flow measurement and...

13
Total pressure uctuations and two-phase ow turbulence in self-aerated stepped chute ows Gangfu Zhang, Hubert Chanson n , Hang Wang The University of Queensland, School of Civil Engineering, Brisbane QLD 4072, Australia article info Article history: Received 9 January 2016 Received in revised form 8 August 2016 Accepted 12 August 2016 Available online 13 August 2016 Keywords: Total pressure uctuations Two-phase ow Turbulence Self-aeration Stepped spillways abstract Current knowledge in high-velocity self-aerated ows continues to rely upon physical modelling. Herein a miniature total pressure probe was successfully used in both clear-water and air-water ow regions of high-velocity open channel ows on a steep stepped channel. The measurements were conducted in a large size facility (θ ¼45°, h ¼0.1 m, W¼0.985 m) and they were complemented by detailed clear-water and air-water ow measurements using a Prandtl-Pitot tube and dual-tip phase-detection probe re- spectively in both developing and fully-developed ow regions for Reynolds numbers within 3.3 10 5 to 8.7 10 5 . Upstream of the inception point of free-surface aeration, the clear-water developing ow was characterised by a developing turbulent boundary layer and an ideal-ow region above. The boundary layer ow presented large total pressure uctuations and turbulence intensities, with distributions of turbulence intensity close to intermediate roughness ow data sets: i.e., intermediate between d-type and k-type. The total pressure measurements were validated in the highly-aerated turbulent shear re- gion, since the total pressure predictions based upon simultaneously-measured void fraction and velocity data agreed well with experimental results recorded by the total pressure probe. The results demon- strated the suitability of miniature total pressure probe in both monophase and two-phase ows. Both interfacial and water phase turbulence intensities were recorded. Present ndings indicated that the turbulence intensity in the water phase was smaller than the interfacial turbulence intensity. & 2016 Elsevier Ltd. All rights reserved. 1. Introduction Dams and reservoirs are man-made hydraulic structures built across rivers and streams to provide water storage. During major rainfalls, the large inows into a reservoir induce a rise in water level associated with the risk of dam overtopping, unless a spill- way system is designed. Most dams are equipped with an overow system, consisting of a crest, a steep chute and a downstream energy dissipator [31,37,46]. On the steep chute, the ow is ac- celerated by gravity and a turbulent boundary layer develops at the upstream end. When the outer edge of the boundary layer interacts with the free-surface, the turbulent shear stresses next to the air-water interface may overcome both the surface tension and buoyancy effects, and free-surface aeration takes place [14,24]. This location is called the inception point of free-surface aeration [34,49]. Fig. 1 illustrates the overow down a steep chute, and the inception point of free-surface aeration is clearly seen in Figs. 1A and B. Downstream self-aeration is commonly observed and the process is called 'white waters' [7,23,43,48] (Fig. 1). The physical processes are basically identical for smooth-invert and stepped spillways, although the latters are characterised by a greater rate of energy dissipation [11]. Current knowledge in high-velocity self-aerated ows relies heavily upon physical modelling and measurements, because of the large number of relevant equations and parameters [5,22,30]. Traditional monophase ow metrology may be used in the de- veloping ow region, although velocity measurements are difcult close to the free-surface [1,36,38]. Accurate measurements in the air-water ow region rely upon intrusive phase-detection probes and hot-lm probes. Review papers include Cain and Wood [8], Chanson [13], Chang et al. [10] and Chanson and Carosi [15]. In the present study, it is shown that a miniature total pressure probe may provide detailed informations in both clear-water and air-water ow regions. The metrology was applied to high-velocity open channel ows on a steep stepped channel. The measure- ments were conducted in a large size facility (θ ¼ 45°, h ¼ 0.1 m, W¼ 0.985 m) in which detailed turbulent ow properties were recorded systematically in both developing and fully-developed ow regions for several discharges, corresponding to Reynolds numbers within 3.3 10 5 to 8.7 10 5 . Contents lists available at ScienceDirect journal homepage: www.elsevier.com/locate/flowmeasinst Flow Measurement and Instrumentation http://dx.doi.org/10.1016/j.owmeasinst.2016.08.007 0955-5986/& 2016 Elsevier Ltd. All rights reserved. n Corresponding author. E-mail address: [email protected] (H. Chanson). Flow Measurement and Instrumentation 51 (2016) 820

Upload: others

Post on 28-Oct-2019

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Flow Measurement and Instrumentationstaff.civil.uq.edu.au/h.chanson/reprints/Zhang_et_al_fmi_2016.pdf · number defined in terms of the step height Herein a Froude si-militude was

Flow Measurement and Instrumentation 51 (2016) 8–20

Contents lists available at ScienceDirect

Flow Measurement and Instrumentation

http://d0955-59

n CorrE-m

journal homepage: www.elsevier.com/locate/flowmeasinst

Total pressure fluctuations and two-phase flow turbulence inself-aerated stepped chute flows

Gangfu Zhang, Hubert Chanson n, Hang WangThe University of Queensland, School of Civil Engineering, Brisbane QLD 4072, Australia

a r t i c l e i n f o

Article history:Received 9 January 2016Received in revised form8 August 2016Accepted 12 August 2016Available online 13 August 2016

Keywords:Total pressure fluctuationsTwo-phase flowTurbulenceSelf-aerationStepped spillways

x.doi.org/10.1016/j.flowmeasinst.2016.08.00786/& 2016 Elsevier Ltd. All rights reserved.

esponding author.ail address: [email protected] (H. Chanson

a b s t r a c t

Current knowledge in high-velocity self-aerated flows continues to rely upon physical modelling. Hereina miniature total pressure probe was successfully used in both clear-water and air-water flow regions ofhigh-velocity open channel flows on a steep stepped channel. The measurements were conducted in alarge size facility (θ¼45°, h¼0.1 m, W¼0.985 m) and they were complemented by detailed clear-waterand air-water flow measurements using a Prandtl-Pitot tube and dual-tip phase-detection probe re-spectively in both developing and fully-developed flow regions for Reynolds numbers within 3.3�105 to8.7�105. Upstream of the inception point of free-surface aeration, the clear-water developing flow wascharacterised by a developing turbulent boundary layer and an ideal-flow region above. The boundarylayer flow presented large total pressure fluctuations and turbulence intensities, with distributions ofturbulence intensity close to intermediate roughness flow data sets: i.e., intermediate between d-typeand k-type. The total pressure measurements were validated in the highly-aerated turbulent shear re-gion, since the total pressure predictions based upon simultaneously-measured void fraction and velocitydata agreed well with experimental results recorded by the total pressure probe. The results demon-strated the suitability of miniature total pressure probe in both monophase and two-phase flows. Bothinterfacial and water phase turbulence intensities were recorded. Present findings indicated that theturbulence intensity in the water phase was smaller than the interfacial turbulence intensity.

& 2016 Elsevier Ltd. All rights reserved.

1. Introduction

Dams and reservoirs are man-made hydraulic structures builtacross rivers and streams to provide water storage. During majorrainfalls, the large inflows into a reservoir induce a rise in waterlevel associated with the risk of dam overtopping, unless a spill-way system is designed. Most dams are equipped with an overflowsystem, consisting of a crest, a steep chute and a downstreamenergy dissipator [31,37,46]. On the steep chute, the flow is ac-celerated by gravity and a turbulent boundary layer develops atthe upstream end. When the outer edge of the boundary layerinteracts with the free-surface, the turbulent shear stresses next tothe air-water interface may overcome both the surface tension andbuoyancy effects, and free-surface aeration takes place [14,24].This location is called the inception point of free-surface aeration[34,49]. Fig. 1 illustrates the overflow down a steep chute, and theinception point of free-surface aeration is clearly seen in Figs. 1Aand B. Downstream self-aeration is commonly observed and theprocess is called 'white waters' [7,23,43,48] (Fig. 1). The physical

).

processes are basically identical for smooth-invert and steppedspillways, although the latters are characterised by a greater rateof energy dissipation [11].

Current knowledge in high-velocity self-aerated flows reliesheavily upon physical modelling and measurements, because ofthe large number of relevant equations and parameters [5,22,30].Traditional monophase flow metrology may be used in the de-veloping flow region, although velocity measurements are difficultclose to the free-surface [1,36,38]. Accurate measurements in theair-water flow region rely upon intrusive phase-detection probesand hot-film probes. Review papers include Cain and Wood [8],Chanson [13], Chang et al. [10] and Chanson and Carosi [15].

In the present study, it is shown that a miniature total pressureprobe may provide detailed informations in both clear-water andair-water flow regions. The metrology was applied to high-velocityopen channel flows on a steep stepped channel. The measure-ments were conducted in a large size facility (θ¼45°, h¼0.1 m,W¼0.985 m) in which detailed turbulent flow properties wererecorded systematically in both developing and fully-developedflow regions for several discharges, corresponding to Reynoldsnumbers within 3.3�105 to 8.7�105.

Page 2: Flow Measurement and Instrumentationstaff.civil.uq.edu.au/h.chanson/reprints/Zhang_et_al_fmi_2016.pdf · number defined in terms of the step height Herein a Froude si-militude was

Nomenclature

C time-averaged void fraction defined as the volume ofair per unit volume of air and water;

DH hydraulic diameter (m) also called equivalent pipediameter;

d clear water flow depth (m) measured normal to thepseudo-bottom formed by the step edges;

dc critical flow depth (m) : = ( )gd Q / Wc2 23 ;

F bubble count rate (Hz) or bubble frequency defined asthe number of detected air bubbles per unit time;

g gravity constant: g¼9.80 m/s2 in Brisbane, Australia;H1 upstream head above crest (m);h vertical step height (m);Lxx air-water advection integral length scale (m): Lxx¼Vx

Txx;(Lxx)max maximum advection air-water length scale (m) in a

cross-section;ks step cavity roughness height (m): ks¼h� cosθ;

′ks equivalent sand roughness height (m);Lcrest crest length (m);l horizontal step length (m);N power law exponent;Pk kinetic pressure (Pa);Ps static pressure (Pa);Pt total pressure (Pa): Pt¼Pkþ Ps;pk kinetic pressure fluctuation (Pa);ps static pressure fluctuation (Pa);pt total pressure fluctuation (Pa);pk

2 variance of kinetic pressure (Pa2);ps

2 variance of static pressure (Pa2);pt

2 variance of total pressure (Pa2);Q water discharge (m3/s);Re Reynolds number defined in terms of the hydraulic

diameter;Rxx normalised auto-correlation function;rpu correlation between static pressure and streamwise

velocity fluctuations: = ⎜ ⎟⎛⎝

⎞⎠r p v / v ppu s x x

2s2

Tu interfacial turbulence intensity: =Tu v /Vaw2

aw;Tup turbulence intensity in the water phase defined as:

=Tu v /Vp x2

x;

TX integral turbulent time scale (s) characterising largeeddies advecting the air bubbles;

Txx auto-correlation time scale (s): ∫= ( )= tT R dtxx 0

txx

Rxx 0

t time lag (s);Vaw interfacial velocity (m/s);Vc critical flow velocity (m/s);Vx streamwise velocity component in water phase (m/s);V90 characteristic interfacial velocity (m/s) where C¼0.90;vaw fluctuation of interfacial velocity (m/s);vx fluctuation of streamwise velocity component in water

phase (m/s);vaw

2 variance of longitudinal component of interfacial ve-locity (m2/s2);

vx2 variance of longitudinal component of water phase

velocity (m2/s2);W channel width (m);x distance along the channel bottom (m);Y90 characteristic depth (m) where the void fraction is

90%;y distance (m) measured normal to the invert (or

channel bed);z transverse distance (m) from the channel centreline;

Greek symbols

δ boundary layer thickness (m);μw water dynamic viscosity (Pa s);θ angle between the pseudo-bottom formed by the step

edges and the horizontal;ρ density (kg/m3);ρw water density (kg/m3);s surface tension between air and water (N/m);τ time lag (s);∅ diameter (m);

Subscript

aw interfacial flow data;p total pressure data;w water properties;xx auto-correlation;50 flow conditions where C¼0.50;90 flow conditions where C¼0.90.

G. Zhang et al. / Flow Measurement and Instrumentation 51 (2016) 8–20 9

2. Physical modelling, experimental facility andinstrumentation

2.1. Presentation

Steep chute flows are characterised by intense turbulence andinterfacial interactions. Physical modelling is typically performedin a down-sized version of the prototype (Fig. 1). A full dynamicsimilarity is necessary for the laboratory model (Fig. 1B) to accu-rately predict a range of prototype characteristics (Fig. 1A). On astepped chute, a simplistic dimensional analysis implies that theflow properties in the developing flow region must satisfy:

ρ ρ ρ ρ

ρμ

μ

ρ σθ

=′

…( )

⎛⎝⎜⎜

⎞⎠⎟⎟

dd

VV

vV

Pg d

Pg d

p

g d

p

g d

Fx

dy

dz

ddh

V D g Wd

kd

, , , , , ,

, , , , , , , , ,1

x w

c

x

c

x

c

t

w c

s

w c

t

w c

s

w c

1c c c

c w H

w

4

3c

s

c

where d is the water depth, Vx is the mean streamwise water velo-city, vx is the streamwise water turbulent velocity fluctuation, Pt andpt are the mean and fluctuating total pressure, Ps and ps are the meanand fluctuating static pressure, g is the gravity constant, dc is thecritical depth: dc¼(Q2/(g W))1/3 with Q the water discharge and Wthe chute width, Vc is the critical velocity: Vc¼(g dc)1/2, ρw is thewater density, x, y and z are respectively the streamwise, normal andtransverse coordinates, h is the step height, DH is the hydraulic dia-meter, μw is the dynamic viscosity of water, s is the surface tension ofwater, θ is the chute slope, and ′ks is the equivalent sand roughness ofthe step surface (Fig. 1C). In the fully-developed air-water flow re-gion, dimensional analysis yields a different expression:

ρ ρ ρ ρ…

μμ

ρ σθ

′…

( )

⎛⎝⎜⎜

⎞⎠⎟⎟

C,VV

,vV

,F dV

,Pg d

,Pg d

,p

g d

p

g d,

Fxd

,yd

,zd

,dh

,V D

,g

,Wd

, ,kd

,2

w

aw

c

aw

c

c

c

t

w c

s

w c

t

w c

s

w c

2c c c

c w aw H

w

4

3c

s

c

Page 3: Flow Measurement and Instrumentationstaff.civil.uq.edu.au/h.chanson/reprints/Zhang_et_al_fmi_2016.pdf · number defined in terms of the step height Herein a Froude si-militude was

Fig. 1. High-velocity self-aerated flows down stepped chutes. (A) Paradise Damstepped spillway in operation on 5 March 2013 (θ¼57.4°, h¼0.62 m, Q¼2,320 m3/s, dc/h¼2.85, Re¼2.9�107), (B) Physical model in operation (θ¼45°, h¼0.10 m,Q¼0.085 m3/s, dc/h¼0.9, Re¼3.6�105), (C) Definition sketch.

G. Zhang et al. / Flow Measurement and Instrumentation 51 (2016) 8–2010

where C is the void fraction, Vaw is the interfacial velocity, vaw is thestreamwise interfacial velocity fluctuation, and F is the bubble countrate.

In Eqs. (1) and (2), the developing clear-water and fully-de-veloped air-water flow properties at a given location (x, y, z) areexpressed as functions of the Froude, Reynolds and Mortonnumbers (4th, 5th and 6th terms) and channel properties. Notethat the dimensionless discharge dc/h is equivalent to a Froudenumber defined in terms of the step height Herein a Froude si-militude was adopted based upon Eqs. (1) and (2). The depen-dence on the Reynolds number implied potential scale effectssince the Reynolds numbers in laboratory were smaller than inprototypes. The facility was operated at relatively high Reynoldsnumbers up to 8.7�105 to minimise potential scale effects. Lastlythe fluid properties were invariant; thus the Morton number wasan invariant.

2.2. Experimental facility and instrumentation

Physical experiments were conducted in a large stepped spill-way model (1 V:1 H) located at the University of Queensland(Fig. 1B). The test section had a footprint of 4.6�0.985 m2. Itconsisted of a 0.6 m long and 0.985 m wide broad crest with a1.2 m high vertical upstream wall, an upstream rounded nose(0.058 m radius), and a downstream rounded edge (0.012 m ra-dius) The weir was followed by a steep chute consisting of 12steps: each step was 0.1 m long, 0.1 m high and 0.985 m wide. Asmooth and stable discharge was delivered by three pumps drivenby adjustable frequency AC motors. Water was fed into a 1.7 mdeep, 5 m wide intake basin with a footprint of 2.7�5 m2, leadingto a 2.8 m long side-wall convergent with a contraction ratio of5.08:1, resulting in a smooth and waveless inflow into the testsection. The water discharge was deduced from detailed velocityand pressure measurements above the broad crested weir [50]:

= +( )

⎜ ⎟⎛⎝⎜

⎞⎠⎟

⎛⎝

⎞⎠

QW

HL

g H0.8966 0.24323 3

1

crest1

3

where W is the crest width (W¼0.985 m), H1 is the total upstreamhead above crest, and Lcrest is the crest length (Lcrest¼0.60 m)(Fig. 1C).

Clear-water flow depths were measured with a pointer-gaugeon the channel centreline, as well as using photographic ob-servations and performing cheques with a phase detection probe.Velocity measurements were performed in the clear water de-veloping flow region with a Dwyers 166 Series Prandtl-Pitot tube(ؼ3.18 mm). The tube was equipped with a hemispherical totalpressure tapping and four equally spaced static pressure tappingslocated 25.4 mm behind the tip. The longitudinal separation be-tween the total and static tappings was taken into account, byrepeating independently measurements at each location.

The instantaneous total pressures were recorded with a micro-electro-mechanical-system (MEMS) MeasureX MRV21 miniaturepressure transducer with an inner diameter of 1 mm and an outerdiameter of 5 mm. The sensor featured a silicon diaphragm, withminimal static and thermal errors. The transducer was custom-designed to measure relative pressures ranging between 0 and0.15 bars with a precision of 0.5% full scale (FS). Note that thesensor could not measure sub-atmospheric pressures with anydegree of reliability. The signal was amplified and low-pass filteredat a cut off frequency of 2 kHz. A sampling frequency of 5 kHz wasselected to minimise information loss [47]. The sampling durationwas 60 s in the clear-water developing flow region and 180 s in theair-water fully-developed flow region during simultaneous sam-pling with the phase-detection probe.

Page 4: Flow Measurement and Instrumentationstaff.civil.uq.edu.au/h.chanson/reprints/Zhang_et_al_fmi_2016.pdf · number defined in terms of the step height Herein a Froude si-militude was

Fig. 2. Phase-detection and total pressure probes mounted side-by-side with 6.5 mm between total pressure sensor centreline and leading tip of phase-detection probe -Flow conditions: dc/h¼1.3, Re¼5.8�105, arrow points to the flow direction. (A) Probes located above the clear-water flow region, (B) Probes located in the upper spray abovethe air-water flow region (view in elevation).

G. Zhang et al. / Flow Measurement and Instrumentation 51 (2016) 8–20 11

Air-water flow measurements were conducted using dual-tipphase detection probes developed and built at the University ofQueensland [13]. The probe consisted of two tips, each having aninner diameter of 0.25 mm and an outer diameter of 0.8 mm. Theinner and outer electrodes of each tip were respectively made ofsilver and stainless steel. The longitudinal separation between thetips for each probe was between 6.9 mm and 8.0 mm dependingupon the probe. The probe was excited by an electronic system(Ref. UQ82.518) designed with a response time less than 10 μs.Each probe tip signal was recorded at a sampling rate of 20 kHz fora duration of 45 s, selected based upon previous sensitivity ana-lyses [26,45]. The sampling rate and duration were 5 kHz and 180 srespectively during simultaneous sampling with the pressuresensor (Fig. 2). In that case, the total pressure probe tip was at thesame elevation as and 6 mm aside the leading tip of the double-tipphase-detection probe.

A trolley system was used to support and position the probes(Pitot, total pressure, phase-detection). The longitudinal move-ment was fixed by steel rails parallel to the pseudo-bottom formedby the step edges and the normal movement was controlled by aMitutoyo™ digital ruler to achieve an accuracy of 70.01 mm.

2.3. Total pressure signal analysis

The total pressure sensor measured the instantaneous totalpressure in the direction aligned with the sensor:

ρ˜ = ˜ ˜ + ˜( )P V P

12 4t x

2s

where Pt is the instantaneous total pressure, ρ is the instantaneousfluid density, Vx is the instantaneous streamwise fluid velocitydetected by the sensor, and Ps is the instantaneous static pressure.

In the followings capital and lower case letters are used to denotemean and fluctuating quantities; for example, ˜ = +P P pt t t.

In clear water flows, the relationship between turbulence in-tensity and root mean square of total pressure fluctuation may bederived:

ρ=

( )Tu

p

V 5p

t2

w2

x4

where =Tu v V/p x2

x and Vx is the streamwise velocity componentin the water. Eq. (5) is similar to an expression derived by Arndtand Ippen [3] (see discussion in Appendix I).

In a two-phase flow, the total pressure output showed a dis-tinct bimodal distribution because of the effects of air bubbles.This is illustrated by the probability density functions (PDF) of thetotal pressure probe signals (Fig. 3). In Fig. 3, the first signalprobability distribution function (PDF) shows a unimodal dis-tribution because of the small number of air bubbles (C¼0.008,purple line), where C is the time-averaged void fraction and F thebubble count rate. The other two PDFs exhibit large peaks next tozero, likely linked to interfacial and capillary effects during airbubble impacts on the total pressure sensor. Neglecting the airdensity and capillary effects and for Tupr0.4–0.5, the relationshipbetween total pressure fluctuation and turbulence intensity yields:

( )=−

( − ) + ( )

ρ

( − )

TuC1 1 6

p

V

C C

Cp

14

2

t2

w2

x4

where Tup is the water-phase turbulence intensity, C is the time-averaged void fraction and Vx was assumed to be equal to theinterfacial velocity: VxEVaw. More generally, the higher order

Page 5: Flow Measurement and Instrumentationstaff.civil.uq.edu.au/h.chanson/reprints/Zhang_et_al_fmi_2016.pdf · number defined in terms of the step height Herein a Froude si-militude was

Fig. 3. Bimodal distributions of total pressure probe output – Flow conditions:θ¼45°, h¼0.1 m, dc/h¼1.7, Re¼8.7�105, step edge 12.

G. Zhang et al. / Flow Measurement and Instrumentation 51 (2016) 8–2012

terms (i.e. Tup3, Tup

4) may not be neglected, and a more completerelationship between total pressure fluctuation and turbulenceintensity is (Appendix I):

ρ= ( − ) + + ( − )

− ( − ) + ( − ) +

+ ( − ) ( )

⎛⎝⎜

⎛⎝⎜

⎞⎠⎟

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

p

VC

CC Tu

C Tu C C Tu

C C

112 4

0.34 1

0.116 1 1 112

14

1 7

t2

w2

x4

2p4

2p3

p2

These analytical expressions (Eqs. (6) and (7)) were derived toestimate the water-phase turbulence intensity from total pressurefluctuations in both clear-water and aerated flows. Eq. (6) char-acterises the turbulent fluctuations in the water phase of a high-velocity air-water flow, which simplifies into Eq. (5) for a clear-water flow. Both Equations may be used with reasonable accuracy,except when Tu exceeds O(1) (Appendix I).

2.4. Experimental flow conditions

Total pressure and two-phase flow measurements were con-ducted for a range of discharges with a focus on the skimmingflow regime (dc/hZ0.9). Both clear-water and two-phase mea-surements were undertaken at each step edge for 0.9rdc/hr1.7,corresponding to Reynolds numbers between 3.3�105 and8.7�105. The experimental flow conditions are summarised inTable 1.

3. Flow patterns and developing flow region

3.1. Presentation

Visual observations indicated that the overflow consisted of a

Table 1Experimental flow conditions.

Ref. Q (m3/s) dc/h Location Comments

Series 1 0.001–0.211 0.045–1.67 Step edges 1-12 Clear waterSeries 2 0.08F3–0.216 0.9–1.7 Step edges 3-9 Clear waterSeries 3 0.057–0.216 0.7–1.7 Step edges 5-12 Air-water fl

succession of free-falling nappes (i.e. nappe flow regime) for smalldischarges: i.e., dc/ho0.4. For a range of intermediate flow rates,the flow motion appeared pseudo-chaotic with strong spray andsplashing, and a combination of filled and partially-filled stepcavities: i.e., a transition flow regime observed for 0.4rdc/ho0.9.For larger discharges (dc/h Z0.9), the flow skimmed as a coherentstream above the pseudo-invert formed by the step edges, as seenin Fig. 1. Beneath the pseudo-bottom, cavity recirculation wasmaintained through the transfer of momentum from the mainstream to the recirculating motion. A significant amount of tur-bulent kinetic energy was dissipated to maintain the cavity cir-culation. For the remaining sections, the focus is on the skimmingflow regime, typical of large prototype spillway operation (Fig. 1A).

At the upstream end of the chute, the skimming flow free-surface was smooth and no free-surface aeration took place(Fig. 1B). Once the outer edge of the developing boundary layerinteracted with the free-surface, the flow was characterised bystrong air bubble entrainment [9,12,50]. This location is known asthe inception point of free-surface aeration, and divides the spill-way flow into an upstream clear water developing flow region,and a downstream air-water fully-developed flow region. The flowin step cavities exhibited a pseudo-stable recirculation motioncharacterised by self-sustaining vortices. A close examination ofthe cavity vortical structures showed irregular ejection of fluidfrom the cavity into the mainstream flow next to the upper ver-tical step face, and replacement of cavity fluid next to the stepedge, in manner similar to the observations of Djenidi et al. [21]and Chanson and Toombes [16].

3.2. Velocity, total pressure and turbulence intensity

In the developing flow region, the velocity data indicated aturbulent boundary layer with an ideal flow region above. In theideal flow region, the flow was accelerated by gravity and the free-stream velocities followed closely theoretical estimates basedupon the Bernoulli equation. Fig. 4 presents some typical dis-tribution of time-averaged velocity and total pressure measured atstep edges. Next to the pseudo-bottom, the velocity and totalpressure distributions both showed a steep gradient because ofeffects of form drag. Above, the velocity and total head remainedconstant as predicted by the Bernoulli equation. The length of thedeveloping flow region was seen to increase with increasing dis-charge. For dc/h41.9, no free-surface aeration was observed be-cause the flow was partially-developed over the entire chutelength.

All data showed large fluctuations in total pressure next to thepseudo-bottom formed by the step edges. This is illustrated inFig. 5A, where pt

2 is the variance of the total pressure fluctuations.The total pressure fluctuations were the largest about y/dcE0.1and decreased towards the free-surface. The occurrence of thislocal maximum might be linked to vortices in the wake of thepreceding step edge. Herein the largest dimensionless total pres-sure fluctuations were observed for the smallest discharge (dc/h¼0.9) corresponding to the shortest developing flow region(Fig. 5A). The turbulence intensity was derived from the totalpressure fluctuation and time-averaged velocity data (Eq. (5)).Typical results are presented in Fig. 5B. In the developing

Instrumentation

and air-water flow regions Visual observations.flow region Total pressure probe & Pitot tube.ow region Total pressure probe & Phase-detection probe.

Page 6: Flow Measurement and Instrumentationstaff.civil.uq.edu.au/h.chanson/reprints/Zhang_et_al_fmi_2016.pdf · number defined in terms of the step height Herein a Froude si-militude was

Fig. 4. Dimensionless distributions of time-averaged velocity and total pressure in the developing flow region of skimming flow - Flow conditions: dc/h¼1.7, Re¼8.7�105.(A) Streamwise velocity Vx, (B) Total pressure Pt.

G. Zhang et al. / Flow Measurement and Instrumentation 51 (2016) 8–20 13

boundary layer, the turbulence intensity Tup ranged between 0.05and 0.45, with maxima next to the pseudo-bottom. In the idealfluid flow region above, Tup was typically less than 0.05. Thepresent results were quantitatively of the same order of magni-tude as the results of Djenidi et al. [21] above d-type roughnessusing laser Doppler anemometry (LDA) and Amador et al. [1]

Fig. 5. Dimensionless distributions of total pressure fluctuations and turbulence intensitfluctuations pt

2 , (B) Turbulence intensity Tup.

above a stepped spillway model using particle image velocimetry(PIV).

In the present geometry, cavity recirculation vortices formed inthe step cavities in a manner somehow similar to the classicald-type roughness [20,21,41]. The triangular cavity might howeveraffect the main flow to a greater extent than a typical d-type

y in the developing flow region of skimming flow at step edge 4. (A) Total pressure

Page 7: Flow Measurement and Instrumentationstaff.civil.uq.edu.au/h.chanson/reprints/Zhang_et_al_fmi_2016.pdf · number defined in terms of the step height Herein a Froude si-militude was

Fig. 6. Turbulence intensity distributions in the developing boundary layer (stepedge 3) - Comparison with intermediate roughness data - Dashed line indicate theboundary layer thickness defined in terms of 99% of free-stream velocity.

G. Zhang et al. / Flow Measurement and Instrumentation 51 (2016) 8–2014

ribbed roughness. At each step edge, the interactions between theoverflow and the step edge induced separation and disturbances(eddies) convected by the mean flow. The mean flow properties inthe overflow were thus altered through interactions with sucheddies. When such vortical structures were produced at a fasterrate than they were dissipated, the skimming flow properties werealtered again before they are restored back to the undisturbedstate. For these reasons, the overflow properties above each stepcavity were likely to lack self-similarity and the present config-uration might be more appropriately classified as an intermediateroughness, bearing characteristics similar to both d-type andk-type roughness [39,41]. Intermediate roughness flows were in-vestigated experimentally by Okamoto et al. [40] and numericallyby Cui et al. [18] using large eddy simulations (LES). Their resultsare compared to the present data in Fig. 6, and a good qualitativeand quantitative agreement was achieved. Interestingly the tur-bulence intensity levels were non-negligible up to y/δE1.2 to 1.4,with δ the boundary layer thickness, highlighting the fluctuatingnature of the outer edge of the turbulent boundary layer[2,19,35,42,44].

3.3. Autocorrelation time and length scales

The autocorrelation time and length scales were characteristicscales of the largest eddies advected in the streamwise direction.In the developing flow region, they were estimated as:

∫= ( ) ( )=

T R t dt 8t

xx0

xxRxx 0

= ( )L V T 9xx x xx

where R is the normalised correlation coefficient of the totalpressure, t is the time lag and it is implicitly assumed that theturbulent structures were convected at the same speed as thestreamwise mean water velocity Vx. The auto-correlation timescale is an integral time scale that characterises the longest con-nections in the turbulent behaviour of the streamwise total pres-sure fluctuations. The auto-correlation length scale is an advectionlength scale that is a measure of the longest connection (or cor-relation distance) between the total pressures at two points of theflow field [32].

Dimensionless autocorrelation time and length scale distribu-tions are presented in Fig. 7, where ks denotes the step cavityroughness height: ks¼h cosθ (¼0.071 m herein). At step edge 3,the largest time and length scales were found around the mid-boundary layer (y/δE0.5), except for the largest discharge. Thecharacteristic eddy sizes were about three times the cavityroughness height, close to the data of Okamoto et al. [39] whoobtained Lxx/ksE2 for various types of k-type triangular ribs. Inthe developing region, the flow was accelerated by gravity andvortices were stretched by the streamwise velocity gradient (∂Vx/∂x), explaining possibly the slightly larger results than Okamotoet al. [39]. Outside the boundary layer, the auto-correlation timeand length scales were small. At the next step edge (step edge 4),the data showed a somewhat different pattern (Fig. 7). The loca-tions of maximum time and length scales shifted towards theouter edge of the boundary layer, and larger time and length scaleswere recorded across the water column up to the free-surface. Thislack of similarity might be linked to the highly fluctuating natureof the boundary layer and highlighted the complexity of the step-wake interactions over a stepped invert.

4. Fully-developed air-water flow region

4.1. Presentation

Downstream of the inception point of free-surface aeration, theflow was fully-developed and strong self-aeration was observed.This is illustrated in Fig. 8, showing typical distributions of time-averaged void fraction and interfacial velocity measured with thephase-detection probe. Herein the interfacial velocity was calcu-lated based upon a cross-correlation technique [17]. The voidfraction distributions showed an S-shape typically observed onstepped spillways with flat steps [16,27] (Fig. 8A). The data pre-sented some self-similarity. In the overflow above the pseudo-bottom formed by the step edges, the void fraction data followedclosely a theoretical model proposed by Chanson and Toombes[16] (Fig. 8A, solid line). The interfacial velocity data were ap-proximated by a power law:

( )= < <( )

⎛⎝⎜

⎞⎠⎟

VV

yY

y Yat step edge, 010

Naw

90 90

1/

90

where Y90 is the characteristic distance normal to the pseudo-bottom where C¼0.9, and V90 is the interfacial velocity at y¼Y90.For y4Y90, the interfacial velocity distributions were essentiallyuniform and described by:

( )= ≥( )

VV

y Y1 at step edge, / 111

aw

9090

The above relationships were compared to present experi-mental data, showing a satisfactory agreement for N¼10 for allstep edges (Fig. 8B), despite some scatter next to the inceptionpoint. In the vicinity of the inception point, the flow was subjectedto rapid flow bulking.

4.2. Total pressure and turbulence intensity

In the air-water flows, the total pressure signal showed a dis-tinct bimodal distribution because of the effects of air bubbles(Fig. 3). Typical distributions of time-averaged total pressure arepresented in Fig. 9, and compared with the void fraction dis-tribution. The data showed a maximum total pressure slightlybelow y¼Y50, where Y50 is a characteristic depth for C¼0.5.

Neglecting the air density and the capillary effects of wettingand drying (i.e. during interfacial interactions with probe sensor),

Page 8: Flow Measurement and Instrumentationstaff.civil.uq.edu.au/h.chanson/reprints/Zhang_et_al_fmi_2016.pdf · number defined in terms of the step height Herein a Froude si-militude was

Fig. 7. Autocorrelation time and length scales in the developing flow region (step edges 3 and 4). (A) Streamwise integral time scale Txx, step edge 3, (B) Advection lengthscale Lxx, step edge 3, (C) Streamwise integral time scale Txx, step edge 4, (D) Advection length scale Lxx, step edge 4.

G. Zhang et al. / Flow Measurement and Instrumentation 51 (2016) 8–20 15

the mean total pressure may be expressed as:

( )ρ= + = ( − ) + + ( )P P P C V v P12

1 12t k s w x2

x2

s

where Pk is the mean kinetic pressure and Ps is the time-averagedstatic pressure. If the pressure gradient is hydrostatic, the time-averaged static pressure, Ps, may be deduced from the time-aver-aged void fraction distribution:

∫ρ θ= ( − )( )

P g Ccos 1 dy13y

Y

s w

90

where y is the coordinate normal to the pseudo-bottom and θ isthe angle between the pseudo-bottom and horizontal. Based uponEq. (12), the kinetic pressure term (Pk) may be estimated from thephase-detection probe data (C and Vaw), if the time-averagedstreamwise water velocity component Vx equals the interfacial

Page 9: Flow Measurement and Instrumentationstaff.civil.uq.edu.au/h.chanson/reprints/Zhang_et_al_fmi_2016.pdf · number defined in terms of the step height Herein a Froude si-militude was

Fig. 8. Dimensionless distributions of time-averaged void fraction and interfacial velocity in the air-water fully-developed flow region for dc/h¼1.3, Re¼5.8�105. (A) Voidfraction distributions, (B) Interfacial velocity distributions.

Fig. 9. Dimensionless distributions of total pressure and void fraction in the air-water fully-developed flow region of skimming flows - Comparison between total pressuredata Pt and estimated total pressure Pt,est based upon phase detection probe data - Dashed line indicates elevation Y50 where C¼0.5. (A) dc/h¼0.9, Re¼3.4�105, step edge 5,(B) dc/h¼1.7, Re¼8.7�105, step edge 12.

G. Zhang et al. / Flow Measurement and Instrumentation 51 (2016) 8–2016

velocity Vaw, and the turbulence intensity is small ( < <v V/ 1x2

x2 ).

Combining with the assumption of hydrostatic pressure distribu-tion (Eq. (13)), an estimate of the mean total pressure may bederived from the phase-detection probe data, denoted Pt, est. Fig. 9presents typical comparisons between the total pressure sensordata Pt and the estimate of mean total pressure Pt, est based upon

the phase-detection probe data. The present data set showed agood agreement between the measured and estimated totalpressures (Fig. 9). The finding implied that a hydrostatic pressuredistribution, taking into account the air entrainment (Eq. (13)),may be assumed in fully-developed air-water skimming flows. Theclose agreement between Pt and Pt, est further implied that the

Page 10: Flow Measurement and Instrumentationstaff.civil.uq.edu.au/h.chanson/reprints/Zhang_et_al_fmi_2016.pdf · number defined in terms of the step height Herein a Froude si-militude was

Fig. 10. Dimensionless distributions of total pressure fluctuations in the air-waterfully-developed flow region of skimming flow - Flow conditions: dc/h¼0.9,Re¼3.4�105.

G. Zhang et al. / Flow Measurement and Instrumentation 51 (2016) 8–20 17

water phase turbulence intensity was small in aerated skimmingflows (Tupo0.4–0.5) (see discussion below).

Typical distributions of the root-mean-square (rms) of di-

mensionless total pressure fluctuations pt2 are presented in

Fig. 10. The distributions presented a characteristic shape with amaximum about y/Y90E0.7. Next to the free-surface, moderatetotal pressure fluctuations were recorded (top most data points)but these data might be biased by capillary and interfacial effectsduring the droplet impacts onto the sensor.

Fig. 11. Dimensionless distributions of turbulence intensity in the air-water fully-develintensity Tup and interfacial turbulence intensity Tu - Flow conditions: dc/h¼1.7, Re¼8.

The water phase turbulence intensity Tup may be derived fromthe total pressure signal and two-phase flow properties (Eq. (6))(see Appendix I). Tup characterises the streamwise velocity fluc-tuations of the water phase. Typical data are presented in Fig. 11,where they are compared with interfacial turbulence intensity,calculated based upon cross-correlation of dual-tip phase detec-tion probe signals [16]. Fig. 11A shows the water phase turbulenceintensity Tup ranging between 0.1 and 0.5, irrespective of thedischarge. Local maxima were found next to the pseudo-bottomranging between 0.25–0.3, close to developing flow skimming flowdata [1,38]. The turbulence intensity of the water phase presenteda minimum value about 0.1–0.15 next to y/Y90E0.5 to 0.7. Theseminimum values were higher than those recorded at the outeredge of the developing boundary layer (5%, Section 3). For y/Y9040.5–0.7, Tup increased with increasing distance from thepseudo-bottom. The relatively large values of Tup next to the free-surface might be partially attributed to the breaking surface withsubstantial water spray. The data trends are highlighted with blacksolid lines in Fig. 11. Typical interfacial turbulence intensities Tudeduced from a dual-tip phase-detection probe are presented inFig. 11B. The interfacial turbulence intensities were systematicallylarger, ranging between 0.4 and 3.0. The data trends are alsomarkedly different, as illustrated by the trend lines (thick solidlines) in Fig. 11. A systematic comparison between the two sets ofdata showed opposite trends, as seen in Fig. 11.

The present data trends (Fig. 11) might suggest that the velocityfluctuations of the water phase were damped by the presence of alarge number of air bubbles in the mid-water column, because ofsurface tension and compressible nature of air. The direct contrastwith the large interfacial turbulence intensities in that regionmight hint that the water particles and air-water interfaces fluc-tuated at different scales despite zero slip on average (VxEVaw).Considering the transport of a fluid particle under an in-stantaneous streamwise pressure gradient ∂ ∂− P / xs , assuming thatthe fluid behaves like an incompressible fluid and neglecting allother surface and body forces, the streamwise acceleration equals

ρ ∂− × ˜ ∂P1/ / xs , where ρ is the density of the fluid. Since the density

oped flow region of skimming flow: comparison between water phase turbulence7�105. (A) Total pressure sensor data Tup, (B) Phase-detection probe data Tu.

Page 11: Flow Measurement and Instrumentationstaff.civil.uq.edu.au/h.chanson/reprints/Zhang_et_al_fmi_2016.pdf · number defined in terms of the step height Herein a Froude si-militude was

Fig. 12. Auto-correlation time scale distributions in the air-water fully-developedflow region of skimming flow - Comparison between phase-detection probe dataand total pressure data (Present study, step edge 10), and integral turbulent timescale in air-water flows [15,26]

.

G. Zhang et al. / Flow Measurement and Instrumentation 51 (2016) 8–2018

of water is approximately 800 times larger than that of air, theacceleration of an air particle is also 800 times larger. If thepressure gradient is uniform for a characteristic duration τc, an airparticle moves 800 times the distance of a water particle of thesame size during that time, if both particles start from rest. Inreality, a number of processes including capillary forces andcompressibility may contribute to a reduction in the differencebetween air and water velocity fluctuation scales. This simplisticdiscussion however underlines key distinctions between turbulentfluctuations of air and water phases.

4.3. Autocorrelation time and length scales

In the air-water flow region, the auto-correlation time andlength scales were calculated based upon the instantaneous totalpressure signal (subscript p) and the instantaneous void fraction(subscript aw). Typical results are shown in Fig. 12, where presentdata are compared with integral turbulent time scale data mea-sured during previous relevant studies. Note that the physical in-terpretation of the time scales differs depending upon the datasource. (Txx)aw characterised the time scale of the longitudinalinterfacial structures advecting the air-water interfaces in the flowdirection [15]. Present data showed a bell shape, with maximumvalues at y/Y90E0.6–0.8 (Fig. 12). All data tended to follow a self-similar distribution. In contrast, (Txx)p represented a characteristictime scale of the energy-containing eddies advected in the flowdirection. The distributions of (Txx)p presented maxima next to thepseudo-bottom, and decreased monotonically up to y/Y90E0.5(corresponding to C¼0.3–0.4). At that location, the data trendshowed a distinct change in slope, implying some physical changein the air-water flow structure. This location might approximatelycorrespond to the upper bound of the shear layer created by thepreceding step edge, below which the wake-step interactionsmight be significant.

The present data were compared to the turbulent integral time

scale data of Chanson and Carosi [15], Felder [25] and Felder andChanson [26] obtained in skimming flows on flat to moderateslopes (θ¼8.9° to 26.6°). The results were quantitatively consistentwith present findings, suggesting that the chute slope and dis-charge might have little influence of the interfacial structures.Such a strong self-similarity could be useful for extrapolation ontoprototype structures, given Reynolds numbers sufficiently high toreproduce large-scale turbulent structures.

The distributions of auto-correlation length scales showedsome self-similarity, with a trend close to those of the auto-cor-relation time scales. The data showed that the largest bubbly flowstructures and energy-containing structures were respectivelypresent in the mid-flow region and next to the pseudo-bottom. Alldata followed a self-similar distribution far downstream of theinception point. The length scales satisfied typically Lxx/Y90E0.1–0.6, with (Lxx)pE(Lxx)aw for y/Y9040.6. (data not shown).

5. Conclusion

The total pressure distributions were measured in the clear-water developing flow region and air-water fully-developed flowregion above a steep stepped chute. These measurements werecomplemented by clear-water velocity measurements with a Pitottube and air-water measurements with a dual-tip phase-detectionprobe. Analytical expressions were derived to estimate the water-phase turbulence intensity from total pressure fluctuations in bothclear-water and aerated flows. The results were tested for fiverelatively large discharges corresponding to a skimming flow re-gime, and demonstrated the suitability of miniature total pressureprobe in both clear-water and air-water flows.

Upstream of the inception point of free-surface aeration, theclear-water developing flow was characterised by a developingturbulent boundary layer and an ideal-flow region above. Thepotential flow velocity was well predicted by the Bernoulli equa-tion. The boundary layer exhibited large total pressure fluctuationsand turbulence intensities. The distributions of turbulence in-tensity were close to intermediate roughness flow data sets (i.e.intermediate between d-type and k-type).

The two-phase flow measurements provided time-averagedvoid fraction and interfacial velocity distributions, illustrating theinterfacial air–water exchange next to the free surface. The totalpressure measurements were validated in the highly-aerated tur-bulent shear region, since the total pressure predictions basedupon the void fraction and velocity data agreed well with ex-perimental results recorded by the total pressure probe. The staticpressure exhibited a hydrostatic distribution, taking into accountthe void fraction distribution, in the air-water fully-developed flowregion. The distributions of total pressure fluctuations showed adistinctive shape with a maximum about y/Y90E0.7. The datapresented a contrast between the interfacial turbulence intensitiesand water phase turbulence intensities. The turbulence intensityin the water phase was typically smaller than the interfacial tur-bulence intensity, suggesting that velocity fluctuations of the wa-ter phase were damped by the presence of a large number of airbubbles because of surface tension and compressible nature of air.The auto-correlation time and length scale results were close toprevious findings with different chute slopes and discharges,suggesting that these geometrical parameters might have littleinfluence on the integral turbulent scales of the interfacialstructures.

Acknowledgements

The authors acknowledge the technical assistance of Jason Van

Page 12: Flow Measurement and Instrumentationstaff.civil.uq.edu.au/h.chanson/reprints/Zhang_et_al_fmi_2016.pdf · number defined in terms of the step height Herein a Froude si-militude was

G. Zhang et al. / Flow Measurement and Instrumentation 51 (2016) 8–20 19

Der Gevel and Stewart Matthews (The University of Queensland).They thank the reviewers for their helpful and constructive com-ments. The financial support through the Australian ResearchCouncil (Grant DP120100481) is acknowledged.

Appendix I. Total pressure signal analysis in air-water flows

The total pressure sensor measures the instantaneous totalpressure aligned with the sensor, comprised of mean and fluctu-ating components:

˜ = ρ ˜ + ˜( )P

12

V P I.1t x2

s

where Pt is the instantaneous total pressure, ρ is the instantaneousfluid density, Vx is the instantaneous streamwise fluid velocitydetected by the sensor, and Ps is the instantaneous static pressure.In what follows capital and lower case letters are used to denotemean and fluctuating quantities; for example, ˜ = +P pPt t t. WhileIppen et al. [33] and Arndt and Ippen [3] proposed derivations formonophase flows (see below), the present appendix is focused onthe signal analysis in high-velocity air-water flows.

In a two-phase flow, the total pressure output will show adistinct bimodal distribution because of the effects of air bubbles.This is seen in the probability density functions (PDF) of the totalpressure probe signals, with a number of examples shown in Fig. 3.In Fig. 3, the first sensor signal PDF (C¼0.008, purple line) shows aunimodal distribution because of the small number of air bubbles,while the other two PDFs exhibit large peaks next to zero, likelylinked to air bubble impacts. When a submerged air bubble im-pacts the probe sensor, the effects include a kinetic pressure closeto zero because air density is negligible compared to that of water,a static pressure increase inside the air bubble because of surfacetension which becomes small compared to the kinetic pressure forbubble sizes greater than the millimetre, and capillary effectsduring wetting and drying processes that bias the signal.

Neglecting the air density and capillary effects, Eq. (I.1) may bedecomposed into mean and fluctuating components and writtenseparately for the individual phases. If the time-averaged voidfraction C is interpreted as the probability of one signal samplebeing air, the classical time-average may be redefined as:

( ) = ( − ) × ( )| + × ( )| ( )= =f c t C f c t f c t, 1 , C , I.2c 0 c 1i i

where f(c,t) is a function void fraction and time, and the overbardenotes a time-average operation. Applying Eq. (I.2) on Eq. (I.1),the mean total pressure becomes:

( )= ( − ) ρ + + ( )C VP12

1 v P I.3t w x2

x2

s

where ρw is the water density. Similarly the total pressure fluc-tuation may be derived as:

= ( − ) ρ + ρ + ρ + ( − ) ρ

+ + ρ − ( − )

ρ +ρ + ρ − ( − ) ρ + ρ

+ ( − ) ρ + + +( )

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎛⎝⎜

⎞⎠⎟

⎞⎠⎟

C C

C C

C

C V

p 114

C V V v14

v14

1 v

p12

C V v12

1

V v V v 2 V v p12

1 v v p

C14

1 2 V v v pI.4

t2

w2 2

x4

w2

x2

x2

w2

x4 2

w2

x22

s2

w2

x2

x2

w2

x2

x2

w2

x x3

w x x s w2

x22

w x2

s

2w2

x4

x2

x2

x22

s2

If the static pressure and streamwise velocity fluctuations, re-spectively ps and vx, are each normally distributed with zero

mean, and denoting = ⎜ ⎟⎛⎝

⎞⎠r p v / v ppu s x x

2s2 their normalised

correlation coefficient, Eq. (I.4) may be simplified into [29]:

= ρ + ρ + +ρ +

+ ρ ( − )

+ ( − ) ρ + + +( )

⎛⎝⎜

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

⎞⎠⎟⎟

⎛⎝⎜

⎛⎝⎜

⎞⎠⎟⎞⎠⎟

V

C

C V

p14

C V 112

C v12

C4

v

2 r V v p 1

14

1 2 V v v C pI.5

x

t2

w2 2

x4

w2 2

x2

x2

w2

2

x22

pu w x2

s2

2w2

x4

x2

x2

x22

s2

For an irrotational flow field, Bradshaw [6] demonstrated therelationship between pressure and velocity fluctuations, showingthat the static pressure fluctuations include contributions fromboth irrotational velocity fluctuations and flow convection andyielding a negative correlation coefficient rpu. In the initial regionof a plane turbulent wind jet, Guo and Wood [28] observedrpuE�0.25 in the turbulent zone away from the jet core. Giventhat the presence of air bubbles might provide some ‘cushioning’damping, it is proposed that rpu¼�0.1, which is yet to be ex-perimentally justified. (As part of preliminary tests, no major dif-ference (o5%) was observed for �0.5o rpuo�0.05 in terms ofturbulence intensity.) In a field of homogeneous and isotropicturbulence with very large Reynolds numbers, the mean-squarepressure fluctuation can be expressed in terms of the mean-squarevelocity fluctuation [4]:

= ρ ( )p 0.34 v I.6s2

w2

x22

Considering Eq. (I.6) as a crude approximation, and assumingrpu¼�0.1, the relationship between total pressure fluctuation and

turbulence intensity =Tu v /Vp x2

x become in air-water flows:

ρ= ( − ) + + ( − )

− ( − ) + ( − ) +

+ ( − ) ( )

⎛⎝⎜

⎛⎝⎜

⎞⎠⎟

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

C C

C C

C C

p

V1

12

C4

0.34 1 Tu

0.116 1 Tu 1 112

C Tu

14

1 I.7

t2

w2

x4

2p4

2p3

p2

If the higher order terms (i.e. Tup3, Tup

4) are neglected (e.g.Tupr0.4-0.5), the following approximate form holds:

( )=−

( − ) + ( )

ρ

( − )

CTu

1 1 I.8

C C

p

p

V

14

C2

t2

w2

x4

Eq. (I.7) must be used if Tup approaches or exceeds O(1). Forclear water flow (C¼0), Eq. (I.8) becomes:

ρ=

( )

p

VTu

I.9x

t2

w2 4 p

2

Eq. (I.9) assumed implicitly that the static pressure fluctuationsare negligible, as these are a higher order term in Eq. (I-7).

Following Ippen et al. [33], Arndt and Ippen [3] used a similarapproach and obtained an expression in monophase flows:

ρ< + +

( )

p

VTu Tu

54

TuI.10

t2

w x4 p

2p3

p4

Eq. (I.10) is identical to (I.9) if the higher order terms are ne-glected and the operator 'o ' is replaced by '¼ '. The main differ-ence between their work and the present lies in the treatment ofthe calculations of expected values of the products of the fluctu-ating terms. Arndt and Ippen [3] assumed:

Page 13: Flow Measurement and Instrumentationstaff.civil.uq.edu.au/h.chanson/reprints/Zhang_et_al_fmi_2016.pdf · number defined in terms of the step height Herein a Froude si-militude was

G. Zhang et al. / Flow Measurement and Instrumentation 51 (2016) 8–2020

( )p v2

v I.11as xw

x23/2

( )p v2

v I.11bs x2 w

x22

( )p4

v I.11cs2 w

2

x22

While Eqs. (I.11a) and (I.11b) are correct, Eq. (I.11c) contradictsBatchelor's [4] hypothesis for isotropic turbulence with Re -þ1(Eq. (I.6)).

In summary, analytical expressions were derived to estimatethe water-phase turbulence intensity from total pressure fluctua-tions in both clear water and aerated flows. Eq. (I.8) characterisesthe turbulent fluctuations in the water phase of a high-velocity air-water flow, which simplifies into Eq. (I.9) for a clear water flow.Both Equations may be used with reasonable accuracy, exceptwhen Tu exceeds O(1).

References

[1] A. Amador, M. Sanchez-Juny, J. Dolz, Characterization of the nonaerated flowregion in a stepped spillway by PIV, J. Fluids Eng. ASME 128 (6) (2006)1266–1273.

[2] R.A. Antonia, Conditionally sampled measurements near outer edge of a tur-bulent boundary layer, J. Fluid Mech. 56 (1) (1972) 1–18.

[3] R.E.A. Arndt, A.T. Ippen, Turbulence measurements in liquids using an im-proved total pressure probe, J. Hydraul. Res. 8 (2) (1970) 131–158, http://dx.doi.org/10.1080/00221687009500300.

[4] G.K. Batchelor, Pressure fluctuations in isotropic turbulence, Math. Proc. Camb.Philos. Soc. 47 (2) (1951) 359–374, http://dx.doi.org/10.1017/s0305004100026712.

[5] F.A. Bombardelli, Computational Multi-Phase Fluid Dynamics to Address Flowspast Hydraulic Structures. Proc. 4th IAHR International Symposium on Hy-draulic Structures, APRH - Associação Portuguesa dos Recursos Hídricos(Portuguese Water Resources Association), J. Matos, S. Pagliara & I. MeirelesEds., 9-11 February 2012, Porto, Portugal, Paper 2, 19 pages (CD-ROM).

[6] P. Bradshaw, Turbulence, Topics in Applied Physics, Springer-Verlag, Berlin,Germany 1976, p. 335.

[7] P. Cain, I.R. Wood, Measurements of self-aerated flow on a spillway, J. Hydraul.Div. 107 (1981) 1425–1444.

[8] P. Cain, I.R. Wood, Instrumentation for aerated flow on spillways, J. Hydraul.Div. ASCE 107 (HY11) (1981) 1407–1424.

[9] M.R. Chamani, N. Rajaratnam, Characteristics of skimming flow over steppedspillways, J. Hydraul. Eng. ASCE 125 (4) (1999) 361–368.

[10] K.A. Chang, H.J. Lim, C.B. Su, Fiber optic reflectometer for velocity and fractionratio measurements in multiphase flows, Rev. Sci. Instrum. 74 (2003)3559–3565, Discussion: 2004, 75:284-286.

[11] H. Chanson, Comparison of energy dissipation between nappe and skimmingflow regimes on stepped chutes, J. Hydraul. Res. IAHR 32 (2) (1994) 213–218.

[12] H. Chanson, Hydraulics of skimming flows over stepped channels and spill-ways, J. Hydraul. Res. IAHR 32 (3) (1994) 445–460.

[13] H. Chanson, Air-water flow measurements with intrusive phase-detectionprobes. Can we improve their interpretation ? J. Hydraul. Eng. 128 (2002)252–255.

[14] H. Chanson, Turbulent air-water flows in hydraulic structures: dynamic si-milarity and scale effects, Environ. Fluid Mech. 9 (2) (2009) 125–142, http://dx.doi.org/10.1007/s10652-008-9078-3.

[15] H. Chanson, G. Carosi, Turbulent time and length scale measurements in high-velocity open channel flows, Exp. Fluids 42 (3) (2007) 385–401, http://dx.doi.org/10.1007/s00348-006-0246-2.

[16] H. Chanson, L. Toombes, Air-water flows down stepped chutes: turbulence andflow structure observations, Int. J. Multiph. Flow. 27 (2002) 1737–1761.

[17] C. Crowe, M. Sommerfield, Y. Tsuji, Multiphase Flows with Droplets and Par-ticles, CRC Press, Boca Raton, USA 1998, p. 471.

[18] J. Cui, V.C. Patel, C. Lin, Large-eddy simulation of turbulent flow in a channelwith rib roughness, Int. J. Heat. Fluid Flow. 24 (3) (2003) 372–388, http://dx.doi.org/10.1016/s0142-727x(03)00002-x.

[19] H.W. Daily, D.R.F. Harleman, Fluid dynamics, Reading, Mass, Addison-WesleyPublishing, USA 1966, p. 454.

[20] L. Djenidi, F. Anselmet, R.A. Antonia, LDA measurements in a turbulentboundary layer over a d-type rough wall, Exp. Fluids 16 (1994) 323–329.

[21] L. Djenidi, R. Elavarasan, R.A. Antonia, The turbulent boundary layer overtransverse square cavities, Jl Fluid Mech. 395 (1999) 271–294.

[22] D.A. Drew, S.L. Passman, Theory of multicomponent fluids, in: J.E. Marsden,L. Sirovich (Eds.), Applied Mathemetical Sciences, 135, Springer, New York,USA, 1999, p. 308.

[23] R. Ehrenberger, Wasserbewegung in steilen Rinnen (Susstennen) mit be-sonderer Berucksichtigung der Selbstbelüftung. ('Flow of Water in SteepChutes with Special Reference to Self-aeration.') Zeitschrift des Ös-terreichischer Ingenieur und Architektverein, (15/16 and 17/18 (in German)(translated by Wilsey, E.F., U.S. Bureau of Reclamation), 1926.

[24] D.A. Ervine, H.T. Falvey, Behaviour of turbulent water jets in the atmosphereand in plunge pools, Proc. Inst. Civil. Eng. Lond. 2 83 (1987) 295–314.

[25] S. Felder, Air-Water Flow Properties on Stepped Spillways for EmbankmentDams: Aeration, Energy Dissipation and Turbulence on Uniform, Non-Uniformand Pooled Stepped Chutes. pH.D. thesis, School of Civil Engineering, TheUniversity of Queensland, Brisbane, Australia, 2013.

[26] S. Felder, H. Chanson, Phase-detection probe measurements in high-velocityfree-surface flows including a discussion of key sampling parameters, Exp.Therm. Fluid Sci. 61 (2015) 66–78, http://dx.doi.org/10.1016/j.expthermflusci.2014.10.009.

[27] C.A. Gonzalez, H. Chanson, Interactions between cavity flow and main streamskimming flows: an experimental study, Can. J. Civ. Eng. 31 (1) (2004) 33–44.

[28] Y. Guo, D.H. Wood, Instantaneous velocity and pressure measurements inturbulent mixing layers, Exp. Therm. Fluid Sci. 24 (3–4) (2001) 139–150, http://dx.doi.org/10.1016/s0894-1777(01)00046-2.

[29] J.B.S. Haldane, Moments of the distributions of powers and products of normalvariates, Biometrika 32 (3/4) (1942) 226, http://dx.doi.org/10.1093/biomet/32.3-4.226.

[30] T.J. Hanratty, T. Theofanous, J.M. Delhaye, J. Eaton, J. McLaughlin,A. Prosperetti, S. Sundaresan, G. Tryggvason, Workshop findings, Int. J. Mul-tiph. Flow. 29 (2003) 1047–1059, http://dx.doi.org/10.1016/S0301-9322(03)00068-5.

[31] F.M. Henderson, Open Channel Flow, MacMillan Company, New York, USA,1966.

[32] J.O. Hinze, Turbulence, 2nd edition, McGraw-Hill Publ, New York, USA, 1975.[33] A.T. Ippen, R.S. Tankin, F. Raichlen, Turbulence Measurements in Free Surface

Flow with an Impact Tube-Pressure transducer Combination. Technical Report(20, Hydrodynamics Laboratory, MIT, USA,1955, p. 96.

[34] R.J. Keller, A.K. Rastogi, Prediction of flow development on spillways, J. Hy-draul. Div. ASCE 101 (HY9) (1975) 1171–1184.

[35] P.S. Klebanoff, Characteristics of Turbulence in a Boundary Layer with ZeroPressure Gradient. NACA Report 1247, 1955, p. 20.

[36] I. Meireles, F. Renna, J. Matos, F.A. Bombardelli, Skimming, nonaerated flow onstepped spillways over roller compacted concrete dams, J. Hydraul. Eng. ASCE138 (10) (2012) 870–877.

[37] P. Novak, A.I.B. Moffat, C. Nalluri, R. Narayanan, Hydraulic Structures, 2ndedition, E & FN Spon, London, UK 1996, p. 599.

[38] I. Ohtsu, Y. Yasuda, Characteristics of flow conditions on stepped channels, in:Proceedings of the 27th IAHR Biennial Congress, San Francisco, USA, Theme D,1997, pp. 583–588.

[39] S. Okamoto, S. Seo, S. Nakaso, S. Morishita, S. Namiki, Effect of sectional shapeof rib on turbulent shear flow over rows of two-dimensional ribs on a groundplane, in: Proc. 11th Australasian Fluid Mechanics Conference, Hobart Aus-tralia, 1, 1992, pp. 539–542.

[40] S. Okamoto, S. Seo, K. Nakaso, I. Kawai, Turbulent shear flow and heat transferover the repeated two-dimensional square ribs on ground plane, J. Fluids Eng.Trans. ASME 115 (4) (1993) 631–637, http://dx.doi.org/10.1115/1.2910191.

[41] A.E. Perry, W.H. Schofield, P.N. Joubert, Rough wall turbulent boundary layers,J. Fluid Mech. 37 (2) (1969) 383–413, http://dx.doi.org/10.1017/s0022112069000619.

[42] W.R.C. Phillips, J.T. Ratnanather, The outer region of a turbulent boundarylayer, Phys. Fluids A-Fluid Dyn. 2 (3) (1990) 427–434.

[43] N.S.L. Rao, H.E. Kobus, Characteristics of Self-Aerated Free-Surface Flows.Water and Waste Water/Current Research and Practice, Eric Schmidt Verlag,Berlin, Germany, 1971.

[44] H. Schlichting, Boundary Layer Theory, 7th edition, McGraw-Hill, New York,USA, 1979.

[45] L. Toombes, Experimental Study of Air-Water Flow Properties on Low-Gra-dient Stepped Cascades, pH.D. thesis, Dept of Civil Engineering, The Universityof Queensland,, Brisbane, Australia, 2002.

[46] USBR, Design of Small Dams. Bureau of Reclamation, 1st edition, US Depart-ment of the Interior, Denver, CO, USA, 1965, 3rd printing.

[47] H. Wang, F. Murzyn, H. Chanson, Total Pressure Fluctuations And Two-phaseFlow Turbulence In Hydraulic Jumps, Exp. Fluids 55 (11) (2014) 16, http://dx.doi.org/10.1007/s00348-014-1847-9.

[48] I.R. Wood, Air Entrainment in Free-Surface Flows. IAHR Hydraulic StructuresDesign Manual (4, Hydraulic Design Considerations, Balkema Publ, Rotterdam,The Netherlands 1991, p. 149.

[49] I.R. Wood, P. Ackers, J. Loveless, General method for critical point on spillways,J. Hydraul. Eng. ASCE 109 (2) (1983) 308–312.

[50] G. Zhang, H. Chanson, Hydraulics of the Developing Flow Region of SteppedCascades: an Experimental Investigation. Hydraulic Model Report No. CH97/15, School of Civil Engineering, The University of Queensland, Brisbane, Aus-tralia, 2015, p. 76.