Вы находитесь на странице: 1из 15

SPILLWAY REHABILITATION: A LABYRINTH WEIR DESIGN TOOL

Brian M. Crookston Ph.D., Postdoctoral Researcher, Utah Water Research Laboratory,


Logan, Utah
Blake P. Tullis, Ph.D., Associate Professor, Utah State University, Logan, Utah
ABSTRACT
A design example is presented to illustrate the labyrinth weir hydraulic design and
analysis method developed by the authors. This design method is based upon
experimental results from physical models tested at the Utah Water Research Laboratory.
It is valid for trapezoidal labyrinth weirs with sidewall angles from 6 to 35, quarterround and half-round crest shapes, and it includes various design parameters and
hydraulic conditions that affect flow performance, including weir geometry, orientation
(e.g., reservoir vs. channel), cycle configuration, cycle efficiency, tailwater submergence,
local submergence, nappe aeration conditions, nappe vibration, nappe instability or flow
surging, and artificial aeration (vents, nappe breakers). The validity of this method is
established by comparing predicted results to data from previously published labyrinth
weir studies.
INTRODUCTION
General Overview
Labyrinth weirs (see Fig. 1) provide an increase in crest length for a given
channel or spillway width, relative to linear weirs. The additional crest length increases
weir discharge for a given upstream head. As a result of their versatility and hydraulic
performance (large discharges at relatively low heads) labyrinth weirs have become
widely used globally as headwater control structures, energy dissipaters, flow aerators,
and spillways (including spillway rehabilitation) in a variety of applications.
Due to the unique geometric shape, the flow passing over a labyrinth weir is
three-dimensional and complex, and exact solutions for head-discharge relationships are
not readily determined analytically. However, Eq. (1) and empirically determined
labyrinth weir dimensionless discharge coefficients, Cd(), can be used to determine the
head-discharge relationship.

2
32
Q = C d ( ) Lc 2 g H T
3

(1)

In Eq. (1), Q is the labyrinth weir flow rate, Lc is the centerline length of the weir crest, g
is the acceleration constant of gravity, and HT is the total upstream head defined as HT =
V2/2g + h (V is the average cross-sectional velocity and h is the piezometric head
upstream of the weir relative to the weir crest elevation, see Fig. 1).

Fig. 1. Labyrinth weir schematic, including flow parameters,


geometric parameters, and crest shapes
Previous Labyrinth Weir Studies

Published research on the hydraulic performance of labyrinth weirs began as early


as 1940 and is still of interest to researchers and designers today. A number of
noteworthy research studies have been conducted over the past 50 years that have
contributed to the evolution of labyrinth weir design; a selection is presented in Table 1.
The Tullis et al. (1995) design method is currently widely accepted in the USA
and presents an intuitive spreadsheet-based labyrinth weir design program [the Cd() data
is also used in a spreadsheet presented by Falvey (2003)]. This method is for trapezoidal,
quarter-round crest applications and utilizes an effective weir length, Le, as the
characteristic weir length in Eq. (1) to partially account for apex influence on discharge
capacity. However, apex influence is more accurately examined via nappe interference
and is therefore an unnecessary complexity.
Willmore (2004) found the Tullis et al. (1995) = 8 data to be incorrect and
discovered a minor mathematical error in the geometric calculations. Also, the = 6
Cd() data display a different trend and are significantly lower than the adjacent curves.

Table 1. Notable labyrinth weir design methods


Study
1

Taylor (1968), Hay and Taylor


(1970)

Darvas (1971)

Crest

Type

Sh
HR

Triangular
Trapezoidal
Rectangular

LQR

Trapezoidal
Triangular
Trapezoidal
Triangular
Trapezoidal

Hinchliff and Houston (1984)

Sh
QR

Lux and Hinchliff (1985)

QR

Magalhes and Lorena (1989)

6
7

WES

Trapezoidal

Tullis et al. (1995)

QR

Trapezoidal

Melo et al. (2002)

LQR

Trapezoidal

Tullis et al. (2007)

HR

Trapezoidal

Lopes et al. (2006, 2008)

LQR

Trapezoidal

QR
10
Crookston (2010)
Trapezoidal
HR
QR Quarter-round (Rcrest=tw/2), LQR Large Quarter-round (Rcrest = tw),
HR Half-round, Sh Sharp, WES Truncated Ogee

The experimental data for the Tullis et al. (1995) method were limited to 6 18,
with the = 25 and 35 curves linearly interpolated (with the aid of = 90 weir data).
Objective

The purpose of this study was to improve current labyrinth weir hydraulic design
and analysis tools by providing new insights, information, and experimental results. This
is to be accomplished by utilizing the experimental results from physical modeling to
provide a design optimization and analysis program and supportive hydraulic information
(e.g., artificial aeration, nappe stability). The design program (see Crookston 2010)
developed during this study is similar to the design procedure presented by Tullis et al.
(1995) with the addition of the following: new Cd() data sets for quarter-round and halfround crests, a user-specified footprint size (channel width, W, and apron length, B),
cycle efficiency (), nappe aeration conditions and behaviors, aeration device placement,
and tailwater submergence effects. This method utilizes Lc instead of the effective crest
length presented by Tullis et al. (1995).
EXPERIMENTAL METHOD

32 labyrinth weir physical models were rigorously tested at the Utah Water Research
Laboratory (UWRL), located in Logan, Utah, USA (Crookston, 2010). Labyrinth weirs
were fabricated from High Density Polyethylene (HDPE) sheeting. A laboratory flume
(4 ft x 48 ft x 3 ft) and an elevated headbox (24 x 22 ft x 5 ft deep) were used for
experimental investigations. All models were placed on an elevated horizontal platform
(level to 1/64 in). The flume facility included a ramped upstream floor transition, which
was reported by Willmore (2004) to have no influence on discharge capacity. In the head
box test facility, the discharge channel downstream of the weir was relatively short (~4
in) and terminated with a free overfall. Where a rounded abutment wall inlet was

Table 2. Physical model test program


Model

P
Lc-cycle
Lc-cycle/w
w/P
N
Crest
()
()
()
(mm)
(mm)
()
()
()
()
1
6
0
304.8
4,654.6
7.607
2.008
2
HR
2-3
6
0
304.8
4,654.6
7.607
2.008
2
QR HR
4
6
0
203.2
3,075.5
7.607
2.008
5
HR
5-7
6
10, 20, 30
203.2
3,075.5
7.607
2.008
5
HR
8
6
0
203.2
3,075.5
7.607
2.008
5
HR
9
6
0
203.2
3,075.5
7.607
2.008
5
HR
10-11
8
0
304.8
3,544.9
5.793
2.008
2
QR HR
12-13
10
0
304.8
2,879.1
4.705
2.008
2
QR HR
14-15
12
0
304.8
2,435.1
3.980
2.008
2
QR HR
16
12
0
203.2
634.6
4.705
2.008
5
HR
17-19
12
10, 20, 30
203.2
634.6
4.705
2.008
5
HR
20
12
0
203.2
634.6
4.705
2.008
5
HR
21
12
0
203.2
634.6
4.705
2.008
5
HR
22-23
15
0
304.8
1,991.4
3.254
2.008
2
QR HR
24
15
0
152.4
1,991.4
3.254
4.015
2
QR
25
15
0
152.4
995.7
3.254
2.008
4
QR
26
15
0
304.8
995.7
3.254
1.019
4
QR
27-28
20
0
304.8
1,548.1
2.530
2.008
2
QR HR
29-30
35
0
304.8
983.5
1.607
2.008
2
QR HR
31-32
90
304.8
1,223.8
1.000
4.015
QR HR
Linear cycle configuration was used for all model orientations unless Arced is specified
Based upon the outlet labyrinth cycle geometry

Type .
()
Trap
Trap
Trap
Trap
Trap
Trap
Trap
Trap
Trap
Trap
Trap
Trap
Trap
Trap
Trap
Trap
Trap
Trap
Trap
-

Orientation
()
Inverse
Normal
Projecting
Arced & Projecting
Flush
Rounded Inlet
Normal
Normal
Normal
Projecting
Arced & Projecting
Flush
Rounded Inlet
Normal
Normal
Normal
Normal
Normal
Normal
-

Fig. 2. Tested labyrinth weir orientations


modeled, the radius was set to the cycle width (Rabutment = w). Details of the physical
model test program are summarized in Table 2 and Figure 2.
Model test flow rates were determined from calibrated orifice meters in the flume
supply piping, differential pressure transducers, and a data logger. The test program
evaluated the performance of aeration vents (1 per sidewall) and wedge-shaped nappe
breakers in a variety of locations (upstream apex, weir sidewall, downstream apex).

Experimental data were collected under steady-state conditions. A large number


of head-discharge data points were collected for all tested weir geometries, and a system
of checks was established wherein at least 10% of the data were repeated to ensure
accuracy and determine measurement repeatability. Q measurements were recorded for 5
to 7 minutes with the data logger to determine an average flow rate, and h (0.006 in)
was determined with a stilling well equipped with a point gauge. Digital photography
and high-definition (HD) digital video recording were used extensively to document the
hydraulic behaviors of the tested labyrinth weirs.
Testing included velocity
measurements and a dye wand to examine flow patterns. Observations also noted nappe
aeration conditions and behavior, nappe stability, nappe separation point, nappe
interference, areas of local submergence, and any harmonic or recurring hydraulic
behaviors for all tested.
EXPERIMENTAL RESULTS
Overview

A new hydraulic design method, based upon the experimental results of this
study, was developed for trapezoidal labyrinth weirs with a quarter- or half-round crest
shape (see Crookston 2010). It is based upon Eq. (1) and utilizes experimentally
determined Cd() values to calculate head-discharge relationships for labyrinth weirs
located in a channel or reservoir application. The design table format introduced by
Tullis et al. (1995) has been adapted to incorporate additional design information and
parameters (e.g., nappe breakers, nappe behavior, cycle efficiency, and submergence).
The following example illustrates how to use this new method to hydraulically design a
labyrinth weir. The design example also includes corresponding design information from
Crookston (2010).
Design Example

A labyrinth weir is being considered as the control structure of a spillway. The


allowable spillway width is ~100 ft; the spillway is required to pass ~16,500 cfs (e.g.,
75% of PMP); and due to urban development, it is desirable to limit the upstream pool
elevation. This information is entered into Table 3. This section also includes desired
apron and crest elevations and downstream total head (relative to crest) for labyrinth weir
submergence calculations (e.g., located on a river with high tailwater effects).
Basic labyrinth weir geometry is input into the second section of the design
method, shown in Table 4. Due to the geometric complexities of labyrinth weirs, a
number of iterations will likely be necessary to determine a satisfactory labyrinth weir
design. For this example, the selected initial values are a quarter-round crest shape, a
sidewall angle of 8, a weir height of 12 ft, the number of cycles N = 4, and a channel
width less than 100 ft. The wall thickness (at the crest) and the apex width have been
selected to be geometrically similar to the physical models tested in this study. This
section also provides the option to include an aeration device (e.g., nappe breakers,
vents). Depending upon the nappe aeration behavior and the potential for debris, it may

Table 3. User-defined hydraulic conditions input section of design method


Symbol

Qdesign
H
Hapron
Hcrest
HT
Hd

Design Flow
Design Flow Water Surface Elevation
Approach Channel Elevation
Crest Elevation
Unsubmerged Total Upstream Head
Downstream Total Head

Value

Units

16,557

Notes

ft /s

Input

582.00

ft

Input

563.00

ft

Input

g = 32.174 ft/s2

575.00

ft

Input

7.00

ft

Input (Pies. Head + Vel. Head Losses)

2.75

ft

Input (Pies. Head + Vel. Head)

above weir crest

Table 4. Geometric user-defined input section of design method

Angle of Side Legs


Width of labyrinth (Normal to Flow)
Number of Cycles
Wall Height
Thickness of Wall at Crest
Inside Apex Width
Crest Shape
Nappe Breakers / Vents

Symbol

Value

Units

W
N
P
tw
A

~ 6 - 35

Notes

95.96

ft

Input or W = Nw

Input or N = W/w

12.00

ft

P ~ 1.0HT

1.50

ft

tw ~ P/8

1.50

ft

A ~ tw

Crest Shape

Quarter

Quarter or Half

None

Breakers, Vents, None

0.140 (rad)

or may not be desirable to include a nappe breaker (see Fig. 3). This optional structural
feature will be discussed further in the nappe behavior section of this design method.
The third section of this design method (see Table 5) calculates various labyrinth
weir ratios and geometric dimensions, including the headwater ratio and discharge
coefficient. This study developed Cd() vs. HT/P for 6 35 for trapezoidal
labyrinth weirs with quarter- and half-round crest shapes (see Crookston 2010). For
convenience, curve-fit equations were provided, and linear interpolation is recommended
for values not tested. Fig. 4 presents the experimental Cd() vs. HT/P results, and Eq.
(2) corresponds to the quarter-round = 8 curve fit equation (see Crookston 2010 for
additional Cd() equations).

Cd (8) = 0.03612

HT
P

2.576

H T 0.4104
P

+ 0.1936

QR, = 8

(2)

At the design flow rate, this labyrinth weir has HT/P = 0.58 and Cd(8) = 0.304; a full
rating curve for this proposed labyrinth weir is presented in Fig. 5. The apron length (B)
is ~ 67 ft, a cycle is 24 ft wide (w), and the labyrinth provides nearly 551 ft of total crest
length. As previously mentioned, this method provides hydraulic design information for
quarter- and half-round crest shapes. The increase in discharge capacity associated with
using a half-round crest rather than the quarter-round crest for the 8 labyrinth weir

Fig. 3. Nappe breaker, located on the downstream apex


Table 5. Free-flow calculations section of design method

Headwater Ratio
Labyrinth Weir Discharge Crest Coefficient
Total Centerline Length of Weir
Centerline Length of Sidewall
Length of Apron (Parallel to Flow)
Cycle Width
Outside Apex Width
Cycle Width Ratio
Relative Thickness Ratio
Apex Ratio
Cycle Efficiency
# of Nappe Breakers or Vents
Linear Weir Discharge Coefficient
Length of Linear Weir for equivalent Q

Symbol

Value

Units

HT/P
Cd()
Lc
lc
B
w
D
w/P
P/tw
A/w

Cd(90)
Lc-linear

0.583

Notes

0.304

Cd() = f(HT/P, , Crest Shape, Aeration)

550.77

ft

Lc = 3/2Qdesign/[(Cd()HT3/2)(2g)1/2]

66.04

ft

lc = (B-tw)/cos()

66.90

ft

Input or B = [Lc/(2N)-(A+D)/2]cos()+tw

23.99

ft

w = 2lcsin()+A+D

4.108

ft

D = A+2twtan(45-/2)

2.00

~2 w/P ~4

8.00

0.063

< 0.08

1.742

= Cd()Lc/(wN)

none

Breakers on DS Apex, Vents on Sidewall

0.829

Cd(90) = f(HT/P, , Crest Shape)

201.71

ft

Lc-linear = 3/2Qdesign/[(Cd(90)HT3/2)(2g)1/2]

ranges from 0 to 15%, as shown in Fig. 6, over the range of HT/P values evaluated, Table
5 also presents the length of linear weir (Lc-linear) that would be required to pass the
equivalent discharge as the labyrinth weir (same crest shape) at the same value of HT. In
this example, the labyrinth weir reduces the required channel width (W) by ~106 ft,
relative to the linear weir, and provides nearly 3 times the weir length.
The cycle width ratio (w/P), the relative thickness ratio (P/tw), and the apex ratio
(A/w) correspond to the physical models tested in this study. The predictions of this

Fig. 4. Cd vs HT/P for quarter round trapezoidal labyrinth weirs ( = 8 and 15)

Fig. 5. Predicted head-discharge rating curve for proposed labyrinth weir geometry

Fig. 6. Hydraulic efficiency crest shape comparison for trapezoidal labyrinth weir, = 8
design method may deviate from actual weir performance if these limits, listed in Table
5, are exceeded.
The cycle width ratio (w/P), the relative thickness ratio (P/tw), and the apex ratio
(A/w) correspond to the physical models tested in this study. The predictions of this
design method may deviate from actual weir performance if these limits, listed in Table
5, are exceeded.
Cycle Efficiency

When trying to optimize a labyrinth weir geometry for a given channel width, it is
useful to note that Cd(), which is proportional to the discharge per unit weir length,
decreases with decreasing . Conversely, Lc increases with decreasing . To characterize
the combined influence of these two counter influences on discharge capacity, a new
parameter is introduced: cycle efficiency (), where = Cd()Lc/(wN). compares the
discharge efficiencies of different cycle geometries for a given channel width; discharge
efficiency increases with increasing values of . The proposed 8 labyrinth weir has =
1.742.
As a comparison in the design example, assume that an = 15 labyrinth weir
alternative is also being considered (w = 24 ft, W = 96 ft). At the same HT/P condition
listed in Table 5 (HT/P = 0.583) for the 8 labyrinth weir, = 1.443 for the 15 labyrinth,
which corresponds to a 17% reduction in discharge (Q = 13,723 cfs) relative to the 8
weir. To pass the design flow rate of 16,577 cfs, the 15 labyrinth weir requires an HT/P
value of 0.704 [determined using Eq. (1) and the Cd(15) data presented in Fig. 4], which
represents a 20.8% increase in HT relative to the 8. Although a higher upstream head is
required to pass the design discharge, it is worth noting that the 15 weir length is ~44%
shorter (Lc = 310 ft) than the 8 weir and the apron length is 34% shorter (B = 44 ft). In
addition to the hydraulic performance of the weir, a proper feasibility study would also
include economic and other considerations in selecting the most appropriate labyrinth
weir design.

Approach Flow Influence

The discharge rating curve of a labyrinth weir can be influenced by the approach
flow conditions (e.g., reservoir approach flow, spillway entrance geometry). The Cd()
data used in the labyrinth weir design method are based on channelized approach flow
conditions. If a labyrinth weir is installed in a reservoir application, the head-discharge
relationship will be influenced by how and where the labyrinth weir is laid out (e.g.,
projecting into the reservoir, at the upstream end of the discharge channel with abutment
wall transitions, etc.) Experimental results from Crookston (2010) showed that the outer
1-2 cycles on either end of the labyrinth weir can be affected by the abutments in a
reservoir application, resulting in a reduced discharge capacity relative to the channelized
approach condition.
Assume that the proposed 4-cycle, 8 labyrinth weir is located in a reservoir, and
features rounded abutment walls. According to the data in Fig. 7, at a dimensionless
upstream head (HT/P) of 0.583 (HT = 7ft), the discharge capacity of the reservoir
labyrinth weir (15,729 cfs) will be approximately 5% less than Qdesign (16,557 cfs). Note
that Figure 7 was developed by comparing 5-cycle labyrinth weir data for various
reservoir applications to the same weir geometry in a channel application. The percent
reduction in discharge efficiency will decrease as the cycle number increases. As needed,
the design table can be used to predict a new HT/P ratio and discharge coefficient to
address this adjustment and satisfy the design requirements (Q = 16,557 cfs / 0.95, HT =
7.4 ft, Cd() = 0.294). Relative discharge data (Q-Res/Q-Channel) for additional inlet
geometries and arced labyrinth weirs can be found in Crookston (2010) to make similar
rating curve adjustments.

Fig. 7. Hydraulic efficiency comparison for labyrinth weir


with rounded inlet (reservoir) and located in a channel
Nappe Behavior

In addition to the head-discharge data, nappe behavior should also be considered


in an effort to reduce the potential for undesirable pressure fluctuations, noise, vibrations,
and flow surging. The range of nappe aeration conditions labyrinth weirs can experience
include: clinging, aerated, partially aerated, or drowned. A clinging nappe is attached to
the downstream face of the weir wall and is hydraulically more efficient than an aerated

nappe; the aerated nappe has an air cavity underneath. A partially aerated nappe is not
fully aerated along the entire weir crest and/or the air cavity may be transient (spatially
and temporally) in nature. A drowned nappe occurs at relatively higher HT/P and is
characterized by a thick nappe without an air cavity. The range of HT/P associated with
the various nappe aeration conditions for the 8 and 15 quarter-round crest are illustrated
in Fig. 4 and summarized in Table 6. Nappe aeration behaviors for trapezoidal labyrinth
weirs, 6 35, with quarter- and half-round crest shapes are documented in
Crookston (2010).
Table 6. Labyrinth weir nappe behaviors
Quarter-Round Crest Shape (HT/P)
= 8
= 15
Clinging
none
none
Aerated
0.057-0.288
0.052-0.256
Partially Aerated
0.288-0.364
0.256-0.508
Drowned
>0.364
>0.508
Nappe Vibration
<0.06
<0.06
Instability
none
0.271-0.468

Labyrinth weirs can also develop nappe vibration (HT/P 0.06), or at higher
heads, an unstable nappe (also termed flow surging). Although nappe instability was
observed in the laboratory, the scale of the physical models was not of sufficient size to
produce nappe vibrations (P 3 ft required). Nappe vibration can cause objectionable
noise that may require remediation as occurred with the Avon spillway in Australia. A
laboratory study was conducted of Avon spillway that documented the flow conditions
when vibrations were observed, and various countermeasures were explored
(Metropolitan Water, Sewerage, and Drainage Board, 1980). Some weir elements that
may prevent nappe vibration include a half-round crest, crest roughness elements, nappe
breakers, or staged or notched cycles.
Nappe instability (also shown in Fig. 4) describes a nappe with an oscillating
trajectory and may be accompanied by sudden changes in the nappe aeration condition.
This is a low frequency phenomenon that occurs under constant upstream flow conditions
and produces fluctuations in Q and h locally along the weir, as well as significant noise,
pressure fluctuations, and increased turbulence. The magnitude of nappe instability
increases with increasing . The net effect of nappe instability on prototype structures is
unclear; as a precaution, however, avoiding these unstable nappe operating regions in
labyrinth weir design is recommended.
Submergence

The final section of this design method incorporates the tailwater submergence
procedure developed by Tullis et al. (2007), shown in Table 7. Assume that the proposed
8 labyrinth weir is located on a river and the tailwater elevation at the Qdesign is predicted
to be 2.75 ft higher than the weir crest. The effects of tailwater submergence may be
determined by using the Tullis et al. (2007) submergence piecewise equation, presented

as Eqs. (3) (5). As previously defined, HT represents the free-flow (unsubmerged)


upstream total head. H* represents the total upstream head under submerged conditions
(H* HT). Hd is the total downstream head; all heads are relative to the weir crest
elevation. Calculated results are shown in Table 7, including tailwater submergence
levels (S). Tailwater submergence resulted in H* (7.22 ft) increasing 3.2% relative to HT.
4

H
H*
= 0.0332 d
HT
HT

H
+ 0.2008 d

HT

H
H*
= 0.9379 d
HT
HT

+ 0.2174

+ 1

Hd
1.53
HT

1.53

3.5

H* = Hd

(3)

Hd
3.5
HT

Hd
HT

(4)

(5)

Table 7. Tailwater submergence section of design method

DS/US Unsubmerged Head Ratio


Submerged Head Discharge Ratio
Submerged Upstream Total Head
Submergence
Submerged Weir Discharge Coefficient

Symbol

Value

Units

Hd/HT
H*/HT
H*
S
Cd-sub

0.39

Notes

1.032

7.222

ft

0.381

S = Hd/H*

0.290

Cc()(HT/H*)3/2

Eqs. (3) (5), from Tullis et al. (2007)

The design method and support data are limited to the geometries (Table 2) and
hydraulic conditions tested in this study (e.g., 0.05 HT/P 0.9). However, this method
can be conservatively applied to geometrically comparable labyrinth weir geometries and
similar flow conditions. The design method may be used as a first-order approximation
for HT/P > 1.0, based on the general trends observed in the available supporting data from
the current study [Model 13 tested to HT/P = 2.0 (QR), see Table 2]. Nevertheless, a
hydraulic model study is recommended to confirm design method predictions, especially
for hydraulic conditions and labyrinth weir geometries that deviate from the tested
physical models.
CONCLUSION

The design method developed by Crookston (2010) and illustrated here is a useful
tool for determining labyrinth weir hydraulic performance; it can also be used to estimate
the hydraulic performance for labyrinth weir geometric configurations and/or operating
conditions not specifically included in available labyrinth weir design data. The design
example presented in this paper illustrates how to use the equations, figures, tables, and
information presented by Crookston (2010).
This design method is valid for trapezoidal labyrinth weirs, 6 35, with
quarter-round and half-round crest shapes. It includes channel and reservoir applications,

linear and arced cycle configurations, tailwater submergence from Tullis et al. (2007),
nappe aeration conditions, nappe vibration, nappe instability (flow surging), and artificial
aeration. Crookston (2010) established the validity of this method by comparing
predicted results to data from previously published labyrinth weir studies. This method
may be applied to geometrically comparable labyrinth weir geometries and flow
conditions (not evaluated during physical modeling) as a first order approximation;
however, a hydraulic model study is recommended to confirm design method predictions.
ACKNOWLEDGEMENTS

This study was funded by the State of Utah and the Utah Water Research Laboratory.
AUTHORS

Brian M. Crookston is a postdoctoral researcher at the Utah Water Research


Laboratory, located in Logan, Utah. Labyrinth weirs were the research subject for his
Ph.D. dissertation, and he has over 6 years of experience performing hydraulic model
studies.
Blake P. Tullis is an associate professor at Utah State University. He has over 15
years of experience performing hydraulic model studies at the Utah Water Research
Laboratory.
NOTATION

A
A/w

B
Cd()
Cd(90)
Cd-sub
D

g
H
h
H*
H*/HT
Hapron
Hcrest
Hd
Hd/H*

Inside apex width


Apex ratio
Sidewall angle
Length of labyrinth weir (Apron) in flow direction
Labyrinth weir free-flow discharge coefficient, data from current study
Linear weir discharge coefficient
Labyrinth weir submerged discharge coefficient
Outside apex width
Efficacy
Cycle efficiency
acceleration constant of gravity
Design flow water surface elevation
depth of flow over the weir crest
Submerged upstream total head
Submerged head discharge ratio
Approach channel elevation
Elevation of labyrinth weir crest
Total tailwater head downstream of labyrinth weir (above the weir
crest)
Downstream/upstream unsubmerged head ratio

HT
HT/P
Lc
lc
Lc-cycle
Lc-cycle/w
Lc-linear
Le
N
P
P/tw
Q
Qchannel
Qdesign
QHR
QQR
Qres
Rcrest
Rabutment
S
tw

V
W
w
w/P

Unsubmerged total upstream head on weir


Headwater ratio
Total centerline length of labyrinth weir
Centerline length of weir side wall
Centerline length for a single labyrinth weir cycle
Magnification ratio, M
Centerline length for a linear weir
Total effective length of labyrinth weir
Number of labyrinth weir cycles
Weir height
Relative thickness ratio
Discharge over weir
Discharge specific to a labyrinth weir located in a channel
Design flow rate specified in design method
Discharge specific to a labyrinth weir with a half-round crest
Discharge specific to a labyrinth weir with a quarter-round crest
Discharge specific to a labyrinth weir located in a reservoir
Radius of crest shape
Radius of rounded abutment
Submergence level
Thickness of weir wall
Cycle arc angle
Average cross-sectional flow velocity upstream of weir
Width of channel
Width of a single labyrinth weir cycle
Cycle width ratio
REFERENCES

Crookston, B. M. (2010). Labyrinth weirs. Ph.D. Dissertation. Utah State University,


Logan, Utah.
Darvas, L. (1971). Discussion of performance and design of labyrinth weirs, by Hay
and Taylor. J. of Hydr. Engrg., ASCE, 97(80), 1246-1251.
Falvey, H. (2003). Hydraulic design of labyrinth weirs. ASCE, Reston, Va.
Hay, N., and Taylor, G. (1970). Performance and design of labyrinth weirs. J. of
Hydr. Engrg., ASCE, 96(11), 2337-2357.
Hinchliff, D., and Houston, K. (1984). Hydraulic design and application of labyrinth
spillways. Proc. of 4th Annual USCOLD Lecture.
Lopes, R., Matos, J., and Melo, J. (2006). Discharge capacity and residual energy of
labyrinth weirs. Proc. of the Int. Junior Researcher and Engineer Workshop on

Hydraulic Structures (IJREWHS 06), Montemor-o-Novo, Hydraulic Model Report


No. CH61/06, Div. of Civil Engineering, the University of Queensland, Brisbane,
Australia, 47-55.
Lopes, R., Matos, J., and Melo, J. (2008). Characteristic depths and energy dissipation
downstream of a labyrinth weir. Proc. of the Int. Junior Researcher and Engineer
Workshop on Hydraulic Structures (IJREWHS 08), Pisa, Italy, 51-58.
Lux III, F. and Hinchliff, D. (1985). Design and construction of labyrinth spillways.
15th Congress ICOLD, Vol. IV, Q59-R15, Lausanne, Switzerland, 249-274.
Magalhes, A., and Lorena, M. (1989). Hydraulic design of labyrinth weirs. Report
No. 736, National Laboratory of Civil Engineering, Lisbon, Portugal.
Melo, J., Ramos, C., and Magalhes, A. (2002). Descarregadores com soleira em
labirinto de um ciclo em canais convergentes. Determinao da capacidad de vazo.
Proc. 6 Congresso da gua, Porto, Portugal. (in Portuguese).
Metropolitan Water, Sewerage and Drainage Board. (1980). Investigation into spillway
discharge noise at Avon Dam. ANCOLD Bulletin No. 57, 31-36.
Taylor, G. (1968). The performance of labyrinth weirs. Ph.D. thesis, University of
Nottingham, Nottingham, England.
Tullis, B., Young, J., & Chandler, M. (2007). Head-discharge relationships for
submerged labyrinth weirs. J. of Hydr. Engrg., ASCE, 133(3), 248-254.
Tullis, P., Amanian, N., and Waldron, D. (1995). Design of labyrinth weir spillways.
J.of Hydr. Engrg., ASCE, 121(3), 247-255.
Willmore, C. (2004). Hydraulic characteristics of labyrinth weirs. M.S. report, Utah
State University, Logan, Utah.

Вам также может понравиться