A&A 421, 305-322 (2004)
DOI: 10.1051/0004-6361:20040121
N. S. Dzhalilov1,2,3 - J. Staude2
1 - Institute of Terrestrial Magnetism,
Ionosphere and Radio Wave Propagation of the Russian Academy of Sciences,
Troitsk City, Moscow Region, 142190 Russia
2 - Astrophysikalisches Institut Potsdam, Sonnenobservatorium Einsteinturm,
14473 Potsdam, Germany
3 - Shamakhi Astrophysical Observatory of the Azerbaidjan Academy of
Sciences, Baku 370096, Azerbaidjan Rep.
Received 15 May 2002 / Accepted 13 October 2003
Abstract
The general partial differential equation governing linear adiabatic nonradial
oscillations in a spherical, differentially and slowly rotating non-magnetic
star is derived. This equation describes mainly low-frequency and high-degree
g-modes, convective g-modes, rotational Rossby-like vorticity modes, and
their mutual interaction for arbitrarily given radial and latitudinal
gradients of the rotation rate. Applying to this equation the "traditional
approximation'' of geophysics results in a separation into radial- and
angular-dependent parts of the physical variables, each of which is described
by an ordinary differential equation.
The angular parts of the eigenfunctions are described by the Laplace tidal
equation generalized here to take into account differential rotation.
It is shown that there appears a critical latitude in the sphere where the
frequencies of eigenmodes coincide with the frequencies of inertial modes.
The resonant transformation of the modes into the inertial waves acts as a
resonant damping mechanism of the modes. Physically this mechanism is akin
to the Alfvén resonance damping mechanism for MHD waves. It applies
even if the rotation is rigid.
The exact solutions of the Laplace equation for low frequencies and rigid
rotation are obtained. The eigenfunctions are expressed by Jacobi polynomials
which are polynomials of higher order than the Legendre polynomials for
spherical harmonics. In this ideal case there exists only a retrograde wave
spectrum. The modes are subdivided into two branches: fast and slow modes.
The long fast waves carry energy opposite to the rotation direction, while
the shorter slow-mode group velocity is in the azimuthal plane along the
direction of rotation. It is shown that the slow modes are concentrated
around the equator, while the fast modes are concentrated around the poles.
The band of latitude where the mode energy is concentrated is narrow, and
the spatial location of these band depends on the wave numbers (l, m).
Key words: hydrodynamics - Sun: activity - Sun: interior - Sun: oscillations - Sun: rotation - stars: oscillations
In a recent paper Dzhalilov et al. (2002; Paper I) investigated which
lowest-frequency eigenoscillations can occur in the real Sun and what role
they might play in redistributing angular momentum and causing solar activity.
We found that such waves could only be Rossby-like vorticity modes induced
by differential rotation. However, the general nonradial pulsation theory
adopted from stellar rotation has some difficulties for such low-frequency oscillations. For slow rotation, when
the sphericity of the star is not violated seriously, the high-frequency
spherical p- and g-modes are no longer degenerate with respect
to the azimuthal number m (Unno et al. 1989).
Independent of the spherical modes, non-rotating toroidal flows (called
"trivial'' modes with a zero frequency) become quasi-toroidal with rotation
(called r-modes with a nonzero frequency; Ledoux 1951; Papaloizou &
Pringle 1978; Provost et al. 1981; Smeyers et al. 1981; Wolff 1998).
Although rotation removes the degeneracy of the modes, it also couples the
modes with the same azimuthal order, and this makes the problem more
difficult. For the high-frequency modes (Rossby number
1, where
and
are the angular
frequencies of oscillations and of stellar rotation, respectively) this
difficulty is resolved more or less successfully. For this case the small
perturbation rotation theory is applied: the eigenfunctions are
represented by power series, the angular parts of which are expressed by
spherical harmonic functions Yml (Unno et al. 1989). These power series
are well truncated, unless
,
when the role of the Coriolis force
increases.
Most pulsating stars point out the low-frequency instability (Cox 1980;
Unno et al. 1989). Rotation causes strong coupling of the high-order
and the convective g-modes, and the r-modes with
and
with the same m, but different l (Lee & Saio 1986). Generally the matrix
of the coupling coefficients to be determined is singular (e.g., Townsend
1997). In all papers on the eigenvalue problem of nonradially pulsating
stars, there exists a "truncation problem'' for the serial eigenfunctions, the
angular parts of which are represented by spherical harmonics (e.g., Lee &
Saio 1997; Clement 1998).
The governing partial differential equations (PDEs) of the eigenoscillations of rotating stars are complicated from the point of view of mathematical treatment, even if the motions are adiabatic. This difficulty arises because in a spherical geometry an eigenvalue problem with a singular boundary condition has to be solved. The singularity of the governing equations on the rotation axis is not a result of the applied physical approximations (small amplitude linearization, adiabaticity of the processes, incompressibility of motion, rigid rotation approximation, etc.), but is a property of the spherical geometry, permitting special classes of motion which may become the global eigenmotion of the star. Due to these difficulties it is very important to obtain some analytical solution of the eigenvalue problem with rotation, even if rather special approximations are used.
For
the Coriolis force is the dominant term in the equation
of motion. In this case this force causes the motion to be preferentially
horizontal,
.
With decreasing wave frequency,
,
the flows become practically horizontal,
.
Here vr is the radial component and
are the surface components of the velocity. This is why in many
investigations of global linear or nonlinear motions in differently rotating
astrophysical objects the radial velocity component and the radial changes
are ignored, i.e., vr=0,
.
That is, 2D flows are considered, e.g. Gilman & Fox (1997), Levin & Ushomirsky (2001), Cally (2001), Charbonneu et al. (1999). Nevertheless, even
such a 2D approximation describes the basic physics of the global r-modes
of rotating stars, e.g., Levin & Ushomirsky (2001).
However, real situations (many stars show nonradial pulsations
with rather low frequencies, e.g. Unno et al. 1989) require the inclusion of
.
It is important to study, e.g., the g-mode rotation
interaction, the transport of rotation angular momentum, the mixing problems
of the stellar interior, etc. The consideration of a small nonzero radial
component of velocity is the next important step for improving the 2D
approximation. Essentially, in geophysics the traditional approximation takes
into account
in the governing equations. This
approximation is also often applied to global motions in astrophysical
objects, for example, Bildstein et al. (1996), Lou (2000), Berthomieu et al. (1978), Lee & Saio (1997), Dziembowski & Kosovichev (1987).
To avoid misunderstanding, we will give here a short excursion into this
approximation; see, for example, Unno et al. (1989), Shore (1992).
To show the meaning of the traditional approximation clearly let us use a
local analysis procedure.
In the case of a homogeneous rotating fluid,
is
included into the wave equation only in the term
In the present work for the non-magnetic and non-convective cases we obtain
one PDE in a spherical geometry for the adiabatic pressure oscillations in
the differentially rotating star (
)
with arbitrary
spatial gradients of rotation (Sect. 2). This general equation is split into
- and r-component ODEs, if the traditional approximation is
applied (Sect. 3). The
-component equation is Laplace's tidal
equation generalized for the differentially rotating case. In Sect. 4 we
analyse this equation more qualitatively. We find the general condition
for the shear instability due to differential rotation in latitude.
We find that the smallest rotation gradient is responsible for the prograde
(seen in the rotating frame) vorticity wave instability, while a stronger
gradient causes the retrograde wave instability. For solar data (small
rotation gradients) the m=1 prograde mode instability is possible
(Sect. 4.4). The possible existence of such a global horizontal shear
instability on the Sun has been investigated by Watson (1981) and Gilman &
Fox (1997), that of shear and other dynamic instabilities and of thermal-type
instabilities in stars also by Knobloch & Spruit (1982) and others.
Laplace's tidal equation for low frequencies in the rigid-rotation case is
investigated in detail in Sect. 5. It is shown that the eigenfunctions are
defined by Jacobi polynomials which are of higher order than the
Legendre polynomials.
The motion of the fluid in a self-gravitating star, neglecting magnetic field
and viscosity, may be described in an inertial frame by the hydrodynamic
equations. These equations in conventional definitions are written as
We suppose that the equilibrium state (zero indices of the variables) of the
star is stationary and that its differential rotation is axially symmetric:
![]() |
(5) |
Small amplitude deviations from the basic state of the star may be
investigated by linearizing Eqs. (1)-(4). For Eulerian
perturbations (variables with a prime) the equation of motion becomes
(Unno et al. 1989)
Excluding
from Eq. (7) by using Eq. (12) and
taking its r and
components we get
![]() |
(17) |
![]() |
(18) |
Now adding to these equations the
component of Eqs. (7)
and (9) we get our set of equations:
For the general case the system of Eqs. (21)-(24) has been
reduced in Appendix A to one PDE for the pressure perturbation:
![]() |
(28) |
Let
![]() |
(29) | ||
![]() |
![]() |
|||
![]() |
|||
![]() |
|||
![]() |
|||
![]() |
|||
![]() |
|||
![]() |
|||
![]() |
|||
![]() |
|||
![]() |
|||
![]() |
|||
![]() |
|||
b4 = 1-2b1 + b2 , | |||
![]() |
|||
![]() |
|||
![]() |
|||
![]() |
|||
![]() |
|||
![]() |
|||
![]() |
We assume here an extra condition for the rotation rate:
.
Such a
representation of the variable part of the rotation rate is no strong
restriction. For the Sun, for example,
except for the
solar tachocline. Depending on the problem to be solved we can simplify the
calculations ignoring either
or
.
In that way we may separate the variables
Equation (32) is the generalized Laplace equation if differential rotation is
present,
.
For rigid rotation,
,
Eq. (32) becomes the standard Laplace equation:
In astrophysics the r-component Eq. (33) appears, and the two
equations must be solved together to find both spectral parameters. In the
rigid-rotation case Lee & Saio (1997) looked for
numerically in
an approach similar to that in geophysics, fixing
in
Eq. (34). Here we offer another approach, where we will find
from the r-equation for a given
.
It is convenient to introduce into Eq. (32) the new variable
:
From Eqs. (A.2), (A.3) and (A.5) we can derive in the traditional
approximation the following formulae for the components of the fluid velocity:
Now we will impose a restriction on
const, the logarithmic latitudinal
gradient of the rotation frequency. We might take the linear dependence
,
but for such a profile the structure of the solutions is
not changed qualitatively.
Really
const. corresponds to a rotation profile
Let us introduce the new dependent variable
In our case the physical equations are reduced to Heun's Eq. (47) for which the Riemannian relation (48) is satisfied. This means that the regular eigenfunctions in the hemisphere including the singular pole and the equator exist and we can find them. A direct numerical solution of this equation is practically impossible and we intend to use the P-branches to find the analytical eigenfunctions. However, the main difficulty of Heun's Eq. (47) is connected with the "accessory'' parameter h which is not included in the Riemannian scheme. The arbitrariness of h does not make it possible to represent the solutions by a single P-function. Therefore, hypergeometric and confluent hypergeometric equations, the differential equations of Lamé, Mathieu, Legendre, Bessel, and Weber, those of the polynomials of Jacobi, Chebyshev, Laguerre and Hermite as well as those of Bateman's k-functions are special or limiting cases of Heun's equation. Due to the parameter h the solutions of Heun's equation are usually searched for as series of P-functions. For example, Erdélyi (1942) has represented in such a series the P-branches by the hypergeometric functions. In physical tasks, e.g. the damping of MHD waves in resonant layers, the same equation arises (Dzhalilov & Zhugzhda 1990). The exact solutions of Eq. (47) will be considered in the next paper of this series. Here we will restrict ourselves to a qualitative analysis and to a limiting case.
Generally for low frequencies (
)
three singularities of Heun's
equation may be realized in a hemisphere: 0 < a < 1. Note that the
singularity
in the Riemann scheme does not occur in our problem.
At the equator (x=1), Eq. (47) has two solutions with the exponents 0
and
.
If we put the P-solutions with these exponents into
Eqs. (39)-(41), we can see that the first solution is divergent,
while the second solution provides the limited
and
at
the equator. In the same manner, at the pole (x=0) one of the solutions
with the exponents "0'' and (
)
may provide the limited
and
only for some selected values of
(see the next
subsection). In the construction of the global solution in the hemisphere,
the boundary conditions at the equator and at the pole should remove the
divergent solutions.
To provide regularity of the global solution at the middle singularity
(x=a) the two P-solutions with the exponents 0 and (
)
must be
regular. This is because every retained regular solution at the equator or
at the pole will, after its analytical continuation, be expressed as the
sum of these two independent solutions at x=a. The second exponent (at
x=a)
if
or if
.
If
then
.
This implies, that the second
independent solution of Eq. (47) with the exponent
is
regular at the singular point x=a for all variables, as follows from
Eqs. (39)-(41). For the first solution, however, with the
exponent of "0" at x=a the physical variables
and
are
limited only if a is complex, i.e.
is complex. In this way we
conclude that the solutions of the singular boundary value problem will be
complex and that the process of wave-rotation interaction produces a
mechanical instability.
In our initial adiabatic motion approximation for which the final
Eq. (47) is derived, the complex
should be related to one of
the following physical mechanisms:
1) Shear instability:
The latitudinal differential rotation with a gradient
can
favour the development of mechanical shear instability. This instability
is akin to the classical Rayleigh instability mechanism or to the
Kelvin-Helmholtz instability with gravity. The latitudinal instability cannot
be prevented by gravity in the horizontal direction. The development of
this instability will strongly depend on the values and the sign of
as well as on the wave frequency.
2) Resonant damping:
The existence of a critical latitude
may be the reason for
the damping of the eigenmodes. For the solar rotation the location of this
latitude in the hemisphere depends weakly on
,
see
Eq. (37). The critical latitudes exist even if the rotation is rigid at
the inertial frequency
.
The
appearance of the critical latitudes on the rotating spherical surface and
their role in the formation of a wave guide around the equator is a well
known phenomenon in geophysics (e.g., Longuet-Higgins 1965). A concentration
of the wave amplitude at the singular latitude is very similar to the
damping of adiabatic waves in MHD resonant layers (recall that the
nature of the Coriolis force is very similar to that of
the ponderomotive force). Physically the damping of MHD waves, for example,
in the resonant Alfvén layer where
,
means the
transformation of waves into Alfvén waves which are concentrated along
the magnetic field in the resonant layer; here
is the radial
dependence of the Alfvén velocity, kx is the wave number along the
magnetic field. This mechanism is an important damping mechanism of MHD waves; it has been used tentatively as a corona heating mechanism (e.g., Ionson
1978, 1984; Hollweg 1987).
In our case the Alfvén waves are changed to inertial oscillations.
The transformation of rotation-gravity waves to inertial waves in the narrow
range of critical latitudes acts as a resonant damping mechanism. The resonance
of the local inertial oscillations (frequency
)
with
the eigenmodes (frequency
)
occurs in those places where their
frequencies are very close to each other. For low frequency modes,
,
the resonant layer is located close to the equator, and
for higher frequency modes,
,
near the pole. It can be shown
that generally wave motion with a
frequency of
must be produced to ensure conservation of
potential vorticity in the rotating system.
The width of the critical layer will depend on the values of the imaginary
parts of the eigenfrequencies. Formally, if there exists a continuous
spectrum, then the critical layers may be very wide. In
reality, however, we may expect the formation of some "active''
latitude belts if only some selected discrete modes are excited. In these
belts the greatest part of the mode energy will be concentrated in the
inertial modes. The further interaction of these waves with rotation can
change the rotation velocity. The evolution of the produced resonant
inertial waves in their initial state should be investigated separately on
the base of our general PD Eq. (27). For the further devolopment another
dissipation effect as well as nonlinearity should be taken into
consideration. We are not aware of investigations of the resonant damping of
rotation waves through the critical latitudes in the physics of stellar
pulsations.
3) Tunneling of waves:
It is well known that the tunnel leakage of wave energy from a cavity leads
to a decrease of wave amplitudes in time, i.e. the tunnel effect can
work if the wave frequency is complex. This is very effective for
the p-modes in the solar atmosphere (Dzhalilov et al. 2000). Equation (35)
or Eq. (47) can easily be transformed into a two-terms form:
Z'' +
U(x) Z'=0. Due to the singular points the behavior of the wave potential U(x) is complicated, especially if
and
(we will omit here a detailed analysis of U(x)). However, a series of
changes of the sign of U(x) shows the existence of cavity (oscillation) and
tunnel (non-oscillation) zones in the area 0<x<1. Close to the equator at
for all cases U>0, and for most cases near the pole (
)
U<0. This means that the eigenmodes are formed in the cavity which is
located around the equator. Near the pole we must choose only a decreasing
solution. For example, in the rigid rotation case we have
.
However, for any
(differential
rotation) and complex
we will have a complex wave vector in the
-direction. This means that we will have runnig waves toward the
pole. As the amplitude of these waves is zero at the pole we have no
reflected waves from the pole. Thus we may conclude that the tunnel effect
can work throughout each hemisphere.
These questions are important, but a detailed discussion may be given only after Heun's equation has been solved.
Here we discuss two conditions: the regularity of the solutions at the pole
and the approximate instability condition for the modes.
If we put the P-solutions with the exponents 0 and
into
Eq. (43) we get
.
Then
means that for the regularity of the
solutions at the pole the condition
must be obeyed.
On the other hand, an instability is possible when the eigenfrequencies are
complex, implying a complex ,
as the wavenumber is
complex (for tunneling) if
is complex. Otherwise, the solution of
the eigenvalue problem gives a dispersion relation for
which
becomes complex if the parameter
is complex (of course, this is
one possibility). This is seen, for instance, directly from the
dispersion relation (68, see below) for the lower frequency limit.
Here |m| should be changed to
.
For the instability it
follows from Eq. (44), that the necessary condition is S2>0. It is
clear that the axially-symmetric mode with m=0 is excluded in this case.
For lower values of the rotation gradient
the necessary condition S2>0 demands
for the prograde waves (
)
the condition
,
which
is more realistic for stellar situations (equatorward spinning up at the
surface with radius r). Rayleigh's necessary condition for
instability (Rayleigh 1880; Watson 1981) says that the function
(gradient of
vorticity) must change its sign in the flow. Rewriting this function in our
definitions we get that
![]() |
(51) |
The sufficient condition for instability is obtained from Eq. (44)
and reads
.
The regularity condition at the pole
can be rewritten
as
![]() |
Figure 1:
The domains of validity of the solution regularity and of
the instability conditions in the phase space
![]() ![]() ![]() ![]() ![]() ![]() ![]() |
Open with DEXTER |
For
(strong gradients) the condition Eq. (52) is
met only for retrograde waves in the range
(region II). For
we get for any
that
.
This is the line between regions II and III.
All three regions are shown in Fig. 1a. It is seen that the regularity
condition is working for
and
if
.
For very small
only modes with large mare possible. The smallest
modes may appear in the limit
.
These conclusions are correct only if the
instability occurs.
Now let us consider the second condition, the complex frequency condition
.
This inequality may be rewritten as
The total condition for the existence of spatially stable but temporally
unstable waves reads as follows:
![]() |
(56) |
![]() |
Figure 2:
The wave instability areas (hatched) in the phase space from an overlap of Figs. 1a and b
and from the condition
![]() ![]() ![]() ![]() ![]() |
Open with DEXTER |
Figures 2a,b show the validity ranges of this condition for some
typical values of
.
The hatched areas are places
where mode instability is possible. These figures are obtained
by overlapping Figs. 1a and b. The hatched areas of prograde
waves are strongly restricted also by the condition
because the integer
.
For prograde
waves on both sides these hatched areas become very narrow: with
decreasing
the extent of the hatched area decreases
and tends to the point
.
We will see
that this is the solar case. The hatched instability areas
disappear with decreasing
.
This means we have a
lower limit
.
For retrograde waves it is sufficient to write the condition as
.
In Fig. 2b
is the dash-dot
curve and
is the solid curve. Instability is possible if
for
.
In Fig. 2 only the hatched areas for given
are ranges
of possible solutions, if wave instability occurs. Outside these hatched
areas regular solutions are impossible. The case without instability (neutral
oscillations) must be investigated separately.
Let us consider at which places we might expect mode instability in the Sun.
Unfortunately, it is not clear how the core rotates. Nevertheless some
rotation gradients close to the Sun's centre might exist, and we could expect
mode instability there. It is known from helioseismology that the radiative interior
has a very small
,
but the exact value is unknown. We have
better information on the rotation profile of the solar envelope, including
the tachocline. Helioseismology data may be described by different
approximate formulae. One of these is (Charbonneau et al. 1998)
Using this formula we show in Fig. 3 the
dependence for different
.
We see that in the Sun
,
and the maximum is in the photosphere.
Figure 2 implies that unstable retrograde waves are not present
in the solar case. Instability of prograde waves in the Sun occurs
in the upper right corner of Fig. 2a which is enlarged in Fig. 4.
Here we see that the instability area disappears when
10-4. This boundary is located at the bold
horizontal line in Fig. 3. Thus the prograde waves become unstable in the
Sun in those places where 3
10-4
0.15.
This means that instability is possible in the area
which includes the greater part of the
tachocline, the convective zone, and the photosphere. With
increasing r the instability zone expands from middle to high
latitudes. Figures 2a and 4 show that the instability occurs at high
frequencies (
)
and on global scales
(
). Considering that
is an integer we
get m=1.
However, our instability analysis is based on the general Riemann scheme of
Heun's equation, which is valid only if the middle singular point x=a is
far from the other edges at x=0 and x=1. Thus, the limiting cases
and
(the latter is more
important for the solar case) should be considered separately. In these
limiting cases the regularity condition Eq. (48) may be changed, and the
curve in Fig. 4 limiting the instability areas from below may be shifted. In
this case instability with higher m-modes should be expected.
![]() |
Figure 3:
The local estimate of the logarithmic gradient of the solar
rotation frequency
![]() ![]() ![]() ![]() |
Open with DEXTER |
![]() |
Figure 4:
Enlarged part of Fig. 2a for the
smallest gradients of rotation. The labels 1, 2,..., 6 correspond
to
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
Open with DEXTER |
After the qualitative analysis of Heun's Eq. (47) we can start a
quantitive analysis. Note that the qualitative conclusions drawn above are
valid for the more general Eq. (35) with the term. Heun's
equation with four singularities in the general case is solved by a series
of hypergeometric Gauss functions. A similar task has been considered for the
damping of MHD waves at resonance levels by Dzhalilov & Zhugzhda (1990). We
will start to study Eq. (47) for some simple limiting cases. At
high frequencies (
,
where shear instability is
acting in the Sun) and at low frequencies (
,
where the
waves are stable against shear instability in the Sun) Heun's equation is
much simpler. In these cases the singular level x=a is shifted
either to the pole or to the equator. For both cases the solutions are
expressed by one hypergeometric function.
In the present work we consider particularly the second case. Let
.
Then we have
and
.
Equation (47) is now the hypergeometric equation:
![]() |
(59) |
![]() |
(61) |
Using the conditions
and
the
parameters in the solution Eq. (60) are greatly simplified. Because
only a regular solution at the pole (x=0) will be left.
In the standard definitions of hypergeometric functions (Abramowitz & Stegun
1984) we have
The new dispersion relation Eq. (69) completely differs from the
dispersion relation of the almost toroidal r-modes. Their dispersion relation
can be derived from Eq. (69) if we formally set
.
Then
![]() |
Figure 5:
The spectrum
![]() ![]() |
Open with DEXTER |
![]() |
Figure 6:
The normalized group velocity
![]() |
Open with DEXTER |
![]() |
Figure 7:
The possible domain of the existence of eigenmodes for given
Rossby numbers
![]() ![]() ![]() ![]() |
Open with DEXTER |
From this we get an upper limit for l if
is given:
.
As
we get an approximate condition for the existence of
the modes:
.
For such degrees of l the azimuthal
numbers are also limited:
.
For
we have
,
,
and
.
However,
considering the regularity of the solutions Eq. (66), we must take
.
m1=1 if
.
For
we have m1=m2. This situation is shown in Fig. 7 for different values of
.
A decrease of the frequency decreases the domain of existence of the modes.
Putting Eq. (73) into Eqs. (71), (72) gives
![]() |
Figure 8:
The phase velocities of fast (solid) and slow (dashed curves)
modes versus l. The numbers at the curves are
![]() ![]() ![]() |
Open with DEXTER |
Here
is the phase velocity of the fast modes and
that of the slow modes. For
we have
and
.
In Eq. (73) and in Fig. 7
the m1 branch corresponds to the fast mode, but m2 to the slow modes,
since
.
In Fig. 8 the normalized phase velocities (with inverse
sign) for the selected values of
in Fig. 7 are shown versus l.
Both branches are retrograde modes (
). Using
km s-1 for the Sun, we get from Fig. 8 very slow phase
velocities. The fast wave velocity (solid lines) depends more strongly on l. With increasing
both branches are accelerated.
In Fig. 9 the group velocities are presented in the same way. For fast waves
the group velocity is always parallel to the phase velocity (
), while for the slow waves we have the opposite behavior
.
Slow mode packets carry off energy in the rotation direction.
is valid everywhere. With deceasing
the range
shifts to the right, and it is
seen in Fig. 9 that
for such low
is almost zero.
Note that m=l modes are always fast modes.
![]() |
Figure 9:
Absolute values of the group velocities of fast (solid) and slow
(dashed) modes normalized to
![]() ![]() ![]() ![]() ![]() ![]() |
Open with DEXTER |
Taking into account the quantization condition Eq. (68) in the
solutions Eqs. (63), (67), (39), and (41) we obtain
the eigenfunctions. Turning from complex velocities into the real
displacements,
(remember that
is the
velocity seen in the rotating frame), we get
![]() |
(88) |
![]() |
Figure 10:
The amplitudes of eigenfunctions versus co-latitude for given
pairs (l,|m|), Eqs. (83)-(85). The first row shows the
![]() ![]() ![]() ![]() |
Open with DEXTER |
Table 1:
Results for the 22-year period: frequency deviations
for permitted quantum numbers (l, |m|).
Let us consider some particular cases.
![]() |
(89) | ||
![]() |
(90) | ||
![]() |
(91) |
In Fig. 10 for some typical selected pairs of (l,m) the amplitude
functions, Eqs. (83)-(85), are shown as functions of .
The first row is the pressure function
function normalized to
its maximum. The first l=1 figure represents
for different |m|(|m| increases from left to right). As the eigenfunctions are multiplied
by a
factor, the amplitudes are strongly suppressed around
the pole. Increasing |m| for a given l shifts the maxima toward the
equator. l is the surface node number of the
function.
On the contrary, increasing l for a given |m| (the second figure of the
first row with m=1 in Fig. 10) suppresses the amplitudes around the equator,
and the maximum is shifted toward the pole. l=|m| is the equilibrium case.
In the third figure of the first row the equilibrium latitude with maximum
amplitude is defined by
for all l=|m| modes.
The second and third rows of Fig. 10 are the latitudinal (
)
and azimuthal (
)
eigenfunction amplitudes,
respectively, normalized to the maximum of
,
see
Eqs. (81) and (82). The latitudinal amplitude behaviour is
similar to that of the pressure. The azimuthal amplitudes are smaller than the
latitudinal amplitudes, but with a change of l a redistribution of the
amplitudes will not take place.
has practically the same
amplitude at all latitudes and for all l.
Figure 10 implies that we can expect an interesting behaviour of the eigenfunction amplitude, when both l and |m| are large. Suppression from two sides may give rise to a concentration of wave energy in narrow latitude bands. For example, this is the case for the 22-year solar mode.
For the 22-year modes we take
(1.441 nHz), for
which
.
Then we derive from the equations after
Eq. (73) the limiting values of the integer l,
.
For all lin this range we find from Eq. (73), rounding off, integer azimuthal
numbers m1 and m2 for the fast and slow modes, respectively. Putting
these integer numbers into Eq. (69) we find the deviation
from the central frequency due to the integer
azimuthal numbers.
Additional slow modes are also possible in the interval l=0-10.
The results are given in Table 1 in Appendix A. It is seen that the fast
modes with low l have larger deviations. This table includes all possible
(l,|m|) pairs which correspond to the 22-year period. For some example
pairs of (l,|m|) we plot in Fig. 11 the latitude dependence of the
quantity
,
averaged over the wave
period, which characterizes the energy density of the modes. The hemisphere
is divided into two equal parts: slow modes are located around the equator
(solid lines), fast modes are concentrated around the pole. Each (l,|m|)
pair is located in a narrow latitude band. Note that the slow modes (the
group velocity of which is in the rotation direction) with sunspot-like
spatial scales are at latitudes of
.
The eigenfunctions Eqs. (76)-(79) allow us to discuss the
flow character produced by the waves, even if the solution of the radial
equation Q(r) is unknown. Excluding from these equations the time-dependent
phase we can obtain the trajectory equations of the fluid elements. In the
meridional
plane we have
![]() |
Figure 11: The normalized energy density of the 22-year fast (dashed) and slow (solid) modes versus co-latitude. The numbers on the curves are (l,|m|) pairs taken from Table 1. |
Open with DEXTER |
At the surface of the cone over
the trajectory of each
fluid element is an ellipse around the equilibrium point. The cone
displacement equation is
![]() |
(94) |
On the surface of any sphere with a radius r the motion of the fluid
elements is on trajectories of the ellipse
![]() |
(95) |
![]() |
(96) |
![]() |
(97) |
In the present work we have derived the general PDE governing
non-radial adiabatic long-period (with respect to the rotation
period) linear oscillations of a slowly and differentially
rotating star. This general equation includes all high-order
g-modes and all possible hybrids of rotation modes as well as
their mutual interaction. The main objective of the present paper
was to obtain an analytic solution of this equation for the extreme
low-frequency limit and to show that the solutions cannot be
expressed by the associated Legendre functions. Thus, in the
low-frequency limit the pulsation theory of stars meets serious
difficulities. For this purpose we used some simplifying
approximations. The geophysical "traditional approximation''
considerably simplifies this general equation, and we get two ODEs
for the r- and -components instead of one with arbitrary
gradients of rotation
.
We have obtained a more
stringent condition for the applicability of this approximation to
the pulsation of stars. Only for very low frequencies is this
restriction the same as that for the standard case. We also imposed
some restrictions on the rotation profile
used. All of these restrictions have been compared with
empirical solar rotation profiles and, thus, we justified their use
in our modelling.
The -equation is Laplace's equation generalized to a latitudinal
differential rotation. Without solving this equation we found qualitatively
the condition for the appearance of a global instability. This
instability is driven by the latitudinal shear of rotation.
It is not influenced by buoyancy, unlike the Kelvin-Helmholtz instability,
and therefore it may easily be realized in stars, even if the horizontal
gradient of rotation is small.
We have shown the possible appearance of critical latitudes at which
there occurs a resonant interaction of eigenmodes with the inertial modes
for a frequency
.
The transformation of modes
with a frequency
in narrow latitude belts acts as a
resonant damping mechanism. With decreasing mode frequency these belts are
shifted toward the equator. This mechanism may play an important role
in the redistribution of rotation angular momentum.
The appearance of mode instability strongly depends on the Rossby number, on the azimuthal wave numbers, and on the latitudinal rotation gradients. Very large gradients produce retrograde waves (seen in the rotating frame), while a slower rotation gradient is responsible for the prograde mode instability. The rotation gradient has a lower boundary below which an instability of the modes is impossible for any Rossby number or azimuthal number m.
We have applied the instability condition to helioseismological data. Here a global instability is possible for the m=1 mode at practically all latitudes. Radially the instability zone extends from the greater part of the tachocline up to the photosphere. The shear instability for the Sun was first obtained by Watson (1981). According to his results the instability is possible only in the photospheric layers. Later Gilman & Fox (1997) showed that such an instability is possible in the tachocline too, if strong toroidal magnetic fields are included. Our results show that the instability of the m=1 modes and other modes (m>1) are possible without magnetic fields, in contradiction to Gilman & Fox (1997; see also Charbonneau et al. 1999). This difference is probably connected with the incompleteness of the equations used by Watson (1981) and by Gilman & Fox (1997); their equations are two-dimensional only.
The exact solutions of Laplace's tidal equation for lower frequencies are expressed by Jacobi's polynomials. Just for lower frequencies the numerical calculations of stellar pulsation analysis meet great problems, when one is looking for the eigenfunctions as infinite series of Legendre functions. The eigenfunctions, defined by higher-order Jacobi polynomials, cannot be expressed by convergent series of associated Legendre functions. Every Legendre function is a particular case of a Jacobi polynomial.
It has been shown here that the retrograde (slow and fast) modes with high
surface wave numbers (l,m) are energetically concentrated in narrow
latitude bands. This analysis was done for the 22-year modes as an example.
Such a concentration of mode energy in a narrow spatial area makes such modes
vulnerable to different instability mechanisms such as the
-mechanism
considered in Paper 1.
All of our results have been obtained in the traditional approximation.
This approximation does not work for waves propagating strictly within the
equatorial plane. In our work this case is excluded. Motions which were
considered in our paper can never cross the equator, and our boundary
condition is
when
.
In the paper we
derived all the solutions which obey the traditional condition: high (l,m) - short waves and
.
For the independent variable
and the definitions
![]() |
|||
![]() |
(A.1) | ||
![]() |
|||
![]() |
![]() |
|||
![]() |
|||
![]() |
(A.7) | ||
![]() |
|||
![]() |
|||
![]() |
![]() |
(A.9) |
Acknowledgements
The authors gratefully acknowledge the critical and very different, but constructive comments by two anonymous referees who helped to improve the paper. Moreover, we would like to thank George Isaak and Jet Katgert for careful reviews of the paper and suggestions for improving the language. The present work has been financially supported by the German Science Foundation (DFG) under grants Nos. 436 RUS 113/560/4-1 and 436 RUS 113/689/2-1 and by the Russian Foundation of Basic Research under RFBR No. 04-02-16386.