You are on page 1of 8

Proceedings of 2013 IAHR World Congress

Steady Vortex Flow in Tangential Intake

Q.S. Qiao
Research Assistant, Dept. of Civil and Environmental Engineering, The Hong Kong University of
Science and Technology, Hong Kong, China. Email: qiaoqs@ust.hk
Joseph H.W. Lee
Chair Professor, Dept. of Civil and Environmental Engineering, The Hong Kong University of Science
and Technology, Hong Kong, China. Email: jhwlee@ust.hk
Lam K.M.
Associate Professor, Dept. of Civil Engineering, The University of Hong Kong, Hong Kong, China.
kmlam@hku.hk

ABSTRACT: This paper presents experimental results of a steady tangential vortex intake flow. The
velocity field is measured with Laser Doppler Anemometry for the first time. It is found that the vortex
intake flow pattern is basically divided into two regimes according to the variable drop shaft inflow
velocity field. For a flow less than the free drainage discharge, the inflow does not interfere with the
swirling flow in the drop shaft; the horizontal and vertical velocity of the inflow in the tapering channel
varies linearly and parabolically with the vertical distance respectively. For larger discharges the inflow
velocity field is notably modified by the drop shaft swirling flow. It is observed that the swirling flow
possesses a stable and asymmetrical 'D' shaped air core with a throat located at the lowest level of the
tapering channel. The tangential velocity of the swirling flow near the throat can be well-approximated by
a Rankine vortex model in the region where the flow layer thickness is of the same order as the junction
width; the vertical velocity of the swirling flow is approximately constant in the radial direction.

KEY WORDS: Vortex flow; Swirling flow; Air Core; Drainage system; Drop shaft.

1 INTRODUCTION
Tangential vortex intakes are commonly used in water supply and sewerage systems. A tangential
intake is a compact hydraulic structure which consists of an approach channel with horizontal bottom and
rectangular cross section, a steep tapering channel, a junction, and a drop shaft. The inflow enters
tangentially into the drop shaft via a tapering channel, and the flow swirls down the drop shaft. The strong
centrifugal effect in the swirling flow results in a stable air core which allows any entrained air to escape.
It is well accepted that the tangential intake is a compact design that possesses the hydraulic advantage of
a stable flow with high energy dissipation in the absence of significant air entrainment. However, the 3D
vortex flow in the tangential vortex intake is complicated and current theoretical models are not
sufficiently complete to interpret the flow process reliably.
Extensive experimental and theoretical investigations have revealed the complexity and subtlety of
the tangential vortex intake flow over half a century. Binnie and Hookings (1948) firstly presented an
analytical approach for predicting the head-discharge relation of swirling flow with air core in a
bellmouth. Brooks and Blackmer (1962) performed a comprehensive experiment study of tangential
intake model for San Diego Ocean Outfall. Jain (1984) firstly proposed a one-dimensional theoretical
model with the concept of uniform open-channel flow for predicting depth-discharge relation and air core
sizes of tangential intake. Yet the analytical approach does not account for the convergence of the
approach channel, the interaction of the approaching flow with the swirling vortex flow, and the
asymmetrical feature of swirling vortex flow. Jain and Ettema (1995) reviewed the main shortcomings of
such analytical model of tangential intake. Zhao et al. (2006) proposed a similar 1D model for predicting
the hydraulic performance of tangential intake. Lee et al. (2006) carried out a hydraulic model study of
tangential intake for Lai Chi Kok Transfer Scheme in Hong Kong. They found that the hydraulic
characteristics of tangential intake heavily depends on its geometry features. More importantly, they
pointed out that the blocking effect of the swirling flow can cause unstable inflow in the tapering channel.
Yet such interaction of swirling flow with inflow has not been further evaluated quantitatively. Based on
experiment results of a series of hydraulic models, Yu and Lee (2009) proposed a stable design criterion
for tangential intake and indicated that the unstable inflow phenomenon will not occur if the intake design
satisfies such criterion. This approach is based on 1D model assuming a symmetrical free vortex flow.
In spite of the extensive experimental and theoretical studies, there still exists several unclarified
issues concerning tangential vortex intake flow. First, the actual inflow pattern from the tapering channel
into the drop shaft (at or near the junction) is unknown. The proposed analytical approach assumed a
constant streamwise velocity in the cross-section. Yet the accelerating inflow into the drop shaft is in fact
three dimensional. Second, the interaction of the swirling vortex flow with the inflow is not well
understood. Third, the tangential vortex intake flow velocity field has not been studied theoretically or
experimentally. As far as we aware, detailed tangential intake flow observations have not been reported,
and there is scant data on the air core size. This paper presents detailed velocity measurements on the
structure of a tangential intake vortex flow for the first time.

2 EXPERIMENTAL SETUP
Figure 1 shows the sketch of the experiment setup in lab. The physical model is fabricated from
Perspex. The steady conveying discharge Q through the model is generated by a water recirculating
system. A two-component DANTECH LDA (Dantec 2006) is utilized for measuring the flow velocity
field with the aid of a specially designed measurement window. All experiments are carried out at steady
flow, and the Reynolds number of the approach channel flow is of order of 105, large enough to maintain
turbulent flow conditions in the model. The water temperature is about 20 oC.
The geometry of tangential intake is mainly determined by the following dimensionless parameters:
e/B, D/B, β0, and β1 where e = junction width, B = approach channel width, D = drop shaft diameter, β0 =
bottom slope of tapering section, and β1 = tapering angle of width of tapering section. For this physical
model, D = 0.124 m, e = 0.031 m, B = 0.124 m, β0= 13o, and β1 = 24o, as shown in Figure 2.
For convenience of presenting experimental results, the flow process in the model is divided into
two regions, simply according with directly observed flow characteristics. One is the inflow bounded by
the straight walls of the tapering channel, and the other is the swirling flow bounded by the circular wall
of the drop shaft. In the approach channel, the flow depth ha is measured with a point gauge. In the
tapering channel, the flow depth at the junction hj is also measured. The inflow velocity field is defined in
a Cartesian coordinate system, and the local horizontal velocity Ux, the transverse velocity Uy
perpendicular to the straight vertical boundary wall, and the vertical velocity Uz are measured for Q = 2.0,
4.0, 6.0, 8.0 and 10.0 L/s.

Figure 1 Experimental setup for measuring vortex flow field with LDA

2
The swirling flow velocity field is defined in a cylindrical polar coordinate system, and the local
vertical velocity Uz, and the tangential velocity Uθ are measured for Q = 4.0 and 8.0 L/s. The average air
core radius ra is measured with a specially designed eight-legged ruler at different levels (Yu and Lee
2009).

3 Experimental results

3.1 General observations of tangential vortex intake flow


The general flow features are firstly observed. For Q < 1.5 L/s, the approach channel flow is critical,
the height of shock waves increases along the inclined wall of the tapering channel, and the inflow issues
freely from the junction, following an approximately helical path down the drop shaft. For 1.5 < Q < 5.5
L/s, the approach channel flow becomes subcritical, the inflow also freely issues into the drop shaft, and
the inflow does not strike the swirling flow near the junction as it turns around. A stable air core with a 'D'
shape forms, and the air core throat size decreases with increasing Q. For Q > 5.5 L/s, the free surface of
the inflow is relatively flat, and the swirling flow strikes with the low part of the inflow near the junction.
According to the definition of control shift discharge Qc and free drainage discharge Qf of tangential
intake proposed by Yu and Lee (2009), Qc ≈ 1.5 L/s and Qf ≈ 5.5 L/s, and this model satisfies stable
design criterion.
Figure 3 shows the measured flow depth ha and hj under varying discharge Q, and the prediction
with the theoretical model proposed by Yu and Lee (2009). It is seen that ha approximately varies linearly
with Q for large discharge, and the predicted depth ha agrees well with the measured. The linear
depth-discharge relation for large Q is similar to that of a scroll vortex intake (Acker and Crump 1960).

Figure 2 Main geometry parameters and measured vortex flow variables

3
(a) (b)
Figure 3 (a) depth-discharge relation (b) Air core throat area ratio –discharge relation
Figure 4 shows the measured air core size of swirling vortex flow in the drop shaft. In Figure 4 (a)
it is seen that the air core is significantly asymmetrical about the axis of the drop shaft, especially
the thickness of the flow layer is almost zero in the region of swirling angle θ = 2700 for Q = 4.0 L/s. In
fact no swirling flow passes through this region under small discharge in experiments. To overcome the
difficulty of asymmetrical feature of swirling flow, the concept of average air core area ratio λm is utilized,
which is defined by

1 8 2
λm = ∑r
8 R 2 i =1 ai
(1)

where R = D/2, radius of drop shaft, rai= air core radius at θ = (i−1)π/4. Figure 3 (b) shows the
measurement of air core throat area ratio, and the prediction by the model of Yu and Lee (2009). It is
found that the measured ratio λm is basically consistent with the predicted. Figure 4 (b) shows the
variation of λm with the level z. For Q= 4.0, 6.0, 8.0, 10.0 L/s, there obviously exists a minimum λm at
certain level z - which increases with increasing Q.

(a) (b)
Figure 4 (a) Measured air core shape at z=-0.04 m; (b) Measured air core area ratio

4
3.2 Tapering Channel Inflow into Drop Shaft
Generally speaking, the inflow is a 3D free surface flow that depends on intake geometry and the
boundary condition at the junction. The main streamwise flow is characterized by the horizontal velocity
Ux, which determines the main flow quantities such as flow volume flux and angular momentum about
axis of drop shaft. The transverse velocity Uy is caused by the convergence of the tapering section.
The temporal mean velocities Ux, Uy and Uz are measured at the predetermined measuring point.
This is also the case for the downstream swirling flow in the drop shaft.
Figure 5 shows the measured velocity profiles of Ux(z) for Q = 2.0, 4.0, 6.0, 8.0, and 10.0 L/s at x =
0.0 m, i.e. just at the junction. It is observed that near the junction Ux varies approximately linearly in the
vertical direction, and the variation of Ux in the transverse direction is relatively small for Q = 2.0, 4.0 L/s.
Note that the free drainage condition holds in these two cases. For larger flow the linear velocity profile
of Ux(z) is notably modified close to the bottom of the inlet channel. The bottom boundary layer is
relatively thin; suggesting that the viscous effect on the bulk inflow pattern is negligible.
The distribution of the transverse velocity Uy is measured separately due to the constraints of the
physical model. Figure 6 (a) shows the measured velocity profile Uy in the y direction for Q = 4.0 and 8.0
L/s. It is seen that Uy increases approximately linearly in the positive y direction. Consider the boundary
condition at vertical boundary wall and impose a slip boundary condition at the tapering wall. The
distribution of Uy(y) should satisfy the boundary condition that Uy = 0 at y = 0 and Uy = Ux tan(β1) at y = e.
The measured distribution of Uy can be reasonably approximated by a linear variation
Uy y
= tan β1 (2)
Ux e
Figure 6 (b) shows the measured velocity profile Uz(z) at x = 0.0 and y = 0.009 m for Q = 2.0 , 4.0,
8.0 and 10.0 L/s. Note that the free drainage condition holds only for Q = 2.0 and 4.0 L/s. It is found that
the measured velocity profiles of Uz(z) has an approximately parabolic shape. In the upstream region of
the junction, similar parabolic distributions of Uz(z) have also been observed under these two discharges.
The parabolic variation of the vertical velocity is a direct consequence from the linear variation of the
vertical velocity based on the continuity equation. For Q=8.0 and 10.0 L/s, the velocity profiles Uz(z)
become affected by the drop shaft flow, and deviate from the parabolic shapes.

(a) y=0.009 m, x=0.0 m (b) y =0.02 m, x=0.0 m


Figure 5 Measured velocity distribution Ux(z) at the junction

5
(a) Uy(y) (x=0.045 m, z=0.05 m) (b) Uz(z) at x=0.0 m and y=0.009 m
Figure 6 Measured velocity profiles Uy(y) and Uz(z)

The pressure field of the inflow is also measured by a set of manometers installed on the vertical
wall of the tapering section. Assuming energy conservation, the pressure field can also be inferred from
the measured velocity field by the Bernoulli equation. Figure 7 shows the measured pressure head at y ≈
0.0 m an x=0.0 m and the computed head with the measured inflow velocity for Q = 4.0 and 8.0 L/s. It is
seen that the computed pressure head is reasonably close to the measured. This indicates the observed
inflow pattern is consistent with theory and gives support to the experimental measurements.

Figure 7 Measured pressure head at x=0.0 m (ρ=density, g=gravity acceleration)

3.2 Swirling flow field in drop shaft


Generally, the fluid elements of the swirling flow follow an approximately helical path down the
drop shaft. The local flow velocity can be decomposed into vertical component Uz and the tangential
component Uθ. Both Uz and Uθ are measured with LDA in the neighboring region of the air core throat.
Figure 8 shows the measured tangential and vertical velocity profiles at the levels z = −0.04,-0.06,
−0.08 and −0.10 m at θ=450 for Q = 4.0 L/s. The vortex flow has clearly identifiable features. First, in this
region the thickness of the boundary layer is relatively thin when compared with that of the bulk vortex
flow. Secondly the tangential velocity Uθ approximately satisfies a Rankine-combined vortex model
(allowing slip at the wall boundary). From near the wall to a certain radius rc, Uθ is inversely proportional
to r, and near the air core Uθ is approximately proportional to r. The vertical velocity Uz increases down

6
the drop shaft, and is quite uniformly distributed in the radial direction, and slightly overshoots near the
free surface of air core.
For better identifying the tangential velocity distribution characteristics, Figure 9 shows the vortex
constant Г=rUθ at these four levels and the computed Гa according to the theory proposed by Yu and Lee
(2009). It is seen that Г keeps approximately constant in the bulk flow region, and the value of Гa is
approximate to the measured at this location. At θ=980, it is observed the swirling flow pattern is similar
yet Г becomes larger than Гa (not shown).
The experiment shows the real swirling flow pattern in the drop shaft, which is more complicated
than that assumed in the previous model.

(a) Uθ (b) Uz
Figure 8 Measured tangential and vertical velocity of vortex flow at θ=450 for Q=4.0 L/s

Figure 9 Measured vortex constant at θ=450 for Q=4.0 L/s

4 Discussion
The experiment results reveal several interesting features of the vortex flow pattern in typical
tangential intake for the first time. The real vortex flow pattern is much more complex than the idealized
symmetrical flow pattern normally assumed in 1D analytical models (Jain 1984; Yu and Lee 2009).
The inflow behaves as a half-slot jet (with the straight channel wall as the symmetry boundary) for
large discharge. In the analytical model proposed by Jain (1984), a 1D model with constant cross section
velocity is assumed for the supercritical flow in the tapering channel. However, the effect of the three

7
dimensional shape of the tapering channel and the pressure boundary condition at the junction on the
inflow pattern has thus far not been considered. The experiment shows that the bulk inflow velocity
increases from the low level to the high level near the junction. It indicates that the potential energy of the
fluid elements especially at the low level has been largely transformed into the kinetic energy. As the
issuing flow from the junction starts to swirl down the drop shaft, its outer side is bounded by the circular
boundary wall, while its free surface at the inner side directly attains atmosphere pressure. The pressure
field can be well-approximated by a centrifugal force balance for a free vortex. The inflow pattern is
characterized by a typical slot jet rather than an open-channel flow.
The swirling vortex flow in the drop shaft basically satisfies the Rankine vortex model. Jain (1984)
simply assumed a constant tangential velocity of the swirling flow in the radial direction – which is
clearly not supported by the measurements. Yu and Lee (2009) utilized a free vortex model to denote the
tangential velocity distribution. Experiment results show that this is a good approximation outside of a
thin shearing layer near the boundary wall, and away from the air core.
The viscous effect of the vortex flow is relatively minor for such tangential intake satisfying stable
design criterion. Our experiments show that the energy near the air core throat is reasonably equal to the
energy head of the upstream approach channel flow.

5 Concluding Remarks
This experiment focuses on the vortex flow pattern of a typical tangential intake. The flow velocity
field in the tapering channel and in the drop shaft, and air core sizes are measured for a wide range of
discharge flows. The inflow in the tapering channel can be divided into two regimes, relating with the
free drainage condition at the junction. The inflow from the tapering inlet channel into the drop shaft is
highly three-dimensional, with a linear vertical variation of horizontal (x) velocity, and a parabolic
vertical variation of vertical velocity (z); the transverse velocity varies linearly in the transverse (y)
direction. The tangential velocity of the swirling vortex flow in the drop shaft can be well-approximated
by a Rankine vortex model; the vertical velocity is approximately constant in the radial direction. The
present study offers comprehensive insights of tangential vortex intake flow, and provide the basis for
developing a theory for the hydraulics of such vortex structures.

Acknowledgement: This work was supported by the Hong Kong Research Grants Council (RGC
HKU/HKUST 714309).

References
Acker, P., and Crump, E.S. 1960. The vortex drop. Proceedings of Institute of Civil Engineers, Part I, 16(4), 433-442.
Binnie, A.M., and Hookings, G.A. 1948. Laboratory experiments on whirlpools. Proceedings of the Royal Society of
London, Seires A, Volume 194(1038), 398-415.
Brooks, N.H., and Blackmer, W.H. 1962. Vortex energy dissipator for San Diego ocean outfall - Laboratory
investigation. Rep. No. KH-R-5, W.M. Keck Laboratory of Hydraulics and Water Resources, California Institute of
Technology, Pasadena, Calif.
Dantec Dynamics. 2006. BSA flow software. Publication No. 9040U5734, Tonsbakken 18, DK-2740 Skovlunde,
Denmark.
Jain, S.C. 1984. Tangential vortex-inlet. Journal of Hydraulic Engineering, Volume 110 (12), 1693-1699.
Jain, S.C., and Ettema, R. 1995. Swirling flow problems at intakes. in Vortex Inatkes, IAHR hydraulic structures
design manual 1, Jost Knauss eds., Balkema, Rotterdam, The Netherlands, 125-137.
Yu, D.Y. and Lee, J. H.W. 2009. Hydraulics of tangential vortex intake for urban drainage. Journal of Hydraulic
Engineering, Volume 135 (3), 164-174.
Lee, J. H.W., Yu, D.Y., and Choi, K.W. 2006. Physical hydraulic model tests for Lai Chi Kok transfer scheme – intake
structure. Croucher Laboratory of Environmental Hydraulics, The University of Hong Kong.
Zhao, C.H., Zhu, D., Sun, S.K., and Liu, Z.P. 2006. Experimental study of flow in a vortex drop shaft. Journal of
Hydraulic Engineering, Volume 132 (1), 61-68.

You might also like