You are on page 1of 39

\section{INTERACTION OF ACOUSTIC-VORTICITY WAVES WITH AN ANNULAR CASCADE

IN
A SWIRLING FLOW}
In the final part of this dissertation, the theory of unsteady swirling
motion developed in the previous chapters is applied to a boundary-value
unsteady aerodynamic problem. To this end, we consider an imposed on a
swirling mean flow incident acoustic-vorticity wave (gust) evolving from
a
given upstream point of origin and interacting with an annular cascade of
blades. This interaction produces a near-field unsteady aerodynamic
response
coupled with far-field hydrodynamic and acoustic pressure fields.
For simplicity, we assume a stationary row of flat-plate non-loaded
blades
installed at zero angle of attack to the mean flow.
The blade chords are centered at $x=0$, and the incident gust plane
of origin (or measurement) coincides with the upstream computational
domain
boundary plane at $x=x_{in}.$
The boundary-value problem is governed by a set of equations
(\ref{G11})-(\ref{G12}). Without loss of generality, we consider a single
harmonic component of an upstream unsteady flow excitation, with a given
temporal
frequency $\omega .$ As usual, an arbitraty initial disturbance field at
$%
x=x_{in}$ is viewed as an inverse Fourier transform of the harmonic
solution
over the range of wavenumbers:
\begin{equation}
{\bf u}_{in}(x,r,\theta ;t)=\int_{-\infty }^\infty \sum_{n_g=-\infty
}^\infty {\bf a}(r)\exp \{i[k_g(x-U_0t)+n_g\theta ]\}dk_g \label{B1}
\end{equation}
In (\ref{B1}), $n_g$ is a circumferential wavenumber, and $k_g=\omega
/U_0$
is an axial wavenumber of an initial harmonic perturbation. The amplitude
vector ${\bf a}(r)$ is expanded in terms of Hankel functions of order one
as
indicated in formulae (\ref{I34})-(\ref{I36}) and (\ref{I75})(\ref{I77}):
\begin{equation}
a_{r,\theta }(r)=D_{r,\theta }\{H_1^{(1)}(\mu _nr)-\frac{H_1^{(1)}(\mu
_nr_h)%
}{H_1^{(2)}(\mu _nr_h)}H_1^{(2)}(\mu _nr)\}, \label{B2}
\end{equation}
where $D_{r,\theta }$ are given constants, and $\mu _n$ is the $n$-th
root
of (\ref{I35}).
Our present numerical analysis is restricted to the case of an imposed
upstream solenoidal perturbation. Thus, the axial amplitude component $%

a_x(r) $ is obtained from a divergence-free condition for ${\bf u}_{in}$:


\begin{equation}
ik_ga_x(r)+\frac 1r\frac{\partial (ra_r(r))}{\partial r}+\frac{in_g}%
ra_\theta (r)=0 \label{B3}
\end{equation}
By construction, the total solution of the boundary-value problem is
equal
to the sum of the generalized gust solution ${\bf Z}_g=\{{\bf u}_h+{\bf
u}%
_p,\phi \}_g$ obtained in Chapter 8, and a scattered part ${\bf Z}_s=\
{{\bf u%
},\phi \}_s$ which should have an outgoing-wave behavior far upstream and
downstream from the blade row. Both solutions are coupled on the blade
surface through the impermeability condition.
In view of the solution splitting, the boundary-value problem for the
potential $\phi _s$ is posed for a homogeneous convective wave equation:
\begin{equation}
\frac{D_0}{Dt}\frac 1{c_0^2}\frac{D_0}{Dt}\phi _s-\frac 1{\rho _0}\nabla
\cdot (\rho _0\nabla )\phi _s=0, \label{B4}
\end{equation}
with a set of boundary conditions imposed along the blades and their
wakes,
the upstream and downstream radiation conditions, and blade row passage
quasiperiodicity. Since the scattered solution is initiated in reponse to
a
single upstream harmonic perturbation of (\ref{B1}), it is obtained in
the
frequency domain by taking
\begin{equation}
\phi _s(x,r,\theta ;t)=\hat \phi _s(x,r,\theta )e^{-i\omega t}.
\label{B5}
\end{equation}
Due to the linear character of the problem, the total scattered solution
is
then equal to an integral of (\ref{B5}) over all frequencies involved in
(%
\ref{B1}).
The convective derivative in (\ref{B4}) reduces to
\begin{equation}
\frac{D_0}{Dt}=-i\omega +U_0\frac \partial {\partial x}+U_s\frac \partial
{\partial \theta }, \label{B6}
\end{equation}
and we restrict our numerical analysis with the mean swirl component
\begin{equation}
U_s=\Omega r+\Gamma /r. \label{B7}
\end{equation}
For further discussion, we nondimensionalize the characteristic
quantities
as follows:
- all the lengths, by the blade semichord $c/2$;

- time $t$, by $c/2U_0$;


- the temporal frequency $\omega $, by $2U_0/c$;
- the axial and radial wavenumbers $k_g$ and $\mu _n$, by $2/c$;
- the mean flow velocity components, by $U_0$;
- the vortical gust amplitude vector ${\bf a}(r)$, by $|D_\theta |$;
- the potential functions $\phi _g$ and $\phi _s$, by $c|D_\theta |/2$;
- the disturbance pressure $p^{\prime }$, by $\overline{\rho }_0U_0|
D_\theta
|$;
- the unsteady lift $L^{\prime },$ by $\pi \overline{\rho }_0cU_0|
D_\theta
|. $
The non-dimensional axial wavenumber $k_g=\omega c/2U_0$ is commonly
referred to as the {\it reduced frequency}. Other important input
parameters
of the problem include Mach numbers of the axial and swirl mean velocity
components, $M_0=U_0/\overline{c}_0,\overline{M}_\Gamma =\Gamma /r_m%
\overline{c}_0,$ and $\overline{M}_\Omega =\Omega r_m/\overline{c}_0.$
The
latter two are specified at the mean radius of the annulus, $%
r_m=(r_h+r_t)/2, $ where $r_h$ and $r_t$ are the hub and tip radii of the
annular duct. The quantities $\overline{\rho }_0$ and $\overline{c}_0$
are
the reference mean flow density and speed of sound in the case of no
swirl.
Finally, among the physical parameters of the problem, the
circumferential
wavenumber of the incident perturbation, $n_g,$ and the number of blades
in
the row, $N_b,$ must be specified.
In the current linearized approach, it is advantageous to assume the
cascade
quasiperiodicity condition
\begin{equation}
\hat \phi _s(x,r,\theta )=\hat \phi _s(x,r,\theta +2\pi /N_b)e^{i\sigma },
\label{B8}
\end{equation}
where $\sigma =2\pi n_g/N_b$ is the interblade phase angle. Thus, a
single
interblade passage may be considered in the numerical calculations.
On the whole, the computational domain comprises a single-passage annular

sector with two adjacent blades centered between some $x$-locations


upstream
and downstream. Note that in the physical space, the domain turns as it
extends downstream, to remain aligned with the swirling streamlines of
the
mean flow. Therefore, the numerical solution of the boundary-value
problem
is obtained in the transformed coordinate system ${\bf y}=(\zeta ,\xi
,\vartheta )$, where
\begin{eqnarray}
\zeta &=&x, \nonumber \\
\xi &=&r, \label{B9} \\
\vartheta &=&\theta -(\Omega +\frac \Gamma {r^2})\frac x{U_0} \nonumber
\end{eqnarray}
\subsection{ Frequency-Domain Numerical Formulation.}
The transformation (\ref{B5}) reduces the governing equation (\ref{B4})
to
the following non-dimensional form in the computational domain
(\ref{B9}):
\begin{equation}
c_1\hat \phi _{s\zeta \zeta }+c_2\hat \phi _{s\xi \xi }+c_3\hat \phi
_{s\vartheta \vartheta }+c_4\hat \phi _{s\zeta \vartheta }++c_5\hat \phi
_{s\xi \vartheta }++c_6\hat \phi _{s\zeta }++c_7\hat \phi _{s\xi }+
+c_8\hat
\phi _{s\vartheta }++c_9\hat \phi _s=0. \label{B10}
\end{equation}
The non-constant, complex, coefficients of the elliptic three-dimensional
partial differential equation (\ref{B10}) are defined as follows:
\begin{eqnarray*}
c_1 &=&\widetilde{\beta }^2, \\
c_2 &=&1, \\
c_3 &=&\widetilde{\beta }^2\frac{M_s^2}{r^2M_0^2}+\frac{\widetilde{\delta
}^2%
}{r^2}+2\frac{M_s^2}{r^2\tau }+4\frac{M_s^2x^2}{r^4M_0^2}, \\
c_4 &=&-2\frac{M_s}{rM_0}\widetilde{\beta }^2-2\frac{M_sM_0}{r\tau }, \\
c_5 &=&4\frac{M_sx}{r^2M_0}, \\
c_6 &=&2i\frac{k_gM_0^2}\tau , \\
c_7 &=&\frac 1r(1+\frac{M_s^2}\tau ), \\
c_8 &=&-4\frac{M_sx}{r^3M_0}+2\frac{M_s^3x}{r^3M_0\tau }, \\
c_9 &=&\frac{k_g^2M_0^2}\tau ,
\end{eqnarray*}
where
\begin{eqnarray*}
\widetilde{\beta }^2 &=&1-\frac{M_0^2}\tau , \\
\widetilde{\delta }^2 &=&1-\frac{M_s^2}\tau ,
\end{eqnarray*}
the total swirl Mach number is $M_s=M_\Omega +M_\Gamma ,$ and $\tau \,$
is
related to the isentropic radial variation of the mean flow density and
speed of sound in a swirling flow:
\begin{equation}

\tau =\frac{c_0^2(r)}{\overline{c}_0^2}=[\frac{\rho _0^2(r)}


{\overline{\rho }%
_0^2}]^{\kappa -1}=1-\frac{\kappa -1}2[M_\Gamma ^2-M_\Omega ^2-4M_\Omega
M_\Gamma \ln (\frac r{r_m})], \label{B11}
\end{equation}
where $\kappa $ is the specific heats ratio, $\kappa =c_p/c_v.$
For the present numerical computations, the equation (\ref{B9}) is
discretized using the central finite-difference scheme. This results, in
general, in 15 non-zero entries per row in a sparse, algebraic, complex,
linear system of equations.
In the numerical analysis, the computational domain extends in the axial
direction for $-T_x\leq \zeta \leq T_x$ (we also take $T_x=|x_{in}|$), in
the radial direction, for $r_h\leq \xi \leq r_t$, and in the
circumferential
direction, for $0\leq \vartheta \leq 2\pi /N_b.$ In the following
subsections, the numerical conditions imposed along the boundaries of
this
domain are considered in detail.
\subsubsection{Conditions along the blade surfaces.}
The position of blades surfaces in the computational domain is specified
for
${\cal S}=$\{$-1\leq \zeta \leq 1;$ $r_h<\xi <r_t;$ $\vartheta =0,2\pi
/N_b\}.$ The imposed flow impermeability condition couples incident and
scattered perturbations in order to provide a zero total upwash velocity
component at the surfaces of the blades:
\begin{equation}
\nabla \widehat{\phi }_s\cdot {\bf n}=-(\widehat{{\bf u}}+\nabla
\widehat{%
\phi }_g)\cdot {\bf n},\hspace{1.0in}{\bf x\in }{\cal S} \label{B12}
\end{equation}
where a Fourier component of the generalized gust solution is introduced
as
\begin{equation}
{\bf Z}_g=\widehat{{\bf Z}}_g\exp (-i\omega t), \label{B13}
\end{equation}
and ${\bf n}$ is the normal vector of a blade surface. In the physical
domain, the equation of the surface is defined by
\begin{equation}
g(x,r,\theta )=\theta -\frac{M_s}{rM_0}x, \label{B14}
\end{equation}
and thus the normal vector can be represented as
\begin{equation}
{\bf n=}\frac{\nabla g}{|\nabla g|}=\cos ({\bf n,}x)\widehat{{\bf
e}}_x+\cos
({\bf n,}r)\widehat{{\bf e}}_r+\cos ({\bf n,}\theta )\widehat{{\bf e}}%
_\theta , \label{B15}
\end{equation}
where the direction cosines are
\begin{eqnarray}
\cos ({\bf n,}x) &=&-\frac{M_s}{rM_0|\nabla g|}, \nonumber \\

\cos ({\bf n,}r) &=&\frac{2M_\Gamma }{r^2M_0|\nabla g|}x, \label{B16} \\


\cos ({\bf n,}\theta ) &=&\frac 1{r|\nabla g|}, \nonumber
\end{eqnarray}
and
\begin{equation}
|\nabla g|=\sqrt{\frac 1{r^2}+(\frac{M_s}{rM_0})^2+4(\frac{xM_\Gamma }
{r^2M_0%
})^2}. \label{B17}
\end{equation}
In the computational domain, the condition (\ref{B12}) reduces, after the
transformation of derivatives, to the form
\begin{eqnarray}
\frac{\partial \widehat{\phi }_s}{\partial \zeta }\cos ({\bf n,}x)
+\frac{%
\partial \widehat{\phi }_s}{\partial \xi }\cos ({\bf n,}r)+\frac{\partial
\widehat{\phi }_s}{\partial \vartheta }[\frac{\cos ({\bf
n,}\theta )}r+\frac{%
2xM_\Gamma }{r^2M_0}\cos ({\bf n,}r)-\frac{M_s}{rM_0}\cos ({\bf n,}x)]
&=&
\nonumber \\
\ -[(\widehat{u}+\frac{\partial \widehat{\phi }_g}{\partial x})\cos ({\bf
n,}%
x)+(\widehat{v}+\frac{\partial \widehat{\phi }_g}{\partial r})\cos ({\bf
n,}%
r)+(\widehat{w}+\frac 1r\frac{\partial \widehat{\phi }_g}{\partial \theta
}%
)\cos ({\bf n,}\theta )],\hspace{0.5in}{\bf y} &\in &{\cal S},
\label{B18}
\end{eqnarray}
where the gust quantities on the right hand side (the gust vortical
velocity
and potential derivatives) are calculated beforehand by solving
iteratively
a mixed initial/boundary value problem as indicated in Chapter 8.
For the numerical implementation of the condition (\ref{B18}), a
fourth-order one-sided differencing is applied at the boundary points,
which
adds 13 non-zero elements per row to the resulting matrix.
\subsubsection{Conditions along the wake surfaces.}
The wake conditions are imposed on two surfaces downstream from the
blades
trailing edges, ${\cal W}=\{1<\zeta <T_x;$ $r_h<\xi <r_t;$ $\vartheta
=0,2\pi /N_b\}$. Hence, the wakes coincide with the mean flow
streamlines.
The conditions are derived from the general principles of unsteady flow
motion along the wakes of streamlined bodies. Note that although the
unsteady potential $\phi _s$ is not continuous across the wake, the
corresponding unsteady pressure and normal component of unsteady velocity
are continuous, which can be expressed as follows:

\begin{equation}
\Delta p_s^{\prime }=\frac{D_0}{Dt}\Delta \phi _s=0 \hspace{0.5in} and
\hspace{0.5in} \Delta (\nabla \phi _s\cdot {\bf n})=0,\hspace{0.5in}{\bf
x\in
}{\cal W}, \label{B19}
\end{equation}
where the wake surface normal vector ${\bf n}$ is the same as in
(\ref{B15}%
). Also note that the wake conditions do not couple scattered and
incident
unsteady fields since the gust solution ${\bf Z}_g$ also satisfies the
same
conditions (\ref{B19}).
The first equation in (\ref{B19}) can be rewritten as
\begin{equation}
-ik_g\Delta \widehat{\phi }_s+\frac{\partial \Delta \widehat{\phi }_s}{%
\partial \zeta }=0,\hspace{0.5in}{\bf y\in }{\cal W} \label{B20}
\end{equation}
and integrated for $\zeta >1$ to give
\begin{equation}
\Delta \widehat{\phi }_s=\Delta \widehat{\phi }_{sTE}\exp \{ik_g(\zeta
-1)\},%
\hspace{0.5in}{\bf y\in }{\cal W}, \label{B21}
\end{equation}
where $\Delta \widehat{\phi }_{sTE}$ stands for the potential jump at the
blades trailing edges. Note that the Kutta condition is satisfied
automatically in this formulation.
To apply the jump condition numerically, we recall the cascade
quasiperiodicity relation (\ref{B8}) which allows to calculate the
potential
jump across the wake as
\begin{equation}
\Delta \widehat{\phi }_s(\zeta ,\xi ,\vartheta )=\widehat{\phi }_s(\zeta
,\xi ,\vartheta )-\widehat{\phi }_s(\zeta ,\xi ,\vartheta +2\pi
/N_b)e^{-i\sigma } \label{B22}
\end{equation}
In the numerical analysis, equation (\ref{B21}) is applied along the
surface
$\vartheta =2\pi /N_b$ whereas the second relation in (\ref{B19}) is
implemented for the grid points at $\vartheta =0.$ Since the
quasiperiodicity condition couples the normal velocity components on both
sides of the blade passage domain,
\begin{equation}
\frac{\partial \widehat{\phi }_s(\zeta ,\xi ,\vartheta )}{\partial
n}=\frac{%
\partial \widehat{\phi }_s(\zeta ,\xi ,\vartheta +2\pi /N_b)}{\partial n}
%
e^{-i\sigma }, \label{B23}
\end{equation}
the continuity of{\it \ }$\partial \widehat{\phi }_s/\partial n$ across
the

wake is provided by taking


\begin{eqnarray}
\frac{\partial \widehat{\phi }_s(\zeta ,\xi ,\vartheta )}{\partial
\zeta }%
\cos ({\bf n,}x)+\frac{\partial \widehat{\phi }_s(\zeta ,\xi
,\vartheta )}{%
\partial \xi }\cos ({\bf n,}r)+\frac{\partial \widehat{\phi }_s(\zeta
,\xi
,\vartheta )}{\partial \vartheta }[\frac{\cos ({\bf n,}\theta )}r+\frac{%
2xM_\Gamma }{r^2M_0}\cos ({\bf n,}r)-\frac{M_s}{rM_0}\cos ({\bf n,}x)]
= \\ \nonumber
-\{\frac{\partial \widehat{\phi }_s(\zeta ,\xi ,\vartheta +2\pi /N_b)}{%
\partial \zeta }\cos ({\bf n,}x)+\frac{\partial \widehat{\phi }
_s(\zeta ,\xi
,\vartheta +2\pi /N_b)}{\partial \xi }\cos ({\bf n,}r)+ \hspace{1.5in}
\label{B24} \\
\frac{\partial
\widehat{\phi }_s(\zeta ,\xi ,\vartheta +2\pi /N_b)}{\partial \vartheta }
[%
\frac{\cos ({\bf n,}\theta )}r+\frac{2xM_\Gamma }{r^2M_0}\cos ({\bf
n,}r)-%
\frac{M_s}{rM_0}\cos ({\bf n,}x)]\}e^{-i\sigma }\ ,\hspace{0.5in}{\bf y}
\in {\cal W}. \hspace{1.5in} \nonumber
\end{eqnarray}
The minus sign on the right hand side of (\ref{B24}) accounts for the
fact
that the normal velocity at $(\zeta ,\xi ,\vartheta +2\pi /N_b)$ points
toward the wake surface.
A fourth-order one-sided finite-difference approximation of the condition
(%
\ref{B24}) requires 26 non-zero elements per row while the condition
(\ref
{B21}) results in only 4 non-zero entries per row.
\subsubsection{Conditions from upstream boundary to the blades.}
These conditions are imposed on both sides of the turning annular sector
in
the physical domain, between the upstream boundary and the blade
surfaces.
Therefore, in the computational space we consider the region ${\cal Q}%
=\{-T_x<\zeta <-1;$ $r_h<\xi <r_t;$ $\vartheta =0,2\pi /N_b\}$.
For $\vartheta =2\pi /N_b,$ the cascade quasiperiodicity condition is
imposed on the potential function:
\begin{equation}
\widehat{\phi }_s(\zeta ,\xi ,\vartheta +2\pi /N_b)=\widehat{\phi }
_s(\zeta
,\xi ,\vartheta )e^{i\sigma },\hspace{0.5in}{\bf y\in }{\cal Q}
\label{B25}
\end{equation}
which thus adds only 2 non-zero elements per row in the matrix.

To avoid redundancy in the resulting algebraic system of equations, a


condition different from (\ref{B25}) needs to be specified for $\vartheta
=0$%
. For this reason, we apply the condition (\ref{B24}) of continuity of
the
normal component of unsteady velocity.
\subsubsection{Conditions along the duct walls surfaces.}
The computational region for the duct wall conditions is specified along
the
surfaces ${\cal D}=\{-T_x<\zeta <T_x;$ $0\leq \vartheta \leq 2\pi
/N_b,\xi
=r_h,r_t\}.$ From Chapter 8, the radial component of the gust velocity
vanishes at the duct walls,
\begin{equation}
\widehat{v}+\partial \widehat{\phi }_g/\partial r=0,\hspace{1.0in}{\bf
x}\in
{\cal D}. \label{B26}
\end{equation}
Thus, for the numerical solution of the gust scattering boundary-value
problem we impose
\begin{equation}
\partial \widehat{\phi }_s/\partial r=0,\hspace{1.0in}{\bf x}\in {\cal
D},
\label{B27}
\end{equation}
which transforms to
\begin{equation}
\frac{\partial \widehat{\phi }_s}{\partial \xi }+\frac{2xM_\Gamma }
{r^2M_0}%
\frac{\partial \widehat{\phi }_s}{\partial \vartheta }=0,\hspace{1.0in}
{\bf y%
}\in {\cal D} \label{B28}
\end{equation}
in the computational domain. The fourth-order finite-difference
approximation of (\ref{b28}) fills out 9 non-zero entries per row in the
linear algrbraic matrix.
\subsubsection{Inflow/Outflow conditions.}
Far away upstream and downstream from the blade row, an asymptotic
behavior
of the scattered unsteady disturbances can be obtained from the normal
mode
analysis of the governing equations (\ref{G11})-(\ref{G12}). The {\it
exact}
inflow/outflow conditions then assume a normal mode representation of the
outgoing unsteady pressure waves, and set to zero all the modes entering
the
computational domain. Such a filter provides a non-reflective solution
thoughout the physical space of the annular blade row passage.
In view of this approach, the solution of the boundary-value scattering

problem at the upstream and downstream planes ${\cal F}=\{\zeta =T_x,T_x;$ $%


r_h\leq \xi \leq r_t;$ $0\leq \vartheta \leq 2\pi /N_b\}$ is expanded in
terms of the following infinite double series:
\begin{eqnarray}
\hat \phi _s(x,r,\theta ) =\sum_{m=0}^\infty \sum_{n=1}^\infty [C_{\nu
_m^{+},n}^{(u)}{\cal R}_{\nu _m^{+},n}(r)e^{i(\nu _m^{+}\theta +k_{\nu
_m^{+},n}^{(u)}x)}+ \hspace{1.5in} \nonumber \\
\ C_{\nu _m^{+},n}^{(d)}{\cal R}_{\nu _m^{+},n}(r)e^{i(\nu _m^{+}\theta
+k_{\nu _m^{+},n}^{(d)}x)}+ \hspace{1.0in}\nonumber \\
\ C_{\nu _m^{-},n}^{(u)}{\cal R}_{\nu _m^{-},n}(r)e^{i(\nu _m^{-}\theta
+k_{\nu _m^{-},n}^{(u)}x)}+ \hspace{1.0in} \nonumber \\
C_{\nu _m^{-},n}^{(d)}{\cal R}_{\nu _m^{-},n}(r)e^{i(\nu _m^{-}\theta
+k_{\nu _m^{-},n}^{(d)}x)}],\hspace{0.3in}{\bf x}\in{\cal F},
\hspace{0.5in} \label{B29}
\end{eqnarray}
where $n$ is an integer radial modal number, and $\nu _m^{\pm }=n_g\pm
mN_b$
is an integer circumferential modal number derived from the cascade
quasiperodicity condition in Appendix D. The signs $\pm $ correspond to
non-axisymmetric modes rotating in the direction or opposite to the
direction of the mean swirl. The axial wavenumbers $k_{\nu _m^{\pm
},n}^{(u)} $ and $k_{\nu _m^{\pm },n}^{(d)}$ represent the two families
of
eigenvalues which were obtained from the normal mode analysis of
equations (%
\ref{G11})-(\ref{G12}) in Chapter 7. In the frame of reference moving
with
the mean flow, these two families correspond to upstream and downstream
propagating radial eigenmodes ${\cal R}_{\nu _m^{\pm },n}(r)$,
respectively.
Finally, the unknown constant amplitudes $C_{\nu _m^{\pm },n}^{(u,d)}$
are
defined numerically by matching the asymptotic far field expansion
(\ref{B29}%
) with the near field solution of the boundary-value problem. A special
numerical algorithm developed to implement this matching is discussed in
details in Appendix E.
As a result of the blade surface coupling between the incident and
scattered
acoustic-vorticity waves in equation (\ref{B12}), the far field expansion
of
the scattered solution (\ref{B29}) must contain, in general, the
following
categories of eigenmodes:
(i) sonic or nearly-sonic non-decaying waves propagating away from the
blade
row;
(ii) sonic or nearly-sonic exponentially decaying (evanescent) waves
propagating away from the blade row;

(iii) sonic or nearly-sonic non-decaying waves propagating toward the


blade
row;
(iv) sonic or nearly-sonic exponentially decaying (evanescent) waves
propagating toward the blade row;
(v) convected or nearly-convected non-decaying waves propagating away
from
the blade row;
(vi) convected or nearly-convected non-decaying waves propagating toward
the
blade row.
At the upstream and downstream boundaries ${\cal F}$ of the computational
domain, the inflow non-physical solutions (iii), (iv) and (vi) must be
filtered out. Thus, only outgoing propagating and evanescent waves are
allowed upstream and downstream which satisfy the conditions
\begin{equation}
Re(k_{\nu _m^{\pm },n}^{(u,d)})\leq 0,\hspace{0.2in}Im(k_{\nu _m^{\pm
},n}^{(u,d)})\leq 0,\hspace{1.0in}x\rightarrow -\infty . \label{B30}
\end{equation}
and
\begin{equation}
Re(k_{\nu _m^{\pm },n}^{(u,d)})\geq 0,\hspace{0.2in}Im(k_{\nu _m^{\pm
},n}^{(u,d)})\geq 0,\hspace{1.0in}x\rightarrow +\infty , \label{B31}
\end{equation}
respectively. Note that the evanescent modes correspond to $Im(k_{\nu
_m^{\pm },n}^{(u,d)})\neq 0$.
In the case of a potential mean flow, the existing outgoing convected
eigenmodes are usually discarded in the formulation of outflow conditions
since they do not carry any pressure component and thus do not affect the
overall unsteady pressure response in the domain. Such an approach which
takes into account only outgoing sonic modes was validated by solving the
boundary-value problems for a 2D cascade in a uniform flow (Fang and
Atassi
\cite{FA}) and 3D cascade in a potential swirling flow (Golubev and
Atassi
\cite{GA1}).
However, a strong interaction between acoustic and vortical modes in
presence of a mean flow vorticity prohibits a purely sonic and purely
convected behavior of coupled acoustic-vorticiy waves. As it was shown in
Chapter 7, the nearly-convected eigenmodes form a discrete spectrum of
real
eigenvalues which may cluster with infinitely increasing density
approaching
the accumulation singular points of the critical layer, or vanish in its
neighborhood, depending on the circumferential modal number $m$. Such
vorticity-dominated, in general, modes may possess a significant pressure
component, especially if the corresponding nearly-convected eigenvalues

cluster far enough from the critical layer. In some cases (see examples
below) the nearly-convected modes may carry away the total unsteady
pressure
response pertaining to a certain modal number $m$. Therefore, such modes
should not be neglected in the formulation of the outflow conditions in
terms of the far field expansion (\ref{B29}) but they should also be
filtered to satisfy the outgoing wave criteria (\ref{B30}) and
(\ref{B31}).
A numerical implementation of the inflow/outflow conditions involves the
application of the pseudospectral technique for an eigenvalue analysis of
equations (\ref{G11})-(\ref{G12}) (see Chapter 7 for details). In
contrast
with a typically sparse filling of the matrix elements pertaining to the
governing equation and boundary conditions, the far field conditions may
fill up the corresponding matrix rows densely depending on the number of
terms taken in the truncated series (\ref{B29}). More non-zero entries
have
to be included at the downstream boundary where the far field condition
is
formulated for a scattered pressure function
$\widehat{p}_s=(ik_g-\partial
/\partial \zeta )\hat \phi _s,$ to be consistent with the wake
conditions.
\subsection{\bf \ Validation and Comparison with 2D Linear Cascade
Theory.}
The results of numerical solution for the boundary-value problem are
presented below in comparison with a flat plate linear cascade code based
on
the integral method (Ventres, \cite{Vent}). A brief discussion of the
conventional gust definition applied to model an upstream disturbance
field
in two-dimensional unsteady aerodynamics will be appropriate here (for
details, a reader can refer, e.g., to Goldstein \cite{AEROAC}, p.222).
Following this, an important note on the comparability of the present and
linear cascade analyses will be made.
\begin{figure}
\vspace{4in}
%\centerline{\psfig{figure=../Gvan/FORT/BLADES/AIAA97/lincas.eps,height=4
in,width=5in}}
\caption{Unsteady inflow in a linear cascade model (adapted from
\cite{AEROAC}.}
\label{bvp2}
\end{figure}
An infinite linear cascade can be obtained by ''unrolling'' the blade row
of
a fan at its mean radius $r_m.$ If we assume that the fan is rotating
with
angular velocity $\widetilde{\Omega }$ than the cascade will move with
velocity $U_c=\widetilde{\Omega }r_m\,$between two infinite lines
connecting

the blades leading and trailing edges (Figure \ref{bvp2}). Upstream of


the
unrolled blade row, a specified small-amplitude vortical velocity field $
%
{\bf u}_{-\infty }$ is imposed on a uniform mean flow $U_0$ which has a
flow
agle $\nu $ to the duct axis. The total nonuniform upstream velocity
field
consists of a vortical part ${\bf u}_{-\infty }$ and an acoustic part,
with
the latter representing an outgoing scattered wave.
The resulting mean flow velocity triangle in the cascade is shown in
Figure
\ref{bvp2}. In the frame of reference fixed with the blades, the total
velocity $U_{r}$ comprises an angle $\mu $ with the oncoming uniform
flow, and thus a cascade stagger angle $\chi $ has to be introduced to
provide zero angles of attack of the blades for the present problem. Note
that $\chi =\mu +\nu .$
The vortical velocity ${\bf u}_{-\infty }$ is periodic in the direction
of
motion of the blade row so that the annular circumference $2\pi r_m$ is
equal to an integer multiple of the disturbance wavelength. Since the
cascade radial direction has a finite spacing distance $b$, the upstream
vortical velocity field can be Fourier-decomposed in the coordinates $
{\bf X}%
=(X_1,X_2,X_3)$ aligned with the oncoming mean flow:
\begin{equation}
{\bf u}_{-\infty }=\sum_{m=-\infty }^\infty \sum_{n=-\infty }^\infty
[(A_{m,n}\widehat{{\bf I}}_1+B_{m,n}\widehat{{\bf I}}_2)\cos (\frac{\pi
nX_3}%
b)+C_{m,n}\widehat{{\bf I}}_3\sin (\frac{\pi nX_3}b)]e^{(imX_2/r_m\cos
\nu
)}, \label{B32}
\end{equation}
where $\hat{{\bf I}}_1,\hat{{\bf I}}_{2}$ and $\hat{{\bf
I}}_3$ are unit vectors in the $X_1$-$,X_2$- and $X_3$-directions; and $%
A_{m,n},B_{m,n}$ and $C_{m,n}$ are complex constant amplitudes of the
Fourier harmonics. Notice that the radial component of ${\bf
u}_{-\infty }$
vanishes at the walls $X_3=0,b$ to satisfy the wall impermeability
condition.
Since ${\bf u}_{-\infty }$ is a solenoidal velocity field,
\begin{equation}
\nabla _{{\bf X}}\cdot {\bf u}_{-\infty }=0, \label{B33a}
\end{equation}
which couples the complex amplitudes:
\begin{equation}
\frac{im}{\pi r_m\cos \nu }B_{m,n}+\frac nbC_{m,n}=0. \label{B33}
\end{equation}
Due to the linearity of the problem, it is customary to consider an
unsteady

aerodynamic response of the cascade to a {\it single} $(m,n)$-harmonic of


the disturbance field (\ref{B32}) (a gust). The coupling of the vortical
and
acoustic velocity components at a blade surface allows to formulate a
boundary value problem for a gust-cascade unsteady interaction which for
a
linear cascade reduces to an integral equation for an unsteady pressure
jump
across the blade surface.
The integral equation is solved in \cite{Vent} using a numerical
collocation
technique. The gust is considered in the cascade frame of reference ${\bf
Y}%
=(Y_1,Y_2,Y_3)$ with a unit upwash velocity component. In these
coordinates,
a gust harmonic from (\ref{B32}) can be represented as
\begin{equation}
{\bf u}_{-\infty }=\frac 12[{\bf a}e^{i(-\omega t+{\bf k\cdot Y})}+{\bf
a}%
^{\prime }e^{i(-\omega t+{\bf k}^{\prime }{\bf \cdot Y})}], \label{B34}
\end{equation}
where ${\bf a=\{}a_1,a_2,a_3\}$ is a vector of complex gust amplitudes,
and $%
{\bf k}=\{k_1,k_2,k_3\}$ is the gust wave vector in the ${\bf Y}$%
-coordinates. The components of the primed vectors in equation
(\ref{B34})
must be taken as ${\bf a}^{\prime }{\bf =\{}a_1,a_2,-a_3\}$ and ${\bf k}%
^{\prime }=\{k_1,k_2,-k_3\}.$ Thus, the complex amplitudes and wave
numbers
in the ${\bf Y}$- and ${\bf X}$-coordinates are related by
\begin{eqnarray}
a_1 &=&A_{m,n}\cos \mu +B_{m,n}\sin \mu , \nonumber \\
a_2 &=&-A_{m,n}\sin \mu +B_{m,n}\cos \mu , \label{B35} \\
a_3 &=&-iC_{m,n} \nonumber
\end{eqnarray}
and
\begin{eqnarray}
k_1 &=&m\widetilde{\omega }/U_r, \nonumber \\
k_2 &=&m\widetilde{\omega }\cot \mu /U_r, \label{B36} \\
k_3 &=&\pi n/b, \nonumber
\end{eqnarray}
where $\widetilde{\omega }=m\widetilde{\Omega }.$ Notice that the
divergence-free relation (\ref{B33a}) for the vortical velocity results
in
the condition ${\bf a\cdot k}=0.$
In view of relations (\ref{B35}) and (\ref{B36}), it is easy to see that
the
two vectors in the form (\ref{B34}) for ${\bf u}_{-\infty }$ will have
the
same effect on the unsteady response of the linear cascade. Thus, the
following gust representations in the oncoming flow and blade row frames
of

reference ${\bf X}$ and ${\bf Y}$,


\begin{equation}
{\bf u}_{-\infty }=[(A_{m,n}\widehat{{\bf I}}_1+B_{m,n}\widehat{{\bf I}}%
_2)\cos (\frac{\pi nX_3}b)+C_{m,n}\widehat{{\bf I}}_3\sin (\frac{\pi
nX_3}%
b)]e^{(imX_2/r_m\cos \nu )} \label{B37}
\end{equation}
and
\begin{equation}
{\bf u}_{-\infty }={\bf a}e^{i(-\omega t+{\bf k\cdot Y})}, \label{B38}
\end{equation}
are equivalent for a linear cascade provided the relations (\ref{B35})
and (%
\ref{B36}) are satisfied.
This observation appears to be very important in the comparison of linear
cascade results with results of the present 3D analysis. It follows from
relation (\ref{B37}) that the radial component of the vortical velocity
vanishes at the duct walls, whereas the other two components remain nonzero
there. In the particular case of an axial uniform flow around a rotor
taken
below for comparison, $\nu =0$, and the unit vectors $\widehat{{\bf
I}}_1,%
\widehat{{\bf I}}_{2}$ and $\widehat{{\bf I}}_3$ point in the axial,
circumferential, and radial directions, respectively. Thus, the radial
variations of the radial and circumferential gust components from
(\ref{B37}%
) are given by $v_{-\infty }\sim C_{m,n}\sin (\pi nX_3/b)$ and
$w_{-\infty
}\sim B_{m,n}\cos (\pi nX_3/b).$
However, in the case of a gust imposed on a swirling mean motion, {\it
both }%
the radial and circumferential components {\it must} vanish at the hub
and
tip radii in view of the gust amplitudes (\ref{B2}) and physical
arguments
given in Chapter 8. If we impose this restriction by considering, for
instance, the following gust solution in the mean flow ${\bf X}$%
-coordinates:
\begin{equation}
{\bf u}_{-\infty }=[(A_{m,n}\widehat{{\bf I}}_1+B_{m,n}\widehat{{\bf I}}%
_2)\sin (\frac{\pi nX_3}b)+C_{m,n}\widehat{{\bf I}}_3\sin (\frac{\pi
nX_3}%
b)]e^{(imX_2/r_m\cos \nu )}, \label{B39}
\end{equation}
then in the frame of reference fixed with the blades, we obtain for (\ref
{B34}) ${\bf a}^{\prime }{\bf =-a}$ and ${\bf k}^{\prime }=\{k_1,k_2,k_3\},$
where the ${\bf a}$-components are now defined by
\begin{eqnarray}

a_1 &=&-i(A_{m,n}\cos \mu +B_{m,n}\sin \mu ), \nonumber \\


a_2 &=&i(A_{m,n}\sin \mu -B_{m,n}\cos \mu ), \label{B40} \\
a_3 &=&-iC_{m,n} \nonumber
\end{eqnarray}
Thus, it is easily seen that {\it no} unsteady response will be produced
by
the gust of the form (\ref{B39}) in the linear cascade model. This
results
in a problem of comparability of 3D and 2D cascade effects and manifests
that the 3D case is not readily reducable to a 2D analog.
To validate and compare our numerical analysis, one may consider for the
linear cascade model two types of gusts typically produced in a rotorstator
interaction. The first is induced by a velocity defect in the wakes of
upstream-located blade rows:
\begin{equation}
{\bf u}_{-\infty }=A_{m,n}\widehat{{\bf I}}_1\cos (\frac{\pi nX_3}%
b)e^{(imX_2/r_m\cos \nu )} \label{B41}
\end{equation}
which in the cascade frame of reference translates to ${\bf u}_{-\infty }
=%
{\bf a}e^{i(-\omega t+{\bf k\cdot Y})}$ with
\begin{eqnarray}
a_1 &=&A_{m,n}\cos \mu , \nonumber \\
a_2 &=&-A_{m,n}\sin \mu , \label{B42} \\
a_3 &=&0 \nonumber
\end{eqnarray}
Notice that such a gust is two-dimensional if one considers an
axisymmetric
mode $n=0.$ The linear cascade results for the wake-induced gust will be
compared below with our numerical computations of the unsteady response
of a
3D cascade in uniform and swirling flows.
The second type of gust typical of turbomachinery is associated with
secondary unsteady flows (briefly discussed in Chapter 4):
\begin{equation}
{\bf u}_{-\infty }=[B_{m,n}\widehat{{\bf I}}_2\cos (\frac{\pi nX_3}b)
+C_{m,n}%
\widehat{{\bf I}}_3\sin (\frac{\pi nX_3}b)]e^{(imX_2/r_m\cos \nu )}
\label{B43}
\end{equation}
with corresponding amplitudes in ${\bf Y}$-coordinates:
\begin{eqnarray}
a_1 &=&B_{m,n}\sin \mu , \nonumber \\
a_2 &=&B_{m,n}\cos \mu , \label{B44} \\
a_3 &=&-B_{m,n}mb/\pi nr_m\cos \nu . \nonumber
\end{eqnarray}
These relations follow from the continuity condition (\ref{B33}) and

indicate that the gust produced by secondary flows is always oblique


($n\neq
0$).
In the integral equation solver \cite{Vent} for the linear cascade model,
the gust encountering the blade is always considered in the rotor frame
of
reference and taken of the form (\ref{B38}), with a unit upwash
component.
Thus, a proper normalization of the generalized gust solution ${\bf Z}_g$
from Chapter 8 is necessary to provide a maximum (to the extent indicated
above) compatibility of the results.
Recall that the phase dependence of ${\bf Z}_g$ in the general case of a
vortical swirling flow is determined by an interaction of two
characteristic
frequencies, the temporal $k_gU_0$ and the Coriolis $2\sqrt{\Omega
(\Omega
+\Gamma /r^2)}.$ The latter is typically small and hence modulates a
highly-oscillatory wave propagating with the temporal frequency. Our
primary
concern is then with the behavior of the dominant complex exponent $\exp
\{i[k_gx+m(\theta -(\Omega +\Gamma /r^2)x/U_0)-k_gU_0t]\}.$ Notice that
in
the limit of no swirl we recover the phase variation in (\ref{B38}) but
otherwise the mean swirl will, in general, induce an azimuthal phase
shift
which increases downstream. However, since for our boundary-value problem
the blades are aligned with the turning mean streamlines, this phase
shift
is canceled by the corresponding turn of the physical domain.
In view of the previous remarks of this section and continuity condition
(%
\ref{B3}), the gust imposed on a swirling mean motion will always have a
radial variation of its velocity components. The expansion (\ref{B2}) for
the circumferential and radial amplitudes of the vortical gust velocity
cancels both at the duct walls, in contrast with the amplitudes
(\ref{B35})
for a gust in a linear cascade model. In addition, the representation
(\ref
{B2}) gives complex amplitudes for $a_{r,\theta }(r)$ which for $r\gg 1$
(assumed for the comparison below) can be approximated using expansions
of
Hankel functions as follows:
\begin{equation}
a_{r,\theta }(r)\simeq D_{r,\theta }(\frac 8{\pi \mu _nr})^{1/2}\sin \mu
_n(r-r_h)[\sin (\mu _nr_h+\frac \pi 4)+i\sin (\mu _nr_h-\frac \pi 4)].
\label{B44a}
\end{equation}
The most consistent comparison then requires taking the absolute value of
an
amplitude of the normal to the blade surface velocity component equal to

unity. For the numerical calculations discussed below, we take $\mu


_1\simeq
\pi /(r_t-r_h)$. If no swirl is imposed, the amplitude of normal
component
corresponds to $a_\theta (r).$ The coefficient $D_\theta $ is then
adjusted
to provide the radial distribution of $|a_\theta (r)|$ as shown in Figure
\ref{bvp3} in comparison with radial variation of the normal vortical
velocity component adopted in the linear cascade model.
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/AIAA97/ingust.eps,height=3in
,width=5in}}
\caption{Imposed gust radial variation in linear cascade model (a) and
present analysis (b).}
\label{bvp3}
\end{figure}
If the swirl is imposed, then an additional adjustment of coefficients is
necessary to assure that the normal velocity component equals to unity.
The
normal to the blade surface projection of the {\it imposed} gust velocity
is
\begin{equation}
a_n(r)=a_x(r)\cos ({\bf n,}x)+a_r(r)\cos ({\bf n,}r)+a_\theta (r)\cos
({\bf %
n,}\theta ), \label{B44b}
\end{equation}
where the direction cosines are defined in (\ref{B16})-(\ref{B17}). For $
%
r\gg 1$ and $-1\leq x\leq 1,$ $|\nabla g|\simeq \sqrt{1+(M_s/M_0)^2}/r$
and $%
\cos ({\bf n,}r)$ is small so that (\ref{B44b}) can be approximated by
\begin{equation}
a_n(r)\simeq \frac 1{\sqrt{1+(M_s/M_0)^2}}[a_\theta (r)-\frac{M_s}{M_0}%
a_x(r)]. \label{B44c}
\end{equation}
Suppose $D_r=0,$ then from the continuity condition (\ref{B3}), $%
a_x(r)=-n_ga_\theta (r)/rk_g.$ If, in addition, $M_s=M_0$ at the mean
radius
$r=r_m$, then the magnitude of $a_n(r)$ reaches its maximum at this
position, equal to
\begin{equation}
a_n(r_m)\simeq \frac 1{\sqrt{2}}a_\theta (r_m)(1+\frac{n_g}{r_mk_g}),
\label{B44d}
\end{equation}
so that the coefficient $D_\theta $ can be easily corrected to ensure
that $%
|a_n(r_m)|=1.$
Note, however, that the above considerations apply only to the {\it
imposed}
gust amplitudes and do not take into account the gust evolution which
should

thus reveal its effect on the unsteady cascade response.


\subsection{Results and Discussion.}
The results for the unsteady aerodynamic response of a blade row are
compared for the absolute values of average and strip unsteady lift
coefficients,
\begin{equation}
|\overline{C}_L|=\frac{|L|}{\pi \overline{\rho }_0c(r_t-r_h)U_0|a_n|}
\hspace{1.0in} and \hspace{1.0in}|C_L^s|=\frac{|L^{\prime }|}{\pi
\overline{\rho }_0cU_0|a_n|}, \label{B45}
\end{equation}
calculated for a range of reduced frequencies $0.5\leq k_g\leq 10.$ The
total blade lift in (\ref{B45}) is denoted by $L$, and of the blade lift
at
a strip $s=$ $(r-r_h)/(r_t-r_h),$ by $L^{\prime }.$
For better illustration of aerodynamic cascade responses in an annular
domain, we also present the normalized real parts of unsteady pressure
jump
across the blade,
\begin{equation}
C_p^s(x)=\frac{Re[\Delta p^{\prime }(x,r)]}{\overline{\rho }_0cU_0|
a_\theta |%
}, \label{B46}
\end{equation}
for a set of $k_g=1,3,6,9$. For each reduced frequency, we show the
eigenvalues describing the modes propagating and decaying in the upstream
and downstream directions. These eigenvalues (and corresponding
eigenfunctions)\ are used in the formulation of the inflow/outflow
conditions. Finally, to control the effect of gust evolution on the
cascade
unsteady response, the downstream behavior of nearly-convected vorticity
and
pressure waves including the development of instability modes is followed
from the upstream boundary $x=x_{in}$ for a chosen set of gust parameters
and mean flow Mach numbers.
The mean flow regimes are chosen with account for two major applications
of
the present analysis found in naval and aerospace research. Marine
propellers usually operate at very low Mach numbers whereas the modern
aircraft turbomachines are designed for high subsonic and transonic flow
speeds. Thus for the first regime a mean flow with uniform and swirl flow
components $M_0=0.03$ and $\overline{M}_s=0.03$ is considered (bars over
swirl Mach numbers denote the value estimated at $r=r_m)$. For the second
regime, we take $M_0=0.6$ and $\overline{M}_s=0.6.$
In numerical computations, the cascade is placed in an annular duct with
hub
and tip radii, $r_h=4.25$ and $r_t=5.25$. The cascade has $N_b=30$ blades
to
provide a mid-blade spacing $\widetilde{s}=2\pi r_m/N_b\simeq 1$ taken
for

comparison in the integral method \cite{Vent}. Other input parameters


fixed
in computations after extensive numerical experiments on the code
validation, include the axial extent of the computational domain,
$2T_x=4$,
and numbers $M=3$ and $N=10$ which truncate the double series (\ref{B29})
in
the inflow/outflow conditions (see Appendix E). The total number of grid
points, $N_x\times N_r\times N_\theta =100\times 20\times 15,$ defines
the
total numbers of complex unknowns in the problem equal to 30,000. The
resulting from the finite-difference approximation (Section 9.1) linear
system of equations is solved using the Yale Sparse Matrix Package \cite
{YSMP} on IBM Risk-6000 workstation.
\subsubsection{\bf \ Uniform Flow.}
To validate the numerical results with \cite{Vent}, we first examine the
unsteady response of a blade row to a purely convected sinusoidal gust $
{\bf %
u}_{-\infty }=[a_x(r),0,a_\theta (r)]\exp [i(k_gx+n_g\theta -k_gM_0t)]$
imposed on a uniform flow. In the linear cascade model in Figure
\ref{bvp2},
this representation can be equivalent to a wake-induced gust of the form
(%
\ref{B41}) if the inflow and swirl angles are $\nu \rightarrow \pi /2$
and $%
\mu \rightarrow -\pi /2$, respectively, to provide zero stagger $\chi
=0.$
Thus, this case rather corresponds to an infinite swirl in Figure
\ref{bvp2}%
. In the cascade frame of reference, the corresponding normalized by $|
a_2|$
gust amplitudes are $a_1=a_3=0$ and $a_2=1.$ Note that for $\mu =-\pi
/2,k_2=k_1\cot \mu =0$, and thus an axisymmetric 1-D vortical disturbance
with $%
n_g=0$ is considered. From continuity, this also requires $a_x=0.$
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/clu003.eps,height=3in,wi
dth=5in}}
\caption{Unsteady lift coefficient vs. reduced frequency. Present results
for $|a_\theta (r)|=\sin k_3 s$: -, $|C_L^{0.5}|$; --, $|\bar{C}_L|$;
for $|a_\theta (r)|=\cos\pi s$: -.-, $|C_L^{1}|$. 2-D Cascade: 'o'.
$M_0=0.03,\chi =0$, $n_{g}=0$. }
\label{bvp4}
\end{figure}
Figure \ref{bvp4} compares the unsteady lift coefficients for $M_0=0.03$,
for three cases. In the first case of a {\it transverse} gust, $k_3=0$ so
that $a_\theta (r)=1.$ The results from \cite{Vent} (marked by circles)
appear to be very close to the present calculations of the average and

mid-span lift coefficients $|\overline{C}_L|$ and $|C_L^{0.5}|.$ If now $


%
k_3=\pi $ and $a_\theta (r)=\cos \pi s$ is taken in calculations to
create
analog of the gust (\ref{B38}) in the linear cascade model, the results
compare well again for $|\overline{C}_L|$ (note that $|C_L^{0.5}|\simeq
0$
due to the gust radial variation in Figure \ref{bvp3}a).
In the third case, we take $k_3=\pi $ and $a_\theta (r)$ from (\ref{B2})
(practically equivalent to $|a_\theta (r)|=\sin \pi s$ after
non-dimensionalization) thus considering adopted form of a gust imposed
on a
swirling motion. In the linear cascade model, such a gust produces no
unsteady response, and comparison can be made with the previous case. A
significant discrepancy is noted for low frequencies, with the gap
gradually
diminishing for higher $k_g$ and practically cancelling for
$|\overline{C}%
_L|.$ In fact, this difference may indicate the effect of
three-dimensionality in the problem. The unsteady lift remains obviously
higher at the strip $s=0.5$ where the maximum of upwash velocity is
reached.
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/eig_u003.eps,height=3in,
width=5in}}
\caption{Eigenvalues for upstream/downstream unsteady response. $k_{g}=1$
(a), 3 (b), 6 (c), 9 (d); $n_{g}=0$. $M_0=0.03, \chi =0.$ }
\label{bvp5}
\end{figure}
On the whole, the lift coefficients are shown to decrease as
$k_g\rightarrow
\infty ,$ in accordance with general theory. We also note a smooth
behavior
of solution as no acoustic resonance effects are produced for such a
small
Mach number. The latter is illustrated in Figure \ref{bvp5} which shows a
set of eigenvalues for acoustic modes used in the far field expansion of
the
potential solution, for reduced frequencies $k_g=1,3,6,9$. The blades
interacting with a gust radiate acoustic modes with the frequency of the
gust $\omega =k_gM_0$. There is a single pair of acoustic modes
propagating
in different directions, without any new acoustic cut-ons noticed for the
range of reduced frequences. All evanescent modes decay upstream due to
the
Doppler frequency shift in the stationary frame of reference. The modes
propagating upstream and amplifying are not physical (although they
appear
in the eigenvalue analysis), and therefore are discarded.
For the same set of reduced frequencies, the pressure jump coefficient $%

C_p^{0.5}(x)$ is depicted (for illustration) for the third case in Figure


\ref{bvp6}.
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/cp_u003.eps,height=3in,w
idth=5in}}
\caption{Aerodynamic response at blade strips: --,$s=0.25$;
-,$s=0.5$; -.-,$s=0.75$; for $k_{g}=1$ (a), 3 (b), 6 (c), 9 (d),
$n_{g}=0$. $M_0=0.03, \chi =0.$ }
\label{bvp6}
\end{figure}
The results for high subsonic uniform flow with $M_0=0.6$ are shown in
Figures \ref{bvp7}-\ref{bvp10}. The unsteady lift is calculated for
$k_3=\pi
,$ and the 2-D cascade results are compared with the present method in
Figure \ref{bvp7}. The results show the lift coefficients at the strips
with
unit upwash, $|C_L^1|$ in response to a gust with $a_\theta (r)=\cos \pi
s,$
and $|C_L^{0.5}|$ in response to a gust with $|a_\theta (r)|\simeq
\sin \pi
s.$ Note that {\it both} compare well for reduced frequencies up to $%
k_g\simeq 3.5$ where all three plots remain at almost the same constant
level of $|C_L|\simeq 0.08.$ The case of $a_\theta (r)=\cos \pi s$
actually
keeps this value up to $k_g\simeq 6$ and subsequently follows a resonant
behavior of linear cascade lift coefficient, with a certain phase shift.
This shift may be attributed to a number of causes. It may be connected
to
different acoustic resonance conditions in an annular duct, thus related
to
3-D effects of the problem. On the other hand, the 3-D cascade spacing
actually changes along the blade span within $\widetilde{s}=0.9...1.1$.
In
the linear cascade model, a change of spacing is known to result in a
different phase behavior of $C_L(k_1)$ curve. Hence, a distributed effect
of
such a change along the blade span in the 3-D cascade may cause a similar
phase effect on the lift. The strip pressure jumps corresponding to the
shifted peaks of $|C_L|$ at $k_g=6.5...8$ is illustrated in Figure
\ref{bvp8}
which shows large variations of $C_p^s(x)$ along the blade span obtained
in
numerical calculations.
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/clu06.eps,height=3in,wid
th=5in}}
\caption{Unsteady lift coefficient vs. reduced frequency. Present results
for $|a_\theta (r)|=\sin\pi s$: $'+'$, $|C_L^{0.5}|$; for $|a_\theta
(r)|=\cos\pi s$: 'x', $|C_L^{1}|$. 2-D cascade analysis: 'o'.
$M_0=0.6,\chi =0, n_{g}=0$ }
\label{bvp7}

\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/cp_uvent06.eps,height=3i
n,width=5in}}
\caption{Aerodynamic response to a gust with $|a_\theta (r)|=\cos\pi s$,
at blade strips: --,$s=0.25$;
-,$s=0.5$; -.-,$s=0.75$; -,$s=1,$ compared to 2-D cascade analysis
(dots). $k_{g}=6.5$ (a), 7 (b), 7.5 (c), 8 (d); $n_{g}=0$. $M_0=0.6,
\chi =0.$ }
\label{bvp8}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/eig_u06.eps,height=3in,w
idth=5in}}
\caption{Eigenvalues for upstream/downstream unsteady response. $k_{g}=1$
(a), 3 (b), 6 (c), 9 (d); $n_{g}=0$. $M_0=0.6, \chi =0.$ }
\label{bvp9}
\end{figure}
The gust (\ref{B2}) used henceforth produces an unsteady lift which
matches
well with the linear cascade result up to frequency $k_g\simeq 6$ but
then
does not exhibit any resonant peaks; it appears that the comparison is
improving for higher frequencies as the lift is decreasing.
The complex behavior of $C_L(k_g)$ plots in Figure \ref{bvp7} is
connected
with acoustic cut-on of modes in the far field of the blade row. The
unsteady lift is known to be very sensitive and exhibit large variations
when the frequency approaches the cut-on conditions (see, e.g.,
\cite{FA}).
The near and far field unsteady responses thus strongly affect each
other,
which emphasizes the importance of the correct formulation for the
inflow/outflow conditions. For $k_g=1,4.5,5.5,8.5,$ Figure \ref{bvp9}
shows
a set of eigenvalues defining the modes able to propagate in both
directions. For $k_g=1,$ the acoustic response includes two modes
propagating upstream and downstream, and a set of evanescent waves in the
upstream direction. A new acoustic cut-on occurs at $k_g\simeq 4.5,$ when
a
new {\it pair} of modes starts propagating upstream. This near-cut-on
situation is discussed in Appendix E in connection with ill-posed
upstream
boundary conditons. The pair of modes finally splits in the upstream and
downstream directions at $k_g\simeq 5.5.$ Another acoustic resonance
occurs
at $k_g\simeq 8.5$ when a new-born modal pair appears upstream in
Figure \ref{bvp9}d .
\begin{figure}

\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/cp_u06.eps,height=3in,wi
dth=5in}}
\caption{Aerodynamic response at blade strips: --,$s=0.25$;
-,$s=0.5$; -.-,$s=0.75$. $k_{g}=1$ (a), 3 (b), 6 (c), 9 (d); $n_{g}=0$.
$M_0=0.6, \chi =0.$ }
\label{bvp10}
\end{figure}
The variation of pressure coefficient $C_p^s$ along the mid-span blade
chord
is shown in Figure \ref{bvp10} for the gust (\ref{B2}), for reduced
frequencies $k_g=1,3,6,9.$ These results will be compared below for the
effect of swirl on the cascade unsteady pressure response.
\subsubsection{\bf Swirling Flow}
The effect of a $45^0$-swirl on the unsteady aerodynamic response of the
annular cascade is investigated for two cases corresponding to potential
and
vortical swirling flows. At its initial position $x=-T_x$, an unsteady
vortical disturbance is imposed on a mean flow in the form of a
sinusoidal
gust,
\begin{equation}
{\bf u}_{-\infty }=[a_x(r),a_r(r),a_\theta (r)]\exp [i\{k_gx+n_g(\theta %
\frac{xM_s}{rM_0})-k_gM_0t\}] \label{B47}
\end{equation}
where the amplitudes are defined by (\ref{B2}) and (\ref{B3}). For
simplicity, we take $D_r=0$ and thus $a_r(r)=0.$
In terms of the linear cascade analogy in Figure \ref{bvp2}, one may
think
of a wake-induced convected velocity defect (\ref{B41}) imposed on a
uniform
inflow with $\nu =0$ and interacting with a rotor which induces a swirl
angle $\mu =\pi /4$. Unloaded rotor blades are thus installed at a
stagger
angle $\chi =\pi /4.$ Another analogy suggests a secondary flow - induced
gust (\ref{B43}) imposed on the same mean flow geometry; however, the
continuity condition (\ref{B33}) requires a radial gust component to
enter
the problem. Nevertheless, although {\it initially} $a_r(r)=0$ is
assumed,
the radial component will {\it develop} in a swirling flow as a results
of
the gust evolution downstream (discussed in Chapter 8), which makes the
secondary flow analogy reasonable.
The numerical results are compared below with the integral method
\cite{Vent}
for a linear cascade. Although the gust definitions in both cases are

different in view of the previous discussion, such a comparison should


stress the effect of swirl on the unsteady cascade response.
In the linear cascade analogy, since $\mu \neq 0$, a vortical disturbance
convected along the uniform inflow streamlines, appears as nonaxisymmetric
in the rotating blade frame of reference (Figure \ref{bvp2}), with the
gust
wavenumbers related by (\ref{B36}),
\begin{equation}
k_2=k_1\cot \mu . \label{B47b}
\end{equation}
For a cascade in a swirling flow, a similar requirement should be
satisfied
for a gust which is ''nearly-convected'' along the swirling streamlines
of $%
U_r$ in Figure \ref{bvp2}, so that $\mu (r)=0.$ Thus, from (\ref{B47}),
the
same relation as (\ref{B47b}),
\begin{equation}
k_g-\frac{n_gM_s(r)}{rM_0}=\frac{n_g}r\tan \mu (r)=0 \label{B48}
\end{equation}
guarantees a constant azimuthal phase anlge of the incident vortical
disturbance for any frequency of unsteady excitation, at a given blade
strip
$r=const$.
Since the blades are aligned with turning streamlines, their stagger
angles
vary as $\tan \chi (r)=M_s(r)/M_0,$ and thus, from (\ref{B48}), the
condition
\begin{equation}
n_g=r_mk_g(\overline{M}_\Gamma +r^2\overline{M}_\Omega /r_m^2)/M_0
\label{B49}
\end{equation}
provides constant azimuthal gust phase along the blade surface. However,
for
fixed $k_g$ and $n_g$ in numerical computations, this condition cannot be
satisfied unless $M_s=M_\Gamma .$ In the present calculations for a
$45^0$%
-swirl, we assume $n_g=r_mk_g,$ with $k_3=\pi $ as before. In addition,
the
case of an axisymmetric vortical disturbance, $n_g=0,$ is considered for
comparison.
\subsubsection{\bf Low Mach numbers.}
We start with analysis of the unsteady effect of impinging axisymmetric
(in
the stationary duct frame of reference) gust, imposed with initial form
(\ref{B47})
on a swirling mean flow with axial component $M_0=0.03.$ Figure \ref
{bvp11} compares the lift coefficients $|C_L^{0.5}|$ obtained for
potential

and vortical swirling flows with $M_0=M_\Gamma =0.03$ and


$M_0=0.03,M_\Gamma
=0.01,M_\Omega =0.02,$ respectively. The linear cascade results put lower
bounds to both curves which exhibit significant variations about the
uniform
flow curve. For higher reduced frequencies, the unsteady responses
approach
the uniform flow limit, and thus one may conclude that the swirl effect
is
more significant for lower frequencies of unsteady excitation. The
oscillations of lift coefficients which seem to increase for higher $k_g$
may be related to a standing wave resonance in a blade passage due to
interaction of staggered blades with an axisymmetric incident wave.
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/cls003.eps,height=3in,wi
dth=5in}}
\caption{Unsteady lift coefficient vs. reduced frequency. Present
results: -, $M_0=0.03$ ($\chi =0$); $'+'$, $M_0=\bar{M}_{\Gamma}=0.03$
($\bar{\chi} =45^{o}$); 'x',
$M_0=0.03, \bar{M}_{\Gamma}=0.01,\bar{M}_{\Omega}=0.02 (\bar{\chi}
=45^{o})$. 2-D cascade analysis: 'o', $M_0=0.03, \chi =0$ . $n_{g}=0.$ }
\label{bvp11}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/cp_ug003.eps,height=3in,
width=5in}}
\caption{Aerodynamic response at blade strips: --,$s=0.25$;
-,$s=0.5$; -.-,$s=0.75$. $k_{g}=1$ (a), 3 (b), 6 (c), 9 (d); $n_{g}=0$.
$M_0=0.03, \bar{M}_{\Gamma}=0.03, \bar{\chi} =45^{o}.$ }
\label{bvp12}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/cp_ugo003.eps,height=3in
,width=5in}}
\caption{Aerodynamic response at blade strips: --,$s=0.25$;
-,$s=0.5$; -.-,$s=0.75$. $k_{g}=1$ (a), 3 (b), 6 (c), 9 (d); $n_{g}=0$.
$M_0=0.03, \bar{M}_{\Gamma}=0.01, \bar{M}_{\Omega}=0.02, \bar{\chi}
=45^{o}.$ }
\label{bvp13}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/eig_ug003.eps,height=3in
,width=5in}}
\caption{Eigenvalues for upstream/downstream unsteady response. $k_{g}=1$
(a), 3 (b), 6 (c), 9 (d); $n_{g}=0$. $M_0=0.03,
\bar{M}_{\Gamma}=0.03, \bar{\chi}=45^{o}.$ }
\label{bvp14}
\end{figure}
\begin{figure}

\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/eig_ugo003.eps,height=3i
n,width=5in}}
\caption{Eigenvalues for upstream/downstream unsteady response. $k_{g}=1$
(a), 3 (b), 6 (c), 9 (d); $n_{g}=0$. $M_0=0.03, \bar{M}_{\Gamma}=0.01,
\bar{M}_{\Omega}=0.02, \bar{\chi}=45^{o}.$ }
\label{bvp15}
\end{figure}
The variation of $C_p^s(x)$ along the span of a blade in Figures
\ref{bvp12}
and \ref{bvp13} for potential and vortical swirling flows, is more
significant than in the similar cases of a uniform flow (Figure
\ref{bvp6})
but still moderate compared to large variations for high-speed flows.
Note
that the integral method in fact compares well with the present results
as
far as general trends are concerned; the amplitudes are different due to
different gust normalization in (\ref{B46}). However, the linear cascade
results do not show oscillations along the chord for higher $k_g.$ Such
an
oscillatory behavior, less pronounced for vortical swirl, may at least
partially be attributed to the appearance of downstream evanescent modes
in
the eigenmode analysis of unsteady swirling flows. The wavenumbers
corresponding to propagating or decaying waves in the cases of potential
and
vortical mean swirl components, are shown in Figures \ref{bvp14} and \ref
{bvp15}, respectively. As for the case of no swirl, there are one
upstream
and one downstream propagating acoustic modes. In addition, the Doppler
shift causes infinite sets of evanescent ''pressure-dominated''
eigenmodes
to propagate in both directions, with all radial modes spinning in the
direction of mean swirl, decaying downstream, and those spinning in the
opposite direction, decaying upstream.
The difference between two cases appears in the nearly-convected region
of
eigenvalues. For the potential swirl, the ''vorticity-dominated''
eigenmodes
have {\it convected} double eigenvalues and do not contribute to the
unsteady pressure field far downstream. On the contrary, certain {\it %
nearly-convected} modes in the vortical swirl may produce an unsteady
pressure field which should be accounted in the downstream outflow
conditions. Such modes typically have eigenvalues which cluster far
enough
from the critical layer. A special numerical procedure selects such modes
and includes them in the series expansion (\ref{B29}) based on the
criterion
of pressure norm comparison for evanescent ''pressure-dominated'' and

propagating or slowly amplifying ''vorticity-dominated'' modes. According


to
this criterion, two axisymmetric ($\nu _0=0)$ nearly-convected modes were
included for $k_g=1$ and $k_g=3$ (Figures \ref{bvp15}a,b), no modes
selected
for $k_g=6$ (Figure \ref{bvp15}c), and one non-axisymmetric nearlyconvected
mode with $\nu _{-2}=-60$ was chosen for $k_g=9$ (Figure \ref{bvp15}d).
The gust evolution in the cases of potential and vortical swirling flows
reflects the main differences of the eigenmode analyses. The radial
variations of gust vortical velocities and pressure at the initial, $x=T_x,$
and mid-chord, $x=0,$ stations are shown for $k_g=1$ and $k_g=9$ in
Figures
\ref{bvp16}, \ref{bvp18} for the mean flow $M_0=M_\Gamma =0.03,$ and in
Figures \ref{bvp20},\ref{bvp22} for $M_0=0.03,M_\Gamma =0.01,M_\Omega
=0.02.$
The corresponding plots of the gust downstream evolution along the mean
flow
streamlines ($\vartheta =const$) are presented for radial strips $%
s=1/4,1/2,3/4$ in Figures \ref{bvp17}, \ref{bvp19} and \ref{bvp21}, \ref
{bvp23}, respectively. A 1-D gust imposed initially with amplitudes $%
a_x(r)=a_r(r)=0$ and $|a_\theta (r)|\simeq \sin \pi s$, develops a
secondary
radial perturbation of the order of circumferential disturbance, at
$x=0.$
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gr_ug06_1.eps,height=3in
,width=5in}}
\caption{Radial
variations of gust vortical velocity and pressure amplitudes: -, real
parts, --, imaginary parts, '.', absolute values. Top: initial station;
bottom: mid-chord station. $k_{g}=1, n_{g}=0$. $M_0=0.03,
\bar{M}_{\Gamma}=0.03.$ }
\label{bvp16}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gr_ug06_9.eps,height=3in
,width=5in}}
\caption{Radial
variations of gust vortical velocity and pressure amplitudes: -, real
parts, --, imaginary parts, '.', absolute values. Top: initial station;
bottom: mid-chord station. $k_{g}=9, n_{g}=0$. $M_0=0.03,
\bar{M}_{\Gamma}=0.03.$ }
\label{bvp17}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gr_ugo003_1.eps,height=3
in,width=5in}}

\caption{Radial
variations of gust vortical velocity and pressure amplitudes: -, real
parts, --, imaginary parts, '.', absolute values. Top: initial station;
bottom: mid-chord station. $k_{g}=1, n_{g}=0$. $M_0=0.03,
\bar{M}_{\Gamma}=0.01,
\bar{M}_{\Omega}=0.02.$ }
\label{bvp18}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gr_ugo003_9.eps,height=3
in,width=5in}}
\caption{Radial
variations of gust vortical velocity and pressure amplitudes: -, real
parts, --, imaginary parts, '.', absolute values. Top: initial station;
bottom: mid-chord station. $k_{g}=9, n_{g}=0$. $M_0=0.03,
\bar{M}_{\Gamma}=0.01,
\bar{M}_{\Omega}=0.02.$ }
\label{bvp19}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gx_ug06_1.eps,height=3in
,width=5in}}
\caption{Downstream evolution
of real parts of gust vortical velocity and pressure amplitudes, at
radial strips: --, $s=0.25$, -, $s=0.5$, '.', $s=0.75$. $k_{g}=1,
n_{g}=0$. $M_0=0.03, \bar{M}_{\Gamma}=0.03.$ }
\label{bvp20}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gx_ug06_9.eps,height=3in
,width=5in}}
\caption{Downstream evolution
of real parts of gust vortical velocity and pressure amplitudes, at
radial strips: --, $s=0.25$, -, $s=0.5$, '.', $s=0.75$. $k_{g}=9,
n_{g}=0$. $M_0=0.03, \bar{M}_{\Gamma}=0.03.$ }
\label{bvp21}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gx_ugo003_1.eps,height=3
in,width=5in}}
\caption{Downstream evolution
of real parts of gust vortical velocity and pressure amplitudes, at
radial strips: --, $s=0.25$, -, $s=0.5$, '.', $s=0.75$. Left: total
values; right: amplifying components. $k_{g}=1, n_{g}=0$. $M_0=0.03,
\bar{M}_{\Gamma}=0.01, \bar{M}_{\Omega}=0.02.$ }
\label{bvp22}
\end{figure}
\begin{figure}

\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gx_ugo003_9.eps,height=3
in,width=5in}}
\caption{Downstream evolution
of real parts of gust vortical velocity and pressure amplitudes, at
radial strips: --, $s=0.25$, -, $s=0.5$, '.', $s=0.75$. Left: total
values; right: amplifying components. $k_{g}=9, n_{g}=0$. $M_0=0.03,
\bar{M}_{\Gamma}=0.01, \bar{M}_{\Omega}=0.02.$ }
\label{bvp23}
\end{figure}
The major difference between the cases of potential and vortical flows is
noticed for the gust unsteady pressure evolution. At the mid-chord loci,
the
maximum pressure response for $k_g=1$ in the potential swirl is about two
times smaller than in the vortical swirl, and rapidly decreases for
higher
reduced frequencies. For the vortical swirling flow, such a response
stays
at about the same level for the range of $k_g=1...10,$ with a significant
contribution coming from amplifying nearly-convected modes, for lower
reduced frequencies.
For an incident 2--D\ non-axisymmetric acoustic-vorticity wave with $%
n_g=r_mk_g,$ a similar set of results is presented in Figures
\ref{bvp24}-%
\ref{bvp36}. The unsteady lift coefficients $|C_L^{0.5}|$ show a
significant
divergence for potential and vortical mean swirling flows for small
reduced
frequencies but compare well for $k_g\geq 5.$ Both the phase dependence
and
amplitudes of the curves are different from the linear cascade results,
although at least the amplitude deviation appears to decrease for higher
reduced frequencies. Among the possible reasons for these differences, we
mention the effects of a non-axisymmetric gust evolution in swirling
flows.
Recall that the circumeferential wavenumbers in the 2-D and 3-D cascade
models are not completely compatible, so that different cascade
quasiperiodicity conditions, along with a variable cascade spacing, may
induce a phase deviation for the lift coefficient curves. These results
are
further illustrated in Figures \ref{bvp25}, \ref{bvp26} which compare the
pressure jump coefficients for the potential and vortical mean flows. In
both cases, a large variation of $C_p^s(x)$ along the blade span is noted
at
the leading edge, which occurs only for small reduced frequencies.
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/cls003ng.eps,height=3in,
width=5in}}

\caption{Unsteady lift coefficient vs. reduced frequency. Present


results: $'+'$, $M_0=\bar{M}_{\Gamma}=0.03$; 'x',
$M_0=0.03, \bar{M}_{\Gamma}=0.01,\bar{M}_{\Omega}=0.02$.
2-D cascade analysis: 'o'. $\bar{\chi} =45^{o}, n_{g}=k_{g}r_{m}.$ }
\label{bvp24}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/cp_ug003ng.eps,height=3i
n,width=5in}}
\caption{Aerodynamic response at blade strips: --,$s=0.25$;
-,$s=0.5$; -.-,$s=0.75$. $k_{g}=1, n_{g}=5$ (a), $k_{g}=3, n_{g}=15$ (b),
$k_{g}=5.8, n_{g}=29$ (c), $k_{g}=9, n_{g}=45$ (d). 2-D cascade analysis:
'o', $k_{1}=k_{2}$. $M_0=0.03, \bar{M}_{\Gamma}=0.03, \bar{\chi} =45^{o}.
$ }
\label{bvp25}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/cp_ugo003ng.eps,height=3
in,width=5in}}
\caption{Aerodynamic response at blade strips: --,$s=0.25$;
-,$s=0.5$; -.-,$s=0.75$. $k_{g}=1, n_{g}=5$ (a), $k_{g}=3, n_{g}=15$ (b),
$k_{g}=5.8, n_{g}=29$ (c), $k_{g}=9, n_{g}=45$ (d). 2-D cascade analysis:
'o', $k_{1}=k_{2}$. $M_0=0.03, \bar{M}_{\Gamma}=0.01,
\bar{M}_{\Omega}=0.02, \bar{\chi} =45^{o}. $ }
\label{bvp26}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/eig_ug003ng.eps,height=3
in,width=5in}}
\caption{Eigenvalues for upstream/downstream unsteady response.
$k_{g}=1,n_{g}=5 $ (a), $k_{g}=3, n_{g}=15$ (b), $k_{g}=5.8, n_{g}=29$
(c), $k_{g}=9, n_{g}=45$ (d); $M_0=0.03, \bar{M}_{\Gamma}=0.03,
\bar{\chi}=45^{o}.$ }
\label{bvp27}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/eig_ugo003ng.eps,height=
3in,width=5in}}
\caption{Eigenvalues for upstream/downstream unsteady response.
$k_{g}=1,n_{g}=5 $ (a), $k_{g}=3, n_{g}=15$ (b), $k_{g}=5.8, n_{g}=29$
(c), $k_{g}=9, n_{g}=45$ (d); $M_0=0.03,\bar{M}_{\Gamma}=0.01,
\bar{M}_{\Omega}=0.02, \bar{\chi}=45^{o}.$ }
\label{bvp28}
\end{figure}
The results of eigenvalue analyses in the present cases (Figures
\ref{bvp27}%
, \ref{bvp28}) show the branches of acoustic eigenvalues corresponding to

the circumferential modal numbers $\nu _m^{\pm }=n_g\pm mN_b,m=0...3$,


and
radial modal numbers $n=1...10.$ The modes from the first quadrant of the
complex $k_{\nu _m^{\pm },n}$-plane, propagate or decay downstream,
whereas
those from the third quadrant, appear upstream. The differences in the
two
cases are limited to the Doppler shift which results in an especially
complicated pattern of eigenvalues in the case of the vortical swirl
(Figure
\ref{bvp28}). Both figures indicate that the eigenvalues retain their
patterns almost the same as the reduced frequency changes.
No nearly-convected modes are included in the modal expansion of the
outflow
conditions for the vortical swirling flow. This indicates that the
unsteady
pressure field associated with such modes is small compared to the one
for
evanescent nearly-sonic modes. It can be expected that the corresponding
gust solution of the initial-value problem will also show a moderate
level
of the unsteady pressure response downstream, and indeed this can be seen
from Figures \ref{bvp33}-\ref{bvp36}. At $x=0$, this response is of the
same
order for $k_g=1$ as for the potential swirling flow in Figures
\ref{bvp29}-%
\ref{bvp32}, and then slowly decreases for higher $k_g$. Note that this
drop
occurs much faster in the case of a potential swirl.
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gr_ug003ng_1.eps,height=
3in,width=5in}}
\caption{Radial
variations of gust vortical velocity and pressure amplitudes: -, real
parts, --, imaginary parts, '.', absolute values. Top: initial station;
bottom: mid-chord station. $k_{g}=1, n_{g}=5$. $M_0=0.03,
\bar{M}_{\Gamma}=0.03.$ }
\label{bvp29}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gr_ug003ng_9.eps,height=
3in,width=5in}}
\caption{Radial
variations of gust vortical velocity and pressure amplitudes: -, real
parts, --, imaginary parts, '.', absolute values. Top: initial station;
bottom: mid-chord station. $k_{g}=9, n_{g}=45$. $M_0=0.03,
\bar{M}_{\Gamma}=0.03.$ }
\label{bvp30}
\end{figure}
\begin{figure}

\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gr_ugo003ng_1.eps,height
=3in,width=5in}}
\caption{Radial
variations of gust vortical velocity and pressure amplitudes: -, real
parts, --, imaginary parts, '.', absolute values. Top: initial station;
bottom: mid-chord station. $k_{g}=1, n_{g}=5$. $M_0=0.03,
\bar{M}_{\Gamma}=0.01,
\bar{M}_{\Omega}=0.02.$ }
\label{bvp31}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gr_ugo003ng_9.eps,height
=3in,width=5in}}
\caption{Radial
variations of gust vortical velocity and pressure amplitudes: -, real
parts, --, imaginary parts, '.', absolute values. Top: initial station;
bottom: mid-chord station. $k_{g}=9, n_{g}=45$. $M_0=0.03,
\bar{M}_{\Gamma}=0.01,
\bar{M}_{\Omega}=0.02.$ }
\label{bvp32}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gx_ug003ng_1.eps,height=
3in,width=5in}}
\caption{Downstream evolution
of real parts of gust vortical velocity and pressure amplitudes, at
radial strips: --, $s=0.25$, -, $s=0.5$, '.', $s=0.75$. $k_{g}=1,
n_{g}=5$. $M_0=0.03, \bar{M}_{\Gamma}=0.03.$ }
\label{bvp33}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gx_ug003ng_9.eps,height=
3in,width=5in}}
\caption{Downstream evolution
of real parts of gust vortical velocity and pressure amplitudes, at
radial strips: --, $s=0.25$, -, $s=0.5$, '.', $s=0.75$. $k_{g}=9,
n_{g}=45$. $M_0=0.03, \bar{M}_{\Gamma}=0.03.$ }
\label{bvp34}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gx_ugo003ng_1.eps,height
=3in,width=5in}}
\caption{Downstream evolution
of real parts of gust vortical velocity and pressure amplitudes, at
radial strips: --, $s=0.25$, -, $s=0.5$, '.', $s=0.75$. Left: total
values; right: amplifying components. $k_{g}=1, n_{g}=5$. $M_0=0.03,
\bar{M}_{\Gamma}=0.01, \bar{M}_{\Omega}=0.02.$ }

\label{bvp35}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/gx_ugo003ng_9.eps,height
=3in,width=5in}}
\caption{Downstream evolution
of real parts of gust vortical velocity and pressure amplitudes, at
radial strips: --, $s=0.25$, -, $s=0.5$, '.', $s=0.75$. Left: total
values; right: amplifying components. $k_{g}=9, n_{g}=45$. $M_0=0.03,
\bar{M}_{\Gamma}=0.01, \bar{M}_{\Omega}=0.02.$ }
\label{bvp36}
\end{figure}
In the vortical mean flow, the radial non-axisymmetric vortical velocity
component is non-zero at the duct walls. The impermeability condition is
provided by a corresponding component of the potential gust velocity (not
shown) which cancels the vortical counterpart. All the vortical velocity
components have comparable non-zero amplitudes at the mid-chord position,
so
that an initially 2-D vortical perturbation develops into a 3-D
acoustic-vorticity wave interacting with the blades. The\ resulting
complex
interference of phases and amplitudes of the gust components along the
blade
surfaces may induce an unsteady aerodynamic response which significantly
diverges from the linear cascade results, which was observed above.
\subsubsection{\bf High Mach numbers.}
In the last section of this dissertation, we examine the unsteady
aerodynamic response of a 3-D cascade interacting with axisymmetric 2-D
and
non-axisymmetric 3-D acoustic-vortcity waves in a high-speed swirling
flow
typical of turbomachinery applications. As before, two cases of potential
and vortical swirls are considered, with $M_0=\overline{M}_\Gamma =0.6$
and $%
M_0=0.6,\overline{M}_\Gamma =\overline{M}_\Omega =0.3,$ respectively.
For an axisymmetric impinging gust ($n_g=0$), the unsteady lift
coefficients
$|C_L^{0.5}|$ are compared with the uniform flow and 2-D cascade analyses
in
Figure \ref{bvp37}. All results show good agreement for low reduced
frequencies, up to $k_g\simeq 4.$ In the resonance region of frequencies,
the curves diverge significantly, but some unexpected trends can be
noticed.
The results for the potential swirl in fact remain close to those
obtained
for the uniform flow, and follow the single-peak behavior of the latter
over
the set of $k_g.$ Similarly, the case of the vortical flow happens to
match

the results of the integral method, but misses one peak. Moreover, both
curves occur to be in phase (compare with Figure \ref{bvp7}). Note that
for
higher reduced frequencies, all the analyses again show good comparison.
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/cls06.eps,height=3in,wid
th=5in}}
\caption{Unsteady lift coefficient vs. reduced frequency. Present
results: -, $M_0=0.6$ ($\chi =0$); $'+'$, $M_0=\bar{M}_{\Gamma}=0.6$
($\bar{\chi} =45^{o}$); 'x',
$M_0=0.6, \bar{M}_{\Gamma}=0.3,\bar{M}_{\Omega}=0.3 (\bar{\chi} =45^{o})
$. 2-D cascade analysis: 'o', $M_0=0.6, \chi =0$. $n_{g}=0.$ }
\label{bvp37}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/eig_ug06.eps,height=3in,
width=5in}}
\caption{Eigenvalues for upstream/downstream unsteady response. $k_{g}=1$
(a), 3 (b), 6 (c), 9 (d); $n_{g}=0$. $M_0=0.6, \bar{M}_{\Gamma}=0.6,
\bar{\chi}=45^{o}.$ }
\label{bvp38}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/eig_ugo06.eps,height=3in
,width=5in}}
\caption{Eigenvalues for upstream/downstream unsteady response. $k_{g}=1$
(a), 3 (b), 6 (c), 9 (d); $n_{g}=0$. $M_0=0.6, \bar{M}_{\Gamma}=0.3,
\bar{M}_{\Omega}=0.3, \bar{\chi}=45^{o}.$ }
\label{bvp39}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/cp_ug06.eps,height=3in,w
idth=5in}}
\caption{Aerodynamic response at blade strips: --,$s=0.25$;
-,$s=0.5$; -.-,$s=0.75$. $k_{g}=1$ (a), 3 (b), 6 (c), 9 (d); $n_{g}=0$.
$M_0=0.6, \bar{M}_{\Gamma}=0.6, \bar{\chi} =45^{o}.$ }
\label{bvp40}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/cp_ugo06.eps,height=3in,
width=5in}}
\caption{Aerodynamic response at blade strips: --,$s=0.25$;
-,$s=0.5$; -.-,$s=0.75$. $k_{g}=1$ (a), 3 (b), 6 (c), 9 (d); $n_{g}=0$.
$M_0=0.6, \bar{M}_{\Gamma}=0.3, \bar{M}_{\Omega}=0.3, \bar{\chi} =45^{o}.
$ }
\label{bvp41}
\end{figure}

The mentioned trends appear to be unusual, especially in view of the


eigenmode analyses in the cases of swirling flows. Instead of few
acoustic
pairs cut-on in the uniform flow (Figure \ref{bvp9}), plenty of
propagating
modes rapidly switch on as $k_g$ increases in Figures
\ref{bvp38},\ref{bvp39}.
Most of such modes spin in the direction opposite to the mean swirl, and
propagate upstream from the blade row. The eigenvalue patterns in both
figures are close to each other compared to the corresponding results of
the
preceding section for low-speed flows. In addition, for the vortical
swirling flow, two axisymmetric ($\nu _0=0$) nearly-convected modes are
included in the downstream outflow condition for $k_g=1$ and $k_g=3,$ one
such mode appears for $k_g=6,$ and one non-axisymmetric nearly-convected
mode ($\nu _0=-30$) carries significant pressure component downstream for
$%
k_g=9.$ Note that for high mean flow velocities, the critical layer
appears
closer to the region of acoustic eigenvalues as the speed of convection
approaches the speed of sound.
Solutions of initial-value problem for a gust evolution are normalized by
a
uniform flow velocity component, and thus depend only on the portion of
swirl in the mean flow, and ratio between defferent swirl components. In
both cases of low-speed and high-speed flows, $\overline{M}_s/M_0=1.$
Hence,
for the potential swirl, the results presented in Figures
\ref{bvp16}-\ref
{bvp19} and discussed in the preceding section, apply to the present
case.
For the mixed vortical flow, the ratio $\overline{M}_\Omega /\overline{M}
%
_\Gamma =1$ changes from $\overline{M}_\Omega /\overline{M}_\Gamma =2,$
but
the effect of this change on the gust evolution appeared to be
insignificant, and thus we also refer to Figures \ref{bvp20}-\ref{bvp23}
and
their discussion.
In view of the previous remarks, a highly oscillatory behavior of the
aerodynamic loading $C_p^s(x)$ along the blade chord, observed in Figures
\ref{bvp40} and \ref{bvp41}, should be attributed to the sensitivity of
the
unsteady cascade response to the far field unsteady flow conditions. With
many pressure modes carrying away acoustic energy away from the blade
surface, the unsteady response of different parts of the blade may be
excited or damped depending on the local phase and amplitude of the
incident
acoustic-vorticity disturbance {\it and} resonance conditions for
scattered
acoustic waves. For the low-speed mean flow, such responses are mostly

damped since very few acoustic modes are excited in the duct; as a
result,
the unsteady aerodynamic loading concentrates at the leading edge as seen
in
Figures \ref{bvp12},\ref{bvp13}. In the present cases, the leading edge
response dominates only for small reduced frequencies. This also
corresponds
to the results for the 2-D cascade. In addition, since the impinging gust
in
a swirling flow exhibits large variations of axial and circumferential
phases of complex amplitudes over the span and chord of a blade, the
unsteady loading shows the corresponding pronounced variations in Figures
\ref{bvp40} and \ref{bvp41}.
If an incident non-axisymmetric acoustic-vorticity wave propagates with $
%
n_g=r_mk_g$, its circumferential phase remains almost constant (exactly
constant for a free vortex swirl) along the surface of a blade aligned
with
turning streamlines of the mean swirling flow. The unsteady response
predictions, however, are additionally affected by an axial gust
component,
and effects of cascade quasiperiodicity on blades interference and
inflow/outflow conditions.
Figure \ref{bvp42} compares the results for the unsteady lift predictions
in
the cases of swirling flows and 2-D cascade model. A significant
difference
from the corresponding low Mach number calculations (in Figure
\ref{bvp24})
is easily noticed. The resonance effect of cascade quasiperiodicity is
now
superimposed on the effects of multiple acoustic resonances in the duct.
As
before, the cases of potential and vortical swirling flows diverge for
low
reduced frequencies but show better agreement for higher $k_g.$
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/cls06ngnew.eps,height=3i
n,width=5in}}
\caption{Unsteady lift coefficient vs. reduced frequency. Present
results: $'+'$, $M_0=\bar{M}_{\Gamma}=0.6$; 'x',
$M_0=0.6, \bar{M}_{\Gamma}=0.3,\bar{M}_{\Omega}=0.3$.
2-D cascade analysis: 'o'. $\bar{\chi} =45^{o}, n_{g}=k_{g}r_{m}.$ }
\label{bvp42}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/eig_ug06ng.eps,height=3i
n,width=5in}}

\caption{Eigenvalues for upstream/downstream unsteady response.


$k_{g}=1,n_{g}=5 $ (a), $k_{g}=3, n_{g}=15$ (b), $k_{g}=5.8, n_{g}=29$
(c), $k_{g}=9, n_{g}=45$ (d); $M_0=0.6, \bar{M}_{\Gamma}=0.6,
\bar{\chi}=45^{o}.$ }
\label{bvp43}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/eig_ugo06ng.eps,height=3
in,width=5in}}
\caption{Eigenvalues for upstream/downstream unsteady response.
$k_{g}=1,n_{g}=5 $ (a), $k_{g}=3, n_{g}=15$ (b), $k_{g}=5.8, n_{g}=29$
(c), $k_{g}=9, n_{g}=45$ (d); $M_0=0.6,\bar{M}_{\Gamma}=0.3,
\bar{M}_{\Omega}=0.3, \bar{\chi}=45^{o}.$ }
\label{bvp44}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/cp_ug06ng.eps,height=3in
,width=5in}}
\caption{Aerodynamic response at blade strips: --,$s=0.25$;
-,$s=0.5$; -.-,$s=0.75$. $k_{g}=1, n_{g}=5$ (a), $k_{g}=3, n_{g}=15$ (b),
$k_{g}=5.8, n_{g}=29$ (c), $k_{g}=9, n_{g}=45$ (d). 2-D cascade analysis:
'o', $k_{1}=k_{2}$. $M_0=0.6, \bar{M}_{\Gamma}=0.6, \bar{\chi} =45^{o}. $
}
\label{bvp45}
\end{figure}
\begin{figure}
\
centerline{\psfig{figure=../Gvan/FORT/BLADES/OUT/cp_ugo06ng.eps,height=3i
n,width=5in}}
\caption{Aerodynamic response at blade strips: --,$s=0.25$;
-,$s=0.5$; -.-,$s=0.75$. $k_{g}=1, n_{g}=5$ (a), $k_{g}=3, n_{g}=15$ (b),
$k_{g}=5.8, n_{g}=29$ (c), $k_{g}=9, n_{g}=45$ (d). 2-D cascade analysis:
'o', $k_{1}=k_{2}$. $M_0=0.6, \bar{M}_{\Gamma}=0.3, \bar{M}_{\Omega}=0.3,
\bar{\chi} =45^{o}. $ }
\label{bvp46}
\end{figure}
The results of the eigenvalue analyses for selected upstream- and
downstream-propagating pressure modes are shown in Figures
\ref{bvp43}-\ref
{bvp44}. The differences between two cases of potential and vortical
swirling flows appear, at first look, as rather insignificant. Both
results
show fast cut-on of propagating (mostly upstream) acoustic modes at
higher
reduced frequencies, similar to the previous discussion of an
axisymmetric
incident wave. A more careful study reveals a different effect of Doppler
shift on the acoustic resonance conditions in the two cases of mean
swirl.
For certain $k_g$ and $\nu _m^{\pm }$, a new-born pair of propagating

acoustic modes cuts on for the potential swirl but yet does not appear
for
the vortical. On the other hand, a total pair of modes may propagate
upstream in one case but splits in both directions, in another. These
indicated differences may cause the divergence of unsteady loading
characteristics in the two cases of mean swirl.
The gust evolution calculations in both cases are very similar to the
corresponding low-speed results presented in Figures
\ref{bvp29}-\ref{bvp32}
and \ref{bvp33}-\ref{bvp36}, by the reasons given above. The effect of
different $\overline{M}_\Omega /\overline{M}_\Gamma $ ratio on the gust
pressure response is slightly more pronounced in the present example but
still is not very significant.
Similar to the previous case of an axisymmetric disturbance, the same
incident wave produced a drastically different distribution of unsteady
loading along the chords and span of the blades (Figures \ref{bvp45},
\ref
{bvp46}). In general, the responses occured to be close, in the shape of
amplitude oscillations, to the previous example, but the phases and
amplitudes are clearly different as one would expected. For the free
vortex
swirl in Figure \ref{bvp45}, a sudden change of loading oscillations at
the
mid-span strip is notable.

You might also like