A Mathematical Model For The Analysis and Control PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/267919339

A mathematical model for the analysis and control of ballast tank blowing and
venting operations in manned submarines

Conference Paper · September 2011

CITATIONS READS

2 351

3 authors:

Roberto Font José Alberto Murillo Hernández


Biometric Vox Universidad Politécnica de Cartagena
18 PUBLICATIONS   104 CITATIONS    15 PUBLICATIONS   51 CITATIONS   

SEE PROFILE SEE PROFILE

J. García Peláez
Navantia
35 PUBLICATIONS   58 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Non-Acoustic Submarine Signatures Control: Deployables Magnetic Anomaly Detection Maskers View project

All content following this page was uploaded by Roberto Font on 15 January 2015.

The user has requested enhancement of the downloaded file.


A mathematical model for the analysis and
control of ballast tank blowing and venting
operations in manned submarines
Roberto Font, José Alberto Murillo

Dpto. de Matemática Aplicada y Estadı́stica, Universidad Politécnica de


Cartagena

roberto.font@upct.es,alberto.murillo@upct.es

Javier Garcı́a

Departamento de Ingenierı́a, Navantia S.A.

jgpelaez@navantia.es

Abstract

The blowing and venting of ballast tanks could potentially be used as a


complementary control system to improve the manoeuvrability of manned sub-
marines. To test the feasibility of such a control system, we first propose a
mathematical model for these operations. Then, the model is applied to the
control of an emergency manoeuvre by using only the blowing and venting of
ballast tanks as control mechanism. To this end, we formulate a suitable con-
strained, nonlinear, optimal control problem where controls are linked to the
variable aperture of blowing and venting valves of each of the tanks. The state
law is composed of a system of nonlinear differential equations where the equa-
tions modeling blowing and venting processes are coupled with the Feldman,
variable mass, coefficient based hydrodynamic model for the equations of mo-
tion. Numerical results seem to indicate that, indeed, an appropriate use of
blowing and venting operations may help in the control of an emergency ma-
noeuvre.
Sección en el CEDYA 2011: MAI

1 Introduction
Ballast tanks are distributed along the length of the submarine. When filled
with water, they contribute with the submarine mass, allowing it to submerge.
During an unexpected event or emergence, air is blown into the ballast tanks
from very high pressure bottles, expelling the water out of the tanks. The
submarine loss weight, its buoyancy is higher, and it can emerge quicker. To
fill the tanks with water again, air is vented out of the ballast tanks.
Although the tanks are usually blown only on emergency situations, blowing
and venting could potentially be used to improve the vehicle manoeuvrability.
These operations are currently performed based exclusively on the operator
experience and, due to the high degree of accuracy required, would enormously
benefit from the implementation of a control system. To this end, first, accurate
mathematical models for both processes are needed.
In Section 2 we propose a model for a coupled system of blowing-venting
operations. Then, these two processes are coupled with the usual Feldman’s
coefficient based hydrodynamic model for the equations of motion (see [4] and
more precisely [8]). In Section 3 we formulate a constrained, nonlinear, optimal
control problem which has the aperture of blowing-venting valves of each of the
tanks as the control variables. Finally, Section 4 presents simulation results
obtained by the proposed optimal control scheme in an emergency rising ma-
noeuvre in which the roll angle must be minimized. Results seem to indicate
that blowing-venting operations may help in a significant way to the control of
this type of manoeuvre.

2 Mathematical modeling
The aim of this section is to provide mathematical models for both blowing and
venting processes as well as their influence over the behavior of the submarine.
Our mathematical model for the equations of motion is based on Feldman’s [4]
six degree of freedom (DOF) submarine equations of motion adapted to the
particular characteristics of a prototype developed by the company Navantia
S.A. Cartagena Shipyard (Spain). We refer the reader to [8] for a complete
description of the resulting model.

2.1 Blowing and venting model


The blowing/venting system is composed of ballast tank, pressure bottle, blow-
ing valve and venting valve. When the blowing valve is opened, air flows into
the tank increasing the pressure and forcing the water to flow out through the
flood port located at the bottom of the tank. When the venting valve is opened,
air can flow out from the tank. The model can be divided into four main parts:
Air flow from pressure bottle, air flow through venting valve, water flow through
flood port and evolution of pressure in ballast tank.

2.1.1 Air flow from pressure bottle


When the blowing valve is opened, the air in the bottle is blown into the tank
through a nozzle. At the beginning of blowing, the flow is supersonic due to
the high pressure difference between bottle and tank. As air flows out of the
bottle, this difference decreases and the flow becomes subsonic below the critical
( ) γ−1
γ
pressure ratio Pc = γ+1 2 , with γ the isentropic constant.

Supersonic flow: In this case, it is easy to calculate the mass flow rate at
the nozzle throat, where the Mach number is unity. Precisely,
ρ∗a a∗
ṁF = ρ∗a v ∗ A = ρa,0 a0 A, (1)
ρa,0 a0
where A is the area in the nozzle throat, a is the speed of sound, v is the air
velocity, ρa is the density of air, the asterisk means conditions wherein the Mach
number is equal to unity and the subscript 0 denotes stagnant conditions.
Assuming that in this particular case stagnant conditions are the bottle
conditions, using the relations √ between stagnant and static conditions, the ideal
gas law ρa,0 = Rpg0T0 and a0 = γRg T0 , the mass flow rate is

( ) γ−1
γ+1
ApF (t) 2
ṁF (t) = √ γ , (2)
Rg TF (t) γ+1

where γ is the isentropic constant, Rg is the gas constant for air and TF is the
temperature in the bottle.

Subsonic flow: In this case, the mass flow rate is given by


ρa,e ae
ṁF = ρa,e Me ae Ae = ρa,0 a0 Me Ae , (3)
ρa,0 a0

where Ae is the exit area, in this case Ae = A, M is the Mach number and
the subscript e denotes conditions at the exit. The Mach number at the exit is
given by v
u [( ) γ−1 ]
u 2 p
Me = t
γ
F
−1 . (4)
γ−1 pB

Substituting (4) into (3) and proceeding as in the supersonic case, we get
v [( ]
u ) γ2 ( ) γ+1
ApF (t) ut 2γ pB (t) pB (t) γ
ṁF (t) = √ − . (5)
Rg TF (t) γ − 1 pF (t) pF (t)

As the pressure in the bottle drops, the temperature increases. If the heat
transmission is neglected, however, the process can be considered to be adi-
abatic. Let mF 0 , pF 0 be the initial mass and pressure in the bottle and VF
the bottle volume. Under the above assumptions the momentary pressure and
temperature are given by
( )γ
mF (t) pF (t)VF
pF (t) = pF 0 , TF (t) = . (6)
mF 0 Rg mF (t)

Let si ∈ L∞ (0, tf ; [0, 1]) denote the aperture of the blowing valve of the
i−th ballast tank during the time interval [0, tf ]. Thus, the final equation for
the mass flow rate from bottle i is
( ) 12
mF i (t)γ+1 pF 0
ṁF i (t) = si (t)A µi (pBi (t), mF i (t)) (7)
mγF 0 VF
with
 √
 ( ) γ+1

 2 γ−1
Pc ≤
pF

 γ ,


γ+1 pB





 v 
 u
u 2   γ+1 
µ(pB , mF ) = u 2γ  p
γ
p
γ
 pF

 u  (B )γ  − (B )γ  ,
 t

1< < Pc

 γ−1 pF 0 mF
pF 0 mF pB

 mF 0 mF 0



 pF

 0, ≤1
pB

2.1.2 Air flow through venting valve


Equations for the air flow from the tank can be obtained analogously to the ones
used for the mass flow from the bottle. In this case, the pressures to evaluate
are the pressure in ballast tank, pB , and the pressure outside the vent valve
pipe system, pext . Let Av be the vent pipe section, similarly to (7), we have
( )
Av pB (t) pext pext
ṁv (t) = µ (pB , pext ) , µ = µmax if ≤ (8)
Rg TB (t) pB pB crit

Although the geometric details of the venting system are not available and
therefore pressure losses can not be directly calculated, an evaluation of µ (pB , pext )
for several values of the ppext
B
ratio taking into account the different pressure
losses along the system has been provided. Let Π = ppext B
be the pressure ratio.
A expression for µ(Π) was obtained used a least squares fit of these data.
The variation in the mass of air in the ballast tank is the difference between
the mass flow rate from the bottle and the mass flow rate through the venting
valve. Let si ∈ L∞ (0, tf ; [0, 1]) be the aperture of venting valve of the i−th
ballast tank, then from (8):

si (t)Av pBi (t)


ṁBi (t) + ṁF i (t) = −µi (Πi (t)) √ , (9)
Rg T B

2.1.3 Water flow through flood port


The difference between the tank and outside pressure forces the water to flow
in or out from the tank trough the flood port located at the bottom. Let us
assume, in the first place, that the pressure in ballast tank is greater than the
outside pressure. Then, the water flows out from the tank. The pressure directly
above the flood port is pB (t) + ρghwc (t), where hwc (t) is the height of the water
column in the tank. Assuming that the tank has a regular shape, the height of
the water column can be expressed as
( )
∆m(t)
hwc (t) = 1 − Htk ,
∆mmax
where Htk is the ballast tank height, ∆m(t) is the mass loss in the tank and
∆mmax is the maximum value of the mass loss (see Section 2.2 for more details).
The Bernoulli equation applied at both sides of the port gives
1 1
pB (t) + ρghwc (t) − ∆ploss = pSEA (t) + ρvh2 (t), ∆ploss = ζh ρvh2 (t)
2 2
where vh and ∆ploss are, respectively, the velocity and pressure losses in the
outlet hole. Both major and minor losses have been calculated by a separate
model and can be compressed into one single term.
This way the volume flow from the ballast is given by

2(pB (t) + ρghwc (t) − pSEA (t))
qB (t) = Ch Ah vh (t) = Cn Ah , (10)
ρ(1 + ζh )

where pSEA is the pressure outside the flood port, Ah is the outlet hole area and
Ch is a coefficient that takes into account that, since the outlet hole is actually
a grid, the effective area is smaller than Ah .
If pressure in ballast tank is lower than the outside pressure, then water
flows into the tank. In this case, the expression for the volume flow is (10), but
with a negative value.

2.1.4 Evolution of pressure in ballast tank


When the blowing valve is opened the air is blown into the tank at a very high
velocity, rapidly mixing with water. This promotes good heat transfer from the
water to the expanding air and we can work under the assumption that the air
will immediately adopt the temperature in the tank. The process can then be
considered to be isothermal. As mentioned in [2], experimental results sustain
this assumption.
From the ideal gas law it follows that

ṁB (t)Rg TB − V̇B (t)pB (t)


ṗB (t) = .
VB (t)

The volume occupied by air increases (and in the same quantity) as the
volume occupied by water decreases. Thus, the rate at which it changes, V̇B , will
be the volume flow out of the ballast tank, qB , given by (10). Using the perfect
mB (t)Rg TB
gas equation, the momentary volume of air in the tank is VB (t) = .
pB (t)
The variation in the i−th tank pressure is then

mBi (t) pBi (t)qBi (t)


ṗBi (t) − ṁBi (t) = − . (11)
pBi (t) R g TB
2.2 Coupling of blowing-venting system with a variable
mass model for the equations of motion
As water flows in or out of the tanks there will be mass variations located at
several points of the vehicle. Since the equations of motion assume the mass of
the submarine to be constant, it is necessary to identify which terms, formerly
constant, will become time dependent due to its dependance with mass. We
will need to write mass, moments and products of inertia and the location of
center of gravity as a function of the amount of water in the tanks.
Let m0 be the initial mass of the submarine (with all tanks completely filled
with water) and ∆mi is the mass loss in the i−th tank. It is 0 when the tank is
completely filled with water, and reaches its maximum value when it empties.
The volume of water that has left the tank is equal to the volume occupied
by air except for the initial air volume in the tank, VB0 , which depends on the
initial mass of air in the tank, mB0 , and the initial depth. The mass loss in the
i−th tank is then,
( )
mBi (t)Rg TB
∆mi (t) = ρ (VBi (t) − VB0 ) = ρ − VB0 . (12)
pBi (t)

The momentary mass of the submarine can be expressed as:


N
m(t) = m0 − ∆mi (t). (13)
i=1

The rest of parameters are calculated using an analogous reasoning.

3 Formulation of the control problem


Let x(t) be the vector of state variables (mFi (t), mBi (t), pBi (t) for each of the i
ballast tanks, position, Euler angles and linear and angular velocities of the ve-
T
hicle), and u(t) = [si (t), si (t)] the control vector, si and si being, respectively,
the aperture of blowing and venting valves of the i-th tank. The state law is
composed of equations (7), (9) and (11) for the blowing-venting system, the
six kinematic equations and the six equations of the hydrodynamic forces and
moments (see [4] and [8]), where the time dependent parameters as described in
Subsection 2.2 have been taken into account). In compact form we write these
equations as
A(x(t))ẋ(t) = f (t, x(t), u(t)) , (14)
We are now in a position to formulate the manoeuvrability control problem.
Given an initial state x (0) = x0 and a desired final target xtf , the goal is to
calculate the vector of control u = u(t), which is able to draw our system from
the initial state x0 to (or near to) the final one xtf in a given time tf , also
minimizing a cost functional. In mathematical terms we have the Bolza-type
problem
 ∫


tf

 Minimize in u : tf
J (u) = Φ (x(tf ), x ) + F (t, x (t)) dt




0

 subject to
A(x(t))ẋ(t) = f (t, x(t), u(t)) (15)





 x(0) = x0 ∈ Ω




0 ≤ si (t), si (t) ≤ 1, 1 ≤ i ≤ N, and x (t) ∈ Ω,

where Ω stands for the set of constraints for the state variable.
We refer to [6] for a more detailed description as well as a detailed analysis
of the well-posedness of (15).

4 A numerical experiment
The proposed models have been implemented with the help of the software
Matlab. At each iteration of the gradient algorithm, the numerical resolutions
of the state and adjoint state equations have been carried out by using the
ODE45 function, which is a one-step solver based on an explicit Runge-Kutta
method.
On an emergency situation, like on board fire or flood, submarines may need
to rise to the surface quickly. The typical protocol for this situation is to use the
control planes to pitch the nose of the submarine up, increase the speed, and
blow the ballast tanks to reduce the weight of the submarine and drive it to the
surface with buoyancy. As analyzed in [1, 9], small to medium size submarines
exhibit a roll instability during these kind of manoeuvres. Full scale trials
showed that submarines rose with roll angles up to 25 degrees. It is also known
that if the the submarine emerges with a high roll angle, it may experiment
large roll oscillations on the surface. Of course, this is an undesirable situation,
particularly if the operators are attending to the original problem that required
the emergency rise.
For these reasons, it might be interesting to check if this situation can be
prevented by using our control algorithm. To this end, and emergency rising
manoeuvre has been simulated for two different scenarios:
1) A standard manoeuvre. Initial depth is 100 m, initial speed is 2 m/s. At
t = 0 the stern and bow planes are set to −20 and 20 degrees respectively
and the propeller is set to 150 rpm. At the same time, ballast tanks 2–5 are
simultaneously blown with half the maximum intensity (this corresponds
to si = 0.5 in our model). Vent valves remain closed (si = 0) throughout
all the simulation.
2) Same manoeuvre, with the control algorithm acting from t = 0 to t = 30
s using Scenario 1 as initialization and looking to achieve three main
objectives: a) Submarine must rise in a similar time as it does in the
Figure 1: Depth, pitch and roll angle for scenarios 1 (dashed line) and 2 (solid
line).

standard manoeuvre, b) Rising pitch angle must be around 20 degrees and


never above 25 degrees and c) Roll angle must be as close as possible to
zero throughout all the simulation. After t = 30 s, simulation continues
with fixed controls (values equal to those used in Scenario 1) until the
submarine reaches a depth of 10 m.
Results for Scenario 1 (dashed lines) and Scenario 2 (solid lines) are shown
in figures 1 to 3. Figure 1 shows vehicle depth, pitch angle and roll angle for
both scenarios. The rest of state variables have not been included since they
are not directly relevant for this particular manoeuvre. Comparison between
dashed (standard manoeuvre) and solid (optimal controls) lines shows that the
three objectives have been achieved for Scenario 2: the rising time is only a few
seconds greater than in the standard manoeuvre, the final pitch angle is close to
20 and below 25 degrees and the roll angle has been significantly reduced with
respect to the standard manoeuvre. Indeed, Scenario 1 exhibits roll angles in
the range of 3–4 degrees during most of the simulation time while in Scenario 2,
after an initial peak of 2 degrees, the roll angle is kept within extremely low
values during the whole t = [0, 30] interval.
The optimal (solid lines) and standard (dashed lines) controls are shown in
Figure 2 (blowing valves) and Figure 3 (venting valves). Although the conve-
nience of the use of venting during an emergency rise can be arguable and is
indeed not an usual practice, it is used here to demonstrate the algorithm ca-
pabilities. Very similar, although slightly worse results, were obtained by using
only the blowing valves as control. As we can see in Figure 2, the rolling mo-
ment is compensated by blowing more ballast from the starboard tanks. It may
surprise that tanks 2 and 4 are being vented while there is no blowing air yet,
Figure 2: Blowing valve aperture for standard (dashed lines) and optimal (solid
lines) controls.

Figure 3: Vent valve aperture for standard (dashed lines) and optimal (solid
lines) controls.
but this seems to allow a smoother transition when the two valves are simul-
taneously open. As we said before, the use of venting valves seems to improve
the results obtained by only blowing the tanks.

5 Acknowledgments
This work has been supported by projects 2989/10MAE from Navantia S. A.
and 08720/PI/08 from Fundación Séneca, Agencia de Ciencia y Tecnologı́a de
la Región de Murcia (Programa de Generación de Conocimiento Cientı́fico de
Excelencia, IIPCTRM 2007-10).

Bibliography
[1] M. C. Bettle, A. G. Gerber and G. D. Watt, Unsteady analysis of the six DOF motion
of a buoyantly rising submarine, Computers and Fluids. (2009) 38:1833–1849.
[2] L. Byström, Submarine recovery in case of flooding, Naval Forces. (2004) Vol. XXV.
[3] C. T. Crowe, D. F. Elger and J. A. Roberson, Engineering Fluid Mechanics. John
Wiley & sons, (2004) Washington DC.
[4] J. Feldman, Revised standard submarine equations of motion, David W. Taylor Naval
Ship Research and Development Center (1979) Washington DC, DTNSRDC/SPD-0393-
09.
[5] R. Font, J. Garcı́a and D. Ovalle, Modeling and Simulating Ballast Tank Blowing and
Venting Operations in Manned Submarines, 8th IFAC Conference on Control Applica-
tions in Marine Systems, Rostock-Warnemünde, Germany, 2010.
[6] R. Font, J. Garcı́a, A. Murillo and F. Periago, Controllability of a mathematical model
for blowing and venting operations in submarines, XXII Congreso de Ecuaciones Difer-
enciales y Aplicaciones, Palma de Mallorca, Spain, 2011.
[7] T. I. Fossen, Guidance and control of ocean vehicles. John Wiley & sons (1994) Wash-
ington DC.
[8] D. Ovalle, J. Garcı́a and F. Periago, Analysis and numerical simulation of a nonlin-
ear mathematical model for testing the manoeuvrability capabilities of a submarine.
NonLinear Analysis B: Real World Applications. (2011) 12:1654–1669.
[9] G.D. Watt, Modelling and simulating unsteady six degrees-of-freedom submarine rising
maneuvers, DRDC Atlantic (2007).

View publication stats

You might also like