You are on page 1of 27

1.

OPEN CHANNEL FLOW AND ITS CLASSIFICATIONS

1.1 Introduction

The aim of the course open channel hydraulics is to present a review of the main principles of
hydraulics and to apply them in design of water resources systems. After the course the student
should understand and should be able to apply the principles of open channel hydraulics in
engineering problems and (s)he should be capable to make uniform and non-uniform flow
computations.

After the classification of open channel flow types and a short description of dimensionless
numbers in chapter 1, Chapter 2 starts with review of the basic hydraulic principles, like
continuity, energy and momentum principles. These basic hydraulic principles and equations are
applied to uniform flow computations, flow over sills and contractions in chapter 3. The
momentum principle is also applied to control sections and outflows problems in the same
chapter. The following two chapters (Chapter 4 and 5) deals with GVF and RVF. These two
chapters deal with GVF computation, flow profiles and applications. Particular cases of RVF
problems are treated in Chapter 5. Basic principles of unsteady flow in open channels and
particular applications are covered in the last chapter (chapter 6).

1.2 Classification of Open Channel flows

Open channel flows versus pipe flow

Open channel: is a conduit for flow, which has a free surface, i.e. a boundary, exposed to the
atmosphere.

The flow of water in a conduit may be either open channel flow or pipe flow. Open channel is a
conduit for flow which has a free surface, i.e. a boundary exposed to the atmosphere. The two
kinds of lows, i.e. open channel and pipe flow, are similar in many ways but differ in one
important aspect. Open channel flow must have a free surface, whereas pipe flow has none, since
the water must fill the whole conduit. A free surface is subject to atmospheric pressure. Pipe
flow being confined in closed conduit, exerts no direct atmospheric pressure but hydraulic
pressure. In the case of the flowing fluid in open channel, the motion is usually caused by
gravity effects and the pressure distribution within the fluid is generally hydrostatic1.

1. If the flow were curvilinear or if the slope of channel were large hydrostatic law of
pressure distribution will not hold true, read pp 30- 33 Ven Te Chow.

1
Figure 1-1Comparison between pipe flow and open channel flow.

Questions:
 Where is the HGL in case of the open channel flow?
 What is the influence of the change in cross–section of the pipe?
 Is a half–filled pipe flow open channel flow or pipe flow?

The two kinds of flows are compared in Figure 1-1 shown on the lift side is pipe flow. Two
piezometer tubes are installed on the pipe at section 1 and 2. The water levels in the tubes are
maintained by the pressure in the pipe at elevations represented by the so–called hydraulic grade
line. The pressure exerted by the water in each section of the pipe is indicated in the
corresponding tube by the height (y) of the water column above the centerline of the pipe. The
total energy in the flow of the section with reference to a datum line is the sum of the elevation z
V2
of the pipe center line, the peizometric height of and the velocity head , where v is the mean
2g
velocity of flow. The energy is represented in the figure by what is called the energy grade line
or a simply energy line. The loss of energy that result when water flow from section 1 to section
2 is represented by hf. A similar diagram for open channel flow is shown on the right side of
Figure 1-1. For simplicity, it is assumed that the flow is parallel and has a uniform velocity
distribution and that the slope of the channel is small. In this case, the water surface is the
hydraulic grade line, and the depth of the water corresponds to the peizometric height.

Despite the similarity between the two kinds of flow, it is much more difficult to solve problems
of flow in open channels than in pressure pipes.

Reasons:

2
 In open channels the position of the free surface is likely to change with respect to
time and space,
 Depth of flow (y), discharge (Q), and bottom slope (S) and slope of the free
surface are independent,
 Physical condition of open channels varies more widely than that of pipes,
 Cross section of open channel is widely variable and even might not be ridged,
 Selection of fiction coefficient is difficult and it varied with the position of the
free surface,

Examples of Open channel Flow:

Mention some examples of open channel flows


- storm sewer (partially field sewers),
- flow in rivers,
- flow in irrigation cannels,
- gutters along residential streets and no roof open channel.

Types of flows
Open channels flow can be classified into many types and described in various ways. The
following classification is made according to the change in flow depth with respect to time and
 dy dy 
space  ,  .
 dt dx 

Steady flow and unready flow: Time as the criterion

Flow in an open channel is said to be steady if the depth of flow does not change or if it can be
assumed to be constant during the time interval under consideration at a fixed point. In steady
flow the flow variables (velocity, pressure, density, flow path etc) do not vary with time at the
spatial point in the flow. In steady flow streamline is also the path followed by an individual
water particle. The flow is unsteady if the depth changes with time. In most open channel
problems it is necessary to study flow behavior only under steady conditions. If, however, the
change in flow condition with respect to time is of major concern, the flow should be treated as
unsteady. In floods and surges, for instance, which are typical examples of unsteady flow, the
stage of flow changes instantaneously as the wave pass by, and the time element becomes vitally
important in design of control structures. In unsteady flow the flow variables (velocity, pressure,
density, flow path etc) vary with time at the spatial points in he flow.

Examples of unsteady flow:


 Oscillatory sea waves,
 Dam break flood waves,
 Surges due to gate operation,
 Floods.

For any flow, the discharge Q at a channel section is expressed by

3
Q = VA Equation 1-1

Where v is the mean velocity and A is the flow cross sectional area normal to the direction of the
flow, since the mean velocity is defined as the discharge divided by the cross-sectional area.

In most problems of steady flow the discharge is constant throughout the reach of the channel
under consideration; in other words the flow is continuous. Thus, using equation 1-1.
Q = V 1 A1 = v 2 A2 = - - - - Equation 1-2

Where the subscripts designate different channel sections. This is the continuity equation for a
continuous steady flow.

Figure 1-2 steady and unsteady flow

Equation 1.2 obviously invalid, however, where the discharge of a steady flow is non-uniform
along the channel, that is, where water runs in or out along the course of the flow. This type of
flow is called spatially varied or discontinuous flow. Spatially constant flow occurs when the
density and average velocity are the same in all points in a flow field. If these quantities change
along or across the flow lines the flow is spatially variable. Examples are side channel spillways,
roadside gutters, teh flow in uniform canal of constant slope receiving inflow or having outflow
(e.g. main drainage channels and feeding channels in irrigation systems).

The law of continuity of unsteady flow requires considerations of the time effect. Hence, the
continuity equation for continuous unsteady flow should include time element as a variable
(section 2-3).

Uniform flow and varied flow: space as the criterion

Open channel flow is said to be uniform if the depth of flow is the same at every section of the
channel. A uniform flow may be steady or unsteady, depending on whether or not the depth
changes with time.

4
Steady uniform flow is the fundamental type of flow treated in open channel hydraulics. The
depth of flow does not change during the time interval under consideration. The establishment of
unsteady uniform flow would require that the water surface fluctuate from time to time while
remaining parallel to the channel bottom.

Obviously, this is a practically impossible condition. The term “uniform flow” is therefore, used
here after only to steady uniform flow.

Flow is varied if the depth of flow changes along the length of the channel. Varied flow may be
either steady or unsteady. Since unsteady uniform flow is rare, the term “unsteady flow” is used
hereafter to designate unsteady varied flow exclusively.

Varied flow may be further classified as either rapidly or gradually varied. The flow is rapidly
varied if the depth changes abruptly over a comparatively short distance; otherwise, it is
gradually varied. A rapidly varied flow is also known as a local phenomenon; examples are the
hydraulic jump and the hydraulic drop. For clarity, the classification of open-channel flow is
summarized as:

Uniform flow

Steady Flow
Flow Gradually Varied (non-
uniform) Flow
Varied Flow
Gradually Varied (non-
uniform) Flow
Unsteady uniform flow
(Quasi uniform flow)
Gradually varied
Unsteady Flow unsteady Flow

Unsteady Flow (i.e.


unsteady varied flow)
Rapidly varied unsteady
flow

Figure 1-3 Classification of open channel flows

Various types of open channel flows are sketched in Figure 1-4.

5
Figure 1-4 various types of open channel flows

Real or ideal flow

Real flow implies viscosity and frictional (viscous) effects. Ideal flow is hypothetical, ideal flow
assumes no friction, which means that the velocity is zero. There are no frictional effects
between moving fluid layers or between these layers and boundary walls; there is no energy
dissipation due to friction.

Incompressible or compressible flow

Incompressible flow assumes a constant density in time and space. Also liquids are slightly
compressible, it is assumed that they are incompressible. Compressible flows consider liquids,
that undergo changes in their density due to changes in absolute pressure and/or temperature.

1.3 Dimensionless Numbers and state of flow.

Forces that may act on a fluid element include those due to gravity Pg, pressure Pp, viscosity Pv,
and elasticity PE. if an element is at a liquid gas interface, there are forces due to surface tension
PT, If the summation of all forces on a fluid element does not add up to zero, the element will
accelerate according to Newton’s law. Such an unbalanced force system can be transformed into
a balanced system by adding an inertial forced P1(=m.a), which is equal and opposite to the
resultant of the acting forces.

Thus, generally,

PG + Pp + PE + PT = Resultant = - PI

6
F = m *a = Pi

Thus: PG + Pp + Pv + PE + PT + PI = 0

These forces may be expressed in simplest terms as:

Gravity: PG  mg  gL3
Pressure: Pp  p  A  p  L2
Viscosity:  du  V 
Pv     A     L2   VL
 dy  L
Elasticity: PE  Ev A  Ev L2

Surface tension: PT  L
Inertia: L
PI  ma  L3 2
 L4 T  2
T
In many flow problems some of these forces are either not present or insignificant. The state or
behavior of open channel flow is governed basically by the effects of viscosity and gravity
relative to the inertia forces of the flow. The surface tension, forces due to pressure and
elasticity of water may affect the behavior of flow under certain circumstances, but it does not
play a significant role in most open channel problems encountered in hydraulic engineering.

Effect of viscosity: Depending on the effect of viscosity relative to inertia, the flow may be
laminar, turbulent, or transitional. The basis for this classification is a dimensionless parameter
known as the Reynold number (Re), which is given by the ratio of these two forces:

Inertia P V 2 L2 VL VL
Re=  I   
Vis cos ity PV VL  

Therefore,
VL
Re Equation 1-3

Where v = characteristic velocity of flow, often taken as the average velocity of low,
L = characteristic length. In open channel thus, the characteristic length commonly used
is the hydraulic radius, which is the ratio of the flow area (A) to the wetted
A
perimeter (P), ( R  ) .
P
This dimensionless number Re is more interesting in pipe flow because in pipe flow the
significant forces are inertia and fluid friction due to viscosity. In a flow through a completely
filled pipe, gravity does not change the flow pattern.

The flow is laminar if the viscous forces are so strong relative to the inertial forces that viscosity
plays a significant part in determining flow behavior. In laminar flow, the water particles appear

7
to move in definite smooth paths, or streamlines, and infinitesimally thin layers of fluid seem to
slide over adjacent layers.

The flow is turbulent if the viscous forces are weak relative to the inertial forces. In turbulent
flow the water particles move in irregular paths, which are neither smooth nor fixed but which in
the aggregate still represent the forward motion of the entire stream.

Between the laminar and turbulent status there is a mixed, or transitional state.

An open channel flow is laminar if the Reynolds number Re is small and turbulent if Re is large.

Numerous experiments have shown that the flow in pipe changes from laminar to turbulent in the
range of Re between the critical value 2,000 and a value as high as 50,000. In this experiment the
diameter of the pipe was taken as the characteristic length in defining the Reynolds number.
When the hydraulic radius is taken as the characteristic length, the corresponding range is from
500 to 12,500 since the diameter of a pipe is four times its hydraulic radius.
A  ( D / 2) 2 D
R   
P D 4

For pipe flow

Flow R taken as characteristic length D taken as characteristic length


Laminar R  500 R 2000
Transitional 500  R  12, 500 2000  R  50,000
Turbulent 12,500  R 50,000  R

The laminar, turbulent, and transitional states of open channel flow can be expressed by a
diagram that shows a relation between Reynolds number and the friction factor of the Darcy.
Weisbach formula. Such a diagram, generally known as the Stanton diagram, has been
developed for flow in pipes. The Darcy weisbach formula also developed primarily for flow in
pipes. (further reading Ven Te chow, open channel hydraulics).
For practical purposes, the transitional range of Re for open channel flow may be assumed to be
500 to 2,000.

For open channel flow

Flow R taken as characteristic length


Laminar R  500
Transitional 500  R  2000
Turbulent 2000  R

8
Effect of gravity: the effect of gravity upon the state of flow is represented by a ratio of
inertial forces to gravity forces.

PI L2V 2 V 2
Fr 2   
PV gL3 gL

This ratio is given by the Froude Number, defined as:

V
Fr  Equation 1-4
gL
Where, V is the mean velocity of flow in m/s, g is acceleration due to gravity in m/s 2 and L is a
characteristic length in m. In open channel flow the characteristic length is made equal to the
hydraulic depth (D), which is defined as the cross-sectional area of the water normal to the
direction of flow in the channel divided by the width of the free surface. For rectangular channels
this is equal to the depth of the flow section.

In some handbooks this number is written as Fr2 = V2/gL.


When Fr2 is equal to unity equation 1-4 gives :

V  gD

and the flow is said to be critical state. If Fr2 is less than unity, or V  gD , the flow is sub
critical. In this state the role played by gravity force is more pronounced; so the flow has low
velocity and is often described as tranquil and streaming. If Fr2 is greater than unity, or
V  gD , the flow is supercritical. In this state the inertia forces become dominant; so the
flow has high velocity and is usually described as rapid, shooting, and torrential.

Systems involving gravity and inertia forces include the wave action by a ship, the flow in open
channels, the forces of a stream on a bridge pier, the flow over a spillway or a jet from an orifice
and other cases where gravity is the dominant force.

1.4 Velocity and Pressure Distributions in Open Channel

Owing to the presence of free surface and to the friction along the channel wall, the velocities in
open channel are not uniformly distributed. The measured maximum velocity in ordinary
channels usually appear to occur below the free surface at a distance of 0.05 to 0.25 of the depth,
the closer the banks the deeper the maximum.

As a result of non-uniform distribution of velocities over the channel section, the velocity head
of an open channel flow is generally greater than the value computed according to the expression
V2/2g, where V is the mean velocity. In practice usually average velocity across the flow is taken

9
and correction coefficients are applied. A correction coefficient called Coriolis coefficient
(energy coefficient)  is used in computations when energy principle is used, i.e. the velocity
head will be expressed as V2/2g. Experimental data indicate that the value of  varies from
1.03 to 1.36 for fairly straight prismatic channels. Non-uniform distribution of velocity also
affects computation of momentum (mv = QgV), where  is momentum coefficient or
Boussinesq coefficient generally 1.01 <<1.12 for straight prismatic channels. For channels of
regular cross section and fairly straight alignment the effect of non-uniform velocity on velocity
head and momentum is small especially in comparison with the uncertainty involved in the
computation. Therefore,  and  are assumed to be unity.

10
a) Parallel flow- hydrostatic pressure distribution b) Concave flow- centrifugal force pointing
downward to reinforce gravity c) Convex flow- centrifugal force acts against gravity (upward) as
a result resulting pressure is less than the otherwise hydrostatic pressure of a parallel flow.
Figure 1-5 Velocity and Pressure distribution in open channel flows

2. BASIC HYDRAULICS PRINCIPLES

2.1 Geometry of Open Channels

Open channels can be natural (rivers, streams, estuaries) or artificial (canals, flumes, chutes,
culverts, drops, open-flow tunnels). The channel geometry can be prismatic (constant So and
constant A) or non-prismatic. The geometry is defined by its cross-section and water depth (y).

Some conventions for naming of canal geometric parameters is given underneath:

11
Bs

1 Y

Figure 2-1Geometry of an open channel

Water depth (y) : Vertical distance bottom to surface (measured on a vertical plane)
Section depth (d) : Normal distance from bottom to surface, depth perpendicular to
the bottom (measured on a plane perpendicular to the canal
bottom)
Area (A) : area normal to flow direction
Wetted perimeter (P) : length of line of wetted intersection
Hydraulic radius (R) : R = A/P
Hydraulic depth (D) : D = A/BS = Area A/top width BS
Total energy (E) : E= z + y + αv2/2g (for small slopes θ with y = d)
Specific energy (ES) : ES energy in relation to lowest point in a section ES = y + αv2/2g
Velocity (V) : v = Q/A
Velocity head : αv2/2g = αQ2/2gA2
Froude number : Fr2 = αQ2BS/gA3
Friction Slope : Sf = Q2n2/A2R4/3 (Manning)
Friction Slope : Sf = Q2/C2A2R ( Chezy)

12
13
Figure 2-2 Presents geometric elements of the most common channel sections

2.2 Main Principles

To any situation in hydraulics we can apply two out of three principles:

1-a Principle of conservation of matter (continuity)


1-b Principle of conservation of energy (in case all energy “losses” are known)

2-a Principle of conservation of matter (continuity)


2-b momentum principle (in case all external forces are known)

2.3 Continuity Principle

In open channel flow the continuity principle for a constant discharge Q is:
 
Q  V1 A1  V2 A2

14
A = the cross-sectional area in sections 1 and 2,

V = the mean velocity in sections 1 and 2

If the mean velocity remains constant (V1 = V2) then the cross-sectional area A (A = Q/V)
remains constant. The open channel has a prismatic cross -section. This means for uniform flow
a constant water depth in all section. Uniform flow is a result from an exact balance between the
force of gravity and the frictional resistance.

Application of the continuity principle to unsteady, open channel flow is more difficult. In
unsteady open channel flow the water surface will change over a certain distance X  X 2  X 1 .
and during a certain time t .

During t : Inflow-Outflow = Storage


Q t  y x BS

As the velocity and the discharge will change over a distance.

x : Q  Q2  Q1  Q / x  x.

The discharge Q will vary with time t and with the distance along the canal.

If y is the water depth and Bs the width at the water surface, then the increase of volume between
the sections 1 and 2 during time t : Q t  Vo1  y Bs x.

The two terms derived are equal in magnitude, but different in sign:
Q y
 BS 0
x t
This is the equation of continuity for unsteady open channel flow.

2.4 Energy Principle

The energy equation and the momentum equation are used in addition to the continuity equation
in analyzing fluid-flow situations. They are both derived from Newton’s second law of motion.

15
In this section, the force components on a fluid particle in the direction of its motion are equated
to the product of mass of the particle and acceleration along the streamline. The equation is
obtained in differential form and requires the assumption of a frictionless fluid and steady flow.
The equation may then be integrated for particular assumptions regarding the relation between
density and pressure to obtain the energy equation.

Figure 2-3 Force components on a fluid particle in the direction of streamline.

In figure 2-3 let S be a streamline in steady flow, and consider the forces acting on a fluid
particle in the direction S of the streamline. The element has a prismatic cross-sectional area A
and length s – a frictionless fluid is assumed in order to eliminate all shear in the fluid (i.e. ideal
flow is assumed instead of real flow). On the upstream end the element the force is PA, in
which P is the pressure intensity at the center of the face. The force on the downstream end of the
element is :-

 dP 
P  s  A
 ds 
and acts in the negative direction. Any forces acting on the sides of the element are normal to the
streamline do not enter the equation. The only other force acting is due to gravity and is AS,
acting vertically down ward. The component in the S-direction is: -AS Cos . By
substituting in to  FS  mas .

 dP  
PA   P   S  A  A  S cos   A S as
 ds  g
After dividing through by the weight of the element  AS and simplifying.

16
1 dP a
 cos   S  0
 ds g

Z is the increase in elevation for a displacement S along the stream line and from Figure 2-3.
Z d
 cos   Z
S dS
dv
The acceleration as is . To keep this expression simple, the assumption of steady flow is
dt
introduced, so that V is a function of S only along a streamline. Writing,

dV ds dV
aS  . V
ds dt ds
ds
as the velocity along the streamline is . The equation becomes.
dt
1 dP dz V dV
  0
 ds ds g ds
By multiplying the above equation trough by g and with the additional assumption that the fluid
is incompressible (i.e. assuming a constant density in time and space) and integrating with
respect to S.
p V2
 gz   constant. Equation 2-1
 2
This is Energy equation (Bernoulli’s Equation) for steady flow of a frictionless, incompressible
fluid along a streamline. It states that energy per unit mass of a flowing fluid is constant, the
L2
dimensions for this equation are 2 , i.e. energy per unit mass.
T

Potential energy = mgZ  gZ is potential energy per unit mass.


2
1 V
Kinetic Energy ( mV 2 )  is kinetic energy per unit mass.
2 2
Work/Energy = F.S = ( P  A)  S
We can represent the unit mass as: unit mass =   A  S
PA s P
Therefore, Pressure Energy per unit mass = 
As  

By dividing equation 2-1 through by g we get on expression for energy per unit weight, or more
simply, meters. This expression is particularly convenient form to apply to situations with free
liquid surface, i.e. open channel flows.
p V2
Z    C
 2g

17
For open channel section with steady flow and straight and parallel streamlines, there is no
centripetal acceleration, i.e. where the hydrostatic pressure distribution holds true, the pressure
p
head lies in the water surface.

p
  d cos  - For channel with large slope

and

p
 y - For channel with small slope

Substituting for channel of large slope the total energy may be written as

V2
Z  d cos   C
2g
In general, every streamline passing through a channel section will have a different velocity
head, owing to the non-uniform velocity distribution in actual flow. Only in an ideal parallel
flow of uniform velocity distribution can the velocity head be truly identical for all points on the
cross-section. In the case of gradually varied flow, however, it may be assumed, for practical
purposes, that the velocity heads for all points on the channel section are equal, and energy
coefficient (Coriolis Coefficient ) may be used to correct for the over all effect of the non-
uniform velocity distribution. Thus, the total energy, equation at a channel section takes the form
v2
Z y   C
2g
According to this equation (the principle of conservation of energy), the total energy head at
upstream section 1 should be equal to the total energy head at downstream section 2 plus the loss
of energy hf between the two section ; or
2 2
v1 v2
Z1  y1    Z 2  y2    hf
2g 2g
This equation applies to parallel or gradually varied flow. Consider now a prismatic channel. The
line representing the elevation of the total head of flow is the energy line. The slope of the line is
known as the energy gradient, denoted by Sf. The slope of the water surface is denoted by Sw and
the slope of the channel bottom by So = tan . In uniform flow, Sf = Sw = So = tan .

18
Figure 2-4 Energy in gradually varied open channel flow

2.5 Specific Energy and Critical Depth

The “Specific energy” is the average energy per unit weight of water with respect to the channel
p
bottom. The piezometric head related to the bottom is  Z  y (with Z = 0) which is the

water depth. Therefore, the specific energy is the sum of the water-depth (y) and the velocity
head, if the streamlines are straight and parallel.

V2
ES  y  
2g
For a given section and constant discharge (Q), the specific energy is a function of water-depth
Q
only, since V  .
A
Q2
 Es  y 
2 g y 2 BS2
When the depth of flow is plotted against the specific energy for a given channel section and
discharge, a specific-energy curve is obtained (figure2-3).

19
Figure 2-5 Specific Energy Curve

This curve has to limbs AC and BC. The limb AC approaches the horizontal axis asymptotically
towards the right. The limb BC approaches the line OD as it extends upward and to the right.
Line OD is a line that passes through the origin and has an angle of inclination equal to 45o. At
any point P on this curve, the ordinate represents the depth, and the abscissa represents the
specific energy. Which is equal to the sum of the pressure head y and the velocity head V 2 2 g.

The curve shows that for a certain discharge Q two flow regimes are possible, viz. slow and deep
flow or a fast and shallow flow, i.e. for a given specific energy, there are two possible depths, for
instance, the low stage y. and the high stage y2. The low stage is called the alternate depth of the
high stage, and vice versa. At pint C, the specific energy is minimum. It can be proved that this
condition of minimum specific energy corresponds to the critical state of flow. Thus, at the
critical state the two alternate depths apparently become one, which is known as the critical
depth (YC). When the depth of flow is greater than the critical depth, the velocity of flow is less
than the critical velocity for the given discharge, and, hence, the flow is sub critical. When the
depth of flow is less than critical depth the flow is supercritical. Hence, Y1, is the depth of a
supercritical flow, and Y2 is the depth of a sub critical flow.

If the discharge changes, the specific energy will be changed accordingly. The two curves A’B’
and A”B” (Figure 2-3) represent positions of the specific energy curve when the discharge is less
and greater, respectively than the discharge used for the construction of the curve AB.

The critical state of Flow

20
The critical state of flow is defined as the state of flow at which the specific energy is a
minimum for a given discharge or it is the condition for which the Froude number (Fr2) is equal
to unity.
V2
ES  y 
2g
V
For Q 
A
Q2
ES  y 
2g A2
Differentiating with respect to y and noting that Q is a constant.

d ES Q 2 dA
 1
dy g A 3 dy

dA
The differential water area dA near the free surface (figure 2-5) is equal to BS dy Now  BS .
dy
A
and the hydraulic depth D  . So the above equation becomes.
BS

dE S Q 2 Bs Q2
 1  1 
dy gA3 gA 2 D
Q
But V  . Substituting
A
dE S V 2 A2 V2
 1  1   ES min imum
dy gA 2 D gD
dE
At the critical state of flow the specific energy is a minimum, or  0 . The above equation,
dy
therefore, gives.

v2 D

2g 2
This is the criterion for critical flow, which states that at critical state of flow, the velocity head is
V
equal to half the hydraulic depth. The above equation may also be written  1, which
gD
means Fr = 1; this is the definition of critical flow given previously.

21
2.6 Momentum Principle

The momentum equation is developed from Newtons second low of motion by summing up the
resultant force on a free body of fluid on one side of the equation and by developing the other
side of the equation into an equivalent expression in terms of rates of in flow and outflow of
momentum.

According to Newton's second law of motion the change of momentum (dmv) per unit time, is
equal to the resultant of all external forces acting on a body (body of water flow in a channel in
our case).
d (mv )
F 
dt

The momentum passing a section A per unit time (t = 1sec):

mv = * (A. V) V

dt

v
A

Where, A*V represent the volume passing per unit time.

mv = QV

Or including the momentum coefficient (Boussinesq coefficient)

mv =  *  * Q * v
Q2
mv =  *  *
A

Within the control volume defined in the figure there is an unknown energy loss and/or force
acting on the flow between section 1 and 2; the result is a change in the linear momentum of the
flow. In many cases, this change in momentum is accompanied by a change in depth of flow.
The application of Newton's second law, in a one dimensional for to the control volume, i.e.
equating the sum of all external forces (F) to the rate of change of momentum ( Q V) for any
two cross-sections 1 and 2 gives:

F  Q (V2  V )1

22
Or including the momentum coefficient

F  Q  2V2  1V 

Consider a channel section of mild slope and analyzes the forces acting.

F wsin

Ff F

W 

Thus, F =  (2v2 – 1v2) implies


F = F1 –F2 + w sin  - Ff =  (2 v2 - 1v1)

F1 and F2 are the resultant pressure forces acting on the two sections and w is the weight of the
water between the two sections. Ff is the total friction force acting along the surface of the body.
The slope is mild assume sin   So = 0
 W sin  = 0

and for a flow with parallel flow lines the pressure is assumed to be hydrostatic.

The force F1 is gAy+, where A is the cross- sectional area and of y+ the depth of the center of
gravity of the area A (the depth of centroid of the cross sectional area measured below the
surface of flow) . For rectangular section of small slope

F = g by (½y) (y+ = ½ y)

(Remark for canals with large slope F = g bdcos  ½ dcos  = ½ g bd 2 cos2)

Therefore, substituting

W sin  = o
F1 = g A1y1+
F2 = gA2y2+
F1 – F2 + w sin - Ff =  (2v2 - 1v1)

 g A1y1+ - g A2 y2+ - Ff = Q (2v2 - 1 v1)


  
 - Ff = g A2y2+ - g A1y1+ Q2  2  1 
 A2 A1 

23
  Q 2    Q2 
 - Ff = g  A2 y 2     A1 y1  1 
 A2 g   gA1 
Ff    1Q 2   Q2 
   A
 1 1y     A2 y 2    2 
pq  gA1   gA2 

The term between brackets is called the momentum function (M) or in general:

Q2
M = A y+ + 
qA
Ff
 = M1 – M 2
sq

For a given discharge Q, channel shape and coefficient  the function, M depends only on the
water depth y. Plotting M against ay gives a similar figure as for the specific energy Es against
depth y. This curve is called specific force curve. In the figure two regions can be determined,
namely sub and supercritical flow. For every M > Mmin two water –depths exist, which are
called the initial and sequent depth. Together they are the conjugate depths.
b
1
h
2

h  2b  B 
1   
3  bB 

h  2B  b 
2   
2  bB 

A 
1
b  B h
2
In applying the momentum principle to a short horizontal reach of a prismatic canal, the external
forces of friction and the weight effect of water can be ignored. Thus, with  = 0 and Ff = 0 the
equation becomes:

M1  M 2

24
Q2
A1 y1+ + 
Q2
= A2 y2+ + 
gA1 gA2
 2   2 
The momentum function M = y+A + consists of two terms. The second term   is the
gA  gA 
momentum of the flow passing through the channel section per unit time per unit weight of
water, and the first term (Ay+) is the force per unit weight of water. Since both terms are
essentially force per unit weight of water, their sum may be called the specific force.
Accordingly, it may be expressed as F1 = F2.

Thus, in analogy with the concept of specific energy) for a given value M, the M- y curve
predicts two possible depths of flow – conjugate depths of a hydraulic jump.
2
For a rectangular cross section y = ½ y. Assuming,  = 1 results in m = ½ Ay +
1
qA

The function per unit width (b = 1) reads

q2
M = ½ y2 +
gy

The function has on extreme for :

dM q2
 y  0 q  is disch arg e per unit width
dy gy 2


Y3 =

1
 q2 3 q2
Y = yc =   = 3
 g  g

d 2m q2
 1 2 3
dy 2 gy

d 2m 2q 2 2g
for y = yc   1 1 3
dy 2 q 
2
g
g  
 g 
The minimum value of the specific momentum function can be found under the assumptions of
parallel flow and uniform velocity distribution by taking the first derivative of M with respect to
y and setting the resulting expression equal to zero or

25
dm
 
 2 dA

d y A
0
 
dy gA2 dy dy

 2 dA
 A 0
gA2 dy

Bs ( dy ) 2
Where d (y+A) = [A (y++dy) + ]  y  A  Ady and where it is assumed that (dy)2 = 0 –
2
 A
Then substituting dA/dy = Bs, v = , and D 
A T

V2 D
We get  which is the same criterion developed for the minimum value of specific force
2g 2
(momentum) occurs at minimum specific energy or critical depth.

The given number 3 is positive, so for momentum value for y =yc . It has to be noted that the
given relation for yc is only applicable for open rectangular (prismatic) channels.

For other cross sections the impulse momentum equation should be used

P    2    2 
  A1 y1  12    A2 y2  22 
sg  A g   A g 

P
 M1  M 2
qg

With these equation it is possible to compute one unknown among the four variables (F, Q, A1
,A2) for a given channel shape and 1 =2 = 1; (y+ is also a function of channel shape and water
level).

The impulse momentum principle also follows from Newton’s second law. The flow may be
compressible or incompressible, real (with friction) or ideal (frictionless), steady or unsteady
moreover, the equation is not only valid along a streamline. The advantage of the impulse
momentum principle is that only the conditions at the end sections of the control volume govern
the analysis. It has a special advantage for application to problems involving high internal energy
changes, such as the problem of the hydraulic jump. If the energy equation is applied to such
problems, the unknown internal energy loss represented by hf is indeterminate, and the omission
of this term would result in a considerable errors. If instead the momentum equation is applied to
these problems, since it deals only with external forces, the effects of the internal forces, the
effects of the internal forces will be entirely out of consideration and need not be evaluated. The

26
term for frictional losses due to external forces, on the other hand, is unimportant in such
problems and can safely be omitted, because the phenomenon takes place in a short reach of the
channel and the effect due to external forces is negligible compared with the internal losses.

27

You might also like