Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/272489063

The Design and Analysis of Labyrinth Weirs

Conference Paper · January 2011

CITATIONS READS
2 839

2 authors:

Brian Crookston B. P. Tullis


Utah State University Utah State University
82 PUBLICATIONS   288 CITATIONS    99 PUBLICATIONS   582 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Size Scale Effects of Nonlinear Weir Hydraulics View project

Dam Safety View project

All content following this page was uploaded by B. P. Tullis on 18 April 2019.

The user has requested enhancement of the downloaded file.


Hydraulic Design and Analysis of Labyrinth
Weirs. I: Discharge Relationships
B. M. Crookston, A.M.ASCE 1; and B. P. Tullis, M.ASCE 2

Abstract: A method is presented for the hydraulic design and analysis of labyrinth weirs based upon the experimental results of physical
Downloaded from ascelibrary.org by UTAH STATE UNIVERSITY on 04/11/13. Copyright ASCE. For personal use only; all rights reserved.

modeling. Discharge coefficient data for labyrinth weirs with quarter-round and half-round crest shapes are presented for sidewall angles
ranging from 6 to 35°. Cycle efficiency is also introduced as a design aid, which compares the hydraulic performance of different cycle
geometries. Geometric parameters that affect flow performance are discussed. The predictive accuracy of the design method is evaluated
through comparisons to previously published labyrinth weir head-discharge data. The companion paper examines nappe behaviors that affect
flow performance and presents hydraulic design considerations specific to nappe characteristics. DOI: 10.1061/(ASCE)IR.1943-4774
.0000558. © 2013 American Society of Civil Engineers.
CE Database subject headings: Weirs; Water discharge; Coefficients; Hydraulics; Design; Irrigation.
Author keywords: Labyrinth weir design; Discharge coefficients; Weir crest shape; Cycle efficiency; Local submergence.

Introduction direction) of equivalent length. However, as the head increases,


labyrinth weir discharge efficiency, as quantified by the discharge
A labyrinth weir is a linear weir that is folded in plan-view to in- coefficient value, begins to decline as nappe collision and local
crease the crest length for a given channel or spillway width (Fig. 1). submergence regions develop (Crookston and Tullis 2012c).
There are an infinite number of possible labyrinth weir configura-
tions and design variations; however, labyrinth cycles are typically
placed in a linear fashion (i.e., upstream apexes align at a common Previous Studies
channel cross section; Fig. 1), have a sidewall angle (α) less than Labyrinth weir head-discharge relationships have been described
30°, and are oriented towards the approaching flow. by various empirical equations. These relationships vary based
A labyrinth weir is able to pass large discharges at relatively low on different definitions of the discharge coefficient, the character-
heads compared to traditional linear weir structures of equal width. istic weir length, and the upstream driving head (e.g., the inclusion
As a result of their hydraulic performance and geometric versatility, of the velocity head component V 2 =2g, described in the following).
labyrinth weirs have been placed in streams, canals, rivers, ponds, In the present study, a standard form of the weir equation, Eq. (1),
and reservoirs as headwater control structures, energy dissipaters, was selected with the centerline length of the crest (Lc ) as the char-
flow aerators, and spillways. Labyrinth weirs are well suited for acteristic weir length:
spillway rehabilitation where aging infrastructure, dam safety con-
cerns, freeboard limitations, and revised and larger probable maxi- 2 pffiffiffiffiffi
mum flows have required increased spillway capacity. Recently Q ¼ Cdðα°Þ Lc 2gH3=2 ð1Þ
3 T
constructed examples are: Lake Brazos spillway in Texas (Vasquez
et al. 2007) and Lake Townsend spillway in Greensboro, North where Q = labyrinth weir discharge; Cdðα°Þ = dimensionless dis-
Carolina (Tullis and Crookston 2008). charge coefficient; Lc ¼ Nð2lc þ A þ DÞ where N = number of
cycles, lc = centerline length of the sidewall, A = inside apex length,
and D = outside apex length; g = acceleration constant of gravity;
Flow Characteristics and HT = total upstream head (unsubmerged) measured relative
The geometry of a labyrinth weir produces complex three- to the crest elevation [H T ¼ V 2 =2g þ h (V is the average cross-
dimensional flow patterns. At very low heads, it behaves similar sectional velocity at the gauging location, and h is the piezometric
to a linear weir (α ¼ 90°, oriented normally to the approach flow head upstream of the weir)].
Several earlier labyrinth weir studies resulted in published
1 design methods; a selection is presented and discussed. Hay and
Postdoctoral Researcher, Utah Water Research Laboratory, Dept. of
Civil and Environmental Engineering, Utah State Univ., 8200 Old Main Taylor (1970) presented parameter guidelines, based upon research
Hill, Logan, UT 84322-8200. E-mail: bcrookston@gmail.com by Taylor (1968), for sharp-crested triangular and trapezoidal
2
Associate Professor, Utah Water Research Laboratory, Dept. of Civil labyrinth weirs. Discharge rating curves for h=P < 0.6 were pre-
and Environmental Engineering, Utah State Univ., 8200 Old Main Hill, sented in terms of a labyrinth-to-linear weir discharge ratio (based
Logan, UT 84322-8200 (corresponding author). E-mail: blake.tullis@ on a common channel width, W, and h), requiring discharge infor-
usu.edu
mation for a linear weir (α ¼ 90°) of equivalent weir height (P),
Note. This manuscript was submitted on July 6, 2011; approved on
October 25, 2012; published online on October 29, 2012. Discussion per- wall thickness (tw ), and crest shape. The Bureau of Reclamation
iod open until October 1, 2013; separate discussions must be submitted for (USBR) conducted model studies to aid in the design of Ute Dam
individual papers. This paper is part of the Journal of Irrigation and (Houston 1982). Discrepancies between their experimental results
Drainage Engineering, Vol. 139, No. 5, May 1, 2013. © ASCE, ISSN and the recommendations by Hay and Taylor (1970) were attributed
0733-9437/2013/5-363-370/$25.00. to different definitions of upstream head [h, Hay and Taylor (1970);

JOURNAL OF IRRIGATION AND DRAINAGE ENGINEERING © ASCE / MAY 2013 / 363

J. Irrig. Drain Eng. 2013.139:363-370.


for the Tullis et al. (1995) method were limited to 6° ≤ α ≤ 18°
and α ¼ 90°; the α ¼ 25° and 35° Cdðα°Þ curves were linearly in-
terpolated. Falvey (2003) reviewed many past research studies
and added new interpretations; this document has contributed to an
increase in labyrinth weir use in practice.
Melo et al. (2002) expanded the work of Magalhães and Lorena
(1989) by adding an adjustment parameter for labyrinth weirs lo-
cated in a channel with converging sidewalls. Tullis et al. (2007)
developed a dimensionless submerged head-discharge relationship
(tailwater submergence) for labyrinth weirs that was verified by
Lopes et al. (2009).
Physical models have proven to be highly useful in designing
Downloaded from ascelibrary.org by UTAH STATE UNIVERSITY on 04/11/13. Copyright ASCE. For personal use only; all rights reserved.

and analyzing specific labyrinth weir geometries, and hydraulic de-


sign methods are useful tools to estimate the discharge performance
of labyrinth weir geometric configurations. The purpose of this
study is to provide new insight into the performance and operation
of labyrinth weirs and to improve the currently available design and
evaluation tools. This is to be accomplished by utilizing experimen-
tal results from physical modeling to provide a design optimization
program, an analysis program, and additional hydraulic informa-
tion. Paper II presents additional labyrinth weir hydraulic informa-
tion and design considerations based on nappe behavior (Crookston
and Tullis 2013). The design program developed during this study
is similar to that of Tullis et al. (1995), but accommodates a user-
Fig. 1. Labyrinth weir schematic including geometric parameters
specified footprint size (channel width, W, and apron length, B).
This method contains new, extensive, and accurate Cdðα°Þ data sets
for quarter-round and half-round crest shapes, utilizes Lc of the
H T , USBR] and limited geometric data. From the physical model crest as the characteristic length, and includes additional design
studies of Ute Dam and Hyrum Dam, Hinchliff and Houston (1984) considerations such as cycle efficiency (ε 0 ), aeration device place-
developed new design guidelines. Despite scope limitations, they ment, and tailwater submergence effects. The labyrinth weir design
provided valuable insights regarding labyrinth weir orientation and method presented in these two companion papers is complimented
placement in reservoir and channel applications. by the additional research and design information on: labyrinth
Based upon model studies of Avon and Woronora Dam, Darvas weirs in reservoir applications (Crookston and Tullis 2012b); linear
(1971) simplified labyrinth weir design by introducing an empirical and arced labyrinth weir cycle configurations (Crookston and Tullis
discharge equation and a discharge coefficient to accompany dis- 2012a); and labyrinth weir nappe interference and local submer-
charge rating curves for labyrinth weirs. However, Magalhães and gence (Crookston and Tullis 2012c).
Lorena (1989) juxtaposed this method with their own experimental
results for labyrinth weirs with a truncated ogee crest shape and
reported their curves to be systematically lower than the data Experimental Method
presented by Darvas (1971).
Lux and Hinchliff (1985) and Lux (1984, 1989) developed a Physical modeling of labyrinth weirs (Crookston 2010) was con-
new empirical equation that includes the cycle width ratio, w=P, ducted at the UWRL. Labyrinth weirs were fabricated from high
and an apex shape constant to determine the discharge of a single density polyethylene (HDPE) sheeting and tested in a rectangular
labyrinth cycle. Although this dimensionless equation applies to flume (1.2 × 14.6 × 1.0 m). The influence of acrylic sidewalls of
trapezoidal and triangular weirs, the inclusion of w=P complicates the flume on the sectional model hydraulic performance was con-
the weir equation and was limited to w=P ≥ 2.0. Similar parameter sidered to be negligible, based upon the findings of Johnson (1996).
limits have been set by other design methods that do not explicitly Details of the tests are summarized in Table 1.
include w=P in the head-discharge equation. Calibrated orifice meters in the flume supply piping, differential
Tullis et al. (1995) developed a design method based upon pressure transducers, and a data logger were used to determine flow
Eq. (1) and research conducted at the Utah Water Research Labo- rates. The flume was equipped with a headbox and baffle (to create
ratory (UWRL) by Waldron (1994), Tullis (1993), and Amanian uniform and tranquil approach flow conditions), a stilling well (hy-
(1987). Instead of using W or Lc, Tullis et al. (1995) introduced draulically connected to the flume 3P upstream from the labyrinth
an effective weir length, Le , as the characteristic weir length in weir), and a rolling instrument carriage. The point gauge instru-
Eq. (1) to define the head-discharge relationship for trapezoidal, mentation was carefully referenced to the crest elevation of the
quarter-round labyrinth weirs, Le < Lc , and was intended to ac- weirs. The labyrinth weirs were installed on an elevated horizontal
count for apex influences on discharge efficiency. They also apron with a ramped upstream floor transition. Experimental data
presented a spreadsheet-based labyrinth weir design program for were collected under steady-state conditions. The Q measurements
quarter-round crest applications. Willmore (2004), however, noted were recorded for 5 to 7 min with the data logger to determine an
the following discrepancies: the α ¼ 8° data fall above the α ¼ 9° average flow rate and h was determined with the stilling well
presented by Waldron (1994), and a minor mathematical error equipped with a point gage accurate to 0.15 mm.
was found in the weir length geometric calculations. The α ¼ 6° In an effort to accurately characterize the labyrinth weir behav-
data are also significantly lower and trend differently than the ior, ∼85 head-discharge data points were collected for each tested
adjacent curves. The supporting quarter-round crest shape data weir geometry (∼2,600 total). A system of checks was established

364 / JOURNAL OF IRRIGATION AND DRAINAGE ENGINEERING © ASCE / MAY 2013

J. Irrig. Drain Eng. 2013.139:363-370.


Table 1. Physical Model Test Program The influence of labyrinth weir orientation in the channel [distal
α a
P Lc-cycle Lc-cycle =w w=P N Crest apexes connecting to channel sidewalls as upstream apexes (re-
Model (°) (mm) (mm) (−) (−) (−) (−) ferred to as normal orientation) or as downstream apexes (inverse
orientation)] was evaluated by testing the α ¼ 6° labyrinth weirs
1b 6 304.8 4,654.6 7.607 2.008 2 HR
2–3 6 304.8 4,654.6 7.607 2.008 2 QR, HR
with both orientations. No measurable variations were observed
4–5 8 304.8 3,544.9 5.793 2.008 2 QR, HR between Cdðα°Þ data sets (data not presented). Consequently, the
6–7 10 304.8 2,879.1 4.705 2.008 2 QR, HR data in Figs. 2 and 3 are assumed to be applicable, independent
8–9 12 304.8 2,435.1 3.980 2.008 2 QR, HR of weir orientation. Crookston and Tullis (2012b) present addi-
10–11 15 304.8 1,991.4 3.254 2.008 2 QR, HR tional design information and discussion regarding labyrinth weirs
12 15 152.4 1,991.4 3.254 4.015 2 QR in reservoir applications and abutment effects on discharge.
13 15 152.4 995.7 3.254 2.008 4 QR For convenience, the labyrinth weir Cdðα°Þ data in Figs. 2 and 3
14 15 304.8 995.7 3.254 1.019 4 QR were curve-fit per Eq. (2), and the corresponding coefficients are
15–16 20 304.8 1,548.1 2.530 2.008 2 QR, HR presented in Tables 2 and 3. Eq. (3) was used for α ¼ 90° data
Downloaded from ascelibrary.org by UTAH STATE UNIVERSITY on 04/11/13. Copyright ASCE. For personal use only; all rights reserved.

17–18 35 304.8 983.5 1.607 2.008 2 QR, HR


and the corresponding coefficients are presented in the aforemen-
19–20 90 304.8 1,223.8 1.000 4.015 — QR, HR
tioned tables. The curves have been validated for 0.05 ≤ HT =
a
Linear configuration and normal orientation for all models unless noted. P < ∼0.8–0.9; however, due to the well-behaved nature of the
b
Inverse orientation. data and Eq. (2), the Cdðα°Þ curves have been extrapolated to
HT =P ¼ 1.0. Eqs. (2) and (3) were selected over polynomial rela-
in which at least 10% of the data were repeated to ensure accuracy tionships because of their improved data representation (R2 ≥ 0.99)
and to determine measurement repeatability. and extrapolation performance (they remain well-behaved up to
HT =P ≤ 2.0). Crookston et al. (2012) evaluated Eq. (2) for HT =P <
2 via physical and numerical modeling and found that Eq. 2 may be
Experimental Results used as a good first order approximation. When the Tullis et al.
(1995) polynomial Cdðα°Þ relationships are extrapolated, they incor-
rectly compute Cdðα°Þ (even producing negative values) beyond the
Discharge Rating Curves
upper limit of their experimental data (experimental data limited to
Eq. (1) was used to quantify the labyrinth weir head-discharge re- HT =P < 0.9). For labyrinth weirs:
lationship with Lc representing the characteristic weir length. The
term Cdðα°Þ can be influenced by weir geometry (e.g., P, tw , A, w, α,   b H T  c
HT P
and crest shape), weir abutments, flow conditions (H T , approach- Cdðα°Þ ¼ a þd ð2Þ
P
ing flow angle, local submergence, and nappe interference), and
nappe aeration conditions (clinging, aerated, partially aerated,
For linear weirs:
and drowned). The two-cycle sectional labyrinth weir models
evaluated in this study did not account for the influence of abut- 1
ments on discharge. Data for Cdðα°Þ are presented in terms of Cdð90°Þ ¼ þd ð3Þ
a þ b HPT þ HTc=P
H T =P for nonvented trapezoidal labyrinth weirs for 6° ≤ α ≤ 35°
in Fig. 2 (quarter-round crest shape) and Fig. 3 (half-round crest
shape). The data for α ¼ 90° (linear) weirs are also included for A comparison between the half-round and quarter-round exper-
comparison. imental data is presented in Fig. 4 as the ratio of the half-round over

Fig. 2. Values of Cd versus H T =P for quarter-round trapezoidal labyrinth weirs

JOURNAL OF IRRIGATION AND DRAINAGE ENGINEERING © ASCE / MAY 2013 / 365

J. Irrig. Drain Eng. 2013.139:363-370.


Downloaded from ascelibrary.org by UTAH STATE UNIVERSITY on 04/11/13. Copyright ASCE. For personal use only; all rights reserved.

Fig. 3. Values of Cd versus H T =P for half-round trapezoidal labyrinth weirs

Table 2. Curve-Fit Coefficients for Quarter-Round Labyrinth and Linear


Weirs, Validated for 0.05 ≤ H T =P < 0.9
α a b c d
6° 0.02623 −2.681 0.3669 0.1572
8° 0.03612 −2.576 0.4104 0.1936
10° 0.06151 −2.113 0.4210 0.2030
12° 0.09303 −1.711 0.4278 0.2047
15° 0.10890 −1.723 0.5042 0.2257
20° 0.11130 −1.889 0.5982 0.2719
35° 0.03571 −3.760 0.7996 0.4759
90° −2.3800 6.476 1.3710 0.5300

Table 3. Curve-Fit Coefficients for Half-Round Labyrinth and Linear


Weirs, Validated for 0.05 ≤ H T =P < 0.9 Fig. 4. Comparison of half-round and quarter-round crest shape on
hydraulic performance of labyrinth weirs
α a b c d
6° 0.009447 −4.039 0.3955 0.1870
8° 0.017090 −3.497 0.4048 0.2286
10° 0.029900 −2.978 0.4107 0.2520 HT =P. Improving weir approach flow conditions and using a more
12° 0.030390 −3.102 0.4393 0.2912 efficient crest shape can obtain further gains in efficiency. Brazos
15° 0.031600 −3.270 0.4849 0.3349 Dam (Waco, TX), for example, features an ogee-type crest [modi-
20° 0.033610 −3.500 0.5536 0.3923 fied half-round crest with an upstream radius of 1=3tw and a down-
35° 0.018550 −4.904 0.6697 0.5062 stream radius of 2=3tw (Willmore 2004)].
90° −8.60900 22.650 1.8120 0.6375
Nappe Behavior and Artificial Aeration
Labyrinth weir flow can produce several nappe behavior phenom-
the quarter-round Cdðα°Þ values (Cd-HR =Cd-QR ) versus HT =P. ena that should not be overlooked: nappe aeration conditions,
A crest that is rounded on the downstream face helps the flow stay nappe instability (also termed flow surging), and nappe vibrations.
attached (clinging flow) to the downstream weir wall at smaller These behaviors, conditions, and remedial actions are discussed
H T =P values, thus increasing flow efficiency and discharge capac- in detail in the companion paper (Crookston and Tullis 2013).
ity. As the discharge and the corresponding momentum of the flow The artificial aeration, vent pipes, and nappe breakers are also
passing over the weir increase, the nappe becomes aerated and the discussed.
streamlines will eventually detach from the weir crest, creating a
similar nappe profile to the quarter-round crest. Once nappe detach-
Labyrinth Design and Analyses
ment occurs, the gains in the half-round crest flow efficiency are
lost relative to the quarter-round crest. All of the Cd-HR =Cd-QR The recommended procedure for designing a labyrinth weir is pre-
curves are anticipated to eventually converge to 1.0 with increasing sented in Table 4, which includes a design example for illustration

366 / JOURNAL OF IRRIGATION AND DRAINAGE ENGINEERING © ASCE / MAY 2013

J. Irrig. Drain Eng. 2013.139:363-370.


Table 4. Recommended Design Procedure for Labyrinth Weirs
Parameter Symbol Value Units Notes
Hydraulic conditions—input data
Design flow Qdesign 1,500.00 (m3 =s) g ¼ 9.81 m=s2
Design flow water surface elevation H 1,680.00 (m)
Approach channel elevation H apron 1,674.00 (m)
Crest elevation H crest 1,678.00 (m)
Unsubmerged total upstream head HT 2.00 (m) Piezometric head + velocity head - losses
Downstream total head Hd 0.50 (m)
Labyrinth weir geometry—input data
Angle of side legs α 12 (°) α ∼ 6°–35°
Number of cycles N 9 — Whole or half cycles
Downloaded from ascelibrary.org by UTAH STATE UNIVERSITY on 04/11/13. Copyright ASCE. For personal use only; all rights reserved.

Cycle width w 11.10 (m)


Crest height P 4.00 (m) P ∼ 1.0H T
Thickness of weir wall at the crest tw 0.50 (m) tw ∼ P=8
Inside apex width A 0.50 (m) A ∼ tw
Crest shape Crest shape Quarter — Quarter- or half-round
Aeration device (nappe breakers, vents) — Breakers — Breakers, vents, or none
Calculated data
Headwater ratio H T =P 0.50 —
Labyrinth weir discharge crest coefficient Cdðα°Þ 0.429 — Cdðα°Þ ¼ fðH T =P; α; Crest ShapeÞ
Total centerline length of weir Lc 418.28 (m) Lc ¼ 3=2Qdesign =½ðCdða° Þ H 3=2
T Þð2gÞ
1=2

Centerline length of sidewall lc 2.33 (m) lc ¼ ðB − tw Þ= cosðαÞ


Outside apex width D 1.30 (m) D ¼ A þ 2tw tanð45 − α=2Þ
Width of labyrinth (normal to flow) W 99.87 (m) W ¼ Nw
Length of apron (parallel to flow) B 22.35 (m) B ¼ ½Lc =ð2NÞ − ðA þ DÞ=2 cos ðαÞ þ tw
Magnification ratio M 4.19 — M ¼ Lc =ðwNÞ
Cycle width ratio w=P 2.77 — Normally 2 ≤ w=P ≤ 4
Relative thickness ratio P=tw 8.0 — ∼8
Apex ratio A=w 0.05 — <0.08
Cycle efficiency ε0 1.80 — ε 0 ¼ Cdðα°Þ M
Efficacy ε 2.23 — ε ¼ Cdðα°Þ M=Cdð90°Þ
Number of nappe breakers or vents — 9 — Breaker on ds apex, 1 vent per sidewall
Linear weir discharge coefficient Cdð90°Þ 0.808 — Cdð90°Þ ¼ fðH T =P; α; Crest ShapeÞ
Length of linear weir for same flow Lcð90°Þ 222.33 (m) Lcð90°Þ ¼ 3=2Qdesign =½ðCdð90°Þ H 3=2
T Þð2gÞ
1=2 

Submergence (Tullis et al. 2007)


Downstream/upstream ratio of unsubmerged head H d =H T 0.25 (m)
Submerged head discharge ratio H  =H T 1.013 — Piecewise function Tullis et al. (2007)
Submerged upstream total head H 2.025 (m)
Submergence level S 0.247 — S ¼ H d =H 
Submerged weir discharge coefficient Cd−sub 0.421 — Cdðα°Þ ðH T =H  Þ3=2

Note: Design is limited to the extent of experimental data; designs that exceed these limits may warrant a physical model study. Table 4 can be modified so that
the weir footprint (B and W) and N are independent variables and α and w are dependent variables.

(input data, computed results). Table 4 can be adjusted for SI or variable using the equations provided in Table 4. For comparison,
Imperial units and includes a “Notes” column that presents appli- Cdð90°Þ and the required linear weir length (with the same crest
cable design guidance for each design variable (formula and geo- shape as the labyrinth weir) are reported, which match the labyrinth
metric limitations.). The top section of the design table includes the weir head-discharge condition. The last section of the design
user-defined hydraulic conditions or requirements for the labyrinth method includes the submerged head-discharge relationships de-
weir. The design flow rate (Qdesign ) typically represents a specific veloped by Tullis et al. (2007).
return-period flood discharge from a hydrologic analysis, HT is Per Figs. 2 and 3, Cdðα°Þ decreases with decreasing α. For a
based on upstream freeboard or flood plain constraints, and H d given weir footprint (W and B held constant); however, the
is determined by a backwater curve based on Qdesign . User-defined labyrinth weir crest length increases with decreasing α. The net
labyrinth weir geometric parameters (e.g., α, P, tw , crest shape, N) effect of the α-dependent opposing influences of Cdðα°Þ and Lc
are entered in the second section of the design table. Also, a place is on discharge efficiency should be considered when optimizing a
provided to specify a nappe aeration device if desired. The third labyrinth weir design. To account for the influences of Cdðα°Þ
section contains the labyrinth weir design geometry, design ratios, and Lc on discharge capacity and to provide a guide in the selection
and hydraulic performance output data (Cdðα°Þ , ε 0 ), which are of α, cycle efficiency [ε 0 ¼ Cdðα°Þ Lc-cycle =w] is introduced,
calculated using the input data from the previous sections. Cells representative of the discharge per cycle (at a specific H T value)
are provided to annotate anticipated nappe aeration behavior, for a given labyrinth weir geometry. Values of ε 0 for the experimen-
determined from information presented in Paper II. If a specific tal labyrinth weir data are presented in Figs. 5 (quarter-round) and 6
labyrinth weir footprint size is required, Table 4 can be adjusted (half-round) as a function of H T =P. The data indicate that the in-
by switching B to an independent variable and N to a dependent crease in weir length more than compensates for the decrease in

JOURNAL OF IRRIGATION AND DRAINAGE ENGINEERING © ASCE / MAY 2013 / 367

J. Irrig. Drain Eng. 2013.139:363-370.


Downloaded from ascelibrary.org by UTAH STATE UNIVERSITY on 04/11/13. Copyright ASCE. For personal use only; all rights reserved.

Fig. 5. Cycle efficiency versus H T =P for quarter-round labyrinth weirs

Cdðα°Þ with smaller α values. Figs. 5 and 6 also show that the maxi-
mum ε 0 values occur at relatively low HT =P (as delineated by the
dashed line), ε 0 increases as α decreases, and the benefit of in-
creased ε 0 of smaller α angles decrease with increasing HT =P.
The value of ε 0 only evaluates the discharge efficiency per cycle.
In choosing a labyrinth weir design, a complete feasibility study
(construction cost, environmental impact, and permitting issues) Fig. 7. Recommended procedure for analyses of labyrinth weirs
is recommended.
Beyond the ability to design a labyrinth weir for a particular
head-discharge condition using Table 4, the ability to determine Cdðα°Þ relationships [Eq. (2)] was chosen because of improved
a complete head-discharge relationship for specific labyrinth weir extrapolation behavior for values of 1.0 ≤ HT =P ≤ 2.0, relative
geometries is also important. Such a procedure, which also adapts to the polynomial functions used by Tullis et al. (1995). Although
easily to a spreadsheet program format, is outlined in Fig. 7. The insufficient data are currently available to validate the predictive
known labyrinth weir geometries are entered. Missing geometric accuracy of all of the α-specific Cdðα°Þ relationships over the range
parameters and labyrinth weir ratios are calculated and a head- of 1.0 ≤ H T =P ≤ 2.0, the α ¼ 15° labyrinth weir model (Model 13
discharge rating curve is produced. The effects of tailwater sub- in Table 1) was tested up to HT =P ¼ 2.0 and good agreement
mergence may be determined by solving for Q or submerged was noted between the experimental and predictive Cdðα°Þ data.
upstream total head, H , by iteration (Tullis et al. 2007). The Using the labyrinth weir Cdðα°Þ relationships for 1.0 ≤ H T =P ≤
dimensionless submerged head relationship, based on Tullis et al. 2.0 conditions is recommended as a first-order approximation of
(2007), is presented in Fig. 8, which includes H and tailwater head-discharge performance. Physical modeling is recommended
submergence levels (S). to validate high-head labyrinth weir hydraulic performance.
The design method and support data are limited to the geom- Crookston and Tullis (2012a, b, c, 2013) and Crookston et al.
etries (Table 1) and hydraulic conditions tested in this study (2012) provide additional design information intended to compli-
(e.g., 0.05 ≤ H T =P ≤ 0.9). These results may be conservatively ment the design method presented herein.
applied (with sound engineering judgment) to other labyrinth
weir geometries and flow conditions (design verification with a
hydraulic model study is recommended). Linear interpolation is Data Verification
recommended to determine Cdðα°Þ for α values other than those An uncertainty analysis was performed, as outlined by Kline and
presented. As previously discussed, the form of the predictive McClintock (1953) for single-point experimental data. The percent

Fig. 8. Dimensionless submerged head relationship for labyrinth weirs


Fig. 6. Cycle efficiency versus HT =P for half-round labyrinth weirs (data from Tullis et al. 2007)

368 / JOURNAL OF IRRIGATION AND DRAINAGE ENGINEERING © ASCE / MAY 2013

J. Irrig. Drain Eng. 2013.139:363-370.


uncertainty for each Cdðα°Þ was computed by determining individ- do not agree with the data from the current study. The α ¼ 8° data
ual parameter uncertainties (e.g., measurement accuracy of Q, Lc , from Tullis et al. (1995) is incorrect; it has been replaced in Fig. 10
h, and V) and taking partial derivatives of Eq. (1), following the with the data from Willmore (2004) α ¼ 8° quarter round, which
methodology prescribed by Kline and McClintock. The average shows good agreement with the data from the present study.
single sample uncertainties for each data set were 1–1.5%. Uncer-
tainties were largest (∼4% for HT =P ≤ 0.075) for very small
values of Q and h (instrument accuracy) and smallest (≤0.9% for Summary and Conclusions
H T =P ≥ 0.3) for large values of Q and h.
The Cdðα°Þ data from the current study were also compared with A labyrinth weir design and analysis procedure is presented
the results of previous studies. The data in Fig. 9 show relatively (Table 4) based upon the results of physical modeling in a labora-
good agreement between the nonvented, half-round labyrinth weir tory flume. The discharge is calculated based on the traditional
Cdðα°Þ data and the data from Willmore (2004), with the largest dis- weir equation [Eq. (1)], utilizing H T as the representative upstream
crepancy appearing for large α labyrinth weirs at HT =P ≤ 0.2. head and Lc as the characteristic weir length. Tailwater submer-
Downloaded from ascelibrary.org by UTAH STATE UNIVERSITY on 04/11/13. Copyright ASCE. For personal use only; all rights reserved.

Fig. 10 compares the quarter-round, nonvented labyrinth weir gence for labyrinth weirs is included, as presented by Tullis et al.
Cdðα°Þ data of this study with Cdðα°Þ data from Tullis et al. (2007). The proposed design and analysis method was validated by
(1995) (presented in terms of Lc rather than Le ). There appears juxtaposing the experimental Cdðα°Þ data of this study with other
to be relatively good agreement at large values of HT =P; however, physical model studies presented in Figs. 9 and 10. Paper II is
large differences are present for HT =P ≤ 0.4. This may be attrib- devoted to nappe behavior and discusses aeration conditions, insta-
uted to the smaller sized labyrinth weir models used by the Tullis bility or flow surging, artificial aeration and structure placement,
et al. (1995) method (potential size-scale effects and different val- and nappe vibration.
ues of P), a potentially higher level of data uncertainty (e.g., their Figs. 2 and 3 present a dimensionless discharge coefficient,
flow measurement technique), and the fact that B was held constant Cdðα°Þ , as a function of HT =P for quarter-round and half-round
(not N), which in some cases resulted in partial cycles. The α ¼ 25° labyrinth weirs (6° ≤ α ≤ 35°) and for linear weirs. The test results
and α ¼ 35° Cdðα°Þ curves linearly interpolated by Tullis et al. indicate that the increase in efficiency provided by a half-round
(1995) (based on α ¼ 18° and α ¼ 90° experimental data) also crest shape (relative to a quarter-round crest) is more significant
for HT =P ≤ 0.4 (Fig. 4).
Cycle efficiency, ε 0 , represents the labyrinth weir cycle-specific
discharge (for a given HT condition) and is a tool for examining
the discharge capacity of different labyrinth weir geometries
(Figs. 5 and 6). An ε 0 analysis found that for a fixed cycle width
(w), the cycle efficiency increased with decreasing values of α be-
cause the increase in cycle crest length, Lc-cycle (with decreasing α),
more than compensates for decreasing Cdðα°Þ values.
Although the methods and tools presented in this study are
recommended for use in the design and analyze of labyrinth
spillways, a physical model study is recommended to verify hy-
draulic performance. A model study would include site-specific
conditions that may be outside the scope of this study and may
provide valuable insights into the performance and operation of
the labyrinth weir.
Additional components of this study not presented here include
arced labyrinth weirs (Crookston and Tullis 2012a), reservoir-
Fig. 9. Comparison between Cd values obtained by Willmore (2004) specific labyrinth weir applications (Crookston and Tullis 2012b),
and Crookston (2010) for nonvented, half-round labyrinth weirs local submergence and nappe interference (Crookston and Tullis
2012c), scale effects, 1 ≤ H T =P ≤ 2 (Crookston et al. 2012),
and the influence of the cycle width ratio and labyrinth apexes
on discharge (2=3 ≤ w=P ≤ 4).

Acknowledgments
This study was funded by the State of Utah and the Utah Water
Research Laboratory at Utah State University.

Notation
The following symbols are used in this paper:
A = inside apex width;
A=w = apex ratio;
a = curve-fit coefficient for quarter-round and half-round
labyrinth and linear weirs (Tables 2 and 3);
Fig. 10. Comparison between proposed Cd design curves by Tullis
B = length of labyrinth weir (apron) in flow direction;
et al. (1995), Willmore (2004), and Crookston (2010) for nonvented,
b = curve-fit coefficient for quarter-round and half-round
quarter-round labyrinth weirs (based upon Lc )
labyrinth and linear weirs (Tables 2 and 3);

JOURNAL OF IRRIGATION AND DRAINAGE ENGINEERING © ASCE / MAY 2013 / 369

J. Irrig. Drain Eng. 2013.139:363-370.


Cd = discharge coefficient, data from current study; Crookston, B. M., and Tullis, B. P. (2012b). “Discharge efficiency of
Cdðα°Þ = discharge coefficient for labyrinth weir of sidewall reservoir-application-specific labyrinth weirs.” J. Irrig. Drain. Eng.,
angle α; 138(6), 773–776.
Cdð90°Þ = discharge coefficient for linear weir; Crookston, B. M., and Tullis, B. P. (2012c). “Labyrinth weirs: Nappe
interference and local submergence.” J. Irrig. Drain. Eng., 138(8),
Cd-sub = discharge coefficient for a labyrinth weir with tailwater
757–765.
submergence; Crookston, B. M., and Tullis, B. P. (2013). “Hydraulic design and analysis
c = curve-fit coefficient for quarter-round and half-round of labyrinth weirs. II: Nappe aeration, instability, and vibration.”
labyrinth and linear weirs (Tables 2 and 3); J. Irrig. Drain. Eng., 139(5), 371–377.
D = outside apex width; Darvas, L. (1971). “Discussion of ‘Performance and design of labyrinth
d = curve-fit coefficient for quarter-round and half-round weirs’ by Hay and Taylor.” J. Hydraul. Eng., 97(80), 1246–1251.
labyrinth and linear weirs (Tables 2 and 3); Falvey, H. (2003). Hydraulic design of labyrinth weirs, ASCE, Reston,
g = acceleration constant of gravity; VA.
H = design flow water surface elevation; Hay, N., and Taylor, G. (1970). “Performance and design of labyrinth
Downloaded from ascelibrary.org by UTAH STATE UNIVERSITY on 04/11/13. Copyright ASCE. For personal use only; all rights reserved.

Hapron = approach channel elevation; weirs.” J. Hydraul. Eng., 96(11), 2337–2357.


Hinchliff, D., and Houston, K. (1984). “Hydraulic design and application of
Hcrest = elevation of labyrinth weir crest;
labyrinth spillways.” Proc., 4th Annual USCOLD Lecture, Dam Safety
Hd = total downstream head measured relative to the weir and Rehabilitation, Bureau of Reclamation, U.S. Dept. of the Interior,
crest; Washington, DC.
H d =HT = downstream/upstream ratio of total unsubmerged head; Houston, K. (1982). “Hydraulic model study of Ute Dam labyrinth
HT = total upstream head (unsubmerged) measured relative to spillway.” Rep. No. GR-82-7, U.S. Bureau of Reclamation, Denver.
the weir crest; Johnson, M. (1996). “Discharge coefficient scale effects analysis for weirs.”
HT =P = headwater ratio; Ph.D. dissertation, Utah State Univ., Logan, UT.
H = submerged total upstream head measured relative to the Kline, S., and McClintock, F. (1953). “Describing uncertainties in single-
weir crest; sample experiments.” Am. Soc. Mech. Eng., 75(1), 3–8.
H =HT = submerged head discharge ratio; Lopes, R., Matos, J., and Melo, J. (2009). “Discharge capacity for free-flow
and submerged labyrinth weirs.” Proc., 33rd IAHR Congress: Water
h = depth of flow over the weir crest;
Engineering for a Sustainable Environment, IAHR, Madrid, Spain,
Lc = total centerline length of labyrinth weir; 1054–1061.
Lc-cycle = centerline length for a single labyrinth weir cycle; Lux, F., III (1984). “Discharge characteristics of labyrinth weirs.” Proc.,
Le = total effective length of labyrinth weir; Conf. on Water for Resources Development, ASCE, Reston, VA.
lc = centerline length of weir sidewall; Lux, F., III (1989). “Design and application of labyrinth weirs.” Design
M = magnification ratio, Lc-cycle =w; of hydraulic structures 89, M. Alberson and R. Kia, eds., Balkema,
N = number of labyrinth weir cycles; Rotterdam, Netherlands, 205–215.
P = weir height; Lux, F., III, and Hinchliff, D. (1985). “Design and construction of labyrinth
P=tw = relative thickness ratio; spillways.” 15th Congress ICOLD, Vol. IV, Q59-R15, ICOLD, Paris,
Q = discharge over weir; 249–274.
Magalhães, A., and Lorena, M. (1989). “Hydraulic design of labyrinth
Qdesign = design discharge over labyrinth weir;
weirs.” Rep. No. 736, National Laboratory of Civil Engineering, Lisbon,
Rcrest = radius of rounded crest (e.g., Rcrest ¼ tw =2); Portugal.
S = submergence level; Melo, J., Ramos, C., and Magalhães, A. (2002). “Descarregadores com
tw = thickness of weir wall; soleira em labirinto de um ciclo em canais convergentes. Determinação
V = average cross-sectional flow velocity upstream of weir; da capacidad de vazão.” Proc., 6th Congresso da Água, CD-ROM,
W = width of channel; Associação Portuguesa dos Recursos Hídricos, Lisboa, Portugal,
w = width of a single labyrinth weir cycle; (in Portuguese).
w=P = cycle width ratio; Taylor, G. (1968). “The performance of labyrinth weirs.” Ph.D. thesis,
α = sidewall angle; Univ. of Nottingham, Nottingham, England.
ε = efficacy; and Tullis, B., and Crookston, B. M. (2008). “Lake Townsend Dam spillway
hydraulic model study report.” Utah Water Research Laboratory,
ε 0 = cycle efficiency.
Logan, UT.
Tullis, B., Young, J., and Chandler, M. (2007). “Head-discharge relation-
ships for submerged labyrinth weirs.” J. Hydrol. Eng., 133(3), 248–254.
References Tullis, P. (1993). “Standley Lake service spillway model study.” Hydraulic
Rep. No. 341, Utah Water Research Laboratory, Logan, UT.
Amanian, N. (1987). “Performance and design of labyrinth spillways.” Tullis, P., Amanian, N., and Waldron, D. (1995). “Design of labyrinth weir
M.S. thesis, Utah State Univ., Logan, UT. spillways.” J. Hydrol. Eng., 121(3), 247–255.
Crookston, B. M. (2010) “Labyrinth weirs.” Ph.D. dissertation, Utah State Vasquez, V., Boyd, M., Wolfhope, J., and Garret, R. (2007). “A labyrinth
Univ., Logan, UT. rises in the heart of Texas.” Proc., 28th Annual USSD Conf., USSD,
Crookston, B. M., Paxson, G. S., and Savage, B. M. (2012). “It can be Denver, CO, 813–826.
done! Labyrinth weir design guidance for high headwater and low cycle Waldron, D. (1994). “Design of labyrinth spillways.” M.S. thesis, Utah
width ratios.” Proc., Dam Safety 2012, ASDSO, Denver, CO. State Univ., Logan, UT.
Crookston, B. M., and Tullis, B. P. (2012a). “Arced labyrinth weirs.” Willmore, C. (2004). “Hydraulic characteristics of labyrinth weirs.”
J. Hydraul. Eng., 138(6), 555–562. M.S. Rep., Utah State Univ., Logan, UT.

370 / JOURNAL OF IRRIGATION AND DRAINAGE ENGINEERING © ASCE / MAY 2013

View publication stats J. Irrig. Drain Eng. 2013.139:363-370.

You might also like