1: \documentclass[12pt,preprint]{aastex}
2: %\usepackage{amssymb}
3: %\usepackage{amsmath}
4: %\usepackage[dvips]{graphicx}
5: \begin{document}
6: \title{On Dissipation inside Turbulent Convection Zones from 3D
7: Simulations of Solar Convection}
8: \author{Kaloyan Penev, Dimitar Sasselov}
9: \affil{Harvard--Smithsonian Center for Astrophysics, 60 Garden St., Cambridge, MA 02138}
10:
11: \author{Frank Robinson, Pierre Demarque}
12: \affil{Department of Astronomy, Yale University, Box 208101,
13: New Haven, CT 06520-8101}
14:
15: \begin{abstract}
16: The development of 2D and 3D simulations of solar convection
17: has lead to a picture of convection quite unlike the usually
18: assumed Kolmogorov spectrum turbulent flow. We investigate
19: the impact of this changed structure on the dissipation
20: properties of the convection zone, parametrized by an
21: effective viscosity coefficient. We use an expansion treatment
22: developed by Goodman \& Oh 1997, applied to a numerical model
23: of solar convection (Robinson et al. 2003) to calculate an effective
24: viscosity as a function of frequency and compare this to
25: currently existing prescriptions based on the assumption of
26: Kolmogorov turbulence (Zahn 1966, Goldreich \& Keeley 1977).
27: The results match quite closely a linear scaling with period,
28: even though this same formalism applied to a
29: Kolmogorov spectrum of eddies gives a scaling with power-law index of
30: $5\over3$.
31: \end{abstract}
32: \keywords{solar convection, turbulence, effective viscosity,
33: dissipation}
34:
35:
36: Turbulent (eddy) viscosity
37: is often considered to be the main mechanism responsible for dissipation
38: of tides and oscillations in convection zones of cool stars and
39: planets (Goodman \& Oh 1997, and references therein).
40: Currently existing descriptions have been used, with varying
41: success, to explain circularization cut-off periods for main
42: sequence binary stars (Zahn \& Bouchet 1989, Meibom \& Mathieu
43: 2005), the red edge of the Cepheid instability strip (Gonczi
44: 1982) and damping of solar oscillations (Goldreich \& Keeley 1977).
45: However, this hypothesis has been far more successful in damping oscillations
46: than damping tides, and different mechanisms have been proposed for the
47: latter, especially for planets (see Wu 2004ab; Ogilvie \& Lin 2004, and references
48: therein). In this paper we reconsider the problem of tidal dissipation in stellar convection
49: zones of solar-type stars using the turbulent velocity field from a
50: realistic 3D solar simulation. \\
51:
52: The standard treatment is to assume
53: a Kolmogorov spectrum in the convection
54: zone and apply some prescription to model the effectiveness of eddies
55: in dissipating the given perturbation. Two prescriptions have
56: been proposed to describe the efficiency of eddies in dissipating
57: perturbations with periods smaller than the eddy turnover
58: time.\\
59:
60: Firstly according to Zahn(1966, 1989), when the period of the
61: perturbation(T) is shorter than the eddy turnover time
62: ($\tau$) the dissipation efficiency is decreased because in half
63: a period the eddy only completes $T \over 2\tau$ of its
64: churn, and hence the dissipation (viscosity) should be
65: inhibited by the same factor:
66: \begin{equation}
67: \nu = \nu_{max} \min\left[ \left(T\over2\tau\right),
68: 1\right]
69: \end{equation}
70: Where $\nu_{max}$ is some constant which depends on the mixing length
71: parameter. With this assumption large eddies dominate the
72: dissipation. This prescription has been tested against tidal
73: circularization times for binaries containing a giant star
74: (Verbunt and Phinney 1996), and is in general agreement with
75: observations.\\
76:
77: Secondly, Goldreich \& Nicholson (1989) and Goldreich \& Keely
78: (1977) argue that the viscosity should be severely suppressed
79: for eddies with $\tau\gg T$, and hence the dissipation should
80: be dominated by the largest eddies with turnover times less
81: than $T/2\pi$. From Kolmogorov scaling the viscosity on a
82: given time-scale is quadratic in the time-scale, or:
83: \begin{equation}
84: \nu = \nu_{max} \min\left[\left( T\over 2\pi\tau\right)^2,
85: 1\right]
86: \end{equation}
87: This description has been used successfully by Goldreich \&
88: Keely (1977), Goldreich \& Kumar(1988), Goldreich, Kumar \&
89: Murray (1994) to develop a theory for the damping of the
90: solar $p$-modes. If the more effective dissipation was applied
91: instead, severe changes would be required in the excitation
92: mechanism in order to explain the observed $p$ mode
93: amplitudes. However, this inefficient dissipation is
94: inconsistent with observed tidal circularization for binary
95: stars (Meibom \& Mathieu 2005). Additionally, Gonczi (1982)
96: argues
97: that for pulsating stars the location of the red edge of the
98: instability strip is more consistent with Zahn's description
99: of eddy viscosity than with that of Goldreich and
100: collaborators.\\
101:
102: However, Goodman \& Oh (1997) gave a consistent hydrostatic
103: derivation of the convective viscosity, using a perturbational
104: approach. For a Kolmogorov scaling they obtained a result that is
105: closer to the less efficient Goldreich \& Nicholson viscosity than it is to
106: Zahn's. While providing a more sound theoretical
107: basis for the former scaling, this does not resolve the
108: observational problem of insufficient tidal dissipation.\\
109:
110: Both 2D and 3D numerical simulations of the solar convection
111: zone have revealed that the picture of a Kolmogorov spectrum
112: of eddies is too simplified (Stein \& Nordlund 1989, Robinson
113: et al. 2003). The simulations showed
114: that convection proceeds in a rather different, highly
115: asymmetric fashion. This suggests that the problem of
116: insufficient dissipation may be resolved by replacing the
117: assumption of Kolmogorov turbulence with the velocity
118: field produced from numerical simulations. More importantly,
119: an asymmetric and non-Kolmogorov turbulence might dissipate
120: different perturbations differently, i.e. depending both on
121: the frequency and geometry of the perturbation. Such
122: simulations have been used to develop a better model for the
123: excitation of solar $p$-modes (Samadi et al. 2003).\\
124:
125: Our approach is to apply the Goodman \& Oh (1997)
126: formalism to the velocity field obtained from realistic 3D solar surface convection
127: in a small box. The 3D simulation was
128: able to reproduce the frequency spectrum of
129: solar $p$-modes. The main result is that we find a scaling relation with
130: frequency that is in better agreement with the more
131: efficient scaling proposed by Zahn, albeit for different reasons.
132:
133: \section{Method}
134: We apply the Goodman \& Oh (1997) treatment of convection
135: to the velocity field of a 3D simulation of the outer layers of the sun.
136: Goodman \& Oh
137: assume that a steady state convection zone velocity
138: field ($\mathbf{v}$) is
139: perturbed by introducing an external velocity ($\mathbf{V}$). They also assume
140: that the convection occurs on scales small compared to the
141: perturbation, and further that the convection is approximately
142: incompressible and isentropic. Assuming that the convective
143: length scales are small compared to the perturbation allowed
144: them to consider a volume small enough to
145: accommodate all convective scales, but over that volume the
146: perturbation velocity field can be assumed linear in the
147: Cartesian coordinates ($\mathbf{x}$):
148: \begin{equation}
149: \mathbf{V} = \mathbf{A}(t)\cdot \mathbf{x}
150: \end{equation}
151: In other words we define the matrix $\mathbf{A}$ as the
152: derivative matrix of $\mathbf{V}$:
153: \begin{displaymath}
154: A_{i,j} = \frac{\partial \mathbf{V}_i}{\partial x^j}
155: \end{displaymath}
156: %
157: % Should j be a subscript on x ?
158: And keep only the first term in the Taylor series of
159: $\mathbf{V}$.\\
160:
161: Under this assumption means the results will only be applicable
162: to perturbations that are large compared to the size of the simulation domain.
163: In particular this prevents us from making any statements
164: about the 5 minute solar oscillations, because the
165: penetration depth of those is less than the box we use, and
166: the coarse resolution prevent us from looking at only the
167: upper part of the box.\\
168:
169: Assuming incompressible and isentropic convection allows one
170: to use the Eulerian equations for fluid motion:
171: \begin{eqnarray}
172: &\partial_t\mathbf{v} +
173: \mathbf{V}\cdot\nabla\mathbf{v} +
174: \mathbf{v}\cdot\nabla\mathbf{V} +
175: \mathbf{v}\cdot\nabla\mathbf{v} + \nabla w=0&
176: \label{eq: Euler 1}\\
177: &\nabla\cdot\mathbf{v}=0 \label{eq: Euler 2},&
178: \end{eqnarray}
179: where $\nabla w$ incorporates pressure and gravitational
180: acceleration, assumed to be gradients of scalar fields.\\
181:
182: The problem has two dimensionless parameters: the tidal strain
183: $\Omega^{-1} \left|\mathbf{A}\right|$, and $\left(\Omega
184: \tau_c\right)^{-1}$, where $\Omega$ is the frequency of the
185: perturbation and $\tau_c\equiv \frac{L_c}{V_c}$. The characteristic convective length scale
186: is $L_c$ and $V_c$ is the
187: characteristic convective velocity. In the case of
188: hierarchical eddie structured convection $\tau_c$ is the eddy
189: turnover time.\\
190:
191: So using eq. \ref{eq: Euler 1} and eq. \ref{eq: Euler 2} one
192: can express the perturbation in the convection velocity field in a
193: coordinate system moving with the perturbation. Expanding in
194: powers of the above dimensionless parameters and keeping only
195: first order terms gives:
196: \begin{equation}
197: \delta_{1,1} \mathbf{v'}(\mathbf{k}, \omega)
198: =-\frac{i}{\omega}\mathbf{P_k}\cdot
199: \left[\mathbf{A}(\Omega)\cdot\mathbf{v}_0(\omega-\Omega,\mathbf{k})+
200: \mathbf{A}(-\Omega)\cdot\mathbf{v}_0(\omega+\Omega),\mathbf{k})\right]
201: \label{eq: delta v}
202: \end{equation}
203: The subscripts of $\delta_{1,1} \mathbf{v'}(\mathbf{k},
204: \omega)$ indicate that only first order terms in the
205: dimensionless parameters have been included, primes
206: indicate quantities expressed in a coordinate system moving
207: with the perturbation, and $\mathbf{v}_0$ is the convective
208: velocity field in the absence of the perturbation. All of the
209: above quantities are in Fourier space, because there the
210: incompressibility is simply imposed by the projection
211: operator:
212: \begin{displaymath}
213: \mathbf{P_k}\equiv\mathbf{I}-\frac{\mathbf{kk}}{k^2}
214: \end{displaymath}
215: Eq. \ref{eq: delta v} can then be used to express the energy
216: dissipation rate again as a power series in the two
217: dimensionless quantities. Goodman and Oh s' treatment
218: implicitly assumes the box is small enough
219: for the density not to vary significantly, and so
220: it is sufficient to write the energy per unit mass as $\left<
221: \mathbf{v}\cdot\mathbf{v}\right>$ and assume that to be
222: independent of position.\\
223:
224: In our case the simulation encompasses about 8 pressure scale heights
225: so that the density varies significantly between
226: the top and bottom. This means
227: we need to use the
228: dissipation per unit volume -
229: $\left<\rho\mathbf{v}\cdot\mathbf{v}\right>$ - instead.\\
230:
231: In order to avoid taking a 7 dimensional integral, which would
232: be prohibitive in terms of computation time, we replace
233: the density with its horizontal and temporal average
234: leaving only the most important vertical dimension.
235: Taking the time derivative of the energy per unit volume using
236: that density and the perturbed convective velocity, our
237: expression for the rate of dissipation per unit volume
238: to lowest order becomes:
239: \begin{eqnarray}
240: \dot{\mathcal{E}}_{2,2}=
241: \mathbf{Re}\Bigg\{ \int\frac{d^3\mathbf{k}\;dk'_z}{(2\pi)^4}
242: \rho^*(k_z+k'_z)\Big[ \left<
243: \mathbf{v}_0(\mathbf{k},-\Omega)\cdot\mathbf{A}(\Omega)\cdot
244: \mathbf{P_{k'}}\cdot\mathbf{A}(\Omega)\mathbf{v}_0(\mathbf{k'},
245: -\Omega)\right> &&\nonumber\\
246: + \left.\mathbf{v}_0(\mathbf{k},-\Omega)\cdot\mathbf{A}(\Omega)\cdot
247: \mathbf{P_{k'}}\cdot\mathbf{A}(-\Omega)\mathbf{v}_0(\mathbf{k'},
248: \Omega)\right> \Big] \Bigg\}&&
249: \label{eq: E dot full}
250: \end{eqnarray}
251: Where $k'=(-k_x, -k_y, k'_z)$, and the subscripts, as before, denote
252: the order in the two dimensionless parameters characterizing the tide
253: and the convection respectively. $\rho(k_z)$ is the fourier transform
254: of the density averaged over $x,y,t$. The normalization is such that
255: $\rho(0)$ is the average density over all space and time.\\
256:
257: Eq. \ref{eq: E dot full} gives an anisotropic viscosity, for which
258: we can obtain
259: the different components by setting all terms of $\mathbf{A}$
260: to $0$ except for one, and comparing to the equivalent
261: expression for the molecular viscosity:
262: \begin{equation}
263: \dot{\mathcal{E}}_{visc} = \frac{1}{2}
264: \left<\rho\nu\right>
265: \;Trace\left[\mathbf{A}(\Omega)\cdot\mathbf{A^*}(\Omega)\right]
266: \label{eq: E dot mol}
267: \end{equation}
268: Where the average is over the volume and over time.
269: %Frank Robinson
270: \section{Realistic 3D solar surface convection}
271:
272: The 3D simulation of the Sun is case D in Robinson et al. (2003). This
273: has dimensions 2700 km $\times$ 2700 km $\times$ 2800 km
274: on a $58 \times 58 \times 170$ grid.
275: A detailed one-dimensional (1D)
276: evolutionary model e.g. see Guenther et al. (1992)
277: provided the starting model for the
278: 3D simulation. Full details of the
279: numerical approach and physical assumptions
280: are described in Robinson et al. (2003).
281:
282: The simulation extended from a few hundred km above the photosphere down to a
283: depth of about 2500 km below the
284: visible surface (photosphere). This is about 8 pressure scale heights.
285: The box had periodic side walls and impenetrable top and bottom surfaces with a
286: constant energy flux fed into the base and a conducting top boundary.
287: The flux was computed from the 1D stellar
288: model, thus was not arbitrary, but
289: was the correct amount of energy flux the computation domain should transport
290: outward in a particular star.
291:
292: To get a thermally relaxed system in a reasonable amount of computer time,
293: they used an implicit numerical scheme, ADISM (Alternating Direction Implicit
294: on a Staggered Mesh) developed by Chan \& Wolff (1982).
295: Careful attention was paid to the geometric size of the box. Importantly the domain was
296: deep enough and wide enough
297: to ensure the boundaries had minimal effect on the bulk of the overturning
298: convective eddies
299: (or on the flow statistics).
300: The convection simulation was run using the ADISM code until it
301: reached a statistically steady state. This was checked by
302: confirming that the influx and outflux of the box were within 5 \% of
303: each other and the run of the maximum velocity
304: have reached an asymptotic state.
305:
306: After the model was relaxed they sampled the entire 3D velocity field at 1
307: minute intervals.
308: The data set used in this paper consists of 150 minutes of
309: such solar surface convection. This is about 20 granule turnover
310: life times. An example velocity snapshot of the convective flow is
311: presented in fig. \ref{fig: snapshot}.
312:
313: \begin{figure}[tbp]
314: \begin{center}
315: \includegraphics[angle=270,width=0.4\textwidth]{f1}
316: \caption{A sample snapshot of the convective flow.
317: Blue color indicates downwarad flow, red indicates upward
318: flow. The arrows show the
319: velocity normalized to the sound speed. The yellow
320: line represents the convective surface (i.e. where the
321: entropy gradient is 0).
322: }
323: \label{fig: snapshot}
324: \end{center}
325: \end{figure}
326:
327: % Frank robinson
328:
329: \section{Results}
330: We implement eqs. \ref{eq: E dot full} and \ref{eq: E dot mol}
331: by taking discrete Fourier transforms (FFT) of the velocity
332: field and the averaging density horizontally and over time.
333: In doing so it is important to verify that the
334: windows introduced by the limited time and space extent of the
335: simulation box do not dominate the results. This was done by
336: repeating the calculation with the raw results, without any
337: windowing and with Welch and Bartlett windows applied to all
338: the dimensions simultaneously. As expected this has little or
339: no effect on the frequency scaling (see below).\\
340:
341: \begin{figure}[tbp]
342: \begin{center}
343: % \plottwo{f2a}{f2b}
344: % \plotone{f2c}
345: \includegraphics[width=0.45\textwidth]{f2a}
346: \includegraphics[width=0.45\textwidth]{f2b}
347: \includegraphics[width=0.45\textwidth]{f2c}
348: \caption{
349: a) The off diagonal terms of the viscosity
350: tensor compared to the z-z component.
351: b) The diagonal terms of the viscosity tensor.
352: c) The z-z component of the viscosity tensor (solid line)
353: computed using eq. \ref{eq: E dot full} and eq.
354: \ref{eq: E dot mol} compared to the frequency
355: scalings proposed by Zahn, Goldreich et. al.
356: and Goodman and Oh.
357: The horisontal axis for all the plots is
358: the frequency in cycles per min.
359: }
360: \label{fig: scalings}
361: \end{center}
362: \end{figure}
363:
364: As the viscosity tensor defined by eqs. \ref{eq: E dot full} and
365: \ref{eq: E dot mol} is clearly symmetric, it only
366: contains 6 independent real valued components. Figure
367: \ref{fig: scalings} displays the values of the viscosities we calculated.
368:
369: % you may need to state the expressions as it is not clear
370:
371: Fig. \ref{fig: scalings}a shows that the off
372: diagonal terms are completely insignificant compared to the
373: diagonal terms. Since in all the situations that concern
374: us, the divergence of the perturbation field is never small
375: compared to the other derivatives of the perturbing velocity
376: field, the dissipation will be dominated by the diagonal terms.
377: Hence their scaling with frequency will determine how the
378: dissipation scales. \\
379:
380: Fig. \ref{fig: scalings}b shows that all the
381: diagonal components scale roughly the same way with frequency
382: and are dominated by the z-z component, although
383: not by that dramatic a difference. Furthermore, for perturbations
384: like tides the z derivative of the z component of the
385: perturbation velocity is the largest element of the matrix
386: $\mathbf{A}$ and hence that will be the term that will
387: determine the frequency scaling of the dissipation.\\
388: % Figures need to be in order
389:
390: In Fig. \ref{fig: scalings}c we see the comparison between the
391: different scalings with frequency suggested so far. We also show the
392: scaling that we obtain by applying the Goodman and Oh (1997) method
393: to a simulated 3D convection velocity field. The lines shown are
394: least square fits to the curve we obtain from the simulation
395: velocities. They seem to all intersect at the upper right-hand corner
396: because the fits were done in linear space, not logarithmic, and
397: hence do not tolerate even small deviations in the upper portion
398: of the log-log plot. The best fit slope for our curve (not
399: shown) is:
400: \begin{displaymath}
401: \nu\propto \Omega^{1.1\pm0.1}
402: \end{displaymath}
403: regardless whether we do the fit in linear or
404: logarithmic space.\\
405:
406: What are the possible sources of
407: error in this result? Firstly we have assumed an
408: incompressible flow in order to simplify the treatment.
409: However, the fluid simulations used are not incompressible,
410: because at the top of the convection zone, where most of the
411: driving of the convection occurs, the flow velocities reach
412: very close to the speed of sound and hence the flow is
413: necessarily compressible. However, even though that layer is
414: extremely important for the flow established below, it only
415: contributes insignifficantly to the turbullent dissipation,
416: because it only contains a few percent of the total mass.
417: To
418: verify that only a small fraction mass lies in a compressible
419: region for each grid point, we define a compressibility
420: parameter $\xi \equiv
421: \tau_c\left|\nabla\cdot\mathbf{v}\right|$, where $\tau_c$ is
422: the eddy turnover time in our box. In fig. \ref{fig: M(xi)} we
423: plot the mass fraction with $\xi$ less than certain value. It
424: is clear that the incompressibility assumption is violated
425: only for a negligible fraction of the mass. As we noted before
426: the flow is compressible only near the top of the box. To
427: confirm that the presence of this region does not
428: signifficantly affect our results we repeated the analysis
429: separately for the top and bottom halves of the simulation
430: box. The two new scalings obtained this way were completely
431: consistent with the scaling of viscosity with frequency for
432: the entire box.\\
433:
434: \begin{figure}[tbp]
435: \begin{center}
436: \includegraphics[width=0.45\textwidth]{f3}
437: \caption{The fraction of the total mass residing in a
438: region with compressibility parameter
439: $\xi\equiv\tau_c\nabla\cdot\mathbf{v}$ less than the
440: given value.
441: }
442: \label{fig: M(xi)}
443: \end{center}
444: \end{figure}
445:
446:
447: Next, the fact that we have a finite
448: (small) portion of the convection zone, both in time and in
449: space could be important. We only treat the top portion of the
450: solar convection zone and hope that the result is not very
451: sensitive to depth. Of course it would be ideal to have the
452: entire depth of the convection zone covered, but with current
453: computational resources this is way outside of reach.\\
454:
455: The finite span of the simulations may also be introducing edge
456: effects which can be treated by applying some sort of a window
457: function. We tried Welch, Bartlett and square
458: window (no window). To verify that the time window
459: available is large enough, we tried ignoring the last approximately
460: $1/3$ of the data. We carried all those test on two independent
461: runs of the model. The slopes this produced ranged from
462: $\nu\propto \Omega^{0.98}$ to $\nu\propto \Omega^{1.19}$, where most of
463: the difference originated from the two independent runs.\\
464:
465: In addition the finite resolution
466: might be leading to aliasing that could change our result.
467: In particular make it flatter than it really is, by basically
468: dumping additional power to the frequencies for which the
469: dissipation is smallest (the places with higher value of the
470: dissipation are less likely to be affected significantly).
471: The effects of this can be seen in the diagonal viscosity
472: components. The tails of their curves become flatter toward
473: the end. The fact that this is restricted to the end of
474: the curves is encouraging as it suggests only the high frequency
475: end of the curve is affected. Also we have looked at
476: crossections of the Fourier transformed velocity field and
477: they do tail off at high $|k|$, which gives us confidence
478: that the resolution is sufficient to capture most of the
479: spectral power and that aliasing effects will be small.\\
480:
481: Finally there are statistical errors associated with every
482: point. Those can be estimated by noting the difference between
483: $\nu_{xx}$ and $\nu_{yy}$ in Fig. \ref{fig: scalings}b.
484: Physically one expects that there should be no differences
485: between the two horizontal directions of the simulation box,
486: so the differences between them is some sort of measure of
487: the error. In particular from there one can see that the first
488: few points (at the low frequency end) are significantly less
489: reliable than the rest, but apart from the first few points
490: those errors become small. The average fractional uncertainty is
491: $\sim 3\%$, which leads to an overall error in the slope of
492: $0.01$.\\
493:
494: Abandoning the
495: Kolmogorov picture of turbullence clearly has a large effect on the result.
496: Even though we use the approach of Goodman \& Oh,
497: which gives a power law index of $5/3$ for a Kolmogorov
498: turbulence, our results give a scaling, rather different
499: from the previous prescriptions. We also find that the viscosity
500: is no longer isotropic. This is due to the signifficant
501: difference in scaling between the velocity power spectrum with
502: frequency and wavenumber in our simulation and the Kolmogorov
503: prescription (see fig. \ref{fig: power spectra}).
504: There are two important distictions apparent.
505: First the frequency spectrum of our box is much shallower than
506: the Kolmogorov prescription. This is
507: responsible for the slower loss of efficiency of viscosity
508: with frequency that we observe. Second the radial direction is
509: clearly very different from the two horizontal directions ---
510: $v_x$ and $v_y$behave very differently from $v_z$ and the
511: dependence of $\mathbf{v}$ on $x$ and $y$ is different from
512: the $z$ dependence (fig. \ref{fig: power spectra} a, b) ---
513: of course this results in the anisotropy of the viscosity
514: tensor we calculate. Even though the spatial dependence of the
515: horizontal velocity components is much different from the
516: radial velocity spatial dependence, the frequency power
517: spectrum of all three components scales roughly like
518: $P\propto\Omega^{-1}$ (fig. \ref{fig: power spectra}c). From
519: eq. \ref{eq: E dot full} we see that if all the components of
520: $\mathbf{v}$ have the same scaling with frequency, that same
521: scaling will also apply for the viscosity, which is indeed
522: what we observe.\\
523:
524: \begin{figure}[tb]
525: \begin{center}
526: % \plottwo{f4a}{f4b}
527: % \plotone{f4c}
528: \includegraphics[width=0.45\textwidth]{f4a}
529: \includegraphics[width=0.45\textwidth]{f4b}
530: \includegraphics[width=0.45\textwidth]{f4c}
531: \caption{a) Spatial power spectrum of the
532: horizontal velocities. Only
533: one of the horizontal componenents is plotted but the
534: power spectrum of the other horizontal component is
535: identical.
536: b) Spatial power spectrum of the radial velocity.
537: c) Frequency power spectrum of the three
538: velocity components. The straight solid line -
539: $P\propto\omega^{-1}$ - gives a good approximation to
540: all three scalings.}
541: \label{fig: power spectra}
542: \end{center}
543: \end{figure}
544:
545: \section{Discussion}
546:
547: Our result is somewhat unexpected. It apparently stems from the
548: fact that the structure of the convection velocity field produced by
549: the 3D simulations is very different
550: from simple isotropic Kolmogorov turbulence. The picture that
551: emerges from these simulations consists of large scale slow up-flows
552: penetrated by relatively fast and very localized
553: down-drafts that are coherent over a signifficant portion of the
554: simulation box and persistent for extended periods of time.
555: This is what causes the anisotropy and also seems to conspire
556: to change the scaling with frequency, and make it relatively flat.
557: This makes our results appear closer to Zahn's prescription, which
558: is coincidental, given the different physical assumptions. The
559: question of what exactly is the reason for the shallower
560: frequency dependence of the dissipation is of course a very
561: interesting one. However, using a perturbative approach,
562: limits us in our ability to answer it. To properly address
563: this question one would need to create a consistent
564: hydrodynamical simulation that allows for the perturbation
565: velocity field to be put directly into the equations of motion
566: and not treated by a perturbative approach after the fact.
567: This would also address the question of whether the expansion
568: is actually converging and if taking the first nonzero term is
569: a good approximation, which is currently only our hope.
570:
571: This enhanced dissipation is in better % can you quantify this statement in any way
572: agreement with data on the
573: circularization of the orbits of Sun-like main sequence stars, and the
574: location of the instability strip as discussed earlier. We currently cannot make
575: any statements about the dissipation of p-modes, because those do not satisfy
576: the assumption of linearity and incompressibility of the perturbation
577: velocity over the simulation box. However, we have used a
578: solar 3D convection simulation which is consistent with the
579: solar p-mode spectrum.\\
580:
581: Note that our approach here is more appropriate to tides raised by a planet on a
582: slow (non-synchronized) star (Sasselov 2003). The problem of binary stars
583: circularization will require a detailed treatment and understanding of
584: the feedback on the convection zone. On the other hand, the tidal
585: dissipation in fast-rotating fully-convective planets and stars might
586: be dominated by inertial waves (Wu 2004ab, Ogilvie \& Lin 2004). They
587: are sensitive to turbulent viscosity however, and the linear scaling
588: has a strong effect on their dissipation (Wu 2004b). This issue deserves
589: further study.
590:
591: \begin{thebibliography}{99}
592: \bibitem{chanwolff1982}
593: Chan, K. L. and Wolff, C. L. 1982, J. Comp. Physics, 47, 109
594:
595: \bibitem{goldreich keely 77}
596: Goldreich, P. \& Keely, D. A. 1977, ApJ, 211, 934
597:
598: \bibitem{goldreich kumar 88}
599: Goldreich, P. \& Kumar, P. 1988, ApJ, 326, 462
600:
601: \bibitem{goldreich kumar murray 94}
602: Goldreich, P., Kumar, P. \& Murray, N. 1994, ApJ, 424, 466
603:
604: \bibitem{goldreich nicholson 89}
605: Goldreich, P. \& Nicholson, P.D. 1989, Icarus, 30, 301
606:
607: \bibitem{goodman and oh97}
608: Goodman, J., Oh, S. P. 1997, ApJ, 486, 403
609:
610: \bibitem{guenther1994}
611: Guenther, D.B., Demarque, P., Kim, Y.-C. and Pinsonneault, M.H. 1992, ApJ 387, 372
612:
613: \bibitem{kimchan1998}
614: Kim, Y.-C. and Chan, K.L. 1998, ApJ, 496, L121
615:
616: \bibitem{}
617: Robinson, F.J., Demarque, P., Li, L.H., Sofia, S., Kim, Y.-C., Chan, K.L.,
618: Guenther, D.B. 2003, MNRAS, 340, 923
619:
620: \bibitem{Sasselov 2003}
621: Sasselov, D.D. 2003, Ap.J., 596, 2, pp. 1327
622:
623: \bibitem{stein nordlund 1989}
624: Stein, R. F., Nordlund, \AA. 1989, ApJ, 342, L95
625:
626: \bibitem{}
627: Verbunt, F., \& Phinney, E. S. 1996, A\&A, 296, 709
628:
629: \bibitem{zahn66}
630: Zahn, J. P. 1966, Ann. d'Astrophys., 29, 489
631:
632: \bibitem{zahn77}
633: Zahn, J. P. 1977, A\&A, 57, 383
634:
635: \bibitem{zahn89}
636: Zahn, J. P. 1989, A\&A, 220, 112
637:
638: \bibitem{zahn bouchet 92}
639: Zahn, J. P. and Bouchet, L. 1989, A\&A, 223, 112
640: \end{thebibliography}
641:
642: \end{document}
643: