Você está na página 1de 11

Home

Search

Collections

Journals

About

Contact us

My IOPscience

Validation of mathematical models for predicting the swirling flow and the vortex rope in a
Francis turbine operated at partial discharge

This content has been downloaded from IOPscience. Please scroll down to see the full text.
2010 IOP Conf. Ser.: Earth Environ. Sci. 12 012051
(http://iopscience.iop.org/1755-1315/12/1/012051)
View the table of contents for this issue, or go to the journal homepage for more

Download details:
IP Address: 189.79.68.93
This content was downloaded on 04/05/2014 at 00:07

Please note that terms and conditions apply.

25th IAHR Symposium on Hydraulic Machinery and Systems


IOP Conf. Series: Earth and Environmental Science 12 (2010) 012051

IOP Publishing
doi:10.1088/1755-1315/12/1/012051

Validation of mathematical models for predicting the


swirling flow and the vortex rope in a Francis turbine
operated at partial discharge
P A Kuibin1,2, V L Okulov1,3, R F Susan-Resiga4 and S Muntean5
1

Kutateladze Institute of Thermophysics, Siberian Branch of Russian Academy of


Science Lavrentyev Ave. 1, Novosibirsk, 630090, Russia
2
Novosibirsk State University, Pirogova Str. 2, Novosibirsk, 630090, Russia
3
Department of Mechanical Engineering, Technical University of Denmark, Lyngby,
Denmark
4
Department of Hydraulic Machinery, Politehnica University of Timioara
Bv. Mihai Viteazu 1, RO-300222, Timioara, Romania
5
Centre of Advanced Research in Engineering Sciences, Romanian Academy Timioara
Branch Bv. Mihai Viteazu 24, RO-300223, Timioara, Romania
E-mail: kuibin@itp.nsc.ru
Abstract. The vortex rope in a hydro turbine draft tube is one the main and strong sources of
pulsations in non-optimal modes of hydro turbine operation. We examine the case of a Francis
turbine model operated at partial discharge, where a strong precessing vortex rope is developed
in the discharge cone downstream the runner. Available experimental data provide the
circumferentially averaged axial and circumferential velocity profiles, as well as the vortex
rope geometry, precessing frequency, and the level of pressure fluctuation at the wall. The
mathematical models presented in this paper can correctly recover all this information without
actually computing the full three-dimensional unsteady flow in the hydraulic turbine. As a
result, we provide valuable mathematical tools for assessing the turbine behaviour at off-design
operating regimes in the early stages of runner design, with computational effort several orders
of magnitude less than the current approaches of simulating the complex turbine flow.

1. Introduction
The vortex rope in a hydro turbine draft tube is one of the main and strong sources of pressure pulsations in
non-optimal modes of hydro turbine operation. The development of the vortex rope in the turbine discharge cone
is directly related to the self-induced instability of the decelerated swirling flow downstream the turbine runner.
For the past decade this conclusion was directly supported by the three-dimensional unsteady flow computations
of Ruprecht et al. [1], Sick et al. [2], Ciocan et al. [3], Zhang et al. [4], where an axi-symmetric swirling flow at
runner outlet becomes unstable, with the development of a precessing helical vortex, while being decelerated in
the turbine discharge cone. Although such numerical simulations are in good agreement with experimental data
for both unsteady pressure and velocity fields, the computing effort is too large both in terms of computing
resources and computing time for turbine design and optimization purposes.
A more tractable approach for investigating the swirling flow in the turbine discharge cone is to use a
simplified axi-symmetric flow model since the geometry is axial-symmetric. Such a turbulent, axi-symmetric
swirling flow can recover a circumferentially averaged flow field, [5], in very good agreement with experimental
data for axial and circumferential velocity components. Moreover, this simplified model captures a central
stagnant region in agreement with the qualitative model proposed by Nishi et al. [6], and the vortex rope is
located in the region of large shear between the stagnant region and the annular swirling flow. In comparison
with three-dimensional unsteady flow simulation, the axi-symmetric flow computation is obviously several
orders of magnitude less expensive in terms of computing time and resources. However, the main drawback of
this approach is that the circumferentially averaged flow provides a steady velocity and pressure fields, without
any direct indication on the real flow unsteadiness. A stability analysis of this axi-symmetric base flow could

c 2010 IOP Publishing Ltd




25th IAHR Symposium on Hydraulic Machinery and Systems


IOP Conf. Series: Earth and Environmental Science 12 (2010) 012051

IOP Publishing
doi:10.1088/1755-1315/12/1/012051

eventually provide additional information with respect to the most amplified perturbations and their dominant
frequency, but it cannot estimate the level of pressure fluctuations associated with the flow unsteadiness.
A mathematical model to be used for analysis of swirling flow with spiral vortex core in a pipe has been
developed by Wang and Nishi [7]. Their quasi-three dimensional model considers a superposition of an
axisymmetric base flow with the velocity field induced by a helical vortex filament. The base flow is divided in a
main flow in an annular section, with a central dead water region. The main flow has constant axial velocity and
free-vortex circumferential velocity distributions. On the other hand, Wang and Nishi consider a rigid body
rotation (linear) for circumferential velocity and a parabolic variation for the axial velocity in the central dead
water region. After superimposing the helical vortex on this base flow, the theoretical results for the velocity
profiles, both with vortex rope and circumferentially averaged, as well as the precessing frequency, are shown in
[7] to be in a reasonable agreement with experimental data.
The theory used in our investigations of precessing helical vortices in swirling flows was developed by
Alekseenko et al. [8, 9] up to analytical solutions for velocity and pressure fields in cylindrical and conical
geometries. A further step toward practical applications in hydraulic machines is presented by Kuibin et al. [10],
where it is shown that the vortex rope geometry, precession frequency, as well as the wall pressure fluctuations
can be computed given a set of swirling flow integral quantities.
In this paper we couple a theory for computing the axisymmetric swirling flow at the outlet of Francis turbine
runners operated far from the best efficiency regime with the theory of precessing helical vortices in order to
provide an integrated methodology that can be used in practice for turbine design and optimization, as well as for
rapid evaluation of existing runners. First, we introduce the constrained variational problem for computing the
axial and circumferential velocity profiles at the Francis runner outlet, and provide an example for the FLINDT
Francis turbine operated at 70% from the best efficiency discharge. These velocity profiles are validated against
Laser Doppler Velocimetry measurements. Second, we employ the theory of helical vortices [10] to compute the
vortex rope precessing frequency as well as the pressure fluctuations, and the results are validated against
corresponding experimental data [3].

2. Swirling flow at the Francis runner outlet


The swirling flow at the outlet of the hydraulic turbine runner essentially influences the overall pressure
recovery and hydraulic losses in the turbine draft tube. Susan-Resiga et al. [11] introduced an empirical
analytical representation for both axial and circumferential components of the swirling flow ingested by the draft
tube of a Francis turbine, leading to a simple parametric description of velocity experimental data. However,
these formulae are valid and can accurately represent the experimental data only in the neighborhood of the best
efficiency point, i.e. 10% the best efficiency discharge. This approach is further refined by Tridon et al. [12]
who examine the radial velocity component as well, besides the axial and circumferential components. When the
turbine discharge is further decreased, say at 70% the best efficiency discharge, the swirling flow at runner outlet
changes significantly, with a quasi-stagnant region near the symmetry axis and the flow is confined in an annular
section close to the wall [3, 11]. As a result, the analytical representation of the velocity profiles which uses a
superposition of Batchelor vortices [11, 12] cannot be used for operating regimes with partial discharge when the
precessing vortex rope is developed. Instead, Susan-Resiga et al. [13] introduce a mathematical model that allow
the computation of swirling flow at the Francis runner outlet for a large range of operating points, and we show
that this model is able to capture the main flow features even at low partial discharge.
2.1 Flow kinematics downstream a fixed-pitch runner
Traditionally, the geometry of runner blades in the neighborhood of the trailing edge is chosen such that the
relative flow at runner outlet has a certain distribution from hub to shroud of the relative flow angle, , defined
as the angle between the relative velocity vector and the tangential direction,

tan =

Uz
R U

(1)

where Uz and U are the axial and circumferential absolute velocity components at the radius R, and is the
runner angular velocity. For fixed pitch runners, the relative flow angle remains practically constant over a
rather large operating range, particularly for Francis turbines where the large number of blades prevents severe
flow detachment at trailing edge. From the velocity triangle at runner outlet, the tangential velocity U is positive
at partial discharge, i.e. co-rotating swirl with respect to runner rotation, and is becomes negative at full load, i.e.
the swirl rotates in opposite direction with respect to the runner. In between, there is a regime where the

25th IAHR Symposium on Hydraulic Machinery and Systems


IOP Conf. Series: Earth and Environmental Science 12 (2010) 012051

IOP Publishing
doi:10.1088/1755-1315/12/1/012051

tangential velocity vanishes, thus we call the corresponding axial velocity the swirl-free velocity U0, which is
defined according to eq. (1) as tan = U0/( R ). As a result, instead of using the relative flow angle to describe
the swirling flow kinematics at runner outlet, we are going to use the swirl-free velocity profile U0 ( R ), defined
as

U0 =

Uz
1 U ( R )

(2)

The above definition can be re-written in dimensionless form using a reference radius Rref and a reference
velocity Rref,

u0 = ( ru z ) ( r u )

where u = U /( Rref ) and r = R / Rref are the dimensionless velocity and radius, respectively. From eq. (3), the
circumferential velocity can be written as function of the axial velocity profile, for given swirl-free velocity
profile,
u = r (1 u z u0 )
(4)
Equation (4) has the advantage of avoiding the singularity at the axis, where u0 remains finite.
As an example, the swirl-free velocity profile can be determined using the measured axial and
circumferential velocity at runner outlet for a Francis turbine for several operating points within the discharge
range of 70%...110% from the best efficiency discharge [3, 5, 101]. Figure 1 shows the experimental data for 7
operating points, translated in u0 values using eq. (3). It can be seen that a simple parabolic fit can correctly
represent the u0(r) profile, which depends directly on the runner blades design at the trailing edge.
0.6

0.4

0.5

dimensionless velocity components

dimensionless swirlfree velocity

experimental data
parabolic fit

0.4

0.3

0.2

0.1

0.2

0.4
0.6
dimensionless radius

0.8

0.3

0.2

0.1

Fig. 1 Swirlfree velocity distribution at runner outlet.

axial velocity measured


swirl velocity measured
axial velocity computed
swirl velocity computed

0.2

0.4
0.6
dimensionless radius

0.8

Fig. 2 Axial and circumferential velocity profiles


at runner outlet. LDV measurements [3] correspo
nd to the survey section located at z = 1.426

2.2 Integral quantities for swirling flow


The swirling flow if characterized by two main integral quantities. The first one is the volumetric flow rate Q
defined as
Rw

2R dR

(5)

where Rw is the wall radius. In dimensionless form we introduce the discharge coefficient,

rw

( Rref ) Rref2

= uz d ( r 2 )

(6)

The second integral quantity for swirling flows is the flux of moment of momentum,

Rw

( RU )U

which can be written in dimensionless form using eq. (4) as,

2R dR

(7)

25th IAHR Symposium on Hydraulic Machinery and Systems


IOP Conf. Series: Earth and Environmental Science 12 (2010) 012051

( Rref )

2
ref

IOP Publishing
doi:10.1088/1755-1315/12/1/012051

rw

rw

2
( ru ) uz d ( r ) = uz 1

uz 2
2
r d (r )
u0

(8)

For a given operating point of the turbine, both and m can be determined, and these two values characterize
globally the swirling flow, in addition to the local kinematical condition from eq. (4).
2.3 Constrained variational problem for axial velocity profile
Benjamin [14] introduced the concept of the flow force, which we define for swirling flow in a pipe as,

Rw

( U

+ Pw P )2R dR

2
z

(9)

where P is the static pressure, with the corresponding value at the wall Pw. Benjamin showed that the flow force
defined in eq. (9) is minimized for admissible swirling flow configurations. The radial distribution of the static
pressure is related to the circumferential velocity through the radial equilibrium equation,

1 dP U 2
=
dR R

(10)

which, using eq. (4), can be rewritten in dimensionless form as

dp

1 u
= 1 z d ( r 2 )
2 u0

dP
( Rref )

(11)

By integrating this differential equation we obtain the dimensionless pressure departure with respect to the wall
pressure,

Pw P

( Rref )

1 w u
= 1 z d ( r 2 )
2 r u0

(12)

Using eq. (12), the flow force can be set in dimensionless form involving only the axial velocity profile.
Moreover, one can see from eqs. (6), (8), and (12) that we can introduce a new radial variable corresponding to
the dimensionless radius squared, y r2. In addition, the flow force can be made dimensionless as f F/[(Rref)
2
2
Rref
].
We arrived at the variational formulation for the swirling flow downstream the Francis turbine runner, as
follows: given the and m, as well as the swirl-free velocity profile u0, find the axial velocity profile uz which
minimizes the flow force

f =

yw

2
uz dy +

ys

1
2

yw yw

ys

u
y 1 u0z dx dy

(13a)

subject to the variational constraints,

yw

m=

dy ,

ys

yw

ys

uz
1 y dy
u0

(13b)

Note that in eqs. (13) the lower limit of the integrals has been set to ys instead of 0. This accounts for the
possibility that the swirling flow develops a central stagnant region, and ys 0 is an additional unknown of the
problem.
2.4 Numerical example and validation
The nonlinear constrained variational problem (13) is solved numerically using the NCONF subroutine from
the IMSL library, together with the QDAG and TWODQ subroutines for simple and double integrals evaluation.
The unknown function uz(y), defined on the interval ys y yw, is approximated using a B-spline interpolation
with N equally spaced nodal values.
Figure 2 shows an example of swirling flow configuration for a Francis turbine operated at 70% the
discharge at best efficiency point, and nominal head. The swirl-free velocity profile is the one shown in Fig. 1,
u0 = 0.312+0.096 y, and the integral quantities for the swirl have the values, computed from experimental data,
= 0.264 and m = 0.0483. The dimensionless wall radius is rw = 1.063 since the survey section where velocity
profiles are measured is a bit downstream the turbine throat / runner outlet. The resulting central stalled region

25th IAHR Symposium on Hydraulic Machinery and Systems


IOP Conf. Series: Earth and Environmental Science 12 (2010) 012051

IOP Publishing
doi:10.1088/1755-1315/12/1/012051

has the radial extent of rs = 0.363.


From Fig. 2 we can conclude that the above simplified mathematical model, which does not account for
viscous and three-dimensional effects in the inter-blade channels of the runner, can capture the main features of
the swirling flow at runner outlet. First, in correctly predicts the extent of the stalled region. Second, it agrees
well with the axial and circumferential velocity profiles. The increase in both axial and tangential velocity shown
in measurements is associated with the inter-blade vortex developed at the junction of the blade with the band.
Of course, the present model cannot account for these secondary flows. On the other hand, we can provide a
good approximation of the swirl at runner outlet, without actually computing the runner flow, thus evaluating the
flow behavior in the discharge cone for a wide range of operating range. This is quite useful in the early design
and optimizations stage, since it provides a quantitative tool to assess various design choices with respect to the
blade trailing edge.

3. Precessing vortex rope


The axi-symmetric approach can correctly predict the circumferentially averaged velocity profiles, but it
cannot describe three-dimensional vortex structure arising behind the turbine runner operating at partial load. A
proper model can be developed using the theory of helical vortices [8, 9]. The possibility for realization of such
idea is evident from paper by Kuibin et al. [10] where both frequency and amplitude of pressure pulsations
generated by the precessing vortex rope were found through given integral characteristics: vortex intensity, liquid
flow rate, momentum and moment of momentum fluxes. Here we present a method for evaluating the vortex
rope parameters close to the mentioned one. Instead of using the integral fluxes we will estimate the geometrical
and dynamical vortex characteristics from direct comparison of the measured circumferentially averaged
velocity profiles with dependencies inherent to the model of helical vortex.
3.1 Helical symmetry
The first step in finding the vortex rope configuration for a given circumferentially averaged swirling flow is
to examine the helical symmetry property of the flow. It is shown in [8-1.5.2] that the helical symmetry
condition implies

r
u z + u = u z axis = constant
l

(14)

where uz axis is the axial velocity value at the axis, and the characteristic length l is related to the axial pitch h by
the relationship h = 2l. In our case, both the experimental data and the simplified model display a stagnant
region near the axis, thus the right-hand side in eq. (14) vanishes, uz axis = 0.
Figure 3 shows the plot of ru versus uz in order to determine the geometrical parameter l in eq. (14). In
particular we are interested in the region close to the axis with largest radial gradients of both axial and
circumferential velocity components, where the vortex rope is developed. By fitting the experimental data within
the vortex rope region with a line passing through origin we obtain lexp = 0.329, as shown with the black dashed
line. On the other hand, for the simplified model presented in Section 2 eq. (14) is trivial in the stagnant region
with r < rs. However, we can employ an analytical continuation of the numerical results for the simplified
swirling flow model, as shown with the red dashed line, resulting in lnum = 0.311. The corresponding
dimensionless pitch values of the helical symmetry are hexp = 2.070 and hnum = 1.952, respectively.
On the other hand, the swirling flow with vortex rope modeled in this paper has been investigated experimentally
by Ciocan and Iliescu [15] using a 3D-PIV method. They identified the vortex rope core from velocity
measurements and proposed a geometrical description as a conical logarithmic spiral of the vortex rope shape, in
very good agreement with experimental data. As one can see from Fig. 4 the actual vortex rope develops in the
draft tube cone, and it is wrapped on a cone with half-cone angle of rope = 17. This cone originates on the
runner crown, where it has a radius of r0 = 0.09 for the axial coordinate value z0 = 0.615. As a result, the
distance between the vortex rope core and cone axis increases downstream as rrope (z) = r0 (z z0) tan rope . The
discharge cone shown in Fig. 4 starts at z = 1 and ends at z = 1 3 = 2.732, and has a half-cone angle of
cone = 8.5, [5, 15].
For the survey section located at z = 1.426 where velocity measurements shown in Fig. 2 were performed,
the dimensionless radial distance of the vortex rope core from the axis is rrope = 0.338, quite close to the stagnant
region radius in the simplified model rs = 0.363. This confirms the conclusion of Susan-Resiga et al. [5] that the
vortex rope is located at the boundary between the main swirling flow and the central stagnant region.

25th IAHR Symposium on Hydraulic Machinery and Systems


IOP Conf. Series: Earth and Environmental Science 12 (2010) 012051

IOP Publishing
doi:10.1088/1755-1315/12/1/012051

0.35
experimental data
l = 0.329485 in Eq.(14)
simplified model
l = 0.310629 in Eq.(14)

radius * circumferential velocity

0.30
0.25
0.20
0.15
0.10
0.05
0.00
0.35

0.30

0.25

0.20 0.15
axial velocity

0.10

0.05

0.00

Fig. 4 Vortex rope core measured by Ciocan and


Iliescu [15] and fitted as a conical
logarithmic spiral.

Fig. 3 Helical symmetry in the vortex rope region.


Points correspond to the data in Fig. 2.

The axial pitch of the conical vortex rope shown in Fig. 4 increases as the flow evolves downstream into the
turbine discharge cone. From the mathematical model of the conical vortex rope [15] we find that the pitch
increases with the rope radius as hrope (z) = rrope (z) ln(b) / tan rope = r0, where b = 3.2 is the rate of radial growth
for a complete rotation. With the data above, the rope pitch increases linearly as hrope (z) = | 0.373 + 1.163z |.
However, one should bear in mind that the vortex rope model employed in this paper considers a helical vortex
in a cylindrical pipe, and does not account for the conical shape of the vortex rope actually developed in the
turbine discharge cone. As a result, for the purpose of comparison with the present simplified model it seems
reasonable to consider a mean value for the actual vortex rope pitch, computed within the discharge cone,
hrope = 1.8, which is slightly smaller than the values obtained from the helical symmetry considerations above.
3.2 Model of helical vortex
At the next step we evaluate the vortex rope parameters by the comparison of the measured velocity profiles with
model ones [8, 9]

u =

G ( r ), u z = u z axis +
G (r ) ,
2r
2l

G (r ) =

1
Sz

0, r < r
dS
r r
Sz

1,

(14)

The integral in (14) equals the intersection area of the circle of radius r with the horizontal cross-section Sz of the
helical tube modeling the vortex core. The ratio of areas in (14) does not change, if both figures are projected on
the plane normal to the helical axis of the vortex. The core cross-section in this plane is the circle of radius and
instead of the circle of radius r we obtain the ellipse prescribed in polar coordinates (, ) by the formula

( a + cos )

l 2 + ( sin )

(a

+ l2 ) = r2 l2

Therefore, we find that G(r) = S(0)/2, where S(0) is the intersection area of this ellipse with the circle = .
From the experimental data arrays we extract the interval of points with radii from 0.25 till 0.5 which
corresponds to the zone with pronounced helical symmetry in Fig. 3. The vortex intensity equals the maximum
value of circulation 2ru at the chosen interval, i.e. = 0.531. The remaining two parameters, helix radius, a,
and radius of vortex core, , are determined by minimizing the function
2
2


2l ( u z i u z axis )
2ru
i i
G (ri ; l ; a, )
i G(ri ; l; a, ) +

As result we find the following values: vortex core radius = 0.126; helix radius a = 0.349, which is close to
the value rrope = 0.338 from [15].

25th IAHR Symposium on Hydraulic Machinery and Systems


IOP Conf. Series: Earth and Environmental Science 12 (2010) 012051

IOP Publishing
doi:10.1088/1755-1315/12/1/012051

dimensionless velocity components

0.4
axial velocity measured
swirl velocity measured
axial velocity vortex rope model
swirl velocity vortex rope model

0.3

0.2

0.1

0.2

0.4
0.6
dimensionless radius

0.8

Fig. 5 Axial and circumferential velocity profiles at the runner outlet and computed
with the model of helical vortex. Experimental data are the same as in Fig. 2.
The corresponding velocity profiles and their comparison with the measured ones shown in Fig. 5 give good
enough fitting of the experimental data only in the inner area of the flow. It looks as the flow induced by the
helical vortex immersed into ambient axi-symmetric flow. Strictly speaking, such superposition would be valid
only in the case when the pitch of vortex lines remains the same in whole area of the flow. Unfortunately, one
can see from Fig. 3 that there exist zones with different pitch. Thus, one should consider this model as some
engineering approach rather than a rigorous one.
A further analysis of the velocity profiles shown in Fig. 5 clearly reveals that for radius values larger than the
region with large velocity gradients which corresponds to the vortex rope location the swirling flow quickly
approaches the irrotational free vortex. The thin black dashed lines from Fig. 5 correspond to the constant axial
velocity uz = /2l and u = /2r, according to eq. (14), with = 0.531 and l = 0.329 for r a = 0.349. It is
clear that the helical vortex model correctly captures the region of the steep gradient in both axial and
circumferential velocity components, while in the main swirl up to the wall it captures only the irrotational part
of the swirling flow with constant specific energy. The rotational part of the swirling flow depends on the blade
runner design, described by the swirl-free velocity u0 in Section 2, and it is not accounted for in the helical
vortex model.
3.3 Frequency of helical vortex precession
Now we can calculate the frequency of the vortex precession with formula derived by Kuibin and Okulov
[16] rewritten here in the following form

vzaxis
l

l
1 3 3 9 73
1

2
2
2

3 + 3 log(1 a% )(1+ ) + + g1 ( ) + log + g2 ( )


2 2
2l a% 1 12
2

where g1 ( ) =

(15)

1 + 1.455 + 1.7232 + 0.7113 + 0.616 4 1

+ 0.4862 + 1.1763 + 4
4

The parameter = l/a means the relative torsion of a helical line; = (1 + 2)1/2, = l/r,
= (1 + 2)1/2, a% = exp[( )/ ( )/]( + )/( + ). The function g2() describes the impact into the
frequency from the non-uniform vorticity distribution in the core,

g2 ( ) =

1 42

w2 d
4 2 0

when a fraction of the total vorticity is concentrated inside the core. Here w denotes the local circumferential
velocity in the vortex core. At uniform vorticity distribution g2() 0. For the Scully vortex [9] with local
distribution of type (r2 + 2)2 we have

g2 ( ) =

1 1
+ + log (1 )
4 2 2

(16)

One can consider some other structures of the vortex core. Nonetheless, the Scully vortex is the most apropriate
model for highly turbulent flow [17] which takes place in the hydroturbine draft tube.

25th IAHR Symposium on Hydraulic Machinery and Systems


IOP Conf. Series: Earth and Environmental Science 12 (2010) 012051

IOP Publishing
doi:10.1088/1755-1315/12/1/012051

When all the parameters of the helical vortex found in 3.1 and 3.2 are used in eq. (15) we find that the
uniform core frequency equals 0.184, much smaller than the measured dimensionless frequency rope = 0.3. In
the same time applying eq. (16) with 90% of the total vorticity in the core yields rope = 0.305, which is
practically the same as in the experiment [3]. Note that the dimensionless rope frequency is defined here as
rope rope / , i.e. it represents the ratio between the rope precession angular speed and the runner angular
speed.
3.4 Pressure oscillations induced by the helical vortex
The model approach for evaluation of amplitude of the pressure pulsations due to precession of the helical
vortex in cylindrical tube was developed by Kuibin et al. [10]. Authors used assumptions on inviscid flow in the
tube in the presence of helical vortex with uniform core. This allowed integrating the Euler equations and to
derive the analytical relation between the pressure distribution and the velocity field and stream function. In
particular, the following formula for pressure pulsations at the tube wall was obtained

v 2 + vz2
.
p% = const lvz +

2 r==R0

The final formula for doubled amplitude (i.e. peak-to-peak) of the dimensionless pressure pulsations reads

p% =

2 2a%
2a%2
1 a% 2l
2
2

%
g
,
log
1
g
,
log
1
a

+
+
+

(
)
(
)
(
)

3
3

22l 2 1 a%2
1 a%2
1+ a%

where g 3 ( , ) =

(17)

1 7 3 9 73 9

+

24 3 3

The result of calculation with these formulae is p% = 0.027 .


In practice, the level of pressure fluctuations is related to the turbine head, as
2
P% ( Rref )
P%
p%
p%
=
=
=2
2
1
2
gH
gH gH ( Rref )

2 ( Rref )2

where is the energy coefficient for the turbine, with = 1.18 for the operating point considered in this paper.
As a result, the peak-to-peak wall pressure fluctuation level with respect to the turbine head is 2p% = 4.6% .
The corresponding experimental value is evaluated from the pressure fluctuation measurements performed by
Ciocan et al. and shown in [3, Fig. 10]. The pressure coefficient defined in [3], cp, is related to the
dimensionless pressure fluctuation defined in this paper, %p by the relationship

c p

P%
Q
2
2 Rref

P%

Q
Rref
2
2 Rref Rref

2
P%
p%
=2 2
2
2
( Rref )

and it is evaluated from [3, Fig. 10] to c p 0.9 . With the discharge coefficient = 0.264 corresponding to the
operating point investigated both in [3], as well as in this paper we obtain the experimental value p% exp = 0.028 ,
in excellent agreement with the value p% = 0.027 predicted by the precessing helical vortex model.

4. Conclusions
The approach developed in this paper is the first step in creating new instruments for assessment of turbine
behavior far from the design operating point, in the early design stage, with several orders of magnitude smaller
computing effort than the full three-dimensional unsteady flow simulation.
First, we present a mathematical model for computing the swirling flow at runner outlet, given the discharge
coefficient, the dimensionless flux of moment of momentum, and the swirl-free velocity radial profile. The first
two integral quantities describe the turbine operating point, while the swirl-free velocity corresponds to the
runner blade design at trailing edge. We show that by solving a constrained variational problem we obtain the
axial and circumferential velocity profiles, as well as the radial extent of the central quasi-stagnant region, in

25th IAHR Symposium on Hydraulic Machinery and Systems


IOP Conf. Series: Earth and Environmental Science 12 (2010) 012051

IOP Publishing
doi:10.1088/1755-1315/12/1/012051

good agreement with the experimental data available for a Francis model turbine. It is remarkable that we do not
need to actually compute the three-dimensional flow in the turbine runner in order to assess the swirling flow at
runner outlet.
Second, we employ a precessing helical vortex model in order to assess the precession frequency and wall
pressure fluctuation level associated with the vortex rope developed a Francis turbine discharge cone at partial
discharge. We show that this helical vortex model can correctly predict the swirling flow unsteadiness, with
respect to precession frequency and level of pressure fluctuation at the wall, in comparison with available
experimental data. Further investigations will consider the evaluation of the above parameters for swirling flow
considerations encountered downstream the runners of Francis turbines with other specific speed values.

Acknowledgments
Prof. P. Kuibin and Prof. V. Okulov were supported by RFBR (projects No. 10-08-01096, 10-08-01093) and
Ministry of Education and Science of Russian Federation (Program Scientific potential development of high
school, project No. 2.1.2/1270). Prof. R. Susan-Resiga and Dr. S. Muntean were supported by CNCSIS
UEFISCSU, PNII IDEI 799/2008.

Nomenclature
Symbols
a
Dimensionless distance from axis to vortex
core center (radius of helix)

F, f
G, g1,
g2
h
l
M, m
P, p
Q
R, r
U, u
w

Flow force [N], dimensionless


Functions, dimensionless

Dimensionless
modified
coordinate
(dimensionless
squared)
Discharge coefficient
Relative flow angle

Dimensionless axial pitch of the vortex


rope
Dimensionless characteristic length for
helical symmetry
Flux of moment of momentum [m5/s3],
dimensionless
Pressure [Pa], dimensionless
Volumetric discharge [m3/s]
Radius [m], dimensionless
Velocity [m/s], dimensionless

Dimensionless vortex intensity

Vortex core radius, dimensionless

radial
radius

Angular speed [rad/s], dimensionless

Energy coefficient
Density [kg/m3]
Dimensionless local radius in the core
Relative torsion of helical lines,
dimensionless

Local circumferential velocity in the vortex


core, dimensionless

Subscript and superscript


0
Swirl-free conditions
s
Stagnant region
ref
Reference values

w
z

Wall
Axial direction
Circumferential direction

References
[1]
[2]
[3]

Ruprecht A, Helmrich T, Aschenbrenner T and Scherer 2002 Simulation of Vortex Rope in a Turbine
Draft Tube Proc. of the 21st IAHR Symp. on Hydr. Mach. and Syst. vol 1 (Lausanne, Switzerland) pp
259-76
Sick M, Stein P, Doerfler P, White P and Braune A 2005 CFD Prediction of the Part-Load Vortex in
Francis Turbines and Pump-Turbines Int. J. of Hydropower and Dams 12(85)
Ciocan G D, Iliescu M S, Vu T C, Nennemann B and Avellan F 2007 Experimental Study and Numerical
Simulation of the FLINDT Draft Tube Rotating Vortex J. Fluids Eng. 129 pp 146-58

25th IAHR Symposium on Hydraulic Machinery and Systems


IOP Conf. Series: Earth and Environmental Science 12 (2010) 012051

[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]

[16]
[17]

IOP Publishing
doi:10.1088/1755-1315/12/1/012051

Zhang R K, Mao F, Wu J Z, Chen S Y, Wu Y L and Liu S H 2009 Characteristics and Control of the DraftTube Flow in Part-Load Francis Turbine J. of Fluids Engineering 131 021101-1-13
Susan-Resiga R, Muntean S, Stein P and Avellan F 2009 Axisymmetric Swirling Flow Simulation of the
Draft Tube Vortex in Francis Turbines at Partial Discharge Int. J. Fluid Mach. and Syst. 2(4) 295-302
Nishi M, Matsunaga S, Okamoto M, Uno M and Nishitani K 1988 Measurement of three-dimensional
periodic flow in a conical draft tube at surging condition Flows in Non-Rotating Turbomachinery
Components eds Rohatgi U S et al (FED vol 69) pp 81-88
Wang X and Nishi M 1998 Analysis of Swirling Flow with Spiral Vortex Core in a Pipe JSME Int. J.
B 41(2) 254-61
Alekssenko SV, Kuibin P A, Okulov V L and Shtork S I 1999 Helical vortices in swirl flow J. Fluid
Mech. 382 195-243
Alekseenko SV, Kuibin P A and Okulov V L 2007 Theory of Concentrated Vortices: An Introduction
(Berlin Springer)
Kuibin P A, Okulov V L and Pylev I M 2006 Simulation of the Flow Structure in the Suction Pipe of a
Hydroturbine by Integral Characteristics Heat Transfer Research 37(8) 675-84
Susan-Resiga R, Ciocan G D, Anton I and Avellan F 2006 Analysis of the Swirling Flow Downstream a
Francis Turbine Runner J. Fluids Eng.128 177-89
Tridon S, Barre S, Ciocan G. D and Tomas L 2010 Experimental analysis of the swirling flow in a Francis
turbine draft tube: Focus on radial velocity component determination European J. of Mech. B/Fluids
doi:10.1016/j.euromechflu.2010.02.004
Susan-Resiga R, Muntean S, Avellan F and Anton I 2010 Mathematical Modeling of Swirling Flow at
Francis Runners Outlet for the Full Turbine Operating Range J. of Fluids Eng. (submitted)
Benjamin T J 1962 Theory of the Vortex Breakdown Phenomenon J. Fluid Mech. 14 593-629
Ciocan G D and Iliescu M S 2007 Vortex Rope Investigation by 3D-PIV Method Proc. of the 2nd IAHR
International Meeting of the Workgroup on Cavitation and Dynamic Problems in Hydraulic Machinery and
Systems (Timioara, Romania) (Scientific Bulletin of the Politehnica University of Timisoara,
Transactions on Mechanics, Vol. 52(66), Issue 6) pp 159-72
Kuibin P A and Okulov V L 1998 Self-Induced Motion and Asymptotic Expansion of the Velocity Field in
the Vicinity of Helical Vortex Filament Phys. of Fluids 10(3) 607-14
Murakhtina T O and Okulov V L 2000 The influence of the core vorticity distribution on a possibility of
the spontaneous change of swirling flow regimes Thermophysics and Aeromechanics 7(1) 63-8

10

Você também pode gostar