Previous Article in Journal
Mining Scraper Conveyors Chain Drive System Lightweight Design: Based on DEM and Topology Optimization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Equations of Motion and Navier–Stokes Equations

by
Dušan J. Simjanović
1,
Ivana Djurišić
2,
Aleksandra Penjišević
3,
Marko Stefanović
1 and
Branislav M. Randjelović
4,5,*
1
Faculty of Information Technology, Belgrade Metropolitan University, 11000 Belgrade, Serbia
2
Institute for Multidisciplinary Research, University of Belgrade, 11000 Belgrade, Serbia
3
Faculty of Economics, Union—Nikola Tesla University, 11000 Belgrade, Serbia
4
Faculty of Electronic Engineering, University of Niš, Aleksandra Medvedeva 14, 18000 Niš, Serbia
5
Faculty of Teachers Education, University of K. Mitrovica, 38218 Leposavić, Serbia
*
Author to whom correspondence should be addressed.
Computation 2025, 13(9), 226; https://doi.org/10.3390/computation13090226
Submission received: 13 August 2025 / Revised: 11 September 2025 / Accepted: 17 September 2025 / Published: 19 September 2025
(This article belongs to the Section Computational Engineering)

Abstract

In this research, we present the analogies between variational calculations in cosmology and in classical mechanics. Our approach is based on the invariants for transformations of affine connections defined on N-dimensional manifolds (special cases are the 8-dimensional, 5-dimensional, and 4-dimensional manifolds used in cosmology and 2-dimensional manifolds used in classical mechanics). Any of these transformations represents a class of curves on initial manifolds, which transmits to an another class of curves on the current manifolds. The main results of this paper are general equations of motion, which are obtained from the invariants caused by the transformation rule of an initial affine connection to the current one and the corresponding Navier–Stokes equations, recognized in transformations of curves along which moves a fluid particle.

Graphical Abstract

1. Introduction

Differential geometry is a mathematical discipline that describes magnitudes that may help and support processes and theories in various sciences. For example, the well-known Einstein–Hilbert action d 4 x g R , from which one obtains the energy–momentum tensor [1] of the form T μ ν = R μ ν 1 2 R g μ ν , has many different applications, such as in classical mechanics [2].
The tools for describing the objects are relatively complex (tensors as linear functions combined with commutators or indexed notations), so engineers usually stay in Euclidean spaces, where the position of indices is not important but results obtained from models are close to the experimental ones.
The purpose of this manuscript is to describe a differential geometric methodology that could be very useful for theoretical research, as well as for some computations in physics from cosmology to classical mechanics.
We will present universal equations of motion, which, depending on the dimensions of different Riemannian spaces, describe different general physical laws applied in different subjects of physics.

1.1. Riemannian Spaces

In this part of the manuscript, we will review some definitions regarding tensor calculus and Riemannian spaces that are necessary for this research. They may be found in [3,4,5] and in many other articles and monographs.
An N-dimensional manifold M N = M N ( u 0 , , u N 1 ) equipped with a symmetric metric tensor g ^ , whose components are g μ ν , μ , ν = 0 , , N 1 , is an N-dimensional Riemannian space R N . We assume it is det g μ ν 0 . Hence, the contravariant metric tensor is defined by components as g μ ν = g μ ν 1 .
The affine connection coefficients of space R N are Christoffel symbols
Γ μ ν π = 1 2 g π α g μ α , ν g μ ν , α + g ν α , μ ,
where the comma denotes partial derivation, g μ ν , π = g μ ν / u π , and the Einstein summation convention is applied to the mute index α .
Because an indexed value τ j 1 j q i 1 i p of the type ( p , q ) is a tensor if under a change of reference frame O u α to O u α its transformed value τ j 1 j p i 1 i p satisfies the equation
τ j 1 j p i 1 i p = x i 1 i 1 x i p i p x j 1 j 1 x j q j q τ j 1 j q i 1 i p ,
where x μ π = u π u μ and x μ π = u π u μ .
The Christoffell symbols Γ μ ν π are not components of a tensor (of the type ( 1 , 2 ) ) because the following relation holds:
Γ μ ν π = x π π x μ μ x ν ν Γ μ ν π + x π π x μ ν π ,
where x μ ν π = 2 u π u μ u ν .
Anyhow, the trace Γ μ = Γ μ α α is a tensor of the type ( 0 , 1 ) because
Γ μ = x μ μ Γ μ + x α α x α μ α = x μ μ Γ μ + x α α δ α α u μ = x μ μ Γ μ .
The covariant derivative of a tensor a μ π of the type ( 1 , 1 ) , with respect to the affine connection Γ μ ν π , is
a μ | ν π = a μ , ν π + Γ α ν π a μ α Γ μ ν α a α π ,
and it is a tensor of the type ( 1 , 2 ) .
With respect to the difference a μ | ν | σ π a μ | σ | ν π , and the corresponding Ricci identity, the curvature tensor of space R N is defined as
R μ ν σ π = Γ μ ν , σ π Γ μ σ , ν π + Γ μ ν α Γ α σ π Γ μ σ α Γ α ν π .
The corresponding Ricci tensor and scalar curvature of space R N are
R μ σ = R μ α σ α and R = g α β R α β .
If Γ μ ν π = Γ ¯ μ ν π + P μ ν π , where P μ ν π is a tensor of the type ( 1 , 2 ) symmetric by indices μ and ν and Γ ¯ μ ν π are Christoffell symbols of a space R ¯ N , the curvature tensors R ¯ μ ν σ π of the space R ¯ N and R μ ν σ π of the space R N satisfy the relation
R μ ν σ π = R ¯ μ ν σ π + P μ ν σ π P μ σ ν π + P μ ν α P α σ π P μ σ α P α ν π ,
where “||” denotes covariant derivative with respect to the affine connection Γ ¯ μ ν π .

1.2. Research Purposes

We aim to obtain a geometrical object, invariant under a general transformation of an affine connection, whose trace is a monic polynomial of the Ricci tensor of Riemannian space.
In the next step, we will generalize the Einstein–Hilbert action d 4 x g R to the general form d N x g W , where a scalar W is the invariant from the previous paragraph composed of the contravariant metric g μ ν .
In differential geometry, deformations of affine connections are mostly generated by transformation of a class of curves in an initial affine connection space to a curve from an another initial class of curves in the deformed affine connection space. In this research we will define special curves for deformations. Starting from these curves and their transformations, we will obtain the corresponding Navier–Stokes equations for a fluid that moves along the border of initial and deformed three-dimensional manifolds.
In this way, we will obtain a special geometrical object that will, depending on the dimension of internal geometry, connect two physical paradigms (vanishing of variations in actions and dynamics of particles connected in different models and kinds of physics).

2. Review on Invariants for Geometric Mappings

In this section, we will review the process for obtaining invariants of geometrical mappings.
For the transformations Γ ¯ μ ν π Γ μ ν π = Γ μ ν π + P μ ν π , where P μ ν π = δ μ π ψ ν + δ ν π ψ μ g μ ν g π α ψ α , H. Weyl gave the well-known process for determination of the invariant from the corresponding transformation of curvature tensors R ¯ μ ν σ π R μ ν σ π [6]. Together with H. Weyl, T. Y. Thomas [7] obtained an invariant from the transformation Γ ¯ μ ν π Γ μ ν π , but they did not find any other invariant from the corresponding transformation R ¯ μ ν σ π R μ ν σ π , the same as by using Weyl’s methodology. Mikeš and his research group [3,4], like Sinyukov [5], continued with the application of Weyl’s methodology for obtaining invariants for geometric mappings.
N. O. Vesić [8] preferred Weyl’s and Thomas’s methodology. He obtained one (associated) invariant of Thomas type and two (associated) invariants of Weyl type for the studied mappings. This methodology for obtaining invariants has been applied in many articles [9,10,11,12,13,14,15,16,17]. This methodology will be presented below.
Before we review the results regarding the invariants for the mappings between Riemannian spaces, we need to define an ω -curve of the space R N . Namely, a curve = ( t ) is an ω -curve of the space R N if its tangential vector λ π = d π d t satisfies the relation
d 2 π d t 2 + Γ α β π d α d t d β d t = ρ d π d t + ω α β π d α d t d β d t ,
where ω μ ν π is a symmetric tensor of the type ( 1 , 2 ) .
A mapping f : R ¯ N R N that transmits any ω ¯ -curve of space R ¯ N to an ω -curve of the space R N is the ( ω ¯ , ω ) -mapping.
Starting from the definitions of a ω ¯ -curve and ω -curve, i.e.,
d 2 π d t 2 + Γ ¯ α β π d α d t d β d t = ρ ¯ d π d t + ω α β π d α d t d β d t , d 2 π d t 2 + Γ α β π d α d t d β d t = ρ d π d t + ω α β π d α d t d β d t ,
with respect to ρ = ρ β d β d t and ρ ¯ = ρ ¯ β d β d t , one gets
Γ μ ν π Γ ¯ μ ν π = ψ ν δ μ π + ψ μ δ ν π + ω μ ν π ω ¯ μ ν π .
Consider a ( ω ¯ , ω ) -mapping f : R ¯ N R N between the Riemannian spaces R ¯ N and R N . The basic equation of this mapping is (11). After contracting this equation by π and ν , and denoting ω μ = ω μ α α and ω ¯ μ = ω ¯ μ α α , we obtain
Γ μ = Γ ¯ μ + ( N + 1 ) ψ μ + ω μ ω ¯ μ .
From the previous equation, we have
ψ μ = 1 N + 1 Γ μ 1 N + 1 ω μ 1 N + 1 Γ ¯ μ + 1 N + 1 ω ¯ μ .
After substituting the expression (13) into the basic Equation (11), we obtain
Γ μ ν π = Γ ¯ μ ν π + ω μ ν π + 1 N + 1 Γ ν δ μ π + Γ μ δ ν π 1 N + 1 ω ν δ μ π + ω μ δ ν π ω ¯ μ ν π 1 N + 1 Γ ¯ ν δ μ π + Γ ¯ μ δ ν π + 1 N + 1 ω ¯ ν δ μ π + ω ¯ μ δ ν π .
The relation (14) is equivalent to the equality T μ ν π = T ¯ μ ν π , where
T μ ν π = Γ μ ν π ω μ ν π 1 N + 1 Γ ν δ μ π + Γ μ δ ν π + 1 N + 1 ω ν δ μ π + ω μ δ ν π ,
T ¯ μ ν π = Γ ¯ μ ν π ω ¯ μ ν π 1 N + 1 Γ ¯ ν δ μ π + Γ ¯ μ δ ν π + 1 N + 1 ω ¯ ν δ μ π + ω ¯ μ δ ν π ,
are the associated basic invariants of Thomas type for the mapping f.
From the equality
W μ ν σ π : = T μ ν , σ π T μ σ , ν π + T μ ν α T α σ π T μ σ α T α ν π = T ¯ μ ν , σ π T ¯ μ σ , ν π + T ¯ μ ν α T ¯ α σ π T ¯ μ σ α T ¯ α ν π = : W ¯ μ ν σ π ,
we obtain the following basic associated invariants of Weyl type for the mapping f:
W μ ν σ π = R μ ν σ π ω μ ν | σ π + ω μ σ | ν π + ω μ ν α ω α σ π ω μ σ α ω α ν π + 1 N + 1 δ μ π ω ν | σ ω σ | ν 1 ( N + 1 ) 2 δ ν π ( N + 1 ) Γ μ | σ ω μ | σ + ω μ σ α ( Γ α ω α ) + Γ μ ω μ Γ σ ω σ + 1 ( N + 1 ) 2 δ σ π ( N + 1 ) Γ μ | ν ω μ | ν + ω μ ν α ( Γ α ω α ) + Γ μ ω μ Γ ν ω ν ,
W ¯ μ ν σ π = R ¯ μ ν σ π ω ¯ μ ν σ π + ω ¯ μ σ ν π + ω ¯ μ ν α ω ¯ α σ π ω ¯ μ σ α ω ¯ α ν π + 1 N + 1 δ μ π ω ¯ ν σ ω ¯ σ ν 1 ( N + 1 ) 2 δ ν π ( N + 1 ) Γ ¯ μ σ ω ¯ μ σ + ω ¯ μ σ α ( Γ ¯ α ω ¯ α ) + Γ ¯ μ ω ¯ μ Γ ¯ σ ω ¯ σ + 1 ( N + 1 ) 2 δ σ π ( N + 1 ) Γ ¯ μ ν ω ¯ μ ν + ω ¯ μ ν α ( Γ ¯ α ω ¯ α ) + Γ ¯ μ ω ¯ μ Γ ¯ ν ω ¯ ν .
The geometrical objects W μ ν σ π and W ¯ μ ν σ π are tensors of the type ( 1 , 3 ) .
The invariants W μ ν σ π and W ¯ μ ν σ π may be expressed as
W μ ν σ π = R μ ν σ π + A μ ν σ π + δ μ π X ν σ X σ ν + δ ν π X μ σ δ σ π X μ ν ,
W ¯ μ ν σ π = R ¯ μ ν σ π + A ¯ μ ν σ π + δ μ π X ¯ ν σ X ¯ σ ν + δ ν π X ¯ μ σ δ σ π X ¯ μ ν ,
where A μ ν σ π = ω μ ν | σ π + ω μ σ | ν π + ω μ ν α ω α σ π ω μ σ α ω α ν π , A ¯ μ ν σ π = ω ¯ μ ν σ π + ω ¯ μ σ ν π + ω ¯ μ ν α ω ¯ α σ π ω ¯ μ σ α ω ¯ α ν π , and the corresponding X μ ν and X ¯ μ ν .
The difference 0 = W μ ν σ π W ¯ μ ν σ π , for the invariants W μ ν σ π and W ¯ μ ν σ π given by (19) and (20), has the form
0 = R μ ν σ π R ¯ μ ν σ π + A μ ν σ π A ¯ μ ν σ π + δ μ π X ν σ X σ ν X ¯ ν σ + X ¯ σ ν + δ ν π X μ σ X ¯ μ σ δ σ π X μ ν X ¯ μ ν .
After contracting the last equality by π and μ , like by π and ν , and using R α ν σ α = 0 and R ¯ α ν σ α = 0 in Riemannian spaces, we obtain the following results:
0 = A α ν σ α A ¯ α ν σ α + ( N + 1 ) X ν σ X σ ν X ¯ ν σ + X ¯ σ ν ,
0 = R μ σ R ¯ μ σ + A μ α σ α A ¯ μ α σ α + X μ σ X σ μ X ¯ μ σ + X ¯ σ μ + ( N 1 ) X μ σ X ¯ μ σ .
We obtain directly A α ν σ α = ω ν | σ + ω σ | ν , A ¯ α ν σ α = ω ¯ ν σ + ω ¯ σ ν , A μ α σ α = ω μ | σ + ω μ σ | α α + ω μ β α ω σ α β ω μ σ α ω α , and A ¯ μ α σ α = ω ¯ μ σ + ω ¯ μ σ α α + ω ¯ μ β α ω ¯ σ α β ω ¯ μ σ α ω ¯ α . If we substitute these expressions into the equalities (22) and (23), we obtain
X ν σ X σ ν X ¯ ν σ + X ¯ σ ν = 1 N + 1 ω ν | σ ω σ | ν 1 N + 1 ω ¯ ν σ ω ¯ σ ν ,
X μ σ X ¯ μ σ = 1 N 1 R μ σ R ¯ μ σ + 1 N 1 ω μ | σ ω μ σ | α α ω μ β α ω σ α β + ω μ σ α ω α 1 N 1 ω ¯ μ σ ω ¯ μ σ α α ω ¯ μ β α ω ¯ σ α β + ω ¯ μ σ α ω ¯ α 1 N 2 1 ω μ | σ ω σ | μ + 1 N 2 1 ω ¯ μ σ ω ¯ σ μ .
The relations (24) and (25) transform the equality (21) into
0 = R μ ν σ π R ¯ μ ν σ π ω μ ν | σ π + ω μ σ | ν π + ω μ ν α ω α σ π ω μ σ α ω α ν π + ω ¯ μ ν σ π ω ¯ μ σ ν π ω ¯ μ ν α ω ¯ α σ π + ω ¯ μ σ α ω ¯ α ν π + 1 N + 1 δ μ π ω ν | σ ω σ | ν 1 N + 1 δ μ π ω ¯ ν σ ω ¯ σ ν 1 N 1 δ ν π R μ σ δ σ π R μ ν + 1 N 1 δ ν π R ¯ μ σ δ σ π R ¯ μ ν + 1 N 1 δ ν π ω μ | σ ω μ σ | α α ω μ β α ω σ α β + ω μ σ α ω α 1 N 1 δ σ π ω μ | ν ω μ ν | α α ω μ β α ω ν α β + ω μ ν α ω α 1 N 1 δ ν π ω ¯ μ σ ω ¯ μ σ α α ω ¯ μ β α ω ¯ σ α β + ω ¯ μ σ α ω ¯ α + 1 N 1 δ σ π ω ¯ μ ν ω ¯ μ ν α α ω ¯ μ β α ω ¯ ν α β + ω ¯ μ ν α ω ¯ α 1 N 2 1 δ ν π ω μ | σ ω σ | μ + 1 N 2 1 δ σ π ω μ | ν ω ν | μ 1 N 2 1 δ ν π ω ¯ μ σ ω ¯ σ μ + 1 N 2 1 δ σ π ω ¯ μ ν ω ¯ ν μ .
The relation (26) is equivalent to the equality W μ ν σ π = W ¯ μ ν σ π for
W μ ν σ π = R μ ν σ π 1 N 1 δ ν π R μ σ δ σ π R μ ν ω μ ν | σ π + ω μ σ | ν π + ω μ ν α ω α σ π ω μ σ α ω α ν π + 1 N + 1 δ μ π ω ν | σ ω σ | ν + 1 N 2 1 δ ν π ( N + 1 ) ω μ | σ ω μ σ | α α ω μ β α ω σ α β + ω μ σ α ω α ω μ | σ + ω σ | μ 1 N 2 1 δ σ π ( N + 1 ) ω μ | ν ω μ ν | α α ω μ β α ω ν α β + ω μ ν α ω α ω μ | ν + ω ν | μ ,
W ¯ μ ν σ π = R ¯ μ ν σ π 1 N 1 δ ν π R ¯ μ σ δ σ π R ¯ μ ν ω ¯ μ ν σ π + ω ¯ μ σ ν π + ω ¯ μ ν α ω ¯ α σ π ω ¯ μ σ α ω ¯ α ν π + 1 N + 1 δ μ π ω ¯ ν σ ω ¯ σ | ν + 1 N 2 1 δ ν π ( N + 1 ) ω ¯ μ σ ω ¯ μ σ α α ω ¯ μ β α ω ¯ σ α β + ω ¯ μ σ α ω ¯ α ω ¯ μ σ + ω ¯ σ μ 1 N 2 1 δ σ π ( N + 1 ) ω ¯ μ ν ω ¯ μ ν α α ω ¯ μ β α ω ¯ ν α β + ω ¯ μ ν α ω ¯ α ω ¯ μ ν + ω ¯ ν μ .
With the above, we have practically proved the following theorem.
Theorem 1.
Let f : R ¯ N R N be a ( ω ¯ , ω ) -mapping. The geometrical object T μ ν π given by (15is an invariant for this mapping. The geometrical object W μ ν σ π given by (17is an invariant for the mapping f. The geometrical object W μ ν σ π is also invariant for the mapping f.
The invariant W μ ν σ π is an associated derived invariant for the mapping f. The traces of this invariant are W α ν σ α = 0 , W μ α σ α = 0 , W μ ν α α = W μ α ν α = 0 .

Invariants for Mappings of Spaces R 2 and R 3

The surfaces in 3D Euclidean spaces are vector functions of two parameters. For this reason, their internal geometries correspond to the internal geometries of Riemannian spaces R 2 . They are two-dimensional manifolds M 2 = M 2 ( u 0 , u 1 ) . In relativistic physics, it is most common to take u 0 = t and u 1 to be spatial parameters. In classical mechanics, both parameters u 0 and u 1 are spatial, but the time t is an outside parameter.
In both of these cases, the basic invariants T μ ν π and W μ ν σ π , and the derived invariant W μ ν σ π for a ( ω ¯ , ω ) -mapping f : R ¯ 2 R 2 , are
T μ ν π = Γ μ ν π ω μ ν π 1 3 Γ ν δ μ π + Γ μ δ ν π + 1 3 ω ν δ μ π + ω μ δ ν π ,
W μ ν σ π = R μ ν σ π ω μ ν | σ π + ω μ σ | ν π + ω μ ν α ω α σ π ω μ σ α ω α ν π + 1 3 δ μ π ω ν | σ ω σ | ν 1 9 δ ν π 3 Γ μ | σ ω μ | σ + ω μ σ α ( Γ α ω α ) + Γ μ ω μ Γ σ ω σ + 1 9 δ σ π 3 Γ μ | ν ω μ | ν + ω μ ν α ( Γ α ω α ) + Γ μ ω μ Γ ν ω ν ,
W μ ν σ π = R μ ν σ π δ ν π R μ σ + δ σ π R μ ν ω μ ν | σ π + ω μ σ | ν π + ω μ ν α ω α σ π ω μ σ α ω α ν π + 1 3 δ μ π ω ν | σ ω σ | ν + 1 3 δ ν π 3 ω μ | σ ω μ σ | α α ω μ β α ω σ α β + ω μ σ α ω α ω μ | σ + ω σ | μ 1 3 δ σ π 3 ω μ | ν ω μ ν | α α ω μ β α ω ν α β + ω μ ν α ω α ω μ | ν + ω ν | μ ,
and the corresponding T ¯ μ ν π , W ¯ μ ν σ π , and W ¯ μ ν σ π .
In classical mechanics, it is common to use the direct notation of the form F ̲ ̲ σ ̲ , where the number of lines indicates the number of indices in operators F ^ and σ ^ (for details, see [18]). The covariance or contravariance of indices is not relevant. The curvature of the space and affine connection coefficients are not implemented in the corresponding physical laws. The indices in F ̲ ̲ and σ ̲ can have values 1, 2, and 3. The physical laws describe the action that deforms different objects.
Considering the geometrical literature, we conclude that this kind of classical mechanics describes the physics of a three-dimensional Euclidean space that surrounds different objects. The geometrical literature, i.e., the fact that tensors are defined on the borders of the manifolds, says that physical actions are defined on the border of a three-dimensional Euclidean space. The bad side of this approach is tyhe classical mechanics of leafs [19], flowers [20], etc., which when expressed by direct notation does not explain the internal physics of these objects.
The N-dimensional invariant W μ σ = W μ α σ α will bridge this gap between the approaches. Moreover, we will obtain general equations of motion in next section, which, depending on the dimension of a manifold, express the equations of motion of a common form in cosmology and classical mechanics.
Anyhow, we will now present three invariants for a mapping f: R ¯ 2 R 2 in the case of a space R 2 with a constant metric tensor. The relations will stay the same if the Riemannian space is a Euclidean space. It will be useful to obtain the corresponding results for a 3-dimensional Riemannian space with a constant metric.
If the metric of a space R N , N = 2 , 3 , is constant, the corresponding Christoffell symbols, curvature tensors, and Ricci tensors will vanish. The covariant derivative with respect to the affine connection of this space reduces to the corresponding partial derivative. Hence, in this case, the invariants for a mapping f: R ¯ N R N are
T μ ν π = ω μ ν π + 1 N + 1 ω ν δ μ π + ω μ δ ν π ,
W μ ν σ π = ω μ ν , σ π + ω μ σ , ν π + ω μ ν α ω α σ π ω μ σ α ω α π ν + 1 N + 1 δ μ π ω ν , σ ω σ , ν + 1 ( N + 1 ) 2 δ ν π ( N + 1 ) ω μ , σ + ω μ σ α ω α ω μ ω σ 1 ( N + 1 ) 2 δ σ π ( N + 1 ) ω μ , ν + ω μ ν α ω α ω μ ω ν ,
W μ ν σ π = ω μ ν , σ π + ω μ σ , ν π + ω μ ν α ω α σ π ω μ σ α ω α ν π + 1 N + 1 δ μ π ω ν , σ ω σ , ν + 1 N 2 1 δ ν π ( N + 1 ) ω μ , σ ω μ σ , α α ω μ β α ω σ α β + ω μ σ α ω α ω μ , σ + ω σ , μ 1 N 2 1 δ σ π ( N + 1 ) ω μ , ν ω μ ν , α α ω μ β α ω ν α β + ω μ ν α ω α ω μ , ν + ω ν , μ .

3. Equations of the Motion

Consider the N-dimensional Riemannian space R ¯ N (equipped with the affine connection Γ ¯ μ ν π ) and the space R N (equipped with the affine connection Γ μ ν π ). Let the deformation tensor P μ ν π = Γ μ ν π Γ ¯ μ ν π be of the form
P μ ν π = ψ ν δ μ π + ψ μ δ ν π + ω μ ν π ω ¯ μ ν π ,
for tensors ω μ ν π and ω ¯ μ ν π of the type ( 1 , 2 ) and symmetric by μ and ν .
Based on the deformation tensor P μ ν π given by (35) and the previous considerations, we conclude that the transformation R ¯ N R N is a ( ω ¯ , ω ) -mapping.
The invariant for this mapping, which is necessary for this section, is W μ σ = W μ α σ α . This invariant, with respect to the expression (17), has the form
W μ σ = R μ σ ω μ | σ + ω μ σ | α α + ω μ β α ω σ α β ω μ σ α ω α + 1 N + 1 ω μ | σ ω σ | μ N 1 N + 1 Γ μ | σ + N 1 N + 1 ω μ | σ N 1 N + 1 ω μ σ α Γ α + N 1 N + 1 ω μ σ α ω α N 1 ( N + 1 ) 2 Γ μ Γ σ + N 1 ( N + 1 ) 2 Γ μ ω σ + Γ σ ω μ N 1 ( N + 1 ) 2 ω μ ω σ .
We are going to eliminate variation from the following action:
C 0 S 0 + = d N | g | g α β W α β .
After contracting the deformation rule (35) by π and ν , we obtain that ψ μ is given by Equation (13). For this reason, the Christoffell symbols Γ μ ν π and Γ ¯ μ ν π satisfy Equation (14). This relation between the Christoffell symbols may be rewritten as
Γ μ ν π = Γ ¯ μ ν π + d μ ν π d ¯ μ ν π ,
for
d μ ν π = ω μ ν π + 1 N + 1 Γ ν δ μ π + Γ μ δ ν π 1 N + 1 ω ν δ μ π + ω μ δ ν π ,
d ¯ μ ν π = ω ¯ μ ν π + 1 N + 1 Γ ¯ ν δ μ π + Γ ¯ μ δ ν π 1 N + 1 ω ¯ ν δ μ π + ω ¯ μ δ ν π .
In this case, the invariant W μ σ becomes
W μ σ = R μ σ d μ | σ + d μ σ | α α + d μ β α d σ α β d μ σ α d α .
Before we continue the further calculations, we need to recall that Γ μ = | g | , μ | g | , i.e., | g | , μ = | g | Γ μ . For this reason, for any tensor τ α of the type ( 1 , 0 ) , we obtain
| g | τ | α α = | g | τ , α α + | g | Γ β α α τ β = | g | τ , α α + | g | , α τ α = | g | τ α , α .
Based on the assumption that δ g μ ν is equal 0 at the border of integration, and with respect to the result (42) combined with Stokes’ theorem, we conclude
δ d N x | g | τ | α α = 0 .
Hence, the variation in action (37) for g < 0 reduces to
C 0 δ S 0 + = δ d N x g R + δ d N x g g α β d α δ γ d β γ δ d α β γ d γ .
By the quotient rule, there exists a tensor D β γ μ ν α of the type ( 1 , 4 ) such that the variation in d β γ α is expressed as δ d β γ α = D β γ μ ν α δ g μ ν . Moreover, the variation in δ d α is δ d α = D α β μ ν β δ g μ ν . It is well known what the variation in the first summand is in the previous equation. Hence
C 0 δ S 0 + = d N x g δ g μ ν R μ ν 1 2 R g μ ν + d N x g δ g μ ν d μ β α d ν α β d μ ν α d α 1 2 d N x g δ g μ ν g μ ν g α β d α δ γ d β γ δ d α β γ d γ + d N x g δ g μ ν g α β 2 D α δ μ ν γ d β γ δ D α β μ ν γ d γ d α β γ D γ δ μ ν δ .
After involving the Lagrangian density L M in the action (37), i.e., if we analyze the action d N x g g α β W α β + L M , the energy–momentum tensor is
T μ ν = R μ ν 1 2 R g μ ν + d μ β α d ν α β d μ ν α d α 1 2 g μ ν g α β d α δ γ d β γ δ d α β γ d γ + g α β 2 D α δ μ ν γ d β γ δ D α β μ ν γ d γ d α β γ D γ δ μ ν δ .
The equation of motion (46) holds for any dimension N, 2 N + . The summand in the last row in (46) is a generalization of Einstein’s cosmological constant Λ .

4. Generalized Navier–Stokes Equations

In the previous section, we obtained physical laws by vanishing the variation in the generalized action. The action is generated by the basic Equation (14).
In the case of dimension N = 2 , the basic Equation (14) means that any ω ¯ -curve of the space R ¯ 2 is transformed to an ω -curve of the space R 2 .
In our example, let a fluid particle moves along the ω ¯ -curve of the space R ¯ 2 . After the deformation of the surface, the ω ¯ -curve is transformed to an ω -curve of the space R 2 .
The equations of the ω ¯ -curve of space R ¯ 2 and ω -curve of space R 2 are
d 2 ¯ π d t 2 + Γ ¯ α β π d ¯ α d t d ¯ β d t = ρ ¯ d ¯ π d t + ω ¯ α β π d ¯ α d t d ¯ β d t ,
d 2 π d t 2 + Γ α β π d α d t d β d t = ρ d π d t + ω α β π d α d t d β d t .
The corresponding Navier–Stokes equations are
d 2 ¯ π d t 2 + Γ ¯ α β π d ¯ α d t d ¯ β d t = g ¯ π α p ¯ α + f ¯ d i s s π + ω ¯ α β π d ¯ α d t d ¯ β d t , d 2 π d t 2 + Γ α β π d α d t d β d t = g π α p | α + f d i s s π + ω α β π d α d t d β d t .
The left sides of the last two Navier–Stokes equations are covariant accelerations of the particle. The values g ¯ π α p ¯ α and g π α p | α are gradients of pressures p ¯ and p. The tensors f ¯ d i s s π and f d i s s π are components of viscosity (if they exist). The symmetric tensors ω ¯ α β π and ω α β π may be factored as ω ¯ α β π = g ¯ π γ F ¯ γ α σ ¯ β + F ¯ γ β σ ¯ α + θ ¯ α β φ ¯ γ and ω α β π = g π γ F γ α σ β + F γ β σ α + θ α β φ γ , where F ¯ α β and F α β are non-symmetric tensors of the type ( 0 , 2 ) , θ ¯ α β and θ α β are symmetric tensors of the type ( 0 , 2 ) , and σ α , σ ¯ α , φ ¯ α , and φ α are covariant vectors. These vectors may be expressed as σ ¯ α = σ ¯ α , σ α = σ | α , φ ¯ α = φ ¯ α , and φ α = φ | α for fields σ ¯ , σ , φ ¯ , and φ .
Let d 2 ¯ π / d t 2 = a ¯ π , d π / d t 2 = a π , d ¯ π / d t = v ¯ π , and d π / d t = v π . The Navier–Stokes Equation (49) transforms to
a ¯ π + Γ ¯ α β π v ¯ α v ¯ β = g ¯ π α p ¯ α + f ¯ d i s s π + ω ¯ α β π v ¯ α v ¯ β , a π + Γ α β π v α v β = g π α p | α + f d i s s π + ω α β π v α v β .
The vectors a ¯ π , a π , v ¯ π , and v π correspond to the components of the acceleration (first two vectors) and the velocity (last two vectors) in classical mechanics. The geometrical objects Γ ¯ α β π v ¯ α v ¯ β and Γ α β π v α v β are kinematic terms that represent the geodetic acceleration. They describe the change in velocity due to the curvature of the manifold.
Let us recognize the Newtonian mechanics in Equation (50). Newtonian mechanics corresponds to the case of g α β = δ α β , and analogously for g ¯ α β = δ α β . In this case, Γ μ ν π = Γ ¯ μ ν π = 0 . Newtonian mechanics, i.e., its most simple case, describes movement without actions of additional forces. For this reason, it holds that g π α p | α = 0 and f d i s s π = 0 . After this, we obtain a π = ω α β π d α d t d β d t . The right side of the last equality, which is equal to the force F π by mass m, is equivalent to the second Newtonian law m a π = F π . In the same manner, we get m ¯ a ¯ π = F ¯ π from the first equation in (50).
These expressions are valid for any dimension N, but we aimed to present the case of N = 2 .

5. Conclusions

In this research, we discovered a new common point of view between differential geometry and classical mechanics. To connect these two subjects, we needed to review Vesić’s methodology for obtaining invariants of geometric mappings (Section 2).
Using the new invariant W μ σ , we generalized the Einstein–Hilbert action in an N-dimensional Riemannian space. This result makes it possible to obtain equations of motion in any dimensional space (Section 3).
As a special case, we generalized Navier–Stokes equations for a fluid in a 3 D manifold. In our approach, we defined forces on this manifold, not at the border of Euclidean space. The Navier–Stokes equations are generalized with respect to the affine connection of the manifold, not Euclidean space (Section 4).
Regarding some possible new research in this field, as a future perspective, it would be possible to analyze the Navier–Stokes-type equations of motion, taking into account the presence of the torsion field. In other words, this is useful for an Einstein–Cartan geometry or its extensions.

Author Contributions

Conceptualization, I.D.; methodology, I.D.; software, M.S. and A.P.; validation, M.S. and D.J.S.; formal analysis, I.D.; investigation, I.D. and D.J.S.; resources, D.J.S. and M.S.; data curation, D.J.S. and M.S.; writing—original draft preparation, I.D. and A.P.; writing—review and editing, D.J.S. and B.M.R.; visualization, A.P. and M.S.; supervision, B.M.R.; project administration, A.P. and B.M.R.; funding acquisition, A.P. and B.M.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science, Innovations and Technological Development of Serbia, through the grants 451-03-137/2025-03/200251 and 451-03-137/2025-03/200102 and funded by Faculty of Teacher Education, Leposavić, through the grant IMP-003.

Data Availability Statement

Data is contained within the article.

Acknowledgments

The authors wish to thank Nenad Vesić, who motivated us to realize this research and helped regarding variational calculus.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Dodelson, S. Modern Cosmology, 1st ed.; Elsevier: San Diego, CA, USA, 2003. [Google Scholar]
  2. Deriglazov, A.A. Potential motion in a geometric setting: Presenting differential geometry methods in a classical mechanics course. Eur. J. Phys. 2008, 29, 767–780. [Google Scholar] [CrossRef]
  3. Mikes, J.; Stepanova, E.; Vanžurová, A.; Bácsó, S.; Berezovski, V.E.; Chepurna, O.; Formella, S.; Gavrilchenko, M.L.; Haddad, M.; Hinterleitner, I.; et al. Differential Geometry of Special Mappings, 1st ed.; Olomouc University: Olomouc, Czech Republic, 2015. [Google Scholar]
  4. Mikes, J.; Stepanova, E.; Vanžurová, A.; Bácsó, S.; Berezovski, V.E.; Chepurna, O.; Formella, S.; Gavrilchenko, M.L.; Haddad, M.; Hinterleitner, I.; et al. Differential Geometry of Special Mappings, 2nd ed.; Olomouc University: Olomouc, Czech Republic, 2019. [Google Scholar]
  5. Sinyukov, N.S. Geodesic Mappings of Riemannian Spaces, 1st ed.; Nauka: Moscow, Russia, 1979. (In Russian) [Google Scholar]
  6. Weyl, H. Zur infinitesimal geometrie: Einordnung der projectiven und der konformen auffssung. Nachr. Gött. 1921, 1921, 99–112. [Google Scholar]
  7. Thomas, T.Y. On the projective and equi-projective geometries of paths. Proc. Nat. Acad. Sci. USA 1925, 11, 199–203. [Google Scholar] [CrossRef] [PubMed]
  8. Vesić, N.O. Basic invariants of geometric mappings. Miskolc Math. Notes 2020, 21, 473–487. [Google Scholar] [CrossRef]
  9. Milenković, V.M.; Stanković, M.S.; Vesić, N.O. Invariant Geometric Objects of the Equitorsion Canonical Biholomorphically Projective Mappings of Generalized Riemannian Space in the Eisenhart Sense. Mathematics 2025, 13, 1334. [Google Scholar] [CrossRef]
  10. Stefanović, M.; Vesić, N.; Simjanović, D. Randjelović, B. Special Geometric Objects in Generalized Riemannian Spaces. Axioms 2024, 13, 463. [Google Scholar] [CrossRef]
  11. Vesić, N.O.; Simjanović, D.J.; Randjelović, B.M. Invariants for Second Type Almost Geodesic Mappings of Symmetric Affine Connection Space. Mathematics 2024, 12, 2329. [Google Scholar] [CrossRef]
  12. Vesić, N.O.; Milenković, V.M.; Stanković, M.S. Invariants for equitorsion geometric mappings. Filomat 2023, 37, 8537–8542. [Google Scholar] [CrossRef]
  13. Vesić, N.O.; Milenković, V.; Stanković, M. Two Invariants for Geometric Mappings. Axioms 2022, 11, 239. [Google Scholar] [CrossRef]
  14. Vesić, N.; Mihajlović, A. Invariants for F-Planar Mappings of Symmetric Affine Connection Spaces. Facta Univ. Ser. Math. Inform. 2022, 37, 283–293. [Google Scholar] [CrossRef]
  15. Najdanović, M.; Velimirović, L.; Vesić, N. Geodesic Infinitesimal Deformations of Generalized Riemannian Spaces. Mediterr. J. Math. 2022, 19, 145. [Google Scholar] [CrossRef]
  16. Simjanović, D.; Vesić, N. Novel invariants for almost geodesic mappings of the third type. Miskolc Math. Notes 2021, 22, 961–975. [Google Scholar] [CrossRef]
  17. Vesić, N.O.; Zlatanović, M.L. Invariants for geodesic and F-planar mappings of generalized Riemannian spaces. Quaest. Math. 2021, 44, 983–996. [Google Scholar] [CrossRef]
  18. Holzapfel, G.A. Nonlinear Solid Mechanics A Continuum Approach for Engineering, 1st ed.; John Wiley & Sons: Chichester, UK, 2000. [Google Scholar]
  19. Niklas, K.J. Research review A mechanical perspective on foliage leaf form and function. New Pytol. 1999, 143, 19–31. [Google Scholar] [CrossRef]
  20. Wang, Z.; Wang, C.; Wei, Y. Differential growth and shape formation of a flower-shaped structure. Int. J. Non-Linear Mech. 2024, 167, 104918. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Simjanović, D.J.; Djurišić, I.; Penjišević, A.; Stefanović, M.; Randjelović, B.M. Equations of Motion and Navier–Stokes Equations. Computation 2025, 13, 226. https://doi.org/10.3390/computation13090226

AMA Style

Simjanović DJ, Djurišić I, Penjišević A, Stefanović M, Randjelović BM. Equations of Motion and Navier–Stokes Equations. Computation. 2025; 13(9):226. https://doi.org/10.3390/computation13090226

Chicago/Turabian Style

Simjanović, Dušan J., Ivana Djurišić, Aleksandra Penjišević, Marko Stefanović, and Branislav M. Randjelović. 2025. "Equations of Motion and Navier–Stokes Equations" Computation 13, no. 9: 226. https://doi.org/10.3390/computation13090226

APA Style

Simjanović, D. J., Djurišić, I., Penjišević, A., Stefanović, M., & Randjelović, B. M. (2025). Equations of Motion and Navier–Stokes Equations. Computation, 13(9), 226. https://doi.org/10.3390/computation13090226

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop