Skip to main content
Log in

From Gauss to Riemann Through Jacobi: Interactions Between the Epistemologies of Geometry and Mechanics?

  • Article
  • Published:
Journal for General Philosophy of Science Aims and scope Submit manuscript

Abstract

The aim of this paper is to argue that there existed relevant interactions between mechanics and geometry during the first half of the nineteenth century, following a path that goes from Gauss to Riemann through Jacobi. By presenting a rich historical context we hope to throw light on the philosophical change of epistemological categories applied by these authors to the fundamental principles of both disciplines (which they came to regard as hypotheses or conventions). We intend to show that presentations of the changing status of the principles of mechanics as a mere epiphenomenon of the emergence of non-Euclidean geometries are inaccurate, that the relations between the two disciplines were richer than what is usually considered in the literature. These claims will be based on historical and philosophical arguments, starting from the fact that disciplinary boundaries at the time were not rigid as we are used today. It is widely known that the main figures we target worked in different areas, which is a first piece of evidence for the plausibility of our main thesis.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. We are thinking of cases such as those of Gauss, Plücker, or Kummer in Germany, Legendre, Cauchy, or Liouville in France.

  2. It is in these earlier lectures where Jacobi introduced the analytical formulation of mechanics, usually known as the Hamilton–Jacobi formalism.

  3. This has happened in particular when considering the use of conventions by Poincaré: many interpreters have seen the attribution of the notion of convention to the principles of mechanics and physics as a mere extension of his famous geometric conventionalism. Examples of this view are Friedman (1999), Giedymin (1982), or Zahar (2001). However, as some authors have shown recently (Pulte 2000; de Paz 2018), and as we have quoted above, Poincaré is not the first or the only one to use the term ‘convention’ to characterize the principles of mechanics.

  4. On this topic see especially the rich paper Lützen (1995), to which a referee directed our attention. The works of this same author on Liouville and on Hertz are likewise directly relevant.

  5. We should like to mention another study which is also cited in Lützen’s paper. Ziegler (1985) discusses another area of interaction between geometry and mechanics, the geometrical study of the motion of rigid bodies on the basis of methods belonging to projective geometry and line geometry. This was the work of geometers and practitioners of mechanics who were not satisfied with the ‘rarefied’, strongly analytical orientation of Lagrange, Hamilton and Jacobi.

  6. There is also a letter from Jacobi to Gauss where he writes about following considerations from the Disquisitiones Aritmethicae (Jacobi to Gauss, 27.10.1826, in: Jacobi 1881–1891, Vol. 7, 391–392).

  7. Jacobi arrived at Kögnisberg as a recently appointed Privatdozent but by the end of 1827 he was appointed ‘extraordinary’ professor and in 1832 ‘ordinary’ professor. Details about Jacobi’s life and career can be found in Ahrens (1904), Königsberger (1904) and Dirichlet (1852).

  8. In a letter to the Prussian governor, Theodor von Schön, he criticized the emphasis of the philologists on classical education and stated: “the Greeks could learn [...] a hundred times more from us than we could from them. I mean in the great kingdom of truth, mathematics, and in the equally great kingdom of observation, nature” (quoted in Olesko 1991, 52). With these claims, he wanted to support the reform of Prussian education that other professors were asking for in Königsberg.

  9. He studied in Paris, being a member of Fourier’s circle, and brought French mathematical physics to Germany. At Berlin he could not become a full professor because he lacked the required Latin skills. The close friendship of Jacobi and Dirichlet is expressed in a letter of Rebecca, Dirichlet’s wife, to her nephew after Jacobi’s death: “Jacobi died in the night from Tuesday to Wednesday […] His relationship to Dirichlet was so very sweet, the way they sat together for hours, I called it mathematical silence, and the way they certainly did not treat each other gently, and Dirichlet often uttered the bitterest truths to him, and Jacobi understood so well and his great spirit knew to bow before Dirichlet’s great character” (quoted in Bergman et al. 2012, 49).

  10. See the interesting quote from Eisenstein’s autobiography in Wussing (1984). The very young Eisenstein indicated in 1843 that the “new school founded by Gauss, Jacobi and Dirichlet” avoids “long and involved calculation [...] by the use of a brilliant expedient: it comprehends a whole area in a single main idea, and in one stroke presents the final result with utmost elegance” (quoted in Wussing 1984, 270, fn. 81).

  11. This can be seen as the year of his great triumph with the publication of the ‘Theorie der Abelschen Functionen’, and which also brought a small improvement in his salary and working conditions.

  12. Riemann worked on the unification of light and electrodynamics, see Jungnickel et al. (1986, 176–177); one can mention early papers on mathematical physics, namely on Kohlrausch’s measurements, and on Nobili rings (see Archibald 1991). Very relevant too is the paper handed in 1858, but left unpublished: ‘Ein Beitrag zur Electrodynamik’; Riemann wrote: “My discovery of a connection between electricity and light I have submitted here to the Royal Society [the Göttingen Königliche Gesellschaft]” (see Ferreirós 2006a).

  13. According to Betti, whom Riemann told this in conversation, the idea of employing dissections for characterizing the “order of connection” of a surface (closely linked to the Euler characteristic) came to Riemann in the course of a conversation with Gauss about a topic in physics (cf. Weil 1979). Also Listing was inspired by the links Gauss saw between circumstances affecting physical experiments and topological notions. We know nothing about the influence Listing himself may have had upon Riemann.

  14. From the Akten Alfred Clebsch, 1866 writing to von Warnstedt; quoted in Ferreirós (2006a, 67).

  15. Papers from 1829 and 1833 cited in Jungnickel et al. (1986, 49).

  16. Sartorius von Waltershausen (1856), Klein (1926, 6, 18), Dunnington (2004).

  17. The whole passage in this work is quite interesting; p. 81 ends by saying: “According to his intimate conviction [innersten Ansicht], expressed frequently, Gauss regarded the three dimensions of space as a specific characteristic of the human soul; and people who could not understand this, he called in his humorous vein, Boeotians” (von Waltershausen 1856, 81). Compare with Lotze’s views, mentioned in a later footnote (and see Torretti 1978, Sec. 4.2).

  18. As he defined it, topology studies spatial relations according to their “modality”, and not their “quantity” (Listing 1847, 3); such “modal relations” have to do with spatial domains, their relative position and arrangement.

  19. On Gauss and his role in the mathematization of the study of knots, see the work of Epple (1999).

  20. This is not the place to enter into the well-known question of Gauss’s reaction to Bolyai and the way the young Hungarian officer felt deprived of his contribution. See e.g. Gray (2004, 54–55); also Gray (1989) and almost any text on this history.

  21. By way of comparison, the contemporary textbook of Georg Carl Justus Ulrich, professor at Göttingen (Lehrbuch der reinen Mathematik, 1836), does not even mention the question of parallels—which is remarkable since the previous generation of professors (Kästner, Klügel) had done quite a lot to revive it.

  22. There is a tendency to forget how important Lambert was in the eighteenth century; see his Theorie der Parallellinien, 1786, and the interesting webpage http://www.kuttaka.org/~JHL/Main.html (Collected Works - Sämtliche Werke Online). For his influence on Gauss, see Abardia et al. (2012), and Rodríguez (2006).

  23. An example from Riemann’s letters to his brother is given by Dedekind (1876, 521): Riemann speaks about the “interconnection between electricity, galvanism, light and gravity”, emphasizing the progress he has made on it; but he is worried “that Gauss has been working on the same topic for years, and has told some friends about the matter, e.g. Weber, under the pledge of secrecy”. In this case his fears were unfounded, since Gauss had nothing similar to his ideas on the question.

  24. Perhaps other members of the Philosophy Faculty, such as the prominent philosopher R. H. Lotze, could have been present. As Dedekind tells us (1876, 517), Riemann made great efforts to present his ideas so that they could be understandable [möglichst verständlich] to “all, including also Faculty members who lacked mathematical education”. (But Lotze was certainly not in agreement with Gauss, he was “a Boeotian” and went so far as calling “scientific freaks” [sic in the 1884 English translation] those who “intimidate” the general public talking about spaces of 4 or 5 dimensions; thus “they make sport of logical distinctions” (Lotze 1884, Sec. 172, p. 173).

  25. Gauss expresses his “joy” to see how easily Bessel had come to coincide with his own views in 1829, and remarks again that so few have an “open sense” for the matter.

  26. Even though he cannot prove it, despite all of his ideas and studies of the geometry of surfaces, and in particular of the pseudosphere (a surface of constant negative curvature, regionally like a non-Euclidean plane). Rigorous proof had to wait until the work of Beltrami.

  27. See Riemann (1854/68, 653, 659–661); on p. 661, Riemann wrote: “If we suppose that bodies exist independently of position, the curvature is everywhere constant, and it then results from astronomical measurements that it cannot be different from zero [...].” It might be that Riemann relied on Bessel’s observations, transmitted through Weber or Listing; for it was Bessel who, in 1838, first measured an annual parallax (0.3136″) for the star 61 Cygni, at a distance from the Earth of 657,000 AU. See Kragh (2012).

  28. This is coherent with his remarks in a review published in 1816, emphasizing the role of the “fructifying, living intuition” (see Ferreirós 2006b, 229), and with the fact that Gauss saw the need for a topological branch of the general theory of magnitudes.

  29. That varied along the century, but at the beginning, according to Bos (1980, 329) in many histories of mathematics mechanics was ranged under “mixed mathematics”. During the nineteenth century the distinction between pure and mixed mathematics will be transformed into pure and applied mathematics (see Epple et al. 2013 for more information about this transformation), and mechanics will be usually included under applied mathematics. An exception is Crelle’s Journal, where in 1826 mechanics was listed as a branch of pure mathematics, in the sense that insofar it is analytic mechanics, it is a branch of analysis (cf. Jungnickel et al. 1986, 27).

  30. This is in fact a claim made by Pulte in several of his papers, e.g. Pulte (1998; 2009).

  31. Interestingly, something like this was happening, exactly at the time, in Helmholtz’s introduction to his paper on the Erhaltung der Kraft. See Bevilacqua (1993), Hyder (2006).

  32. This idea, common as it might appear today, was not common at the time. When Neumann implemented mechanics as the introduction to his seminar course in theoretical physics, “his decision […] did not derive from his own educational experiences” (Olesko 1991, 125).

  33. Pulte (1994) points at a similar idea when he talks about the physicalization of Jacobi’s thought. He further develops it in Pulte (2005), in particular Ch. 6, Sect. 2.3.

  34. In the original “ad 1 im kleinen Bären”. This could be a reference to observations of a star (numbered 1), which is likely to be Polaris.

  35. For the relevance of the Besselian experiment to Neumann’s conception of Mechanics see Olesko (1991, Ch. 4).

  36. Herbart replaced Kant in Königsberg before moving to Göttingen in 1833 (until his death in 1841). For his influence on Riemann (who emphasized his dissent about metaphysics) see Scholz (1982a), Ferreirós (2006a). For his general relevance in nineteenth century philosophy, which is greater than usually realized, see Beiser (2015).

  37. See also the Herbart texts cited in Scholz (1982a, 421–22). On continuity and its role in Riemann’s mathematics, see Ferreirós (2006a, 79) and Laugwitz (1999). It may be mentioned here that Discrete/Continuous is precisely the first of Riemann’s antinomies, where those concepts are linked in particular with the constitution of space and time.

  38. We mean in particular the idea of manifolds of non-constant curvature; see Scholz (1982b).

  39. See de Paz (2018, 231) for Jacobi’s conception of simplicity and its relation to conventions and to the principle of inertia.

  40. More on this conception of a ‘third way epistemology’ in Pulte (2000) and de Paz (2014).

  41. More than a reshaping of a previously existing system, the process can be described as the creation of the scientific disciplines that we have known since then (see Stichweh 1984).

References

  • Abardia, J., Reventós, A., & Rodríguez, C. J. (2012). What did Gauss read in the Appendix? Historia Mathematica,39(3), 292–323.

    Article  Google Scholar 

  • Ahrens, W. (1904). C. G. J. Jacobi und die Jacobi-Biographie. Mathematisch-Naturwissenschaftliche Blätter, 1, 165–172.

    Google Scholar 

  • Ahrens, W. (1907). Briefwechsel zwischen C. G. J. Jacobi und M. H. Jacobi. Leipzig: BG Teubner.

    Google Scholar 

  • Archibald, T. (1991). Riemann and the theory of Nobili’s rings. Centaurus,34(3), 247–271.

    Article  Google Scholar 

  • Beiser, F. (2015). Two traditions of idealism. In G. Hartung & V. Pluder (Eds.), From Hegel to Windelband: Historiography of philosophy in the 19th century. Berlin: Walter de Gruyter.

    Google Scholar 

  • Bergman, B., Epple, M., & Ungar, R. (2012). Transcending tradition: Jewish mathematicians in German speaking academic culture. Berlin: Springer.

    Book  Google Scholar 

  • Bessel, F. W. & Gauss, C. F. (1880). Briefwechsel zwischen Gauss und Bessel. Edited by Preussische Akademie der Wissenschaften zu Berlin. W. Engelmann: Leipzig.

  • Bevilacqua, F. (1993). Helmholtz’s Ueber die Erhaltung der Kraft: The emergence of a theoretical physicist. In D. Cahan (Ed.), Hermann von Helmholtz and the foundations of nineteenth-century science (pp. 291–333). Berkeley: University of California Press.

    Google Scholar 

  • Bonola, R. (1906). Non-Euclidean geometry: A critical and historical study of its development (trans. from Italian). Chicago: The Open Court Publishing Company.

    Google Scholar 

  • Borchardt, W. (1875). Correspondance mathématique entre Legrendre et Jacobi. JRAM,80, 205–279.

    Google Scholar 

  • Bos, H. J. M. (1980). Mathematics and rational mechanics. In G. S. Rousseau & R. Porter (Eds.), The ferment of knowledge: Studies in the historiography of eighteenth-century science (pp. 333–351). Cambridge: Cambridge University Press.

    Google Scholar 

  • Caneva, K. L. (1978). From galvanism to electrodynamics: The transformation of German physics and its social context. Historical Studies in the Physical Sciences,9, 63–159.

    Article  Google Scholar 

  • de Paz, M. (2014). The third way epistemology: A re-characterization of Poincaré’s conventionalism. In M. de Paz & R. DiSalle (Eds.), Poincaré, philosopher of science. Problems and perspectives (Western Ontario Series in the Philosophy of Science, Vol. 79, pp. 47–65). New York: Springer.

    Google Scholar 

  • de Paz, M. (2018). From jurisprudence to mechanics: Jacobi, Reech and Poincaré on convention. Science in Context,31(2), 223–250.

    Article  Google Scholar 

  • Dedekind, R. (1876). Bernhard Riemanns Lebenslauf. In Riemann, Gesammelte mathematische Werke, 2nd ed., Leipzig 1892 (pp. 541–558).

  • Dirichlet, G. P. (1852). Gedächtnisrede auf C. G. J. Jacobi. Abhandlungen der Königlichen Akademie der Wissenschaften zu Berlin aus dem Jahre 1852, 1–27. In Dirichlet, Werke (Vol. 2, pp. 227–252).

  • Dirichlet, G. P. (1889/1897). Gesammelte Werke, 2 vols. In L. Kronecker & L. Fuchs (Eds)., Berlin, 1889. Reprinted by Chelsea in one volume, New York, 1969.

  • Dunnington, G. W. (2004). Carl Friedrich Gauss, titan of science. Cambridge: Cambridge University Press.

    Google Scholar 

  • Epple, M. (1999). Die Entstehung der Knotentheorie: Kontexte und Konstruktionen einer modernen mathematischen Theorie. Wiesbaden: Vieweg Verlag.

    Book  Google Scholar 

  • Epple, M., Kjeldsen, T. H., & Siegmund-Schultze, R. (2013). From ‘mixed’ to ‘applied’ mathematics: Tracing an important dimension of mathematics and its history. Oberwolfach Reports, 10(1), 657–733. https://doi.org/10.4171/OWR/2013/12.

    Article  Google Scholar 

  • Ferreirós, J. (1995). De la Naturlehre a la física: factores epistemológicos y factores socioculturales en el nacimiento de una disciplina científica. Arbor,151(596), 9–61.

    Google Scholar 

  • Ferreirós, J. (2006a). Riemann’s Habilitationsvortrag at the crossroads of mathematics, physics and philosophy. In J. Ferreirós & J. J. Gray (Eds.), The architecture of modern mathematics: Essays in history and philosophy (pp. 67–96). Oxford: Oxford University Press.

    Google Scholar 

  • Ferreirós, J. (2006b). Ο θεος άριθμητίζει: The rise of pure mathematics as arithmetic with Gauss. In C. Goldstein, N. Schappacher, & J. Schwermer (Eds.), The shaping of arithmetic: Number theory after Gauss’s Disquisitiones Arithmeticae (pp. 206–240). Berlin: Springer.

    Google Scholar 

  • Ferreirós, J. (2007). Labyrinth of thought. A history of set theory and its role in modern mathematics, Basel/Boston: Brinkhäuser (Science Networks, no. 23).

  • Friedman, M. (1999). Reconsidering logical positivism. Cambrige: Cambrige University Press.

    Book  Google Scholar 

  • Gauss, C. F. (1870–1927). Werke, 12 vols. Leipzig: B. G. Teubner; also in http://dz-srv1.sub.uni-goettingen.de/cache/toc/D38910.html.

  • Gauss, C. F. & Gerling, C. L. (1927). Briefwechsel zwischen Carl Friedrich Gauss und Christian Ludwig Gerling, C. Schaefer (Ed.), Berlin.

  • Gerling, C. L., & Lorenz, J. F. (1820). Grundriss der reinen und angewandten Mathematik, Helmstedt: G. Fleckeisen.

  • Giedymin, J. (1982). Science and convention: Essays on Henri Poincaré’s philosophy of science and the conventionalist tradition. Oxford: Pergamon Press.

    Google Scholar 

  • Gray, J. (1989). Ideas of space (2nd ed.). Oxford: Oxford University Press.

    Google Scholar 

  • Gray, J. (2004). János Bolyai, non-Euclidean geometry, and the nature of space. Burndy Library Publications (new series). Cambridge, MA: The MIT Press.

  • Heilbron, J. (Ed.). (2005). The Oxford guide to the history of physics and astronomy. Oxford: Oxford University Press.

    Google Scholar 

  • Hyder, D. (2006). Kant, Helmholtz and the determinacy of physical theory. In V. F. Hendricks, K. F. Jørgensen, J. Lützen, & S. A. Pedersen (Eds.), Interactions: Mathematics, physics and philosophy from 1860 to 1930 (pp. 1–42). Dordrecht: Kluwer.

    Google Scholar 

  • Jacobi, C. G. J. (1866). Vorlesungen über Dynamik, A. Clebsch (Ed.), Berlin.

  • Jacobi, C. G. J. (1881–1891). Gesammelte Werke. Edited by C. W. Borchardt (Vol. 1), K. Weierstrafi (Vols. 2–7) and E. Lottner (Suppl.). 7 vols. and 1 Suppl., Berlin.

  • Jacobi, C. G. J. (1996). Vorlesungen über analytische Mechanik. Berlin 1847/48, Braunschweig/Wiesbaden: F. Vieweg & Sohn.

  • Jungnickel, Ch., & McCormmach, R. (1986). Intellectual mastery of nature: Theoretical physics from Ohm to Einstein. Volume 1: The torch of mathematics. 1800–1870. Chicago: University of Chicago Press.

    Google Scholar 

  • Klein, F. (1926). Vorlesungen über die Entwicklung der Mathematik im 19. Jahrhundert (Vol. 1). Berlin: Springer Verlag.

    Google Scholar 

  • Königsberger, L. (1904). Carl Gustav Jacobi. Festschrift zur Feier der hundertsten Wiederkehr seines Geburtstages. Leipzig: B. G. Teubner.

    Google Scholar 

  • Kragh, H. (2012). Is space flat? Nineteenth-century astronomy and non-Euclidean geometry. Journal of Astronomical History and Heritage,15(3), 149–158.

    Google Scholar 

  • Kummer, E. E. (1860). Gedächtnisrede auf P. G. L. Dirichlet. In Dirichlet, Werke (Vol. 2, pp. 311–344).

  • Laugwitz, D. (1999). Bernhard Riemann 1826–1866: Turning points in the conception of mathematics. Basel: Birkhäuser.

    Book  Google Scholar 

  • Lipschitz, R. (1872). Untersuchung eines Problems der Variationsrechnung in welchem das Problem der Mechanik enthalten ist. Journal für reine und angewandte Mathematik,74, 116–149.

    Google Scholar 

  • Listing, J. B. (1847). Vorstudien zur Topologie. Göttinger Studien,1, 811–875.

    Google Scholar 

  • Lotze, R. H. (1884). Logic, in three books: Of thought, of investigation, and of knowledge (Vol. 1, p. 1874). Oxford: Clarendon Press (German orig.).

    Google Scholar 

  • Lützen, J. (1995). Interactions between mechanics and differential geometry in the 19th century. Archive for History of Exact Science,49(1), 1–72.

    Article  Google Scholar 

  • Olesko, K. M. (1991). Physics as a calling. Cornell: Cornell University Press.

    Google Scholar 

  • Pieper, H. (2007). A network of scientific philanthropy: Humboldt’s relations with number theorists. In C. Goldstein, N. Schappacher, & J. Schwermer (Eds.), The shaping of arithmetic: Number theory after Gauss’s Disquisitiones Arithmeticae (pp. 201–233). Berlin: Springer.

    Chapter  Google Scholar 

  • Pulte, H. (1994). C. G. J. Jacobis Vermächtnis einer ‘konventionalen’ analytischen Mechanik: Vorgeschichte, Nachschriften und Inhalt seiner letzten Mechanik-Vorlesung. Annals of Science,51(5), 487–517.

    Article  Google Scholar 

  • Pulte, H. (1998). Jacobi’s criticism of Lagrange: The changing role of mathematics in the foundations of classical mechanics. Historia Mathematica,25(2), 154–184.

    Article  Google Scholar 

  • Pulte, H. (2000). Beyond the edge of certainty: Reflections on the rise of physical conventionalism. Philosophia Scientiae,4(1), 47–68.

    Google Scholar 

  • Pulte, H. (2005). Axiomatik und Empirie. Eine wissenschaftstheoriegeschichtliche Untersuchung zur Mathematischen Naturphilosophie von Newton bis Neumann. Darmstadt: Wissenschaftliche Buchgesellschaft.

    Google Scholar 

  • Pulte, H. (2009). From axioms to conventions and hypotheses: The foundations of mechanics and the roots of Carl Neumann’s “Principles of the Galilean-Newtonian theory”. In M. Heidelberger & G. Schiemann (Eds.), The significance of the hypothetical in the natural sciences (pp. 77–98). Berlin: de Gruyter.

    Google Scholar 

  • Riemann, G. F. B. (1854/1868). ‘Ueber die Hypothesen, welche der Geometrie zu Grunde liegen’ (Habilitation lecture, first published posthumously in Abhandlungen der Königlichen Gesellschaft der Wissenschaften zu Göttingen, Vol. 13 (= Werke, 272–287). English translation by W. Clifford (revised by Ewald) in E. Ewald (Ed.), From Kant to Hilbert: A source book in the foundations of mathematics (Vol 2., pp. 652–661). Oxford, 1996: Clarendon Press; page references to Ewald translation.

  • Riemann, G. F. B. (1876). Fragmente philosophischen Inhalts (I. Zur Psychologie und Metaphysik; II. Erkenntnisstheoretisches; III. Naturphilosophie). In Gesammelte mathematische Werke, 507–538. English trans.: Philosophical fragments. 21st CENTURY (Winter 1995–1996), 50–62. Also: On psychology and metaphysics: Being the philosophical fragments of Bernhard Riemann (C. J. Keyser, Trans.). The Monist, 10(2) (January, 1900), 198–215.

  • Rodríguez, C. J. (2006). La importancia de la ‘Analogía’ con una esfera de radio imaginario en el descubrimiento de las geometrías no-euclidianas. Prepublicacions del Departament de Matemàtiques, UAB,30(2006), 1–48.

    Google Scholar 

  • Rowe, D. (2004). Making mathematics in an oral culture: Göttingen in the era of Klein and Hilbert. Science in Context,17(1/2), 85–129.

    Article  Google Scholar 

  • Scholz, E. (1982a). Herbart’s influence on Bernhard Riemann. Historia Mathematica,9, 413–440.

    Article  Google Scholar 

  • Scholz, E. (1982b). Riemann’s frühe Notizen zum Mannigfaltigkeitsbegriff und zu den Grundlagen der Geometrie. Archive for History of Exact Sciences,27(3), 213–232.

    Article  Google Scholar 

  • Scholz, E. (1992). Riemann’s vision of a new approach to geometry. In L. Boi, D. Flament, & J.-M. Salanskis (Eds.), 1830–1930: A Century of geometry: Epistemology, history and mathematics (Lecture Notes in Physics, Vol. 402, pp. 22–34). Berlin: Springer.

    Chapter  Google Scholar 

  • Scholz, E. (2004). C. F. Gauß’ Präzisionsmessungen terrestrischer Dreiecke und seine Überlegungen zur empirischen Fundierung der Geometrie in den 1820er Jahren. In R. Seising, M. Folkerts, & U. Hashagen (Eds.), Form, Zahl, Ordnung. Studien zur Wissenschafts- und Technikgeschichte. Ivo Schneider zum 65. Geburtstag (pp. 355–380) Stuttgart: Steiner.

  • Stäckel, P. (1923). Gauss als Geometer. In Gauss, Werke (Vol. 10.2, pp. 1–123).

  • Stäckel, P., & Engel, F. (1895). Theorie der Parallellinien von Euklid bis auf Gauss. Leipzig: Teubner.

    Google Scholar 

  • Stichweh, R. (1984). Zur Entstehung des modernen Systems wissenschaftlicher Disziplinen: Physik in Deutschland. Frankfurt a. M.: Suhrkamp.

    Google Scholar 

  • Torretti, R. (1978). Philosophy of Geometry from Riemann to Poincaré. Dordrecht: D. Reidel.

  • Ulrich, G. C. J. (1836). Lehrbuch der reine Mathematik. Göttingen.

  • von Waltershausen, S. (1856). Gauss zum Gedächtniss. Leipzig: Hirzel.

    Google Scholar 

  • Wachter, F. L. (1817). Demonstratio axiomatis in Euclideis undecimi. See Stäckel, P. (1900). Friedrich Ludwig Wachter, ein Beitrag zur Geschichte der nichteuklidischen Geometrie. Mathematische Annalen, 54, 49–75, with a translation of the Demonstratio and some letters.

  • Weber, W., & Weber, E. H. (1825). Wellenlehre auf Experimente gegründet. Leipzig: Fleischer; repr. as Vol. 5 of W. Weber, Werke (ed. by Eduard Riecke, Berlin, 1893).

  • Weil, A. (1979). Riemann, Betti and the birth of topology. Archive for History of Exact Sciences,20, 91–96.

    Article  Google Scholar 

  • Wussing, H. (1984). The genesis of the abstract group concept. Cambridge, MA: The MIT Press.

    Google Scholar 

  • Zahar, E. (2001). Poincaré’s philosophy: From conventionalism to phenomenology. Chicago: Open Court.

    Google Scholar 

  • Ziegler, R. (1985). Die Geschichte der geometrischen Mechanik im 19. Jahrhundert. Eine historisch-systematische Untersuchung von Möbius und Plücker bis zu Klein und Lindemann. Stuttgart: Steiner Verlag.

    Google Scholar 

Download references

Acknowledgements

We want to thank two anonymous referees and the editors of this journal for insightful comments. We are very grateful to Warren Schmaus for the English revision of the manuscript. We also thank the research project “The Genesis of Geometric Knowledge” (FFI2017-84524-P) for funding support.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Maria de Paz.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

de Paz, M., Ferreirós, J. From Gauss to Riemann Through Jacobi: Interactions Between the Epistemologies of Geometry and Mechanics?. J Gen Philos Sci 51, 147–172 (2020). https://doi.org/10.1007/s10838-020-09501-x

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10838-020-09501-x

Keywords

Navigation