You are on page 1of 33

CHAPTER 8

HYDRAULIC DESIGN FOR

ENERGY GENERATION

H. Wayne Coleman
C. Y. Wei
James E. Lindell
Harza Company
Chicago, Illinois

8.1 INTRODUCTION
This chapter describes the design aspects of hydraulic structures related to the pro-
duction of hydroelectric power. These structures include headrace channels; intakes;
conveyance tunnels; surge tanks; penstocks; penstock manifolds; draft-tube exits; tail-
tunnels, including tail-tunnel surge tanks and outlets; and tailrace channels. The pro-
cedures provided in this chapter are most suitable for developing the preliminary
designs of hydraulic structures related to the development of the hydroelectric pro-
jects. To finalize designs, detailed studies must be conducted: for example, economic
analysis for the determination of penstock diameters, computer modeling of hydraulic
transients for surge tank design, and studies of physical models of intake and its
approach.

8.2 HEADRACECHANNEL
An open-channel called the headrace channel or power channel (canal) is sometimes
required to connect a reservoir with a power intake when the geology or topography is not
suitable for a tunnel or when an open-channel is more economical. The channel can be
lined or unlined, depending on the suitability of the foundation material and the projects
economics. Friction factors for various linings used for design are as follows:
Manning's n
Lining Minimum. Maximum
Unlined rock 0.030 0.035
Shotcrete 0.025 0.030
Formed concrete 0.012 0.016
Grassed earth 0.030 0.100
Headrace channels are generally designed and sized for a velocity of about 2 m/s
(6.6 ft/s) at design flow conditions. Economic considerations may result in some variation
from this velocity, depending on actual project conditions.
Channel sections are normally trapezoidal because this shape is easier to build for
many different geologic conditions. The bottom width should be at least 2 m (6.6 ft) wide.
Side slopes are determined according to geologic stability as follows: earth, 2H: 1V or flat-
ter; and rock, IH: IV or steeper. The channel's proportions—bottom width versus depth—
are largely a matter of construction efficiency. In general, the minimum bottom width
reduces excavation, but geologic conditions may require a wider, shallower channel. The
channel slope will result from the conveyance required to produce design velocity for
design flow.
Channel bends should have a center-line radius of 3W to 5W or more, where W is the
water surface width of the design flow. For this radius, head loss and the rise in the water
surface at the outer bank (superelevation) will be minimal. If the radius must be reduced,
the following formula can be used to estimate head loss hL:

hL = K6^ (8.1)

where Kb = 2 (W/RC), W = channel width, Rc = center-line radius, and V = mean velocity.


Superelevation will be as follows (Chow, 1959):
2W V2
AZ =T¥ (8.2)

where AZ = rise in water surface above mean flow depth.


Freeboard must include allowances for the following conditions: (1) static condi-
tions with maximum reservoir level (unless closure gates are provided to isolate the
channel from the reservoir), (2) water surface rise (superelevation) caused by flow
around a curve, and (3) surge resulting from shut-off of flow downstream or sudden
increase of flow upstream. A forebay is provided at the downstream end of the head-
race channel to facilitate one or more of the following: (1) low approach velocity to
intake, (2) surge reduction, (3) sediment removal (desanding), or (4) storage. The fore-
bay should be designed to maintain the approach flow conditions to the intake as
smoothly as possible. As the minimum requirement, a small forebay should be provid-
ed to facilitate good entrance conditions to the intake. It should include a smooth tran-
sition to a section with a velocity not exceeding 0.5 m/s (1.64 ft/s) at the face of the
intake structure
A larger forebay could be required for upsurge protection during rapid closure of tur-
bine gates for load rejection. The size would be determined on the basis of the freeboard
allowance for the entire headrace channel and on a hydraulic transient analysis of the
channel, if necessary.
Surge calculations should consider maximum and minimum friction factors, depend-
ing on which is more critical for the case under study. Hydraulic transient (surge) studies
are generally performed using a one-dimensional, unsteady open-channel-flow simulation
program. The computer model developed should be capable of simulating the operation
of various hydraulic structures, the effect of the forebay, and operation of the power plant.
Several advanced open-channel flow-simulation programs have been described by Brater
et al. (1996).
Exhibit 8.1 Sun Koshi hydroelectric project, Nepal.
(a) A view of the desanding basin (looking upstream) showing concrete
guide vanes.(
(b) Layout Of the desanding basin.

A large forebay is required if it will be used for diurnal storage-say, for a power peak-
ing operation. In such a case, maximum and minimum operating levels would include the
required water volume, with the intake located below the minimum level. Such a forebay
also could accommodate the other three functions described above.
When the flow carries too much sediment and its removal is required to protect the tur-
bines, a still larger forebay would be provided to function as a desanding basin (also
known as a desilting basin or desander). However, the desanding basin is more likely to
be located at the upstream end of the headrace channel. Exhibit 8.1 Illustrates a desend-
ing basin. The basin can be sized using the following equation (Vanoni, 1977):
Fall velocity i-f (m s ')

Particle diameter d (mm)


FIGURE 8.1 Settling velocity as a function of particle diameter.
(Dingman, 1984)

_ LVs
P = (I- eVD ) X 100% (8.3)
where P = percentage of sediment of a particular size to be retained by the basin, L =
basin length, Vs = settling (fall) velocity of suspended particles, V = mean flow velocity,
and D = depth of the desanding basin. The settling velocity Vs for each particular sand
particle size can be estimated from Fig. 8.1. A separate sluicing outlet (or outlets) would
be provided to flush the desanding basin intermittently.

8.3 INTAKES
Most power intakes are horizontal, a few are vertical, and very few are inclined. Figures
8.2, 8.3, and 8.4 are examples of the three types of intakes. The horizontal intake is usu-
ally connected to a tunnel or penstock on a relatively small slope (up to 2-3 percent). The
vertical intake is frequently used in pumped-storage projects when the upper reservoir is
on high ground, such as a mountain top, and a vertical shaft-tunnel is the obvious choice.
An inclined intake is used when the topography, geology, or type of dam dictate a steep-
er slope for the downstream tunnel or penstock. Exhibits 8.2 and 8.3 are examples of
intakes on an arch dam and on a pump-storage upper reservoir.
El 54 O Intake
Max. operating
T.W El 50 O-
Min. operating
TW El. 43 5 - Units (150 o.c )
Flow
GENERA TOR ISO T bridge crane
El 3325 HALL
-Draft tube gantry crone
Service Bay
Trashrock guide EI.S75
Penstock 5.O ID.
Emergency gate slot Max. T.W. El 6.5
Service gate slot Normal T.W. £1.61
Min. T.W
EI.48
£ distributor
El. €.5
Note: Dimensions in meters.
Draft tubs gate
FIGURE 8.2 A typical horizontal intake. (Harza Engineering Co.)
£7 1217.25

e/ 73727

€/. 64728

FIGURE 8.3 A typical vertical intake. (Harza Engineering Co.)


Dasigrt flood Axis of dam
surcharge
El. 781 £ El.785.0
Max. power pool
El. 770.0
Normal min. El. 700.0
Extreme min. pool Service gate blackout
(Drawdown limit)
El. 6OO O
Appro*. 22O ft. (Unit I) Units (65 -O" o. C )
Appro*. 190 ft. (Unit 2)
Trashrack blocKout 450 TPH. crone
Porcelain Al. insulated wall panel
penstock Glass block panel
E1.512 ± Porctlain Al. insulated wall panel
Stair house Max Operating
El 4750 T.W. El. 475.0

Normal Operating (2 Units)


T.W. El 430 O
MIn Operating
T.W. El. 427.0
distributor El. 417O

Draft tube gats

FIGURE 8.4 A typical inclined intake. (Harza Engineering Co.)


Spillway

Powerhouse

Main dam

Exhibit 8.2 (a)


Exhibit 8.2 Karun hydroelectric project, Iran
(a) A view of the dam (looking downstream) showing spillway
crest, radial gates, power intakes, and diversion tunnel entrace
structure.
(b) Layout of dam showing spillway, intake and powerhouse.

A variation on the three basic intake types is a tower structure, sometimes required for
selective withdrawal of water. The tower includes openings with tmshmcks and bulk-
heads at various levels, which permit water to be withdrawn from different depths to con-
trol temperature or water quality. Computer modeling of a reservoir's temperature and
water-quality structure is generally required to finalize the required opening sites.
Descriptions of several reservoir-simulation models can be found in Brater et al. (1996).
Figure 8.5 is an example of a multilevel intake tower structure for selective withdraw-
al.Exhibit 8.3 illustrates the intake structure for a pumped storage project.
Trashracks for power intakes are designed for a velocity of about 1 m/s (3.3 ft/s) when
the intake is accessible for cleaning. If a trashrack is not accessible for cleaning, the allow-
able velocity is approximately 0.5 m/s (1.6 ft/s). Trashrack bar spacing is dictated by tur-
bine protection requirements, but clear spacing of 5cm (2 in) is typical. Although head
loss through trashracks depends heavily on the amount of clogging, the following can be
used for a clean trashrack, (U.S. Bureau of Reclamation, 1987);
EL.2057 MIGMMT LIFT
33 TON CRANE FOLLOWER
TRAiHWACK INTAKE HOUSE

NORMAL MAX
OPERATING WL.EL.20OO POST'TCNSIONCO
ANCHOM(TYR)
HEATED ICE BOOM

3'-00IA. HOLE (TYR)

SHUT ten

AMCHOMS (TYP.)
TRASHRACK(TYP)

FLOW HYDRAULIC HOIST

MINIMUM OPERATING
W.L.EL.I8SO
•INTAKE OATC
12-0 W «24-0H

ENDOFPIER
BULKHEAD GATE
I4'-OW» 32'-OH
FIGURE 8.5 A typical multi-level intake tower structure for selective with-
drawal. (Harza Engineering Co.)

hL = K,^ (8.4)

A (A Y
where Vn = velocity based on the net area, K = 1.45 — 0.45 -r- A
— hp L An = net area of
s {AgJ
trashrack and support structure, and Ag = gross area of trashrack and support structure.
An intake gate is generally provided when the power tunnel or penstock is long or
when a short penstock does not have a turbine inlet valve. This gate is provided for emer-
gency closure against flow in case of runaway conditions at the turbine or penstock rup-
ture. The effective area of the gate is usually about the same as that of the power tunnel
or penstock, but it is rectangular in shape, with a height that is the same as the conduit's
diameter and a width that is 0.8 X the conduit diameter. A bulkhead (or stop-logs) is pro-
vided upstream of the intake gate for servicing the gate. The trashrack slot might be used
for this function by first pulling the trashrack.
Exhibit 8.3 Rocky mountain pumped storage project, Georgia.
(a) Intake structure of the upper reservoir.
(b) Close-up view of the upper reservoir intake structure.
(c) Profile of the project including upper reservoir intake, power tunnel, and power
house.
Ar , ,at, Md access 6uiMn»
€ H 737.27
0,rt,ub.g«,

Exhibit 8.3 (c)


A hydraulic study is generally conducted for emergency closure of the intake gate.
The maximum turbine flow or runaway flow should be considered. The runaway flow
may be 50 percent higher than the normal turbine flow for a propeller turbine. In the
hydraulic study, the water levels and pressures, as well as flow into and from the gate
well, as a function of gate position are investigated (Fig. 8.6). With this information, crit-
ical gate loads can be determined for the gate and hoist. The gate also may be used for
penstock filling. A minimum gate opening of 10 to 15 cm (4-6 in) is usually specified
for this, but a special hydraulic study must be made to determine potential gate load and
vibration if the gate opens continuously by accident. In such cases, a generous gate well
or air vent must be provided downstream of the gate to provide relief once the tunnel or
penstock fills.
The head loss for a bulkhead or gate slot, including top opening, is generally about 0.1
of the local velocity head at the slot. The transition length (m or ft) Lt from gate section
to tunnel or penstock should be approximately:

WIRE ROPES

FACE OF
ROLLER TRACK

TRASHRACK SEAL
TOP SEAL EL.545

PENSTOCK

EL.5I2

FEET

FIGURE 8.6 A typical intake gate arrangement. (Harza Engineering Co.)


A =^ (8-5)

where V = tunnel/penstock velocity (m/s or ft/s), D = tunnel/penstock diameter


(m or ft), and C = 3.00 for units in metric systems or = 9.84 for units in English sys-
tems. The variation of velocity in the transition section should be as close to linear as
practicable.
Overall head loss for an intake includes trashrack, bellmouth (0.1 X V1IIg), gate slots,
and transition. The potential vortex formation for an intake should be checked using Fig.
8.7. Note that when the intake Froude number (V/VgD) exceeds 0.5, submergence
requirements increase dramatically, and the vortex formation is difficult to predict. In this
case, a physical model study should be carried out.

8.4 TUNNELS
When the powerhouse is situated a considerable distance from the intake and when geo-
logic conditions permit, a tunnel is often used to convey the flow for power generation.
The size of the tunnel is dictated by economics: that is, construction cost is added to the
cost of head loss (loss of generating revenue) to obtain the minimum combined cost. This
determination is usually obtained by trial and error because the process does not lend itself

Horizontal intake
Vertical intake
Intakes with
vortex problems
Intakes with
no problems

FIGURE 8.7 Intake submergence and vortex formation.


(Gulliver and Arndt, 1991)
to a simple formula. The resulting tunnel velocity with the economic diameter is usually
in the range of 3 to 5 m/s (10 to 17 ft/s).
The shape of the excavated tunnel normally will approximate a square bottom and a cir-
cular top. The diameter of the circular top (or the width of the square bottom) should be larg-
er than the required diameter. If the tunnel is lined with concrete, its cross section is likely to
be circular or have a square or trapezoidal bottom. If it is unlined or lined with shotcrete, the
excavated shape will remain, with some smoothing by filling the larger overbreak sections.
Lining is an economic consideration, balancing the cost of the lining with the power
loss caused by friction. Even an unlined tunnel will have lined sections, such as portals,
and sections where rock needs extra support for geologic stability. Friction factors for
design are as follows:
Manning's n
Lining Minimum Maximum
Unlined 0.030 0.035
Shotcrete 0.025 0.030
Formed concrete 0.012 0.016
Minimum friction corresponds to new conditions and is used for turbine-rating and
pressure-rise calculations. Maximum friction corresponds to aging and is used for eco-
nomic-diameter and pressure-drop calculations. Tunnel slope is dictated by construction
suitability and geology, with a minimum of 1:1000 for drainage during dewatered condi-

Exhibit 8.4 (a) Bath County pumped storage project, Virginia.


(a) Surge tank openings during construction (44-ft inside diamenter and 300-ft deep)
imping

Bl*2J^o"PO°'

Sta'sssSfcr'oiBaisM

200 7. mm.

Exhibit 8.4 Bath County pumped storage project, Virginia. (Continued)


(b) Profile of the project including upper reservor intake, gate structure, surge tanks, power tunnel and powerhouse
tions. Tunnel horizontal bends generally have large radii for convenience of construction.
Vertical bends at shafts usually have a minimum radius of 3D to minimize head loss and
to provide constructibility.

8.5 SURGETANKS
Surge tanks generally are used near the downstream end of tunnels or penstocks to reduce
changes in pressure caused by hydraulic transients (waterhammer) resulting from load
changes on the turbines (ASCE, 1989; Chaudhry, 1987; Gulliver and Arndt, 1991; Moffat
et al., 1990; Parmakian, 1955; Rich, 1951; Wylie and Streeter, 1993; Zipparro and Hasen,
1993). A surge tank should be provided if the maximum rise in speed caused by maximum
load rejection cannot be reduced to less than 60 percent of the rated speed by other prac-
tical methods, such as increasing the generator's inertia or the penstock's diameter or by
decreasing the effective closing time of the wicket gates. In general, the provision of a
surge tank should be investigated if

I>,-
—jj— > 3 to 5 for units in m/s and m or
n
> 10 to 20 for units in ft/s and ft, (8.6)
where L1. is the length of a penstock segment and V1 is the velocity for the segment
(Dingman, 1984). The term ^L1V1 is computed from the intake to the turbine and Hn is the
minimum net head.
Surge tanks normally are located as close as possible to the powerhouse for maximum
effectiveness and may be free-standing or excavated in rock. The tanks are usually vent-
ed to atmosphere or can be pressurized as air chambers. The latter is not used frequently
because of requirements of size, air compressors, and air tightness. Exhibit 8.4 illustrates
a pumped storage project with a surge tank. Figure 8.8 shows typical installations of surge
tanks for controlling hydraulic transients.
Surge tanks usually are simple cylindrical vertical shafts or towers, but other geomet-
ric designs are used when the surge amplitude is to be limited. For instance, an enlarged
chamber can be used at the top if upsurge might cause the water level to rise above the
ground surface. Similarly, an enlargement or lateral tunnel or chamber is sometimes used
near the bottom of the shaft if downsurge would caused the water level to drop below the
tunnel crown. When the geometry is a cylinder, analysis is relatively simple and can be
performed using design charts. If the geometry is more complicated, a hydraulic transient
simulation model is required to carry out the study (Chaudhry, 1987; Wylie and Streeter,
1993; Brater et al., 1996).
Hydraulic stability for a surge tank assures that surging is limited and brief after load
changes (Rich, 1951; Parmakian, 1955; U.S. Bureau of Reclamation 1980; Zipparro and
Hasen, 1993). The minimum cross-sectional area of a simple cylindrical surge tank
required for stability can be determined using the Thoma formula:
AL
A (8 7)
^=2^H '

where AST = minimum tank area, A = tunnel area between reservoir and surge tank, L =
tunnel length between reservoir and surge tank, g = gravitational acceleration, c =
2100 »»WBATH COUNTY FOWCRPUUtT
funtf^Ster»99 f»«i«tap»inf t1985)
MNOMU on Buck Cw»*

FIGURE 8.8(a) Typical vented surge tank installation. Bath County powerplant (1985): 2100 MW pumped storage development on Back Creek, Virginia.
lnttkt control building
Cl. 1035.0
HW £1. 10250
r>«»n/wc*
FlOKf
InIfHt oil* Hot
EI tee o

30'-0' cti» tit chtmt»r

0. »940
ElBtSOQ T.W. El. »900
£1 674.1
O. 159.53
4.600' O-
MOOSE RIVER POWERPLANT (1987)
12 MW Development on Moose River
NEW YORK
FIGURE 8.8(b) Typical pressurized surge tank installation. Moose River powerplant (1987): 12 MW
development on Moose River, New York. (Harza Engineering Co.)

A# 1 ( Mi \
head loss coefficient = -F^- = y- T/2/^~~ , where A// = minimum head loss from
reservoir to surge tank, including tunnel velocity head W2g, and H - minimum net oper-
ating head on turbine.
For a simple surge tank (without an orifice), increase the diameter obtained from the
Thoma formula by 50 percent. For a typical surge tank with a restricted orifice, increase
the diameter by 25 percent. These increases are necessary to provide damping of the oscil-
lation in a reasonable period of time.
Maximum upsurge in a cylindrical surge tank can be determined from Fig. 8.9. For a
given tank size, the optimum size of the orifice is based on the balanced head design so
that the maximum tunnel pressure below the surge tank equals the maximum upsurge
level.
Maximum downsurge in a cylindrical surge tank can be determined from Fig. 8.10.
Here again, the size of the orifice should be based on balanced head design as a first
attempt. However, since downsurge may differ from upsurge, and the required orifice size
may be different for the two purposes, shaping the orifice (i.e., changing the discharge
coefficient) by rounding the top or bottom may satisfy the two area requirements approx-
imately.
For maximum upsurge, use the maximum normal headwater, minimum head loss
between reservoir and surge tank, and maximum plant flow. Assume full plant load-rejec-
tion (tripout) in the shortest reasonable time.
For maximum downsurge, use the minimum normal headwater, maximum head loss,
and accept load from 50 percent to 100 percent in the shortest reasonable time. At some
projects, such as pumped-storage plants, the load acceptance is criterion is more extreme;
full load acceptance, is O percent to 100 percent in the shortest reasonable time. The con-
trolling criterion will be used to design the orifice on downsurge. When the surge tank
geometry is complex (noncylindrical), a computer model should be used to determine the
limiting surge levels. (See Brater et al., 1996, for available computer models). Freeboard
F » Cross-sectionol oreo of surge tonk...(sq.ft.)
L -Length of pipeline between reservoir ond surge tonk...(ft.)
ft • Cross-sectionoi oreo of pipe.._(sq ft)
SAg »- Moi
Accelerotion
imum surgeo<from
grovity...(ft
operotmgperlevel$ecdueperIosec.)
instontoneous stopping of flow Oo—(ft)
For turbine operotion__.5«« Upsurge (ft.) For pump operotion.._S»» Oownsurge (ft.)
(Noreturn flow permitted through the pump)
Hf • Pipe f net i on loss •+velocity heod •«• ony other pipe losses between surge tonk ond
reservoir ossocioted with OQ--(M.)
H|? « Throttling loss for flow into 01 out of surge tonk ossocioted with Q0. .(ft.)
SURGE RATIO

BALANCED DESIGN

MAXIMUM SURGE IN SURGE TANK DUE TO


INSTANTANEOUS STOPPING OFFLOW Qo
FIGURE 8.9 Maximum surge in surge tank due to instantaneous stopping of flow.
(Parmakian, 1955)

F - Cross- iectionol oreo ot surge tonk...(sq.ft.)


A » Cross-sectionol oreo of pipe...(sq.ft.)
Sgg "• Moiimum
Accelerotionsurgeoffrom
grovity...(ft. perdue
stotic level sec.topermstonfoneous
sec.) storting of Oe (ft.)
For turbine operotion...Sg • Oownsurge (ft) For pump operotion...S»- Upsurge (ft)
OMf 8 -• Pipe
Flow demonded by the turbine or dischorged by the pump...l
friction loss + velocity heod + ony other pipe losses between eu. ft. persurge
sec.)tonk ond
1
reservoir lojs
Hf. • Throttling ossocioted
for flowwith
intoQore—.(ft.)
out of surge tonk ossocioted with 0( (ft.)
SURGE RATIO

BALANCEODESIGN

MAXIMUM SURGE IN SURGE TANK DUE TO


INSTANTANEOUS STARTING OFFLOW Ot
FIGURE 8.10 Maximum surge in a surge tank resulting from instantaneous starting
of flow. (Parmakian, 1955)
for the surge tank is 10 percent of the computed rise in the water level in the surge tank
for upsurge and 15 percent of the drop in the water level for downsurge to maintain, sub-
mergence of the tank invert or the orifice to avoid admitting air into the penstock.
Pressurized air chambers are often used in pumping plants for surge protection. They
are used occasionally for power plants when the generating flow is not excessive. The
hydraulic characteristics of the chambers are complicated by the compressibility effects
of air and temperature, and the analysis does not lend itself to simple formulas and charts.
A computer model is required to verify performance. Fig. 8.8(B) shows a typical air cham-
ber design for a hydropower plant.

8.6 PENSTOCKS
A penstock generally refers to a steel conduit or steel-lined tunnel connecting a reservoir
or surge tank to a powerhouse (ASCE, 1989, 1993; U.S. Bureau of Reclamation, 1967;
Chaudhry, 1987; Gulliver, and Arndt, 1991; Warnick et al., 1984; Wylie and Streeter,
1993; Zipparro and Hasen, 1993). It is used when the internal pressure is high enough to
make a concrete-lined tunnel or unlined rock tunnel uneconomical, particularly where
rock cover is low.
Penstock size is usually governed by project economics. The economical diameter is
determined by the minimum combined cost of construction and energy reduction caused
by head loss in the penstock. The energy loss decreases as the diameter of the penstock
increases while construction cost increase. As with tunnels, the most economical diame-
ter can be determined more accurately by a trial-and-error procedure. The following vari-
ables are generally considered (U.S. Bureau Reclamation, 1967; Gulliver and Arndt,
1991):
1. Cost of pipe 7. Surface roughness (friction factor)
2. Value of energy loss 8. Weight of steel penstock
3. Plant efficiency 9. Design discharge
4. Minor loss factor 10.Allowable hoop stress
5. Average head
6. Waterhammer effect

For the assessment of a preliminary design or a feasibility level, the most economical
diameter can be estimated using the following formula (Moffat et al., 1990).
CP0-43
D. = -JpW (8.8)

where De = the most economical penstock diameter (m or ft), H = the rated head (m or
ft), P = the rated capacity of the plant (kW or hp), and C = 0.52 (for metric units) or =
3.07 (for English units). If the project is a small hydropower installation, the following
simple equation can be used (Warnick et al., 1984).
D6 = CQ0* (8.9)
where De = the most economical penstock diameter (m or ft), Q = the design discharge
(m3/s or ftVsec), and C = 0.72 (for metric units) or = 0.40 (for English units).
For large hydroelectric projects with heads varying from approximately 60 m (190 ft)
to 315 m (1,025 ft) and power capacities ranging from 154 MW to 730 MW, the follow-
ing equation can be used (Warnick et al., 1984).

D
O?0-43
. = -%r (8-10)

where D6 = the most economical penstock diameter (m or ft), p = the rated turbine capac-
ity (kW or hp), h = the rated net head (m or ft), and C = 0.72 (for metric units) or = 4.44
(for English units).
The maximum velocity in the penstock is normally kept lower than 10 m/s (33 ft/s).
To determine the minimum thickness of the penstock, based on the need for stiffness, cor-
rosion protection, and handling requirements, the following formula can be used (U.S.
Bureau of Reclamation, 1967; Warnick et al., 1984).
D+K
(8 U)
*min = 400 '

where tmin = the minimum thickness of the penstock (mm or in), D = penstock diameter
(mm or in), and K = 500 (for metric units) or = 20 (for English units). After determining
the economic diameter, check for the operating stability of the generating unit-penstock
combination using the following steps (Chaudhry, 1987; U.S. Bureau of Reclamation,
1980; Warnick et al., 1984).
1. Determine the mechanical starting time in seconds for the unit T1n as
JGD^V-_
lm
36XlO4P ^'^

or
(WR2W
T (8 13)
™ ~ 1.6 X 106P7 '
2
where GD = flywheel effect of the turbine and generator rotating parts used in
metric system (kg-m2), WR2 = flywheel effect of the turbine and generator rotating
parts in English system (lb-ft2) = 5.932 GD2, G = weight of rotating parts (kg), D
= 2 X radius of gyration of the rotating parts (m), W = weight of rotating parts
(Ib), R = radius of gyration of the rotating parts (ft), N = turbine speed (rpm), P =
maximum turbine output (kW), and P1 = maximum turbine output (hp).
Tm is the time for torque to accelerate the rotating mass from zero to rotational
speed. Together, the turbine runner in water, connecting shafts, and the generator
develop the flywheel effect WR2 or GD2. The WR2 can be determined using on the
following formulas:

WRL^ = 23,800 (^) (8.14)

and
KVA\5/4
^ -JJST) (8.15)
'
where Pd = turbine rated output (hp) and kVA = generator rated output (kilovolt-
amperes).
2. Determine the water column starting time for the penstock Tw as follows:
Z(LV)
T
» = ~W~ <8-16)
where ^LV) = summation of product of length (measured from nearest open water
surface) and velocity for each segment of penstock from intake or surge tank to tail-
race (m2/s or ftVsec), g = gravitational acceleration (m/s2 or ft/sec2), and H = min-
imum net operating head (m or ft).
3. In general, TJTj should be maintained greater than 2 for good operating stability
and to have reasonably good responses to load changes. If Tm/Tw2 is less than 2,
there are three possible solutions:
• Increase WR2 or GD2 for the generator; this is relatively inexpensive for
increases of up to 50%.
• Increase the penstock diameter; this is probably not economical, except for a
narrow range.
• Add a surge tank or move the surge tank closer to the powerhouse.
A combination of these three possible solutions may be the most cost-effective solu-
tion. The following friction factors are recommended for designing steel penstocks:

Exhibit 8.5 A typical steel penstock branch structure


being fabricated
Penstock Age Manning's n
New 0.012
Old 0.016
Use the value for new penstock to calculating turbine-rating and pressure-rise. To cal-
culate pressure drop use the higher values.
Design pressure is determined on the basis of the turbine's characteristics and the clo-
sure rates of the wicket gates or needle valves. For Pelton turbines, closure rates are slow,
and design pressure rise is usually of the order of 20 percent of the static pressure head.
For Francis turbines, design pressure rise is usually 30 to 40 percent of the static pres-
sure head, depending on the cost of steel lining required. A fast closure is desirable to
minimize speed rise and the potential for runaway conditions in the turbine. Detailed
pressure conditions are determined by a computer model that includes the water con-
ductors and surge tank as well as the turbine discharge-speed characteristics and gener-
ator inertia. Many computer programs capable of simulating hydraulic transients are
described in Wylie and Streeter, 1993. Such computer simulation studies are often
required of turbine or governor manufacturers now as a part of the specifications.
Ultimately, the predicted pressure conditions are verified in the load rejection tests dur-
ing unit start-up.
The profile for a free-standing penstock is based on the topographic and geologic con-
ditions of the ground. In other cases, the penstock may consist of shaft and tunnel sections
that are largely lined with concrete, with a relatively short section of steel-lined penstock
near the powerhouse.
If the penstock is free-standing, the risk of penstock rupture is greater than it is for the
shaft and tunnel system. If there is a long tunnel section upstream of the free-standing pen-
stock, an emergency closure valve is often added near the tunnel outlet. A hydraulic tran-
sient study is necessary to determine closure conditions (by accident or because of pen-
stock rupture). A vent must be provided to admit air just downstream of the valve for pen-
stock rupture and must be large enough to prevent collapse of the penstock from internal
subatmospheric pressure caused by water-column separation. A free-standing penstock
also requires small air inlet-outlet valves at local high points to remove air during filling
and admit air during dewatering.

8.6.1 Penstock Branches


A penstock often delivers water to more than one turbine. In such cases, the penstock is
branched in various ways to subdivide the flow.Exhibit 8.5 illustrates a typical steel pen-
stock branch structure. When the powerhouse is normal to the penstock, several configu-
rations are possible (Fig. 8.11). If the powerhouse is at an angle with the penstock, a man-
ifold is used (Fig. 8.12).
Head losses in branches and manifolds depend on precise geometry and often are
developed by model studies. However, for a typical well-designed layout, the following
head loss coefficients can be used to estimate the head loss hb from the main into a
branch:

hb = Kb^ (8.17)

where V = branch velocity (m/s or ft/s); g = gravitational acceleration (m/s2 or ft/sec2); and
Kb = head loss coefficient 0.2 for symmetrical bifurcation, 0.3 for symmetrical trifurca-
tion, and 0.2 for manifold branch.
MAIN PENSTOCK MAIN PENSTOCK
TRIFURCATION
SINGLE
BIFURCATION
BRANCH BRANCH
(TYPICAL)
(TYPICAL)

2 UNITS 3 UNITS

MAIN PENSTOCK MAIN PENSTOCK

BOUBLE
Y-BRANCHING

DOUBLE
•BRANCH BIFURCATION
(TYPICAL)
BRANCH
(TYPICAL)

3 UNITS 4 UNITS
FIGURE 8.11(a) Schematic penstock branch configurations for powerhouse normal to the penstock.

The diameters of branched penstocks are usually determined so that the velocity is
increased significantly relative to the main penstock. Here again, the branch size is deter-
mined by economics so that construction and material costs added to cost of energy loss
are at a minimum. The lower limit for the size of the branch is the size of the turbine inlet
that is normally provided by the turbine manufacturer. If a turbine inlet valve is provided,
its diameter will either be equal to the inlet diameter or be between the inlet diameter and
the penstock branch diameter. This valve is usually a spherical type, and, as such, no head
loss occurs in the fully open position. Friction losses in the branch penstocks are calcu-
lated using the same friction factors used for the main penstock and the conduit lengths
up to the net head taps in the turbine inlet.
SINGLE BIFURCATION
TRIFURCATION

DOUBLE Y-BRANCHING

FIGURE 8.11(b) Configurations for single bifurcated, double y-branching, and trifurcated penstocks.
(Harza Engineering Co.).

8.7 DRAFT-TUBEEXITS
Draft-tubes are designed by considering the turbine's characteristics. The net head for the
turbine is based on pressure taps at the spiral-case inlet and near the draft-tube exit.
Therefore, any head losses which occur after the draft-tube pressure taps are subtracted
from the turbine net head. Because the exit head loss is generally considered to be the
average velocity head at the end of the draft-tube, a longer draft-tube with expansion to a
larger area would, in theory, reduce this loss. In actuality, however the flow is not uniform
MAIN PENSTOCK
MAIN PENSTOCK

BRANCH BRANCH
(TYPICAL) (TYPICAL)

2 UNITS 3 UNITS

MAIN PENSTOCK

BRANCH
(TYPICAL)

4 UNITS
FIGURE 8.12(a) Schematic penstock manifold configurations for a powerhouse
oriented at an angle with the penstock.

FIGURE 8.12(b) Penstock manifold for an installation with six units. (Harza Engineering Co.)
at this point; it is highly turbulent and swirling, and the true exit loss is difficult to define.
Current thinking is that further extension of the draft-tube is not economical. The rule of
thumb is to end the draft-tube when the mean velocity is about 2m/s and to base the exit
head loss on this velocity.
A trashrack is usually provided at the end of the draft-tube at a pumped-storage pro-
ject to prevent entry of coarse debris during the pumping mode. However, during the gen-
erating mode with the trashrack in place, the trashrack is subject to vibration caused by
the concentration of flow and by swirling. This complicates the design of the trashrack
and increases its cost. The analysis of the rack is a combined hydraulic and structural one.
The hydraulic loadings consist of drag forces on rack bars that are dependent on velocity
patterns along with pulsation of pressure caused by swirling flow. The data on hydraulic
conditions can be obtained from a physical model (usually the model from the pump-tur-
bine manufacturer) because fully developed mathematical models are not readily avail-
able to predict these forces. A structural mathematical model is then applied using the
hydraulic loadings obtained from the hydraulic model tests. By trial and error, the
trashrack is designed to withstand the flow-induced vibrations.

8.8 TAIL-TUNNELS
An underground power plant will have a tail-tunnel to deliver the flow to the downstream
river or lake. For a pumped-storage project, this tunnel provides flow both ways, because
it acts as the inlet tunnel during pumping.
For a conventional hydroelectric plant with generating only, the tunnel is usually pres-
surized.. However, if the turbines are the Pelton type, the tunnel is likely to be free flow
to maintain freeboard on the turbine. For a pumped-storage plant, the tunnel is most like-
ly to be pressurized, because it must deliver water both ways. If the tunnel is pressurized
and is long enough, a surge chamber will be required to prevent large fluctuations of pres-
sure on the turbines during load changes.
The number of tail-tunnels, usually one or two, is based on economics and con-
structability. From an operational standpoint, two tunnels are desirable to allow partial
operation of the plant even during maintenance or inspection of one of the tunnels.
However, two tunnels are usually more expensive than one, and usually only one will be
used unless its size becomes unmanageable. The limiting size is dictated by available
equipment and tunneling methods. These factors must be evaluated carefully when esti-
mating the costs of one tunnel versus two tunnels.
A manifold is used to collect the flow from the individual draft-tubes and guide the
flow through a transition section to the tail-tunnel proper. This manifold is similar in con-
cept to the penstock manifold, but generally the velocities are much lower. The velocity
at the end of the draft-tube is typically 2 m/s (7 ft/s) and 3 m/s (10 ft/s) at the tail-tunnel.
Therefore, head losses are not significant and the flow conditions are generally accept-
able. A typical tail-tunnel manifold design is shown on Fig. 8.13.

8.8.1 Tail-Tunnel Surge Tanks


When an underground power plant has a significant length of pressurized tail-tunnel, a
surge tank is likely to be required. The procedures for sizing and determining extreme
surges are similar to the procedures used for surges in the head-tunnel, using the hydraulic
characteristics of the tail-tunnel instead of the head-tunnel. (Refer to Sec. 8.5). Figure 8.14
shows a typical tail-tunnel surge chamber.
GENERAL PLAN

FIGURE 8.13 A typical tail-tunnel manifold arrangement. (Harza Engineering Co.)

8.8.2 Tail-Tunnel Outlet Structures


The tail-tunnel outlet structure is typically a bulkhead structure, which might incorporate
some flow spreading for energy recovery. The spreading of the flow is an economic deci-
sion based on construction costs and the value of energy loss. Figure 8.15 shows a typi-
cal structure of a tail-tunnel outlet. If the project is the pumped-storage type, the outlet
structure will incorporate trashracks at the face of the structure, and the velocity at the
trashracks will be approximately 1.0 m/s (3.3 ft/s), because the racks tend to be self-clean-
ing during the generating mode.

8.9 TAILRACECHANNELS
If the outlet structure is a significant distance from the receiving waterway, a tailrace
channel will be required (Fig. 8.16). The sizing of the channel will be similar to that of the
headrace channel. (Refer to Sec. 8.2).
TAIL-TUNNEL SURGE CHAMBER

FIGURE 8.14 A typical tail-tunnel surge chamber. (Harza Engineering Co.)


STEEL CMATMC COVEK

PLAN EL. 1670.0

ROAO
STEEL CHATMC COVCH
2.0 FREEBOARD

(SLOT FON TRASHRACK


SEE NOTE 31
BULKHEAD SLOT

BULKHEAD SHAFT SECTION INTAKE /OUTLET SECTION

BULKHEAD SHAFT PLAN


INTAKE /OUTLET PLAN
LOWER RESERVOIR
NOTCS:
L ALL DIMENSIDATUM
ONS ANO2000ELEVATONS AAE M KTERS.SCA LEVEL.
2. PKOJCCT
ALL CONCKETEWLLFORCL.HAVE
CONDUCTORS
• UCOUCRBANEAN
HEADRACE
A MMUUUANO COUPRESSI
TAlRACE VNAfER
E STRENClH
3. THE TRASHRACK BAR SP*CNC SHALL BC EQUAL TO 20 CU.
SCALE O IO 20 30 METERS
1 i SOO
THE ISRAEL ELECTRIC CORPORATION LIMITED
PARSA PUMPED-STORAGE PROJECT
WATER CONDUCTORS
INTAKE /OUTLET STRUCTURFQ

FIGURE 8.15 A typical tail-tunnel outlet structure. (Harza Engineering Co.)


General Plan of the Guri Project
Left embankment dam
(Final Stag*)
Left gravity dorr
(Final Stag*)
Exi
gravistitnygdam
loft
Flow
Powerhouse No. I
Raising of right Substoae U
Substag*
Existingdam
gravity right

Powerhouse Na 2
Final Stage
Extensiondamto main
gravity

Ridam
ght (Final
embankment-
Stage)
Trestle

FIGURE 8.16 Tailrace channels of the Guri Project. (EDELCA, Venezuela)

REFERENCES
American Society of Civil Engineer (ASCE), Civil Engineering Guidelines for Planning and
Designing Hydroelectric Developments: Vol. 2 Waterways, American Society of Civil Engineers,
New York, 1989.
American Society of Civil Engineer (ASCE), Steel Penstock, ASCE Manuals and Reports on
Engineering Practice No. 79, American Society of Civil Engineers, New York, 1993.
Brater, E. R, King,
H. W., J. E. Lindell, and C. Y. Wei, Handbook of Hydraulics, 7th ed., McGraw-Hill, New York,
1996.
Chaudhry, M. H., Applied Hydraulic Transients, 2nd ed., Van Nostrand Reinhold, New York, 1987.
Chow, V. T., Open-Channel Hydraulics, McGraw-Hill, New York, 1959.
Dingman, S. L., Fluvial Hydrology, W. H. Freeman, New York, 1984.
Gulliver, J. S., and R. E. A. Arndt, Hydropower Engineering Handbook, McGraw-Hill, New York,
1991.
Henderson, F. M., Open Channel Flow, Macmillan, New York, 1966.
Moffat, A. I. B., C. Nalluri, and R. Narayanan, Hydraulic Structures, Unwin Hyman, London, UK,
1990.
Parmakian, J., Waterhammer Analysis, Dover Publications, New York, 1955.
Rich, G. R., Hydraulic Transients, Dover Publications, New York, 1951.
U. S. Army Corps of Engineer (USAGE), Hydraulic Design Criteria, U.S. Army Corps of
Engineers, Waterways Experiment Station, Vicksburg, MS, 1988.
U.S. Bureau of Reclamation, Selecting Hydraulic Reaction Turbines, Engineering Monograph
No.20, Department of the Interior, 1980.
U. S. Bureau of Reclamation, Design of Small Dams, U.S. Department of the Interior, Denver, Co,
1987.
U. S. Bureau of Reclamation Welded Steel Penstocks, Engineering Monograph No.3, U.S.
Department of the Interior, Denver, Co, 1967.
Vanoni, V. A., ed., Sedimentation Engineering, American Society of Civil Engineers, New York
1977.
Warnick, C. C., H. A. Mayo Jr., J. L. Carson, and L. H. Sheldon, Hydropower Engineering,
Prentice-Hall, NJ, 1984.
Wylie, E. B., and V. L. Streeter, Fluid Transients in Systems, Prentice-Hall, Englewood Cliffs, NJ,
1993.
Zipparro, V. J., and H. Hasen, Davis' Handbook of Applied Hydraulics, 4th ed., McGraw-Hill, New
York, 1993.

You might also like