hilbert and physics (1900-1915)corry/publications/articles/pdf/symbolic.pdf · hilbert and physics...

51
Hilbert and Physics (1900-1915) by Leo Corry 1 Some mathematical-physical theories look to me like a toy, that a child has completely messed up and that every three minutes needs to be fixed again, in order to keep it working. David Hilbert, Tagebuch 2 1. Introduction 1 2. Physics and Geometry in Hilbert’s Early Career 5 3. Die Grundlagen der Geometrie 10 4. Physics in Hilbert’s 1900 List of Problems 14 5. Hilbert, Minkowski, and Physics in Göttingen: 1900-1909 16 6. Mechanics, Kinetic Theory, Radiation Theory: 1910-1914 32 7. Electrodynamics and General Relativity: 1913-1915 36 8. Concluding Remarks 42 Bibliography 1. Introduction The name of David Hilbert (1862-1943) arises in connection with the history of physics usually in three different, and rather circumscribed contexts. The first of these contexts is the famous list of twenty three problems that Hilbert presented in 1900 in Paris, with occasion of the Second International Congress of Mathematicians. The sixth problem of this list deals specifically with physics. Only one year previous to the Paris Congress Hilbert had published his seminal study on the foundations of geometry, Die Grundlagen der Geometrie. The axiomatic approach adopted by Hilbert in this book was to have an enormous influence on the development of twentieth- 1. Sections 1 to 5 of the present preprint include condensed versions of the contents of Corry 1997, 1997a. 2. “Manche math-physikalische Theorie erscheint mir wie ein Kinderspielzeug, dass in Unordnung geraten ist und alle 3 Minuten wieder aufgerichtet werden muss, damit es weiter geht.” The date of this quotation is unknown, but may be somewhat before 1900. The manuscript of Hilbert’s Tagebuch is found in the Hilbert Nachlass, Staats- und Universitätsbibliothek Göttingen, Cod MS Hilbert 600/3, see p. 134.

Upload: vanngoc

Post on 28-Aug-2018

225 views

Category:

Documents


2 download

TRANSCRIPT

Hilbert and Physics (1900-1915)by Leo Corry1

Some mathematical-physical theories look to me like a toy,that a child has completely messed up and that every three minutes

needs to be fixed again, in order to keep it working.

David Hilbert, Tagebuch2

1. Introduction 1

2. Physics and Geometry in Hilbert’s Early Career 5

3. Die Grundlagen der Geometrie 10

4. Physics in Hilbert’s 1900 List of Problems 14

5. Hilbert, Minkowski, and Physics in Göttingen: 1900-1909 16

6. Mechanics, Kinetic Theory, Radiation Theory: 1910-1914 32

7. Electrodynamics and General Relativity: 1913-1915 36

8. Concluding Remarks 42 Bibliography

1. Introduction

The name of David Hilbert (1862-1943) arises in connection with the history of physics usually in three different, and rather circumscribed contexts. The first of these contexts is the famous list of twenty three problems that Hilbert presented in 1900 in Paris, with occasion of the Second International Congress of Mathematicians. The sixth problem of this list deals specifically with physics. Only one year previous to the Paris Congress Hilbert had published his seminal study on the foundations of geometry, Die Grundlagen der Geometrie. The axiomatic approach adopted by Hilbert in this book was to have an enormous influence on the development of twentieth-

1. Sections 1 to 5 of the present preprint include condensed versions of the contents of Corry 1997, 1997a.

2. “Manche math-physikalische Theorie erscheint mir wie ein Kinderspielzeug, dass in Unordnung geraten ist und alle 3 Minuten wieder aufgerichtet werden muss, damit es weiter geht.” The date of this quotation is unknown, but may be somewhat before 1900. The manuscript of Hilbert’s Tagebuch is found in the Hilbert Nachlass, Staats- und Universitätsbibliothek Göttingen, Cod MS Hilbert 600/3, see p. 134.

2Hilbert and Physics:

century mathematics and on the way mathematicians looked at their science. Hilbert’s sixth problem was the suggestion that an analysis similar to that performed for the case of geometry in the Grundlagen should also be applied to individual physical disciplines.

The problems in this list became an object of great mathematical and historical interest over the following decades, with mathematicians of all specialties and of all countries attempting to solve them, and with periodical accounts of the current state of research on some, or on all of them. However, this particular problem, the sixth one, received much less attention than any of the other twenty two in the list. Working

mathematicians directed in general relatively little effort to solve it.3 Neither have historians greatly contributed to understand both its roots and whatever attempts have

been made to address it after 1900.4 This problem seems to have been largely considered as an accessory surplus, that only artificially can be seen as part of Hilbert’s otherwise comprehensive and harmonious image of mathematics.

The second context is the kinetic theory of gases. In 1912, at a time when most of Hilbert’s energies were directed to his work on the theory of linear integral equations, he solved the so-called Boltzmann equation. Although this represented a major contribution to the development of this particular physical discipline, Hilbert is often considered to have had no real interest in the kinetic theory as such. Rather, his solution of the equation has been considered as an isolated, if important, furtive incursion into this field. In his authoritative account of the development of the kinetic theory, Stephen G. Brush dedicated one short section to describing Hilbert’s contribution. Brush’s assessment of Hilbert’s motivations is expressed in the following passage:

When Hilbert decided to include a chapter on kinetic theory in his treatise on integral equations, it does not appear that he had any particular interest in the physical problems associated with gases. He did not try to make any detailed calculations of gas properties, and did not discuss the basic issues such as the nature of irreversibility and the validity of mechanical interpretations which had exercised the mathematician Ernst Zermelo in his debate with Boltzmann in 1896-97. A few years later, when Hilbert presented his views on the contemporary problems of physics, he did not even mention kinetic theory. We must therefore conclude that he was simply looking for another possible application of his mathematical theories, and when he had succeeded in finding and characterizing a special class of solutions (later called “normal”) ... his interest in the Boltzmann equation and in kinetic theory was exhausted. (Brush 1976, 448)

Brush added that Hilbert did encourage some of his students to work on mathematical problems connected with the theory, and that he seems to have also taught courses on this issue. Yet these qualifications did not change Brush’s overall evaluation of Hilbert’s motivations. It seems to me that Brush’s assessment manifests the widely held beliefs about the nature of Hilbert’s contribution to physics.

The third context in which Hilbert’s name has been associated to a significant development in physics is general relativity. On November 20, 1915, Hilbert presented to the Royal Scientific Society in Göttingen his version of the field equations of gravitation, in the framework of what he saw as an axiomatically formulated

3. See, e.g., Wightman 1976, Gnedenko 1979.

4. For a general historical account of the problems in the list, their roots, and their place in the development of twentieth centruy mathematics see Rowe 1996.

3Hilbert and Physics:

foundation for the whole of physics. During that same month of November, Einstein had been struggling with the final stages of his own effort to formulate the generally covariant equations that lie at the heart of the general theory of relativity. His struggle had spanned at least three years of intense work and included the publication of several

previous versions, each of which Einstein found inadequate for different reasons.5 Only in November of 1915 he presented three different versions at the weekly meetings of the Prussian Academy of Sciences in Berlin, before attaining his final version, on November 25, that is, five days after Hilbert had presented his theory.

Einstein had visited Göttingen in the summer of 1915, to lecture on the progress and the difficulties encountered in his work. Hilbert was then in the audience and Einstein was greatly impressed by him. He felt that in Göttingen his work had been

fully understood to the details.6 Hilbert’s involvement in the problems associated to general relativity has in general been traced back not earlier than to this visit of Einstein or, at best, to the years immediately preceding it. Although the axiomatic character of Hilbert’s work on relativity has often been stressed—and by implication it has been associated to his work on the foundations of geometry and of mathematics at large—its actual place as part of his overall conception of mathematics and science has not been fully clarified. Like in the case of the kinetic theory, this contribution of Hilbert has mainly been seen as a furtive incursion in physics, aimed at illustrating the power and the scope of validity of the “axiomatic method” (whatever is meant by that)

and as a test of Hilbert’s mathematical abilities.7

A source quoted very often when describing the various aspects of Hilbert’s career—and which also refers to his work on physics—is a passage taken from Hermann Weyl’s obituary of Hilbert. According to this passage, Hilbert’s mathematical career covered five, clearly discernible periods during which his attention focused on a single issue: (1) Theory of invariants (1885-1893); (2) Theory of algebraic number fields (1893-1898); (3) Foundations, (a) of geometry (1898-1902), (b) of mathematics in general (1922-1930); (4) Integral equations (1902-1912); (5) Physics (1910-1922) (Weyl 1944, 619). According to Weyl, the passage from one period to the next always signified a sharp departure from Hilbert’s past topic of interest in order to move into a completely new one. For instance, concerning his research on the foundations of geometry Weyl wrote (1944, 635): “[T]here could not have been a more complete break than the one dividing Hilbert’s last paper on the theory of number fields from his classical book Grundlagen der Geometrie.” Thus, even if Weyl conceded that Hilbert’s scientific interests focused during a relatively long period of time on physics alone (twelve years, two of which overlap with his work on linear integral equations), his account still suggests a clear separation between this period and the rest of his long career.

5. See Norton 1984.6. As he wrote to Sommerfeld upon returning from Göttingen. See Hermann (ed.) 1968, 30.7. See, e.g., Earman and Glymour 1978, 293; Mehra 1974, 17 & 83. For a recent reinterpretation of the actual contents of Hilbert’s lecture in November 1915, see Corry, Renn & Stachel 1997.

4Hilbert and Physics:

The periodization manifest in Weyl’s account reflects indeed rather faithfully the distribution of Hilbert’s published work along the years, and what constituted his main domains of interest at different times. However, the actual scope of Hilbert current interests is much broader than such an account may suggest. A clear and balanced perception of Hilbert’s mathematical world, and in particular of the actual place he accorded to physics as part of it, necessitates a deeper examination of his docent and institutional activities in Göttingen: his lectures, his seminars, the doctoral

dissertations he advised, the activities of the Göttingen Mathematical Society (GMG).8 As even the most cursory examination of the lists included in the appendix at the end of this article immediately indicates, these activities included a long-standing interest in physical issues that covered all his years at Göttingen. Physics was always accorded a central place in the scientific agenda implemented at Göttingen by Hilbert, by his colleagues and by his students.

Given the astonishing breadth of Hilbert’s scope of scientific interests and knowledge, an examination of his scientific activities other than pure research and standard publishing becomes fundamental in any attempt to understand his mathematical views. Hilbert had the ability to attract extremely gifted students and to communicate them the kind of deep open questions that in his opinion should be investigated in various fields. He also provided very often the inspiration to solve those problems. Hilbert directed no less than sixty eight doctoral dissertations, sixty of them

in the relatively short period between 1898 and 1914.9 Four or five of these dissertations deal with issues that can be directly related to physics as well as to Hilbert’s current scientific interests. Also his lectures and seminars in Göttingen were never mere systematic presentations of well-established knowledge. On the contrary, Hilbert considered the classroom and the seminar meetings as ideal settings for putting forward new, untried ideas and to benefit from the feedback of his students. Later in life Hilbert described the central place he conceded to his teaching in the following terms:

The closest conceivable connection between research and teaching became a decisive feature of my mathematical activity. The interchange of scientific ideas, the communication of what one found by himself and the elaboration of what one had heard, was since my early years at Königsberg a pivotal aspect of my scientific work ... In my lectures, and above all in the seminars, my guiding principle was not to present material in a standard and as smooth as possible way, so as to help the student keeping clean and ordered notebooks. Above all, I always tried to illuminate the problems and difficulties and to offer a bridge leading to the present open questions. It often happened, that in the course of a semester the program of an advanced lecture was completely changed, because I wanted to discuss issues in which I was currently involved as a researcher and which in no way had yet attained their definite formulation. (Hilbert 1971, 79)

Thus, the notes that Hilbert prepared for his courses provide an essential source for understanding the development of his ideas.

8. In fact, Weyl himself acknowledged in some occassions the centrality of Hilbert’s docent activities and the impact of his influence as a teacher as main traits of his sceintific legacy. See Sigurdsson 1994, 356-358.9. Hilbert’s doctoral students are listed in Hilbert GA Vol. 3, 431-433. This list, however, is incomplete.

5Hilbert and Physics:

A detailed and comprehensive account of the contents of Hilbert’s unpublished

lectures on physics, their historical context and their influence is yet to be done.10 Such an account will probably bring to light many unfinished ideas that Hilbert presented in the classroom to his students, and which the latter further developed in their own works thus leading to central contributions in the evolution of particular physical disciplines. It will perhaps likewise bring to light many ideas that led nowhere. The present paper surveys in general lines—using both published an unpublished sources—the evolution of Hilbert’s interest in various physical domains until 1915, and attempts to explain the place that Hilbert attributed to physical issues within his overall conception of mathematics. I will claim that understanding these aspects of his work is essential for understanding his view on mathematics. I will also suggest that, given the centrality of Hilbert in the scientific world of Göttingen, and given the centrality of Göttingen in the scientific world at large, understanding Hilbert’s influence on other scientists working at that center may appear as fundamental for better understanding the historical context of many central developments in early twentieth-century physics.

Hilbert was undoubtedly among the most influential mathematicians of the beginning of this century, if not the most influential one. His name is associated with many results and concepts that play fundamental roles in several, distant fields of mathematics such as algebra, number theory, functional analysis, mathematical physics, metamathematics, and foundations of geometry. Even more pervasive is the view that associates his name with the formalist approach that came to dominate a considerable part of mathematical practice throughout the twentieth century. It was only relatively recently, however, that historians of mathematics undertook a careful analysis of his work and of his mathematical world, and a more balanced image of the latter has begun to arise from these studies. The present article makes use of that recent scholarship and is also intended as a further contribution to it.

2. Physics and Geometry in Hilbert’s Early Career

Crucial for understanding the place of physics in Hilbert’s overall conception of science is his view of geometry as a natural science, similar in most essential aspects to other physical disciplines. This conception—opposed perhaps to the largely accepted view of Hilbert as the champion of a formalistic interpretation of the essence of mathematics—is consistently manifest throughout his career, in spite of many other changes recognizable in the latter. An interest in physics, in geometry and in the relationship among them was present in Hilbert’s mind from very early on and until well the end of his career.

Hilbert’s studied in his native city of Königsberg and there he also spent his first years as a young professor. The University of Königsberg had a long tradition of experimental physics and of mathematical analysis that went back to the days of Franz

Ernst Neumann (1798-1895) and Carl Gustav Jacobi (1804-1851).11 We have no direct

10. But for an initial contribution in that direction, see Corry 1997.

6Hilbert and Physics:

evidence, however, to decide whether Hilbert took courses of physics as a student at all. We do know that Hilbert attended courses given by the versatile mathematician

Heinrich Weber (1842-1913),12 whose fields of interest covered mathematical physics as well, but it is unlikely that Hilbert came under his influence in this respect. Hilbert’s doctoral advisor was Ferdinad Lindemann (1852-1939), who introduced him mainly into the study of algebraic invariants; Hilbert’s interest in physics apparently developed somewhat later. The main sources of overall influence on Hilbert during his Königsberg years came from his friends Adolf Hurwitz (1859-1919) and Hermann Minkowski (1864-1909); the scientific scope of interest of these three young colleagues focused above all on pure mathematical domains, but gradually it also extended as to concede physics an important place in it.

In 1893 Hilbert wrote his last paper on the theory of invariants and his major research effort shifted now towards a new domain, the theory of algebraic number fields which, however, was not completely apart from the former one. Yet, as Michael Toepell’s study of the origins of Hilbert’s Grundlagen der Geometrie has shown, Hilbert was simultaneously pursuing at this time in a systematic way an additional field of interest, namely, the study of the foundations of geometry. This interest was

reflected in the courses he taught at that time.13 Hilbert lectured on geometry in Königsberg for the first time in 1891 and planned to do so again in 1893. For lack of students registered, however, this latter course was postponed until 1894. A remarkable change occurred in Hilbert’s approach to geometry between 1891 and 1893, whereby he gradually moved towards the axiomatic treatment as an appropriate way to address the foundational problems of this discipline. But even in his first lectures of 1891 Hilbert clearly expressed his conception of geometry as a natural science. In its essence, this view was not unlike that of other German geometers—even those who, like Moritz Pasch (1843-1930), had adopted an axiomatic approach in their

presentations from very early on (Pasch 1882).14 The similarity between geometry and physics, and the different character of the former as compared to other mathematical domains recurrently appears in Hilbert’s lectures. In the introduction to his 1891 course, for instance, it is expressed as follows:

Geometry is the science dealing with the properties of space. It differs essentially from pure mathematical domains such as the theory of numbers, algebra, or the theory of functions. The results of the latter are obtained through pure thinking ... The situation is completely different in the case of geometry. I can never penetrate the properties of space by pure reflection, much the same as I can never recognize the basic laws of mechanics, the law of gravitation or any other physical law in this way. Space is not a product of my reflections. Rather, it is given to me through the senses.15

11. On the Königsberg school see Klein 1926-7 Vol. 1, 112-115 & 216-221; Lorey 1916, 59-64; Volk 1967. The workings of the Königsberg physics seminar—initiated in 1834 by Franz Neumann—and its enormous influence on nineteenth-century physics education in Germany are described in great detail in Olesko 1991.12. For more details on Weber (especially concerning his contributions to algebra) see Corry 1996, §§ 1.2 & 2.2.4. 13. See Toepell 1986. Toepell also describes Hilbert’s only gradual adoption of an axiomatic approach to deal with these questions.14. See Contro 1976, 284-289.15. The German original is quoted in Toepell 1986, 21. Similar testimonies can be found in many other manuscripts of Hilbert’s lectures. Cf., e.g., Toepell 1986, 58.

7Hilbert and Physics:

When rewriting his lecture notes of 1893 for teaching in 1894, Hilbert added some remarks from which we learn that in the meantime he became acquainted with the ideas put forward in Heinrich Hertz’s textbook on The Principles of Mechanics, published that same year. Hilbert referred once again to the natural character of geometry and explained the possible role of axioms in elucidating its foundations, by making reference to Hertz’s characterization of a “correct” scientific image (Bild) or theory. Thus Hilbert wrote:

Nevertheless the origin [of geometrical knowledge] is in experience. The axioms are, as Hertz would say, images or symbols in our mind, such that consequents of the images are again images of the consequences, i.e., what we can logically deduce from the images is itself valid in nature.16

In the same context, Hilbert also pointed out the need of establishing the independence of the axioms of geometry, while alluding to the kind of demand stipulated by Hertz. Stressing once more the objective and factual character of geometry, Hilbert wrote:

The problem can be formulated as follows: What are the necessary, sufficient, and mutually independent conditions that must be postulated for a system of things, in order that any of their properties correspond to a geometrical fact and, conversely, in order that a complete description and arrangement of all the geometrical facts be possible by means of this system of things.17

It is difficult to determine whether Hilbert had indeed mastered the contents of Hertz’s book at this stage, and even whether he actually got to read thoroughly its introduction. However, it is likely that whatever acquaintance he made with the basic ideas presented in this book so soon after its publication, this was suggested to Hilbert by his friend Minkowski. Minkowski spent three semesters in Bonn still as a student, and he returned to that city after receiving his doctorate in Königsberg in 1885. He stayed in Bonn until 1894. While at Bonn, Minkowski became involved in both mathematical

and experimental physics.18 In 1888 he published an article on hydrodynamics that was submitted to the Berlin Academy by Hermann von Helmholtz (Minkowski 1888). During this period of his life, Minkowski’s greatest scientific source of inspiration

came from Hertz19 and he certainly must have communicated this enthusiasm to Hilbert as well.

Hilbert arrived in Göttingen in 1895, at a time when he was completing his major work on algebraic number theory, the Zahlbericht (Hilbert 1897). Algebraic number fields continued to be his major field of publication during his first years at this university, but at the same time he organized seminars and lectured on other topics. Between 1895 and 1898 he held joint seminars with Klein on number theory,

mechanics and function theory.20 In 1899 he lectured in Göttingen on the foundations of geometry for the first time. The notes of this course provided the basis for the Grundlagen der Geometrie, published in June 1899 as part of a Festschrift issued in

16. Hilbert 1893/94, 10: “Dennoch der Ursprung aus der Erfahrung. Die Axiome sind, wie Herz [sic] sagen würde, Bilde[r] oder Symbole in unserem Geiste, so dass Folgen der Bilder wieder Bilder der Folgen sind d.h. was wir aus den Bildern logisch ableiten, stimmt wieder in der Natur.”17. Translated from the original passage quoted in Toepell 1986, 58-59.18. See Rüdenberg and Zassenhaus (eds.) 1973, 39-42, and Hilbert 1909, 355.19. At least according to Hilbert’s report (Hilbert 1909, 355)..20. See Lorey 1916, 128.

8Hilbert and Physics:

Göttingen with occasion of the unveiling of the Gauss-Weber monument. But before that, back in the winter semester of 1898/99 Hilbert taught his first course on a physical issue: mechanics. In the introduction to this course, Hilbert stressed once again the essential affinity between geometry and the natural sciences, and also explained the role that axiomatization should play in the mathematization of the latter. He compared the two domains in the following terms:

Also geometry [like mechanics] emerges from the observation of nature, from experience. To this extent, it is an experimental science.... But its experimental foundations are so irrefutably and so generally acknowledged, they have been confirmed to such a degree, that no further proof of them is deemed necessary. Moreover, all what is needed is to derive these foundations from a minimal set of independent axioms and thus to construct the whole building of geometry by purely logical means. In this way [i.e., by means of the axiomatic treatment] geometry is turned into a pure mathematical science. In mechanics it is also the case that all physicists recognize its most basic facts. But the arrangement of the basic concepts is still subject to a change in perception ... and therefore mechanics cannot yet be described today as a pure mathematical discipline, at least to the same extent that geometry is. We must strive that it becomes one. We must ever stretch the limits of pure mathematics further afar, on behalf not only of our mathematical interest, but rather of the interest of science in general.21

For Hilbert, then, the difference between geometry and other physical sciences—mechanics in this case—was not one of essence, but rather one of historical stage of development. He saw no reason of principle why an axiomatic analysis of the kind he was then developing for geometry could not eventually be applied to mechanics with similar, useful consequences. Eventually, that is to say, when mechanics would attain a degree of development similar to the present one of geometry, in terms of the quantity and certainty of known results, and in terms of an appreciation of what are really the “basic facts” on which the theory is based.

Hilbert’s first course on mechanics was an elementary one, based on standard presentations, in which he attempted to give a basic overview of the discipline. The long bibliographical list that Hilbert recommended to his students for further reading (Hilbert 1898-9, 4-5) may be helpful for reconstructing and assessing the degree of his acquaintance with current knowledge on the field. However, one must exercise some care before relying too heavily on these lists as a faithful indicator, given Hilbert’s general tendency not to study thoroughly and comprehensively all the existing literature on a topic he was pursuing. On the contrary, Hilbert—as David Rowe has argued— “measured the quality of a mathematician’s work by the number of earlier

21. Hilbert 1898/9, 1-3 (Emphasis in the original): “Auch die Geometrie ist aus der Betrachtung der Natur, aus der Erfahrung hervorgegangen und insofern eine Experimentalwissenschaft. ... Aber diese experimentellen Grundlagen sind so unumstösslich und so allgemein anerkannt, haben sich so überall bewährt, dass es einer weitere experimentellen Prüfung nicht mehr bedarf und vielmehr alles darauf ankommt diese Grundlagen auf ein geringstes Mass unabhängiger Axiome zurückzuführen und hierauf rein logisch den ganzen Bau der Geometrie aufzuführen. Also Geometrie ist dadurch eine rein mathematische Wiss. geworden. Auch in der Mechanik werden die Grundthatsachen von allen Physikern zwar anerkannt. Aber die Anordnung der Grundbegriffe ist dennoch dem Wechsel der Auffassungen unterworfen ... so dass die Mechanik auch heute noch nicht, jedenfalls nicht in dem Maasse wie die Geometrie als eine rein mathematische Disciplin zu bezeichnen ist. Wir müssen streben, dass sie es wird. Wir müssen die Grenzen echter Math. immer weiter ziehen nicht nur in unserem math. Interesse sondern im Interesse der Wissenschaft überhaupt.”

9Hilbert and Physics:

investigations it rendered obsolete”.22 Still, given the fact that Hilbert had no made no specific contributions of his own to mechanics, a detailed account of this bibliographical list, which include several rather uncommon titles, seems to be instructive about his views and his knowledge.

First, Hilbert included in the bibliography four “classical works:” Lagrange’s Méchanique analytique (1788); Jacobi’s, Dynamik (1843); Kirchhoff’s, Mechanik (1877) and Thomson and Tait’s, Theoretische Physik (in German translations of 1871 and 1886). The textbooks he recommended were the following: Mechanik (1880), by

someone called Schell,23 a book he described as “somewhat antiquated but rich in contents”; Kinematik, Statik, Dynamik (1884), by some Petersen, “short and easily comprehensible”; Cours de Mechanique 2 Vols. (1884-86) by Despeyrous and

Darboux, “like Schell’s”; and Analytische Mechanik (1888), by Otto Rausenberg.24 In a different section of the bibliographical list, Hilbert mentioned various “courses”: Elementare Mechanik (1889), by the Göttingen physicist Woldemar Voigt, “illuminating from the physical point of view”; Mach’s Principien der Mechanik (1889); Mechanik (1890), by some Büdele, “like Schell and Desperyous-Darboux”; Hertz’s Prinzipien der Mechanik (1894); “a memorial (Denkmal), in which this young and brilliant physicists presented classical mechanics with Euclidean rigor”; Helmtoltz’s Dynamik diskreter Massenpunkte (1894), reportedly a manuscript of the latter’s university lectures; Boltzmann’s Prinzipien der Mechanik (1897), “develops the atomistic point of view; opposed to Hertz”; Dynamik der Systeme starrer Körper (1897-98), by Routh; and Traité de méchanique rationelle 3 Vols. (1893-98), by Appell, “a comprehensive and systematic handbook”. Hilbert also included a section with “historical” texts: Düring’s Principien der Mechanik (1873), and again Mach’s book. Finally, his list included three collections of exercises: Aufgaben aus der analytischen Mechanik (1879), by Fuhrmann; Aufgabe aus der theor. Mechanik

(1891), by Zieh; and Problèmes de méchanique (1867), by Tullien.25

The list is certainly impressive and interesting in itself but, as already said, it is very difficult to know to what extent Hilbert was actually familiar with all the texts and relied on them for his course. In any case, following his lectures of 1899 on the foundations of geometry, Hilbert was to concentrate over the following two years exclusively in this latter topic. Only in the winter semester of 1901-02, he taught again a course on potential theory. But before retaking that thread, it is relevant to discuss very briefly the contents of the Grundlagen and the views put forward in it.

22. Rowe 1994, 192.23. In his well-informed history of the teaching of mathematics in the nineteenth century, Wilhelm Lorey (1916, p. 135), mentions the name of Wilhelm Schell (1826-1904), who taught in Marburg and Karlsruhe, and published a book on the theory of curves. Lorey doesn’t mention any textbook on mechanics published by him.24. Lorey 1916, 179, asserts that Rausenberg was born in 1852.25. Consulting several books on German mathematics an physics in the nineteenth century (Jungnickel & McCormmach 1986, Lorey 1916, Olesko 1991), I haven’t been able to gather any infromation about the other authors mentioned by Hilbert: Petersen, Büdele, Routh, Appell, Fuhrmann, Zieh or Tullien,

10Hilbert and Physics:

3. Die Grundlagen der Geometrie

As already said, Hilbert had been involved since 1891 with questions pertaining the foundations of geometry. In 1898 Friedrich Schur (1856-1932) proved the so-called theorem of Pappus without using any continuity assumptions. In doing this he reinforced a line of attack initiated in 1891 by Hermann Ludwig Wiener (1857-1939) in this context, which Hilbert had been closely attentive to. Wiener had proved in 1891 that this theorem, together with the so-called theorem of Desargues, suffice to prove the fundamental theorem of projective geometry. The role of continuity assumption in his proofs, however, remained to be fully elucidated. Wiener’s ideas had been a main factor in attracting Hilbert’s attention to the study of the foundations of geometry in the first place. Schur’s proof, on the other hand, seems to have provided the definitive

trigger that made him redirect all his efforts to this field of research.26 Thereafter Hilbert undertook a detailed elucidation of the logical interdependence of the various fundamental theorems of projective and Euclidean geometry and, more generally, of the structure of the various kinds of geometries that can be produced under various sets of assumptions. A main question that stood at the center of Hilbert’s investigation was one that had arisen from the attempts of Felix Klein (1849-1925), beginning in 1871, to coordinatize projective geometry using ideas originally formulated by Arthur

Cayley (1821-1895).27 This coordinatization was essential for providing a connecting link between synthetic and analytic geometry, and its realization depended on the same assumptions necessary for proving the already mentioned fundamental theorems. Following the work of Wiener and Schur, Hilbert focused on understanding the specific role of continuity in those proofs. These kinds of questions provided the main motivations behind the investigation undertook by Hilbert in the Grundlagen. The experience gained while thinking on these issues when preparing his lectures on geometry, on the one hand, and his acquaintance with the ideas put forward in Hertz’s

treatment of physics,28 on the other hand, indicated to Hilbert that the axiomatic method would provide a powerful and effective tool with which to address these crucial issues in the most effective way.

It would be well beyond the scope of the present article to discuss all the details of the Grundlagen, and how the main foundational questions of geometry were

addressed therein.29 Neither can we discuss here the criticisms aroused by its publication and the successive versions of the book, beginning in 1902, which corrected certain gaps of the original edition and added new results. However, in order to understand the basic ideas behind Hilbert’s axiomatic approach and its bearing on physical theories, we must nevertheless discuss some of the main features of the analysis put forward in the Grundlagen. Of particular interest are the kinds of questions that Hilbert systematically pursued here for the first time, thus establishing a standard for his future work.

26. For Hilbert’s acquaintance with the works of Wiener and Schur, and the events surrounding the publication of Schur’s proof and its effect on Hilbert, see Toepell 1986, 114-122.27. For an account of Cayley’s contributions see Klein 1926-7 Vol. 1, 147-151.28. In fact, Hertz’s was perhaps only one among several sources from within the physical literature that influenced Hilbert’s initial inclination to, and then his fully adoption of, the axiomatic method. See Corry 1997, 92-103.29. Such a discussion can be found in Toepell 1986. See especially, pp. 143-236.

11Hilbert and Physics:

The declared aim of the Grundlagen was to present a “simple” and “complete”

system of “mutually independent” axioms,30 from which all known theorems of geometry might be deduced. The axioms were defined for three systems of undefined objects (“points”, “lines” and “planes”), and they postulated relations that should be satisfied by these objects. The axioms are divided into five groups (axioms of incidence, of order, of congruence, of parallels and of continuity), but the groups have no pure logical significance in themselves. Rather they reflect Hilbert’s actual conception of the axioms as an expression of our spatial intuition: each of the groups expresses a particular way in which these intuitions manifest themselves.

The requirements imposed by Hilbert on his system of axioms are remarkably similar to the criteria put forward by Hertz as a basis for constructing and assessing

physical theories: permissibility, correctness, and appropriateness.31 Consider, in the first place, Hilbert’s requirement for independence of the axioms. This requirement is the most direct manifestation of the kind of foundational concerns that motivated Hilbert’s research. When analyzing independence, Hilbert’s interest focused mainly on the axioms of congruence, of continuity and of parallels, since this independence would specifically explain how the various basic theorems of projective and of Euclidean geometry are logically interrelated. In Hilbert’s early lectures on geometry, this requirement had already appeared, albeit more vaguely formulated, as a direct echo of Hertz’s demand for appropriateness. Now, in the Grundlagen, independence of axioms not only appeared as a more clearly formulated requirement, but Hilbert also provided the tools to prove systematically the mutual independence among the individual axioms within the groups and among the various groups of axioms in the system. He did so by introducing the method that has since become standard: he constructed models of geometries which fail to satisfy a given axiom of the system but satisfy all the others.

Also the requirement of simplicity had been explicitly put forward by Hertz and it complements that of independence. It means, roughly, that an axiom should contain ‘no more than a single idea.’ Beyond mentioning it in the introduction, this requirement was neither explicitly formulated nor otherwise realized in any clearly identifiable way in the Grundlagen. It appeared, however, in an implicit way and remained here—as well as in other, later works—an aesthetic desiderata for axiomatic

systems, which was not transformed into a mathematically controllable feature.32

30. See Hilbert 1899, 1 (Italics in the original): “... ein einfaches und vollständiges System von einander unabhängiger Axiome aufzustellen ...”31. See Jesper Lützen’s contribution to this volume.32. In his 1905 lectures on the axiomatization of physics, Hilbert explicitly demanded the simplicity of the axioms for physical theories. It should also be remarked that in a series of investigations conducted in the USA in the first decade of the present century under the influence of the Grundlagen, a workable criterion for simplicity of axioms was still sought after. For instance, Edward Huntington (1904, p. 290) included simplicity among his requirements for axiomatic systems, yet he warned that “the idea of a simple statement is a very elusive one which has not been satisfactorily defined, much less attained.”

12Hilbert and Physics:

The “completeness” that Hilbert demanded for his system of axioms runs

parallel to Hertz’s demand for correctness.33 Very much like Hertz’s stipulation for correct images, Hilbert required from any adequate axiomatization to allow for a derivation of all the known theorems of the discipline in question. The axioms formulated in the Grundlagen purportedly allowed to show how all the known results of Euclidean, as well as of certain non-Euclidean, geometries could be elaborated from

scratch, depending on which groups of axioms were admitted.34 Thus, Hilbert discussed in great detail the role of each of the groups of axioms in the proofs of two crucial results: the theorems of Desargues and the theorem of Pappus (also called by Hilbert the theorem of Pascal). Hilbert’s analysis allowed a clear understanding of the actual premises necessary for coordinatizing projective geometry, which, as already stressed, was a key step in building the bridge between the latter and other kinds of

geometry and a main concern of Hilbert.35

Although not explicitly mentioned in the introduction to the Grundlagen, the question of the consistency of the various kinds of geometries was an additional concern of Hilbert’s analysis, which he addressed in the Festschrift right after having introduced all the groups of axioms and discussed their immediate consequences. Although seen from the point of view of Hilbert’s later metamathematical research and the developments that followed it the question of consistency appears as the most important one undertaken in the Grundlagen, in the historical context of the evolution of his ideas it certainly was not. In fact, the consistency of the axioms is discussed in barely two pages, and it is not immediately obvious why Hilbert addressed this question at all. It doesn’t seem likely that in 1899 Hilbert envisaged the possibility that the body of theorems traditionally associated with Euclidean geometry might contain contradictions. Euclidean geometry, after all, was for Hilbert a natural science whose subject matter is the properties of physical space. Hilbert seems rather to have been echoing here Hertz’s demands for scientific theories, in particular his demand for the permissibility of images. In fact, a main point that Hilbert will stress in future opportunities, following Hertz, is that the axiomatic analysis of physical theories was meant to clear away any possible contradictions brought about over time by the

gradual addition of new hypotheses into a specific theory.36 Although this was not likely to be the case for the well-established discipline of geometry, it might still happen that the particular way in which the axioms had been formulated in order to account for the theorems of this science led to statements that contradict each other.

33. This completeness should not be confused with the later, model-theoretical notion of completeness, which is totally foreign to Hilbert’s early axiomatic approach.34. Several important changes concerning the derivability of certain theorems appeared in the successive editions of the Grundlagen. These are, however, not directly relevant to the main concerns of this article.35. However, there were many subsequent corrections and additions, by Hilbert as well as by others, that sharpened still further the picture put forward by Hilbert in the first edition of the Grundlagen. Toepell 1986, 252, presents a table summarizing the interconnections between theorems and groups of axioms as known by 1907. See also Freudenthal 1957 for later developments.36. See for instance Hertz 1956, 7-8: “It is not by finding out more and fresh relations and connections that [the question of the nature of force] can be answered; but by removing the contradictions existing between those already known, and thus perhaps by reducing their number.”

13Hilbert and Physics:

The recent development of non-Euclidean geometries made this possibility only more patent. Thus, Hilbert believed that in the framework of his system of axioms for geometry he could also easily show that no such contradictory statements would appear.

The publication of the Grundlagen was followed by many further investigations into Hilbert’s technical arguments, as well as by more general, methodological and philosophical discussions. One important such discussion appeared in the oft-cited correspondence between Hilbert and Gottlob Frege (1846-

1925).37 This interchange has drawn considerable attention of historians and philosophers, especially for the debate it contains between Hilbert and Frege concerning the nature of mathematical truth. But this frequently-emphasized issue is only one side of a more complex picture advanced by Hilbert in his letters. In particular, it is interesting to notice Hilbert’s explanation to Frege, concerning the main motivations for undertaking his axiomatic analysis: the latter had arisen, in the first place, from difficulties Hilbert had encountered when dealing with physical, rather than mathematical theories. Echoing once more ideas found in the introduction to Hertz’s textbook, Hilbert stressed the need to analyze carefully the process whereby physicists continually add new assumptions to existing physical theories, without properly checking whether or not the former contradict the latter, or consequences of the latter. In a letter written on December 29, 1899, Hilbert wrote to Frege:

After a concept has been fixed completely and unequivocally, it is on my view completely illicit and illogical to add an axiom—a mistake made very frequently, especially by physicists. By setting up one new axiom after another in the course of their investigations, without confronting them with the assumptions they made earlier, and without showing that they do not contradict a fact that follows from the axioms they set up earlier, physicists often allow sheer nonsense to appear in their investigations. One of the main sources of mistakes and misunderstandings in modern physical investigations is precisely the procedure of setting up an axiom, appealing to its truth (?), and inferring from this that it is compatible with the defined concepts. One of the main purposes of my Festschrift was to avoid this mistake.38

In a different passage of the same letter, Hilbert commented on the possibility of substituting the basic objects of an axiomatically formulated theory by a different system of objects, provided the latter can be put in a one-to-one, invertible relation with the former. In this case, the known theorems of the theory are equally valid for the second system of objects. Concerning physical theories, Hilbert wrote:

All the statements of the theory of electricity are of course valid for any other system of things which is substituted for the concepts magnetism, electricity, etc., provided only that the requisite axioms are satisfied. But the circumstance I mentioned can never be a defect in a theory [footnote: it is rather a tremendous advantage], and it is in any case unavoidable. However, to my mind, the application of a theory to the world of appearances always requires a certain measure of good will and tactfulness: e.g., that we substitute the smallest possible bodies for points and the longest possible ones, e.g., light-rays, for lines. At the same time, the further a theory has been developed and the more finely articulated its structure, the more obvious the kind of application it has to the world of appearances, and it takes a very large amount of ill will to want to apply the more subtle propositions of [the theory of surfaces] or of Maxwell’s theory of electricity to other appearances than the ones for which they were meant ...39

37. The relevant letters between Hilbert and Frege appear in Gabriel et al. (eds.) 1980, esp. pp. 34-51. For comments on this interchange see Boos 1985; Peckhaus 1990, 40-46; Resnik 1974.38. Quoted in Gabriel et al. (eds.) 1980, 40. The question mark “(?)” appears in the German original (after the word “Wahrheit”).

14Hilbert and Physics:

Hilbert’s letters to Frege help understanding the important role played in the development of his axiomatic point of view by his conception of the relation between physical and mathematical theories. Hilbert’s axiomatic approach clearly did not involve either an empty game with arbitrary systems of postulates nor a conceptual break with the classical entities and problems of mathematics and empirical science. Rather it sought after an improvement in the mathematician’s understanding of the latter. This motto was to guide much of Hilbert’s incursions into several domains of physics over the years to come.

4. Physics in Hilbert’s 1900 List of Problems

In 1900, speaking before the Second International Congress of Mathematicians in Paris, Hilbert presented his famous list of twenty three problems. This list implied an overarching research program that Hilbert was suggesting for the entire mathematical community for years to come. Hilbert declared that a wealth of significant open problems is a necessary condition for the healthy development of any mathematical

branch and, more generally, of the living organism that mathematics constitute.40 Empirical motivations appear in his conception of mathematics as a main source of life for that organism. Stressing once more at this opportunity the close interrelation between mathematics and the physical sciences, Hilbert stated that the quest for rigor in analysis and arithmetic should be extended to cover geometry and the physics, not only because it would perfect our understanding of the latter, but also because it would eventually provide mathematics with ever new and fruitful ideas. Commenting on the opinion that geometry, mechanics and other physical sciences are beyond the possibility of a rigorous treatment, he wrote:

But what an important nerve, vital to mathematical science, would be cut by the extirpation of geometry and mathematical physics! On the contrary I think that whenever from the side of the theory of knowledge or in geometry, or from the theories of natural or physical science, mathematical ideas come up, the problem arises for mathematical science to investigate the principles underlying these ideas and so to establish them upon a simple and complete system of axioms, that the exactness of the new ideas and their applicability to deduction shall be in no respect inferior to those of the old arithmetic concepts. (Hilbert 1902, 442)

Hilbert described the development of mathematical ideas as an ongoing dialectical interplay between the two poles of thought and experience; an interplay that brings to

light a “pre-established harmony” between nature and mathematics.41 Hilbert also expressed here his celebrated opinion that every mathematical problem can indeed be solved: “In mathematics there is no ignorabimus”. (p. 445)

The sixth problem of the list was a calling to undertake the axiomatization of physical science. Hilbert formulated it as follows:

39. Quoted in Gabriel et al. (eds.) 1980, 41. I have substituted here “theory of surfaces” for “Plane geometry”, which was the English translator’s original choice, since in the German original the term used is “Flächentheorie.”40. See especially the opening remarks in Hilbert 1902, 438. See also his remarks on p. 480.41. The issue of the “pre-established harmony” between mathematics and nature was a very central one among Göttingen scientists. This point has been discussed in Pyenson 1975.

15Hilbert and Physics:

The investigations on the foundations of geometry suggest the problem: To treat in the same manner, by means of axioms, those physical sciences in which mathematics plays an important part; in the first rank are the theory of probabilities and mechanics. (Hilbert 1902, 454)

Hilbert mentioned several existing works as examples of what he had in mind here: the fourth edition of Mach’s Die Mechanik in ihrer Entwicklung, Hertz’s Principles, Boltzmann’s 1897 Vorlesungen über die Principe der Mechanik, and also Einführung in das Studium der theoretischen Physik (1900) by the Königsberg physicist Paul Volkmann (1856-1938), with whom Hilbert, while still at his native city, may have had

the opportunity to discuss the question of the role of axioms in physics.42 Together with these well-known works on mechanics, Hilbert also mentioned a recent work by the Göttingen actuarial mathematician Georg Bohlmann (1869-1928) on the

foundations of the calculus of probabilities.43 The latter was important for physics, he said, for its application to the method of mean values and to the kinetic theory of gases.

Modeling this research on what had already been done for geometry meant that not only such theories which are considered as closer to “describing reality” should be investigated, but also other, logically possible ones. The mathematician undertaking the axiomatization of physical theories should obtain a complete survey of all the results derivable from the accepted premises. Moreover, echoing the concern already found in Hertz and in Hilbert’s letters to Frege, a main task of the axiomatization would be to avoid that recurrent situation in physical research, in which new axioms are added to existing theories without properly checking to what extent the former are compatible with the latter. This proof of compatibility, concluded Hilbert, is important not only in itself, but also because it compels us to search for ever more precise formulations of the axioms (p. 445).

Although the sixth problem is the only one in the list to refer directly to physics as such, three additional ones concern mathematical issues intimately connected to the classical problems of mathematical physics. The nineteenth problem concerns the question whether all the solutions of the Lagrangian equations that arise in the context of certain typical variational problems are necessarily analytic. The twentieth, closely related to the former and at the same time to Hilbert’s long-standing interest in the domain of validity of the Dirichlet principle, deals with the existence of solutions to partial differential equations with given boundary conditions. Finally, the twenty-third problem of the list is an appeal to extend and refine the existing methods of variational calculus. All these three problems are also strongly connected to physics, though at

variance with the sixth, they are part of mainstream, traditional research concerns.44 The role of variational principles in Hilbert’s program for axiomatizing physics will be further discussed below.

42. See Corry 1997, 101-103.43. Bohlmann 1900. This article reproduced a series of lectures delivered by Bohlmann in a Ferienkurs in Göttingen. In his article Bohlmann referred the readers, for more details, to the chapter he had written for the Encyclopädie der mathematischen Wissenschaften on insurance mathematics.44. A detailed account of the place of variational principles in Hilbert’s work, see Blum 1994 (unpublished).

16Hilbert and Physics:

But what has been said above suffices to recognize the natural place of the sixth problem of Hilbert’s 1900 list in his overall conception of science. The task of axiomatizing physical theories had arisen organically in conjunction with the very consolidation of Hilbert’s view of the centrality of the axiomatic method for studying the foundations of geometry. By 1900 his interest in physical theories had found a natural place among his overall views of mathematics, its main methods and its problems.

5. Hilbert, Minkowski, and Physics in Göttingen: 1900-1909

Following the publication of the Grundlagen, Hilbert’s main focus of attention remained in the study of the foundations of geometry, until 1903. That year Bertrand Russell published his discovery of a paradox arising in Frege’s logical system. Although contradictory arguments of the kind discovered by Russell had been made

known in Göttingen a couple of years earlier by Ernst Zermelo (1871-1953),45 it seems that Russell’s publication led Hilbert to attribute to the axiomatic analysis of logic and of the foundations of set theory a much central role in establishing the consistency of arithmetic than it had been the case until then. Beginning in 1903, an intense activity

was developed in Göttingen in this direction,46 whereby the systematic study of logic and set theory as a central issue in the foundations of mathematics was initiated in Hilbert’s mathematical circle. Zermelo was then working on the proof of the consistency of arithmetic and on the axiomatization of set theory. After 1905 Hilbert dedicated no further efforts to such foundational studies, and Zermelo was left alone at this. In 1908 Zermelo published his well-known paper on the foundations of set-

theory.47 But as early as 1902 Hilbert had begun publishing in the new domain of research that would concentrate his best efforts until 1912: the theory of linear integral equations.

Still, during all these years, Hilbert interest in physical issues became only more sustained. In 1901 and 1902 he lectured on potential theory and in 1902 and 1903 on continuum mechanics. That Hilbert considered these courses to have some original and interesting content, rather than being a simple repetition of existing presentations, is evident from the fact that the only two talks he gave in 1903 at the meetings of the

GMG were dedicated to report on what he did in those courses.48 In the winter semester of 1904-05 he taught an exercise course on mechanics and later gave a seminar on mechanics. Then, in the summer semester of 1905, in the framework of a

45. Peckhaus 1990, 48-49.46. Peckhaus 1990, 56-57.47. Zermelo 1908. A comprehensive account of the background, development and influence of Zermelo’s axioms see Moore 1982. For an account of the years preceding the publication, see esp. pp. 155 ff.48. See the announcements in the Jahresbericht der Deutschen Mathematiker-Vereinigung (JDMV). Vol. 12 (1993), 226 & 445. Earlier volumes of the JDMV do not contain announcements of the activities of the GMG, and therefore it is not known whether he also discussed his earlier courses there.

17Hilbert and Physics:

course on the “Logical Principles of Mathematical Thinking”, he gave a long and detailed overview of how the axiomatic approach should be applied in various individual physical disciplines. In the next winter semester (1905-06) he lectured again on mechanics, and then two more semesters on continuum mechanics.

In 1902 Minkowski arrived in Göttingen, following the creation of a third chair of mathematics in that university, under Hilbert’s pressure on Klein to convince the Prussian ministry. The renewed encounter between the two old friends was an enormous source of intellectual stimulation for both. As usual, their mathematical walks were an opportunity to discuss a wide variety of mathematical topics. This time, however, physics became a more prominent common interest that it had been in the past. Teaching in Zürich since 1894, Minkowski had kept alive his interest in

mathematical physics, and in particular in thermodynamics.49 While at Göttingen, he further developed this interest. In 1906 Minkowski published an article on capillarity, commissioned for the physics volume of the Encyclopädie der mathematischen Wissenschaften, edited by Arnold Sommerfeld (Minkowski 1906). At several meetings of the GMG he lectured on this, as well as on other physical issues such as Euler’s

equations of hydrodynamics and Nernst’s work on thermodynamics.50 He also taught advanced seminars on physical topics and more basic courses on continuum

mechanics, exercises on mechanics and heat radiation.51 In 1905 Hilbert and Minkowski organized, together with other Göttingen professors, an advanced seminar that studied recent progress in the theories of the electron. This seminar—whose

details Lewis Pyenson has helped reconstructing52—exemplifies the vitality of physical research at that university, and the role Hilbert and Minkowski played in fostering it. Again in 1907, the two conducted a joint seminar on the equations of

electrodynamics.53 Finally, as it is well known, during the last years of his life, Minkowski’s efforts were intensively dedicated to electrodynamics and the principle of relativity. In order to gain a deeper understanding of Hilbert’s views on physical issues during these years it is useful to discuss his 1905 course on the axiomatic method and the work of Minkowski in electrodynamics.

In order to understand correctly the context of Hilbert’s course of 1905 it is important to stress, that it started with a detailed treatment of the axioms of arithmetic and of geometry, and that in its last section it attempted to develop a formalized

calculus for prepositional logic.54 At this time the reconstruction of the logical foundations of mathematics was increasingly drawing Hilbert’s attention, and the overall aim of the course was to discuss issues related to it. In exactly what fashion

49. See Rüdenberg and Zassenhaus (eds.) 1973, 110-114.50. As registered in the JDMV, Vol. 12 (1903), 445 & 447; Vol. 15 (1906), 407.51. See the announcement of his courses in JDMV Vol. 13 (1904), 492; Vol. 16 (1907), 171; Vol. 17 (1908), 116.52. Pyenson 1979.53. Notes of this seminar were taken by Hermann Mierendorff, and they are kept in Hilbert’s Nachlass (Cod Ms 570/5).54. For a discussion of this part of the course see Peckhaus 1990, 61-75.

18Hilbert and Physics:

Hilbert conceived the possible role of the axiomatic approach as part of this reconstruction and as part of his general views on physics and on mathematics at this time is illuminatingly condensed in the following passage taken from one of the lectures:

The building of science is not raised like a dwelling, in which the foundations are first firmly laid and only then one proceeds to construct and to enlarge the rooms. Science prefers to secure as soon as possible comfortable spaces to wander around and only subsequently, when signs appear here and there that the loose foundations are not able to sustain the expansion of the rooms, it sets to support and fortify them. This is not a weakness, but rather the right and healthy path of development.55

Hilbert’s discussion of the axioms of physical science covered a surprisingly varied range of domains: mechanics, thermodynamics, probability calculus, kinetic theory of gases, insurance mathematics, electrodynamics, psychophysics. Without entering into a detailed account of what Hilbert did for each and every domain he considered, and of the background of his treatment, it is nevertheless convenient to describe the general aims he pursued in his presentation and what he did for certain

domains.56 The first domain of physics that Hilbert discussed was mechanics. He started with the axioms defining the addition of vectors, the main ideas of which he took from recent works of Gaston Darboux (1842-1917), of Georg Hamel (1877-

1954), and of the Göttingen student Rudolf Schimmack (1881-1912).57 The first three of the axioms adopted by Hilbert stipulate the existence of a sum of any two vectors, its commutativity and its associativity. The fourth axiom connects the sum-vector with the directions of the factors as follows:

4. Let aA denote the vector (aAx,aAy, aAz), having the same direction as A. Then every real number a defines the sum:

A + aA = (1 + a)A.

i.e., the addition of two vectors having the same direction is defined as the algebraic addition of the longitudes along the straight line on which both vectors lie.58

The fifth one establishes the interchangeability of addition and rotation of vectors. The sixth axiom concerns continuity:

6. Addition is a continuous operation, i.e., given a sufficiently small domain G around the end-point of A + B one can always find domains G1 and G2, around A and B respectively, such that the end-point of the sum of any two vectors belonging to each of these domains will always fall inside G. (p. 124)

55. Hilbert 1905, 102: “Das Gebäude der Wissenschaft wird nicht aufgerichtet wie ein Wohnhaus, wo zuerst die Grundmauern fest fundiert werden und man dann erst zum Auf- und Ausbau der Wohnräume schreitet; die Wissenschaft zieht es vor, sich möglichst schnell wohnliche Räume zu verschaffen, in denen sie schalten kann, und erst nachträglich, wenn es sich zeigt, dass hier und da die locker gefügten Fundamente den Ausbau der Wohnräume nicht zu tragen vermögen, geht sie daran, dieselben zu stützen und zu befestigen. Das ist kein Mangel, sondern die richtige und gesunde Entwicklung.”

Other places where Hilbert uses the “building metaphor” are Hilbert 1897, 67; Hilbert 1918, 148.56. For a fully detailed account see Corry 1997. The following passages condensate the arguments put forward in that article.57. Darboux 1875, Hamel 1905, Schimmack 1903. An additional related work, also mentioned by Hilbert in the manuscript, is Schur 1903.58. I have transcribed the relevant, original passages of this manuscript in Corry 1997, 132.

19Hilbert and Physics:

This last axiom of continuity plays a very central role in Hilbert’s overall conception of the axiomatization of natural science—geometry, of course, included. It is part of the essence of things—said Hilbert in his lecture—that the axiom of continuity should appear in every geometrical or physical system. Therefore it can be formulated not just with reference to a specific domain, as was the case here for vector addition, but in a much more general way. A very similar opinion had been advanced by Hertz, who in the introduction to his textbook described the continuity of nature as “an experience of the most general kind” ... “an experience which has crystallized into firm conviction in the old proposition—Natura non facit saltus” (Hertz 1956, 36-37). Hilbert formulated a general principle of continuity in the following terms:

If a sufficiently small degree of accuracy is prescribed in advance as our condition for the fulfillment of a certain statement, then an adequate domain may be determined, within which one can freely choose the arguments [of the function defining the statement], without however deviating from the statement, more than allowed by the prescribed degree. (p. 125)

Experiment—continued Hilbert—compels us to place this axiom on the top of every

natural science, since it allows to assert the validity of our assumptions and claims. In every special case, this general axiom must be given the appropriate version, as was done in an earlier part of the lectures for geometry and here for vector addition.

An interesting example of Hilbert’s detailed treatment of a physical theory appears in the section on thermodynamics, a domain whose

central concern Hilbert described as the elucidation of the two main theorems of the theory of heat. Hilbert declared that the system of axioms he was about to discuss provided a new foundation of thermodynamics, which would resemble closely the kind of axiomatic treatment used earlier in his discussion of mechanics. His stress on the mathematical elegance of the presentation led to an unusual order in the introduction of the concepts, in which the immediate physical motivations are not directly manifest. For simplicity he considered only homogenous bodies (a gas, a metal), denoting by v the reciprocal of the density. If H denotes the entropy of the body, then these two magnitudes are meant to fully characterize the elastic and the thermodynamical state of the body. Hilbert introduced the energy function , meant to describe the state of matter. The various possible states of a certain amount of matter are represented by the combinations of values of v and H, and they determine the corresponding values of the function e. This function then allows to provide a foundation of thermodynamics by means of five axioms, as follows:

I. Two states 1,2 of a certain amount of matter are in elastic equilibrium with one another if

i.e., when they have the same pressure. By pressure we understand here the negative partial derivative of the energy with respect to v

.

II. Two states 1,2 of matter are in thermal equilibrium when

� �� ( , )v H

20Hilbert and Physics:

.

i.e., when they have the same temperature q. By temperature we understand here the derivative of the energy with respect to the entropy:

The definitions of pressure and temperature are among the examples of Hilbert’s subordinating physical meaning of concepts to considerations of mathematical convenience. Assume now that v and H are functions of time t, and call the set of points in the v,H plane between any two states a path. One introduces two new functions of the parameter t: Q(t) (heat) and A(t) (work). Given two states and a

path between them, the total heat acquired between the two states is , and similarly for the work. Hilbert added the following axiom involving these functions:

III. The sum of acquired work and heat on a given path between 1 and 2 equals the difference of the energy-functions at the endpoints:

This the law of conservation of energy, or of the mechanical equivalent.

The remaining axioms are:

IV. On a path with H = const., the total heat acquired is zero. A path of this kind (parallel to the v-axis is called adiabatic).

V. On a path with v=const. the total work introduced is zero.

To these five the Hilbert added—as he had done before for geometry, for vector addition, and for mechanics—the continuity axiom. For thermodynamics this axiom is formulated as follows:

VI. Given two paths connecting the points 1,2, the quantities of heat added when moving along those two paths may be made to diverge from one another less than any arbitrarily given quantity, if the two paths are sufficiently close to one another in a uniform way (i.e., the two lie in a sufficiently narrow strip around the other).

As an important feature of this system of six axioms Hilbert stressed its symmetric treatment of work and heat. This clearly appears as a very convenient feature from the perspective of Hilbert’s mathematical account of the theory, but one wonders how physicists may have reacted to it, being opposed as it to a basic dichotomy of thermodynamics, namely, that between reversible and irreversible processes. Hilbert also discussed briefly the logical interdependence of the axioms. From axioms VI. and III., for instance, one can deduce a continuity condition similar

���

���

( , ) ( , )v H

H

v H

Hv vH H

v vH H

���

���

� ���

���

1

1

2

2

���

��� �

( , )( , )

v H

Hv H

dQdQ

dtdt

1

2

1

2

� ��

dQ dA v H v Ht

t

t

t

� � � �� �2

1

2

1

1

2

2 2 1 1� � �( , ) ( , ).

21Hilbert and Physics:

to VI., but valid for work rather than for heat. Finally, Hilbert showed how some of the basic results of thermodynamics, such as the entropy formula, can actually be derived from this system. Consider the curves of constant temperature (isothermals) Q(v,H) = const. In order to move along one of this curves from the point H = 0, to the point H, one uses a certain amount of heat, which depends only on the temperature Q and on H:

The quantity of heat invested in moving through an isothermal line is given by the

function . But what is the exact form of this function? Its determination, Hilbert said in this lecture, is typical of the axiomatic method. It is the same problem as, in the case of geometry, the determination of the function that represents the straight line; or, in the addition of vectors, the proof that the components of the vector that represents the addition equal the additions of the components of the factors. In all these cases, the idea is to decompose the properties of a certain function into small, directly evident axioms, and from them to obtain its precise, analytical representation. In this way—he concluded—we

obtain directly from the axioms, the basic laws of the discipline investigated.59

At least one well-known, published work on the foundations of thermodynamics was directly influenced by these lectures of Hilbert: an article of 1909 by Constantin Carathéodory (1873-1950). An examination of Carathéodory’s axioms and of his more general remarks in the article indicate very clearly the influence of Hilbert’s ideas. In 1925 Carathéodory presented a second axiomatic treatment of thermodynamics in which this influence is even more visible. Elaborating on a suggestion of Planck, Carathéodory discussed the place of irreversible processes in thermodynamics. He referred here once again to this earlier paper and explained what he had tried to do in it. He thus wrote:

If one believes that geometry should be seen as the first chapter of mathematical physics, it seems judicious to treat other portions of this discipline in the same manner as geometry. In order to do so, we are in possession since ancient times of a method that leaves nothing to desire in terms of clarity, and that is so perfect that it has been impossible ever since to improve essentially on it. Newton felt this already when trying to present also his mechanics in an external appearance that would fit the classical model of geometry. It is quite remarkable that with even less effort than in mechanics, the classical thermodynamics can be treated by the same methods as geometry.

This method consists in the following:

1. Create thought experiments, as in the case of geometry, constructing figures or moving around spaces figures already constructed.

2. Apply to these thought experiments the axioms that the objects considered are supposed in general to satisfy.

3. Extract the logical conclusion that follow from the given premises. (Carathéodory 1925, 176-177)

59. Hilbert 1905, 163: “Allemal handelt es sich darum, die Eigenschaften einer gewissen Funktion im kleine unmittelbarer evidente Axiome zu zerlegen, und aus ihnen dann die anlitysch Darstellung der Funktion herzuleiten; diese läßt dann die wesentlichen Eigenschaften Sätze der vorliegenden Disziplin unmitelbar zu erkennen.”

dQ f HH

H

v H

��

��

�� �

0 T T

�( , )

( , ).

f H( , )�

22Hilbert and Physics:

Carathéodory explained that in his 1909 article he had proceeded exactly in this way, but, in his opinion, only in this later paper of 1925 the parallel application of the axiomatic method to thermodynamics and geometry was more clearly manifest.

Carathéodory work had itself little impact among contemporary physicists. This becomes evident from a paper published in 1921 by Max Born in the Physikalische Zeitschrift, aimed precisely at making Carathéodory’s point of view more widely known than it was. Born’s article, in turn, interestingly makes manifest the influence of Hilbert on his own conception of the link between physics and mathematics. In its traditional presentation, Born said, thermodynamics had not attained the logical separation—so desirable, and in fact necessary, in the eyes of this disciple of the Göttingen school—between the physical content and the mathematical representation of the theory. Born’s characterization of the litmus test for identifying the point when this separation is achieved brings us back directly to Hilbert’s 1905 lecture: a clear specification of the way to determine the form the entropy function (Born 1921, 218).

In his article, Born reelaborated Carathédory’s presentation of thermodynamics, in a way he thought more amenable to appreciation by physicists. His article seems to have had no more noticeable influence than the one that inspired it. But for the purposes of the present account it is very helpful for understanding the way Hilbert conceived of the role of axiomatization on physical theories: starting from the basic facts of experience, to strive after the formulation of an elaborate mathematical theory in which the physical theorems are derived from simple axioms. This theory may itself be different from the classical, more physically intuitive one, but the mathematical presentation helps providing a more unified view of physics as a whole.

Hilbert next discussed the calculus of probability, which, as we saw above, he considered to be among the natural sciences that deserved an axiomatic treatment. Hilbert presented a system of axioms that he said to have taken from an article on insurance mathematics that Georg Bohlmann published in the Encyclopädie der mathematischen Wissenschaften (1901). Like Bohlmann in his article, beyond stating the axioms as such Hilbert did not go any further. He did not comment on the independence, consistency or “completeness” of these axioms. This system was a rather crude one from the point of view of Hilbert’s conception; more elaborate ones

were attempted after Bohlmann.60

Much more interesting than the calculus of probabilities as such, however, were for Hilbert its applications, among which he mentioned three: the theory of compensations of errors (Ausgleichungsrechnung), the kinetic theory of gases, and insurance mathematics. Of special interest for the present account is Hilbert’s discussion of the kinetic theory. This theory will occupy much of Hilbert’s efforts during the years 1911 to 1913, but his lectures of 1905 is perhaps the first place where we find a clear evidence to the deep attraction that the theory exerted on him. Hilbert

60. For a review of later attempts to axiomatize the calculus of probabilities until 1933, see Schneider (ed.) 1988, 353-358. A more detailed account appears in Von Plato 1994; see especially pp. 179-278, for the foundational works of von Mises, Kolmogorov, and De Finetti.

23Hilbert and Physics:

admired the way this theory combined the postulation of far-reaching assumptions concerning the structure of matter with the use of the theory of probabilities, thus leading to new and interesting physical results. Moreover, the particular historical development of this theory offered a remarkable example of the kind of problematic situation that Hilbert’s axiomatic analysis was meant to help overcoming. In fact, along the years additional assumptions had been gradually added to the existing body of knowledge related to the theory without properly checking the possible logical

difficulties that would arise from this addition.61 Also the question of the role of probability arguments in physics was not a settled one in this context. In Hilbert’s view, the axiomatic treatment was the proper way to restore order to this whole system of knowledge, so crucial to the contemporary conception of physical science.

Hilbert accepted without any further qualification the controversial atomistic assumptions underlying the classical approach to this theory as developed by Ludwig Boltzmann. What Hilbert saw as especially problematic was the status of the probabilistic arguments in a physical theory. Even if we know the exact position and velocities of the particles of a gas—Hilbert explained—it is impossible in practice to integrate all the differential equations describing the motions of these particles and their interactions. We know nothing of the motion of individual particles, but rather consider only the average magnitudes that constitute the subject of the probabilistic, kinetic theory of gases. In an oblique reference to the arguments that Boltzmann had devised as replies to the objections raised against his theory, Hilbert stated that the combined use of probabilities and infinitesimal calculus in this context is a very original contribution of mathematics, which may lead to deep and interesting consequences, but which at this stage has in no sense been fully justified. As an example Hilbert mentioned the equations of the vis viva. In the probabilistic version of the theory, Hilbert said, the solution of the corresponding differential equation is not derived using the differential calculus alone and yet it is correctly determined. It could well be the case, however, that the application of the probability calculus could have led here to results that contradict well-known results of the theory. Such results would clearly be considered false. Hilbert explained what he meant with this warning, by showing how a fallacious probabilistic argument could lead to contradiction in the

theory of numbers.62

Following the detailed discussion of these issues, Hilbert formulated a very interesting, and surprisingly pragmatist, opinion concerning the role of probabilistic arguments in mathematical and in physical theories. According to this opinion the calculus of probability is not an exact mathematical theory, but one that may appropriately be used as a first approximation, provided we are dealing with immediately perceivable mathematical facts. Otherwise it may lead to significant contradictions. The use of the calculus of probabilities is justified—Hilbert concluded—inasmuch as it leads to results that are correct and in accordance with the facts of experience or with the accepted mathematical theories (p. 182).

61. Two classical, detailed accounts of the development of the kinetic theory of gases and the conceptual problems involved in it (particularly during the late nineteenth century) can be consulted: Brush 1976 and Klein 1970 (esp. 95-140).62. See Hilbert 1905, 178-180. For more details on these arguments see Corry 1997, 167.

24Hilbert and Physics:

An in important feature of Hilbert’s discussion of insurance mathematics is the methodological comparison he drew between this domain and thermodynamics. In the

latter domain, the state of matter had been expressed in terms of a function . A similar thing Hilbert intended to do here for describing the relevant state of a person: an individual person is characterized, for the purposes of insurance, by means of a function p(x,y), defined for y > x. This function expresses the probability that a person of age x will reach age y, and it is required to satisfy the following axiom:

The probabilities p(x,y) p’(x',y') associated with two different individuals are independent for all pairs x,y x',y' of positive numbers.

A collection of individuals, such for that any two of them p(x,y) = p’(x,y), is called an equal-risk group. From the point of view of insurance, the individuals of any of these collections are identical, since the function p wholly characterizes their relevant behavior. Now, very much like in thermodynamics, where the main accomplishment achieved in the lectures was the explicit derivation of the form of the function

, using to this end only the particular axioms postulated, Hilbert declared aim for this domain was the determination of a certain function of one variable starting from the axioms. For every equal-risk group associated with a function of probability p(x,y) Hilbert defined a “virtual mortality-order” (fingierte Absterbeordnung). In this way, with every such group one can associate a function of the continuous variable x, l(x), called the “number of living people of age x” or “life function”, satisfying some simple properties. Without entering here into the details of the axioms postulated by Hilbert for this function, I will only say that Hilbert did actually not prove any of the results pertaining to this theory and to the functions p and l. He just stated that such proofs would involve a kind of deductions similar to those used in the other domains. He added, however, that also in these deductions, an axiom of continuity of the kind assumed in the former domains—the particular version of which he would not formulate explicitly in this case—plays a central role.

The last example I want to mention here concerns Hilbert’s treatment of psychophysics. Hilbert referred in this section to a recent work on the theory of color perception published by Egon Ritter von Oppolzer, a psychologist from Innsbruck (Oppolzer 1902-3). Oppolzer’s article was a classical representative of the German school of experimental psychology, going back to the work of Gustav Fechner (1801-

1887).63 Its declared aim was to characterize the sensation of light in “total colorblind systems” by means of a single, purely psychological parameter—the brightness (Helligkeit)—as opposed to the physically characterizable concept of intensity (Intensität). The problem addressed by Oppolzer, as Hilbert characterized it in his lectures, was to express the magnitude of this parameter as a function of the intensity

and wave-length of light.64 The derivation of the analytic expression of this function, starting from axioms fitted very well with the other domains considered earlier in the lectures.

63. On Fechner’s contributions see Boring 1929, 265-287. More generally, on the German school, see there, pp. 237-401. Oppolzer is mentioned neither in Boring’s classical account, nor in other, standard similar works.64. Hilbert 1905, 189: “Das Hauptproblem ist, diese Helligkeit x als Funktion der Bestimmungstücke der das Licht physisch (sic) zusammensetzenden homogenen Lichter (d.i. Intensität und Wellelänge eines jeder) darzustellen.”

� � ( , )v H

f H( , )�

25Hilbert and Physics:

As in the case of Bohlmann’s work on probabilities, the axioms mentioned by Hilbert for the case of psychophysics can be only retrospectively found in Oppolzer’s own article. Oppolzer himself described his basic assumptions discursively, sometimes loosely, and not only in the opening sections, but rather throughout the article. Needless to say, he didn’t analyze the independence, consistency or any other property of his “axioms”. Yet, precisely for the unsystematic way in which Oppolzer discussed principles and ideas drawn from works as diverse as those of Goethe and the German physiologists, Newton and Thomas Young, this work seems to have presented Hilbert with a further, unexplored territory in which the axiomatic approach could usefully be applied. In fact, Oppolzer’s article was in this sense symptomatic of a more

generalized situation in contemporary research in psychophysics,65 and was therefore well-suited to exemplify Hilbert’s claims concerning the unmindful introduction of new assumptions into existing physical theories.

The manuscript of the lectures makes no mention of the differences between Hilbert’s formulation and Oppolzer’s own one. Hilbert simply put forward his axioms, which are defined for a collection of “brightnesses” x1, x2, ..... The axioms postulate the following properties that the brightnesses are demanded to satisfy:

1. To every pair of brightnesses x1, x2, a third one [x1,x2] can be associated, called “the brightness of the mixed light of x1,x2.” Given a second pair of brightnesses x3,x4, such that x1 = x3 and x2 = x4, then [x1,x2] = [x3,x4].

2. The “mixing” of various brightnesses is associative and commutative.

3. By mixing various homogeneous lights of equal wave-lengths, the brightness of the mixed light has the same wave light, while the intensity of the mixed light is the sum of the intensities.

Experience, said Hilbert, amply confirms these three axioms. The first one contains what Hilbert called the law of Grassmann, namely, that intensities that are psychically equal (but which may be physically different), after undergoing the operation of physical mixing, remain equivalent at the psychical level.

If one calls the uniquely determined number [x1,x2], x12, one can write it as a function of the two parameters

.

From the second axiom, one can derive the functional equation:

.

One can then introduce a new function F that satisfies the following relation:

.

65. As Kremer 1993, 257, describes it: “For a variety of philosophical, institutional and personal reasons, color researchers between 1860 and 1920 simply could not agree on which color experiences are quintessential or on what criteria are appropriate to evaluate hypothetical mechanisms for a psychoneurophysiological system of sensation.”

x f x x( ) ( , )12 1 2�

f f x x f x x f f x x f x x f f x x f x x( ( , ), ( , )) ( ( , ), ( , )) ( ( , ), ( , ))1 2 3 4 1 3 2 4 1 4 2 3� �

F x F f x x F x F x( ) ( ( , )) ( ) ( )12 1 2 1 2� � �

26Hilbert and Physics:

From axiom 3, and assuming the by now well-known general postulate of continuity, it follows that the function F , for homogeneous light, is proportional to the intensity. This function is called the “stimulus value” (Reizwert), and once it is known, then the whole theory becomes, so claimed Hilbert, well-established. One notices immediately, Hilbert went on to say, the analogy with the previously studied domains, and especially with the theorem of existence of a function l(x) in the life-insurance mathematics. This very analogy could suffice to acknowledge, he concluded, that also in this latter domain, so far removed from the earlier ones, the approach put forward in the whole

course would become fruitful.66

Hilbert’s treatment of psychophysics, at least as it appears in the manuscript, was rather sketchy and its motivation was far from obvious, since he did not provide any background to understand the actual research problems of this domain. Moreover, as in the case of probabilities, Hilbert did not examine the logical interrelations among the axioms, beyond the short remarks quoted in the preceding paragraphs. Yet, in the context of his treatment of other physical domains and of the confused state of affairs in contemporary psychological research, one can grasp the width of application that Hilbert envisaged for the axiomatic method in science. Hilbert’s ideas don’t seem to have influenced in any tangible way the current research of German psychologists, and one wonders whether or not there was some personal contact between him an his psychologist colleagues, at least in Göttingen.

There is no direct evidence to judge what was the reaction of the students who attended these lectures of Hilbert, that were listed among the elementary courses offered in Göttingen in that semester. Before those students stood the great Hilbert, quickly overviewing many different physical theories, together with arithmetic, geometry and even logic, all in the framework of a single course. Hilbert moved form one theory to the other, and from one discipline to the next, without providing motivations or explaining the historical background to the specific topics addressed, without giving explicit references to the sources, without stopping to work out any particular idea, without proving any assertion in detail, but claiming all the while to have a unified view of all these matters. The impression must have been thrilling, but one wonders to what extent his students could really appreciate the ideas presented to them. In fact, it is hard to determine with exactitude to what extent Hilbert himself dominated the physical subtleties of the issues he discussed, though there can be no doubt that his unusual abilities helped him overcoming very easily what perhaps presented great mathematical difficulties to others. Still, the picture of Hilbert’s knowledge of physics that arises from these lectures is impressive both in its breadth and its incisiveness.

In Hilbert’s treatment of physical theories of 1905 we come across diverse kinds of axioms. In the first place, every theory is assumed to be governed by specific axioms that characterize it. These axioms usually express mathematical properties establishing relations among the basic magnitudes involved in the theory. Then, there are certain general mathematical principles that Hilbert thought should be valid of all

66. Hilbert 1905, 190: “Das mag zur Kennzeichnung genügen, wie auch in diesem von den früheren so ganz verschiedene Gebiete unsere Gedankengänge fruchtbar werden.”

27Hilbert and Physics:

physical theories. In the lectures he stressed above all the “continuity axiom”, providing both a general formulation and more specific ones for each theory. As an additional general principle of this kind he suggested the assumption that all functions appearing in the natural sciences should have at least one continuous derivative. Furthermore, the universal validity of variational principles as the key to deriving the main equations of physics, especially in mechanics and electrodynamics (which were not described in detail above) was a central underlying assumption of all of Hilbert’s work on physics, and they appear throughout these lectures as well. In each of the theories he considered in his 1905 lectures, Hilbert attempted to show how the exact analytic expression of a particular function that condenses the contents of the theory in question could be effectively derived from the specific axioms of the theory, together with more general principles. On some occasions he elaborated this more thoroughly, while on others he simply declared that such a derivation should be possible.

There is yet a third type of axiom for physical theories, however, which Hilbert avoided addressing in his 1905 lectures. They comprise claims about the ultimate nature of physical phenomena, an issue which was particularly controversial during the years preceding these lectures. Although Hilbert’s sympathy for the mechanistic world-view is apparent throughout the manuscript of the lectures, his axiomatic analysis of physical theories contain no direct reference to it. The logical structure of the theories should be fully understood independently of any particular position one would assume in this debate. As will be seen below, this position would change around 1913, when Hilbert wholeheartedly adopted the view that all physical phenomena should be explained in terms of electrodynamic processes. I discuss this issue below.

After the summer semester of 1905, Hilbert lectured again on mechanics (WS

1905-06)67 and continuum mechanics (SS 1906 and WS 1906-07). Since then and until 1910, he taught no additional courses on physics. But at the same time, from 1907 and until his death in 1909, Hilbert’s colleague, Minkowski, dedicated much efforts to the study of electrodynamics and the principle of relativity. It would seem as if Hilbert had left the stage open to Minkowski alone to work out his ideas on these issues. As a matter of fact, however, Minkowski’s work can be seen to a large extent as a realization of Hilbert’s program of axiomatization, in which the specific role of the newly adopted principle of relativity in various physical theories was thoroughly investigated for the first time. Moreover, Hilbert’s lectures on physical issues over the years following Minkowski’s death indicate that the former fully adopted the point of view, the notations, and the concepts introduced by the latter. Thus a description of Minkowski’s work on electrodynamics may also serve as a parallel description of how Hilbert’s conceptions applied to the new situation created in physics by the introduction of the principle of relativity, and at the same time, as evidence of what Hilbert’s own ideas on physical issues looked like over that period of time of which we have no additional, direct evidence.

67. A deatiled analysis of this course, with especial emphasis on the role of variational principles in Hilbert’s conception of physics appears in (the unpublished) Blum 1994.

28Hilbert and Physics:

Elsewhere I have presented a comprehensive picture of Minkowski’s work on electrodynamics, in which his motivations and the details of his work are interpreted in

the lines suggested in the preceding paragraph.68 For the purposes of the present article it may suffice to discuss briefly only one of Minkowski’s articles, dealing with “The Basic Equations of Electromagnetic Processes in Moving Bodies” (Minkowski 1908). This was Minkowski’s only publication on this issue that appeared in print during his lifetime. This was also the article that Hilbert considered as his friend’s most significant contribution to this field. In his obituary of Minkowski, Hilbert stressed the importance and the innovative character of the axiomatic analysis presented in that article, as well as of Minkowski’s derivation of the equations for moving matter starting from the so-called “World-postulate”, and from three additional axioms. The correct form of these equations had been theretofore an extremely controversial issue among physicists, but this situation had totally changed—so Hilbert believed—thanks to Minkowski’s work (Hilbert 1909, 93-94).

Minkowski’s article opened with an analysis of current developments in the theories of the electron and of the role played by the principle of relativity in them. Minkowski distinguished three possible different meanings of this principle. First, one has the plain mathematical fact that the Maxwell equations, as formulated in Lorentz’s theory of electrodynamics, are invariant under the Lorentz transformations. Minkowski called this latter fact the “theorem of relativity.” The “postulate of relativity” differs from the theorem in that it expresses more of a confidence (Zuversicht) than an objective assessment concerning some actual state of affairs: a confidence that the domain of validity of the theorem may be extended to cover all laws governing ponderable bodies, including laws that are still unknown. He compared this postulate to the postulation of the validity of the principle of conservation of energy, which we assume even for forms of energy that are not yet known. Finally, there is the “principle of relativity” which expresses the assertion that the expected Lorentz covariance actually holds as a relation between purely observable magnitudes relating to a moving body (Minkowski 1908, 353).

Minkowski’s declared aim was to deduce the exact expression of the equations for moving bodies from the principle of relativity. He claimed that his formulation of the principle had never before been articulated the way he did and that under earlier formulations the equations were not truly invariant. Minkowski believed that his axiomatic analysis of the principle of relativity and of the electrodynamics theories of moving bodies was the best approach for unequivocally obtaining the correct equations. Minkowski started by presenting the equations for pure ether. Then, in the second part of his article, he discussed how the equations change when matter is added to the ether. For the case of a body at rest in the ether, Minkowski relied on Lorentz’s version of Maxwell’s equations, and analyzed the symmetry properties of the latter. He formulated the equations as follows:

(I)

(II)

68. See Corry 1997a.

curlme

s� ���t

div e � �

29Hilbert and Physics:

(III)

(IV)

Here M and e are called the magnetic and electric intensities (Erregung) respectively, E and m are called the electric and magnetic forces, is the electric density, s is the electric current vector (elektrischer Strom). The properties of matter, in the case of isotropic bodies, are characterized by the following equations:

(V) ,

where � is the dielectric constant, � is the magnetic permeability, and � is the conductivity of matter.

From the basic properties of the equations for bodies at rest, Minkowski deduced the fundamental equations for the case of a body in motion. This deduction is where the detailed axiomatic derivation is realized: Minkowski assumed the validity of the previously discussed equations for matter at rest and to them he added three more axioms. He then sought to derive the equations for matter in motion exclusively from the axioms, together with the equations for rest. Minkowski’s axioms are:

1. Whenever the velocity v of a particle of matter equals 0 at x, y, z, it in some reference system, then equations (I)-(V) also represent, in that system, the relations among all the magnitudes: , the vectors s, m, e, M, E, and their derivatives with respect to x, y, z, it.

2. Matter always moves with a velocity which is less than the velocity of light in empty space (i.e., �v�= v < 1).

3. If a Lorentz transformation acting on the variables x, y, z, it, transforms both m,-ie and M,-iE as space-time vectors of type II, and s,i as a space-time vector of type I, then it transforms the original equations exactly into the same equations written for the transformed magnitudes.69

This last axiom is what Minkowski called the principle of relativity.

Minkowski deduced the equations for moving bodies in a section were he showed in detail how every step of the deduction is allowed by one of the axioms. On first reading, his straightforward argument may seem somewhat out of place amidst all the elaborate mathematical and physical arguments displayed throughout his long article. But when seen in the light of the kind of axiomatic conceptual clarification promoted by Hilbert in his lectures on physics, it becomes clear that Minkowski was simply stressing this deduction as a central task of his exposition of the theory. Moreover, Minkowski went on to check to what extent different existing versions of the equations satisfy the principle as stated in his axioms. Minkowski’s implicit assumption was that only such equations can be accepted as correct, which comply to his own version of the principle. Thus Minkowski showed that the equations for moving media formulated by Lorentz in his Encyclopädie article (Lorentz 1904) are in certain cases incompatible with his principle. Minkowski also discussed the equations

69. See Minkowski 1908, 369. For the sake of simplicity, my formulation here is slightly different yet essentially equivalent to the original one.

curlEM

� ���t

0

div M � 0

e E M m s E� � �� � �, ,

30Hilbert and Physics:

formulated in 1902 by Emil Cohn, pointing out that they agree with his own ones up to terms of first order in the velocity (Minkowski 1908, 372). After having formulated the equations and discussed their invariance properties, Minkowski dealt in detail, in three additional sections, with the properties of electromagnetic processes in the presence of matter.

Minkowski’s article also contained an appendix discussing the relations between mechanics and the postulate of relativity. In this appendix the similarity of Minkowski’s and Hilbert’s treatment of physical theories is more clearly manifest: it explores the consequences of adding the postulate of relativity to the existing building of mechanics, and the compatibility of the postulate with the already established principles of this discipline. The extent to which this addition can be successfully realized provides, in Minkowski’s view, a standard to assess the status of Lorentz covariance as a truly universal postulate of all physical science.

Using the four-vector formalism that he had introduced in the earlier sections of his article, Minkowski showed that the equations of motion of classical mechanics are invariant under the Lorentz group only under the assumption that c = �. It would be embarrassing or perplexing (verwirrend), he said, that the laws of transformation of the basic expression

into itself would necessitate a certain finite value of c in a certain domain of physics and a different, infinite one, in a second domain. In view of this situation, the postulate of relativity (i.e., our confidence in the universal validity of the theorem) compels us to see Newtonian mechanics only as a tentative approximation initially suggested by experience, which however must be corrected to make it invariant for a finite value of c. Minkowski thought that reformulating mechanics in this direction was possible, and moreover— expressing himself in terms that could have been equally found in Hilbert’s lecture notes— he asserted that such a reformulation would seem only to perfect, to a considerable extent, the axiomatic structure of mechanics as currently

conceived.70

Clearly, the universal validity of the postulate of relativity could only be asserted if one could show that it does not contradict the observable phenomena related to gravitation. To this effect, in the last section of his article, Minkowski sketched a proposal for a theory of gravitation that would also be Lorentz covariant. This sketch was no more than a preliminary attempt in this direction, that neither Minkowski himself, nor Hilbert went on to pursue.

A main motivation of Minkowski’s work in electrodynamics, then, is a systematic investigation of the logical consequences of assuming the universal validity of Lorentz covariance for all physical disciplines. This is exactly the formulation used by Hilbert in his future lectures to describe the contents of the “new mechanics.” The

70. Minkowski 1908, 393 (Italics in the original.): “Ich möchte ausführen, daß durch eine Reformierung der Mechanik, wobei an Stelle des Newtonschen Relativitätspostulates mit c = � ein solches für ein endliches c tritt, sogar der axiomatische Aufbau der Mechanik erheblich an Vollendung zu gewinnen scheint.”

� � � �x y z c t2 2 2 2 2

31Hilbert and Physics:

postulate of relativity had been strongly suggested by experimental results obtained during the late nineteenth century, and its theoretical implications had been investigated from different perspectives in recent works, noticeably those of Lorentz, Poincaré and Einstein. Yet, in a spirit similar to that underlying Hilbert’s program, Minkowski believed that the logical structure of the physical theories built on the principle of relativity had not yet been satisfactorily elucidated. The postulate of relativity should be taken as a further axiom appearing at the basis of each and every physical theory, together with the particular axioms of that theory. In his work Minkowski was able to prove for certain domains of physics that the ensuing theory indeed produced a consistent logical building. Concerning gravitation, he was less successful, but always declared his conviction to have indicated in principle how a consistent, Lorentz covariant theory of gravitation could eventually be elaborated in detail.

For Minkowski the postulate of relativity was not simply an additional axiom, with perhaps a wider domain of validity in physics than others. It was an axiom of a different nature: a principle that should be valid for every conceivable physical theory, even those theories that were yet to be discovered or formulated. As I mentioned above, also Hilbert laid a great stress on the importance of this kind of universal physical principle, and focused especially on the “principle of continuity.” Minkowski’s comparison in this context, of the status of the postulate of relativity with

that of the principle of conservation of energy, had been drawn earlier by Einstein.71 But Einstein’s and Minkowski’s comparisons were basically different. Einstein spoke in his article of two “open” principles of physics, with a strong heuristic character. Unlike Minkowski and Hilbert, Einstein did not see the principle of relativity and the principle of energy conservation as parts of strictly deductive systems from which the

particular laws of a given domain could be derived.72 More generally, although Einstein introduced the principle of relativity together with the constancy of light at the beginning of his 1905 article as “postulates” of the theory (in some sense of the word), it is necessary to draw a clear difference between what he did and what Minkowski and Hilbert had in mind when speaking of an axiomatic analysis of the postulate of relativity. As a matter of fact, one of the main explicit aims of Hilbert’s program was to address situations similar to the one created here by Einstein, which had the potential of turning out to be problematic, namely, that faced with conflict between an existing theory and new empirical findings, it was common for physicists to add new hypothesis that apparently settle down the disagreement in question but which perhaps contradict some other consequences of the existing theory. Hilbert thought that an adequate axiomatic analysis of the principles of a given theory would help clear away possible contradictions and superfluous additional premises created by the gradual introduction of new hypotheses into existing theories. This was also what Minkowski was pursuing with his analysis: to verify that the introduction of the principle of relativity will not create such a problematic situation.

71. Einstein 1907. Minkowski was very likely aware of this specific article of Einstein, if only because it appeared in the Annalen der Physik as a reply to an earlier article of Paul Ehrenfest who at that time was in Göttingen.72. Cf. Einstein 1907, 411: “Es handelt sich hier also keineswegs um ein ‘System’, in welchem implizite die einzelnen Gesetze enthalten wären, und nur durch Deduktion daraus gefunden werden könnten, sondern nur um ein Prinzip, das (ähnlich wie der zweite Hauptsatz der Wärmetheorie) gewisse Gesetze auf andere zurückzuführen gestattet.”

32Hilbert and Physics:

Minkowski’s analysis of the place of the postulate of relativity in physics avoided any assumption about, and certainly any commitment to, any particular

conception of the ultimate nature of physical phenomena.73 Perhaps Minkowski had some kind of clear position of his own on these issues, though we have no direct evidence of it. Minkowski’s admiration for Hertz and the fact that in 1910 Hilbert sided with the mechanistic world-view when lecturing on mechanics under the

declared influence of Minkowski’s ideas74 might seem to suggest that this was also the latter’s view. Hilbert was initially very attracted to the mechanistic reductionism of Hertz or Boltzmann and around 1913, as will be seen below, he changed his position diametrically and adopted an electromagnetic, reductionistic world-view based on Gustav Mie’s theory of matter. Yet, both Minkowski and Hilbert, at least by this time considered that the much needed task of axiomatically analyzing the logical building of physical theories should be carried out while leaving aside this kind of unsettled physical issues.

6. Mechanics, Kinetic Theory, Radiation Theory: 1910-1914

Until 1912, Hilbert’s mathematical efforts concentrated on the theory of linear integral equations, publishing the successive installments of what finally constituted his classical treatise on this topic (Hilbert 1912). At the same time, however, after Minkowski’s death, Hilbert returned to teach courses on physical issues. Hilbert now moved into disciplines that he had never taught in the past. Thus, after teaching mechanics and continuum mechanics in 1910 and 1911, Hilbert taught statistical mechanics for the first time in the winter semester of 1910-11. This course marked the starting point of Hilbert’s definitive involvement with a wider variety of physical theories. In December of 1911 he presented to the GMG an overview of his recent

investigations on the kinetic theory of gases, that were soon to be published.75 The kinetic theory was also the topic of his course during the winter semester of 1911-12, and in the following semester he taught a course on radiation theory (Hilbert 1911-12). In 1912 Hilbert enrolled an assistant for physics, who was commissioned with the task of keeping him abreast of current developments in the various branches of physics. Paul P. Ewald (1888-1985), who had recently finished his dissertation in Munich, was the first to hold this position, which Hilbert maintained for many years to come. In April 1912, Hilbert was among the lecturers that took part in a twelve-day seminar on

the axiomatic foundations of physics, opened for high-school teachers.76

73. This claim, which differs from most existing interpretations of Minkowski’s work, is argued in detail in Corry 1997.74. As manifest, i.e., in Hilbert’s lecture notes: see Hilbert 1910-11, 295.75. See the announcement in the JDMV Vol. 21, p. 58.76. See the announcement in the JDMV Vol. 21, p. 166. Unfortunately, no recorded evidence of what Hilbert did in this Ferienkurs seems to have been kept.

33Hilbert and Physics:

Hilbert’s increasingly deep involvement in physics led him to ponder again, now from a wider perspective, some basic questions concerning the foundations of this discipline. By 1910 Hilbert’s approach, as already said, was dominated by the view that all physical phenomena could be reduced to mechanics. This view was clearly manifest in the courses he taught. Between 1910 and 1913, however, although his reductionistic inclinations did not change, he moved from the mechanistic to the electromagnetic point of view. Electromagnetic reductionism became the basis of all of Hilbert’s work in physics thereafter.

From the manuscripts of Hilbert’s courses between 1911 and 1913, as well as from his publications, it becomes evident that Hilbert was very much impressed by recent developments in quantum theory. The importance of these developments had particularly been discussed and highlighted during the First Solvay Conference, held in

Brussels in October 1911,77 and whose echoes must have reached Hilbert through his physicist colleagues. “Never has been a most proper and challenging time than now”—said Hilbert in the opening lecture of a course taught in 1912—“to undertake the research of the foundations of physics.” What seems to have impressed him more than anything else were the recently discovered, deep interconnections, “of which formerly no one could have even dreamed, namely, that optics is nothing but a chapter of the theory of electricity, that electrodynamics and thermodynamics are one and the same, that also energy possesses inertial properties, that physical methods have been

introduced into chemistry as well.”78 And above all, the “atomic theory”, the “principle of discontinuity”, as Hilbert said, which today is not hypothesis anymore,

but rather, “like Copernicus’s theory, a fact confirmed by experiment.”79 Very much like the unification of apparently distant mathematical domains, which played a leading role throughout his career, the unity of physical laws exerted a strong attraction on Hilbert.

The kind of interests manifest in Hilbert’s lectures of 1911 and 1912 were directly reflected in his publications of 1912 as well. The latter include articles dealing

with these two issues: the kinetic theory and radiation theory.80 Hilbert’s first major physical publications bring to the fore an interesting mixture of all the themes formerly found in his lectures and published works in both mathematics and physics: an attempt to connect apparently distant issues by uncovering their underlying, structural similarities; an appeal to the axiomatic analysis of physical theories in order to redesign their logical building and to clarify the specific roles of their basic principles; a use of deeply sophisticated mathematical tools; an attempt to clarify the interrelation between probabilistic and analytic reasoning in physics.

77. See Kormos Barkan 1993. 78. Hilbert 1912c, 2: “Nun kommen wir aber zu eigentlicher Physik, welche sich auf der Standpunkt der Atomistik stellt und da kann man sagen, dass keine Zeit günstiger ist und keine mehr dazu herausfordet, die Grundlagen dieser Disziplin zu untersuchen, wie die heutige. Zunächst wegen der Zusammenhänge, die man heute in der Physik entdeckt hat, wovon man sich früher nichts hätte träumen lassen, dass die Optik nur ein Kapitel der Elektrizitätslehre ist, dass Elektrodynamik und Thermodynamik dasselbe sind, dass auch die Energie Trägheit besitzt, dann dass auch in der Chemie (Metalchemie, Radioaktivität) physikalische Methoden in der Vordergrund haben.”79. Hilbert 1912c, 2: “... wie die Lehre des Kopernikus, eine durch das Experimente bewiesene Tatsache.”80. Hilbert 1912a, 1912b, 1913,1913a, 1914.

34Hilbert and Physics:

This is not the place to describe the details of Hilbert’s involvement with kinetic theory

and with raditation theory.81 Still, some aspects of it are worth mentioning here, since they help understanding properly the actual place of physics in Hilbert’s world. Thus, for instance, shortly after the publication of his first article on kinetic theory, Hilbert organized a seminar on this topic, together with his former student Erich Hecke. The seminar was also attended by the Göttingen docents Max Born, Paul Hertz, Theodor von Kármán and Erwin Madelung. A contemporary report in a journal published by German students of mathematics provides an invaluable source of information

concerning the issues discussed in that seminar, which included the following:82 the ergodic hypothesis and its consequences; on Brownian motion and its theories; electron theory of metals in analogy to Hilbert’s theory of gases; report on Hilbert’s theory of gases; on dilute gases; temperature split by the walls; theory of dilute gases using Hilbert’s theory; on the theory of chemical equilibrium, including a referat on the related work of Sakur [sic]; diluted solutions. It is not clear whether the term “Hilbert’s theory of gases” was meant here as what Hilbert published in his article or as the ideas he had developed in his courses, but in any case the very use of the terms suggests that Hilbert considered to have in his hands more than a simple local addition to the theory as then conceived. Moreover, the fact that the seminar discussions involved the younger colleagues mentioned above indicates that the deep physical issues that were at stake, and that are listed in detail in the report, could not have easily been ignored or discussed only superficially. Especially indicative of Hilbert’s surprisingly broad spectrum of interests is the reference to the work of Otto Sackur (1880-1914). Sackur was a physical chemist from Breslau, whose work dealt mainly with the laws of chemical equilibrium in ideal gases, on Nernst law of heat, and who wrote a widely used textbook on thermochemistry and thermodynamics (Sackur 1912). Sackur work included also an important experimental side, and in general his work was far from the typical kind of mathematical physics with which Hilbert and the Göttingen school may

more easily be associated.83

Also significant is the fact that Hilbert’s ideas on kinetic theory seem to have influenced the work of several of his students. In the first place one can mention two doctoral dissertations written under his supervision on issues related to the kinetic theory: one by Hans Bolza (concluded in July of 1913) and one by Bernhard Baule

(concluded in February of 1914).84 Second, other young Göttingen scientists, like Max Born, Theodor von Kármán and Erich Hecke, who had attended Hilbert’s seminar,

published in this field under its influence.85 But perhaps of a much greater impact on the development of the theory was the work of the Swedish physicist David Enskog

(1884-1947), who attended Hilbert’s lectures of 1911-12.86 Building on ideas

81. Such a discussion can be found in Corry 1998.82. Reference to this seminar appears in Lorey 1916, 129. Lorey took this information from the Semesterberichte des Mathematischen Verereins, but he does not give the exact year of the seminar.83. See Sackur’s obituary in Physikalische Zeitschrift Vol. 16, 1915, 113-115. Sackur died in an explosion at the Kaiser-Wilhelm-Institut in Berlin-Dahlem, while taking part in current research on explosives, related with the war effort.84. This was published as Baule 1914.85. Cf. for instance: Bolza, Born & van Kármán 1913; Hecke 1918; Hecke 1922.86. See Mehra 1973, 178. I know of no evidence to decide whether Enskog attended the above mentioned seminar.

35Hilbert and Physics:

contained in Hilbert’s article, Enskog developed what has come to constitute, together with the work of Sydney Chapman, the now standard approach to the whole issue of

transport phenomena in gases.87 Although a detailed analysis of Hilbert’s influence on Enskog is yet to be written, there can be little doubt that such an influence can indeed be traced down to the 1911-12 lectures. Finally, a further issue that should be investigated in this context, is the possible influence of Hilbert on the publication of Paul and Tatyana Ehrenfest’s famous Encyclopädie article on the conceptual foundations of statistical mechanics (Ehrenfest 1956 [1912]). Ehrenfest studied in Göttingen between 1901 and 1903, and returned there in 1906 for one year, before moving with his wife Tatyana to St. Petersburg. Tatyana had studied mathematics in Göttingen. The idea of writing this Encyclopädie article arised following a seminar

talk in Göttingen, to which Paul Ehrenfest was invited by Felix Klein.88 the Ehrenfests’s style of theory clarification, as manifest in this article, is strikingly reminiscent of Hilbert’s lectures in many respects, and strongly suggests a possible influence of the latter.

Worth of mention are also the critical reactions to Hilbert’s published papers on radiation theory. Such reactions came especially from Ernst Pringsheim (1859-1917),

and they led to a somewhat heated debate between the two.89 This debate is very instructive as to the way in which a physicist could have reacted to Hilbert’s approach to physical issues, and to how Hilbert’s treatment, rather than presenting the systematic and finished structure characteristic of the Grundlagen, was piecewise, ad-hoc and

sometimes confused or unilluminating.90

A further, compelling testimony of Hilbert’s deep interest in physics over this period of time is furnished by the proceedings of a meeting held at the Royal Academy of Sciences in Göttingen in May of 1913. The meeting consisted of a series of lectures on the current state of research in the kinetic theory, and the lecturers included some of the finest physicists of the time. Max Planck, whose work on radiation Hilbert had studied in great detail when writing his own articles, lectured on the significance of the quantum hypothesis for the kinetic theory. Peter Debye (1884-1966) had become in 1914 professor of physics in Göttingen; his talk dealt with the equation of state, the quantum hypothesis and heat conduction. Walther Nernst (1864-1941), whose work on

thermodynamics Hilbert had been following with interest,91 spoke about the kinetic theory of rigid bodies. Marian von Smoluchowski (1872-1917) came from Krakow and lectured on the limits of validity of the second law of thermodynamics. Arnold Sommerfeld (1868-1931) came from Munich to talk about problems of free trajectories. Hendrik A. Lorentz (1853-1928) was invited form Leyden; he spoke on the applications of the kinetic theory to the study of the motion of the electron. That the

87. See Brush 1976, 449-468.88. See Klein 1970, 81-83.89. Pringsheim published his objections in Pringsheim 1913, 1913a. 90. For more details on this debate see Corry 1998.91. In January 1913, Hilbert had lectured on Nernst’s law of heat at the Göttingen Physical Society. The manuscript of the lecture is preserved in Hilbert’s Nachlass, Ms Cod 590. See also a remark added in Hilbert’s handwritting in Hilbet 1905, 167.

36Hilbert and Physics:

meeting was an initiative of Hilbert is clear from the fact that it was sponsored by the Wolfskehlstiftung, whose chair was Hilbert himself. Hilbert wrote a report on the

lectures delivered in the meeting,92 as well as the introduction to the published collection of lectures, in which expressed his hope that it would stimulate further interest, especially among mathematicians, and lead to additional involvement with the

exciting world of ideas created by the new physics of matter.93

7. Electrodynamics and General Relativity: 1913-1915

By the year 1913 Hilbert’s interest in a wide variety of physical disciplines had become a truly central feature of his current research and teaching concerns. In 1912 Hilbert solved the Boltzmann equation, which was intimately connected to his current research in the theory of linear integral equations, but which at the same time raised intriguing physical problems that directly attracted his attention. Hilbert’s student Lüdwig Föppl completed in March 1912 his dissertation on atomic stability (Föppl 1912), and so did Hans Bolza with his own one dealing with the theory of gases. Hilbert also gave public lectures on Maxwell’s theory of gases, statistical mechanics and Nernst’s law of heat. The meetings of the GMG, clearly with Hilbert’s approval, if not under his direct initiative, discussed Gustav Mie’s recent work on an electromagnetic theory of matter (in December 1912) as well as Albert Einstein and Marcel Grossmann’s first attempt to formulate a relativistic theory of gravitation (in

December 1913), the famous Entwurf paper.94

This increased interest in physics is also manifest in Hilbert’s lectures of 1913. The topics of his lectures on physics had expanded way beyond the more traditional ones of classical mechanics and continuum mechanics and now covered also statistical mechanics, radiation theory and the molecular theory of matter. In the summer of 1913 Hilbert returned to an old field of interest that would occupy his thoughts for several years now: electromagnetism. He taught a course on electron theory and, at the same time, a second course on the principles of mathematics, very similar to the earlier one of 1905, where he included again a long section on the axiomatization of physics. Over the following years he would also lecture on electromagnetic oscillations, statistical mechanics, the structure of matter, and, in 1916-17, on the general theory of relativity and the theory of the electron. These years were also characterized by Hilbert’s unqualified adoption of an electromagnetic, reductionist view of physical phenomena. More generally, his views on physical issues during this period became more articulate and, in many respects, much more dogmatic.

Hilbert’s 1913 course on the theory of the electron was conceived as an axiomatic treatment of electrodynamics. Following Minkowski’s work, Hilbert laid special stress on the role of the (special) relativity principle, i.e., on the assumption that all laws of nature are expressible as formulas that are invariant with respect to

92. See JDMV Vol. 22 (1913), pp. 68-69.93. See Planck et al. 1914.94. For more details on the relevant activities of the GMG see the appendix to this article.

37Hilbert and Physics:

simultaneous, homogenous (orthogonal) transformations of the four variables x,y,z,t.95 Hilbert dedicated some effort to highlight the importance of the specific contributions of his Göttingen colleagues: Minkowski’s “World vector analysis”, and Born’s rigid body. The Maxwell equations and the concept of energy, he explained, do not suffice to provide a complete foundation of electrodynamic. An additional concept is thus needed: the concept of rigidity. Electricity must be attached to a steady scaffold. This scaffold is what we call an electron. The electron, he explained, embodies the concept of a rigid body of Hertz’s mechanics. All of the laws of mechanics can be derived, in principle at least, from these three ideas: Maxwell’s equations, the concept of energy, and rigidity. From them also all the forces of physics can be derived, and in particular the molecular forces. Only gravitation, he concluded, has evaded until now every

attempt at an electrodynamic explanation.96

Hilbert also stressed very much the mathematical difficulty involved in solving the n-electron problem. This difficulty, he asserted, provides additional justification for studying the movements of the electron based on the principles of statistical

mechanics.97 It is interesting to notice that Gustav Mie and his electromagnetic theory of matter were not mentioned explicitly in the lectures, despite the fact that back in December 1912 this theory had been discussed in the meeting of the Göttingen Mathematical Society.

Parallel to this course, Hilbert also lectured in the summer of 1913 on the “Elements and Principles of Mathematical Thinking”. A glance at the contents of the course indicates the extent to which Hilbert still considered physics and mathematics as tightly interconnected at their most fundamental and essential aspects, and how he thought that the axiomatic method should be similarly applied for the benefit of both domains. The contents of the course also illustrate, like was the case with the similar 1905 course, how a typical Hilbert lecture could look like: several, not obviously interconnected, mathematical issues could be discussed in successive lectures, in very general terms and without entering into any kind of details. The initial program of the course could always be changed, according to the way in which it developed.

One significant difference between this course and the one taught earlier, in 1905, is the addition of a new section dealing with the axioms of the theory of radiation, which essentially repeated what had appeared in his published work on this topic. Like in the 1905 course, a main motto that appears in Hilbert’s treatment of each

95. Hilbert 1913b, 1 (Emphasis in the original): “Der Inhalt des Relativitätsprinzip ist nun die Behauptung, dass die gesamten Naturgesetze mathematischen Formal ihren Ausdruck finden, die invarianten gegen einen simultanen homogenen (orthogonale) Transformation der vier Variablen x,y,z,t sind.”96. Hilbert 1913b, 61-62 (Emphasis in the original): “Auf die Maxwellschen Gleichungen und den Energiebegirff allein kann man die Elektrodynamik nicht gründen. Es muss noch der Begriff der Starrheit hinzukommen; die Elektrizität muss an ein festes Gerüst angeheftet sein. Dies Gerüst bezeichen wir als Elektron. In ihm ist der Begriff der starrer Verbindung der Hertzschen Mechanik verwirklicht. Aus den Maxwellschen Gleichungen, dem Energiebegriff und dem Starrheitsbegriff lassen sich, im Prinzip wenigstens, die vollständigen Sätze der Mechanik entnehmen, auf sie lassen sich die gesamten Kräfte der Physik, im Besonderen die Molekularkräfte zurückzuführen. Nur die Gravitation hat sich bisher dem Versuch einer elektrodynamischen Erklärung widersetzt.”97. Hilbert 1913b, 83: “Je komplizierte aber das Problem, mit umsomehr Recht wenden wir später das Grundprinzip der statistischen Mechanik auf die Elektronenbewegung an.”

38Hilbert and Physics:

separate field is the need to determine the exact analytic form of a certain function, which lies at the heart of the theory, starting from the axioms alone. The main laws that must be derived here are Kirchhoff´s laws of emission and absorption. We are given an empty closed sphere at a given temperature. Independently from the material composition of the sphere, the law establishes that the radiation energy of a well-determined color depends only on the wave-length and the temperature. The function whose expression must be deduced in this case is the wave function. Hilbert formulated axioms from which he believed the precise analytic expression of the function might be derived as a logical consequence, in a way that in a state of thermal equilibrium, and independently of the kind of matter involved and of its location, the

equation would hold.98 A second innovation found in these lectures, compared to those of 1905, is an elaborate treatment of special relativity or, as Hilbert called it, of the new conceptions of space and time.

In the winter semester of 1913-14 Hilbert taught his last course on a physical issue before eventually turning to general relativity. It dealt with electromagnetic oscillations. In the introductory lecture of this course he referred once again to the example of geometry as a model of an experimental science, our thorough knowledge of which has transformed into a mathematical, and therefore a “theoretical science”. In a passage that characterizes very aptly his conception of the relation between physical disciplines and mathematics, Hilbert said:

From antiquity the discipline of geometry is a part of mathematics. The experimental grounds necessary to build it are so suggestive and generally acknowledged, that from the outset it has immediately appeared as a theoretical science. I believe that the highest glory that such a science can attain is to be assimilated by mathematics, and that theoretical physics is presently on the verge of attaining this glory. This is valid, in the first place for the relativistic mechanics, or four dimensional electrodynamics, of whose belonging to mathematics I am already convinced for a long time.99

But, here, for the first time in Hilbert’s lectures, we come across an explicit suggestion that electrodynamics is the field that might provide the correct foundation for all of physics. In fact, Hilbert seems to have conceived a much more comprehensive, unified picture of science that also covered all of mathematics and physics. In a somewhat unclear passage, Hilbert claimed:

98. Hilbert 1913c, 107: “Die Axiome müssen die Tatsache als logisch Folgerung ergeben, daß die thermischen Gleichgewicht, unabhängig von Stoff und Ort, die Beziehung gilt:

m=f(J, z).”

99. Hilbert 1913-14, 1 (Emphasis in the original): “Seit Alters her ist die Geometrie eine Teildisziplin der Mathematik; die experimentelle Grundlagen, die sie benutzen muss, sind so naheliegend und allgemein anerkannt, dass sie von vornherein und unmittelbar als theoretische Wissenschaft auftrat. Nun glaube ich aber, dass es der höchste Ruhm einer jeden Wissenschaft ist, von der Mathematik assimiliert zu werden, und dass auch die theoretische Physik jetzt im Begriff steht, sich diesen Ruhm zu erwerben. In erster Linie gilt dies von der Relativitätsmechanik oder vierdimensionalen Elektrodynamik, von deren Zuhörigkeit zur Mathematik ich seit langem überzeugt bin.”

39Hilbert and Physics:

It appears, however, as if theoretical physics has finally and totally been absorbed by the electrodynamics, to the extent that every special question should be solved, in the last instance, by appealing to electrodynamics. Following the methods prevailingly used in the individual mathematical disciplines, one could also classify mathematics—more from the point of view of contents than from a formal one—into one-dimensional mathematics (i.e., arithmetic), then function theory (which in essence is limited to two dimensions), geometry, and finally four-dimensional mechanics.100

This more general, unified conception would remain basic to Hilbert’s subsequent dealings with physics, and particularly to his involvement with the problems associated with the general theory of relativity.

While teaching his 1913-14 course, Hilbert’s interest in the work of Gustav Mie must have already become evident. As already mentioned, Max Born lectured before the GMG on this theory in December 1912. On October 22, 1913, Mie wrote a letter to Hilbert expressing his satisfaction for the interest that the latter had manifested on Mie’s recent work (presumably in an earlier letter which has not been

preserved).101 Then, on December 12, 1913, Born lectured again at the meeting of the GMG, this time on his own contributions to Mie’s theory. The reformulation of Mie’s theory at the hands of Born was to become a crucial turning-point in Hilbert’s way to

the problems of general relativity.102 Incidentally, Born’s lecture was communicated to

the society by Hilbert himself,103 At the same time, recent work on gravitation and relativity began to be discussed more intensely in Göttingen. Einstein and Grossmann's Entwurf paper, as already mentioned, was discussed in the GMG on December 12, 1913. But also the works of Max Abraham and Gunnar Nordström were studied with

great interest.104

Hilbert’s physical activities in 1914 were less intense than in previous years. He published the third installment of his work on the foundations of the theory of radiation, and lectured once again on statistical mechanics. Also, in June 1914, his student Kurt Schellenberg presented a dissertation dealing with the applications of integral equations to the theories of electrolysis (Schellenberg 1915). The beginning of the war obviously altered the normal course of activities in Göttingen, and in particular the presence of students and young docents there over the next years. Already in November 3, 1914, Hilbert discussed in the meeting of the GMG the consequences of

war on the society’s activities.105

100. Hilbert 1913-14, 1: “Es scheint indessen, als ob die theoretische Physik schliesslich ganz und gar in der Elektrodynamik aufgeht, insofern jede einzele noch so spezielle Frage in letzter Instanz an die Elektrodynamik appellieren muss. Nach den Methoden, die die einzelnen mathematischen Disziplinen vorwiegend benutzen, könnte man alsdann—mehr inthaltlich als formel—die Mathematik einteilen in die eindimensionale Mathematik, die Arithmetik, ferner in die Funktionentheorie, die sich im wesentlichen auf zwei Dimensionen beschränkt, in die Geometrie, und schliesslich in die vierdimensionale Mechanik.”101. Mie’s letter is in Hilbert’s Nachlass, NSUB Göttingen - Cod Ms David Hilbert 254 - 1.102. For more details see Corry 1998a.103. See the announcement in Jahrb. DMV Vol. 22, p. 207. Born’s work appeared later as Born 1913.104. See Born 1922, 593. For an historical account of Nordström’s work in relativity see Norton 1992.105. See the announcements of the meetings of the GMG for that year in the JDMV Vol. 23 (1914).

40Hilbert and Physics:

In the summer semester of 1915 Hilbert lectured on the structure of matter. The notes of this course illuminate yet another significant aspect of Hilbert’s wide range of physical knowledge and interests. The course focused, in fact, on Born’s theory of crystals, which was based on the study of the potential energy of a lattice of particles. For Hilbert, this theory, in the one hand, and the theory of dilute gases, on the other hand, were complementary to each other in accounting for the properties of matter (Hilbert 1915, 1). Hilbert discussed briefly the mathematical aspects of the theory (crystallographic groups), and also, in much more detail, its physical aspects: wave displacement inside the lattice, crystal elasticity, specific heat in the lattice, the piezo-electric effect, etc.

The announcement of meetings at the GMG for the summer semester of 1915 record a lecture on structure of crystals by Felix Klein (together with Hilbert and Mügge) in May, and a lecture by Sommerfeld “On Modern Physics” in June. Then from June 29 to July 7, 1915, Einstein gave a series of lectures on the current state of his research on gravitation and relativity. Unfortunately not much is known about the visit itself, except that, as I mentioned in the introduction to this article, Einstein felt

that his work had been understood to the details.106 Over the months following that visit, and especially in October and November, Hilbert devoted most of his efforts to what he later called “The Foundations of Physics”, namely, the formulation of a

unified field theory, based on Mie’s electromagnetic theory of matter.107 He presented the results of his investigations in the meeting of the Göttingen Scientific Society on

November 20, 1915.108

Hilbert’s work during these months and the publication of his field equations of gravitation opened a new phase in his career, during which he and his colleagues in

Göttingen would dedicate much efforts to the general theory of relativity.109 From that time on, Hilbert became increasingly enthusiastic about the significance of general covariance. In a lecture held in 1921, for instance, he asserted that no other discovery in history had aroused so much interest and excitement as Einstein’s relativity theory (“the highest achievement of human spirit”) did. This excitement was indeed justified in Hilbert’s view since, whereas all former laws of physics were provisory, inexact and special, the principle of relativity (and Hilbert meant by this the general covariance of physical laws) signified “for the first time, since the world exists, a definitive, exact

and general expression of the nature laws that hold in reality.”110 At the same time, Hilbert’s involvement with general relativity provided additional support to his

106. I have made some efforts to gather documents related to this visit, so far without much success. What I did find in Hilbert’s Nachlass in Göttingen, nevertheless, are the handwritten notes taken from the first of Einstein’s lectures (Staats- und Universitätsbibliothek Göttingen, Cod Ms D Hilbert 724). These notes have been published meanwhile in Kox et al (eds.) 1996, App. B, 586-590.107. For an account of this period and the details of Hilbert’s theory, see Corry 1998a.108. They were published in March of the following year as Hilbert 1916.109. In Corry 1998, 1998a, I give detailed historical accounts of Hilbert’s way of general relatvity, as well as the more immediate context of his publications. David Rowe’s contribution to the present volume discusses the developments in Göttingen around general relativity between 1916 and 1918. 110. Hilbert 1921, 1: “... denn das Relativitätsprinzip bedeutet, wie mir scheint, zum ersten Mal, seit die Welt steht, eine definitive, genaue und allgemeine Aussage über die in der Wirklichkeit geltenden Naturgesetze.”

41Hilbert and Physics:

empiricist conception of geometry. Although a detailed analysis of the place of geometry in Hilbert’s work on relativity would be much beyond the scope of the present account, it is pertinent to quote here from the manuscript of his course on this issue, taught in the winter semester of 1916-17, where Hilbert expressed this connection as part of a more general, unified view of all branches of human knowledge. Hilbert thus said:

In the past, physics adopted the conclusions of geometry without further ado. This was justified insofar as not only the rough, but also the finest physical facts confirmed those conclusions. This was also the case when Gauss measured the sum of angles in a triangle and found that it equals two right ones. That is no longer the case for the new physics. Modern physics must draw geometry into the realm of its investigations. This is logical and natural: every science grows like a tree, of which not only the branches continually expand, but also the roots penetrate deeper.

Some decades ago one could observe a similar development in mathematics. A theorem was considered according to Weierstrass to have been proved if it could be reduced to relations among integer numbers, whose laws were assumed to be given. Any further dealings with the latter were laid aside and entrusted to the philosophers. Kronecker said once: ‘The good Lord created the integer numbers.’ These were at that time a touch-me-not (noli me tangere) of mathematics. That was the case until the logical foundations of this science began to stagger. The integer numbers turned then into one of the most fruitful research domains of mathematics, and especially of set theory (Dedekind). The mathematician was thus compelled to become a philosopher, for otherwise he ceased to be a mathematician.

The same happens now: the physicist must become a geometer, for otherwise he runs the risk of ceasing to be a physicist and viceversa. The separation of the sciences into professions and faculties is an anthropological one, and it is thus foreign to reality as such. For a natural phenomenon does not ask about itself whether it is the business of a mathematician or of a physicist. On these grounds we should not be allowed to simply accept the axioms of geometry. The latter may be the expression of certain facts of experience that further experiments would contradict.111

111. Hilbert 1916-17, 2-3 (Emphasis in the original): “Früher übernahm die Physik die Lehren der Geometrie ohne weiteres. Dies war berechtigt, solange nicht nur die groben, sondern auch die feinsten physikalischen Tatsachen die Lehren der Geometrie bestätigen. Dies war noch der Fall, als Gauss die Winkelsumme im Dreieck experimentell mass und fand, dass sie zwei Rechte beträgt. Dies gilt aber nicht mehr von der neuesten Physik. Die heutige Physik muss vielmehr die Geometrie mit in den Bereich ihrer Untersuchungen ziehen. Das ist logish und naturgemäss: jede Wissenschaft wächst wie ein Baum, nicht nur die Zweige greifen weiter aus, sondern auch die Wurzeln dringen teifer.

Vor einigen Jahrzehnten konnte man in der Mathematik eine analoge Entwicklung verfolgen; einen Satz hielt man damals nach Weierstrass dann für bewiesen, wenn er auf Beziehungen zwischen ganzen Zahlen zurückführbar war, deren Gesetz man als gegeben hinnahm. Sich mit diesen zu beschäftigen, wurde abgelehnt und den Philosophen überlassen. Kronecker sagte einmal: ‘Die ganzen Zahlen hat der liebe Gott geschaffen.’ Diese waren damals noch einen noli me tangere der Mathematik. Das ging so fort, bis die logischen Fundamente dieser Wissenschaft selbst zu wanken begannen. Nun wurden die ganzen Zahlen eines der fruchtbarsten Arbeitfelder der Mathematik uns speziell der Mengenlehre (Dedekind). Der Mathematiker wurde also gezwungen, Philosoph zu werden, weil er sonst aufhörte, Mathematiker zu sein.So ist es auch jetzt wieder: der Physiker muss Geometer werden, weil sonst Gefahr läuft, aufzuhören, Physiker zu sein und umgekehrt. Die Trennung der Wissenschaften in Fächer und Fakultäten ist eben etwas Antropologisches, und der Wirklichkeit Fremdes; denn eine Naturerscheinung frägt nicht danach, ob sie es mit einem Physiker oder mit einem Mathematiker zu tun hat. Aus diesem Grunde dürfen wir die Axiome der Geometrie nicht übernehmen. Darin könnten ja Erfahrungen zum Ausdruck kommen, die den ferneren Experimenten widersprächen.”

42Hilbert and Physics:

But as was the case with his earlier contributions to physics, Hilbert’s work in general relativity did not attain wide recognition. Einstein himself, like many other physicist, disliked both Hilbert’s derivation of the equations from a variational principle and his excessive reliance on Mie’s theory of matter. Nevertheless, seen in the context of his other contributions to physics, there can be no doubt that, like in other disciplines, Hilbert’s involvement in general relativity was far from being a short-range, opportunistically motivated endeavor.

8. Concluding Remarks

Hilbert had a sustained interest in physics that can be traced throughout his career. In the present article I have shown how this interest is manifest from his early dealings with geometry, around 1894, and up until 1915. This interest, however, continued to occupy a central place in Hilbert’s overall conception of science until the end of his career. Although Hilbert’s published work on physical issues covers only a small portion of his overall output, he actually dedicated a much more significant part of his efforts to various physical disciplines than the amount of his publications could indicate. This effort can properly be understood only by examining Hilbert’s docent and organizational activities in Göttingen, as I have done in the preceding pages.

In order to determine whether Hilbert’s ideas had any actual influence on the development of twentieth century physics it is necessary to undertake a more detailed examination of his published and unpublished writings in each individual discipline, and of the people who were exposed to these ideas. I have suggested above that this influence may have been instrumental, in one way or another, in shaping the ideas of several persons who contributed significant work in physics: Max Born, Hermann Minkowski, Paul Ehrenfest, David Enskog, as well as other, relatively minor figures. Other names come to mind which were not mentioned in this article but which certainly were influenced by Hilbert in different domains of physics, such as Hermann Weyl in relativity and John von Neumann in quantum theory.

Many physicist reacted with lack of enthusiasm, to say the least, to Hilbert’s incursions into physics. Pringsheim’s reaction to Hilbert’s articles on radiation offers a good example of this. Also Einstein expressed many reservations concerning Hilbert’s approach to relativity. In a letter to Hermann Weyl, Einstein criticized Hilbert’s use of the variational principle in this context and judged his approach to the general theory of relativity to be “childish ... in the sense of a child that recognizes no malice in the

external world.”112 Weyl himself considered that Hilbert’s work in physics was of rather limited value, especially when compared to his work in pure mathematics. In particular Weyl considered that Hilbert’s application of the axiomatic method was of little significance for physics. Weyl thought, that a valuable contribution to physics required a different kind of skills than those in which Hilbert excelled. In an obituary of Hilbert, Weyl wrote:

112. In a letter of November 23, 1916. Quoted in Seelig 1954, 200.

43Hilbert and Physics:

The maze of experimental facts which the physicist has to take in account is too manifold, their expansion too fast, and their aspect and relative weight too changeable for the axiomatic method to find a firm enough foothold, except in the thoroughly consolidated parts of our physical knowledge. Men like Einstein and Niels Bohr grope their way in the dark toward their conceptions of general relativity or atomic structure by another type of experience and imagination than those of the mathematician, although no doubt mathematics is an essential ingredient.113

Max Born was perhaps the physicist that expressed a more consistent enthusiasm for Hilbert’s physics. He seems also to have truly appreciated the exact nature of Hilbert’s program for axiomatizing physical theories and the potential contribution that the realization of that program could bear. On the occasion of Hilbert’s sixtieth birthday, the journal Die Naturwissenschaften dedicated one of its issues to celebrate the achievements of the master. Several of his students were commissioned with articles summarizing Hilbert’s contributions in different fields. Born, who as a young student in Göttingen attended many of Hilbert’s courses, and later on as a colleague continued to participate in his seminars, wrote about Hilbert’s physics. Besides his extolling summary of Hilbert’s achievements, Born also explained why in his view Pringsheim had misunderstood Hilbert and why the former’s reproaches to the latter were unjustified. In doing so he also clarified, in a very succinct formulation, the nature of Hilbert’s axiomatic treatment and why, in general, physicist tended not to appreciate it. Born put it in the following words:

The physicist set outs to explore how things are in nature; experiment and theory are thus for him only a means to attain an aim. Conscious of the infinite complexities of the phenomena with which he is confronted in every experiment, he resists the idea of considering a theory as something definitive. He therefore abhors the word “Axiom”, which in its usual usage evokes the idea of definitive truth. The physicist is thus acting in accordance with his healthy instinct, that dogmatism is the worst enemy of natural science. The mathematician, on the contrary, has no business with factual phenomena, but rather with logic interrelations. In Hilbert’s language the axiomatic treatment of a discipline implies in no sense a definitive formulation of specific axioms as eternal truths, but rather the following methodological demand: specify the assumptions at the beginning of your deliberation, stop for a moment and investigate whether or not these assumptions are partly superfluous or contradict each other. (Born 1922, 591)

In fact Hilbert never performed for a physical theory exactly the same kind of axiomatic analysis he had done for geometry, though he very often declared this to be the case. In no case in the framework of his lectures, did Hilbert actually prove the independence, consistency or completeness of the axiomatic systems he introduced. In certain cases, like vector addition, he quoted works in which such proofs could be found. In other cases there were no such works to mention, and Hilbert limited himself—as in the case of thermodynamics—to state that his axioms are indeed independent. Still in other cases, he barely mentioned anything about independence or other properties of his axioms. Also his derivations of the basic laws of the various disciplines from the axioms are rather sketchy, when they appear at all. Many times Hilbert simply declared that such a derivation was possible. Among his published works, his last article on radiation theory contains—perhaps under the pressure of criticism—his more detailed attempt to prove independence and consistency of a system of axioms for a physical theory. But what is clear in all cases is that Hilbert always considered that an axiomatization along the lines he suggested was plausible

113. Quoted in Sigurdsson 1994, 363.

44Hilbert and Physics:

and could eventually be fully performed following the standards established in the Grundlagen. Be that as it may, there can be no doubt that the kind of conceptual clarification attained in Hilbert’s examination of physical theories, as well as in works of those who followed his lead, provided an important contribution in and of itself.

Whether or not physicists should have looked more closely at Hilbert’s ideas than they actually did, and whether or not Hilbert’s program for the axiomatization of physics had any influence on subsequent developments in this discipline, it is nevertheless important to stress, that a full picture of Hilbert’s own conception of mathematics cannot be complete without taking into account his views on physical issues and on the relationship between mathematics and physics. More specifically, a proper understanding of Hilbert’s conception of the role of the axioms in physical theories—a conception condensed in the above quoted passage of Born, and illustrated throughout this article—helps us understanding his conception of the role of axioms in mathematical theories as well. The picture that arises from such an understanding is obviously very far away from the somewhat widespread image of Hilbert as the champion of a formalistic conception of the nature of mathematics.

Acknowledgments

The research and archival work that preceded the writing of this article was conducted as part of a larger project on the historical context of the rise of the general theory of relativity at the Max-Planck-Institut für Wissenschaftsgeschichte in Berlin, where I worked during the academic year 1994-95. I would like to thank the staff for their warm hospitality and diligent cooperation, and especially to Jürgen Renn for inviting me to participate in the project and for his constant encouragement. During my stay in Berlin, I benefited very much from illuminating discussions on the history of modern physics with Jürgen Renn, Tilman Sauer and John Stachel.

This article was partly written at the Dibner Institute during the academic year 1995-96. I wish to thank the Directors and staff of the DI, as well as the other fellows with whom I was fortunate to share my time. Special thanks to Ulrich Majer and David Rowe for long and interesting discussions on the issues analyzed in this article.

I also thank Jeremy Gray and the participants of the meeting on Geometry and Physics, 1900-1930, at the Open University, Milton Keynes (March 14-19, 1996) for their illuminating lectures and for their fruitful remarks and criticism following my own talks.

Arne Schirrmacher read an earlier version of this article and provided helpful corrections and suggestions for improvement. I thank him very much.

Original manuscripts are quoted in the text by permission of the Staats- und Universitätsbibliothek Göttingen (Handschriftenabteilung), the library of the Mathematisches Institut Universität Göttingen. I thank Ralf Haubrich and the Hilbert Edition staff, for allowing me to quote the motto of this article from their provisory typescript draft of the Tagebuch. Special thanks I owe to Peter Damerow for allowing me to read and quote from the manuscript of Hilbert 1913c, belonging to his private collection.

45Hilbert and Physics:

Appendix

All the lists included in this appendix have been compiled from various sources.Among them: the manuscripts of the lecture notes in the Lesezimmer of themathematical institute in Göttingen; Hilbert’s Nachlass (Niedersächsiche Staats- undUniversitätsbibliothek Göttingen); announcements in the Jahresbericht der DeutschenMathematiker Vereinigung and in the Physikalische Zeitschrift; Wilhelm Lorey’s bookquoted in the bibliography. It is conceivable that additional sources may add furtheritems to each of these lists.

Table 1:

Hilbert’s Courses on Physics (1898-1927)

1. Mechanics SS 18982. Lectures on Potential Theory WS 1901/023. Selected Chapters of Potential Theory SS 19024. Continuum Mechanics - Part I WS 1902-035. Continuum Mechanics - Part II SS 19036. Exercises on Mechanics - WS 1904/057. Logical Principles of Mathematical Thinking (and of Physics) SS 19058. Mechanics WS 1905/069. Continuum Mechanics SS 190610. Lectures on Continuum Mechanics WS 1906/0711. Mechanics WS 1910/1112. Continuum Mechanics SS 191113. Statistical Mechanics WS 1911/1214. Radiation Theory SS 191215. Molecular Theory of Matter WS 1912/1316. Foundations of Mathematics (and the axiomatizaion of Physics) SS 1913 17. Electron Theory SS 191318. Electromagnetic Oscillations WS 1913/1419. Statistical Mechanics SS 191420. Lectures on the Structure of Matter SS 191521. The Foundations of Physics I (General Relativity) SS 191622. The Foundations of Physics II (General Relativity) WS 1916/1723. Electron Theory SS 191724. Space and Time WS 1917/1825. Space and Time WS 1918/1926. Mechanics and the New Theory of Gravitation SS 1920 27. Basic Principles of the Theory of Relativity SS 192128. Statistical Mechanics SS 1922

46Hilbert and Physics:

29. Mathematical Foundations of Quantum Theory WS 1922/2330. On the Unity of Science WS 1923-2431. Mechanics and Relativity Theory SS 192432. Mathematical Methods of Quantum Theory WS 1926/27

Table 2:Hilbert’s Doctoral Students on Physical Issues

1. Lüdwig Föppl: “Stabile Anordnungen von Elektronen im Atom” (March 1, 1912)2. Hans Bolza: “Anwendungen der Theorie der Integralrechnungen auf die

Elektronentheorie und die Theorie der verdünnten Gasen.” (July 2, 1913)3. Bernhard Baule: “Theoretische Behandlung der Erscheinungen in verdünnten

Gasen.” (Feb. 18, 1914)4. Kurt Schelenberg: “Anwendungen der Integralgleichung auf die Theories der

Elektrolysie.” (June 24, 1914)

Table 3:

Hilbert’s Seminars and Public Lectures on Physics

A. Seminars:1. Mechanics (together with Klein) 1896 (?) 2. Mechanics WS 1904-053. Electron Theory (together with H. Minkowski et al) SS 19054. The Equations of Electrodynamics (with Minkowski) SS 19075. Hydrodynamics ??? 6. Electrodynamics ???7. Kinetic Theory of Gases 1912 (?)

B. Public Lectures:1. Maxwell’s Theory of Gases 1912 (?)2. Statistical Mechanics 19123. On Nernst’s Law of Heat (G_ttingen) 19134. Space and Time (Bucharest) 19185. Nature and Mathematical Knowledge (G_ttingen) 1919-20

47Hilbert and Physics:

6. On the Laws of Chance 19207. Nature and Mathematical Knowledge (Copenhagen) 19218. Science and Mathematical Thinking (G_ttingen) 1922-23

48Hilbert and Physics:

Table 4:

Lectures on Physical Issues delivered at the meetings of the Mathematical Society in Göttingen: 1904-1918

(According to the announcements in the Jahrb. DMV)

A. Hilbert’s Lectures:1. Continuum Mechanics Feb. 24, 19032. Continuum Mechanics Aug. 4, 19033. The relations between variational principles and the theory of partial

differential equations, with applications to the integral principles of mechanics. Jan. 18, 1910

4. Kinetic theory of gases Dec 19, 19115. Axiomatic Foundations of Physics (Ferienkurs for high-school

teachers) April 15-27, 19126. Theory of Radiation July 30, 19127. Theory of Radiation Jan. 21, 19138. Theory of Radiation July 28, 19149. The Fundamental Equations of Physics (General Relativity) Nov. 16, 191510. Theory of Invariants and the Energy Principle Jan. 25, 191611. The Causality Principle in Physics Nov. 21 & 28, 191612. Non-Euclidean geometry and the new Gravitation Theory Jan. 23, 191713. Laue’s Theorem June 12, 191714. Reply to Klein’s “On Hilbert’s first note on the

Foundations of Physics” Jan. 29, 191815. The Energy Principle for the Motion of Planets in the New

Theory of Gravitation June 4, 191816. On Weyl’s Communication to the Berlin Academy (May 2, 1918)

“The Energy Principle in the General Theory of Relativity” July 15, 1918

B. Lectures by Others:1. On the Axioms of Vector Addition (R. Schimmack) June 9, 19032. Molecular Theory of Heat Conduction (G. Prasad) June 9, 19033. Capillarity (H. Minkowski) June 23, 19034. Linear Heat Conduction in Surfaces (G. Prasad) June 23, 19035. Maxwell’s Work on Stress Systems (F. Klein) June 23, 19036. Euler’s Equations of Hydrodynamics (H. Minkowski) June 28, 19037. Electromagnetic Quatity of Motion (M. Abraham) July 14, 19038. Gibb’s Thermodynamical Surfaces (H. Happel) Dec. 8, 1903

49Hilbert and Physics:

9. Variational Principles in Electrodynamics (K. Schwarzschild) Jan. 26, 190410. Can the Electron Reach the Speed of Light (P. Hertz) Jan. 26, 190411. Overview of a Seminar on Hydrodynamics and Hydraulics (F. Klein) Feb. 9, 190412. Motion of a Material Particle on a Uniformly Moving

Plane (P. Ceresole) May 17, 190413. On the Elasticity of the Earth (G. Herglotz) June 28, 190414. On Sommerfeld’s Works on Electron Theory (G. Herglotz) Dec. 6, 190415. Motion of a Fluid with Little Friction (L. Prandtl) Dec. 13, 190416. On a Talk by Poincaré on the Future of Mathematical

Physics (C.H. Müller) Jan. 24, 190517. On Gases ans Vapors (L. Prandtl) May 23, 190518. On Poincarés Published Lectures on Mathematical

Physics (M. Abraham) Feb. 6, 190619. On Poincarés Investigations on Rotating Fluid Masses (H. Müller) Feb. 13, 190620. On Gibb’s Book on Statistical Mechanics (E. Zermelo) Feb. 20, 190621. Graphical Methods in Mechanics and Physics (C. Runge) Feb. 27, 190622. On Painlevé’s Work on the Foundations of Mechanics

(C. Carathéodory) May 28, 190623. On W. Nernst’s “On Chemical Equilibrium” (H. Minkowski) June 26, 190624. Problems of Aeronavigation (L. Prandtl & E. Wiechert) Oct. 30, 190625. The Mathematical Theory of Elasticity (C.H. Müller) Nov. 6, 190626. On Botzmann’s H-Theorem (P. Ehrenfest) Nov. 13, 190627. The Evolution of the Theory of Radiation through the Works

of Lorentz, Rayleigh, W. Wien and Planck (H. Minkowski) Dec 1, 190628. On H. Witte’s “On the Possiility of a Mechanical Explanation

of Electromagnetic Phenomena” (M. Abraham) Dec. 18, 190629. On the Application of Probability Calculus to Astronomy

(K. Schwarzschild) Jan. 8, 190730. Theories of the Effects of Air Resistance (L. Prandtl) Jan. 22, 190731. Seismic Waves (E. Wiechert) Jan. 29, 190732. Statistical Stellar Astronomy (K. Schwarzschild) Feb. 19, 190733. Seismic Rays (G. Herglotz) May 14, 190734. Solutions of Differential Equations for Gas Spheres (Gaskügeln)

(K. Schwarzschild) July 30, 190735. On the Equations of Electrodynamics (H. Minkowski) Nov. 5, 190736. Graphical Methods in Fluid Mechanics (C. Runge) Nov. 26, 190737. Applications of Quaternions to Electron Theory (F. Klein) Dec. 10, 190738. A New, Simple General Proof of the Second Law of

Thermodynamics (C. Carathéodory) Dec. 17, 190739. An Overview of Man’s Attempts to Fly (C. Runge) March 3, 190840. Report on a Joint Seminar on Hydrodynamics (F. Klein,

L. Prandtl, C. Runge, E. Wiechert) May 5, 190841. An Experiment on Stabilization of Air Balloons (L. Prandtl) May 12, 1908

50Hilbert and Physics:

42. On Lanchester’s Book “Aerodynamics” (C. Runge) May 12, 190843. On the Equations of Electrodynamics (H. Minkowski) July 28, 190844. On Recent French Research on Aviation (C. Runge) Nov. 3, 190845. Recent Works on Earth Pressure (Th. Van Kármán) Nov. 24, 190846. Theory of Earth Pressure (A. Haar & Th. Van Kármán) Dec. 8, 190847. Position Determination from and Air Balloon (C. Runge) May 11 & 18, 190948. Defintion of a Rigid Body on the “Einstein-Minkowski”

System of Electrodynamics (M. Born) June 8 & 15, 190949. Average Motion in the Theory of Perturbations and

Applications of Probability to Astronomy (F. Bernstein) June 22, 190950. On Minkowski’s Nachlass (Electrodynamics) (M. Born) Feb. 8, 1910 51. On the Definition of a Rigid Body (M. Born) Juni 21, 191052. Stable Orderings of Electrons in the Atom

(L. Föppl - PhD Dissertation supervised by Hilbert) Nov. 21, 191153. On Herglotz Work on Deformable Bodies in the

Theory of Relativity (M. Born) Dec. 12, 191154. The Behavior of Solid Bodies and Hooke’s Law (L. Prandtl) Jan. 16, 191255. A Newly Discovered Relation Between Elasticity of Crystals

and Optical Oscillations (M. Born & Th. van Kármán) Feb. 13, 191256. Mollecular Oscillations and Specific Heat (Born & van Kármán) May 14, 191257. Theory of Disperssion in Crystals (P.P. Ewald - PhD Dissertation) June 4, 191258. New Works of Poincaré end Ehrenfest on the Axiomatic

Foundation of Quantum Theory (Th. van Kármán) Juli 16, 191259. Statistical Mechanics (P. Hertz) Nov. 26, 191260. On Sommerfeld’s Article on the Theory of

Oscillating Equations (H. Weyl) Dec. 10, 191261. Mie´s Theory of Matter (M. Born) Dec. 17, 191262. Motion of Fluids (L. Prandtl) Feb. 4, 191363. Reports on the Solvay Conference, Brussels 1911

(Born & van Kármán) Feb. 25 & March 4, 191364. An Application of Diophantine Approximations to a

Question in Statistical Mechanics (E. Hecke) May 20, 191365. On the Structure of Crystals (M. Born) June 7, 191366. Recent Work of J.J. Thomson on Canal Waves (Kanalstrahlen)

(C. Runge) Juni 24, 191367. An Application of Quantum Theory to Capillarity

(M. Born & R. Courant) July 1, 191368. On a Recent Work of E. Noether on Turbulences in a Fluid

(Th. van Kármán) July 30, 191369. On Poincaré’s Book on Cosmogonic Hypotheses (L. Föppl) July 30, 191370. Propagation of Light in Transparent Media (W. Behrens) Nov. 4, 191371. On Mie’s theory of Matter (M. Born) Nov. 25, 191372. The Solution of an Integral Equation of Spectroscopy (C. Runge) Dec. 2, 1913

51Hilbert and Physics:

73. Recent Work of Einstein and Grossmann on Gravitation (F. Böhm) Dec. 9, 191374. On Mie’s Theory of Matter (M. Born) Dec. 16, 191375. Theoretical Treatment of Phenomena in Diluted Gases

(B. Baule - PHD Dissertation Supervised by Hilbert) Feb 24, 191476. Review of Recently Published Works by von Smoluchowski

(Brownian Movement), and Einstein (On Light Deflection; On the Determination of Molecular Dimensions) (P. Hertz) Feb. 24, 1914

77. Lattice Theory of Diamonds (M. Born) March 3, 191478. Intensity Distribution in Spectral Lines (P. Debye) Dec. 18, 191479. Foundation and Problems of Quantum Theory (P. Debye) Feb. 23, 191580. Dynamics of Crystal Lattices (M. Born) Feb. 25, 191581. Structure of Crystals (F. Klein, with Hilbert and Mügge) May 18, 191582. On Herglotz’s Research on Potentials in the Interior of

Attracting Masses (Wiarda) June 1, 191583. On Modern Physics (A. Sommerfeld) June 15, 191584. On Gravitation (A. Einstein) June 29, 191585. Theory of Distant Forces (Uhlich-Pirna) July 20, 191586. History of Mechanics up until Galileo (C. Müller) March 2, 191687. Four-dimensional Vectorial Analysis (C. Runge) Dec. 5, 191688. Foundations of a Theory of Matter (G. Mie) June 5-8, 191789. On the Riemannian Curvature (F. Klein) Oct. 30, 191790. On the Riemannian Curvature (F. Klein) Nov. 6, 191791. On G. Herglotz’s Paper on Curvature and Gravitation (F. Klein) Dec. 4, 191792. On Liquid Crystals (M. Born) Dec. 11, 191793. On Invariants of Arbitrary Differential Expressions (E. Noether) Jan. 15, 191894. On Hilbert’s First Note on the Foundations of Physics (F. Klein) Jan. 22, 191895. On Einstein’s Cosmological Ideas of 1917 (F. Klein) May 7, 191896. Four Lectures on Quantum Theory (M. Planck) May 13-17, 191897. On Einstein’s “On Gravitational Waves” (C. Runge) Jun. 31, 191898. On Einstein’s Cosmological Ideas of 1917 (F. Klein) June 11, 191899. On the Three-body Problem (C. Carath_odory) June 24, 1918100. On Einstein’s “Energy Principle in General Relatitvity” (F. Klein) July 4, 1918101. Hilbert’s Energy Vector (F. Klein) July 22, 1918102. Invariant Variational Problems (E. Noether) July 23, 1918