+ All Categories
Home > Documents > An experimental and numerical study on hydraulic characteristics...

An experimental and numerical study on hydraulic characteristics...

Date post: 10-Apr-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
8
An experimental and numerical study on hydraulic characteristics and theoretical equations of circular weirs Benyamin Naghavi, Kazem Esmaili, Jafar Yazdi, and Fatemeh Koorosh Vahid Abstract: Due to streamline curvature, non-hydrostatic pressure distribution, and nappe adherence on the weir wall, the dis- charge coefficient of weirs is less than 1.0. However, in cylindrical weirs, the discharge coefficient extends to values more than 1.0 due to the different pressures and velocity distributions. In this study, pressure and velocity distributions of different circular weirs are simulated by computational fluid dynamics (CFD) and Fluent software and compared with experimental results. The numerical results of CFD show a significant correlation with measured data from manometers and a laser Dop- pler velocimeter (LDV). Experimental observations show that the critical flow depth forms before and after the crest for greater and smaller discharge coefficients, respectively. Nappe separation depends on overflow discharge and will shift to the downstream face of the cylinder in high discharges. To recognize the location of critical flow and nappe separation, ana- lytical formulations are proposed depending on weir size and inflow conditions. The results show good agreement between the analytical predictions and experimental observations. Key words: laboratory model, CFD, critical depth location. Résumé : En raison de la courbe dans la ligne découlement, de la distribution non hydrostatique de la pression et de lad- hérence de la lame déversante sur le mur du déversoir, le coefficient de débit des déversoirs est inférieur à 1,0. Toutefois, dans les déversoirs cylindriques, le coefficient de débit atteint des valeurs supérieures à 1,0 en raison des pressions et des ré- partitions des vitesses différentes. La présente étude simule les répartitions des pressions et des vitesses de différents déver- soirs circulaires en utilisant Fluent, un logiciel de dynamique numérique des fluides, et les compare aux résultats expérimentaux. Les résultats numériques de la dynamique numérique des fluides indiquent une corrélation importante avec les données mesurées à partir de manomètres et dun vélocimètre Doppler à laser. Les constatations des expériences mon- trent que la plus profondeur critique découlement se forme respectivement avant et après la crête pour des profondeurs cri- tique respectivement supérieurs et inférieurs. La séparation de la lame deau dépend de la surverse et se déplacera vers la face aval du cylindre lors de grands débits. Afin de reconnaître lemplacement de lécoulement critique et de la séparation de la lame deau, des formules analytiques sont proposées selon la dimension du déversoir et les conditions du débit entrant. Les résultats montrent une bonne corrélation entre les prévisions analytiques et les observations expérimentales. Motsclés : modèle de laboratoire, dynamique numérique des fluides, emplacement de la profondeur critique. [Traduit par la Rédaction] 1. Introduction Circular-crested or cylindrical weirs are one of the oldest hydraulic structures for discharge measurement. The advan- tages of the circular weirs include their stable overflow pat- tern compared to sharp-crested weirs, their ease in passing floating debris, their simplicity of design compared to the ogee crest design, and their associated lower cost (Chanson and Montes 1997). These advantages lead to abundant appli- cation of circular crested weirs for water level control in irri- gation systems (Heidarpour et al. 2008). Much research on circular weirs, including research by Verwoerd (1941), Sar- ginson (1972), Escande and Sananes (1959), Cassidy (1965), Hager (1985, 1993, 1995), Bos (1989), Ramamurthy et al. (1992), Ramamurthy and Vo (1993a, 1993b), Chanson and Montes (1998), Chanson (2006), and Heidarpour and Chamani (2006), show that the discharge coefficient, C d , is primarily a function of upstream head to crest radius ratio, and is close to and usually greater than 1.0. Studies on nappe suction and nappe ventilation show the nappe suction prevents flow separation, which subsequently raises coeffi- cients by 15% to 20% (Escande and Sananes 1959). As water passes over the dam crest, the nappe free-surface re- mains smooth and clear, and the falling nappe adheres to the weir face. This phenomenon (i.e., the two-dimensional nappe flowing past the convex curved wall and adhering to the wall) is called a Coanda effect. The resulting Coanda force, induced by convex curvature of the wall and suction Received 2 February 2011. Revision accepted 17 September 2011. Published at www.nrcresearchpress.com/cjce on XX November 2011. B. Naghavi, K. Esmaili, and F.K. Vahid. Department of Water Engineering, Ferdowsi University of Mashhad, Iran. J. Yazdi. Water Research Institute, Tehran, Iran. Corresponding author: Benyamin Naghavi (e-mail: [email protected]). Written discussion of this article is welcomed and will be received by the Editor until 30 April 2012. 1327 Can. J. Civ. Eng. 38: 13271334 (2011) doi:10.1139/L11-092 Published by NRC Research Press Can. J. Civ. Eng. Downloaded from www.nrcresearchpress.com by University of Oklahoma Libraries on 11/29/11 For personal use only.
Transcript
Page 1: An experimental and numerical study on hydraulic characteristics …profdoc.um.ac.ir/articles/a/1024597.pdf · 2020-03-07 · flow, which induces the nappe adherence (Chanson and

An experimental and numerical study on hydrauliccharacteristics and theoretical equations ofcircular weirs

Benyamin Naghavi, Kazem Esmaili, Jafar Yazdi, and Fatemeh Koorosh Vahid

Abstract: Due to streamline curvature, non-hydrostatic pressure distribution, and nappe adherence on the weir wall, the dis-charge coefficient of weirs is less than 1.0. However, in cylindrical weirs, the discharge coefficient extends to values morethan 1.0 due to the different pressures and velocity distributions. In this study, pressure and velocity distributions of differentcircular weirs are simulated by computational fluid dynamics (CFD) and Fluent software and compared with experimentalresults. The numerical results of CFD show a significant correlation with measured data from manometers and a laser Dop-pler velocimeter (LDV). Experimental observations show that the critical flow depth forms before and after the crest forgreater and smaller discharge coefficients, respectively. Nappe separation depends on overflow discharge and will shift tothe downstream face of the cylinder in high discharges. To recognize the location of critical flow and nappe separation, ana-lytical formulations are proposed depending on weir size and inflow conditions. The results show good agreement betweenthe analytical predictions and experimental observations.

Key words: laboratory model, CFD, critical depth location.

Résumé : En raison de la courbe dans la ligne d’écoulement, de la distribution non hydrostatique de la pression et de l’ad-hérence de la lame déversante sur le mur du déversoir, le coefficient de débit des déversoirs est inférieur à 1,0. Toutefois,dans les déversoirs cylindriques, le coefficient de débit atteint des valeurs supérieures à 1,0 en raison des pressions et des ré-partitions des vitesses différentes. La présente étude simule les répartitions des pressions et des vitesses de différents déver-soirs circulaires en utilisant Fluent, un logiciel de dynamique numérique des fluides, et les compare aux résultatsexpérimentaux. Les résultats numériques de la dynamique numérique des fluides indiquent une corrélation importante avecles données mesurées à partir de manomètres et d’un vélocimètre Doppler à laser. Les constatations des expériences mon-trent que la plus profondeur critique d’écoulement se forme respectivement avant et après la crête pour des profondeurs cri-tique respectivement supérieurs et inférieurs. La séparation de la lame d’eau dépend de la surverse et se déplacera vers laface aval du cylindre lors de grands débits. Afin de reconnaître l’emplacement de l’écoulement critique et de la séparationde la lame d’eau, des formules analytiques sont proposées selon la dimension du déversoir et les conditions du débit entrant.Les résultats montrent une bonne corrélation entre les prévisions analytiques et les observations expérimentales.

Mots‐clés : modèle de laboratoire, dynamique numérique des fluides, emplacement de la profondeur critique.

[Traduit par la Rédaction]

1. Introduction

Circular-crested or cylindrical weirs are one of the oldesthydraulic structures for discharge measurement. The advan-tages of the circular weirs include their stable overflow pat-tern compared to sharp-crested weirs, their ease in passingfloating debris, their simplicity of design compared to theogee crest design, and their associated lower cost (Chansonand Montes 1997). These advantages lead to abundant appli-cation of circular crested weirs for water level control in irri-gation systems (Heidarpour et al. 2008). Much research oncircular weirs, including research by Verwoerd (1941), Sar-ginson (1972), Escande and Sananes (1959), Cassidy(1965), Hager (1985, 1993, 1995), Bos (1989), Ramamurthy

et al. (1992), Ramamurthy and Vo (1993a, 1993b), Chansonand Montes (1998), Chanson (2006), and Heidarpour andChamani (2006), show that the discharge coefficient, Cd, isprimarily a function of upstream head to crest radius ratio,and is close to and usually greater than 1.0. Studies onnappe suction and nappe ventilation show the nappe suctionprevents flow separation, which subsequently raises coeffi-cients by 15% to 20% (Escande and Sananes 1959). Aswater passes over the dam crest, the nappe free-surface re-mains smooth and clear, and the falling nappe adheres tothe weir face. This phenomenon (i.e., the two-dimensionalnappe flowing past the convex curved wall and adhering tothe wall) is called a Coanda effect. The resulting Coandaforce, induced by convex curvature of the wall and suction

Received 2 February 2011. Revision accepted 17 September 2011. Published at www.nrcresearchpress.com/cjce on XX November 2011.

B. Naghavi, K. Esmaili, and F.K. Vahid. Department of Water Engineering, Ferdowsi University of Mashhad, Iran.J. Yazdi. Water Research Institute, Tehran, Iran.

Corresponding author: Benyamin Naghavi (e-mail: [email protected]).

Written discussion of this article is welcomed and will be received by the Editor until 30 April 2012.

1327

Can. J. Civ. Eng. 38: 1327–1334 (2011) doi:10.1139/L11-092 Published by NRC Research Press

Can

. J. C

iv. E

ng. D

ownl

oade

d fr

om w

ww

.nrc

rese

arch

pres

s.co

m b

y U

nive

rsity

of

Okl

ahom

a L

ibra

ries

on

11/2

9/11

For

pers

onal

use

onl

y.

Page 2: An experimental and numerical study on hydraulic characteristics …profdoc.um.ac.ir/articles/a/1024597.pdf · 2020-03-07 · flow, which induces the nappe adherence (Chanson and

pressure, acts on the wall surface in the same direction as theflow, which induces the nappe adherence (Chanson andMontes 1997). Suction pressure and the resulting nappe ad-herence cause the streamlines to become more curved andthe flow velocity to increase. Therefore, the discharge coeffi-cient of a circular weir is greater than that of sharp andbroad-crested weirs. To predict the wall suction pressure,Chanson (1996) proposed an analytical method in which apressure correction coefficient accounts for the non-hydro-static pressure distribution on the crest. The literature reviewshows that the pressure distribution on a cylindrical weir dif-fers from conventional weirs, and by increasing the upstreamhead to crest radius ratio to more than 0.8, the pressure dis-tribution on a cylindrical weir changes to non-hydrostatic(Bos 1989; Ramamurthy and Vo 1993a; Heidarpour et al.2008). The dimensionless velocity distribution is only afunction of upstream flow to the radius of a circular weir,while the maximum crest velocity, the normalized crest pres-sure, and the pressure correction coefficient are functions ofthe average upstream velocity and a velocity correction fac-tor (Heidarpour et al. 2008). Ramamurthy and Vo (1993b)used the expression of Dressler (1978) for the velocity pro-file of a shallow irrotational flow over a curved bed to pre-dict the velocity distribution over a cylindrical weir.Heidarpour and Chamani (2006) present a method to predictthe velocity distribution over the crest of the cylindrical weiras well as the discharge coefficient of the cylindrical weir,using potential flow past a cylindrical obstacle. Assumingan inviscid flow, the maximum velocity occurs at the edgeof the thin boundary layer and closer to the crest (Heidar-pour et al. 2008).In this research, computational fluid dynamics (CFD) is

used to simulate the water surface profiles and the pressureand velocity distributions on the cylindrical weirs. To vali-date the fluid velocity field predicted by the CFD software,a laser Doppler velocimeter (LDV) is used. Using an LDV,which does not interfere with the flow, enables us to obtainlocalized flow measurements and record a wide range of ve-locities. The CFD model’s results respecting the pressure dis-tribution are compared with experimental data from themanometers and a new interpretation of pressure distributionon cylindrical weirs is obtained. Finally, methods for findingthe location of critical depth and nappe separation are given,and some analytical equations are introduced.

2. Experimental worksThe experiments were performed on a rectangular flume of

10 m in length, 0.3 m in width, and 0.5 m in depth. Theoverflow characteristics of cylindrical weirs were investigatedfor several configurations, i.e., six cylinder sizes with radiusof 0.03 < R < 0.08 m, in supported and non-supported con-ditions. The height of weir supports were in a range of0.025 < P < 0.1 m. All cylinders were made of PVC pipes.In each run, and for all different discharges, the water surfaceprofiles and the pressure and velocity distributions weremeasured. In all runs, the longitudinal slope of the flumewas constant and equal to 0.005. To simulate the flow in theflume, a flow circulation system was set up, and dischargeswere measured with a prior calibrated rectangular sharpcrested weir (Hosseini 2003).

To investigate the pressure distribution on the weir, eightmanometers were installed on the weir wall at multiples of30 degrees from 0° to 210° (Fig. 1). The water surface pro-files were measured by a point gauge in 3 mm distances overthe weirs from upstream to downstream. The experimental er-ror due to the apparatuses, i.e., point gauge and manometers,were ±0.1 mm and ±0.1 cm, respectively.To measure the horizontal and vertical components of ve-

locity, a 2D LDV was used.

3. Numerical modelFluent was used to simulate weir flow. To measure the free

surface profile, the Hirt-Nichols’ VOF (volume of fluid) andYoungs’ method are used (Hirt and Nichols 1981; Youngs1982; Tadayon and Ramamurthy 2009). In modeling the freesurface between water and air, a transport equation can beconsidered for the water phase:

½1� @

@tðawÞ þ rðawvÞ ¼ 0

where aw is the volume fraction of water and v is flow velo-city. In this case, as the volume fraction for other phases canbe inferred from the constraint, the transport equation shouldonly be solved:

½2� aa ¼ 1� aw

This equation is solved in the entire domain and volumefraction computed for all cells. For each cell, aw = 1 when itis full of water and aw = 0 when it is full of air. For cellsthat span the interface between air and water, aw is between0 and 1. The fluid properties in each cell are determined ac-cording to the local volume fraction. For example, the densityin each cell is as follows:

½3� r ¼ awrw þ ð1� awÞrawhere rw and ra are water and air densities, respectively.In this research, both structured and unstructured meshes

were used. Because of the complex geometry of circularweirs, the area around the weir was meshed in unstructuredform, and elsewhere the structured mesh was used. To definewater and air flow into the domain, two different inletsshould be used as boundary conditions for upstream flow.To estimate the wall’s effect on flow, empirical wall func-tions known as standard wall functions were used. For strongcurvatures like spillways, applying the RNG k–3 model ismore appropriate (Tadayon and Ramamurthy 2009). Thusthe RNG k–3 turbulence model was used with standard wall

Fig. 1. Manometers joint on weir at angles 0° to 210°.

1328 Can. J. Civ. Eng. Vol. 38, 2011

Published by NRC Research Press

Can

. J. C

iv. E

ng. D

ownl

oade

d fr

om w

ww

.nrc

rese

arch

pres

s.co

m b

y U

nive

rsity

of

Okl

ahom

a L

ibra

ries

on

11/2

9/11

For

pers

onal

use

onl

y.

Page 3: An experimental and numerical study on hydraulic characteristics …profdoc.um.ac.ir/articles/a/1024597.pdf · 2020-03-07 · flow, which induces the nappe adherence (Chanson and

functions (Yakhot and Orszag 1986, Yakhot and Smith1992). Accordingly, for the upper boundary above the airphase, symmetry conditions were considered, which enforcea zero normal velocity and a zero shear stress. For down-stream flow, gradients of all variables were assumed to bezero, and since the outflow of the flume was assumed to bea jet, pressure was considered to be zero, i.e., atmosphericpressure.To complete the CFD modeling, the PRESTO pressure

discretization, which showed the best convergence, was used.The flow was simulated to be unsteady. Thus, the pressure–velocity coupling scheme was achieved using the pressure-implicit with splitting of operators (PISO) algorithm (Issa1986).Calculating an unsteady free surface profile requires small

initial time steps. Therefore, a time step ranging from 0.01 to0.001 was selected. During calculations and simulations, theconvergence of the solution and the water surface profileswas monitored. The convergence occurred whenever the nor-malized residuals of each variable were equal to or less than0.001. In this paper the free surface is defined by VOF = 0.5,which is a common value for this purpose (Yazdi et al.2010).Computations were conducted by a Pentium IV system

with an AMD 3800+ processor. Each run of the numericalmodel took 1 to 2 h, depending on model dimensions, meshnumbers, and flow characteristics.

4. ResultsExperimental results related to water surface profiles, pres-

sure distributions, and streamwise velocity distributions forflow over the circular crested weir were used to validate thenumerical simulation predictions. For water surface profilesover the crest, Fig. 2 shows that there is a good agreementbetween CFD predictions and experimental results, and themaximum difference is about 2.5%, which is mainly due tomeasurement inaccuracy in the experimental model and smallvariations in the numerical model. The CFD model indicatedthat the streamlines near the weir surface adhere to the weirwall, and thus the water surface profile follows the curvatureshape of the weir. In the experimental model, the water sur-face fluctuations are negligible before the weir crest, and thewater surface slope is roughly constant. However, after thecrest, depending on discharge values and water adherence tothe weir surface, the water surface slope changes.Both experimental and predicted pressure head distribu-

tions over the circular weir are shown in Fig. 3. For all dis-charge rates, the pressure head decreases with increasingangle. However, the curves corresponding to all dischargerates cross at an angle of 70°. That means that the pressurehead at an angle of 70° is independent of discharge rate. Be-fore this region, higher discharges show higher pressureheads, and after that, the reverse is true, so that the lowestdischarges show the highest pressure heads. Like the physicalmodel, the numerical model shows that the lowest pressureoccurs at point 120° for the highest discharge, and as the dis-charges decrease, the lowest pressure point shifts towards thedownstream flow. The minimum pressure occurs in a regionafter the crest, where the velocity head is at maximum. Inthis region, because of weir curvature, the pressure drops

and the sub-atmospheric pressure increases the velocity. Thedifferences between the experimental and numeical pressuredistributions are due to severe pressure gradients on the cir-cular face. These pressure gradients cannot be predicted pre-cisely by the K–3 turbulence model using standard wallfunctions (Tadayon and Ramamurthy 2009; Yazdi et al.2010). Therefore, a mean error of about 10% is seen in theresults.The measured and predicted fluid velocity field is plotted

in a contour form shown in Figs. 4 and 5, for horizontal andvertical components, respectively. In this case, the averagevelocity of water is approximately 0.65 m/s. According toFig. 4, a good agreement between the CFD simulation resultsand the LDV data can be seen for the horizontal fluid veloc-ity (0 < Vx < 1.5). The main differences are in high veloc-ities near the crest, where the CFD simulation is predictedhigher than LDV data. In all points close to the weir surfaceVx = 0; in upper layers and towards the downstream flow, Vxis gradually increased, and in the region between 90° and120°, it reaches maximum value. After this region, a decreas-ing trend is formed.For the vertical fluid velocity (Vy), the simulation also pro-

vides accurate predictions in most areas, especially in the up-stream flow. Vy lies in the range between 0 and 0.8; on theweir wall Vy = 0, and in upper layers it is increased. More-over, on the weir crest and in all upper layers of this region,Vy = 0, while upstream and downstream of it, Vy extends tomaximum values (the region between 30° and 60° and also120° and 150°, symmetrically). However, the upstream flowis the inverse of the downstream; the vertical component ofthe velocity vector is in a positive direction.

Fig. 2. Comparison between experimental (Exp) and numerical(Num) water surface profiles in different discharges.

Fig. 3. Comparison between experimental (Exp) and numerical(Num) pressure head in different discharges.

Naghavi et al. 1329

Published by NRC Research Press

Can

. J. C

iv. E

ng. D

ownl

oade

d fr

om w

ww

.nrc

rese

arch

pres

s.co

m b

y U

nive

rsity

of

Okl

ahom

a L

ibra

ries

on

11/2

9/11

For

pers

onal

use

onl

y.

Page 4: An experimental and numerical study on hydraulic characteristics …profdoc.um.ac.ir/articles/a/1024597.pdf · 2020-03-07 · flow, which induces the nappe adherence (Chanson and

5. Theoretical equations in cylindrical weirs

5.1. Critical depth locationThe critical flow principle is a useful approach for the hy-

draulic analysis of round-crested weirs due to their singlehead discharge relationships (Castro-Orgaz 2008). At the crit-ical point, the effects of the streamline curvature and slopeare significant because the velocity distribution is non-uni-form and the pressure distribution is non-hydrostatic. Theseeffects must be taken into account to obtain accurate results(Hager 1985; Chanson 2006; Castro-Orgaz et al. 2008a).To calculate the critical flow over a round crested weir,

different models are proposed. Castro-Orgaz et al. (2008b)

generalized the Jaeger theory, assuming a linear law for theradius of curvature of the streamlines along a normal depth,and developed some potential flow equations to measure thecritical flow depth. He imposed a minimum specific energyhead for a given discharge to generalize the critical flow con-dition for curved channel flows, and proposed a model basedon critical flow conditions developed by Hager (1985) to pre-dict the critical depth at the weir crest. According to Castro-Orgaz (2008), in circular crested weirs, the flow is critical atthe crest (Fr = 1) for E/R around 0.5–0.6, where E is the spe-cific energy. For larger heads, the flow at the weir crest issupercritical, and for smaller heads, the flow at the crest issubcritical. Chanson (2006) used the continuity and Bernoulli

Fig. 4. Horizontal fluid velocity, (a) CFD result vs. (b) LDV data.

1330 Can. J. Civ. Eng. Vol. 38, 2011

Published by NRC Research Press

Can

. J. C

iv. E

ng. D

ownl

oade

d fr

om w

ww

.nrc

rese

arch

pres

s.co

m b

y U

nive

rsity

of

Okl

ahom

a L

ibra

ries

on

11/2

9/11

For

pers

onal

use

onl

y.

Page 5: An experimental and numerical study on hydraulic characteristics …profdoc.um.ac.ir/articles/a/1024597.pdf · 2020-03-07 · flow, which induces the nappe adherence (Chanson and

equations in terms of the critical flow depth and depth-aver-aged velocity, and proposed a new analytical expression ofthe critical flow depth for ideal-fluid flows. He implied thatdischarge coefficients larger than 1.0 may be obtained onlywhen the crest pressure distribution is less than hydrostatic.Using the proposed methods for calculating the critical

depth in curved flow, a new method is introduced herein toidentify the location of critical depth in a circular crestedweir. To determine the location of critical depth, a geometricrelation is proposed herein. In Fig. 6, the lines CE and HIrepresent the water depths at the crest (dcrest) and the criticaldepth (dc), respectively.

According to water surface profiles, the water surfaceslopes around points 60° and 90° are the same and thus

½4� tana ¼ AB

R¼ d

Rcos q) d ¼ AB� cos q

where AB = Hw – dcrest, Hw is the flow depth upstream, and dis as follows:

½5� d ¼ ðdc � dcrestÞ � Rð1� sin qÞConsidering the equality of eq. [4] with eq. [5] and using

(dc – dcrest) = U, the following equation can be obtained:

Fig. 5. Vertical fluid velocity, (a) CFD result vs. (b) LDV data.

Naghavi et al. 1331

Published by NRC Research Press

Can

. J. C

iv. E

ng. D

ownl

oade

d fr

om w

ww

.nrc

rese

arch

pres

s.co

m b

y U

nive

rsity

of

Okl

ahom

a L

ibra

ries

on

11/2

9/11

For

pers

onal

use

onl

y.

Page 6: An experimental and numerical study on hydraulic characteristics …profdoc.um.ac.ir/articles/a/1024597.pdf · 2020-03-07 · flow, which induces the nappe adherence (Chanson and

½6� U � R ¼ AB� cos q � R� sin q

According to trigonometric functions and applying x = tan(q/2):

½7� sin q ¼ 2x

1þ x2; cos q ¼ 1� x2

1þ x2

Accordingly, eq. [6] can change to a nonlinear equation.The resulting equation for calculating the critical depth loca-tion is as follows:

½8� q ¼ 2� tan�1 �RþffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiðABÞ2 � U2 þ 2UR

pABþ U � R

!

Figure 7 shows the variation of critical depth location q

with discharge coefficient Cd. In this figure, two differenttrends are distinguishable, in regions before and after thecrest, respectively. It is seen that the discharge coefficient isalways more than 1.0. Higher values of Cd form before thecrest (1.4 < Cd <1.6) and the highest ones are closest to thecrest. After the crest, again higher values of Cd are closer tothe crest, and these values decrease downstream of the crest(1.2 < Cd <1.4). Overall, a L trend forms for Cd variations

over the cylindrical weirs in which the highest point is near-est to the crest.Moreover, the variation of Cd with the ratio of dcrest to dc

reported in Fig. 8, indicates an increase of dcrest/dc with de-creasing Cd. The higher Cd forms before the point 0.9 ofdcrest/dc. After this point, there are two decreasing branchesof Cd in which the higher one relates to 2 < Hw/R < 8 andthe lower one relates to Hw/R < 2.Comparison between Fig. 7 and Fig. 8 indicates that when

dcrest/dc are less and more than 0.9, the critical depths are lo-cated before and after the crest, respectively. Higher values ofCd near the crest are the ones with lower dcrest/dc. The steadybranch around point 120° in Fig. 7 expresses the lowerbranch in Fig. 8, hence for Hw/R < 2, the critical depth loca-tion is stable.

5.2. Separation locationTo locate the separation point, an imaging technique is

used. For the sake of simplicity, we shall consider the anal-ogy with the overflow above a two-dimensional circular weir(Fig. 6). According to the images, it is seen that when theflow passes towards downstream, it separates because of theflow inertia and weir curvature. Results of this study showthe pressure decreases as the flow passes towards down-stream. For the range of experimental discharges investigated

Fig. 6. Geometrical parameters for critical depth and separation location.

Fig. 7. Critical depth location in different Cd. Fig. 8. Variation of Cd with dcrest/dc.

1332 Can. J. Civ. Eng. Vol. 38, 2011

Published by NRC Research Press

Can

. J. C

iv. E

ng. D

ownl

oade

d fr

om w

ww

.nrc

rese

arch

pres

s.co

m b

y U

nive

rsity

of

Okl

ahom

a L

ibra

ries

on

11/2

9/11

For

pers

onal

use

onl

y.

Page 7: An experimental and numerical study on hydraulic characteristics …profdoc.um.ac.ir/articles/a/1024597.pdf · 2020-03-07 · flow, which induces the nappe adherence (Chanson and

herein, the minimum pressure occurs at point 120° for higherdischarges, and for average and lower discharges, it occursaround points 150° and 180°, respectively.Suction pressure at the weir wall causes nappe adherence,

and thus the pressure head decreases towards the downstreamflow (Chanson and Montes 1998). However, pressure in-creases again after the nappe separation, and as the dischargedecreases, the nappe separation location shifts towards down-stream and farther away from the crest. Basic ideal-fluid flowcalculations provide an estimate of the wall pressure andnappe separation location induced by cavitation on the down-stream face of inflatable flexible membrane dams (Chanson1997):

½9� cos4sep ¼ 1� 1

sC

�DP

P� 2� dcrest

R

� �

where 4sep is separation location, sc is a critical cavitationnumber at which cavitation starts to appear in the flow whichis typically about 0.5 (Knapp et al. 1970), DP = Patm – PV,Patm is the atmospheric pressure, PV = Ps is the absolutepressure at the cylinder surface, P = rw × g × R, rw is thewater density, and g is the gravity acceleration. Other para-meters have been defined previously. The angle 4 is equal tozero at the crest (Fig. 6). Figure 9 shows the variation of DP/P on the weir for different dcrest/R (sC = 0.5). The line DP/P = 1 represents the atmospheric pressure where the nappeseparation occurs. According to this figure, as dcrest/R in-creases, the nappe separation location approaches the crest.In other words, for a given radius size, as the discharge in-creases, the separation point gets closer to the crest. In lowerdischarges, there is a suction pressure around the point 210°,and past this point, the atmospheric pressure leads to separa-tion.

6. Discussion

To calculate the dimensionless critical depth in non-hydrostatic pressure condition, Chanson (2006) gives anexpression as a function of the pressure correction coeffi-cient Lcrest, momentum correction coefficient bcrest, anddischarge coefficient Cd as follows:

½10� bcrest � C2d �L2

crest � 1

As bcrest is equal to or larger than 1.0, eq. [10] implies thatdischarge coefficients larger than 1.0 may be obtained onlywhen the crest pressure distribution is less than hydrostatic(Chanson 2006).

In experimental works, it is seen that in the region beforethe crest, higher discharges show higher pressure heads(Fig. 3), and thus according to eq. [10], Lcrest decrease leadsto Cd increase. As it was mentioned before, when the dis-charge increases, the suction pressure gets closer to the crest.Hence, dcrest decreases and the critical depth forms near thecrest. For lower discharges, the effect of suction pressure ondcrest is lower, and dc forms far from the crest. However, thepressure head is high enough that Cd > 1.4. For lower pres-sure heads in which Cd < 1.4, the value of dcrest/dc is greaterthan 0.9, and dc forms downstream of the crest. In this situa-tion, dcrest is nearly close to dc, and the water surface profileis uniform. Referring to Fig. 3, lower discharges show higherpressure heads after the crest. Thus, dc moves downstream.However, for Cd < 1.2, in which Hw/R < 2, the critical depthis located around the point 120°. It seems that in this rangethe head pressure effect is negligible.

7. ConclusionsThe CFD model is used to predict the hydraulic character-

istics of cylindrical weirs. Numerical simulations were ingood agreement with experimental data, which confirmedthe CFD capability in prediction of flow conditions on struc-tures such as rubber dams. Moreover, some theoretical for-mulations were derived to determine the critical depths andseparation locations in cylindrical weirs. According to theseequations, the critical depth points are a function of weir ra-dius and upstream flow depth. As the flow passes over theweir, the pressures always decrease towards the downstreamflow. However, depending on discharge values and afternappe separation, the pressure head increases. According toanalytical results, nappe separation usually occurs after thecrest, and approaches it in higher discharges.

ReferencesBos, M.G. 1989. Discharge measurement structures, 3rd ed. Institute

of Land Reclamation and Improvement ILRI, Wageningen,Netherlands.

Cassidy, J.J. 1965. Irrotational flow over spillways of finite height.Journal of the Engineering Mechanics Division, 91(6): 155–173.

Castro-Orgaz, O. 2008. Curvilinear flow over round-crested weirs.Journal of Hydraulic Research IAHR, 46(4): 543–547. doi:10.3826/jhr.2008.3251.

Castro-Orgaz, O., Giraldez, J.V., and Ayuso, J.L. 2008a. Higher ordercritical flow condition in curved streamline flow. Journal ofHydraulic Research IAHR, 46(6): 849–853. doi:10.1080/00221686.2008.9521931.

Castro-Orgaz, O., Giraldez, J.V., and Ayuso, J.L. 2008b. Critical flow

Fig. 9. Separation location for different dcrest/R ratios.

Naghavi et al. 1333

Published by NRC Research Press

Can

. J. C

iv. E

ng. D

ownl

oade

d fr

om w

ww

.nrc

rese

arch

pres

s.co

m b

y U

nive

rsity

of

Okl

ahom

a L

ibra

ries

on

11/2

9/11

For

pers

onal

use

onl

y.

Page 8: An experimental and numerical study on hydraulic characteristics …profdoc.um.ac.ir/articles/a/1024597.pdf · 2020-03-07 · flow, which induces the nappe adherence (Chanson and

over circular crested weirs. Journal of Hydraulic Engineering,134(11): 1661–1664. doi:10.1061/(ASCE)0733-9429(2008)134:11(1661).

Chanson, H. 1996. Some hydraulic aspects during overflow aboveinflatable flexible memberane dam. Queensland (AU), Departmentof Civil Engineering at the University of Queensland. Report No.CH47/96. pp. 1–60.

Chanson, H. 1997. A review of the overflow of inflatable flexiblemembrane dams. Journal of Australian Civil/Structure Engineer-ing Transactions, 39(2–3): 107–116.

Chanson, H. 2006. Minimum specific energy and critical flowconditions in open channels. Journal of Irrigation and DrainageEngineering, 132(5): 498–502. doi:10.1061/(ASCE)0733-9437(2006)132:5(498).

Chanson, H., and Montes, J.S. 1997. Overflow characteristics ofcircular weirs. Queensland (AU): Department of Civil Engineeringat the University of Queensland. Report No. CE154. pp. 1–98.

Chanson, H., and Montes, J.S. 1998. Overflow characteristics ofcircular weir. Journal of Irrigation and Drainage Engineering,124(3): 152–162. doi:10.1061/(ASCE)0733-9437(1998)124:3(152).

Dressler, R.F. 1978. New nonlinear shallow flow equations withcurvature. Journal of Hydraulic Research, 16(3): 205–222. doi:10.1080/00221687809499617.

Escande, L., and Sananes, F. 1959. Etude des seuils déversants à fenteaspiratrice. Houille Blanche, 14: 892–902. doi:10.1051/lhb/1959005. [In French.]

Hager, W.H. 1985. Critical flow condition in open channelhydraulics. Acta Mechanica, 54(3-4): 157–179. doi:10.1007/BF01184843.

Hager, W.H. 1993. Abfluss über zylinderwehr. Wasser und Boden,44(1): 9–14. [In German.]

Hager, W.H. 1995. Discussion to characteristics of circular crestedweirs. Journal of Hydraulic Engineering, 120(12): 1494–1495.doi:10.1061/(ASCE)0733-9429(1994)120:12(1494).

Heidarpour, M., and Chamani, M.R. 2006. Velocity distribution overcylindrical weirs. Journal of Hydraulic Research IAHR, 44(5):708–711. doi:10.1080/00221686.2006.9521719.

Heidarpour, M., Habili, J.M., and Haghiabi, A.H. 2008. Applicationof potential flow to circular-crested weir. Journal of HydraulicResearch IAHR, 46(5): 699–702. doi:10.3826/jhr.2008.3181.

Hirt, C.W., and Nichols, B.D. 1981. Volume of fluid (VOF) methodfor the dynamics of free boundaries. Journal of ComputationalPhysics, 39(1): 201–225. doi:10.1016/0021-9991(81)90145-5.

Hosseini, M. 2003. Investigation of efficiency and dischargecoefficient of culverts, M.Sc. thesis, Ferdowsi University ofMashhad, Iran.

Issa, R.I. 1986. Solution of the implicitly discretised fluid flowequations by operator-splitting. Journal of Computational Physics,62(1): 40–65. doi:10.1016/0021-9991(86)90099-9.

Knapp, R.T., Daily, J.W., and Hammit, F.G. 1970. Cavitation.McGraw-Hill Book Company, New York, N.Y.

Ramamurthy, A.S., and Vo, N.D. 1993a. Characteristics of circular-crested weirs. Journal of Hydraulic Engineering, 119(9): 1055–1062. doi:10.1061/(ASCE)0733-9429(1993)119:9(1055).

Ramamurthy, A.S., and Vo, N.D. 1993b. Application of dresslertheory to weir flow. Journal of Applied Mechanics, 60(1): 163–166. doi:10.1115/1.2900740.

Ramamurthy, A.S., Vo, N.D., and Vera, G. 1992. Momentum modelof flow past weir. Journal of Irrigation and Drainage Engineering,

118(6): 988–994. doi:10.1061/(ASCE)0733-9437(1992)118:6(988).

Sarginson, E.J. 1972. The influence of surface tension on weir flow.Journal of Hydraulic Research, 10(4): 431–446. doi:10.1080/00221687209500034.

Tadayon, R., and Ramamurthy, A.S. 2009. Turbulence modeling offlows over circular spillways. Journal of Irrigation and DrainageEngineering, 135(4): 493–498. doi:10.1061/(ASCE)IR.1943-4774.0000012.

Verwoerd, A.L. 1941. Calcul du débit de déversoirs dénoyés et noyésà crête arrondie. L’Ingénieur des Indes Néerlandaises, 8: 65–78.[In French.]

Yakhot, V., V, and Orszag, S.A. 1986. Renormalization-groupanalysis of turbulence. Physical Review Letters, 57(14): 1722–1724. doi:10.1103/PhysRevLett.57.1722. PMID:10033528.

Yakhot, V., and Smith, L.M. 1992. The renormalization group, the 3-expansion and derivation of turbulence models. Journal ofScientific Computing, 7(1): 35–61. doi:10.1007/BF01060210.

Yazdi, J., Sarkardeh, H., Azamathulla, H., and Ghani, A.A. 2010. 3Dsimulation of flow around a single spur dike with free-surfaceflow. International Journal of River Basin Management, 8(1): 55–62. doi:10.1080/15715121003715107.

Youngs, D.L. 1982. Time-dependent multi-material flow with largefluid distortion. In Numerical methods for fluid dynamics. Editedby K.W. Morton and M.J. Baines. Academic Press, London.pp. 273–285.

List of symbols

AB differences between upstream flow and water depths atthe weir crest, m

Cd discharge coefficientdc critical flow depth, m

dcrest water depths at the weir crest, mi specific energy, m

Fr Froude numberg gravity acceleration, m/s2

Hp pressure head, mHw flow depth in the upstream of the weir crest, mP pressure head on the weir, Pa

Patm atmospheric pressure, PaPV absolute pressure at the cylinder surface, PaR radius of cylindrical weir, mU differences between critical depth and water depths at

the weir crest, mVx horizontal fluid velocity, m/sVy vertical fluid velocity, m/sv flow velocity, m/saa volume fraction of airaw volume fraction of water

bcrest momentum correction coefficientDP differences between atmospheric and absolute pressures,

Pad flow depth in the critical depth location, mq critical depth location

Lcrest pressure correction coefficientr density of each cell in CFD modelra air density, kg/m3

rw water density, kg/m3

sC critical cavitation number (= 0.5)4sep separation location

1334 Can. J. Civ. Eng. Vol. 38, 2011

Published by NRC Research Press

Can

. J. C

iv. E

ng. D

ownl

oade

d fr

om w

ww

.nrc

rese

arch

pres

s.co

m b

y U

nive

rsity

of

Okl

ahom

a L

ibra

ries

on

11/2

9/11

For

pers

onal

use

onl

y.


Recommended