0710.3868/ms.tex
1: %\documentclass[12pt,preprint]{aastex} 
2: \documentclass{emulateapj}
3: 
4: \newcommand{\be}{\begin{equation}}
5: \newcommand{\ee}{\end{equation}}
6: \newcommand{\beqn}{\begin{eqnarray}}
7: \newcommand{\eeqn}{\end{eqnarray}}
8: \newcommand{\bld}[1]{\mbox{\boldmath$#1$\unboldmath}}
9: \newcommand{\mycite}[1]{\citeauthor{#1}\ \citeyear{#1}}
10: 
11: \begin{document}
12: 
13: 
14: \title{Formation, Survival, and Destruction of Vortices in Accretion Disks}
15: 
16: 
17: \author{Yoram Lithwick\altaffilmark{1}}
18: \altaffiltext{1}{CITA. Toronto, Ontario, Canada; yoram@cita.utoronto.ca}
19: 
20: 
21: \begin{abstract}
22: 
23: 
24: Two dimensional hydrodynamical disks
25:  are nonlinearly unstable to the formation of vortices.  
26:  Once formed, these vortices essentially survive forever.  
27:  What happens in three dimensions?
28:  We show with incompressible shearing box simulations that  in 3D a vortex in
29:  a short box forms and survives just as in 2D.  But a vortex
30:  in a tall box is unstable and is destroyed. 
31: In our simulation, the unstable vortex decays into a transient turbulent-like state that 
32: transports angular momentum outward at a nearly constant rate
33: for hundreds of orbital times.
34: The 3D  instability that destroys vortices is a generalization of  the 2D instability
35: that forms them.
36: We derive  the conditions for these nonlinear instabilities to act
37: by calculating the coupling between linear modes, and
38: thereby derive the criterion
39: for a vortex to survive in 3D as it does in 2D: {\it the azimuthal extent of the
40: vortex must be larger than the scale height  of the accretion disk}. 
41:  When this criterion is violated, the
42: vortex is unstable and decays.
43: Because vortices are longer in azimuthal than
44: in radial extent by a factor that is inversely proportional to their
45: excess vorticity, a vortex with given radial extent will only survive in a 3D disk
46: if it is sufficiently weak.  This counterintuitive result explains why previous
47: 3D simulations always yielded decaying vortices: their 
48: vortices were too strong.
49:  Weak vortices behave two-dimensionally even if their width is much less than their height
50:  because they are stabilized by rotation, and behave as Taylor-Proudman columns.
51: We conclude that in protoplanetary disks  weak vortices can trap dust and serve
52: as the nurseries of planet formation.  Decaying strong vortices might be responsible
53: for the outwards transport of angular momentum that is required to make accretion
54: disks accrete. 
55: 
56:  
57:  \end{abstract}
58: \keywords{accretion, accretion disks --- instabilities --- solar system: formation ---turbulence }
59: 
60: \section{Introduction}
61: 
62: Matter accretes onto a wide variety of objects, such as young stars, black holes, and white
63: dwarfs, through accretion disks.  In highly ionized disks  magnetic fields are important, 
64: and they trigger turbulence via 
65: the magnetorotational instability \citep{BH98}.  However, many disks,
66:  such as those around
67: young stars or dwarf novae, are nearly neutral  \citep[e.g.,][]{SMUN00,GM98}. 
68: In these disks, 
69: the fluid motions
70: are well described by hydrodynamics. 
71: 
72: Numerical simulations of hydrodynamical 
73: disks in two-dimensions---in the plane of the disk---often
74: produce long-lived vortices \citep{GL99,UR04,JG05}.
75: If vortices really exist in accretion disks, 
76: they can have important consequences.  First and foremost,
77: they might generate turbulence.
78: Since turbulence naturally transports angular momentum outwards\footnote{
79: Energy conservation implies that turbulence transports angular momentum outwards;
80: see \S \ref{sec:pseudo}.  
81:  Nonetheless, if an external energy source  (e.g., the radiative
82: energy from the central star) drives the turbulence, then angular momentum could in principle be transported 
83: inwards.
84: }, 
85: as is required for mass to fall inwards, it might be vortices 
86: that cause accretion disks to accrete.
87: Second, in disks around young stars, long-lived vortices can trap solid particles
88: and initiate the formation of planets \citep{BS95}.
89: 
90: 
91: Why do vortices naturally form in 2D simulations?
92: Hydrodynamical disks are stable to linear perturbations.  
93: However, they are nonlinearly unstable, despite some claims to the 
94: contrary in the
95: astrophysical literature.
96: In two dimensions, the incompressible hydrodynamical
97: equations of a disk are equivalent to those of a non-rotating linear
98: shear flow \citep[e.g.,][hereafter L07]{L07}.
99: And it has long been known that such flows are nonlinearly unstable 
100: (\mycite{Gill65}; \mycite{LK88}; L07).  This nonlinear instability is just 
101: a special case of the Kelvin-Helmholtz instability. 
102: Consider a linear shear flow extending throughout the $x$-$y$ plane
103: with velocity profile $\bld{v}=-qx\bld{\hat{y}}$, where $q>0$ 
104: is the constant shear rate, so that $-q$ is the flow's vorticity.
105: (In the equivalent accretion disk, the local angular speed is 
106: $\Omega=2q/3$.)  This shear flow is linearly stable to 
107: infinitesimal perturbations.  But if the shear profile is altered
108: by a small amount, the alteration can itself be unstable
109: to infinitesimal perturbations.  To be specific, let the alteration
110: be confined
111: within a band of width $\Delta x$, and let it 
112: have vorticity $\omega=\omega(x)$ (with $|\omega|\lesssim q$), so that it induces a velocity field
113: in excess of the linear shear
114: with components $u_y\sim \omega\Delta x$ and $u_x=0$.  
115: Then this band is unstable to infinitesimal nonaxisymmetric
116: (i.e. non-stream-aligned) perturbations provided roughly that
117: \be
118: \left|k_y
119: \right|
120: \lesssim
121: {1\over q}{|\omega|\over \Delta x}
122: \ \ \Rightarrow 
123: {\rm 2D\ instability}
124: \label{eq:2dinst}
125: \ee
126: where $k_y$ is the wavenumber of the nonaxisymmetric perturbation.\footnote{
127: More precisely, the necessary and sufficient condition for instability
128: in the limit $|\omega|\ll q$ is that $|k_y|<{1\over 2 q}\int_{-\infty}^\infty {d\omega/dx\over x-x_0}dx$, where
129: $x_0$ is any value of $x$ at which $d\omega/dx=0$ \citep[][L07]{Gill65,LK88}.
130: For arbitrarily large  $\omega$, 
131: Rayleigh's inflection point theorem and Fj\o rtoft's theorem give necessary (though
132: insufficient) criteria for instability \citep{DR04}.  The former states that for 
133: instability, it is required that $d\omega/dx=0$ somewhere in the flow, i.e. that the velocity
134: field must have an inflection point.  \cite{Love99} generalize Rayleigh's inflection
135: point theorem to compressible and nonhomentropic disks.
136: \label{foot:inst}
137: }
138: For any value of $|\omega|$ and $\Delta x$, the band is always unstable 
139: to perturbations with long enough wavelength.
140: Remarkably, instability even occurs when $|\omega|$ is infinitesimal.
141: Hence we may regard this  as a true nonlinear instability.  
142: \cite{BH06} assert that detailed numerical simulations have not shown evidence for nonlinear
143: instability. The reason many simulations fail to see it
144: is that their boxes are not long enough in the $y$-direction to encompass a small
145: enough non-zero $|k_y|$.
146: 
147: 
148: In two dimensions, the outcome of this instability is a long-lived vortex (e.g., L07).
149: A vortex that has been studied in detail is the Moore-Saffman vortex, 
150: which is a localized patch of spatially
151: constant vorticity  superimposed on a linear shear flow  \citep{Saffman95}.
152: When $|\omega|\lesssim q$, 
153: where $\omega$ here refers to the spatially constant
154: excess vorticity within the patch, and when the vorticity within
155: the patch ($\omega-q$) is stronger than that of the background shear,
156: then the patch forms a stable vortex that is elongated in $y$ relative
157: to $x$ by the factor 
158: \be
159: {\Delta y\over \Delta x} \sim {q\over |\omega|} \ . \label{eq:ms}
160: \ee
161: This relation applies not only to Moore-Saffman vortices,
162: but also to vortices whose $\omega$ is not spatially constant.
163: It may be understood as follows.
164: A patch with characteristic excess vorticity $\sim\omega$ and with $\Delta y\gg \Delta x$
165: induces a velocity field in the $x$-direction with amplitude $u_x\sim|\omega|\Delta x$, 
166: independent of the value of $\Delta y$ (e.g., \S 6 in L07). As long as $|\omega|\lesssim q$,
167: the $y$-velocity within the vortex is predominantly due to the background shear, 
168: and is $\sim q\Delta x$. Therefore the time to cross the width of the vortex is
169: $t_x\sim \Delta x/u_x \sim 1/|\omega|$,
170: and the time to cross its length is 
171: $t_y\sim \Delta y/(q\Delta x)$.  Since these times must be comparable in a vortex,
172: equation (\ref{eq:ms}) follows.
173: Equation (\ref{eq:ms}) is very similar to equation (\ref{eq:2dinst}). 
174: The 2D instability
175: naturally forms into a 2D vortex.  
176: Futhermore, the exponential growth rate of the instability is $\sim |\omega|$, 
177: which is comparable to the rate at which fluid circulates around the vortex.
178: 
179: More generally,
180: an arbitrary axisymmetric profile of $\omega(x)$ tends to evolve into a 
181: distinctive
182: banded structure.  Roughly speaking, bands where $\omega<0$ contain vortices, and these
183: are interspersed 
184: with bands where $\omega>0$, which  contain no vortices.   (Recall that we take
185: the background vorticity to be negative; otherwise, the converse would hold.)
186: The reason for this  is that only regions that have $\omega<0$ can be unstable,
187:  as may be inferred either from the integral
188: criterion for instability given in footnote \ref{foot:inst}, or from Fj\o rtoft's theorem.
189: For more detail on vortex dynamics in shear flows, see
190: the review by \cite{Marcus93}.
191: 
192: 
193: What happens in three dimensions? To date, numerical simulations of 
194: vortices in 3D disks have been reported in two papers.
195: \cite{BM05b} 
196: initialized their simulation with a Moore-Saffman vortex, and
197: solved the anelastic equations in a stratified disk.
198: They found that this vortex decayed.  As it decayed,
199: new vortices were formed
200: in the disk's atmosphere, two scale heights above the midplane. The new vortices
201:  survived for the duration of the simulation.
202: \cite{SSG06} performed both 2D and 3D simulations of the compressible 
203: hydrodynamical equations in an unstratified disk, initialized with 
204: large random fluctuations.  They found that whereas the 2D simulations
205: produced long-lived vortices, 
206: in three dimensions  vortices rapidly decayed.
207: 
208:  Intuitively, it seems
209: clear that a vortex in a very thin disk will behave as it does in  2D.
210: And from the  3D simulations described above it may be inferred that placing this vortex in a
211: very thick disk will induce its decay. 
212: Our main goal in this paper is to understand  these two behaviors, and the transition
213: between them.
214: A crude explanation of our final result is that vortices decay when the 2D
215: vortex motion
216: couples resonantly to 3D modes, i.e., to modes that have vertical wavenumber
217: $k_z\ne 0$.  As described
218: above, a vortex with excess vorticity $|\omega|$ has circulation frequency $\sim|\omega|$, 
219: and $k_y/k_x\sim |\omega|/q$, where $k_x$ and $k_y$ are its ``typical'' wavenumbers.
220: Furthermore, it is well-known that the frequency of axisymmetric ($k_y=0$) inertial
221: waves is $\Omega k_z/\sqrt{k_x^2+k_z^2}$  (see eq.  [\ref{eq:axi3d}]).
222: Equating the two frequencies, and taking the $k_x$ of the 3D mode to be comparable
223: to the $k_x$ of the vortex, as well as setting $q=3\Omega/2$ for a Keplerian disk, we find
224: \be
225: k_z\sim k_y \label{eq:res}
226: \ee
227: as the condition for resonance.
228: Therefore a vortex with length $\Delta y$  will survive in a box
229: with height $\Delta z\lesssim \Delta y$, because in such a box
230: all 3D modes have too high a frequency to couple with the vortex, i.e., all
231:  nonzero $k_z$
232: exceed the characteristic $k_y\sim 1/\Delta y$.
233: But when $\Delta z\gtrsim \Delta y$, there exist $k_z$ in the box that satisfy
234: the resonance condition (\ref{eq:res}), leading to the vortex's destruction.
235: This conclusion suggests that vortices live indefinitely in disks with scale height
236: less than their length ($h\lesssim \Delta y$)
237: because in such disks all 3D modes have too high a frequency for resonant coupling.
238: This conclusion is also consistent  with the simulations of
239: \cite{BM05b} and \cite{SSG06}. Both of these works
240:  initialized their simulations with strong excess vorticity
241: $|\omega|\sim q$, corresponding to nearly circular vortices.  
242: Both had vertical domains that were comparable to the vortices' width.
243: Therefore both saw that their vortices decayed.  Had they initialized their simulations
244: with smaller $|\omega|$, and increased the box length $L_y$ to encompass
245: the resulting elongated vortices, both would have found long-lived 3D vortices.
246:  \citeauthor{BM05b}'s discovery of long-lived vortices in the disk's
247: atmosphere is simple to understand because  the local scale height is reduced
248: in inverse proportion to the height above the midplane.   Therefore 
249: higher up in the atmosphere the dynamics becomes more two-dimensional, and
250: a given vortex is better able to survive the higher it is.\footnote{
251: However, \cite{BM05b} also include buoyancy forces in their simulations,
252: which we ignore here.  How buoyancy affects the stability of vortices is a topic
253: for future work.
254: }
255: 
256: 
257: \subsection{Organization of the Paper}
258: 
259: In  \S \ref{sec:eom} we introduce the equations of motion, and
260: in \S \ref{sec:pseudo} we present
261:  two pseudospectral simulations.
262:  One illustrates the
263: formation and survival of a vortex in a short box, and the other illustrates
264: the destruction of a vortex in a tall box. 
265: 
266: In \S\S \ref{sec:lin}-\ref{sec:nonlin} we develop a theory explaining
267: this behavior.
268: The reader who is satisfied by the qualitative description leading
269: to equation (\ref{eq:res}) may
270: skip those two sections.
271: The theory that we develop is indirectly related to the transient
272: amplification scenario for the generation of turbulence.
273:  Even though 
274: hydrodynamical disks are linearly stable, linear perturbations 
275: can be transiently amplified before they decay, often by a 
276: large factor.  It has been proposed that sufficiently amplified modes might
277: couple nonlinearly, leading to turbulence \citep[e.g.,][]{CZTL03,Yecko04,AMN05}.
278: However, to make this proposal more concrete, one must work out
279: how modes couple nonlinearly.  In  L07, we did that in two dimensions.  
280: We showed that the 2D nonlinear instability of equation (\ref{eq:2dinst}) is a consequence
281: of the coupling of an axisymmetric mode with a transiently amplified mode, 
282: which may be called a ``swinging mode'' because its phasefronts are swung around
283: by the background shear. 
284: In \S \ref{sec:nonlin} we show that the 3D instability responsible
285: for the destruction of vortices is a generalization of this 2D instability. 
286: It may be understood by examining the coupling of a 
287: 3D swinging mode 
288: with an axisymmetric one.    3D modes become increasingly unstable as $|k_z|$
289: decreases, and in the limit that $k_z\rightarrow 0$, the 3D instability matches
290: smoothly onto the 2D one.
291: Thicker disks are more prone to 3D instability because they encompass
292: smaller $|k_z|$.
293: 
294: 
295: 
296: 
297: \section{Equations of Motion}
298: \label{sec:eom}
299: 
300: 
301: 
302: 
303: We solve the  ``shearing box'' equations,
304: which approximate the dynamics in an accretion disk on lengthscales
305: much smaller than the distance to the disk's center. 
306: We assume incompressibility, which is a good approximation
307: when relative motions are subsonic.
308: We also neglect vertical gravity, and hence stratification and buoyancy, which is 
309: an oversimplification. 
310: To fully understand vortices in astrophysical disks, one must consider the effects
311: of  vertical gravity as well as of shear and rotation. 
312: In this paper, we consider only two pieces of this puzzle---shear and
313: rotation.  Adding the third piece---vertical gravity---is a topic that 
314: we leave for future investigations.  See also the Conclusions for some 
315: speculations.
316: 
317: 
318: 
319: An unperturbed Keplerian disk has angular velocity profile $\Omega(r)\propto r^{-3/2}$.
320: In a reference
321: frame rotating at constant angular speed $\Omega_0\equiv \Omega(r_0)$, where $r_0$ is
322: a fiducial radius, 
323: the incompressible shearing box equations of motion  read
324: \begin{eqnarray}
325: {\partial_t {\bld v}}+{\bld{ v\cdot\nabla v}}&=&-
326: 2\Omega_0{\bld{\hat z}}{\bld{\times v}}+2q\Omega_0x{\bld{ \hat{x}}} -{\bld\nabla} P/\rho\ \ ,
327: \label{eq:bigeom}
328: \\
329: \bld{\nabla\cdot v}&=&0  
330: \end{eqnarray}
331: adopting
332:  Cartesian coordinates $x,y,z$, which are related to the disk's cylindrical $r,\theta$ 
333: via $x\equiv r-r_0$ and $y\equiv r_0(\theta-\Omega_0t)$;
334: $\bld{\hat{x}}$ and $\bld{\hat{z}}$ are unit vectors, 
335: and
336: \be
337: q\equiv -{d \Omega\over d\ln r}\Big\vert_{r_0}={3\over 2}\Omega_0 \ ,
338: \ee
339: We retain $q$ and $\Omega_0$ as independent parameters because they 
340: parameterize different effects: shear and rotation, respectively. 
341: The first term on the right-hand side of equation (\ref{eq:bigeom}) is the Coriolis force, 
342: and the second is what remains after adding centrifugal and gravitational forces.
343: Decomposing  the velocity  into
344: \be
345: \bld{v}=-qx\bld{\hat{y}}+\bld{u} \ , \label{eq:dec}
346: \ee
347: where the first term is the shear flow of the unperturbed disk,  yields
348: \begin{eqnarray}
349: \left( {\partial_t}-qx{\partial_y}\right) {\bld u}
350: +{\bld{u\cdot\nabla u}}
351: &=&
352:  q u_x\bld{\hat{y}}
353:  -2\Omega {\bld{\hat z}}{\bld{\times u}}
354:  -\bld{\nabla} P/\rho \label{eq:eom2}
355: \\
356: {\bld{\nabla\cdot u}} &=&
357: 0   \  ,
358: \label{eq:eom1}
359: \end{eqnarray}
360: dropping the subscript from $\Omega_0$, as we shall do 
361: in the remainder of this paper.  An unperturbed disk has 
362: $\bld{u}=0$.
363: 
364: In addition to the above ``velocity-pressure'' formulation, an
365: alternative ``velocity-vorticity'' formulation will prove convenient.  It
366: is given by 
367: the curl of equation (\ref{eq:eom2}), 
368: \be
369: (\partial_t-qx\partial_y)\bld{\omega}=-q\bld{\hat{y}}\omega_x+(2\Omega-q)\partial_z\bld{u}+
370: \bld{\nabla\times}(\bld{u\times \omega})
371: \label{eq:eomo}
372: \ee
373: where
374: \be
375: \bld{\omega}\equiv \bld{\nabla\times u}
376: \label{eq:omdef}
377: \ee
378: is the vorticity of $\bld{u}$.
379:  Equation (\ref{eq:eomo}),  together with the inverse of equation (\ref{eq:omdef})
380: \be
381: \bld{u}=-\nabla^{-2}\bld{\nabla\times\omega} \label{eq:ominv} \ ,
382: \ee
383: form a complete set. 
384: 
385: Equation (\ref{eq:eomo}) implies that the total vorticity field
386: is frozen into the fluid, because it is equivalent to 
387: \be
388: \label{eq:omtot}
389: \partial_t\bld{\omega}_{\rm tot}=\bld{\nabla\times}(\bld{v\times\omega}_{\rm tot}) \ ,
390: \ee
391: where
392: \be
393: {\bld \omega}_{\rm tot}\equiv (2\Omega-q)\bld{\hat{z}}+\bld{\omega} \ 
394: \ee
395: is the total vorticity; note that
396:  $-q\bld{\hat{z}}$ is the vorticity of the unperturbed shear flow in the rotating frame, 
397: and hence $(2\Omega-q)\bld{\hat{z}}$ is the unperturbed vorticity in the non-rotating frame.
398: The vorticity-velocity picture is similar to MHD, where it is the
399: magnetic field that is frozen-in because it satisfies equation (\ref{eq:omtot})
400: in place of $\bld{\omega}_{\rm tot}$.  However, in MHD the velocity field has its own dynamical
401: equation, whereas in incompressible hydrodynamics it is determined directly from the vorticity field
402: via equation (\ref{eq:ominv}).
403: 
404: 
405: %\clearpage
406: %%%%%%%%%%%%%begin figure
407: \begin{figure*}
408:   \centering
409:   \includegraphics[width=.32\textwidth]{f1a.eps}
410:   \includegraphics[width=.33\textwidth]{f1b.eps}
411:   \includegraphics[width=.33\textwidth]{f1c.eps}  
412:  \includegraphics[width=.32\textwidth]{f1d.eps}
413:   \includegraphics[width=.32\textwidth]{f1e.eps}  
414:   \includegraphics[width=.33\textwidth]{f1f.eps}  
415:   \caption{{\bf Vortex Formation and Survival in a Short Box:} 
416:    Color depicts  $\omega_z$.
417:   The initial state is unstable to vertically symmetric ($k_z=0$) perturbations,
418:   and forms into a vortex.  But it is stable to 3D ($k_z\ne 0$) perturbations, and the evolution remains
419:   two-dimensional.  The bottom panels show  horizontal slices through the boxes in the upper
420:   panels, midway through the boxes.
421:    At time=150, the vortex has already
422:   formed. Only fluid with $\omega_z<-0.08$ is shown  in the middle panels to
423:   highlight the vortex, and to illustrate that surfaces of constant
424:   $\omega_z$ remain purely vertical.  At time=500,
425:   the vortex still survives.  Its amplitude is slowly decaying by viscosity, which
426:   acts on timescale=1130.  We set  $\Omega=1$ and $q=3/2$.
427:    The number of modes in the simulation is
428:   $n_x\times n_y\times n_z=64\times 64\times 32$, and the size of the simulation box is $(L_x,L_y,L_z)=({1\over 15},1,{1\over 2})$. In this figure, $L_z$ is to scale relative to
429:   $L_y$, but $L_x$ has been expanded by a factor of 5 for clarity.
430: }
431:   \label{fig:shortsim}
432: \end{figure*}
433: %%%%%%%%%%%%%end figure
434: 
435: 
436: 
437: 
438: 
439: %%%%%%%%%%%%%begin figure
440: \begin{figure*}
441:   \centering
442:   \includegraphics[width=.32\textwidth]{f2a.eps}
443:   \includegraphics[width=.33\textwidth]{f2b.eps}
444:   \includegraphics[width=.33\textwidth]{f2c.eps}  
445:   \includegraphics[width=.32\textwidth]{f2d.eps}
446:   \includegraphics[width=.32\textwidth]{f2e.eps}  
447:   \includegraphics[width=.33\textwidth]{f2f.eps}  
448:   \caption{{\bf Vortex Destruction in a Tall Box:} The setup 
449:   is identical to the short-box simulation of Figure \ref{fig:shortsim}, except that the height
450:   $L_z$ has been increased by a factor of 4, so that it now exceeds $L_y$. 
451:    The resulting evolution is dramatically different.
452:   The initial state is now unstable not only to 2D perturbations, but to 3D ones as well.  In the
453:   middle panels, surfaces of constant $\omega_z$ are warped, and the evolution is
454:   no longer vertically symmetric. 
455:     In the right panels, the flow looks turbulent.}
456:   \label{fig:tallsim}
457: \end{figure*}
458: %%%%%%%%%%%%%end figure
459: 
460: 
461: 
462: 
463: 
464: %%%%%%%%%%%%%begin figure
465: \begin{figure}
466:   %\centering
467:   \hspace{-1cm}\includegraphics[width=.55\textwidth]{f3.eps}
468:   \caption{{\bf Energy in the Short Box:}
469:   The three contributions to the energy budget, $E_{u^2}$, $\Delta E_{\rm shear}$,
470:   and $\Delta E_{\rm visc}$, are defined in equations (\ref{eq:e1})-(\ref{eq:e3}).
471:   $E_{u^2}$ initially decays, and then rises to a peak
472:   near $t\sim 200$ as nonaxisymmetric perturbations turn the axisymmetric
473:   mode into a vortex. 
474:   Subsequently, the vortex decays due to viscosity.  The spikiness
475:   of the evolution is due to the boundary conditions, as explained in the text.
476:   Also shown in the bottom panel is the error due to numerical effects,
477:   $\Delta E_{\rm error}$ (eq. [\ref{eq:ee}]).
478:   It is unlabelled because it is mostly obscured by $\Delta E_{\rm shear}$.
479:   But it is nearly equal to zero everywhere, showing that the code accurately
480:   tracks the components of the energy budget.
481:   }
482:   \label{fig:shorten}
483: \end{figure}
484: %%%%%%%%%%%%%end figure
485: 
486: %%%%%%%%%%%%%begin figure
487: \begin{figure}
488:   %\centering
489:   \hspace{-1cm}\includegraphics[width=.55\textwidth]{f4.eps}
490:   \caption{{\bf Energy in the Tall Box:} The initial evolution is almost the same
491:   as that seen in the short box (Fig. \ref{fig:shorten}).  But 3D perturbations are unstable and force
492:   the destruction of the vortex.   In the time interval $300\lesssim t\lesssim 600$,
493:   while the initial axisymmetric disturbance decays in a turbulent-like state,  the value of $E_{u^2}$ is significantly larger than its
494:   initial value, and $\Delta E_{\rm shear}$ rises nearly linearly in time, corresponding
495:   to nearly constant outwards transport of angular momentum in a disk.  
496:   The contribution of numerical errors to the time-integrated energy budget, $\Delta E_{\rm error}$ (eq. [\ref{eq:ee}]),
497:   remains small throughout.
498:   }
499:   \label{fig:tallen}
500: \end{figure}
501: %%%%%%%%%%%%%end figure
502: %\clearpage
503: 
504: 
505: \section{Two Pseudospectral Simulations}
506: \label{sec:pseudo}
507: 
508: 
509: 
510: The pseudospectral code is described in detail in the Appendix of L07.  
511: It solves the velocity-pressure equations of motion with an explicit viscous term
512: \be
513: \nu\nabla^2{\bld{u}}
514: \label{eq:visc}
515: \ee
516: added to equation (\ref{eq:eom2}).
517: (In L07, we did not include this term because we
518: only considered inviscid flows.)
519: The equations are solved 
520: in Fourier-space by decomposing fields into
521:  spatial Fourier modes whose wavevectors
522: are advected by the background flow $-qx\bld{\hat{y}}$.
523: As a result, the boundary conditions are periodic in the $y$ and $z$ dimension, and
524: ``shearing periodic'' in $x$.
525: Most of our techniques are standard \citep[e.g.,][]{MG01,Rogallo81,BM06}.  One exception is 
526: our method for
527: remapping 
528: highly trailing wavevectors into highly leading ones, which is both simpler
529: and more accurate than the usual method.  In addition, our remapping 
530: does not introduce power into leading modes, because a mode's amplitude
531: has always been set to zero before the remap.
532: The code was extensively tested on 2D flows in L07.
533: A number of rather stringent 3D tests are performed in this  paper.  We shall show 
534: that the code correctly reproduces the linear evolution of 3D modes (\S \ref{sec:lin}), as
535: well as the nonlinear coupling between them (\S \ref{sec:nonlin}). We also show in 
536: the present section
537: that it tracks
538: the various contributions to the energy budget, and that the sum
539: of the contributions vanishes to high accuracy.
540: 
541: Figures \ref{fig:shortsim}-\ref{fig:tallen} show results 
542:   from two pseudospectral 
543: simulations.  One simulation illustrates the formation
544: and survival of a vortex, and the other illustrates
545:  vortex destruction.
546: In the first (the ``short box''),
547:  the number of Fourier modes  used 
548: is $n_x\times n_y\times n_z=64\times 64\times 32$, 
549: and the simulation box has dimensions
550: $L_x={1\over 15}$, $L_y=1$, and $L_z={1\over 2}$.  
551: In the second (the ``tall box''), the setup is identical,
552: except that it has $L_z=2$ instead of $1/2$.
553: Both simulations are initialized by setting
554: \be
555: \omega_z\Big\vert_{t=0}=-0.1\cos\left({2\pi\over L_x}x\right) \ .
556: \label{eq:initz}
557: \ee
558: In addition,  small perturbations are added to long-wavelength modes.
559: Specifically, labelling the wavevectors as
560: \be
561: (k_x,k_y,k_z)=2\pi\left({j_x\over L_x},{j_y\over L_y},{j_z\over L_z}  \right) \ ,
562: \label{eq:kj}
563: \ee 
564: with integers $(j_x,j_y,j_z)$,
565: we select all modes that satisfy
566:  $|j_x|\leq 3$, $|j_y|\leq 3$,
567: and $|j_z|\leq 3$, and set the Fourier amplitude of their $\omega_z$  to 
568:  $10^{-4}e^{i\phi}$, where $\phi$ is a random phase. 
569: But
570: we exclude the  $(j_x,j_y,j_z)=(0,0,0)$ mode, as well as 
571: $(j_x,j_y,j_z)=(\pm 1,0,0)$, which is  given by equation (\ref{eq:initz}).
572: Finally, we set $\Omega=1$, $q=3/2$, $\nu=10^{-7}$, and integration 
573: timestep $dt=1/30$.
574: 
575: With our chosen initial conditions, the mode given by equation (\ref{eq:initz})
576: is nonlinearly unstable to vertically symmetric ($k_z=0$) perturbations, and hence
577: it tends to wrap up into a vortex. From the approximate criterion for instability 
578:  (eq. [\ref{eq:2dinst}]),
579: we see that to illustrate the wrapping up into a vortex
580: of a mode with a small amplitude, one must make the simulation box
581: elongated in the $y$-direction relative to the $x$-scale of the mode in equation
582: (\ref{eq:initz}).
583: 
584: 
585: In the short box (Fig. \ref{fig:shortsim}), the evolution proceeds just as it would in two dimensions. 
586: The initial mode indeed wraps up into a vortex, and 
587:  the evolution remains vertically
588: symmetric throughout.  Once formed, the vortex can live for ever in the absence
589: of viscosity.
590: But in our simulation, there is a slow viscous decay.  The timescale for viscous decay
591: across
592: the width of the vortex is
593: $\sim 1/\nu k_x^2=1130$, taking $k_x=2\pi/L_x$.
594: 
595: In the tall box (Fig. \ref{fig:tallsim}), the evolution is dramatically different.  
596: In this case, the initial state is unstable not only to 2D  perturbations,
597:  but to 3D $(k_z\ne 0)$
598: ones as well. In the middle panel of that figure, we see 
599: that instead of forming a vertically symmetric vortex as in the short box,
600:  surfaces of constant $\omega_z$ are 
601: warped. 
602: By the third panel, the flow looks turbulent.
603: 
604: 
605: Figures \ref{fig:shorten}-\ref{fig:tallen} shows the evolution of the
606:  energy in these simulations.
607: Projecting ${\bld u}$
608: onto
609: the Navier-Stokes equation (eq. [\ref{eq:eom2}]  with viscosity included),
610: and spatially averaging, we arrive at the energy equation
611: \be
612: {d\over dt}
613: {\left\langle u^2
614: \right\rangle\over 2}=q
615: \langle u_xu_y\rangle 
616: +\nu\langle
617: {\bld u\cdot}\nabla^2\bld{u}
618: \rangle \ ,
619: \ee
620: after applying the shearing-box boundary conditions, 
621: where angled brackets denote a spatial average.
622: The time integral of this equation is
623: \be
624: E_{u^2}-E_{u^2}\vert_{t=0}=\Delta E_{\rm shear}+\Delta E_{\rm visc} \ , 
625: \label{eq:inte}
626: \ee
627: where
628: \beqn
629: E_{u^2}&\equiv& {\left\langle u^2
630: \right\rangle\over 2}
631: \label{eq:e1}
632:  \\
633: \Delta E_{\rm shear}&\equiv&
634: q\int_0^t \langle u_xu_y\rangle dt' \\
635: \Delta E_{\rm visc}&\equiv&
636: \nu\int_0^t
637: \langle
638: {\bld u\cdot}\nabla^2\bld{u}
639: \rangle dt' \ . \label{eq:e3}
640: \eeqn
641: The pseudospectral code records  each of these terms, and
642: Figure \ref{fig:shorten} shows the result in the short box 
643: simulation.  At very 
644: early times, $E_{u^2}$ decays from its initial value due to viscosity. 
645: At the same time, the small vertically symmetric perturbations are growing exponentially, 
646: and they start to give order unity perturbations by $t\sim 150$, by which time
647:  a vortex has been formed (Fig. \ref{fig:shortsim}).  
648:  As time evolves, $E_{u^2}$ gradually decays due to viscosity on the viscous
649:  timescale $=1130$.   The evolution  is very spiky.  
650:  We defer a discussion of this spikiness to the end of this section.
651:  
652:  Figure \ref{fig:tallen} shows the result in the tall box. The early evolution of 
653:  $E_{u^2}$ is similar to that seen in the short box.  Both start with the same $E_{u^2}$,
654:  and an initial period of viscous decay is interrupted by exponentially growing 
655:  perturbations. But in the tall box, not only are vertically symmetric modes
656:  growing, but modes with $k_z\ne 0$  are growing as well. By $t\sim 150$, there is a distorted
657:  vortex that subsequently decays into a turbulent-like state.  The energy $E_{u^2}$
658:  rises to a value significantly larger than its initial one, and it continues to rise
659:  until $t\sim 600$, when it starts to decay.  Throughout the time interval
660:  $300\lesssim t\lesssim 600$, $\Delta E_{\rm shear}$ rises nearly linearly in 
661:  time, showing that $\langle u_xu_y\rangle$ is positive and  nearly constant.
662:  
663: It is intriguing that  $\langle u_x u_y\rangle$ is positive for hundreds of orbits,
664: because
665:  it suggests that decaying vortices might transport
666: angular momentum outwards in disks and hence drive accretion.
667: Understanding the level of the turbulence,
668:  its lifetime, 
669:  and its nature are  topics for future work.
670:  Here we merely address the sign of $\langle u_xu_y\rangle$.
671: The quantity  $\langle u_xu_y\rangle$ is the 
672: flux of $y$-momentum in the $+x$-direction (per unit mass and spatially averaged).
673: It corresponds to the flux of angular momentum
674: in a disk.  A positive
675: $\langle u_xu_y\rangle$ implies an outwards flux of angular momentum, 
676: as is required to drive matter inwards in an accretion disk.
677:  (Even though the shearing box cannot distinguish
678:  inwards from outwards,
679: the sign of the angular momentum within a box depends on which side of the shearing
680: box one calls inwards.  Therefore, outwards transport of (positive) angular momentum
681: is well-defined in a shearing box.)
682: In the shearing box, 
683: any force that tends to diminish the background shear
684: flow $-qx\bld{\hat{y}}$ necessarily transports $y$-momentum in the $+x$-direction.
685: Hence the fact that $\langle u_xu_y \rangle >0$ in Figure \ref{fig:tallen} shows
686: that the turbulence exerts forces that resist the background shear, as one 
687: might
688: expect on physical grounds.
689: One can also understand why   $\langle u_xu_y \rangle >0$ from
690: energy considerations. 
691: Since  $\Delta E_{\rm visc}<0$, as may be seen explicitly by
692: an integration by parts, i.e. $\langle \bld{u\cdot}\nabla^2\bld{u} \rangle
693: =-\sum_{i,j}\langle(\partial_j u_i)^2\rangle$, equation (\ref{eq:inte})
694: may be rearranged to read
695: \be
696: \Delta E_{\rm shear} = \left\vert \Delta E_{\rm visc}\right\vert
697: +E_{u^2}-E_{u^2}\vert_{t=0} \ .
698: \label{eq:ee1}
699: \ee
700: If  the turbulence reaches a steady state---as it approximately does in 
701: Figure \ref{fig:tallen} during the time interval
702: $300\lesssim t\lesssim 600$---then the last two terms in the 
703: above equation are nearly constant, whereas $\left\vert \Delta E_{\rm visc}\right\vert$
704: increases linearly with time.   Hence $\Delta E_{\rm shear}$ must also increase.
705: The fact that energy dissipation implies outwards transport of angular momentum
706: is a general property of accretion disks \citep[e.g., ][]{LP}.   Since turbulence always
707: dissipates energy, it must also transport angular momentum outwards.  
708: However, this argument can be violated if  an external energy source drives
709: the turbulence,
710: in which case one would have to add this energy to the left-hand side of equation (\ref{eq:ee1}).
711: For example, the simulations of \cite{SB96} show that convective disks can transport angular momentum
712: inwards when an externally imposed heat source drives the convection. 
713: 
714:  
715:  
716:  
717: Also shown in the bottom panels of Figures \ref{fig:shorten}-\ref{fig:tallen}
718: is the integrated energy error
719: \be
720: \Delta E_{\rm error}\equiv \Delta E_{\rm shear}+\Delta E_{\rm visc}+E_{u^2}\vert_{t=0}-E_{u^2}
721: \label{eq:ee}
722: \ee
723:  due to 
724: numerical effects, which is seen to be small. 
725:  (In Figure \ref{fig:shorten}, $\Delta E_{\rm error}$ is not labelled because
726: the curve is mostly obscured by $\Delta E_{\rm shear}$; it can be seen near $t\sim 200$,
727: and is everywhere very nearly equal to zero.)
728: The fact that $\Delta E_{\rm error}$ nearly vanishes throughout the simulations is 
729: not guaranteed by the pseudospectral algorithm.  Rather, we have chosen
730: $\nu$ to be large enough that the algorithm introduces negligible  error into the
731: energy budget.  To be more precise, at each timestep in the pseudospectral code,
732: modes that have $|j_x|>n_x/3$ or $|j_y|>n_y/3$ or $|j_z|>n_z/3$, where $j_{x,y,z}$
733: are defined via equation (\ref{eq:kj}), have their amplitudes set to zero (``dealiased'').
734: This introduces an error that is analogous to
735: grid error in grid-based codes.  By choosing  $\nu$ to be sufficiently large, it
736: is the explicit viscosity that forces modes
737: with large $k$ to have small amplitudes, in which case
738: the dealiasing procedure has little effect on the dynamics.
739: Increasing the resolution $n_x\times n_y\times n_z$
740:  would allow a smaller $\nu$ to be chosen---implying
741: a larger effective Reynolds number---while keeping the energy error small.
742: 
743: 
744: 
745: The curves of $E_{u^2}$ show
746:  sharp narrow spikes every time interval $\Delta t=10$, with width $\sim 1$.
747:  Similar spikes have been seen in other simulations \citep{UR04,SSG06}, but
748:  they are stronger and narrower in our simulations because our simulation box
749:  is elongated.  These spikes are due to the
750:  shearing-periodic boundary conditions.  It is perhaps simplest to understand
751:  them by following the evolution in $k$-space, as we shall do in 
752:   \S \ref{sec:nonlin} (see also L07).  But for now, we explain their origin
753:    in real-space.
754: By the  nature of shearing-periodic boundary conditions,   associated with the simulation 
755:  box centered  at $x=0$ are  ``imaginary boxes'' centered
756:     at $x=jL_x$ with
757:  integer $j=\pm 1,\pm 2,\cdots$.  These imaginary boxes completely tile the $x-y$ plane,
758:  and each
759:   contains a virtual copy of the conditions inside the simulation box.
760:  The boxes move relative to the simulation box in 
761:  the $y$-direction, with the speed of the mean shear at the center of each box,
762:  $-qjL_x$.
763:  Therefore, in  the  time interval 
764:  $\Delta t = L_y/(qL_x)=10$, all the boxes line up.  When this happens, 
765:  the velocity field $\bld{u}$ that is induced by the vorticity within all the boxes 
766:  (via eq. [\ref{eq:ominv}]) becomes large, because all the boxes reinforce each other,
767:  and therefore $E_{u^2}$ exhibits a spike.
768: Even though the shearing-periodic boundary conditions that we use are 
769: somewhat artificial, 
770: we are confident that using more realistic open boundary conditions would not
771: affect the main results of this paper---and particularly not the stability of
772:  axisymmetric modes to 3D perturbations.
773: In L07, where we considered 2D dynamics, 
774: we investigated both open and shearing-periodic boundary conditions, and
775: showed explicitly that both give similar results.
776: We also feel that the boundary conditions likely do not affect the level
777: and persistence of the ``turbulence'' seen in Figure \ref{fig:tallen}.  However, 
778: this is less certain.  Future investigations should more carefully address the role
779: of boundary conditions.  
780: 
781: 
782: \section{Linear Evolution}
783: \label{sec:lin}
784: 
785: 
786: %\clearpage
787: %%%%%%%%%%%%%begin figure
788: \begin{figure*}
789:   \centering
790:   \includegraphics[width=1\textwidth]{f5.eps}
791:   \caption{{\bf Evolution of Wavevectors:} Modes have constant $k_y$ and $k_z$, 
792:   and $k_x=qtk_y+$const.  The three spheres depict modes that
793:    play important roles in nonlinear
794:   instability.  The mode at $(1,0,0)$ does not move in $k$-space.  The other two modes
795:   are swinging modes that are depicted in the leading phase of their swing. They will
796:   become trailing after crossing through the radially symmetric plane. The mode
797:   crossing through $(1,-1,0)$ is responsible for 2D instability that forms vortices.
798:   The one crossing through $(1,-1,1)$ is responsible for 3D instability that destroys
799:   vortices.}
800:   \label{fig:kspace}
801: \end{figure*}
802: %%%%%%%%%%%%%end figure
803: %\clearpage
804: 
805: In the remainder of this paper, we develop a theory explaining the
806: stability of vortices seen in the above numerical simulations.  
807: We first consider the linear evolution of individual modes, 
808: and then proceed to show how nonlinear coupling between
809: linear modes can explain vortex stability.
810: 
811: 
812: The linear evolution has been considered previously \citep{AMN05,JG05,BH06}.
813: Only two aspects of our treatment are new.
814: First, we give the solution in terms
815: of variables that allow the simple reconstruction of
816: the full vectors $\bld{\omega}$ and $\bld{u}$.
817: And second, we give the analytic expression 
818:   for matching a leading mode 
819: onto 
820: a trailing mode that is
821: valid for all $k_y$ and $k_z$,
822: 
823: The linearized equation of motion is (eq. [\ref{eq:eomo}])
824: \be
825: (\partial_t-qx\partial_y)\bld{\omega}=-q\bld{\hat{y}}\omega_x+(2\Omega-q)\partial_z\bld{u} \ .
826: \label{eq:elin}
827: \ee
828:  A single mode may be written as
829:  \be
830: \bld{\omega}(\bld{x},t)=\bld{\hat{\omega}}(\bld{k_0},t)e^{i 
831: \left[
832: {\bld{k}}
833: \left(
834: {\bld{k_0}},t
835: \right)
836: \right]
837: {\bld\cdot}
838: \bld{x}
839: } \ , \label{eq:omk}
840: \ee
841: where $\bld{k_0}$ is a constant vector that denotes the wavevector at time $t=0$, and
842: the  wavevector $\bld{k}=\bld{k}(\bld{k_0},t)$ has components
843: \beqn
844: k_y &=& k_{0y}={\rm const} \\
845: k_z &=& k_{0z} ={\rm const}\\
846: k_x &=& {k_{0x}}+qtk_y\ne {\rm const} \ ,
847: \eeqn
848: so that upon insertion into equation (\ref{eq:elin}), the time-derivative of the exponential
849: cancels the term $-qx\partial_y{\bld{\omega}}$.
850: The velocity field induced by such a mode 
851:  is (eq. [\ref{eq:ominv}])
852: \beqn
853: \bld{u}(\bld{x},t)=
854: \bld{\hat{u}}(\bld{k_0},t)
855: e^{i \bld{k}\left[
856: \left(
857: {\bld{k_0}},t
858: \right)
859: \right]
860: {\bld\cdot}
861: \bld{x}} 
862: \label{eq:uk}
863: \eeqn
864: where
865: \be
866: \bld{\hat{u}}=
867: i{\bld{k\times\hat{\omega}}\over k^2}
868: \label{eq:udef}
869: \ee
870: 
871: Figure \ref{fig:kspace} sketches the evolution of wavevectors. 
872: Axisymmetric modes ($k_y=0$) do not move in $k$-space, as depicted by
873: the sphere at $(1,0,0)$ in Figure \ref{fig:kspace}.
874:  ``Swinging modes'' have $k_y\ne 0$, and their $k_x$ is time-dependent.  Their
875: fronts of constant phase are advected by the background shear.  Swinging modes
876: with $k_x/k_y<0$, as depicted by the two spheres 
877: near $(1,-1,0)$ and $(1,-1,1)$ in Figure \ref{fig:kspace},
878:  have phasefronts tilted into the background shear, i.e., they are leading modes.
879:    As time evolves, 
880: the shear first swings their $k_x$ through $k_x=0$, at which point their phasefronts
881: are radially symmetric.
882: Subsequently, they become
883: trailing modes ($k_x/k_y>0$), and approach alignment with the azimuthal direction
884:  ($k_x/k_y\rightarrow \infty$).  
885: 
886: 
887: We turn now to the evolution of the Fourier amplitudes.
888: In the remainder of this paper we drop the hats
889: \beqn
890: \bld{\hat\omega}\rightarrow \bld{\omega} \ ,
891: \ \ 
892: \bld{\hat u}\rightarrow \bld{u} \ .
893: \eeqn
894: To distinguish real-space fields, we shall explicitly write their
895: spatial dependence, e.g. $\bld{\omega}(\bld{x})$.
896: 
897: Because $\bld{\omega}(\bld{x})$ is divergenceless, $\bld{{\omega}}$
898: only has two degrees of freedom, which we select to be
899:  ${\omega}_x$ and
900: \beqn
901: {\omega}_{yz}\equiv 
902: {\bld{\hat{x}\cdot}
903: \left({{\bld{k\times{\omega}}}}\right)\over k_{yz}} = 
904: \left\{
905:  \begin{array}
906:  {r@{\quad,\quad}l}
907:  -{\omega}_y & {\rm if\ } k_y=0 \\
908: {\omega}_z & {\rm if\ } k_z=0
909:  \end{array}
910:  \right.
911: \eeqn
912: where
913: \be
914: k_{yz}\equiv \sqrt{k_y^2+k_z^2} 
915: \ee
916: Our variable $\omega_{yz}$ is proportional to the variable $U$ of
917: \cite{BH06}. Adopting $\omega_x$ as the second degree of freedom
918: enables
919:   the full vectors to 
920: be reconstructed  as
921: \beqn
922: \bld{{\omega}} &=& 
923: -{\omega}_x
924: {\bld{k\times}(\bld{k\times \hat{x}})\over k_{yz}^2}
925: -{{\omega}_{yz}}{\bld{k\times\hat{x}}\over k_{yz}}
926: \label{eq:omrec}
927: \\
928: \bld{{u}} &=& -i\omega_{yz}{\bld{k\times}(\bld{k\times \hat{x}})\over k^2 k_{yz}}+
929: i\omega_x {\bld{k\times\hat{x}}\over k_{yz}^2}
930: \label{eq:urec}
931: \eeqn
932: 
933: The linearized equation  (\ref{eq:elin}) is expressed in terms of these degrees of freedom as
934: \be
935: {k_{yz}\over qk_y}{d\over dt}
936: \left(\begin{array}{c}{\omega}_x \\ {\omega}_{yz}\end{array}\right)
937: =\beta{\Omega\over \kappa}
938: \left(\begin{array}{cc}0 & -{1\over 2}{\kappa^2\over\Omega^2}{1\over 1+\tau^2} \\ 2 & 0\end{array}\right)
939: \left(\begin{array}{c}{\omega}_x \\ {\omega}_{yz}\end{array}\right) \ ,
940: \label{eq:linlim}
941: \ee
942: after introducing the  epicyclic frequency,
943: \beqn
944: \kappa \equiv \sqrt{2\Omega(2\Omega-q)}  \ ,
945: \eeqn
946: with $\kappa=\Omega$ in a Keplerian disks, and
947: \beqn
948: \tau&\equiv& {k_x\over k_{yz}} \\
949: \beta&\equiv&{\kappa\over q}{k_z\over k_y} \ .
950: \eeqn
951: As long as $k_y\ne 0$,
952: $\tau$ varies in time through its dependence on
953: $k_x=k_{0x}+qtk_y$.
954: 
955: For axisymmetric modes ($k_y=0$), $\tau$ is constant and
956: \be
957: {d^2\over dt^2}\omega_{yz}+\kappa^2{k_z^2\over k_x^2+k_z^2}\omega_{yz}=0 \ ,
958: \label{eq:axi3d}
959: \ee
960: the solution of which is sinusoidal with frequency $\kappa k_z/\sqrt{k_x^2+k_z^2}$.
961: Axisymmetric modes with phasefronts aligned with the plane of the disk ($k_x=k_y=0$) have
962: in-plane fluid velocities, and they oscillate at
963: the epicyclic frequency of a free test particle, $\kappa$.
964:   But axisymmetric modes with tilted
965: phasefronts have slower frequencies, because fluid pressure causes deviations from
966: free epicycles.  In the limit of vertical axisymmetric phasefronts ($k_z=k_y=0$), the effects of
967: rotation disappear entirely, and this zero-frequency mode merely alters the mean shear
968: flow's velocity profile.
969: 
970: 
971: For swinging modes
972: ($k_y\ne 0$),  it is convenient to employ $\tau$ as  the time variable.
973: Since 
974: \be
975: {k_{yz}\over qk_y}{d\over dt}={d\over d\tau} \ ,
976: \ee
977: we have
978: \be
979: {d^2\over d\tau^2}\omega_{yz}+{\beta^2\over 1+\tau^2}\omega_{yz}=0 \ .
980: \label{eq:bh}
981: \ee
982: \citep{BH06}.
983: Figure \ref{fig:lin} plots numerical solutions of this equation, and shows that it matches 
984: the output from the pseudospectral code, as well as the analytic theory described below.
985: Given $\omega_{yz}$, it is trivial to construct $\bld{\omega}$ and $\bld{{u}}$
986: from
987: \be
988: \omega_x={\kappa\over 2\beta\Omega}{d\omega_{yz}\over d\tau}
989: \label{eq:oxeq}
990: \ee
991: and equations (\ref{eq:omrec}) and (\ref{eq:urec}).
992: 
993: %\clearpage
994: %%%%%%%%%%%%%begin figure
995: \begin{figure} %old \begin{figure*} .... \end{figure*}
996:   %\centering
997:  \hspace{-1.cm} \includegraphics[width=.55\textwidth]{f6.eps}
998:   \caption{{\bf Linear Evolution of Mode Amplitudes for Three Values 
999:   of $\beta$:} Time runs from right to left.
1000:   Solid curves show the exact, numerically integrated solution of equation (\ref{eq:bh}).
1001:   The initial value of $d\omega_{yz}/d\tau$ was chosen so that $\omega_B=0$ initially
1002:   (eq. [\ref{eq:powerlaw}]). 
1003:     Dashed lines show the analytic solution (eq. [\ref{eq:powerlaw}])
1004:   with constant $\omega_A$ and $\omega_B=0$ for $\tau>0$; while for $\tau<0$, the
1005:   normal mode amplitudes are set to
1006:   different constants that are given by equation (\ref{eq:trans}). 
1007:   We exclude the domain $|\tau|<1$ from the dashed curve, because the analytic
1008:   approximation does not apply there.
1009:    Circles show output
1010:   from the pseudospectral code, integrated with a timestep $dt=1/15$ and with the viscosity
1011:   set to zero. 
1012:     }
1013:   \label{fig:lin}
1014: \end{figure}
1015: %%%%%%%%%%%%%end figure
1016: %\clearpage
1017: 
1018: For highly leading or trailing modes ($|\tau|\gg 1$), equation (\ref{eq:bh}) has simple power-law solutions,
1019: \beqn
1020: \omega_{yz}&=&\omega_A|\tau|^{1-\delta\over 2}+\omega_B|\tau|^{1+\delta\over 2} \ , 
1021: \ \ |\tau|\gg 1 \label{eq:powerlaw}
1022: \eeqn
1023: \citep{BH06}, where  $\omega_A$ and $\omega_B$ are constants that
1024: we shall call the ``normal-mode'' amplitudes, and
1025: \be
1026: \delta \equiv \sqrt{1-4\beta^2} \ ,
1027: \ee
1028: which is imaginary for $|\beta|>1/2$.
1029: As a mode's wavevector evolves along a line in $k$-space, its
1030: amplitude is oscillatory if this line is much closer to the $k_z$ axis
1031: than to the $k_y$ one, and  non-oscillatory if the converse is true.
1032: The transition occurs at $|\beta|=1/2$.     This
1033: behavior may be understood as a competition between shear and epicyclic oscillations.  
1034: The timescale for $k_x$ to change by an order-unity factor due to the
1035: shear is $t_{\rm shear}\sim |k_x/\dot{k}_x|=|k_x/qk_y|$, and
1036: the timescale for epicyclic oscillations of axisymmetric modes
1037:  is $t_{\rm epi}\sim \kappa^{-1}|k_x/k_z|$ for 
1038: $|k_x|\gg |k_z|$.  
1039: Therefore $|\beta|\sim t_{\rm shear}/t_{\rm epi}$, and
1040: when $|\beta|\gg 1$ the epicyclic time is shorter and
1041: so the mode's amplitude oscillates as its wavevector is slowly advected by the shear.  But
1042:  when $|\beta|\ll 1$ the shear changes the wavevector faster
1043: than the amplitude can oscillate.
1044: 
1045: 
1046: The solution (\ref{eq:powerlaw}) breaks down in mid-swing.  
1047: As a swinging wave changes from leading to trailing, its ``normal-mode
1048: amplitudes''  change abruptly  on the timescale that $\tau$ changes from
1049: $\pm 1$ to $\mp 1$ via
1050: \be
1051: \left(\begin{array}{c} \omega_A \\ \omega_B\end{array}\right)_{\rm trail}
1052: =
1053: \left(\begin{array}{cc}T_{AA} & T_{AB} \\T_{BA} & T_{BB}\end{array}\right)
1054: \left(\begin{array}{c} \omega_A \\ \omega_B\end{array}\right)_{\rm lead}
1055: \label{eq:trans}
1056: \ee
1057: where the transition matrix has components
1058: \beqn
1059: T_{AA}&=&-T_{BB}=\csc\left({\delta\pi/ 2}\right) \\
1060: T_{BA}&=&
1061: -{\cot(\delta\pi/2)^2\over T_{AB}}
1062: =-2^{\delta+1}{1\over\delta}{1-\delta\over 1+\delta}
1063: {\Gamma(1+\delta/2)^2\over \Gamma(1/2+\delta/2)^2}
1064: \eeqn
1065: and determinant $=-1$, and hence is its own inverse.
1066:  The components are complex when $|\beta|>1/2$.
1067: To derive these components, we took advantage of the fact that equation (\ref{eq:bh}) has
1068: hypergeometric solutions   \citep{JG05,BH06}, and matched
1069: these onto the normal-mode solution given above.  We omit
1070: the unenlightening details.  
1071: 
1072: 
1073: \section{Nonlinear Evolution: Formation and Destruction of Vortices}
1074: \label{sec:nonlin}
1075: 
1076: 
1077: \subsection{Qualitative Description}
1078: 
1079: The instability that destroys  vortices is a generalization of 
1080: the one that forms them. 
1081:  We review here how vortices form,
1082: before describing the instability that destroys them.  In \S \ref{subsec:stab},
1083: we make this description quantitative. 
1084: 
1085: 
1086: Vortices form out of a nonlinear instability that involves vertically symmetric ($k_z=0$)
1087: modes. (See L07 for more details of the 2D dynamics than are presented here.) 
1088:   Consider the two vertically symmetric modes shown in Figure \ref{fig:kspace}: the
1089: ``mother'' mode at $(1,0,0)$ and the ``father'' mode that is depicted crossing through $(1,-1,0)$.
1090: Triplets of integers 
1091: $(j_x,j_y,j_z)$
1092: label  values of wavevectors $(k_x,k_y,k_z)$
1093: (for example, via [\ref{eq:kj}]).
1094: The mother is both axi- and vertically-symmetric, and the father is a leading swinging mode.
1095: 
1096: 
1097: As the father swings through radial symmetry, i.e. as it crosses through
1098: the point $(0,-1,0)$, its
1099:  velocity
1100: field  is
1101: strongly amplified by the background shear.  This can be seen from  
1102:  \S \ref{sec:lin}, which shows that swinging modes with $k_z=0$ have
1103:  $\omega_{yz}$=const., and hence $u_x=i(\omega_{yz}/k_{yz})/(1+\tau^2)$, which
1104:  becomes largest when $\tau$ crosses through 0.
1105: When the father is near the peak of its transient amplification ($|\tau|\lesssim 1$), 
1106: it couples most strongly with the mother, and they produce a ``son''
1107: near $(1,-1,0)=(1,0,0)+(0,-1,0)$.  The son will then swing through radial symmetry
1108: where it will  couple (oedipally) with the mother to produce a grandson near $(1,-1,0)$,
1109: which can repeat the cycle.
1110: We summarize this 2D instability feedback loop  as
1111: \beqn
1112: {\rm linear\ amplification:}& (1,-1,0)\rightarrow (0,-1,0)  \nonumber \\
1113: {\rm nonlinear\ coupling:}& (0,-1,0)+(1,0,0)\rightarrow(1,-1,0) \nonumber
1114: \eeqn
1115: The criterion for instability is simply that the amplitude of the son's $\omega_{yz}$ be
1116: larger than that of the father.  
1117: As shown in L07, if instability is triggered, its nonlinear outcome 
1118: in two dimensions is a long-lived vortex.
1119: 
1120: The three-dimensional instability that is responsible for destroying vortices
1121: is a straightforward generalization.   
1122: The mother mode is still at $(1,0,0)$, but now the father mode starts near 
1123: $(1,-1,1)$.
1124: Symbolically, the feedback loop is
1125: \beqn
1126: {\rm linear\ amplification:}& (1,-1,1)\rightarrow (0,-1,1)  \nonumber \\
1127: {\rm nonlinear\ coupling:}& (0,-1,1)+(1,0,0)\rightarrow(1,-1,1) \nonumber
1128: \eeqn
1129: The 2D instability described above is just a special case of this 3D one in the limit that $k_z=0$.
1130: In general, the stability of a mother  mode at $(1,0,0)$ with given $k_x=\bar{k}_x$ and 
1131: ${\bld\omega}=\bld{\bar{\omega}}$ 
1132: depends on the $k_y$ and $k_z$ of the 
1133:  father-mode perturbations (as well as on the parameters
1134: $q$ and $\Omega$). Which $k_y$ and $k_z$ are accessible in turn depends
1135:  on the dimensions $L_y\times L_z$ of the simulation box---or equivalently,
1136: on the circumferential distance around a disk and the scale-height.
1137: In \S \ref{subsec:stab}, we map out quantitatively  the  region in the
1138: $k_y-k_z$ plane that leads to instability.
1139: For now, it suffices to note that the unstable region has $|k_y|\lesssim |\bar{k}_x \bar{\omega}|/q$ and
1140: $|k_z|\lesssim |k_y|$.  
1141: We conclude that a given mother mode suffers one of three
1142: possible fates, depending on $L_y$ and $L_z$.
1143: \begin{enumerate}
1144: \item
1145: If $L_y$ is less than a critical value 
1146: ($\sim q/|\bar{\omega}\bar{k}_x|$), then the mother mode is
1147: stable to all perturbations. 
1148: \item
1149:  If $L_y$ is larger than this critical value, then the 
1150: mother mode is unstable to vertically symmetric ($k_z=0$) perturbations;  if in addition $L_z$ is sufficiently
1151: small that all modes with $k_z\ne 0$ are stable, then the  mother mode turns
1152: into a long-lived vortex (Figure \ref{fig:shortsim}).
1153:  \item If both $L_y$ and $L_z$ are sufficiently large, the mother mode
1154: is unstable both to vertically symmetric and to 3D perturbations.  When this happens, the mother
1155: starts to form a vortex, but this vortex is 3D-unstable.  The result is turbulence (Figure \ref{fig:tallsim}).
1156: \end{enumerate}
1157: There is also a possibility that is intermediate between numbers 2 and 3: if the conditions
1158: described in number 2 hold, the essentially 2D dynamics that results
1159: can nonlinearly produce new mother modes that are unstable to 3D perturbations.
1160: In this paper, we shall not consider this possibility further, since it did not occur in
1161: the pseudospectral simulations of \S \ref{sec:pseudo}.  
1162: We merely note that in our simulations
1163: of this possibility (not presented in this paper), we found that when the new mother modes
1164: decayed, they also destroyed the original mother mode.
1165: 
1166: 
1167: \subsection{The Stability Criterion}
1168: \label{subsec:stab}
1169: 
1170: %\clearpage
1171: %%%%%%%%%%%%%begin figure
1172: \begin{figure}
1173:   %\centering
1174:   \hspace{-1cm}\includegraphics[width=.55\textwidth]{f7.eps}
1175:   \caption{{\bf Nonlinear Evolution of 3D Instability:} Time runs from right to left.  In the two left panels, lines show numerical 
1176:   solutions of equations (\ref{eq:lin2}) and (\ref{eq:soneq}), as well as the grandson's
1177:   equation. Also shown as circles
1178:   are  the output from a pseudospectral simulation, showing excellent agreement
1179:   with the ``exact'' solutions.    The
1180:   following parameters have been chosen: $\Omega=1$, $q=3/2$, $\bar{\omega}=0.005$, $\bar{k}_x=2\pi\cdot 15$,
1181:   $k_y=-2\pi/30$, $k_z=0.45|k_y|\Rightarrow \beta=-0.3$. The small disagreement
1182:   between pseudospectral and exact solutions  for $\omega_x'$ at
1183:   $\tau<-2\bar{\tau}$ is due to the conjugate modes that, for simplicity,  we have not included in equations 
1184:   (\ref{eq:lin2}) and (\ref{eq:soneq}); see footnote \ref{foot:cc}.   The two right panels show the mode amplitudes, defined via equation (\ref{eq:sonnm})
1185:   for the son, and similarly for the father and grandson. Although these two right
1186:   panels contain the same information as the left ones, they are helpful
1187:   in constructing the analytic form of the growth factor $\chi$ (see Appendix).  With the parameters
1188:   chosen for this figure, equation (\ref{eq:amp}) predicts $\chi=-2.2$ for the amplification
1189: factor between successive generations, in agreement with that seen in the figure.
1190:   }
1191:   \label{fig:nonlin}
1192: \end{figure}
1193: %%%%%%%%%%%%%end figure
1194: %\clearpage
1195: 
1196: 
1197: To quantify the previous discussion, we choose an initial state as
1198: in Figure \ref{fig:kspace}, with the mother
1199: mode at $(1,0,0)$ and the father a leading mode crossing through $(1,-1,1)$. 
1200:  The son mode, not depicted in the figure,
1201:  is initially
1202: crossing through the point $(2,-1,1)$.  We set its initial vorticity---as well as the
1203: initial vorticity of all modes other than the mother and father---to 
1204: zero.\footnote{\label{foot:cc}We
1205: ignore the complex conjugate modes for simplicity.
1206: Since 
1207: $\bld{\omega(x)}$ is real-valued, each mode with wavevector 
1208: and amplitude $(\bld{k},\bld{\omega)}$ is accompanied by a
1209: conjugate mode that has $(-\bld{k},\bld{\omega}^*)$.
1210: In our initial state, there are really four modes with non-zero amplitudes: the mother
1211: at $(1,0,0)$ and its conjugate at $(-1,0,0)$, and the father and its conjugate.  We may ignore the conjugate
1212: modes because they do not affect the instability described here.  
1213: As shown in L07 for the 2D case, 
1214: their main effect  is that when the son swings through $(0,-1,1)$, not only 
1215: does it couple with the mother at $(1,0,0)$ to produce a grandson at $(1,-1,1)$, but
1216: it also couples with the conjugate mother at $(-1,0,0)$ to partially kill its father,
1217:  which is then at $(-1,-1,1)$ (bringing to mind the story of Oedipus).
1218: But since the father is a trailing mode at this time, it no longer participates
1219: in the instability.  Nonetheless, the conjugate modes do play a role in the nonlinear
1220: outcome of the instability.
1221: }
1222: The father's wavevector and Fourier amplitude are labelled as in \S\ref{sec:lin}, and
1223: the mother's and son's  are labelled with bars and primes:
1224: \beqn
1225: {\rm father:\ }&{\bld{{k}}}& \ , \
1226: {\bld{\omega}}
1227:  \nonumber \\
1228: {\rm mother:\ }&{\bld{\bar{k}}}&\equiv \bar{k}{\bld{\hat{x}}}  \ \  , \ 
1229: {\bld{\bar{\omega}}}\equiv{\bar{\omega}}{\bld{\hat{z}}} 
1230: \nonumber \\
1231: {\rm son:\ }&{\bld{k'}}& \equiv \bar{k}{\bld{\hat{x}}} +{\bld{k}} \  , \ 
1232: {\bld{\omega'}}
1233: \nonumber
1234: \eeqn
1235: Note that $\bar{k}$=const., and $\bar{\omega}_x=0$ because the vorticity
1236: must be transverse to the wavevector.
1237: We also set $\bar{\omega}_y=0$; otherwise $\bar{u}_z\ne 0$, which corresponds to
1238:  a mean flow out the top of the box  and  in through the bottom.
1239: 
1240: At early times, the father mode swings through the point $(0,-1,1)$. 
1241: Since the only other nonvanishing mode at this time is the mother, there 
1242: are no mode couplings  that can nonlinearly change the father's
1243: amplitude.
1244: Therefore its amplitude
1245: is governed by the linear equation (\ref{eq:linlim}), which we reproduce here as
1246: \be
1247: {d\over d\tau}
1248: \left(\begin{array}{c}{\omega}_x \\ {\omega}_{yz}\end{array}\right)
1249: =\beta{\Omega\over \kappa}
1250: \left(\begin{array}{cc}0 & -{1\over 2}{\kappa^2\over\Omega^2}{1\over 1+\tau^2} \\ 2 & 0\end{array}\right)
1251: \left(\begin{array}{c}{\omega}_x \\ {\omega}_{yz}\end{array}\right) \ .
1252: \label{eq:lin2}
1253: \ee
1254: During its swing, it couples with the mother to change the amplitude of the son. 
1255:  The linear part  of the son's evolution is given by the
1256: above equation with primed vorticity and wavevector in place of unprimed.
1257: The nonlinear part is given by \be
1258: {d\over dt}{\bld{\omega'}}\Big\vert_{\rm nonlin}=i\bld{
1259: k'\times(\bar{u}\times \omega+u\times\bar{\omega})
1260: }
1261: \ee
1262: (eq. [\ref{eq:eomo}])
1263: where  $\bld{\bar{u}}=-i(\bar{\omega}/\bar{k}){\bld{\hat{y}}}$ and
1264:  $\bld{u}=i\bld{k\times\omega}/k^2$ (eq. [\ref{eq:udef}]).
1265: Adding the linear and nonlinear parts, and re-expressing in terms of our chosen
1266: degrees of freedom, we find
1267: \beqn
1268: {d\over d\tau}
1269: \left(\begin{array}{c}{\omega}_x' \\ {\omega}_{yz}'\end{array}\right)
1270: =\beta{\Omega\over \kappa}
1271: \left(\begin{array}{cc}0
1272: \ \ \ %spaces
1273:  & -{1\over 2}{\kappa^2\over\Omega^2}{1\over 1+(\tau+\bar{\tau})^2} \\ 
1274: 2\ \ \ %spaces
1275:   & 0\end{array}\right)
1276: \left(\begin{array}{c}{\omega}_x' \\ {\omega}_{yz}'\end{array}\right) 
1277: \nonumber
1278: \\
1279: -{\bar{\omega}\over q}
1280: \left(\begin{array}{cc} {1\over \bar{\tau}}
1281: \ \ \ %spaces
1282: &{\beta q\over\kappa}{1\over 1+\tau^2} \\
1283:  0\ \ \  %spaces
1284:  &{1\over\bar{\tau}}- {\bar{\tau}\over 1+\tau^2}\end{array}\right)
1285:  \left(\begin{array}{c}{\omega}_x \\ {\omega}_{yz}\end{array}\right)  \ ,
1286:  \label{eq:soneq}
1287: \eeqn
1288: where the dimensionless constant
1289: \be
1290: \bar{\tau}\equiv {\bar{k}_x\over k_{yz}} \ 
1291: \ee
1292: depends on both the mother's and father's wavevectors.
1293: It is the father's $\tau\equiv k_x/k_{yz}$ that is being used as the time-coordinate
1294: for evolving the son's amplitude.
1295: The grandson's equation is the obvious extension: 
1296: denoting
1297: the grandson's amplitudes with double primes, one need only
1298: make the following replacements
1299: in equation (\ref{eq:soneq}):
1300:  $\bld{\omega'}\rightarrow\bld{\omega''}$,
1301: $\bld{\omega}\rightarrow\bld{\omega'}$, and
1302: $\tau\rightarrow \tau+\bar{\tau}$.
1303: Subsequent generations
1304: evolve  analogously.
1305: 
1306: The father's equation  (\ref{eq:lin2}) is easily solved, as shown in \S \ref{sec:lin}. 
1307: Inserting this solution into equation (\ref{eq:soneq}) produces a linear inhomogeneous
1308: equation for the son's amplitude, and similarly for the grandson's.
1309: Figure \ref{fig:nonlin} plots numerical solutions of these equations.  Also shown
1310:  as circles are output from a pseudospectral simulation,
1311: showing excellent agreement.
1312: 
1313: 
1314: %\clearpage
1315: %%%%%%%%%%%%%begin figure
1316: \begin{figure*}
1317:   \centering
1318:   \hspace{-3cm}\includegraphics[width=.4\textwidth]{f8a.eps}
1319:  \hspace{1cm} \includegraphics[width=.4\textwidth]{f8b.eps}  
1320:   \caption{{\bf Curves of Marginal Stability for a Mother Mode
1321:   With  $\bld{\bar{\omega}=0.005}$ (left panel) and
1322:   $\bld{\bar{\omega}=0.05}$ (right panel):}  Left panel corresponds to Figure \ref{fig:nonlin} 
1323:   and right panel corresponds to the pseudospectral simulations of Figures \ref{fig:shortsim}-\ref{fig:tallsim}.
1324:   We set $\Omega=1$ and $q=3/2$.
1325: To make the solid lines in these plots (the ``exact solutions''), 
1326: we repeated the integrations that produced the lines in Figure \ref{fig:nonlin}, but varying
1327: $k_z$ for each $k_y$ until perturbations neither grew nor decayed.  The dashed line
1328: in the left panel
1329: shows that the analytic approximation of equation (\ref{eq:amp}) agrees reasonably well
1330: with the exact solution.  We do not show equation (\ref{eq:amp}) in the right panel because
1331: the agreement is poorer there. Right panel shows two X's for the values of the
1332: smallest non-zero $|k_y|$ and $|k_z|$ in the simulations of Figures \ref{fig:shortsim}-\ref{fig:tallsim},
1333: i.e., $|k_y/\bar{k}|=L_x/L_y=0.067$ for both simulations, and $|k_z/\bar{k}|=L_x/L_z=0.13$
1334: for the short box and $=0.033$ for the tall box.  The tall box contains a 3D-unstable mode
1335: that leads to the destruction of the vortex into a turbulent-like state.  The short box contains
1336: no such mode, and is stable to 3D perturbations.
1337:  }
1338:   \label{fig:marg}
1339: \end{figure*}
1340: %%%%%%%%%%%%%end figure
1341: %\clearpage
1342: 
1343: 
1344: In the Appendix, we solve  equation (\ref{eq:soneq}) analytically to
1345:  derive the amplification factor $\chi$, which is the ratio of the son's amplitude
1346: at any point in its evolution (e.g., when it is radially symmetric), to the father's amplitude at
1347: the same point in its evolution. 
1348: We find 
1349: \be
1350: \chi=-{\bar{\omega}\over q}\bar{\tau}^\delta\sqrt{\pi}{1+\delta\over\delta^2}
1351: \left(
1352: 1+{q\Omega\over\kappa^2}(1-\delta)
1353: \right)
1354: {\Gamma(1+\delta/2)\over\Gamma(1/2+\delta/2)}
1355: \label{eq:amp}
1356: \ee
1357: where $\delta=\sqrt{1-4\beta^2}$.
1358: Equation (\ref{eq:amp}) is applicable in the limit $|\bar{\omega}|/q\ll 1$.
1359: For 2D modes ($\beta=0\Rightarrow \delta=1$), it recovers  equation
1360: 42  of L07 (see also eq. [\ref{eq:2dinst}] of this paper): 
1361: \be
1362: \chi_{2D}=-\pi{\bar{\omega}\over q}{\bar{k}\over k_y} \ .
1363: \ee
1364: 
1365: Marginally stable modes have $|\chi|=1$.  
1366: Figure \ref{fig:marg} plots curves of marginal stability. 
1367: The left panel is for the case $\bar{\omega}=0.005$, as in Figure \ref{fig:nonlin}, and the right
1368: panel is for $\bar{\omega}=0.05$, as in the pseudospectral simulations presented 
1369: at the outset of this paper (eq. [\ref{eq:initz}]; Figs. \ref{fig:shortsim}-\ref{fig:tallsim}).
1370: The left panel shows that equation (\ref{eq:amp}) gives a fair reproduction of the exact curve.
1371: We do not show equation (\ref{eq:amp}) in the right panel, because it gives poorer
1372: agreement there (since $|\bar{\omega}|/q$ is too large). In the right panel we also plot X's for the values
1373: of the smallest nonvanishing 
1374: 3D wavenumbers in the simulations of Figures \ref{fig:shortsim}-\ref{fig:tallsim}.
1375: In the short-box simulation, all 3D modes lie in the stable zone.  Therefore the dynamics remains
1376: two-dimensional.  But in the 3D box, there is a 3D mode in the unstable zone that destroys the vortex
1377: and gives rise to turbulent-like behavior.
1378: 
1379: It is interesting to consider briefly how the instability described here connects with the
1380: Rayleigh-unstable case, which occurs when $\kappa^2<0$ . 
1381:   At small $|k_y|$, the  marginally stable curves in Figure \ref{fig:marg}
1382:  are
1383: given by $|\beta|=1/2$, where $\beta=(\kappa/q)(k_z/k_y)$.  Hence if one decreases
1384: $\kappa$ from its Keplerian value $\Omega$, the marginally stable curve becomes
1385: steeper in the $k_z-k_y$ plane, and an increasing number of 3D modes become
1386: unstable.  As $\kappa\rightarrow 0$, if a 2D mode with some $k_y$ is unstable, then
1387: so are all 3D modes with the same $k_y$.  Therefore any 2D-unstable state is also
1388: 3D-unstable, and any forming vortex would decay into turbulence.
1389: 
1390: 
1391: 
1392: \section{Conclusions}
1393: \label{sec:conc}
1394: 
1395: 
1396: Our main result follows from Figure \ref{fig:marg}, which maps
1397: out the stability of a 
1398: ``mother mode'' (i.e., a mode with wavevector $\bar{k}\bld{\hat{x}}$
1399:  and amplitude $\bar{\omega}$)
1400: to nonaxisymmetric 3D perturbations. 
1401: A mother mode is unstable 
1402: provided that the $k_y$ and $k_z$ of the
1403:  nonaxisymmetric perturbations  satisfy both
1404: $|k_y|\lesssim \bar{k}\bar{\omega}/q$ and $|k_z|\lesssim |k_y|$, dropping
1405: order-unity constants.
1406: Based on this result, we may understand the formation, survival, 
1407: and destruction of vortices.
1408: Vortices form out of mother modes that are unstable to 2D ($k_z=0$)
1409: perturbations. Mother modes that are unstable to 2D modes but stable
1410: to 3D ($k_z\ne 0$) ones, form into long-lived vortices.  Mother modes
1411: that are unstable to both 2D and 3D modes are destroyed.
1412: Therefore a mother mode with given $\bar{k}$ and $\bar{\omega}$
1413: will form into a vortex if the disk has a sufficiently large circumferential
1414: extent and a sufficiently small scale height, i.e., if $r\gtrsim \bar{k}^{-1}q/\bar{\omega}$
1415: and $h\lesssim \bar{k}^{-1}q/\bar{\omega}$, where $r$ is the distance
1416: to the center of the disk, and $h$ is the scale height.
1417: Alternatively, 
1418: the mother mode will be destroyed in a turbulent-like state
1419: if  both $r$ and $h$ are sufficiently large
1420:  ($r\gtrsim \bar{k}^{-1}q/\bar{\omega}$
1421:   and
1422: $h\gtrsim \bar{k}^{-1}q/\bar{\omega}$).
1423: 
1424: 
1425: Our result
1426: has a number of astrophysical consequences.
1427: In protoplanetary disks that do not contain any vortices,
1428: solid particles drift inward. 
1429: Gas disks orbit at sub-Keplerian speeds,  
1430: $v_{\rm gas}\sim \Omega r (1-\eta)$, where 
1431: $\Omega r$ is the Keplerian speed and $\eta\sim  (c_s/\Omega r)^2$, with
1432: $c_s$  the sound speed.  Since solid particles  would orbit
1433: at the Keplerian speed in the absence of gas, the mismatch of speeds
1434: between solids and gas produces a drag on the solid particles, removing
1435: their angular momentum and causing them to fall into the star.
1436: For 
1437: example, in the minimum mass solar nebula, meter-sized
1438: particles fall in from 1 AU in around a hundred years.  
1439: This rapid infall presents
1440: a serious problem for theories of planet formation, since it is
1441: difficult to produce planets out of dust in under a hundred years.
1442: Vortices can solve this problem \citep{BS95}.
1443: A vortex that has 
1444: excess vorticity $-\bar{\omega}$ 
1445: and radial width $1/\bar{k}$ can halt the infall of particles
1446: provided that $\bar{\omega}/\bar{k}\gtrsim (\Omega r)\eta$, 
1447: because the gas speed induced by such a vortex more than 
1448: compensates for
1449: the sub-Keplerian speed induced by gas pressure.\footnote{
1450: We implicitly assume here that the stopping time of the particle
1451: due to gas drag
1452: is comparable to the orbital time, which is true for meter-sized particles
1453: at 1 AU in the minimum mass solar nebula.  A more careful treatment
1454: shows that
1455: a vortex can stop a particle with stopping time $t_s$ provided
1456: that  $\bar{\omega}/\bar{k}\gtrsim (\Omega r)(\Omega t_s)\eta$ 
1457: \citep{Youdin08}.
1458: }
1459: Previous simulations implied that 3D vortices rapidly 
1460: decay, and so cannot  prevent the rapid infall of solid
1461: particles \citep{BM05b,SSG06}.  Our result
1462: shows that vortices can survive within disks, and so restores
1463: the viability of vortices as a solution to the infall problem. 
1464: 
1465: 
1466: A more important---and more speculative---application of our result is
1467: to the transport of angular momentum within neutral accretion disks. 
1468: In our simulation of a vortex in a tall box, we found that as the vortex 
1469: decayed it transported angular momentum outward at a nearly
1470: constant rate for hundreds of orbital times.
1471:  If decaying vortices transport a significant amount of angular momentum
1472: in disks, they would resolve one of the most important outstanding questions
1473: in astrophysics today: what causes hydrodynamical accretion disks to accrete?
1474: To make this speculation more concrete, one must understand the amplitude
1475: and duration of the ``turbulence'' that results from decaying vortices.  
1476: This is a topic for future research.
1477: 
1478: In this paper, we considered only the effects of rotation and shear on
1479: the stability of vortices, while we neglected the
1480: effect of vertical gravity.  
1481: There has been a lot of research in the geophysical community 
1482: on the dynamics of fluids in the presence of vertical gravity, since 
1483: stably stratified fluids are very common on Earth---in the atmosphere, oceans, 
1484: and lakes.  In numerical and laboratory experiments of strongly stratified
1485: flows, thin horizontal ``pancake vortices'' often form, and fully developed 
1486: turbulence is characterized by thin horizontal layers. 
1487: \citep[e.g., ][]{BBLC07}.  Pancake vortices are stabilized by vertical gravity,
1488: in contrast to the vortices studied in this paper which are stabilized by rotation.
1489: Gravity inhibits vertical motions because of buoyancy: it costs gravitational energy
1490: for fluid to move vertically.
1491: The resulting quasi-two-dimensional
1492: flow can form into a vortex.\footnote{\cite{BC00} 
1493: show that
1494: a vertically uniform
1495: vortex column in a stratified (and non-rotating and non-shearing) 
1496:  fluid suffers an instability (the ``zigzag instability'') that is 
1497: characterized by a typical vertical lengthscale $\lambda_z\sim U/N$, where $U$ is 
1498: the horizontal speed induced by the vortex, $N$ is the Brunt-V\"ais\"al\"a frequency,
1499: and the horizontal lengthscale of the vortex $L_h$ is assumed to be much
1500: greater that  $\lambda_z$ (hence the pancake structure).
1501:  We may understand
1502: Billant \& Chomaz's result in a crude fashion with an argument similar to that employed in the introduction
1503: to explain the destruction of rotation-stabilized vortices: since the frequency
1504: of buoyancy waves is $Nk_x/k_z$ (when $|k_x|\ll |k_z|$), and since the frequency
1505: at which fluid circulates around a vortex is $U/L_h\sim k_xU$, there is a resonance between
1506: these two frequencies for vertical lengthscale $1/k_z\sim U/N$.} 
1507: We may speculate that in an astrophysical disk vertical gravity provides an 
1508: additional means to stabilize vortices, in addition to rotation.  
1509: But to make this speculation concrete, the theory presented in this paper
1510: should be extended to include vertical gravity.
1511: 
1512: 
1513: 
1514: We have not addressed in this paper the origin of the axisymmetric structure
1515: (the mother modes) that give rise to surviving or decaying vortices.
1516: One possibility is that decaying vortices can produce more axisymmetric
1517: structure, and therefore they can lead to self-sustaining turbulence.
1518: This seems to us unlikely.
1519: We have not seen evidence for it in our simulations, but this could
1520: be  because
1521: of the modest resolution of our simulations.  
1522: Other possibilities for the generation of axisymmetric structure include thermal instabilities, such
1523: as the baroclinic instability, or convection, or
1524: stirring by planets.  This, too, is a topic for future
1525: research.
1526: 
1527: 
1528: 
1529: 
1530: 
1531: \appendix
1532: \section{Analytic Expression for Growth Factor $\chi$}
1533: 
1534: In this Appendix, we derive equation (\ref{eq:amp}) by analytically
1535: integrating equation (\ref{eq:soneq}) for the son's vorticity, given the father's vorticity
1536: as a function of time (\S \ref{sec:lin}),
1537: and taking the mother's vorticity $\bar{\omega}$
1538:  to be constant, which is valid when the father's amplitude is 
1539: small relative to the mother's.
1540: The numerical integral of equation (\ref{eq:soneq})  is shown in Figure \ref{fig:nonlin}.
1541: Recall that initially  $\tau=\bar{\tau}>0$ and $\tau$ decreases in time, and typically
1542: $\bar{\tau}\gg 1$.
1543: It simplifies the analysis to work with the son's ``normal-mode'' amplitudes $\omega_A'$ and
1544: $\omega_B'$, defined from $\omega_x'$ and $\omega_{yz}'$ via 
1545: (eqs. [\ref{eq:oxeq}] and [\ref{eq:powerlaw}])
1546: \be
1547: \label{eq:sonnm}
1548: \left(\begin{array}{c}{\omega}_x' \\ {\omega}_{yz}'\end{array}\right)
1549: =
1550: \left(\begin{array}{cc}
1551: {\kappa\over 2\beta\Omega}{1-\delta\over 2\tau'}|\tau'|^{1-\delta\over 2} & 
1552: {\kappa\over 2\beta\Omega}{1+\delta\over 2\tau'}|\tau'|^{1+\delta\over 2} 
1553:  \\
1554: |\tau'|^{1-\delta\over 2}   
1555:   & 
1556: |\tau'|^{1+\delta\over 2}  
1557:   \end{array}\right)
1558: \left(\begin{array}{c}{\omega}_A' \\ {\omega}_B'\end{array}\right) \ .
1559: \ee
1560: where
1561: \be
1562: \tau'\equiv\tau+\bar{\tau}
1563: \ee
1564: Substituting this into equation (\ref{eq:soneq}), the time derivative of the above
1565: matrix cancels the homogeneous term in that equation if we approximate
1566: $1+\tau^{'2}\simeq  \tau^{'2}$, which holds until just before the time
1567: that $\tau=-\bar{\tau}$.  The inhomogeneous term produces
1568: \beqn
1569: {d\over d\tau}\omega_A'={\bar{\omega}\over q}{1\over\delta}|\tau'|^{1+\delta\over 2}
1570: \left(
1571: \omega_{yz}
1572: \left[
1573: {1\over 1+\tau^2}
1574: \left(
1575: {2\beta^2q\Omega\over\kappa^2}+{1+\delta\over 2}{\bar{\tau}\over\tau'}
1576: \right)
1577: -{1+\delta\over 2}{1\over \bar{\tau}\tau'}
1578: \right]
1579: +\omega_x{2\beta\Omega\over\kappa}{1\over{\bar{\tau}}}
1580: \right) \ \ , \ \ \tau\gtrsim -\bar{\tau}
1581: \label{eq:aprime}
1582: \eeqn
1583: Since $\omega_x$ and $\omega_{yz}$ are known (\S \ref{sec:lin}), 
1584: a straightforward integration yields $\omega_A'$ just before $\tau=-\bar{\tau}$.
1585: To perform this integral, we resort to some approximations, guided by the solution
1586: shown in Figure \ref{fig:nonlin}.
1587: For the first term, we need
1588: \beqn
1589: \int_{\bar{\tau}}^{-\bar{\tau}}\omega_{yz}{(\tau+\bar\tau)^{1+\delta\over 2}\over 1+\tau^2}d\tau
1590: &\approx&
1591: %-{2\Omega\over\beta\kappa}\int_\epsilon^{-\epsilon}{d\omega_x\over d\tau}
1592: %(\tau+\bar\tau)^{1+\delta\over 2}d\tau+
1593: -{1\over \beta^2}\int_{\bar{\tau}}^{-\epsilon\bar\tau}
1594: {d^2\omega_{yz}\over d\tau^2}(\tau+\bar\tau)^{1+\delta\over 2}
1595: d\tau
1596: +
1597: \omega_B\int_{-\epsilon\bar\tau}^{-\bar\tau}
1598: |\tau|^{{1+\delta\over 2}-2}
1599: (\tau+\bar\tau)^{1+\delta\over 2}
1600: d\tau 
1601: \\
1602: &\approx&
1603: \omega_B\bar{\tau}^\delta{1+\delta\over 2\beta^2}\epsilon^{-1+\delta\over 2}
1604: -\omega_B\bar{\tau}^\delta\int_{\epsilon}^1s^{-3+\delta\over 2}(1-s)^{1+\delta\over 2}ds\\
1605: &\approx&
1606: \omega_B{\bar\tau}^\delta2^{-\delta}\sqrt{\pi}{1+\delta\over1-\delta}
1607: {\Gamma(1/2+\delta/2)\over \Gamma(1+\delta/2)}
1608: \eeqn
1609: where $\epsilon$ is a parameter that satisfies $1\gg \epsilon\gg 1/\bar\tau$.  In the first line,
1610: we used equation (\ref{eq:bh}),
1611: and
1612: we discarded the $\omega_A$ mode from the second integral 
1613: because the $\omega_B$ mode increases faster
1614: with increasing $|\tau|$. 
1615: From Figure \ref{fig:nonlin}, $\omega_A'$ nearly vanishes until $\tau\simeq 0$.  Therefore,
1616: in the second line we approximated the first integral as 
1617: $-(\bar{\tau}^{1+\delta\over 2}/\beta^2){d\omega_{yz}/d\tau}\vert_{-\epsilon}$.
1618: The third line holds in the limit of small $\epsilon$.
1619: The other three terms in equation (\ref{eq:aprime}) are integrated similarly, yielding
1620: \be
1621: \omega_A'=\omega_B{\bar{\omega}\over q}{1\over\delta}{\bar\tau}^\delta 2^{-\delta-1}\sqrt{\pi}
1622: {\Gamma(1/2+\delta/2)\over \Gamma(1+\delta/2)}
1623: \left(
1624: 1+{q\Omega\over\kappa^2}(1-\delta)
1625: \right){(1+\delta)^2\over 1-\delta}
1626: \label{eq:omap}
1627: \ee
1628: just before the time when $\tau=-\bar{\tau}$, i.e., just before the son is radially symmetric. 
1629: At this time, Figure \ref{fig:nonlin} shows that
1630: $\omega_B'$ very nearly vanishes.  Therefore, just after the son is radially symmetric, 
1631: it will have $\omega_B'=T_{BA}{\omega_A'}$ (eq. [\ref{eq:trans}]), with $\omega_A'$
1632: given by equation (\ref{eq:omap}). This gives the amplification factor $\chi\equiv\omega_B'/\omega_B$
1633: that is displayed in equation (\ref{eq:amp}).
1634: 
1635: 
1636:  
1637:  
1638: 
1639: 
1640: \bibliographystyle{apj}
1641: %\bibliography{shearbib}
1642: 
1643: \begin{thebibliography}{20}
1644: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1645: 
1646: \bibitem[{{Afshordi} {et~al.}(2005){Afshordi}, {Mukhopadhyay}, \&
1647:   {Narayan}}]{AMN05}
1648: {Afshordi}, N., {Mukhopadhyay}, B., \& {Narayan}, R. 2005, \apj, 629, 373
1649: 
1650: \bibitem[{{Balbus} \& {Hawley}(1998)}]{BH98}
1651: {Balbus}, S.~A. \& {Hawley}, J.~F. 1998, Reviews of Modern Physics, 70, 1
1652: 
1653: \bibitem[{{Balbus} \& {Hawley}(2006)}]{BH06}
1654: ---. 2006, \apj, 652, 1020
1655: 
1656: \bibitem[{{Barge} \& {Sommeria}(1995)}]{BS95}
1657: {Barge}, P. \& {Sommeria}, J. 1995, \aap, 295, L1
1658: 
1659: \bibitem[{{Barranco} \& {Marcus}(2005)}]{BM05b}
1660: {Barranco}, J.~A. \& {Marcus}, P.~S. 2005, \apj, 623, 1157
1661: 
1662: \bibitem[{{Barranco} \& {Marcus}(2006)}]{BM06}
1663: ---. 2006, Journal of Computational Physics, 219, 21
1664: 
1665: \bibitem[{{Billant} \& {Chomaz}(2000)}]{BC00}
1666: {Billant}, P. \& {Chomaz}, J.-M. 2000, Journal of Fluid Mechanics, 419, 29
1667: 
1668: \bibitem[{{Brethouwer} {et~al.}(2007){Brethouwer}, {Billant}, {Lindborg}, \&
1669:   {Chomaz}}]{BBLC07}
1670: {Brethouwer}, G., {Billant}, P., {Lindborg}, E., \& {Chomaz}, J.-M. 2007,
1671:   Journal of Fluid Mechanics, 585, 343
1672: 
1673: \bibitem[{{Chagelishvili} {et~al.}(2003){Chagelishvili}, {Zahn}, {Tevzadze}, \&
1674:   {Lominadze}}]{CZTL03}
1675: {Chagelishvili}, G.~D., {Zahn}, J.-P., {Tevzadze}, A.~G., \& {Lominadze}, J.~G.
1676:   2003, \aap, 402, 401
1677: 
1678: \bibitem[{{Drazin} \& {Reid}(2004)}]{DR04}
1679: {Drazin}, P.~G. \& {Reid}, W.~H. 2004, {Hydrodynamic Stability} (Hydrodynamic
1680:   Stability, by P.~G.~Drazin and W.~H.~Reid, pp.~626.~ISBN
1681:   0521525411.~Cambridge, UK: Cambridge University Press, September 2004.)
1682: 
1683: \bibitem[{{Gammie} \& {Menou}(1998)}]{GM98}
1684: {Gammie}, C.~F. \& {Menou}, K. 1998, \apjl, 492, L75+
1685: 
1686: \bibitem[{{Gill}(1965)}]{Gill65}
1687: {Gill}, A.~E. 1965, Journal of Fluid Mechanics, 21, 503
1688: 
1689: \bibitem[{{Godon} \& {Livio}(1999)}]{GL99}
1690: {Godon}, P. \& {Livio}, M. 1999, \apj, 523, 350
1691: 
1692: \bibitem[{{Johnson} \& {Gammie}(2005)}]{JG05}
1693: {Johnson}, B.~M. \& {Gammie}, C.~F. 2005, \apj, 635, 149
1694: 
1695: \bibitem[{{Lerner} \& {Knobloch}(1988)}]{LK88}
1696: {Lerner}, J. \& {Knobloch}, E. 1988, Journal of Fluid Mechanics, 189, 117
1697: 
1698: \bibitem[{{Lithwick}(2007)}]{L07}
1699: {Lithwick}, Y. 2007, \apj, 670, 789
1700: 
1701: %\bibitem[{{Lithwick} \& {Chiang}(2007)}]{LC}
1702: %{Lithwick}, Y. \& {Chiang}, E. 2007, \apj, 656, 524
1703: 
1704: \bibitem[{{Lovelace} {et~al.}(1999){Lovelace}, {Li}, {Colgate}, \&
1705:   {Nelson}}]{Love99}
1706: {Lovelace}, R.~V.~E., {Li}, H., {Colgate}, S.~A., \& {Nelson}, A.~F. 1999,
1707:   \apj, 513, 805
1708: 
1709: \bibitem[{{Lynden-Bell} \& {Pringle}(1974)}]{LP}
1710: {Lynden-Bell}, D. \& {Pringle}, J.~E. 1974, \mnras, 168, 603
1711: 
1712: \bibitem[{{Marcus}(1993)}]{Marcus93}
1713: {Marcus}, P.~S. 1993, \araa, 31, 523
1714: 
1715: \bibitem[{{Maron} \& {Goldreich}(2001)}]{MG01}
1716: {Maron}, J. \& {Goldreich}, P. 2001, \apj, 554, 1175
1717: 
1718: \bibitem[{{Rogallo}(1981)}]{Rogallo81}
1719: {Rogallo}, R.~S. 1981, NASA STI/Recon Technical Report N, 81, 31508
1720: 
1721: \bibitem[{{Saffman}(1995)}]{Saffman95}
1722: {Saffman}, P.~G. 1995, {Vortex Dynamics} (Vortex Dynamics, by P.~G.~Saffman,
1723:   pp.~325.~ISBN 0521477395.~Cambridge, UK: Cambridge University Press, February
1724:   1995.)
1725: 
1726: \bibitem[{{Sano} {et~al.}(2000){Sano}, {Miyama}, {Umebayashi}, \&
1727:   {Nakano}}]{SMUN00}
1728: {Sano}, T., {Miyama}, S.~M., {Umebayashi}, T., \& {Nakano}, T. 2000, \apj, 543,
1729:   486
1730: 
1731: \bibitem[{{Shen} {et~al.}(2006){Shen}, {Stone}, \& {Gardiner}}]{SSG06}
1732: {Shen}, Y., {Stone}, J.~M., \& {Gardiner}, T.~A. 2006, \apj, 653, 513
1733: 
1734: \bibitem[{{Stone} \& {Balbus}(1996)}]{SB96}
1735: {Stone}, J.~M. \& {Balbus}, S.~A. 1996, \apj, 464, 364
1736: 
1737: 
1738: \bibitem[{{Umurhan} \& {Regev}(2004)}]{UR04}
1739: {Umurhan}, O.~M. \& {Regev}, O. 2004, \aap, 427, 855
1740: 
1741: \bibitem[{{Yecko}(2004)}]{Yecko04}
1742: {Yecko}, P.~A. 2004, \aap, 425, 385
1743: 
1744: \bibitem[{{Youdin}(2008)}]{Youdin08}
1745: {Youdin}, A. 2008, ArXiv e-prints, 807
1746: 
1747: \end{thebibliography}
1748: 
1749: 
1750: \end{document}
1751: 
1752: 
1753: