1: \documentclass[12pt,preprint]{aastex}
2:
3: \begin{document}
4:
5:
6: \newtheorem{theorem}{Theorem}[section]
7: \newtheorem{lemma}{Lemma}[section]
8: \newtheorem{corollary}{Corollary}[section]
9: \newtheorem{proposition}{Proposition}
10: \newtheorem{example}{Example}
11:
12:
13: \title{Magnetoelliptic Instabilities}
14: \author{Norman R. Lebovitz}
15: \affil{Mathematics Department, University of Chicago}
16: \affil{Chicago, IL 60637, USA}
17: \email{norman@math.uchicago.edu}
18: \and
19: \author{Ellen Zweibel}
20: \affil{Astronomy Department and Center for Magnetic Self-Organization}
21: \affil{University of Wisconsin, Madison, WI 53706 USA}
22: \email{zweibel@astro.wisc.edu}
23:
24:
25: \begin{abstract}
26: We consider the stability of a configuration consisting of a vertical
27: magnetic field in a planar flow on elliptical streamlines in ideal
28: hydromagnetics. In the absence of a magnetic field the elliptical flow
29: is universally unstable (the ``elliptical instability''). We find this
30: universal instability persists in the presence of magnetic fields of
31: arbitrary strength, although the growthrate decreases somewhat. We
32: also find further instabilities due to the presence of the magnetic
33: field. One of these, a destabilization of Alfven waves, requires the
34: magnetic parameter to exceed a certain critical value. A second,
35: involving a mixing of hydrodynamic and magnetic modes, occurs for all
36: magnetic-field strengths. These instabilities may be important in tidally
37: distorted or otherwise elliptical disks. A disk of finite thickness is stable
38: if the magnetic fieldstrength exceeds a critical value, similar to
39: the fieldstrength which suppresses the magnetorotational instability.
40:
41: \end{abstract}
42:
43: \keywords{accretion discs -- instabilities: elliptical, hydromagnetic}
44:
45: \section{Introduction}\label{intro}
46: The problem of momentum transport in accretion disks is widely
47: believed to require hydrodynamic or hydromagnetic turbulence for its
48: resolution. The origin of this turbulence may be sought in the
49: instability of laminar solutions of the equations of hydromagnetics,
50: solutions that are compatible with the geometry of accretion
51: disks. The recent history of these efforts has taken the form of first
52: recognizing such an instability mechanism, and then trying to
53: incorporate that mechanicsm into realistic disk models.
54:
55: The magnetorotational instability (MRI) mechanism, originally discovered by
56: \citet{vel} and \citet{cha} and first applied to accretion disks in
57: \citet{bah}, is of this kind (see
58: \citet{mri} for a review). It appears in rotating, magnetized systems in
59: which the specific angular momentum increases outward and in which the
60: magnetic field is weak enough that rotational effects are not overwhelmed
61: by magnetic tension.
62:
63: A second mechanism, that of the elliptical instability considered by
64: \citet{jg} and others \citep{lpk,ryug, ryugv}, is also consistent with the
65: accretion-disk setting. This instability mechanism has been reviewed by \citet{rk02}. In the setting considered by Goodman et. al., it
66: appears to require a secondary in order to enforce departure from
67: rotational symmetry of the streamlines via a tidal potential. This is
68: certainly appropriate for binary systems but it is likely that, even
69: in the absence of a secondary, the laminar motion in the plane of the
70: disk would not be accurately circular, so the elliptical-instability
71: mechanism would appear to be a candidate of considerable
72: generality. It does not require a magnetic field. One of the
73: conclusions of the present paper is that it further persists in the
74: presence of a magnetic field. In the idealized setting of the present
75: problem, the latter may be of arbitrarily large strength. However, we
76: also argue that in the setting of a disk geometry, there may indeed be
77: a limit on the field strength.
78:
79: In this paper we therefore investigate the interaction of a vertical
80: magnetic field with flow on elliptical streamlines, on the ground that
81: both magnetic fields and noncircular streamlines are likely
82: ingredients in accretion-disk settings. There are similarities with
83: and differences from previous work on effect of magnetic fields on the
84: elliptical instability \citep{rk94}, which are
85: discussed in \S \ref{discussion}.
86:
87:
88: \section{Formulation}\label{form}
89: We consider flow on elliptical streamlines together with a magnetic
90: field and investigate linear stability theory. The underlying
91: equations are the Euler equations of fluid dynamics
92: \begin{equation}\label{euler}
93: {\mathbf u}_t + {\mathbf u} \cdot \nabla {\mathbf u} = - \nabla p +
94: \left(\mbox{curl}\,{\mathbf B} \right) \times {\mathbf B}
95: \end{equation}
96: and the induction equation
97: \begin{equation}\label{induc}
98: {\mathbf B}_t + {\mathbf u} \cdot \nabla {\mathbf B} = {\mathbf B}
99: \cdot \nabla {\mathbf u}.
100: \end{equation}
101: We shall assume that
102: div$\,{\mathbf u} =0$ and div$\,{\mathbf B} =0$ and that the fluid is
103: unbounded.
104:
105: It is easy to check that the following steady fields represent a
106: solution of the preceding system:
107: \begin{equation}\label{unpert.sol}
108: {\mathbf U} = \Omega \left(- \frac{a_1}{a_2}x_2,
109: \frac{a_2}{a_1}x_1,0\right),\;{\mathbf B} = \left(0,0,B\right), \; P =
110: \frac{\Omega ^2}{2}\left(x_1^2 + x_2^2 \right).\end{equation} Here
111: $\Omega$ and $B$ are constants, and a constant may also be added to
112: the pressure term. More general exact solutions of the combined
113: fluid/magnetic equations exist in an unbounded domain \citep{addc88};
114: the case in hand is probably the simplest of these.
115:
116: \subsection{The perturbed system}\label{pert.sys}
117: Let ${\mathbf u}, {\mathbf B}, p$ be replaced by ${\mathbf U}+{\mathbf
118: u}, {\mathbf B}+{\mathbf b}, P + p$ in equations (\ref{euler},
119: \ref{induc}) above, and linearize. The resulting perturbation
120: equations are
121: \begin{equation}\label{pert.euler}
122: {\mathbf u}_t + {\mathbf U} \cdot \nabla {\mathbf u} + {\mathbf u} \cdot
123: \nabla {\mathbf U} = - \nabla p + \left(\mbox{curl} {\mathbf b} \right)
124: \times {\mathbf B}\end{equation} and
125: \begin{equation}\label{pert.induc}
126: {\mathbf b}_t + {\mathbf U} \cdot \nabla {\mathbf b} = {\mathbf B}
127: \cdot \nabla {\mathbf u} + {\mathbf b} \cdot \nabla {\mathbf U}, \end{equation}
128: together with the conditions that ${\mathbf u}$ and ${\mathbf b}$ be
129: solenoidal. These equations allow {\em rotating-wave} solutions of the
130: form
131: \begin{equation}\label{r-w.ansatz}
132: {\mathbf u}={\mathbf v}\left(t\right) \exp i \left({\mathbf
133: k}\left(t\right),{\mathbf x}\right),\;{\mathbf b}={\mathbf
134: w}\left(t\right) \exp i \left({\mathbf k}\left(t\right),{\mathbf
135: x}\right),\; p = \phi \left(t\right) \exp i \left({\mathbf
136: k}\left(t\right),x\right)\end{equation} where the expression $\left({\mathbf
137: k},{\mathbf x}\right)$ denotes the inner product.
138: Because ${\mathbf u}$ and ${\mathbf b}$ are solenoidal, the conditions
139: \begin{equation}\label{solenoidal}
140: \left({\mathbf k}\left(t\right),{\mathbf v}\left(t\right)\right) =0,\;\left({\mathbf k}\left(t\right),{\mathbf w}\left(t\right)\right) =0\end{equation}
141: must be satisfied.
142:
143: Write
144: \begin{equation}\label{U.matrix}
145: {\mathbf U} = A{\mathbf x}\;\mbox{where}\;A=\Omega
146: \left(\begin{array}{ccc}0&-E&0\\E^{-1}&0&0\\0&0&0
147: \end{array}, \;\right), E = a_1/a_2.\end{equation} Then on substituting the
148: rotating-wave expressions from equation (\ref{r-w.ansatz}) into the
149: perturbation equations one finds
150: \begin{eqnarray}
151: \dot{\mathbf k}&=& -A^t{\mathbf k},\label{keq}\\
152: \dot{\mathbf v}&=& -A{\mathbf v} -i{\mathbf k}\phi + iB\left({\mathbf k}\times {\mathbf w}\right)\times {\mathbf e}_3 \label{veq}\\
153: \dot{\mathbf w}&=&i\left(k_3 B\right){\mathbf v} + A{\mathbf w}.\label{weq}\end{eqnarray}
154: Equation (\ref{keq}) can be solved to give
155: \begin{equation}\label{ksol}
156: {\mathbf k} = \left(\kappa \cos \left(\Omega t - \chi\right),
157: E\kappa \sin \left(\Omega t - \chi\right), k_3\right),\end{equation}
158: where $\kappa, k_3 $ and $\chi$ are constants.
159: The pressure coefficient $\phi$ can be eliminated with the aid of the
160: solenoidal condition (\ref{solenoidal}). One finds
161: \begin{equation}\label{phieq}
162: -i\left(\phi + Bw_3\right) = 2k^{-2}\left(A^t{\mathbf k},{\mathbf
163: v}\right).\end{equation} The equation for ${\mathbf v}$ now takes the
164: form
165: \begin{equation}\label{new.veq}
166: \dot{\mathbf v}= C\left(t\right){\mathbf v} +i\left(k_3 B\right){\mathbf w},\end{equation}
167: where
168: \begin{equation}
169: C\left(t\right) =-2\left(\Omega /k^2\right)
170: \left(\begin{array}{ccc}-E^{-1}k_1k_2&Ek_1^2&0\\-E^{-1}k_2^2&Ek_1k_2&0\\-E^{-1}k_3k_2&Ek_1k_3&0\end{array}\right)
171: + \Omega \left(\begin{array}{ccc}0&E&0\\-E^{-1}&0&0\\0&0&0\end{array}\right).\end{equation}
172: It's convenient to break the six-dimensional system consisting of
173: equations (\ref{new.veq}) and (\ref{weq}) into two, one of size four and the other of size two:
174: \begin{eqnarray*}
175: \dot{v_1} &=&- \left(2\Omega /k^2\right)k_1\left(E k_1v_2 -
176: E^{-1}k_2v_1\right) + \Omega E v_2 +im w_1,\\
177: \dot{v_2} &=&
178: -\left(2\Omega /k^2\right)k_2 \left(E k_1v_2 - E^{-1}k_2 v_1\right)
179: -\Omega E^{-1}v_1 + im w_2,\\
180: \dot{w_1} &=& im v_1 -\Omega
181: Ew_2,\\ \dot{w_2} &=& im v_2 + \Omega E^{-1}w_1,\end{eqnarray*}
182: where $m = k_3 B.$ These four equations are self-contained and the remaining equations,
183: \begin{eqnarray}\label{v3w3}
184: \dot{v_3}&=&-\left(2\Omega /k^2\right)k_3\left(Ek_1v_2-E^{-1}k_2v_1\right) +im w_3,\\
185: \dot{w_3}&=&im v_3
186: \end{eqnarray}
187: may be integrated once the expression
188: \begin{equation}\label{c1}
189: c_1 \equiv Ek_1v_2 - E^{-1}k_2v_1
190: \end{equation}
191: is found by solving the four-dimensional system above. Equations
192: (\ref{keq}), (\ref{veq}),(\ref{weq}) and (\ref{phieq}) imply that
193: \[ \frac{d}{dt} \left({\mathbf k}, {\mathbf v}\right)= im
194: \left({\mathbf k}, {\mathbf w}\right) \;\mbox{and}\;\frac{d}{dt} \left({\mathbf k}, {\mathbf w}\right)= im
195: \left({\mathbf k}, {\mathbf v}\right).\]
196: Thus in solving this system we need to impose the conditions
197: that these inner products are zero initially; this will thereafter
198: maintain the incompressibility conditions (\ref{solenoidal}).
199:
200: The incompressibility condition provides an alternative way of finding $v_3$ and $w_3$ once the
201: equations for $v_1,v_2,w_1,w_2$ have been solved, provided that
202: $k_3 \ne 0.$ The only cases for which $k_3$ can vanish are those for which the combinations
203: $k_1 v_1 + k_2 v_2$ and $k_1 w_1 + k_2 w_2$ are also found to vanish on
204: solving the four-dimensional system above. It is not difficult to show
205: that there can be no instability associated with such a solution (see
206: in particular the equivalent system (\ref{c_eq}) below). Accordingly,
207: we henceforth consider only perturbations with vertical wave number
208: $k_3 \ne 0.$
209:
210: \subsection{Change of variables}\label{new-variables}
211: We change to new variables to facilitate subsequent
212: calculations\footnote{The origin of this change of variables is
213: related to the existence of the rotating-wave solutions.}.
214: \begin{eqnarray}\label{new_vars}
215: c_1&=&Ek_1v_2 - E^{-1}k_2v_1, \nonumber \\ c_2&=&k_1v_1 + k_2v_2
216: \left(=-k_3v_3\right), \nonumber \\ c_3&=&Ek_1w_2 -E^{-1}k_2w_1,
217: \nonumber \\ c_4&=&k_1w_1 +k_2w_2 \left(=-k_3w_3\right).
218: \end{eqnarray}
219: This is a time-dependent (periodic) change of variables since
220: ${\mathbf k}$ is periodic in $t$.
221: \noindent The equations to be solved take the form
222: \begin{equation}\label{c_eq}\dot{c}=D\left(t\right)c\end{equation} in these variables, with (we have put $\Omega = 1$ here to agree with earlier conventions)
223: \begin{equation}\label{D_matrix}
224: D\left(t\right) =
225: \left(\begin{array}{cccc}-2\left(E-E^{-1}\right)k^{-2}k_1k_2&-2&im&0\\2k^{-2}k_3^2&0&0&im\\im&0&0&0\\0&im&0&0\end{array}\right).
226: \end{equation}
227:
228:
229: \subsection{General Considerations}\label{general}
230:
231: The coefficients of the matrix $D$ depend on the phase angle $\chi$
232: appearing in the expressions (\ref{ksol}) above for the wave vector
233: $k$. For purposes of studying stability, we may set $\chi =0$\footnote{On the
234: other hand, for purposes of solving the initial-value problem, which
235: involves integrating over initial wave vectors, we would need to retain
236: it.}. This is easily seen by making the substitution $t^\prime =
237: t-\chi$, which eliminates $\chi$ from the equation. For the
238: remainder of this work we take $\chi =0.$
239:
240: \medskip
241:
242: The system (\ref{c_eq}) presents a Floquet problem (cf. \citet{ys}): the stability of
243: the chosen steady solution ${\mathbf U}, {\mathbf B}$ depends on
244: whether there are solutions of this system that grow exponentially
245: with time. This is settled by finding the Floquet multiplier matrix
246: $M$. The latter is defined as follows. Let $\Phi \left(t\right)$ be
247: the fundamental matrix solution of equation (\ref{c_eq}) that reduces
248: to the identity at $t=0$. Then, since the periodicity of $D$ is
249: $2\pi$, $M= \Phi \left(2\pi \right).$ If any eigenvalue $\lambda$
250: of $M$ has modulus exceeding one, this implies that there is indeed an
251: exponentially growing solution.
252:
253: It is familiar in conservative problems that
254:
255: \begin{proposition}\label{quartet}
256: Whenever $\lambda$ is an
257: eigenvalue of the Floquet matrix, so also are its inverse $\lambda
258: ^{-1}$ and its complex conjugate $\overline{\lambda}$.
259: \end{proposition}
260: The first statement of this proposition is a typically a consequence of
261: canonical Hamiltonian structure, the second a consequence of the
262: reality of the underlying problem. However, the system (\ref{c_eq}) is
263: not canonical, and the matrix appearing in it is not real. We can
264: nevertheless establish these familiar properties of the eigenvalues
265: directly from the system (\ref{c_eq}), as follows. The time-reversal
266: invariance of the physical problem is reflected in the existence
267: of a reversing symmetry $R = \mbox{diag}\,\left(1,-1,-1,1\right)$ of
268: the matrix $D$ above: $RD\left(-t\right) =- D\left(t\right)R$,
269: implying that whenever $c\left(t\right)$ is a solution so also is
270: $Rc\left(-t\right)$. Since the solutions of the Floquet problem have
271: the structure $c(t)=p(t) \exp \left(\sigma t\right)$, there must also
272: be a solution $p(-t) \exp \left(-\sigma t\right).$ But the eigenvalues
273: of the Floquet multiplier matrix $M$ are the values $\lambda = \exp
274: \left(2\pi \sigma \right)$. This shows that if $\lambda$ is an
275: eigenvalue of $M$, so also is $\exp \left(-2 \pi \sigma \right) =
276: \lambda ^{-1}$.
277:
278: Similarly, under matrix transformation $S = \mbox{diag}\,
279: \left(1,1,-1,-1\right)$, $D$ goes to its complex conjugate. This shows
280: that $M = S \overline{M} S^{-1}$, i.e., that $M$ and its conjugate
281: have the same eigenvalues.
282:
283: Immediate consequences of Proposition \ref{quartet} are the following:
284: first, in the stable case, eigenvalues of $M$ lie on the unit circle; second,
285: if, as parameters change, an eigenvalue is at the onset of
286: instability, it must have multiplicity two (or higher). The latter
287: conclusion is because, in the complex $\lambda$-plane, the dangerous
288: eigenvalue $\lambda$ must leave the unit circle simultaneously with
289: $\overline{\lambda}^{-1}$, which lies along the same ray as $\lambda
290: $ and therefore coincides with it when they both lie on the unit
291: circle. Thus a necessary condition for the onset of linear instability is a
292: resonance where two Floquet multipliers coincide.
293:
294: \subsection{Parameters}\label{parameters}
295: There are three dimensionless parameters that figure in this
296: problem. We call them $\epsilon, \mu,$ and $\eta$:
297: \begin{equation}\label{equ:definitions}
298: \epsilon = \frac{1}{2}\left(E - E^{-1}\right), \; \mu = k_3/k_0, \;
299: \mbox{and} \; \eta = k_0B.\end{equation}
300: Thus $\epsilon$ represents
301: the departure of the streamlines of the unperturbed flow from axial
302: symmetry. In these equations $k_0 = \sqrt{\kappa ^2 + k_3^2}$ and
303: represents the length of the wave vector if $\epsilon = 0$. The
304: magnetic parameter $\eta$ depends not only on the strength of the
305: unperturbed magnetic field but also on the wavelength of the
306: perturbation.
307:
308: In the matrix $D$ above, the magnetic field enters through the
309: parameter $m=k_3 B$ (which is a
310: measure of the magnetic tension
311: force), and we shall continue to use this notation for
312: the present, on the understanding that $m=\mu \eta$. It is also clear
313: that we can use $E$ rather than $\epsilon$ to measure the departure
314: from rotational symmetry, and we shall do this in some cases.
315:
316:
317: \section{Analysis}
318:
319: The Floquet matrix is $$
320: M\left(\epsilon , \mu, \eta \right) = \Phi \left(2\pi,\epsilon , \mu
321: , \eta \right).$$
322: One could map out the stability and instability
323: regions in the $\epsilon \mu \eta$ parameter space numerically by
324: integrating the system (\ref{c_eq}) systematically for many values of
325: these parameters. We in fact do this for a selection of parameter
326: values in \S \ref{numerical} below. However, in this and the
327: following section we present the outlines and results of an asymptotic
328: analysis based on regarding $\epsilon$ as a small parameter (details
329: are presented in the Appendices). This is more revealing than the
330: numerical reults on their own. It is also of considerable importance
331: in interpreting the numerical results, and is quite accurate even
332: for values of $\epsilon$ that are not very small (cf. Figure \ref{hydromode} below).
333: The calculation proceeds in two steps; finding $M$, and calculating its eigenvalues.
334:
335: \subsection{The Floquet Matrix}\label{floquet}
336:
337: The asymptotic
338: analysis is facilitated by the circumstance that, if we put $\epsilon
339: =0$, the coefficient matrix $D_0$ (say) of equation (\ref{D_matrix})
340: becomes constant. We find the eigenvalues and eigenvectors of $D_0$ in Appendix
341: \ref{subsec:zop}. There are two complex conjugate pairs of modes. One pair
342: reduces to the ordinary hydrodynamic modes in the limit $\eta\rightarrow 0$. The
343: second pair are magnetic modes with zero frequency at $\eta = 0$. In the weak
344: field limit, the ratio of magnetic to kinetic energy is ${\mathcal{O}}(\eta^2/4)$ for the modified hydrodynamic modes and ${\mathcal{O}}(4/\eta^2)$ for the
345: magnetic modes. If $\eta\gg 1$, the kinetic and magnetic energies are near
346: equipartition for both types of mode.
347:
348: We now turn to a brief description of the asymptotic (or
349: perturbation) procedure. In what follows the parameter $\eta$ will
350: be held fixed so, to simplify the notation, we suppress the dependence
351: of the Floquet matrix on this parameter: $M = M\left(\epsilon ,\mu
352: \right) = \Phi \left(2 \pi , \epsilon, \mu \right)$. We shall need the
353: Taylor expansion
354: \begin{equation}\label{M_exp}
355: M\left(\epsilon , \mu \right) = M\left(0,\mu _0 \right) + M_\epsilon
356: \left(0,\mu _0 \right) \epsilon + M_\mu \left(0, \mu _0 \right) \left(\mu -
357: \mu _0 \right) + \cdots ,
358: \end{equation}
359: where the dots indicate higher-order terms in $\epsilon$ and $\mu -
360: \mu _0$. The reason for allowing variations with $\mu$ as well as
361: variations with $\epsilon$ is that the region in the $\epsilon \mu $
362: plane where instability occurs is typically a wedge with apex at a
363: point $\left(\epsilon, \mu \right) = (0, \mu _0)$ and boundaries $\mu
364: = \mu_0 + \nu _{\pm}\epsilon$, where the slopes $\nu _+$ and $\nu _-$
365: are to be found. (cf. Figure \ref{hydromode} below). We will therefore
366: consider $\mu$ of the form
367: \begin{equation}\label{mu_expansion}
368: \mu = \mu _0 + \nu \epsilon + \cdots
369: \end{equation}
370: where $\nu$ may be regarded as a fixed parameter to be chosen
371: later. This will lead us to the widest of wedges in the $\epsilon \mu$
372: plane, of width $O\left(\epsilon\right)$, excluding other wedges of
373: width $O\left(\epsilon ^m\right)$ with $m \ge 2$. These higher-order
374: wedges typically occupy a tiny fraction of the parameter space
375: (see Figure \ref{ho} below). As a result of the representation
376: (\ref{mu_expansion}) we may write
377: \begin{equation}\label{M_exp_2}
378: M = M_0 + M_1 \epsilon + \cdots
379: \end{equation}
380: where
381: \begin{equation}\label{M0andM1}
382: M_0 = M\left(0, \mu _0 \right) \; \mbox{and}\; M_1 = M_\epsilon
383: \left(0,\mu _0\right) + \nu M_\mu \left(0 , \mu _0 \right).
384: \end{equation}
385:
386:
387: The matrix $M_0$ is given in Appendix \ref{subsec:zop}, together with an
388: expression for $M_{\epsilon}$. We effect a change of basis, which diagonalizes $M_0$
389: and simplifies the calculation of $M_{\epsilon}$, in Appendix \ref{distinct}. The elements
390: of $M_1$ are calculated in Appendix \ref{subsec:elements}.
391:
392: \subsection{The characteristic polynomial}\label{char_poly}
393: Denote by
394: \begin{equation}\label{charpol}
395: p\left(\lambda, \epsilon \right) = \left|M\left(\epsilon
396: \right) - \lambda I \right|\end{equation}
397: the characteristic polynomial of the
398: Floquet multiplier matrix $M\left(\epsilon \right)$ and its roots by
399: $\Lambda _1, \Lambda _2, \cdots$. A necessary
400: condition for stability is that each root lie on the unit circle. To
401: explore the onset of instability for various parameter values, we
402: shall obtain the characteristic polynomial $p$ in the form
403: \begin{equation}\label{p_expansion}
404: p\left(\lambda , \epsilon \right) = p_0\left(\lambda \right) +
405: p_1\left(\lambda \right) \epsilon + p_2\left( \lambda \right) \epsilon
406: ^2 + \cdots \end{equation} and exploit our knowledge of the roots of
407: $p_0$ to obtain the roots of $p\left(\lambda , \epsilon \right)$ in
408: the form of a Puiseux expansion in $\epsilon$ \citep{hil}).
409:
410: The nature of this expansion depends on the
411: multiplicities of the roots $\left\{\lambda _k \right\}$ of $p_0$, the
412: characteristic polynomial of the unperturbed Floquet matrix
413: $M_0$. These roots are given by the expressions $\lambda _k = \exp
414: 2\pi \sigma _k$ where the $\left\{\sigma _k \right\}$ are the
415: eigenvalues of the matrix $D_0$ given in Appendix \ref{subsec:zop} (equation \ref{evals})
416: ; they are all distinct. However, it is possible for the multipliers $\left\{\lambda _k
417: \right\}$ to be repeated even when, as in the present case, the $\left\{\sigma _k \right\}$ are
418: distinct: if $\sigma _k - \sigma _l = i k$ for an integer $k \ne 0$,
419: then $\lambda _k = \lambda _l$.
420:
421: \medskip
422: A necessary condition for the onset of instability is that there be a
423: double (or higher) root of the characteristic equation, and we henceforth
424: restrict consideration to the case of double roots\footnote{Higher
425: order zeros are not ruled out in this problem, since there are three
426: independent parameters, but we do not pursue this here.}. For
427: definiteness, we suppose $\lambda _2 = \lambda _1$. Then the Puiseux
428: expansion takes the form
429: \begin{equation}\label{puiseux} \Lambda _1 \left(\epsilon \right) = \lambda _1 + \epsilon ^{1/2}
430: \beta _{1/2} + \epsilon \beta _1 + \cdots .\end{equation}
431: Substituting into the characteristic equation (and taking into account
432: that $p_0$ and $p_0^\prime$ both vanish at $\lambda _1$) yields for
433: the coefficient $\beta _{1/2}$ of the leading-order correction the
434: equation
435: \begin{equation}\label{mu1/2} \beta _{1/2} ^2 = -2p_1\left(\lambda _1 \right)/p_0 ^{\prime \prime}
436: \left(\lambda _1\right).\end{equation}
437: The two values of $\beta _{1/2}$ give the generic expressions for the change in a double eigenvalue, yielding a pair of roots branching
438: from the double root $\lambda _1$. However, if $p_1 \left(\lambda _1
439: \right) =0,$ this expression is inadequate and one must proceed to the
440: next term in order to determine the effect of the perturbation on the
441: stability. Under the present assumptions, it is indeed the case that $p_1$ vanishes at
442: $\lambda _1$, as we show in Appendix \ref{subsec:charpoly}\footnote{This
443: ``nongeneric'' behavior can be traced to the circumstance that the perturbation expansion takes place at a codimension-two point, i.e., where two
444: relations must hold among the parameters.}.
445:
446:
447: We must suppose then that the expansion of $p\left(\lambda , \epsilon
448: \right)$ is carried out to second order in $\epsilon$:
449: \begin{equation}\label{p_exp_2}
450: p\left(\lambda , \epsilon \right) = p_0\left(\lambda \right) +
451: \epsilon p_1\left(\lambda\right) +
452: \epsilon ^2 p_2\left( \lambda \right) + \cdots
453: \end{equation}
454: Then in the Puiseux expansion above $\beta _{1/2}
455: = 0$ so $\Lambda _1 = \lambda _1 + \beta _1 \epsilon + \cdots,$ and $\beta _1$ is found by solving the quadratic equation
456: \begin{equation}\label{beta1_eq}
457: \frac{1}{2} p_0^{\prime \prime}\left(\lambda _1\right)\beta _1 ^2 +
458: p_1^\prime \left( \lambda _1\right) \beta _1 + p_2\left(\lambda _1 \right) =0.\end{equation}
459:
460: \medskip
461: In the case at hand, the common value of $\lambda _1$ and $\lambda _2$
462: lies on the unit circle. In order for the perturbed values of the
463: Floquet multipliers to lie {\em off} the unit circle (and therefore imply
464: instability), it is easy to verify that it is necessary and sufficient that $\beta _1/ \lambda
465: _1$ have a nonvanishing real part. Thus if we define
466: \begin{equation}\label{ratio}
467: \alpha = \beta _1 /\lambda _1 ,
468: \end{equation}
469: we have the following criterion:
470: \begin{proposition}\label{stab_crit}
471: Either $\alpha $ is pure-imaginary and we infer stability
472: (to leading order in $\epsilon$), or $\mbox{Re}\, \alpha \ne 0$ and
473: we infer instability.\end{proposition}
474: The magnitude of the real part of $\alpha$ is also related to the growthrate of the instability. If we define an instability increment
475: \begin{equation}\label{incr}
476: \Delta = \left| \Lambda \right| -1,\end{equation}
477: then $\Delta = \epsilon \mbox{Re}\, \alpha$ and the growthrate is equal to $\Delta /2 \pi$, to leading order in $\epsilon$.
478:
479: The long calculations that lead to the coefficients appearing in
480: equation (\ref{beta1_eq}) are carried out in the Appendices. In the
481: notation employed there, equation (\ref{beta1_eq}) therefore takes
482: the form
483: \begin{equation}\label{beta1_eq_3}
484: \alpha ^2 - \left\{\tilde{J}_{11} + \tilde{J}_{22} +\frac{2\pi
485: i}{\mu}\nu \left[\omega _1 + \omega _2\right] \right\} \alpha + \left|
486: \begin{array}{cc} \tilde{J}_{11}+ 2\pi \nu \sigma _1/\mu
487: &\tilde{J}_{12}\\ \tilde{J}_{21}&\tilde{J}_{22} +2\pi \nu
488: \sigma _2/\mu
489: \end{array} \right| = 0,
490: \end{equation}
491: In equation (\ref{beta1_eq_3}), $\alpha$
492: has the meaning of Proposition \ref{stab_crit} above, and the symbols
493: $\tilde{J}_{ij}$ are defined in Appendix \ref{distinct} (equation \ref{Jij}).
494: There are obvious modifications of this formula if $\lambda _k =
495: \lambda _l$ instead of $\lambda _1 = \lambda _2$.
496:
497: \subsection{The Resonant Cases}\label{kne2}
498: The resonant cases for $\epsilon = 0$ (circular streamlines) are those
499: parameter values $(\mu,\eta)$ such that $\omega _j - \omega _l =k$,
500: where $k$ is an integer. We'll find that these can be written in the
501: form $\mu = f\left(\eta\right)$ (e.g., equation (\ref{pure_hydro_res}
502: below). Since $\epsilon = 0$, the $\mu$ values in question are those
503: that were designated $\mu _0$ in equation (\ref{mu_expansion})
504: above. We no longer need the designation $\mu _0$ and, in the relations
505: below, use the symbol $\mu$ in its place.
506:
507: If $k \ne \pm 2$ the matrix $\tilde{J}$ is
508: diagonal (see equation \ref {newJtilde} below) and equation
509: (\ref{beta1_eq_3}) has the roots $\alpha = J_{11} + 2\pi i \nu \omega
510: _1/\mu$ and $\alpha = J_{22}+ 2\pi i \nu \omega _2/\mu$ where $\sigma
511: _j = i \omega _j$ for $j=1,2,3,4).$ It is easy to check that these
512: diagonal entries are pure-imaginary (cf. equations \ref{diagonal_eq}
513: and \ref{Tinverse} below) and therefore, in accordance with
514: Proposition \ref{stab_crit}, there is no instability to leading order
515: in $\epsilon$. We therefore now consider the only cases $(k=\pm 2)$ that
516: can lead to instability to this order.
517:
518: \medskip
519: Recall that the original parameters of the problem are $\epsilon$ --
520: representing the departure from axial symmetry of the undisturbed
521: streamlines -- $\mu = k_3/k_0$ -- representing the vertical wavenumber
522: --and $\eta = k_0B$. The auxiliary parameters $m=\mu \eta$ and $q=\mu
523: \sqrt{1 + \eta ^2}$ are introduced to simplify the notation. With the
524: frequencies taken in the order
525: \begin{equation}\label{four_freqs} \left( \omega _1, \omega _2, \omega _3, \omega _4 \right) = \left( \mu + q, -\mu -q, \mu -q, -\mu +q \right) \end{equation}
526: the replacement $\mu \to -\mu$ results in the same frequencies in the opposite order. We may therefore assume without loss of generality that
527: $\mu > 0$. Further scrutiny of the formula (\ref{four_freqs}) shows that we need only consider the following four, distinct, $k=2$ resonances.
528:
529: \subsubsection{Case 1. $\omega _1 - \omega _2 = 2$}\label{resonance1}
530: The resonant modes are those that reduce, when $\eta =0$, to the
531: purely hydrodynamic modes. This case therefore represents the
532: modification, due to the presence of the vertical magnetic field, of
533: the universal elliptical instability. In this case $\omega _1 = -
534: \omega _2 =\mu + q=1$, implying that
535: \begin{equation}\label{pure_hydro_res} \mu = \frac{1}{1+ \sqrt{1 + \eta ^2}}.\end{equation}
536: This ratio changes from $1/2$ at $\eta =0$ to $0$ as $\eta \to \infty$.
537: Evaluating the integrals defining $\left(\tilde{J}_{ij}\right)$ (equation \ref{Jij} below) and
538: solving equation (\ref{beta1_eq_3}) above, we find
539: \begin{equation}\label{alpha_res1}
540: \alpha ^2 = \left[\frac{\pi}{2}\left(1+\mu \right)^2\right]^2 - \pi ^2
541: \left[\frac{2\nu }{\mu} - \mu \left(1+\mu
542: \right)\right]^2. \end{equation} This has a maximum instability
543: increment (when $\nu = \mu ^2 \left(1+\mu \right)/2$) given
544: (suppressing a factor of $\epsilon$) by
545: \[ \alpha _{max} = \frac{\pi}{2}\left(1+\mu \right)^2.\]
546: In the pure-hydrodynamic limit $\eta = 0$ we find $\mu = 1/2$ and
547: therefore $\alpha _{max} = 9\pi/8$. Since the growthrate is given by
548: $\alpha \epsilon /2\pi$, this gives a maximum growthrate of $9\epsilon
549: /16$, in agreement with previous results obtained by other methods \citep{wal,jg}. In the limit $\eta \rightarrow \infty$ this
550: maximum instability increment tends to the finite limit $\pi /2$,
551: about half its value in the pure-hydrodynamic limit.
552:
553: This instability has a {\em bandwidth} $\left(\nu _+ - \nu _-
554: \right)\epsilon$ which is, for given $\epsilon$ and $\eta$, the length
555: of the $\mu$-interval for which the unperturbed configuration is
556: unstable. It is determined by the values of $\nu$ that make the real
557: part of $\alpha$ vanish. These may be read off equation
558: (\ref{alpha_res1}) above:
559: \[ \nu _+ = \mu \left(1+\mu \right) \left(1+3\mu \right)/4,\; \nu _- = - \mu \left(1 -\mu ^2\right)/4.\]
560: In the limit $\eta \to 0$ (the pure-hydrodynamic case) these give $\nu
561: _+ = 15/32$ and $\nu _- = -3/32$. These values of $\nu _\pm$ can also
562: be inferred from Waleffe's treatment of the pure-hydrodynamic case. In
563: the limit of large magnetic parameter $\eta$, the width of the band
564: tends to zero.
565:
566:
567: \subsubsection{Case 2. $\omega _1 - \omega _3 = 2$}\label{resonance2}
568: The resonant modes consist of a hydrodynamic mode and a purely
569: magnetic mode (one frequency would vanish if $\eta = 0$).
570: In this case $q=1$, implying that $ \omega _1= \mu +1$, $ \omega _ 3= \mu -1$
571: and
572: \begin{equation}\label{equ:reshm}
573: \mu = \frac{1}{\sqrt{1+\eta ^2}}.
574: \end{equation}
575:
576: Thus the ratio $\mu$ changes from $1$ at $\eta=0$ to $0$ as $\eta \to
577: \infty$.
578: This represents a 'mixed mode', i.e., the resonance is between a
579: purely hydrodynamic mode and a purely magnetic one. Evaluating the
580: integrals and solving the quadratic (\ref{beta1_eq_3}), we find
581:
582: \begin{equation}\label{alpha_res2}
583: \alpha= i\pi \nu \left( \frac{2 \nu }{\mu} -1+\mu ^2 \right) \pm \sqrt{D}
584: \end{equation}
585: where
586: \[ D = -\pi ^2 \left[ \frac{2\nu }{\mu} - \frac{1}{2}\left(1-\mu ^4\right)\right]\left[\frac{2\nu }{\mu} - \frac{1}{2}\left(1-\mu ^2\right)\left(3\mu ^2 -1 \right)\right].\]
587: If $D<0$ then $\alpha$ is pure-imaginary and the unperturbed configuration is stable, so instability prevails if and only if $D > 0$.
588:
589: Instability can indeed occur and has its maximal increment when $\nu = \mu ^3\left(1-\mu ^2 \right)/2$. This maximal increment is (except for a factor of $\epsilon$)
590: \[ \left( \mbox{Re}\, \alpha \right)_{max} = \frac{\pi}{2}\left(1-\mu
591: \right)^2 .\]
592: This ranges from $0$ when $\eta = 0$ to $\pi /2$ as $\eta \to
593: \infty$.
594:
595: The width of instability band can be calculated as in the preceding case by finding the values of $\mu$ for which $D=0$. These are
596: \[ \nu _+ = \frac{1}{4}\mu \left(1 - \mu ^4 \right) \; \mbox{and}\; \nu
597: _- = \frac{1}{4}\mu \left(1-\mu ^2 \right)\left(3\mu ^2 -1 \right).\]
598:
599: \subsubsection{Case 3. $\omega _4 - \omega _3 =2$}\label{resonance3}
600: These are the purely magnetic modes, which play no role if $\eta = 0$.
601: For this case we have $q - \mu =1$ implying, for fixed $\eta$ that
602: \begin{equation}\label{equ:resmm}
603: \mu = \frac{1}{\sqrt{1+\eta ^2} -1}.
604: \end{equation}
605:
606: Since $\mu $ cannot exceed $1$, this resonance can only occur for
607: sufficiently large values of $\eta $, namely
608: \begin{equation}\label{eta_big}
609: \eta > \sqrt{3}.\end{equation}
610: For a given wavelength $k_0$ this would require a sufficiently large
611: magnetic field $B$. When this
612: condition is satisfied, we have $ \omega _4= 1 =- \omega _3 .$
613: The formula for the instability increment $\alpha$ becomes
614: \begin{equation}\label{alpha_res3}
615: \alpha ^2=\left[\frac{\pi }{2}\left(1-\mu \right)^2 \right]^2 - \pi ^2 \left[\frac{2\nu }{\mu} + \mu \left(1 -\mu \right) \right]^2.\end{equation}
616: The maximum instability increment for given $\epsilon$ is (again
617: suppressing a factor of $\epsilon$)
618: \[ \alpha _{max} = \frac{\pi }{2}\left(1-\mu \right)^2,\]
619: which occurs for $\nu = -\mu ^2 \left(1-\mu \right)/2.$
620: It vanishes in the limit $\eta =0$ and tends to $\pi/2$ as $\eta \to \infty$. The upper and lower edges of the band of instability are expressed by the formulas $\mu = \mu _0 + \nu _\pm \epsilon$ where $\nu _\pm$ may be determined from equation (\ref{alpha_res3}) by requiring $\alpha$ to vanish. One finds
621: \[ \nu _+ = \frac{1}{4}\mu \left(1-\mu \right) \left(1-3\mu \right) \; \mbox{and} \; \nu _- = -\frac{1}{4}\mu \left(1-\mu ^2\right).\]
622: The bandwidth is therefore
623: \[ \nu _+ - \nu _- = \frac{1}{2}\mu \left(1-\mu \right)^2.\]
624: This vanishes both in the limit as $\eta \to 0$ and in the limit as $\eta \to \infty$. The maximum bandwidth occurs when $\mu = 1/3$ or $\eta = \sqrt{15}$.
625:
626: \subsubsection{Case 4. $\omega _1 - \omega _4 =2$}\label{resonance4}
627:
628: For this $\mu =1$. It is clear from the expressions for the matrix $J$
629: that the latter vanishes in this case. This leads to pure-imaginary values of $\alpha$ and therefore there is no
630: instability associated with this resonance.
631:
632:
633: \section{Numerical results}\label{numerical}
634: In this section we present a selection of numerical results. These are
635: obtained by integrating the system (\ref{c_eq}) to obtain the
636: fundamental matrix solution $\Phi \left(t,\epsilon,\mu,\eta\right)$
637: and evaluating it at $t=2\pi$ to get the Floquet matrix
638: $M\left(\epsilon,\mu ,\eta,\right)$. For fixed $\eta$, the eigenvalue
639: of maximum modulus is found as a function of $E$ (rather than
640: $\epsilon$) and $\mu$, and the regions of the $E \mu$ plane where this
641: maximum modulus exceeds one are distinguished. We have carried this
642: out to $E = 1.6$ ($\epsilon = 0.4875)$ in the figures although there is no
643: limitation on the size of $E$, or of $\epsilon$, in this method.
644:
645: In Figure \ref{hydromode}, we have taken the magnetic parameter equal
646: to zero in the left-hand panel, so this represents the purely
647: hydrodynamic case studied originally by \citet{bay} and \citet{rtp} and subsequently by others. For $\eta > 0$, a mixed mode of
648: interaction involving both hydrodynamic and hydromagnetic modes
649: comes into existence, and this is shown in the right-hand panel of
650: Figure \ref{hydromode}, where $\eta =1$. This is too small for the remaining
651: leading-order instability to appear.
652: \placefigure{hydromode}
653: \begin{figure}
654: %\figurenum{hydromode}
655: \plotone{f1.eps}
656: \caption{\small On the left is the case $\eta =0$, i.e., the purely
657: hydrodynamic case considered earlier by several other authors. It is
658: shown here to contrast it with the case when the magnetic field is not
659: zero. The lines of asterisks indicate the same stability boundaries obtained from the asymptotic formulas $\mu = \mu_0 + \nu _{\pm}\epsilon$; this approximation is seen to be quite good. It is typical of the other cases considered. On the right is the case $\eta =1$, showing both the effect of
660: the magnetic field on the hydrodynamic mode and the existence of a new
661: mode of instability that is due to the presence of the magnetic
662: field.}\label{hydromode}
663: \end{figure}
664: That remaining instability, which represents a resonant interaction
665: between two modes that owe their existence to the presence of the
666: magnetic field, is indicated in Figure \ref{mixedmode} for a magnetic
667: parameter $\eta = 2$ (right-hand panel), slightly greater than the minimum value
668: ($\sqrt{3}$) for the existence of this instability.
669: \placefigure{mixedmode}
670: \begin{figure}
671: %\figurenum{mixedmode}
672: \plotone{f2.eps}
673: \caption{\small On the left is the case $\eta =2$. Three regions of
674: instability occur here. The lowest of these is the modification of the
675: hydrodynamic-mode instability indicated in Figure \ref{hydromode}. The
676: middle region refers to the mixed hydrodynamic-magnetic mode. The
677: uppermost region, very thin and labeled MAGNETIC MODE, exists only for
678: values of $\eta$ exceeding $\sqrt{3}$, so is poorly developed for this
679: value of $\eta$. On the right is the case $\eta = \sqrt{15} \approx
680: 3.873$, for which the width of the uppermost, purely magnetic, mode
681: band is at its greatest. Note however that the vertical scale is
682: compressed relative to the diagram on the left, exaggerating the
683: widths of these bands by about a factor of two.}\label{mixedmode}
684: \end{figure}
685:
686: The asymptotic formulas imply that the maximal growthrate (or
687: equivalently the maximal instability increment $\Delta$) for each of
688: the wedges of instability tends to a fixed value as the magnetic
689: parameter $\eta$ increases. This is illustrated in Figure
690: \ref{grra}. However, the asymptotic formulas for the growthrates are
691: less accurate than those for the stability boundaries, for the larger
692: values of $E$.
693: \placefigure{grra}
694: \begin{figure}
695: %\figurenum{grra}
696: \plotone{f3.eps}
697: \caption{The left-hand panel shows the instability increment $\Delta$ as a function
698: of $\mu$ for a fixed value of the magnetic parameter ($\eta =0$) and the ellipticity ($E = 1.3$). The right-hand panel does the same except that $\eta = \sqrt{15}$.}\label{grra}
699: \end{figure}
700:
701: In identifying these tongues, we have made repeated use of the
702: asymptotic formulas presented earlier. The numerical procedure also
703: picks up some further tongues, related to higher-order resonances, that are excluded by the procedure leading to the asymptotic formulas. These we
704: have mostly ignored on the ground that they are too weak and occupy too small
705: a region of the parameter space to be significant, but we show one such resonance tongue for two values of $\eta$ in Figure \ref{ho}
706: \placefigure{ho}
707: \begin{figure}
708: %\figurenum{ho}
709: \plotone{f4.eps}
710: \caption{These show higher order wedges of instability. Both are cases
711: of resonance between the two hydrodynamic modes but with
712: $\omega_1-\omega_2 =4$, rather than $2$. The left-hand panel shows the
713: instability wedge when $\eta = 1$, the right-hand panel shows the
714: corresponding wedge when $\eta = \sqrt{15}$.}\label{ho}
715: \end{figure}
716:
717:
718:
719: \section{Upper bound on fieldstrength}\label{sec:upperbound}
720:
721: The results of \S\S \ref{kne2} and \ref{numerical} show that magnetoelliptic
722: instability occurs in vertically
723: unbounded systems whatever the fieldstrength, and that the growth rates are
724: relatively insensitive to the magnetic fieldstrength parameter $\eta$. The
725: loci of instability in the $(E, \mu)$ plane, however, do depend on $\eta$. As
726: $\eta\rightarrow\infty$, the resonant $\mu$, in all three cases,
727: scales as $\eta^{-1}$ and the magnetic tension parameter $m\rightarrow 1$. This
728: is a consequence of the resonant character of the instability. Since
729: the mode frequency must be close to the rotational frequency, the Alfv\'en
730: frequency $m$ cannot be too large. As $B\rightarrow\infty$, $k_3$ must approach
731: zero.
732:
733: In a system of finite vertical thickness $H$, $k_3$ cannot drop below $\chi/H$,
734: where $\chi$ is a factor of order unity. Therefore, if $v_A$ exceeds $\Omega
735: H/\chi$, we expect the magnetoelliptic instability to be suppressed. A precise
736: upper bound on $v_A$ follows from eqn. (A3) and application of the resonance
737: conditions; equations (\ref{pure_hydro_res}), (\ref{equ:reshm}), and
738: (\ref{equ:resmm}).
739: In Cases 1 and 2, instability requires $m^2\le 1$. Case 3 requires
740: $m^2\le 3$. Therefore, magnetoelliptic instability requires $v_A\le\sqrt{3}
741: \Omega H/\chi$ (there is a correction of order $\epsilon$ due to the finite
742: bandwidth). The corresponding fieldstrength
743: is comparable to the maximum value of the field at
744: which the magnetorotational instability can operate \citep{mri}.
745:
746: \section{Discussion}\label{discussion}
747: We have explored the effect of a uniform, vertical magnetic field on
748: the stability of planar, incompressible flow with elliptical
749: streamlines in an unbounded medium, in the approximation of ideal
750: magnetohydrodynamics. In the absence of magnetic fields, flows with
751: elliptical streamlines having ellipticity parameter $\epsilon$ [see
752: equation (\ref{equ:definitions})] are known to be unstable to
753: perturbations with wavevectors that are transverse to the plane of the
754: flow (the ``elliptical instability"). Our first conclusion is that
755: this elliptical instability persists in the presence of a vertical
756: magnetic field: the latter decreases the maximum growthrate but fails
757: to suppress the instability, no matter how large the magnetic-field
758: parameter becomes. It can be compared with the conclusion of
759: \citet{rk94} that a toroidal magnetic field has a stabilizing
760: influence. Kerswell's analysis holds for small $\epsilon$ only and
761: shows that the growthrate decreases with magnetic field in that
762: limit. Our result, which holds for a vertical magnetic field, shows a
763: similar trend for small values of the magnetic-field parameter, but
764: this trend never results in complete stabilization of the elliptical
765: instability with increasing magnetic field.
766:
767: A second conclusion is that there are further instabilities associated
768: with the presence of the magnetic field. One of these, for which the
769: eigenvector is a mixture of hydrodynamic and magnetic modes, occurs
770: for all values of the magnetic field parameter. Another, for which the
771: eigenvector is a combination of magnetic modes only, sets in for
772: values of the magnetic-field parameter exceeding a certain threshold
773: value ($\eta > \sqrt{3}$).
774: In all three of these instabilities, for large magnetic fields, the
775: wave vector makes only a small angle with the plane of the
776: unperturbed flow, reflecting the familiar tendency for dynamics to become
777: nearly 2-dimensional in a strong, well ordered magnetic field. This is
778: reflected in Figure \ref{mixedmode}, which shows that as $\eta$ increases,
779: the unstable wedges are pushed to smaller $\mu$. Although the
780: unstable fraction of
781: the $(E,\mu)$ plane decreases with increasing $\eta$ (except for a very
782: slight maximum at $\eta\sim$ 2.18, reflecting the onset of instability
783: between magnetic modes), the separation between the unstable wedges also
784: decreases. While
785: the nonlinear evolution of the unstable
786: system is beyond the scope of this work, the destabilization of a nearly
787: continuous swath of parameter space may have consequences for the
788: interactions between unstable
789: modes.
790: In all three
791: cases, the maximum instability increment tends
792: to $\epsilon \pi/2$, i.e., the maximum growthrate of the unstable
793: modes tends, in dimensional units, to $\epsilon\Omega /4$.
794:
795: All three of these instabilities may be relevant in accretion-disk settings.
796: In systems of finite thickness $H$, however, the instability is suppressed if
797: the Alfv\'en speed $v_A$ exceeds a critical value of order $\Omega H$.
798: Magnetorotational instabilities are quenched at approximately the same
799: fieldstrength \citep{mri}.
800: The growth rate of magnetoelliptical instabilities is smaller than
801: that of magnetorotational instabilities by a factor of order $\epsilon$, and
802: thus they are not necessarily the primary instability in magnetized disks.
803: They may well play a secondary
804: role by breaking up eddies or vortices generated by other
805: mechanisms.
806:
807: Magnetoelliptic instabilities may also occur in the inner parts of
808: barred galaxies, in which the gas flow is slightly elliptical and the
809: magnetic field, at least in the Milky Way, has a vertical component
810: \citep{mor}. In such settings, the instabilities could be a source of
811: turbulence, possibly affecting the mass supply to a central compact
812: object.
813:
814: \acknowledgements
815:
816: We are happy to acknowledge the referee for useful comments. Material
817: support for this work was provided by NSF grants AST-0098701,
818: AST-0328821, PHY-0215581, and the Graduate School of the University of
819: Wisconsin, Madison.
820: \appendix
821:
822: \pagebreak
823: \noindent {\Huge \bf Appendices}
824:
825: \noindent We carry out calculations leading to
826: the coefficients appearing in equation (\ref{beta1_eq}) in a series of
827: steps. Since $p(\lambda)= |M - \lambda I|$, with $M$ given by
828: equations (\ref{M_exp_2}) and (\ref{M0andM1}), we begin with the expression for $M$.
829:
830: \section{The Expansion for M}
831: \subsection{Zero-order Problem and Solution for $M$}\label{subsec:zop}
832: The matrix $D\left(t, \epsilon , \mu\right)$ has only two entries
833: that depend on $\epsilon$: $D_{11}$ and $D_{21}$. If, therefore, we
834: write the Taylor expansion
835: \begin{equation}\label{D_expansion} D\left(t, \epsilon, \mu \right) = D_0 \left(t, \mu \right) + \epsilon D_\epsilon \left( t,\mu \right) + \cdots \end{equation}
836: then
837: \[ D_0 \left(t,\mu \right) = D\left(t, 0, \mu \right) \; \mbox{and}\; D_\epsilon \left(t, \mu \right) = D_\epsilon \left(t, 0, \mu \right), \]
838: where $D_\epsilon = \partial D/\partial \epsilon$.
839: For $D_0$ we find the constant matrix
840: \begin{equation}\label{D0_matrix}
841: D_0\left(t, \mu \right) = D_0\left(\mu \right)= \left(
842: \begin{array}{cccc}0&-2&im&0\\2\mu ^2&0&0&im\\im&0&0&0\\0&im&0&0
843: \end{array}\right).\end{equation} Its eigenvalues are
844: \begin{equation}\label{evals}
845: \sigma _1 = i\left(\mu + q\right),\sigma _2 = -i\left(\mu +
846: q\right),\sigma _3 = i\left(\mu - q\right),\sigma _4 = -i\left(\mu -
847: q\right),\end{equation} where $q = \sqrt{\mu ^2 + m^2} = \mu \sqrt{1 +
848: \eta ^2}.$ These are distinct and nonzero as long as $\mu \ne 0$ and
849: $\eta \ne 0$, which we assume to be the case. The first two
850: correspond to ``hydrodynamic modes'' since they reduce, when $\eta=0$,
851: to the eigenvalues of the purely hydrodynamic case. The second two
852: refer to ``magnetic modes'' since they are zero in that limit. These
853: are all of stable type, corresponding to frequencies $\omega _k$, $k=
854: 1, \ldots, 4$. Regarding the matrix $D_\epsilon$, one can easily work it out
855: from the expression (\ref{D_matrix}) above: all its entries vanish
856: except $\left(D_\epsilon\right)_{11}$ and $\left(D_\epsilon\right)_{21}.$ One finds
857: \begin{equation}\label{D11}\left(D_\epsilon\right)_{11} = i\left(1-\mu
858: ^2\right)\left(e^{2it} - e^{-2it}\right),\end{equation} and
859:
860: \begin{equation}\label{D21}\left(D_\epsilon\right)_{21} = \mu ^2 \left(1-\mu ^2 \right)
861: \left(e^{2it} + e^{-21t} -2 \right) + 2\mu \nu.
862: \end{equation}
863:
864:
865:
866: \bigskip
867:
868: tFrom the matrices $D_0$ and $D_\epsilon$ we can construct the matrices
869: $M_0\left(\mu \right)$ and $M_\epsilon\left(\mu\right)$ needed in the
870: formula (\ref{M0andM1}) for $M_1$. For $M_0$ we simply have $\exp {2\pi
871: D_0}$. For $M_\epsilon$ we proceed as follows.
872: On the finite time-interval $[0,2\pi]$ we may write
873: \[ \Phi \left(t,\epsilon ,\mu \right) = \Phi _0\left(t,\mu\right) + \epsilon \Phi _1 \left(t, \mu \right) + \ldots, \; \Phi _1 \left(0, \mu \right)=0. \]
874: Substituting this in the
875: differential equation (\ref{c_eq}), expanding to first order in
876: $\epsilon$, using the variation of constants formula (cf. \citet{cl}) and setting $t=2\pi$, we get
877:
878: \begin{equation}\label{floquet_matrix}
879: M\left(\epsilon , \mu \right) = M_0\left(\mu \right)\left(I + \epsilon \int _0^{2\pi} \Phi_0^{-1}\left(s, \mu \right)D_\epsilon\left(s,\mu \right) \Phi _0
880: \left(s, \mu \right)ds\right).\end{equation}
881: This expresses the Floquet matrix correctly to linear order in $\epsilon$, and this will turn out to be sufficient for our purpose. The formula above identifies $M_\epsilon\left(0,\mu\right)$:
882: \begin{equation}\label{M_epsilon}
883: M_\epsilon \left(0,\mu \right) = M_0 \left(\mu \right) \int _0^T \Phi
884: _0^{-1}\left(s, \mu \right)D_\epsilon\left(s,\mu \right) \Phi _0
885: \left(s, \mu \right)ds .\end{equation}
886: We next proceed to simplify this expression.
887:
888:
889: \subsection{A further transformation}\label{distinct}
890: The characteristic polynomial given in equation (\ref{charpol}) above
891: is the same in any coordinate system, so we shall choose one to simplify the unperturbed Floquet matrix $M_0\left(\mu\right)$.
892:
893: If $\mu \ne 0$ and $\eta \ne 0$ the eigenvalues $\{\sigma _k\}$ given
894: by equation (\ref{evals}) are all distinct, so the
895: eigenvectors are linearly independent and the matrix $T\left(\mu\right)$
896: formed from their columns diagonalizes $D_0$:
897: \[\tilde{D}_0 =
898: \mbox{diag}\left(\sigma_1,\sigma_2,\sigma_3,\sigma_4\right),\]
899: where the tilde indicates the transformed matrix: $\tilde{D}=T^{-1}DT$.
900: We shall need to know $T$ and $T^{-1}$ explicitly.
901: It is a
902: straightforward matter to show that any eigenvector of $D_0$ must have
903: the structure (up to a constant multiple)
904: \[ \xi = \left(\begin{array}{c} \sigma \\-\left(\sigma ^2 +
905: m^2\right)/2\\im \\ -i m\left(\sigma ^2 + m^2 \right)/2\sigma
906: \end{array} \right).\]
907: Substituting the particular values of $\sigma$ given in equation
908: (\ref{evals}) gives the four columns of the matrix $T$, and from this
909: we can construct its inverse. One finds
910: \begin{equation}\label{Tmatrix}
911: T = \left(\begin{array}{cccc}\sigma _1&\sigma _2 & \sigma _3 &\sigma
912: _4\\ -i\mu \sigma_1 & i\mu \sigma_2 & -i\mu \sigma_3 & i\mu \sigma_4
913: \\ im & im & im& im \\m\mu & -m\mu & m\mu & -m\mu \end{array}\right),
914: \end{equation}
915: and
916: \begin{equation}\label{Tinverse}
917: T^{-1} = \frac{1}{4mq} \left( \begin{array}{cccc} -im & m/\mu & \sigma
918: _3 & i\sigma _3/\mu \\ im & m/\mu & \sigma _3 & -i\sigma _3/\mu \\ im
919: & -m/\mu & -\sigma _1 & -i \sigma _1 /\mu\\ -im & -m/\mu & -\sigma _1 &
920: i\sigma _1/ \mu
921: \end{array}\right).\end{equation}
922: The matrices $T$ and $T^{-1}$ depend on $\mu$ and on $\eta$ through the parameters
923: \begin{equation}\label{mqdef}m=\mu \eta \; \mbox{and} \; q=\sqrt{\mu ^2 + m^2} = \mu \sqrt{1+\eta ^2}.\end{equation}
924:
925: \medskip
926: In place of equation (\ref{floquet_matrix}) we now obtain
927: \begin{equation}\label{Mtilde}
928: \tilde{M} = \tilde{M}_0\left(I + \epsilon \tilde{J}\right),
929: \end{equation}
930: where
931: \begin{equation}\label{Jeq}
932: \tilde{J}\left(\mu \right) = \int _0^{2\pi}
933: \tilde{\Phi}^{-1}\left(t,\mu \right)
934: \tilde{D}_1\left(t,\mu \right) \tilde{\Phi} \left(t,\mu \right)
935: \,dt.\end{equation}
936: Because the eigenvalues $\{\sigma _k \}$ are distinct, the matrix
937: $\tilde{\Phi} = \exp \left\{ \tilde{D}_0 t \right\}$ takes the simple,
938: diagonal form
939: \begin{equation}\label{fundmatdistinct}
940: \tilde{\Phi}\left(t\right) = \mbox{diag}\, \left(\exp \left(\sigma _1 t
941: \right),\exp \left(\sigma _2 t
942: \right),\exp \left(\sigma _3 t
943: \right),\exp \left(\sigma _4 t
944: \right) \right)
945: \end{equation}
946: The $ij$ entry of the matrix
947: $\tilde{D}_1$
948: is (since $D_\epsilon$ has only two nonzero entries)
949: \begin{equation}\label{D1ij}
950: \left(\tilde{D}_1\right)_{ij} = T_{1j}\left(\left( T^{-1}\right)_{i1}
951: \left(D_\epsilon\right)_{11} + \left(T^{-1}\right)_{i2} \left(D_\epsilon\right)_{21}\right).
952: \end{equation}
953: As a result, the $ij$ entry of the matrix $\tilde{J}$ providing the
954: leading-order perturbation of the Floquet matrix is
955: \begin{eqnarray}\label{Jij}
956: \tilde{J}_{ij} &=& T_{1j}\left(T^{-1}\right)_{i1}\int _0^{2\pi}
957: e^{\left(\sigma _j - \sigma _i
958: \right)t}\left(D_\epsilon\right)_{11}\left(t\right) \,dt \nonumber \\ &+&T_{1j}\left(T^{-1}\right)_{i2}\int _0^{2\pi}
959: e^{\left(\sigma _j - \sigma _i
960: \right)t}\left(D_\epsilon\right)_{21}\left(t\right) \,dt .
961: \end{eqnarray}
962: This enables us to find $\tilde{M}_\epsilon \left(0,\mu\right)$.
963:
964: For the matrix $\tilde{M}_0\left(\mu\right)$ we have the expression
965: \begin{equation}\label{tildeM0}
966: \tilde{M}_0\left(\mu\right)=\tilde{M}\left(0,\mu\right) = \mbox{diag} \left( \lambda _1 \left(\mu \right), \ldots , \lambda _4\left(\mu \right) \right)
967: \end{equation}
968: with $\lambda _k = 2\pi \sigma _k$. Recall that in order to construct $\tilde{M}_1$ we need also the derivative of this matrix with respect to $\mu$,
969: \begin{equation}\label{tildeM0prime}
970: \tilde{M}_0^\prime \left(\mu\right)=\tilde{M}_\mu \left(0,\mu\right) = \mbox{diag} \left( \lambda _1^\prime \left(\mu \right), \ldots , \lambda _4^\prime \left(\mu \right) \right).
971: \end{equation}
972: According
973: to equations (\ref{evals}) and (\ref{mqdef}), each eigenvalue $\sigma _k$ of $D_0$ is
974: linear in $\mu$. Therefore $\sigma _k ^\prime \left(\mu \right) =
975: \sigma _k \left(\mu \right)/\mu $. Since $\lambda _k = \exp \left(2\pi
976: \sigma _k \right)$, we have
977: \begin{equation}\label{lambda_prime}
978: \lambda _k ^\prime \left(\mu \right) = \lambda _k \left(\mu \right) 2
979: \pi \sigma _k \left(\mu \right) /\mu = \lambda _k 2\pi i \omega _k /\mu .\end{equation}
980:
981: The formulas of this section allow one to determine the matrices
982: $\tilde{M}_0$ and $\tilde{M}_1$. To produce from these the
983: coefficients $p_j\left(\lambda\right)$ appearing in equations
984: (\ref{p_exp_2}) and (\ref{beta1_eq}) above, we need formulas for the
985: derivatives of a determinant. These are presented in Section \ref{detderivs} below and applied in the following section.
986:
987: \subsection{The Expansion for $p\left(\lambda,\epsilon\right)$}\label{subsec:charpoly}
988:
989: We can now find the required expansion for $p\left(\lambda , \epsilon
990: \right)$ by identifying the matrix $A$ of Section \ref{detderivs} below with $\tilde{M} -\lambda I$ and
991: the coefficients $q_k$ with the coefficients $p_k \left(\lambda
992: \right)$ of equation (\ref{p_exp_2}). We obtain these coefficients by
993: writing $a_k = \lambda _k - \lambda,$ $A_{kl} ^\prime\left(0\right) =
994: \left(\tilde{M}_1\right) _{kl}$ and $A_{kl} ^{\prime
995: \prime}\left(0\right) = 2\left(\tilde{M}_2\right) _{kl}$. This gives
996: for $p_1$ the following expression (with $n=4$):
997:
998: \begin{eqnarray}\label{p1}
999: p_1 \left(\lambda\right) &=&
1000: \left(\tilde{M}_1\right)_{11}\left(\lambda _2- \lambda \right)
1001: \left(\lambda _3 - \lambda \right) \left(\lambda _4 - \lambda
1002: \right) \\&+& \left(\tilde{M}_1\right)_{22}\left(\lambda _1 - \lambda
1003: \right) \left(\lambda _3 - \lambda \right) \left(\lambda _4 -
1004: \lambda \right) + {\nonumber}\\&+&
1005: \left(\tilde{M}_1\right)_{33}\left(\lambda _1- \lambda \right)
1006: \left(\lambda _2 - \lambda \right)\left(\lambda _4 -
1007: \lambda \right){\nonumber}\\&+&
1008: \left(\tilde{M}_1\right)_{44}\left(\lambda _1- \lambda \right)
1009: \left(\lambda _2 - \lambda \right)\left(\lambda _3 -
1010: \lambda \right){\nonumber}
1011: ,\end{eqnarray} and there is a similar,
1012: lengthier expression for $p_2$ obtained by making the corresponding substitutions in equation (\ref{det_der_2}) below.
1013:
1014: The development thus far has required
1015: no assumptions regarding the multipliers $\left\{\lambda _k \right\}$.
1016: We now suppose that $\lambda _1 = \lambda _2$. It is then clear from
1017: the expression above that $p_1\left(\lambda _1 \right) =0,$ as
1018: asserted in Section \ref{char_poly}. The coefficients appearing in
1019: equation (\ref{beta1_eq}) are now easily found to be
1020: \begin{equation}\label{p0pp_eq}
1021: p_0^{\prime \prime }\left(\lambda _1\right) = 2\left(\lambda _3 -
1022: \lambda _1\right) \left( \lambda _4 - \lambda _1 \right),
1023: \end{equation}
1024: \begin{equation}\label{p1p_eq}
1025: p_1^\prime \left(\lambda _1\right) =
1026: -\left\{\left(\tilde{M}_1\right)_{11} +
1027: \left(\tilde{M}_1\right)_{22}\right\} \left(\lambda _3 - \lambda
1028: _1\right)\left( \lambda _4 - \lambda _1 \right) \end{equation}
1029: and
1030: \begin{equation}\label{p2_eq}
1031: p_2 \left(\lambda _1\right) = \left| \begin{array}{cc}
1032: \left(\tilde{M}_1\right)_{11}&\left(\tilde{M}_1\right)_{12}\\\left(\tilde{M}_1\right)_{21}&\left(\tilde{M}_1\right)_{22}
1033: \end{array} \right| \left(\lambda _3 - \lambda
1034: _1\right) \left( \lambda _4 - \lambda _1
1035: \right). \end{equation} Equation (\ref{beta1_eq}) therefore takes the
1036: form
1037: \begin{equation}\label{beta1_eq_2}
1038: \beta _1 ^2 + \left\{\left(\tilde{M}_1\right)_{11} -
1039: \left(\tilde{M}_1\right)_{22}\right\} \beta _1 + \left|
1040: \begin{array}{cc}
1041: \left(\tilde{M}_1\right)_{11}&\left(\tilde{M}_1\right)_{12}\\\left(\tilde{M}_1\right)_{21}&\left(\tilde{M}_1\right)_{22}
1042: \end{array} \right| = 0 .
1043: \end{equation}
1044:
1045: We note that the calculation of the perturbation of $\lambda _1$ to first order in $\epsilon$, which
1046: requires expanding
1047: $p$ to second order, requires the expansion of the
1048: Floquet matrix $\tilde{M}\left(\epsilon\right) = \tilde{M}_0 +
1049: \tilde{M}_1 \epsilon + \cdots $ only to first order.
1050:
1051: We have assumed that the coincident roots are the first two, $\lambda
1052: _1 = \lambda _2$. If instead we should have $\lambda _k = \lambda _l$,
1053: equation (\ref{beta1_eq_2}) is modified by the replacement $(1,2) \to
1054: (k,l)$.
1055:
1056: Equation (\ref{beta1_eq_2}), together with equations (\ref{Mtilde}),
1057: (\ref{tildeM0}), (\ref{tildeM0prime}) and (\ref{lambda_prime}) leads
1058: to equation (\ref{beta1_eq_3}) of the text. What remains is to
1059: evaluate the integrals defining $\tilde{J}$, and we now turn to this.
1060:
1061: \subsection{Calculating the Elements of $\tilde{M}_1$}\label{subsec:elements}
1062: By equations (\ref{M0andM1}) and (\ref{M_epsilon}) above (see also
1063: equation \ref{Jeq}), the matrix $\tilde{M}_1$ is given by the formula
1064: \begin{equation}\label{M1tilde_def}
1065: \tilde{M}_1 = \tilde{M}_0 \left(\mu \right)\tilde{J} + \nu \tilde{M}_\mu \left(0,\mu \right)
1066: \end{equation}
1067: where the entries of $\tilde{J}$ are given by equation
1068: (\ref{Jij}). From equations (\ref{D11}) and (\ref{D21})
1069: it is a straightforward matter to carry out the integrations. We'll use for $T$ the matrix given above in equation (\ref{Tmatrix}). Since for this matrix $T_{1j} = \sigma _j= i \omega _j$, the formula for the entries of $\tilde{J}$ becomes
1070: \begin{equation}\label{newJtilde}
1071: \tilde{J}_{ij} = \sigma _j \left\{ \left(T^{-1}\right) _{i1} \int _0
1072: ^{2\pi} e^{\left(\sigma _j - \sigma _i \right) t }
1073: \left(D_\epsilon\right)_{11}\left(t\right) \,dt + \left(T^{-1}\right)_{i2}
1074: \int _0^{2\pi} e^{\left(\sigma _j - \sigma _i \right) t
1075: }\left(D_\epsilon\right)_{21} \,dt\right\} .
1076: \end{equation}
1077:
1078: \medskip
1079:
1080: For the diagonal entries the exponential factors in the integrand reduce to unity and one finds
1081: \begin{equation}\label{diagonal_eq}
1082: \tilde{J}_{jj} = -4\pi \mu^2 \left(1-\mu ^2 \right) \sigma _j \left(T^{-1} \right) _{j2},
1083: \; j=1,2,3,4.\end{equation}
1084:
1085:
1086:
1087: \medskip
1088: For the off-diagonal entries, the formulas may be found generally, but
1089: we need them only in the resonant cases where, for some pair of
1090: indices $(i,j)$, $\sigma _i - \sigma _j = ki$ for a non-zero integer
1091: $k$.\footnote{Recall that $\sigma _i \ne \sigma _j$ for any pair
1092: $i,j$ so $k=0$ is excluded.} It is clear from the formulas (\ref{D11}) (\ref{D21}) and
1093: (\ref{newJtilde}) that only resonances with $k=\pm 2$ contribute
1094: off-diagonal terms to leading order in $\epsilon$ since for any other
1095: choice of $k$ the integrals vanish: for $k \ne 2$ the matrix
1096: $\tilde{J}$ is diagonal.
1097:
1098:
1099: \section{Determinantal Derivatives}\label{detderivs}
1100:
1101: Consider an $n
1102: \times n$ matrix $A\left(\epsilon \right)$ having the properties that
1103: it is a smooth function of $\epsilon$ and is diagonal at $\epsilon =
1104: 0$: $A\left(0\right) = \mbox{diag}\,\left(a_1,a_2, \ldots ,
1105: a_n\right)$.
1106:
1107:
1108: We need the coefficients in the Taylor expansion of
1109: det$A\left(\epsilon\right) \equiv q\left(\epsilon\right)$:
1110: \begin{equation}\label{q_exp}
1111: q\left(\epsilon \right) = q_0 + q_1 \epsilon + q_2 \epsilon ^2 +
1112: \cdots
1113: \end{equation}
1114: where
1115: \begin{equation}\label{detA}
1116: q_0= \left|A\left(0\right)\right|, \; q_1 = \frac{d
1117: \left|A\left(\epsilon\right)\right|}{d\epsilon}|_{\epsilon =0}, \; q_2 =
1118: \frac{1}{2} \frac{d^2
1119: \left|A\left(\epsilon\right)\right|}{d \epsilon ^2}|_{\epsilon =0}, \;
1120: \cdots\end{equation}
1121: Straightforward applications of the formula for the derivative of a
1122: determinant show that
1123: \begin{eqnarray}\label{det_der_1}
1124: q_1 &=& A_{11}^\prime a_2 a_3 \cdots a_n +
1125: A_{22}^\prime a_1 a_3 \cdots a_n + \cdots +
1126: A_{nn}^\prime a_1 a_2 \cdots a_{n-1},\\
1127: q_2 &=& \frac{1}{2}\left[ A_{11}^{\prime \prime} a_2 a_3 \cdots a_n +
1128: A_{22}^{\prime \prime} a_2 a_3 \cdots a_n + \cdots +
1129: A_{nn}^{\prime \prime} a_1 a_2 \cdots a_{n-1}\right] \nonumber \\ &+& \left|
1130: \begin{array}{cc}A_{11}^\prime &A_{12}^\prime
1131: \\A_{21}^\prime & A_{22}^\prime
1132: \end{array}\right| a_3 a_4 \cdots a_n +\left|
1133: \begin{array}{cc}A_{11}^\prime &A_{13}^\prime
1134: \\A_{31}^\prime & A_{33}^\prime
1135: \end{array}\right|a_2 a_4 \cdots a_n \\ &+& \cdots +\left|
1136: \begin{array}{cc}A_{(n-1)(n-1)}^\prime &A_{(n-1)n}^\prime
1137: \\A_{n(n-1)}^\prime & A_{nn}^\prime
1138: \end{array}\right|a_1 a_2 \cdots a_{n-2} \label{det_der_2},
1139: \end{eqnarray}
1140: where the terms involving two-by-two determinants represent the sum over $k < l$ of the
1141: product of the $\left\{a_j\right\}$, $a_k$ and $a_l$ omitted, with the
1142: determinant
1143: \[ \left| \begin{array}{cc} A_{kk}^\prime & A_{kl}^\prime
1144: \\A_{lk}^\prime & A_{ll}^\prime
1145: \end{array}\right|\]
1146: and all derivatives are evaluated at the origin.
1147:
1148: \begin{thebibliography}{}
1149:
1150: \bibitem[Balbus \& Hawley(1991)]{bah} Balbus, S. \& Hawley, J. 1991, \apj,
1151: 376, 214
1152:
1153: \bibitem[Balbus \& Hawley(1998)]{mri} Balbus, S. and Hawley, J. 1998,
1154: Rev. Mod. Phys., 70, 1
1155:
1156: \bibitem[Bayly(1986)]{bay} Bayly, B.J. 1986, \prl, 57, 2160
1157:
1158: \bibitem[Chandrasekhar(1960)]{cha} Chandrasekhar, S. 1960, Proc. Nat. Acad.
1159: Sci. 46, 253
1160:
1161: \bibitem[Coddington and Levinson(1955)]{cl} Coddington, E. and Levinson, N. 1955,
1162: Theory of Ordinary Differential Equations, New York: Addison-Wesley
1163:
1164: \bibitem[Craik(1988)]{addc88} Craik, A.D.D. 1988, Proc. Roy. Soc. A 417, 235
1165:
1166: \bibitem[Kerswell(1994)]{rk94} Kerswell, R.R. 1994, J. Fluid Mech. 274, 219
1167:
1168: \bibitem[Kerswell(2002)]{rk02} Kerswell, R.R. 2002, Annu. Rev. Fluid Mech 34, 83
1169:
1170: \bibitem[Goodman(1993)]{jg} Goodman, J. 1993 \apj, 406, 596
1171:
1172: \bibitem[Hille(1972)]{hil} Hille, E.1962, Analytic Function Theory, volume II, New York: Ginn
1173:
1174: \bibitem[Lubow, Pringle, and Kerswell(1993)]{lpk} Lubow, S.H., Pringle, J.E.,
1175: and Kerswell, R.R. 1993, \apj, 419, 758
1176:
1177: \bibitem[Morris and Serabyn(1996)]{mor} Morris, M. \& Serbyn, E. 1996, \araa,
1178: 34, 645
1179:
1180: \bibitem[Pierrehumbert(1986)]{rtp} Pierrehumbert, R.T. 1986, \prl, 57, 2157
1181:
1182: \bibitem[Ryu and Goodman(1994)]{ryug} Ryu, D. and Goodman, J. 1994 \apj, 422, 269
1183:
1184: \bibitem[Ryu, Goodman, and Vishniac(1996)]{ryugv} Ryu, D., Goodman, J. \&
1185: Vishniac, E.T. 1996, \apj, 461, 805
1186:
1187: \bibitem[Velikhov(1959)]{vel} Velikhov, E.P. 1959, Sov. Phys. JETP, 36, 995
1188:
1189: \bibitem[Waleffe(1990)]{wal} Waleffe, F. 1990, Phys. Fluids 2, 76
1190:
1191: \bibitem[Yakubovich and Starzhinsii(1975)]{ys} Yakubovich, V.A. and Starzhinsky, V.M. 1975,
1192: Linear Differential Equations with Periodic Coefficients, New York:Wiley
1193:
1194: \end{thebibliography}
1195:
1196: \end{document}
1197: