Sunteți pe pagina 1din 14

Open-Channel Flow

Introduction
The flow of water in an open channel is a familiar sight, whether in a natural channel like that of a
river, or an artificial channel like that of an irrigation ditch. Its movment is a difficult problem when
everything is considered, especially with the variability of natural channels, but in many cases the
major features can be expressed in terms of only a few variables, whose behavior can be described
adequately by a simple theory. The principal forces at work are those of inertia, gravity and viscosity,
each of which plays an important role.
We shall consider only water as the fluid, as it is by far the most common one, and for which the most
experimental data are available. Other fluids will behave in about the same way, under conditions that
we shall specify below. Water has a density of about 1 g/cc or 62.3 to 62.4 pounds per cubic foot
depending on the temperature, and does not vary significantly for the temperatures and pressures that
we shall consider. Sea water weighs 64 pcf. In this article, I shall use mainly metric units, but may slip
into foot-pound-second units now and then. The acceleration of gravity will be taken as 980 cm/s 2 or
32.2 ft/s2. Variations in gravity with latitude and place will have very little effect, so we shall neglect
them. The kinematic viscosity of water under normal conditions is about 0.15 cm/s 2. It decreases
from about 0.20 at freezing to 0.0028 at the boiling point, or to 0.008 at 90F. The figure 0.15
corresponds to about 55F. The kinematic viscosity is the ratio of the dynamic viscosity to the density,
and can be considered as the diffusion constant of momentum due to molecular motion. Since the
density of water is 1, the dynamic viscosity is 0.01 g/cm-s, numerically the same. The shear force on
a surface is then = (dv/dz), where dv/dz is the velocity gradient at the surface. The modulus of
compressibility of water is about 300,000 psi, so water can be considered incompressible for the
present purposes.
All of our results will be based on the conservation of mass, momentum and energy (in the form of
Bernoulli's theorem), and the Manning formula for frictional resistance. It is amazing how far one can
go on this simple basis, without appeal to the general equations of hydrodynamics or the structure of
boundary layers.
The assumptions, methods and formulas that we will present here work for problems of water flows
on the usual engineering scale. The flows are long compared to their cross-sections, so that a single
velocity can describe the situation adequately at any cross-section. Flow depths and dimensions are
larger than the thickness of boundary layers, so that the flow is completely turbulent. Sizes are large
enough that capillary forces play a negligible role. Under these conditions, the analysis will normally
give very good results, of an accuracy sufficient for engineering purposes. The References will lead
into the extensive literature of the field.

Channel Flow
Consider the flow of water in a channel of arbitrary shape. The area between the free surface of the
water and the bed of the channel is A, which depends on the height of the water y above the bottom of
the channel. An average depth d can be defined by d = A/y. It is the depth which, when multiplied by a
certain height of the water, gives the area. Let the total amount of flow be Q cm 3/s. A more practical
unit is m3/s or ft3/s, often written cfs (the unit "second-foot" means the same as cfs, but should be
deprecated). 1 m3/s = 35.31 cfs = 106 cm3/s. An acre-foot is 43,560 ft3, so a flow of an acre-foot per
day is 0.504 cfs. A U.S. gallon is 231 in3, so 1 cfs = 7.48 gal/s or 449 gal/min.
If we know the discharge Q and the area A, we can calculate a quantity V = Q/A with the dimensions
of a velocity. If the velocity were uniform over the cross-section A, then this would be the velocity

that would give the observed discharge. The velocity is actually zero at the wetted boundary of the
channel, and increases to a maximum at the centre of the channel and a little distance beneath the
surface. The region near the wetted boundary is called the boundary layer, and most of the change in
velocity takes place across it. This means that the quantity V is close to the velocity of most of the
water, and we use it as if this were true.
In fact, the kinetic energy per unit mass, v2/2g, when averaged over the cross-section, is V2/2g,
where is a correction factor a little larger than unity (1.1 is typical), which we can use if we require
more accuracy in expressing the kinetic energy in terms of V. Similarly, the average momentum per
unit volume is V, where is a correction factor usually less than . The approximate independence
of our results on the actual velocity distribution is a consequence of the tbinness of the boundary
layer. We replace all the complexity of the velocity distribution by one number, the average velocity
V, and in most of our work it can even be taken to be the actual water velocity.
The surface velocity of, say, a river can be measured by timing a float placed in the center of the
stream. The distance between two points established by poles on either side of the river, divided by
the time required for the float to cover the distance, gives the surface velocity. The old rule of thumb
was that V is 0.8 of the surface velocity. Then, if you measure the depth of the river at good points in a
cross section, its flow area A is known, from which the discharge is Q = AV. Discharge is more
accurately measured by a sharp-ceested weir, and approximate gauging is now done by a propeller or
other device similar to an anemometer that measures the velocity at some selected point and is
calibrated for the particular stream.

Flow Velocity Formulas


It's clear that water flows more rapidly the steeper the slope. On a constant slope, the velocity reaches
a steady value when the gravitational force is equal to the resistance to flow. For solid bodies, the
frictional force is about independent of velocity, so they just go on accelerating if they move at all,
and never reach a steady state, ideally. Fluids are different. The shear force between the liquid and its
bed depends on the square of its velocity, which is suggested by dimensional analysis. The frictional
force per unit area is approximately = CV2/2, where C/2 is just a constant that must be determined
empirically. There is no physics in this; we have just used dimensional analysis and the wonder is that
it turns out to be almost right. Therefore, we are pretty sure water will reach a steady velocity in
flowing down an incline.
The slope S is the ratio of fall h to distance along the channel, S = h/L. In surveying, slope is usually
referred to horizontal distance. In fact, the slopes of water channels are usually so small that the
difference is negligible. A slope of 0.01, or 1 in 100, is a rather steep slope for a channel, and 0.001 or
even less is more typical. The weight of a length L of water with cross-sectional area A is gAL. If the
channel slopes, the component of the weight along the channel is gAL(h/L) = gAh. This has to be
balanced by the force exerted by the bed of the channel, PL, where P is the wetted perimeter of the
channel. The frictional force is exerted only along this perimeter. Therefore, gAh = PL = CPLV2/2,
and so V2 = 2gCAh/PL = 2gCRS, where S is the slope and R = A/P is called the hydraulic radius.
Then, V = C'(RS), which is Ch&eaucte;zy's formula (1775). Note that the coefficient C' has
dimensions, and so differs when different units are used. Let us agree that in the basic formula, R will
be in m and V in m/s. Then, C' is on the order of 100. If we use cm, then C' = 10, and if we use feet, C'
= 181.
Chzy's formula is not bad, but did not provide results that satisfied engineers. The Swiss engineers E.
Ganguillet and W. R. Kutter (1869) showed that much better results could be obtained if the constant
C' depended on R, S and a constant n that was characteristic of the roughness of the channel as
follows: C' = (23 + 0.00155/S + 1/n)/[1 + n(23 + 0.00155/S)n/R]. This, of course, is for metres. For
feet, it is multiplied by 1.811, and the parenthesis in the denominator is also multiplied by 1.811. As
an example, consider a smooth iron pipe for which n = 0.011, S = 0.001 and R = 1.0 m. Kutter's

formula gives C' = 90.9, and V = 2.87 m/s. For many channels of metal, concrete and wood, n = 0.011
to 0.017. For canals in earth and smooth natural channels, n = 0.025 to 0.033, while for weedy
streams, n = 0.075 to 0.150. This will give some idea of the variation of n. Better values can be found
in engineering references.
Kutter's formula works well, but is very difficult to use in iterative and theoretical calculations
because it cannot be solved for R or S easily. Manning found a formula that gives results very close to
Kutter's, but is much simpler, and which is now used almost exclusively. Manning's formula (1890) is
V = R2/3S1/2/n, where n is the same as in Kutter's formula. It is absolutely remarkable that such a
simple formula gives such good results, and that it uses the same roughness parameter n. Working the
same example as in the preceding paragraph, we find V = 2.87 m/s, so the agreement is excellent.
Manning's formula is easily solved for R or S, a great advantage. If you use feet, multiply by 1.49.
Let us use a simple rectangular channel as an example. Many channels are actually rectangular, or
close to it, so this is a useful example. We will show the modifications necessary when the channel is
not rectangular, and will note that the results are pretty much the same qualitatively. Let y be the depth
of the water above the bottom of the channel, and b the width of the channel. Then its area is yb, and
its wetted perimeter is b + 2y, so its hydraulic radius is yb/(b + 2y). The discharge is Q = AV = byV, or
the discharge per unit width is q = Q/b = yV. Using Manning's formula, q = y[yb/(b + 2y)] 2/3S1/2/n.
Now, if we know y, R and S we can just substitute in the equation and find q. In many problems,
however, we are given some desired discharge, and want to find the corresponding y and V. This was
once an arduous process, involving an iterative solution. Now, with a tool like the HP-48G, it is
straightforward. Use the SOLVE facility. All you have to do is express the equation in terms of
numbers and the variable y, type it in, and press SOLVE. Even in Daugherty's excellent and accurate
text, on p. 313 this is done manually with an incorrect result, 2.17 ft instead of 2.47 ft. The HP-48 not
only saves work, but more importantly it is less subject to error.
Many of our illustrations will be for an even simpler example, the wide rectangular channel in which
only bottom friction is present. If b is the width of the channel (b >> y) then the discharge per unit
width is q = Q/b. The area A = by, and the wetted perimeter is b, so R = y. Results for the wide
rectangular channel will be similar to those for a channel of any shape.
Given S and Q, and the shape of the channel, we can find y and V; given S and y, we can find Q. If we
do not have a rectangular channel, but one of arbitrary shape, we must work with Q instead of q, and y
is usually the depth from the lowest point in the channel. Then the area A occupied by fluid and the
wetted perimeter P are functions of y, which may be expressed analytically for channels of a
geometric shape. Then, R can be expressed as a function of y, and used in the Manning equation. We
have, in fact, done this for the rectangular channel. All we have to do is replace the square bracket
above by R(y). If R is expressed as, say, a table of values, the best way to proceed is to plot or tabulate
q as a function of y, and then enter the table with q instead of y. You should feel comfortable with
these calculations after a little practice. They can solve many problems all by themselves. The value
of y corresponding to a given Q and S is called the normal depth for those conditions.
Williams and Hazen's formula, V = 1.32CR0.63S0.54 is for nonferrous pipe. C = 120 for bad, 130 for
average, and 140 for new pipe. The formula also applies to rock tunnels, with C = 38 for bad, 45 for
ordinary and 50 for best. For steel pipe, Scobey's formula is V = CR 0.58S0.526, with C = 154 for smooth
welded pipe, 120 for riveted. These empirical formulas are quite close to Manning's, and not too far
even from Chzy's.

Specific Energy
Bernoulli's theorem expresses the energy content of unit volume of fluid as U = gz + p + V 2/2, and
states that it is constant along a streamline in the absence of dissipation. It is usually more convenient
to express the terms as lengths, or "heads," by dividing by = g: H = z + p/ + V 2/2g. The three

energy components are elevation, pressure, and velocity. All play a role in open-channel flow. For any
flow, there is an energy grade line that can be imagined above the flow, and its slope is S'. The water
surface is the hydraulic grade line (HGL), which is below the energy grade line by the velocity head
V2/2g. Below this is the bed grade line, with slope S, and (usually) below that is the horizontal datum,
the reference surface.
The streamlines of the flow are parallel. Along any streamline, z + p/ + V 2/2g is a constant. Let us
now assume z is the elevation of the bottom streamline, so that if the flow depth is y, the elevation of
the surface streamline is z + y. The gauge pressure here is zero, so Bernoulli's Equation for this
streamline is z + y + V2/2g = C. Now for the bottom streamline, the gauge pressure is y, so that p/ =
y, and Bernoulli's equation is z + y + V2 = C, where C has the same value as for the top streamline. At
any intermediate height y', z + y' + (y - y')/ + V2/2g = C. Therefore, C, the energy per unit weight,
has the same value at any depth. The part y + V2/2g is called the specific energy E, and is the energy
per unit weight referred to the stream bed.
When a closed channel runs full, then the depth can no longer vary to accommodate the discharge,
and the pressure becomes different from the atmospheric pressure, and must be taken into account in
using Bernoulli's theorem. This is the fundamental difference between open channel flow and pipe
flow.
We can express E as a function of Q easily, using Q = AV: E = y + Q 2/2gA2. For simplicity, consider a
rectangular channelof width b, for which A = by. Then E = y + q 2/2gy2, expressing E as a function of
the discharge q and the depth y, or q = y[2g(E - y)], expressing q as a function of E and y.
The curve of q as a function of y for a fixed E is plotted at the right. We notice that q is a doublevalued function of y, and has a maximum possible value qm. The corresponding depth y can be found
by differentiating q with respect to y and setting the derivative equal to zero. The result is y c = 2E/3,
called the critical depth. The corresponding value of q, qm = (gyc3) is the critical flow, and Vc =
qc/yc is the critical velocity. For depths greater than the critical depth, the velocity is smaller than the
critical velocity. Flow in this region is called subcritical. For depths smaller than the critical depth, the
velocity is greater than the critical velocity. Flow in this region is called supercritical. Note that the
sub- and super- refer to the velocity of flow. The same discharge q is possible with given E in either
region. In the upper region, we have greater flow area, in the lower region greater flow velocity.
Because the frictional resistance varies rapidly with
velocity, subcritical uniform flow is associated with
gentle slopes, supercritical uniform flow with steep
slopes.
Note that the curve is plotted with respect to
dimensionless variables, so the same curve can be
used for any E or qm. Consider flows described by
points a and c. Since they are on the same vertical
line, the discharge is the same for each. The
distance from the y = 0 axis to point a corresponds
to the static part of E, while the distance from a to
the y = E line corresponds to the kinetic energy,
which sum to E. The same holds for point c, but
here the static part is much smaller and the
dynamic part larger. At the point of maximum
discharge for this value of E, point e, the static
energy is twice the dynamic energy.
Suppose that there is a lateral constriction in the channel, reducing its area so that q/q m increases from
0.6 to 0.8. Assume there is no head loss, the specific energy does not change, so the flows in the

constriction are represented by points b and d. We note that in subcritical flow, the depth of flow
decreases, while in supercritical flow the depth increases.
We can also plot y/yc as a function of E for a constant discharge q, as shown at the left. Again, this is
for a rectangular channel but diagrams for other channel shapes are similar. This curve is also plotted
with dimensionless variables, E/yc vs. y/yc, so it can be used for any discharge. The critical depth y c =
(q2/g)1/3, where q is the discharge per unit width. If x = y/yc, then E/yc = x + 1/2x2. Critical depth
corresponds to y/yc = 1, for which E/yc = 3/2. The point a corresponds to an upper-stage or tranquil
flow. Line segment hf represents y/yc, while line segment fa represents the kinetic energy V2/2gyc.
Note that the specific energy is a minimum at the critical depth. This minimum value will, of course,
depend on the discharge.
Let us suppose there is a hump in the
bed of the channel that decreases the
specific energy from E1 to E2. The
height of the hump will be the decrease
in the specific energy. If the flow is
subcritical, we see that depth will
decrease slightly to point b. If the flow
is supercritical, the depth will, on the
other hand, increase slightly to point e.
This is exactly the same as the response
to a lateral constriction. If the hump is
high enough, the flow may become
critical at it. For any larger hump, the
specific energy cannot decrease further,
and instead the upstream depth must
increase to keep the flow critical over
the hump. For this reason, such a point
may be called a control section, since it
controls the upstream depth. The
vertical tangent at critical depth means
that small changes in E will cause large
changes in y, so the surface may appear disturbed.
On any declining slope S, uniform flow will be established (if possible) at a depth y called the normal
depth for a given Q. If the normal depth is greater than the critical depth, then the slope S is
called mild and the flow tranquil orupper-stage. If it is smaller, then the slope S is called steep and the
flow is called rapid or lower-stage. (or, of course, subcritical or supercritical.) Whether a slope is mild
or steep depends on Q and n, the discharge and roughness. A slope giving exactly the critical velocity
is called, unsurprisingly, critical, and is not often found. There are also horizontal (S = 0)
and adverse (S < 0) slopes, on which uniform flow is not possible, but such slopes are always of
limited extent.
Since yc = 2E/3 and E = y + V2/2g, it is easy to see that V2/2g = yc/2, or Vc = (gyc). Since q = y[2g(E
- y)], we find qmax = (gyc3). That this is indeed the maximum q for the given E can be found by setting
dq/dy = 0 and solving for y, with the result y = y c. Maximum discharge occurs for critical flow.
We also note that Vc2/gyc = 1. The dimensionless term V/(gy) is called the Froude number F. It
expresses the relative strength of inertial and gravitational forces, and was first used in ship modelling
in the estimation of wave drag effects. If the Froude numbers of two flows are the same, then these
effects will be similar. In the present case, F = 1 corresponds to critical flow, F < 1 to subcritical flow,
and F > 1 to supercritical flow.

The Reynolds number of an open-channel flow is defined as Re = 4RV/, where the kinematic
viscosity has been given above. It expresses the ratio of inertial forces to viscous forces. With R in
m and V in m/s, Re = 2667RV. In nearly all practical cases, Re >> 1, so the flow is turbulent.
When we use specific energy, we are really using Bernoulli's principle, where the total energy H = z +
E. Energy is conserved except for friction, and this causes H to decrease with distance at a rate s', so
the loss of head in a distance L is hL = Hs'L. s' can be found from Manning's formula: s' = [nV/R2/3]2.
For a fixed Q, the velocity V is found from the depth y by continuity, Q = VA(y). Therefore, for any
depth of water we can find the frictional head loss. If the bottom of the channel varies in elevation,
then E, which is the difference between H and y + z, varies accordingly.
These ideas are summarized in the diagram at the right, which shows a reach between 1 and 2 in
which the flow is uniform. The direction of flow is, of course, the direction in which the energy line
(EL) falls. Note that the slope of the EL (s) is the same as the bed slope (s'). The average velocity V
has adjusted itself to make this so, determining the depth y from the known specific energy E or
discharge Q. This depth is the standard depth, and the velocity is the standard velocity, for this flow.
The hydraulic grade line HGL is the water surface, which is also parallel to the EL and the bed. The
critical depth yc is the lower boundary of the region for which the specific energy decreases with
increasing depth (subcritical flow) and the upper boundary of the region where the specific energy
increases with a decrease in depth (supercritical flow). The critical depth is the solution of the
equation Q2/g = A3/b, where b is the width of the water surface. For a rectangular channel of width b,
Q2/g = b2y3, or yc = [q2/g]1/3, where q is the discharge
per unit width. In the diagram, y > yc, so the flow is
subcritical.

Nonuniform Flow
The slope s' of the channel may be different in
different reaches; the channel may change width or
shape, there may be humps and hollows in the
channel, or weirs and sudden drops, and other factors
that change the flow conditions. The resulting flow
will be steady, although the elements of the water
will experience acceleration from point to point. We
can usually make a good approximation to the flow by using Bernoulli's theorem and dividing the
problem into lengths of approximately uniform conditions. We will want to know how the depth y and
the velocity V vary with position, as well as the other characteristics of the flow. These problems are
quite interesting, have many practical applications, and show the power of engineering hydraulics.
Consider a hump in the water surface, a surface wave travelling down the channel with some velocity
V. That such waves exist is an experimental fact. Now suppose the water in the channel is moving
with velocity V in the other direction; the wave will appear to stand still. For simplicity, suppose the
channel is level, S = 0, at least for a short distance. Then, from Bernoulli's theorem, E = E', where E is
the specific energy approaching the wave, and E' is the specific energy at the centre of the wave. Let
the depth increase from y to y' = y + y, so that y is the amplitude of the wave. From E = E' we
easily find that V2/2g = (y' - y)/[1 - (y/y')2]. Now substitute y + y for y', and expand in powers of y.
The result is V = (gy)(1 + 3y/4y + ...). For waves of small amplitude, the phase velocity is c =
(gy).
The general expression for the velocity of surface waves of wavelength is c =
[(g/2)tanh(2y/)]1/2 (vide Lamb, p. 364), where y is the depth of the water. The two limiting cases
are shallow water (or long wavelengths), y << , when c 2 = gy, independently of the wavelength, and
deep water, y >> , when c2 = g/2. In deep water (or short wavelengths), the velocity is proportional
to the square root of the wavelength, as in the ocean, where waves of different periods separate

themselves from each other when propagating long distances. The group velocity in this case is half
the phase velocity. The propagation of gravity waves on a water surface is said to be dispersive.
We observe both kinds of waves on the surface of the water in our channels. The most important thing
to us is that the speed of long surface waves is exactly the critical velocity! If the water is moving
faster than the critical velocity, as in rapid flow, then wave disturbances cannot propagate upstream.
An obstacle sticking up throught the surface creates a spreading wave like a supersonic bullet, with a
half-angle of = sin-1(Vc/V). On the other hand, in tranquil flow waves from a disturbance can
propagate upstream as well as down. This is one way to determine whether an observed flow is
supercritical or subcritical, by simple observation.
Suppose that the slope increases from mild to steep at a certain point. The upstream normal depth y is
greater than yc, while the downstream normal depth y' is less than yc. The depth decreases as E
decreases on the upper-stage curve, then continues to decrease as E goes through a minimum at the
critical depth near the break point of the profile, and then E increases again until the normal depth for
the steep slope is reached. If we try to decrease E below the minimum by raising the bed of the
channel by z, of course E cannot decrease further, and the result will be to raise the upstream water
level by z. Because of this, the point at which critical depth is reached is called a control section,
because it controls the upstream depth.
If the slope decreases from steep to mild, something very different takes place. As y increases toward
the critical depth, a flow instability occurs at some point, and the flow becomes turbulent until the
new normal depth is attained downstream in tranquil flow. This is called a hydraulic jump, which will
be analyzed below.
If H = z + E, then dH/dx = dz/dx + dE/dx, where x is distance along the flow. Now dH/dx = s, the
slope of the energy line, and dz/dx = s', the slope of the bed of the channel. Therefore, dE/dx = s - s',
where dE/dx = dy/dx + d(V2/2g)/dx. Since q = Vy = constant, Vdy + ydV = 0. Then d(V 2/2g) =
(V2/gy)(-dy), so s - s' = (1 - V2/gy)dy/dx = (1 - F2)dy/dx. Hence, dy/dx = (s - s')/(1 - F2). This equation
tells us whether the water surface is rising or falling in the direction of motion. In uniform motion, s' =
s and dy/dx = 0. For upper-stage flow, the sign of dy/dx is the same as the sign of s - s'. For lowerstage flow, the sign of dy/dx is opposite. The dimensionless parameter F is the Froude Number. F > 1
corresponds to supercritical flow, F < 1 to subcritical flow. It is analogous to the Mach Number in
compressible fluid flow. It was originally defined by Froude as the speed of a ship divided by the
square root of its water level length (not dimensionless). The relative wave resistances of hulls of
different sizes and speeds are the the same if the Froude numbers are equal, permitting resistance tests
on model ships.
The equation for dy/dx must be used with care near critical depth, since it predicts an infinite slope
there. When water is flowing at critical depth, the surface is typically disturbed. However, the
equation cannot account for the details of flow in this region, and, of course, the infinte slopes are not
observed.
The figure at the right shows a reach in which s &neq; s'. In the case shown, s > s', so the specific
energy decreases from 1 to 2. Since the flow is subcritical, this means that the depth decreases while
the velocity increases. The flow is no longer uniform, but is still steady, and the discharge is constant
at any cross-section. It is easy to get an expression for the change in specific energy by equating the
vertical distances at 1 and 2: s'L + V12/2g + y1 = y2 + V22/2g + sL. This relation is usually rearranged to
give the distance L, when the depths at each end of the reach are assumed: (s' - s)L = E 2 - E1.

Suppose we have uniform subcritical flow on a mild


slope, and let us modify the downstream end of the
channel. One modification would be to create a dam
that would raise the depth in front of it. The
discharge must get by the dam somehow, either by
flowing over the top or through a gate, for example.
A backwater deeper than the normal depth would
form that would slowly approach the normal depth as
we pass upstream. Or, we could allow a free
discharge from the end of the channel that would
then fall as a free jet. In this case, we would have
a drawdown water surface with a depth less than the
normal depth, approaching normal depth as we go upstream. These surface profiles are denoted
M1 and M2, respectively, by Bakhmeteff. If the water discharges through a gate at the bottom of a dam
with water behind it at greater than critical depth, its velocity will be greater than critical. The depth
will then rise as the water decelerates because of the large resistance. Before it reaches critical depth,
a hydraulic jump will occur, making the transition to subcritical flow. The resulting profile is denoted
M3. The water profile can be calculated by the formula just derived, starting from some point where
the depth is known and finding the distances to points where the depth takes a series of increments
approaching the normal depth.
On a steep slope, the normal depth is less than the critical depth, so the water profiles are different
from those on a mild slope. If the depth is greater than the normal depth (but less than the critical
depth), it will approach the normal depth as it accelerates downstream, while if the depth is less than
normal depth, the water will decelerate with increasing depth, approaching normal depth
asymptotically. These are the profiles S2 and S3, respectively. A supercritical flow approaching a dam
will undergo a hydraulic jump and become subcritical, then rise on the steep slope with a backwater
curve designated S1. Downstream influences do not affect upstream flows when the velocity is
supercritical. In particular, there will be no drawdown curve approaching a free exit.
These six flow regimes cover most applications, but special profiles can be identified for critical
slopes (normal and critical depths equal), and for horizontal and adverse slopes (normal depth
infinite). The profiles for horizontal and adverse slopes are similar to those for a mild slope with an
infinite depth, but of course normal uniform flow is not possible. Diagrams of these profiles are
shown in Daugherty and Franzini (p. 306), and in Urquhart (p. 4-90).
When the channel slope changes from mild to steep, the initially uniform subcritical flow must change
to the finally uniform supercritical flow, while the flow velocity accelerates from its initial value to its
final value. This transition is smooth and efficient, with little additional loss of head, like converging
flow in a pipe. In the region of acceleration on the mild slope, the velocity is higher than its normal
value, so the rate of head loss is greater. The energy line now approaches the profile line, so that the
specific energy (which is the vertical distance between them) decreases. From the specific energy
diagram, we see that the flow depth decreases as a result. At the break in bed slope, the flow velocity
is less than the normal value, so the resistance is less, and the energy line and profile line now diverge,
and the specific energy increases. The point of the break, then, is the point of minimum specific
energy, and the corresponding depth of flow is the critical depth for the given discharge. Because of
the rapid change of depth with specific energy (vertical tangent to the specific energy curve) there
may be surface disturbances at this point. As the specific energy increases further on, the depth of
flow decreases until it finally approaches the normal depth for the profile gradient. Similar flows
occur at the outflow of reservoirs and spillways, but the location of critical depth may be somewhat
further upstream. This transition is always characterized by a decrease in depth of flow in the
direction of flow.

At a free outfall from a mild slope, the specific energy decreases up to the end of the channel.
Nevertheless, critical depth is reached before this point, typically a distance 4y c from the lip. The
depth at the lip is about 0.7yc. The reason for this is in the change in the flow pattern near the end. The
kinetic energy factor may decline to near 1.0 at the exit, and the curvature of the streamlines may
reduce the pressure by centrifugal force. In the free jet, the velocity may well be uniform across the
jet, very different from the velocity distribution in the channel. It is
usually accurate enough to assume that critical depth is reached near
a crest of the flow.

Water Profiles
Let's use the relation we have just obtained to find the water levels in
a practical problem. Suppose we have a rectangular channel of width
b = 6 m and slope S = 0.002, which is 1 in 500, or 10.5 ft/mile. The
discharge is Q = 100 cms (cubic metres per second), and the roughness is n = 0.012. The depth of
steady flow yo can be found from Manning's formula. We have 100 = (6y)(1/0.012)[6y/(6 +
2y)]2/3(0.002)1/2, which we solve for y = yo. The HP-48G gives us yo = 3.3066 m. Then Vo = 100/(6yo)
= 5.040 m/s, so Eo = 4.6028 m. The critical depth yc = (2/3)E = 3.0685 m. Since this is less than the
actual depth, the flow is tranquil or upper-stage, and the slope is mild.
It is easy to repeat the calculation for any Q. Simply edit the constant value in the equation, and use
SOLVE again. For example, if Q = 10 cms, then yo = 0.669 m, V = 2.492 m/s, E = 0.986 m and yc =
0.657 m.
Now suppose we have placed a dam with a spillway height of 8 m above datum at the lower end of
the channel. We assume that the width is the same, 6 m, for simplicity, and that the height of the water
over the spillway sill will be the critical depth at that point. Since the critical velocity V c = (gyc) and
Q = bVc = b(gyc), we can solve for yc, with the result yc = 3.05 m, and Vc = 5.467 m/s. The height of
the energy line at this point is then 8 m + (3/2)yc = 12.57 m.
Now we can estimate conditions at the end of the channel at the dam. We assume that y 2 = 12.57 m,
neglecting the small contribution of the velocity, which is only V = 100/(12.57)(6) = 1.326 m/s. This
is not quite correct, but the error will not be large, so a more detailed estimate is not worth the effort.
Make a table with columns for y, A, P, R, V, V2/2g and E, for calculating the specific energy at the
ends of a series of reaches upstream. For this station, the numbers are 12.57, 75.42, 31.14, 2.422,
1.326, 0.0897 12.660. Instead of trying to find the conditions at some distance L upstream, it is much
easier to assume a new depth, and then find the distance L that corresponds.
Therefore, take a new depth at y1 = 12.00 m. The columns of the table can be filled in with little
trouble. The numbers are: 72.00, 30.00, 2.400, 1.389, 0.0984, 12.094. The specific energy has
decreased from 12.660 m to 12.094 m in the reach, so the change in E is -0.566 m. We now calculate
the slope S' of the energy line using Manning's formula. Average the velocities and the hydraulic radii
at the ends of the reach. We find V = 1.358 m/s and R = 2.411 m. The energy slope that corresponds to
this velocity is S' = [nV/R2/3]2 = [(0.012)(1.358)/2.4112/3]2 = 8.214 x 10-5. Because the velocity is so
low, this is really a small frictional loss. Then S - S' = 1.917 x 10 -3. Then L = 0.566/1.917 x 10-3 = 295
m. The minus sign just means we are going upstream.
Conditions in the last reach of the channel are shown in the figure at the left. The elevation of the
water surface has dropped only 2 cm in this distance, though y has decreased by 57 cm. The depth is
still 12 m at the entrance of the reach, so we have about 4.5 km to go before it gets down to 3.31 m,
and will thereafter be parallel to the grade line. The water surface is called abackwater curve. We have
calculated only the first section, but it is clear how to proceed. We could, for example, take depths of
11.5, 11.0, 10.5, ... and get to normal depth in about 18 reaches. This is a good job for a computer,

which makes calculating water profiles easy and error-free. Why not compute the next step yourself,
and find out that this is really not hard at all.
Putting in the dam created a backwater, and it is
easy to see why the dam is called a control section.
Since we know that the depth will be the critical
depth (at least somewhere in the vicinity) we can
calculate the critical depth from the discharge Q,
and the height of the energy line is determined at
that point. Nothing we do to the grade line upstream will make the slightest difference to this. Raising
the dam will push the backwater upstream, and lowering it will bring the backwater downstream. Also
note that we can find the discharge by simply measuring the height of the water on the spillway. This
is the principle of using weirs to measure flow quantity.
If, instead of a dam, we simply made a free outfall at the end of the channel, we would again get
critical depth in the vicinity. Because of the curvature of the streamlines at the lip of the outfall, the
point of critical depth is moved upstream by 3 to 10 times yc. The actual depth at the lip is about 0.72
yc. Now the depths are less than the normal depth, and the depth approaches the normal depth from
below as we move upstream, instead of from above, as in a backwater. The water surface is now
called a drop-down curve, and can be calculated in exactly the same way as a backwater curve.
The backwater curve and the drop-down curve on a mild slope are the most commonly observed
water profiles. The flow is tranquil or upper-stage. The water level rises when the velocity is retarded,
and falls when the velocity is increased. If the initial flow is rapid or lower-stage, then the water level
is certainly less than the normal level, as well as the critical depth, and the velocity much greater. The
velocity must decrease, since it cannot be supported by the mild slope, and the water level must rise.
In this case, however, the water does not placidly go through the critical depth, but before this
happens a hydraulic jump occurs that makes a turbulent transition to mild flow at a higher level.
Water profiles can be classified by the relative values of the depth, the normal depth and the critical
depth, and whether the slope is mild, steep, critical, horizontal or adverse. There are 12 possibilities,
of which we have mentioned the 3 occurring on a mild slope, which are by far the most common.
There are 3 kinds of water profile on a steep slope, all of which involve a hydraulic jump here or
there. There are only 2 kinds of profiles for the special cases of the slope, which do not differ greatly
from those on mild or steep slopes. For example, the drop-down curve for a horizontal grade is just
like the one for a mild slope. Daugherty gives sketches of the 12 nonuniform flow types (p. 306).

Hydraulic Jump
The transition from supercritical to subcritical flow where the profile changes from steep to mild is
very different from its inverse. Because the incident flow is more rapid than the wave velocity,
information of the coming transition cannot move upstream. In essence, the incident flow collides
inelastically with the slowly-moving ahead. Momentum is conserved, but energy is not, and there is
usually a significant head loss, and the production of large-scale turbulence. The collision is called
a hydraulic jump. A hydraulic jump is characterized by a depth of flow that increases in the
downstream direction.
A hydraulic jump is analogous to a shock wave in aerodynamics. It is a turbulent, non-energyconserving process that passes from rapid flow to tranquil flow. Unlike a shock wave, its front is not
vertical, but the depth increases to its final value in a distance about equal to six times the final depth.
Since energy is not conserved, the before and after states do not correspond to the intersection of a
line E = const. with the specific energy curve. However, momentum is conserved, and this permits us
to find the final state if we know the initial state. We will assume the channel is horizontal, since the
hydraulic jump takes place in a limited length of channel, and gravitational energies will not be

important. Let the water enter the jump with velocity V1 and height y1, and leave with velocity V2 and
height y2. In one second, the momentum lost by the fluid in passing through the jump is Q(V 2 - V1),
and this must be equal to the difference in pressure forces on the two cross sections. This difference is
gh1A1 - gh2A2, where h is the distance from the surface to the centroid of area A. This is one place
where the pressure plays an explicit role. Equating the rate of momentum change with the net force
acting, we find that the combination hA + QV/g is conserved. For a rectangular channel, A = by and h
= y/2m so a conserved quantity is f = q2/yg + y2/2.
We can write this in dimensionless form by setting x = y/yc, as we did for the specific energy above.
Then, f/yc = 1/x + x2/2. We shall call a plot of f/yc vs y/yc the hydraulic jump curve. Curiously, it is just
the specific energy curve with 1/x in place of x. The hydraulic jump curve for a wide rectangular
channel is shown at the right. The dashed curve is the specific energy curve for the same discharge q.
An initial state a is taken into state b by the hydraulic jump. We can combine the hydraulic jump curve
with the specific energy curve to find the energy loss in the jump. The connection between the curves
is through y/yc. Horizontal lines through a and b will intersect the specific energy curve at the
corresponding energies. It is easy to see that there will always be a decrease in specific energy in the
hydraulic jump, since the curves cross at x = 1. Hydraulic jump curves for other channel shapes will
be similar.
We can solve the equation f1 = f2 to
obtain x2 = (1/2)[-x1 + (x12 + 8/x1)],
relating the heights on the two sides of
the hydraulic jump, which are
called conjugate depths. The length of
the jump is about 6 times the greater
height y. If the channel is not horizontal,
in the momentum balance we must add
the weight of the water between y1 and
y2 times the sine of the slope, which is
S, approximately, since the angle is
small.
When the slope changes from steep to
mild, a hydraulic jump occurs at a point
such that y2 is the normal depth on the
downstream end. Depending on the
corresponding value of y1, the jump may
occur either after or before the break in
slope, wherever the required value of
y1 can be found. If y1 is greater than the
depth on the steep slope, then the jump will occur on the mild slope when the height of the rapid flow
reaches y1. If it is not, then the jump will occur on the steep slope to a depth that becomes the normal
depth on the mild slope.
This may be illustrated by the jump shown in the figure at the left. The water comes through a gate
with high velocity and a depth less than the critical depth, so the flow is rapid. The depth conjugate to
y1' is y2'. Since this is greater than the normal depth on the mild slope, the jump will not occur
immediately, because there is insufficient energy. The rapid flow decelerates on the mild slope, and
the water depth increases, bringing down the conjugate depth. Both the energy line and the water
surface are slightly concave upwards in this region, though drawn as straight. When the depth reaches
y1 conjugate to the normal depth on the mild slope, the jump begins, at the first point where it is
possible. It does not occur at a single section, but extends for the length of the turbulent jump
(foreshortened in the diagram). It is easy to see that the momentum on the left is balanced by the
pressure force on the right. The energy line falls rapidly through the region of the jump, then resumes

with the mild slope, S' = S. There is also a small energy loss at the gate, from the level of the slack
water upstream. There are other cases, but this is the most common one. A good place to see a jump is
at the overflow spillway of a dam, with a rapid flow down the inclined face of the spillway, and a
jump on the apron beyond in the tail water.
A hydraulic jump is probably most easily seen
if a jet of water issuing from a spout falls
normally on a stainless-steel kitchen sink (or
in any similar case). Immediately around the
point where the jet impacts there is a circular
region of smooth flow, then a hydraulic jump
when the velocity of the radially diverging
water falls below the wave speed. Beyond the
jump is a noticeably turbulent flow.
Related to the hydraulic jump is the tidal bore on a river, occurring when the river is shallow, has a
funnel-shaped estuary, and the tidal range is large. The jump is not stationary, but moves up the river
at a considerable speed, faster than that of surface waves, driven by the pressure of the higher water
behind it. The turbulent front may be followed by undulatory disturbances which move with it and
are solitons, waves in which nonlinear effects and dispersion combine to create solitary waves.
Hydraulic jumps and solitons may often be seen on smooth beaches in wave backwash.
Perhaps the most well-known bores are the equinoctial undulatory bores on the River Severn in
western England, easily visible from Sheerness up to the Over bridge near Gloucester. They can be up
to 2 m high. The tidal range at Sheerness can be 15 m, and bores occur only when the tide is
sufficiently high, at equinoctial spring tides. They were observed by Sir George Airy, the Astronomer
Royal, who initiated their theoretical explanation. The largest bore in the world appears to be on the
River Qiangtang, which flows past Hangzhou into Hangzhou Wan (Bay). It is up to 9 m high and
travels at 25 mph. In this stretch the river is called the Fuchun. The bore on the Amazon is called
the pororoca, and can be 4 m high. Several English rivers have bores, known by various names, but
the only other prominent one is the eagre on the Trent. There are significant bores on the Brahmaputra
and Indus, which flow into the Indian Ocean.
When a bore passes the observer and moves upstream, the direction of the current in the river
reverses, and for a while the river flows backwards. At some later time, the flow again reverses and
resumes its normal direction.
The famous bore called le mascaret on the Seine has been destroyed by dredging and deepening the
river, and the bore on the Petitcodiac off the Bay of Fundy has been similarly mostly eliminated by
construction of a causeway. Bores are very sensitive to artificial modifications of the channel. They
require rather special conditions to exist, and are not by any means common.

Channel Shapes
So far we have worked only with rectangular channels to avoid unnecessary complication. Although
artificial channels may often be rectangular, other shapes are used, mainly for convenience in
construction. The wooden flume is indeed usually rectangular, but circular pipes are easy to obtain in
metal and are often used, especially for water supply, sewers, drains and culverts. A canal in earth is
usually trapezoidal, to avoid caving of the banks, and concrete channels often imitate this shape.
Natural channels we must take as they present themselves.
The most efficient channel will have the minimum wetted perimeter for a given area, or the greatest
hydraulic radius for a given area. A rectangular channel of width b and height a has A = ab and P = 2a
+ b, or R = ab/(2a + b). If A is kept constant, then a db + b da = 0. To find the condition for a

minimum P, set dP = 2da + db = (-2a/b)db + db = 0, whence b = 2a, and R = a/2. A semicircular


channel of radius r has A = r2/2 and P = r, or R = r/2. To be equivalent to the most efficient
rectangular channel, 2a2 = r2/2, or r = 2a/. Hence R = a/ = a/1.77, which is greater than the R for
the most efficient rectangular channel. The semicircular channel is the most efficient of all. For a
circular pipe running full (to which our analysis applies), R = r/2, but now r = (2/)a, so R = a/(2)
= a/2.51. A circular pipe is not more efficient than the most efficient rectangular channel. Remember
that there is no friction on a free surface, so a free surface on the boundary of the flow area increases
the efficiency.
A simple channel shape that could be used as an example in our analysis is the 90 triangular flume. It
is not often found in practice, but is not a bad cross-section at all. If y is the depth from the bottom, A
= y2 and P = 22 y, so that R = y/22 = y/2.83. To be equivalent to the most efficient rectangular
channel, y = 2 a, or R = a/2. The triangular flume is just as efficient. The specific energy for a
triangular flume is E = y + Q2/2gy4. Expressing Q as a function of E and y and setting dQ/dy = 0, we
find that yc = (4/5)E. The critical depth is then yc = (2Q2/g)1/5. The dimensionless form of E is E/yc = x
+ 1/4x4. The derivation of the hydraulic jump curve is left to the reader. The centroidal distance of a
triangular area is 1/3 of its altitude.
The circular channel running part-full is very commonly met with. The geometry of the channel is
shown at the right. The flow occupies a segment of the circle of total angle 2, related to the depth y
as shown. There are tables of the properties of a segment, though the calculations are easy enough on
a calculator. In the formulas, is expressed in radians, but the modifications for using degrees are
obvious. Unfortunately, the properties are not suitable for easy algebra, so analytic results are
difficult. Solving for normal flow depth in terms of Q and S is not simple. One method would be to
tabulate Q as a function of y for the given S, and then find the value of y that corresponds to the given
Q. For smooth metal pipes, n = 0.011 to 0.015. For corrugated pipes, the value of n is about twice this.
Galvanized corrugated pipes are often used as culverts.
The trapezoidal section is used for canals, especially those
excavated in earth, to avoid caving of the banks. The
cross-section can be specified by top width t, bottom width
b and depth y. The area is A = (t + b)y/2, and the wetted
perimeter is P = b + 2[y2 + (t - b)2/4]. By the same
method as for the rectangular channel, the most efficient
trapezoidal channel is a half-hexagon, with sides of 60
slope. If b is the bottom width, then the top width is 2b,
the area A = 3by/2, and the wetted perimeter P = 3b, so the hydraulic radius is y/2. Of course, y = b
cos &30 = b3/2. This section is as bad algebraically as the circular one, since this simple expression
for R does not hold when the cross-section is not full.
A canal in earth loses water very readily. This was discovered by many American canal builders in the
early days, who thought that making a canal was just a matter of digging a ditch. Their canals rapidly
ran out of water, especially when crossing ridge lines. The secret to canals was puddling, or coating
the wetted perimeter with an impervious layer of clay tempered with sand that prevented water loss.
The sand is to prevent cracking, and good puddle is almost impervious. This, however, requires
expensive labor, as does lining with concrete, wood or brick, so most irrigation canals to this day are
not puddled or lined, relying on a copious source of water instead. Earth dams, however, often contain
puddle in their core wall, that prevents leakage of water through the dam.

Weirs
A weir (pronounced "we're") is a dam placed across a stream to raise or regulate the water level
upstream. The discharge of the stream flows through an opening in the weir, whose lowest level is
the sill height. A depth of at least this height is maintained, which becomes a little greater, depending

on the actual discharge. The water levels in the River Thames, from Richmond to Oxford, are
maintained by numerous weirs, with locks in them to pass river traffic. The Ohio River is also
maintained as a slack-water navigation by about 50 dams along its length, which are really weirs. In
hydraulic engineering, weirs are used for measuring discharge, for which they are very convenient.
Hydraulic weirs are usually thin and sharp-crested, made from sheet metal, because this shape is more
amenable to theoretical analysis and reproducibility. Weirs for other purposes usually have broad
crests. When we say "weir," we may be referring to the opening for the discharge (as here), or to the
whole dam. The word "weir" is from Anglo-Saxon werian, to defend, cognate with the German wehr.
We will not give an exhaustive account of hydraulic weirs here, but only explain the principles.
A convenient form of weir for small discharges is the 90 V-notch weir, shown at the left (the dropdown curve is exaggerated). We will assume that the velocity in the approach is negligible, so that the
water level is also the energy level, and that the discharge is free, not partially or wholly submerged.
We also assume that at the weir, the water depth will be the critical depth y c. We have already shown
that for a triangular channel, Q = (g/2) yc 5/2. The velocity at the weir is V = Q/yc2, so the velocity
head there is yc/4. Since the specific energy is equal to H (this is the usual symbol for the upstream
water depth above the sill), yc = 4H/5. Expressing the discharge in terms of H, we have Q = (g/2)
(4H/5)5/2 = 1.267 H2.5 cms. Using feet instead, we have Q = 2.30 H2.5 cfs. To find Q, all we have to do
is read H from the gauge at the upper
end of the drop-down curve and perform
a simple calculation.
The two effects we have neglected are
the influence of upstream velocity and
the contraction at the crest, as in the case
of orifices. Upstream velocity can be
taken into account by increasing H by
the velocity head. We won't make this
correction, although it is usually made
when precise results are necessary, and
will be found in handbooks. Accurate measurments gave Q = 2.52 H 2.47 cfs and 2.48 H2.48 cfs, the
former for commercial sheet metal, the latter for a polished brass plate. The heads varied from 0.2 to
1.8 ft (Q = .05 to 11 cfs, or 0.4 to 82 gal/s). It is clear that our analysis is close to the truth, differing
by less than 10% from experimental results. The notch weir is very suitable for laboratory
experiments or in a scientific garden.
If we apply the same analysis to a wide rectangular weir, we find that H = 3yc/2, and the discharge per
unit width is q = g(2/3)1.5 H1.5 = 1.70 H1.5. In feet, q = 3.08 H1.5. Francis gave the formula q = 3.33
H1.5, so again our analysis is not far wrong, in spite of its simplicity.

References
R. L. Daugherty and J. B. Franzini Fluid Mechanics, 6th ed. (New York: McGraw-Hill, 1965).
Chapter 10, pp. 276-326.
J. C. Trautwine, The Civil Engineer's Pocket-Book (Philadelphia: Wm. Butter & Co., 1888). pp. 236281. Kutter's formula, together with much practical information.
L. C. Urquhart, ed., Civil Engineering Handbook, 4th ed. (New York: McGraw-Hill, 1959). Section 4.
H. Lamb, Hydrodynamics, 6th ed. (New York: Dover, 1945).

S-ar putea să vă placă și