Open access peer-reviewed chapter

Gravitational Waves and Parametrizations of Linear Differential Operators

Written By

Jean-Francois Pommaret

Submitted: 14 November 2022 Reviewed: 20 November 2022 Published: 19 January 2023

DOI: 10.5772/intechopen.1000851

From the Edited Volume

Gravitational Waves - Theory and Observations

Carlos Frajuca

Chapter metrics overview

87 Chapter Downloads

View Full Metrics

Abstract

When D:ξη is a linear differential operator, a “direct problem” is to find the generating compatibility conditions (CC) in the form of an operator D1:ηζ such that Dξ=η implies D1η=0. Similarly, D1η=ζ may imply D2ζ=0 and so on. Conversely, when D1 is given, a much more difficult “inverse problem” is to look for an operator D:ξη with generating CC D1η=0. If this is possible, one shall say that the operator D1 is parametrized by D. The parametrization is “minimum” if the differential module defined by D does not contain any free differential submodule. The systematic use of the adjoint of a differential operator provides a constructive test. The parametrization of the Cauchy stress operator in arbitrary dimension n has attracted many famous scientists (G.B. Airy in 1863 for n=2, J.C. Maxwell in 1863, G. Morera and E. Beltrami in 1892 for n=3, A. Einstein in 1915 for n=4). We prove that all these works are already explicitly using the self-adjoint Einstein operator, which cannot be parametrized, and are thus all based on a confusion between the Cauchy operator, (adjoint of the Killing operator D), and the div operator induced from the Bianchi operator D2 CC of the Riemann operator D1 parametrized by D. This purely mathematical result deeply questions the origin and existence of gravitational waves that are solutions of the adjoint of the Ricci operator. We do believe that Einstein was aware of these previous works as the comparison needs no comment. The same methods are also used in order to revisit the mathematical foundations of electromagnetism.

Keywords

  • differential sequence
  • Killing operator
  • Riemann operator
  • Bianchi operator
  • general relativity
  • gravitational waves
  • Maxwell equations

1. Introduction

The problem of parametrizing the Einstein operator or, equivalently and by analogy with Maxwell equations for electromagnetism (EM), to decide about the existence of a potential for Einstein equations in vacuum, has been proposed for the first time as a 1000 dollars challenge by J. Wheeler while the author of this paper was a visiting student of D. C. Spencer in 1970 at Princeton university. No progress at all has been done during the next 25 years, till the author gave a negative answer in 1995, contrary to what the general relativity (GR) community was believing [1, 2]. Indeed, after teaching elasticity for 25 years to high-level students in some of the best french civil engineering schools, the author of this paper proposed an exercise explaining why a dam made with concrete is always vertical on the water side with a slope of about 42 degrees on the other free side in order to obtain a minimum cost and the auto-stability under cracking of the surface underwater (See ([3], p. 108) and the introduction of [4] for more details). Surprisingly, the main tool involved is the approximate computation of the Airy function inside the dam in this two-dimensional elasticity problem. The author discovered at that time that no one of the other teachers did know that the Airy parametrization was nothing else than the adjoint of the linearized Riemann operator used as generating CC for the deformation tensor by any engineer. Being involved in GR with A. Lichnerowicz at that time, he got for the first time the idea of using the adjoint of an operator in a systematic way. Giving a seminar in Paris in order to present this result, somebody in the audience told him about a possible link with the recently published master thesis of the Japanese student M. Kashiwara [5]. It has been a shock to discover this mixing up of differential geometry [6, 7] and homological algebra [8, 9, 10], now called “Differential Homological Algebra”, in particular the introduction of the Differential Extension Modules (See [4, 5, 11, 12, 13] for extensive references) (See also Zbl 1079.93001 for comments). It is only recently that he discovered GR could be considered as a way to parametrize the Cauchy operator and to introduce gravitational waves (GW) [14, 15]. It follows that exactly the same confusion has been done by Maxwell, Morera, Beltrami and Einstein because, in all these cases, the operator considered is self-adjoint. However, most of the pure mathematicians involved were proud not to be interested by applications and in any case unable to do any computation. Accordingly and until now, the GR community has never wanted to take these new tools into account and Ref. [16] is providing a good example of such a poor situation. This is the reason for which we have not been able to provide any other reference and why all the results presented are new.

For example, the fact that the Cauchy operator is the adjoint of the Killing operator for the Euclidean metric is apparently in the chapter “variational calculus” of any textbook of continuum mechanics, and the parametrization problem has been quoted by many famous authors, as we said in the Abstract, but only from a computational point of view. The same comment can be done for the two sets of Maxwell equations in electromagnetism [4, 13]. However, it is still not known that the adjoint of the 20 components of the Bianchi operator has been introduced by C. Lanczos between 1939 and 1962 [17] as we explained with details in [18] by using Spencer cohomology. The main trouble is that these two problems have never been treated in an intrinsic way and, in particular, changes of coordinates have never been considered. The same situation can be met for the Vessiot structure equations but is out of the scope of this paper [12, 19].

Lemma 1.1. When yk=fkx is invertible with Δx=detifkx0 and inverse x=gy, then we have n identities yk1Δifkgy=0 (See [4] p 490 and [20] for other applications).

Proof: Using the chain rule for derivatives, we get iΔ=ijfk cofactor jfk=ijfkΔgjyk and thus:

1Δijfkgjyk1Δ2ifkgjykjΔ=1Δijfkgjyk1Δ2iΔ=0

Proposition 1.2. The Cauchy operator is the adjoint of the Killing operator in arbitrary dimension n, up to sign. Similarly, when n=4, the Maxwell operator 4T2Tadd4TT:FijiFij=Jj is the adjoint of the parametrizing operator Td2T:AdA=F in electromagnetism (EM), independently of the Minkowski constitutive relations FF.

Proof: Let X be a manifold of dimension n with local coordinates x1xn, tangent bundle T, and cotangent bundle T. If ωS2T is a metric with detω0, we may introduce the standard L, that is, derivative in order to define the first order Killing operator ξLξω, namely:

D:ξTΩ=Ωij=ωrjxiξr+ωirxjξr+ξrrωijxS2TE1

Here starts the problem because, in our opinion at least, a systematic use of the (formal) adjoint operator has never been done in mathematical physics (continuum mechanics, EM, …) apart from a variational procedure. As will be seen later on, the purely intrinsic definition of the adjoint can only be done in the theory of differential modules by means of the so-called side-changing functor. From a purely differential geometric point of view, the idea is to associate to any vector bundle E over X, a new vector bundle adE=nTE, where E is obtained from E by patching local coordinates while inverting the transition matrices, exactly like T is obtained from T in tensor calculus. It follows that the stress tensor σ=σijadS2T=nTS2T is not a tensor but a tensor density, that transforms like a tensor up to a certain power of the Jacobian matrix. When n=4, the fact that such an object is called stress-energy tensor does not change anything as it cannot be related to the Einstein tensor which is a true tensor indeed. Of course, it is always possible in GR to use detω12 but, as we shall see, the study of contact structures must be done without any reference to a background metric. In any case, we may define as usual:

adD:nTS2TnTT:σφE2

Multiplying Ωij by σij and integrating by parts, the factor of 2ξk is easily seen to be:

iσik=iσik+γijkσij=φkE3

with well known Christoffel symbols γijk=12ωkriωrj+jωirrωij.

However, if the stress should be a tensor, we should get for the covariant derivative:

rσij=rσij+γrsiσsj+γrsjσisiσik=iσik+γrirσik+γijkσij

The difficulty is to prove that we do not have a contradiction because σ is a tensor density.

If we have an invertible transformation like in the lemma, we have successively by using it:

τklfx=1Δifkxjflxσijxτklyk=yk1Δifkjflσij+1Δifkykjflσij+1Δifkjflykσijτkuyk=1Δijfuσij+1Δjfuiσij

Now, we recall the transformation law of the Christoffel symbols, namely:

rfuxγijrx=ijfux+ifkxjflxγ¯klufx1Δrfuγijrσij=1Δijfuσij+γ¯kluyτkl

Eliminating the second derivatives of f, we finally get:

ψu=τkuyk+γ¯kluτkl=1Δrfuiσir+γijrσij=1Δrfuφr

This tricky technical result, which is not evident, explains why the additional term we had is just disappearing, in fact, when σ is a density.

The case of EM is even simpler because ijfuFij=jifuFji=ijfuFij=0 and γ is not needed. The two sets of Maxwell equations are thus separately invariant by any diffeomorphism. Though surprising it may look like, the conformal group of space–time is only the maximum group of invariance of the Minkowski constitutive law in vacuum. Indeed, this law is not at all Fij=μ0ωirωjsFrs, where μ0 is the magnetic constant because such a relation is not tensorial as F is a 2-form, that is, a 2-covariant tensor, but F is a 2-contravariant tensor density. Hence, introducing the metric density ω̂ij=detω1/nωij, we must set Fij=μ0ω̂irω̂jsFrs. Accordingly, this constitutive law is only invariant by diffeomorphisms preserving ω̂, and this is exactly the definition of the Lie pseudogroup of conformal transformations [13].

By chance, the control community has been interested during a while by these new techniques for dealing with PD control theory but mostly restricting to operators with constant coefficients [4, 20, 21]. The following example, coming from partial differential (PD) control theory, will allow the reader to become familiar with these new tools and to understand why they are related to the mathematics of GW. Accordingly, the “relative” parametrization of the Cauchy stress operator has thus nothing to do with the mathematical background of elasticity theory. In particular, the way to simplify the GW equations by bringing them back to the d’Alembert operator while adding a few differential constraints (Compare [14, 15, 22, 23, 24]) through a reference to the so-called “gauge invariance” is rather a physical argument and not a mathematical one.

Example 1.3. Let us consider the first order operator with two independent variables x1x2:

D1:η1η2d2η1d1η2+x2η2=ζE4

The non-commutative ring of differential operators involved in a formal study is D=Kd1d2 with K=Qx1x2=Qx and the characters of this involutive system are α11=2α12=1 with β12=21=1 as there is only one equation. Multiplying on the left by a test function λ and integrating by parts, the corresponding adjoint operator is described by:

adD1:λd2λ=μ1d1λ+x2λ=μ2

Using crossed derivatives, this operator is injective because λ=d2μ2+d1μ1+x2μ1 and we even obtain a lift id:λμλ. Substituting, we get the second order involutive operator adD:μ1μ2ν1ν2 with characters α21=1α22=1, namely:

d22μ2+d12μ1d12μ2+d11μ1+x2d2μ1+2μ1=ν1+2x2d1μ1+x2d2μ2+x22μ1μ2=ν2121

allowing to define a second order operator D by using the fact that adadD=D. This operator is involutive and the only corresponding generating CC is d2ν2d1ν1x2ν1=0. Therefore, ν2 is differentially dependent on ν1 but ν1 is also differentially dependent on ν2. Multiplying on the left by a test function θ and integrating by parts, the corresponding adjoint operator is:

D1:θd1θx2θ=ξ1d2θ=ξ2

Multiplying now the first equation of adD by the test function ξ1, the second equation by the test function ξ2, adding and integrating by parts, we get the second order operator:

D:ξ1ξ2d22ξ1+d12ξ2d12ξ1+d11ξ2x2d2ξ22ξ2=η2x2d2ξ12x2d1ξ2+ξ1+x22ξ2=η1121E5

which is easily seen to be a parametrization of D1. This operator is involutive and the kernel of this parametrization has differential rank equal to 1 because ξ1 or ξ2 can be given arbitrarily.

As we are using the Lagrange multiplier ξ1ξ2, we consider in fact the PD equation ξ1ν1+ξ2ν2. Hence, we could indeed consider each term separately, that is using independently each equation as we shall see later on, ν1 for a certain ξ and ν2 for a certain ξ'. Equivalently (exactly like Morera and Maxwell did as we shall see later on), keeping ξ1=ξ while setting ξ2=0, we now obtain the first second order minimal parametrization

ξd22ξ=η2d12ξx2d2ξ+ξ=η1E6

This system is again involutive, and the parametrization is minimal because the kernel of this parametrization has differential rank equal to 0. With a similar comment, setting now ξ1=0 while keeping ξ2=ξ, we get the second second order minimal parametrization:

ξd12ξx2d2ξ2ξ=η2d11ξ2x2d1ξ+x22ξ=η1,E7

which is again easily seen to be involutive by exchanging x1 with x2.

With again a similar comment, setting now ξ1=d1ϕ,ξ2=d2ϕ in the canonical parametrization, we obtain the third different second order minimal parametrization:

ϕx2d22ϕ+2d2ϕ=η2x2d12ϕx22d2ϕ+d1ϕ=η1E8

We are now ready for understanding the meaning and usefulness of what we have defined and called “relative parametrization” in [24] by imposing the differential constraint d2ξ1+d1ξ2=0. First of all, we have to prove that such a constraint is compatible. For this, taking into account the constraint, we have the following first order system defined over K:

d2ξ1+d1ξ2=0d2ξ2+2x2ξ2=1x2η2d1ξ2x2ξ21x2ξ1=1x2η112121

We let the reader prove that this first order system is involutive with full class 2 and characters α11=1α12=0 and that the only CC involved is the initial system for η (The reader will discover that this checking is quite harder that what one could believe on such an elementary example).

We obtain therefore the new first order relative parametrization:

ξ1ξ2x2d2ξ22ξ2=η2x2d2ξ1+x22ξ2+ξ1=η1modd2ξ1+d1ξ2=0

In a different way, we may add the differential constraint d1ξ1+d2ξ2=0 but we have to check similarly that it is compatible with the previous parametrization. For this, we have to consider the following second order involutive system with five equations, which are easily seen to be involutive, obtained by adding the constraint and its two derivatives to the system like before:

d22ξ2+d12ξ1=0d22ξ1+d12ξ2x2d2ξ22ξ2=η2d12ξ2+d11ξ1=0d12ξ1+d11ξ2x2d2ξ12x2d1ξ2+ξ1+x22ξ2=η1d2ξ2+d1ξ1=0121211

The four generating CC only produce the desired system for η1η2 as we wished.

We could not impose the condition D1θ=ξ already found as it should give the identity 0=η.

It is, however, also important to notice that the long strictly exact sequence [25]:

0θD1ξDηD1ζ0

splits because we have a lift id:ζ1ζ+x2ζ=η12η2=η22η11η2+x2η2=ζ.

All the differential modules defined from the operators involved are projective, thus torsion-free, and we notice that D1 is parametrized by D which is again parametrized by D1, exactly like div is parametrized by curl which is again parametrized by grad in vector geometry. Needless to say that such an approach has nothing to do with Lorenz gauge invariance in electromagnetism (EM) and we shall arrive to the same conclusion for GW in GR.

Going on along the historical survey 50 years ago, while the author of this paper was working in GR under the leadership of Prof. A. Lichnerowicz, he became familiar with the Lanczos problems. Since that time, he had no wish at all to enter this kind of game as this domain became the private garden of a few persons, each one writing after another one alternatively, claiming to have the full solution. Also, the papers were covered with “computations” involving awful technical formulas, one paper using Gröbner bases, another computer algebra, another Cartan exterior calculus or Janet bases and so on during these 50 years.

Later on, in 2001 and for different reasons, namely control theory as we just explained, being more familiar with differential homological algebra and the so-called “parametrization problem,” the way towards the Lanczos problems became easier as follows [18]. In dimension four, the only considered by Lanczos, the Lanczos “potential” Lij,k=Lji,k has 6×4=24 components. As they must be related by the four relations Lij,k+Ljk,i+Lki,j=0, we get 20 independent components, namely the number of (second) Bianchi identities. However, who is speaking about “potential” means “parametrization”, … of what?. Here comes the confusion of Lanczos, too much familiar with electromagnetism (EM) while using mainly quadratic Lagrangians with the Riemann tensor in place of the EM field F and the Bianchi identities as differential constraint in the corresponding variational calculus with constraint. The operator to be parametrized was thus the adjoint of the Riemann operator that we called Beltrami operator while the parametrizing operator, that we called Lanczos operator, was just the adjoint of the Bianchi operator, going now backwards, that is from right to left in the adjoint sequence of the Killing resolution presented in the abstract. As for the extension to the conformal framework, it is clear that Lanczos did not even know the Weyl tensor when he lectured in France in 1962, invited by Lichnerowicz. Moreover, it is only recently in 2016 that the author of this paper proved that the analogue of the Bianchi identities is made by an operator of order two when n=4, such a result being tested through computer algebra by his former PhD student A. Quadrat (See [12, 13] and arXiv:1603.05030 for more details). Such a construction, based on difficult results of homological algebra, has been missed by Lanczos and all followers as such tools were only available after 1995 through the works of pure mathematicians not interested by applications. Therefore, the main idea is to replace technical formulas by diagram chasing without any formula. As a byproduct, we shall understand, without any computation, the confusion done between the Cauchy operator, adjoint of the Killing operator, and the Bianchi operator as explained in the abstract. We shall explain its historical origin as the names of many celebrated scientists are involved in this confusion as we also said in the abstract.

The story ended with a strange letter sent by J. Wheeler back to the author with a one dollar bill attached, refusing to admit the negative answer to his parametrization challenge and claiming that future quantum GR should find a positive answer (!). As a byproduct, the impossibility to parametrize Einstein equations in vacuum can only be found in books of control theory [20, 26].

After this rather historical introduction, the content of the paper is clear:

The second section presents the mathematical tools that are absolutely needed while the third section is dealing with the solution of the parametrization problem. The applications to Einstein equations and the corresponding GW equations is finally presented in the fourth section before concluding. In any case, we want to point out that no one of these methods have ever been used in GR, in particular for the study of GW.

Advertisement

2. Mathematical tools

We start recalling the basic tools from the formal theory of systems of ordinary differential (OD) or partial differential (PD) equations and differential modules needed in order to understand and solve the parametrization problem presented in the abstract. Then we provide the example of the system of infinitesimal lie equations defining contact transformations and conclude the paper with the general parametrization problem existing in continuum mechanics and general relativity for an arbitrary dimension of the ground manifold. As these new tools are difficult and not so well known as we already said, we advise the interested reader to follow them step by step on the explicit motivating examples illustrating this paper, while referring to [4, 12, 20, 22, 25, 27, 28, 29, 30] for more details, even though the paper is rather self-contained and uses standard notations. Parts of the present paper have already been published independently with slight differences in [2, 12, 13, 18, 20, 21, 22, 23, 24, 31] but the present paper is mainly revisiting the mathematical foundations of GW in the framework of differential homological algebra.

2.1 System theory

If X is a manifold of dimension n with local coordinates x=x1xn, we denote as usual by T=TX the tangent bundle of X, by T=TX the cotangent bundle, by rT the bundle of r-forms and by SqT the bundle of q-symmetric tensors. More generally, let E be a vector bundle over X with local coordinates xiyk for i=1,,n and k=1,,m simply denoted by xy, projection π:EX:xyx and changes of local coordinate x¯=φx,y¯=Axy. We shall denote by E the vector bundle obtained by inverting the matrix A of the changes of coordinates, exactly like T is obtained from T. We denote by f:XE:xxy=fx a global section of E, that is a map such that πf=idX but local sections over an open set UX may also be considered when needed. Under a change of coordinates, a section transforms like f¯φx=Axfx and the changes of the derivatives can also be obtained with more work. We shall denote by JqE the q-jet bundle of E with local coordinates xiykyikyijk=xyq called jet coordinates and sections fq:xxfkxfikxfijkx=xfqx transforming like the sections jqf:xxfkxifkxijfkx=xjqfx where both fq and jqf are over the section f of E. For any q0, JqE is a vector bundle over X with projection πq, while Jq+rE is a vector bundle over JqE with projection πqq+r,r0.

Definition 2.A.1: A linear system of order q on E is a vector sub-bundle RqJqE and a solution of Rq is a section f of E such that jqf is a section of Rq. With a slight abuse of language, the set of local solutions will be denoted by ΘE.

Let μ=μ1μn be a multi-index with length μ=μ1++μn, class i if μ1==μi1=0,μi0 and μ+1i=μ1μi1μi+1μi+1μn. We set yq=yμk1km0μq with yμk=yk when μ=0. If E is a vector bundle over X and JqE is the q-jet bundle of E, then both sections fqJqE and jqfJqE are over the section fE. There is a natural way to distinguish them by introducing the Spencer operator d:Jq+1ETJqE with components dfq+1μ,ikx=ifμkxfμ+1ikx. The kernel of d consists of sections such that fq+1=j1fq=j2fq1==jq+1f. Finally, if RqJqE is a system of order q on E locally defined by linear equations Φτxyqakτμxyμk=0 and local coordinates xz for the parametric jets up to order q, the r-prolongation Rq+r=ρrRq=JrRqJq+rEJrJqE is locally defined when r=1 by the linear equations Φτxyq=0,diΦτxyq+1akτμxyμ+1ik+iakτμxyμk=0 and has symbol gq+r=Rq+rSq+rTEJq+rE if one looks at the top order terms. If fq+1Rq+1 is over fqRq, differentiating the identity akτμxfμkx0 with respect to xi and substracting the identity akτμxfμ+1ikx+iakτμxfμkx0, we obtain the identity akτμxifμkxfμ+1ikx0 and thus the restriction d:Rq+1TRq. More generally, we have the restriction:

d:sTRq+1s+1TRq:fμ,IkxdxIifμ,Ikxfμ+1i,IkxdxidxIE9

with standard multi-index notation for exterior forms and one can easily check that dd=0. The restriction of d to the symbol is called the Spencer map δ in the sequences:

s1Tgq+r+1δsTgq+rδs+1Tgq+r1E10

because δδ=0 similarly, leading to the purely algebraic δ-cohomology Hq+rsgq at sTgq+r (See [4, 6, 7, 12, 13, 25, 27, 28, 29, 31] for details and examples).

Definition 2.A.2: A system Rq is said to be formally integrable (FI) when all the equations of order q+r are obtained by r prolongations only, r0 or, equivalently, when the projections πq+rq+r+s:Rq+r+sRq+rsRq+r are such that Rq+rs=Rq+r, r,s0.

Finding an intrinsic test has been achieved by D.C. Spencer in 1965–1970 [6, 7] along coordinate dependent lines sketched by M. Janet in 1920 [20, 25, 29]. The next procedure providing a Pommaret basis and where one may have to change linearly the independent variables if necessary, is intrinsic though it must be checked in a particular coordinate system called δ-regular [4, 25, 27, 28, 29].

  • Equations of class n: Solve the maximum number βqn of equations with respect to the jets of order q and class n. Then, call x1xn multiplicative variables.

  • Equations of class i1: Solve the maximum number βqi of remaining equations with respect to the jets of order q and class i. Then, call x1xi multiplicative variables and xi+1xn non-multiplicative variables.

  • Remaining equations equations of order q1: Call x1xn non-multiplicative variables.

In actual practice, we shall use a Janet tabular where the multiplicative “variables” are in upper left position while the non-multiplicative variables are represented by dots in lower right position.

Definition 2.A.3: A system of PD equations is said to be involutive if its first prolongation can be obtained by prolonging its equations only with respect to the corresponding multiplicative variables. In that case, we may introduce the characters αqi=mq+ni1!q1!(ni!βqi for i=1,,n with αq1αqn0 and we have dimgq=αq1++αqn while dimgq+1=αq1++nαqn.

Remark 2.A.4: As long as the prolongation/projection (PP) procedure allowing to find two integers r,s0 such that the system Rq+rs is involutive, has not been achieved, nothing can be said about the CC (Fine examples can be found in [12, 16]).

When Rq is involutive, the operator D:EjqJqEΦJqE/Rq=F0 of order q is said to be involutive. Introducing the Janet bundles Fr=rTJqE/rTRq+δSq+1TE, we obtain the linear Janet sequence (Introduced in [25]):

0ΘEDqF0D11F1D21Dn1Fn0E11

where each other operator is first order involutive and generates the CC of the preceding one.

Similarly, introducing the Spencer bundles Cr=rTRq/δr1Tgq+1, we obtain the linear Spencer sequence induced by the Spencer operator, that can be linked to the Janet sequence [25, 29]:

0ΘjqC0D11C1D21Dn1Cn0E12

It must be noticed, as we shall see in Section 4, that the Killing operator/system is FI if and only if the metric ω has constant Riemanniann curvature (for example the flat Minkowski metric) but is not involutive and the Janet sequence cannot be exhibited. As for the conformal Killing operator/system obtained by eliminating the arbitrary function Ax in the inhomogeneous system Lξω=Axω or by considering simply the system Lξω̂=0, it is FI if and only if ω has vanishing Weyl tensor but is not involutive and the order of the successive CC operators may change with the ground dimension n, the worst situation being for n=4 as the analogue of the Bianchi operator is now an operator of order two indeed [12, 13]. As these results highly depend on the Spencer δ-cohomology, it is clear that they are neither known nor acknowledged and it follows that the mathematical foundations of conformal geometry must be entirely revisited.

2.2 Module theory

Let K be a differential field with n commuting derivations 1n and consider the ring D=Kd1dn=Kd of differential operators with coefficients in K with n commuting formal derivatives satisfying dia=adi+ia in the operator sense. If P=aμdμD=Kd, the highest value of μ with aμ0 is called the order of the operator P and the ring D with multiplication PQPQ=PQ is filtred by the order q of the operators. We have the filtration 0K=D0D1DqD=D. As an algebra, D is generated by K=D0 and T=D1/D0 with D1=KT if we identify an element ξ=ξidiT with the vector field ξ=ξixi of differential geometry, but with ξiK now. It follows that D=DDD is a bimodule over itself, being at the same time a left D-module by the composition PQP and a right D-module by the composition PPQ. We define the adjoint functor ad:DDop:P=aμdμadP=1μdμaμ and we have adadP=P both with adPQ=adQadP,P,QD. It follows that any operator can be considered as the adjoint of its own adjoint. Such a definition can be extended by linearity to any matrix of operators by using the transposed matrix of adjoint operators (See [4, 5, 11, 12, 13, 18, 20, 21, 24] for more details and applications to control theory or mathematical physics).

Accordingly, if y=y1ym are differential indeterminates, then D acts on yk by setting diyk=yikdμyk=yμk with diyμk=yμ+1ik and y0k=yk. We may therefore use the jet coordinates in a formal way as in the previous section. Therefore, if a system of OD/PD equations is written in the form Φτakτμyμk=0 with coefficients aK, we may introduce the free differential module Dy=Dy1++DymDm and consider the differential module of equations I=DΦDy, both with the residual differential module M=Dy/DΦ or D-module and we may set M=DM if we want to specify the ring of differential operators. We may introduce the formal prolongation with respect to di by setting diΦτakτμyμ+1ik+iakτμyμk in order to induce maps di:MM:y¯μky¯μ+1ik by residue with respect to I if we use to denote the residue DyM:yky¯k by a bar like in algebraic geometry. However, for simplicity, we shall not write down the bar when the background will indicate clearly if we are in Dy or in M. As a byproduct, the differential modules we shall consider will always be finitely generated (k=1,,m<) and finitely presented (τ=1,,p<). Equivalently, introducing the matrix of operators D=akτμdμ with m columns and p rows, we may introduce the morphism DpDDm:PτPτΦτ over D by acting with D on the left of these row vectors while acting with D on the right of these row vectors by composition of operators with imD=I. The presentation of M is defined by the exact cokernel sequence DpDDmM0. We notice that the presentation only depends on K,D and Φ or D, that is to say never refers to the concept of (explicit local or formal) solutions. It follows from its definition that M can be endowed with a quotient filtration obtained from that of Dm which is defined by the order of the jet coordinates yq in Dqy. We have therefore the inductive limit 0M0M1MqM=M with diMqMq+1 and M=DMq for q0 with prolongations DrMqMq+r,q,r0. It may be sometimes quite difficult to work out Iq or Mq from a given presentation which is not involutive [4].

Definition 2.B.1: An exact sequence of morphisms finishing at M is said to be a resolution of M. If the differential modules involved apart from M are free, that is isomorphic to a certain power of D, we shall say that we have a free resolution of M.

Having in mind that K is a left D-module with the action DKK:diaia and that D is a bimodule over itself with PQQP, we have only two possible constructions:

Definition 2.B.2: We may define the right (care) differential module homDMD with fPm=fmPfPQm=fPmQ=fmPQ=fmPQ.

Definition 2.B.3: We define the system R=homKMK and set Rq=homKMqK as the system of order q. We have the projective limit R=RRqR1R0. It follows that fqRq:yμkfμkK with akτμfμk=0 defines a section at order q, and we may set f=fR for a section of R. For an arbitrary differential field K, such a definition has nothing to do with the concept of a formal power series solution (care).

Proposition 2.B.4: When M is a left D-module, then R is also a left D-module.

Proof: As D is generated by K and T as we already said, let us define:

afm=afm=fam,aK,mMξfm=ξfmfξm,ξ=aidiT,mM

In the operator sense, it is easy to check that dia=adi+ia and that ξηηξ=ξη is the standard bracket of vector fields. We finally get difμk=difyμk=ifμkfμ+1ik and thus recover exactly the Spencer operator of the previous section though this is not evident at all. We also get didjfμk=ijfμkifμ+1jkjfμ+1ik+fμ+1i+1jkdidj=djdi,i,j=1,,n and thus diRq+1RqdiRR induces a well defined operator RTR:fdxidif. This operator has been first introduced, up to sign, by F.S. Macaulay as early as in 1916 but this is still not ackowledged. For more details on the Spencer operator and its applications, the reader may look at [12, 13, 25, 27, 28, 29, 30].

Definition 2.B.5: With any differential module M, we shall associate the graded module G=grM over the polynomial ring grDKχ by setting G=q=0Gq with Gq=Mq/Mq1, and we get gq=Gq where the symbol gq is defined by the short exact sequences:

0Mq1MqGq00gqRqRq10

We have the short exact sequences 0Dq1DqSqT0 leading to grqDSqT and we may set as usual T=homKTK in a coherent way with differential geometry.

The two following definitions, which are well known in commutative algebra, are also valid (with more work) in the case of differential modules (See [4, 20] for more details or the references [4, 8, 9, 10, 12, 13, 29] for an introduction to homological algebra and diagram chasing).

Definition 2.B.6: The set of elements tM=mM0PDPm=0M is a differential module called the torsion submodule of M. More generally, a module M is called a torsion module if tM=M and a torsion-free module if tM=0. In the short exact sequence 0tMMM0, the module M is torsion-free. Its defining module of equations I is obtained by adding to I a representative basis of tM set up to zero and we have thus II.

Definition 2.B.7: A differential module F is said to be free if FDr for some integer r>0 and we shall define rkDF=r. If F is the biggest free differential module contained in M, then M/F is a torsion differential module and homDM/FD=0. In that case, we shall define the differential rank of M to be rkDM=rkDF=r. Accordingly, if M is defined by a linear involutive operator of order q, then rkDM=αqn.

Proposition 2.B.8: If 0MMM0 is a short exact sequence of differential modules and maps or operators, we have rkDM=rkDM+rkDM.

In the general situation, let us consider the sequence MfMgM of modules which may not be exact. Then, we may define the coboundary submodule B=imfM, the cocycle submodule Z=kergM and the cohomology module H=Z/B.

Using the last (delicate) proposition, we may provide the following definitions that will be in the heart of the parametrization problem, successively in the operator and module frameworks.

Definition 2.B.9: When D=Φjq:EF is a linear differential operator of order q with coefficients in a differential field K, between the sections of two vector bundles E with dimE=m and F with dimF=p, we shall define in a formal way the differential rank rkDD=mαqn=βqn=rkDimDp by introducing the characters of the corresponding linear system Rq=kerΦJqE of order q over E. We have thus rkDDinfmp (See [4, 20] for details). The order of an operator will be indicated under it arrow.

Definition 2.B.10: When a differential module M is defined by the presentation DpDqDmM0, we shall introduce the differential module I=imDDm and set rkDD=rkDI=mrkDM in a coherent way with the last proposition and definition.

We obtain the important theorem which is generalizing to operators the rank property of a m×p matrix, even when D and adD are neither FI nor involutive ([4, 20, 29], p. 340):

Theorem 2.B.11: One has rkDD=rkDadDinfmp.

In order to conclude this section, we may say that the main difficulty met when passing from the differential framework to the algebraic framework is the “inversion” of arrows. Indeed, when an operator is injective, that is when we have the exact sequence 0EDF with dimE=m, dimF=p, like in the case of the operator 0EjqJqE, on the contrary, using differential modules, we have the epimorphism DpDDm0. The case of a formally surjective operator, like the div operator, described by the exact sequence EDF0 is now providing the exact sequence of differential modules 0DpDDmM0 because D has no CC.

Advertisement

3. Parametrization problem

In this section, we shall set up and solve the minimum parametrization problem by comparing the differential geometric approach and the differential algebraic approach. In fact, both sides are essential because certain concepts, like “torsion”, are simpler in the module approach while others, like “involution” are simpler in the operator approach. However, the reader must never forget that the “extension modules” or the “side changing functor” are pure product of differential homological algebra with no system counterpart. The link between “differential duality” and the “adjoint operator” may not be evident at all, even for people familiar with mathematical physics [2, 4, 12, 13, 18, 22, 29].

Let us start with a given linear differential operator ηD1ζ between the sections of two given vector bundles F0 and F1 of respective fiber dimension m and p. Multiplying the equations D1η=ζ by p test functions λ considered as a section of the adjoint vector bundle adF1=nTF1 and integrating by parts as we did in the introduction, we may introduce the adjoint vector bundle adF0=nTF0 with sections μ in order to obtain the adjoint operator μadD1λ, writing on purpose the arrow backwards. More generally, let us consider a differential sequence:

ξDηD1ζE13

such that D1 generates the CC of D or, equivalently, such that D1 is parametrized by D.

We may introduce the adjoint differential sequence:

νadDμadD1λE14

As we have D1D=0, we obtain adDadD1=0. However, if D1 generates the CC of D, then adD may not generate the CC of adD1. Such a situation may not be satisfied as we saw and the so-called extension modules have been introduced in order to measure these “gaps.”

In order to pass to the differential module framework, let us introduce the free differential modules Dl,Dm,Dp. We have similarly the adjoint free differential modules Dl,Dm,Dp, because dimadE=dimE and homDDmDDm. Of course, in actual practice, the geometric meaning is totally different because we have volume forms in the dual framework. We have thus obtained the formally exact sequence of differential modules:

DpD1DmDDlM0M100E15

with rkDD+rkDD1=mrkDM+rkDM1=l. We have the adjoint sequence:

DpadD1DmadDDl

with rkDadD+rkDadD1=m and we may thus state [4, 9, 10, 20]:

Definition 3.1: We may define the zero differential extension module ext0M=keradD and the first differential extension module ext1M to be the cohomology of this sequence at Dm. The latter is a torsion module because it has vanishing differential rank. They only depend on M.

Theorem 3.2: There is a constructive test in order to find out whether a differential operator D1 can be parametrized or not (Example 1.3 or the div operator with n=3).

Proof: The test has five steps along with the following diagram in operator language:

1 Start with D1:ηζ,

2 Construct adD1:λμ,

3 Construct its CC as an operator adD:μν,

4 Construct its adjoint D=adadD:ξη,

5 Construct the CC D1 of D:ηζ.

ζ5D14ξDηD1ζ13νadDμadD1λ2

We have adDadD1=0D1D=0, that is D1 is surely among the CC of D but other CC may also exist. The parametrization is thus existing if and only if we may have D1=D1.

Corollary 3.3: Each new CC eventually brought by D which is not already a differential consequence of D1 is providing a torsion element of the differential module M1 determined by D1.

Example 3.4: If D:ξd22ξ=η2d12ξ=η1 we have D1=η1η2d1η2d2η1=ζ and the only first order generating CC of adD1:λd2λ=μ1d1λ=μ2 is d1μ1+d2μ2=ν' while adD:μ1μ2d12μ1+d22μ2=ν is a second order operator like D. Hence, if we should like to parametrize adD, we should successively find D, D1, adD1 and finally get the additional first order CC d1μ1+d2μ2=ν' which is such that ν=0d2ν'=0, that is ν' is a torsion element for the system d12μ1+d22μ2=0.

Example 3.5: Many other examples can be found in ordinary differential control theory because it is known that a linear control system is controllable if and only if it is parametrizable (See [4, 20, 29] for more details and examples). In our opinion, the simplest one is provided by the double pendulum in which a rigid bar is able to move horizontally with reference position x and has two pendulums attached with respective length l1 and l2 making the (small) angles θ1 and θ2 with the vertical. The corresponding control system does not depend on the mass of each pendulum and is:

d2x+l1d2θ1+gθ1=0,d2x+l2d2θ2+gθ2=0

where g is the gravity. The classical approach is to prove that this control system is controllable if and only if l1l2 by using a tedious computation through the standard Kalman test [20]. However, equivalently, but this way is still not acknowledged by the control community, the idea is to prove that the corresponding second order operator adD1 is injective. We let the reader realize the experiment, prove this result as an exercise and apply the previous theorem in order to work out the parametrizing operator D of order 4, namely:

l1l2d4ϕgl1+l2d2ϕg2ϕ=xl2d4ϕ+gd2ϕ=θ1l1d4ϕ+gd2ϕ=θ2

Example 3.6: As a less academic example, the following diagram is proving that Einstein equations cannot be parametrized [1, 2]:

10Riemann20Bianchi204Killing10Einstein10div4004Cauchy10Einstein10

and that the Cauchy and Killing operators (left side) have strictly nothing to do with the Bianchi and div operators (right side). According to the last corollary, the 2010=10 new CC are generating the torsion submodule of the differential module defined by the Einstein operator. In the last section we shall explain why such a basis of the torsion module is made by the 10 independent components of the Weyl tensor, a result which is not evident, leading to the so-called Lichnerowicz waves (in France) [12, 13, 22, 32, 33].

In continuum mechanics, the Cauchy stress tensor may not be symmetric in the so-called Cosserat media where the Cauchy stress equations are replaced by the Cosserat couple-stress equations which are nothing else than the adjoint of the first Spencer operator, totally different from the third [27, 28, 34, 35]. When n=2, we shall see that the single Airy function has strictly nothing to do with any perturbation of the metric having three components.

Example 3.7: A similar comment can be done for electromagnetism through the exterior derivative as the first set of Maxwell equations can be parametrized by the EM potential 1-form because dA=FdF=0 while the second set of Maxwell equations (adjoint of this parametrization) can be parametrized by the EM pseudo-potential, a 3 - skew-symmetric contravariant tensor density in 4T3T through the adjoint of the exterior derivative 2Td3T [12, 13, 27, 28]. These results, which are deeply supporting the conformal origin of EM [31], are also strengthening the comments we shall make in Section 4 on the origin and existence of gravitational waves [13, 22, 23].

As a summarizing comment, we discover that not only it is sometimes not possible to parametrize a linear differential operator but that, whenever it is possible, not only it is not easy to have an idea about the number of potential functions needed but even more difficult to have any idea about the order of the parametrizing operator that may be unexpectedly quite high indeed for the double pendulum. In addition, we have provided in [4], examples showing that the case of variable coefficients is even much more difficult than the case of constant coefficients. Moreover, the mathematical tools involved are sometimes not accessible to intuition like this theorem (See [4, 5, 20] for details):

Theorem 3.8: If M is a differential module, we have the exact sequence of differential modules:

0tMMεhomDhomDMDDE16

where the map ε is defined by εmf=fm,mM,fhomDMD.

Theorem 3.9: When D1 can be parametrized or, equivalently, when M1 is torsion-free and can be thus embedded into a free module Dl, we have thus rkDM1=ll. There is a constructive procedure in order to embed M1 into Dl, that is to obtain a minimum parametrization.

Proof: The procedure with 4 steps is as follows in the operator language (Example 1.3):

1 Start with the formally exact parametrizing sequence already constructed by differential biduality. We have thus imD=kerD1 and the corresponding differential module M1 defined by D1 is torsion-free by assumption.

2 Construct the adjoint sequence which is also formally exact by assumption.

3 Find a maximum set of differentially independent CC adD':μν' among the generating CC adD:μν of adD1 in such a way that imadD is a maximum free differential submodule of imadD that is any element in imadD is differentially algebraic over imadD.

4 Using differential duality, construct D=adadD.

Let us prove that D1 generates the CC of D in the following double diagram:

4ξ'D'ξDηD1ζ1νadDμadD1λ2adD3ν00

First of all, we have by construction imadD=imadD in the bottom diagram and thus:

rkDadD+rkDadD1=rkDadD+rkDadD1=m

Passing to the upper diagram, we have, therefore:

rkDD'+rkDD1=rkDD+rkDD1=m

We have adDadD1=adD1D=0D1D=0 and D1 is surely among the CC of D. Therefore, the differential sequence ξDηD1ζ on the operator level or the sequence DpD1DmDDl on the differential module level may not be exact. In the latter, we have now B=imD1=kerDkerD'=ZDm. But we have also rkDB=mrkDD,rkZ=mrkDDrkDH=rkDDrkDD=0. Using the fact that M1=cokerD1 while setting M1=imDDl, we get the commutative and exact diagram:

000BDmM100ZDmM10H000

A snake chase shows that the kernel of the induced epimorphism M1M10 is isomorphic to H and is thus a torsion module because we have just proved that rkDH=0. However, we know that M1 is a torsion-free module and we reach a contradiction unless H=0M1M1.

In actual practice, as shown in the contact case below, things are not so simple and we shall use the following commutative and exact diagram of differential modules:

0keradDDladDDmcokeradD0L0Dl00

Setting L=Dl/keradD and introducing the biggest free differential module DlL we have l=rkDDl=rkDLrkDDlll, we may define the injective (care) operator adD by the composition of monomorphisms DlLDm where the second is obtained by picking a basis of Dl, lifting it to Dl and pushing it to Dm by applying adD. We notice that L can be viewed as the differential module defined by the generating CC of adD.

Example 3.10: Contact transformations.

With m=n=3,K=Qx1x2x3=Qx, we may introduce the so-called contact 1-form α=dx1x3dx2. The system of infinitesimal Lie equations defining the infinitesimal contact transformations is obtained by eliminating the factor ρx in the equations Lξα=ρα where L is the standard Lie derivative. This system is thus only generated by η1 and η2 below but is not involutive and one has to introduce η3 defined by the first order CC operator:

η1η2η3d3η1d2η2x3d1η2+η3=ζ

in order to obtain the following involutive system with two equations of class 3 and one equation of class 2, a result leading to β13=2,β12=1,β11=0 and the characters α13=32=1<α12=31=2,α11=30=3 with sum equal to 1+2+3=6=dimg1=3×33.

d3ξ3+d2ξ2+2x3d1ξ2d1ξ1=η3d3ξ1x3d3ξ2=η2d2ξ1x3d2ξ2+x3d1ξ1x32d1ξ2ξ3=η112312312

The characters are thus α13=32=1<α12=31=2,α11=30=3 with sum equal to 1+2+3=6=dimg1=3×33=6 and we get dimg2=3+2×2+3×1=10=dimS2TT8 along the Janet tabular [20].

In this situation, if M is the differential module defined by this system or the corresponding operator D, we know that rkDM=α13=1=32=rkDrkDD. Of course, a differential transcendence basis for D can be the operator D':ξη2η3 but, in view of the CC, we may equally choose any couple among η1η2η3 and we obtain rkDD'=rkDD=2 in any case, but now D' is formally surjective, contrary to D. The same result can also be obtained directly from the unique CC or the corresponding operator D1 defining the differential module M1. Finally, we have rkDM1=31=2=rkDrkDD1 and we check that we have indeed rkDM+rkDM1=1+2=3=rkD.

It is well known that such a system can be parametrized by the injective parametrization (See [19, 29] for more details and the study of the general dimension n=2p+1):

x3d3ϕ+ϕ=ξ1,d3ϕ=ξ2,d2ϕ+x3d1ϕ=ξ3ξ1x3ξ2=ϕ

Noticing that D is generated by D:ξη1η2, we obtain the operator adD':

d3μ2d2μ1x3d1μ1=ν1,x3d3μ2+μ2+x3d2μ1+x32d1μ1=ν2,μ1=ν3

It follows that μ1=ν3,μ2=ν2+x3ν1 and, substituing, the only CC:

x3d3ν1+d3ν2+d2ν3+x3d1ν3=0

which is exactly adD1 as can be seen by multiplying by a test function ϕ and integrating by parts. No such computations can be found in the literature on contact structures.

The associated differential sequence is:

0ϕD1ξDηD1ζ0013310

with Euler-Poincaré characteristic 13+31=0 but is not a Janet sequence because D1 is not involutive, its completion to involution being the trivially involutive operator j1:ϕj1ϕ.

Introducing the ring D=Kd1d2d3=Kd of linear differential operators with coefficients in the differential field K, the corresponding differential module MD is projective and even free, thus torsion-free or 0-pure [24], being defined by the split exact sequence of free differential modules:

0DD1D3DD3D1D0

We let the reader prove as an exercise that the adjoint sequence:

0θadD1νadDμadD1λ0013310

starting from the Lagrange multiplier λ is also a split exact sequence of free differential modules.

We invite the reader to study, as a delicate exercise, the system of infinitesimal Lie equations defining the infinitesimal unimodular contact transformations preserving the 1-form α=dx1x3dx2, thus also both the 2-form β==dx2dx3 and the volume 3-form αβ=dx1dx2dx3. Surprisingly, the Lie operator D:T1T×X2T:ξLξα=ALξβ=B for the geometric object ω=αβ is involutive if and only if =cβ,=cαβ with cc=0. It provides the differential Janet sequence 36410 with Euler-Poincaré characteristic rkDM=36+41=0. It follows that D cannot be parametrized. We have D1:ABdAcB=UdBcAβ+αB=V and D2:UVdU+cV=W because cc=0. (See [12] and the recent [19] for more details).

Advertisement

4. Einstein equations

If g1TT with dimg1=nn1/2 is the symbol of the Killing system R1J1T with dimR1=nn+1/2, its first prolongation is g2=0. We may introduce the Riemann tensor ρ=ρl,ijkF1=H12g1 with n2n21/12 components in the short exact sequence [13, 18, 25, 27, 28, 29]:

0F12Tg1δ3TT0E17

Linearizing the Ricci tensor ρij=ρi,rjr=ρjiS2T over the Minkowski metric ω, we obtain the usual second order homogeneous Ricci operator ΩR with 4 terms (care):

2Rij=ωrsdrsΩij+dijΩrsdriΩsjdsjΩri=2RjitrΩ=ωijΩijtrR=ωijRij=ωijdijtrΩωruωsvdrsΩuv

We may define the Einstein operator by setting εij=ρij12ωijωrsρrsEij=Rij12ωijtrR and obtain the 6 terms (care) [9]:

2Eij=ωrsdrsΩij+dijΩrsdriΩsjdsjΩriωijωrsωuvdrsΩuvωruωsvdrsΩuv

We have the (locally exact) differential sequence of operators acting on sections of vector bundles:

TKilling1S2TRiemann2F1Bianchi1F2nDnn+1/2D1n2n21/12D2n2n21n2/24E18

where F2=H13g1 is similarly defined by the short exact sequence:

0F23Tg1δ4TT0E19

Our purpose is now to study the differential sequence onto which its right part is projecting:

S2TEinstein2S2Tdiv1T0nn+1/2nn+1/2n0

and the following adjoint sequence where we have set [13, 18, 25, 27, 28, 29]:

adTCauchyadS2TBeltramiadF1LanczosadF2E20

In this sequence, if E is a vector bundle over the ground manifold X with dimension n, we may introduce the new vector bundle adE=nTE, where E is obtained from E by inverting the transition rules exactly like T is obtained from T. We have for example adT=nTTnTTn1T because T is isomorphic to T by using the metric ω. The 10×10 Einstein operator matrix is induced from the 10×20 Riemann operator matrix and the 10×4 div operator matrix is induced from the 20×20 Bianchi operator matrix. We advise the reader not familiar with the formal theory of systems or operators to follow the computation in dimension n=2 with the 1×3 Airy operator matrix, which is the formal adjoint of the 3×1 Riemann operator matrix, and n=3 with the 6×6 Beltrami operator matrix which is the formal adjoint of the 6×6 Riemann operator matrix which is easily seen to be self-adjoint up to a change of basis.

  • n=2: The stress equations become d1σ11+d2σ12=0,d1σ21+d2σ22=0. Their second order parametrization σ11=d22ϕ,σ12=σ21=d12ϕ,σ22=d11ϕ has been provided by George Biddell Airy in 1863 [36] and is well known [4]. We get the second order system:

σ11d22ϕ=0σ12d12ϕ=0σ22d11ϕ=01211

which is involutive with one equation of class 2, 2 equations of class 1 and it is easy to check that the 2 corresponding first order CC are just the Cauchy equations. Of course, the Airy function (1 term) has absolutely nothing to do with the perturbation of the metric (3 terms). With more details, when ω is the Euclidean metric, we may consider the only component:

trR=d11+d22Ω11+Ω22d11Ω11+2d12Ω12+d22Ω22=d22Ω11+d11Ω222d12Ω12

Multiplying by the Airy function ϕ and integrating by parts, we get Airy=adRiemann and Cauchy=adKilling in the following differential sequences:

2Killing13Riemann21002Cauchy13Airy21

n=3: It is quite more delicate to parametrize the 3 PD equations:

d1σ11+d2σ12+d3σ13=0,d1σ21+d2σ22+d3σ23=0,d1σ31+d2σ32+d3σ33=0

A direct computational approach has been provided by Eugenio Beltrami in 1892 [37, 38], James Clerk Maxwell in 1870 [39] and Giacinto Morera in 1892 [38, 40] by introducing the 6 stress functions ϕij=ϕji in the Beltrami parametrization. The corresponding system:

σ11d33ϕ22+d22ϕ332d23ϕ23=0σ12d33ϕ12+d12ϕ33d13ϕ23d23ϕ13=0σ22d33ϕ11+d11ϕ332d13ϕ13=0σ13d23ϕ12+d12ϕ23d22ϕ13d13ϕ22=0σ23d23ϕ11+d11ϕ23d12ϕ13d13ϕ12=0σ33d22ϕ11+d11ϕ222d12ϕ12=0123123123121212

is involutive with 3 equations of class 3, 3 equations of class 2 and no equation of class 1. The three characters are thus α23=1×63=3<α22=2×63=9<α21=3×60=18 and we have dimg2=α21+α22+α23=18+9+3=30=dimS2TS2TdimS2T=6×66 [22]. The 3 CC are describing the stress equations which admit therefore a parametrization … but without any geometric framework, in particular without any possibility to imagine that the above second order operator is nothing else but the formal adjoint of the Riemann operator, namely the (linearized) Riemann tensor with n2n21/2=6 independent components when n=3 [12, 13].

Breaking the canonical form of the six equations which is associated with the Janet tabular, we may rewrite the Beltrami parametrization of the Cauchy stress equations as follows, after exchanging the third row with the fourth row, keeping the ordering 11<12<13<22<23<33:

d1d2d30000d10d2d3000d10d2d3000d332d23d220d33d230d13d120d23d22d13d120d3302d1300d11d23d13d120d110d222d120d11000

as an identity where 0 on the right denotes the zero operator. However, if Ω is a perturbation of the metric ω, the standard implicit summation used in continuum mechanics is, when n=3:

σijΩij=σ11Ω11+2σ12Ω12+2σ13Ω13+σ22Ω22+2σ23Ω23+σ33Ω33=Ω22d33ϕ11+Ω33d22ϕ112Ω23d23ϕ11++Ω23d13ϕ12+Ω13d23ϕ12Ω12d33ϕ12Ω33d12ϕ12+

because the stress tensor density σ is supposed to be symmetric. Integrating by parts in order to construct the adjoint operator, we get:

ϕ11d33Ω22+d22Ω332d23Ω23ϕ12d13Ω23+d23Ω13d33Ω12d12Ω33

and so on. The identifications Beltrami=adRiemann,Lanczos=adBianchi in the diagram:

3Killing16Riemann26Bianchi13003Cauchy16Beltrami26Lanczos13.E21

prove that the Cauchy operator has nothing to do with the Bianchi operator [12, 13].

When ω is the Euclidean metric, the link between the two sequences is established by means of the elastic constitutive relations 2σij=λtrΩωij+2μΩij with the Lamé elastic constants λμ but mechanicians are usually setting Ωij=2εij. Using the standard Helmholtz decomposition ξ=φ+ψ and substituting in the dynamical equation diσij=ρd2/dt2ξj where ρ is the mass per unit volume, we get the longitudinal and transverse wave equations, namely Δφρλ+2μd2dt2φ=0 and Δψρμd2dt2ψ=0, responsible for earthquakes, with respective speeds cL2=λ+2μ/ρ and cT2=μ/ρ.

Then, taking into account, the factor 2 involved by multiplying the second, third and fifth row by 2, we get the new 6×6 operator matrix with rank 3 which is clearly self-adjoint:

000d332d23d2202d332d2302d132d1202d232d222d132d120d3302d1300d112d232d132d1202d110d222d120d1100

Following Maxwell, we keep ϕ11=A,ϕ22=B,ϕ33=C, set ϕ12=ϕ23=ϕ31=0 and get:

σ11d33B+d22C=0σ22d33A+d11C=0σ23d23A=0σ33d22A+d11B=0σ13d13B=0σ12d12C=0123123121211

This system may not be involutive and no CC can be found “a priori” because the coordinate system is surely not δ-regular. Effecting the linear change of coordinates x¯1=x1,x¯2=x2,x¯3=x3+x2+x1 and taking out the bar for simplicity, we obtain the involutive system:

d33C+d13C+d23C+d12C=0d33B+d13B=0d33A+d23A=0d23C+d22Cd13Cd13Bd12C=0d23Ad22C+d13B+2d12Cd11C=0d22A+d22C2d12C+d11C+d11B=0123123123121212

and it is easy to check that the 3 CC obtained just amount to the desired 3 stress equations when coming back to the original system of coordinates. However, the three characters are different as we have now α23=33=0<α22=2×33=3<α21=3×30=9 with sum equal to dimg2=6×36=186=12. We have thus a minimum parametrization. Of course, if there is a geometrical background, this change of local coordinates is hiding it totally. Moreover, we notice that the stress functions kept in the procedure are just the ones on which d33 is acting. The reason for such an apparently technical choice is related to very general deep arguments in the theory of differential modules that will only be explained at the end of the paper.

Following Morera, we keep ϕ23=L,ϕ13=M,ϕ12=N, set ϕ11=ϕ22=ϕ33=0, and get:

d23L=0d33Nd13Ld23M=0d13M=0d22Md23Nd12L=0d11Ld12Md13N=0d12N=0

Using the same change of coordinates, we obtain the following involutive system:

d33N+d23N+d13N+d12N=0d33M+d13M=0d33L+d23L=0d23N+d23Md23L+d13Nd13M+d13L+d12N=02d23M+d13Nd13Md13L+d12Md11L=0d22M+d12Nd12Md12L+d11L=0123123123121212

After elementary but tedious computations (that could not be avoided!), one can prove that the 3 CC corresponding to the 3 dots are effectively satisfied and that they correspond to the 3 Cauchy stress equations which are therefore parametrized. The parametrization is thus provided by an involutive operator defining a torsion module because the character α23 is vanishing in δ-regular coordinates, just like before for the Maxwell parametrization. We have thus another minimum parametrization. Of course, such a result could not have been understood by Beltrami in 1892 because the work of Cartan could not be adapted easily in the language of exterior forms and the work of Janet appeared only in 1920 with no explicit reference to involution because only Janet bases are used while the Pommaret bases have only been introduced in 1978 (See [25, 29] for historical facts).

We may finally keeep ϕ11ϕ12ϕ22, set ϕ13=ϕ23=ϕ33=0 and get the different involutive system with the same characters, in particular with α23=0:

σ1133ϕ22=0σ1233ϕ12=0σ2233ϕ11=0σ1323ϕ1213ϕ22=0σ2323ϕ1113ϕ12=0σ3322ϕ11+11ϕ22212ϕ12=0123123123121212E22

So far, we have thus obtained three explicit local minimum parametrizations of the Cauchy stress equations with nn1/2=3 stress potentials but there may be others.

n=4: It just remains to explain the relation of the previous results with Einstein equations. The first suprising link is provided by the following technical proposition:

Proposition 4.1: The Beltrami parametrization is just described by the Einstein operator when n=3. The same confusion existing between the Bianchi operator and the Cauchy operator has been made by both Einstein and Beltrami because the Einstein operator is self-adjoint in arbitrary dimension n3, contrary to the Ricci operator.

Proof: The number of components of the Riemann tensor is dimF1=n2n21/12. We have the combinatorial formula n2n21/12nn+1/2=nn+1n+2n3/12 expressing that the number of components of the Riemann tensor is always greater or equal to the number of components of the Ricci tensor whenever n>2. Also, we have shown in many books [12, 13, 25, 27, 28, 29] or papers [16, 18] that the number of Bianchi identities is equal to n2n21n2/24, that is 3 when n=3 and 20 when n=4. Of course, it is well known that the div operator, induced as CC of the Einstein operator, has n components in arbitrary dimension n3.

Accordingly, when n=3 we have n2n21/12=nn+1/2=6 and it only remains to prove that the Einstein operator reduces to the Beltrami operator and not just to the Ricci operator.

The following formulas can be found in any textbook on general relativity. In particular, the difference existing between Rij (4 terms only) and Eij (6 terms) can only be seen when ωij=0. In our situation with n=3 and the Euclidean metric, we obtain successively:

2R12=2E12=d11+d22+d33Ω12+d12Ω11+Ω22+Ω33d11Ω12+d12Ω22+d13Ω23d12Ω11+d22Ω12+d23Ω13=d33Ω12+d12Ω33d13Ω23d23Ω13
2R11=d11+d22+d33Ω11+d11Ω11+Ω22+Ω332d11Ω11+d12Ω12+d13Ω13=d22+d33Ω11+d11Ω22+Ω332d12Ω12+d13Ω13
trR=d11Ω22+d11Ω33+d22Ω11+d22Ω33+d33Ω11+d33Ω222d12Ω12+d13Ω13+d23Ω232E11=d22Ω33+d33Ω222d23Ω23

In the light of modern differential geometry, comparing these results with the works of Maxwell, Morera, Beltrami and Einstein, it becomes clear that they have been confusing the div operator induced from the Bianchi operator with the Cauchy operator. However, it is also clear that they both obtained a possibility to parametrize the Cauchy operator by means of 3 arbitrary potential like functions in the case of Maxwell and Morera, 6 in the case of Beltrami explaining the previous choices, and 10 in the case of Einstein. Of course, as they were ignoring that the Einstein operator was self-adjoint whenever n3, they did not notice that we have Cauchy=adKilling and they were unable to compare their results with the Airy operator found as early as in 1870 for the same mechanical purpose when n=2. To speak in a rough way, the situation is similar to what could happen in the study of contact structures if one should confuse D1 with D1 or D with D2 in the Killing sequence [16]. Finally, using the double differential duality test, we can choose a differential transcendence basis with nn1/2 potentials that can be indexed by ϕij=ϕji with i<j or 1i,jn1 or even 2i,jn when the dimension n2 is arbitrary (See [24, 29] for more details on differential algebra).

Remark 4.2: The author of this paper is not an historian of sciences but a specialist of mathematical physics interested by the analogy existing between electromagnetism (EM), elasticity (EL), and gravitation (GR) by using the conformal group of space–time along the idea of H. Weyl [41] (See [4, 13, 18, 19, 28, 29, 30, 31] for related works). It is thus difficult to imagine that Einstein could not have been aware of the works of Maxwell and Beltrami on the foundations of EL and tensor calculus as they were quite famous when he started his research on GR. We also notice that the Mach-Lippmann analogy [4, 13, 28, 29, 35] was introduced at the same time and that the phenomenological law of piezzo-electricity has been discovered by… Maxwell [19, 20]. We do believe that classical variational calculus must be considered as a kind of “duality theory” that should only depend on new purely mathematical tools, namely “group theory” on one side and “differential homological algebra” on the other side (See [12, 13, 28] for the theory and [4] for the applications).

The two following crucial results, still neither known nor acknowledged today, are provided by the next proposition and corresponding corollary:

Proposition 4.3: The cauchy operator can be parametrized by the adjoint of the Ricci Operator

Proof: The Einstein operator ΩE is defined by setting Eij=Rij12ωijtrR that we shall write Einstein=CRicci where C:S2TS2T is a symmetric matrix only depending on ω, which is invertible whenever n3. Surprisingly, we may also introduce the same linear transformation C:ΩΩ¯=Ω12ωtrΩ and the unknown composite operator X:Ω¯ΩE in such a way that Einstein=XC where X is defined by (See [15], 5.1.5, p. 134):

2Eij=ωrsdrsΩ¯ijωrsdriΩ¯sjωrsdsjΩ¯ri+ωijωruωsvdrsΩ¯uv

Now, introducing the test functions λij, we get:

λijEij=λijRij12ωijtrR=λij12λrsωrsωijRij=λ¯ijRij

Integrating by parts while setting as usual =ωrsdrs, we obtain:

λ¯rs+ωrsdijλ¯ijωsjdijλ¯riωridijλ¯sjΩrs=σrsΩrsE23

Moreover, suppressing the “bar” for simplicity, we have:

drσrs=ωijdrijλrs+ωrsdrijλijωsjdrijλriωridrijλsj=0E24

As Einstein is a self-adjoint operator (contrary to the Ricci operator), we have the identities:

adEinstein=adCadXEinstein=CadXadX=RicciX=adRicci

Indeed, adC=C because C is a symmetric matrix and we know that adEinstein=Einstein. Accordingly, the operator ad(Ricci) parametrizes the Cauchy equations, without any reference to the Einstein operator, which has no mathematical origin, in the sense that it cannot be obtained by any diagram chasing. The three terms after the Dalembert operator factorize through the divergence operator diλri. We may thus add the differential constraints diλri=0 without any reference to a gauge transformation in order to obtain a (minimum) relative parametrization (see [24] for details and explicit examples). When n=4 we finally obtain the adjoint sequences:

4Killing10Ricci1004Cauchy10adRicci10E25

without any reference to the Bianchi operator and the induced div operator.

Finally, using Theorem 2.1 or Proposition 2.2, we may choose a differential transcendence basis made by λiji<j or λij1<ij<n1 or even λij2<ij<n when n2.

According to Theorem 3.2 and Example 3.6, the Einstein and thus also the Ricci operators cannot be parametrized. Now, according to Corollary 3.3, the differential module N that they define is not torsion-free.

Corollary 4.4: tN is generated by the 10 components of the Weyl tensor and each component is killed by the Dalembert operator.

Proof: We first recall the 5 steps of the double differential duality test in this framework:

1 Start with the Einstein operator D1:10Einstein10.

2 Consider its formal adjoint: adD1:10Einstein10.

3 Compute the generating CC, namely the Cauchy operator: adD:4Cauchy10.

4 Consider its formal adjoint: D=adadD:4Killing10.

5Compute the generating CC, namely the Riemann operator: D1:10Riemann20.

With a slight abuse of language, we have the direct sum Riemann=RicciWeyl with 20=10+10. It follows from differential homological algebra that the 10 additional CC in D'1 that are not in D1, are generating the torsion submodule tN of the differential module N defined by the Einstein or Ricci operator. If K is a differential field and we consider the ring D=Kd1dn=Kd of differential operators with coefficients in K, we know that rkDD=rkDadD for any operator matrix D with coefficients in K. In the present situation, as the Minkowski metric has coefficients equal to 0,1,1, we may choose K=Q. Hence, there must exist operators P and Q with rkDP=10 and:

PWeyl=QRicciE26

One may also notice that rkDEinstein=rkDRicci with:

rkDEinstein=nn+12n=nn12,rkDRiemann=nn+12n=nn12

It is a pure chance that the differential ranks of the Einstein and Riemann operators are equal. Indeed, rkDEinstein has only to do with the div operator induced by contracting the Bianchi operator, while rkDRiemann has only to do with the classical Killing operator and the fact that the corresponding differential Killing module is a torsion module because we have a Lie group of transformations having n+nn12=nn+12 parameters (translations + rotations). Hence, as the Riemann operator is a direct sum of the Weyl operator and the Einstein or Ricci operator according to the previous theorem, each component of the Weyl operator must be killed by a certain operator whenever the Einstein or Ricci equations in vacuum are satisfied. Equivalently, we have to prove that we obtain a torsion differential module if we set the 10 constraints Rij=0 in the 20 equations Rkl,ij=0. With more details, we may start from the long exact sequence:

0Θ4Killing110Riemann220Bianchi120160E27

This resolution of the set of Killing vector fields is not at all a Janet sequence because the Killing operator is not involutive as it is an operator of finite type with symbol of dimension nn1/2=6 and one should need one prolongation for getting an involutive operator with vanishing symbol at order two. Splitting the Riemann operator, we get the commutative and exact diagram:

00001016604Killing10Riemann20Bianchi206010Einstein10div40000E28

Passing to the differential module point of view, we have the long exact sequence:

0D6D20BianchiD20RiemannD10KillingD4M0E29

Which is a resolution of the Killing differential module M=cokerKilling, and we have indeed the vanishing of the Euler-Poincaré characteristic with rkDM=410+2020+6=0.

Accordingly, we have N=cokerRiemannimKillingD4 and thus N is torsion-free with rkDN=40=4=n because rkDM=0. The kernel L of the epimorphism NN is a torsion module because rkDL=rkDNrkDN=44=0. As D is an integral domain and ND4 is torsion-free, we have thus L=tN in the following commutative and exact diagram where N=cokerEinstein is the differential module defined by Einstein equations in vacuum:

0000tN0D4divD10EinsteinD10N00D6D20BianchiD20RiemannD10N00D6D16D1000000

A snake chase in the previous diagram allows to exhibit the long exact connecting sequence:

0D6D16D10tN0

We point out that, for n=4 (only), the CC of the Weyl operator are of order 2 and not 1 like the Bianchi CC for the Riemann operator (See [12, 18], Appendix 2 for a computer algebra checking by A. Quadrat). Accordingly, we have the conformal sequence in which D̂ is the conformal Killing operator when n=4, with rkDM̂=49+109+4=0:

0Θ̂4D̂19D̂1210D̂229D̂3140E30

and one cannot find canonical morphisms between the classical and the conformal resolutions (!).

It follows from the last theorem that the short exact sequence 0D10D20D100 splits with D20D10D10 but the existence of a canonical lift D20D100 in the above diagram does not allow to split the right column and thus NNtN as N is not even free. As for the torsion elements, we have tN=cokerD16D10 and we may thus represent them by the 10 components of the Weyl tensor. It is not at all evident that is killing each component of the Weyl tensor whenever Rij=0. A tricky technical computation can be found in ([32], p. 206), ([33], Exercise 7.7) and ([13], p. 95).

Indeed, according to the double differential duality test, each additional CC in D'1 which is not already in D1 is a torsion element as it belongs to this module. Now, as rkDD=rkD(imD the differential ranks of the Einstein and Riemann operators are thus equal to nn1/2, but this is a pure coincidence because rkDEinstein has only to do with the div operator induced by contracting the Bianchi identities, while rkDRiemann has only to do with the classical Killing operator and the fact that the corresponding differential module is a torsion module because we have a Lie group of transformations having n+nn12=n(n+12 parameters (translations + rotations). Hence, as the Riemann operator is a direct sum of the Weyl operator and the Einstein or Ricci operator according to the previous theorem, each component of the Weyl operator must be killed by a certain operator whenever the Einstein or Ricci equations in vacuum are satisfied in arbitrary dimension n4.

We prove directly that each component of the Weyl tensor is killed by the Dalembertian.

With Christoffel symbols γ and the corresponding covariant derivative we know that ω=0 and we may thus move up and down the indices as needed. We provide this tricky computation using essentially the (second) Bianchi identities. We have successively:

rρl,ijk+iρl,jrk+jρl,rik=0rρrl,ijiρlj+jρli=0rrρkl,ij+riρkl,jr+rjρkl,ri=0rrρkl,ij+irρkl,jr+jrρkl,ri+riρkl,jr+rjρkl,ri=0rrρkl,ij+irρkl,jr+jrρkl,ri+quadratic=0

But we have also ρkl,ij=ρij,kl and thus rρij,rl=iρljjρlirρjr12jρ=0.

ρkl,ij=ikρljlρkjij+quadratic

Linearizing at the Euclidean metric for n=2,3 or at the Minkowski metric for n=4, we get:

Rkl,ij=didkRljdlRkjdjdkRlidlRkiE31

We may finally use the splitting formula for defining the Weyl tensor σl,ijk with σl,rjr=0, namely:

σl,ijk=ρl,ijk1n2δikρljδjkρli+ωksωljρsiωliρsj+1n1n2δikωljδjkωliρσl,ijk=ρl,ijkρrsΣl,ijk=Rl,ijkRrs

At any moment, we could have used the Ricci operator in place of the Einstein operator.

Finally, for the sake of completeness, we compute directly the characters of the Einstein system.

Using a direct checking with the ordering 11<12<13<14<22<<34<44, we obtain:

E33=ω44d44Ω33+lower  terms,E23=ω44d44Ω23+

We are in the position to compute the characters of the Einstein operator but a similar procedure could be followed with the Ricci operator. We obtain at once:

β24=6,β23=4,β22=0,β21=0α24=4,α23=16,α23=30,α21=40

a result leading to dimg2=α21+α22+α23+α24=90 and dimg3=α21+2α22+3α23+4α24=164 in a coherent way with the long exact sequences:

0g2S2TS2TS2T0,0g3S3TS2TTS2TT0

Now, we have by definition div=d1d2d3d4 and divEinstein=0,0,0,0. However, the Einstein operator is a 10×10 operator matrix which is self-adjoint up to a change of basis [2] because it is made with homogeneous second order terms. It is thus of rank 6 and we obtain therefore, with a slight abuse of language, detEinstein=0. This result which is not evident at first sight must be compared with the Poincaré situation when n=3:

d1d2d30d3d2d30d1d2d10=000

This module interpretation may thus question the proper origin and existence of gravitational waves because the div operator on the upper left part of the diagram has strictly nothing to do with the Cauchy=adKilling operator which cannot even appear anywhere in this diagram. Also, looking back to Example 3.3, if we use ad(Einstein) or ad(Ricci) in order to parametrize the Cauchy operator, the 10 potentials involved are similar to the Airy or Maxwell/Morera potentials and have thus strictly nothing to do with a perturbation of the metric. Such a result is explaining the conceptual confusion announced in the abstract.

Corollary 4.5: When D is a Lie operator of finite type, that is when ΘΘΘ under the ordinary bracket of vector fields and gq+r=0 for r large enough, then the Spencer sequence is locally isomorphic to the tensor product of the Poincaré sequence for the exterior derivative by a finite dimensional Lie algebra G, namely [29]:

0Θ0TGd1TÂGd2TGddnTG0.E32

It is thus formally exact both with its adjoint sequence. As it is known that the extension modules do not depend on the resolution used, this is the reason for which not only the Cauchy stress operator can be parametrized but also the Cosserat couple-stress operator adD1 can be parametrized by adD2, a result not evident at all (see [34, 35] for explicit computations). Similarly, in the case of the conformal Killing system R̂1J1T for n=4, the symbols do not depend on any conformal factor because ĝ1 is defined by ωrjξir+ωirξjr2nωijξrr=0, ĝ2T (the so-called elations of E. Cartan) is defined by ξijk1n(δikξrjr+δjkξrirωijωksξrsr)=0 (with the Kronecker notation) and finally ĝ3=0,n3. The EM field is thus a section FδTĝ2=δTT=2T killed by the exterior derivative d. It follows that EM only depends on the conformal group and not on U1 [30, 31] in a coherent way with the dream of H. Weyl because TTS2T2T=RijFij [29, 41].

Remark 4.6: A similar situation is well known for the Cauchy- Riemann equations when n=2. Indeed, any infinitesimal complex transformation ξ must be solution of the linear first order homogeneous system ξ22ξ11=0,ξ21+ξ12=0 of infinitesimal Lie equations which is defining a torsion differential module because we have ξ111+ξ221=0,ξ112+ξ222=0, that is ξ1 and ξ2 are separately killed by the Laplace operator Δ=d11+d22.

Remark 4.7: A similar situation is also well known for the EM field F in electromagnetism. Indeed, starting with the first set of Maxwell equations dF=0 (M1) and using the Minkowski constitutive law in vacuum with electric constant ε0 and magnetic constant μ0, such that ε0μ0c2=1 for the second set of Maxwell equations (M2) for the induction F, a standard tricky differential elimination allows to avoid the Lorenz (no “t”) gauge condition for the EM potential. Indeed, from M1 we get formally drFij+diFjr+djFri=0 and from M2 we get drFri=0drFri=0. We finally obtain directly F=drdrFij=drdiFrjdjFri=0 [4].

Example 4.8: (Lorenz condition for EM)

We prove that the Lorenz gauge condition for EM is just amounting to a relative minimum parametrization. Using an Euclidean metric instead of the Minkowski metric in order to have the Minkowski constitutive relation F=F in vacuum between the EM field and the EM induction, we have the parametrization diAjdjAi=Fij and obtain the conservation of current through the composition:

didiAjdjAi=didiAjdjdiAi=JjdjdidiAjdjdiAi=djJj=0

with implicit summation on i and j. The differential module defined by the involutive system diiAjdijAi=0 has a differential rank equal to 1 as there is only one CC. Adding the Lorenz condition diAi=0 is bringing the rank to zero. The idea is just to prove that the inhomogeneous system didiAj=Jj,diAi=0 has again the only CC djJj=0 like in Example 1.3.

Advertisement

5. Conclusion

After teaching elasticity theory during 25 years to students in some of the best french civil engineering schools, the author of this paper still keeps in mind one of the most fascinating but most difficult exercises that he has set up. The purpose was to explain why the dome of the cathedral St Paul in London is called “whispering cupola”, that is why if you go up to the gallery at one point, you can listen to a friend whispering 30 meters away on the opposite side. This striking phenomena has been first studied by Lord Rayleigh in 1878 and he introduced in 1910 the “surface elastic waves” now called Rayleigh waves (See [3] p. 199). In fact the molecules close to the free surface are moving along small ellipses as a combination of standard longitudinal (L) and transverse (T) waves while propagating at a slightly smaller speed. Such waves cannot propagate in liquid but may travel all around the earth surface many times after earthquakes. If λ and μ are the elastic Lamé constants of the material with mass ρ per unit volume, the respective speeds are such that cL2=λ+2μ/ρ,cT2=μ/ρ and the speed of the Rayleigh wave is cR2=χ12ν/21ν where ν=λ/2λ+μ is the Poisson coefficient. After (very) tedious computations one may find that χ must be the only real root of the cubic equation χ38χ2+82ν1νχ81ν=0 existing in the interval [0, 1] because 0<ν<0,5.

Accordingly, elasticity can be considered as a way to parametrize the Cauchy operator when n=3 while GR can be considered as a way to parametrize the Cauchy operator when n=4. Hence, the situations existing with the Cauchy stress equations, with the Cosserat couple-stress equations and with the Maxwell equations are similar, only the constitutive laws are different. Meanwhile, we have shown why the mathematical foundations of conformal geometry must be revisited in this new framework which is valid in arbitrary dimension and could provide an intrinsic way to unify EM and GR along the dream of Weyl [12, 31].

The situation of the gravitational waves equations seems quite different but the least that can be said is that it is not coherent with differential double duality. However, it follows that exactly the same confusion has been done by Maxwell, Morera, Beltrami, and Einstein because, in all these cases, the operator considered is self-adjoint. Like the Michelson and Morley experiment, we do believe that Einstein already knew the previous works of all these researchers who were quite famous at the time he was active. In any case, the comparison of the various parametrizations described in this paper needs no comment.

References

  1. 1. Pommaret J-F. Dualité Différentielle et Applications. Comptes Rendus Académie des Sciences Paris, Série I. 1995;320:1225-1230
  2. 2. Pommaret J-F. The mathematical foundations of general relativity revisited. Journal of Modern Physics. 2013;4:223-239. DOI: 10.4236/jmp.2013.48A022
  3. 3. Bellet D, Duc J. Problèmes d’ Elasticité. Toulouse: Cepadues-Edition. 1977. ISBN 2.85428.018.0
  4. 4. Pommaret J-F. Partial Differential Control Theory. Dordrecht: Kluwer; 2001 (Zbl 1079.93001)
  5. 5. Kashiwara M. Algebraic study of systems of partial differential equations. mémoires de la société mathématique de france. 1995;63 (Transl. from Japanese of his 1970 [master thesis]). p.1-72
  6. 6. Goldschmidt H. Prolongations of linear partial differential equations: I inhomogeneous equations. Ann. Scient. Ec. Norm. Sup. 1968;4:617-625. DOI: 10.24033/asens.1173
  7. 7. Spencer DC. Overdetermined systems of partial differential equations. Bulletin of the American Mathematical Society. 1965;75:1-114
  8. 8. Hu S-T. Introduction to Homological Algebra. San Francisco: Holden-Day; 1968
  9. 9. Northcott DG. An Introduction to Homological Algebra. Cambridge: CambridgeUniversity Press; 1966
  10. 10. Rotman JJ. An Introduction to Homological Algebra. Urbana: Academic Press; 1979
  11. 11. Schneiders J-P. An introduction to D-modules. Bulletin de la Société Royale des Sciences de Liège. 1994;63:223-295
  12. 12. Pommaret J-F. Deformation Theory of Algebraic and Geometric Structures. Saarbrucken, Germany: Lambert Academic Publisher (LAP); 2016. http://arxiv.org/abs/1207.1964
  13. 13. Pommaret J-F. New Mathematical Methods for Physics. Mathematical Physics Books: Nova Science Publishers, New York; 2018. p. 150
  14. 14. Bieri L, Garfinkle D, Hill CD, Nurowski P, Yunes N. The Mathematics of Gravitational Waves. Notices of the AMS. 2017;64(7):686-707
  15. 15. Foster J, Nightingale JD. A Short Course in General Relativity. New York: Longman; 1979
  16. 16. Pommaret J-F. Killing operator for the Kerr metric. Available from: https://arxiv.org/abs/2211.00064)
  17. 17. Lanczos C. The splitting of the Riemann tensor. Reviews of Modern Physics. 1962;34:379-389
  18. 18. Pommaret J-F. Homological solution of the lanczos problems in arbitrary dimension. Journal of Modern Physics. 2021;12:829-858. DOI: 10.4236/jmp.2020.1110104
  19. 19. Pommaret J-F. How many structure constants do exist in Riemannian geometry. Mathematics in Computer Science. 2022;16:23. DOI: 10.1007/s11786-022-00546-3
  20. 20. Pommaret J-F. Algebraic analysis of control systems defined by partial differential equations. In: “Advanced Topics in Control Systems Theory”, Lecture Notes in Control and Information Sciences. Vol. 311. London: Springer; 2005. Chapter 5. pp. 155-223
  21. 21. Pommaret J-F, Quadrat A. Localization and parametrization of linear multidimensional control systems. Systems & Control Letters. 1999;37:247-260
  22. 22. Pommaret J-F. Why gravitational waves cannot exist. Journal of Modern Physics. 2017;8:2122-2158. Available from: https://arxiv.org/abs/1708.06575. DOI: 104236/jmp.2017.813130
  23. 23. Pommaret J-F. Minimum parametrization of the cauchy stress operator. Journal of Modern Physics. 2021;12:453-482. Available from: https://arxiv.org/abs/2101.03959. DOI: 10.4236/jmp.2021.124032
  24. 24. Pommaret J-F. Relative parametrization of linear multidimensional systems. Multidimensional Systems and Signal Processing. 2015;26:405-437. DOI: 10.1007/s11045-013-0265-0
  25. 25. Pommaret J.-F.: Systems of Partial Differential Equations and Lie Pseudogroups, Gordon and Breach, New York (1978); Russian translation: MIR, Moscow. 1983
  26. 26. Zerz E. Topics in multidimensional linear systems theory. In: Lecture Notes in Control and Information Sciences, LNCIS. Vol. 256. London: Springer; 2000
  27. 27. Pommaret J-F. Differential Galois Theory. New York: Gordon and Breach; 1983
  28. 28. Pommaret J-F. Lie Pseudogroups and Mechanics. New York: Gordon and Breach; 1988
  29. 29. Pommaret J-F. Partial Differential Equations and Group Theory. Kluwer; 1994. DOI: 10.1007/978-94-017-2539-2
  30. 30. Pommaret J-F. Spencer operator and applications: From continuum mechanics to mathematical physics. In: Continuum Mechanics-Progress in Fundamentals and Engineering Applications. Gan Y, editor. London: InTech; 2012. Available from: DOI: 10.5772/35607. ISBN: 978-953-51-0447–6
  31. 31. Pommaret J-F. Nonlinear conformal electromagnetism. Available from: https://arxiv.org/abs/2007.01710. DOI: 10.4236/jmp.2022.134031
  32. 32. Choquet-Bruhat Y. Introduction to General Relativity, Black Holes and Cosmology. Oxford: Oxford University Press; 2015
  33. 33. Hughston LP, Tod KP. An Introduction to General Relativity, London Math. Soc. Students Texts 5. Cambridge: Cambridge University Press; 1990
  34. 34. Cosserat E, Cosserat F. Théorie des Corps Déformables. Paris: Hermann; 1909
  35. 35. Pommaret J-F. Parametrization of Cosserat Equations. Acta Mechanica. 2010;215:43-55. DOI: 10.1007/s00707-010-0292-y
  36. 36. Airy GB. On the strains in the interior of beams. Philosophical Transactions. Royal Society of London. 1863;153:49-80
  37. 37. Beltrami E. Osservazioni sulla Nota Precedente. Atti della Accademia Nazionale dei Lincei Rend. 1892;1(5):137-141 + 233-234 Collected Works, t IV
  38. 38. Landriani GS. A note about an exchange of opinions between Beltrami and Morera on elastic equilibrium equations. Meccanica. 2017;52:2801-2806. DOI: 10.1007/s11012-016-0611-z
  39. 39. Maxwell JC. On reciprocal figures, frames and diagrams of forces. Transactions of the Royal Society of Edinburgh. 1870;26:1-40
  40. 40. Morera G. Soluzione Generale della Equazioni Indefinite di Equilibrio di un Corpo Continuo. Atti della Academia Nazionale dei Lincei Rend. 1892;1(5):137-141 + 233-234
  41. 41. Weyl H. Space, Time, Matter, (1918). New York: Dover; 1952

Written By

Jean-Francois Pommaret

Submitted: 14 November 2022 Reviewed: 20 November 2022 Published: 19 January 2023