Keywords

12.1 Introduction

In the nineteenth and twentieth centuries German scientists and engineers have developed a special taste for the composition of impressive encyclopaedias and so-called “Handbucher” (Handbooks). A typical Handbuch (in fact a Taschenbuch) in mechanical engineering has been the very popular one by Hütte with many foreign translations, but this was more a catalogue of prescriptions, standards, and elementary formulas of mathematics and strength of materials. Famous collections of the “Handbuch der Physik” have been edited by Geiger and Scheel between 1926 and 1933 [18] and by Flügge between 1955 and 1988 [17]. Mathematicians did not escape this trend. In particular, renowned mathematicians such as Felix Klein (1849–1925) and Conrad H. Mueller (1857–1914) contributed their wide experience and many friendly connections to the creation and edition of a monumental encyclopaedia of mathematics under the German title “Encyklopädie der mathematischen Wissenschaften mit Einschluss ihrer Anwendungen”—in brief: Enz. Math. Wiss. (EmW)—published by B. G. Teubner (Verlag) in Leipzig between 1907 and 1914 [31].Footnote 1 Various mathematicians and physicists were called to contribute to this vast enterprise. Part Four of Volume Four was devoted to Mechanics (Mechanik).Footnote 2 In that Volume the burden of writing Article 30 on the General bases/formulation (principles) of continuum mechanics (“Die allgemeinen Ansätze der Mechanik der Kontinua”) fell on Ernst Hellinger then in Marburg. The Encyclopaedia is well-documented with scholarly articles. It is aimed at the specialist. Concerning the whole EmW, it is salient to note the following appreciation of I. Grattan–Guiness (2009), the famous historian of sciences: “Many of the articles were the first of their kind on their topic, and several are still the last or the best. Some of them have excellent information on the deeper historical background. This is especially true of articles on applied mathematics, including engineering, which was stressed in its title”. This particularly applies to Hellinger’s contribution.

Ernst Hellinger (1883–1950) had been educated in Heidelberg, Breslau and Göttingen and was a doctoral student of David Hilbert. This indicates that he was a rather pure mathematician whose most famous mathematical accomplishments were in integral and spectral theories. He became a professor in Frankfurt am Main but he left Germany for the USA in 1939 and then taught at Evanston, Illinois. The writing of this contribution in continuum mechanics in 1913 [26]Footnote 3 may have been a parenthetical episode in his career. Nonetheless, he was much interested in variational formulations (as shown by the forthcoming perusal of his contribution) and even introduced the notion of two-field variational principle now referred to as the Hellinger-Reissner variational principle in elasticity (cf. [53]). Nonetheless, we surmise that his formation with Hilbert led him to view continuum mechanics as one of the physical sciences to be formalised and given an axiomatic framework, an orientation that will be materialized later on by the Truesdellian school with Noll (cf. [48]). Although not a full time mechanician, Hellinger was able to capture in a rather concise contribution all recent and promising advances by keeping a sufficiently high standpoint, a balanced neutrality, and an acute insight, and this, in our opinion, much more than some professional mechanicians who kept too much with well established subject matters. In order to help the reader not accustomed with reading in German, a partial translation in English of Hellinger’s contribution is provided in an Appendix.

12.2 The Scientific Environment

Although Hellinger was essentially foreign to the engineering spirit, in writing his opus of 1914 he gathered a rich past and contemporary documentation and accounted for most of the recent works in the field of theoretical continuum mechanics. He was not building in a vacuum, but this voluntary embedding in a medium other than his own is altogether remarkable. Of course the influence of his mentor David Hilbert may have played a fundamental role in his clear interest for the general and somewhat axiomatic aspects, so that he must have been familiar with the then recent attempt of Hamel [24] to delineate the structure and principles of mechanics (as of the beginning of the twentieth century), and the recently published treatise on “energetics” by Duhem [12] with its pre-Truesdellian flavour which may have been to his taste. This is corroborated by his frequent citations of these two authors. But he also knows the impressive treatise of Appell [1] on the rational mechanics of deformable bodies and the German synthetic texts of Heun [29, Voss [52], and Voigt [51].

Being basically a mathematician and a great admirer of Lagrange, Hellinger is also very much concerned with variational formulations in the works of W. Thomson (Lord Kelvin), Kirchhoff and, above all, the Cosserat brothers [7, 8, 40]. The last connection may have been through his reading of the Third volume of Appell’s treatise [1] in which there is a supplement written by the Cosserats. The appeal to group symmetries in the line of Sophus Lie and Henri Poincaré by the Cosserats may have been very attractive to him. But he also considered the possible occurrence of dissipation with the notion of dissipation potential introduced by Rayleigh, and even time-dependent (memory-like) behaviours in the manner of Boltzmann [2]. The recent work of Hadamard [23] on wave propagation has also left a strong print. Finally, in contrast with many other writers of the period who remain in the classical (Newtonian) framework, Hellinger has already integrated in his views the revolutionary ideas of Einstein in 1905 on relativity and Minkowski [42] on space-time. All these remarks are based on the citations of these authors by Hellinger as checked in the many footnotes to his contribution. This is the general background and favourable scientific environment in which this perspicacious author has framed his article.

12.3 The Contents of Hellinger’s Article

12.3.1 Introductory Remark

Hellinger’s article is only ninety two pages in print. Nonetheless, it succeeds in providing a rather complete survey of the field both with its established bases, its recent successes and some view of things to come. This friendly neutrality with which the author looks upon his assigned duty—in principle perusing a vast domain of knowledge with about a hundred fifty years of history and a vivid contemporary activity—is conducted with no a priori prejudice as a result—we surmise—of Hellinger being a somewhat outside observer. Hellinger is rather generous but also very accurate with citations. He cites many authors, whatever their nationality, but is clearly most influenced by works published within the thirty years before his synthesis, say in the period 1880–1910. Perhaps because of this “actuality”, he does not confine himself to the well established fields (linear elasticity and Eulerian fluids), but he often venture in newly expanded fields of interest such as finite deformations, oriented (Cosserat) bodies, capillarity, formulation of thermo-mechanics, analogy with electrodynamics, and even relativistic continuum mechanics.

From a historical viewpoint, our perusal of this beautiful contribution should not be influenced by our own education in the field (rough period 1960s–1970s) and our knowledge accumulated over an active professional period of some forty five years that witnessed many developments. But it happened that many of these rich developments in a vivid period of research more or less coincide with many of the points touched upon by Hellinger. We do not think that this kind of resonance between Hellinger’s approach to our field of interest and our own view is so much due to an influence that this author would have exerted on the generations that followed his own. Indeed, Hellinger’s text may have been read by German scientists between the two World Wars. But we must notice that his article was published in an encyclopaedia of mathematics, in a style that is permeated by the rigorous thinking of a mathematician—far from engineering interests—and that the text was not translated in any foreign language. It just happened that a spirit close to that of Hellinger re-appeared in our period of activity, and this of course greatly facilitates our apprehending of his exposition.

12.3.2 The Layout and Articulation of the Contribution

Every synthetic work in a field has to respect a definite agenda. This particularly applies to an article in an encyclopaedia of which the readership is not so well delineated. In the present case a tradition has settled that the progression in the presentation of the subject matter follows an almost fixed order (as exemplified in many textbooks on continuum mechanics), geometric background being introduced first, followed by kinematics and the theory of deformations, then kinetics and the general laws of mechanics, general classes of mechanical behaviours and a few more specific examples, and finally (but not always) some more exotic extensions. Hellinger’s approach is more difficult to grasp because he is ahead of his time while simultaneously following some masters such as Kirchhoff, Helmholtz, Clebsch and Barré de Saint-Venant, and he has thoroughly gone through the then recent works by W. Voigt, J. V. Boussinesq, E. and F. Cosserat, H. Poincaré, and P. Appell, authors who are very often accurately cited. In reason of the imposed exercise, Hellinger’s text is extremely dense. Instead of perusing his contribution just in its order of presentation—the easy way—we have preferred to examine various points, that recur in the whole text and seem to emphasize Hellinger’s repeated interest in some specific aspects as an exemplary mathematician (obviously not the point of view of an engineer).

12.4 The Identified Fields of Marked Interest of Hellinger

12.4.1 On General Principles of Mechanics and General Equations

This is not an original point of departure in Sect. 12.2. Hellinger builds on the commonly admitted bases of Newtonian mechanics in the tradition set forth by Euler, Lagrange, Cauchy, but with modern references to Brill [5]; Duhem [12]; Voigt [51], and other contributions to the same encyclopaedia by, e.g., Voss [52] and Heun [29]. He clearly indicates his favoured view of Hilbert and Hamel [24, 25] —later on formalized in Hamel's contribution to the Handbuch der Physik in 1927 —for axiomatization and the consideration of a general thermodynamic framework by Duhem [12]. He also heavily borrows from the treatise of Appell [1] and the recent works by the Cosserat brothers ([7, 8], and their numerous notes in the Comptes-Rendus of the Paris Academy of Sciences). But Hellinger does not hesitate to introduce the relativistic Einstein–Minkowski’s vision in the last section of his contribution.

Formally, Hellinger is much more attached to the Lagrangian-Hamiltonian variational formulation than to the classical Newtonian type of approach that relies on a statement of laws of equilibrium or dynamics. This he shows even for the bases of statics where he readily implements the principle of virtual work (Sects. 12.3 and 12.4). This may be one of the reasons why this work is not so much cited in the “Anglo-Saxon” literature dominated by Newton’s vision and made popular in continuum mechanics by the Truesdellian school in the 1960s. But Hellinger cannot avoid discussing the notion of force as a polar vector (p. 613) and the clever introduction of the concept of stress by Cauchy (Cauchysche “Drucktheorem”; p. 615). On this occasion, Hellinger, above all a mathematician, acknowledges the usefulness of the notions of vector analysis and dyadics—linear vector functions—in the line of J. W. Gibbs (cf. [21]) and the matrix calculus of Cayley (p. 613). He also refers to “tensor components of a dyad” (Tensorenkomponenten) after Voigt’s lectures on the physics of crystals (p. 624). This is to be contrasted with the rather shy attitude of contemporary authors (e.g., Appell [1]; see my own appraisal in Maugin [41]).

Hellinger’s presentation of equilibrium equations in the Eulerian framework with the associated natural boundary conditions (reflecting Cauchy’s postulate)—Eqs. (5a) and (5b), p. 617—is rather modern. But he also gives what may be considered the Piola-Kirchhoff format as Eqs. (9a) and (9b) in p. 618, after what looks like a Piola transform for the stress in Eq. (8). He indeed refers to the work of Piola [44] in p. 620. For the symmetry of the Cauchy stress, he refers (p. 619) to Hamel who calls this the “Boltzmannsches Axiom” for “die Symmetrie der Spannungsdyade”. Reductions to the two-dimensional (e.g., plates) and one-dimensional cases (e.g., rods, filaments)—in Eqs. (18a) and (18b) in p. 622 for this last case—are given following the Cosserats.

12.4.2 On Variational Formulations

For a mathematician like Hellinger the attraction to the beauty, economy of thought, and efficacy of variational formulations is inevitable. Hellinger, a follower of Lagrange, Piola, Hamilton, Kirchhoff, Helmholtz and the Cosserats, in fact starts by emphasizing the exploitation of the principle of virtual perturbations (“virtuellen Verrückungen”; p. 611 on)—virtual work (a weak formulation in the modern jargon). To the risk of creating an anomalous connection with modern standards, we perceive in these perturbations the notion of test functions (see Maugin [35, 37]). Note that Hellinger gives a mathematically correct definition of what is a material variation by considering an infinitesimal parameter noted σ (and not ε like in modern treatments; cf. pp. 607–608). As a matter of fact Hellinger’s statement (7) in p. 612 is, but for different symbols, just the same as in a modern formulation where the principle of virtual powers (for statics) is written for a massive body as

$$P_{vol}^{*} + P\,_{int}^{*} + P_{surf}^{*} = 0 ,$$
(12.1)

where the three terms refer to volume, internal and surface forces, respectively. The Cauchy stress is introduced in the second term as a co-factor. A power of inertial (acceleration) forces is added in the right-hand side of Eq. (12.1) in the dynamical case. The second term is transformed with the help of Green’s divergence theorem [22] to yield a divergence term in the bulk and Cauchy’s natural boundary condition at the surface. Hellinger emphasizes the equivalence of the statement (1) with Newton’s laws (cf. p. 630).

In dynamics we have D’Alembert’s principle per se (d’Alembertschen Prinzips, p. 629) and this yields the looked for equations such as Eq. (2) in p. 630. On introducing the kinetic energy, Hellinger is led to the principle of least action (p. 633) of Maupertuis and Hamilton. Gauss’ principle of least constraint is also evoked in the same page with the possibility to account for non-holonomic constraints. The general nature of such formulations is clearly acknowledged including with due reference to the Cosserats. With the assumed existence of a strain potential Hellinger touches upon the favourite subject matter of Kirchhoff, Boussinesq, Duhem [10], Poincaré and the Cosserats (pp. 643–651). This led him to examine some questions related to stability in agreement with Dirichet and above all Born [3], as also Italian authors such as Menabrea and Castigliano. He introduces appropriately the notions of canonical transformation (p. 657) and Legendre transformation (function H in p. 654). This leads him to say a few words about minimum principles and stability. Unknown multipliers (interpreted sometimes as stresses or “reaction forces”) are introduced wherever a mathematical constraint is imposed (e.g., incompressibility), following ideas of the French mathematician J. Bertrand and also D. Hilbert (see pp. 661–663). Ideal fluids accept a characteristic equation \(p = p\left( \rho \right)\) when, following Hadamard [23], the potential reduces to a function of the Jacobian of the deformation. In presence of some dissipation Hellinger follows an idea of Rayleigh to consider a potential of dissipation (p. 657). This will later be formalized even for plasticity (dissipation function homogeneous of order one only) in works of the 1970s–1980s (see, e.g., Maugin [36]).

Apart from the extensions to oriented media (his Paragraph 4b, and Paragraph 4.4 below), Hellinger touches two other extensions of the principle of virtual perturbations that were to bear fruits later on. One is the possibility of considering higher-order space derivatives of perturbations. This was envisaged early by Le Roux [33]—apparently unknown to Hellinger—to account for effects of spatially non-uniform strains (such as in torsion) in small-strain elasticity. This would later on be expanded in the so-called gradient theory of continua [Cf. works by R. D. Mindlin in the 1960s, and above all: Germain [19], for the second gradient, and Maugin [35] for a general framework, using the principle of virtual power without knowledge of Hellinger’s contribution]. The other is the possibility to account for the existence of unilateral constraints during the variation (Cf. Paragraph 4c). This was to be expanded in the theory of variational inequalities in the mechanics of continua (Cf. e.g., Duvaut and Lions [13, 14]).

Finally, it is often said (cf. Washizu [53]) that Hellinger contributed to the variational formulation of continuum mechanics (elasticity) by introducing before Reissner [45] the notion of two-field variational principles. In these both displacement and stresses are varied, allowing a relatively easy accounting of boundary conditions of mixed type. Reissner—educated in Germany and himself the son of a reputed mathematical physicist—must have heard of, if not studied, Hellinger’s contribution. However, he proudly told the present writer that “he did not see why Hellinger’s name was attached to his own name for this notion”. It is true that we could not locate where Hellinger introduced this notion. But the association may come from the fact that—as noted above—Hellinger duly considers Legendre transformations of the energy potential, introducing a kind of complementary energy.

12.4.3 On Finite Strains and Elasticity

Hellinger follows the tradition established by Piola, Kirchhoff, Boussinesq and the Cosserats by always considering the case of finite strains, linear elasticity being only an approximation. This is exemplified at many stages in his contribution. First both actual (noted \(x,y,z\)) and referential or material (Lagrangian) coordinates (noted \(a,b,c\)) are introduced. This later on allows for the introduction of the Piola transformation (8) in p. 618 with a clear algorithm in spite of the absence of tensor notation. The Piola-Kirchhoff form of the equilibrium equations follows at once as Eq. (9a, b). This also applies to Cosserats’ media (p. 624–625). In the case of Green elasticity for which there exits a strain potential, the Cauchy-Green’s finite strain is duly introduced (cf. Eq. (12.1) in p. 663). An example of higher order (than quadratic) strain energy function is given in p. 665. The exact constitutive equations for Cauchy’s stress tensor in finite strains are given as Eq. (5) in p. 645 in Boussinesq’s form while Piola’s form is given in p. 654 together with Max Born’s Footnote 4 equations in terms of a potential in stresses—Eqs. (22a, b) in p. 654—after introduction of the complementary energy by means of a Legendre transformation. The resulting compatibility condition for the finite deformation gradient is given in Eq. (24) in p. 655 in a form due to von Kármán. In the case of isotropic materials Hellinger rightfully calls for the invariance under orthogonal transformations (p. 664) and the resulting dependency of the strain energy on the basic invariants that are factors in the Cayley-Hamilton theorem. He evokes on this occasion the possible existence of self-stresses. Citations to Boussinesq, Duhem, Poincaré, the Cosserats, Helmholtz and J. Finger abound. All these now seems quite familiar to students who followed the masters of continuum mechanics in the 1960s–1980s—e.g., in the books of Truesdell and Toupin, Green and Zerna, Leigh, Eringen, etc., in the USA and those of Goldenblatt, Novozhilov, Lurie, Sedov, Ilyushin and others in the Soviet Union—this includes the present writer.

As a true mathematician, Hellinger views small-strain elasticity as a theory of perturbations introducing wherever necessary a small parameter (noted sigma and not epsilon) that indeed indicates the smallness of strains about an undeformed state [cf. Eq. (3’) in p. 608].

12.4.4 On Oriented Media

From the very beginning of his exposition Hellinger envisages the possible existence of internal degrees of freedom of the type proposed by the Cosserats in 1909. For instance, introducing the basic physical parameters of a continuum, together with the notion of density (p. 609), he feels quite natural to consider the possible attachment to each material point (the “Quantum der Materie” with material coordinates in his own language; p. 606) of an oriented trihedron or triad of rigid vectors (ein “rechtwinkliges Axenkreus”) likely to represent the varying orientation of “molecules”—as proposed by Voigt [50] and possibly by S.D. Poisson much earlier in 1842 (cf. footnote in p. 609). This yields the notion of “Medien mit orientierten Teilchen” (pp. 609–610) in the manner of the Cosserat brothers (and perhaps Duhem [11], p. 206; see Maugin [39]).

Then in considering a variational formulation (principle of virtual work), Hellinger naturally generalizes it to the case including local orientational kinematic properties (pp. 623–627) with specialization to two-dimensional and one-dimensional cases. The concept of couple-stress tensor [“Drehmoment” (dyade)] then appears naturally. The author recurs to this framework of “generalized continua” on many occasions, in particular when considering the Green type of elasticity based on the existence of a potential for strains (pp. 646–651) with the application of the Cosserats’ concept of “Euclidean action”. He returns to the notion of “generalized continuum” while dealing with analogies with the equations of light propagation and electrodynamics (the MacCullagh “ether” of 1839 [34]—an elastic medium able to transmit only transverse waves (light) in agreement with Fresnel’s observations—the deduction of Maxwell equations by identifying elastic displacement and electric field on the one hand and vorticity with magnetic induction on the other and as done by authors such as Kelvin or J. Larmor—see pp. 675–681). Of course this is now rather obsolete and was already evaporating in thin air at the time of Hellinger after the works of Lorentz, Poincaré and Einstein. But Hellinger’s attitude is above all witness of a marked interest in the rich modelling potentiality offered by continuum mechanics—although sometimes along paths with dead-end—leaving the final choice to true physicists.

We must note that, just like most authors until 1966, Hellinger does not see that, similar to density with its conservation law (cf. Eq. (12.7) in p. 609 in the Lagrange-Piola format), there must exist a conservation law associated with the inertia of the new orientational degrees of freedom. This missed step was resolved much later by Eringen [15].

12.4.5 On One-Dimensional and Two-Dimensional Bodies

Hellinger always considers two-dimensional and one-dimensional material bodies (“Platten und Drähte”) as special cases. In this he does not follow the Cosserat brothers who work more with an increase in spatial dimensions than with a successive reduction. Much more than that, in pp. 658–660, he shows his apprehension of the true mathematical problem at the basis of this reduction in dimensions by introducing small parameters (this time noted ε) that are representative of the slenderness in thickness or of two (small) lateral dimensions of the considered material structure. Equation (12.2) in p. 659 is typical of this “asymptotic” approach that will later on be the source of an efficient asymptotic derivation of equations for plates, shells and rods in the expert hands of Gold’denveizer, S.A. Ambartsumian, V.L. Berdichevsky, Ph. Ciarlet and others. Furthermore, Hellinger does not hesitate to introduce the Gaussian parametrization of curved surfaces to treat two-dimensional bodies [cf. pp. 620–621; in particular Eq. (14a, b)]. For one-dimensional elastic bodies, he is naturally led to mentioning the Bernoulli-Euler problem of the elastica (pp. 667–668) with the only surviving material coordinate taken as the arc-length along the curve. One had to await the remarkable work by A. E. H. Love (later perfected by R. D. Mindlin) to correctly deduce a quasi-one dimensional dynamical theory of rods with the strange lateral inertia term (the print left by the asymptotic procedure in passing from three dimensions to the rod-like picture).

12.4.6 On Thermodynamics and Dissipative Behaviours

In his introduction (p. 604) Hellinger clearly expresses his opinion that the “mechanics of deformable media, as an autonomous discipline, comprises under formal statements, next to the usual theory of elasticity and hydrodynamics, all the related physical manifestations in the considered continuously extending bodies”. The development of these ideas has certainly been influenced by the discipline of thermodynamics which, in principle, tries “to embrace the totality of physics” (my translation). Here Hellinger is obviously influenced by his recent reading of “energetists” such as Pierre Duhem with his magisterial treatise of 1911 [12]. The latter may have been read by a handful of happy few.Footnote 5 What Hellinger tries in his Sect. 15 (pp. 682–695) is to incorporate the dual notions of entropy and thermodynamic temperature in his fundamental variational formulation. Entropy is considered as an extensive quantity (i.e., proportional to the quantity of matter). Then a term \(\delta Q\)—Equation (12.1) in p. 683—representing the “Wärmezufuhr” with variation of the entropy and co-factor none other than the temperature is to be added to the purely mechanical variation mentioned above at point 4.2. With the introduction of a potential for thermoelastic processes this yields the thermal definition of the temperature (in modern terms: the derivative of internal energy with respect to entropy) and, more surprisingly for the period, Maxwell’s compatibility condition for second-order derivatives of the energy in thermoelasticity in finite strains (Eq. (5) in p. 684; in modern notation this reads

$$\frac{{\partial {\mathbf{T}}}}{\partial S} = \frac{\partial \theta }{{\partial {\mathbf{F}}}},$$
(12.2)

where F is the deformation gradient and T is the first Piola-Kirchhoff stress).

In a more general context Hellinger comments on other coupled effects such as temperature and capillarity, pyro-electricity and thermo-chemical processes as considered by J. W. Gibbs in his original works of 1876–1878. He does not mention piezoelectricity although this is already more than thirty years old (experimental discovery by the Curie brothers in 1881) when he writes his contribution.

The above mentioned variational formulation that includes the notion of entropy and temperature is seldom considered. However, Sedov’s [46] generalized variational principle—also discussed in Maugin [35]—is along the same line.

For truly dissipative phenomena such as viscosity, in spite of his familiarity with Duhem’s treatise which does not propose yet a solution (the future “theory of irreversible processes”), Hellinger is reduced to invoking the notion of dissipation potential à la Rayleigh, as in the case of G.G. Stokes’ viscous fluids (cf. p. 671). But he is aware of the existence of more sophisticated models of viscosity. Such a model is the one proposed by Boltzmann [2] in the form—“elastischen Nachwirkung”—of hereditary integrals [see p. 641 and Eq. (5) in p. 672] for which Hellinger also cites very recent works, in particular by Vito Volterra up to year 1913 (the year Hellinger completed his manuscript). This shows the concern of this author to be up to date until the last moment. Finally, he also mentions the possible occurrence of a plastic behaviour with a simple plasticity criterion in terms of principal stresses which recalls the Tresca criterion—Inequalities (7) in p. 673—although he refers for these to a work of 1909 by A. Haar and Th. von Kármán. More general or singular behaviours are simply referred to as “halbplatische” oder “vollplastische” Zustände (no need for translation).

What is strange is that Hellinger does not comment on the then recent Caratheodory [6] axiomatization of thermodynamics as suggested by his own friend M. Born, a contribution that is purely in the analytical line and would certainly had been to Hellinger’s liking.

12.4.7 On Newly Studied Behaviours

This is just mentioned for the sake of completeness since hereditary materials, half plastic or fully plastic materials, are already evoked in the preceding paragraph. But Hellinger also pays some attention to the phenomenon of capillarity in pp. 674–675 for which a rather not commonly referenced work is by the mathematician of “relativity fame”, Herrmann Minkowski (see below).

Hellinger, although not pursuing the line further, gives the exact mathematical definition of material inhomogeneity (dependence of material properties on the material coordinates; see top of p. 639). General anisotropic elastic materials (crystals) with at most twenty one independent elasticity coefficients are mentioned for the linear case. In the case of finite strains, like all authors since Cauchy he focuses on the case of isotropy with the resulting introduction of the principal invariants of strains (p. 664) in the strain-energy function. This is purely academic as Hellinger and all other authors of the period could not guess that only rubber-like materials and then finitely-deformable soft biological tissues would provide in time the realm of application of this material description (see Maugin [38]).

12.4.8 On Relativistic Continuum Mechanics

In his attempt at a large conspectus of the State of the Art in 1913, Hellinger included (Sect. 16, pp. 685–694) comments on the most recent developments concerning the relativistic mechanics of continua. This is rather exceptional for the period; in particular if we compare with other well established authors in mechanics (e.g., Appell). This may have aroused his sensibility of mathematician. He seems to be well aware of the original developments by Voigt, Lorentz, and Poincaré on the group structure of special-relativistic transformations. The Lorentz-Poincaré group was a good subject of interest with the works of Minkowski [42], A. Sommerfeld, and F. Klein. His friend Max Born may have had some influence on Hellinger’s interest in the field since Born (especially, [4]) and Herglotz [28] seem to be his main sources for the basic definitions and the problem of the possibility of “rigid-body motion” in relativity.

Most of Hellinger’s discussion is about the essential differences between the Lorentz-Poincaré group and the Galilean-Newtonian group of space-time transformations. But he is also particularly interested in two points. One is the possible re-formulation of the Cosserats’ action principle in space-time in agreement with Minkowski and Herglotz (cf. Eq. (13a, b) in p. 693) with a space-time parametrization that combines material coordinates and a propertime (a parameter along the world line following Minkowski’s description) and a total virtual variation for internal forces (components of the energy-momentum tensor). The second point is the possible definition of the notion of rigid-body motion, a much discussed matter being given the existing bound on velocities, with the possible local (i.e., differential) solution given by Born and Herglotz in space-time. Allusion to relativistic continuum mechanics will later be given in a bibliographic appendix by Truesdell and Toupin [49, pp. 790–793]. The present writer is one of the very few to have devoted a full albeit brief chapter to relativistic continuum mechanics in a treatise (cf. Eringen and Maugin [16], Vol. 2, Chap. 15; see also the historical perspective in Maugin [38], Chap. 15).

12.5 Conclusion

In his introduction—written in 1913—Hellinger claims that there exists no textbook or monograph in the literature on the specific subject treated in his contribution although there do exist textbooks and treatises of a general nature, but the latter do not emphasize the bases and various possibilities offered by the scheme of continuous matter. He does not intend to treat applications and specific problems. He confines himself to the essentials, “die allgemeinen Ansätze” in his own words. His viewpoint is that analytical mechanics (exploitation of variational formulations) is “the most uniform and efficient manner to approach the general problem of describing a large variety of descriptions of deformable media” in agreement with recent authors like the Cosserats and the initial standpoint of G. Green with an energy potential. This is the type of approach (principle of virtual work, d’Alembert’s principle [9], Lagrange-Hamilton action principle [32], etc.) that suits best his essentially mathematical vision. The pregnant brevity of this approach possesses a “high heuristic value for the exploration of new areas. This is particularly stressed through the intimate relation of such variational principles with thermodynamics”. Furthermore, this allows one to place in evidence the invariant theoretical nature of the considered problems with the notion of transformation groups. This gives a very “modern” print that helps us understand his exposition without too much effort. This “modernity” is striking in spite of the somewhat obsolete notation. Its opens up horizons to many models that will have to wait progress in some branches of pure and applied mathematics for a full blossom (e.g., large deformations, media with internal degrees of freedom, capillarity, hereditary processes, multi-field phenomena).

His mathematical inclination leads him to accept unhesitatingly all new mathematical tools of the period (vector and tensor analysis, matrix calculus, differential geometry, perturbations). The only part that is still missing is convex analysis to be much developed in the 1950s–1970s. But, overall, Hellinger is very successful in his endeavour. This is our appraisal one hundred years later. Unfortunately, we were not able to locate any substantial review or criticism of his contribution in the few years following its publication so that we have no precise idea of the quality and extent of its reception among professional circles, mechanicians and mathematicians. This may exist in some periodical bulletin of a mathematical society. It is therefore with modern eyes, perhaps themselves influenced by Hellinger’s writing—a kind of feedback—that we evaluate it. This is an inevitable bias that we willingly acknowledge.

Following the Cosserat brothers, Hellinger’s view of the domain of interest of continuum mechanics is essentially the mechanics of deformable bodies, by which must be understood the case of deformable solids. This is in contrast with treatises by famous authors such as Appell [1], where most contents rather deal with fluids. At the time fluid mechanics has become a rather autonomous field of study limited to perfect fluids and the Navier-Stokes equations, with specific mathematical techniques of which the use of complex variables has become endemic. But some recent developments of theoretical fluid mechanics such as the asymptotic method involved in the theory of the boundary layer by L. Prandtl could have been to the taste of an analyst like Hellinger. It is only with the birth of the science of rheology (concerning whatever can flow to a larger or smaller extent) and the notion of non-Newtonian fluids in the 1920s in the expert hands of E. C. Bingham (1878–1945) and M. Reiner (1888–1976) that fluids will return to the general stage of continuum mechanics. Liquid crystals, with a behaviour clearly classified in 1922 by Georges Friedel (1865–1933) and exhibiting mixed crystal (ordered state) and fluid (flow) characteristics with directional properties, will also enter this general framework with a natural connection with generalized continua of the Duhem-Cosserat type established in the 1960s–1970s. This could not be imagined by Hellinger who remains essentially an analyst, as shown by his other very successful contribution to the same Encyclopaedia of mathematics in co-operation with a friend of student days in Breslau and Göttingen, O. Toeplitz (cf. Hellinger and Toeplitz [27]).

In conclusion, we find in Hellinger’s brilliant and very informative contribution all elements and remarks that we would like to deliver—even though superficially—to our mathematically oriented students in an introductory course of high level (e.g., something similar to what Germain tried to do in his course at Ecole Polytechnique, 1986 [20]); many students then thought that this was too much superficial, although all aspects of further developments in specialized short courses were outlined.