font Info: Checking defaults for T1/cmr/m/n on input line 590. LaTeX Font Info: ... okay on input line 590. LaTeX Font Info: Checking defaults for OT1/cmr/m/n on input line 590. LaTeX Font Info: ... okay on input line 590. LaTeX Font Info: Checking defaults for OMS/cmsy/m/n on input line 590. LaTeX Font Info: ... okay on input line 590. LaTeX Font Info: Checking defaults for OMX/cmex/m/n on input line 590. LaTeX Font Info: ... okay on input line 590. LaTeX Font Info: Checking defaults for U/cmr/m/n on input line 590. LaTeX Font Info: ... okay on input line 590. LaTeX Font Info: Checking defaults for PD1/pdf/m/n on input line 590. LaTeX Font Info: ... okay on input line 590. (/usr/share/texlive/texmf-dist/tex/context/base/supp-pdf.mkii [Loading MPS to PDF converter (version 2006.09.02).] \scratchcounter=\count155 \scratchdimen=\dimen171 \scratchbox=\box69 \nofMPsegments=\count156 \nofMParguments=\count157 \everyMPshowfont=\toks29 \MPscratchCnt=\count158 \MPscratchDim=\dimen172 \MPnumerator=\count159 \makeMPintoPDFobject=\count160 \everyMPtoPDFconversion=\toks30 ) *geometry* driver: auto-detecting *geometry* detected driver: pdftex *geometry* verbose mode - [ preamble ] result: * driver: pdftex * paper: * layout: * layoutoffset:(h,v)=(0.0pt,0.0pt) * hratio: 1:1 * modes: includehead includefoot twoside * h-part:(L,W,R)=(9.03374pt, 325.215pt, 9.03375pt) * v-part:(T,H,B)=(0.54498pt, 486.46004pt, 0.81747pt) * \paperwidth=343.28249pt * \paperheight=487.8225pt * \textwidth=325.215pt * \textheight=424.58624pt * \oddsidemargin=-63.23625pt * \evensidemargin=-63.23624pt * \topmargin=-71.725pt * \headheight=12.0pt * \headsep=19.8738pt * \topskip=12.0pt * \footskip=30.0pt * \marginparwidth=98.0pt * \marginparsep=7.0pt * \columnsep=10.0pt * \skip\footins=10.8pt plus 4.0pt minus 2.0pt * \hoffset=0.0pt * \voffset=0.0pt * \mag=1000 * \@twocolumnfalse * \@twosidetrue * \@mparswitchtrue * \@reversemarginfalse * (1in=72.27pt=25.4mm, 1cm=28.453pt) \AtBeginShipoutBox=\box70 (/usr/share/texlive/texmf-dist/tex/latex/graphics/color.sty Package: color 2005/11/14 v1.0j Standard LaTeX Color (DPC) LaTeX Info: Redefining \color on input line 76. (/usr/share/texlive/texmf-dist/tex/latex/latexconfig/color.cfg File: color.cfg 2007/01/18 v1.5 color configuration of teTeX/TeXLive ) Package color Info: Driver file: pdftex.def on input line 130. ) Package hyperref Info: Link coloring ON on input line 590. (/usr/share/texlive/texmf-dist/tex/latex/hyperref/nameref.sty Package: nameref 2012/10/27 v2.43 Cross-referencing by name of section (/usr/share/texlive/texmf-dist/tex/generic/oberdiek/gettitlestring.sty Package: gettitlestring 2010/12/03 v1.4 Cleanup title references (HO) ) \c@section@level=\count161 ) LaTeX Info: Redefining \ref on input line 590. LaTeX Info: Redefining \pageref on input line 590. LaTeX Info: Redefining \nameref on input line 590. (./50880-t.out) (./50880-t.out) \@outlinefile=\write4 \openout4 = `50880-t.out'. Overfull \hbox (10.53983pt too wide) in paragraph at lines 604--604 []\OT1/cmtt/m/n/8 to check the laws of the country where you are located before using this ebook.[] [] LaTeX Font Info: Try loading font information for U+msa on input line 620. (/usr/share/texlive/texmf-dist/tex/latex/amsfonts/umsa.fd File: umsa.fd 2013/01/14 v3.01 AMS symbols A ) LaTeX Font Info: Try loading font information for U+msb on input line 620. (/usr/share/texlive/texmf-dist/tex/latex/amsfonts/umsb.fd File: umsb.fd 2013/01/14 v3.01 AMS symbols B ) [1 {/var/lib/texmf/fonts/map/pdftex/updmap/pdftex.map}] [2] <./images/device.pdf, id=138, 159.71669pt x 135.8676pt> File: ./images/device.pdf Graphic file (type pdf) Package pdftex.def Info: ./images/device.pdf used on input line 676. (pdftex.def) Requested size: 108.405pt x 92.2208pt. [1 <./images/device.pdf>] [2 ] [3 ] [4] [5] [6] (./50880-t.toc [7 ]) \tf@toc=\write5 \openout5 = `50880-t.toc'. [8] LaTeX Font Info: Try loading font information for OMS+cmr on input line 983. (/usr/share/texlive/texmf-dist/tex/latex/base/omscmr.fd File: omscmr.fd 1999/05/25 v2.5h Standard LaTeX font definitions ) LaTeX Font Info: Font shape `OMS/cmr/bx/n' in size <12> not available (Font) Font shape `OMS/cmsy/b/n' tried instead on input line 983. LaTeX Font Info: Font shape `OMS/cmr/m/n' in size <8> not available (Font) Font shape `OMS/cmsy/m/n' tried instead on input line 1016. LaTeX Font Info: Font shape `OMS/cmr/m/n' in size <7> not available (Font) Font shape `OMS/cmsy/m/n' tried instead on input line 1019. [1 ] [2] LaTeX Font Info: Font shape `OMS/cmr/m/n' in size <12> not available (Font) Font shape `OMS/cmsy/m/n' tried instead on input line 1083. [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] <./images/fig1.p df, id=371, 340.27126pt x 585.18625pt> File: ./images/fig1.pdf Graphic file (type pdf) Package pdftex.def Info: ./images/fig1.pdf used on input line 1622. (pdftex.def) Requested size: 252.94499pt x 435.0044pt. LaTeX Warning: Float too large for page by 10.41815pt on input line 1622. [17] [18 <./images/fig1.pdf>] [19] [20] [21] [22] [23] [24] [25] [26 ] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38 ] [39] [40] [41] [42] [43] [44] [45 ] [46] [47] [48] [49] [50] [51] [52] [53] [54 ] [55] [56] [57] [58] [59] [60] <./images/fig2.pdf, id=662, 236.885pt x 263.986 25pt> File: ./images/fig2.pdf Graphic file (type pdf) Package pdftex.def Info: ./images/fig2.pdf used on input line 3129. (pdftex.def) Requested size: 236.88441pt x 263.9856pt. [61] [62 <./images/fig2.pdf>] [63] [64] [65] [66] [67] Overfull \hbox (2.41357pt too wide) in paragraph at lines 3382--3387 \OT1/cmr/m/n/12 Assuming that only the laws of Boyle, Gay-Lussac, and []Avogadr o [] [68] [69] [70] [71] [72] [73] [74] <./images/fig3.pdf, id=821, 242.9075pt x 264 .99pt> File: ./images/fig3.pdf Graphic file (type pdf) Package pdftex.def Info: ./images/fig3.pdf used on input line 3665. (pdftex.def) Requested size: 242.90689pt x 264.98935pt. [75] [76 <./images/fig3.pdf>] [77] [78] [79] [80 ] [81] [82] [83] [84] [85] [86] [87] [88] [89] [90] [91] [92 ] [93] [94] [95] [96] [97] [98] [99] [100] [101] [102 ] [103] [104] [105] [106] [107] [108] [109] [110] [111] [112] [113] [114] [115] [116] [117] LaTeX Font Info: Font shape `OMS/cmr/m/n' in size <10> not available (Font) Font shape `OMS/cmsy/m/n' tried instead on input line 5246. [118] [119] [120] [121] [122] [123] [124] [125 ] [126] [127] [128] [129] [130] [131] [132] [133] [134] [135] [136] [137] [138] [139] [140] [141] [142 ] [143] [144] [145] [146] [147] [148] [149] [150] [151] [152] [153] [154] [155] [156] [157] [158] [159 ] [160] [161] [162] [163] [164] [165] [166] [167] [168] [169] [170] [171] [172] [173] [174] [175] [176] [177] [178] [179] [180] [181] [182] [183] [184] [185] [ 186] [187] [188] [189] Overfull \hbox (5.85457pt too wide) in paragraph at lines 7833--7833 [] [] [190] [191] <./images/fig4.pdf, id=1826, 386.44376pt x 370.38374pt> File: ./images/fig4.pdf Graphic file (type pdf) Package pdftex.def Info: ./images/fig4.pdf used on input line 7908. (pdftex.def) Requested size: 307.14749pt x 294.38585pt. [192] [193 <./images/fig4.pdf>] [194] [195] [196] [197] [198] [199] [200] [201] [202] [203] [204] [205] [206] [207] [208] [209] [210] [211] [212 ] [213] [214] [215] [216] [217] [218] [219] [220] [221] [222] [223] [224] [225] [226] [227] [228] [229] [230] [231] [232] [233] [234] [235] [236] [237] [238] [ 239] [240] [241] [242] [243] [244] [245] [246] [247] [248] [249] [250] [251] [2 52] [253 ] [254] [255] [256] <./images/fig5.pdf, id=2464, 201.75375pt x 254.9525pt> File: ./images/fig5.pdf Graphic file (type pdf) Package pdftex.def Info: ./images/fig5.pdf used on input line 10304. (pdftex.def) Requested size: 201.75325pt x 254.95186pt. [257] [258 <./images/fig5.pdf>] [259] [260] [261] [262] [263] [264] [265] [266] [267] [268] [269] [270] [271] [272 ] [273] [274] [275] [276] [277] [278] [279] [280] [281] [282] [283] [284] [285] [286] [287] [288] [289] [290] [291] Overfull \hbox (1.07619pt too wide) in paragraph at lines 11644--11650 []\OMS/cmsy/b/n/12 x\OT1/cmr/bx/n/12 264. []Two In-de-pend-ent Con-stitu-ents i n Two Phases. [] [292] [293] [294] [295] Overfull \hbox (5.88911pt too wide) detected at line 11837 []\OML/cmm/m/it/12 ; []; []; [] \OMS/cmsy/m/n/12 j []\OML/cmm/m/it/12 : [] [296] [297] [298] [299] [300] [301] [302] [303] [304] [305] [306] [307] [308] [ 309] [310] [311] [312] [313] [314] [315] [316] [317] [318] [319] [320] Package babel Info: Redefining german shorthand "f (babel) in language on input line 12881. Package babel Info: Redefining german shorthand "| (babel) in language on input line 12881. Package babel Info: Redefining german shorthand "~ (babel) in language on input line 12881. Package babel Info: Redefining german shorthand "f (babel) in language on input line 12881. Package babel Info: Redefining german shorthand "| (babel) in language on input line 12881. Package babel Info: Redefining german shorthand "~ (babel) in language on input line 12881. [321 ] Package babel Info: Redefining german shorthand "f (babel) in language on input line 12933. Package babel Info: Redefining german shorthand "| (babel) in language on input line 12933. Package babel Info: Redefining german shorthand "~ (babel) in language on input line 12933. Package babel Info: Redefining german shorthand "f (babel) in language on input line 12933. Package babel Info: Redefining german shorthand "| (babel) in language on input line 12933. Package babel Info: Redefining german shorthand "~ (babel) in language on input line 12933. [322] (./50880-t.ind [323] [324 ] [325] [326] [327] [328] [329] [330] [331] [332] [333]) [1 ] [2] [3] [4] [5] [6] Overfull \hbox (6.28976pt too wide) in paragraph at lines 14091--14091 []\OT1/cmtt/m/n/8 Section 3. Information about the Project Gutenberg Literary A rchive Foundation[] [] [7] [8] Package atveryend Info: Empty hook `BeforeClearDocument' on input line 14184. [9] Package atveryend Info: Empty hook `AfterLastShipout' on input line 14184. (./50880-t.aux) Package atveryend Info: Executing hook `AtVeryEndDocument' on input line 14184. *File List* book.cls 2007/10/19 v1.4h Standard LaTeX document class bk12.clo 2007/10/19 v1.4h Standard LaTeX file (size option) inputenc.sty 2008/03/30 v1.1d Input encoding file latin1.def 2008/03/30 v1.1d Input encoding file babel.sty 2013/12/03 3.9h The Babel package germanb.ldf 2013/12/13 v2.7 German support for babel (traditional orthograp hy) english.ldf 2012/08/20 v3.3p English support from the babel system ifthen.sty 2001/05/26 v1.1c Standard LaTeX ifthen package (DPC) amsmath.sty 2013/01/14 v2.14 AMS math features amstext.sty 2000/06/29 v2.01 amsgen.sty 1999/11/30 v2.0 amsbsy.sty 1999/11/29 v1.2d amsopn.sty 1999/12/14 v2.01 operator names amssymb.sty 2013/01/14 v3.01 AMS font symbols amsfonts.sty 2013/01/14 v3.01 Basic AMSFonts support array.sty 2008/09/09 v2.4c Tabular extension package (FMi) alltt.sty 1997/06/16 v2.0g defines alltt environment footmisc.sty 2011/06/06 v5.5b a miscellany of footnote facilities mhchem.sty 2014/02/01 v3.15 for typesetting chemical formulae calc.sty 2007/08/22 v4.3 Infix arithmetic (KKT,FJ) twoopt.sty 2008/08/11 v1.5 Definitions with two optional arguments (HO) keyval.sty 1999/03/16 v1.13 key=value parser (DPC) graphics.sty 2009/02/05 v1.0o Standard LaTeX Graphics (DPC,SPQR) trig.sty 1999/03/16 v1.09 sin cos tan (DPC) graphics.cfg 2010/04/23 v1.9 graphics configuration of TeX Live pdftex.def 2011/05/27 v0.06d Graphics/color for pdfTeX infwarerr.sty 2010/04/08 v1.3 Providing info/warning/error messages (HO) ltxcmds.sty 2011/11/09 v1.22 LaTeX kernel commands for general use (HO) pdftexcmds.sty 2011/11/29 v0.20 Utility functions of pdfTeX for LuaTeX (HO) ifluatex.sty 2010/03/01 v1.3 Provides the ifluatex switch (HO) ifpdf.sty 2011/01/30 v2.3 Provides the ifpdf switch (HO) expl3.sty 2014/01/07 v4646 L3 Experimental code bundle wrapper l3names.sty 2014/01/04 v4640 L3 Namespace for primitives l3bootstrap.sty 2014/01/04 v4640 L3 Experimental bootstrap code etex.sty 1998/03/26 v2.0 eTeX basic definition package (PEB) l3basics.sty 2014/01/04 v4642 L3 Basic definitions l3expan.sty 2014/01/04 v4642 L3 Argument expansion l3tl.sty 2013/12/27 v4625 L3 Token lists l3seq.sty 2013/12/14 v4623 L3 Sequences and stacks l3int.sty 2013/08/02 v4583 L3 Integers l3quark.sty 2013/12/14 v4623 L3 Quarks l3prg.sty 2014/01/04 v4642 L3 Control structures l3clist.sty 2013/07/28 v4581 L3 Comma separated lists l3token.sty 2013/08/25 v4587 L3 Experimental token manipulation l3prop.sty 2013/12/14 v4623 L3 Property lists l3msg.sty 2013/07/28 v4581 L3 Messages l3file.sty 2013/10/13 v4596 L3 File and I/O operations l3skip.sty 2013/07/28 v4581 L3 Dimensions and skips l3keys.sty 2013/12/08 v4614 L3 Experimental key-value interfaces l3fp.sty 2014/01/04 v4642 L3 Floating points l3box.sty 2013/07/28 v4581 L3 Experimental boxes l3coffins.sty 2013/12/14 v4624 L3 Coffin code layer l3color.sty 2012/08/29 v4156 L3 Experimental color support l3luatex.sty 2013/07/28 v4581 L3 Experimental LuaTeX-specific functions l3candidates.sty 2014/01/06 v4643 L3 Experimental additions to l3kernel multicol.sty 2011/06/27 v1.7a multicolumn formatting (FMi) makeidx.sty 2000/03/29 v1.0m Standard LaTeX package graphicx.sty 1999/02/16 v1.0f Enhanced LaTeX Graphics (DPC,SPQR) fancyhdr.sty geometry.sty 2010/09/12 v5.6 Page Geometry ifvtex.sty 2010/03/01 v1.5 Detect VTeX and its facilities (HO) ifxetex.sty 2010/09/12 v0.6 Provides ifxetex conditional hyperref.sty 2012/11/06 v6.83m Hypertext links for LaTeX hobsub-hyperref.sty 2012/05/28 v1.13 Bundle oberdiek, subset hyperref (HO) hobsub-generic.sty 2012/05/28 v1.13 Bundle oberdiek, subset generic (HO) hobsub.sty 2012/05/28 v1.13 Construct package bundles (HO) intcalc.sty 2007/09/27 v1.1 Expandable calculations with integers (HO) etexcmds.sty 2011/02/16 v1.5 Avoid name clashes with e-TeX commands (HO) kvsetkeys.sty 2012/04/25 v1.16 Key value parser (HO) kvdefinekeys.sty 2011/04/07 v1.3 Define keys (HO) pdfescape.sty 2011/11/25 v1.13 Implements pdfTeX's escape features (HO) bigintcalc.sty 2012/04/08 v1.3 Expandable calculations on big integers (HO) bitset.sty 2011/01/30 v1.1 Handle bit-vector datatype (HO) uniquecounter.sty 2011/01/30 v1.2 Provide unlimited unique counter (HO) letltxmacro.sty 2010/09/02 v1.4 Let assignment for LaTeX macros (HO) hopatch.sty 2012/05/28 v1.2 Wrapper for package hooks (HO) xcolor-patch.sty 2011/01/30 xcolor patch atveryend.sty 2011/06/30 v1.8 Hooks at the very end of document (HO) atbegshi.sty 2011/10/05 v1.16 At begin shipout hook (HO) refcount.sty 2011/10/16 v3.4 Data extraction from label references (HO) hycolor.sty 2011/01/30 v1.7 Color options for hyperref/bookmark (HO) auxhook.sty 2011/03/04 v1.3 Hooks for auxiliary files (HO) kvoptions.sty 2011/06/30 v3.11 Key value format for package options (HO) pd1enc.def 2012/11/06 v6.83m Hyperref: PDFDocEncoding definition (HO) hyperref.cfg 2002/06/06 v1.2 hyperref configuration of TeXLive url.sty 2013/09/16 ver 3.4 Verb mode for urls, etc. hpdftex.def 2012/11/06 v6.83m Hyperref driver for pdfTeX rerunfilecheck.sty 2011/04/15 v1.7 Rerun checks for auxiliary files (HO) supp-pdf.mkii color.sty 2005/11/14 v1.0j Standard LaTeX Color (DPC) color.cfg 2007/01/18 v1.5 color configuration of teTeX/TeXLive nameref.sty 2012/10/27 v2.43 Cross-referencing by name of section gettitlestring.sty 2010/12/03 v1.4 Cleanup title references (HO) 50880-t.out 50880-t.out umsa.fd 2013/01/14 v3.01 AMS symbols A umsb.fd 2013/01/14 v3.01 AMS symbols B ./images/device.pdf omscmr.fd 1999/05/25 v2.5h Standard LaTeX font definitions ./images/fig1.pdf ./images/fig2.pdf ./images/fig3.pdf ./images/fig4.pdf ./images/fig5.pdf 50880-t.ind *********** Package atveryend Info: Executing hook `AtEndAfterFileList' on input line 14184 . Package rerunfilecheck Info: File `50880-t.out' has not changed. (rerunfilecheck) Checksum: B36BC6583614C4B13F36B6DFCFD8827A;2043. Package atveryend Info: Empty hook `AtVeryVeryEnd' on input line 14184. ) Here is how much of TeX's memory you used: 14542 strings out of 493304 235562 string characters out of 6139871 345829 words of memory out of 5000000 16775 multiletter control sequences out of 15000+600000 17567 words of font info for 67 fonts, out of 8000000 for 9000 957 hyphenation exceptions out of 8191 38i,20n,46p,10391b,1066s stack positions out of 5000i,500n,10000p,200000b,80000s < /usr/share/texlive/texmf-dist/fonts/type1/public/amsfonts/cm/cmr6.pfb> Output written on 50880-t.pdf (352 pages, 1106477 bytes). PDF statistics: 4010 PDF objects out of 4296 (max. 8388607) 3564 compressed objects within 36 object streams 1011 named destinations out of 1200 (max. 500000) 263 words of extra memory for PDF output out of 10000 (max. 10000000)