Está en la página 1de 9

Evaluation of optimal constant-volume control for

irrigation canals
J. Mohan Reddy
Department

of Agricultural

Engineering,

University

of Wyoming,

Laramie,

WY,

USA

A lumped-parameter
model was used for simulating the dynamics of an irrigation canal using the
concepts of optimal control theory. Using the state-space approach presented by Corriga et al., the
problem was formulated as a discrete optimal control problem, and the solutions for gate opening in
the presence of known variations in water withdrawal rates were obtained by solving the algebraic
Riccati equation. An example problem was considered, and the variations in the depth offlow obtained
by using the optimal control theory were compared with the results obtained from an unsteady openchannel flow model. For small and slow changes in the j7ow rates, the difference between the two
models in predicting the changes in water surface elevation was not significant. However, as the
variations in withdrawal rates increased, the accuracy of the optimal control model (which is based
upon small perturbations around the uniform flow) in predicting the water surface elevations decreased.
Kevwords: automation

mode1, state-space

of canals, constant-volume
model

control, discrete optimal control, lumped-parameter

Introduction
A significant benefit of a demand- or user-oriented
system is that such a system should be easier to manage
than other systems. By definition, such a system should
automatically
respond to new turnout discharges,
thereby decreasing the need for calculations and adjustments by canal operators to ensure smooth canal
operation. In social conditions in which a limited demand system is more desirable because of the likelihood of farmers stealing water or some other reason,
the farmers themselves might not be allowed to turn
their water supply on or off. An automatic downstream-controlled system makes operation by the canal
personnel much simpler than conventional methods.
Although adoption of a user-oriented water delivery
schedule is desirable, the existing solutions usually involve very detailed hydraulic modelling using a distributed parameter model. 2,3 To make the conversion
from upstream oriented to downstream oriented, there
is a need to develop relatively simple mathematical
models for improved operation of the canal systems
using downstream control. Burt developed a simple
control logic based upon regression analysis of data
obtained from extensive simulations of the unsteady
open-channel flow equations. However, the rules on
amount of gate opening appear a bit arbitrary.4

Lumped-parameter
models are much simpler than
distributed-parameter
models. Corriga et al. presented a continuous-time lumped parameter model for
optimal operation of irrigation canals based upon constant-volume control. However, the control was suboptimal in nature, since the effect of disturbances was
neglected in deriving the control law. This resulted in
a steady-state error in the volume of water maintained
in the reaches.
Rapid developments in digital computing technology coupled with sign&ant
reductions in their cost
make real-time data processing and control more attractive. This calls for discrete-time control rather than
continuous-time. Therefore the objective of this paper
is to present a discrete-time control technique by incorporating the effect of known external disturbances
(changes in withdrawal rates) acting on the system and
compare the performance of the linearized lumpedparameter model with the results obtained using an
unsteady nonlinear open-channel flow model. The case
of random (unknown) disturbances will be dealt with
in a separate paper.

Mathematical

model

The flow in open channels is described


Venant equations:

by the Saint-

_
Address
cultural
USA.
Received

450

reprint requests
to Dr. Reddy at the Department
of AgriEngineering,
University
of Wyoming,
Laramie,
WY 82071,

Q
;It+B$=o
a6
vav
--$+--+*=&_G

3 January

1989; accepted

20 March

Appl. Math. Modelling,

1990

1990, Vol. 14, September

gdx

gat

(continuity)

(1)

(momentum)

(2)

0 1990 Butterworth-Heinemann

Optima,

%(hb(i- 1)

bi

hai)

(3)

in which Ui = opening size of gate i (m*), { = gate


coefficient, and hbci_1j and hai = water level values
upstream and downstream of gate i, respectively. Assuming that the turnouts are located at the downstream
end of the reach, the continuity equation across the
gate is given by
1) =

qu(i+

qbi

the transformed

.___

J. M. Reddy

z=o

$0

equations are given as

dQ

z+BsH=O

(7)

and

in which qbi is the flow rate at the downstream end of


reach i and qri is the sum of the withdrawal rates in
reach i.
By considering a small perturbation about the steady
condition, equations (1) and (2) can be written in the
following form?
z+B$=O

q=o

h=O

(4)

qci

control for Irrlgatlon:

in which h(x, t) = level variation relative to a uniform


flow reference configuration with level ho, u. = initial
mean water velocity, q(x, t) = flow rate variation relative to a uniform flow configuration with flow rate qo,
co = initial celerity of perturbation; p = 2gs0/uo, y =
socakluo, and k = (b/ho + z)l(blh, + 22) in which b =
canal bottom width, and z = side slope. Since equations (5) and (6) are linear, the Laplace transform technique, with appropriate initial and boundary conditions, can be used to solve them. With the following
initial conditions:

in which q(x, t) = flow rate (m3/s), &x, t) = depth of


flow (m), u(x, 1) = velocity of flow (m/s), So = slope
of the channel, x = distance (m), t = time (seconds),
g = gravitational acceleration (m/se& k = conveyance factor commonly expressed through Mannings
equation, and B = water surface width (m). Equations
(1) and (2) are applicable in all the reaches of a canal.
The reaches are interconnected
by gates that control
the flow rate from upstream reach to the downstream
reach. The flow rate through gate i(q,J is governed by
4ai =

*r

constant-volume

(cg -

u$$ - (2~

dH
+ y)- s(s + P)H = 0
dx
(8)

in which H(x, s) and Q(x, s) are the Laplace transforms


of h(x, t) and q(x, t), respectively. Equations (7) and
(8) have solutions of the following form:
p,x

(5)

Q(x,s)=

and

-Bs

C,-+c,-ffl

&w

(9)
a2

>

H(x, s) = Cl ealr + C2e


(10)
in which the constants C, and C2 are determined by
considering the boundary conditions for the specific
canal portion and the (YSare given by the roots of the
characteristic equation of equation (8):

S*h
3 + 2&l

(2suo + y) t ~(2suo + $2 + 4(&j - u$s (s + p)


a1.2

The boundary

conditions

for the ith reach, using equation (IO), are

Hi(O, S) = Cli + C2; = HAi


Hi(L;, S) = CliealJl + C2ieazJl = HBi

Solution of equations
C,; = 2;:::

Derivation

(11)

2(cS - u;,

iaZF

forx; = 0

(12)

fOrXi = Li

(13)

(12) and (13) yields


c,i = HBi - e"""H.4i
e%iL1- e%L,

(14)

of the state equations

Let us consider the ith canal reach belonging to a Ncascaded reach open-channel system as shown in Figwe 1, where the reaches are joined by controlled gates.

Let us assume that the first reach is fed by a reservoir


whose water level can be considered practically constant and that the last reach is joined, through a fixed
noncontrolled gate, to a constant head reservoir. Since
the objective is to develop a constant-volume control,
the continuity and momentum equations can be used
to derive an expression for volume variation as a function of the variation in the depth of flow at the upstream
and downstream ends. These equations, as derived by
Corriga et alA in more detail, are concisely presented
here.

Appl. Math. Modelling,

1990, Vol. 14, September

451

Optimal constant-volume

control for irrigation:

J. M. Reddy

GATE i+l

GATE i

-LiGATE 1

Vi(s) = B
GATE N

GATE 2

REACH 1

hb(i-l)

hai

(Clieall~ + CIiea2f) dx

hbi ha(i+l)

R?cH

BCli(elL~ -

1) + BC2i(eaz~Lr - 1)

qb(i-l)

qoi
I
%(i-1)

(15)

ff2i

ali

By using equation (14) the expressions for Cl; and Cli


are derived for both of the following cases:

qbi
-1

qa(i+l)
HA;

rf

HBi = 0

0,

and

qci

Figure 1. An irrigation canal with cascaded reaches

HAi = 0,

The variation of the stored volume in the ith reach


(V;(s)) can be expressed, in the frequency domain, as
follows:

HgifO

These expressions are then substituted into equation


(15) to derive the transfer functions between the volume variation and the level variation at the upstream
and downstream ends of the ith reach:

(17)

Application

of the superposition

principle results in

V;(S) = GII(S)HAI(S) + Gzi(s)Hsi(s)

(18)
which expresses the volume variation in the ith reach
as a function of the level variations at the upper and
lower ends of the reach. It can be used in designing
the feedback control law that operates on the upperend control gate in order to keep a constant stored
volume in the reach (Figure 1). Corriga et al. used a
proportional (P) control law (the expression for the gate
opening) of the following form in the analysis:
Vi(s) = - G,i(s)Vi(s)

By using equation (9), with x = 0, an expression


for the flow rate variation is given by
Qi(O,s)

= -Bs(%

+ 2)

= Q,z,;

(22)

in which QAi = flow rate variation at the ith reach


upper end. With HAi # 0 and Hsi = 0, one obtains

(19)

where z/i(S) = the Laplace transform of the function


representing the variation of gate opening (pi) and
GJs) = the transfer function of the controller for gate
i and is given as

and with HA; = 0 and Hei # 0, one obtains

Wzi(S)
=

E
BI

G,i(s) = &

(20)

in which Kci = proportional gain of the controller for


gate i. The larger the gain &, the higher the sensitivity
of controllers actuating signal to deviations in the output. This control law, when applied to stabilize a system in the presence of constant disturbances, will yield
a nonzero steady-state value for the output. This value
can be reduced to zero by using the proportional-plusintegral (PI) control. In this case the transfer function
of the controller takes the following form:
G,i(s) = K,i(l + l/(Tis))

Appl.

Math.

Modelling,

1990, Vol.

14, September

(25)

QA~(s>= WI;(S)HAi(S) + W~i(S)HBi(S)

In a similar fashion, from (9) with x = Li,


e4L

(21)

where Ti = integral time constant or reset time. The


integral control action causes the controller output to
change as long as a nonzero value exists in the system
output.

452

By taking into account the level variations both at the


upper and lower ends of the reach and using the superposition principle the flow rate variation at the ith
reach upper end can be expressed as

Qi(L;y

S) =

-Bs

eW,Li

+ Czj -

Cli ali

a2i

>

= QB;

(26)

and
QB~(s) = WX(S)HA~(S) + W4i(s)HBi(s)

(27)

Optimal

where QRj = flow rate variation at the ith reach lower


end and
W&)
Equations
under the
at the ith
variations
ith reach.

(2%

W&(S) = gg

Al

BI

(25) and (27) express the flow rate variation


ith control gate and the flow rate variation
reach lower end as a function of the level
at the upstream and downstream ends of the

where Poi is the initial point, corresponding to the uniform flow conditions.
To derive the state equations, equations (25) and
(27) are written in the following form, after using a
low-order approximation for W(S):
h!Aj(s) = T

QA;(s) + +

(I -

Zl;s)QBi(S)

constant-volume

control

for irrigation:

J. M. Reddy

To complete the reach model for the ith reach, one


must add to the above relations the continuity equation
(equation (4)) for the variation in the flow rates:
QB~(s) = Qci(S) + QA(~+I~s)

(2%

and the equation for the flow through the upstream end
control gate of the ith reach which, in the time domain,
is given by equation (3). In equation (29), Qcj is the
variation in the water withdrawal rate in the ith reach.
Equation (3) can be linearized and then Laplace-transformed, leading to

of gates (m + 1). The parameters of the matrices G(s)


and W(s) are absorbed into the above matrices. Once
the elements of the above equations are obtained, the
control theory principles can be applied to derive the
optimal control law for the system.

(32)

Discrete optimal control


Hi;

= $

(1 -

where the constants

TziS)QA;(S) + y

QBi(s)

(33)

verify the following conditions:

Kri = -Kzi

x(k + 1) = Fx(k) + Gu(k) - Cq,.(k)

K3i = - Kdi
and the constants Tj; are the time lags in the propagation of an infinitesimal perturbation in the upstream
and downstream directions. By taking the inverse Ltransforms of equations (29), (30), (32), and (33) and
expressing the volume variation in a given reach by
f
u(t) =

(34)

(qaib) - qbi(T)) dT

the problem can be transformed, after including the


disturbances, to the following form in the time domain:5
G(r) = Av(t) + Bu(t) - Zq,

(35)

in which v(t) = variation in the volume of water in a


given reach, and this is the state variable in the system;
and u = variation of gate opening, which is the control
vector; A = state feedback matrix; B = input matrix;
and I = identity matrix. The matrices A and B represent the characteristics of the system at steady state
(or initial state). The dimensions for the matrices are
as follows:
A:m x m,

Since the objective is to derive a discrete optimal control law, the state equation (equation (35)) must be
converted to a discrete-time version of the following
form (after setting x = v):

B: m

1,

I:mxm,
v:m X 1,

u:l x I

where m = the number of reaches and I = the number

(36)

and the output equation is given by


Y&l

= Wk)

(37)

in which F and G are the discrete-time versions of


matrices A and B, respectively; C is the disturbance
matrix and is equal to TZ; q<.(k) is the disturbance (variation in withdrawal rate); k is the sampling instant; T
is the sampling period; y is the system output; and D
is the output matrix.
To derive the control law, let us first assume that
q,. = 0 and the system is at rest, but owing to some
impulsive disturbance the state at k = 0 is displaced
to x(k = 0) # q,. The regulator problem is to apply a
state feedback control u(k) (Figure 2) such that the
state is returned to the original state, x(k) = Q,, as
quickly as possible. There are two ways of achieving
the quick response. If {F, G} is controllable, an impulsive input, u(k), can be used to instantaneously restore the state to zero. The price that one pays for this
solution is, of course, very high (infinite). Conversely,
one can obtain a finite energy input as u(k) = - Kx(k),
so that
x(k + 1) = (F - GK)x(k)

(38)

and by choosing K suitably, x(k) can be made to decay


to x, as fast as desired. The rate of decay depends on
how negative the real parts of the eigenvalues of (F -

Appl. Math. Modelling,

1990, Vol. 14, September

453

Optimal constant-volume

control for irrigation:

J. M. Reddy
stant error between the actual and desired operating
condition of the canal. Depending upon the magnitude
of the disturbance, the error may become unacceptable. However, the effect of these disturbances can often
be eliminated by using the so-called integral feedback
(Figure 2). This is achieved by appending additional
state variables of the formlo

qua

4
C

w(k + 1) = Dx(k)T + w(k)

(44)

to the system dynamic equation (equation (36)). This


modifies equation (36) to the following form:
Figure 2.

Schematic

of the feedback control loop

X(k + 1) = FX(k) + Gu(k) - Cq,.(k)

GK) are. The more negative these values are, the larger
will be the value of K and therefore the higher the
required signal energy and the associated cost.9
The above discussion suggests that we should try
to make a tradeoff between the rate of decay (response
of the system) of x(k) to its initial value and the energy
of the input. In the quadratic regulator problem this is
done by choosing u(k) to minimize the cost function

2 (x(k)TQx(k)

J =

+ u(k)=Ru(k))

(39)

k=l

subject to equation (36). In equation (39) the first term


is a measure of the extent to which the deviation in a
state variable at time period k is penalized; the weight
matrix Q determines how much weight is attached to
each of the state variables. The second term (with R)
in the cost function penalizes the input (gate opening)
amplitudes. By applying the Pontryagins maximum
principle to the above problem, the optimal control
(input) law is given by
u(k) = -Kx(k)

(40)

where K is the steady-state


is defined by
K = R-G=FTpl(P,

feedback

F[Pk+l

(41)

pk+ ,G(GTpk+
,G +

WGTPk+

x F + Q

,)I

(42)

with
(43)

where H is the terminal penalty matrix and IV is the


number of time periods considered. In this case, H was
assumed to be zero. After selection of the number of
time periods (which should be large for steady-state
values), equation (42) is solved backward in time.
Irrigation canals are always acted upon by periodic
disturbances. These must be taken into account in the
operation of canals. Otherwise, there would be a con-

454

Appl.

Math. Modelling,

WV = [x(k) w(k)lT
F=

OT

[ TD

I 1

G = [G

GJT

C = [C

O]T

where 0 = a null matrix and G, = a nonzero square


matrix. The dimensions of the Z and G1 matrices are
selected so that the matrices are compatible with the
other matrices (or vectors). With the integral feedback
the new control law is given by
u(k) = -KX(k)

(46)

u(k) = [K,

(47)

or
K,]

The second term in equation (47) accounts for the disturbances. In the absence of any disturbances the outer
feedback loop in Figure 2 is not required (i.e., K, =
0) to bring the system to equilibrium.

Evaluation

- Q)

PN = H

in which

gain matrix and

in which F-l denotes the inverse of the transpose of


F and P, is the steady-state solution of the discretetime matrix Riccati equation, which is given below:
Pk

(45)

1990, Vol. 14, September

of the constant-volume

method

In the evaluation of the model, first the model and the


example problem presented by Corriga et a1.j. were
used to obtain values for variations in the volume of
water stored in each reach and the variations in gate
openings. Then a discrete-time version of the above
model was derived, including a proportional-plus-integral (PI) controller to account for the known disturbances acting on the system. The variations in the
volume of water and the gate openings predicted by
the discrete-time model with the PI controller were
compared with the results obtained by Corriga et a1.5
These variations in the volume of water and the gate
opening were then used to predict the variations in
flow rates and depths of flow at the upstream and
downstream ends of each reach. Equations (29), (30),
(32), and (33) (after taking the inverse Laplace transform) along with equation (34) were used in the above
calculations.
Since the rate of discharge into the lateral canals
depends upon the depth of flow in the main canal,

Optimal constant-volume

whether it is a constant-volume control or a constantlevel control, we are always interested in knowing the
expected deviations in the depth of flow at the downstream end of the reach where the lateral canals are
usually located. This is helpful in deciding the appropriate structures to discharge the required quantity of
water into the lateral canals. To accomplish the above,
the variations in flow rates and depths of flow predicted
by the above model were added to the initial flow rates
and depths of flow, respectively, in the channel. An
unsteady open-channel flow model developed by Hromadka et al. ** was used to simulate the canal dynamics
using the same data. This model needs, as boundary
K12 = 2.85 x 10P5m-*s

K3, = 6.38 x 10P5m-*s

K3* = 7.51

z-1, = 1739s

T,* = 2475 s

T,, = 917 s

Tz2 = 1334 s

cri = 2.41 m/s

10-5mP2s

/3; = Om*/s

ff; = 1.91 m/s


/3; = 2.99m2/s

y; = 3.76m2/s

r;

a; = 0.61 m/s
p; = 5.08 m*/s

= - 2.99 m*/s

~4 = 0 m2/s

The discrete-time matrices F and G were computed by


using the above values and a sampling interval of 10
seconds. The elements of the matrices are given below:
F=

[ 0.9955
0.0038
and

0.9933
0.0024

J. M. Reddy

conditions, the inflow rate at the upstream-end and the


outflow rate at the downstream-end of the given reach
in addition to the initial flow rate and the depth of flow
at the downstream end of the canal. The variations in
flow rates predicted by the optimal control technique
were used in the nonlinear unsteady model to predict
the depths of flow at the upstream and downstream
ends of the given reach. These predicted depths were
then compared with the depths obtained by using the
optimal control technique.
The geometry of the canal, along with the initial
conditions, is shown in Figure 3. The values of the
constants used in the simulation study are as follows:

K,, = 3.76 x 10P5mP2s

control for irrigation:

The gain matrix K was obtained by solving the steadystate Riccati equation:

K =

-5.160

1.525 x lop3

x lop4

1O-4
-5.875
5.002 x lop3

[ - 1.570
1.943 x 10-4
IO-3

These values are almost equal to the values obtained


by Corriga et ~1.~ using the continuous-time control.
After freezing the last gate position at its equilibrium
value, the integral feedback elements (equation (44))
were appended to the original state equation, resulting
in the following values for matrices F and G:

G=

-8.817
33.700
- 7.402
1.870 1
-- 27.905
30.200
The weighting matrices are diagonal in nature, and the
elements of the matrices are usually found by trial and
error. The selection is based upon the desired closedloop dynamics of the given system. The elements of
the weighting matrices were as follows:

0.9955
F = [ 0.0038
9.9773
0.0191

0.0024
0.9933
0.0121
9.9663

0 0
0 0
1 0
0 1

and
30.203 1
G = I - 7.4020
151.1451
1 -37.1480

- 27.9052
33.7322
- 139.7006
168.9403

The elements of the new gain matrix were as follows:


K, =

Figure 3.
lation

Geometry

of the irrigation

0.0290
[ 0.0014

0.0192
0.0268

0.0011
-0.0002

0.0005
0.0010

These values were used in simulating the system dynamics by using equation (45).
Initially, a 10% change (increase) in the farmers
withdrawal rate (qc(l> = 0.295 m3/s and qc(2) = 0.25

canal used in the simu-

Appl.

Math.

Modelling,

1990,

Vol.

14,

September

455

Optimal constant-volume

control for irrigation:

J. M. Reddy

m3/s) was introduced in both the reaches, and the variation of the gate openings, changes in the volume of
water (state of the system), and changes in the depth
of flow and the flow rates at the upstream and downstream ends of both the reaches were simulated. Simulations were also done by increasing the changes in
the water withdrawal rates to 20% of the original withdrawal rates.
Results and discussion
First the volume variations in both the reaches were
simulated for a 10% change in the withdrawal rates.
Figure 4 presents the volume variations obtained by
using the proportional (P) and proportional-plus-integral (PI) control laws. The predicted variation in volumes using the PI control was less than 3 m3 in contrast
to 80 m3 obtained by using Corriga et al.s5 method.
This shows the advantage of the PI controller over the
proportional (P) controller in keeping the volume variation at almost zero. But it should be pointed out that
the percent variation in the volume of water was still
negligible in both cases for a 10% change in withdrawal
rate.
The gate openings predicted by the discrete-time
model with the PI controller (Figure 5) were compared
with the values predicted by the continuous-time model
with the P controller (Figure 6). The gate openings
predicted by the PI controller (in the presence of disturbances) were slightly higher than the values predicted by the P controller. This resulted in negligible
volume variation in the case of the PI control. Also,
the PI control has some overshoot at the beginning and
end of the disturbance period. This can be smoothed
out by increasing the value of R in the cost function
(equation (39)).
The variations in flow rates at the upstream and
downstream ends of the reaches are depicted in Figure
7. These variations in flow rates in both the reaches
were less than the required values. The total change
required in flow rate in the first reach was 0.55 m3/s
(0.298 m3/s and 0.250 m3/s). But the actual variation
obtained by using the linear optimal control model was
approximately 0.40 m3/s. This resulted in a negative
change in the volume of water stored in both the reaches.
In Figure 8 the predicted variations in the depth of
flow at the upstream (hai) and downstream (hbi) ends
of the reaches are presented. The change in the depth
of flow was positive at the upstream ends of the reaches
because of the increase in flow rates into the reaches.
The maximum increase in the upstream depth of flow
was 0.027 m and occurred in the first reach. Conversely, the change in the depth of flow at the downstream end of the reaches was negative with a maximum deviation of - 0.025 m. This also occurred in the
first reach. These predicted variations in depth are
negligible.
Later on, the variations in the volume due to a 20%
change in withdrawal rate were simulated. Again, the
volume variations in the first and second reaches were
found to be negligible (Figure 9). The variations in gate

456

Appl.

Math. Modelling,

1990, Vol. 14, September

-1001
0

500

1000
duration

1500
2000
of disturbance.
see

3000

2500

Figure 4. Deviations in volume of water stored for a 10% change


in demand: continuous-time
model with P controller and discrete-time model with PI controller

2
duration

Figure 5. Variation
mand: discrete-time

-0.1

I
0

3
of simulation.

4
set

5
(thousands)

in gate opening for a 10% change in demodel with PI control

I
duration

3
of disturbance.

sec(thousands)

Figure 6. Variation in gate openings for a 10% change in demand: continuous time model with P control

openings, the maximum depth of flow, and the flow


rates were almost twice as much as the first case. But
these variations were not significant.
In general, the predicted variations in the stored

Optimal constant-volume

2
duration

4
5
sec(thousands)

3
of stmulatmn.

i
a
n
9
e

c------,
0

2
duration

-7,
3
of simulatmn,

J. M. Reddy

be within acceptable limits even when the deviations


in withdrawal rates approach 40%. However, our interest is basically in knowing how closely these predicted values for depth of flow match the actual depth
of flow in the canal for the given deviations in withdrawal rates. In lieu of actual field data the predicted
depths of flow values obtained from the lumped-parameter model were compared with the depths of flow
values obtained from the Saint-Venant equations of
open-channel flow.
For a 10% variation in the withdrawal rates, the
difference in the depth of flow predicted by the two
models was moderately large (Figure 10). For the linearized model, the maximum predicted deviations in
depth of flow at the upstream and downstream ends
of the first reach were 0.027 m and -0.025 m, respectively. When the nonlinear flow model was used,
the maximum predicted deviation at the upstream end
of the first reach was 0.05 m, whereas at the downstream end the variation in the depth of flow reached
a maximum value of - 0.16 m. Though the difference
between the two models was large, the maximum deviations were still within acceptable limits.

Figure 7. Variations in flow rates at upstream and downstream


ends of reaches for a 10% change in demand

-0.04

control for irrigation:

I
4
5
s&thousands)

Figure 8. Variations in depth of flow at the upstream (h,) and


downstream (hb) ends of reaches for a 10% change in demand

l-

m
e

2
duration

3
of simulation,

4
5
secfthousands)

Figure 10. Variations in flow depth obtained from the optimal


control model and the unsteady flow model for a 10% variation
in demand: Reach 2

.__

a
r

-1

i
0
n

0.2

-2

:
a

reach 2

react! I

L
m

0.1

:
e

-3

-4A

O-0.1

f
0

2
duration

3
of simulatmn.

4
5
sec(thousands)

0
w

-0.2

downslream depth. Slrady model

-0.3

Figure 9. Variation in volume of water stored for a 20% change


in demand

volumes, gate openings, flow depths, and flow rates


were within acceptable limits as long as the deviations
in water withdrawal rates were less than or equal to
20%. One can extrapolate and say that the predicted
values from using the lumped-parameter
model would

-0.4

7
h

-0.5

-0.6

L
0

r
1

2
duration

3
of simulation.

4
5
secfthousands)

Figure 11. Variations in flow depth obtained from the optimal


control model and the unsteady flow model for a 20% variation
in demand: Reach 1

Appl.

Math. Modelling,

1990, Vol. 14, September

457

Optimal constant-volume

control for irrigation:

J. M. Reddy

The difference between the two models in predicting


the depth of flow was even higher when the change in
the water withdrawal rate was increased to 20% (Figure II). The maximum predicted deviations in depth
of flow at the upstream and downstream ends of the
first reach were approximately 0.05 m and -0.05 m,
respectively, using the linearized model. The nonlinear
model predicted a maximum deviation of 0.10 and
-0.48 m, respectively,
at the upstream and downstream ends of the same reach.
As the changes in withdrawal rates increased, the
nonlinear model simulated the dynamics much better
than the linearized model, which has the exponentially
decaying smooth behavior. Hence as large, sudden flow
rate changes were introduced in the canal, the accuracy
of prediction of depth of flow in the reaches using the
linearized model diminished. However, since the above
linear optimal control model directly solves for gate
openings in the presence of known disturbances and
maintains a constant volume in each reach, independent of the percent deviations in the water withdrawal
rates, where large deviations in withdrawal rates are
expected, the outlet structures must be fitted with discharge regulators 12-14to deliver constant flow rate into
the laterals under variable flow depths in the main
canal.
Summary

Appl.

References
1

The Saint-Venant equations of open-channel flow were


linearized about the initial steady-state profile. A methodology to derive the transfer functions and the statespace equations, as presented by Corriga et ~l.,~ was
briefly discussed. A discrete optimal PI controller for
operating the gates along a canal reach was derived.
By considering an example problem the variations in
the depth of flow obtained by using the linearized
lumped-parameter
model along with the optimal control theory were compared with the results obtained
from an unsteady open-channel flow model.
For small and slow changes in the flow rates, the
difference between the two models in predicting the
changes in depth of flow was within acceptable limits.
However, as the transfers between the reaches increased, the accuracy of the lumped-parameter
model

458

(which is based upon the linearized Saint-Venant equations) in predicting the deviations in depth of flow decreased. By using the distributed-parameter
(full hydrodynamic) model, the range in which the performance
of the optimal control model is acceptable can be identified for any given system. Since the computational
effort involved in the optimal control model is much
less, it can be used for operation of irrigation canals.

Math. Modelling,

1990, Vol. 14, September

9
IO
11

12
13

14

Burt, C. M. Canal automation


for rapid demand deliveries.
Proceedings
of the ASCE Irrigation
and Drainage Division
Specialty Conference,
Flagstaff,
Arizona,
1984
Zimbelman,
D. D. and Bedworth,
D. D. Computer
control for
irrigation canal systems. J. Irrigation Drainage Engrg. ASCE
1983, 109(l), 43-59
Rogier, D., Coeuret,
C. and Bremond,
J. Dynamic regulation
on the Canal de Provence.
Proceedings
of ASCE Symposium
on Planning, Operation, Rehabilitation
and Automation
of Irrigation Water Delivery Systems,
Portland,
Oregon,
1987
Clemmens,
A. J. and Replogle, J. A. Control of irrigation canal
networks.
J. Irrigation Drainage Engrg. ASCE 1989, 115(l),
96-110
Corriga, G., Sanna, S. and Usai, G. Sub-optimal
constant volume control for open-channel
networks.
Appl. Math. Modelling 1983, I, 262-267
Hancu, S. and Rus, E. Stabilitt de mouvement
de leau dans
les canaux dadduction
des systkmes dirrigation
a fonctionement automatique.
Rev. RoumaineSci.
Tech. 1976,19,431-450
Corriga,
G., Fanni, A., Sanna, S. and Usai, G. A constantvolume control method for open-channel
operation.
Internat.
J. Modelling Simulation
1982, 2(2), 108-l 12
Corriga, G., Sanna, S. and Usai, G. Frequency
response and
dynamic behavior of canal networks
with self-levelling
gates.
Appl. Math. Modelling
1980, 4, 125-129
Kailath, T. Linear Systems.
Prentice-Hall,
Englewood
Cliffs,
N.J., 1980
Kwakemaak,
H. and Sivan, R. Linear Optimal Control Systems. John Wiley & Sons, New York, 1972
Hromadka,
T. V., Durbin, T. J. and DeVries, J. J. Computer
Methods in Water Resources.
Lighthouse
Publications,
Mission Viejo, Calif., 1985
Alsthom Atlantic,
Catalog and Technical Description
of Automatic Control Gates. Grenoble,
France,
1980
Clemmens,
A. J. and Replogle,
J. A. Mechanical-hydraulic
dual-acting
controllers
for canal discharge
rates. J. Irrigation
Drainage Engrg. ASCE 1987, 113(l), 69-85
Kermo;e,
I. The bladder valve: A self-regulating
constant discharge irrigation turnout. Winchmore
Irrigation Research
Station, Ashburton,
New Zealand

También podría gustarte