Maneuvering and Control of Marine Vehicles
Maneuvering and Control of Marine Vehicles
Maneuvering and Control of Marine Vehicles
OF MARINE VEHICLES
Michael S. Triantafyllou
Franz S. Hover
Department of Ocean Engineering
Massachusetts Institute of Technology
Cambridge, Massachusetts USA
These notes were developed in the instruction of the MIT graduate subject
13.49: Maneuvering and Control of Surface and Underwater Vehicles. We plan
many enhancements; your comments are welcome!
i
3.3.3 Stationary Frame . . . . . . . . . . . . . . . . . . . . . 24
3.4 Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . 24
3.5 Example: Spinning Book . . . . . . . . . . . . . . . . . . . . . 26
3.5.1 x-axis . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.5.2 y-axis . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.5.3 z-axis . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.6 Parallel Axis Theorem . . . . . . . . . . . . . . . . . . . . . . 28
3.7 Basis for Simulation . . . . . . . . . . . . . . . . . . . . . . . 28
4 HYDRODYNAMICS: INTRODUCTION 30
4.1 Taylor Series and Hydrodynamic Coefficients . . . . . . . . . . 30
4.2 Surface Vessel Linear Model . . . . . . . . . . . . . . . . . . . 31
4.3 Stability of the Sway/Yaw System . . . . . . . . . . . . . . . . 32
4.4 Basic Rudder Action in the Sway/Yaw Model . . . . . . . . . 34
4.4.1 Adding Yaw Damping through Feedback . . . . . . . . 36
4.4.2 Heading Control in the Sway/Yaw Model . . . . . . . . 36
4.5 Response of the Vessel to Step Rudder Input . . . . . . . . . . 37
4.5.1 Phase 1: Accelerations Dominate . . . . . . . . . . . . 37
4.5.2 Phase 3: Steady State . . . . . . . . . . . . . . . . . . 37
4.6 Summary of the Linear Maneuvering Model . . . . . . . . . . 38
4.7 Stability in the Vertical Plane . . . . . . . . . . . . . . . . . . 38
5 SIMILITUDE 40
5.1 Use of Nondimensional Groups . . . . . . . . . . . . . . . . . 40
5.2 Common Groups in Marine Engineering . . . . . . . . . . . . 42
5.3 Similitude in Maneuvering . . . . . . . . . . . . . . . . . . . . 44
5.4 Roll Equation Similitude . . . . . . . . . . . . . . . . . . . . . 46
6 CAPTIVE MEASUREMENTS 48
6.1 Towtank . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
6.2 Rotating Arm Device . . . . . . . . . . . . . . . . . . . . . . . 48
6.3 Planar-Motion Mechanism . . . . . . . . . . . . . . . . . . . . 49
ii
7.3.2 Speed Loss . . . . . . . . . . . . . . . . . . . . . . . . 53
7.3.3 Heel Angle . . . . . . . . . . . . . . . . . . . . . . . . . 53
7.3.4 Heeling in Submarines with Sails . . . . . . . . . . . . 54
8 STREAMLINED BODIES 56
8.1 Nominal Drag Force . . . . . . . . . . . . . . . . . . . . . . . 56
8.2 Munk Moment . . . . . . . . . . . . . . . . . . . . . . . . . . 56
8.3 Separation Moment . . . . . . . . . . . . . . . . . . . . . . . . 57
8.4 Net Effects: Aerodynamic Center . . . . . . . . . . . . . . . . 57
8.5 Role of Fins in Moving the Aerodynamic Center . . . . . . . . 58
8.6 Aggregate Effects of Body and Fins . . . . . . . . . . . . . . . 60
8.7 Coefficients Zw , Mw , Zq , and Mq for a Slender Body . . . . . . 60
9 SLENDER-BODY THEORY 62
9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
9.2 Kinematics Following the Fluid . . . . . . . . . . . . . . . . . 62
9.3 Derivative Following the Fluid . . . . . . . . . . . . . . . . . . 63
9.4 Differential Force on the Body . . . . . . . . . . . . . . . . . . 64
9.5 Total Force on a Vessel . . . . . . . . . . . . . . . . . . . . . . 64
9.6 Total Moment on a Vessel . . . . . . . . . . . . . . . . . . . . 65
9.7 Relation to Wing Lift . . . . . . . . . . . . . . . . . . . . . . . 67
9.8 Convention: Hydrodynamic Mass Matrix A . . . . . . . . . . 67
iii
12.2.2 Solution for Steady Conditions . . . . . . . . . . . . . 81
12.2.3 Engine/Motor Models . . . . . . . . . . . . . . . . . . 81
12.3 Unsteady Propulsion Models . . . . . . . . . . . . . . . . . . . 83
12.3.1 One-State Model: Yoerger et al. . . . . . . . . . . . . . 83
12.3.2 Two-State Model: Healey et al. . . . . . . . . . . . . . 84
13 TOWING OF VEHICLES 85
13.1 Statics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
13.1.1 Force Balance . . . . . . . . . . . . . . . . . . . . . . . 85
13.1.2 Critical Angle . . . . . . . . . . . . . . . . . . . . . . . 87
13.2 Linearized Dynamics . . . . . . . . . . . . . . . . . . . . . . . 89
13.2.1 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . 89
13.2.2 Damped Axial Motion . . . . . . . . . . . . . . . . . . 91
13.3 Cable Strumming . . . . . . . . . . . . . . . . . . . . . . . . . 94
13.4 Vehicle Design . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
iv
15.4.2 Proportional-Derivative Only . . . . . . . . . . . . . . 104
15.4.3 Proportional-Integral-Derivative . . . . . . . . . . . . . 105
15.5 Heuristic Tuning . . . . . . . . . . . . . . . . . . . . . . . . . 105
15.6 Block Diagrams of Systems . . . . . . . . . . . . . . . . . . . . 106
15.6.1 Fundamental Feedback Loop . . . . . . . . . . . . . . . 106
15.6.2 Block Diagrams: General Case . . . . . . . . . . . . . . 106
15.6.3 Primary Transfer Functions . . . . . . . . . . . . . . . 107
v
19 KALMAN FILTER 128
19.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
19.2 Problem Statement . . . . . . . . . . . . . . . . . . . . . . . . 128
19.3 Step 1: An Equation for Σ̇ . . . . . . . . . . . . . . . . . . . . 129
19.4 Step 2: H as a Function of Σ . . . . . . . . . . . . . . . . . . 131
19.5 Properties of the Solution . . . . . . . . . . . . . . . . . . . . 132
19.6 Combination of LQR and KF . . . . . . . . . . . . . . . . . . 133
19.7 Proofs of the Intermediate Results . . . . . . . . . . . . . . . . 134
19.7.1 Proof that E(eT W e) = trace(ΣW ) . . . . . . . . . . . 134
∂
19.7.2 Proof that ∂H trace(−ΛHCΣ) = −ΛT ΣC T . . . . . . . 135
∂
19.7.3 Proof that ∂H trace(−ΛΣC T H T ) = −ΛΣC T . . . . . . 135
19.7.4 Proof of the Separation Principle . . . . . . . . . . . . 136
23 REFERENCES 159
24 PROBLEMS 163
vi
1
1 MATH FACTS
1.1 Vectors
1.1.1 Definition
We use the overhead arrow to denote a column vector, i.e., a number with a
direction. For example, in three-space, we write
2
~a = 1 .
7
1
2
y
x
~a + ~b = ~
c
2
3
5
1 + 3 = 4 .
7
2
9
2
−4
−2 × 1 = −2 .
7 −14
v
um
uX
||~x|| = t x2i ;
i=1
m
~x · ~y = ~xT ~y =
X
xi yi .
i=1
~x · ~y = ||~x||||~y || cos θ,
x2 y3 − x3 y2
~x × ~y = x3 y1 − x1 y3 .
x1 y2 − x2 y1
1.2 Matrices
1.2.1 Definition
A matrix, or array, is equivalent to a set of row vectors, arranged side by side,
say
2 3
~
A = [~a b] = 1 3 .
7 2
This matrix has three rows (m = 3) and two columns (n = 2); a vector is
a special case of a matrix with one column. Matrices, like vectors, permit
pointwise addition and scalar multiplication. We usually use an upper-case
symbol to denote a matrix.
n
X
ci = Ai1 v1 + Ai2 v2 + · · · + Ain vn = Aij vj ,
j=1
1 0 0
I3×3 = 0 1 0
.
0 0 1
Diagonal Matrices . A diagonal matrix is square, and has all zeros off the
diagonal. For instance, the following is a diagonal matrix:
4 0 0
0 −2 0 .
A=
0 0 3
The product of a diagonal matrix with another diagonal matrix is diagonal,
and in this case the operation is commutative.
1.2.5 Transpose
The transpose of a vector or matrix, indicated by a T superscript results
from simply swapping the row-column indices of each entry; it is equivalent to
“flipping” the vector or matrix around the diagonal line. For example,
1
~a = 2 −→ ~a T = {1 2 3}
3
1 2 " #
1 4 8
A = 4 5 −→ AT = .
2 5 9
8 9
A very useful property of the transpose is
(AB)T = B T AT .
1.2.6 Determinant
The determinant of a square matrix A is a scalar equal to the volume of the
parallelepiped enclosed by the constituent vectors. The two-dimensional case
is particularly easy to remember, and illustrates the principle of volume:
y
Area = det(A) = 2
A:2 A:1
x
-1 1
j=n
Akj (−1)k+j ∆kj ,
X
det(A) =
j=1
where ∆kj is the determinant of the sub-matrix formed by neglecting the k’th
row and the j’th column. The formula is symmetric, in the sense that one
could also target the k’th column:
j=n
Ajk (−1)k+j ∆jk .
X
det(A) =
j=1
1.2.7 Inverse
The inverse of a square matrix A, denoted A−1 , satisfies AA−1 = A−1 A = I.
Its computation requires the determinant above, and the following definition
of the n × n adjoint matrix:
T
(−1)1+1 ∆11 · · · (−1)1+n ∆1n
adj(A) =
··· ··· ··· .
n+1 n+n
(−1) ∆n1 · · · (−1) ∆nn .
1.2 Matrices 7
adj(A)
A−1 = .
det(A)
If A is singular, i.e., det(A) = 0, then the inverse does not exist. The inverse
finds common application in solving systems of linear equations such as
A~x = ~b −→ ~x = A−1~b.
1.2.8 Trace
The trace of a matrix is simply the sum of the diagonals:
n
X
tr(A) = Aii .
i=1
A~x = λ~x,
where A is an n × n matrix, ~x is a column vector with n elements, and λ is
a scalar. We ask for what nonzero vectors ~x (right eigenvectors), and scalars
λ (eigenvalues) will the equation be satisfied. Since the above is equivalent to
(A − λI)~x = ~0, it is clear that det(A − λI) = 0. This observation leads to the
solutions for λ; here is an example for the two-dimensional case:
" #
4 −5
A = −→
2 −3
" #
4−λ −5
A − λI = −→
2 −3 − λ
det(A − λI) = (4 − λ)(−3 − λ) + 10
= λ2 − λ − 2
= (λ + 1)(λ − 2).
8 1 MATH FACTS
(A − λ1 I)~x1 = ~0 −→
" #
5 −5
~x1 = ~0 −→
2 −2
n√ √ oT
~x1 = 2/2, 2/2 .
(A − λ2 I)~x2 = ~0 −→
" #
2 −5
~x2 = ~0 −→
2 −5
n √ √ oT
~x2 = 5 29/29, 2 29/29 .
Eigenvectors have arbitrary magnitude and sign; they are often normalized to
have unity magnitude, and positive first element (as above). A set of n eigen-
vectors is always linearly independent. The condition that rank(A − λi I) =
rank(A)−1 indicates that there is only one eigenvector for the eigenvalue λi . If
the left-hand side is less than this, then there are multiple unique eigenvectors
that go with λi .
The above discussion relates only the right eigenvectors, generated from the
equation A~x = λ~x. Left eigenvectors, also useful for many problems, pertain
to the transpose of A: AT ~y = λ~y . A and AT share the same eigenvalues λ,
since they share the same determinant. Example:
(AT − λ1 I)~y1 = ~0 −→
" #
5 2
~y1 = ~0 −→
−5 −2
n √ √ oT
~y1 = 2 29/29, −5 29/29 .
(AT − λ2 I)~y2 = ~0 −→
" #
2 2
~y2 = ~0 −→
−5 −5
n√ √ oT
~y2 = 2/2, − 2/2 .
1.2 Matrices 9
then we have
Y T X = I, or
Y T = X −1 .
λ1 0
Λ=
· ;
0 λn
it follows that
AV = V Λ −→
A = V ΛW T
n
~ iT .
X
= λi~vi w
i=1
G = U ΣV ∗ ,
σ1 0 0 0
0 · 0 0
Σ= ,
0 0 σp 0
0 0 0 0
where p = min(m, n). Each nonzero entry on the diagonal is a real, positive
singular value, ordered such that σ1 > σ2 > · · · σp . The notation is common
that σ1 = σ, the maximum singular value, and σp = σ, the minimum singular
value. The auxiliary matrices U and V are unitary, i.e., they satisfy X ∗ =
X −1 . Like eigenvalues, the singular values of G are related to projections. σi
represents the Euclidean size of the matrix G along the i’th singular vector:
σ = max||x||=1 ||Gx||
σ = min||x||=1 ||Gx||.
• σ(AB) ≤ σ(A)σ(B).
q
• σ(A) = λmax (A∗ A).
q
• σ(A) = λmin (A∗ A).
• σ(A) = 1/σ(A−1 ).
• σ(A) = 1/σ(A−1 ).
1.3 Laplace Transform 11
Z ∞
Y (s) = L(y(t)) = y(τ )e−sτ dτ,
0
The last two properties are of special importance: for control system design,
the differentiation of a signal is equivalent to multiplication of its Laplace
transform by s; integration of a signal is equivalent to division by s. The other
terms that arise will cancel if y(0) = 0, or if y(0) is finite, so they are usually
ignored.
1.3.2 Convergence
We note first that the value of s affects the convergence of the integral. For
instance, if y(t) = et , then the integral converges only for Re(s) > 1, since
the integrand is e1−s in this case. Convergence issues are not a problem for
evaluation of the Laplace transform, however, because of analytic continuation.
This result from complex analysis holds that if two complex functions are
equal on some arc (or line) in the complex plane, then they are equivalent
everywhere. This fact allows us to always pick a value of s for which the
integral above converges, and then by extension infer the existence of the
general transform.
y(t) ←→ Y (s)
(Impulse) δ(t) ←→ 1
1
(Unit Step) 1(t) ←→
s
1
(Unit Ramp) t ←→ 2
s
−αt 1
e ←→
s+α
ω
sin ωt ←→ 2
s + ω2
s
cos ωt ←→ 2
s + ω2
ω
e−αt sin ωt ←→
(s + α)2 + ω 2
s+α
e−αt cos ωt ←→
(s + α)2 + ω 2
1 1
(e−at − e−bt ) ←→
b−a (s + a)(s + b)
1 1 1
1+ (be−at − ae−bt ) ←→
ab a−b s(s + a)(s + b)
ωn −ζωn t
q ωn2
√ e sin ωn 1 − ζ 2 t ←→
1 − ζ2 s2 + 2ζωn s + ωn2
1 ωn2
q
−ζωn t
1− √ e sin ωn 1 − ζ 2t + φ ←→
1 − ζ2 s(s2 + 2ζωn s + ωn2 )
√ !
−1 1 − ζ2
φ = tan
ζ
driven by an input signal x(t); the output is y(t) = g(t)∗x(t). The Convolution
Theorem is
Z t
y(t) = g(t − τ )x(τ )dτ ⇐⇒ Y (s) = G(s)X(s).
0
Z ∞
Y (s) = y(t)e−st dt
0
Z ∞ Z t
= g(t − τ ) x(τ ) dτ e−st dt
0 0
Z ∞ Z ∞
= g(t − τ ) 1(t − τ ) x(τ ) dτ e−st dt
0 0
Z ∞ Z ∞
−st
= x(τ ) g(t − τ ) 1(t − τ ) e dt dτ
Z0∞ 0
= x(τ ) G(s)e−sτ dτ
0
= G(s)X(s)
When g(t) is the impulse response of a dynamic system, then y(t) represents
the output of this system when it is driven by the external signal x(t).
z=z’ z’’
z’’’
y’’’ x’
y x
y’=y’’ x’’=x’’’
A first question is: what is the coordinate of a point fixed in inertial space,
referenced to a rotated body frame? The transformation takes the form of
a 3×3 matrix, which we now derive through successive rotations of the three
2.1 Rotation of Reference Frames 15
Euler angles. Before the first rotation, the body-referenced coordinate matches
that of the inertial frame: ~xb0 = ~x. Now rotate the movable frame yaw axis
(z) through an angle φ. We have
cos φ sin φ 0
~xb1 = − sin φ cos φ 0 ~xb0 = R(φ)~xb0 . (1)
0 0 1
Rotation about the z-axis does not change the z-coordinate of the point; the
other axes are modified according to basic trigonometry. Now apply the second
rotation, pitch about the new y-axis by the angle θ:
cos θ 0 − sin θ
2
~xb =
0 1 0
~xb1 = R(θ)~xb1 . (2)
sin θ 0 cos θ
Finally, rotate the body system an angle ψ about its newest x-axis:
1 0 0
3
~xb = 0 cos ψ sin ψ0
~xb2 = R(ψ)~xb2 . (3)
0 − sin ψ cos ψ
This represents the location of the original point, in the fully-transformed
body-reference frame, i.e., ~xb3 . We will use the notation ~xb instead of ~xb3 from
here on. The three independent rotations can be cascaded through matrix
multiplication (order matters!):
1 δφ −δθ
R ' −δφ
1 δψ (6)
δθ −δψ 1
0 δφ −δθ
= −δφ 0 δψ + I3×3 .
δθ −δψ 0
~xb = ~x − δ E~ × ~x (7)
~ × ~xb .
~x = ~xb + δ E (8)
We now fix the point of interest on the body, instead of in inertial space,
calling its location in the body frame ~r (radius). The differential rotations
occur over a time step δt, so that we can write the location of the point before
and after the rotation, with respect to the first frame as follows:
~x(t) = ~r (9)
T ~ × ~r.
~x(t + δt) = R ~r = ~r + δ E
δ~x δE~
= × ~r (10)
δt δt
= ω × ~r,
~
where the rotation rate vector ω ' dE/dt because the Euler angles for this
infinitesimal rotation are small and decoupled. This same cross-product rela-
tionship can be derived in the second frame as well:
~ × ~r
~xb (t) = R~r = ~r − δ E (11)
~xb (t + δt) = ~r.
such that
δ~xb δE~
= × ~r (12)
δt δt
= ω × ~r,
On a rotating body whose origin point is fixed, the time rate of change of a
constant radius vector is the cross-product of the rotation rate vector ω~ and
the radius vector itself. The resultant derivative is in the moving body frame.
In the case that the radius vector changes with respect to the body frame, we
need an additional term:
d~xb ∂~r
= ω × ~r + . (13)
dt ∂t
Finally, allowing the origin to move as well gives
∂~r
~v = ω × ~r + + ~vo , (15)
∂t
where ~vo is the body-referenced velocity of the origin. The total velocity of the
particle is equal to the velocity of the reference frame origin, plus a component
18 2 KINEMATICS OF MOVING FRAMES
due to rotation of this frame. The velocity equation can be generalized to any
body-referenced vector f~:
df~ ∂f
= ~ × f~.
+ω (16)
dt ∂t
dψ
0
0
dt
dθ
ω
~ = R(ψ)R(θ) 0 + R(ψ) dt
+ 0 (17)
dφ
dt
0
0
dψ
1 0 − sin θ dt
dθ
0
= cos ψ sin ψ cos θ
dt
.
dφ
0 − sin ψ cos ψ cos θ
dt
~ 1 sin ψ tan θ cos ψ tan θ
dE
= 0
cos ψ − sin ψ ω
~ (18)
dt
0 sin ψ/ cos θ cos ψ/ cos θ
~ ω.
= Γ(E)~
ẋ = U cos φ
ẏ = U sin φ.
Needless to say, currents and vehicle sideslip will cause this to be in error.
Nonetheless, some of the most remarkable feats of navigation in history have
depended on dead reckoning.
20 3 VESSEL INERTIAL DYNAMICS
d
F~ = (m~v ) (19)
dt
A rigid body consists of a large number of these small particles, which can be
indexed. The summations we use below can be generalized to integrals quite
easily. We have
~ i = d (mi~vi ) ,
F~i + R (20)
dt
where F~i is the external force acting on the particle and R ~ i is the net force
exerted by all the other surrounding particles (internal forces). Since the
collection of particles is not driven apart by the internal forces, we must have
equal and opposite internal forces such that
N
~ i = 0.
X
R (21)
i=1
z, w, φ
x, u, ψ
y, v, θ
N N
d
F~i =
X X
~ × ~ri ))
(mi (~vo + ω (23)
i=1 i=1 dt
N
" #
∂~vo d X
= m + ~×
ω mi~ri ,
∂t dt i=1
!
∂~vo d~ω
F~ =
X
N =m ~ × ~vo +
+ω × ~rG + ω
~ × (~ω × ~rG ) . (26)
i=1 ∂t dt
Now we list some conventions that will be used from here on:
" #
∂u dq dr
X = m + qw − rv + zG − yG + (qyG + rzG )p − (q 2 + r2 )xG (27)
∂t dt dt
" #
∂v dr dp 2 2
Y = m + ru − pw + xG − zG + (rzG + pxG )q − (r + p )yG
∂t dt dt
" #
∂w dp dq 2 2
Z = m + pv − qu + yG − xG + (pxG + qyG )r − (p + q )zG .
∂t dt dt
Note that about half of the terms here are due to the mass center being in a
different location than the reference frame origin, i.e., ~rG 6= ~0.
!
∂~vo d~ω
F~ = m ~ × ~vo +
+ω × ~rG + ω
~ × (~ω × ~rG ) .
∂t dt
3.3 Example: Mass on a String 23
The force of the string pulls in on the mass to create the circular motion.
N N N
" # !
X
~ i + ~ri × F~i ) =
X ∂~vo X ∂~ω
(M mi~ri × ~ × ~vo +
+ω mi~ri × × ~ri +
i=1 i=1 ∂t i=1 ∂t
N
X
mi~ri × (~ω × (~ω × ~ri )).
i=1
The summation in the first term of the right-hand side is recognized simply as
m~rG , and the first term becomes
" #
∂~vo
m~rG × +ω~ × ~vo . (29)
∂t
The second term expands as (using the triple product)
N N
! ! !
X ∂~ω X ∂~ω ∂~ω
mi~ri × × ~ri = mi (~ri · ~ri ) − · ~ri ~ri (30)
i=1 ∂t i=1 ∂t ∂t
PN 2 2
Pi=1 mi ((yi + zi )ṗ − (yi q̇ + zi ṙ)xi )
= N 2 2
i=1 mi ((xi + zi )q̇ − (xi ṗ + zi ṙ)yi ) .
PN 2 2
i=1 mi ((xi + yi )ṙ − (xi ṗ + yi q̇)zi )
3.4 Angular Momentum 25
Ixx Ixy Ixz
Iyx Iyy Iyz
I = (inertia matrix)
the second term of the angular momentum right-hand side collapses neatly
into I∂~ω /∂t. The third term can be worked out along the same lines, but
offers no similar condensation:
N
X N
X
mi~ri × ((~ω · ~ri )~ω − (~ω · ω
~ )~ri ) = mi~ri × ω
~ (~ω · ~ri ) (31)
i=1 i=1
PN
P i=1 mi (yi r − zi q)(xi p + yi q + zi r)
N
= i=1 mi (zi p − xi r)(xi p + yi q + zi r)
N
mi (xi q − yi p)(xi p + yi q + zi r)
P
i=1
2 2
Iyz (q − r ) + Ixz pq − Ixy pr
= I (r2 − p2 ) + I rq − I pq
xz xy yz +
Ixy (p2 − q 2 ) + Iyz pr − Ixz qr
(Izz − Iyy )rq
(Ixx − Izz )rp .
(Iyy − Ixx )qp
26 3 VESSEL INERTIAL DYNAMICS
Letting M~ = {K, M, N } be the total moment acting on the body, i.e., the left
side of Equation 28, the complete moment equations are
dp
Ixx + (Izz − Iyy )rq = 0
dt
dq
Iyy + (Ixx − Izz )pr = 0
dt
dr
Izz + (Iyy − Ixx )qp = 0.
dt
We consider in turn the stability of rotations about each of the main axes,
with constant angular rate Ω. The interesting result is that rotations about
the x and z axes are unstable, while rotation about the y axis is not.
3.5.1 x-axis
In the case of the x-axis, p = Ω + δp, q = δq, and r = δr, where the δ prefix
indicates a small value compared to Ω. The first equation above is uncoupled
3.5 Example: Spinning Book 27
from the others, and indicates no change in δp, since the small term δqδr can
be ignored. Differentiate the second equation to obtain
∂ 2 δq ∂δr
Iyy 2
+ (Ixx − Izz )Ω =0
∂t ∂t
Substitution of this result into the third equation yields
∂ 2 δq
Iyy Izz + (Ixx − Izz )(Ixx − Iyy )Ω2 δq = 0.
∂t2
√
A simpler expression is δ q̈ + αδq = 0, which has response δq(t) = δq(0)e −αt ,
when δ q̇(0) = 0. For spin about the x-axis, both coefficients of the differential
equation are positive, and hence α > 0. The imaginary
√ exponent indicates
that the solution is of the form δq(t) = δq(0)cos αt, that is, it oscillates but
does not grow. Since the perturbation δr is coupled, it too oscillates.
3.5.2 y-axis
Now suppose q = Ω + δq: differentiate the first equation and substitute into
the third equation to obtain
∂ 2 δp
Izz Ixx + (Iyy − Ixx )(Iyy − Izz )Ω2 δp = 0.
∂t2
Here the second coefficient has negative sign, and therefore α < 0. The ex-
ponent√ is real now, and the solution grows without bound, following δp(t) =
δp(0)e −αt .
3.5.3 z-axis
Finally, let r = Ω + δr: differentiate the first equation and substitute into the
second equation to obtain
∂ 2 δp
Iyy Ixx + (Ixx − Izz )(Iyy − Izz )Ω2 δp = 0.
∂t2
The coefficients are positive, so bounded oscillations occur.
28 3 VESSEL INERTIAL DYNAMICS
where I¯ represents an MMOI in the axes of the mass center, and δx, for ex-
ample, is the translation of the x-axis to the new frame. Note that translation
of MMOI using the parallel axis theorem must be either to or from a frame
resting exactly at the center of gravity.
• ω
~ , body rotation rate vector.
The derivatives of body-referenced velocity and rotation rate come from Equa-
tions 27 and 32, with some coupling which generally requires a 6 × 6 matrix
inverse. The Cartesian position propagates according to
~x˙ = RT (E)~
~ vo , (34)
while the Euler angles follow:
~˙ = Γ(E)~
E ~ ω. (35)
30 4 HYDRODYNAMICS: INTRODUCTION
4 HYDRODYNAMICS: INTRODUCTION
The forces and moments on a vessel are complicated functions of many fac-
tors, including water density, viscosity, surface tension, pressure, vapor pres-
sure, and motions of the body. The most important factors for large ocean
vehicles are density and motion, and we can make simplifications to param-
eterize the most prominent relationships. This section pertains to the use of
hydrodynamic coefficients for predicting hydrodynamic response.
∂f (xo ) 1 ∂ 2 f (xo )
f (x) = f (xo ) + (x − xo ) + (x − xo )2 + · · · . (36)
∂x 2! ∂x2
We introduce the notation
∂f (xo )
fx =
∂x
1 ∂ 2 f (xo )
fx x = ,
2! ∂x2
and so on, so that a two-variable Taylor expansion would have the form
Note that all of the factorials are included in the coefficients. This notation
covers some instances where the formal Taylor series is meaningless, but the
notation is still clear. As one example, fluid drag is often written as
1
F = ρCd Au|u| = Fu|u| u|u|.
2
4.2 Surface Vessel Linear Model 31
!
∂u
X = m − rv − xG r2 (39)
∂t
!
∂v ∂r
Y = m + ru + xG
∂t ∂t
!
∂r ∂v
N = Izz + mxG + ru .
∂t ∂t
∂u
X = m (40)
∂t !
∂v ∂r
Y = m + rU + xG
∂t ∂t
!
∂r ∂v
N = Izz + mxG + rU .
∂t ∂t
terms, or higher time derivatives. It should be noted that some nonlinear terms
related to those we have eliminated above are not zero. For instance, Yuu = 0
because of hull symmetry, but in general Xvv = 0 only if the vessel is bow-stern
symmetric.
We have so far, considering only the linear hydrodynamic terms,
The right side here carries also the imposed forces from a thruster(s) and
rudder(s) {X 0 , Y 0 , N 0 }. Note that the surge equation is decoupled from the
sway and yaw, but that sway and yaw themselves are coupled, and there-
fore are of immediate interest. With the state vector ~s = {v, r} and exter-
nal force/moment vector F~ = {Y 0 , N 0 }, a state-space representation of the
sway/yaw system is
" # " #
m − Yv̇ mxG − Yṙ d~s Yv Yr − mU
= ~s + F~ , or (44)
mxG − Nv̇ Izz − Nṙ dt Nv Nr − mxG U
M ~s˙ = P~s + F~
~s˙ = M −1 P~s + M −1 F~
~s˙ = A~s + B F~ . (45)
s̈1 + (−A11 − A22 )ṡ1 + (A11 A22 − A12 A21 )s1 = 0. (47)
Note that these operations are allowed because the derivative operator is linear;
in the language of the Laplace transform, we would simply use s. A necessary
and sufficient condition for stability of this ODE system is that each coefficient
must be greater than zero:
The denominators are identical, and can be simplified. First, let xG ' 0;
valid for many vessels with the origin is at the geometric center. If the origin
is at the center of mass, xG = 0. Next, if the vessel is reasonably balanced
with regard to forward and aft areas with respect to the origin, the terms
{Nv̇ , Yṙ , Nv , Yr } take very small values in comparison with the others. To wit,
the added mass term −Yv̇ is of the order of the vessel’s material mass m,
and similarly Nṙ ' −Izz . Both Yv̇ and Nṙ take large negative values. Linear
drag and rotational drag are significant also; these are the terms Yv and Nr ,
34 4 HYDRODYNAMICS: INTRODUCTION
both large and negative. The denominator for A’s components reduces to
(m − Yv̇ )(Izz − Nṙ ), and
Yv
A11 = m−Yv̇
< 0
Nr
A22 = Izz −Nṙ
< 0.
Hence the first condition for stability is met: −A11 − A22 > 0. For the second
condition, since the denominators of the Aij are identical, we have only to look
at the numerators. For stability, we require
The first term is the product of two large negative and two large positive
numbers. The second part of the second term contains mU , which has a large
positive value, generally making stability critical on the (usually negative) Nv .
When only the largest terms are considered for a vessel, a simpler form is
common:
δ x
r
Figure 3: Convention for positive rudder angle in the vessel reference system.
lift coefficient Cl is normally linear with α near α = 0, but the rudder stalls
when the angle of attack reaches a critical value, and thereafter develops much
less lift. We will assume that α is small enough that the linear relationship
applies:
∂Cl
Cl (α) = α. (53)
∂α α=0
Since the rudder develops force (and a small moment) far away from the body
origin, say a distance l aft, the moment equation is quite simple. We have
1 ∂Cl
Yα = ρA U2 (54)
2 ∂α α=0
1 ∂Cl
Nα = − ρA lU 2 . (55)
2 ∂α α=0
Note the difference between the rudder angle expressed in the body frame, δ,
and the total angle of attack α. Angle of attack is influenced by δ, as well
as v/U and lr. Thus, in tank testing with v = 0, δ = α and Nδ = Nα ,
etc., but in real conditions, other hydrodynamic derivatives are augmented to
capture the necessary effects, for example Nv and Nr . Generally speaking, the
hydrodynamic characteristics of the vessel depend strongly on the rudder, even
when δ = 0. In this case the rudder still opposes yaw and sway perturbations
and acts to stabilize the vessel.
A positive rudder deflection (defined to have the same sense as the yaw angle)
causes a negative yaw perturbation, and a very small positive sway perturba-
tion.
36 4 HYDRODYNAMICS: INTRODUCTION
Since all coefficients are positive (recall Nδ < 0), the equation gives a stable θ
response, settling under second-order dynamics to θ(∞) = 0. The control law
δ = kφ (φ − φdesired ) + kr r is the basis for heading autopilots, which are used
to track φdesired . This use of an error signal to drive an actuator is in fact the
essence of feedback control. In this case, we require sensors to obtain r and φ,
a controller to calculate δ, and an actuator to implement the corrective action.
4.5 Response of the Vessel to Step Rudder Input 37
The radius goes up directly with C, indicating that too stable a ship has poor
turning performance. We see also that increasing the rudder area increases
Nδ , decreasing R as desired. Increasing the deflection δ to reduce R works
only to the point of stalling.
!
∂w
m − Uq = Zẇ ẇ + Zw w + Zq̇ q̇ + Zq q + (B − W ) (65)
∂t
dq
Iyy = Mẇ ẇ + Mw w + Mq̇ q̇ + Mq q − Blb sin θ. (66)
dt
The last term in each equation is a hydrostatic effect induced by opposing net
buoyancy B and weight W . lb denotes the vertical separation of the center
of gravity and the center of buoyancy, creating the so-called righting moment
which nearly all underwater vehicles possess. Because buoyancy effects do not
4.7 Stability in the Vertical Plane 39
change with speed, the dynamic properties and hence stability of the vehicle
may change with speed.
40 5 SIMILITUDE
5 SIMILITUDE
5.1 Use of Nondimensional Groups
For a consistent description of physical processes, we require that all terms
in an equation must have the same units. On the basis of physical laws,
some quantities are dependent on other, independent quantities. We form
nondimensional groups out of the dimensional ones in this section, and apply
the technique to maneuvering.
The Buckingham π-theorem provides a basis for all nondimensionalization.
Let a quantity Qn be given as a function of a set of n − 1 other quantities:
[X(·)] = m
[m] = kg
[F ] = kgm/s2
[T ] = s,
Q = fQ (h, d, ρ, µ, g),
where the water density is ρ, and its absolute viscosity µ; g is the acceleration
due to gravity. No other parameters affect the flow rate. We have n = 6, and
the (MKS) units of these quantities are:
[Q] = m3 /s
[h] = m
[d] = m
[ρ] = kg/m3
[µ] = kg/ms
[g] = m/s2
There are only three units that appear: [length, time, mass], and thus k = 3.
Hence, only three non-dimensional groups exist, and one is a unique func-
tion of the other two. To arrive at a set of valid groups, we must create
three nondimensional quantities, making sure that each of the original (di-
mensional) quantities is represented. Intuition and additional manipulations
come in handy, as we now show. √
Three plausible first groups are: π1 = ρQ/dµ, π2 = dρ gh/µ, and π3 = h/d.
Note that all six quantities appear at least once. Since h and d have the same
units, they could easily change places in the first two groups. However, π1 is
recognized as a Reynolds number pertaining to the orifice flow. π2 is more
awkward, but products and fractions of groups √ are themselves valid groups,
2
and we may construct π4 = π1 /π2 = Q/d gh √ to nondimensionalize Q with
a pressure velocity, and then π5 = π1 /π4 = ρd gh/µ to establish an orifice
Reynolds number independent of Q. We finally have the useful result
π4 = fπ (π5 , π2 ) −→
√ !
Q ρd gh h
√ = fπ , .
d2 gh µ d
The uncertainty about where to use h and d, and the questionable importance
of h/d as a group are remnants of the theorem. Intuition is that h/d is im-
42 5 SIMILITUDE
material, and the other two terms have a nice physical meaning, e.g., π5 is a
Reynolds number.
The power of the π-theorem is primarily in reducing the number of parameters
which must be considered independently to characterize a process. In the flow
example, the theorem reduced the number of independent parameters from
five to two, with no constraints about the actual physics taking place.
1. Froude number:
U
Fr = √ , (69)
gL
2. Cavitation number:
P∞ − P v
δ= 1 , (70)
2
ρU 2
3. Reynolds number:
Ul
Re = , (71)
µ/ρ
5.2 Common Groups in Marine Engineering 43
ρU 2 l
W = , (72)
σ
where σ is the surface tension of a fluid. Given that [σ] = N/m (MKS),
ρU 2 normalizes pressure, and l normalizes length. The Weber number
indicates the importance of surface tension.
To appreciate the origins of these terms from a fluid particle’s point of view,
consider a box having side lengths [dx, dy, dz]. Various forces on the box scale
as
∂v ∂v
(inertia) Fi = ρ dxdydz + ρv dxdydz ' ρU 2 l2
∂t ∂x
3
(gravity) Fg = ρgdxdydz ' ρgl
(pressure) Fp = P dxdy ' P l2
∂v
(shear) Fs = µ dxdy ' µU l
∂z
(surface tension) Fσ = σdx ' σl.
Thus the groups listed above can be written as
Fi U2
Fr = '
Fg gl
Fp P
δ= '
Fi ρU 2
Fi ρU l
Re = '
Fs µ
Fi ρU 2 l
W = '
Fσ σ
44 5 SIMILITUDE
D
Cr = 1 = fQ (ρ, µ, g, σ, U, l)
2
ρAU 2
= fπ (Re, F r, W ).
First, since l is large, W is very large, and hence
√ surface tension plays no role.
Next, we look at Re = U l/ν and F r = U/ gl, both of which are important
for surface vessels. Suppose that lship = λlmodel , so that usually λ >> 1;
additionally, we set gmodel = gship , i.e., the model and the true vessel operate
in the same gravity field. √
Froude number similitude requires Umodel = Uship / λ. Then Reynolds number
scaling implies directly νmodel = νship /λ3/2 . Unfortunately, few fluids with this
property are workable in a large testing tank. As a result, accurate scaling of
Re for large vessels to model scale is quite difficult.
For surface vessels, and submarines near the surface, it is a routine procedure
to employ turbulence stimulators to achieve flow that would normally occur
with ship-scale Re. Above a critical value Re ' 500, 000, Cf is not sensitive
to Re. With this achieved, one then tries to match F r closely.
time, and mass, and furthermore are readily accessible to the user. First, we
create nondimensional states, denoted with a prime symbol:
L
v̇ 0 = v̇ (73)
U2
1
v0 = v
U
L2
ṙ0 = ṙ
U2
L
r0 = r.
U
We follow a similar procedure for the constant terms as follows, including a
factor of 1/2 with ρ, for consistency with our previous expressions:
m
m0 = 1 (74)
2
ρL3
0 Izz
Izz = 1
2
ρL5
xG
x0G =
L
0 U
U = =1
U
Yv̇
Yv̇0 = 1
2
ρL3
Yv
Yv0 = 1
2
ρU L2
Yṙ
Yṙ0 = 1
2
ρL4
Yr
Yr0 = 1
2
ρU L3
Y
Y0 = 1
2
ρU 2 L2
Nv̇
Nv̇0 = 1
2
ρL4
Nv
Nv0 = 1
2
ρU L3
46 5 SIMILITUDE
Nṙ
Nṙ0 = 1
2
ρL5
Nr
Nr0 = 1
2
ρU L4
N
N0 = 1 .
2
ρU 2 L3
Note that every force has been normalized with 12 ρU 2 L2 , and every moment
with 21 ρU 2 L3 ; time has been also nondimensionalized with L/U . Thus we
arrive at a completely equivalent set of nondimensional system equations,
(m0 − Yv̇0 )v̇ 0 − Yv0 v 0 + (m0 U 0 − Yr0 )r0 + (m0 x0G − Yṙ0 )ṙ0 = Y 0 (75)
0
(Izz − Nṙ0 )ṙ0 + (m0 x0G U 0 − N 0 r0 )r0 + (m0 x0G − Nv̇0 )v̇ 0 − Nv0 v 0 = N 0 .
Since fluid forces and moments generally scale with U 2 , the nondimensionalized
description holds for a range of velocities.
0 Ixx
Ixx = 1 (77)
2
ρL5
Kṗ
Kṗ0 = 1
2
ρL5
Kp
Kp0 = 1
2
ρU L4
Kψ
Kψ0 = 1
2
ρU 2 L3
5.4 Roll Equation Similitude 47
0 L2
ṗ = ṗ
U2
L
p0 = p,
U
leading to the equivalent system
2gL
0
(Ixx − Kṗ0 )ṗ0
− Kp0 p0
− ∇0 (GM 0 )φ = K 0 . (78)
U2
Note that the roll angle φ was not nondimensionalized. The Froude number
has a very strong influence on roll stability, since it appears explicitly in the
nondimensional righting moment term, and also has a strong influence on Kp0 .
In the case of a submarine, the righting moment has the form Kψ = −Bh,
where B is the buoyant force, and h is the righting arm. The nondimensional
coefficient becomes
Bh
Kψ0 = − 1 .
2
ρU 2 L3
Kψ0 again depends strongly on U , since B and h are fixed; this Kψ0 needs to be
maintained in model tests.
48 6 CAPTIVE MEASUREMENTS
6 CAPTIVE MEASUREMENTS
Before making the decision to measure hydrodynamic derivatives, a prelimi-
nary search of the literature may turn up useful estimates. For example, test
results for many hull-forms have already been published, and the basic lift-
ing surface models are not difficult. The available computational approaches
should be considered as well; these are very good for predicting added mass
in particular. Finally, modern sensors and computer control systems make
possible the estimation of certain coefficients based on open-water tests of a
model or a full-scale design.
In model tests, the Froude number F r = √UgL , which scales the influence
of surface waves, must be maintained between model and full-scale surface
vessels. Reynolds number Re = UνL , which scales the effect of viscosity, need
not be matched as long as the scale model attains turbulent flow (supercritical
Re). One can use turbulence stimulators near the bow if necessary. Since the
control surface(s) and propeller(s) affect the coefficients, they should both be
implemented in model testing.
6.1 Towtank
In a towtank, tow the vehicle at different angles of attack, measuring sway
force and yaw moment. The slope of the curve at zero angle determines Yv
and Nv respectively; higher-order terms can be generated if the points deviate
from a straight line. Rudder derivatives can be computed also by towing with
various rudder angles.
and the transverse forces on the posts are measured and approximated as
We have v = (ẏa + ẏa )/2 and r = (ẏb − ẏa )/2l. When a = b, these become
aω
v = − (sin ωt(1 + cos ψ) + cos ωt sin ψ) (82)
2
aω 2
v̇ = − (cos ωt(1 + cos ψ) − sinωt sin ψ)
2
aω
r = − (sin ωt(cos ψ − 1) + cos ωt sin ψ)
2l
aω 2
ṙ = − (cos ωt(cos ψ − 1) − sin ωt sin ψ) .
2l
Equating the sine terms and then the cosine terms, we obtain four independent
equations:
50 6 CAPTIVE MEASUREMENTS
aω 2
!
(m − Yv̇ ) − (1 + cos ψ) − (83)
2
aω
Yv − sin ψ +
2
aω
(mU − Yr ) − sin ψ +
2l
aω 2
!
(mxG − Yṙ ) − (cos ψ − 1) = Fa cos θa + Fb cos θb
2l
aω 2
!
(m − Yv̇ ) − (− sin ψ) − (84)
2
aω
Yv − (1 + cos ψ) +
2
aω
(mU − Yr ) − (cos ψ − 1) +
2l
aω 2
!
(mxG − Yṙ ) − (− sin ψ) = −Fa sin θa − Fb sin θb
2l
aω 2
!
(Izz − Nṙ ) − (cos ψ − 1) + (85)
2l
aω
(mxG U − Nr ) − sin ψ +
2l
aω 2
!
(mxG − Nv̇ ) − (1 + cos ψ) −
2
aω
Nv − sin ψ = l(Fb cos θb − Fa cos θa )
2
aω 2
!
(Izz − Nṙ ) − (− sin ψ) + (86)
2l
aω
(mxG U − Nr ) − (cos ψ − 1) +
2l !
aω 2
(mxG − Nv̇ ) − (− sin ψ) −
2
aω
Nv − (1 + cos ψ) = l(−Fb sin θb + Fa sin θa )
2
6.3 Planar-Motion Mechanism 51
In this set of four equations, we know from the imposed motion the values
[U, ψ, a, ω]. From the experiment, we obtain [Fa , Fb , θa , θb ], and from the rigid-
body model we have [m, Izz , xG ]. It turns out that the two cases of ψ = 0
(pure sway motion) and ψ = 180o (pure yaw motion) yield a total of eight
independent equations, exactly what is required to find the eight coefficients
[Yv̇ , Yv , Yṙ , Yr , Nv̇ , Nv , Nṙ , Nr ]. Remarkably, we can write the eight solutions
directly: For ψ = 0,
aω 2
!
(m − Yv̇ ) − (2) = Fa cos θa + Fb cos θb (87)
2
aω
−Yv − (2) = −Fa sin θa − Fb sin θb
2!
aω 2
(mxG − Nv̇ ) − (2) = l(Fb cos θb − Fa cos θa )
2
aω
−Nv − (2) = l(−Fb sin θb + Fa sin θa ),
2
to be solved respectively for [Yv̇ , Yv , Nv̇ , Nv ]. For ψ = 180o , we have
aω 2
!
(mxG − Yṙ ) − (−2) = Fa cos θa + Fb cos θb (88)
2l
aω
(mU − Yr ) − (−2) = −Fa sin θa − Fb sin θb
2l !
aω 2
(Izz − Nṙ ) − (−2) = l(Fb cos θb − Fa cos θa )
2l
aω
(mxG U − Nr ) − (−2) = l(−Fb sin θb + Fa sin θa ),
2l
to be solved for [Yṙ , Yr , Nṙ , Nr ]. Thus, the eight linear coefficients for a surface
vessel maneuvering, for a given speed, can be deduced from two tests with
a planar motion mechanism. We note that the nonlinear terms will play a
significant role if the motions are too large, and that some curve fitting will be
needed in any event. The PMM can be driven with more complex trajectories
which will target specific nonlinear terms.
52 7 STANDARD MANEUVERING TESTS
The Dieudonné maneuver has the vessel path following a growing spiral, and
then a contracting spiral in the opposite direction. The test reveals if the vessel
has a memory effect, manifested as a hysteresis in yaw rate r. For example,
suppose that the first 15◦ rudder deflection causes the vessel to turn right, but
that the yaw rate at zero rudder, on the way negative, is still to the right. The
vessel has gotten “stuck” here, and will require a negative rudder action to
pull out of the turn. But if the corrective action causes the vessel to turn left
at all, the same memory effect may occur. It is easy to see that the rudder in
this case has to be used excessively driving the vessel back and forth. We say
that the vessel is unstable, and clearly a poor design.
the turn, both initially and in the steady state. The heel angle declines as the
speed of the submarine drops.
56 8 STREAMLINED BODIES
8 STREAMLINED BODIES
8.1 Nominal Drag Force
A symmetric streamlined body at zero angle of attack experiences only a drag
force, which has the form
1
FA = − ρCA Ao U 2 . (93)
2
The drag coefficient CA has both pressure and skin friction components, and
hence area Ao is usually that of the wetted surface. Note that the A-subscript
will be used to denote zero angle of attack conditions; also, the sign of FA is
negative, because it opposes the vehicle’s x-axis.
1
Mm = − (Azz − Axx )U 2 sin 2α (94)
2
' −(Azz − Axx )U 2 α.
Azz > Axx for a slender body, and the negative sign indicates a negative
pitch with respect to the vehicle’s pitch axis. The added mass terms Azz and
Axx can be estimated from analytical expressions (available only for regular
shapes such as ellipsoids), from numerical calculation, or from slender body
approximation (to follow).
8.3 Separation Moment 57
1
Fn = ρCn Ao U 2 (95)
2
1 ∂Cn
' ρ αAo U 2 .
2 ∂α
With a positive angle of attack, this force is in the positive z-direction. The
zero-α drag force FA is modified by the vortex shedding:
1
Fa = − ρCa Ao U 2 , where (96)
2
Ca = CA cos2 φ.
The last relation is based on writing CA (U cos φ)2 as (CA cos2 φ)U 2 , i.e., a
decomposition using apparent velocity.
Fa = −γa
Fn = γn α
Mm = −γm α.
Each constant γ is taken as positive, and the signs reflect orientation in the
vehicle reference frame, with a nose-up angle of attack. The Munk moment is a
58 8 STREAMLINED BODIES
pure couple which does not depend on a reference point. We pick a temporary
origin O for Fn however, and write the total pitch moment about O as:
M = Mm + Fn ln (97)
= (−γm + γn ln )α.
where ln denotes the (positive) distance between O and the application point
of Fn . The net moment about O is zero if we select
γm
ln = , (98)
γn
and the location of O is then called the aerodynamic center or AC.
The point AC has an intuitive explanation: it is the location on the hull where
Fn would act to create the total moment. Hence, if the vehicle’s origin lies in
front of AC, the net moment is stabilizing. If the origin lies behind AC, the
moment is destabilizing. For self-propelled vehicles, the mass center must be
forward of AC for stability. Similarly, for towed vehicles, the towpoint must
be located forward of AC. In many cases with very streamlined bodies, the
aerodynamic center is significantly ahead of the nose, and in this case, a rigid
sting would have to extend at least to AC in order for stable towing. As a
final note, since the Munk moment persists even in inviscid flow, AC moves
infinitely far forward as viscosity effects diminish.
1
L = ρAf U 2 Cl (α) ' γL α (99)
2
1
D = − ρAf U 2 Cd ' −γD (constant)
2
These forces act somewhere on the fin, and are signed again to match the
vehicle frame, with γ > 0 and α > 0. The summed forces on the body are
thus:
8.5 Role of Fins in Moving the Aerodynamic Center 59
γm = γn ln + lf (γL + γD ). (102)
The Munk moment γm opposes the aggregate effects of vorticity lift γn and
the fins’ lift and drag γL + γD . With very large fins, this latter term is large, so
that lf might be very small; this is the case of AC moving aft toward the fins.
A vehicle with excessively large fins will be difficult to turn and maneuver.
Equation 102 contains two length measurements, referenced to an arbitrary
body point O. To solve it explicitly, let lf n denote the (positive) distance that
the fins are located behind Fn ; this is likely a small number, since both effects
usually act near the stern. We solve for lf :
γm + γn lf n
lf = . (103)
γn + γL + γD
This is the distance that AC is located forward of the fins, and thus AC can
be referenced to any other fixed point easily. Without fins, it can be recalled
that
γm
ln = .
γn
Hence, the fins act directly in the denominator to shorten lf . Note that if
the fins are located forward of the vortex shedding force Fn , i.e., lf n < 0, lf is
reduced, but since AC is referenced to the fins, there is no net gain in stability.
60 8 STREAMLINED BODIES
1
X = F̂a ' − ρĈa Âo U 2 (104)
2
1 0
Z = F̂n ' ρĈn Âo U 2 α
2
1
M = −F̂n xAC ' − Ĉn0 Âo U 2 xAC α.
2
w
−1
α = − tan (105)
u
w
' − for U >> w.
U
We can then write several linear hydrodynamic coefficients easily:
1
Zw = − ρĈn0 Âo U (106)
2
1 0
Mw = ρĈ Âo U xAC .
2 n
The rotation of the vessel involves complex flow, which depends on both w and
q, as well as their derivatives. To start, we consider the contribution of the fins
only – slender body theory, introduced shortly, provides good results for the
hull. The fin center of pressure is located a distance lf aft of the body origin,
and we assume that the vehicle is moving horizontally, with an instantaneous
pitch angle of θ. The angle of attack seen by the fin is a combination of a part
due to θ and a part linear with q:
lf q
αf ' −θ + (107)
U
8.7 Coefficients Zw , Mw , Zq , and Mq for a Slender Body 61
and so lateral force and moment derivatives (for the fin alone) emerge as
1
Zq = − ρCl0 Af U lf (108)
2
1
Mq = − ρCl0 Af U lf2 .
2
62 9 SLENDER-BODY THEORY
9 SLENDER-BODY THEORY
9.1 Introduction
Consider a slender body with d << L, that is mostly straight. The body
could be asymmetric in cross-section, or even flexible, but we require that
the lateral variations are small and smooth along the length. The idea of
the slender-body theory, under these assumptions, is to think of the body as
a longitudinal stack of thin sections, each having an easily-computed added
mass. The effects are integrated along the length to approximate lift force and
moment. Slender-body theory is accurate for small ratios d/L, except near
the ends of the body.
As one example, if the diameter of a body of revolution is d(s), then we can
compute δma (x), where the nominal added mass value for a cylinder is
π
δma = ρ d2 δx. (109)
4
The added mass is equal to the mass of the water displaced by the cylinder.
The equation above turns out to be a good approximation for a number of
two-dimensional shapes, including flat plates and ellipses, if d is taken as the
width dimension presented to the flow. Many formulas for added mass of two-
dimensional sections, as well as for simple three-dimensional bodies, can be
found in the books by Newman and Blevins.
∂zb
wn = cos α − U sin α. (110)
∂t
The first component is the time derivative in the body frame, and the second
due to the deflection of the particle by the inclined body. If the body reference
frame is rotated to the flow, that is, if w 6= 0, then ∂zb /∂t will contain w. For
small angles, sin α ' tan α = ∂zb /∂x, and we can write
∂zb ∂zb
wn ' −U .
∂t ∂x
!
dzb ∂ ∂
wn = = −U zb .
dt ∂t ∂x
d 1
[µ(x, t)] = lim [µ(x + δx, t + δt) − µ(x, t)]
dt δt→0 δt
" #
∂µ ∂µ
= −U .
∂t ∂x
The second equality can be verified using a Taylor series expansion of µ(x +
δx, t + δt):
∂µ ∂µ
µ(x + δx, t + δt) = µ(x, t) + δt + δx + h.o.t.,
∂t ∂x
and noting that δx = −U δt. The fluid is convected downstream with respect
to the body.
64 9 SLENDER-BODY THEORY
d
δF = − [ma (x, t)wn (x, t)] δx. (111)
dt
Note that we could here let the added mass vary with time also – this is the
case of a changing cross-section! The lateral velocity of the point zb (x) in the
body-reference frame is
∂zb
= w − xq, (112)
∂t
such that
wn = w − xq − U α. (113)
Taking the derivative, we have
!
δF ∂ ∂
=− −U [ma (x, t)(w(t) − xq(t) − U α(x, t))] .
δx ∂t ∂x
We now restrict ourselves to a rigid body, so that neither ma nor α may change
with time.
δF ∂
= ma (x)(−ẇ + xq̇) + U [ma (x)(w − xq − U α)] . (114)
δx ∂x
Z xN
Z = δF dx (115)
xT
where xT represents the coordinate of the tail, and xN is the coordinate of the
nose. Expanding, we have
9.6 Total Moment on a Vessel 65
Z xN ∂ Z xN
Z = ma (x) [−ẇ + xq̇] dx + U [ma (x)(w − xq − U α)] dx
xT xT ∂x
= −m33 ẇ − m35 q̇ + U ma (x)(w − xq − U α)|x=x
x=xT .
N
Z xN
m33 = ma (x)dx
xT
Z xN
m35 = − xma (x)dx.
xT
Additionally, for vessels with pointed noses and flat tails, the added mass ma
at the nose is zero, so that a simpler form occurs:
Zẇ = −m33
Zq̇ = −m35
Zw = −U ma (xT )
Zq = U xT ma (xT )
Zα(xT ) = U 2 ma (xT ).
It is interesting to note that both Zw and Zα(xT ) depend on a nonzero base
area. In general, however, potential flow estimates do not create lift (or drag)
forces for a smooth body, so this should come as no surprise. The two terms are
clearly related, since their difference depends only on how the body coordinate
system is oriented to the flow. Another noteworthy fact is that the lift force
depends only on α at the tail; α could take any value(s) along the body, with
no effect on Z.
Z xN
M = − xδF dx
xT
!
Z xN∂ ∂
= x −U [ma (x)(w − xq − U α)] dx
xT ∂t ∂x
Z xN Z xN
∂
= xma (x)(ẇ − xq̇)dx − U x [ma (x)(w − xq + U α)] dx.
xT xT ∂x
Then we make the further definition
Z xN
m55 = x2 ma (x)dx,
xT
Z xN
U ma (x)(w − xq − U α)dx.
xT
The integral above contains the product ma (x)α(x), which must be calculated
if α changes along the length. For simplicity, we now assume that α is in fact
constant on the length, leading to
Mẇ = −m35
Mq̇ = −m55
Mw = U xT ma (xT ) + U m33
Mq = −U x2T ma (xT ) + U m35
Mα = −U 2 xT ma (xT ) − U 2 m33 .
The derivative Mw is closely-related to the Munk moment, whose linearization
would provide Mw = (m33 − m11 )U . The Munk moment (an exact result) may
therefore be used to make a correction to the second term in the slender-body
approximation above of Mw . As with the lift force, Mw and Mα are closely
related, depending only on the orientation of the body frame to the flow.
9.7 Relation to Wing Lift 67
Z = −ma (xT )U w
π
= − ρs2 U w.
4
Slender-body theory is thus able to recover exactly the lift of a low-aspect
ratio wing. Where does the slender-body predict the force will act? Recalling
that Mw = U m33 + U xT ma (xT ), and since m33 = 0 for a front-back symmetric
wing, the estimated lift force acts at the trailing edge. This location will tend
to stabilize the wing, in the sense that it acts to orient the wing parallel to
the incoming flow.
FN
CN = 1 (119)
2
ρU 2 Ar
FA
CA = 1
2
ρU 2 Ar
10.2 Jorgensen’s Formulas 69
M xm
CM = 1 .
2
ρU 2 Ar L
The moment Mxm is taken about a point xm , measured back from the nose; this
location is arbitrary, and appears explicitly in the formula for Cm . Jorgensen
gives the coefficients as follows:
Ab α Ap
CN = sin 2α cos + Cd sin2 α (120)
Ar 2 Ar n
CA = CAo cos2 α (121)
∇ − Ab (L − xm ) α Ap x m − x c
CM = − sin 2α cos − Cdn sin2 α.(122)
Ar L 2 Ar L
We have listed only the formulas for the special case of circular cross-section,
although the complete equations do account for more complex shapes. Further,
we have assumed that L >> D, which is also not a constraint in the complete
equations. The parameters used here are
• Ap : planform area.
• CAo : axial drag at zero angle of attack, both frictional and form. CAo '
0.002 − 0.006 for slender streamlined bodies, based on wetted surface
area. It depends on Re = U L/ν.
• ∇: body volume.
• xc : distance from the nose backwards to the center of the planform area.
70 10 PRACTICAL LIFT CALCULATIONS
In the formula for normal force, we see that if Ab = 0, only drag forces act
to create lift, through a sin2 α-dependence. Similarly, the axial force is simply
the zero-α result, modified by cos2 α. In both cases, scaling of U 2 into the
body principle directions is all that is required.
There are several terms that match exactly the slender-body theory approxi-
mations for small α. These are the first term in the normal force (CN ), and
the entire first term in the moment (CM ), whether or not Ab = 0. Finally, we
note that the second term in CM disappears if xm = xc , i.e., if the moment is
referenced to the center of the planform area. The idea here is that the fore
and aft components of crossflow drag cancel out.
The aerodynamic center (again referenced toward the stern, from the nose)
can be found after the coefficients are computed:
CM
xAC = xm + L. (123)
CN
As written, the moment coefficient is negative if the moment destabilizes the
body, while CN is always positive. Thus, the moment seeks to move the AC
forward on the body, but the effect is moderated by the lift force.
Y
CY =
DLq
Y
CYd = 2 .
D q
The first version uses the rectangular planform area as a reference, while the
second uses the square frontal area. Hence, CY = CYd D/L. Two moment
coefficients are:
10.4 Slender-Body Theory vs. Experiment 71
N
CN =
LD2 q
N1
CN1 = ,
LD2 q
where N is the moment taken about the body mid-point, and N1 is taken
about the nose. Note that the reference area for moment is the square frontal
projection, and the reference length is body length L. The following relation
holds for these definitions
1
CN1 = CN + CYd .
2
The lift and moment coefficients are strongly dependent on angle of attack;
Hoerner uses the notation
∂CN
Cnb =
∂b
∂CN1
Cn·b =
∂b
∂Cy
Cyb = ,
∂b
and so on, where b is the angle of attack, usually in degrees. It follows from
above that Cn·b = Cnb + Cydb /2.
reduces to −0.009 for L/D = 4. Note that the negative sign here is consistent
with Hoerner’s convention that destabilizing moments have negative sign.
The experimental lift force is typically given by CY b ' 0.003/deg; this acts at
a point on the latter half of the vehicle, stabilizing the angle. Because this
coefficient scales roughly with wetted area, proportional to LD, it changes
little with L/D. It can be compared with a low-aspect ratio wing, which
achieves an equivalent lift of π(AR)/2 = 0.0027 for (AR) = 10 ' D/L.
The point at which the viscous forces act can then be estimated as the following
distance aft of the nose:
x Cn·b − C̃nb
= (124)
L Cydb
The calculation uses experimental values of Cydb and Cn·b , the moment slope
referenced to the nose. In the table following are values from Hoerner (p. 13-2,
Figure 2) for a symmetric and a blunt-ended body.
symmetric blunt
L/D 6.7 6.7
C̃nb -0.012 -0.012
Cyb 0.0031 0.0037
Cydb 0.021 0.025
Cn·b 0.0012 (stable) 0.0031 (stable)
Cnb -0.0093 (unstable) -0.0094 (unstable)
x/L 0.63 0.60
In comparing the two body shapes, we see that the moment at the nose is
much more stable (positive) for the body with a blunt trailing edge. At the
body midpoint, however, both vehicles are equally unstable. The blunt-tailed
geometry has a much larger lift force, but it acts too close to the midpoint to
add any stability there.
The lift force dependence on the blunt tail is not difficult to see, using slender-
body theory. Consider a body, with trailing edge radius r. The slender-body
lift force associated with this end is simply the product of speed U and local
added mass ma (xT ) (in our previous notation). It comes out to be
1
Z = ρU 2 (πr2 )(2α), (125)
2
such that the first term in parentheses is an effective area, and the second is
a lift coefficient. With respect to the area πr2 , the lift curve slope is therefore
10.5 Slender-Body Approximation for Fin Lift 73
Z = πs2 ρU 2 α. (126)
Letting the aspect ratio be (AR) = (2s)2 /Af , where Af is the total area of the
fin pair, substitution will give a lift curve slope of
π
Cl0 = (AR).
2
This is known as Jones’ formula, and is quite accurate for (AR) ' 1. It is
inadequate for higher-aspect ratio wings however, overestimating the lift by
about 30% when (AR) = 2, and worsening further as (AR) grows.
74 11 FINS AND LIFTING SURFACES
and then L = −ρU Γ. Circulation is the integral of velocity around the cross-
section, and a lifting surface requires circulation in order to meet the Kutta
condition.
span span2
AR = = . (128)
chord area
The second representation is useful for non-rectangular control surfaces. The
effective span is taken to be the length between the free ends of a symmetric
wing. If the wing is attached to a wall, the effective span is twice the physical
value, by reflection, and in this case the effective aspect ratio is therefore twice
the physical value.
The aspect ratio is a strong determinant of wing performance: for a given
angle of attack, a larger aspect ratio achieves a higher lift value, but also stalls
earlier.
Lift is written as
1
L = ρU 2 ACl , (129)
2
where A is the single-side area of the surface. For angles of attack α below
stall, the lift coefficient Cl is nearly linear with α: Cl = Cl0 α, where Cl0 is called
the lift coefficient slope, and has one empirical description
1
Cl0 = 1 1 1 , (130)
2π ᾱ
+ π(AR)
+ 2π(AR)2
where α is in radians, ᾱ ' 0.90, and AR is the effective aspect ratio. When
AR → ∞, the theoretical and maximum value for Cl0 is 2π.
The lift generated on a surface is the result of a distributed pressure field; this
fact creates both a net force and a net moment. A single equivalent force acts
at a so-called center of action xA , which depends mainly on the aspect ratio.
For high AR, xA ' c/4, measured back from the front of the wing. For low
AR, xA ' c/2.
Table 1: Nomenclature
ne = λnp (135)
Qp = ηg λQe ,
To
KT = (thrust coefficient) (136)
ρn2p D4
Qp
KQ = 2 o 5 (torque coefficient). (137)
ρnp D
12.2 Steady Propulsion of Vessels 79
1.0 0.10
η0
0.8 0.08
K
T 0.6 0.06 K Q
η
0 0.4 K 0.04
Q
0.2 0.02
KT
0.0 0.00
0.0 0.2 0.4 0.6 0.8 1.0
J
Figure 4: Typical thrust and torque coefficients.
To U J(U )KT
ηo = = . (138)
2πnp Qpo 2πKQ
This efficiency divides the useful thrust power by the shaft power. Thrust and
torque coefficients are typically nearly linear over a range of J, and therefore
fit the approximate form:
KT (J) = β1 − β2 J (139)
KQ (J) = γ1 − γ2 J.
a vessel moving at speed U , and with the propeller spinning at the same
np , the prop creates some extra thrust. w scales U at the prop and thus
J; w is then chosen so that the open water thrust coefficient matches
what is observed. The wake fraction can also be estimated by making
direct velocity measurements behind the hull, with no propeller.
• Rt = Rsp (1 − t). Often, a propeller will increase the resistance of the
vessel by creating low-pressure on its intake side (near the hull), which
makes Rsp > Rt . In this case, t is a small positive number, with 0.2 as a
typical value. t is called the thrust deduction even though it is used to
model resistance of the hull; it is obviously specific to both the hull and
the propeller(s), and how they interact.
The thrust deduction is particularly useful, and can be estimated from
published values, if only the towed resistance of a hull is known.
• Qpo = ηR Qpsp . The rotative efficiency ηR , which may be greater than
one, translates self-propelled torque to open water torque, for the same
incident velocity Up , thrust T , and rotation rate np . ηR is meant to
account for spatial variations in the wake of the vessel which are not
captured by the wake fraction, as well as the turbulence induced by the
hull. Note that in comparison with the wake fraction, rotative efficiency
equalizes torque instead of thrust.
Rt U
ηQP = (140)
2πnp Qpsp
To (1 − t)Up ηR
=
2πnp (1 − w)Qpo
(1 − t)
= ηo ηR .
(1 − w)
To and Qpo are values for the inflow speed Up , and thus that ηo is the open
propeller efficiency at this speed. It follows that To (Up ) = Tsp , which was used
to complete the above equation. The quasi-propulsive efficiency can be greater
than one, since it relies on the towed resistance and in general Rt > Rsp . The
ratio (1 − t)/(1 − w) is often called the hull efficiency, and we see that a
12.2 Steady Propulsion of Vessels 81
small thrust deduction t and a large wake fraction w are beneficial effects, but
which are in competition. A high rotative efficiency and open water propeller
efficiency (at Up ) obviously contribute to an efficient overall system.
Tsp = Rsp
To = Rt /(1 − t)
1
KT (J(Up ))ρn2p D4 (1 − t) = ρCr Aw U 2
2
1 Up2
(β1 − β2 J(Up ))ρn2p D4 (1 − t) = ρCr Aw
2 (1 − w)2
Cr Aw
β1 − β2 J(Up ) = J(Up )2
2D (1 − t)(1 − w)2
2
| {z }
δ
q
−β2 + β22 + 4β1 δ
J(Up ) = . (142)
2δ
The last equation predicts the steady-state advance ratio of the vessel, depend-
ing only on the propeller open characteristics, and on the hull. The vessel speed
can be computed by recalling that J(U ) = U/np D and Up = U (1 − w), but it
is clear that we need now to find np . This requires a torque equation, which
necessitates a model of the drive engine or motor.
Q e /Q
m
f3
f2
f1
n e /n m
! !
f ne f
F (f /ff m, ne /nm ) = − a +b + c +d (144)
fm nm fm
np
= −α1 + α2 .
nm /λ
then a closed-form solution for ne (and thus np ) can be found. The manipula-
tions begin by equating the engine and propeller torque:
where Ip is the total (material plus fluid) inertia reflected to the prop;the
propeller spins at ωp radians per second. The differential equation in ωp pits
the torque delivered by the motor against a quadratic-drag type loss which
depends on rotation speed. The thrust is then given as a static map directly
from the rotation speed.
This model requires the identification of three parameters: Ip , Kω , and Ct .
It is a first-order, nonlinear, low-pass filter from Qm to T , whose bandwidth
depends directly on Qm .
84 12 PROPELLERS AND PROPULSION
13 TOWING OF VEHICLES
Vehicles which are towed have some similarities to the vehicles that have been
discussed so far. For example, towed vehicles are often streamlined, and usu-
ally need good directional stability. Some towed vehicles might have active
lifting surfaces or thrusters for attitude control. On the other hand, if they
are to be supported by a cable, towed vehicles may be quite heavy in water,
and do not have to be self-propelled. The cable itself is an important factor in
the behavior of the complete towed system, and in this section, we concentrate
on cable mechanics more than vehicle characteristics, which can generally be
handled with the same tools as other vehicles, i.e., slender-body theory, wing
theory, linearization, etc.. Some basic guidelines for vehicle design are given
at the end of this section.
Modern cables can easily exceed 5000m in length, even a heavy steel cable with
2cm diameter. The cables are generally circular in cross section, and may
carry power conductors and multiple communication channels (fiber optic).
The extreme L/D ratio for these cables obviates any bending stiffness effects.
Cable systems come in a variety of configurations, and one main division may
be made simply of the density of the cable. Light-tether systems are character-
ized by neutrally-buoyant (or nearly so) cables, with either a minimal vehicle
at the end, as in a towed array, or a vehicle capable of maneuvering itself, such
as a remotely-operated vehicle. The towed array is a relatively high-velocity
system that nominally streams out horizontally behind the vessel. An ROV,
on the other hand, operates at low speed, and must have large propulsors to
control the tether if there are currents. Heavy systems, in contrast, employ a
heavy cable and possibly a heavy weight; the rationale is that gravity will tend
to keep the cable vertical and make the deployment robust against currents
and towing speed. The heavy systems will generally transmit surface motions
and tensions to the towed vehicle much more easily than light-tether systems.
We will not discuss light systems specifically here, but rather look at heavy
systems. Most of the analysis can be adapted to either case, however.
13.1 Statics
13.1.1 Force Balance
For the purposes of deriving the static configuration of a cable in a flow, we
assume for the moment that that it is inextensible. Tension and hydrostatic
86 13 TOWING OF VEHICLES
y s+ T+dT
ds
x ds Rt φ U
T ds Rn
ds Wn
pressure will elongate a cable, but the effect is usually a small percentage of
the total length.
We employ the curvilinear axial coordinate s, which we take to be zero at the
bottom end of the cable; upwards along the cable is the positive direction.
The free-body diagram shown has the following components:
Force balance in the tangential and normal coordinates gives two coupled
equations for T and φ:
dT
amp; = amp; Wn sin φ − Rt (151)
ds
dφ
T amp; = amp; Wn cos φ + Rn . (152)
ds
The external forces are primarily fluid drag; the tangential drag is controlled
by a frictional drag coefficient Ct , and the normal drag scales with a crossflow
drag coefficient Cn . In both cases, the fluid velocity vector, U horizontal
toward the left, is to be projected onto the relevant axes, leading to
13.1 Statics 87
1
Rt = − ρCt dU 2 cos2 φ (153)
2
1
Rn = − ρCn dU 2 sin2 φ. (154)
2
Note that we simplified the drag laws from a usual form v|v| to v 2 , since as
drawn, 0 ≤ φ ≤ π/2.
The equations for T and φ can be integrated along the cable coordinate s to
find the cable’s static configuration. Two boundary conditions are needed, and
the common case is that a force balance on the vehicle, dominated by drag,
weight, and the cable tension, provides both T (0) and φ(0). For example, a
very heavy but low-drag vehicle will impose φ(0) ' π/2, with T (0) equal to
the in-water weight of the vehicle.
With regard to Cartesian coordinates x, y, the cable configuration follows
dx
amp; = amp; cos φ (155)
ds
dy
amp; = amp; sin φ. (156)
ds
The simultaneous integration of all four equations (T, φ, x, y) defines the cable
configuration, and current dependency may be included, say U is a function
of y.
Wn
δ= ,
ρCn dU 2
88 13 TOWING OF VEHICLES
85
13 TOWING OF VEHICLES
Vehicles which are towed have some similarities to the vehicles that have been
discussed so far. For example, towed vehicles are often streamlined, and usu-
ally need good directional stability. Some towed vehicles might have active
lifting surfaces or thrusters for attitude control. On the other hand, if they
are to be supported by a cable, towed vehicles may be quite heavy in water,
and do not have to be self-propelled. The cable itself is an important factor in
the behavior of the complete towed system, and in this section, we concentrate
on cable mechanics more than vehicle characteristics, which can generally be
handled with the same tools as other vehicles, i.e., slender-body theory, wing
theory, linearization, etc.. Some basic guidelines for vehicle design are given
at the end of this section.
Modern cables can easily exceed 5000m in length, even a heavy steel cable with
2cm diameter. The cables are generally circular in cross section, and may
carry power conductors and multiple communication channels (fiber optic).
The extreme L/D ratio for these cables obviates any bending stiffness effects.
Cable systems come in a variety of configurations, and one main division may
be made simply of the density of the cable. Light-tether systems are character-
ized by neutrally-buoyant (or nearly so) cables, with either a minimal vehicle
at the end, as in a towed array, or a vehicle capable of maneuvering itself, such
as a remotely-operated vehicle. The towed array is a relatively high-velocity
system that nominally streams out horizontally behind the vessel. An ROV,
on the other hand, operates at low speed, and must have large propulsors to
control the tether if there are currents. Heavy systems, in contrast, employ a
heavy cable and possibly a heavy weight; the rationale is that gravity will tend
to keep the cable vertical and make the deployment robust against currents
and towing speed. The heavy systems will generally transmit surface motions
and tensions to the towed vehicle much more easily than light-tether systems.
We will not discuss light systems specifically here, but rather look at heavy
systems. Most of the analysis can be adapted to either case, however.
13.1 Statics
13.1.1 Force Balance
For the purposes of deriving the static configuration of a cable in a flow, we
assume for the moment that that it is inextensible. Tension and hydrostatic
1 2
δ cos φc − sin φc = 0.
2
We are considering the case of 0 < φc < π/2. Substituting sin2 φc = 1−cos2 φc ,
we solve a quadratic equation and keep only the positive solution:
√
cos φc = δ 2 + 1 − 1. (157)
In the case of a √
very heavy cable, δ is large, and the linear approximation of
the square root 1 + ≈ 1 + /2 gives
1
cos φc amp; ' amp; −→
2δ
π 1
φc amp; ' amp; − . (158)
2 2δ
For a very light cable, δ is small; the same approximation gives
13.2 Linearized Dynamics 89
√
cos φc ' 1 − δ −→ φc ' 2δ.
The table below gives some results of the exact solution, and the approxima-
tions.
We also write the axial deflection with respect to the static configuration as
p(s, t), and the lateral deflection q(s, t). It follows that φ̃ = ∂q/∂s. Now
augment the two static configuration equations with inertial components:
!2
∂2p ∂ T̄ ∂ T̃ 1 ∂p
m 2 amp; = amp; + − Wn sin(φ̄ + φ̃) − ρCt d U cos φ +
∂t ∂s ∂s 2 ∂t
∂2q
!
∂ φ̄ ∂ φ̃
(m + ma ) 2 amp; = amp; (T̄ + T̃ ) + − Wn cos(φ̄ + φ̃) +
∂t ∂s ∂s
!2
1 ∂q
amp; amp; ρCn d U sin φ − .
2 ∂t
90 13 TOWING OF VEHICLES
Here the material mass of the cable per unit length is m, and its transverse
added mass is ma . Note that avoiding the drag law form v|v| again, we have
implicitly assumed that U cos φ > |∂p/∂t| and U sin φ > |∂q/∂t|. If it is not
the case, say U = 0, then equivalent linearization can be used for the quadratic
drag.
Now we perform the trigonometry substitutions in the weight terms, let φ ' φ̄
for the calculation of drag, and substitute the constitutive (Hooke’s) law
∂ T̃ ∂2p
= EA 2 .
∂s ∂s
The static solution cancels out of both governing equations, and keeping only
linear terms we obtain
∂2p ∂2p ∂q ∂p
m 2
amp; = amp; EA 2
− Wn cos φ̄ − ρCt dU cos φ̄
∂t ∂s ∂s ∂t
2 2
∂ q ∂ q ∂p ∂ φ̄ ∂q ∂q
(m + ma ) 2 amp; = amp; T̄ 2 + EA + Wn sin φ̄ − ρCn dU sin φ̄ .
∂t ∂s ∂s ∂s ∂s ∂t
The axial dynamics (p) couples with the lateral equation through the weight
term −Wn cos φ̄φ̃. The lateral dynamics (q) couples with the axial through the
term T̃ ∂ φ̄/∂s. An additional weight term Wn sin φ̄∂q/∂s also appears. The
uncoupled dynamics are both in the form of damped wave equations
∂2p ∂p ∂2p
m + b t amp; = amp; EA
∂t2 ∂t ∂s2
2 2
∂ q ∂q ∂ q ∂q
(m + ma ) 2 + bn amp; = amp; T̄ 2 + Wn sin φ̄ ,
∂t ∂t ∂s ∂s
where we made the substitution bt = ρCt dU cos φ̄ and bn = ρCn dU sin φ̄. To a
linear approximation, the out-of-plane vibrations of a cable are also governed
by the second equation above.
Because of light damping in the tangential direction, heavy cables easily trans-
mit motions and tensions along their length, and can develop longitudinal
resonant conditions (next section). In contrast, the lateral cable motions are
heavily damped, such that disturbances only travel a few tens or hundreds of
meters before they dissipate. The nature of the lateral response, in and out
13.2 Linearized Dynamics 91
∂p(0, t)
EA = M p̈(0, t). (161)
∂s
These top and bottom behaviors comprise the boundary conditions for the
wave equation. We let p(s, t) = p̃(s) cos ωt, so that
92 13 TOWING OF VEHICLES
mω 2 − iωbt
!
00
p̃ + p̃ = 0. (162)
EA
This admits the solution p̃(s) = c1 cos ks + c2 sin ks, where
s
mω 2 − iωbt
k= . (163)
EA
6 0. The top and bottom boundary conditions
Note that k is complex when bt =
give, respectively,
• The maximum tension is T̄ + |T̃ | and must be less than the working load
of the cable. This is normally problematic at the top of the cable, where
the static tension is highest.
• If |T̃ | > T̄ , the cable will unload completely and then reload with ex-
tremely high impulsive forces. This is known as snap loading; it occurs
primarily at the vehicle, where T̄ is low.
q
kL << 1, but we find that a first-order approximation yields ω = EA/LM ,
which is only a correct answer if M >> mL, i.e., the system is dominated by
the vehicle mass. Some higher order terms need to be kept. We start with
better approximations for sin() and cos():
(kL)2 (kL)2
! !
δ 1− = kL 1 − .
2 6
(kL)2 (kL)2
! !
mL
1− = (kL)2 1− .
M 2 6
v
EA/L
u
u
ω=t .
M + mL/2
This has the familiar form of the square root of a stiffness divided by a mass:
the stiffness of the cable is EA/L, and the mass that is oscillating is M +mL/2.
In very deep water, the effects of mL/2 dominate; if ρc = m/A is the density
of the cable, we have the approximation
s
1 2E
ω' .
L ρc
A few examples are given below for a steel cable with E = 200 × 109 P a, and
ρc = 7000kg/m3 . The natural frequencies near wave excitation at the surface
vessel must be taken into account in any design or deployment. Even if a
cable can withstand the effects of resonance, it may be undesirable to expose
the vehicle to these motions. Some solutions in use today are: stable vessels
(e.g., SWATH), heave compensation through an active crane, a clump weight
below which a light cable is employed, and an S-shaped length of cable at the
bottom formed with flotation balls.
94 13 TOWING OF VEHICLES
1. The towpoint must be located above the vehicle center of in-water weight,
for basic roll and pitch stability.
3. The combined center of mass (material and added mass) should be longi-
tudinally between the towpoint and the aerodynamic center, and nearer
13.4 Vehicle Design 95
Meeting all of these criteria simultaneously is no small feat, and the perfor-
mance of the device is very sensitive to small perturbations in the geometry.
For this reason, full-scale experiments are commonly used in the design pro-
cess.
96 14 TRANSFER FUNCTIONS & STABILITY
K(s + z1 )(s + z2 ) · · · (s + zm )
Y (s) = . (166)
(s + p1 )(s + p2 ) · · · (s + pn )
Although for the moment we are discussing the signal Y (s), later we will see
that dynamic systems are described in the same format: in that case we call
the impulse response G(s) a transfer function. A system transfer function is
identical to its impulse response, since L(δ(t)) = 1.
The constants −zi are called the zeros of the transfer function or signal, and
−pi are the poles. Viewed in the complex plane, it is clear that the magnitude
of Y (s) will go to zero at the zeros, and to infinity at the poles.
Partial fraction expansions alter the form of Y (s) so that the simple transform
pairs can be used to find the time-domain output signals. We must have
m < n; if this is not the case, then we have to divide the numerator by the
denominator as necessary to find a simple form.
How to find ai ? A simple rule applies: multiply the right-hand sides of the
two equations above by (s + pi ), evaluate them at s = −pi , and solve for ai ,
the only one left.
s(s + 6)
G(s) = e−2s .
(s + 4)(s − 1)
Since we have a pure delay and m = n, we can initially work with G(s)/se−2s .
We have
s+6 a1 a2
= + , giving
(s + 4)(s − 1) s+4 s−1
h i
(s+6)(s+4) 2
a1 = (s+4)(s−1) s=−4
= −
5
h
(s+6)(s−1)
i 7
a2 = (s+4)(s−1) s=1
=
5
Thus
2 7
L−1 (G(s)/se−2s ) = − e−4t + et −→
5 5
8 7
g(t) = δ(t − 2) + e−4(t−2) + et−2 .
5 5
The impulse response is needed to account for the step change at t = 2. Note
that in this example, we were able to apply the derivative operator s after
expanding the partial fractions. For cases where a second derivative must be
taken, i.e., m ≥ n + 1, special care should be used when accounting for the
signal slope discontinuity at t = 0. The more traditional method, exemplified
by Ogata, may prove easier to work through.
The case of repeated real roots may be handled elegantly, but this condition
rarely occurs in applications.
98 14 TRANSFER FUNCTIONS & STABILITY
b1 s + b2 a3
Y (s) = + + ···, (168)
(s + p1 )(s + p2 ) s + p3
where p1 = p∗2 (complex conjugate). As before, multiply through by (s +
p1 )(s + p2 ), and then evaluate at s = −p1 .
s+1 b1 s + b2 a3
G(s) = = + :
s(s + j)(s − j) (s + j)(s − j) s
s+1
= [b1 s + b2 ]s=−j −→
s s=−j
1 + j = −b1 j + b2 −→
b1 = −1
b2 = 1; also
" #
s+1
= a3 = 1.
(s + j)(s − j) s=0
Working out the inverse transforms from the table of pairs, we have simply
(noting that ζ = 0)
1. Exponential. If ~u(t) = ~0 and ~y (0) = ~yo , then |yi (t)| < αe−γt , for finite
α and γ > 0. The output asymptotically approaches zero, within a
decaying exponential envelope.
15 CONTROL FUNDAMENTALS
15.1 Introduction
15.1.1 Plants, Inputs, and Outputs
ẋ = Ax + Bu + Gw (169)
y = Cx + Du + v
where
• Du is a direct map from input to output (usually zero for physical sys-
tems).
w G
v
+ +
+ x’ x +
u B 1/s C y
+ +
y(s)
G(s) = = C(sI − A)−1 B + D, (170)
u(s)
where I is the identity matrix with the same size as A. A similar equation
holds for y(s)/w(s), and clearly y(s)/v(s) = I.
disturbance d(t). If the linear position of the mass is y(t), and it is perfectly
measured, we have the plant
Suppose that the desired condition is simply y(t) = 0, with initial conditions
y(0) = yo and y 0 (0) = 0.
" #
−kd
t kd
y(t) = yo e 2m cos ωd t + sin ωd t ,
2mωd
15.5 Heuristic Tuning 105
q
where ωd = 4mkp − kd2 /2m. This response is exponentially stable as desired.
Note that if the mass had a very large amount of natural damping, a negative
kd could be used to cancel some of its effect and speed up the system response.
Now consider what happens if d(t) has a constant bias do : it balances exactly
the proportional control part, eventually settling out at y(t = ∞) = do /kp . To
achieve good rejection of do with a P D controller, we would need to set kp very
large. However, very large values of kp will also drive the resonant frequency
ωd up, which is unacceptable.
15.4.3 Proportional-Integral-Derivative
Rt
Now let u(t) = −kp y(t) − ki 0 y(τ )dτ − kd y 0 (t): we have
Z t
00
my (t) = −kp y(t) − ki y(τ )dτ − kd y 0 (t) + d(t).
0
ke−T s
G(s) ' (172)
τs + 1
The following rules apply only if the plant contains no dominating, lightly-
damped complex poles, and has no poles at the origin:
106 15 CONTROL FUNDAMENTALS
P kp = 1.0τ /T
PI kp = 0.9τ /T ki = 0.27τ /T 2
PID kp = 1.2τ /T ki = 0.60τ /T 2 kd = 0.60τ
Note that if no pure time delay exists (T = 0), this recipe suggests the pro-
portional gain can become arbitrarily high! Any characteristic other than a
true first-order lag would therefore be expected to cause a measurable delay.
u(s) y(s)
C(s) P(s)
noise properties can ruin the performance of a control system, no matter how
perfectly understood are the other components.
du dy
+ e u + + y
r C P
- + +
+
n
+
e 1
= =S
r 1 + PC
y PC
= =T
r 1 + PC
u C
= = U.
r 1 + CP
e/r = S relates the reference input and noise to the error, and is known as the
sensitivity function. We would generally like S to be small at low frequencies,
so that the tracking error there is small. y/r = T is called the complementary
sensitivity function. Note that S + T = 1, implying that these two functions
must always trade off; they cannot both be small or large at the same time.
Other systems we encounter again later are the (forward) loop transfer function
P C, the loop transfer function broken between C and P : CP , and
e −P
=
du 1 + PC
y P
=
du 1 + PC
u −CP
=
du 1 + CP
e −1
= = −S
dy 1 + PC
108 15 CONTROL FUNDAMENTALS
y 1
= =S
dy 1 + PC
u −C
= = −U
dy 1 + CP
e −1
= = −S
n 1 + PC
y −P C
= = −T
n 1 + PC
u −C
= = −U.
n 1 + CP
If the disturbance is taken at the plant output, then the three functions S, T ,
and U (control action) completely describe the system. This will in fact be
the procedure when we address loopshaping.
109
16 MODAL ANALYSIS
16.1 Introduction
The evolution of states in a linear system occurs through independent modes,
which can be driven by external inputs, and observed through plant output.
This section provides the basis for modal analysis of systems. Throughout, we
use the state-space description of a system with D = 0:
In the scalar case, the response is x(t) = χeat , giving a decaying exponential
if a < 0. The same notation holds for the case of a vector ~x, and matrix A:
~x(t) = eAt χ
~ , where
(At)2
eAt = I + At + + ···
2!
eAt is usually called the matrix exponential.
This is called the modal canonical form, since the states are simply the modal
amplitudes. These states are uncoupled in Λ, but may be coupled through the
input (V −1 B) and output (CV ) mappings. The modal form is numerically
robust for computations.
(At)2
eAt = I + At + + ···
2!
(Λt)2
!
= V I + Λt + + ··· WT
2!
= V eΛt W T
n
eλi t~vi w
~ iT .
X
=
i=1
n
eλi t~vi (w
~ iT χ
X
~x(t) = ~ ).
i=1
The product w ~ iT χ
~ is a scalar, the projection of the initial conditions onto the
i’th mode. If χ~ is perpendicular to w ~ i , then the product is zero and the i’th
mode does not respond. Otherwise, the i’th mode does participate in the
response. The projection of the i’th mode onto the states ~x is through the
right eigenvector ~vi .
For stability of the system, the eigenvalues of A, that is, λi , must have neg-
ative real parts; they are in fact the poles of the equivalent transfer function
description.
16.3 Forced Response and Controllability 111
Z t
~x(t) = eA(t−τ ) B~u(τ )dτ
0
n Z t
eλi (t−τ )~vi w
~ iT B~u(τ )dτ.
X
=
i=1 0
Suppose now that there are m inputs, such that B = [~b1 , ~b2 , · · · , ~bm ]. Then
some rearrangement will give
n m Z t
~ iT~bk ) eλi (t−τ )~uk (τ )dτ.
X X
~x(t) = ~vi (w
i=1 k=1 0
The product w ~ iT~bk , a scalar, represents the projection of the k’th control chan-
nel onto the i’th mode. We say that the i’th mode is controllable from the
k’th input if the product is nonzero. If a given mode i has w ~ iT~bk = 0 for all
input channels k, then the mode is uncontrollable.
In normal applications, controllability for the entire system is checked using
the following test: Construct the so-called controllability matrix:
n
eλi t~vi w
~ iT χ
X
~x(t) = ~.
i=1
~c1T
C = · .
~crT
n
(~ckT ~vi )eλi t (w
~ iT χ
X
yk (t) = ~ ).
i=1
The i’th mode is observable in the k’th output if the product ~ckT ~vi 6= 0. We
say that a system is observable if every mode can be seen in at least one output
channel. The usual test for system observability requires computation of the
observability matrix:
h i
Mo = C T , AT C T , · · · , (AT )n−1 C T . (175)
This matrix has size n × (rn); the system is observable if Mo has rank n.
113
e(s) 1
S(s) = = ,
r(s) 1 + P (s)C(s)
where P (s) represents the plant transfer function, and C(s) the compensator.
The closed-loop characteristic equation, whose roots are the poles of the closed-
loop system, is 1 + P (s)C(s) = 0, equivalent to P (s)C(s) + P (s)C(s) =
0, where the underline and overline denote the denominator and numerator,
respectively. The Nyquist criterion allows us to assess the stability properties
of a system based on P (s)C(s) only. This method for design involves plotting
the complex loci of P (s)C(s) for the range ω = [−∞, ∞]. There is no explicit
calculation of the closed-loop poles, and in this sense the design approach is
quite different from the root-locus method (see Ogata).
114 17 CONTROL SYSTEMS – LOOPSHAPING
P = CCW. (176)
In words, stability requires that the number of unstable poles in F (s) is equal
to the number of CCW encirclements of the origin, as s sweeps around the
entire right-half s-plane.
This is in fact the complete Nyquist criterion for stability. It is a necessary and
sufficient condition that the number of unstable poles in the loop transfer func-
tion P (s)C(s) must be matched by an equal number of CCW encirclements
of the critical point (−1 + 0j).
There are several details to keep in mind when making Nyquist plots:
• If neither the plant nor the controller have unstable modes, then the loci
of P (s)C(s) must not encircle the critical point at all.
• Because the path taken in the s-plane includes negative frequencies (i.e.,
the negative imaginary axis), the loci of P (s)C(s) occur as complex
conjugates – the plot is symmetric about the real axis.
• The requirement that the number of poles in P (s)C(s) exceeds the num-
ber of zeros means that at high frequencies, P (s)C(s) always decays such
that the loci go to the origin.
• For the multivariable (MIMO) case, the procedure of looking at individ-
ual Nyquist plots for each element of a transfer matrix is unreliable and
outdated. Referring to the multivariable definition of S(s), we should
count the encirclements for the function [det(I + P (s)C(s)) − 1] instead
of P (s)C(s). The use of gain and phase margin in design is similar to
the SISO case.
These notions lead to definition of the gain margin kg and phase margin γ for
a design. As the figure shows, the definition of kg is different for stable and
unstable P (s)C(s). Rules of thumb are as follows. For a stable plant, kg ≥ 2
and γ ≥ 30◦ ; for an unstable plant, kg ≤ 0.5 and γ ≥ 30◦ .
116 17 CONTROL SYSTEMS – LOOPSHAPING
Re(s) Re(s)
γ
1/kg
P(s)C(s)
|1 + P C| > |W2 P C| ←→
|W2 P C|
1 > ←→
|1 + P C|
1 > |W2 T |, (180)
1. At low frequency, |W1 S| ' |W1 /P C|, since |P C| is large. This leads
directly to the performance condition |P C| > |W1 | in this range.
1Z∞ d
angle(H(jω0 )) = (ln|H(jω)) · ln(coth(|ν|/2))dν, (182)
π −∞ dν
where ν = ln(ω/ω0 ). The integral is hence taken over the log of a frequency
normalized with ω0 . It is not hard to see how the integral controls the angle:
the function ln(coth(|ν|/2)) is nonzero only near ν = 0, implying that the
angle depends only on the local slope d(ln|H|)/dν. Thus, if the slope is large,
the angle is large.
Example: Suppose H(s) = ω0n /sn , i.e., it is a simple function with n poles at
the origin, and no zeros; ω0 is a fixed constant. It follows that |H| = ω0n /ω n ,
and ln|H| = −nln(ω/ω0 ), so that d(ln|H|)/dν = −n. Then we have just
nZ ∞ nπ
angle(H) = − ln(coth(|ν|/2))dν = − . (183)
π −∞ 2
17.7 The Recipe for Loopshaping 119
This integral is trivial to look up or compute. Each pole at the origin clearly
induces 90◦ of phase loss. In the general case, each pole not at the origin
induces 90◦ of phase loss for frequencies above the pole. Each zero at the
origin adds 90◦ phase lead, while zeros not at the origin at 90◦ of phase lead
for frequencies above the zero. In the immediate neighborhood of these poles
and zeros, the phase may vary significantly with frequency.
The Nyquist loci are clearly susceptible to these variations is phase, and
the phase margin can be easily lost if the slope of P C at crossover (where
the magnitude is unity) is too steep. The slope can safely be first-order
(−20dB/decade, equivalent to a single pole), and may be second-order
(−40dB/decade) if an adequate phase angle can be maintained near crossover.
C = L/P. (184)
This extraordinarily simple step involves a plant inversion.
The overall idea is to first shape L as a stable transfer function meeting the
requirements of stability and robustness, and then divide through by the plant.
• When the plant is stable and has stable zeros (minimum-phase), the
division can be made directly.
When a control system involves multiple inputs and outputs, the ideas from
scalar loopshaping can be adapted using the singular value. We list below
some basic properties of the singular value decomposition, which is analogous
to an eigenvector, or modal, analysis. Useful properties and relations for the
singular value are found in the section MATH FACTS.
The condition for MIMO robust performance can be written in many ways,
including a direct extension of our scalar condition
B1 C1
A B2 C2 D
B3 C3 n=3
s=2
for example, there are many paths to choose from! Assume that one way or
another we have found the best path, and that a Point B lies along this path,
say Las Vegas. Let X be an arbitrary point east of Las Vegas. If we were to
now solve the optimization problem for getting from only Las Vegas to Boston,
this same arbitrary point X would be along the new optimal path as well.
The point is a subtle one: the optimization problem from Las Vegas to Boston
is easier than that from Los Angeles to Boston, and the idea is to use this
property backwards through time to evolve the optimal path, beginning in
Boston.
Example: Nodal Travel. We now add some structure to the above experi-
ment. Consider now traveling from point A (Los Angeles) to Point D (Boston).
Suppose there are only three places to cross the Rocky Mountains, B1 , B2 , B3 ,
and three places to cross the Mississippi River, C1 , C2 , C3 .3 By way of no-
tation, we say that the path from A to B1 is AB1 . Suppose that all of the
paths (and distances) from A to the B-nodes are known, as are those from the
B-nodes to the C-nodes, and the C-nodes to the terminal point D. There are
nine unique paths from A to D.
A brute-force approach sums up the total distance for all the possible paths,
and picks the shortest one. In terms of computations, we could summarize that
this method requires nine additions of three numbers, equivalent to eighteen
additions of two numbers. The comparison of numbers is relatively cheap.
The dynamic programming approach has two steps. First, from each B-node,
pick the best path to D. There are three possible paths from B1 to D, for
example, and nine paths total from the B-level to D. Store the best paths
as B1 D|opt , B2 D|opt , B3 D|opt . This operation involves nine additions of two
3
Apologies to readers not familiar with American geography.
18.4 Dynamic Programming and Full-State Feedback 123
numbers.
Second, compute the distance for each of the possible paths from A to D,
constrained to the optimal paths from the B-nodes onward: AB1 + B1 D|opt ,
AB2 + B2 D|opt , or AB3 + B3 D|opt . The combined path with the shortest
distance is the total solution; this second step involves three sums of two
numbers, and total optimization is done in twelve additions of two numbers.
Needless to say, this example gives only a mild advantage to the dynamic
programming approach over brute force. The gap widens vastly, however, as
one increases the dimensions of the solution space. In general, if there are s
layers of nodes (e.g., rivers or mountain ranges), and each has width n (e.g., n
river crossing points), the brute force approach will take (sns ) additions, while
the dynamic programming procedure involves only (n2 (s − 1) + n) additions.
In the case of n = 5, s = 5, brute force requires 6250 additions; dynamic
programming needs only 105.
i.e., the integral over time of the instantaneous penalty. Finally, the optimal
return is the cost of the optimal trajectory remaining after time t:
Z ∞
V (x(t), u(t)) = l(x(τ ), u(τ ))dτ. (188)
t
.
We have directly from the dynamic programming principle
V (x(t), u(t)) = min {l(x(t), u(t))δt + V (x(t + δt), u(t + δt))} . (189)
u
The minimization of V (x(t), u(t)) is made by considering all the possible con-
trol inputs u in the time interval (t, t + δt). As suggested by dynamic pro-
gramming, the return at time t is constructed from the return at t + δt, and
124 18 LINEAR QUADRATIC REGULATOR
the differential component due to l(x, u). If V is smooth and has no explicit
dependence on t, as written, then
∂V dx
V (x(t + δt), u(t + δt)) = V (x(t), u(t)) + δt + h.o.t. −→ (190)
∂x dt
∂V
= V (x(t), u(t)) + (Ax(t) + Bu(t))δt.
∂x
Now control input u in the interval (t, t + δt) cannot affect V (x(t), u(t)), so
inserting the above and making a cancellation gives
( )
∂V
0 = min l(x(t), u(t)) + (Ax(t) + Bu(t)) . (191)
u ∂x
We next make the assumption that V (x, u) has the following form:
1
V (x, u) = xT P x, (192)
2
where P is a symmetric matrix, and positive definite.45 It follows that
∂V
= xT P −→ (193)
∂x n o
0 = min l(x, u) + xT P (Ax + Bu) .
u
We finally specify the instantaneous penalty function. The LQR employs the
special quadratic form
1 1
l(x, u) = xT Qx + uT Ru, (194)
2 2
where Q and R are both symmetric and positive definite. The matrices Q
and R are to be set by the user, and represent the main “tuning knobs” for
the LQR. Substitution of this form into the above equation, and setting the
derivative with respect to u to zero gives
4
Positive definiteness means that xT P x > 0 for all nonzero x, and xT P x = 0 if x = 0.
5
This suggested form for the optimal return can be confirmed after the optimal controller
is derived.
18.5 Properties and Use of the LQR 125
0 = uT R + x T P B (195)
uT = −xT P BR−1
u = −R−1 B T P x.
The gain matrix for the feedback control is thus K = R−1 B T P . Inserting
this solution back into equation 194, and eliminating u in favor of x, we have
1 1
0 = xT Qx − xT P BR−1 B T P + xT P Ax.
2 2
1 1
xT P Ax = xT P Ax + xT AT P x,
2 2
leading to the final matrix-only result
0 = Q + P A + AT P − P BR−1 B T P. (196)
This equation is the matrix algebraic Riccati equation (MARE), whose solution
P is needed to compute the optimal feedback gain K. The MARE is easily
solved by standard numerical tools in linear algebra.
X
φ /2
n=4
X φ = 45 degrees
φ
φ
φ
X
φ /2
X
128 19 KALMAN FILTER
19 KALMAN FILTER
19.1 Introduction
In the previous section, we derived the linear quadratic regulator as an optimal
solution for the full-state feedback control problem. The inherent assumption
was that each state was known perfectly. In real applications, the measure-
ments are subject to disturbances, and may not allow reconstruction of all the
states. This state estimation is the task of a model-based estimator having
the form:
y
C
Bu
y - +
+ x x
H 1/s
+
+
ẋ = Ax + Bu + W1 (198)
y = Cx + W2 .
There are n states, m inputs, and l outputs, so that A has dimension n × n, B
is n × m, and C is l × n. The plant subject to two random input signals, W1
19.3 Step 1: An Equation for Σ̇ 129
Σ = E(eeT ). (204)
There are two main steps for deriving the optimal gain H.
J 0 = trace(ΣW + ΛF ), (208)
where F is an n × n matrix of zeros, and Λ is an n × n matrix of unknown
Lagrange multipliers. Note that since F is zero, J 0 = J, so minimizing J 0 solves
the same problem. Lagrange multipliers provide an ingenious mechanism for
drawing constraints into the optimization; the constraint we invoke is the
evolution of Σ, Equation 206:
J 0 = trace ΣW + Λ(−Σ̇ + AΣ − HCΣ + ΣAT − ΣC T H T + V1 + HV2 H T )
(209)
If J 0 is an optimal cost, it follows that ∂J 0 /∂H = 0, i.e., the correct choice of
H achieves an extremal value. We need the following lemmas, which give the
derivative of a trace with respect to a constituent matrix:
∂
trace(−ΛHCΣ) = −ΛT ΣC T
∂H
∂
trace(−ΛΣC T H T ) = −ΛΣC T
∂H
∂
trace(ΛHV2 H T ) = ΛT HV2 + ΛHV2 .
∂H
Proofs of the first two are given at the end of this section; the last lemma uses
the chain rule, and the previous two lemmas. Next, we enforce Λ = ΛT , since
the values are arbitrary. Then the condition ∂J 0 /∂H = 0 leads to
Equations 210 and 211 represent the practical solution to the Kalman filtering
problem, which minimizes the squared-norm of the estimation error. The
evolution of Σ is always stable, and depends only on the constant matrices
[A, C, V1 , V2 ]. Notice also that the result is independent of the weighting matrix
W , which might as well be the identity.
The limiting closed-loop poles of the Kalman filter are similar, and dual to
those of the LQR:
ẋ = Ax + Bu + W1
y = Cx + W2 ,
C(s)
B
+ +
y H φ -K u
- +
φ = (sI-A) -1
i=1 j=1
n
XXn
= Σij Wji ,
i=1 j=1
19.7 Proofs of the Intermediate Results 135
the transpose of W being valid since it is symmetric. Now consider the diagonal
elements of the product ΣW :
Σ11 W11 + Σ12 W21 + · · · · ·
ΣW =
· Σ21 W12 + Σ22 W22 + · · · · →
· · ···
n X
X n
trace(ΣW ) = Σij Wji . 2
i=1 j=1
∂
19.7.2 Proof that ∂H
trace(−ΛHCΣ) = −ΛT ΣC T
n l
Hjk Bkl , the il0 th element
X X
trace(AHB) = trace Aij
j=1 k=1
n X
X n l
X
= Aij Hjk Bki ,
i=1 j=1 k=1
where the second form is a sum over i of the ii’th elements. Now
n
∂ X
trace(AHB) = Aijo Bko i
∂Hjo ko i=1
= (BA)ko jo
= (BA)Tjo ko −→
∂
trace(AHB) = (BA)T
∂H
= AT B T . 2
∂
19.7.3 Proof that ∂H
trace(−ΛΣC T H T ) = −ΛΣC T
n
trace(AH T ) = trace Aij HjlT , the il0 th element
X
j=1
n
X
= trace Aij Hlj
j=1
n X
X n
= Aij Hij ,
i=1 j=1
136 19 KALMAN FILTER
where the last form is a sum over the ii’th elements. It follows that
∂
trace(AH T ) = Aio jo −→
∂Hio jo
∂
trace(AH T ) = A. 2
∂H
( ) " #( )
d x A −BK x
= .
dt x̂ HC A − BK − HC x̂
Using the definition of estimation error e = x − x̂, we can write the above in
another form:
( ) " #( )
d x A − BK BK x
= .
dt e 0 A − HC e
If A0 represents this compound A-matrix, then its eigenvalues are the roots of
det(sI − A0 ) = 0. However, the determinant of an upper triangular block
matrix is the product of the determinants of each block on the diagonal:
det(sI −A0 ) = det(sI −(A−BK))det(sI −(A−HC)), and hence the separation
of eigenvalues follows.
137
and the figure. Note that we have not included the factor Bu as part of the
figure, since it does not affect the error dynamics of the filter.
As noted previously, the KF loop has good robustness properties, specifically
to perturbations at the output ŷ, and further is amenable to output tracking.
In short, the KF loop is an ideal candidate for a loopshaping design. Supposing
that we have an estimator gain H which creates an attractive loop function
L(s), we would like to find the compensator C(s) that establishes
It will turn out that the LQR can be set up so that the an LQG-type com-
pensator achieves exactly this result. The procedure is termed Loop Transfer
138 20 LOOP TRANSFER RECOVERY
y ^x
H φ (s) C
Recovery (LTR), and has two main parts. First, one carries out a KF design for
H, so that the Kalman filter loop itself has good performance and robustness
properties. In this regard, the KF loop has sensitivity function S(s) = (I +
Cφ(s)H)−1 and complementary sensitivity T (s) = (I + Cφ(s)H)−1 Cφ(s)H.
The condition σ(W1 (s)S(s)) + σ(W2 (s)T (s)) < 1 is sufficient for robust per-
formance with multiplicative plant uncertainty at the output. Secondly, we
pick suitable parameters of the LQR design, so that the LQG compensator
satisfies the approximation of Equation 215.
LTR is useful as a SISO control technique, but has a much larger role in
multivariable control.
√
lim ( ρK) = W C,
ρ→0
where K is the LQR gain matrix, and W is an orthonormal matrix, for which
W T W = I.
First recall the gain and Riccati equations for the LQR:
K = R−1 B T P
0 = Q + P A + AT P − P BR−1 B T P.
In the limit as ρ → 0, it must be the case that P → 0 also, and so in this limit
20.3 The Loop Transfer Recovery Result 139
ρ(W C)T W C ≈ P BB T P
= (B T P )T B T P
= (R−1 B T P )T RR(R−1 B T P )
= ρ2 K T K −→
√
WC ≈ ρK. 2
• The design plant has no unstable zeros. The LTR method can be in fact
be applied in the presence of unstable plant zeros, but the recovery is
not to the Kalman filter loop transfer function. Instead, the recovered
function will exhibit reasonable limitations inherent to unstable zeros.
See Athans for more details and references on this topic.
The proof of the LTR result depends on some easy lemmas, given at the end
of this section. First, we develop C(s), with the definitions φ(s) = (sI − A)−1
and X(s) = (φ−1 (s) + HC)−1 = (sI − A + HC)−1 .
140 20 LOOP TRANSFER RECOVERY
Next we invoke the result from the LQR design, with ρ → 0, to eliminate
√
ρK:
= (CX(s)B)−1 CX(s)H.
In the last expression, we used the assumption that W is square and invertible,
both properties of orthonormal matrices. Now we look at the product CX(s):
= (Cφ(s)B)−1 Cφ(s)H
= P −1 (s)Cφ(s)H.
• Scale the plant outputs (and references), so that one unit of error in one
channel is as undesirable as one unit of error in another channel. For
example, in depth and pitch control of a large submarine, one meter of
depth error cannot be compared directly with one radian of pitch error.
• Scale the plant inputs in the same way. One Newton of propeller thrust
cannot be compared with one radian of rudder angle.
Proof:
142 20 LOOP TRANSFER RECOVERY
(A + BCD)(A + BCD)−1 = I
A(A + BCD)−1 = I − BCD(A + BCD)−1
(A + BCD)−1 = A−1 − A−1 BCD(A + BCD)−1
= A−1 − A−1 BCD(I + A−1 BCD)−1 A−1
= A−1 − A−1 B(D−1 C −1 + A−1 B)−1 A−1
= A−1 − A−1 B(C −1 + DA−1 B)−1 DA−1 . 2
Proof:
21 SYSTEM IDENTIFICATION
21.1 Introduction
Model-based controller design techniques, such as the LQR and LTR, require
a plant model. The process of generating workable models from observed
data is the goal of system identification. Good controllers usually have some
reasonable robustness guarantees, which motivates identification with simple
methods. We discuss in this section four fundamental but useful techniques
for approximate system identification of single-input, single-output plants. It
should be noted that the area of system identification is a very rich one, and
that the methods are only a small subset of what is available.
Except for the last approach, time-domain simulation, the methods are limited
to linear models. If different inputs give different linear model coefficients, then
it is likely that nonlinear terms are playing a role. The user then has the choice
of ignoring the nonlinearity, for example if the operating point is controlled
very closely, or developing a controller which takes specific account of the it.
In any event, simulations with the nonlinear plant should always be performed
to assess the robustness of the control strategy.
1.5 1
0.8
1
0.6
0.5
0.4
0
0.2
−0.5 0
0 0.2 0.4 0.6 0.8 1 0 1 2 3 4 5
seconds seconds
functions P (s), assuming that the step input was of magnitude one.
The time trace of the first system looks like the Laplace transform pair:
1 1
y(t) = (e−at − e−bt ) ↔ y(s) =
b−a (s + a)(s + b)
for the values b = 0 and a = −1. Thus for the experimental data,
0.58
y(s) = .
s(s − 1)
Since y(s) = P̂ (s)u(s), where u(s) = 1/s, we have P̂ (s) = 0.58/(s − 1). The
second trace looks like a second-order response of the form
y(s) kωn2
= 2 ,
u(s) s + 2ζωn s + ωn2
ωd
ωn = √ ' 3.75rad/s.
1 − ζ2
Inserting these into the template above, we have
7.0
P̂ (s) = .
s2 + 1.28s + 14.1
y(s)
P̂ (s) = . (216)
u(s)
In applications, the input signal u(s) is known quite accurately because it is
generated by a computer. Two sources of error can corrupt the output signal
y(s), however: real disturbances and sensor noise. In some cases, disturbances
may be band-limited (e.g., water waves), and if these occur in a frequency
range far away from the dominant dynamics, the estimated transfer function
approach will succeed. A similar argument holds for sensor noise, which in
many devices is negligible for low frequencies. A sensor which has noise in the
frequency range of the system dynamics is problematic for obvious reasons.
This section discusses the use of periodic inputs to create a Bode plot, while
the next section generalizes to broadband input. Bode plots are figures of
transfer function magnitude and phase as functions of frequency, which can
be either parameterized in terms of poles, zeros, and a gain, or used directly
in a loopshaping or Nyquist plot approach.
Certain plants which cannot admit a step or impulsive input will tolerate a
sinusoidal input. The idea then is to drive the plant with a narrow-band, i.e.,
periodic, input signal u(s). For each such test at a specific frequency, compute
the magnitude and phase relating the input to the output. Conduct as many
tests as are necessary to build a Bode plot of the plant estimate P̂ (s).
Example: Sinusoidal rudder angles trajectories, of 10deg amplitude (|u(t)| =
10deg), were implemented on a vessel operating near its cruising speed. The
data are plotted, along with the magnitude and phase of the transfer function
146 21 SYSTEM IDENTIFICATION
0.093
P̂ (s) = ,
s2 + 0.83s + 0.21
which holds reasonably well at low frequencies. The phase angle of y(s) is taken
with respect to the input signal u(s). Note that the experimental magnitude
and phase in this example deteriorate at the higher frequencies. This is a
property of almost all physical systems, and an indicator that the plant model
cannot be trusted above a certain frequency range.
−10
|y(s)/u(s)|, dB
−20
−30
−40
0
angle(y(s)), deg
−50
−100
−150
−200
−3 −2 −1 0 1 2
10 10 10 10 10 10
rad/s
N −1
x(n)e−j2πnm/N ,
X
X(m) = (217)
n=0
for m = [1, N ]. We review several points for dealing with the DFT calculation:
2π m − 1
ωm = , (218)
dt N
i.e., the first frequency is 0, and the last frequency is slightly less than
the sampling rate.
• The sampling theorem limits our usable frequency range to only one-half
of the sampling rate, called the Nyquist rate. The points higher than
π/dt correspond to negative frequencies, and a complex-conjugacy holds:
X(2) = X ∗ (m), X(3) = X ∗ (m − 1), and so on. What happens near the
Nyquist rate depends on whether N is odd or even.
An additional multiplication of the DFT result by 2 will give peaks which
are of about the right magnitude to be compared with the time-domain
signal. With this scaling, the value X(1) is twice the true DC value.
• The Nyquist rate depends only on the sampling time step dt, but the
frequency vector accompanying the DFT can be made arbitrarily long
by increasing the number of data points. To improve the frequency
resolution of the DFT, a common approach is to zero-pad the end of the
real data,
148 21 SYSTEM IDENTIFICATION
" !#
1 2πk
w(k) = 1 − cos
2 N
• The DFT of a given signal is subject to bias and variance. These can
be reduced by taking separate DFT’s of sub-sections of data (perhaps
zero-padded to maintain frequency resolution), and then averaging them.
It is common to use segments which do not overlap, but it is not a
necessity. This averaging idea is especially important in transfer function
verification; see below.
• If possible, the DFT should be run on a number of samples N which is
a multiple of a large power of two. If N is prime, the DFT will take a
long time to execute.
The above steps, all of which are available through the Matlab function
spectrum() for example, will ensure a fair spectral analysis of a time series.
y(s)
P̂ (s) = . (219)
u(s)
As noted above, spectral analysis of a signal benefits by using multiple seg-
ments and averaging; we now want to include the same approach in our esti-
mation of P (s). The procedure is:
• For the p’th segment of data, compute the transfer function estimate
Yp (m)Yp∗ (m)
P̂p (m) = .
Up (m)Up∗ (m)
• For the p’th segment of data, compute two covariances and a cross-
covariance:
• Construct average values of the transfer function estimate and the other
quantities:
• Compute the coherence function, which assesses the quality of the final
estimate P̂ (m):
If the coherence is near zero, then the segmental cross-covariances Γpyu (m)
are sporadic and have cancelled out; there is no clear relation between
the input and the output. This result could be caused by either dis-
turbances or sensor noise, both of which are reasonably assumed to be
random processes, and uncoupled to the input signal. Alternatively, if
the coherence is near one, then the cross-covariances are in agreement
and a real input-output relationship exists. With real data, the coher-
ence will deteriorate at high frequencies and also at any frequency where
disturbances or noise occur.
• In this example, the same simulation generates the “observed” data and
the simulated response. Since the program simulate always uses the
global variable theta as the parameters, we must be careful about setting
theta in the calling programs.
21.5 Time-Domain Simulation 151
• The initial guess for theta is a random vector; the Simplex method will
take over from this point. In many instances, however, theta is roughly
known, and a better starting value can be given.
• The Simplex method may head into invalid parameter space, e.g., nega-
tive mass. The error calculation, however, can be easily augmented by
a term which penalizes invalid parameters, e.g.,
err = err + 1000*(1-sign(mass)).
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
dt = .3 ; % time step
theta = [1 2] ; % true parameter vector
t_obs = 0:dt:20*dt ; % observed time vector
u_obs = ones(length(t_obs),1) ; % observed input
[t_raw, y_raw] = ode45(’simulate’, [0 max(t_obs)], 0) ;
y_obs = spline(t_raw, y_raw, t_obs) ; % observed output
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
theta = theta_arg ;
[t_raw, y_raw] = ode45(’simulate’, [0 max(t_obs)], 0) ;
y_sim = spline(t_raw, y_raw, t_obs) ;
err = norm(y_sim - y_obs) ;
disp(sprintf(’error: %f.’, err)) ;
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
k = theta(1) ;
tau = theta(2) ;
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
153
22 CARTESIAN NAVIGATION
The bulk of our discussion on maneuvering and control has assumed that the
necessary system states can be measured. The marine engineer is in fact faced
with choices between many different basic sensor packages, notably compasses,
paddle wheels, inertial navigation units, rate gyros, and depth guages, for ex-
ample. These listed sensors are self-contained and rely primarily on the phys-
ical properties of the natural environment. There is also a class of distributed
sensor systems; these generally involve an array of communicating elements, lo-
cated remotely from the vehicle. We present the fundamental concepts behind
two methodologies in this second class: the global positioning system (GPS)
and acoustic navigation, both of which can provide high-accuracy absolute
Cartesian navigation.
T = Tt + 2R/cw ,
where cw is the speed of sound in water, about 1450m/s. The range R follows
by inversion:
(T − Tt )c2
R= .
2
Suppose that the location of B is known; then a given measurement of R
places A on a sphere around B. It is a case of three unknowns (x, y, z) and one
equation:
B C
A3
A2
B C
A1
A*2
A*3
condition, false signals can be caused by reflections off the seafloor and the
surface; these signals are typically eliminated by rejecting those receptions
which are outside a very tight and slowly moving window on travel time T .
Sometimes, bottom topography can shield a direct acoustic path, and the only
receptions available are via multipath!
r + r̃ = c(t + t̃).
The estimated range circles are too large, and the estimated location of the
target too far from the baseline. Introduce a third satellite now so that the
intersections of the circle pairs now occur at six points. Three of these are
very close, and indicate the approximate true solution. The trick is to find the
correction for t̃ that puts the three close solutions onto a single point. This
single point is near the centroid of the three approximate points, and represents
the best position solution. The target time base can be kept up to date by
22.2 Global Positioning System (GPS) 157
performing this check for every set of signals. Hence, the two-dimensional
timing problem is solved with three satellites; the three-dimensional timing
problem is solved with four satellites.
The position specification for GPS is 25m, at the 95’th percentile. This is a re-
markable feat, given that it is an absolute measure over the entire surface and
atmosphere of our 6000km-radius Earth. Major sources of error include: clock
base and satellite navigation (≈ 2.5m), ionosphere and troposphere electro-
magnetic variations (≈ 2.5m), and inaccuracies in the receiver and multipath
(≈ 2m). GPS is also subject to selective availability or S/A, the addition of
a slow-varying random component in the satellite ephemeris data. Selective
availability is controlled by the United States Department of Defense, and de-
grades the position specification to 100m (typical). The advent of differential
GPS solved most of the S/A concerns of American allies, by providing high-
accuracy navigation in local areas. The idea here is that a stationary target
can detect the effects of S/A (since it is not moving) and then transmit the
corrections over a small geographical area. Many current GPS receivers are
capable of decoding these local corrections, which are then applied to their
own satellite navigation processing. Differential GPS typically provides 1-2m
accuracy.
Finally, as with the acoustic navigation systems described, an independent
altitude measurement will enhance the accuracy of GPS, essentially reducing
a three-dimensional to a two-dimensional problem.
158 22 CARTESIAN NAVIGATION
159
23 REFERENCES
Short notes indicate how the reference is related to the text.
Optimal Spacecraft Rotational Maneuvers, J.L. Junkins and J.D. Turner, El-
sevier: New York, 1986. Euler angles and quaternions.
Guidance and Control of Ocean Vehicles, T.I. Fossen, Wiley: New York, 1994.
Comprehensive modern treatment of marine dynamics and control.
Formulas for Natural Frequency and Mode Shape, R.D. Blevins, Krieger: Mal-
abar, FL, 1984. Extensive added mass formulas.
Fluid Dynamic Lift, S.F. Hoerner and H.V. Borst, Hoerner Fluid Dynamics:
Bakersfield, CA, 1975. Extensive data on lift of bodies and wings.
Fluid Dynamic Drag, S.F. Hoerner and H.V. Borst, Hoerner Fluid Dynamics:
Bakersfield, CA, 1975. Data on drag of bodies and wings.
incorporation into control system design, D. R. Yoerger, J.G. Cooke, and J.-
J. E. Slotine, IEEE J. Oceanic Engineering, 15:167-178, 1990. First-order
nonlinear thruster model.
Governing ship propulsion gas turbine engines, C.J. Rubis and T.R. Harper,
SNAME Trans., 94:283-308, 1986.
Circuits, Signals, and Systems, W.M. Siebert, MIT Press: Cambridge, MA,
1986. General reference on transforms.
Feedback Control Theory, J.C. Doyle, B.A. Francis, and A.R. Tannenbaum,
Macmillan: New York, 1992. Loopshaping concepts.
Optimal Control, F.L. Lewis and V.L. Syrmos, Wiley: New York, 1995. Gen-
eral reference on LTR.
Numerical Recipes in C, 2nd Edition, W.H. Press, S.A. Teukolsky, W.T. Vet-
terling, and B.P. Flannery, Cambridge University Press: Cambridge, UK,
1992. Function minimization.
NAVSTAR: Global positioning system – Ten years later, B.W. Parkinson, S.W.
Gilbert, Proc. IEEE, 71:1177-1186, 1983. The whole issue is in fact dedicated
to global positioning systems. There is another such issue: IEEE Proc. Vol.
87, No. 1, 1999.
162 23 REFERENCES
163
24 PROBLEMS
Many of the following problems use data presented by Fossen and in Principles
of Naval Architecture.
φ
y
π/4 φ
the condition for stability is that A, B, and C must have the same
sign. Express this requirement in terms of the derivatives in the
previous question. Give physical interpretations for what would
make such a device stable or unstable.
(d) Create a numerical model of this system, using the MATLAB ODE
solver ode45. The system equation can be written as two first-order
equations:
( ) " #( )
d θ̇ −B/A −C/A θ̇
= .
dt θ 1 0 θ
O θ
x
Uο
165
F F F
O O
u u u
O A A A
mẍ + bẋ + kx = 0,
wherein all the coefficients are positive and real (i.e., physical).
(a) Set x(t) = est , and find two solutions for s in terms of m, b, and k,
using the quadratic formula. The solution pairs depend on whether
b exceeds a critical value: what is the critical value and how do the
solutions change?
(b) From the initial condition x(0) = xo , and ẋ(0) = 0, find the coeffi-
cients for the general solution
s
k
ωn = (undamped natural frequency)
m
b
ζ = √ (damping ratio)
2 km
q
ωd = ωn 1 − ζ 2 (damped natural frequency).
mẍ + bẋ + kx = u.
(c) If the speed suddenly drops to 2.5m/s, how does the range of sta-
bilizing k1 change?
7. Steering. Draw curves of rudder angle δ vs. yaw rate r for a dynamically
stable surface vessel, and then for an unstable vessel. Indicate areas
where the vessel turns against the rudder action.
(a) Derive the equations of motion for the submarine moving at speed
U.
(b) Derive the hydrostatic, restoring pitch moment for small pitch θ.
(c) Linearize the inertial terms in the equations of motion.
(d) Expand the fluid forces and moments in terms of the motions. Omit
nonlinear and memory effects, and be sure to include the hydrostatic
moment. Explain your choices.
(e) Write out the complete linearized equations of motion. Does surge
decouple from pitch and heave? Write out the coupled equations in
matrix form.
(f) Can the submarine be stable in pitch without feedback control?
Can the submarine be stable in depth without feedback control?
the radius is 0.5m, length is 17m, and speed is 2m/s, use slender-
body theory to express the following coefficients: Zẇ , Mq̇ , and Mw ,
all referenced to the body midpoint.
(b) It turns out that in order to match the lift and moment of the hull
in experiments, a flat tail of radius 0.4m has to be included in the
slender-body estimate. Determine the corrected value for Mw , and
in addition compute Zw . Where is the aerodynamic center with
respect to the center of the hull, and is this hull stable in pitch?
Recall that the aerodynamic center is the equivalent location of the
observed lift force that creates the observed moment.
(c) A fin with pre-determined lift slope of 3.6/rad is to be applied at
the tail to make the device just stable in pitch. How big is the
required fin area that brings the net Mw to zero, as desired?
10. Lifting surfaces. Using the lifting surface formulas, estimate the coef-
ficient Kv of a submarine sail as shown below, for U = 10m/s.
7m
10m
15m
11. Slender body theory. Consider a long ellipsoidal body of length l and
diameter d.
(a) With l/d = 7.0, approximate the cross-body added mass, using
slender-body theory, and compare it with the exact results from
the table below (Blevins).
(b) Perform the slender body calculation also for a sphere, and compare
again to the exact result.
(c) What is the added mass in the in-line direction?
169
2b
2a
is appended with two fins as shown in the figure: each has an area of
1m2 , geometric aspect ratio of 3, and has its longitudinal center of action
1m forward of the body stern.
(b) Apply slender body theory to estimate Mw for the body alone.
(e) Use the experimental data from Hoerner (p. 13.2) below to estimate
Mw for the body.
(f) Use the Hoerner result and your result from wing theory to write a
net Mw for the body plus fins.
(h) What is the diameter of a flat stern that will allow the slender body
theory to give the same moment as the Hoerner (experimental)
data, for the body alone?
(i) What is the diameter of a flat stern that allows the approximate
formulas of Jorgensen to give the same moment as the Hoerner data,
for the body alone?
Hoerner p. 13.2:
Symmetric body of revolution with L/D = 6.7.
Force and moment referenced to body midpoint:
Z = −0.5ρU 2 D2 Cydb tan(w/u). Cy = 1.20.
M = 0.5ρU 2 D2 LCnb tan(w/u). Cm = 0.53. DESTABILIZING
171
L/D = 7
x y
L = 35m
14. Cable mechanics (hard!). You are asked to assess the operational
envelope of a cable/vehicle system which has been installed on a vessel.
The cable is steel-jacketed, with diameter 3cm, an effective extensibility
of 100 × 109 P a, and a density of 5000kg/m. The vehicle is streamlined,
and has a mass (material plus added) of 100kg.
(a) The towing angle at the vessel cannot exceed 25deg from vertical,
for reasons of deck and crane geometry. If this angle is considered
equal to the critical angle φc , what is the maximum speed for the
system? Assume the normal drag coefficient of the cable is Cn =
1.2.
172 24 PROBLEMS
∂2q ∂2q ∂q
(m + ma ) = T̄ + wn sin φ̄ ,
∂t2 ∂s ∂s
where ma is the added mass of the cable per unit length, and wn is
the in-water weight of the cable. Separation of variables q(s, t) =
q̃(s) cos ωt gives
∂ 2 q̃ ∂ q̃
T̄ + wn sin φ̄ + wn γ q̃ = 0,
∂s2 ∂s
where γ = (m + ma )ω 2 /wn . Simplify this expression by writing
T̄ = wn s (the vehicle weight effect is small), sin φ̄ ' 1 (the cable is
√
nearly vertical), and then use the substitution z = 2 γs to arrive
at the Bessel equation
∂ 2 q̃ ∂ q̃
z2 2
+ z + z 2 q̃ = 0.
∂z ∂z
The derivatives can be transformed from s to z coordinates via the
chain rule:
∂ q̃ ∂ q̃ ∂z
= ·
ds ∂z ∂s
!2
∂ 2 q̃ ∂ 2 q̃ ∂z ∂ q̃ ∂ 2 z ∂z
= + · · .
ds2 ∂z 2 ∂s ∂z ∂z∂s ∂s
(d) The equation above has a solution of the form q̃ = cJ0 (z), where
c is a constant to be found, and J0 (z) is the zero’th order Bessel
function of the first kind. If q̃(s = L) = Q, that is, we impose a
harmonic input at the top, find c in terms of Q and z(s = L). What
condition makes q̃ blow up, indicating a resonance frequency ωn ?
173
" # ( )
−0.90 −0.42 −0.13
A= , B= .
−4.8 −2.3 1.4
r0 (s) s + z1
=K ,
δ(s) (s + p1 )(s + p2 )
where K is a constant gain, −z1 is the zero (there is one real zero),
and [−p1 , −p2 ] are the two poles of the second-order system. Are
both poles stable?
(b) For the purposes of autopilot design, the transfer function φ(s)/δ(s)
is crucial. Fortunately, all you have to do is divide r0 (s)/δ(s) by an-
other factor of s to account for the fact that dφ/dt0 = r0 . Sketch all
the poles and zeros of the plant transfer function P (s) = φ(s)/δ(s)
on a root locus plot.
174 24 PROBLEMS
(c) A PID controller C(s) will work to stabilize the heading angle,
and brings to P (s)C(s) one additional pole at the origin, and two
arbitrary zeros. Demonstrate a rough PID design by devising two
controller zero locations that will attract the closed-loop poles into
the left-half plane. Draw the path that you expect each closed-loop
pole to follow, as the control gain increases from zero to very large
values.
You do not need to calculate the PID gains that go with your pro-
posed zero locations.
0
Ixx = 0.000118 Mq0 = −0.0113 Zq0 = −0.0175
0
Izz = 0.00193 Mq̇0 = −0.00157 Zq̇0 = −0.000130
x0G = 0 Mw0 = 0.0112 Zw0 = −0.0439
m0 = 0.0364 Mẇ0 = −0.000146 Zẇ0 = −0.0315
U = 2.0m/s Mδ0 = −0.0128 Zδ0 = −0.0277
L = 15.0m Mθ0 = −0.156/U 2
17. Nyquist stability. The inverted pendulum shown below is often used
as a simple model for rocket flight, and can also illustrate the dynamic
behavior of an unstable ocean vessel which is propelled from the stern,
e.g., a barge being pushed by a tugboat.
For this problem, we assume that all of the mass (m) is concentrated
at the distal end of an arm of length l; a high-performance positioning
system sets the horizontal position of the cart uo .
g l
uo
g 1
φ̈ = sin(φ) − cos(φ)üo .
l l
Note that the acceleration of the cart is now considered to be the input
to the plant, e.g., u = üo .
(b) Show that this plant is fully state observable and controllable.
(c) Now make the assumption that l = 1, and g = 1 (it can be achieved
through a proper nondimensionalization) What is the transfer func-
tion of the plant, P (s)? Use this plant model for the rest of this
question.
(d) Where are the plant poles and zeros, if any?
(e) A stabilizing controller is not hard to find, and a suggested one is:
−4(2s + 1)
C(s) = .
s+2
Create a rough Nyquist plot of the loci of P (s)C(s), for the frequen-
cies ω = [0, 2.2, ∞]rad/s. These three frequencies are all that you
need to sketch the overall shape; there are no hidden loops or other
features. In the case of ω = 2.2, for which you will need to make
explicit calculations, you may find the following identity useful:
18. Root locus and loopshaping. The parameters governing the surface
maneuvering of a high-speed container ship are given below for reference:
m0 0.00792 0
Izz 0.000456 x0G −0.05
L 175.0m U 8m/s Yv̇0 −0.00705
Yṙ0 0.0000 Yv0 −0.0116 Yr0 0.00242
Yδ0 −0.00258 Nv̇0 0.0000 Nṙ0 −0.000419
Nv0 −0.00385 Nr0 −0.00222 Nδ0 0.00126
177
Note that the center of vessel mass is located aft of the origin; for this
model, the origin coincides with the center of added mass, so that Yṙ =
Nv̇ = 0. The nondimensional system with states ~x0 = [v 0 , r0 ] evolves
according to d~x0 /dt0 = A~x0 + Bδ, where
" # ( )
−0.90 −0.42 −0.13
A= , B= .
−4.8 −2.3 1.4
(a) Use the Matlab command tf(), or ss(), to create a system model
of the open-loop transfer function P (s)C(s), using the plant above
and a PID-type controller:
1
C(s) = kp 1 + τd s + .
τi s
" #
0 −1
A = ,
1 0
B = [1 0]T
C = [0 1]
D = 0
Q = CT C
R = ρ.
(a) What is the control gain K in terms of ρ? Hints: There are two
solutions for p12 ; choose the positive one. Also, the expression for
p22 is messy; luckily, you won’t need to use it.
179
(a) Solve the LQR Riccati equation for P , for the case Q = C T C and
R = ρ. Your answer should give a positive P as a function of ρ
alone.
(b) Consider the cheap-control problem, with ρ → 0. Write the leading-
term approximations for P and then K, and compute the eigenvalue
(there is only one) of the closed-loop system. Note that eig(M ) =
M , if M is a scalar instead of a matrix.
(c) Now look at the expensive-control problem, with ρ → ∞. Again,
write the approximations for P and K, and find the closed-loop
system eigenvalue.
21. LQG/LTR. For the inverted pendulum plant (see previous question)
with state-space matrices
" #
0 1
A= , C = [0 1] ,
1 0
(a) Compute the open-loop transfer function for the Kalman filter loop:
L(s) = C(sI − A)−1 H.
180 24 PROBLEMS
(b) Even though this L(s) is unstable, its magnitude plot confirms that
it is a reasonable loopshaping design. What is the low-frequency
gain for this L(s), and what kind of tracking performance should
we expect from the associated LTR design?
(c) Now consider the closed-loop transfer function, which you can write
as S(s) = L(s)/(1 + L(s)) (the sensitivity). What is the character-
istic equation for the closed-loop system, and about what damping
ratio has the Kalman filter provided?
22. LQG/LTR design. The parameters for the linearized sway/yaw mo-
tions of a swimmer delivery vehicle are given below.
You are asked to develop an LQG/LTR controller for this plant, and it is
suggested that you compose a single Matlab script to perform the steps
in sequence. Please make sure you answer all the questions, and include
a listing of your code. This entire design is made in nondimensional
coordinates.
i.e., the input to the closed-loop system is ~xdesired and the out-
put is ~x. Your plot should show specifically the output φ, for an
input of ~xdesired = [vdesired = 0, rdesired = 0, φdesired = 1]. This
compression can be achieved in one step by premultiplying the
system by C T , and post-multiplying it by C:
ii. The Matlab command lqe() can be used to generate the Kal-
man filter gain H, given design matrices A, C, V1 , and V2 . For
the choices V1 = I3×3 and V2 = 0.01, compute H, and make
a plot of the closed-loop step response. Be sure to give the
numerical values of H.
182 24 PROBLEMS
Note that the lqe() command asks for a disturbance gain ma-
trix G; you should set this to I3×3 . The closed-loop KF system
is as follows:
x̂˙ = (A − HC)x̂ + Hy
ŷ = C x̂,
i.e., the input is the measurement y and the output is an esti-
mated version of it, ŷ.
(c) Loop Transfer Recovery
The LQG compensator is a combination of the KF and LQR designs
above. With normal negative feedback, the compensator C(s) has
the following state space representation:
~z˙ = (A − BK − HC)~z + He
u = K~z,
so that the input to the compensator is the tracking error e =
r − y, and its output u is the control action to be applied as
input to the plant. The total open-loop transfer function is the
P (s)C(s); in Matlab, you may simply multiply the systems, e.g.,
sysPC = sysP * sysC ;.
i. Make a log(magnitude) plot of the KF open-loop transfer func-
tion L(s) = C(sI − A)−1 H, versus log(frequency). You may
find the Matlab command freqresp() helpful. |L(s)| should
be large at low frequencies, and small at high frequencies, con-
sistent with the rules of loopshaping.
ii. As ρ → 0, the product P (s)C(s) → L(s). Demonstrate this by
computing P (s)C(s) for the three different values of ρ above,
and overlaying the respective |P (s)C(s)| over the plot of part
3a).
iii. Make a closed-loop step response plot for the smallest value of
ρ. How does it compare with the KF step response of part 2b)?
(a) Ignoring the coupling of Euler small angles, what estimates can
you give for the roll and pitch angles? I suggest that you draw the
measurement vector in the coordinates of the vehicle frame, and
then consider the orientation of the body which would lead to it.
(b) Given that the acceleration due to gravity is 9.81m/s, how do we
know that these are reasonable (but not foolproof!) angle calcula-
tions to make?
184 24 PROBLEMS
24. LQG/LTR design: term project. You are asked to carry out mod-
eling and control system design for the surge dynamics of an oceanic
cable-laying vessel. This vessel needs to have very good speed control
so that the cable, which is being paid out constantly and sinks slowly
under its own weight, lies properly on the seafloor. The bathymetry
for the path is well-known but variable, so the vessel will need to slow
down and speed up frequently. Unexpected deviations in speed from the
calculated trajectory are likely to create loops or kinks in the cable, or
induce large tensions.
! !
Qe f ne f
=− a +b + c +d ,
Qm fm nm fm
(a) Make a map of the gas turbine characteristics. For instance, make
a contour plot which gives Qe /Qm (the fraction of maximum torque
developed) as a function of f /fm (the fraction of the maximum fuel
rate) and ne /nm (the fraction of the maximum rotation rate of the
engine).
Be sure to include clipping on your contours, where the calculated
torque exceeds Qm . Similarly, the engine cannot develop negative
torque.
(b) Make a table of some steady operating conditions for fo /fm in the
range of 0.05–0.95, in which you show: advance ratio seen by the
propeller, engine torque, propeller rotational speed, engine rota-
tional speed, and vessel linear speed.
Plot the pairs [fo /fm , neo /nm ] on the characteristic plot from Part
1, and assess whether the vessel, engine, and propeller are well
matched to operate over a range of fo .
(c) Construct a linear approximation for the plant dynamics around
the operating condition fo /fm = 0.8. If the plant input is δf , and
the output is δu, list the A, B, C matrices for your plant, and the
eigenvalues. δ here indicates the perturbed value from the steady
state, e.g., f = fo + δf .
186 24 PROBLEMS
(d) Create an LTR-type controller for the plant. First, choose the filter
gains V1 and V2 to make the crossover frequency (where the magni-
tude of L(s) = C(sI −A)−1 H passes through unity) about 0.6rad/s.
This choice corresponds with the closed-loop system being able to
overcome waves with period 10seconds or longer. Note that for this
scalar design, you will be able to move the curve |L(s)| up or down
on the Bode plot, but you cannot easily change its shape, or move
it side-to-side.
Secondly, recover this loop shape with an LQG-type controller, us-
ing a small control penalty.
You should prepare for this part: a plot of log(|L(s)|) versus log(ω),
a listing of your V1 and V2 choices, and the A, B, C matrices of your
LQG compensator.
(e) In the event of a current or wind disturbance, the actual vessel
speed will vary from the nominal value. This is not acceptable in
the long term, given that the vessel is laying out a cable.
An integral action can be added quite easily in the LTR design
technique. First, add an integrator to the input of the plant, so
that an “augmented plant” is created. This augmented plant has
one more state than the original plant:
" #
A B
Aaug =
0 0
Baug = [0 1]T
Caug = [C 0]
The idea is to then carry out the KF design and LTR as before, with
the augmented plant model. When this is done, move the integrator
from the augmented plant into the compensator. The addition of an
integrator channel to the plant or compensator can be accomplished
with the command sysPaug = sysP * tf([0 1],[1 0]) ;, for ex-
ample.
Present a figure of |L(s)| and the other quantities as you gave them
in Part 4. This is to be kept as a separate controller design.
(f) For the purpose of demonstrating your controllers, construct a non-
linear simulation of the plant system, with input f . You don’t
187