A publishing partnership

The following article is Free article

THE ROLE OF MODE MIXING IN THE ABSORPTION OF p-MODES

, , and

Published 2009 March 25 © 2009. The American Astronomical Society. All rights reserved.
, , Citation M. Gordovskyy et al 2009 ApJ 694 1602 DOI 10.1088/0004-637X/694/2/1602

0004-637X/694/2/1602

ABSTRACT

Observations show that a p-mode may lose up to 70% of its energy flux when it interacts with a sunspot. Part of the absorbed energy is assumed to be converted into other types of waves, while part of it is re-emitted into modes with different radial orders n. In the present paper, we investigate absorption of p-modes with the azimuthal order m = 0 due to their interaction with magnetic flux tubes and attempt to determine the role of mode mixing in this phenomenon. We consider the linearized magnetohydrodynamic equations in two-dimensional, cylindrical geometry, with all the model parameters depending only on radius r and depth z. It is assumed that the wave field may be decomposed into incoming and outgoing components that separately satisfy the governing equations. These components are calculated numerically using a second-order Runge–Kutta finite difference scheme. The calculations reveal substantial scattering from higher-to-lower radial orders n, predominantly into the f-mode (n = 0). Only weak scattering occurs from lower-to-higher radial orders. At the same time, the amount of energy transferred from the p-modes to the f-mode can account for 25%–30% of the energy lost by an incoming p-mode.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Observations show that p-modes interacting with sunspots and other magnetic field concentrations in the solar atmosphere may lose up to 70% of their energy flux (Braun et al. 1987; Braun 1995; Braun & Birch 2008). Various mechanisms have been proposed to explain the effect. Hollweg (1988) suggested that resonant absorption may occur when the sunspot boundary has a thickness comparable to the p-mode's wavelength. It was found that such an effect may cause noticeable loss of power but cannot fully explain the observed absorption coefficients. Later Rosenthal (1990) suggested that the fibril structure of the magnetic field may lead to a substantial increase in resonant absorption.

Another possible mechanism behind the absorption of p-modes is the conversion of acoustic waves into fast and slow magnetoacoustic waves in a thin layer where the local sound speed equals the local Alfvén speed (Spruit & Bogdan 1992; Cally & Bogdan 1993). Crouch & Cally (2003) studied the effect of magnetic field inclination on mode conversion and calculated eigenfunctions and eigenvalues for a polytropic atmosphere with an embedded uniform magnetic field. Based on these results, Cally et al. (2003) considered the interaction between p-modes and flux tubes with a piecewise-uniform configuration of the magnetic field and showed that such a model can quantitatively explain the absorption coefficients measured using Fourier–Hankel decomposition.

Mode mixing is another mechanism that can contribute to the absorption of p-modes. In this process, a p-mode interacts with an inhomogeneity, losing part of its energy by scattering into a mode with the same temporal frequency but different radial order n. Although there is observational evidence and theoretical arguments supporting this idea (e.g., Braun 1995; Barnes & Cally 2000), this mechanism has not been thoroughly investigated as a possible explanation for observed p-mode absorption. Obviously, in isolation, mode mixing does not result in conversion of acoustic energy into other forms of energy, i.e., the total energy flux carried by the outgoing p-mode packet is the same as that of the incoming p-mode packet. However, the observed power of individual outgoing p-modes can be lower than the power of corresponding incoming modes.

Gordovskyy & Jain (2008) considered the absorption of p2-modes by a magnetic flux tube through the solution of the linearized magnetohydrodynamic (MHD) equations for harmonic oscillations in a two-dimensional, cylindrical model. They compared the power of the incoming and outgoing wave components for a given degree l. In the present paper, we utilize a similar approach and investigate what happens to the power lost by the incoming wave through an analysis of the outgoing wave spectra.

2. PROBLEM DESCRIPTION

2.1. Magnetohydrostatic Model

We consider wave propagation within a two-dimensional model with cylindrical geometry, in which the ambient medium is axisymmetric and depends only on the radial distance r and depth z (Figure 1). In the absence of the magnetic field, the pressure and density in the model grow with depth z as

Equation (1)

Equation (2)

where z increases downward, L is a characteristic scale length, p0 is a characteristic pressure, g is the constant gravitational acceleration, and the parameter q is the polytropic index, which is related to the specific heat ratio as γ = (q + 1)/q.

Figure 1. Refer to the following caption and surrounding text.

Figure 1. Considered model scheme. The vertical magnetic field (dashed lines) concentrated at the r = 0 axis causes local deviation of the pressure and density from their unperturbed values. All of the atmospheric, magnetic, and wave field parameters lack dependence on the angle ϕ.

Standard image High-resolution image

In the present calculations, we adopt a model (see Figure 2(a)) with the characteristic pressure p0 = 1.8 × 104 Pa, hydrodynamic scale height L = 0.5 Mm, polytropic index q = 2.5, and gravity g = 2.7 × 102 m s−2.

Figure 2. Refer to the following caption and surrounding text.

Figure 2. Panel (a): sound speed vs. depth z in the unperturbed plane–parallel atmosphere (rR) (solid line), at r = R (dashed line) and at the axis r = 0 (dot-dashed line) in the flux tube with the magnetic field B0 = 1.2 kG. Panel (b): the upper turning depths (circles) and lower turning depths (squares) vs. frequency ν in the undisturbed atmosphere with the adopted pressure and density profiles (Section 2.1). Dashed lines are for n = 0 and solid lines are for n = 2.

Standard image High-resolution image

The magnetic field is purely vertical (B = {0, 0, Bz(r)}) and in the form of a tube with field strength, varying with radius as $B_z(r)=B_0 e^{-r^2/R^2}$, where B0 is the characteristic magnetic field strength, and R is the characteristic radius of the flux tube. In the presence of such a magnetic field, the gas pressure may be determined to be

Equation (3)

Since the density inside a purely vertical flux tube is only a function of height (see, e.g., Gordovskyy & Jain 2007), the sound speed is given by

Equation (4)

2.2. The Governing Equations

Let us consider linear harmonic waves in terms of complex pressure and velocity perturbations, p'(r, z) and v'(r, z), respectively, which vary with time as eiωt and do not depend on the angle ϕ (i.e., all the considered modes have azimuthal order m = 0). Their propagation through a magnetized nonuniform plasma may be described by the following set of linearized MHD equations (see, e.g., Gordovskyy & Jain 2008):

Equation (5)

Equation (6)

Equation (7)

Here, T is a vector representing the magnetic forces:

Equation (8)

Let us introduce new dimensionless variables as follows: r = LS and z = LH for radius and depth, respectively; $\omega = \sqrt{\frac{g}{L}}\, \Omega$ for frequency; $v_r^{\prime } = i V_r e^{i\omega t}\,\sqrt{gL}$, $v_z^{\prime } = i V_ze^{i\omega t}\,\sqrt{gL}$, c2 = C2gL for velocity components and sound speed, respectively; p' = p0Qeiωt and p = p0f for pressures; $\rho = \frac{p_0}{gL}\, w$ for density; $B_r = \sqrt{\mu p_0} b_r$ and $B_z = \sqrt{\mu p_0} b_z$ for magnetic field components. Substitution of these variables into Equations (5)–(8) yields three equations for dimensionless wave perturbations weighted by the nondimensional radial distance S:

Equation (9)

Equation (10)

Equation (11)

with the magnetic force term given by

Equation (12)

We assume that the wave perturbations SQ, SVr, and SVz can be decomposed into incoming and outgoing components, and that both components satisfy Equations (9)–(12). Far from the flux tube, at S = Smax , the pressure and velocity perturbations for the incoming component correspond to those within the nonmagnetic plane–parallel atmosphere. This provides a boundary condition for SQin, SVr in, and SVz in:

Equation (13)

Equation (14)

Equation (15)

where M is the Kummer function. Here, l is the p-mode degree that defines the horizontal wavenumber $k_h = \sqrt{l(l+1)/R^2_{\odot }}$, and ξ is a dimensionless parameter, ξ = L/R.

It can be seen that the first term on the right-hand side of Equation (9) asymptotically grows when the radial distance S tends to zero. Hence, the wave function cannot be calculated in the vicinity of the axis S = 0. Therefore, we introduce a boundary condition very close to the axis at r = rmin . This does not significantly affect the solution as far as the distance Smin  = rmin /L is much smaller than the considered radius of the flux tubes and the considered p-mode wavelength. In the present calculations, the distance Smin  is chosen to be 0.2 (i.e., rmin  = 0.1 Mm). Hence, for the smallest considered flux tube radius of 2 Mm the ratio R/rmin  = 20, while for the shortest considered wavelength of ∼3.1 Mm (n = 0, ν = 4 mHz) the ratio is λ/rmin  = 33.

Thus, at the distance S = Smin , the following conditions apply:

Equation (16)

Equation (17)

Equation (18)

Equation (19)

The upper and lower boundaries of the numerical domain are chosen such that the upper and the lower turning points of the considered p-modes are well within the computational domain (see Figure 2(b)); thus, we can set these boundaries as open-type boundaries.

2.3. Method of Solution and the Parameter Set

The set of Equations (9)–(12) together with the boundary conditions (Equations (13)–(19)) is solved using a finite difference approach. A second-order Runge–Kutta scheme is used to integrate the equations in the radial direction, marching from S = Smax  toward S = Smin  for the incoming wave component and marching from S = Smin  toward S = Smax  for the outgoing wave component.

The step in the radial direction is δS = 0.002, and the step in the vertical direction is δH = 0.2. The radial distance S ranges from Smin  = 0.2 to Smax  ≈ 205, while the depth H ranges from Hmin  = −0.6 to Hmax  ≈ 205. Hence, the considered numerical domain has physical dimensions of ∼ 100 × 100 Mm.

The flux tube's characteristic magnetic field B0 is set to be 1.4 kG, and the flux tube's radius is either 2 or 8 Mm.

We carry out calculations for different sets of incoming wave frequency ν = 1–4 mHz and radial order n0 = 0–2. Throughout the paper, n0 denotes the radial order of the incoming mode. The corresponding harmonic degrees may be calculated using the following expression (Lamb 1932):

Equation (20)

The p-mode degrees corresponding to our chosen sets of frequency ν and order n are shown in Table 1.

Table 1. p-Mode Degrees l for Different Sets of Frequency ν and Order n for a Complete Polytrope with Index q = 2.5.

  Order n
ν, mHz 0 1 2 3 4
1 100 55 43 29 24
2 400 222 154 118 95
3 900 500 346 265 214
4 1600 889 615 471 380

Download table as:  ASCIITypeset image

The open upper and lower boundaries of the domain are implemented using asymmetrical schemes for the appropriate derivatives. Thus, the vertical derivatives $\frac{\partial f}{\partial H}$ for the internal (I =  2, ..., N − 2 ) grid points are evaluated using the fourth-order symmetrical schemes (known as CD4):

Equation (21)

while for the near-boundary grid points, the asymmetrical scheme is used

Here, I is the grid number (I = 0, ..., N), and δH is the grid step in the vertical direction. Therefore, no external grid points are required.

Physically, these boundary conditions mean that waves can leave the domain through upper and lower boundaries without reflection.

3. RESULTS AND DISCUSSION

3.1. Wave Functions of Outgoing Components

Figures 3 and 4 show typical wave perturbations (or wave functions) for the dimensionless pressure, calculated using the approach described in Section 2. From these figures, it can be seen, for the large degree l, that the amplitude of the outgoing component is markedly lower in comparison to the amplitude of the corresponding incoming component.

Figure 3. Refer to the following caption and surrounding text.

Figure 3. Dimensionless pressure wave perturbations Q (wave functions, thereafter) weighted by the radial distance S at the depth H = 4 vs. the radial distance S(=r/L). The magnetic flux tube axis corresponds to S = 0. The flux tube characteristic induction and radius are B0 = 1.2 kG and R = 8 Mm (corresponding to 16 in nondimensional units). The p-mode frequency is ν = 2 mHz, panel (a) is for an incident radial order of n0 = 0 (corresponding to degree l0 = 400) and panel (b) is for n0 = 2 (l0 = 154). Dashed lines denote the corresponding incoming component.

Standard image High-resolution image
Figure 4. Refer to the following caption and surrounding text.

Figure 4. The same as in Figure 3 but for the frequency ν = 3 mHz. (Here n0 = 0 corresponds to the degree l0 = 900 and n0 = 2 corresponds to l0 = 346.)

Standard image High-resolution image

This indicates that the mode lost a portion of its energy as it traversed the magnetic inhomogeneity. This finding is qualitatively consistent with observations and with our previous work (Gordovskyy & Jain 2008). Furthermore, most of the wave functions exhibit modes with nn0 in the outgoing component.

Figures 3(a) and 4(a) for n0 = 0 show that the initial oscillations are slightly modulated by a longer wavelength. However, in the case of n0 = 0, the additional mode order has a much weaker amplitude than in n0 = 2 case: the amplitude of the mode-mixed wave is about 10–20 times smaller than the amplitude of the incoming mode.

Figures 3(b) and 4(b) for n0 = 2 clearly show that the solution is a combination of the incident wavelength with oscillations of a shorter wavelength (i.e., a wave with lower order n). Thus, in the case of ν = 2 mHz and n0 = 2, the amplitude of the overtone is about 10% of the amplitude of the incoming mode. In the case of ν = 3 mHz and n0 = 2, the amplitude of the overtone is approximately the same as that of the n = n0 mode.

In order to study the composition of the outgoing component, we calculate power spectra using the Fourier–Hankel analysis described in Section 3.2.

3.2. Power Spectra of Outgoing Waves

We analyze the calculated wave functions using an approach analogous to the decomposition method used in helioseismic observations (e.g., Braun 1995). The dimensionless pressure wave perturbation Q is decomposed into a Hankel series for m = 0 at some fixed depth H1:

Equation (22)

Here, $\mathcal {H}^{(I)}_0$ and $\mathcal {H}^{(II)}_0$ are zeroth-order Hankel functions of the first and second kinds, respectively, and Al and Bl are complex amplitudes of the incoming and outgoing wave components, respectively. Therefore, |Al|2 and |Bl|2 are powers of the incoming and outgoing components which are proportional to the energy fluxes carried by these wave components.

The decomposition is done outside the magnetized region between S = 48 (corresponding to r ≈ 24 Mm) and Smax .

The resolution of these spectra in l depends on the width (or radius) of the domain as δl ≈ πR/rmax . For the domain with radius rmax  = 102 Mm, resolution in l is about 40. Hence, for the frequency of 2 mHz, peaks corresponding to a different radial order may be resolved up to n = 3 and up to n = 4 for ν = 3 mHz.

Due to the finite number of harmonics used in the discrete Fourier–Hankel transform, any δ-like peak in the power spectrum |Bl|2 is distributed among neighboring mesh points (the spectral leakage effect) and, as a result, the amplitude of the peak is reduced. In order to achieve the correct power values, we use a power-correction procedure that selects peaks higher than 4% of the maximum value and sums power over the five nearest points: the central point and two points on either side.

Power spectra versus degree l for the incoming and outgoing wave components (|Al|2 and |Bl|2, respectively) are shown in Figures 5 for a few sets of parameters. It is evident that some of the spectra reveal a number of easily resolved peaks. Comparison of the peak locations in degree l with those in Table 1 show that all of them correspond to modes with a different radial order for the same frequency. This suggests that scattering across the radial order n occurs when the incident wave interacts with the magnetic flux tube.

Figure 5. Refer to the following caption and surrounding text.

Figure 5. Power spectra of the wave function vs. degree l obtained using Fourier–Hankel decomposition. The frequency is ν = 3 mHz. Panels (a) and (b) are for the incident radial order n0 = 0; panels (c) and (d) are for n0 = 2. Incoming and outgoing component spectra are shown by the dashed and solid lines, respectively. Panels (a) and (c) are for the flux tube radius R = 2 Mm; and panels (b) and (d) are for R = 8 Mm.

Standard image High-resolution image

The power-correction procedure gives good results if the distance between two peaks is more than 100–120, otherwise the powers may be substantially overestimated. Therefore, this method may not provide correct powers for peaks corresponding to high orders n.

In order to avoid this problem, we use an alternative approach by fitting the obtained wave functions with sets of Whittaker functions, which are the vertical part of the eigenfunction for the case of a nonmagnetized atmosphere:

Equation (23)

where Φ(n, k(n), z) = zqekzM(−n, q, 2kz). Here, the function k(n) taken for a complete polytrope with a constant polytropic index and for a constant frequency. The fitting is done using the following two steps. At first, we use the least-squares method to fit the pressure wave perturbation of the outgoing wave p'out(r1, z) with the series of Φ(n, k, z) functions as $p^{\prime }_{\rm out}(r_1,z) = \sum \limits _n \tilde{b}(n,r_1) \Phi (n,k,z)$ for a wide range of the radial distance r1 outside the magnetized region (S > 48). In the numerical domain with zmax  ≈ 100 Mm the spectra can be easily resolved up to n = 10. The obtained complex coefficients are functions of radius: $\tilde{b}(n,r) = b(n) \mathcal {H}^{(II)}_0(kr)$. Second, we calculate the absolute values of these complex coefficients $|\tilde{b}|^2(n,r) = |b|^2(n) |\mathcal {H}^{(II)}_0|^2(kr)$. Since $|\mathcal {H}^{(II)}_0|^2(kr) \sim \frac{2}{\pi k r}$ for large kr, we can easily "eliminate" the Hankel function and, as a result, we deduce powers |b|2(n) for each mode in the outgoing wave function.

In order to estimate the error of this method, we use wave functions calculated for plane–parallel nonmagnetic atmosphere. In this case, obviously, the amplitudes of the outgoing modes should be |b|2(n0) = |a|2(n0) and |b|2(nn0) = 0. The characteristic error in the power |b|2 is about 0.5%–0.8%. It should be noted that this error contains the error of the method as well as the error of the numerical scheme.

We now calculate absorption coefficients defined as α(n) = (|a(n)|2 − |b(n)|2)/|a(n0)|2, where a(n) and b(n) are the amplitudes of the incoming and outgoing wave of the mode n, respectively. It should be noted that α(n) represents absorption of an individual p-mode and includes "real" absorption as well as redistribution of energy between modes with a different radial order. Since the coefficients |a(n)|2 are given, the characteristic error in α is the same as the characteristic error in |b(n)|2.

Absorption coefficients α versus order n are shown in Figures 6 and 7.

Figure 6. Refer to the following caption and surrounding text.

Figure 6. Absorption coefficients α = (|a|2(n) − |b|2(n))/|a|2(n0) vs. order n obtained using Whittaker decomposition. The frequency is ν = 2 mHz. Panels (a) and (b) are for the initial radial order n = 0; panels (c) and (d) are for the initial order n0 = 2. Panels (a) and (c) are for the flux tube radius R = 2 Mm; and panels (b) and (d) are for R = 8 Mm.

Standard image High-resolution image
Figure 7. Refer to the following caption and surrounding text.

Figure 7. Same as in Figure 3, but for ν = 3 mHz.

Standard image High-resolution image

The obtained absorption coefficients α(n) reveal two interesting features. First, the absorption coefficients for modes with n = n0 are positive, which suggests a partial absorption of the incident wave in the magnetized region. However, the absorption coefficients appear to be substantially smaller here than in Gordovskyy & Jain (2008); the maximum absorption coefficient for the f-mode is about 75%, whereas for p2, it is only about 10%. This arises because the pressure is smaller near the top boundary of the model (≈0.3p0), which, in turn, results in a smaller characteristic magnetic field B0. Therefore, the equipartition layer, where the sound speed and Alfvén speed are equal, is shallow, resulting in less efficient interaction between the wave and magnetic field.

The second feature is that α is negative for nn0, which suggests that significant scattering of acoustic energy across the radial order n occurs. Thus, in cases where the incoming wave consists solely of p2, the outgoing wave contains not only p2, but also the f-mode (Figures 6(c)–(d)) and p1-mode (Figures 7(c)–(d)). The power of the f-mode excited by p2 is about 30% of the power of an incident p2-mode (see also Figure 5(d)).

The calculated powers do not directly represent the amount of energy carried by different modes in the outgoing wave function; there is a multiplicative function of frequency and mode order n as well. In the following section, we evaluate energy fluxes based on the calculated powers in order to understand how the scattering across n affects actual energy fluxes carried by different modes.

3.3. Redistribution of Energy Flux Between Modes

Let us consider the outgoing component of a p-mode with the radial order n, wavenumber k, azimuthal order m = 0, amplitude b(n, k), and phase δ.

The pressure wave perturbation is described by Equation (23), while the radial velocity wave perturbations can be found as

Equation (24)

The energy flux carried by the mode outside the magnetized region through a cylindrical surface r = r1 can be written as

Equation (25)

Using Equations (22)–(23), one can rewrite the integrand in Equation (24) as follows:

Equation (26)

which, in turn, after averaging over the period of oscillations 2π/ω, can be rewritten using the Wronskian of Bessel functions (Ym(kr)Jm + 1(kr) − Jm(kr)Ym + 1(kr) = 2/(πkr)) as follows:

Equation (27)

The ratio $\mathcal {F} /b^2(n,k)$ versus order n is plotted in Figure 8. This function rapidly grows with the radial order n, and for the energy flux carried by p2 is approximately 10–20 times higher than the energy flux carried by the f-mode.

Figure 8. Refer to the following caption and surrounding text.

Figure 8. Energy flux $\mathcal {F}$ carried by a mode, normalized by the energy flux of the f-mode. The solid line with squares is for frequency ν = 3 mHz and the dashed line with circles is for ν = 2 mHz.

Standard image High-resolution image

Further, we use Equation (26) to evaluate the energy flux scattered from the incoming mode with the radial order n = n0 into other modes (nn0). This is done by adding up the energy fluxes of modes with the radial order n from 0 to 5 (except n = n0) in the outgoing wave. The resulting values, κ, are shown as dashed lines in Figure 9 as a function of frequency compared with the total energy flux lost by the incoming wave (α, solid lines). Hence, the difference between the two curves is the "real absorption."

Figure 9. Refer to the following caption and surrounding text.

Figure 9. The overall energy flux lost by mode with n0 normalized by initial energy flux (i.e., absorption coefficient α, solid lines with squares) and energy flux lost to other modes with order nn0 (κ, dashed lines with circles). Left column panels (ac) are for flux tube radius R = 2 Mm and right column panels (df) are for R = 8 Mm, the appropriate radial order of incoming wave n0 is shown in each panel.

Standard image High-resolution image

It can be seen that similar to the absorption coefficient α the energy scattered due to mode mixing, κ increases with frequency and slightly decreases with order n0. Furthermore, the coefficient κ is slightly higher for larger flux tube radii. However, it is interesting to note that the fraction of energy flux lost due to mode mixing (i.e., the ratio κ/α) apparently sharply decreases with frequency. Thus, in the case of n0 = 0 (Figures 9(a) and (d)) at frequency ν = 2 mHz, the amount of energy flux lost due to mode mixing is about κ/α ∼ 0.25–0.30 of the overall energy flux lost by the incoming mode. At the same time, at ν = 4 mHz, this fraction is only 0.05–0.15. In the case of higher radial orders n0, this contrast is not so strong: the κ/α ratio for n0 = 2 and ν = 2 mHz is about 0.3–0.5 decreasing to 0.2–0.3 at ν = 4 mHz.

However, it can be noted that the error in energy fluxes calculated for higher orders may be large. Indeed, if we assume that the characteristic error in power is the same for all orders n and is about Δb2 ∼ 2% then the characteristic error in the energy flux calculations can be found as $\Delta \mathcal {F} \approx \Delta b^2 \mathcal {F} /b^2(n,k) \Phi ^2(n,k,z)$. If the error Δb2 is constant and the ratio $\mathcal {F} /b^2(n,k)$ substantially increases with the order n, the error in the energy flux calculations also quickly increases with order n. Therefore, the contribution of the high-n modes into the energy flux of outgoing waves and, consequently, the part of energy lost due to mode mixing may be underestimated.

4. CONCLUSION

In the present paper, we consider the interaction of p-modes with a magnetic flux tube and attempt to determine the role of mode mixing in absorption of individual p-modes. We assume that there are at least two mechanisms responsible for reduction of p-mode amplitudes: conversion of the incident acoustic energy into slow magnetoacoustic waves (i.e., "real" absorption) and the redistribution of incident acoustic energy between p-modes with the same frequency but different radial orders due to mode mixing (see also Jain & Gordovskyy 2008).

Based on the results discussed in the previous section, one can make the following conclusions. The interaction of a p-mode with a magnetic flux tube results in the excitation of modes with a lower radial order, mostly the f-mode. However, this effect cannot account for all of the observed loss of energy by the incident p-mode; other effects must play a role.

Additionally, there is evidence that part of the incident acoustic energy is scattered into modes with a higher radial order. However, because the mode mass of a mode grows radically with n (see Section 3.3), even if a substantial amount of energy was transferred into p-modes with high radial orders, the resulting surface response would be weak and perhaps difficult to measure.

One of the weaknesses of this calculation is that it ignores the acoustic-jacket modes (Bogdan & Cally 1995) necessary to match the displacement of the tube oscillations and the p-modes at the interface of the magnetic and nonmagnetic regions. These modes may be important for estimating the correct amount of scattering. This issue will be addressed in our future work.

This work is supported by the Engineering and Physical Sciences Research Council, grant EP/C548795/1. B.W.H. acknowledges support from NASA through grants NNG05GM83G, NNX08AJ08G, NNX08AQ28G.

Please wait… references are loading.
10.1088/0004-637X/694/2/1602