Foundations of Physics manuscript No. (will be inserted by the editor) The Mechanics of Spacetime – A Solid Mechanics Perspective on the Theory of General Relativity T G Tenev · M F Horstemeyer arXiv:1603.07655v6 [gr-qc] 1 Jun 2017 Received: date / Accepted: date Abstract We present an elastic constitutive model of gravity where we identify physical space with the mid-hypersurface of an elastic hyperplate called the “cosmic fabric” and spacetime with the fabric’s world volume. Using a Lagrangian formulation, we show that the fabric’s behavior as derived from Hooke’s Law is analogous to that of spacetime per the Field Equations of General Relativity. The study is conducted in the limit of small strains, or analogously, in the limit of weak and nearly static gravitational fields. The Fabric’s Lagrangian outside of inclusions is shown to have the same form as the Einstein-Hilbert Lagrangian for free space. Properties of the fabric such as strain, stress, vibrations, and elastic moduli are related to properties of gravity and space, such as the gravitational potential, gravitational acceleration, gravitational waves, and the energy density of free space. By introducing a mechanical analogy of General Relativity, we enable the application of Solid Mechanics tools to address problems in Cosmology. Keywords modified gravity · constitutive model · spacetime · cosmic fabric PACS 04.50.Kd · 46.90.+s Mathematics Subject Classification (2000) 83D05 · 74L99 1 Introduction In 1678, Robert Hooke, a contemporary of Isaac Newton, published what later became known as Hooke’s Law [13]. In 1827, Cauchy [2] advanced Hooke’s Law by defining the tensorial formulation of stress. For an isotropic linear elastic material, T G Tenev Mississippi State University, Starkville, MS 39759, USA E-mail: [email protected] M F Horstemeyer Mississippi State University, Starkville, MS 39759, USA E-mail: [email protected] 2 T G Tenev, M F Horstemeyer Hooke’s Law states in tensorial form that, ν Y ij kl ik jl kl g g +g g εij σ = 1 + ν 1 − 2ν (1.1) where σ kl , εij , and g ij are the stress, strain, and the metric tensors, respectively, Y is Young’s elastic modulus, and ν is the Poisson’s ratio. Latin indexes, i, j, k, l = 1 . . . 3, run over the three spatial dimensions, and Einstein summation convention is employed. In 1916 Einstein published the field equations of General Relativity [8], which can be written as, 1 1 Tµν = Rµν − Rgµν (1.2) κ 2 where Tµν , Rµν , and gµν are the stress-energy tensor, Ricci curvature tensor, and µ spacetime metric tensor, respectively; R ≡ Rµ is the Ricci scalar, κ ≡ 8πG/c4 is the Einstein constant as c and G are the speed of light and gravitational constant, respectively. Greek indexes, µ, ν = 0 . . . 3, run over the four dimensions of spacetime with the 0th dimension representing time. For the purposes of this paper, we have omitted the Cosmological Constant, which is sometimes included in Equation (1.2), because its value is negligible for length-scales below the size of the observable universe. Einstein’s Gravitational Law (1.2) suggests a material-like constitutive relation, similar to Hooke’s Law (1.1), because it relates stress, on the left-hand side, to deformation on the right-hand side. At first glance, the similarity appears imperfect because the right-hand sides differ in dimensionality: whereas the strain term, εij , is dimensionless, the curvature terms, Rµν and R, have dimensions of Length−2 . However, this problem is resolved once bending deformation is considered instead of purely longitudinal deformation (stretch or contraction). In the equations for bending, stress is proportional to the second spatial derivative of strain. In this paper, we develop a formal analogy between Solid Mechanics and General Relativity by identifying physical space with the mid-hypersurface of a four dimensional hyperplate, called the “cosmic fabric,” which has a small thickness along a fourth spatial dimension and exhibits a constitutive stress-strain behavior. Matter-energy fields act as inclusions within the fabric causing it to expand longitudinally and consequently to bend. The effect, illustrated on Fig. 1, is analogous to the result from General Relativity in which matter causes space to bend resulting in gravity. We conduct our study in the limit of weak and nearly static gravitational fields, and demonstrate that outside of inclusions, the fabric’s action SF , assumes the form of the Einstein-Hilbert action SEH , Z p Z p YL 1 4 SF = R |g| dx vs. SEH = R |g| dx4 (1.3) 48 2κ where L is the reference thickness of the fabric, g ≡ det gµν , and the integral is taken over a large enough volume of spacetime sufficient to ensure convergence. The action integral of any physical system fully determines its dynamics, because the system’s equations of motion can be derived from the variation of the action with respect to the metric. Therefore, once we recognize SF as analogous to SEH , The Mechanics of Spacetime 3 Fig. 1 A plate bending from flat geometry (a) into a curved geometry (b) because of an inclusion that prescribes uneven strain field, as indicated by the concentric dashed lines and the diverging arrows. The strain is larger near the center and tapers off with the distance from it. For the geometry of the plate to accommodate the prescribed strain, the plate must bend into the transverse dimension. we can interpret various attributes of the cosmic fabric, such as its shape, strain, vibrations, and elastic moduli as analogous to properties of gravity and space, such as curvature, gravitational potential, gravitational waves, and the zero point energy density of space. In some aspects, our approach resembles the Arnowitt-Deser-Misner (ADM) formulation of gravity. Under the ADM approach, spacetime is foliated into spacelike hypersurfaces related to each other via shift and lapse functions. ADM has been used in computational models of gravity, because it reformulates the equations of General Relativity (1.2) as an initial value problem (IVP). Likewise, the Cosmic Fabric model discussed here could also be used in the future to construct an IVP. Nevertheless, the Cosmic Fabric model differs from the ADM formulation in that it associates constitutive behavior with the geometric description of gravity and derives its governing equations from a material-like constitutive relation. Since as early as Isaac Newton [16] there have been various scientific theories about an all pervasive cosmic medium, also known as “ether,” through which light and matter-matter interactions propagate. These theories culminated with the Lorentz Ether Theory (LET) [20, 21], which postulated length contractions and time dilations for objects moving through ether [10, 20] in order to explain the negative outcome of the Michelson and Morleys ether detection experiment [23]. Although the theory of Special Relativity (SR) [7] appeared to obviate the need for an ether, in reality SR and LET are mathematically equivalent and experimentally indistinguishable from one another. After the development of General Relativity [8], which attributed measurable intrinsic curvature to spacetime, Einstein conceded [9] that some notion of an ether must remain. More recently, theoretical predictions from Quantum Field Theory [30] include zero-point energy density of space, which further supports the material view of space. After Einstein’s publication of General Relativity [8], a number of researchers have investigated the relationship between Mechanics notions and General Relativity. One category of publications dealt with generalizing the equations of Solid Mechanics to account for relativistic effects. Synge [31] formulated a constitutive relationship in relativistic settings. Rayner [28] extended Hooke’s Law to a relativistic context. Maugin [22] generalized the special relativistic continuum mechanics theory developed by Grot and Eringen [12] to a general relativistic context. More recently, Kijowski and Magli [18] presented the relativistic elasticity theory as a gauge theory. A detailed review of relativistic elasticity can be found in Karlovini and Samuelsson [17]. 4 T G Tenev, M F Horstemeyer Another category of publications interprets General Relativity in Solid Mechanical terms. Kondo [19] mentions an analogy between the variation formalism of his theory of global plasticity and General Relativity. Gerlach and Scott [11] introduce a “metric elasticity” tensor in addition to the elasticity of matter itself and “stresses due to geometry.” However, these stress and strain terms are not a constitutive model of gravity, because they are not expected to apply in the absence of ordinary matter. Tartaglia [32] attempted to describe spacetime as a four-dimensional elastic medium in which one of the spatial dimensions has been converted into a time dimension by assuming a uniaxial strain. However, many of the ideas in Tartaglia’s paper appear to be incomplete. Antoci and Mihich [3] explored the physical meaning of the straightforward formal extension of Hooke’s Law to spacetime but did not consider the possibility, which is explored in this paper, that Einstein’s Gravitational Law may be related to Hooke’s Law. Beau [5] pushed the material analogy further by interpreting the cosmological constant Λ as related to a kind of a spacetime bulk modulus, but the analogy is to a fluid-like material and not a solid. A set of recent publications, for which Rangamani [27] presents a literature review, explore the applicability of the Navier-Stokes equations of Fluid Dynamics to gravity. While a fluid analogy is useful for some applications, it does not account for shear waves in space, such as gravity waves, because fluids are only capable of propagating pressure waves and not shear waves. In contrast to the prior literature recounted above, the work presented here begins with the premise that space exhibits material-like behavior subject to a constitutive relationship. The Cosmic Fabric model of gravity allows General Relativity problems to be formulated as Solid Mechanics problems, solved within the Solid Mechanics domain, and the solution interpreted back in General Relativity terms. The reverse is also true. Thus, ideas, methodologies and tools from each field become available to the other field. Over the past century, Solid Mechanics and General Relativity have advanced independently from each other with few researchers having expertise in both. Consequently, significant terminology and focus gaps exist between these two fields, which obscure their underlying physical similarities. Our research attempts to bridge these gaps. The remainder of this paper is organized as follows: In Section 2 we develop the Solid Mechanics analogy of gravity by specifying a material body whose behavior, determined solely based on Hooke’s Law (1.1), is demonstrably analogous to the behavior of spacetime. In Section 3 we discuss the implications of the resulting model, and summarize and conclude in Section 4. 2 Formulation of the Cosmic Fabric Model of Gravity Consider a four dimensional hyperplate, called here the “cosmic fabric,” which is thin in the fourth spatial dimension, x4 . We show that, for a suitably chosen initial geometry and Poisson’s ratio p of the fabric, its Lagrangian density outside of inclusions is LF = (Y L/48)R |g|, where LF is the integrand in Equation (1.3). This result enables us to subsequently analyze how the remaining kinematic properties of the cosmic fabric correspond to properties of gravity. For the remainder of this paper, we will use the following notation and conventions: Lower case Latin indexes, i, j, k, l = 1 . . . 3 run over the three ordinary spatial The Mechanics of Spacetime 5 Fig. 2 Multi-axial stress state (a), and a uniaxial deformation of an object (b) from the transparent to the opaque shape. Each component σij represents the stress through the ith surface in the j th direction. The Poisson’s ratio ν measures the effect of the longitudinal stress along the ith direction on the longitudinal strain along the j th direction, for j 6= i. In the case of uniaxial stress state, εjj = (−ν/Y )σii = −νεii . dimensions. Upper case Latin indexes, I, J, K, L = 1 . . . 4 run over the four hyperspace dimensions, while Greek indexes, µ, ν, α = 0 . . . 3 run over the four spacetime dimensions, where indexes 0 and 4 represent, respectively, the time dimension and the extra spatial dimensions. Indexes appearing after a comma represent differentiation with respect to the indexed dimension. For example, ui,j ≡ ∂ui /∂xj . For spacetime, we adopt the space-like metric signature (−, +, +, +) and denote the flat metric tensor as ηµν , where [ηµν ] ≡ diag[−1, 1, 1, 1]. Furthermore, we will use geometric units where the gravitational constant and speed of light are set to unity: G = 1 and c = 1, in which case Einstein’s constant κ = 8π. 2.1 Deformation Basics Let xI be material coordinates assigned to the Cosmic Fabric, and let gIJ be the metric tensor of the fabric. Material coordinates are those that remain attached to the fabric during deformation and displace along with it. The metric tensor defines how coordinate differences relate to distances. Thus, the distance ds between two nearby material points is given by, ds2 = gIJ dxI dxJ (2.1) The distance ds between the same two material points prior to deformation is given by, ds2 = g IJ dxI dxJ (2.2) where g IJ is the metric tensor of the undeformed fabric. The strain tensor εIJ describes the amount of relative length change during deformation. By definition, εIJ is such that, 2εIJ dxI dxJ = ds2 − ds2 = (gIJ − g IJ )dxI dxJ 1 ∴ εIJ = (gIJ − g IJ ) 2 (2.3) 6 T G Tenev, M F Horstemeyer The tensors in the above equations can be related to familiar Solid Mechanics tensors. If we used Cartesian coordinates for the undeformed configuration, then the xI would be the so called “reference coordinates.” The elements of the undeformed metric would be g IJ = δIJ , where δIJ is the Kronecker delta while the elements of the deformed metric would be gIJ = CIJ , where CIJ is the CauchyGreen deformation tensor. Consequently, the definition of the strain tensor εIJ , per the above equation, would correspond to the definition of the Green-Saint Venant strain tensor from Solid Mechanics, EIJ = (CIJ − δIJ )/2. The Young’s elastic modulus Y , which figures in Hooke’s Law (1.1), is the amount of longitudinal stress (force per unit of cross section area) σii in the ith direction needed to produce a unit amount of longitudinal strain εii in the same direction under a uniaxial stress condition (no summation intended over the index i). The effect of longitudinal stress along a given orientation on the longitudinal strains in the transverse orientations is known as the Poisson effect and is measured by the Poisson’s ratio ν (see Fig. 2). 2.2 Postulates We postulate the cosmic fabric to be a (1) thin hyperplate, (2) exhibiting isotropic hyperelastic constitutive behavior, with (3) matter-energy fields as inclusions, and (4) advancing in time at a rate inversely proportional to its thickness. Each of these postulates is described and motivated in the sections below. 2.2.1 Hyperplate Cosmic space is identified with the mid-hypersurface of a hyperplate called the Cosmic Fabric that is thin along the fourth spatial dimension. Because of its correspondence to cosmic space, the intrinsic curvature, R3D , of the fabric’s midhypersurface corresponds to that of three-dimensional (3D) space. Likewise, the intrinsic curvature R of the fabric’s world volume, corresponds to that of fourdimensional (4D) spacetime. The term “world volume” refers to the four-dimensional shape traced out by an object in spacetime as it advances in time. From a Solid Mechanics perspective, the fabric must have thickness so it can resist bending. At the same time, the thickness must be very small so that the fabric can behave as an essentially 3D object at ordinary length scales and be an appropriate analogy of 3D physical space. The thickness itself defines a microscopic length scale at which the behavior of the physical world would have to differ significantly from our ordinary experience. A value equal to or comparable to Planck’s length lp meets this criteria. However, the exact value of the thickness is not essential to the model as long as it is small but not vanishingly so. We imagine the cosmic fabric as immersed in a 4D hyperspace and able to bend along the x4 dimension. The 4D hyperspace outside of the cosmic fabric is not directly observable, but the bending of the fabric into x4 can be observed indirectly by measuring the intrinsic curvature R3D of the fabric’s mid-hypersurface. F GR Let gIJ be the fabric’s metric and let gµν be the spacetime metric of General Relativity. We can identify the 3D spatial components of the two metrics with each other as follows, F GR gij = gij ≡ gij (2.4) The Mechanics of Spacetime 7 so from here on, we can drop the superscripts F and GR , and understand from the context and from the use of subscripts which metric tensor components are being referenced. 2.2.2 Isotropicity and Hyperelasticity The cosmic fabric exhibits an isotropic hyperelastic constitutive behavior subject to Hooke’s Law (1.1) and the Poisson effect (see Fig. 2b). The meaning of stress, strain, and the Poisson effect is extended from 3D to 4D as follows. First, we observe that both strain and Poisson effect are geometric in nature and naturally extend to the 4D hyperspace. Next, we define stress σIJ to be the thermodynamic conjugate of the strain εIJ , 1 ∂U σ IJ = p |g| ∂εIJ (2.5) where U is the elastic energy density, which is an intensive scalar quantity and thus also readily generalizable to 4D. Note the use of script symbols, such as p U, to denote tensor or scalar densities. A density quantity includes the factor |g| p 3D 3D (or |g | for 3D) as part of its definition, where g (or g ) is the determinant of the metric. This factor represents the ratio between the coordinate volume element d4 x, which depends on the choice of coordinates, and the proper volume element, which is invariant under coordinate transformations. The 3D components σij coincide with the conventional definition of stress as a traction force per unit of cross-section area, while the additional components σI0 = σ0I are stress-like quantities. 2.2.3 Inclusions Matter-energy fields behave as inclusions in the fabric inducing membrane strains leading to transverse displacements and hence bending (Fig. 1). Thus, matter is a source of volumetric strain, so: ( ε3D ,kk ε3D ,kk >0 =0 within a matter-energy inclusion outside of inclusions (2.6) 2 3D where ε3D ≡ εii is the volumetric strain, and ε3D is its Laplacian. ,kk ≡ ∇ ε The term “membrane” strain (or stress) refers to strains (or stresses) that change in-plane but are uniform across the thickness of the fabric as opposed to bending strains (or stresses) that switch sign through the thickness across the midhypersurface. Note that it is not the spatial extent of matter that causes it to displace the fabric material, but rather it is the mass content of matter. In Section 3.5 we attempt to quantify more precisely the effect of inclusions on the fabric. 8 T G Tenev, M F Horstemeyer 2.2.4 Lapse Rate The lapse rate of proper time is postulated to be the inverse of the fabric’s transverse stretch. Therefore, in orthonormal coordinates as we have adopted here, dτ = (−g00 )1/2 = (g44 )−1/2 dx0 ∴ g00 = −(g44 )−1 (2.7) where τ designates proper time. Under the weak field condition, which is analogous to a small strain condition, the metric tensors can be approximated as, gIJ = δIJ + 2εIJ , |εIJ | 1 gµν = ηµν + 2εµν , |εµν | 1 (2.8) where we have identified εij with the spatial components of a gravitational gauge as per [26, Ch.18], and we have also applied the notation to the space-time components 2ε0µ and 2εi0 of the gauge. Thus, we have fixed the gauge to be 2εµν . This gauge has the physical meaning of being the fabric’s material strain and for that reason will be used throughout the paper. Note that, except under special conditions, this gauge does not necessarily comply with the harmonic gauge condition, which is often used in Linearized Gravity. Under weak field conditions, the postulated equation (2.7) can be approximated as follows, dτ = −1 + 2ε00 = −(1 + 2ε44 )−1 ≈ −1 + 2ε44 dx0 ∴ε00 ≈ ε44 (2.9) 2.3 Weak and Nearly Static Fields Condition To keep the math tractable, we conduct our study under the assumption of weak and nearly static fields. We believe that this assumption is not fundamental to the model, and that it could be relaxed or removed in the future. As will be shown in Section 3.1, the fabric strain is analogous to the gravitational potential, so the weak field condition, which is the subject of Linearized Gravity (see [26, Ch. 18]) is analogous to the small strain condition, which is the subject of Solid Mechanic’s Infinitesimal Strain Theory. We consider a gravitational potential Φ to be weak if Φ/c2 1. By this definition, most gravitational fields that we experience on an everyday basis are weak. For example, the values for Φ/c2 at the Earth’s surface due to the gravitational fields of the Earth, Sun, and Milky way are 6.7 × 10−10 , 1.0 × 10−8 , and 1.4 × 10−6 [33], respectively. As such, we consider these gravity fields to be weak. Except in regards to gravity waves (Section 3.3), we will assume nearly static (or slow moving) fields in addition to weak fields. A field is considered slow moving if its velocity v satisfies, v 2 /c2 1, which is the case for most gravitational or strain fields that we experience. The nearly static field assumption means that differentiation with respect to time results in negligibly small values. The Mechanics of Spacetime 9 Fig. 3 The cosmic fabric is treated as a stack of three-dimensional hypersurfaces Σξ each parameterized by ξ ≡ x4 = const. The reference thickness L ≡ ∆ξ is the difference between the ξ coordinates of the two faces of the fabric. The actual thickness is N L, where the contraction N = N (xi ) varies from point to point xi along the mid-hypersurface Σ0 . 2.4 Bending Energy Density Rather than treating the fabric as a 4D hyperplate, it is convenient to approximate it as a 3D hypersurface. This can be done once we have averaged the fabric’s elastic energy density across its thickness and assigned it to its mid-hypersurface. At that point, we can use the fabric’s mid-hypersurface as a proxy instead of the fabric in future calculations. To compute UB , we adapt the work of Efrati et al. [6] concerning the bending of conventional thin plates. For ease of notation, let ξ ≡ x4 denote the coordinate offset from the mid-hypersurface of the fabric. The fabric, having a reference thickness L, is regarded as foliated into infinitely many hypersurfaces Σξ each parameterized by ξ = const. (Fig. 3). We carry over the simplifying assumption from Kirchoff-Love thin plate theory [1] to thin hyperplates and assume that the set of material points along any given hypersurface that were along a normal prior to bending remain along the normal after bending. Consequently, the fabric’s four-dimensional metric gIJ can be expressed in terms of the three-dimensional hypersurface metrics gij = gij (ξ) as follows, gIJ 2 N 0 ≈ 0 gij (2.10) where gij = gij (ξ) is the metric of a given hypersurface Σξ that is offset by ξ from the mid-hypersurface, and N is the fabric’s contraction along the x4 dimension (Fig. 3). It can be shown [6] that the metric gij of each Σξ takes the form, gij = aij − 2N bij ξ + N 2 cij ξ 2 (2.11) where aij = aij (xi ) and bij = bij (xi ) are, respectively, the first and second fundamental forms of the mid-hypersurface, and cij = akl bik bjl . In the same way, let g IJ represent the undeformed metric of the fabric, and g ij (ξ) be the undeformed 10 T G Tenev, M F Horstemeyer metric of each hypersurface Σξ . g µν ≈ 2 N 0 0 g ij where (2.12) 2 g ij = aij − 2N bij ξ + N cij ξ 2 The total elastic energy density of any solid is half of the inner product of its stress and strain tensors. Applying Hooke’s Law (1.1), the total elastic energy density Uξ of each hypersurface Σξ is given by, Uξ = 1 ij σ εij 2 q C ijkl such that q 1 ijkl C εij εkl |g 3D | 2 Y ν ij kl ik jl ≡ g g +g g 1 + ν 1 − 2ν |g 3D | = (2.13) where σij = σij (ξ) and εij = p εij (ξ) are,p respectively, the stress and strain at each hypersurface Σξ . The factor |g 3D | ≡ det gij converts coordinate unit volume into proper volume, which is required for Uξ to be a density. Next, we compute the total elastic energy density U averaged across the fabric’s thickness, and we separate it into a bending term UB and a membrane stretch term UM . For this purpose, we split the strain at each surface, εij into a membrane strain B εM ij and a bending strain εij as follows: 1 B (gij − g ij ) = εM ij + εij 2 1 = (aij − aij ) 2 = − N bij − N bij ξ + O(ξ 2 ) εij = εM ij εB ij U= 1 L L 2 Z 1 = 2L (2.14) Uξ N dξ −L 2 Z L 2 −L 2 M B B M B B M C ijkl (εM ij εkl + εij εkl + [εij εkl + εij εkl ]) dξ = UM + UB Z L 2 1 M C ijkl εM UM = ij εkl dξ 2L − L 2 Z L 2 1 B UB = C ijkl εB ij εkl dξ 2L − L (2.15) 2 The mixed terms inside the square brackets in Equation (2.15) vanish under integration because the bending strain reverses sign across the mid-hypersurface; B M M hence εB ij = εij (ξ) is an odd function, while εij = εij (ξ) is an even function. For the remainder of this subsection, we focus on evaluating the term UB . The term UM will be addressed in the following subsection where we show that it vanishes under appropriately chosen material properties and deformation kinematics. The Mechanics of Spacetime 11 Evaluating UB from Equation (2.15), we obtain, q UB = N bij − N bij N bkl − N bkl N |g 3D | + O(L3 ) (2.16) 3D 3D Using the identity, Rlijk = bik bjl − bij bkl , where Rlijk is the Riemann curvature 3 tensor of the mid-hypersurface, and setting O(L ) = 0, we can express UB in terms of the intrinsic three-dimensional spatial curvature R3D as follows, q L2 Y 1−ν i k UB = − N 2 R3D + bi bk N |g 3D | 24(1 + ν) 1 − 2ν (2.17) q 2 L ijkl 2 3D + C −N N bij bkl − N bij N bkl + N bij bkl N |g | 24 The undeformed geometry and Poisson’s ratio of the cosmic fabric had remained unspecified as freedoms to be fixed at a later time such as now. Requiring that the cosmic fabric’s mid-hypersurface has a flat undeformed geometry, bij = 0, causes the last term in Equation (2.17) to vanish. The bii bkk term would also vanish if we chose Poisson’s ratio ν = 1. In this case, the bending energy becomes the following, q Y L2 3 3D UB = − N R |g 3D | (2.18) 48 subject to the conditions, bij = 0, and ν=1 (2.19) 2.5 Membrane Energy Density We now show that for any given small-strain deformed configuration, we can identify a material displacement field that results in no membrane energy. Consequently, we conclude that the bending energy UB is the only contribution to the total elastic energy of the fabric for the case of slow moving fields. Since General Relativity (GR) is only concerned with the curvature of the deformed body, in developing the material analogy of GR we have freedom to prescribe a specific material displacement field for the deformation. Let us consider a displacement field where each point of the mid-hypersurface, x4 = 0, of the fabric is displaced by the amount w = w(xi ) along a geodesic normal to the hypersurface. It should be evident that using such a displacement field, one can deform a flat body into any given shape that represents a small deviation from flatness and does not contain folds. Flamm’s paraboloid, which is used to illustrate the geometry of space around a static gravitating body, is an example of such a shape. Let y I be the coordinate of the position to which the material point at xi was displaced. Thus, y i = xi and y 4 = w. The metric tensor of the deformed hypersurface can be computed from the dot product of the position differentials as follows, K K gij = y,i y,j = xk,i xk,j + w,i w,j = δij + w,i w,j 1 1 ∴ εij = (gij − δij ) = w,i w,j 2 2 (2.20) 12 T G Tenev, M F Horstemeyer Using the formula for elastic energy density, UM = σ kl εkl /2 and applying Hooke’s Law (1.1) to Equation (2.20) with ν = 1, we find, UM ∝ σ kl εkl ∝ (g ik g jl − g ij g kl )εij εkl = = εkj εjk − εjj εkk ∝ w,k w,j w,j w,k − w,j w,j w,k w,k = 0 (2.21) ∴ UM = 0 Fixing the fabric’s deformation to material displacements only along the hypersurface normals is a valid approximation under the assumption of nearly static fields. In such cases, the reason for the deformation would have been to geometrically accommodate inclusions by bending into the x4 dimension. Once bending has taken place, the material points of the fabric can shift within the plane of the fabric to minimize its membrane energy without affecting the geometrical constraints imposed by the inclusion. For nearly static situations, we have shown that the membrane energy can be minimized to where it vanishes. In such cases, the net displacement would have taken the form described in this subsection. 2.6 Lagrangian Density Ignoring the kinetic energy component, under the simplifying assumption of slow moving fields, the Lagrangian density is LF = −UB ∝ R3D . We now derive an expression for LF in terms of the Ricci curvature R of the fabric’s world volume. Let us adopt the conditions (2.19), and choose orthonormal coordinates, such that g ij = δij , where δij is the Kronecker delta. Then, gij = δij + 2εij , and so 2εij plays the role of a gauge for the purposes of the linearized curvature equations per [26, Ch. 18]. We extend this gauge to spacetime by defining additional strain-like components ε0µ and εk0 such that gµν = ηµν +2εµν , where gµν is the metric of the fabric’s world volume, and ηµν is the Minkowski metric. Using the gauge-invariant linearized expression for R, R = 2 −εµµ,αα + εαµ,αµ = 2 −εii,kk + εik,ik − ε00,kk − εkk,00 + 2ε0k,0k (2.22) ≈ R3D + 2ε00,kk In the last step of the above derivation, we have recognized that the purely spatial terms add up to the gauge-independent linearized expression for R3D . Furthermore, the terms differentiated with respect to x0 are negligible because of the slow moving fields assumption, and lowering or raising a single 0 index, which is done using ηµν , changes the sign of the term. Next, we show that ε00,kk = 0 in free space. Because the fabric is very thin outside of inclusions, there are only in-plane stresses and hence plane-stress conditions apply. Therefore, from Hooke’s Law (1.1) and considering that ν = 1, ( Y ε3D = (1 − 2ν)σ 3D Y ε44 = −νσ 3D (2.23) ν 3D 3D ∴ ε44 = − ε =ε 1 − 2ν The Mechanics of Spacetime 13 where σ 3D ≡ σkk and ε3D ≡ εkk are the 3D volumetric stress and strain, respectively, and Y is the Young’s modulus of elasticity. Combining the above result with the Inclusions and Lapse Rate postulates of the Cosmic Fabric Model, we conclude that, ε00 ≈ ε44 = ε3D ∴ ε00,kk = ε3D ,kk = 0 (2.24) Based on the results form the previous subsections, the Lagrangian density of the fabric for the case of weak and slow moving fields is simply, LF = −UB = Y L2 3 3D N R 48 q |g 3D | (2.25) Let g ≡ det gµν . For the case of slow and weak fields, g ≈ g 3D g00 = −g 3D (g44 )−1 = −g 3D N −2 q p ∴ |g 3D | = N |g| (2.26) Combining Equations (2.25), (2.26), (2.24), and (2.22) while also letting N ≈ 1, we finally arrive at, LF = Y L2 p R |g| 48 (2.27) which has the same form as the Einstein-Hilbert Lagrangian density. 3 Discussion In the previous section, we postulated a material body, which we named the “cosmic fabric” whose constitutive behavior outside of inclusions is analogous to the behavior of gravity. For the analogy to be useful, it should allow us to map between notions in Solid Mechanics and General Relativity. Such a mapping is possible on the basis of identifying the fabric Lagrangian density LF with the Lagrangian density from the Einstein-Hilbert action, LEH , as applying to free space. Specifically, LF = Y L2 p 1 p R |g| = LEH = R |g| 48 2κ (3.1) where κ is the Einstein constant. In the subsections below, we discuss the correspondence between mechanical properties of the cosmic fabric and known properties of gravity. 14 T G Tenev, M F Horstemeyer 3.1 Fabric Strain and Gravitational Potential A General Relativity result is that the ratio by which clocks slow down in gravitational field, also known as “time dilation,” is given by, p ∆tΦ = 1 + 2Φ/c2 ≈ 1 + Φ/c2 ∆t0 (3.2) where Φ is the gravitational potential, ∆t0 is a time interval measured away from gravitational fields, while ∆tΦ is a corresponding time interval measured within a gravitational field of potential Φ. In the coordinate system we have adopted, the values of x0 correspond to the clock measurements away from gravitational fields, so by the Lapse Rate postulate (2.7) and applying Equation (2.23) we conclude that, ∆tΦ dτ = ∆t0 dx0 1 + Φ/c2 = 1 − ε00 (3.3) ∴ −Φ/c2 = ε00 = ε3D Thus, the gravitational potential corresponds to the volumetric expansion of the fabric. The above result is consistent with the Schwarzchild metric in the exterior of a non-rotating body, because the space metric coefficients are the inverses of the time coefficient. 3.2 Poisson’s Ratio and the Substructure of Space Known materials with a Poisson’s ratio of ν = 1 have a fibrous substructure, which suggests that the cosmic fabric is, in fact, a fabric! For ν = 1, the bulk modulus is K = Y /[3(1−2ν)] < 0. A negative bulk modulus means that compressing the fabric results in an overall increase of the material volume and vice versa. Although such behavior is unusual for most conventional materials, there are recently discovered compressive dilatant [29] and stretch densifying [4] materials, for which ν = 1 in either compression or tension, respectively. Compressive dilatant materials are artificially manufactured and their substructure consists of entangled stiff wires. Stretch densifying materials, have textile-like substructure comprised of woven threads each consisting of twisted fibers. 3.3 Fabric Vibrations and Gravitational Waves Having Poisson’s ratio ν = 1 also implies that there can only be transverse (shear) waves in the fabric but no longitudinal (pressure) waves. The shear modulus µ and the p-wave modulus M are as follows, Y Y = 2(1 + ν) 4 1−ν M =Y =0 (1 − 2ν)(1 + ν) µ= (3.4) The Mechanics of Spacetime 15 p implying that the transverse (shear) wavepvelocity vs = µ/ρ 6= 0, while the longitudinal (pressure) wave velocity vp = M/ρ = 0. This result shows why the speed of light is the fastest entity of the universe, given that a longitudinal wave is typically faster than a shear wave. For a shear wave to be the fastest, the Poisson’s Ratio must be 1.0. This conclusion is consistent with observations, because all known waves that propagate in free space, such as gravity or electromagnetic waves, are transverse. Let us consider the analogy between shear waves in the fabric and gravitational waves. Such an analogy depends on demonstrating that the fabric’s behavior parallels that of spacetime for high velocity fields as well. We leave the rigorous proof for a future article, and for the rest of this subsection we assume that the fabric’s behavior implied by the Lagrangian (2.27) also holds for high velocities. Based on this assumption, we proceed to investigate in-plane shear waves propagating through the fabric and their correspondence to gravitational waves. First, we show that if static fields are negligible and in the absence of torsion, then the gravitational gauge defined by the strain 2εµν satisfies the harmonic gauge condition, εµα,α = (1/2)εαα,µ . For shear waves, ε3D = 0, and by Equation (2.24), e00 = 0, implying that εαα = 0, so proving the condition reduces to demonstrating that, εµα ,α = 0. Furthermore, the shear time-space components must vanish, ε4j = εj4 = 0 = ε0j = εj0 , because we are assuming negligible static fields and in-plane shear waves. Therefore, in order to prove that the harmonic gauge condition holds, we just need to show that εik,k = 0. Let ui be the material displacement field. In terms of the displacement field, the strain is 2εij = ui,j + uj,i , and so, 2εij = 2uj,i + [ui,j − uj,i ] 2εik,k = 2uk,ki + [ui,k − uk,i ],k (3.5) But, uk,ki = 0 since εkk = uk,k = 0. The difference in the square brackets corresponds to material torsion and must vanish too, thus concluding, eik,k = 0 (3.6) Since εµν satisfies the harmonic gauge condition, we can apply the linearized α approximation for the Ricci tensor, Rµν ≈ −εµν,α . After substituting into the Einstein Field Equations (1.2), and taking into account that R ≈ εαα,µµ = 0, and that in empty space Tµν = 0, we arrive at, α εµν,α = εij,kk − εij,00 = 0 ∴εij,00 = εij,kk (3.7) which is a wave equation with solutions that are traveling waves at the speed of light c. To see this clearly, let us re-write Equation (3.7) in terms of the time variable t, where x0 ≡ ct, and using the canonical form derivative operators ∂, and ∇, 1 ∂2 εij = ∇2 εij (3.8) c2 ∂t2 The above equation can be related to the Solid Mechanics equation for the propagation of a shear wave in elastic medium with density ρ and shear modulus µ. In the absence of body forces, the equation of motion is the following, ρ ∂2 ui = σij,j ∂t2 (3.9) 16 T G Tenev, M F Horstemeyer Applying Hooke’s Law (1.1) and recognizing that, εij = (ui,j + uj,i )/2, µ = Y /[2(ν + 1)], and uk,k = εkk = 0, we arrive at, ρ ρ ∂2 ui = µ∇2 ui ∂t2 ∂2 (ui,j + uj,i ) = µ∇2 (ui,j + uj,i ) ∂t2 ∂2 ∴ ρ 2 εij = µ∇2 εij ∂t (3.10) The parallel between Equations (3.8) and (3.10) confirms that gravitational waves are analogous to shear waves propagating in a solid material and that furthermore the speed of propagation, which is the speed of light c, is related to the shear modulus and density of the medium per c2 = µ/ρ. Although Equation (3.7) suggests that there are ostensibly ten strain components, εαβ , oscillating independently, in reality only two are independent and the rest are coupled to the two. To show this, consider a traveling wave, which corresponds to a gravity wave, propagating along the x3 direction. Then ε3α = εα3 = 0 for the wave to be a shear wave. Furthermore, as shown earlier, ε00 = ε3D = 0 and εj0 = ε0j = 0. Finally, we have ε3D = ε11 + ε22 = 0, because ε33 = 0 already. Therefore, ε11 = −ε22 ε12 = ε21 (3.11) are the only two independent degrees of freedom left implying just two types of wave polarizations. The fact that we are using the material strain, which has concrete physical meaning, ensures that the waves must also be physical as opposed to being mere coordinate displacements. This result, derived from a Solid Mechanic’s perspective, is consistent with the analogous result from General Relativity about the polarization of gravitational waves [26, Ch. 35]. 3.4 Elastic Modulus and Density of Free Space From the result in Equation (3.1), the fabric’s elastic modulus Y could be computed given an estimate p for the fabric’s thickness L. As reasoned in Section 2.2.1, Planck’s length lp ≡ ~G/c3 is a suitable estimate for L, where ~ is the reduced Planck’s constant. So, assuming L ∼ lp , we can estimate Y to be, Y ∼ 24 = 4.4 × 10113 N m−2 lp2 κ (3.12) The density of the fabric ρ is related to the wave speed and shear modulus, as shown in Section 3.3, and can now be computed, ρ= Y µ = 2 ∼ 1.3 × 1096 kg m−3 c2 4c (3.13) In accordance with the Cosmic Fabric analogy, the density of the fabric corresponds to the density of free space, which is also known as the zero-point energy The Mechanics of Spacetime 17 density. The computed value for ρ agrees to an order of magnitude with the predictions of Quantum Field Theory (∼ 1096 kg m−3 ) for the energy density of free space [30]. Note that the predictions of Quantum Field Theory are also based on using Planck’s length lp as a length-scale parameter. 3.5 Generalizing the Cosmic Fabric Model The Cosmic Fabric analogy to cosmic space was demonstrated subject to certain simplifying conditions such as small strain (weak gravity), nearly static equilibrium (slow fields) and outside of inclusions (free space). In this subsection, we consider how each of these conditions might be removed or relaxed in a future generalization of the model. We also sketch how we could quantify the effect of inclusions on the fabric, and investigate the significance of non-flat unstrained geometry. The small strain (weak gravity) condition was imposed so we could use the linearized equations for strain and, analogously, the linearized equations for gravity. To relax this condition, we will need to account for the higher order terms in the equations of strain and also use covariant derivatives instead of conventional differentiation for all field variables. Some materials exhibit nonlinear elastic behavior under large strains. By requiring that the fabric behaves analogously to cosmic space at large strains (strong gravity), we would determine whether the fabric is one such material. Nonlinear elastic behavior is related to the substructure of the material, and discovering such behavior would suggest clues about the substructure of cosmic space. Imposing the nearly static (slow fields) condition allowed us to ignore the kinetic energy term in the fabric’s Lagrangian. It also let us assume specific bending kinematics that minimizes the membrane energy of the fabric and, in Section 2.5, we showed that such kinematics result in zero membrane energy. Without the condition of nearly static equilibrium (slow fields), we will need to take into account the kinetic and membrane energies of the fabric and also consider a more complex deformation state. One possible simplification would be to concentrate on the condition without bending (away from static gravitational fields), and derive a closed-form result for the fabric’s Lagrangian. This is the condition under which we would study gravitational waves as discussed in Section 3.3. The mathematical complexities resulting in the general non-static (fast fields) case will probably require that it be studied using numerical techniques. The reason we focused on deriving the fabric’s Lagrangian outside of inclusions was so that we could eliminate the ε00,kk term in Equation (2.22), which we showed equaled the Laplacian of the volumetric strain per Equation (2.24). The latter had to vanish because of the Inclusions Postulate (Section 2.2.3). Consequently we could conclude that R ≈ R3D and show that the Fabric Lagrangian outside of inclusions (2.27) takes a form analogous to the Einstein-Hilbert Lagrangian for free space (3.1). It maybe possible to show that in the case of inclusions, the term ε00,kk in Equation (2.22) yields an additional Lagrangian term that is analogous to the energy-matter Lagrangian term LM in the generalized Einstein-Hilbert action p R SEH = (R/2κ + LM ) |g| dx4 . We propose the idea below and leave the rigorous derivation for future research. 18 T G Tenev, M F Horstemeyer First, we quantify the effect of inclusions as a source of strain by restating the Inclusions Postulate (2.6) as follows, ε3D ,kk = O(1)κρ (3.14) where ρ is the matter-energy density of the inclusion, κ is the Einstein constant, and O(1) stands for some dimensionless coefficient of order unity. In the context of General Relativity, mass can be related to geometry through its Schwarzchild radius. Thus, one meter of mass is the amount of mass whose Schwarzchild radius is two meters. This is in fact the mass-geometry relationship implied when using geometric units, where the speed of light c and the gravitational constant G are set to unity. In the same way, the geometric significance of a matter-energy field, represented by the right hand side of Equation (3.14), can be understood as the Schwarzchild radius density and κ as a units conversion factor (in geometric units κ = 8π). Thus, the meaning of Equation (3.14) is to postulate that the Schwarzchild radius density of a matter-energy field is a source of volumetric strain in the cosmic fabric. Next, we relate ε3D ,kk to ε00,kk via the Poisson effect and the Lapse Rate postulate. Outside of inclusions, we had shown in Equations (2.23) and (2.24) that ε3D ,kk = ε00,kk , but with an inclusion we will have to account for the transverse stress, so in general, ε00,kk = O(1)ε3D ,kk = O(1)κρ (3.15) Substituting the above results into Equation (2.22) and retracing the steps in the derivation of the fabric’s Lagrangian LF in Section 2.6, we conclude that, p Y L2 (R − 2O(1)κρ) |g| 48 p κY L2 R = |g| − O(1)ρ 24 2κ LF = (3.16) We note that, within General Relativity, the matter p Lagrangian density LM for a nearly static fluid of density ρ is LM = −O(1)ρ |g| [25]. Therefore, by identifying Y L2 /24 = 1/κ as was suggested already by Equation (3.1) and choosing appropriately the various O(1) factors mentioned above, we can identify the full fabric Lagrangian LF that accounts for inclusions with the full Einstein-Hilbert Lagrangian which in turn accounts for matter energy fields as follows, LF = LEH = LB + LM , where Rp LB = |g| is due to bending 2κ p LM = −O(1)ρ |g| is due to matter-energy fields (3.17) Assuming that the above result can be demonstrated rigorously, it would imply that matter-energy is analogous to the strain energy associated with introducing an inclusion into the cosmic fabric. The Mechanics of Spacetime 19 3.6 Possible Explanation for Dark Energy and Dark Matter An initial analysis suggests that if we accounted for non-flat unstrained geometry of the fabric (inherent curvature of cosmic space not due to matter-energy fields), then we may be able to explain the observational results that are currently attributed to dark energy and dark matter. For a non-flat unstrained geometry, the last term in Equation (2.17) yields an additional Lagrangian density term, LB , that must be included in the fabric’s Lagrangian density (3.16) as follows, LF = LB + LB + L M (3.18) By the Cosmic Fabric analogy, the LB term is due to the inherent curvature of cosmic space. Suppose that cosmic space had a nearly constant inherent global curvature of radius comparable to the Hubble radius, that is c/H0 (about 13.7 billion light years), where H0 is the Hubble parameter. We may be able to show that such curvature can explain the presence of a Cosmological Constant as well as phenomena currently attributed to dark matter. The main distinction between these two explanations depends on whether the effect of the global curvature is considered away from or near matter-energy fields. Below is a brief sketch of the idea, whose full development is left for future research. Away from inclusions (matter-energy fields), LB would remain constant and, upon variation of the fabric’s action, would yield a Cosmological Constant Λ term as part of the resulting field equations, which is associated with dark energy. The value of Λ would be such that Λ−1/2 ∼ c/H0 , which is consistent with the observationally determined value of the known Cosmological Constant. Near inclusions (matter-energy fields), the inherent curvature of the fabric would interact with the bending caused by the matter-energy field as represented by the mixed bij and bij terms in Equation (2.17). The overall effect would be to boost the gravitational acceleration due to matter, which is observationally equivalent to the presence of fictitious invisible mass, also known as dark matter. The boost in gravitational acceleration due to the inherent curvature of cosmic space, would only be appreciable for exceedingly small accelerations a where a ∼ cH0 . Such conclusion would be consistent with the calculations of the Modified Newtonian Dynamics theory [24], which introduces an acceleration scale parameter a0 whose empirically determined value happens to be such that a0 ∼ cH0 . 4 Summary and Conclusion In this paper, we showed that the behavior of spacetime per Einstein’s Field Equations (1.2) is analogous to that of an appropriately chosen material body, termed the “cosmic fabric” that is governed by a simple constitutive relation per Hooke’s Law (1.1). In Section 2, we postulated several basic properties of the fabric and how they correspond to physical space and matter in space. Other properties, such as the elastic moduli and undeformed geometry, were left unconstrained as model parameters. These were fixed once we computed the Lagrangian of the fabric and required that it be identified with the Einstein-Hilbert Lagrangian of General Relativity. After the Cosmic Fabric model was calibrated in this way, in Section 3, it was applied to interpret various properties of gravity in terms of the fabric’s mechanics and vice versa. To a great extent, the interpretations seemed physically 20 T G Tenev, M F Horstemeyer Table 1 Comparison between the General Relativity and Solid Mechanics Perspectives. General Relativity Perspective Solid Mechanics Perspective Physical space Mid-hypersurface of a hyperplate called “cosmic fabric”. Spacetime The world volume of the cosmic fabric’s mid-hypersurface Intrinsic curvature of space Intrinsic curvature of the fabric’s midhypersurface Intrinsic curvature of spacetime Intrinsic curvature of the fabric’s world volume Gravitational potential Φ in free space Volumetric strain ε3D outside of inclusions, such that ε3D = −Φ/c2 Gravitational waves Shear waves traveling at the speed of light Matter curves spacetime Matter induces prescribed strain causing the fabric to bend Action integral in free space, Z p 1 R |g| d4 x S= 2κ Action integral outside of inclusions, Constants of Nature: Elastic constants: G, ~, c S= L2 Y 24 Z R p |g| d4 x Y = 6c7 /2π~G2 , ν=1 meaningful from both perspectives of General Relativity and Solid Mechanics. Table 1 summarizes the correspondence between concepts from one field to analogous concepts in the other. The research presented in this paper suggests an equivalence between postulating the field equations of General Relativity and postulating a cosmic fabric having material-like properties as described here. We believe that these are two different approaches for studying the same underlying reality. The Cosmic Fabric model introduces a new paradigm for interpreting cosmological observations based on well-established ideas from Solid Mechanics. In recent decades, Solid Mechanics has made significant advancements in describing the structure of materials at various length and time scales ranging from electrons to large scale engineering structures [14, 15]. Such advances, in conjunction with the advances in high-performance computing, have made possible the construction of multiscale models that accurately simulate the behavior of metals, ceramics, polymers, and biomaterials. The Cosmic Fabric model should enable the application of these techniques to simulate both the fine and large scale structures of the cosmos, and consequently, to address some of the outstanding problems in Cosmology, such as those pertaining to the density of free space, dark energy, and dark matter. Our research in developing and applying the Cosmic Fabric model is still ongoing, and the results we have shared are subject to some simplifying assumptions. In the future, we hope to relax these assumptions if not eliminate them completely. Nevertheless, even at the current stage of the work the results appear promising and useful. The Mechanics of Spacetime 21 Acknowledgements MFH would like to acknowledge the Center for Advanced Vehicular Systems (CAVS) at Mississippi State University for support of this work. The authors would also like to thank Russ Humphreys and Anzhong Wang for their insight with respect to Cosmology, and Shantia Yarahmadian for his feedback with respect to Mathematics. Conflict of Interest The authors declare that they have no conflict of interest References 1. A.E.H.Love: The Small Free Vibrations and Deformation of a Thin Elastic Shell on JSTOR. Philosophical Transactions of the Royal Society of London. A 179, 491–546 (1888) 2. A.L.Cauchy: De la pression ou tension dans un corps solide, [On the pressure or tension in a solid body]. Exercices de Mathématiques 2, 42 (1827) 3. Antoci, S., Mihich, L.: A four-dimensional Hooke’s law can encompass linear elasticity and inertia. Nuovo Cimento B p. 8 (1999) 4. Baughman, R.H., Fonseca, A.F.: Architectured materials: Straining to expand entanglements. Nature materials 15(1), 7–8 (2016) 5. Beau, M.: Théorie des champs des contraintes et déformations en relativité générale et expansion cosmologique: Theory of stress and strain fields in general relativity and cosmological expansion. Annales de la Fondation Louis de Broglie 40 (2015) 6. Efrati, E., Sharon, E., Kupferman, R.: Elastic theory of unconstrained non-Euclidean plates. Journal of the Mechanics and Physics of Solids 57(4), 762–775 (2008) 7. Einstein, A.: Zur Elektrodynamik bewegter Körper. Annalen der Physik 322(10), 891–921 (1905) 8. Einstein, A.: Die Grundlage der allgemeinen Relativitätstheorie. Annalen der Physik 354(7), 769–822 (1916) 9. Einstein, A.: Ether and the Theory of Relativity (1922). URL http://www-history.mcs. st-and.ac.uk/Extras/Einstein{_}ether.html 10. FitzGerald, G.F.: The Ether and the Earth’s Atmosphere. Science (New York, N.Y.) 13(328), 390 (1889) 11. Gerlach, U.H., Scott, J.F.: Metric elasticity in a collapsing star: Gravitational radiation coupled to torsional motion. Physical Review D 34(12), 3638–3649 (1986) 12. Grot, R., Eringen, A.: Relativistic continuum mechanics: Part IIelectromagnetic interactions with matter. International Journal of Engineering Science (1966) 13. Hooke, R.: Lectures de Potentia Restitutiva, Or of Spring Explaining the Power of Springing Bodies. - 1678. John Martyn, London (1678) 14. Horstemeyer, M.: From Atoms to Autos: Part 1 Microstructure-Property Modeling. Tech. rep., Sandia National Laboratories (2000) 15. Horstemeyer, M.F.: Integrated Computational Materials Engineering (ICME) for Metals: Using Multiscale Modeling to Invigorate Engineering Design with Science, 1st edn. WileyTMS (2012) 16. Isaac Newton: The Third Book of Opticks (1718) (1718) 17. Karlovini, M., Samuelsson, L.: Elastic Stars in General Relativity: I. Foundations and Equilibrium Models (2002). URL http://arxiv.org/abs/gr-qc/0211026 18. Kijowski, J.: Relativistic Elastomechanics is a Gauge–Type Theory. Leonardo p. 12 (1994) 19. Kondo, K.: On the analytical and physical foundations of the theory of dislocations and yielding by the differential geometry of continua. International Journal of Engineering Science 2(3), 219–251 (1964) 20. Lorentz, H.A.: De relatieve beweging van de aarde en den aether [The Relative Motion of the Earth and the Aether]. Zittingsverlag Akad. V. Wet. 1, 74–79 (1892) 21. Lorentz, H.A.: Versuch einer Theorie der electrischen und optischen Erscheinungen in bewegten Körpern [Attempt of a Theory of Electrical and Optical Phenomena in Moving Bodies]. E.J. Brill, Leiden (1895) 22. Maugin, G.A.: Magnetized deformable media in general relativity. Ann. Inst. Henri Poincare 4, 275–302 (1971) 23. Michelson, A.A., Morley, E.W.: On the relative motion of the Earth and the luminiferous ether. American Journal of Science s3-34(203), 333–345 (1887) 24. Milgrom, M.: A Modification of the Newtonian Dynamics - Implications for Galaxy Systems. The Astrophysical Journal 270, 384 (1983) 22 T G Tenev, M F Horstemeyer 25. Minazzoli, O., Harko, T.: New derivation of the Lagrangian of a perfect fluid with a barotropic equation of state. Physical Review D - Particles, Fields, Gravitation and Cosmology 86(8), 1–5 (2012) 26. Misner, C.W., Thorne, K.S., Wheeler, J.A.: Gravitation. W. H. Freeman (1973) 27. Rangamani, M.: Gravity and hydrodynamics: lectures on the fluid-gravity correspondence. Classical and Quantum Gravity 26(22), 224,003 (2009) 28. Rayner, C.B.: Elasticity in General Relativity. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 272(1348), 44–53 (1963) 29. Rodney, D., Gadot, B., Martinez, O.R., du Roscoat, S.R., Orgéas, L.: Reversible dilatancy in entangled single-wire materials. Nature materials 15(1), 72–7 (2016) 30. Rugh, S., Zinkernagel, H.: The quantum vacuum and the cosmological constant problem. Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics 33(4), 663–705 (2002) 31. Synge, J.L.: A theory of elasticity in general relativity. Mathematische Zeitschrift 72(1), 82–87 (1959) 32. Tartaglia, A.: Four Dimensional Elasticity and General Relativity (1995). URL http: //arxiv.org/abs/gr-qc/9509043 33. World Heritage Encyclopedia: Gravitational potential (2014). URL http://www. worldheritage.org/article/WHEBN0000579026/Gravitationalpotential
© Copyright 2025 Paperzz