You are on page 1of 20

Chapter 3

Dynamics

Newtons second law of mechanics is a monumental achievement. It vali-


dates the concept of point particles and shows the relevance of forces; maybe
more importantly, it brings mathematics (calculus) to the center of our un-
derstanding of mechanical phenomena (a program initiated by Archimedes
of Syracuse in his palimpsest, and evolving through much of modern mathe-
matical physics).
Essential mathematical concepts are vectors and differentials; additional
mechanical ideas include stresses (in common with solid mechanics, under
the umbrella of continuum mechanics).

3.1 Newtonian dynamics of continua


The extension of Newtons second law from point particles to continua is
basis for both elasticity theory and advanced fluid dynamics. Euler took
the first step in this direction, by combining the use of material derivatives
with the knowledge (from fluid statics) that a pressure gradient is an internal
force. Dividing by the uniform density in an incompressible flow1
1
t u + u u = p (3.1)

1
The distinction between an incompressible flow, in which there are no significant effects
of compressibility, and the flow of an incompressible fluid, is important. Even water is not
incompressible, and a study of incompressible fluids would run head-on into the second
law of thermodynamics; on the other hand, the compressibility of air can be ignored in
low-speed aerodynamics: the weaker assumption of incompressible flow is sufficient to
allow for simple solutions. See Section 3.7 for details.

83
84 CHAPTER 3. DYNAMICS

Figure 3.1: Free-body diagram of an elementary volume, and a simpler 2D


variant used in the analysis

or
1
t ui + u j j ui = i p (3.2)

Eulers equation represents a leap away from material particles to the velocity
field in space, already discussed in the previous chapter. In hindsight, Eulers
equation is seen as ignoring the effects of viscosity: an approximation.
The next step was the work of Cauchy, who adopted as the basic object of
analysis the collection of particles included in a volume element specified in
some cartesian coordinates as dV = dx dy dz. The Lagrangian description is
implied for now. The free-body diagram (Fig. 3.1) entails body forces (such
as gravity, or the Lorentz force) applied at the center of mass, and surface
forces, applied at the center of each surface, that account for interactions
with the neighboring elements.
The body and surface forces can be decomposed into their cartesian com-
ponents, normalized by the size of the volume or surface to which they are
applied. In the case of surface forces, taking the outward normal as one of
the local axes, the normal component for force per unit area is a familiar
concept: negative pressure2 . The tangential forces per unit area on each
surface are the components of stress, one of the cornerstones of continuum
2
The negative sign comes from the simple consideration that the force in the equation
is the force applied to the fluid element, rather than by the fluid to the surface at which
it is measured.
3.1. NEWTONIAN DYNAMICS OF CONTINUA 85

mechanics. Note the direction of local axes to ensure right-handed local axes
on each surface element, which is specified by its direction and its area; for
each possible direction, there are 3 components of force. Thus there are 9
components of stress at each point. A vector such as position has 3 compo-
nents, that can be arranged as a vector for which the single index can have
3 possible values; whereas the stress requires two such indices, one for the
direction of the surface on which the stress is acting, and one for the com-
ponent of force. Thus the stress will be represented as a 3x3 matrix, with
intrinsic, component and index representations respectively as

Txx Tyx Tzx
T or Txy Tyy Tzy or Tij (3.3)

Txz Tyz Tzz

The first index denotes the direction of the surface, the second index the
direction of the force. Normal forces (per unit area) are therefore the diagonal
elements of stress. The resultant force on a surface with a given direction n
or ni is given by Cauchys relation

F =nT (3.4)

or
    Txx Tyx Tzx
Fx Fy Fz = nx ny nz Txy Tyy Tzy

(3.5)
Txz Tyz Tzz
or
Fi = nj Tji (3.6)
In this framework, rotational equilibrium of the material element gives an
important result. With the body forces applied at the center of mass, their
moment about the center of mass is zero. Taking the centroidal moment of
the components of stress (see Panton p120), we obtain

Tij = Tji (3.7)

This for static conditions, no rotational body forces. The symmetry of the
stress tensor is very important in all areas of continuum mechanics.
Then, momentum balance includes the net effect of all surface and body
forces (Fig. 3.2). The derivation is carried out in 2D, with the 3D version
similar (but with more terms). Consider the rectangular volume element
86 CHAPTER 3. DYNAMICS

Figure 3.2: The 2-D free body diagram for a fluid element

centered at (x, y) and of sides dx and dy. Let us focus on the x-component
of force dFx :
dFx = Bx dxdy + Txx |x+dx/2 dy Txx |xdx/2 dy
+Tyx |y+dy/2 dx Tyx |ydy/2 dx (3.8)
Expanding the various terms in Taylor series, we see that the leading terms
cancel out in pairs of opposite signs, and that the terms of order dx dy
remain. Dividing throughout by the volume gives for the x-component
Fx = Bx + x Txx + y Tyx (3.9)
or in intrinsic notations
F =B+T (3.10)
or again
Fi = Bi + j Tij (3.11)
(summation over j is implied). Cauchys idea that the divergence of stress is
equivalent to a force, has implications throughout continuum mechanics.

3.2 Stress at a point, Newtonian fluids


While Cauchys equation represents Newtons equation in continua, the ex-
pression for the stress tensor is not governed by fundamental laws (although
3.2. STRESS AT A POINT, NEWTONIAN FLUIDS 87

continuum mechanics imposes rational constraints on the possible forms of


the stress as function of other variables). It belongs in the group of phe-
nomenological (or constitutive) equations (or again equations of state), which
include Newtons law of viscosity, Ohms law of electrical resistance, Hookes
law of elasticity or Newtons law (again?!!) of heat exchange, and many
other empirical approximations: assuming a simple relation between shear
stress and shear rate (respectively: voltage and current, strain and deforma-
tion, heat flux and temperature difference), a truncated Taylor series yields
a sensible result. A complete theory is provided within the framework of
continuum mechanics; only the simplest variants need to be considered here.
First, we note that pressure is independent of direction, separate the
isotropic part of the stress:
Tij = pij + ij (3.12)
where (the deviatoric of stress, or viscous stress in the simple cases of in-
terest at this level) includes the non-isotropic contributions. In this instance,
we need to model the dependence of stress on flow variables. It is clear that
the viscous stress tensor cannot depend on velocity, because the addition of a
uniform velocity cannot modify the stress (Galilean invariance). The obvious
option is that the stress could depend on the velocity gradient; alternatives
include higher derivatives, time-derivatives (visco-elastic materials) or mem-
ory effects, or even non-local combinations of properties (required in some
polymer and biological flows). Here, we assume
= F (u) (3.13)
or
ij = Fij (k um ) (3.14)
Because is symmetric, linear algebra dictates that it can only be a
function of the symmetric part of the velocity gradient, i.e. the rate-of-strain
ij = Fij (skm ) (3.15)
(Note that, in index notations, the dependence of the i-j-component of stress
on any or all of the k-m-components of rate-of-strain is made explicit: writing
Fij (sij ) would imply that the x-y component of stress could not depend e.g.
on sxx , would be quite restrictive.) Assuming that the function F can be
expanded in Taylor series, the first few terms are
ij = C 0 ij + Cijkm
1 2
skm + Cijml smk skl + ... (3.16)
88 CHAPTER 3. DYNAMICS

where the coefficients could, in general, depend on the scalar invariants of


skm 3 .
Because the viscous stress has no isotropic contribution, we have C 0 = 0.
We will limit ourselves in this course to the first term in the series. The
student should be aware that relatively simple substances, such as power-
law fluids and visco-elastic fluids, are excluded. Also, we assume that the
coefficients C 1 , ... in the series are independent of the scalar invariants of
the rate-of-strain. It turns out that these assumptions allow for accurate
descriptions of many common gases and liquids.
Then, the simplest approximation is to neglect non-isotropic stresses
ij = 0 (3.17)
The corresponding force per unit volume is then
fi0 = i p (3.18)
which corresponds to Eulers inviscid dynamics. In intrinsic notations, we
have
T = p1 (3.19)

If terms linear in skm are retained, we obtain the viscous approximation


worked out rigorously by Stokes (1845) but first included in the equations of
motion by Navier (1821). For an incompressible fluid, we have
Tij = pij + 2sij (3.20)
See Panton for the additional terms for compressibility. It is easy to see
that, in a uniform shear flow (Couette flow) between two flat plates, this
expression reduces to Newtons definition of viscosity; hence the Newtonian
fluid.

3.2.1 The Navier-Stokes equations


On this basis, we obtain our basic form of momentum balance (Newtons
second law) for incompressible Newtonian fluids:
1 2
t ui + uj j ui = bi i p + jj ui (3.21)

3
First, second and third invariants of sij . See a textbook on linear algebra for details.
3.2. STRESS AT A POINT, NEWTONIAN FLUIDS 89

Figure 3.3: The stress tensor


90 CHAPTER 3. DYNAMICS

or, in intrinsic notations


1
t u + (u )u = b p + 2 u (3.22)

The body force per unit mass bi = Bi / is assumed to derive from a
potential from now on. Specifically,

bi = gi3 (3.23)

where the index 3 corresponds to the vertical upward direction. The corre-
sponding potential energy per unit mass is gx3 .

3.2.2 Boundary conditions


Impermeable walls moving at velocity uw . Then, the normal velocity must
match the motion of the wall:

u n = uw n (3.24)

The no-slip condition:


u n = uw n (3.25)
rather than merely
u = uw (3.26)
For an additional discussion, see Tritton Section 5.7 p.63. In addition to
the rarefied gas dynamics case mentioned there, it should be noted that the
no-slip condition (introduced by Stokes) is not necessarily exact, as recent
microfluidics studies have shown.
Boundaries may also be associated with the application of forces driving
the flow. Depending on the problem, it may be necessary to to impose
pressure or components of tangential stress.

3.2.3 Pressure
From experience in thermodynamics and pipe flows, pressure may appear
to be a simple term. As a matter of fact, if pressure is known, taking its
gradient and seeing its effect on velocity can be rather simple; but the inverse
problem, when the velocity field is known and pressure is the unknown, is
quite different. The analogy with the vorticity/velocity relationship in Ch.3
3.2. STRESS AT A POINT, NEWTONIAN FLUIDS 91

Figure 3.4: About pressure

(if velocity is known, take derivatives to get vorticity; if vorticity is known,


the Biot-Savart relation is far from trivial!) is very deep and the student
should ponder this.
Indeed, mass conservation for the incompressible flows
i ui = 0 (3.27)
gives, by application of the divergence:
2 p = i uj j ui = ij2 (ui uj ) = rhs (3.28)
(Note that, in intrinsic notations, the divergence of the nonlinear term is
ambiguous, whereas the index notation is quite clear). This is a Poisson equa-
92 CHAPTER 3. DYNAMICS

tion, similar to the relation between vector potential and vorticity, between
velocity and flexion. Here, the velocity gradients are the sources of pressure
variations, and the use of Greens functions gives the general solution for 3D
in the absence of boundaries:
1 rhs(x0 )
Z
p= dV 0 (3.29)
4 | x x0 |

We recognize the property of elliptic pdes, by which the solution at any


point depends on sources in the entire field, with vanishing awareness for
remote sources. Just as a vortex induces velocity throughout the flow domain,
velocity gradients induce pressure variations at large distances. We can hear
wind noise when driving a car, and it has been shown that vorticity is a
dominant source of pressure variations (hence the term: vortex noise). The
compressibility of air results in a finite speed of propagation of the pressure
fluctuations, captured in Kirchhoffs theory of delayed potentials.

3.2.4 Vorticity in NS
This brings up the idea of looking for vorticity in the Navier-Stokes dynamics.
We make use of two identities. First

( u) = = 2 u + ( u) (3.30)

in which the last term vanishes; second, the Lamb vector (see Ch. 2) reap-
pears since
1
( u) u = u = u u (u u). (3.31)
2
Then, the momentum equation can be rewritten as

p u2
t u + u = ( + + gz) (3.32)
2

So vorticity is present in the Navier-Stokes equations, affecting both the


viscous term and part of the nonlinear term when moved to the right-hand
side and multiplied by density, the Lamb vector is also called the Magnus
force. This term is a major contributor to pressure sources (above), i.e vortex
noise. The u2 -term on the r.h.s. is obviously related to Bernoullis equation,
a topic to be pursued in Ch.5.
3.2. STRESS AT A POINT, NEWTONIAN FLUIDS 93

Figure 3.5: Nonlocal effects in incompressible fluid dynamics


94 CHAPTER 3. DYNAMICS

3.3 Vorticity equation


Sensing the need to now more about vorticity, we turn to its dynamics. Tak-
ing the curl of the Navier-Stokes equations, we get after some rearrangement
2
t i + uj j i = j j ui + jj i (3.33)

On the l.h.s., we have the material derivative of the vorticity; on the r.h.s.,
the inviscid term is interpreted as a stretching term. We will return to it also
in Ch. 5.
While pressure is non explicitly present in the vorticity equation, its ef-
fects are implied through the velocity field. When reconstructing the velocity
from vorticity with the Biot-Savart formula, the irrotational part (with local
pressure: see Ch. 5) is missing and is uniquely determined by matching the
boundary conditions; the rotational part included in Biot-Savart is induced
by vorticity distribution throughout the field.

3.4 Energy and dissipation


Finally, the kinetic energy equation is obtained by taking the dot product
of NS on u it turns out that index notations work out better. Calling
u2 = ui ui , we have

u2 u2
t + uj j = bi ui + ui j Tij
2 2
= bi ui + j (ui Tij ) j ui Tij
= bi ui + j (ui Tij ) sij Tij
2
= bj uj + j (j u2 1 puj ) (j ui )2 , (3.34)

where we made use of incompressibility and of the renaming of summation


indices. The rate-of-strain appears because of the symmetry of Tij , which
cancels out the non-symmetric part of the velocity derivative. We recognize
the material derivative (as for all balance equations: Reynolds Transport
Theorem) on the l.h.s., then on the r.h.s. the rate of work from body forces,
rate of work from stresses (isotropic and viscous), and finally a term that
cannot be positive: the energy dissipation rate. Note that the dissipation
rate depends (quadratically) on the rate of strain, not vorticity.
3.4. ENERGY AND DISSIPATION 95

Figure 3.6: Energy balance: mechanics and thermodynamics

In the purely mechanical perspective afforded by Newtons second law,


energy is conserved in the sense that all changes are accounted for, but with
a net loss assoicated with dissipation. This is rather different from the ther-
modynamic perspective. The two viewpoints are reconciled by including a
thermal part of internal energy and energy exchange (Fig. 3.6). Then, the
dissipation term corresponds to the rearrangement of energy from mechani-
cal (as in the equation above) to internal energy, and does not show in the
overall budget.
96 CHAPTER 3. DYNAMICS

3.4.1 Bernoulli with losses


We can relate the energy considerations to Bernoullis equation with losses.
Strictly speaking, Bernoullis equation is a momentum equation (see Ch. 5),
derived under the assumption of inviscid flow. Also the spirit of Bernoullis
equation is to establish the conditions under which the Bernoulli term B is
constant, while Croccos theorem (Ch. 5), deriving how B varies, is closer to
the present topic. But the effect of viscosity is known as Bernoullis equation
with losses, and can be studied at this time in the context of energy.
We start from the NS equations in the form

t u + u = B + 2 u (3.35)

and evaluate the changes of B along a streamline (note the subtle meanings
of the u operation):

u B = u t u + u (2 u)
2 2
= (t u2 + 2 u2 ) (u)2 (3.36)

Two contributions are recognized on the r.h.s. First is the unsteady diffusion
of kinetic energy. For steady pipe flow, the time derivative term cancels out,
and the velocity profile shows a maximum at the centerline in the direction
of the flow, so the Laplacian of energy is negative, and kinetic energy diffuses
from the centerline toward the walls. Second, we see the dissipation term,
always negative. Both contributions (diffusion and dissipation) are negative
in steady flow, leading to a gradual loss in the value of B as we follow a
streamline. These effects are modeled in practice (see undergraduate texts)
in terms of an empirical formula based on the Darcy friction factor.

3.5 Enstrophy
Similar considerations apply to the square vorticity 2 = i i . The enstro-
phy 2 /2 plays an important role in geophysical applications and turbulence
theory. Starting from the vorticity equation and multiplying (with summa-
tion) by i , we get
2 2 2 2
t + uj j = i j j ui + jj 2
(j i )2
2 2
2 2
= i j sij + jj 2
(j i )2 (3.37)
3.6. STRUCTURE OF THE EQUATIONS 97

Beside the familiar convection and diffusion terms, we see a vortocity dissi-
pation term similar to its counterpart in the energy equation. The vortex
stretching term is quadratic in vorticity, and retains only the symmetric part
of velocity derivatives (rate-of-strain). We will return to this term in Ch. 5.

3.6 Structure of the equations


Our equations can be categorized in several ways:

the conservation laws (mass, momentum, and the pressure, vorticity,


energy, enstrophy and related equations as corollaries), as distinct from
the constitutive equations and from kinematic definitions, relations
and/or assumptions. The distinction is physical in nature.

Mathematical distinctions impact how equations can be solved, and


the similarities betwen the momentum, vorticity, energy and enstrophy
equations on one side, as distinct from the Poisson/Laplace equations
for pressure and other properties as seen later in these notes, and from
the definitions and constitutive relations and kinematic relations.

In this section, we focus on the latter.


A common feature of the momentum, vorticity, energy and enstrophy
equations is that they include (Fig. 3.7)

a convection-diffusion core, of the form t + u 2 applicable to


each of these properties;

some terms (including the convective derivative and diffusion term) of


divergence form (also called transport terms), which Gauss theorem
associates with boundary terms (what is the physics here? how about
numerical implications?)

the remaining terms, called dissipation or production if their signs are


known, or simply source/sink terms otherwise.

Other equations are somewhat subsidiary: continuity or pressure or stream-


function (they are not independent) and may be divided among

the Poisson and related equations (Laplace, Biot-Savart), for which the
Greens function introduces non-local effects.
98 CHAPTER 3. DYNAMICS

kinematic relations between properties

constitutive equations (some already implied, e.g. by diffusion, above)

other definitions

It is important to note that the use of Greens functions in relation with


the nonlocal effects relies on linearity (Laplacian) of either kinematic or dy-
namical relations: superposition of effects from distributed sources is essential
for Greens functions to be applicable. In the Navier-Stokes equations, the
nonlinear terms preclude the use of this approach: solutions based on Greens
function for diffusion cannot accommodate the convective terms.

3.7 Incompressible flow approximation


These notes deal exclusively with incompressible flows, where the simplified
form of the continuity equation applies:

i ui = 0. (3.38)

This is not the same as assuming that the fluid density is constant, which
would be an equation of state. For example, it is well known that ideal gas
behavior
pv = RT (3.39)
is consistent with incompressible aerodynamics. The resolution of this ap-
parent paradox rests on the observation that we only need to assume that the
effects of density variations are dynamically unimportant a much weaker
assumption. Two important classes of incompressible flows are established:
the small-Mach-number (low speed) flows, and the small-buoyancy natural
convection flows (Boussinesq approximation).
The analysis is based on the adoption of some reference density 0 and
small departures from it (/0  1); and the subtraction of the refer-
ence hydrostatic balance
3 p0 = 0 g (3.40)
from the momentum equation. The remaining terms are
1
t ui + u j j ui = i p gi3 + jj ui (3.41)
0
3.7. INCOMPRESSIBLE FLOW APPROXIMATION 99

Figure 3.7: Categories of equations and their structure


100 CHAPTER 3. DYNAMICS

neglecting corrections of order p. Note that the body force becomes


the buoyancy force responsible for natural convection and also relevant in
centrifuges and other applications. Compressibility effects are then captured
by = p
dp or
1
= p = p (3.42)
p
where is the (isentropic) compressibility coefficient.

3.7.1 Small Mach number flows


What would be the order of magnitude of pressure variations? The NS
equation shows that it can scale as the convective term or as the viscous
term. Focussing first on the former, we see that x p 12 x u2 , so that
Bernoulli-type scaling applies (p u2 ) and


u2 (3.43)

Introducing the general relation for the speed of sound a
p 1
a2 = |s (3.44)

we obtain
u2
2 M a2 . (3.45)
a
Consequently, density variations can be neglected if M a  1.
If viscous scaling is relevant, the incompressibility condition becomes

M a2  Re (3.46)

3.7.2 Boussinesq approximation


The Boussinesq approximation seems paradoxical at first: in natural con-
vection, the buoyancy force (dependent on thermal expansion) cannot be
neglected, so how come we can neglect density variations in mass balance?
In the full continuity equation

(t + ui i ) + i ui = 0 (3.47)
3.8. SUMMARY 101

the first group of terms (including derivatives of density) scale as U/L,


whereas the last group of 3 terms (summation...) scale as 0 U/L. So, as long
as /0 is small, we can keep the simpler form i ui = 0.
Consistency with momentum balance results form the fact that the buoy-
ancy term (variable density) is at most comparable to the other terms. It
can be shown that the corresponding scaling requirement is

gL < 1 (3.48)

See Tritton Section 15.1 p198, for additional details.

3.8 Summary
The Navier-Stokes equations are the cornerstone of fluid dynamics. They
embody Newtons second law (F = ma) for incompressible Newtonian fluids,
with more general forms available in the literature. They have also been
shown to be the statistical limit of kinetic theory (Chapman-Enskog) for
gases slightly out of equilibrium. In conjunction with kinematic constraints,
they represent the only analytic basis for the study of fluid motion.
Thus, it is humbling to realize that, nearly two centuries after they were
derived, so few solutions have been obtained. The undergraduate student
is exposed to Poiseuille and Couette flows, in which stationary flow and
geometry eliminate the time dependence and nonlinearities. Oseens vortex
is time-dependent, but there are no convective effects; the same holds for
Stokes two classic problems (see Ch. 6). See Pantons Chs. 7 and 11 for list
of solutions; see e.g. Acheson for the Burgers vortex.
But, as the leading idea through the remainder of the course, one can learn
from the equations without solving them. By understanding their physical
content, we will be able to work out rational approximations, keeping in mind
when they might fail and what the telltale signs might be.

3.9 Advanced topics and ideas for further read-


ing
Compressible flows, acoustics.

Non-Newtonian fluids, rheology


102 CHAPTER 3. DYNAMICS

Suspensions, colloids, foams, granular flows

Problems
1. Show that the energy dissipation rate depends on the rate-of-strain
(and not on vorticity). Show that, in the case of a flow enclosed by
rigid boundaries, the vanishing boundary flux terms (divergence) allow
the rewriting of dissipation rate as proportional to square vorticity.
(Adapted from Batchelor, p.263)

2. Consider a one-dimensional unsteady flow for which u(y, t) is the only


non-vanishing velocity component. Show that the equations of motion
take the form of unsteady diffusion. (from Tritton, p 471.)

3. Consider an infinite flat plate oscillating it its own plane with velocity
U = U0 sin t . Assume that the fluid oscillates at the same frequency
(but not in phase). How do the amplitude and phase of fluid motion
vary with distance from the plate (Stokes second problem).

4. Write any of the major equations (momentum, vorticity, energy, en-


strophy, temperature...) to emphasize transport terms (i.e. make di-
vergence terms appear wherever possible). Are these forms unique?

5. Helicity is defined as ui i . It is of some interest because it vanishes


in 2-D, and can serve as an indicator of 3-dimensionality. Using index
notations, derive the evolution equation governing helicity and label its
various terms.

6. Using index notations, derive the evolution equation governing the


Lamb vector and label its various terms.

7. (From Currie, p.60) Using index notations, prove the incompressible


kinematic relation
[(u )u] = 21 2 (u u) u (2 u) , or k (uj j uk ) = 12 kk
2
uj uj
2
uj kk uj j j , and discuss the nature of the various terms as sources
in the pressure equation.

You might also like