Sunteți pe pagina 1din 40

VORTEX DOMINATED FLOWS 123

Chapter 3. MATHEMATICAL AND NUMERICAL


ANALYSIS OF AXISYMMETRIC SWIRLING FLOWS
Romeo SUSAN-RESIGA1, Sebastian MUNTEAN2, Alexandru BAYA3,
Liviu Eugen ANTON4, Teodor MILOŞ5, Adrian STUPARU6

Chapter 3. MATHEMATICAL AND NUMERICAL ANALYSIS OF AXISYMMETRIC


SWIRLING FLOWS........................................................................................................................... 123
3.1 Axisymmetric swirling flows............................................................................................. 124
3.2 Numerical methods for axisymmetric swirling flows........................................................ 128
3.2.1 The finite element method for steady axisymmetric swirling flows.............................. 128
3.2.2 Extension of the variational principle to swirling flows with vortex breakdown.......... 131
3.2.3 Numerical methods for unsteady axisymmetric inviscid swirling flows....................... 134
3.2.4 Numerical methods for axisymmetric turbulent swirling flows .................................... 135
3.3 Stability analysis for axisymmetric swirling flows............................................................ 140
3.3.1 Stability analysis using axisymmetric Navier-Stokes equations ................................... 140
3.3.2 Stability analysis of the Swirling Flow Downstream a Francis Turbine Runner........... 141
3.4 Axisymmetric turbomachinery swirling flow .................................................................... 146
3.4.1 Summary of turbomachinery swirling flow equations .................................................. 151
3.4.2 Blade-to-blade flow on axisymmetric stream-sheet ...................................................... 152
3.4.3 Thin hydrofoil cascade design ....................................................................................... 154
3.5 References.......................................................................................................................... 160

1
Professor, PhD, “Politehnica” University of Timişoara
2
Senior Researcher, PhD, Romanian Academy – Timişoara Branch
3
Professor, PhD, “Politehnica” University of Timişoara
4
Professor, PhD, “Politehnica” University of Timişoara
5
Associate Professor, PhD, “Politehnica” University of Timişoara
6
Assistant Professor, “Politehnica” University of Timişoara
124 AXISYMMETRIC SWIRLING FLOWS

A complete description of a fluid flow


∂Vθ ∂V ∂V V V
presumes the specification of some magnitudes. + Vz θ + Vr θ + r θ = 0 (3-6)
First of all, a description of the velocity of the ∂t ∂z ∂r r
fluid at all points in space and time is required. Because the flow is axisymmetric, we can reduce
However, the velocity field does not contain the momentum equations (3-4), (3-5) and (3-6) to
enough information to define the state of the fluid two equations in swirl function C ≡ rVθ and
in full, as other static properties of the fluid must
circumferential vortex density [56],
be known.
1 ⎛ ∂V ∂V ⎞
χ ≡ ⎜ r − z⎟=
3.1 Axisymmetric swirling r ⎝ ∂z ∂r ⎠
(3-7)
flows −
1 ⎛ ∂ 2 Ψ ∂ 2 Ψ 1 ∂Ψ ⎞
+ 2 −
⎜ ⎟ ≡ LΨ
r 2 ⎝ ∂z 2 ∂r r ∂r ⎠
Axisymmetric swirling flows are most
conveniently described in terms of cylindrical Note that the circumferential vortex density χ
polar coordinates ( z , r,θ ) , i.e. axial, radial and corresponds to the circumferential component of
circumferential coordinates, respectively. The the vorticity vector,
velocity vector is written in component form as ω=
V = Vz e z + Vr er + Vθ eθ , where the assumption of 1 ∂ ( rVθ ) ∂V ⎛ ∂V ∂V ⎞ , (3-8)
axial-symmetry implies that the velocity e z − θ er + ⎜ r − z ⎟ eθ
r ∂r ∂z ⎝ ∂z ∂r ⎠
components are functions of z and r only. The
velocity field for any incompressible flow can be divided by the radius r . The vorticity transport
written in terms of the vector potential β as equation for a three-dimensional, inviscid flow is
V = ∇ × β , where for axisymmetric flows the Dω ∂ω
circumferential component of β is related to the ≡ + (V ⋅ ∇ ) ω = ( ω ⋅ ∇ )V . (3-9)
Dt ∂t
Stokes’ streamfunction by βθ = Ψ / r . Since
In expressing (3-9) in component form, it is
Ψ depends only on the axial and radial important to note that the unit vectors er and eθ
coordinates, i.e. Ψ ( z, r ) , we have
have nonzero gradient with respect to θ such
that
⎛ Ψ ⎞ 1 ∂Ψ 1 ∂Ψ
∇ × ⎜ eθ ⎟ = ez − er (3-1)
⎝ r ⎠ r ∂r r ∂z ∂er ∂eθ
= eθ and = − er . (3-10)
∂θ ∂θ
As a result, the Stokes’ streamfunction is related
to the axial and radial velocity components, Using (3-10) as well as (3-8), the component
form of the vorticity transport equation for
1 ∂Ψ 1 ∂Ψ axisymmetric flow can be written as [35],
Vz = and Vr = − (3-2)
r ∂r r ∂z
Dωz ∂V ∂V
With (3-2) the continuity equation for = ωz z + ωr z , (3-11)
Dt ∂z ∂r
axisymmetric flows, with or without swirl,
∂Vz 1 ∂ ( rVr ) Dωr ∂V ∂V
+ =0 (3-3) = ω z r + ωr r , (3-12)
∂z r ∂r Dt ∂z ∂r
is identically satisfied.
Dωθ Vrωθ − 2Vθ ωr
The axisymmetric flow of an inviscid and = . (3-13)
incompressible fluid is governed by the Euler Dt r
momentum equations with the following axial, The transport equation for ωθ , Eq.(3-13), can be
radial and circumferential projections: written alternatively as
∂Vz ∂V ∂V 1 ∂p
+ Vz z + Vr z = − (3-4) D ⎛ ωθ ⎞ 2Vθ ωr 1 ∂Vθ2
∂t ∂z ∂r ρ ∂z ⎜ ⎟ = − = 2 . (3-14)
Dt ⎝ r ⎠ r2 r ∂z

∂Vr ∂V ∂V V 2 1 ∂p For the special case of axisymmetric flow


+ Vz r + Vr r − θ = − (3-5) without swirl, Vθ = 0 , the only non-zero vorticity
∂t ∂z ∂r r ρ ∂r
VORTEX DOMINATED FLOWS 125

component is in the circumferential direction, and symmetry axis. For steady flow Eq. (3-16)
Eq. (3-14) implies that ωθ / r is conserved on any reduces to {Ψ, C } = 0 . Since the jacobian
material point. vanishes, we must have a functional dependence
The axial and radial components of the of the form C ( Ψ ) , i.e. the swirl function C
vorticity transport equation, Eqs. (3-11) and
(3-12) reduce to the flow invariant remains constant along a streamline. The swirl
term in the right-hand side of (3-17) can be
D ( rVθ ) DC written in this case as
≡ = 0, (3-15)
1 ∂ (C ) 1 d ( C ) ∂Ψ
Dt Dt 2 2

which states that the circulation is constant about = =


4 y 2 ∂z 4 y 2 dΨ ∂z
a circle enclosing the symmetry axis when the (3-21)
circle moves with the surrounding fluid particles. d (C 2 ) ⎧ 1 ⎫ ⎧⎪ 1 d ( C ) ⎫⎪
2

This result is a direct consequence of the − ⎨ , Ψ ⎬ = ⎨ Ψ, ⎬


dΨ ⎩ 4 y ⎭ ⎩⎪ 4 y dΨ ⎭⎪
Kelvin’s circulation theorem, and it can be
directly obtained from the circumferential
As a result, in the case of steady axisymmetric
component of the momentum equation (3-6).
swirling flows the transport equation for χ
The evolution equations for C and χ can
(3-17) becomes
be written as [56],
⎧⎪ 1 d ( C ) ⎫⎪
2
∂C
+ {Ψ, C } = 0 , (3-16) ⎨ Ψ , χ − ⎬=0, (3-22)
∂t 4 y d Ψ
⎩⎪ ⎭⎪
∂χ 1 ∂C 2
+ {Ψ, χ } = 2 , (3-17) which implies that χ − ⎡⎣d ( C 2 ) / dΨ ⎤⎦ / ( 4 y )
∂t 4 y ∂z
depends only on Ψ , say F ( Ψ ) . It remains to
where { f , g } is the canonical Poisson bracket or interpret this functional dependence. In order to
jacobian, do so, we start from the vector product of the
velocity and vorticity,
∂f ∂g ∂f ∂g
{ f , g} = − , (3-18) ⎡V ∂ ( rVθ ) ⎤
∂y ∂z ∂z ∂y V ×ω = ⎢ θ + Vrωθ ⎥ e z
⎣ r ∂z ⎦
and y = r 2 / 2 is the new radial coordinate. If
⎡V ∂ ( rVθ ) ⎤ (3-23)
viscous flows are to be considered, with +⎢ θ − Vzωθ ⎥ er
ν the kinematic viscosity, then Eqs. (3-16) ⎣ r ∂r ⎦
and (3-17) have additional terms [56], + (Vzωr − Vrωz ) eθ
∂C From the circumferential component of the
+ {Ψ, C } = −2ν y LC , (3-19)
∂t momentum equation (3-15),
1 D ( rVθ ) DVθ VrVθ
∂χ = +
+ {Ψ, χ } = r Dt Dt r ,
∂t (3-24)
. (3-20) ∂Vθ
1 ∂C 2 ∂χ = + Vrωz − Vzωr = 0
− 2ν y Lχ + 4ν ∂t
4 y 2 ∂z ∂y
we see that for steady flows the circumferential
Because of Eq. (3-7), equations (3-16) and (3-17) component of V × ω , i.e. (Vzωr − Vrωz ) ,
constitute a pair of coupled equations for the
dependent variables C and χ . Pressure has vanishes. On the other hand, using the functional
disappeared as a dynamical variable, essentially dependence C ( Ψ ) , as well as Eq. (3-2), we can
because of taking the curl to obtain χ . write the gradient of the swirl function as
In a steady axisymmetric flow, the fluid dC
particles initially lying on a given streamline will ∇C = r (Vz er − Vr e z ) . (3-25)

advect over time on a surface of revolution
formed by rotating the streamline about the As a result, we can write V × ω as
126 AXISYMMETRIC SWIRLING FLOWS

Using the transformed radial coordinate


⎛ C dC ⎞
V ×ω = ⎜ − ωθ ⎟ (Vz er − Vr e x ) y = r 2 / 2 , Eq. (3-31) can also be written as
⎝ r dΨ ⎠
. (3-26)
⎛ 1 dC2 ⎞ ∂2Ψ 1 ∂2Ψ
= r⎜ − χ ⎟ (Vz er − Vr e x ) − − −
⎝ 4 y dΨ ⎠ ∂y 2 2 y ∂z 2
, (3-33)
C ( Ψ ) C ′ (ψ )
Since the first paranthesis depends only on the − + E′( Ψ ) = 0
streamfunction, according to Eq.(3-22), and this 2y
dependence has been denoted as − F ( Ψ ) , we can where the prime denotes the differentiation with
simply re-write Eq. (3-26) as respect to Ψ .
The steady, axisymmetric, Euler equation
V × ω = − rF ( Ψ )(Vz er − Vr e x ) . (3-27) for swirling flows, written for the Stokes’
streamfunction, in either (3-31), (3-32) or (3-33)
For any steady flow with conservative body force
form, is known in literature either as Bragg-
b = −∇λ , the momentum equation can be
Hawthorne (1950) equation, or Long(1953)-
expressed as
Squire(1956) equation. In fact, Goldshtik and
V × ω = ∇E , (3-28) Hussain [20] argued that this celebrated equation
has been in fact derived much earlier by Meissel
where E = p / ρ + V 2 / 2 + λ is the Bernoulli (1873).
function, with the physical meaning of a specific An elegant and intuitive derivation of the
mechanical energy per unit mass. Since E is streamfunction equation for axisymmetric,
invariant on a fluid particle in steady flow, E steady, incompressible and inviscid swirling
must depend only on the streamfunction, i.e. we flows is given by Benjamin [9]. In a meridional
have E ( Ψ ) . Similar to the gradient of the swirl half-plane, the resultant of the velocity
function, Eq. (3-25), we can write now the components Vz and Vr is directed along the
gradient of E as being streamline Ψ = constant , Figure 3-1, and has
the magnitude
dE
∇E = r (Vz er − Vr e z ) . (3-29) 1 ∂Ψ
dΨ Vm = − , (3-34)
r ∂n
Combining Eqs. (3-27), (3-28) and (3-29) finally
gives the physical interpretation of the functional obviously derived from the streamfunction
dependence F ( Ψ ) obtained as a result of Eq. definition, Eq.(3-2).
(3-22) in terms of the Bernoulli function (specific r
total mechanical energy),
dE ( Ψ ) ψ =constant τ
F (ψ ) = − , (3-30)
dΨ meridional
n
streamline
which substituted back into Eq. (3-22) yields a
nonlinear equation for streamfunction in a steady z
axisymmetric flow,

1 d ⎡⎣C ( Ψ ) ⎤⎦ dE ( Ψ )
2
Figure 3-1. Normal and tangential unit vectors
χ ≡ LΨ = − . (3-31)
4y dΨ dΨ on a streamline in the meridional half-plane,
with n × τ = eθ .
Given the definition of the linear operator L ,
one can re-write Eq. (3-31) in the form Analogously, one can find the meridional
∂ Ψ ∂ Ψ 1 ∂Ψ
2 2 component of the vorticity vector as
+ 2 −
∂z 2 ∂r r ∂r
. 1 ∂ ( rVθ ) 1 ∂C
(3-32) ωm = − =− . (3-35)
2 dE ( Ψ ) dC ( Ψ ) r ∂n r ∂n
=r − C (Ψ)
dΨ dΨ Moreover, according to Kelvin’s circulation
theorem in a steady flow a vortex line must lie on
a streamsurface and when applied to a circuit
VORTEX DOMINATED FLOWS 127

around a particular streamsurface Ψ = constant


we obtain C = constant , resulting in the
n
functional dependence C ( Ψ ) . As a result, the β

meridional vorticity component can be expressed ω V


as
1 dC ∂Ψ
ωm = − = VmC ′ ( Ψ ) . (3-36)
r dΨ ∂n
ωm Vm
The total vorticity ω is the rezultant of the
perpendicular components ωm and ωθ , and the α
velocity vector is the rezultant of perpendicular β

components Vm and Vθ = C / r . The V
ωθ ω
considerations above show that both vorticity and
velocity vectors are directed along a
streamsurface.
Figure 3-2. Velocity V and vorticity ω vectors in
Now, Eq. (3-28) shows the gradient of the
an axisymmetric streamsurface.
specific total mechanical energy, ∇E , to be
perpendicular to both V and ω , which implies When considering swirling flows in a axi-
that it is normal to the streamsurface and hence symmetrical diffusers with large increase of
E is expressible as a function of Ψ alone. We cross-sectional area, an elementary vortex flow
can rewrite Eq. (3-28) as threory based on Eq. (3-33) leads to a „diffuser
∂E paradox” [8]. For large swirl numbers it appears
= V ω sin β , (3-37) to be impossible to find steady axisymmetric
∂n
solutions. However, this apparent paradox can be
where β is the angle between V and ω . resolved [28] by including the possibility of a
Analogously to (3-36) we can put (3-37) in the loss-free transition to a hollow-core vortex,
form Figure 3-3.
V sin β
rE ′ ( Ψ ) = − ω sin β = −ω . (3-38)
Vm cos α
From Figure 3-2 the following identities are
readily obtained,
ω sin β = ωθ cos α − ωm sin α , and
C (Ψ) (3-39)
tan α = Figure 3-3. Loss-free transition in a diffuser
rVm with inlet swirling flow represented as a solid
Finally, Eq. (3-38) gives body vortex and constant axial velocity [29].

ωθ cos α − ωm sin α In situations involving rapid changes of cross-


rE ′ ( Ψ ) = − =
cos α sectional area it is more convenient to consider y
C (Ψ) as the dependent variable in Eq. (3-33) and Ψ
−ωθ + ωm tan α = −ωθ + VmC ′ ( Ψ ) (3-40) and z as the independent variables. Keller et al.
rVm
[28][29] introduced and used this approach to
ωθ C ( Ψ ) C′( Ψ ) study the swirling flow dynamics, and in
or − = E′( Ψ ) − , particular the vortex breakdown phenomenon, in
r r2
ducts with constant or varying cross-sectional
which is precisely the Bragg-Hawthorne or Long- area.
Squire equation. Interchanging variables Ψ ( z, y ) → y ( z, Ψ )
in the equation of motion (3-33) yields,
128 AXISYMMETRIC SWIRLING FLOWS

⎛ 1 ⎡ ∂y ⎤ ⎞ ∂ 2 y 1 ⎡ ∂y ⎤ ∂ 2 y
2 2
Ψ ( z, Y ( z ) ) = Q / 2π ,
⎜⎜ y + ⎢ ⎥ ⎟⎟ + − (3-43)
⎝ 2 ⎣ ∂z ⎦ ⎠ ∂Ψ 2 2 ⎢⎣ ∂Ψ ⎥⎦ ∂z 2 on the duct wall Γ w ,
⎡ ∂y ⎤ dE ( Ψ )
3
∂y ∂y ∂ 2 y (3-41)
− + y⎢ −
⎣ ∂Ψ ⎥⎦ dΨ
y
∂z ∂Ψ ∂z ∂Ψ Ψ ( 0, y ) = ∫ Vzinlet ( y ) dy ,
dC ( Ψ ) (3-44)
3
1 ⎡ ∂y ⎤ 0
− ⎢ ⎥ C (Ψ) =0
2 ⎣ ∂Ψ ⎦ dΨ on the inlet section Γi ,

∂Ψ
3.2 Numerical methods for ( y, L ) = 0 ,
∂z (3-45)
axisymmetric swirling
flows on the outlet section Γo ,

where Y ( z ) = R 2 ( z ) / 2 describe the wall shape,


Section 3.1 introduced the mathematical models
governing the axisymmetric swirling flows, with Q is the volumetric flow rate, and Vzinlet ( z ) is the
emphasis on inviscid flow. Various numerical inlet velocity profile, and L is the length of the
methods have been specifically designed to duct. In addition to these conditions on the
compute the hydrodynamic field (velocity and domain boundary, one should also prescribe the
pressure) for such flows, ranging from general functions C ( Ψ ) and E ( Ψ ) from the upstream
finite volume methods for primary variables
swirling flow configuration. Instead of
formulation in cylindrical coordinates, to
homogeneous Neumann condition (3-45) at
specialized methods for either steady and
outlet, corresponding to vanishing radial velocity,
unsteady streamfunction-circulation-vorticity
one can consider ∂Vr / ∂z = 0 . In this case one
formulation.
should impose a Dirichlet outlet condition
3.2.1 The finite element method for corresponding to the solution of the non-linear
steady axisymmetric swirling flows differential equation derived from Eq.(3-33),
d 2 Ψ C ( Ψ ) C ′ (ψ )
Let us consider the following boundary value − − + E′( Ψ ) = 0 , (3-46)
problem for the steady axisymmetric swirling dy 2 2y
flow, governed by Eq. (3-33), in a duct with with boundary conditions (3-42) and (3-43). In
variable radius R ( z ) , Figure 3-4. situations where pipe boundary shape varies only
slowly in the axial direction, it may also be
r n
wall appropriate to use the cylindrical form (3-46).
τ For solutions calculated at some downstream
Γw
τ station using (3-46) the axial dependence of the
n
outflow

Γi D flow enters only through the pipe radius in the


* Γo n
inflow

τ wall condition (3-43). This model is called quasi-


axis τ z cylindrical approximation.
Solutions to the boundary value problem (3-33),
Γa
n (3-42)...(3-45) correspond to stationary points of
the following functional [58]:
Figure 3-4. Computational domain in a
meridional half-plane. Normal unit vector n is F (Ψ) =
positive toward exterior, and tangential unit
⎡ 1 ⎛ ∂Ψ ⎞ 2 1 ⎛ ∂Ψ ⎞ 2 ⎤
vector τ is positive counter-clockwise.
L Y ( z) ⎢ ⎜ ⎟ + ⎜ ⎟ ⎥
⎢ 2 ⎝ ∂y ⎠ 4 y ⎝ ∂z ⎠ ⎥ dy dz (3-47)
Typical boundary conditions for this boundary- ∫0 ∫0 ⎢ C 2 ( Ψ ) ⎥
value problem are: ⎢− + E (Ψ) ⎥
⎢⎣ 4y ⎥⎦
Ψ ( z,0 ) = 0
(3-42) In order to prove the equivalence between the
on the symmetry axis Γ a , variational and differential formulations, let us
VORTEX DOMINATED FLOWS 129

compute the Frêchet differential of the functional result, if a curve in the plane ( y , Ψ ) is to
F ( Ψ ) . At a stationary point the differential must represent a physically realisable cylindrical
vanish, i.e. (columnar) flow, then it must be an extremal of
the flow force.
F ( Ψ + εψ ) − F ( Ψ )
lim =0, (3-48) The minimization of the functional F ( Ψ )
ε →0 ε
can be performed numerically using the Finite
where ψ is an arbitrary perturbation of Ψ . In Element Method (FEM). This is the most
order for Ψ + εψ to satisfy the Dirichlet convenient approach, since it combines the
conditions, the perturbation ψ must vanish on geometrical and meshing flexibility with a well
established and mature implementation
the boundary segments where Dirichlet
procedure. The first step is to approximate the
conditions are prescribed. Equation (3-48) leads
solution as [26]
to
Ψ ( z, y ) = ∑ Ψ j N j ( z, y ) ,
⎡ ∂ψ ∂Ψ 1 ∂ψ ∂Ψ ⎤ (3-52)
⎢ ∂y ∂y + 2 y ∂z ∂z ⎥
j

L Y (z)
⎢ ⎥ where Ψ j are the nodal values of the unknown
⎢ C ( Ψ ) C′( Ψ ) ⎥
∫0 ∫0 ⎢ −ψ 2y ⎥ dy dz = 0 . (3-49) function and N j ( z, y ) are known shape
⎢ ⎥
⎢ +ψ E ′ ( Ψ ) ⎥ functions associated with the nodes. After
⎢ ⎥ introducing (3-52) into (3-47) the conditions for
⎣ ⎦ extremum become
After integrating by parts the first two terms we ∂F
obtain =0=
∂Ψ i
⎡ ∂2Ψ 1 ∂2Ψ ⎤ ⎡ ∂N i ⎛ ∂N j ⎞ ⎤
L Y (z) ⎢ − ∂y 2 − 2 y ∂z 2 ⎥ ⎢ ⎜ ∑Ψ j ⎟ ⎥

∫0 ∫0 ⎢ C ( Ψ ) C ′ ( Ψ )
ψ ⎥ dy dz ⎢ ∂y ⎝ j ∂y ⎠ ⎥
⎥ ⎢
⎢− + E ′ ( Ψ )⎥ ⎢ + 1 ∂N i ⎛⎜ ∑ Ψ ∂N j ⎞ ⎥⎥ , (3-53)
⎣ 2y ⎦ ⎟ dy dz
∫∫ ⎢⎢ 2 y ∂z ⎝ j ⎠⎥
j
∂z
L Y (z) (3-50) ⎥
∂ ⎛ ∂Ψ ⎞
+∫ ∫ ψ dy dz . ⎢ C ( Ψ ) C′( Ψ ) ⎥
∂y ⎜⎝ ∂y ⎟⎠ ⎢− Ni ⎥
0 0 2y
⎢ ⎥
1 ⎡ ∂ ⎛ ∂Ψ ⎞ ⎤
Ymax
⎢⎣ + E ′ ( Ψ ) N i
L
⎥⎦
+ ∫ 0
⎢ ∫ ⎜ψ ⎟ dz ⎥ dy = 0
2 y ⎣⎢ z ( y ) ∂z ⎝ ∂z ⎠ ⎦⎥
which can be rewritten as
The last two integrals vanish thanks to the ⎛ ∂N i ∂N j 1 ∂N i ∂N j ⎞
boundary conditions. Since the first integral must ∑ Ψ ∫∫ ⎜j +
∂y ∂y 2 y ∂z ∂z ⎠
⎟ dy d z
vanish for arbitrary ψ , it follows that Eq. (3-33) j ⎝
Aij
must be satisfied by the streamfunction (3-54)
Ψ corresponding to a stationary point of the ⎛ C ( Ψ ) C′( Ψ ) ⎞
+ ∫∫ N i ⎜ E ′ ( Ψ ) − ⎟ dy dz = 0
functional F ( Ψ ) . ⎝ 2y ⎠
For the problem (3-46) the functional bi

reduces to or in matrix form


⎡ 1 ⎛ dΨ ⎞ C 2 ( Ψ ) ⎤
2
AΨ + b ( Ψ ) = 0 . (3-55)
⎟ − 4 y ⎥ dy
Ywall
⎢ ⎜
F (Ψ) = ∫ ⎢ 2 ⎝ dy ⎠ ⎥ (3-51)
The nonlinear system of equations (3-55) is
0 ⎢+ E ( Ψ ) ⎥
⎣ ⎦ solved using Newton’s iterative method [44]. Let
us assume that at iteration m the solution vector
Noting that the integrand in (3-51) can be written
is Ψ ( m ) . Let ψ ( m ) be a small correction to
as Vz2 + p / ρ , Benjamin [9] introduces the
Ψ ( m ) such that the solution at the next iteration
appelation „flow force” for the functional F ( Ψ )
associated with cylindrical swirling flows. As a
130 AXISYMMETRIC SWIRLING FLOWS

step will be Ψ ( m +1) = Ψ ( m ) + ψ( m ) . The nonlinear dE ( Ψ ) 1 dE (Y )


term b ( Ψ ) can then be linearized about Ψ ( m ) as E′( Ψ ) ≡ = , (3-59)
dΨ Vz (Y ) dY
∂b ( m )
b ( Ψ( m ) + ψ( m ) ) = b ( Ψ ( m ) ) + ψ , (3-56) where Vz (Y ) = dΨ / dY is the inlet axial velocity
∂Ψ
profile. Using the radial momentum equation we
where the ∂b / ∂Ψ matrix is can write E ′ as
⎛ ∂b ⎞ ∂bi dE ( Ψ ) 1 dE (Y )
Bij ≡ ⎜ ⎟ ≡ E′( Ψ ) ≡ = , (3-60)
⎝ ∂Ψ ⎠ij ∂Ψ j dΨ Vz (Y ) dY
⎛ ⎞ with C ′ being the derivative with respect to Ψ ,
⎜ E ′′ ( Ψ ) − ⎟
⎜ ⎟ . (3-57) 1 dC (Y )
⎜ ⎡⎣C ′ ( Ψ ) ⎤⎦ C ′ (Y ) =
2
⎟ . (3-61)
= ∫∫ N i N j ⎜ − ⎟ dy dz V z ( Y ) dY
⎜ 2y ⎟
⎜ C ( Ψ ) C ′′ ( Ψ ) ⎟ Finally, the integrand for the vector bi is
⎜⎜ − ⎟⎟ computed as
⎝ 2y ⎠
Obviously, the duble prime denotes the second ( Ψ , y ) → (Y ( Ψ ) , y ) →
order differentiation with respect to Ψ . As a dVz (Y ) ⎛ 1 1 ⎞. (3-62)
result, at each iteration step we have to solve a + C (Y ) C ′ (Y ) ⎜ − ⎟
dY ⎝ 2Y 2 y ⎠
linear system of equations,
The integrand for the Jacobian matrix Bij
⎡ A + B ( Ψ( m ) )⎤ ψ( m ) =
⎣ ⎦ requires the evaluation of the expression
, (3-58)
− ⎡⎣ AΨ( m ) + b ( Ψ ( m ) )⎤⎦
E ′′ ( Ψ ) − ( C ′ ( Ψ ) ) + C ( Ψ ) C ′′ ( Ψ )⎤ / 2 y . Using
⎡ 2

⎣ ⎦
in order to obtain the solution correction ψ( m ) . the same approach as above we obtain
The left-hand-side matrix is the Jacobian of the ( Ψ , y ) → (Y ( Ψ ) , y ) →
non-linear system (3-55) and the right-hand-side
1 ⎡ d Vz (Y ) C (Y ) C ′ (Y ) ⎤
2
of (3-58) is the residual of the non-linear system ,
⎢ − ⎥
(3-55), with negative sign, both computed with Vz (Y ) ⎣ dY 2 2Y 2 ⎦ (3-63)
the solution approximation Ψ ( m ) . Since ψ ( m )
satisfies only homogeneous essential and natural { 2
}
⎛ 1
+ ⎣⎡C ′ (Y ) ⎦⎤ + C (Y ) C ′′ (Y ) ⎜ −
1 ⎞

⎝ 2Y 2 y ⎠
conditions, the correction vanishes as the residual
goes to zero, provided the symmetric matrix where
A + B ( Ψ ( m ) ) is non-singular. This is the major
1 dC ′ (Y )
limitation of the method, essentially originating C ′′ (Y ) =
Vz ( Y ) d Y
from the steady flow assumption.
In order to evaluate the entries in the 1 d ⎡ 1 dC (Y ) ⎤
= ⎢ ⎥ . (3-64)
residual vector we need to compute Vz (Y ) dY ⎣ Vz ( Y ) d Y ⎦
E ′ ( Ψ ) − C ( Ψ ) C ′ ( Ψ ) / 2 y in (3-54). Since both
1 ⎡ d C (Y ) dV (Y ) ⎤
2

C ( Ψ ) and E ( Ψ ) functions are given for the = ⎢ − C ′ (Y ) z ⎥


Vz (Y ) ⎣ dY
2 2
dY ⎦
incoming columnar swirling flow, let us denote
by Y the transformed radial coordinate The function Y ( Ψ ) on the inlet section can be
corresponding to the inlet section. For a given
computed numerically from the boundary
streamfunction value, the inverse map Ψ → Y
condition (3-44). Using the Newton’s method, we
allows the change of variable from E ( Ψ ) to can correct iteratively the Y value for a given
E (Y ) . The derivative with respect to Ψ can Ψ as follows,
accordingly be written
VORTEX DOMINATED FLOWS 131

Y ( m +1) = Y ( m ) + F (Ψ) =
z2 Ywall ( z )
⎛ ⎞,
(m)
Y
1 ⎛ ∂Ψ ∂Ψ ⎞
⎜ Ψ − ∫0 V inlet
( y ) dy ⎟ ∫ ∫ F ⎜ z , y , Ψ, , dy dz
Vzinlet (Y ( m ) ) ⎜⎝ ⎟ ∂z ∂y ⎟⎠
z
⎠ (3-65) z1 Ya ( z ) ⎝
where
Ψ (3-66)
with Y (0) = . 2 2
Vzinlet 1 ⎛ ∂Ψ ⎞ 1 ⎛ ∂Ψ ⎞
F= ⎜ ⎟ + ⎜ ⎟
2 ⎝ ∂y ⎠ 4 y ⎝ ∂z ⎠
Obviously, the initial approximation for
Y corresponds to the constant velocity profile C2 (Ψ)
− + E (Ψ)
with Vzinlet = Q / (π R 2 (0) ) . 4y
We have detailed above the Finite Element where Ywall ( z ) is the given wall shape. Following
Method for solving the boundary value problem
for steady, axisymmetric, incompressible and the theory of calculus of variations with variable
inviscid swirling flows. On the inlet section we end-points, see for example Riley et al.
have assumed that the axial and circumferential [45]§20.3.4., we have the Euler-Lagrange
velocity components are given, and the radial equation,
velocity vanishes. It is quite clear that the main
∂F ∂ ⎛ ∂F ⎞ ∂ ⎛ ∂F ⎞
limitation of the above approach is that the axial − ⎜ ⎟− ⎜ ⎟=0, (3-67)
velocity cannot vanish or the Jacobian matrix in ∂Ψ ∂z ⎝ ∂Ψ z ⎠ ∂y ⎜⎝ ∂Ψ y ⎟⎠
(3-58) cannot be singular. However, stagnant or
recirculating regions may develop in swirling and an auxiliary equation to be satisfied on the
flows with large enough swirl, a phenomenon unknown part of the boundary,
called vortex breakdown. ∂F ∂F
F − Ψz − Ψy =0. (3-68)
∂Ψ z ∂Ψ y
3.2.2 Extension of the variational
principle to swirling flows with In Eqs. (3-67) and (3-68) we have denoted by
vortex breakdown Ψ z and Ψ y the partial derivatives of the
Let us go back to the functional (3-47), but we dependent variable Ψ ( z , y ) with respect to z and
consider now the case from Figure 3-3 where y , respectively. Inserting the expression for
part of the boundary is initially unknown. More F ( z, y , Ψ, Ψ z , Ψ y ) into the equation (3-67)
precisely, when the swirl intensity is large
enough, the flow decelerates severely along the yields,
symmetry axis, reaches a stagnation point, then ∂F C ( Ψ ) C′( Ψ )
an annular swirling flow develops around a = E′( Ψ ) − ,
central stagnant region. Keller et al. [28] extends ∂Ψ 2y
the variational formulation by taking into account
∂ ⎛ ∂F ⎞ 1 ∂ 2 Ψ
the near-axis boundary shape Ya ( z ) ⎜ ⎟= ,
∂z ⎝ ∂Ψ z ⎠ 2 y ∂z 2
corresponding to the radius of the stagnant region
ra from Figure 3-3. Obviously, as long as the ∂ ⎛ ∂F ⎞ ∂2Ψ
⎜ ⎟⎟ = 2 , (3-69)
axial velocity on the axis is positive, we have ∂y ⎝⎜ ∂Ψ y
Ya = ra2 / 2 = 0 , i.e. the computational domain ⎠ ∂y
extends up to the symmetry axis and we have the ∂2Ψ 1 ∂2Ψ
boundary condition (3-42). However, when the − − −
∂y 2 2 y ∂z 2
annular flow develops, the determination of its and ,
C ( Ψ ) C ′ (ψ )
shape Ya ( z ) becomes part of the problem and − + E′( Ψ ) = 0
2y
requires an additional equation. The functional
(3-47) becomes, which is precisely the Long-Squire equation
(3-33). The additional equation to be satisfied on
the free-surface Ya ( z ) > 0 results from Eq.(3-68)
as
132 AXISYMMETRIC SWIRLING FLOWS

2
thus solving a problem in the simple rectangular
∂F 1 ⎛ ∂Ψ ⎞ domain [ z1 , z2 ] × [ 0, Q / 2π ] . In this case, the
Ψz = ⎜ ⎟ ,
∂Ψ z 2 y ⎝ ∂z ⎠ functional for the variational principle becomes
2 z2 Q / 2π
∂F ⎛ ∂Ψ ⎞ ⎛ ∂y ∂y ⎞
Ψy = , F ( y) = ∫ ∫ F ⎜ z , Ψ, y , , ⎟ dΨ dz
∂Ψ y ⎜⎝ ∂y ⎟⎠ z1 0 ⎝ ∂z ∂Ψ ⎠
and (3-70) where (3-71)

E (Ψ) − 1 y ⎡2
C (Ψ) ⎤ 2
F= + + yΨ ⎢ E ( Ψ ) −
z

2yΨ 4 yyΨ ⎣ 4y ⎦
⎡ ⎤
1 1 ⎛ ∂Ψ ⎞ ⎛ ∂Ψ ⎞ C ( Ψ ) ⎥
⎢ 2 2 2
The Euler-Lagrange equation in this case is
− ⎢ ⎜ ⎟ + + ⎥=0
2 ⎢ 2 y ⎝ ∂z ⎠ ⎜⎝ ∂y ⎟⎠ 2y ⎥
⎢ ⎥ ∂F ∂ ⎛ ∂F ⎞ ∂ ⎛ ∂F ⎞
⎣ Vr2 Vz2 Vθ2 ⎦ − ⎜ ⎟− ⎜ ⎟ = 0, (3-72)
∂y ∂z ⎝ ∂y z ⎠ ∂Ψ ⎝ ∂yΨ ⎠
Eq. (3-70) holds on the streamline
which, after laborious calculations, leads to Eq.
Ψ ( z, Ya ( z ) ) = Ψ 0 , and implies that the static (3-41). Let us first check the simplified problem
fluid pressure remains constant on the breakdown for columnar flows, i.e. the case when
bubble region. This is the natural condition on ∂y / ∂z = 0 . The Euler-Lagrange equation
the breakdown bubble surface. If there is no becomes,
central body upstream, and the streamline on
which Eq. (3-70) is applied originates from the ∂F ∂ ⎛ ∂F ⎞
− ⎜ ⎟ = 0,
symmetry axis as in Figure 3-3, then the velocity ∂y ∂Ψ ⎝ ∂yΨ ⎠
magnitude vanishes at the stagnation point on the
axis and remains zero further downstream on the with F ( Ψ, y , yΨ ) = . (3-73)
free-surface. As a result, the breakdown bubble is 1 ⎡ C (Ψ) ⎤ 2

a stagnant region, a remarkable result taking into + yΨ ⎢ E ( Ψ ) − ⎥


2yΨ ⎣ 4y ⎦
account that we have considered an inviscid
model. Goldshtik and Hussain [20] advocate this We have successively,
stagnant separation zone model, and imagine that
if a solid body would replace the stagnant ∂F C2 (Ψ)
breakdown bubble then this axisymmetric body = yΨ ,
∂y 4 y2
would experience a very low drag since the
velocity on its surface vanishes even in the ∂F 1 C2 (Ψ)
= − 2 + E (Ψ) − ,
absence of viscosity. However, in real flow the ∂yΨ 2 yΨ 4y
viscous effects do not allow such a well-defined
∂ ⎛ ∂F ⎞ yΨΨ
sharp shape of the central stagnant region for ⎜ ⎟= + E′( Ψ ) −
vortex breakdown in diffusers, as observed by ∂Ψ ⎝ ∂yΨ ⎠ yΨ3 (3-74)
Brücker [12]. C ( Ψ ) C′( Ψ ) C (Ψ) 2

The numerical implementation of the above − + yΨ ,


2y 4 y2
variational formulation, coupled with the
determination of the unknown surface of the resulting in
breakdown bubble would imply that at each yΨΨ C ( Ψ ) C′( Ψ )
iteration the domain boundary, or at least part of 3
+ E′( Ψ ) − = 0,
yΨ 2y
it, changes and a new discretization is required.
Moreover, the generating functions of the vortex which is in fact Eq. (3-46) after the change of
flow C ( Ψ ) and E ( Ψ ) are defined only on the variables Ψ ( y ) → y ( Ψ ) .
interval [0, Q / 2π ] and reverse flow with The terms in Eq. (3-72) can be computed
streamfunction values outside this interval may using the expression F ( z , Ψ, y , y z , yΨ ) from
occur during the iteration process. As a result, (3-71) as follows:
Keller [28] came up with the idea of
interchanging the variables Ψ ( z, y ) → y ( z, Ψ ) ,
VORTEX DOMINATED FLOWS 133

∂F y2 C2 (Ψ) are longer in longer diffusers, their relative


= − 2z + yΨ , lengths with respect to the diffusers are smaller.
∂y 4 y yΨ 4 y2
∂ ⎛ ∂F ⎞ 1 y zz 1 y z2 1 y z y Ψz
⎜ ⎟ = − −
∂z ⎝ ∂y z ⎠ 2 y yΨ 2 y yΨ 2 y yΨ2
2

∂ ⎛ ∂F ⎞ yΨΨ 1 y z yΨz y z2 1 y z2 yΨΨ


⎜ ⎟= 3 − + +
∂Ψ ⎝ ∂yΨ ⎠ yΨ 2 y yΨ2 4 y 2 yΨ 2 y yΨ3
C ( Ψ ) C′( Ψ ) C2 (Ψ)
+ E′( Ψ ) − + yΨ ,
2y 4 y2
leading immediately to Eq.(3-41).
The boundary conditions for the variational
problem (3-71), are derived from (3-42)...(3-45)
on the fixed (given) boundary segments. For
example, on the duct wall we have
y ( z , Q / 2π ) = Ywall ( z ) . On the inlet section, for a
prescribed axial velocity profile we have
y ( z1 , Ψ ) = Y ( Ψ ) as computed by the iterative
approach (3-65). On the outlet section, the zero
radial velocity condition (3-45) can be rewritten
as ( ∂y / ∂z ) z = z / ( ∂y / ∂Ψ ) z = z = 0 . The streamline
2 2

Ψ = 0 will partly correspond to the symmetry


axis and the rest will be the breakdown bubble
surface, with the corresponding conditions:
Figure 3-5. Swirling flow in a diffuser, with
y ( z ,0 ) = 0 on the axis (a ) vortex breakdown, for increasing swirl intensity,
∂y computed by Keller et al.[29].
→ ∞ as Ψ → 0 . (3-75)
∂Ψ ( b)
on the free-surface When a central body is present upstream, with
the shape given by y ( z , Ψ 0 ) = Y0 ( z ) , the swirling
Keller et al. [29] computed the vortex breakdown flow may detach and a free surface is formed on
in a diffuser with sinusoidal shape given by which the natural condition (3-70) holds in the
R12 2 form,
Ywall ( z ) = s ( z ) with
2 ⎡ ⎛ ∂y 2

⎧1 if z ≤ 0 ⎢ ⎜ ⎟
⎪ (3-76) 1 ⎢ 1 ⎜ ∂z Ψ=Ψ0
⎪ 1⎡ ⎛ π ⎞⎤ E ( Ψ0 ) − ⎟ +
s ( z ) = ⎨1 + ⎢1 − cos ⎜ z ⎟ ⎥ if 0 < z < 6 2 y0 ( z ) ⎢ 2 ⎜ ∂y ⎟
⎪ 2⎣ ⎝ 6 ⎠⎦ ⎢ ⎜ ⎟
⎪⎩2 if z ≥ 6 ⎢⎣ ⎝ ∂Ψ Ψ=Ψ0 ⎠
. (3-77)

with R1 = 2 . For a solid-body swirl at inlet, ⎥
with constant axial velocity Vzinlet and a y0 ( z ) C 2 ( Ψ0 ) ⎥
+ 2
+ ⎥=0
ξ r inlet ⎛ ∂y ⎞ 2 ⎥
circumferential velocity Vθinlet = Vz , Keller ⎜⎜ ⎟⎟ ⎥
2 R1 ⎝ ∂Ψ Ψ=Ψ 0 ⎠ ⎦
et al. [29] obtain the streamline maps for various
swirl intensity ξ as shown in Figure 3-5.
Beginning and end of the diffuser are marked by
arrows. Comparing flow fields in diffusers of
different lenghts shows that, although transitions
134 AXISYMMETRIC SWIRLING FLOWS

3.2.3 Numerical methods for unsteady incompressibility condition (3-78)(b) is global,


axisymmetric inviscid swirling flows and thus it is solved by the standard Poisson
solver for the streamfunction ψ using
Although the steady axisymmetric swirling continuous finite elements. The main advantage
flow model reduces to a single non-linear partial of the method proposed by Liu and Shu is that
differential equation, its numerical solution there are no matching conditions needed for the
brings implementation difficulties when vortex two finite element spaces for the vorticity ω and
breakdown is present. Rusak, Wang and Whiting for the streamfunction ψ . The incompressibility
[46] advocate for the use of the unsteady swirling
condition, represented by the streamfunction ψ ,
flow model (3-7)(3-16)(3-17) to follow the
nonlinear dynamics of small- or large-amplitude is exactly satisfied pointwise and it is naturally
disturbances and, specifically, the unstable matched with the convective terms in the
axisymmetric disturbance as it evolves in time momentum equation. The normal velocity V ⋅ n
into the axisymmetric breakdown. Since the is automatically continuous along any element
swirling flow in a pipe is dominated by the boundary, allowing for correct upwinding for the
propagation of disturbances in the axial direction convective terms and still maintaining a total
(both upstream and downstream depending on energy conservation and enstrophy stability. Let
the swirl level) they choose a finite difference us briefly describe the continuous-discontinuous
scheme based on the Lax methods for the Galerkin method of Liu and Shu [33]. We start
transport equations (3-16) and (3-17), with a with a triangulation T h of the 2D computational
small enough time step to ensure that the scheme domain Ω , consisting of polygons of maximum
is numerically stable. At each time level, an size (diameter) h , and two approximation
iterative point relaxation algorithm based on a spaces: Vhk = {v : v K ∈ P k ( K ) , ∀K ∈ T h } and
central-difference scheme has been implemented
to solve the Poisson equation (3-7) for the W0,kh = Vhk ∩ C0 ( Ω ) , where P k ( K ) is the set of
streamfunction. all polynomials of degree at most k on the cell
Saiac [47] develops and analyses a Finite K , and W0,kh is the subspace of continuous
Element Method to solve the elliptic equation
(3-7) in Ψ coupled with the transport equations functions over the domain. For the Euler
(3-16) and (3-17) for C = rVθ and χ = ωθ / r , equations (3-78), the numerical method is defined
as follows: find ωh ∈ Vhk and ψ h ∈ W0,kh such that
respectively. Two time discretization schemes,
the Leap-frog scheme and the semi-implicit ∂ tωh v − ωhVh ⋅ ∇v +
Crank-Nicolson scheme, are investigated with K K

respect to their stability properties. In order to get + ∑ Vh ⋅ n ωˆ h v − = 0, ∀v ∈ Vhk ,


(a )
e (3-79)
a stationary solution of the Euler equations one e∈∂K

can march in time or, alternatively, a simple − ∇ψ h ⋅ ∇ϕ = ωhϕ , ∀ϕ ∈ W0,kh , (b)


characteristic method can be employed thus
taking advantage on knowing the streamlines. with the velocity field obtained from the
A very promising method for the 2D streamfunction by Vh = ∇ ⊥ψ h . Here ⋅ is the
incompressible unsteady Euler equation in usual integration over either the whole domain
streamfunction-vorticity formulation has been Ω or a subdomain denoted by a subscript. Notice
introduced by Liu and Shu [33]. They solve the that the normal velocity Vh ⋅ n is continuous
equations across any element boundary e , buth both the
∂ω solution ωh and the test function v are
+ ∇ ⋅ (V ω ) = 0 (a )
∂t , (3-78) discontinuous there. The values of the test
Δψ = ω (b) function from within the element K are denoted
by v − . The solution at the edge is taken as a
with V = ∇ ⊥ψ , where ∇ ⊥ = ( −∂ / ∂y , ∂ / ∂x ) . The single-valued flux ωˆ h , which can be either a
momentum equation (3-78)(a) is treated central or an upwind-biased average. The
explicitly in the discontinuous Galerkin continuous-discontinuous Galerkin method is
framework, and as a result there is no global proved to be efficient for inviscid or high
matrix to invert unlike conventional finite Reynolds number flows.
element methods. This makes the method highly
efficient for parallel implementation. The
VORTEX DOMINATED FLOWS 135

In our oppinion, the extension of Liu and Shu the best numerical prediction for separation and
approach to axisymmetric swirling flows, with reattachment positions, with reasonable CPU
continuous finite elements for Eq. (3-7) and time. Lu and Semião [34] propose an improved
discontinuous finite elements for (3-16) and anisotropic model for the dissipation rate – ε – of
(3-17) will lead to a very efficient and accurate the turbulent kinetic energy – k – to be used
numerical algorithm. together with a non-linear pressure-strain
correlations model. An important conclusion is
3.2.4 Numerical methods for that for the case of strongly swirling flows the
axisymmetric turbulent swirling results are very sensitive to the pressure-strain
flows model, which has to be non-linear in this case.
Analysis of real flows requires a full unsteady,
Axisymmetric swirling flow models for axisymmetric turbulent model. The general
incompressible Navier-Stokes equations have equations to be solved are the continuity equation
been successfully used in numerical simulation of with an eventual source term,
vortex breakdown, with very good accuracy in
∂ρ ∂ ∂ ρV
comparison with experimental data [18]. + ( ρVz ) + ( ρVr ) + r = Sm (3-80)
However, for high Reynolds number turbulent ∂t ∂z ∂r r
flows the turbulence modelling introduces and the momentum equations
additional difficulties. The state of the art in
modelling of turbulent swirling flow two decades ∂ 1 ∂ 1 ∂
ago is presented by Sloan et al. [50] in a
( ρVz ) + ( r ρVzVz ) + ( rρVrVz ) =
∂t r ∂z r ∂r
comprehensive review focused on combustion ∂p 1 ∂ ⎡ ⎛ ∂Vz 2 ⎞⎤
systems. There are two main viewpoints: the first − + Fz + ⎢ rμ ⎜ 2 − ( ∇ ⋅ V ) ⎟ ⎥ (3-81)
is focused on isotropic eddy viscosity and a ∂z r ∂z ⎣ ⎝ ∂z 3 ⎠⎦
modified Boussinesq hypothesis, while the 1 ∂ ⎡ ⎛ ∂Vz ∂Vr ⎞ ⎤
+ rμ ⎜ + ⎟
second assumes that the eddy viscosity approach r ∂r ⎢⎣ ⎝ ∂r ∂z ⎠ ⎥⎦
is inherently inadequate and a redistribution of
stress magnitudes is necessary via high-order
closure. Armfield et al. [3], [4] and [5] develop
∂ 1 ∂ 1 ∂
and evaluate both k − ε and algebraic Reynolds ( ρVr ) + ( rρVzVr ) + ( rρVrVr ) =
stress models for swirling flows in conical ∂t r ∂z r ∂r
diffusers, but validation was performed against ∂p V 2 1 ∂ ⎡ ⎛ ∂Vr ∂Vz ⎞ ⎤
− + Fr + ρ θ + rμ ⎜ + ⎟
Clausen et al. [17] experiments where no ∂r r r ∂z ⎢⎣ ⎝ ∂z ∂r ⎠ ⎥⎦
separation or vortex breakdown were present. (3-82)
1 ∂ ⎡ ⎛ ∂Vr 2 ⎞⎤
Yaras and Grosvenor [61] test several turbulence + ⎢ rμ ⎜ 2 − ( ∇ ⋅V ) ⎟⎥
models to establish the prediction accuracy with r ∂r ⎣ ⎝ ∂r 3 ⎠⎦
respect to axisymmetric separating flows and Vr 2 μ
flows with high streamline curvature. They −2 μ + (∇ ⋅V )
r2 3 r
conclude that the prediction of strongly swirling
confined flow was rather poor, with all models
(Rodi’s k-ε, Menter’s two-equation shear-stress- ∂ 1 ∂ 1 ∂
transport SST model, and the one-equation model ( ρVθ ) + ( r ρVzVθ ) + ( r ρVrVθ ) =
of Spalart and Allmaras) significantly ∂t r ∂z r ∂r
overestimating the radial diffusive transport. V V 1 ∂ ⎡ ∂Vθ ⎤ 1 ∂ ⎡ 3 ∂ ⎛ Vθ ⎞⎤
−ρ r θ + rμ + rμ ⎜ ⎟⎥
Amongst these models, the SST model yielded r r ∂z ⎢⎣ ∂z ⎥⎦ r 2 ∂r ⎢⎣ ∂r ⎝ r ⎠⎦
the worst prediction. Use of the streamline-
curvature corrections in the turbulence transport (3-83)
equations has little impact on the prediction where ∇ ⋅ V = ∂Vz / ∂z + ∂Vr / ∂r + Vr / r .
accuracy. Xu et al. [60] compare various
The FLUENT commercial code includes a
commonly used general k-ε turbulence models
Finite Volume solver for axisymmetric turbulent
with respect to their ability to predict the
swirling flows, with a large variety of turbulence
separation and reattachment of turbulent flows
models. Our numerical experiments and
inside an axisymmetric diffuser with a curved
comparison with experimental data showed that a
surface centre-body. They conclude that the high-
good compromise between computational effort
Reynolds number k-ε + one-equation model gives
136 AXISYMMETRIC SWIRLING FLOWS

and accuracy can be achieved with the realizable of the full 3D flow simulation, with huge savings
k-ε model [48]. The term „realizable” means that in both computing time and resources.
the model satisfies certain mathematical The actual draft tube of Francis turbines,
constraints on the Reynolds stresses, consistent Figure 3-6, has a complex three-dimensional
with the physics of turbulent flows. An shape, with a discharge cone, followed by an
immediate benefit of the realizable k-ε model is elbow and a downstream pier dividing the
that it more accurately predicts the spreading rate hydraulic passage toward the outlet. The only
of both planar and round jets. It is also geometrically axisymmetric part is the discharge
conjectured that the realizable k-ε model provides cone, but because of the elbow the flow departs
superior performance for flows involving from axial-symmetry even in this region. In
rotation, boundary layers under strong adverse Francis turbines operated at partial discharge, the
pressure gradients, separation, and recirculation. swirling flow downstream the runner becomes
unstable inside the draft tube cone, with the
development of a precessing helical vortex (also
called vortex rope) and associated severe
pressure fluctuations. Numerical simulations of
the precessing vortex rope have reached the level
of accuracy where the main features of the
phenomenon (vortex rope shape, precession
frequency, pressure fluctuation) are accurately
described [49][51][16]. However, such 3D
unsteady turbulent flow computations are very
expensive in terms of computing time and
resources.
Our main goal is to develop a methodology
for analyzing the swirling flows with helical
vortex breakdown by using an axisymmetrical
swirling flow model. Obviously, the axial-
symmetry hypothesis is a major simplification
having the main benefit of dramatically reducing
the computational cost, but introduces important
limitations as far as the three-dimensionality and
Figure 3-6. The actual draft tube of a Francis unsteadiness of the flow are concerned.
turbine, with the equivalent axisymmetric Essentially, an axisymmetric flow solver
computational domain from Figure 3-7 . provides a circumferentially averaged, or time
averaged, velocity and pressure fields, to be
As a first example of turbulent axisymmetric further used as a base flow for linear stability
swirling flow analysis we consider the flow in the analysis. From practical point of view, the
draft tube cone of a Francis turbine operated at ultimate goal is to reach stable axisymmetric
partial discharge. This is a complex swirling flow configurations which avoid the
hydrodynamic phenomenon where an incoming development of the helical vortex breakdown,
steady axisymmetric swirling flow evolves in a through suitable flow control techniques [52].
three-dimensional unsteady flow field generated In order to perform an axisymmetric flow
by the precessing helical vortex (also called simulation, an equivalent axisymmetric
vortex rope), with associated pressure computational domain must be considered. Given
fluctuations. There is a fundamental question to the three-dimensional geometry of the actual
be answered: is it possible to compute the draft tube, where only the cone is geometrically
circumferentially avergaed flow field induced by axisymmetric, we consider the domain shown in
the precessing vortex rope by using an Figure 3-7 with the wall radius computed as the
axisymmetric swirling flow model? In other equivalent hydraulic radius where the cross-
words, instead of averaging the (measured or section is no longer circular. The upstream part of
computed) 3D velocity and pressure fields we the domain corresponds to the actual draft tube
would like to use the circumferentially averaged cone, while the downstream part corresponds to
governing equations (3-80)...(3-83). As a result, the elbow, up to the pier.
one could use a 2D axi-symmetric model instead
VORTEX DOMINATED FLOWS 137

All lengths are made dimensionless with respect used in [51] for 3D unsteady flow simulation
to the turbine throat radius, and the coordinate with vortex rope in the real draft tube shown in
system origin is chosen such that the throat Figure 3-6. Two survey sections where velocity
dimensionless abscissa is equal to unity. The inlet components radial profile was measured using
section for the computational domain is located Laser Doppler Anemometry [16] are also shown
just downstream the runner blade trailing edge, as in Figure 3-7, futher denoted here as S1 and S2.

1.5

1
r [-]

0.5

00 1 2 3 z [-] 4 5 6 7

e wall
be con
dimensionless radial coordinate

1.2 draft tu

survey section S2

cone outlet
survey section S1

0.8
on
e cti
turbine throat

0.6
e ts
inl
0.4

0.2
crown
cone
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 3
dimensionless axial coordinate
Figure 3-7. Axisymmetric computational domain for flow analysis in a hydraulic turbine draft tube.

The circumferentially averaged velocity


components and turbulence quantities 0.4
inlet meridian velocity
corresponds to the data from [51], Figure 3-8 , for inlet circumferential velocity
a partial discharge operating point
dimensionless velocity [−]

Q / Qopt = 0.692 . The meridian velocity 0.3

component is defined as Vm = Vz2 + Vr2 , and


0.2
the absolute/relative flow angles are
α = arctan (Vm / Vθ ) and
0.1
β = arctan ⎡⎣Vm / ( Ωr − Vθ ) ⎤⎦ , respectively, with
crown

band

Vz ,Vr ,Vθ the axial, radial and circumferential 0


0 0.2 0.4 0.6 0.8 1 1.2
velocity components, Ω the runner angular dimensionless curvilinear coordinate [−]
velocity. The velocity is made dimensionless
Figure 3-8. Meridian and circumferential
with respect to 2gH , where H is the turbine velocity profile on the inlet section shown in
head. The absolute and relative flow angle on the Figure 3-7 .
inlet section are shown in Figure 3-9.
138 AXISYMMETRIC SWIRLING FLOWS

60 The streamline pattern obtained without and


with the stagnation region model (SRM) is
50 shown in Figure 3-10. The plain axisymmetric
swirling flow model predicts a recirculation
40 region, with strong reverse flow near the axis.
This strong recirculation disagrees with
flow angle [dgr]

30
experimental data, and the streamline bounding
the main stream does not correspond to the
20
relative flow angle location of the vortex rope (shown with circles)
absolute flow angle
as found experimentally [15][31]. However, with
the stagnation region model the vortex rope
10
crown

band
location practically coincides with the boundary
of the central dead water region.
0
0 0.2 0.4 0.6 0.8 1 1.2 0.4
dimensionless curvilinear coordinate [−] LDV measurement
without stagnation model
with stagnation model
Figure 3-9. Meridian and circumferential 0.3

dimensionless axial velocity [−]


velocity profile on the inlet section shown in
Figure 3-7 .
0.2

One important issue to be addressed is the effect


of neglecting the actual precessing vortex rope on 0.1

the circumferentially averaged flow. Both


experimental data [39][15][31], as well as the 0.0
theoretical analyses [29] suggest that when a
vortex breakdown occurs a quasi-stagnation
region is developed near the axis. As a result, we −0.1
1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2
have enforced a stagnation condition on top of dimensionless radial coordinate [−]

the FLUENT axisymmetric flow solver whenever 0.4


a negative (i.e. upstream oriented) axial velocity LDV measurement
0.3 without stagnation model
dimensionless circumferential velocity [−]

occurs. with stagnation model


0.2

0.1

−0.1

−0.2

−0.3

−0.4
1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2
dimensionless radial coordinate [−]

Figure 3-11. Axial and circumferential velocity


profiles in the survey section S1.

Figure 3-11 shows the measured and


computed axial and circumferential velocity
profiles on the survey section S1. One can see
that the plain axisymmetric swirling flow model
(dashed lines) predicts negative axial velocity
near the axis, and correspondingly the axial
Figure 3-10. Streamline pattern and the location velocities increases well above the measured
of the vortex rope. Left/righ half-plane values at the boundary between the main flow
without/with stagnation model . and the central region. On the other hand, when
the stagnation model is activated the numerical
VORTEX DOMINATED FLOWS 139

results are in much better agreement with discharge is that the circulation function displays
experimental data. The radial extent of the central a significat increase near the shroud with respect
stagnation region is very well predicted, as well to the practically linear variation in the rest of the
as the velocity distribution in the main annular main stream. The same conclusion holds for the
stream. The same conclusions hold in survey total pressure, with respect to a quasi-constant
section S2, Figure 3-12. However, one should value for 75% of the main stream in the inlet
keep in mind that the downstream survey section section. Besides the dissipation of the circulation
is close to the draft tube elbow, and the velocity and total pressure excess near the cone wall, there
field departs significantly from the axisymmetric is a strong dissipation at the boundary of the
hypothesis due to the additional distortions central stagnation region as well.
induced by the elbow secondary flows. 0.35
0.4 inlet section
0.3 survey section 1
LDV measurement
survey section 2
without stagnation model
cone outlet
with stagnation model
0.3 0.25
dimensionless axial velocity [−]

dimensionless r*Vu [−]


0.2
0.2

0.15

0.1
0.1

0.0 0.05

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
−0.1
1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 normalized streamfunction [−]
dimensionless radial coordinate [−]

0.4
Figure 3-13. Circulation function C ≡ rVθ
versus the normalized streamfunction .
0.3
dimensionless circumferential velocity [−]

LDV measurement
without stagnation model
0.2 with stagnation model

0.2
0.1
inlet section
survey section 1
0.0 0.15 survey section 2
cone outlet
total pressure coefficient [−]

−0.1
0.1
−0.2

0.05 mass weighted average


−0.3
inlet total pressure

−0.4 0
1.2 1.0 0.8 0.6 0.4 0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2
dimensionless radial coordinate [−]
−0.05
Figure 3-12. Axial and circumferential velocity
profiles in the survey section S2. −0.1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
normalized streamfunction [−]
As we have shown in Section 3.1, for
inviscid swirling flows both the circulation Figure 3-14. Total pressure ptot ≡ p + ρV 2 / 2
function C ≡ rVθ and the total pressure versus the normalized streamfunction .
ptot ≡ p + ρV 2 / 2 should remain constant along
streamlines in steady swirling flows. However,
when viscous flows are considered one should
expect a decay of both quantitites as the flow
evolves downstream. This decay is shown in
Figure 3-13 and Figure 3-14 within the draft tube
cone. An interesting feature of the flow
downstream a Francis runner operated at partial
140 AXISYMMETRIC SWIRLING FLOWS

3.3 Stability analysis for that breakdown occurs at the location where the
flow first supports standing waves (critical flow).
axisymmetric swirling
flows 3.3.1 Stability analysis using
axisymmetric Navier-Stokes
Huerre and Monkewitz [25] reviewed the equations
developments in the hydrodynamic stability
theory of spatially developing flows pertaining to The stability analysis of axisymmetric swirling
absolute/convective and local/global instability flow is generally performed within the parallel
concepts. They consider only open flows, where flow approximation. This assumption implies
the fluid particles do not remain within the zero radial velocity, thus the mean, steady
physical domain of interest but are advected swirling flow is described by the following radial
through downstream flow boundaries. The terms profiles:
„local” and „global” refer to instability of the
local velocity profile and of entire flow, Vz = Vz ( r ) , Vr = 0,
. (3-84)
respectively. At the local level of description, (i) Vθ = Vθ ( r ) , P = P ( r )
if localized disturbances spread upstream and
downstream and contaminate the entire parallel Usually all flow parameters are made
flow, the velocity profile is said to be locally dimensionless using the characteristic axial
absolutely unstable, (ii) if disturbances are swept velocity Vz C . The flow is characterized by two
away from the source, the velocity profile is said non-dimensional parameters: a Reynolds number,
to be locally convectively unstable. It is argued
the the notions of local absolute/convective rC Vz C
Re = , (3-85)
instability provide a rigorous justification for ν
selecting the spatial theory in open flows (given
real frequency and unknown complex wave where rC is a characteristic radius (e.g. the
number) rather than temporal theory (with given dispersion radius of the vorticity) and ν is the
real wave number and unknown complex kinematic viscosity; and a swirl parameter,
frequency).
Vθ C
Khorrami et al. [30] derive the stability Sw = , (3-86)
equations for swirling flows, and develop a Vz C
spectral collocation method for studying both
temporal and spatial stability of incompressible where Vθ C is a characteristic azimuthal velocity.
swirling flows. To analyse the linear stability of the above basic
In general, stability theories use linearized flow, the velocity and pressure fields are
flow equations about a base steady flow, then use decomposed into their mean parts (3-84) and
a Fourier series in the circumferential direction small perturbations. The perturbations are
and apply a Fourier transform in the axial decomposed in the standard form,
direction. However, Cary [14] argue that for
spatially developing flow it is more rigorous to ⎧Vr ⎫ ⎧ F ( r ) ⎫
⎪ ⎪ ⎪ G(r) ⎪
consider only the Fourier decomposition of the ⎪Vθ ⎪ ⎪ ⎪
Navier-Stokes equations in the circumferential ⎨ ⎬=⎨ ⎬ exp ⎡⎣ a z + i ( nθ − ωt ) ⎤⎦
V
⎪ ⎪ ⎪
z
H ( r ) ⎪ (3-87)
direction.
In this section we briefly review the model ⎩⎪ P ⎭⎪ ⎪⎩ Π ( r ) ⎪⎭
currently used for investigating the spatial = S( r ) exp ⎣⎡ a z + i ( nθ − ωt ) ⎦⎤
stability of swirling flows using primary
variables formulation, then we focus on the so- The complex amplitude of the perturbation S ( r )
called wave theory [9] first employed to explain depends only on the radial coordinate in the
the vortex breakdown phenomenon. parallel-flow approximation. The non-
Experimental measurements and observations dimensional order-of-unity complex axial
indicate that flows upstream of breakdown wavenumber is defined as a ≡ γ + iα , where the
cannot support standing or upstream travelling
waves (supercritical flow), while these waves ale real part γ is the exponential growth rate, and
allowed downstream of breakdown (subcritical the imaginary part α is the axial wavenumber.
flow). The wave theories of breakdown claim The circumferential wavenumber n is equal to
VORTEX DOMINATED FLOWS 141

zero for axisymmetric perturbations, and real part of the eigenvalue, γ , is positive,
different from zero for non-axisymmetric provided that the group velocity is also positive.
perturbations. The non-dimensional frequency of Herrada and Barrero [24] use a more general
the perturbation is ω . version of the local stability analysis by allowing
This approach has been employed, for a variation with z of both the mean flow as well
example, by Parras and Fernandez-Feria [40] as the perturbations. The resulting parabolized
who substitute (3-87) into the incompressible and stability equations can be written in matrix form
axisymmetric Navier-Stokes equations. After as
neglecting the second-order terms in the small
perturbations, one obtains the following set of 1 ∂S
M⋅ =
linear stability equations [40]: Re ∂z
, (3-90)
1 a2 ⎡ 1 a2 ⎤
= ⎢ L1 + a L 2 + L3 + L4 ⋅ S
Re ⎥⎦
L ⋅ S = 0, where L = L1 + a L 2 + L3 + L4 ,
Re Re ⎣ Re
⎛ d 1 n ⎞
⎜ dr + r i
r
0 0⎟
and the axial derivatives are discretized using
⎜ ⎟
⎜ i ⎛ nVθ − ω ⎞ −
2Vθ
0
d ⎟ first order differences ∂S / ∂z = ( S j +1 − S j ) / Δz ,
⎜ ⎜⎝ r ⎟
⎠ r dr ⎟
L1 = ⎜ ⎟ where j is the index in the axial direction and
⎜ dVθ Vθ ⎛ nVθ ⎞ n⎟
⎜ dr + r i⎜ −ω⎟ 0 i ⎟ Δz is the incremental step.
⎜ ⎝ r ⎠ r

⎜ dVz ⎛ nVθ ⎞ ⎟
⎜ 0 i ⎜ − ω ⎟ 0 ⎟ 3.3.2 Stability analysis of the Swirling
⎝ dr ⎝ r ⎠ ⎠
⎛ 0 0 1 0⎞ ⎛ 0 0 0 0⎞
Flow Downstream a Francis Turbine
⎜V 0 0 0 ⎟ ⎜ −1 0 0 0 ⎟ (3- Runner
L2 = ⎜ z ⎟, L4 = ⎜ ⎟ 88)
⎜ 0 Vz 0 0 ⎟ ⎜ 0 −1 0 0 ⎟
⎜ ⎟ ⎜ ⎟ While most theoretical studies are performed
⎝ 0 0 Vz 1 ⎠ ⎝ 0 0 −1 0 ⎠
on simple swirling flow configurations, practical
⎛ 0 0 0 0⎞
⎜ ⎟ applications require the analysis of real
⎜ −D2 + n + 1
2
2n
i 2 0 0⎟ turbomachinery flows. Susan-Resiga et al. [53]
⎜ r r2 r ⎟
L3 = ⎜ ⎟ investigate the real flow at the outlet of a Francis
⎜ 2n n2 + 1
−i 2 − Dr2 + 2 0 0⎟ turbine runner, in order to elucidate the causes of
⎜ r r ⎟
⎜ ⎟ a sudden drop in the draft tube pressure recovery
n2
⎜ 0 0 − Dr2 + 2 0⎟ coefficient at a turbine discharge near the best
⎝ r ⎠
d2 1 d
efficiency point. The main idea is that the
Dr ≡ 2 +
2
. swirling flow ingested by the conical diffuser
dr r dr
located downstream the runner determines the
When solving for a swirling flow in a pipe, at behaviour and performances of the draft tube.
the pipe wall the velocity vanishes and as a result The axial and circumferential velocity
the velocity perturbations should vanish as well, components are measured with a two-component
F = G = H = 0 . At the axis, r = 0 , the following probe Laser Doppler Anemometer, using back-
conditions should be satisfied: scattered light and transmission by optical fiber,
F = G = 0, dH / dr = 0, ( n = 0 ) , within the FLINDT project [6]. The survey
section is located downstream the turbine throat,
F ± iG = 0, dF / dr = 0, H = 0, ( n = ±1) , (3-89) in the upper part of the draft tube cone, Figure
F = G = H = 0, ( n > 1) . 3-15. All velocity values are made dimensionless
by the runner speed × throat radius = ω × Rref ,
In the spatial stability analysis carried out by and length are made dimensionless with respect
Parras and Fernandez-Feria [40], for a given real to the Rref = throat radius . Measurements were
frequency ω , azimuthal wavenumber n ,
Reynolds number Re , and flow configuration made at optimum head and variable discharge
around the best efficiency point. The discharge
(3-84), the system of equations (3-88) with
boundary conditions (3-89) constitutes a coefficient ϕ = Q / ⎡⎣ (ω Rref ) (π Rref
2
)⎤⎦ was in the
nonlinear eigenvalue problem for the complex range 0.34...0.41, with a value of 0.368 at
eigenvalue a . The flow is considered unstable turbine’s best efficiency point.
when the disturbance grows with z , i.e. when the
142 AXISYMMETRIC SWIRLING FLOWS

⎡ ⎛ R02 ⎞ ⎤
ϕ = U 0 R02 + U1 R12 ⎢1 − exp ⎜ − 2 ⎟⎥
⎣ ⎝ R1 ⎠ ⎦
(3-93)
Spiral case ⎡ ⎛ R2 ⎞⎤
+ U 2 R ⎢1 − exp ⎜ − 02 ⎟ ⎥
2
2
Runner blades ⎣ ⎝ R2 ⎠ ⎦
The comparison of the discharge coefficient
computed with the measured turbine discharge
Guide vanes
and the value obtained form (3-93) is a good
Stay vanes
Rref indicator of the accuracy of velocity profile
Survey section representation. All eight parameters in (3-91) and
Optical access R0 (3-92), namely the characteristic radii R1 and
Window
R2 , characteristic angular velocities Ω0 , Ω1 and
Figure 3-15. Sketch of the Francis turbine model Ω 2 , and characteristic axial velocities U 0 , U1
and LDA setup for the flow survey section. and U 2 , respectively, are determined using a
least-squares method such that the analytical
The velocity field at the survey section from representation is the best fit of experimental data.
Figure 3-15 is practically axisymmetric, but it has Figure 3-16 through Figure 3-21 show the
a rather complicated structure which cahnges experimental data for axial and circumferential
with the operating point. Before attempting any velocity profiles measured for constant head and
stability analysis, an analytical representation of variable discharge operating points, together with
the radial profiles for axial and circumferential the analytical velocity profiles (3-91) and (3-92)
velocity components is necessary. A simple fitted with a least-squares procedure [53].
swirl, with uniform axial velocity Vz ( r ) = U 0 and
solid body rotation Vθ ( r ) = Ω0 r proved to be a 0.5

rather crude approximation of the experimental 0.45


data. Susan-Resiga et al. [53] suggested that two 0.4
Batchelor vortices [7], one co-rotating and the
0.35
other counter-rotating with respect to Ω0 , and
0.3
co-flowing/counter-flowing with respect to U 0 ,
0.25
respectively, should be superimposed for
Dimensionless Velocity

consistency with experimental data for 0.2

circumferential and axial velocity profiles: 0.15

0.1
R2 ⎡ ⎛ r 2 ⎞⎤
Vθ ( r ) = Ω 0 r + Ω1 1 ⎢1 − exp ⎜ − 2 ⎟⎥ 0.05
r ⎣ ⎝ R1 ⎠ ⎦
(3-91) 0
R ⎡
2
⎛ r 2 ⎞⎤
+ Ω2 2
⎢1 − exp ⎜ − 2 ⎟⎥
−0.05
r ⎣ ⎝ R2 ⎠ ⎦ −0.1

−0.15
⎛ r2 ⎞ axial velocity
Vz ( r ) = U 0 + U1 exp ⎜ − 2 ⎟ −0.2 circumferential velocity
⎝ R1 ⎠ −0.25
(3-92) −1.0 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.6 1.0
⎛ r ⎞ 2
Dimensionless Radius
+ U 2 exp ⎜ − 2 ⎟
⎝ R2 ⎠ Figure 3-16. Axial and circumferential velocity
If R0 is the dimensionless survey section radius, profiles at discharge ϕ = 0.340 .
then the discharge coefficient can be obtained by
integrating the axial velocity profile (3-92),
VORTEX DOMINATED FLOWS 143

0.5 0.5

0.45 0.45

0.4 0.4

0.35 0.35

0.3 0.3

0.25 0.25
Dimensionless Velocity

Dimensionless Velocity
0.2 0.2

0.15 0.15

0.1 0.1

0.05 0.05

0 0

−0.05 −0.05

−0.1 −0.1

−0.15 −0.15
axial velocity axial velocity
−0.2 −0.2 circumferential velocity
circumferential velocity
−0.25 −0.25
−1.0 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.6 1.0 −1.0 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.6 1.0
Dimensionless Radius Dimensionless Radius

Figure 3-17. Axial and circumferential velocity Figure 3-19. Axial and circumferential velocity
profiles at discharge ϕ = 0.360 . profiles at discharge ϕ = 0.380 .

0.5 0.5

0.45 0.45

0.4 0.4

0.35 0.35

0.3 0.3

0.25 0.25
Dimensionless Velocity

Dimensionless Velocity

0.2 0.2

0.15 0.15

0.1 0.1

0.05 0.05

0 0

−0.05 −0.05

−0.1 −0.1

−0.15 −0.15
axial velocity axial velocity
−0.2 circumferential velocity −0.2 circumferential velocity
−0.25 −0.25
−1.0 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.6 1.0 −1.0 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.6 1.0
Dimensionless Radius Dimensionless Radius

Figure 3-18. Axial and circumferential velocity Figure 3-20. Axial and circumferential velocity
profiles at discharge ϕ = 0.368 . profiles at discharge ϕ = 0.390 .
144 AXISYMMETRIC SWIRLING FLOWS

0.55 4

0.5 3.5

0.45 3

Vortex Angular Velocity [−]


0.4 2.5
Vortex 0
0.35 2
Vortex 1
1.5 Vortex 2
0.3
Dimensionless Velocity

1
0.25
axial velocity 0.5
0.2 circumferential velocity
0
0.15
−0.5
0.1
−1
0.05 0.33 0.35 0.37 0.39 0.41
Discharge Coefficient [−]
0

−0.05 Figure 3-23. Characteristic angular velocities


−0.1 versus the discharge coefficient.
−0.15

−0.2
−1.0 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.6 1.0 0.5
Dimensionless Radius

Figure 3-21. Axial and circumferential velocity 0.4


profiles at discharge ϕ = 0.410 .
Vortex Core Radius [−]

0.3
The excellent agreement between analytical Vortex 1
formulae and experiments led to the conclusion Vortex 2
0.2
that the flow physics is correctly captured. The
parameters describing the velocity profiles have a
rather simple variation with the discharge 0.1
coefficient, in general linear or at most parabolic,
as shown in Figure 3-22 for characteristic axial
0
velocities, Figure 3-23 for characteristic angular 0.33 0.35 0.37 0.39 0.41
velocities and Figure 3-24 for characteristic Discharge Coefficient [−]

Batchelor vortex radii. Figure 3-24. Characteristic vortex radii versus


0.5 the discharge coefficient.
0.4
0.3 The flow parametrization shown above provide
the complex axial and circumferential velocity
Vortex Axial Velocity [−]

0.2
components profile as function of a single
0.1
parameter, the discharge coefficient. Among the
0 simplifications assumed one can easily notice the
−0.1 lack of boundary layer near the wall. Moreover,
−0.2 for the stability analysis we also neglect the flow
Vortex 0 evolution along the axis, even if the draft tube
−0.3
Vortex 1 cone has an included angle of 17 .
−0.4 Vortex 2 Since the swirling flow is considered steady
−0.5
0.33 0.35 0.37 0.39 0.41 and the wall boundary layer is neglected, the
Discharge Coefficient [−] mathematical model corresponds to the Bragg-
Hawthorne equation (3-33). After re-writing the
Figure 3-22. Characteristic axial velocities axial velocity profile as
versus the discharge coefficient.
VORTEX DOMINATED FLOWS 145

Vz ( y ) = U 0 + U1 exp ( − y / Y1 )
, (3-94)
+ U 2 exp ( − y / Y2 )

with Y1 = R12 / 2 and Y2 = R22 / 2 , one can integrate


(3-94) to obtain the streamfunction distribution
Ψ ( y ) = U 0 y + U1Y1 ⎡⎣1 − exp ( − y / Y1 ) ⎤⎦
(3-95)
+ U 2Y2 ⎡⎣1 − exp ( − y / Y2 )⎤⎦

If Ψ = 0 on the axis, on the wall we have εψ


Ψ wall = ϕ / 2 . Axial
For a negligible radial velocity, i.e. parallel Velocity
flow, the last two terms in (3-33) can be
evaluated as follows: ψ
C ( Ψ ) C′( Ψ )
E′( Ψ ) − := Ψ ⎯⎯⎯⎯
Eq.(3.91)
→y ϕ/2 ψ
2y Circumferential
(3-96) Velocity
U1 ⎛ y⎞ U ⎛ y⎞
→− exp ⎜ − ⎟ − 2 exp ⎜ − ⎟ r r
Y1 ⎝ Y1 ⎠ Y2 ⎝ Y2 ⎠
The map Ψ → y has to be computed Figure 3-25. Synoptic view of the model for
numerically, for example using the Newton swirling flow downstream of a Francis turbine
iterative method as shown in (3-65). runner.
In order to investigate the flow stability we
consider in (3-33) a perturbed streamfunction of For example, by using finite elements, [53], we
the form obtain the eigenvalue problem in matrix form
Ψ ( y ) + εψ ( y ) exp ( iκ z ) , (3-97) A ψ = κ 2B ψ ,
dN i ( y ) dN j ( y )
where Ψ ( y ) is the base flow given by (3-97) , Aij = − ∫ dy
dy dy
ψ ( y ) is a perturbation of the base flow, Figure . (3-100)
− ∫ N i ( y ) K ( y ) N j ( y ) dy ,
3-25, and κ is the axial wavenumber of this
perturbation. Introducing (3-97) into (3-33) one Ni ( y) N j ( y)
obtains the linearized equation Bij = ∫ dy .
2y
⎡ E ′′ ( Ψ ) − ⎤
d 2ψ ⎢ 2 ⎥ In (3-100) K ( y ) denotes the expression in
− ⎢ C ′ ( Ψ ) + C ( Ψ ) C ′′ ( Ψ ) ⎥ψ
dy 2 square brackets from (3-98). The GVCSP
⎢ 2y ⎥ (3-98) subroutine from the International Math and
⎣ ⎦
κ2 Statistics Libraries (IMSL) can be used to
= ψ compute all of eigenvalues and eigenvectors of
2y the generalized real symmetric eigenvalue
In order to preserve the overall flow rate for the problem (3-100).
perturbed flow as well, the perturbation must Let us summarize now the swirling flow
satisfy homogeneous boundary conditions model according to the synoptic Figure 3-25.
Once the analytical representation for axial and
ψ ( 0 ) = ψ (Y0 ) = 0 . (3-99) circumferential velocity components has been
established, the mean flow streamfunction can be
Equations (3-98) and (3-99) define a generalized computed. A streamtube Ψ = constant may be
eigenvalue problem. The eigenvalues κ 2 can be subject to axisymmetric perturbations εψ ( y )
computed numerically once the problem is
discretized. which are eigenfunctions of problem (3-98)
(3-99). Such a perturbation can be sustained
146 AXISYMMETRIC SWIRLING FLOWS

or not depending on the sign of the eigenvalue becomes positive, followed by the
corresponding eigenvalue κ 2 . If κ 2 < 0 then next eigenvalues as ϕ is further decreased, and
κ is imaginary and the exponential factor in the flow is subcritical with standing waves
(3-97) will be exp ( ± κ z ) . As we move described by the corresponding eigenvectors ψ .
The critical state occurs at the discharge
downstream the current section (increasing ϕ = 0.365 , according to Figure 3-26, and this
z ) the only physically acceptable solution
value is quite close (only 1.3% smaller) to the
corresponds to exp ( − κ z ) showing an value of ϕ = 0.37 where the sudden drop in the
exponential damping of ψ . A swirling flow draft tube pressure recovery coefficient is
for which all eigenvalues are negative is observed. As a result, it seems reasonable to
unable to sustain axisymmetric small- assume that the critical state is directly related to
disturbance standing waves and it was this experimentally observed phenomenon, since
by trying several draft tube geometries while
termed supercritical by Benjamin [9]. On the
keeping the same runner (and the swirling flow)
other hand, if at least one eigenvalue κ 2 is the same behaviour has been observed practically
positive, then the perturbation will take the at the same discharge value.
form of a standing wave exp ( ± i κ z ) and the While reaching the critical swirl
corresponding swirling flow is termed configuration seems to be the cause, the actual
physical mechanism by which the pressure
subcritical. All physical interpretations
recovery suffers an abrupt change cannot be
attempted for the distinction between inferred from the present analysis. Experimental
supercritical and subcritical states were [6] as well as numerical [36] investigation offer a
mainly focused at the vortex breakdown comprehensive analysis of the Francis turbine
phenomena. draft tube flow.
0.34 0.35 0.36 0.37 0.38 0.39 0.4 0.41 The previous analysis leads to the
300 0.8
conclusion that when designing or optimizing
250 turbine runners one should avoid reaching a
Wall Pressure Recovery Coefficient

pressure recovery coeff.


Eigenvalues (wavenumber squared)

200
0.7
critical state for the swirl at the runner outlet
150 1st eigenval.
2nd eigenval.
within the normal operating range.
100 3rd eigenval.

3.4 Axisymmetric
4th eigenval.
50 0.6
0
turbomachinery swirling
−50
−100
0.5 flow
−150
The preliminary analysis and design of
−200 0.4
0.34 0.35 0.36 0.37 0.38 0.39 0.4 0.41 turbomachinery flow can be performed within the
Discharge Coefficient
inviscid fluid assumption, since losses can be
Figure 3-26. The first four eigenvalues and the considered negligible at the design operating
point. Of course, maximizing the machine
pressure recovery coefficient function of the
efficiency requires the evaluation of viscous
discharge coefficient.
losses, but the first step is to get a preliminary
Once the eigenvalues computed we can examine design within the loss-free framework.
the transition of the swirling flow downstream a For the absolute steady flow (stay vanes,
Francis turbine runner from subcritical to guide vanes, bladeless regions) the continuity and
supercritical as the discharge coefficient momentum equations are
increases. Moreover, we can correlate the critical ∇⋅V = 0 continuity,
state with the experimentally observed sudden ( ∇ × V ) × V = −∇E momentum,
drop in the draft tube pressure recovery (3-101)
coefficient. From Figure 3-26 one can see that for p V2 specific energy
E≡ +
ϕ > 0.365 all eigenvalues defined by (3-98) ρ 2 per unit mass.
(3-99) are negative, thus the flow is supercritical Obviously, the momentum equation gives
and it cannot sustain axisymmetric standing V ⋅ ∇E = 0 , i.e. the specific mechanical energy
waves. However, for ϕ < 0.365 the larges
VORTEX DOMINATED FLOWS 147

remains constant along a streamline for the axisymmetric model for the bub-to-shroud
absolute velocity field. turbomachinery flow considers that the
For the runner bladed region it is streamsurfaces retain axial symmetry within the
convenient to consider the relative flow blade regions as well.
equations, where the relative velocity is The simplified axisymmetric computation
W = V − Ω × r and the absolute specific energy retains only the average inter-blade pressure
E is replaced by the relative specific since no circumferential gradient is allowed. As a
energy E R = E − Ω ( rVθ ) . The corresponding result, we need an artificial quantity, the blade
body force B , to account for the blade-flow
steady relative flow equations are
interaction. This body force introduced in the
∇⋅W = 0 continuity, axisymmetric flow model should turn the flow
( ∇ × V ) × W = −∇E R momentum, equivalently with the actual blades. The
relative (3-102) axisymmmetric absolute flow equations (3-101)
W 2 ( Ωr )
2
p become
ER ≡ + − specific
ρ 2 2
energy. ∇⋅V = 0 continuity,

From the momentum equation for relative flow


( ∇ × V ) × V =
momentum,
− ∇E + B
we have W ⋅ ∇ER = 0 , i.e. the relative specific (3-103)
specific energy
energy ER remains constant along relative flow p V2
E≡ + per unit mass
streamlines. ρ 2 in absolute flow

while the corresponding axisymmetric


relative flow equations are
∇⋅W = 0 continuity,
(∇ × V ) × W = momentum,
− ∇E R + B
specific (3-104)
W 2 ( Ωr )
2
p energy per
ER ≡ + −
ρ 2 2 unit mass in
relative flow
Within the bladed regions, the flow is considered
Figure 3-27. Three-dimensional effects of blade to take place on an absolute or relative
loading and the axisymmetric flow assumption. streamsurface defined as
α ( z, r,θ ) ≡ θ − f ( z, r ) = constant , (3-105)
For the three-dimensional flow in the bladed
regions, the blade produces a pressure difference where f ( z, r ) defines the blade wrap angle,
between suction and pressure sides, which in turn
within the concept of infinite number of blades.
produces a circumferential pressure gradient that
Conceptually, the constant α surface
deflects the flow. Figure 3-27 shows that the flow
corresponds to the S2-surface concept proposed
inside the interblade channel is essentially three-
by Wu [59]. For example, a surface described as
dimensional since the streamlines originating on
in Eq.(3-105) may be seen as corresponding to
a circle (normal to the machine axis) do not
the thin blade from Figure 3-27, in the limit of
remain on an axisymmetric surface. Instead, the
zero thickness. With the unit normal vector to the
streamlines close to the blade pressure side are
streamsurface,
pushed radially towards the hub, while the
streamlines in the neighborhood of blade suction r ∂f r ∂f
eθ − er − ez
side are deflected toward the shroud. It is obvious ∇α ∂r ∂z
n≡ = , (3-106)
from Figure 3-27 that at the blade trailing edge ∇α 2 2
⎛ r ∂f ⎞ ⎛ r ∂f ⎞
there is a significant radial departure between 1+ ⎜ ⎟ +⎜ ⎟
streamlines originating at the same radius ⎝ ∂r ⎠ ⎝ ∂z ⎠
upstream the blade. However, a simplified
the flow tangency condition requires
148 AXISYMMETRIC SWIRLING FLOWS

U = Ωr . After using the Gauss theorem to


n ⋅ V = 0 absolute flow
. (3-107) transform the volume integral into an integral
n ⋅ W = 0 relative flow over the domain boundary, we have taken into
The blade body force B is normal to either the account that V ⋅ n = 0 on the hub and shroud
absolute or relative velocity vector (i.e. along the walls. Obviously, on the inlet section we have
normal direction to the absolute/relative V ⋅ n IN < 0 , while on the outlet section we have
streamsurface), V ⋅ nOUT > 0 . If mass weighted average values for
B ⋅ V = 0 absolute flow UVθ are considered on the inlet/outlet section,
, (3-108) one can write the fundamental turbomachinery
B ⋅ W = 0 relative flow
equation in the simple form
and it can be written as,
P = m (UVθ )OUT − (UVθ ) IN , with m the mass
⎡ ⎤
⎣ ⎦
B = B ( z , r ) ∇α , with B = V ⋅ ∇ ( rVθ ) . (3-109) flow rate through the turbomachine.
In Eq.(3-109) the function B ( z, r ) is related to We can now replace the three projections
of the momentum equations by
the change in the fluid’s angular momentum i) Flow tangency condition on the absolute or
imparted by the blade rows. One can easily check relative streamsurface, (3-107);
from (3-109) that the blade body force is ii) Projection on the streamsurface, along the
measured in ⎡⎣ m / s 2 ⎤⎦ , as it should in velocity vector, i.e. V ⋅ ∇E = 0 for absolute
absolute/relative momentum equations. Since the flow and W ⋅ ∇ER = 0 for relative flow;
constant α surface may be seen as an iii) Projection on the streamsurface, normal to
approximation for a thin blade surface, ∇α is a the velocity vector. This projection leads to
vector normal to the blade surface. The blade the main scalar equation for steady absolute
body force B vanishes outside the blade region or relative turbomachinery axisymmetric
since rVθ does not change anymore along the flow.
absolute streamlines. The definition (3-109) is The projection of the momentum equation ii)
consistent with the Euler equation of shows that E R does not change along a
turbomachinery. The torque generated by the streamline in relative flow. Since by definition
circumferential component of the blade body we have ER = E − UVθ , the fundamental
force, Bθ = ⎡⎣ V ⋅ ∇ ( rVθ )⎤⎦ / r is turbomachinery equation can be rewritten as
P = m ⎡⎣ EOUT − E IN ⎤⎦ , thus showing that, if no
T= ∫ r ρ Bθ dVol
Vol
viscous losses are accounted for, the mechanical
. (3-110) power transferred from/to the runner corresponds
= ∫ ⎡⎣ ρ V ⋅ ∇ ( rVθ )⎤⎦ dVol to the increase/decrease of the fluid specific
Vol
energy E .
Note that thanks to the continuity equation, the
term in the square brackets can be written as
∇ ⋅ ⎡⎣ ρ V ( rVθ )⎤⎦ . The power transferred by the
runner, rotating with the angular speed Ω ,
from/to the fluid is
P = TΩ = ∫ ∇ ⋅ ⎣⎡ ρ V ( rVθ )⎦⎤ dVol
Vol

= ∫ ( ΩrVθ ) ρ V ⋅ n dS
S
Figure 3-28. Unit vectors on a streamsurface, as
(3-111) defined in Eqs. (3-106) and (3-112).
= ∫ (UVθ ) ρ V ⋅ n IN dS
S IN
Let us focus now on iii), and follow the
+ ∫ (UVθ ) ρ V ⋅ nOUT dS derivation of the main scalar equation for the
SOUT relative flow. The direction on the streamsurface,
normal to the velocity vector is given by the
Eq.(3-111) is the fundamental turbomachinery following unit vector
equation. We have denoted the transport velocity
VORTEX DOMINATED FLOWS 149

W ∂Vr ∂Vz ∂f ∂ ( rVθ ) ∂f ∂ ( rVθ )


N≡ ×n , (3-112) − = −
W ∂z ∂r ∂z ∂r ∂r ∂z
shown in Figure 3-28. After a straightforward 1 ⎡ ∂E ⎛ r ∂f ⎞ ∂E ⎛ r ∂f ⎞⎤
+ 2 ⎢ R ⎜ Wr + Wθ ⎟ − R ⎜ Wz + Wθ ⎟ ⎥
computation, and using the flow tangency W ⎣ ∂z ⎝ ∂r ⎠ ∂r ⎝ ∂z ⎠⎦
condition (3-107), (3-116)
⎛W ⎞
⎜ × n ⎟ ⋅ ⎣⎡ − W × ( ∇ × V ) ⎦⎤ Although currently used since ‘70s, Eq.(3-116) is
⎝W ⎠
still the basic mathematical model for hydraulic
⎧ W⎫ turbomachinery primary design, [42][41].
= ⎨ ⎣⎡ − W × ( ∇ × V )⎦⎤ × ⎬ ⋅ n
⎩ W⎭ Bosman [11] has a word of caution concerning
n Eq.(3-116): where the streamsurface tangent
{
= ⋅ W × ⎡⎣ W × ( ∇ × V ) ⎤⎦
W
} plane subtends a large angle to the meridional
n plane, i.e. ( nr / nθ ) and ( nz / nθ ) become large,
{
= ⋅ −W 2 ( ∇ × V ) + ⎡⎣ W ⋅ ( ∇ × V ) ⎤⎦ W
W
} then Eq.(3-116) fails to converge numerically due
to the amplification of the right-hand side by
= −Wn ⋅ ( ∇ × V )
these factors. This situation is anticipated to be
we obtain the projection along N of the left-hand more prevalent in radial-flow machines.
side of the momentum equation, For practical use it is convenient to re-
write (3-116) using the Stokes’ streamfunction
N ⋅ ⎡⎣ ( ∇ × V ) × W ⎤⎦ = −Wn ⋅ ( ∇ × V ) . (3-113) Ψ for axisymmetric flows. Within the blade
regions of a turbomachine the finite blade
For the right-hand side of the momentum
thickness is accounted for through the
equation projection along N we have dimensionless blade blockage coefficient b < 1 .
( W × n ) ⋅ ∇E R = Within blade-less regions we obviously have
b = 1 . From the circumferentially averaged
∂ER
− ( nθWz − nzWθ ) (3-114)
continuity equation [27]
∂r
∂E ∂ ( brVz ) ∂ ( brVr )
+ R ( nθWr − nrWθ ) + =0, (3-117)
∂z ∂z ∂r
After recalling the vorticity vector in cylindrical we can introduce the streamfunction as
coordinates,
∂V ⎛ ∂V ∂V ⎞ 1 ∂ ( rVθ ) ∂Ψ ∂Ψ
∇ × V = − θ e r + ⎜ r − z ⎟ eθ + ez = rbVz and = − rbVr
∂z ⎝ ∂z ∂r ⎠ r ∂r ∂r ∂z
we use (3-113) and (3-114) to obtain the for absolute flow, and
(3-118)
N − projection of the momentum equation, also ∂Ψ ∂Ψ
= rbWz and = − rbWr
called the principal equation of motion: ∂r ∂z
for relative flow.
⎛ ∂V ∂V ⎞ 1 ∂ ( rVθ ) 1 ∂ ( rVθ )
nθ ⎜ r − z ⎟ + nz − nr As a result, the principal equation for
⎝ ∂z ∂r ⎠ r ∂r r ∂z turbomachinery swirling flow Eq.(3-116) can be
.
1 ⎡ ∂E R ∂E R ⎤ re-written as
= 2⎢ ( nθWr − nrWθ ) − ( nθWz − nzWθ )⎥
W ⎣ ∂z ∂r ⎦ ∂ ⎛ 1 ∂Ψ ⎞ ∂ ⎛ 1 ∂Ψ ⎞
⎜ ⎟+ ⎜ ⎟=
(3-115) ∂z ⎝ br ∂z ⎠ ∂r ⎝ br ∂r ⎠
From (3-106) we have the streamsurface normal ∂f ∂ ( rVθ ) ∂f ∂ ( rVθ )
orientation −
∂r ∂z ∂z ∂r
nz r ∂f n r ∂f (3-119)
=− and r = − , ⎡ ∂E R ⎛ r ∂f ⎞ ⎤
nθ ∂z nθ ∂r ⎢ ⎜ Wz + Wθ ⎟ − ⎥
1 ∂r ⎝ ∂z ⎠ ⎥
+ 2⎢
and the principal equation (3-115) can be W ⎢ ∂E R ⎛ r ∂f ⎞⎥
rewritten as ⎢ − ∂z ⎜ Wr + ∂r Wθ ⎟ ⎥
⎣ ⎝ ⎠⎦
150 AXISYMMETRIC SWIRLING FLOWS

In this form, one has to solve an elliptic partial flow streamsurface shape f . The term dE R / dΨ
differential equation for the streamfunction Ψ , is known as function of Ψ from the upstream
with the corresponding source term in the right- conditions. The second relationship between Ψ
hand side. Usually, the source term is treated and f is given by the flow tangency condition
explicitly within an iterative procedure. However,
(3-107) rewritten as
for loss-free flows the term
1 ⎡ ∂E R ⎛ r ∂f ⎞ ∂E ⎛ r ∂f ⎞⎤ ⎛ 1 ∂Ψ ⎞ ∂f ⎛ 1 ∂Ψ ⎞ ∂f Vθ
2 ⎢ ⎜ Wz + Wθ ⎟ − R ⎜ Wr + Wθ ⎟ ⎥ , ⎜ ⎟ −⎜ ⎟ = −Ω. (3-121)
W ⎣ ∂r ⎝ ∂z ⎠ ∂z ⎝ ∂ r ⎠⎦ ⎝ rb ∂r ⎠ ∂z ⎝ rb ∂z ⎠ ∂r r
can be treated implicitly. Since E R depends in The system of partial differential equations
this case only on the streamfunction (being (3-120) and (3-121), with appropriate boundary
constant along the streamlines in the meridian conditions, allows the computation of both
half-plane), we have Ψ ( z, r ) and f ( z, r ) in the runner region. Borges
∂E R dE R ∂Ψ dE [10] uses this mathematical model to design
= = rbWz R , and
∂r dΨ ∂r dΨ mixed-flow pumps. By assuming a uniform non-
∂E R dE R ⎛ ∂Ψ ⎞ dE R swirling inlet flow, the term dE R / dΨ vanishes.
− = ⎜− ⎟ = rbWr
∂z dΨ ⎝ ∂z ⎠ dΨ Also, when thin blades are considered in a first
Note that when losses are accounted for, E R (or approximation, the blade blockage coefficient is
b = 1 . Borges discretizes the corresponding
E ) does not remain constant along a relative simplified equation (3-120) with a second-order
(absolute) axisymmetric streamtube; it actually accurate finite difference scheme, resulting in a
decreases downstream due to viscous dissipation. nine-point stencil on a structured quadrilateral
From the flow tangency condition (3-107) grid of the meridional domain. Equation (3-121)
we have is a first order partial differential equation with
W ∂f ∂f characteristic lines coincident with the
W ⋅ ∇α = θ − Wz − Wr = 0,
r ∂z ∂r streamlines Ψ = constant in the meridian half-
r ∂f r ∂f plane. In order to integrate this differential
or Wθ = Wz + Wr
∂z ∂r equation, some initial data must be specified
A straightforward computation gives along a line roughly perpendicular to these
1 ⎡ ∂ER ⎛ r ∂f ⎞ ∂E ⎛ r ∂f ⎞⎤ characteristic lines and extending from hub to
2 ⎢ ⎜ Wz + Wθ ⎟ − R ⎜ Wr + Wθ ⎟ ⎥ shroud. This initial data on f are the stacking
W ⎣ ∂r ⎝ ∂z ⎠ ∂z ⎝ ∂r ⎠⎦
condition of the blade. Borges [10] implements
⎡ dE ⎛ r ∂f ⎞ ⎤ this condition by giving, as input, the values of
⎢ rbWz R ⎜ Wz + Wθ ⎟ + ⎥
1 d Ψ ⎝ ∂ z ⎠ ⎥ the blade angular coordinate f along the
= 2⎢
W ⎢ dE R ⎛ r ∂f ⎞⎥ impeller blade leading edge. Zangeneh [62] also
⎢ + rbWr dΨ ⎜ Wr + ∂r Wθ ⎟ ⎥ uses the system of equations (3-120) and (3-121),
⎣ ⎝ ⎠⎦
but the blade camber surface is no longer
dE R 1 ⎡ 2 ⎛ r ∂f r ∂f ⎞ ⎤ approximated by the constant α streamsurface.
= rb ⎢Wz + Wr + Wθ ⎜ Wz ∂z + Wr ∂r ⎟ ⎥
2

dΨ W 2 ⎣ ⎝ ⎠⎦ In this case, the velocity field is decomposed into


dE W + Wr + Wθ
2 2 2
dE circumferentially averaged and periodic
= rb R z = rb R components, by using the Clebsh formulation of
dΨ W 2
dΨ steady rotational flow. The blade shape is
As a result, the principal equation for loss-free determined by imposing the inviscid slip
axisymmetric turbomachinery swirling flows condition (i.e. blade shape aligned with the local
becomes velocity vector), and it is different from the
∂ ⎛ 1 ∂Ψ ⎞ ∂ ⎛ 1 ∂Ψ ⎞ dE R constant α streamsurface.
⎜ ⎟+ ⎜ ⎟ − rb = For blade-less regions, the Kelvin’s
∂z ⎝ br ∂z ⎠ ∂r ⎝ br ∂r ⎠ dΨ
(3-120) theorem states that the circulation function
∂f ∂ ( rVθ ) ∂f ∂ ( rVθ ) C ≡ rVθ is constant on axisymmetric
= −
∂r ∂z ∂z ∂r streamtubes, therefore we have C = C ( Ψ ) . As a
For a given ( rVθ ) distribution in the bladed result, the right-hand side of Eq.(3-120) becomes
region, Eq.(3-120) provides a relationship
between the streamfunction Ψ and the relative
VORTEX DOMINATED FLOWS 151

∂f ∂C ∂f ∂C dC ∂f dC ∂f ∂f ∂C ∂f ∂C dC ∂f dC ∂f
− = − rWr − rWz − = − rVr − rVz
∂r ∂z ∂z ∂r dΨ ∂r dΨ ∂z ∂r ∂z ∂z ∂r dΨ ∂r dΨ ∂z
dC ⎛ ∂f ∂f ⎞ dC Wθ dC ⎛ ∂f ∂f ⎞ dC Vθ C dC
= −r ⎜ Wr + Wz ⎟ = − r = −r ⎜ Vr + Vz ⎟ = − r =−
dΨ ⎝ ∂r ∂z ⎠ dΨ r dΨ ⎝ ∂r ∂z ⎠ dΨ r r dΨ
dC Vθ − Ωr dC C d C thus recovering (3-122) from (3-123) as well.
= −r = rΩ −
dΨ r dΨ r dΨ
3.4.1 Summary of turbomachinery
where b = 1 in the blade-less regions (there are
swirling flow equations
no blades, thus there is no blade blockage).
Moreover, since E R = E − ΩC , we have Let us summarize now the mathematical model
dE R d E dC for turbomachinery swirling flow, derived within
= −Ω .
dΨ dΨ dΨ the following simplified assumptions:
Finally, Eq.(3-120) becomes o Incompressible and inviscid fluid; no
∂ ⎛ 1 ∂Ψ ⎞ ∂ ⎛ 1 ∂Ψ ⎞ dE hydraulic losses are accounted for
⎜ ⎟+ ⎜ ⎟−r + o Axi-symmetrical steady swirling flow;
∂z ⎝ r ∂z ⎠ ∂r ⎝ r ∂r ⎠ dΨ
this assumption can be seen as
dC dC C dC considering an infinite number of zero
+ Ωr − Ωr + =0
dΨ dΨ r dΨ thickness blades.
or In order to make the equations more convenient
for numerical computations, the radial
1 ⎡ ∂ ⎛ 1 ∂Ψ ⎞ ∂ ⎛ 1 ∂Ψ ⎞ ⎤ independent variable r is replaced by the new
⎜ ⎟+ ⎜ ⎟ −
r ⎢⎣ ∂z ⎝ r ∂z ⎠ ∂r ⎝ r ∂r ⎠ ⎥⎦ variable y = r 2 / 2 . There are three dependent
(3-122)
dE C dC variables: the streamfunction Ψ ( z, y ) , the
− + =0
dΨ r 2 dΨ circulation function C ( z, y ) ≡ rVθ , and the so-
which is the Long-Squire or Bragg-Hawthorne called blade wrap angle f ( z , y ) which describe
equation for inviscid, incompressible,
the shape of a S2-surface.
axisymmetric steady swirling flows.
If the real blades thickness is to be taken into
The above derivation started from the
account, the dimensionless blade blockage
principal equation of turbomachinery swirling
flow written for relative motion (in the rotating coefficient b ( z , y ) must be know within the
blade regions). Obviously, the same result can be bladed regions. Moreover, the function dE / dΨ
obtained starting with Eq.(3-120) written for (for absolute flow), or dER / dΨ (for relative
absolute flow, flow) must be known as function of Ψ from
∂ ⎛ 1 ∂Ψ ⎞ ∂ ⎛ 1 ∂Ψ ⎞ dE flow configuration upstream the computational
⎜ ⎟+ ⎜ ⎟ − rb = domain.
∂z ⎝ br ∂z ⎠ ∂r ⎝ br ∂r ⎠ dΨ The governing equations for axisymmetric
(3-123)
∂f ∂ ( rVθ ) ∂f ∂ ( rVθ ) turbomachinery swirling flow are
= −
∂r ∂z ∂z ∂r 1 ∂ ⎛ 1 ∂Ψ ⎞ ∂ ⎛ 1 ∂Ψ ⎞
⎜ ⎟+ −
where f denotes the streamsurface for absolute 2 y ∂z ⎝ b ∂z ⎠ ∂y ⎜⎝ b ∂y ⎟⎠
flow and we use the specific energy E instead of (a )
∂E ∂f ∂C ∂f ∂C
E R . Together with −b R = − (3-125)
∂Ψ ∂y ∂z ∂z ∂y
⎛ 1 ∂Ψ ⎞ ∂f ⎛ 1 ∂Ψ ⎞ ∂f Vθ 1 ⎛ ∂Ψ ∂f ∂Ψ ∂f ⎞ C
⎜ ⎟ −⎜ ⎟ = (3-124) − = −Ω ( b)
⎝ rb ∂r ⎠ ∂z ⎝ rb ∂z ⎠ ∂r r b ⎜⎝ ∂y ∂z ∂z ∂y ⎟⎠ 2 y
we have a system of two partial differential where f ( z , y ) corresponds to a streamsurface
equations for the streamfunction Ψ and the shape
for relative flow, and Ω is the runner angular
of absolute streamsurface f . Using the Kelvin’s
speed. By setting Ω = 0 we obtain the
theorem, the right-hand side of Eq.(3-123) corresponding equations for absolute flow.
becomes Since we have only two equations and three
unknown functions, either C ( z , y ) or f ( z, y )
152 AXISYMMETRIC SWIRLING FLOWS

must be specified, while Ψ ( z , y ) is allways half plane along a Ψ = constant curve. On the
computed. As a result we have two alternatives axisymmetric S1-surface one has to design a
for using Eqs.(3-125): hydrofoil cascade that produces the scheduled
o Analysis mode: for a given streamsurface flow turning as prescribed when solving the
axisymmetric swirling flow. The absolute flow is
(blade) shape, f ( z, y ) , compute the
assumed irrotational throughout so that the only
corresponding axisymmetric velocity field; vorticity is the bound vorticity in the blade
the velocity components in a meridian half- surfaces and there is no trailing vorticity due to
plane are obtained from the streamfunction the spanwise variation of circulation about each
Ψ ( z, y ) , while the circumferential velocity blade [57].
component is given by the circulation
function C ( z , y ) ;
o Design mode: for a given distribution fo the
circulation function C ( z , y ) , compute the
shape of S2-streamsurfaces, f ( z, y ) , and the
streamfunction Ψ ( z, y ) ; a stacking curve
from hub to shroud should be given on the
inlet section as initial condition to integrate
(3-125)(b). Actual blade sections are further
designed on axisymmetric S1-streamsurfaces
obtained by revolving Ψ = constant curves
about the symmetry axis.
Efficient numerical approaches for solving
directly the three-dimensional inviscid or viscous
flow are now readily available, thus making the
above “analysis mode” obsolete. However, the
design mode is still the first choice for hydraulic
turbines and pumps preliminary design.
An iterative algorithm for solving
Eqs.(3-125) starts with solving the homogeneous
version of (3-125)(a) and obtain a first Figure 3-29. Streamlines Ψ = constant in a
approximation for Ψ ( z, y ) . In design mode, this meridian half-plane, and the flow surface
approximation together with the prescribed generated by revolving such a curve about the z-
circulation function C ( z , y ) is used to integrate axis, [38].
(3-125)(b) and obtain an approximation for
In bladeless regions, the vorticity vector is
f ( z , y ) . The right-hand side in (3-125)(a) can tangent to the axisymmetric streamsurface, i.e. its
now be evaluated, and the next iteration can be meridian component is directed along the
started by computing a new approximation for meridian streamline Ψ = constant . On the other
Ψ ( z, y ) . hand, in the bladed regions there is a vorticity
component normal to the S1-surface,
3.4.2 Blade-to-blade flow on γ ≡ ω⋅n =
axisymmetric stream-sheet
⎛ ∂Vθ ⎛ ∂V ∂V ⎞ 1 ∂ ( rVθ ) ⎞
⎜− e r + ⎜ r − z ⎟ eθ + ez ⎟
Once the streamfunction Ψ has been ⎝ ∂z ⎝ ∂z ∂r ⎠ r ∂r ⎠
computed, axisymmetric flow surfaces ⎛V V ⎞ (3-126)
corresponding to the S1-surface concept, ⋅⎜ r e z − z er ⎟ =
⎝ Vm Vm ⎠
proposed by Wu [59], can be considered by
revolving Ψ = constant curves about the 1 1 ⎡ ∂ ( rVθ ) ∂ ( rVθ ) ⎤ 1 ∂C
= ⎢Vz ∂z + Vr ∂r ⎥ = r ∂m
symmetry axis. Within such a surface, the natural r Vm ⎣ ⎦
coordinates are m and θ , Figure 3-29, where m
denotes the curvilinear coordinate in the meridian
VORTEX DOMINATED FLOWS 153

This vorticity is responsible for turning the flow corresponding circumferentially averaged values
between an upstream station m = m1 and a Vm and Vθ used to compute the flow on the S2-
downstream station m = m2 . Indeed, integrating surface. The circumferentially averaged meridian
γ on the axisymmetric streamsurface we have and circumferential velocity components satisfy
∂ ∂V
m2 m2
1 d ( rVθ ) ( rbhVm ) = 0 and θ = 0 . (3-131)
∫m γ 2π r dm = m∫ r dm 2π r dm = (3-127)
∂m ∂θ
1 1
The first equation in (3-131) corresponds to the
= ( 2π rVθ )2 − ( 2π rVθ )1 = Γ 2 − Γ1 circumferentially averaged continuity equation
on the S1-surface, stating that the volumetric
showing the change in the circulation
flowrate 2π rbhVm remains constant on the
Γ ≡ 2π rVθ = 2π C from the upstream to the
stream sheet, and it is used to determine the
downstream location. The circumferentially variation of the stream sheet thickness,
averaged vorticity γ ( m ) is going to be
rbhVm = r 0h 0Vm0
distributed only on the N B blades surface,
h 0 r 0 Vm0 , (3-132)
2π r 2π dC ⇒ h(m) =
γ (m) = γ (m) = b ( m ) r ( m ) Vm ( m )
N B N B dm (3-128)
where the superscript 0 indicates an upstream
(circulation per unit lenght) location on the S1-surface where b = 1 , chosen as
leaving the blade-to-blade flow in the S1-surface origin for the ( m,θ ) coordinate system as in
irrotational. Hawthorne et al. [22] developed an Figure 3-29.
analytical method for designing two-dimensional In terms of flow relative to the blades the
cascades with given γ distribution, and this equations become
method has been further extended to axial
machines [57] and radial-axial machines [62]. continuity
The runner cascades are designed for relative ∂ ∂
flow, with the blade bound vorticity ( rhwm ) + ( hwθ ) = 0
∂m ∂θ
(3-133)
2π d ( rWθ )
γ R (m) ≡ = irrotationality
N B dm
. (3-129) ∂ ∂ dr
2π dr ( wm ) − ( rwθ ) = 2Ωr
= γ (m) − 2Ω r ∂θ ∂m dm
NB dm
where wm ( m,θ ) = vm ( m,θ ) and
As a result, one has to solve the incompressible,
inviscid and irrotational flow on the surface of wθ ( m,θ ) = vθ ( m,θ ) − Ωr are the meridian and
revolution, with variable stream sheet thickness, circumferential relative velocity components,
h( m) , [37]. The governing equations for the respectively, and Ω is the angular velocity of the
absolute blade-to-blade flow on the S1-surface blades. The stream sheet thickness h(m ) is
are obviously computed as in Eq.(3-132).
By choosing the reference upstream
continuity
location on the S1-surface at a point with r 0
∂ ∂ radius and Vm0 = Wm0 meridian velocity, we can
( rhvm ) + ( hvθ ) = 0
∂m ∂θ rewrite (3-130) and (3-133) in a dimensionless
(3-130) form. First, we introduce the conformal mapping
irrotationality
∂ ∂ dm
( vm ) − ( rvθ ) = 0 dx1 = and dx2 = dθ . (3-134)
∂θ ∂m r

where vm and vθ are the meridian and Second, we define the dimensionless velocities
circumferential absolute velocity components, Absolute flow (3-135)
respectively, and h is the stream sheet thickness.
Note that vm and vθ do not coincide with the
154 AXISYMMETRIC SWIRLING FLOWS

rvm rv ∂ ⎛ 1 ∂ψ ⎞ ∂ ⎛ 1 ∂ψ ⎞
v1∗ ≡ 0 0
and v2∗ ≡ 0 θ 0 ⎜ ⎟+ ⎜ ⎟=
r Vm r Vm ∂x1 ⎝ h∗ ∂x1 ⎠ ∂x2 ⎝ h∗ ∂x2 ⎠
Relative flow Ωr ∂ ( r / r )
0

=2 0
rwm rw Wm ∂x1
w1∗ ≡ 0 0
and w2∗ ≡ 0 θ 0
r Wm r Wm The blade bound vorticity (3-128) becomes in
The dimensionless stream sheet thickness dimensionless form
variation h∗ ≡ h / h 0 is given by Eq.(3-132). dv2∗ 2π γ r
Equations (3-130) will be written in γ∗ ≡ = . (3-140)
dx1 N B Vm0 r 0
dimensionless form as
continuity to be used for designing stationary blades (e.g.
stay vanes or guide vanes for hydraulic turbines).
∂ ∗ ∗ ∂
∂x1
( h v1 ) +
∂x2
( h∗v2∗ ) = 0 A similar relationship holds for relative flow. The
dimensionless circulation around a blade section
(3-136) is
irrotationality
x2 x2 m2
γ r 1
∫ γ dx = ∫V ∫ γ dm

∂v1∗ ∂v2∗ dx =
− =0 0
m r 0 0 0
r Vm
∂x2 ∂x1 x1 x1 m1
. (3-141)
1 Γ 2 − Γ1
and the dimensionless form of Eqs.(3-133) is =
N B r 0Vm0
continuity
Once the function h∗ ( x ) known, Eqs.(3-139) can
∂ ∗ ∗ ∂
∂x1
( h w1 ) +
∂x2
( h∗w2∗ ) = 0 be used either to analyse the flow around blade
sections in S1-surfaces, or to design blade
irrotationality
(3-137) sections for given γ ∗ ( x ) .

Ωr ∂ ( r / r )
0
∂w1∗ ∂w2∗ 3.4.3 Thin hydrofoil cascade design
− =2 0
∂x2 ∂x1 Wm ∂x1
The cascade model is defined as an infinite
Obviously, the continuity equation, either row of equidistant similar airfoil-shaped bodies,
for absolute of for relative flow in the and plays a central role in turbomachines design,
( x1 , x2 ) plane, can be identically satisfied by analysis and optimization. As pointed by
introducing the streamfunction ψ defined as Gostelow [21], the cascade is merely a model. In
fluid mechanics the process of modelling is one
∂ψ of physical and mathematical simplification and
h ∗v1∗ = , is intended to result in a rationalization and
∂x2
absolute flow deeper understanding of the flow field behaviour.
∗ ∗ ∂ψ As a result, the cascade flow analysis benefits
h v =−
2
∂x1 from a large body of literature dealing with both
(3-138) experimental as well as theoretical studies. The
∗ ∂ψ

hw = 1 , Timişoara School of Turbomachinery
∂x2
relative flow Hydrodynamics was particularly focused on
∗ ∗∂ψ experimental investigation of linear [2] and radial
h w =−2
∂x1 [19] cascades, as well as on developing analytical
[43] and numerical [13] methods for cascade
The governing equation for this streamfunction is flow computation. Both energetical and
Absolute flow cavitational [1] characteristics of the hydrofoil
cascades were investigated in order to provide the
∂ ⎛ 1 ∂ψ ⎞ ∂ ⎛ 1 ∂ψ ⎞ required information for improving the hydraulic
⎜ ⎟+ ⎜ ⎟=0 (3-139)
∂x1 ⎝ h ∗ ∂x1 ⎠ ∂x2 ⎝ h∗ ∂x2 ⎠ machines design. The classical methods for
cascade flow analysis have been summarized
Relative flow more than two decades ago in monographies such
VORTEX DOMINATED FLOWS 155

as [21] and [63], and experimental results have far upstream, α1 , and downstream, α 2 , the
been collected in catalogs. More recently, the spacing s , as well as the vortex strength
focus has switched from using available aero-
distribution γ ( x ) .
hydrofoil shapes and data toward numerical
design and optimization such that carefully
chosen flow specifications are met directly in the

y
design stage. The main idea is to prescribe either V∞
the flow on the hydrofoil shape or the pitchwise
averaged flow and compute the corresponding α∞
streamlined body shapes.
A particularly important requirement when V2
designing a cascade is to insure the so-called V1 c
“shock-free” inflow condition. This terminology s
f(x) α2
is a little unfortunate since it has nothing to do α1 αs
with the shock waves caused by compressibility
0 1
effects in high speed flow. It is instead frequently x
used to refer to the particular inlet angle for
which the leading edge stagnation point is located
precisely on the end of the profile camber line.
For greater or smaller inlet angles the stagnation
point will move instead onto the pressure or Figure 3-30. Thin hydrofoil cascade notations.
suction surfaces, respectively. Shock-free inlet
When employing a quasi three-dimensional
flow thus ensures the smoothest entry conditions
approach for turbomachinery design and analysis,
into the cascade and is thus likely to be close to
the axial symmetry assumption implies the use of
the minimum loss situation. However, the shock-
pitch-averaged velocity vector,
free inlet angle may not necessarily coincide
exactly with that for minimum loss, which will 1
s
V ( x) = V ( x, y ) dy .
s ∫0
usually be one or two degree greater [32]. The (3-142)
methodology introduced by Hawthorne et al. [23]
directly enforces the shock-free condition at the If the velocities upstream and downstream the
leading edge of thin hydrofoil cascades. It is this cascade are V1 and V2 , respectively, we have
methodology we have revisited in [54], with the
development of a numerical design method using Vx1 = Vx 2 = Vx and
a piecewise cubic polynomial for the thin blade 1
1 (3-143)
camberline representation. The results of this V y 2 − V y 1 = ∫ γ ( x ) dx .
approach were further used to validate and assess s0
the accuracy of our Finite Element code for
Obviously, for blades with zero thickness the
inviscid cascade hydrodynamics [55].
pitchwise mean value of the axial velocity
Hawthorne et al. [23] developed a quasi-
remains constant within the bladed region as
analytical method for designing two-dimensional
well. We can also introduce a pitchwise averaged
highly loaded cascades. The fluid is assumed to
boundary vorticity,
be incompressible and inviscid and the blades are
assumed to have zero thickness and incidence so γ ( x) dV y
that there are no stagnation points at the leading γ ( x) ≡ = , (3-144)
s dx
edges. In analysing the potential flow through the
cascade, one assumes that the thin blades may be and the pitch-averaged tangential velocity can be
represented by bound vortices of strength γ ( x ) written as
distributed along camber lines given by x

y − f ( x ) = 0, ± s, ±2 s, , where s is the blade V y ( x ) = V y 1 + ∫ γ ( t ) dt . (3-145)


0
spacing, or pitch, Figure 3-30. All lengths are
made dimensionless by the cascade axial extent. One can see from (3-145) that for cascade design
The goal of the cascade design is to find the we can alternatively provide the V y schedule
camberline shape f ( x ) given the flow direction instead of γ . An average streamline, starting at
origin, can also be found as,
156 AXISYMMETRIC SWIRLING FLOWS

γ (t ) s d ⎧ 2π 2π
x
ln ⎨cosh ( x − t ) − cos ⎡⎣ f ( x ) − f ( t )⎤⎦ ⎫⎬
f ( x ) = ∫ V y ( t ) dt . (3-146) 2 2π dx ⎩ s s ⎭
0 we can integrate (3-150). If f = 0 at the leading
The above average streamline, f ( x ) , does not edge x = 0 , i.e. the camberline starts at the
origin, we obtain
correspond to any actual streamline in a finite
pitch cascade. However, it might be seen as the x ⎛ V y1 V y 2 ⎞ s γ ( t )
1
f ( x) =
2 ⎝ Vx Vx ⎠ 4π ∫0 Vx
representative streamline in the limit s → 0 , or ⎜ + ⎟ + ×
for infinitely dense csacade.
In conventional cascade theory, the velocity ⎧ 2π 2π ⎫
field V ( x, y ) is written as the superposition of ⎪ cosh s ( x − t ) − cos s ⎡⎣ f ( x ) − f ( t ) ⎤⎦ ⎪
ln ⎨ ⎬ dt
2π 2π
the vector mean of the upstream and downstream ⎪ cosh t − cos f (t ) ⎪
⎩ s s ⎭
flow, V∞ ≡ ( V1 + V2 ) / 2 , and the velocity
ˆ ( x, y ) induced by the vortices,
V (3-151)
The determination of the blade profile f ( x )
V y1 + V y 2
V ( x, y ) = e xVx + e y + requires the integration of equations (3-150) and
2 (3-147) (3-151), each of which has a singularity at t = x .
ˆ ( x, y ) ,
+V As t → x , we have f ( x ) − f ( t ) → ( x − t ) f ' ( x )
where e x and e y are unit vectors in x - and y - and we can separate the integrand singularities.
With Vx = 1 , we can rewrite (3-150) as,
direction, respectively. The vorticity induced
velocity can be written using the Biot-Savart df tan α1 + tan α 2
approach, ( x) = +
dx 2
, (3-152)
Vˆx ( x, y ) − iVˆy ( x, y ) = 1
sγ ( x ) x
2π 2π
+ ∫ F1 ( x, t ) dt + ln
sinh ( x − t ) − i sin ⎡⎣ y − f ( t )⎤⎦ 2π 1− x
i
1
s s (3-148) 0
()
2 ∫0
− γ t dt
2π 2π
cosh ( x − t ) − cos ⎡⎣ y − f ( t )⎤⎦ γ (t )
s s F1 ( x, t ) = ×
2
On the blade, the normal velocity must 2π 2π
vanish. The normal direction on the blade is sinh ( x − t ) + f ' ( x ) sin ⎡⎣ f ( x ) − f ( t )⎤⎦
given by the vector s s −
2π 2π
∇ ( y − f ( x ) ) = −e x f ' ( x ) + e y . Using (3-147) cosh ( x − t ) − cos ⎡⎣ f ( x ) − f ( t )⎤⎦
s s
and (3-148), the condition of flow tangency on sγ ( x ) 1
the blade can be written −
2π x − t
V y1 + V y 2 Now the integrand F1 vanishes as t → x and the
− f ' ( x )V x + −
2 , (3-149) integral can be performed with standard
− f ' ( x )V x ( x , f ( x ) ) + V y ( x , f ( x ) ) = 0
ˆ ˆ quadrature rules. The last term in the right-hand
side of (3-152) is the Cauchy principal value of
leading to the following integral equation for the the singularity integral. The same approach can
camberline shape, be used to remove the singularity in (3-151),

df 1 ⎛ V y1 V y 2 ⎞ 1
( x) = ⎜ + ⎟+ × tan α1 + tan α2
1
dx 2 ⎝ Vx Vx ⎠ Vx f ( x) = x + ∫ F2 ( x, t ) dt +
2π 2π 2
γ ( t ) sinh s ( x − t ) + f ' ( x ) sin s ⎡⎣ f ( x ) − f ( t )⎤⎦
0
1

∫0 2 dt ⎡ ⎛ 2π ⎞ 1 ⎛ 1 + f '2 ( x ) ⎞ ⎤ , (3-153)
cosh
2π 2π
( x − t ) − cos ⎣⎡ f ( x ) − f ( t )⎦⎤ sγ ( x ) ⎢ln ⎜ ⎟ + ln ⎜ ⎟ +⎥
s s + ⎢ ⎝ s ⎠ 2 ⎝ 2 ⎠ ⎥
2π ⎢ ⎥
(3-150) ⎣ + x ln x + (1 − x ) ln (1 − x ) − 1 ⎦
By writing the integrand in (3-150) as
VORTEX DOMINATED FLOWS 157

sγ ( t )
F2 ( x, t ) = × f h ( t ) = N1 ( t ) f1 + N 3 ( t ) f 2

+ N 2 ( t ) f '1 + N 4 ( t ) f '2 ,
⎧ 2π 2π ⎫
⎪ cosh s ( x − t ) − cos s ⎡⎣ f ( x ) − f ( t )⎤⎦ ⎪ ( x2 − t )
2
⎡⎣ h + 2 ( t − x1 ) ⎤⎦
ln ⎨
2π 2π ⎬− N1 ( t ) = ,
⎪ cosh t − cos f (t ) ⎪ h3
⎩ s s ⎭
( t − x1 )( t − x2 )
2

sγ ( x ) ⎧⎪ ⎡ 2π ⎤ 1 + f ' ( x ) ⎫⎪
2 2 N2 (t ) = , (3-154)
− ln ⎨ ⎢ ( x − t )⎥ ⎬ h2
4π ⎪⎩ ⎣ s ⎦ 2 ⎪⎭ ( t − x1 ) ⎡⎣ h + 2 ( x2 − t ) ⎤⎦
2

In order to solve the equations (3-152) and N3 (t ) = ,


h3
(3-153), we assume a first guess for the unknown
( t − x1 ) ( t − x2 ) ,
2
function f ( x ) and its derivative f ' ( x ) , and N4 (t ) =
evaluate the right-hand sides. A suitable first h2
guess is provided by the average streamline where x1 ≤ t ≤ x2 , and h = x2 − x1 .
(3-146). The new values for f ( x ) and f ' ( x ) This piecewise cubic approximation for f ( t ) is
are further used to evaluate again the right-hand
used to evaluate F1 ( x, t ) and F2 ( x, t ) ,
sides in (3-152) and (3-153), and the iterative
process continues until the solution change from respectively, at quadrature points. There is no
one iteration to the next one is smaller than a need for an approximation of f ' on the sub-
given threshold. intervals. Regular quadrature rules can be
Since the integrals of F1 and F2 cannot be employed, since neither F1 ( x, t ) nor F2 ( x, t )
evaluated analytically, we consider a discrete have singularities as t → x .
representation of the unknown function. First, a Once the camberline shape f ( x)
set of points 0 = x1 < x2 < … < xn +1 = 1 are computed, one can compute the velocity field. It
considered on the unit interval. Our goal is to is convenient to re-write the velocity vector, V ,
find the function values as the sum of the pitch averaged velocity V and
f1 ≡ f ( x1 ) , f 2 ≡ f ( x2 ) ,…, f n +1 ≡ f ( xn +1 ) , as
an y-periodic velocity V ,
well as the function derivatives
f '1 ≡ f ' ( x1 ) , f '2 ≡ f ' ( x2 ) ,…, f 'n +1 ≡ f ' ( xn +1 ) u V ( x, y ) = V ( x ) + V ( x, y ) , where
sing the above iterative method. The function Vx ( x, y ) = Vˆx ( x, y ) and
approximation f h is restricted on each sub- (3-155)
interval [ xi , xi +1 ] to a cubic polynomial, thus V y ( x, y ) =
V y1 + V y 2
− V y ( x ) + Vˆy ( x, y )
allowing a continuously differentiable 2
approximation f ∈ C . For notational simplicity,
h 1
Using the identity
we write x1 and x2 in place of xi and xi +1 , V y1 + V y 2 1
γ (t )
− Vy ( x ) = − ∫ sgn ( x − t ) dt
respectively. On a subinterval f h may be written 2 0
2
as [26] as well as (3-148), we have the y-periodic
velocity components
Vx ( x , y ) =

1
γ (t ) [ y − f (t ) ]
sin
−∫ s dt
2 cosh 2π x − t − cos 2π y − f (t )
0 ( ) [ ]
s s
158 AXISYMMETRIC SWIRLING FLOWS

V y ( x, y ) = 1 sγ ( x)
Vblup ( x ) = Vbl ( x ) −
⎡ 2π ⎤ 2 1 + f '2 ( x )
⎢ sinh (x − t) ⎥
1
γ (t ) ⎢ s
⎥ The pressure follows from Bernoulli’s theorem.
∫0 2 ⎢ cosh 2π ( x − t ) − cos 2π [ y − f (t )] ⎥ dt The pressure coefficient is defined with respect to
⎢ s s ⎥
⎢⎣ − sgn ( x − t ) ⎥⎦ the upstream conditions,
2
p − p1 ⎛V ⎞
(3-156) cp = = 1 − ⎜ ⎟ cos2 α1 . (3-160)
1 ⎝ Vx ⎠
For y → f ( x ) we obtain the vector average of ρV12
2
the velocity on the upper and lower sides of the
Let us consider an example for the cascade
thin hydrofoil, Vbl ( x ) ≡ V ( x, f ( x ) ) . The design method presented above. The key
magnitude of the average velocity on the blade ingredient is the vortex strenght distribution
can be obtained by taking the scalar product with function, which is taken here of the form:
the tangent unit vector,
15
γ ( x ) = ( tan α 2 − tan α1 ) x (1 − x ) . (3-161)
V lo ( x ) + Vblup ( x ) 4
Vbl ( x ) = bl =
2 This function is in agreement with usual
V + f ' ( x )V y ( x ) 1 considerations concerning hydrofoils currently
Vbl ( x ) ⋅ τ ( x ) = x + ×
1+ f ' ( x)2
1 + f '2 ( x ) used in hydraulic machinery blading. One can
2π 2π
immediately see that x (1 − x ) vanishes at both

⎢ f ' ( x ) sinh ( x − t ) − sin ⎡⎣ f ( x ) − f ( t )⎤⎦ ⎥⎤ leading edge and trailing edge, thus providing a
s s
1
γ (t ) ⎢ ⎥
∫0 2 ⎢ cosh 2π ( x − t ) − cos 2π ⎣⎡ f ( x ) − f ( t )⎦⎤ ⎥ dt shock-free flow and complying with the Kutta-
Joukowsky condition, respectively. Of course,
⎢ s s ⎥
⎢ − f ' ( x ) sgn ( x − t ) ⎥ this is not the only choice available, and in
⎣ ⎦
practice one should use a parametric vortex
(3-157) strength distribution to provide the required
Note that the integrand vanishes as t → x , thus flexibility for optimization studies.
The pitch-averaged tangential velocity can
we have a regular integral.
be computed from (3-145),
In addition to the average velocity (3-157),
we can write the velocity jump using the Vy ( x )
elementary circulation as = tan α1 +
Vx
γ ( x ) dx = (Vbl − Vbl ) dl , where the circulation is
lo up
(3-162)
x x
positive counterclockwise. As a result, the + ( tan α 2 − tan α1 ) ( −3x + 5)
velocity jump from the lower to the upper side of 2
the thin foil is The average streamline that originates at
the origin is, according to (3-146),
dx
Vbllo ( x ) − Vblup ( x ) = γ ( x )
dl f ( x ) = x tan α1 +
γ ( x) (3-158)
⎛ 3 ⎞ (3-163)
= + ( tan α 2 − tan α1 ) x 2 x ⎜ − x + 1⎟
1+ f ' ( x)
2
⎝ 7 ⎠

Equations (3-157) and (3-158) give now the Figure 3-31 shows the vortex strength (3-161),
velocity magnitude on the foil lower and upper tangential velocity (3-162), and an average
sides as, streamline (3-163), dashed line in Figure 3-31c,
for a turbine cascade with inlet/outlet angles
1 sγ ( x) α1 = 35 and α 2 = 55 .
Vbllo ( x ) = Vbl ( x ) + (3-159)
2 1 + f '2 ( x )
VORTEX DOMINATED FLOWS 159

1.2 60

downstream flow angle


1
50

camberline slope [dgr]


0.8
vortex strenght

0.6 40
upstream flow angle
0.4
30
0.2

0 20
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x x

a) Pitchwise averaged vortex strength γ ( x ) d) Camberline slope arctan ( f ' ( x ) ) .


1.5
Figure 3-31. Data for the design of a turbine
cascade with α1 = 35 and α 2 = 55 , and the
computed camberline for cascade pitch s = 1 .
dimensionless tangential velocity

The primary design data is γ ( x ) , but


1

alternatively one could specify the tangential


velocity schedule V y ( x ) . The mean streamline
f ( x ) , shown with dashed line in Figure 3-31c,
0.5

does not depend on the cascade pitch, but the


camberline shape, f ( x ) , does. As a result, in
0 addition to the above flow schedule, we need to
0 0.2 0.4 0.6 0.8 1
x specify the cascade pitch, s . Since the thin blade
shape (camberline) is not know yet, the blade
b) Pitchwise averaged tangential velocity V y ( x ) spacing is made dimensionless by the cascade
axial extent since we do not know apriori the
V2 blade chordlenght. Let us consider a numerical
1.2 example for a cascade with s = 1 . The computed
camberline is shown Figure 3-31c with solid line.
1 Although the difference between the camberline
camberline and average streamline

mean streamline
camberline
and the average streamline might seem rather
0.8 small from Figure 3-31c, one can see that the
difference in slope is quite large, particularly near
0.6 the leading edge. Figure 3-31d shows that the
camberline direction at leading edge should be
0.4
21.7 , quite different from the far upstream flow
0.2 angle α1 = 35 , in order to obtain zero incidence
and shock-free flow. The angle of attack, i.e. the
V1 0
0 0.2 0.4 0.6 0.8 1 angle between the chordline and the
x upstream/downstream vector average velocity
V∞ ≡ ( V1 + V2 ) / 2 is α s − α ∞ = 48.59 − 46.78
c) Mean streamline f ( x ) and camberline f ( x ) .
= 1.81 . Near the trailing edge the camberline
angle is sligthly larger than the downstream flow
angle α 2 = 55 . An interesting feature is the
160 AXISYMMETRIC SWIRLING FLOWS

reverse curvature near the trailing edge, as noted Flow, Turbulence and Combustion, 69, pp.
by Hawthorne et al. [23] as well. 63-78.
[13] Carte, I.N., 1986, “Contributions to the
3.5 References Study of Radial-Axial Cascades and
Applications to Francis Runner Design”, (in
[1] Anton I., 1985, Cavitaţia, Romanian Romanian), PhD Thesis, Polytechnic Institute
Academy Press, Vol. II, Ch. 7. “Traian Vuia”, Timişoara.
[2] Anton, Viorica, 1972, “Experimental [14] Cary, A.W., 1997, “The Onset of Non-
Investigations on the Influence of Hydrofoil Axisymmetric Vortex Breakdown”, PhD
Geometry on Energetical and Cavitational Thesis, University of Michigan.
Characteristics of Hydrofoil Cascades” (in [15] Ciocan G.D., and Iliescu M.S., 2007,
Romanian), PhD Thesis, Polytechnic Institute “Vortex Rope Investigation by 3D-PIV
“Traian Vuia”, Timişoara. Method”, 2nd IAHR International Meeting of
[3] Armfield, S.W., and Fletcher, C.A.J., the Workgroup on Cavitation and Dynamic
“Numerical Simulation of Swirling Flow in Problems in Hydraulic Machinery and
Diffusers”, International Journal for Systems, Timisoara, Romania, Scientific
Numerical Methods in Fluids, 6, pp.541-556. Bulletin of the Politehnica University of
[4] Armfield, S.W., and Fletcher, C.A.J., Timisoara. Transactions on Mechanics, Tom
1989, “Comparison of k-ε and Algebraic 52(66), Fasc. 6, pp. 159-172.
Reynolds Stress Models for Swirling [16] Ciocan G.D., Iliescu M.S., Vu T.C.,
Diffuser Flow”, International Journal for Nennemann B., and Avellan F., 2007,
Numerical Methods in Fluids, 9, pp. 987- “Experimental Study and Numerical
1009. Simulation of the FLINDT Draft Tube
[5] Armfield, S.W., Cho, N.-H., Clive, A., Rotating Vortex”, Journal of Fluids
and Fletcher, J., 1990, “Prediction of Engineering, Vol. 129, pp. 146-158.
Turbulence Quantities for Swirling Flow in [17] Clausen, P.D., Koh, S.G., and Wood,
Conical Diffusers”, AIAA Journal, 28(3), pp. D.H., 1993, “Measurements of a Swirling
453-460. Turbulent Boundary Layer Developing in a
[6] Avellan, F., 2000, “Flow Investigation Conical Diffuser”, Experimental Thermal
in a Francis Draft Tube: the FLINDT and Fluid Science, 6, pp. 39-48.
Project”, in Proceedings of the 20th IAHR [18] Darmofal, D.L., 1996, “Comparisons of
Symposium on Hydraulic Machinery and Experimental and Numerical Results for
Systems, Paper DES-11. Axisymmetric Vortex Breakdown in Pipes”,
[7] Batchelor, G.K., 1964, “Axial Flow in Computers & Fluids, 25(4), pp.353-371.
Trailing Line Vortices”, J. Fluid Mech., [19] Gheorghiu, Monica, 1976, “Theoretical
20(4), pp. 645-658. and Experimental Study of Energetical
[8] Batchelor, G.K., 1967, An Introduction Characterisitics of Radial Cascades for
to Fluid Dynamics, Cambridge University Turbine Guide Vanes”, (in Romanian), PhD
Press. Thesis, Polytechnic Institute “Traian Vuia”,
[9] Benjamin, T.B., 1962, “Theory of Timişoara.
Vortex Breakdown Phenomenon”, Journal of [20] Goldshtik, M., and Hussain, F., 1998,
Fluid Mechanics, 14, pp. 593-629. “Analysis of Inviscid Vortex Breakdown in a
[10] Borges, J.E., 1993, “A Proposed Semi-infinite Pipe”, Fluid Dynamics
Through-Flow Inverse Method for the Research, 23, pp. 189-234.
Design of Mixed-Flow Pumps”, International [21] Gostelow, J.P., 1983, Cascade
Journal for Numerical Methods in Fluids, 17, Aerodynamics, Oxford, New York, Toronto,
pp. 1097-1114. Pergamon Press.
[11] Bosman, C., and El-Shaarawi, M.A.I., [22] Hawthorne, W.R., Wang, C., Tan, C.S.,
1977, “Quasi-three-dimensional numerical and McCune J.E., 1984, “Theory of Blade
solution of flow in turbomachines”, J. Fluids Design for Large Deflections: Part I – Two-
Engineering, March 1977, pp. 133-140. Dimensional Cascade”, Journal of
[12] Brücker, Ch., 2002, “Some Engineering for Gas Turbines and Power,
Observations of Vortex Breakdown in a 106, pp.346-353.
Confined Flow with Solid Body Rotation”, [23] Hawthorne, W.R., Wang, C., Tan, C.S.,
and McCune J.E., 1984, “Theory of Blade
VORTEX DOMINATED FLOWS 161

Design for Large Deflections: Part I – Two- Turbulent Swirling Confined Flows”,
Dimensional Cascade”, Journal of International Journal for Numerical Methods
Engineering for Gas Turbines and Power, in Fluids, 41, pp. 133-150.
106, pp.346-353. [35] Marshall, J.S., 2001, Inviscid
[24] Herrada, M.A., and Barrero A., 2004, Incompressible Flow, John Wiley & Sons,
“Nonparallel linear stability analysis of Canada.
unconfined vortices”, Physics of Fluids, [36] Mauri, S., Kueny, J.-L., and Avellan,
16(10), pp. 3755-3764. F., 2004, “Werlé-Legendre Separation in a
[25] Huerre, P., and Monkewitz, P.A., 1990, Hydraulic Machine Draft Tube”, Journal of
“Local and Global Instabilities in Spatially Fluids Engineering, 126, pp. 976-980.
Developing Flows”, Annu. Rev. Fluid Mech., [37] McFarland, E.R., 1984, “A Rapid
22, pp. 473-537. Blade-to-Blade Solution for Use in
[26] Hughes, J.R., 2000, The Finite Element Turbomachinery Design”, Journal of
Method. Linear static and dynamic Finite Engineering for Gas Turbines and Power,
Element Analysis, Dover Publications, 106, pp. 376-382.
Reprint of The Finite Element Method, [38] Miller, P.L., 2000, “Blade Geometry
originally published by Prentice-Hall in Description using B-splines and General
1987. Surfaces of Revolution”, PhD Thesis, Iowa
[27] Jennions, I.K., and Stow, P., 1985, “A State University.
quasi-three-dimensional blade design system: [39] Nishi M., Matsunaga S., Okamoto S.,
Part I: Throughflow analysis”, Journal of Uno M., and Nishitani K., 1988,
Engineering for Gas Turbines and Power, “Measurement of three-dimensional periodic
107, pp. 301-307. flow in a conical draft tube at surging
[28] Keller, J.J., Egli, J., and Exley, J., 1985, condition”, in Rothagi U.S., et al., (eds),
“Force- and Loss-Free Transitions Between Flow in Non-Rotating Turbomachinery
Flow States”, Journal of Applied Components, FED, Vol. 69, pp.81-88.
Mathematics and Physics (ZAMP), 36, pp. [40] Parras, L., and Fernandez-Feria, R.,
854-889. 2007, “Spatial stability and the onset of
[29] Keller, J.J., Egli, W., and Althaus, R., absolute instability of Batchelor’s vortex for
1988, “Vortex Breakdown as a Fundamental high swirl numbers”, J. Fluid Mechanics,
Element of Vortex Dynamics”, Journal of 583, pp.27-43.
Applied Mathematics and Physics (ZAMP), [41] Peng, G., 2005, “A Practical Combined
39, pp. 404-440. Computation Method of Mean Through-Flow
[30] Khorrami, M.R., Malik, M.R., and Ash, for 3D Inverse Design of Hydraulic
R.L., 1989, “Application of Spectral Turbomachinery Blades”, J. Fluids
Collocation Techniques to the Stability of Engineering, 127, pp. 1183-1190.
Swirling Flows”, Journal of Computational [42] Peng, G., Cao, S., Ishizuka, M., and
Physics, 81, pp. 206-229. Hayama, S., 2002, “Design optimization of
[31] Kirschner O., and Ruprecht A., 2007, axial flow hydraulic turbine runner. Part I –
“Vortex Rope Measurement in a Simplified an improved Q3D inverse method”, Int. J.
Draft Tube”, 2nd IAHR International Numer. Meth. Fluids, 39, pp. 517-531.
Meeting of the Workgroup on Cavitation and [43] Popa, O., 1980, “Potential Flows and
Dynamic Problems in Hydraulic Machinery Cascade Hydrodynamics”, (in Romanian)
and Systems, Timisoara, Romania, Scientific Polytechnic Institute “Traian Vuia”,
Bulletin of the Politehnica University of Timişoara.
Timisoara. Transactions on Mechanics, Tom [44] Press, W.H., Teukolsky, S.A.,
52(66), Fasc. 6, pp. 173-184. Vetterling, W.T., and Flannery, B.P., 1992,
[32] Lewis, R.I., 1996, Turbomachinery Numerical Recipes, Cambridge University
Performance Analysis, Ch. 2. Press.
[33] Liu, J.-G., and Shu, C.-W., 2000, “A [45] Riley, K.F., Hobson, M.P., and Bence,
High-Order Discontinuous Galerkin Method S.J., 1997, Mathematical Methods for
for 2D Incompressible Flows”, Journal of Physics and Engineering, Cambridge
Computational Physics, 160, pp. 577-596. University Press.
[34] Lu, P., and Semião, V., 2003, “A New [46] Rusak, Z., Wang., S., and Whiting,
Second-Moment Closure Approach for C.H., 1998, “The Evolution of a Perturbed
162 AXISYMMETRIC SWIRLING FLOWS

Vortex in a Pipe to Axisymmetric Vortex [56] Szeri, A., and Holmes, P., 1988,
Breakdown”, Journal of Fluid Mechanics, “Nonlinear Stability of Axisymmetric
366, pp. 211-237. Swirling Flows”, Phil. Trans. R. Soc. Lond.,
[47] Saiac, J.-H., 1989, “Finite Element A 326, pp. 327-354.
Solutions of Axisymmetric Euler Equations [57] Tan, C.S., Hawthorne, W.R., McCune,
for an Incompressible and Inviscid Fluid”, J.E., and Wang, C., 1984, “Theory of Blade
INRIA Rapports de Recherche No. 999. Design for Large Deflections: Part II –
[48] Shih, T.-H., Liou, W.W., Shabbir, Z., Annular Cascades”, Journal of Engineering
Yang Z., and Zhu, J., 1995, “A New k-ε for Gas Turbines and Power, 106, pp. 354-
Eddy-Viscosity Model for High Reynolds 365.
Number Turbulent Flows – Model [58] Wang, S., and Rusak, Z., 1997, “The
Development and Validation”, Computers & Dynamics of a Swirling Flow in a Pipe and
Fluids, 24(3), pp. 227-238. Transition to Axisymmetric Vortex
[49] Sick M., Stein P., Doerfler P., Breakdown”, Journal of Fluids Mechanics,
Sallaberger M., and Braune A., 2005, “CFD 340, pp. 177-223.
prediction of the part-load vortex in Francis [59] Wu, C.-H., 1952, “A General Theory of
turbines and pump-turbines”, International Three-Dimensional Flow in Subsonic and
Journal of Hydropower and Dams, Vol. 12, Supersonic Turbomachines of Axial-, Radial-
No. 85. , and Mixed-Flow Types”, NACA Technical
[50] Sloan, D.G., Smith, P.J., and Smoot, Note 2604.
L.D., 1986, “Modeling of Swirl in Turbulent [60] Xu, D., Khoo, B.C., and Leschziner,
Flow Systems”, Progress in Energy and M.A., 1998, “Numerical Simulation of
Combustion Science, 12, pp. 163-250. Turbulent Flow in an Axisymmetric Diffuser
[51] Stein P., Sick M., Doerfler P., White P., with a Curved Surface Centre-Body”,
and Braune A., 2006, “Numerical Simulation International Journal of Numerical Methods
of the Cavitating Draft Tube Vortex in a for Heat & Fluid Flow, 8(2), pp. 245-255.
Francis Turbine”, Proceedings of the 23rd [61] Yaras, M.I., and Grosvernor, A.D.,
IAHR Symposium on Hydraulic Machinery 2003, “Evaluation of One- and Two-Equation
and Systems, Yokohama, Japan, paper F228, Low-Re Turbulence Models. Part I –
(on CD-ROM, ISBN 4-8190-1809-4). Axisymmetric Separating and Swirling
[52] Susan-Resiga R., Vu T.C., Muntean S., Flow”, International Journal for Numerical
Ciocan G.D., Nennemann B., 2006, “Jet Methods in Fluids, 42, pp. 1293-1319.
Control of the Draft Tube Vortex Rope in [62] Zangeneh, M., 1991, “A Compressible
Francis Turbines at Partial Discharge”, Three-Dimensional Design Method for
Proceedings of the 23rd IAHR Symposium on Radial and Mixed Flow Turbomachinery
Hydraulic Machinery and Systems, Blades”, International Journal for Numerical
Yokohama, Japan, paper F192, (on CD- Methods in Fluids, 13, pp. 599-624.
ROM, ISBN 4-8190-1809-4). [63] Zidaru, Gh. 1981, Potential Flows and
[53] Susan-Resiga, R., Ciocan, G.D., Anton, Cascade Hydrodynamics, (in Romanian), Ed.
I., Avellan, F., 2006, “Analysis of the Didactică şi Pedagogică, Bucureşti.
Swirling Flow Downstream a Francis
Turbine Runner”, Journal of Fluids
Engineering, 128, pp. 177-189.
[54] Susan-Resiga, R., Muntean, S., Bernad,
S., Frunză, T., and Balint, D., 2006, “Thin
Hydrofoil Cascade Design and Numerical
Flow Analysis. Part I: Design”, Proc.
Romanian Academy, Series A, 7(2), pp. 117-
126.
[55] Susan-Resiga, R., Muntean, S., Bernad,
S., Frunză, T., and Balint, D., 2006, “Thin
Hydrofoil Cascade Design and Numerical
Flow Analysis. Part II: Analysis”, Proc.
Romanian Academy, Series A, 7(3), pp. 199-
208.

S-ar putea să vă placă și