Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

3.13.

KINEMATICS AS A DYNAMICAL SYSTEM 101

3.13 Kinematics as a dynamical system


Let us apply some standard notions from dynamical systems theory (see Powers and Sen
(2015), Ch. 9.6; Powers (2016), Ch. 6.1; or Paolucci (2016), Ch. 3.2.6) to fluid kinematics.
Let us imagine that we are given a time-independent flow field, where the fluid velocity is
known and is a function of position only. Then the motion of an individual fluid particle is
governed by the following autonomous system of non-linear ordinary differential equations:

dx
= v(x(t)), x(0) = xo . (3.235)
dt
Here, the initial position of the fluid particle is given by the constant vector xo . The solution
of Eq. (3.235) can be expressed in general form

x = x(t; xo ), (3.236)

a function of time parameterized by the initial condition of the fluid particle. Such a solution
is certainly a pathline, streamline, and streakline. It is also known as a trajectory in the
dynamical systems literature.
Let us analyze Eq. (3.235) in some more detail. From the definition of the total derivative,
see Eq. (3.94), we have

dv = (∇vT )T ·dx, (3.237)


! "# $
LT
T
dv = L · dx. (3.238)

This gives the acceleration vector as

dv dx
= LT · , (3.239)
dt dt
= LT · v. (3.240)

Example 3.5
Study the following non-linear autonomous system, that could describe the steady three-dimensional
kinematics of a fluid:
dx1
= v1 (x1 , x2 , x3 ) = 1 + x1 x2 x3 , xo1 = 0, (3.241)
dt
dx2
= v2 (x1 , x2 , x3 ) = x1 + x22 + x1 x32 , xo2 = 0, (3.242)
dt
dx3
= v3 (x1 , x2 , x3 ) = 2 − x1 + x2 x3 , xo3 = 0. (3.243)
dt

CC BY-NC-ND. 06 February 2022, J. M. Powers.


102 CHAPTER 3. KINEMATICS

When considering dynamic systems, one should always consider equilibrium points. Such points
exist when v1 = v2 = v3 = 0; in fluid mechanics, they are known as stagnation points. They are found
by solving the nonlinear algebra problem:

0 = v1 (x1 , x2 , x3 ) = 1 + x1 x2 x3 , (3.244)
0 = v2 (x1 , x2 , x3 ) = x1 + x22 + x1 x32 , (3.245)
0 = v3 (x1 , x2 , x3 ) = 2 − x1 + x2 x3 . (3.246)

Numerical solution reveals three roots. Two of them are complex, and a third real. As we are generally
only concerned with real solutions, we focus only on that root, which is

x1o = 1, x2o = −1.46557, x3o = 0.682328. (3.247)

The “o” as a subscript denotes a stagnation condition, in contrast to “o” as a superscript, that denotes
an initial condition. Direct substitution into the equations for velocity confirm this is a stagnation
point. Flow in the neighborhood of a stagnation point can be understood by considering the locally
linear behavior. Taylor4 series of the velocity in the neighborhood of the stagnation point reveals the
local kinematics are given by the locally linear system
)
)
 dx1   ∂v1 ∂v1 ∂v1 ))  
dt ∂x1 ∂x2 ∂x3 ) x1 − x1o
 dx2  =  ∂v2 ∂v2 ∂v2 ))  x2 − x2o  , (3.248)
dt ∂x1 ∂x2 ∂x3 )
dx3 ∂v3 ∂v3 ∂v3 ) x − x
3 3o
dt
! ∂x1 ∂x "#2 ∂x3 $))
=LT xio
)
)
 ))  
x2 x3 x1 x3 x1 x2 ) x1 − x1o
)
=  1 + x32 2x2 + 3x1 x22 0 )  x2 − x2o  , (3.249)
)
−1 x3 x2 ) x3 − x3o
! "# $))
=LT xio
)
)
 ))  
−1. 0.682328 −1.46557 ) x1 − x1o
=  −2.1479 3.51255 0 ))  x2 − x2o  . (3.250)
)
−1 0.682328 −1.46557 ) x3 − x3o
! "# $))
=L T
xio

As discussed in standard texts on applied mathematics, e.g. Powers and Sen (2015), Ch. 9, the local
dynamics in the neighborhood of the stagnation point are dictated by the eigenvalues of the coefficient
matrix. Those are easily numerically evaluated as λ = 3.25643, −2.20944, and 0. The positive, negative,
and zero eigenvalues are associated with unstable, stable, and neutrally stable modes, respectively.
Because of the presence of both stable and unstable modes, this stagnation point is a so-called saddle
node. Had all eigenvalues been real and positive, it would have been an unstable source node. Had
all eigenvalues been real and negative, it would have been a stable sink node. Had any eigenvalues
contained an imaginary component, the solution could take on a locally oscillatory behavior.
4
Brook Taylor, 1685-1731, English mathematician and artist, educated at Cambridge, published on cap-
illary action, magnetism, and thermometers, adjudicated the dispute between Newton and Leibniz over
priority in developing calculus, contributed to the method of finite differences, invented integration by parts,
has his name ascribed to Taylor series of which variants were earlier discovered by Gregory, Newton, Leibniz,
Johann Bernoulli, and de Moivre.

CC BY-NC-ND. 06 February 2022, J. M. Powers.


3.13. KINEMATICS AS A DYNAMICAL SYSTEM 103

x1 x3
1.2
1.5
1.0

0.8
1.0
0.6

0.4
0.5
0.2

t t
0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0

x2 x1
x2
0.6

0.5

0.4

0.3

0.2
x3
0.1

0.2 0.4 0.6 0.8 1.0


t

Figure 3.16: Plot of x1 (t), x2 (t), x3 (t), along with the coincident pathline, streamline, and
streakline for a steady three-dimensional fluid particle that commences at the origin.

Numerical solution of this nonlinear system of ordinary differential equations yields x1 (t), x2 (t),
x3 (t), that for this time-independent velocity field induces the particle pathlines, streamlines, and
streaklines. All are plotted in Fig. 3.16. We could also apply the complete mathematical theory of
dynamic systems to understand the system better.
We can use Eq. (3.240) to calculate the acceleration vector field:

 dv1   ∂v1 ∂v1 ∂v1   dx1 


dt ∂x1 ∂x2 ∂x3 dt
 dv2  =  ∂v2 ∂v2 ∂v2   dx2  , (3.251)
dt ∂x1 ∂x2 ∂x3 dt
dv3 ∂v3 ∂v3 ∂v3 dx3
dt ∂x1 ∂x2 ∂x3 dt
  
x2 x3 x1 x3 x1 x2 1 + x1 x2 x3
=  1 + x32 2x2 + 3x1 x22 0   x1 + x22 + x1 x32  , (3.252)
−1 x3 x2 2 − x1 + x2 x3

CC BY-NC-ND. 06 February 2022, J. M. Powers.


104 CHAPTER 3. KINEMATICS

 
2x1 x2 − x21 x2 + x21 x3 + x2 x3 + 2x1 x22 x3 + x21 x32 x3 + x1 x22 x23
=  1 + 2x1 x2 + 3x21 x22 + 3x32 + 5x1 x42 + 3x21 x52 + x1 x2 x3 + x1 x42 x3  . (3.253)
−1 + 2x2 − x1 x2 + x1 x3 − x1 x2 x3 + 2x22 x3 + x1 x32 x3
If we know the kinematics of a fluid particle, we know everything about its motion, including its
acceleration. We shall soon, Ch. 4.2, discuss things like Newton’s second law of motion that relates
acceleration to forces. If we know the acceleration, it is possible to induce what the force was that
generated it by simply multiplying the acceleration by the mass. Rarely is this the case however. It is
more common to know something about the forces and to use this to deduce what the motion is.

Now, we seek to analyze a particular pathline. Note that the velocity vector is tangent
to the fluid particle trajectory. Let us study a unit vector that happens to be tangent to the
velocity field:
v
αt = . (3.254)
|v|
Next, use the quotient rule to examine how the unit tangent vector evolves with time:
dαt 1 dv v d|v|
= − . (3.255)
dt |v| dt |v|2 dt
We can scale Eq. (3.240) by |v| to get (1/|v|)dv/dt = LT · v/|v| = LT · αt . Thus Eq. (3.255)
can be rewritten as
dαt v d|v|
= LT · αt − , (3.256)
dt |v|2 dt
1 d|v|
= LT · αt − αt . (3.257)
|v| dt
Next consider the following series of operations starting with Eq. (3.240):
dv
= LT · v, (3.258)
dt
T dv
v · = vT · LT · v, (3.259)
* T dt +
d v ·v
= vT · LT · v, (3.260)
dt 2
* 2+
d |v|
= vT · LT · v, (3.261)
dt 2
d
|v| (|v|) = vT · LT · v, (3.262)
dt
1 d vT T v
(|v|) = ·L · , (3.263)
|v| dt |v| |v|
1 d
(|v|) = αTt · LT · αt . (3.264)
|v| dt

CC BY-NC-ND. 06 February 2022, J. M. Powers.


3.13. KINEMATICS AS A DYNAMICAL SYSTEM 105

Now substitute Eq. (3.264) into Eq. (3.257) to get

dαt , -
= LT · αt − αTt · LT · αt αt . (3.265)
dt

As an aside, take the dot product of Eq. (3.265) with αt to get

dαt , -
αTt · = αTt · LT · αt − αTt · LT · αt αTt · αt , (3.266)
dt ! "# $
=1
= αTt T
· L · αt − αTt T
· L · αt , (3.267)
= 0. (3.268)

This must be an identity, because αTt · αt = 1, and its time derivative gives αTt · dα/dt = 0.
Now recalling Eq. (3.99), and employing αTt · RT · αt = 0, because of the anti-symmetry of
R, and DT = D, because of the symmetry of D, Eq. (3.265) can be rewritten as

dαt , -
= LT · αt − αTt · D · αt αt . (3.269)
dt

Let us consider how a volume stretches in a direction aligned with the velocity vector.
We first specialize the general differential arc length to that found along the particle path:
ds = ds. Now, recall from geometry that the square of the differential arc length must be

ds2 = dxT · dx, (3.270)

where dx is also confined to the particle path. Consider now how this quantity changes with
time when we move with the particle:

d d , T -
(ds)2 = dx · dx , (3.271)
dt dt
* +T
T d d
= dx · (dx) + (dx) · dx, (3.272)
dt dt
* + * * ++T
T dx dx
= dx · d + d · dx, (3.273)
dt dt
= dxT · dv + dvT · dx, (3.274)
= 2 dxT · dv, (3.275)
= 2 dxT · LT · dx, (3.276)
d
2 ds (ds) = 2 dxT · LT · dx, (3.277)
dt
1 d dx T T dx
(ds) = ·L · . (3.278)
ds dt ds ds
CC BY-NC-ND. 06 February 2022, J. M. Powers.
106 CHAPTER 3. KINEMATICS

Recall now that


v
αt = , (3.279)
|v|
dx
dt
= ds
, (3.280)
dt
dx
= . (3.281)
ds
So, Eq. (3.278) can be rewritten as

1 d
(ds) = αTt · LT · αt , (3.282)
ds dt
d
(ln ds) = αTt · (D + R)T · αt , (3.283)
dt
= αTt · D · αt , (3.284)
= D : αt αTt . (3.285)

This relative tangential stretching rate is closely related to the result of Eq. (3.125) for
extensional strain rate. Specializing Eq. (3.125) for a particle pathline, and combining, we
can say

dv(es) = (αTt · D · αt )αt ds, (3.286)


dv(es)
= (αTt · D · αt )αt , (3.287)
ds
dv(es)
αTt · = (αTt · D · αt ) αTt · αt , (3.288)
ds ! "# $
=1
1 d
= αTt · D · αt = (ds), (3.289)
ds dt* +
1 ds
= αTt · D · αt = d , (3.290)
ds dt
d|v|
= αTt · D · αt = , (3.291)
ds
d|v|
= D : αt αTt = . (3.292)
ds
Here, we invoked Eq. (3.284) to obtain Eq. (3.290). The quantity αTt · D · αt = D : αt αTt
is a measure of how the magnitude of the velocity changes with respect to arc length along
the particle path.
We can gain further insight into how velocity magnitude changes by a diagonal decom-
position of D = Q · Λ · QT , where Q is an orthogonal rotation matrix with the normalized
eigenvectors of D in its columns, and Λ is the diagonal matrix with the eigenvalues of D in

CC BY-NC-ND. 06 February 2022, J. M. Powers.


3.13. KINEMATICS AS A DYNAMICAL SYSTEM 107

its diagonal. Thus


d|v|
= αTt · Q · Λ · QT ·αt , (3.293)
ds ! "# $
D
= (Q · αt ) · Λ · (QT · αt ).
T T
(3.294)
The operation QT · αt ≡ αs generates a new rotated unit vector αs = (αs1 , αs2 , αs3)T . Thus
we can state
d|v| 2 (1) 2 (2) 2 (3)
= αs1 λ + αs2 λ + αs3 λ , (3.295)
ds
2 2 2
1 = αs1 + αs2 + αs3 . (3.296)
The rate of change of the velocity magnitude along a particle pathline can be understood
to be a weighted average of the eigenvalues of the deformation tensor D. In the special
case in which αt is the ith eigenvector of D, we simply get d|v|/ds = λ(i) , where λ(i) is the
corresponding eigenvalue.
If we extend Eq. (3.184) to differential material volumes, we could say the relative ex-
pansion rate is
1 d
(dV ) = tr D, (3.297)
dV dt
d
(ln dV ) = tr D. (3.298)
dt
Now our differential volume can be formed by
dV = dA ds, (3.299)
where dA is the cross-sectional area normal to the flow direction. Thus
ln dV = ln dA + ln ds, (3.300)
ln dA = ln dV − ln ds, (3.301)
d d d
(ln dA) = (ln dV ) − (ln ds) . (3.302)
dt dt dt
Substitute from Eqs. (3.284,3.298) to get the relative rate of change of the differential area
normal to the flow direction:
d
(ln dA) = tr D − αTt · D · αt . (3.303)
dt
This relation, while not identical, is similar to the expression for shear strain rate, Eq. (3.135).
We can also use Eq. (2.98) to rewrite Eq. (3.303) as
d
(ln dA) = D : I − D : αt αTt , (3.304)
dt , -
= D : I − αt αTt . (3.305)

CC BY-NC-ND. 06 February 2022, J. M. Powers.


108 CHAPTER 3. KINEMATICS

Now the matrix I − αt αTt has some surprising properties. It is singular and has rank two.
Because it is symmetric, it has a set of three orthogonal eigenvectors that can be normalized
to form an orthonormal set. Its three eigenvalues are 1, 1, and 0. Remarkably, the eigenvector
associated with the zero eigenvalue must be parallel to and can be selected as αt , the unit
tangent to the curve. Thus the other two eigenvectors can be thought of as unit normals to
the curve, that we label αn1 and αn2 . These eigenvectors are not unique; however, a set can
always be found. We can summarize the decomposition in the following steps:

I − αt αTt = Q · Λ · QT , (3.306)
 . .. ..    
.. . . 1 0 0 · · · αTn1 ···
 
=  αn1 αn2 αt  0 1 0   · · · αTn2 ···, (3.307)
.. .. .. 0 0 0 · · · αTt ···
. . .
= αn1 αTn1 + αn2 αTn2 . (3.308)

The two unit normals are orthogonal to each other, αTn1 · αn2 = 0. Thus, we have
d , -
(ln dA) = D : αn1 αTn1 + αn2 αTn2 , (3.309)
dt
= D : αn1 αTn1 + D : αn2 αTn2 , (3.310)
= αTn1 · D · αn1 + αTn2 · D · αn2 . (3.311)

Comparing to Eq. (3.284) that has one mode associated with αt available for stretching of
the one-dimensional arc length in the streamwise direction, there are two modes associated
with αn1 , αn2 available for stretching the two-dimensional area.
The form αTn1 · D · αn1 suggests it determines the relative normal stretching rate in the
direction of αn1 ; a similar rate exists for the other normal direction. One might imagine
that there exists a normal direction that yields extreme values for relative normal stretching
rates. It is easily shown this achieved by the following. First, define a rectangular matrix,
0 whose columns are populated by αn1 and αn2 :
Q,
 . .. 
.. .
0  
Q =  αn1 αn2  . (3.312)
.. ..
. .
0 associated with
Then project the 3 × 3 matrix D onto this basis to form the 2 × 2 matrix D
stretching in the directions normal to the motion:
0=Q
D 0 T · D · Q.
0 (3.313)
0 give the maximum and minimum values of the relative normal stretching
The eigenvalues of D
rates, and the eigenvectors give the associated directions of extremal normal stretching.
Looked at another way and motivated by standard results from differential geometry,
we can make special choices, αn1 = αnp , αn2 = αnb , where αnp is the so-called “principal

CC BY-NC-ND. 06 February 2022, J. M. Powers.


3.13. KINEMATICS AS A DYNAMICAL SYSTEM 109

normal unit vector” and αnb is the so-called “bi-normal unit vector.” The following results
are described in more detail in many sources, e.g. Powers and Sen (2015), p. 89. We have
the so-called “Frenet-Serret”5 relations:
dαt
= καnp , (3.314)
ds
dαnp
= −καt − τ αnb , (3.315)
ds
dαnb
= τ αnp . (3.316)
ds
Here κ is the so-called “curvature,” of the curve and τ is the so-called “torsion” of the curve.
One can, with effort show that κ and τ are given by
1
) d2 x )2 ) dx )2 2 dx T d2 x 32 )
) dx d2 x )
)
) 2) ) ) − · ) dt × dt2 )
dt dt dt dt2
κ = ) dx )3 = ) )3 , (3.317)
) ) ) dx )
dt dt
2 2
3T 3
− dx dt
× ddt2x · ddt3x
τ = ) ) ) ) 2 32 . (3.318)
) d2 2x )2 ) dx )2 − dx T · d2 2x
dt dt dt dt

Note κ and τ are expressed here as functions of time. This is certainly the case for a particle
moving along a path in time. But just as the intrinsic curvature of a mountain road is
independent of the speed of the vehicle traveling on the road, despite the traveling vehicle
experiencing a time-dependency of curvature, the curvature and torsion can be considered
more fundamentally to be functions of position only, given that the velocity field is known
as a function of position. Analysis reveals in fact that
4
(vT · L · LT · v) (vT · v) − (vT · LT · v)2
κ = . (3.319)
(vT · v)3/2
One could also develop an expression for torsion that is explicitly dependent on position.
The expression is complicated and requires the use of third order tensors to capture the
higher order spatial variations.
We can also use this intrinsic orthonormal basis to get
d , -
(ln dA) = D : αnp αTnp + αnb αTnb , (3.320)
dt
= D : αnp αTnp + D : αnb αTnb , (3.321)
= αTnp · D · αnp + αTnb · D · αnb . (3.322)
The following example, adapted from Powers and Sen (2015), illustrates how kinematics
illuminates the general field of nonlinear dynamical systems.
5
Jean Frédeŕic Frenet, 1816-1900, and Joseph Alfred Serret, 1819-1885, French mathematicians.

CC BY-NC-ND. 06 February 2022, J. M. Powers.


110 CHAPTER 3. KINEMATICS

Example 3.6
Consider the example of Mengers6

dx1 1
= (1 − x21 ), (3.323)
dt 20
dx2 35
= −2x2 − x2 + 2(1 − x21 )x3 , (3.324)
dt 16
dx3
= x2 + x3 , (3.325)
dt
and identify so-called heteroclinic trajectories and their attractiveness.

There are only two finite equilibria for this system, a saddle at (−1, 0, 0)T and a sink at (1, 0, 0).
Because the first equation is uncoupled from the second two and is sufficiently simple, it can be inte-
grated exactly to form x1 = tanh(t/20). This, coupled with x2 = 0 and x3 = 0, satisfies all differential
equations and connects the equilibria, so the x1 axis for x1 ∈ [−1, 1] is what is known as the heteroclinic
trajectory. One then asks if nearby trajectories are attracted to it. This can be answered by a local
geometry-based analysis. Our system is of the form dx/dt = v(x). Let us consider its behavior in the
neighborhood of a generic point x0 that is on the heteroclinic trajectory, but is far from equilibrium.
We then locally linearize our system as

d
(x − x0 ) = v(x0 ) + L|x0 · (x − x0 ) + . . . , (3.326)
dt ! "# $ ! "# $
translation deformation+rotation
= v(x0 ) + D|x0 · (x − x0 ) + R|x0 · (x − x0 ) + . . . . (3.327)
! "# $ ! "# $ ! "# $
translation deformation rotation

Here, we have employed the local velocity gradient L as well as its symmetric (D) and anti-symmetric
(R) parts:

∂v L + LT L − LT
L= = D + R, D= , R= . (3.328)
∂x 2 2
The symmetry of D allows definition of a real orthonormal basis. For this three-dimensional system,
the dual vector ω of the anti-symmetric R defines the axis of rotation, and its magnitude ω describes
the rotation rate. Now the relative volumetric stretching rate is given by tr L = tr D = div v. And it is
not difficult to show that the linear stretching rate D associated with any direction with unit normal
α is D = αT · D · α.
For our system, we have
 
− x101 0 0
L =  −4x1 x3 −2 − 16 35
+ 2(1 − x21 )  . (3.329)
0 1 1

We see that the relative volumetric expansion rate is


x1
tr L = −1 − . (3.330)
10
6
J. D. Mengers, 2012, “Slow invariant manifolds for reaction-diffusion systems,” Ph.D. Dissertation,
University of Notre Dame, Notre Dame, Indiana.

CC BY-NC-ND. 06 February 2022, J. M. Powers.


3.13. KINEMATICS AS A DYNAMICAL SYSTEM 111

Because on the heteroclinic trajectory x1 ∈ [−1, 1], we always have a locally shrinking volume on that
trajectory. Now by inspection the unit tangent vector to the heteroclinic trajectory is αt = (1, 0, 0)T .
So the tangential stretching rate on the heteroclinic trajectory is
x1
Dt = αTt · L · αt = − . (3.331)
10
So near the saddle we have Dt = 1/10, and near the sink we have Dt = −1/10. Now we are concerned
with stretching in directions normal to the heteroclinic trajectory. Certainly two unit normal vectors
are αn1 = (0, 1, 0)T and αn2 = (0, 0, 1)T . But there are also infinitely many other unit normals. A
detailed optimization calculation reveals however that if we 1) form the 3 × 2 matrix Qn with αn1 and
αn2 in its columns:
 
0 0
Qn =  1 0  , (3.332)
0 1

and 2) form the 2×2 matrices Dn and Rn associated with the plane normal to the heteroclinic trajectory

Dn = QTn · D · Qn , Rn = QTn · R · Qn , (3.333)

that a) the eigenvalues of Dn give the extreme values of the normal stretching rates Dn1 and Dn2 ,
and the normalized eigenvectors give the associated directions for extreme normal stretching and b)
the magnitude of extremal rotation in the hyperplane normal to αt is given by ω = ||Rn ||2 . On the
heteroclinic trajectory, we find
 x1 
− 10 0 0
D= 0 −2 19
− 32 + 1 − x21  . (3.334)
19 2
0 − 32 + 1 − x1 1

The reduced deformation tensor associated with motion in the normal plane is
* +
T −2 − 19
32 + 1 − x21
Dn = Qn · D · Qn = . (3.335)
− 19 2
32 + 1 − x1 1

Its eigenvalues give the extremal normal stretching rates that are
5
1 2473 − 832x21 + 1024x41
Dn,1,2 = − ± . (3.336)
2 32
For x1 ∈ [−1, 1], we have Dn,1 ≈ 1 and Dn,2 ≈ −2. Because of the presence of a positive normal
stretching rate, one cannot guarantee trajectories are attracted to the heteroclinic trajectory, even though
volume of nearby points is shrinking. Positive normal stretching does not guarantee divergence from the
heteroclinic trajectory; it permits it. Rotation can orient a collection of nearby points into regions where
there is either positive or negative normal stretching. There are two possibilities for the heteroclinic
trajectory to be attracting: either 1) all normal stretching rates are negative, or 2) the rotation rate is
sufficiently fast and the overall system is volume-decreasing,7 so that the integrated effect is relaxation
to the heteroclinic trajectory. For the heteroclinic trajectory to have the additional property of being
restricted to the slow dynamics, we must additionally require that the smallest normal stretching rate
be larger than the tangential stretching rate.
We illustrate these notions in the sketch of Fig. 3.17. Here we imagine a sphere of points as initial
7
Such systems have ∇T · v < 0. In the dynamic systems literature, this is known as a dissipative
system; however, in fluid mechanics we reserve the word “dissipative” for systems that have thermodynamic
irreversibilities.

CC BY-NC-ND. 06 February 2022, J. M. Powers.


112 CHAPTER 3. KINEMATICS

sink

saddle

Figure 3.17: Sketch of a phase volume with ∇T · v < 0 showing heteroclinic connection
between a saddle and sink equilibria along with the evolution of a set of points initially
configured as a sphere as they move into regions with some positive normal stretching rates.

conditions near the saddle. We imagine that the system is such that the volume shrinks as the sphere
moves. While the overall volume shrinks, one of the normal stretching rates is positive, admitting
divergence of nearby trajectories from the heteroclinic trajectory. Rotation orients the volume into a
region where negative normal stretching brings all points ultimately to the sink.
For our system, families of trajectories are shown in Fig. 3.18a, and it is seen that there is divergence
from the heteroclinic trajectory. This must be attributed to some points experiencing positive normal
stretching away from the heteroclinic trajectory. For this case, the rotation rate is ω = −51/32 + 1 − x21.
Thus, the local rotation has a magnitude of near unity near the heteroclinic orbit, and the time scales
of rotation are close to the time scales of normal stretching.
We can modify the system to include more rotation. For instance, replacing Eq. (3.325) by dx3 /dt =
10x2 +x3 introduces a sufficient amount of rotation to render the heteroclinic trajectory to be attractive
to nearby trajectories. Detailed analysis reveals that this small change 1) does not change the location
of the two equilibria, 2) does not change the heteroclinic trajectory connecting the two equilibria, 3)
modifies the dynamics near each equilibrium such that both have two stable oscillatory modes, with
the equilibrium at (−1, 0, 0)T also containing a third unstable mode and that at (1, 0, 0)T containing
a third stable mode, 4) does not change that the system has a negative volumetric stretch rate on
the heteroclinic trajectory, 5) does not change that a positive normal stretching mode exists on the
heteroclinic trajectory, and 6) enhances the rotation such that the heteroclinic trajectory is locally
attractive. This is illustrated in Fig. 3.18b.
Had the local velocity gradient been purely symmetric, interpretation would be much easier. It is
the effect of a non-zero anti-symmetric part of L that induces the geometrical complexities of rotation.
Such systems are often known as non-normal dynamical systems.

CC BY-NC-ND. 06 February 2022, J. M. Powers.


3.13. KINEMATICS AS A DYNAMICAL SYSTEM 113

x3 1
0
-1
-1 -1

saddle
equilibrium
with one unstable
x2
0 mode sink
0
x2
sink

1
-1 1
1
-1 0
0 x3
a) x1 0
x1
1 b) 1 -1

Figure 3.18: Plots of trajectories near the heteroclinic connection between equilibria with
one unstable mode and a sink illustrating a) divergence of nearby trajectories due to positive
normal stretching with insufficiently rapid rotation and b) convergence of nearby trajectories
in the presence of positive normal stretching with sufficiently rapid rotation.

CC BY-NC-ND. 06 February 2022, J. M. Powers.

You might also like