Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

Journal Pre-proof

A revisit of Navier–Stokes equation

Wanan Sheng

PII: S0997-7546(19)30203-1
DOI: https://doi.org/10.1016/j.euromechflu.2019.12.005
Reference: EJMFLU 103578

To appear in: European Journal of Mechanics / B Fluids

Received date : 15 April 2019


Revised date : 4 December 2019
Accepted date : 5 December 2019

Please cite this article as: W. Sheng, A revisit of Navier–Stokes equation, European Journal of
Mechanics / B Fluids (2019), doi: https://doi.org/10.1016/j.euromechflu.2019.12.005.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the
addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive
version of record. This version will undergo additional copyediting, typesetting and review before it
is published in its final form, but we are providing this version to give early visibility of the article.
Please note that, during the production process, errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier Masson SAS.


Journal Pre-proof

2 A revisit of Navier-Stokes equation


3
4 Dr. Wanan Sheng
5 SW MARE Marine Technology, Ireland
6 Email: wanan_sheng@outlook.com

of
7
8

Abstract

pro
9
10 An effort has been recently paid to derive and to better understand the Navier-Stokes (N-S) equation,
11 and it is found that, although the N-S equation has been proven to be correct by numerous examples,
12 some concepts and principles behind the equation may not be correct or consistent. For instance, from
13 an analysis of the simple classic Couette flow, the requirement of the symmetric stress tensor is in fact
14 conflicting with the solution of the Couette flow.
re-
15 To solve the inconsistencies identified in this research, a reformulation of the total tensor is suggested
16 for accommodating the fluid friction which bears a solid physics, and the new total tensor could resolve
17 all the inconsistencies and conflicts identified. The newly defined fluid friction tensor is then used to
18 derive N-S equation, and as expected, the same N-S equation as the original form of N-S equation for
lP

19 incompressible flows is obtained. For compressible flows, to achieve the same N-S equation as the
20 original N-S equation, a slightly different assumption but yet in a very similar manner as Stokes made
21 in 1845 is needed.
22 It is the author’s intention that the N-S equation under the new defined total tensor has different, but yet
23 more physical background concepts and principles. It is hoped that the revisit of the N-S equation could
a

24 shed some light to better understand the dynamic flows and lead to establish new and better approaches
25 to solve the complicated flow problems in future.
urn

26 1 Introduction
27 Fluid dynamics is an ancient topic and people have been trying to solve the fluid mechanics/dynamics
28 problems since the great Greek philosophers and scientists Aristotle (384-322 BC) and Archimedes
29 (287 – 212 BC) [1]. The first partial differential equation for fluid dynamics was much later formulated
Jo

30 by Euler in 1752 when he considered only for inviscid fluids (that is why the fluid dynamic equation
31 for inviscid fluids is called Euler equation). After that, by adopting the Newton’s definition of friction
32 due to the velocity gradient and fluid viscosity, Navier and Stokes could independently include the
33 viscous forces into the equation, thus now it is called Navier-Stokes equation. According to Anderson
34 [2], an interesting fact is that Navier did get the correct equation in 1822, but his derivation has been
35 greatly flawed. Stokes re-derived the fluid dynamic equation in a more delicate manner in 1845 [3].
36 Darrigol [4] has indicated that in the period between Navier (in 1822) to Stokes (in 1845), other three
37 main scholars: Cauchy, Poisson and Saint-Venant, had made their contributions to establish the fluid

1
Journal Pre-proof

38 dynamic equation under different assumptions. Without particular reasons, the final settlement of the
39 fluid dynamics equation might be the successful applications of Stokes’ pendulum theory in the Hagen’s
40 Bessalian pipe flow and the Poiseuille’s capillary-tube flow (see [4]). Now it is generally accepted that
41 the establishment of the fluid dynamics equation was finished with the work of Stokes in 1845, and the
42 fluid dynamic equation was later named as the Navier-Stokes equation, even though Navier and Stokes
43 published their equations independently in a gap of more than 20 years.
44 The Navier-Stokes (N-S) equation is the fundamental equation for governing fluid motion and dynamics,
45 and so far numerous examples have proven the correctness of the N-S equation for fluid dynamics.
46 However, it has been well recognised that seeking an analytical solution to the N-S equation has been

of
47 proven too difficult and analytical solutions can only be obtained for some simple laminar flows,
48 therefore, turbulence is frequently referred as the major unsolved problem of classical physics [5]. In
49 2000 the Navier-Stokes equation was selected to be one of seven Millennium Problems by the Clay

pro
50 Mathematics Institute of Cambridge, U.S. (http://www.claymath.org/millennium-problems), and a
51 special award of $1 million is provided for the answer to each of the 7 millennium questions. In 2008
52 the U.S. Defense Advanced Research Projects Agency (DARPA) listed it as one of 23 DARPA
53 Mathematical Challenges- “Mathematical Challenge Four: 21st Century Fluids”. The challenge
54 statement is as following: although classic fluid dynamics and the Navier-Stokes Equation were
55 extraordinarily successful in many practical problems, including the understanding of shock waves and
re-
56 turbulent flows, new methods (and understandings) are still needed to tackle the complex fluids such as
57 foams, suspensions, gels, and liquid crystals (https://www.britannica.com/science/Navier-Stokes-
58 equation).
59 Due to the difficulties in obtaining analytical solutions to N-S equation, especially for those real
lP

60 complicated turbulent flows, modern numerical methods, especially the Computation Fluid Dynamics
61 (CFD), have been developed and advanced with the increased computer power and the advanced
62 numerical algorithms, and nowadays the computational fluid dynamics (CFD) have been widely used
63 to numerically solve the N-S equation, including some very complicated fluid dynamics problems, such
64 as airplanes, air engines, ships, fire modelling, heat transfer, chemical reaction and so on. To date, CFD
a

65 have been very successful for studying various fluid dynamic problems, and great achievements have
66 been attained. Good examples are the CFD applications in development of Airbus 380 in Figure 1 [6]
urn

67 and of Boeing 787 in Figure 2 [7], from which we can already see how much computer modelling work
68 can be already carried out using CFD. With the increase of computer powers, the advancement of
69 numerical algorithms and the understanding of the complicated flows (for instance, the better turbulence
70 models), CFD will get more and more applications and become more and more capable in solving the
71 problems in our daily and in the complicated fluid dynamics.
72 Though CFD have been very successful in solving N-S equation, there are still many challenges in using
Jo

73 CFD for complicated flows. This includes the examples mentioned above, where CFD is still limited to
74 some specific problems and the explorations are still on going to use more CFD in the design of the
75 whole aircrafts and the off-design situations. In addition, the needs to use CFD to solve the multi-
76 physical problems [6-10] and seek global optima [7] are the current challenges.
77 The current approaches in CFD include the most computational demanding approach: direct numerical
78 simulation (DNS, [11, 12]); relatively less (compared to DNS) but still very computational demanding
79 method: large-eddy simulation (LES, [13-17]) and the most practical method: the Reynolds-average N-
80 S (RANS, [5, 18-30]). The RANS method resolves the mean flow numerically, and models the turbulent

2
Journal Pre-proof

81 flow via turbulence models, and it has the advantages of the moderate requirements in grid and temporal
82 stepping (thus the generally accepted computational burden), and of a good numerical convergence.
83 The method has been proven to be very useful in many practical flows, including the examples above.
84 However, it is also generally accepted the RANS method may not be very reliable in predicting the
85 complicated flows since different turbulence models may lead to different flow simulations. In the past
86 decades, many researchers have tried to tune/modify/extend the turbulence models, but mostly for their
87 specific problems, other than for more general turbulence models. As a result of such difficulties, so far
88 there is hardly any universal turbulent model for different flows, and we may even have difficulties in
89 many general flow problems, such as flow transition from laminar to turbulent; the flows with adverse

of
90 pressure gradients; and flow separations and re-attachments. The author thinks that it may be a good
91 idea to go back to the very fundamentals in fluid dynamics, and the enhancement and better
92 understanding of some concepts and principles behind the N-S equation may be needed, and believes

pro
93 that a better physical understanding and significance to the fundamentals of fluid dynamic equation
94 could pave a path to construct better turbulence models for and thus the better solutions to the
95 complicated flows. This is the main intention of the research.
re-
a lP

96
97 Figure 1 CFD applications in the design of Airbus A380 [6]
urn
Jo

98
99 Figure 2 Impact of CFD at Boeing. Green areas have strong CFD penetration; purple areas have some
100 penetration; red areas present future opportunities [7]

3
Journal Pre-proof

101 This research examines the very fundamentals of the N-S equation, for instance, the symmetric viscous
102 stress tensor in fluids introduced by Stoke in 1845 (named as the Stokes’ symmetric stress tensor which
103 originated from the Cauchy’s symmetric stress tensor for elastic materials). This Stokes’ symmetric
104 stress tensor was so well accepted in the community that it has been taken as a law that all fluids must
105 follow, in a similar manner as the Newton’s laws of motion as the universal law of motion. However,
106 as stated in [31], the symmetry in the stress tensor is violated in an electric field on polarized fluid
107 molecules, in which antisymmetric stresses must be included in such circumstances. A similar
108 contradiction has been reported in [32]. The author is wondered why the Stokes’ symmetric stress tensor
109 is violated in some special flows?

of
110 To find an answer to the question above, the author has recently made an effort to derive and better
111 understand the N-S equation. It is found that, although the N-S equation is proven to be correct for
112 governing all various flows, including the compressible flows with the Stokes’ assumption [31] (page

pro
113 128), some concepts and principles behind the equation may not be correct or consistent. For instance,
114 an analysis to the classic Couette flow has shown that the requirement of the symmetric stress tensor is
115 conflicting with the solution of the Couette flow when a real physics is applied in the analyses (more
116 details on these conflicts/inconsistencies can be found in Section 4).
117 To understand the problems mentioned above, the forces acting on the fluids are re-examined which
118 could show the inconsistencies for the Stokes’ symmetric stress tensor and a reformulation has been
re-
119 made to the total tensor for accommodating the total surface force of more physical significance. Using
120 the newly defined friction tensor, an N-S equation exactly same as the original form of N-S equation is
121 obtained for incompressible flows. It is also possible to obtain a same N-S equation for compressible
122 flows using a slightly different assumption as Stokes made. Therefore, under the newly defined total
lP

123 tensor, the N-S equation is maintained same, however, the concepts and principles behind the equation
124 are different but of more physical significance.
125 To present the research work, following arrangement is made for the rest of the paper. Section 2 presents
126 the simple introduction and derivation of the original form of N-S equation, with some discussions for
127 the equation; in Section 3 viscous stress tensor and surface forces are illustrated, providing better
a

128 understandings and physical significance to the viscous stress tensor and its components; Section 4 lays
129 out 3 inconsistencies behind the N-S equation, all based on the solid physical understanding to the force
urn

130 analyses for the fluids; in Section 5, more details of fluid motions and forces are presented and the
131 asymmetric friction tensor is proposed; in Section 6 Navier-Stokes equation is derived using the new
132 total tensor for both incompressible and compressible flows; and in Section 7, some more discussions
133 are made for the newly defined asymmetric stress tensor. In the last section, conclusions are provided
134 for the research work.
Jo

135 2 The original form of N-S equation


136 In this section, the original N-S equation is first derived briefly, to provide a base for further analyses
137 to what problems may have behind the N-S equation.

138 2.1 Conservation of momentum


139 The N-S equation can be easily derived from the transport theorem and the conservation of momentum
140 (note: the fluid dynamic equation can also be derived in different ways, see [33, 34]).

4
Journal Pre-proof

141 Based on the Newton second law of motion, the momentum of the fluid dynamics can be expressed as,
D 
Dt   
VdV  f dV  T  ndS (1)
V V S

142 In the equation, V and V represent different variables: the former is the fluid velocity (vector) and the

143 latter the fluid volume (scalar); T is the total tensor (the double-sided arrow means a tensor) for surface
144 stress (a force per unit surface area); f the body force (force per unit volume). For many applications,
145 the simplest body force is the gravitational force of the fluid, that is, f = ρgk, with g being the
146 gravitational acceleration, k the unit vector along z-axis.

of
147 By invoking the Gauss divergence theorem (see Appendix A2) on the last term in the right-hand-side
148 of Eq. (1), the momentum conservation can be expressed as,

D  f    T dV

pro
Dt  
VdV  (2)
V V
 
149 invoking the transport theorem, the conservation of momentum (see Appendix A3) leads to,
  V  
    VV   f    T (3)
t
150 Using the Einstein summation convention, the conservation of momentum can be simply expressed as
re-
 ui   ui u j  Tij
  fi  (4)
t x j x j
151 where fi is the body force component and hereafter the subscripts i, j =1, 2, 3 mean the components
152 along x-, y- and z-axes respectively.
lP

153 2.2 Total and stress tensors in fluids


154 Following the general convention, the total stress tensor has a general form as,

 T11 T12 T13 


T  T21 T22 T23 
a

(5)

T31 T32 T33 


urn

155 As such, the surface force (per unit surface area) acting on a surface is calculated as

F1   T11 T12 T13  n1  T11 n1  T12 n2  T13 n3 



 
F  F2   T  n  T21 T22 T23  n2   T21 n1  T22 n2  T23 n3  (6)

F3  T31 T32 T33  n3  T31 n1  T32 n3  T33 n3 


Jo

156 where F1, F2,and F3 are three components along x-, y- and z-axes; n is the normal vector of the surface,
157 and n1, n2 and n3 are the components of the normal vector along x1-, x2- and x3-axes, respectively (see
158 Figure 3).

5
Journal Pre-proof

S
n

ΔS
x3

x2

of
x1
o
159
160 Figure 3 A stress force (vector) on a small surface, ΔS

pro
161 In details, the definition of the total tensor was proposed by Stokes [3] as

  p 0 0   11  12  13    V 0 0 
T   0  p 0    21  22  23    0   V 0  (7)

 0 0  p  31  32  33   0 0   V 
re-
162 where λ is so called the secondary viscosity coefficient (see [33]), which is different from the physical
163 viscosity coefficient μ of the fluid.
164 Based on the conservation of angular momentum (from the Cauchy’s second law of motion), the viscous
165 stress tensor must be symmetric, and the currently accepted stress components are given as
lP

 u1 u 2 u 3
 11  2  x ;  22  2 x ;  33  2  x
 1 2 3

  u u 
 12   21    1  2 

  x 2 x1 
 (8)
      u1  u 3 
a

 13 31  x x1 
  3

  u 2 u 3 
urn

 23   32     

 
 3x x 2 

166 The first term in the right-hand-side of Eq. (7) is pressure tensor, and the diagonal tensor means the
167 pressure could produce only normal forces on the surfaces (the negative sign means the pressure force
168 pointing into the body); the second term is the stress tensor due to the fluid viscosity, which is very
169 similar to pressure and has the same unit as the pressure. The difference between the pressure force and
Jo

170 the stresses is that the pressure is a scalar at one point in the fluid, but the stress is a tensor, of which
171 the components have orientations when act on a surface. For example, the stress component,  21 , denotes
172 a stress acting in the y-direction on a surface of constant x; and  12 , denotes a stress acting in the x-
173 direction on the surface of constant y. Similarly,  11 denotes a normal stress acting on a surface of
174 constant x (see [35, 36]).
175 The third term is a term of the fluid volume dilatation contributing to the total stress tensor, and Stokes
176 [3] added for accommodating the fluid compressibility. But it is still a controversial term [37, 38].

6
Journal Pre-proof

177 Obviously, this term disappears for incompressible flows. It is reported that it is be very important in
178 the analysis of the shock wave structure where the pressure and temperature could change dramatically
179 in short time and short distances [33]. Stokes [3] also proposed a relation between the physical and the
180 secondary viscosity coefficients as
2
  (9)
3
181 It should be noted that this relation is still controversial based on the textbook ([39], page 114).

of
182 2.3 Navier-Stokes equation
183 Based on the principles and assumptions above, the full Navier-Stokes equation can be derived as

  u1   u12   u1 u 2   u1 u 3  p 1   u1 u 2 u 3 


    f1    2 u1      

pro

 t x1 x 2 x 3 x1 3 x1  x1 x 2 x 3 
  u 2   u1 u 2   u 2   u 2 u 3 
 2
p 1   u1 u 2 u 3  (10)
     f2    2 u 2      
 t x1 x 2  x 3  x 2 3 x 2  x1 x 2 x 3 
  u 3    u1 u 3    u 2 u 3    u 3   f  p   2 u  1    u1  u 2  u 3 
 2

 t x1 x 2 x 3 x 3 3 x 3  x1 x 2 x 3 
3 3
re-
184 This is the widely accepted N-S equation for both compressible and incompressible flows.
185 In a vector form, the Navier-Stokes equation is expressed as
 V 
   VV   f B  p   2 V    V 
1 (11)
t 3
lP

186 or in the Einstein summation convention,

 ui   ui u j  p 1   u j



  fi    2 ui   (12)
t x j xi 3 xi  x 
 j 
a

187 3 Viscous stress tensor and surface forces


urn

188 Here the viscous stress tensor means the part of the stress tensor due to the fluid viscosity, written as

  12  13 
  11
τ   21  22  23  (13)

 31  32  33 
Jo

189 From the textbooks and research papers, it is frequently seen that the stress tensor components are
190 shown in Figure 4 (note: on surface S2 the normal points opposite to x2-direction). However, to better
191 understand the figure, some delicate explanations must be given to the illustration. Taking the small
192 cube (enlarged for illustration) and considering the special orientations of the chosen plates, S1, S2 and
193 S3, using Eq. (6), we can easily show that there are only three simple surface force components on each
194 of those specific planes, see Eq. (14) and Figure 5(a). Here the normal vectors for these specific planes
195 are still indicated by n1, n2 and n3 for a clarification.

7
Journal Pre-proof

τ33
τ23
τ13 τ
s3 31
τ21
τ11
x3 τ12
τ22 s1
s2 τ32
x2

of
o x1
196
197 Figure 4 Illustration of the stress tensor

pro
  11  12  13  n1   11n1 
 
Fs1   21  22  23   0    21n1 
  31  32  33   0   31n1 

  11  12  13   0   12 n2  (14)
 
Fs 2   21  22  23  n2    22 n2 
re-
  31  32  33   0   32 n2 

  11  12  13   0   13n3 
 
Fs 3   21  22  23   0    23n3 

lP

  31  32  33  n3   33n3 


198 For the specific faces S1, S2 and S3, their normal vectors are actually n1 =1; n2 = -1; n3 =1, therefore,

  11  12  13  n1   11n1   11 
 
Fs1   21  22  23   0    21n1    21 
a

  31  32  33   0   31n1   31 



  11  12  13   0   12 n2     12 
urn

(15)
 
Fs 2   21  22  23  n2    22 n2     22 
  31  32  33   0   32 n2    32 

  11  12  13   0   13n3   13 
 
Fs 3   21  22  23   0    23n3    23 
  31  32  33  n3   33n3   33 
Jo


199 The transformed result is shown in Figure 5(b), which is actually the same as those in Figure 4. This
200 explains why the surface forces can be specified using the stress tensor components (on S2, the negative
201 signs mean the opposite directions of the force components).

8
Journal Pre-proof

(a) (b)
τ33n3 τ33
n1=1 on S1
τ23n3 τ23
n2= -1 on S2 τ13
τ13n3
s3 τ31n1 s3 τ31
n3=1 on S3
τ32n2 τ21n1 τ21
τ22n2
τ11
x3 τ11n1 x3 τ12
τ12n2
τ22 s1
s1 τ32
s2 s2

of
x2 x2
x1 x1
o o
202

pro
203 Figure 5 Surface forces and stress tensor components

204 4 Inconsistencies on shear stress for N-S equation


205 4.1 Fluid definition and shear stress in fluids
re-
206 Let us go back to the very fundamental concept for fluid: what is fluid? Here a stricter technical or
207 physical definition of fluid is referred: based on Britannica Encyclopaedia
208 (https://www.britannica.com/science/fluid-physics), Fluid, including any liquid or gas or generally any
209 other material, can not sustain a tangential/shear force when at rest and could undergo a continuous
210 change in shape under such a stress. In other word, under a shear stress (regardless how small it could
lP

211 be), a continuous and irrecoverable change of position of the material forms a flow, which is a very
212 basic property of fluids. In contrast, the shear stresses can be maintained within a deformed elastic solid,
213 and the deformed solid could spring back to its original shape when the stresses are removed.
214 While in Bertin and Smith [33], it states that the technical distinction between a fluid and a solid lies
215 with their reaction to an applied shear/tangential stress acting to them: a solid can resist a shear stress
a

216 by a static deflection; a fluid can not. Any shear stress applied to a fluid, no matter how small, will
217 result in a flow of that fluid, a continuous deformation of the shape. A very similar statement can be
urn

218 found in White [40].


219 In both definitions, a significant distinction between a fluid and a solid is that a fluid can not resist a
220 shear stress without a permanent deformation (i.e., flowing) regardless how small of the shear stress
221 could be, while a solid could spring back to original form after the removal of the shear stress acting on
222 the solid (if the solid deformation is within the elastic deformation). In a word, fluids can only have
Jo

223 shear stress components when it flows. This is a very important statement which will be used for
224 discussing the inconsistencies on the shear stress of fluids.

225 4.2 Inconsistency 1: symmetric shear stress in fluids


226 The Cauchy’s second law of motion is based on the assumption of the conservation of angular
227 momentum. This law was proposed by Cauchy for elastic materials that the stress tensor must be
228 symmetric in the equilibrium state due to the force balance.

9
Journal Pre-proof

229 The Cauchy’s concept of the symmetric stress tensor was adopted for fluid dynamics by Stokes in 1845
230 [3]. Analogous to the Newton’s formula for fluid friction, Stokes formulated the symmetric stress tensor,
231 i.e., the viscous stress tensor in fluid dynamics, and the tensor components given in Eq. (8). To date,
232 this is well accepted definition, and hardly there are any doubts for this definition since 1845, even
233 though some research results have shown that the viscous stress tensor may be asymmetric for some
234 special flows, such as compressible flows as seen in the cases of rarefied gases, shock waves and
235 gaseous flows through micro fluidic channels [32, 37, 41, 42], but these research results have been only
236 taken as the exceptional cases in the fluid dynamics community. For instance, Wilcox [24] has
237 especially stated that the symmetric stress tensor is for simple viscous fluids, but not for some

of
238 anisotropic liquids (see page 39). However, no reason is given for the statement.
239 Another example for the inconsistent symmetrical stress tensor is given in the textbook of Kundu etc
240 ([31], page 126), with a statement as: the stress tensor symmetry is violated in the electric field on

pro
241 polarized fluid molecules, where antisymmetric stresses must be included in the analysis.
242 A question would be if the Cauchy’s law of angular moment conservation is a universal law for fluids,
243 which requires the fluid viscous shear stress must be symmetric. Then why there are so many exceptions?
244 After all, a universal law must be satisfied for all cases as we do not see any exceptions for Newton’s
245 laws of motion. Obviously, there is an inconsistency in some practical examples when the fluid viscous
246 stress tensor is required to be symmetric. Hence the inconsistency is termed as Inconsistency 1 here.
re-
247 4.3 Inconsistency 2
248 According to the friction definition of a Newtonian fluid, if a fluid velocity gradient is between adjacent
249 fluid particles, a fluid friction occurs due to the fluid viscosity, as
lP

u
 12 ( xy )   (16)
y
250 This is actually a formula for defining the fluid viscosity coefficient, and an illustration can be seen in
251 Figure 6(a).
a

(a) (b)
u
urn

Δu
 11
Δy  12 u1 u2
y y
Δx
o o x
Jo

252 x
253 Figure 6 Shear stress due to the velocity gradient
254 From this example in Figure 6(a), it can be seen obviously that the shear stress or friction given by
255 Eq.(16), the shear stress tensor component, τ12, has a direction along with the velocity increment, Δu,
u
256 that is,  12   has a direction on x-axis. Similarly, in Figure 6(b), the normal stress, τ11, has a
y
u
257 direction with the velocity increment, Δu, hence the normal stress  11   has a direction on x-axis.
x

10
Journal Pre-proof

258 Hence, τ11 and τ12 both have a direction in x-axis, but acting on different surfaces of constant x and
259 constant y, respectively, see Figure 5.
260 It will be seen later in Eq. (27) that the fluid friction direction has always the same direction with the
261 velocity increment, regardless of whether the velocity increment is on x-, y- or z-directions.
262 Next the symmetric shear stress is examined. Taking τ12 as an example, it is a stress acting in the x-
263 direction (subscript ‘1’) upon a surface of constant y (subscript ‘2’) (see [33, 35] and Figure 5(b)). Based
264 on the Stokes’ symmetric stress tensor, the tensor component is defined as,
 u v  u v
 12          (17)

of
 y x  y x
265 This tensor component consists of two terms in the right-hand-side in Eq.(17). In physics as shown
266 above, the first shear stress term has a direction along x-axis due to the increment Δu while the second

pro
267 shear stress term has a direction along y-axis due to Δv. These two orthogonal stress terms (along x- and
268 y-axes respectively) together make a single stress tensor component τ12. In addition, following the
269 definition of stress tensor definition in [35], the tensor component, τ12, is acting in x-axis on a plane of
270 constant y, while its symmetric pair τ21 (of a same definition) is acting along the y-direction on a plane
271 of constant x. From the physical standpoint, the symmetric shear stress component is ambiguously
272 defined, which has no clear physical significance.
re-
273 Therefore, a question will be: how a shear stress acting in one direction could physically hold two terms
274 of different orientations of frictions? Here this conflict is termed as Inconsistency 2.

275 4.4 Inconsistency 3


lP

276 Here we examine the classic Couette flow. The Couette flow is confined between two large parallel
277 plates: the lower plate is fixed and the upper plate moves at a constant speed (u0). The speed of the
278 upper plate is taken to be relatively small, so that the flow between the plates can be regarded as a
279 laminar flow, see Figure 7(a).

u=u0 u0
a
urn

(a) (b) h
y

u=0
o
280 x
281 Figure 7 Couette flow: Fluid flow between two plates
Jo

282 The dynamic problem can be simplified mathematically as a 2D flow as shown in Figure 7(b). From
283 the plot, we have: u = u(y), and v = w = 0. As such, the N-S equation (zero pressure gradient here) is
284 degraded to,
 2u
0 (18)
y 2
285 Applying the boundary conditions: u = 0 at y =0; u = u0 at y = h, the solution of the Eq. (18) is

11
Journal Pre-proof

y
u u0 (19)
h
286 This is the analytical solution of the Couette flow, and the flow velocity distribution is shown in Figure
287 7(b).
288 Based on the definition of the symmetric stress tensor components in Eq. (8), we have
u v u
 12     0 (20)
y x h
289 Following the definition of the orientation of the stress tensor component (see Figure 5), it is a stress in

of
290 x-axis, which is caused due to the different velocities of u in y-direction.
291 From the symmetrical stress tensor, it requires

 21   12   
u0

pro
(21)
h
292 It must be noted that this is a stress in y-direction.
293 As shown in Figure 8, for a small rectangular fluid element, on its four sides, there are 4 stresses: τ12 &
294  making two pairs of stresses due to the symmetric stress tensor. For the Couette flow, the
τ21,  12 &  21
295 horizontal viscous stresses τ12 and  12 on the upper and lower sides can be easily understood, and they
re-
296 have an obvious physical foundation, since the stress in x-direction can cause a flow of the fluid (a
297 constant deformation) in the same direction. However, from the requirement of the symmetric stress
298  must exist as shown in Figure 8. The question here is
tensor the vertical viscous stresses τ21 and  21
299  , why
what causes the vertical stresses in the Couette flow? Or if there exist vertical stresses τ21 and  21
lP

300 they do not cause a flow along y-direction?


301 As it is already shown in the fluid definition, a shear stress in a fluid can only exist when it flows. But
302 for the Couette flow, there is only a flow in x-direction, but no flow in y-direction, then how the vertical
303 shear stresses are caused?
304 The similar conflict can be seen from the solution of the classic flow in a horizontal pipe (see examples
a

305 in [34]).
306 The shear stress acting on a fluid causes a flow of the fluid, and as a result of the flow, the flow induced
urn

307 fluid friction could balance the shear stress. This is very different from the solids in which a force
308 balance must be required on the conservation of the angular moment, that is, the symmetric stress tensor
309 in solids.
310 Therefore, this is a conflict, termed as Inconsistency 3 in this research.
Jo

12
Journal Pre-proof

'

'

of
311
312 Figure 8 Stresses in the simple Couette flow based on Eqs. (20) and (21)

pro
313 5 New understanding of fluid motion and force
314 5.1 Relative velocity between two points in fluid
315 At a certain time t, the fluid point A is taken as a reference point, given by a position vector rA=(x0, y0,
re-
316 z0)T, and the corresponding velocity components as (u0, v0, w0)T. At the same time, one of its
317 neighbouring point B is positioned by a vector, rB =(x0+Δx, y0+Δy, z0+Δz)T. So the vector from A to B
318 is calculated as dr =(Δx, Δy, Δz)T.

B
lP

z dr
rB
y A
rA

o x
a

319
320 Figure 9 The fluid particle A and its neighbouring point B
urn

321 The corresponding velocity components at B can be simply put as

u1  u 0  u

v1  v 0  v (22)
 w  w  w
 1 0

322 The fluid friction (from the friction definition of a Newtonian fluid) is caused due to the velocity
Jo

323 difference between the two neighbouring points, A and B, i.e., the relative velocity components (B to A)
324 can be calculated as

13
Journal Pre-proof

 u u u   u u u 
 x x  y y  z z   x y z  x 
 u    
 v    v x  v y  v z    v v v   
y (23)
   x y z   x y z   
w  w w w   w w w   z 
 x  y  z   
 x y z   x y z 
325 From Eq. (23), it can be seen that the directions of fluid friction is always in the same direction with the
326 increment of velocity component, that is, Δu in x-direction, Δv in y-direction and Δw in z-direction.

of
327 5.2 Fluid friction tensor
328 The relative velocity components between the joint fluid points are the requisite condition for inducing

pro
329 viscous forces (frictions) in fluid. Based on this physical understanding and the definition of fluid
330 friction (note: friction is used here to avoid using the word ‘stress’), it is reasonable to define a friction

331 tensor σ as

 u u u 
   
x y z 
  v v v 
σ    a
re-
 (24)
 x y z 
 w w w 
   
 x y z 
332 with the friction tensor components being given as
lP

 u u u
 11   x ;  12   y ;  13   z

 v v v (25)
 21   ;  22   ;  23  
 x y z
 w w w
a

 31   ;  32   ;  33  
 x y z
urn

333 Obviously, the friction tensor is an asymmetric tensor, which is different from the symmetric stress
334 tensor, shown in Eq. (8).
335 And the fluid velocity gradient tensor (an asymmetric tensor) defined as
 u u u 
 
x y z  a a12 a13 
  v v v  
11

a a 23 
Jo

 a 21 a 22 (26)
 x y z 
 w w w  a31 a32 a33 
 
 x y z 
336 The corresponding friction force on a unit surface is given as

14
Journal Pre-proof

 u u u   u u u 
      n1   n 2   n 3 
Fx 
  x y z  n   x
1
y z 
v v v v v v
F  Fy   σ  n      n2     n1   n2   n3 

 (27)
 x y z   x y z 
 Fz   w w  n
w  3    w w w 
     n1   n2   n3 
 x y z   x y z 
337 Now we come back to the Couette flow, the friction tensor component  12 give the friction along x-axis
338 as

of
u u
 12    0 (28)
y h
339 but the corresponding vertical shear stress component is

pro
v
 21   0 (29)
x
340 Obviously, there is no friction on y-direction. Therefore in this way the physical conflict of the stresses
341 in the Couette flow is resolved.
re-
342 5.3 Analysis of velocity gradient tensor
343 The velocity gradient tensor can be rewritten as

S S12 S13   0  3 2 
  11
a  S21 S22 S23    3 0  1  (30)
lP

S31 S32 S33   2 1 0 


344 with the symmetric part as following,

 u v w
S11  x ; S 22  y ; S 33  z ;

a

 1  u v 
S12  S 21    
 2  y x  (31)

urn

S  S  1  u  w 
 13 31
2  z x 

 1  v w 
 S  S    
 
32 23
 2  z y 
345 and the anti-symmetric part with the tensor components defined as
Jo

1  w v  1  u w  1  v u 
1    ;  2    ; 3     (32)
2  y z  2  z x  2  x y 
346 here ω1, ω2, ω3 are the components of the angular velocity vector or vorticity vector, ω, following the
347 definition given in [34],
1
ω V (33)
2

15
Journal Pre-proof

348 From Eq. (30), it can be seen that the velocity gradient tensor can be written into two parts: the
349 symmetric part and the anti-symmetric part. The symmetric strain-rate has been adopted in Stokes stress
350 tensor for deriving the N-S equation.
351 The basic difference between the symmetric stress tensor and the asymmetric friction tensor is the last
352 pure rotation term in Eq. (30). If we consider a special case of an irrotational flow, we have
 V  0 (34)

353 Under such a flow condition, the last term in Eq. (30) disappears. Therefore, these two tensors: the
354 symmetric stress tensor and the asymmetric friction tensor; are exactly same.

of
355 For the Couette flow, we can show that it is rotational, since
1 u0 (35)
3   0
2 h

pro
356 meaning,
 V  0 (36)

357 From the practical point of view, the rotational flow condition holds in most flows within the viscous
358 boundary layers, in which the perfect pairs of the viscous stress tensor components do not exist.
re-
359 6 New concepts for Navier-Stokes equation

360 6.1 Surface force for fluid dynamics


361 For fluid dynamics, the forces acting on the fluid include both the body and surface forces, with the

lP

362 surface force being calculated based on the total tensor,  , as shown in Eq. (1). However, based on
T
363 the discussions above, a change in the total tensor is proposed to make it more physical.

364 Now the new total tensor T  is defined as

 u u u 
  
a


   p 0 0      V 0 0   x y z 
v v v
T   0  p 0    0   V 0       
urn

(37)
 x y z 
 0 0  p  0 0   V   w w w 
   
 x y z 
365 In this new definition of the total tensor, the symmetric stress tensor is replaced by the friction tensor,
366 such a new definition (the friction from the stress tensor) could avoid the conflicts behind the original
367 N-S equation, including:
Jo

368 - In fluids, shear stresses may only exist when the fluids are flowing. This is very different from the
369 solids. Hence, the requirement of the symmetric stress tensor developed for solids may not be correct
370 for fluids, because we have already had exceptions (see [24, 32, 37, 41, 42]). In [31] (see page 126),
371 it states that the stress tensor symmetry is violated in the electric field on polarized fluid molecules,
372 where antisymmetric stresses must be included in the analysis. In fact, this problem can be solved
373 using the asymmetric friction tensor proposed in the research, Eq. (30), because both the symmetric
374 and anti-symmetric stresses have been automatically included. As such, Inconsistency 1 is resolved.

16
Journal Pre-proof

375 - All components of the fluid friction tensor have a correctly-defined physical property, including the
376 consistent and unique orientations of the frictions, hence Inconsistency 2 is resolved.
377 - Using the new friction tensor in Eq.(24), rather than the symmetric viscous stress tensor defined in
378 Eq. (8), Inconsistency 3 for the Couette flow is removed, shown by Eqs. (28) and (29).

379 6.2 Incompressible fluids


380 For an incompressible fluid, its density is constant. Together with   V  0 , Eq. (3) can be rewritten as
V 1 
 V   V   f B    T  (38)

of
t  
381 with
  p    

pro
  T     11  12  13 i
 x x1 x2 x3 
 p    
    21  22  23  j (39)
 y x1 x2 x3 
 p    
    31  32  33 k
re-
 z x1 x2 x3 
382 Using the new friction tensor of Eq. (25),

 11  12  13  2u1  2u1  2u1


    2   2   2   2u1 (40)
x1 x2 x3  x1  x2  x3
lP

383 Similarly, we have

  21  22  23
 x  x  x   u2
2

 1 2 3
(41)

  31   32   33   2u
 x1 x2 x3
3
a

384 Putting these together, we have the Navier-Stokes equation for the incompressible fluid as
urn

 u1 u u u 1 1 p
  u1 1  u2 1  u3 1  f 1    2u1
 t x1 x2 x3   x1
 u2 u u u 1 1 p (42)
  u1 2  u2 2  u3 2  f 2     2u 2
 t x1 x2 x3   x2
 u3 u u u 1 1 p
  u1 3  u2 3  u3 3  f 3     2u 3
Jo

 t x1 x2 x3   x3


385 In a vector form, it is
V 1 1
 V   V  f B  p   2 V (43)
t  
386 This is exactly same N-S equation for incompressible flows as that of original form of N-S equation.
387 This means that the N-S equation is exactly same as the original form as Stokes established in 1845,
388 however, some concepts and principles are different and of more physics.

17
Journal Pre-proof

389 6.3 Compressible fluids


390 For compressible flows, the Navier-Stokes equation can be derived as
V
  
 V   V   f B  p   2 V    V  (44)
 t 
391 To make the equation exactly same as the original Navier-Stokes equation for compressible flows as
392 Eq. (12), it is assumed as
1
  (45)

of
3
393 which is slightly different from the proposal that Stokes made in 1845, see Eq.(9). This difference is
394 caused because of the different symmetric stress tensor and the asymmetric friction tensor.

pro
395 The choice of Eq. (45) seems reasonable since the rate of volume dilatation appears 3 times in the
396 Navier-Stokes equation (in each component in the Cartesian coordinate system), so the coefficient of
397 1/3 is taken. However, more work must be done on how the second viscosity coefficient can be decided.

398 7 Discussions
re-
399 7.1 A shear stress applied on a solid element and a fluid element
400 Here we will see what will happen if a shear stress is applied on a solid element and a fluid element,
401 and will explain why the asymmetrical stress tensor will not cause a fluid spinning in the fluid.
If the element is an elastic material (e.g., a solid), under the shear stress τxz, there will be a deformation
lP

402
403 of the element. However, due to the elasticity, the element will produce a shear stress τzx to balance the
404 applied shear stress and avoid an infinite rotation of the element as seen in Figure 10. Also, τzx = τxz (this
405 is required for solids by the Cauchy symmetric stress tensor).
a
urn

406
407 Figure 10 A shear stress τ23 is applied on a solid element

408 If the element is a fluid, the applied shear stress, τxz, will simply cause a flow of the element (see based
Jo

409 on the fluid definition). Due to the flow motion, a fluid friction is induced which would balance the
410 applied shear stress. Therefore, this shows that there will be no shear stress pair to balance the applied
411 shear stress in fluids, but a fluid friction to balance the applied shear stress (see Figure 11), calculated
412 as,
u
 xz   (46)
z

18
Journal Pre-proof

413
414 Figure 11 A shear stress τ23 is applied on a fluid element

of
415 7.2 A proof for no vertical shear stress in the Couette flow
416 It may be argued that there may be a non-zero vertical shear stress τyx in the Couette flow, but the upper

pro
417 plate stops the vertical flow. Imagine if we make an open on the upper plate, there will be a flow coming
418 out from the opening because of the vertical shear stress (see Figure 12). Is this true?

?
re-
lP

419
420 Figure 12 An opening on the upper plate, is there flow coming out due to the vertical shear stress τyx?

421 However, the answer is NO. Because in reality, whether there will be a flow flowing out from the
422 opening totally depending on the pressure of the fluid between the plates. If the flow speed is large
423 enough, there may be a suction of flow in the opening.
a

424 Another proof is the experiment setup similar to the Bernoulli’s principle. At the bottom of the water
425 tank, two different size pipes are connected. Different sizes of the pipes would allow different flow
urn

426 velocities, as V1 and V2. Based on the Bernoulli’s principle, we could have different water heights in the
427 small tubes, as h1 and h2.
428 If the non-zero vertical stress, τyx, exists, it surely induce flows in the vertical small tubes (based on the
429 fluid definition). Therefore, there will be outflows in the small tubes regardless of the height of the
430 tubes because of the non-zero vertical shear stress. Obviously, this is not true. The heights of the fluid
431 in the small tubes, h1 and h2, totally depend on the pressures of the flow in the pipes.
Jo

19
Journal Pre-proof

of
432
433 Figure 13 An experiment setup for proving that the vertical shear stresses is zero

pro
434 8 Conclusions
435 The research has examined some fundamental issues behind the Navier-Stokes equation, and it is
436 identified that there exist some conflicts on the concepts and principles in deriving the N-S equation.
437 To solve these conflicts, the forces acting on the fluids have been re-analysed based on a solid physics.
438 By considering the fundamental properties of fluids, it can simply show that the symmetric stress tensor
re-
439 is violated in the classic Couette flow.
440 To resolve these conflicts behind N-S equation, it is proposed that the fluid friction is regarded as the
441 special surface force, not a part of the Stokes’ symmetric stress. Under this new concept, the friction
442 tensor is independent of the stress tensor, and more importantly, the friction tensor is not required to be
443 symmetric if the real physics of the fluid frictions are referred to.
lP

444 From the study and the understanding of the concepts and principles behind N-S equation in this
445 research work, the following conclusions can be drawn:
446 - Identify the conflicts of the concepts and principles in deriving the N-S equation.
447 - Illustrate a conflict on the requirement of the symmetric stress tensor using the well-known
a

448 Couette flow.


449 - Reformulate the total tensor by replacing the symmetric stress tensor with the fluid friction tensor,
urn

450 and the friction tensor is more based on the real physics of the friction acting on the fluids.
451 - The formulation of new friction tensor could resolve all identified inconsistencies, including the
452 shear stress in the Couette flow and the special problem as in the electric field on polarized fluid
453 molecules, since the anti-symmetric tensor is automatically included in the new total tensor
454 formulation.
455 - The exactly same N-S equation as the original N-S equation can be derived using the newly
Jo

456 defined total tensor for the incompressible flows. However, to obtain the same N-S equation for
457 compressible flows, a slight different yet very similar assumption from the Stokes’ assumption
458 should be taken.
459 The new friction tensor does lead to the exactly same N-S equation, hence it is hoped that this revisit of
460 N-S equation with the newly proposed friction tensor could provide a better physics in understanding
461 the N-S equation. With the better physics as proposed in this research, it may pave a way to understand
462 the fluid dynamics better, especially for those complicated flows. For instance, currently, solving the
463 N-S equation uses turbulence models in most practical applications, and the turbulence models have

20
Journal Pre-proof

464 been frequently constructed using the symmetric stress tensor following the requirement of the
465 symmetric strain-rate tensor, see [21, 43]. Under the new physical significance presented in the research
466 work, the asymmetric friction tensor could allow different views on the complex flows and it may shed
467 some light for reconstructing different turbulence models so to help to solve the fluid dynamic problems
468 better.

469

470 Appendix: Useful mathematical tools

of
471
472 A1. Material derivative
473 In the Euler expression, a given physical function, f, in the fluid motion can be expressed as the function

pro
474 of the independent variables: the coordinates (x, y, z) and time, t, as
f  f ( x, y , z ; t ) (A.1)

475 Note: the arbitrary function can be either a scalar, such as the fluid density, ρ; pressure, p, etc) or a
476 vector, such as, the fluid velocity, V, or the force, F and so on (hereafter the bold letters denote vectors).
477 The material derivative of the function can be easily obtained using the chain rule of derivative as
re-
Df f f f f (A.2)
 u v w
Dt t x y z
478 In a more compact manner, it is
Df f (A.3)
  V    f
Dt t
lP

479 The flow acceleration is the material derivative of the flow velocity vector with regard to time and it
480 can be derived in similar manner as above, so
DV V (A.4)
  V   V
Dt t
a

481 The first term of the right-hand-side of Eq. A.4 is the local acceleration (the independent coordinate
482 variables, (x, y, z), are all kept constant), and the second term is the convective acceleration due to the
urn

483 different velocities of the fluid particles over different positions.


484
485 A2. Gauss convergence theorem
486 Gauss divergence theorem is
 
 T  n dS   dV
  T (A.5)
Jo

S V

487
488 A3. Transport theorem
489 The transport theorem is a very important tool to derive the equation for fluid motion. For a given
490 physical function, f, in the fluid domain, V, the transport theorem gives

 f 
fdV        f V dV
D

Dt V V  t 
(A.6)

21
Journal Pre-proof

491 The details of the derivation of the transport theorem can be found in [35] (pages 57-59).
492

493 References
494 1. Anderson, J., A history of Aerodynamics, and its Impact on Flight Machines. Cambridge Aerospace Series
495 Vol. 8. 1997: Cambridge University Press.
496 2. Anderson, J., Brief History of the Early Development of Theoretical and Experimental Fluid Dynamics, in
497 Encyclopedia of Aerospace Engineering. 2010, John Wiley & Sons, Ltd.
498 3. Stokes, G.G., 1845. On the theories of the internal friction of fluids in motions, and the equilibrium and
499 motion of elastic solids. Transaction of the Cambridge Philosophical Society, 8: pp. 75-129.

of
500 4. Darrigol, O., 2002. Between Hydrodynamics and Elasticity Theory: The First Five Births of the Navier-
501 Stokes Equation. Archive for History of Exact Sciences, 56(2): pp. 95-150.
502 5. Speziale, C.G., 1991. Analytic methods for the development of Reynolds stress closures in turbulence.
503 Annual Review of Fluid Mechanics, 23: pp. 107-157. doi: 10.1146/annurev.fl.23.010191.000543.
504 6. Witherden, F.D. and A. Jameson. 2017, Future directions of computational fluid dynamics. 23rd AIAA

pro
505 Computational Fluid Dynamics Conference, AIAA 2017-3791. Denver, Colorado, US.
506 7. Spalart, P.R. and Venkatakrishnan, 2016. On the role and challenges of CFD in the aerospace industry. The
507 Aeronautical Journal 120(1223): pp. 209-232. doi: 10.1017/aer.2015.10.
508 8. McDaniel, D.R., R.H. Nichols, and S.A. Morton. 2017, Capabilities and Challenges in CFD: A Perspective
509 from the DoD HPCMP CREATETM-AV Kestrel Development Team. 23rd AIAA Computational Fluid
510 Dynamics Conference, 5-9 June 2017. Denver, Coorado, USA.
511 9. Cozzi, L., et al. 2017, Facing the challenges in CFD modelling of multistage axial compressors. Proceedings
512 of ASME Turbo Expo 2017: Turbomachinery Technical Conference and Exposition, June 26-30, 2017.
re-
513 Charlotte, NC, USA.
514 10. Garmann, D.J. and M.R. Visbal. 2018, Challenges & Perspectives on CFD and Experimental Interactions
515 for Complex Unsteady Flows (Invited). AIAA Aviation Forum, June 25-29, 2018. Atlanta, Georgia, US.
516 11. Moin, P. and K. Mahesh, 1998. DIRECT NUMERICAL SIMULATION: A Tool in Turbulence Research.
517 Annual Review of Fluid Mechanics, 30: pp. 539-578.
518 12. Sengupta, T.K. and S. Bhaumik, DNS of Wall-Bounded Turbulent Flows: A First Principle Approach. 2019:
lP

519 Springer.
520 13. Geogiadis, N.J., D.P. Rizzetta, and C. Fureby. 2009, Large-Eddy Simulation: Current Capabilities,
521 Recommended Practices, and Future Research. Proceedings of 47th Aerospace Sciences, Jan. 5-8, 2009.
522 Orlando, Florida, US.
523 14. Fureby, C. 2012, Large Eddy Simulation: A Useful Tool for Engineering Fluid Dynamics. Proceedings of
524 18th Australasian Fluid Mechanics Conference, 3-7 Dec. 2012. Launceston, Australia.
525 15. Pope, S.B., 2004. Ten questions concerning the large-eddy simulation of turbulent flows. New Journal of
526 Physics, 6: pp. 1-24. doi: 10.1088/1367-2630/6/1/035.
a

527 16. Rogallo, R. and P. Moin, 1984. Numerical simulation of turbulent flows. Annual Review of Fluid Mechanics,
528 16: pp. 99-137. doi: 10.1146/annurev.fl.16.010184.000531.
urn

529 17. Yang, Z., 2015. Large-eddy simulation: Past, present and the future. Chinese Journal of Aeronautics, 28(1):
530 pp. 11-24. doi: 10.1016/j.cja.2014.12.007.
531 18. Hanjalic, K. and B.E. Launder, Modelling Turbulence in Engineering and the Environment: Second-Moment
532 Routes to Closure. 2011: Cambridge University Press, UK.
533 19. Launder, B.E. and D.B. Spalding, 1974. The numerical computation of turbulence flows. Computer Methods
534 in Applied Mechanics and Engineering, 3: pp. 269-289.
535 20. Launder, B.E., G.J. Reece, and W. Rodi, 1975. Progress in the development of a Reynolds-stress turbulence
536 closure. Journal of Fluid Mechanics, 68: pp. 537-566.
537 21. Spalart, P.R. and S.R. Allmaras. 1992, A one-equation turbulence model for aerodynamic flows. 30th
Jo

538 Aerpspcae Science Meeting & Exhibit, Jan. 6-9, 1992. Reno, Nevada, USA.
539 22. Speziale, C.G., S. Sarkar, and T.B. Gatski, 1991. Modelling the pressure–strain correlation of turbulence: an
540 invariant dynamical systems approach. Journal of Fluid Mechanics, 227: pp. 245-272. doi:
541 10.1017/S0022112091000101.
542 23. Wilcox, D.C., 1988. Reattachment of the scale-determining equation for advanced turbulence models. AIAA
543 Journal, 26(1): pp. 1299-1310. doi: 10.2514/3.10041.
544 24. Wilcox, D.C., Turbulence Modelling for CFD (3rd Edition). 2006: DCW Industries, Inc.
545 25. Wilcox, D.C., 2008. Formulation of the k-w Turbulence Model Revisited. AIAA Journal, 46(11): pp. 2823-
546 2838. doi: 10.2514/1.36541.
547 26. Menter, F.R., 1994. Two-Equation Eddy-Viscosity Turbulence Models for Engineering Applications. AIAA
548 Journal, 32(8): pp. 1598-1605. doi: 10.2514/3.12149.

22
Journal Pre-proof

549 27. Menter, F.R., M. Kuntz, and R. Langtry, Ten Years of Industrial Experience with the SST Turbulence Model,
550 in Turbulence, Heat and mass Transfer, K. Hanjalic, Y. Nagano, and M. Tummers, Editors. 2003.
551 28. Menter, F.R., 2009. Review of the shear-stress transport turbulence model experience from an industrial
552 perspective. International Journal of Computational Fluid Dynamics, 23(4): pp. 305-316. doi:
553 10.1080/10618560902773387.
554 29. Menter, F.R., 2011, Turbulence Modeling for Engineering Flows: a Technical paper, ANSYS INC. Cited at:
555 30. Durbin, P.A., 2018. Some recent development in turbulence closure modeling. Annual Review of Fluid
556 Mechanics, 50: pp. 77-103. doi: 10.1146/annurev-fluid-122316-045020.
557 31. Kundu, P.K., I.M. Cohen, and D. Rowling, Fluid Mechanics (6th Edition). 2016: Elsevier, Inc.
558 32. Rigelesaiyin, J., et al., 2018. Asymmetry of the atomic-level stress tensor in homogeneous and
559 inhomogeneous materials. Proceedings of Royal Society A, 474: pp. 20180155. doi: 10.1098/rspa.2018.0155.
560 33. Bertin, J., J. and M.L. Smith, Aerodynamics for Engineers (2nd Edition). 1989, US: Prentice-Hall, Inc.

of
561 34. Schlichting, H. and K. Gersten, Boundary-Layer Theory (9th Edition). 2017: Springer-Verlag, Berlin
562 Heidelberg.
563 35. Newman, J.N., Marine Hydrodynamics. 1977: The MIT Press, Cambridge, Massachusetts, USA.
564 36. Bertin, J., J. and R.M. Cummings, Aerodynamics for Engineerings (6th Edition). 2013: Pearson.

pro
565 37. Brenner, H., 2005. Navier-Stokes revisited. Phyisca A, 394: pp. 60-132. doi: 10.1016/j.physa.2004.10.034.
566 38. Brenner, H., 2012. Beyond Navier–Stokes. International JournaL of Engineering Science, 54. doi:
567 10.1016/j.ijengsci.2012.01.006.
568 39. Houghton, E.L., et al., Aerodynamics for Engineering Students (6th Edition). 2013: Elsevier, Ltd.
569 40. White, F.M., Fluid Mechanics (8th Edition). 2016, New York, US.: McGraw-Hill Education.
570 41. Zhang, E., et al., 2009. Asymmetric tensor analysis for flow visualization. IEEE Transactions on
571 Visualzation and Computer Graphics, 15(1): pp. 106-122. doi: 10.1109/TVCG-2007-12-0188.
572 42. Vedan, M.J. and S.M. Panakkal, 2018. Vorticity and Stress Tensor. Journal of Informatics and mathematical
573 Science, 10: pp. 351-357.
re-
574 43. Wilcox, D.C. 2007, Formulation of the k-w turbulence model revisited. 45th AIAA Aerospace Sciences
575 Meeting and Exhibit, Jan. 8-11, 2007. Reno, Nevada, USA.
576
a lP
urn
Jo

23
Journal Pre-proof

Highlights for ‘A revisit of Navier-Stokes equation’

- Identify the conflicts of the concepts and principles behind the original N-S equation.
- Explain the conflicts using better physical understandings and illustrate the conflicts using the
simple classic Couette flow.
- Reformulate the total tensor by replacing the symmetric stress tensor using the fluid friction
tensor. The friction tensor is not required to be symmetric, better fitting to the real physics.

of
- The newly defined total tensor can be used to derive N-S equation, and lead to an exactly same
N-S equation.
- The formulation of new friction tensor could solve all identified inconsistencies completely.

pro
re-
a lP
urn
Jo
Journal Pre-proof

I would like to confirm that there is no conflict of interest in this research work.

Dr. Wanan Sheng

SW MARE Marine Technology

of
pro
re-
a lP
urn
Jo

You might also like