Hostname: page-component-848d4c4894-x5gtn Total loading time: 0 Render date: 2024-05-21T15:29:22.067Z Has data issue: false hasContentIssue false

The Moral Freedom of Man and the Determinism of Nature: The Catholic Synthesis of Science and History in the Revue des Questions Scientifiques

Published online by Cambridge University Press:  05 January 2009

Mary Jo Nye
Affiliation:
Department of the History of Science, The University of Oklahoma, 601 Elm, Norman, Oklahoma 73069, U.S.A.

Extract

In 1877 the first issue of the Revue des questions scientifiques, published by the Scientific Society of Brussels, appeared in France and Belgium. The new journal was greeted with disdain and hostility by Emile Littrè and George Wyrouboff, the disciples of Auguste Comte and editors of La philosophie positive. The Scientific Society of Brussels was a Catholic organization, and the positivists' opinion was that ‘If science is spoken of in this assembly, it is in order to organize a veritable crusade against it’. But the highly prejudiced assessment by Littré and Wyrouboff completely misread the goals of the society. At the time, the Catholic Church was en pleine crise both in France and Belgium. Church attendance had declined dramatically in recent years, as had the number of young people entering religious orders. Many Catholic laymen and church officials were becoming convinced that some rapprochment with the modern world and modern science was essential. It was to this difficult task that the Catholics of the Brussels Scientific Society addressed themselves with determination.

Type
Research Article
Copyright
Copyright © British Society for the History of Science 1976

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

REFERENCES

1 ‘X’, ‘Science et religion’, La philosophie positive, xix (1877), 321–38 (321, 324).Google Scholar
2 See the excellent study of the Catholics in France under the Third Republic by McManners, John, Church and state in France, 1870–1914 (New York, 1973), pp. 412.Google Scholar
3 An ideological link between Catholic theology and Newtonianism is noted by Francis Oakley, who describes Newtonianism as a reassertion of the nominalist and voluntarist strain of Christian theology developed in the late thirteenth century. See Oakley, , ‘Christian theology and Newtonian science: the rise of the concept of Laws of Nature’, Church history, xxx (1961), 433–57.Google Scholar
4 See, in addition to McManners, Moody, Joseph N., ‘French anti-clericalism: image and reality’, The Catholic historical review, lvi (1970), 630–48 (642), andGoogle Scholar
Rummel, Leo L., ‘The anti-clerical program as a disruptive factor in the solidarity of the late French Republics’, Catholic historical review, xxxiv (19481949), 119 (1415).Google Scholar
5 Pope Pius IX denounced in the Syllabus of Errors (1864) ‘progress, liberalism and modern civilization’ and enunciated, on the very day that the French declared war against Prussia, the dogma of papal infallibility.Google Scholar
6 See McManners, , op. cit. (2), pp. 27–9.Google Scholar
7 But Ferry's law of 1879 limited the name ‘university’ to state-controlled institutions, and afterwards the Catholic universities of Paris, Lille, Lyons, Angers, and Toulouse renamed themselves Instituts Catholiques. See McManners, , op. cit. (2), pp. 3940, 4952;Google Scholar
Park, Julian, ‘Education’, in Park, (ed.), The culture of France in our time (New York, 1954), p. 244; andGoogle Scholar
Prost, Antoine, Histoire de l'enseignement en France 1800–1967 (Paris, 1963).Google Scholar
8 See, for example, ‘Le vrai terrain de la lutte entre croyants et incroyants’, discourse of 30 01 1884, quoted inGoogle Scholar
Paul, Harry W., ‘Science and the Catholic Institutes in 19th-century France’, Societas, i (1971), 271–85 (273).Google Scholar
9Carbonelle, Ignace, ‘Léon XIII et la Société Scientifique de Bruxelles’, Revue des questions scientifiques, v (1879), 353–60 (353).Google Scholar
10 Mansion originally named the group Cercle Leibniz, and then replaced the name of the Protestant Leibniz by that of Mansion's fellow-Catholic Cauchy. See de Locht, Lagasse, ‘Paul Mansion’, Revue des questions scientifiques, lxxvii (1920), 726 (1516 and 22).Google Scholar
11 After controversy in 1835–6 the Belgian university system was established with two state universities (Liège and Ghent) and two ‘free’ universities, a Catholic one at Louvain and a Liberal-Party sponsored one at Brussels. See Ensor, Robert C. K., Belgium (London, n.d. [c. 19141915]), p. 174.Google Scholar
12 See the list of the editorial board for the Revue des questions scientifiques in its first volume of 1877.Google Scholar
Also Van den Gheyn, J., , R. P., ‘Le jubilé de Société Scientifique de Bruxelles’, Revue des questions scientifiques, 1 (1901), 929 (1112), and ‘Le Ve Congrès Scientifique international des Catholiques, Munich, 24–28 septembre 1900’,Google Scholar
Van den Gheyn, J., , R. P., xlix (1901), 134–64 (135). In 1879 the Society had 700 members, according toGoogle Scholar
Carbonelle, , op. cit (9), p. 355. Membership required the presentation of a name by two members of the Society and approval by a majority vote at one of the two meetings of the Society that were held in each year. The Society elected annually a governing council of twenty members. The five sections of the Society included the mathematical, physical, natural, medical, and economic sciences. See the Society's statutes, on pp. 346–8 ofGoogle Scholar
Revue des questions scientifiques, iii (1878).Google Scholar
13Carbonelle, , op. cit. (9), pp. 353, 356.Google Scholar
14 See Carbonelle, , ‘L'encyclique et la science’, Revue des questions scientifiques, vi (1879), 353411 (357), and the translation therein of the encyclical, pp. 390–3;Google Scholar
also Carbonelle, 's ‘Une entrée en campagne’, Revue des questions scientifiques, iii (1878), 225–47, especially p. 234.Google Scholar
15Draper, John W., History of the conflict between religion and science (New York, 1899). The first edition appeared in 1874.Google Scholar
16 See Draper, , ‘The restoration of science in the South’, pp. 102–18 of Conflict, op. cit. (15); also pp. 148–50, 298–300; andGoogle Scholar
‘Science et religion’, La philosophie positive, xix (1877), 334–5. and xx (1878), 4056 (41).Google Scholar
17 For an account of investigations on Brownian motion through the experimental work of Jean Perrin and the mathematical contributions of Einstein, see my Molecular reality. A perspective on the scientific work of Jean Perrin (New York and London, 1972).Google Scholar
18 From 1862 to 1891 Delsaulx taught at the Collège Notre-Dame de la Paix at Namur and then at the Collège de Louvain. His scientific researches included the mathematical theory of capillarity, geometrical optics, and physical optics. See ‘Le R. P. Delsaulx, S. J.’, Revue des questions scientifiques, xxix (1891), 584–8.Google Scholar
19Carbonelle, , notes of 06-10 1874, in Bulletin de l'Académie Royale de Belgique, xli (1874), 410 cited inGoogle Scholar
Thirion, Julien, ‘Les mouvements moléculaires’, Revue des questions scientifiques, vii (1880), 555 (43); andGoogle Scholar
Delsaulx, Joseph, ‘Thermodynamic origin of the Brownian motions’, Monthly microscopical journal, xviii (1877), 17 (1).Google Scholar
20 See Thirion, , op. cit. (19), pp. 5155, andGoogle Scholar
Delsaulx, , op. cit. (19), pp. 1 and 56.Google Scholar
21Delsaulx, , op. cit. (19), p. 16, and his résumé,Google Scholar
‘Origine thermique des mouvements browniens’, Revue des questions scientifiques, ii (1877), 319–20.Google Scholar
22Schaffers, V., ‘Le R. P. Thirion’, Revue des questions scientifiques, lxxvii (1920), 2752.Google Scholar
Also Thirion, , ‘Le mouvement brownien’, Revue des questions scientifiques, xv (1909), 250–66 (260–1).Google Scholar
23Thirion, , op. cit. (19), p. 43. Jean Perrin won the Nobel Prize for physics in 1926, largely for his experimental investigations of Brownian motion. In an influential memoir of 1909 he cited the work of Thirion, as well as of Desaulx and Carbonelle, in a historical discussion of the few previous thermodynamic approaches to the problem.Google Scholar
See Perrin, , ‘Mouvement brownien et réalité moléculaire’, Annales de chimie et de physique, xvii (1909), 1114; translated byGoogle Scholar
Soddy, Frederick as Brownian movement and molecular reality (London, 1910).Google Scholar
24 See Heimann, P. M., ‘Molecular forces, statistical representation and Maxwell's demon’, Studies in the history and philosophy of science, i (1970), 189211 (203–4).Google Scholar
25Maxwell, , ‘Science and free will’, on pp. 434–44 ofGoogle Scholar
Campbell, Lewis and Garnett, William, The life of James Clerk Maxwell (London, 1882), p. 439.Google Scholar
26Delsaulx, , ‘La probabilité philosophique et la nature cinétique de la chaleur’, Revue des questions scientifiques, xxviii (1890), 484516. He also argued that the widely accepted ether was likewise incapable of absolute confirmation:Google Scholar
‘It is demonstrated, not shown’ (p. 514).Google Scholar
27 Paul De Broglie became well known in France for two books which he organized around the themes of an imminent demise of positivism and a rejuvenation of religion. See Le présent et l'avenir du catholicisme en France (Paris, 1892) andGoogle Scholar
La réaction centre le positivisme (Paris, 1894).Google Scholar
28Carbonelle, , ‘La physique moderne’, Revue des questions scientifiques, i (1877), 512–61, especially p. 554, andGoogle Scholar
‘La théorie atomique’, Revue des questions scientifiques, ii (1877), 236–73 (247–51). Barré de Saint-Venant had introduced Boscovichean ideas into a discussion at the Paris Academy of Sciences in 1877, and Maxwell discussed Boscovichean molecules in his article,Google Scholar
‘Atom’, in the Encyclopædia Britannica (9th edn., Edinburgh, 1878), iii. 3648.Google Scholar
29Carbonelle, , ‘La théorie atomique’, op. cit. (28), pp. 253, 271–3; andGoogle Scholar
Carbonelle, , ‘Réponse à M. l'Abbé de Broglie’, Revue des questions scientifiques, xi (1882), 225–60, andGoogle Scholar
de Broglie, Abbé, ‘Dynamisme et atomisme’, Revue des questions scientifiques, x (1881), 353412; and xi (1882) 169224, quotingGoogle Scholar
Carbonelle, , on pp. 370–1.Google Scholar
30De Broglie, , Revue des questions scientifiques, pp. 361, 364–5.Google Scholar
Also d'Estienne, Jean, ‘La constitution de l'espace céleste d'après M. Hirn et d'après la théorie atomique moderne’, Revue des questions scientifiques, xxvi (1889), 542–64, especially pp. 457, 549, 564.Google Scholar
31Carbonelle, , ‘Les actions vitales’, Revue des questions scientifiques, v (1879), 234–86, especially pp. 241–2, 262.Google Scholar
32 Ernst Cassirer argues that this is a false problem. See his Determinism and indeterminism in modern physics, trans. by Benfey, O. Theodor (New Haven, 1956), pp. 39.Google Scholar
Dubois-Reymond, 's lecture was translated as ‘Les bornes de la philosophie naturelle’, in Revue scientifique, 2nd ser. ix (10 10 1874), 337–45.Google Scholar
33Carbonelle, , ‘Actions vitales’, op. cit. (31), pp. 244–5:Google Scholar
‘The internal liberty of our will is an essential part of our [individual] nature; that alone destroys it which can destroy ourselves … this certitude does not allow the least doubt; it results from the clear and distinct view of truth’ (p. 246).Google Scholar
34Boussinesq, Joseph, ‘Conciliation du véritable déterminisme mécanique avec l'existence de la vie et de la vérité morale, précédé d'un rapport à l'Académie des Sciences morales et politiques par Paul Janet’,Google Scholar
ed. by Janet, Paul (Paris: Ernest Colas, 1878), 65 pp., p. 42. This is an extract from the longer work published byGoogle Scholar
Gauthier-Villars, in 1878.Google Scholar
35 See Picard, Emile, ‘La vie et l'oeuvre de Joseph Boussinesq’, Mémoires de l'Académie des Sciences, xxviii (1933), 44 pp. Boussinesq supported the mechanical viewpoint in lectures at the Sorbonne, under heavy criticism from Henri Saint-Claire Deville and Marcelin Berthelot. See, for example, hisGoogle Scholar
‘Recherches sur les principes de la mécanique, sur la constitution moléculaire des corps et sur une nouvelle théorie des gaz parfaits’, Mémoires de l'Académie des Sciences et Lettres de Montpellier, 8 juillet 1872, especially pp. 1, 18. In lectures given almost sixty years later, Boussinesq criticized modern physics and non-Euclidean geometry, the latter as depending on algebraic symbols too far divorced from reality. God, he said, is the ‘Eternal geometer’Google Scholar
(in ‘Cours de physique mathématique de la Faculté des Sciences. Epilogue’, 5th edn. [Paris: Gauthier-Villars, 1928], especially pp. 8, 11.) Boussinesq here recalls the plea of Joseph de Maistre for a genius who would unite science and religion, reason and faith (p. 12).Google Scholar
36Picard, , op. cit. (35), pp. 32–3.Google Scholar
37Boussinesq, , ‘Conciliation’, op. cit. (34), p. 39.Google Scholar
38Boussinesq, , pp. 3942. Maxwell's solution for free will was even less general. He identified human conduct with unstable systems ‘where an infinitely small variation in the present state may bring about a finite (rather than infinitely small) difference in the state of the system in a finite time. Free will can have no more than an infinitesimal range’.Google Scholar
See ‘Science and free will’, op. cit. (25), p. 441.Google Scholar
In arriving at the role of the ‘directing principle’, Boussinesq reveals the influence of the mathematician A. A. Cournot and physiologist Claude Bernard. Cournot's ‘directing power’ is less well known than Bernard's ‘directing idea’, which Bernard invoked especially to deal with the problems of inheritance and growth. Cournot had suggested that man might be able so to arrange the gears of a machine that the work produced by the machine would be infinitely small. The direction of the machine might be the result of the imprint of a ‘directing power’, rather than of a finite physical force. Venant, who was a friend and mentor of Boussinesq, revived this idea in the mid 1870s. See Picard, , op. cit. (35), p. 12, andGoogle Scholar
Janet, 's preface to Boussinesq, 's ‘Conciliation’, op. cit. (34), pp. 1315.Google Scholar
39 See Comptes rendus de l'Académie des Sciences morales et politiques, cix (1878), 719–20.Google Scholar
40Carbonelle, , ‘Les actions vitales’, op. cit. (31), pp. 282, 284. He discusses Boussinesq's memoir, pp. 273–9, andGoogle Scholar
Bertrand, 's remarks (from the Journal des savants, 09 1878), on p. 280.Google Scholar
41Boutroux, Emile, The contingency of the laws of nature, trans. by Rothwell, Fred (Chicago and London, 1920). It is important in understanding Boutroux to note that he was a religious man who sought to establish that the individual is responsible for the course of his life (p. 172). He wrote that ‘the understanding, through its category of necessity, is the middle term between the world and God’ (p. 179).Google Scholar
42 These generalizations result from my analysis of the contents of these journals during the period 1877–1914. The shared major interests of the three journals during these years were medicine, pathology, and physiology; then technology; and, to a lesser extent, descriptive zoology and anatomy. Certainly the missionary activities of the Catholic Church must account to some degree for the overwhelming interest of the Revue des questions scientifiques in anthropology and ethnology from about 1877 to 1896 and in the social sciences (and economics) from 1896 to 1914. The activities of Leopold II in the Congo also help account for the interest of the Brussels-based Revue des questions scientifique in Africa.Google Scholar
43Tyndall, John, ‘L'évolution historique des idées scientifiques’, Revue scientifique, 2nd ser. x. (19 09 1874), 265–82. Tyndall recommends Draper's History of the intellectual development of Europe, which contains similar views to the very recently published Conflict. Draper was a medical doctor who taught at Johns Hopkins and Harvard Universities.Google Scholar
44 ‘X’, op. cit. (1), p. 327; andGoogle Scholar
Draper, , Conflict, op. cit. (15), p. 301.Google Scholar
45 My italics. Carbonelle, , ‘Introduction’ to ‘L'aveuglement scientifique’, Revue des questions scientifiques, i (1877), 153.Google Scholar
46 The latter remarks are in Thirion, , ‘L'analyse des radiations’, Revue des questions scientifiques, xlii (1898), 524–41: and xliv (1898), 488520.Google Scholar
‘Les Jésuites astronomes’ was published in Précis historiques (Brussels, 1880).Google Scholar
Schaffers, V. compiled a bibliography of Thirion's writings for the obituary of 1920, ‘Le R. P. Thirion’, op. cit. (22).Google Scholar
47Thirion, , ‘Pour l'astronomie grecque’, Revue des questions scientifiques, xlv (1899), 547, 435–75; and xlvi (1899), 111–58, especially pp. 43, 436, 440, 473–4.Google Scholar
See also Schaffers, , op. cit. (22), pp. 41–3.Google Scholar
See also Thirion, 's astute analysis of the arguments of 1906–7 among historians concerning Pascal's work on the void and atmospheric pressure: ‘Pascal. L'horreur du vide et la pression atmosphérique’, Revue des questions scientifiques, lxii (1907), 384450; and lxiii (1907), 149251.Google Scholar
Tannery, Paul cites Thirion's analysis as thorough and meticulously fair in his Mémoires scientifiques, vi. 316.Google Scholar
Duhem, Pierre also cited Thirion's work on Greek astronomy in the Systéme du monde, i. 418n. and 448n.Google Scholar
48 See de Locht, , ‘Paul Mansion’, op. cit. (10), especially p. 20.Google Scholar
De Locht, cites a bibliography of Mansion's work in Liber memorialis, Université de Gand, p. 349, including publications to 1912.Google Scholar
See also the notice by Demoulin, A. in Annuaire de l'Académie royale de Belgique, xcv (1929), 77147.Google Scholar
In his Système du monde, ii. 68n.,Google Scholar
Duhem cites Mansion, 's ‘Note sur le caractère géométrique de l'ancienne astronomie’, Abhandlungen zur Geschichte der Mathematik, ix (1899). George Sarton studied at Ghent while Mansion was teaching there; on their intellectual relationships,Google Scholar
see Gillis, J., ‘Paul Mansion en George Sarton’, in Mededelingen van der Koninklijke Academie … van België, Klasse der Wetenschappen, xxxv (1973), 321, cited inGoogle Scholar
Thackray, Arnold and Merton, Robert K., ‘George Sarton’, Dictionary of scientific biography, vol. xiii (New York, 1975), 114.Google Scholar
49 ‘Définition d'un nombre incommensurable’, in Mathesis (1885), p. 2; quoted inGoogle Scholar
Carbonelle, Ignace, ‘Les nombres et la philosophie’, Revue des questions scientifiques, xvii (1885), 467505, especially pp. 490–1.Google Scholar
50 See Paul, Harry W., ‘The debate over the bankruptcy of science in 1895’, French historical studies, v (1968), 299327 (325–7). Professor Paul offers a more thorough analysis ofGoogle Scholar
Tannery, 's candidacy for the chair in ‘Scholarship vs. ideology: the chair of the general history of science at the Collège de France (1892–1913)’, Isis, forthcoming.Google Scholar
51 In reports by his rector to the Ministry of Education in Paris, Duhem was branded as an intransigent Catholic, the companion of monks and religious orders, and a fanatical ultra-montane. See Paul, Harry W.. ‘The crucifix and the crucible: Catholic scientists in the Third Republic’, Catholic historical review, lviii (1972), 195219 (202–11).Google Scholar
52Duhem, , ‘Quelques réflexions au sujet des théories physiques’, Revue des questions scientifiques, xxxi (1892), 139177.Google Scholar
53Duhem, , ‘Notation atomique et hypothèses atomiques’, Revue des questions scientifiques, xxxi (1892), 391454.Google Scholar
54Duhem, , ‘Physique et metaphysique’, Revue des questions scientifiques, xxxiv (1893), 5583.Google Scholar
55Duhem, , ‘Quelques réflexions au sujet de la physique expérimentale, Revue des questions scientifiques, xxxvi (1894), 179229.Google Scholar
56 During the 1890s Duhem reviewed for the Revue des questions scientifiques Poincaré, 's Thermodynamics and Mathematical theory of light,Google Scholar
Boltzmann, 's Lectures on Maxwell's theory of electricity and light,Google Scholar
Lorentz, 's Electromagnetic theory of Maxwell and its application to moving bodies,Google Scholar
Kirchoff, 's Lectures on mathematical physics, andGoogle Scholar
Hertz, 's Researches on the propagation of electrical force.Google Scholar
See Revue des questions scientifiques, xxxi (1892), 603–6; xxxiii (1893), 257–65; and xxxvi (1894), 265–8. In addition to Maxwell's view of physical theory as ‘consistent representation’, Joseph Larmor's view of theory as ‘working representation’ was widely read at the turn of the century in FranceGoogle Scholar
(on their views, see the unpublished dissertation by Heimann, P. M., ‘James Clerk Maxwell: his sources and influence’ [University of Leeds Ph.D. thesis, 1970]). An analysis of Duhem's philosophy must also take into account the critiques of Kant's writings which were beginning to figure heavily in French philosophy in the mid-1870s. The Thomist Society of Paris, for example, devoted discussions to Thomism and Kantianism. Boutroux was a neo-Kantian, and one of Duhem's best friends, Victor Delbos, was a student of Boutroux who wrote his principal doctoral thesis on Kant (and his Latin thesis on Schelling and Hegel).Google Scholar
57 Indeed, critics charged that his Catholic metaphysics likewise determined his physics. In his Physique du croyant (see extract from the Annales de philosophie chrétienne [Paris, 1905]), Duhem drew analogies between Aristotelian metaphysics and thermodynamics in regard to (1) the equal importance to each of the categories of quality and quantity; (2) in each, local movement is merely a form of general movement; and (3) transformations are not simply due to local movement, but to generation and corruption, i.e. the (chemical) creation of new substances (p. 47). He regards Aristotle's principle of movement towards natural place as a statement analogous to the principle of increasing entropy and movement towards equilibrium in the second law of thermodynamics (pp. 49–50.)Google Scholar
58 See Picard, , La vie et l'oeuvre de Pierre Duhem (Paris, 1922), especially pp. 3451, on Duhem as an historian.Google Scholar
59 About 1900, articles in the Revue des questions scientifiques described a complex picture of atoms and molecules, reflecting both the recent philosophical critiques of atomic theory and the new investigations on radiations and kinetic theory. For example, Albert de Lapparent referred, in 1902, to Duhem's critiques and to Jean Perrin's article of 1901 (in Revue scientifique) ‘Molecular hypotheses’. Lapparent expressed hope for a reconciliation between Duhem's energetics and Perrin's molecular hypothesis, where the latter conception would be ‘representative’ of the ‘distribution of local centres of energy’. See Lapparent, , ‘Atomes et molécules’, Revue des questions scientifiques, li (1902), 353–87, especially pp. 353, 354, 377, 383, 384. V. Schaffers concluded that contemporary investigations of matter, ether, and electricity were not contradictory only ‘if we are content with … a species of symbol or image, proper to coordinate our sparse knowledge and put us on the way to new researches’. InGoogle Scholar
Schaffers, S. J., ‘Électrons’, Revue des questions scientifiques, liii (1903), 65119, especially pp. 116, 118.Google Scholar
See also Thirion, 's ‘L'analyse des radiations’, op. cit. (46), andGoogle Scholar
Adhemar, Vte Robert, ‘Les principes de la mécanique et les idées de Hertz’, Revue des questions scientifiques, li (1902), 173204.Google Scholar
60LeRoy, , ‘Science positive et les philosophies de la liberté’, cited inGoogle Scholar
Lachalas, Georges, ‘Les confins de la science et de la philosophie’, Revue des questions scientifiques, xlix (1901), 491505 and 567–84, quotingGoogle Scholar
LeRoy, on pp. 495, 497, 498. Boutroux presided over the meeting, and Paul Mansion sat on the organizing committee (p. 491).Google Scholar
See also LeRoy, 's earlier article, ‘Science et philosophie’, Revue de métaphysique et de morale, vii (1899), 375425, 503–62, 706–73; viii (1900), 3772.Google Scholar
61 See note 57 above.Google Scholar
62Sabatier, Auguste, Outlines of a philosophy of religion (U.S.A., n. d., no translator); from the 5th edn., 1898, pp. 230–1. In his autobiography Jean Piaget remarked that Sabatier's book was one of the most influential on Piaget's own youthful intellectual development.Google Scholar
See Evans, Richard I. (ed.), Jean Piaget. The man and his ideas, trans. by Duckworth, Eleanor (New York, 1973).Google Scholar
63McManners, , op. cit. (2), pp. 105–6.Google Scholar
64 See Coumet, E., ‘Edouard LeRoy’, Dictionary of scientific biography, vol. viii (New York, 1973). pp. 256–8.Google Scholar
65 With the death of Leo XIII in 1903, the moves of the modernists were cut short. Pius X excommunicated Loisy and condemned modernist writings in the encyclical of 1907, Pascendi. LeRoy, 's later Problème de Dieu (1921) was placed on the Index in 1931.Google Scholar
See McManners, , op. cit. (2), p. 107, andGoogle Scholar
Coumet, , op. cit. (64).Google Scholar
66 A student of Boutroux and a comrade of Duhem's friend Victor Delbos, Blondel submitted his thesis in 1893. It was criticized by Renan's colleagues as too religious and by Catholic theologians as religiously inadequate, according to Ratté, John, ‘Church, Modernism in the Christian’, pp. 418–27 in theGoogle Scholar
Dictionary of the history of ideas, vol. i (New York, 1968), 422.Google Scholar
See Blondel, Maurice, L'Action. Essai d'une critique de la vie et d'une science de la pratique (Paris, 1893), pp. 424–5.Google Scholar
67Poincaré, Henri said of LeRoy, Edouard: ‘Nominalist in doctrine but realist at heart, he appears to escape absolute nominalism only by an act of desperate faith’;Google Scholar
see Poincaré, , ‘Sur la valeur objective de la science’, Revue de métaphysique et de morale, x (1902), 263–93 (263).Google Scholar