cond-mat0203040/pre.tex
1: \documentclass[pre,twocolumn,showpacs,preprintnumbers,floatfix]{revtex4}
2: \usepackage{graphicx,citesort,psfrag,amsmath,amssymb}
3: 
4: \newcommand{\rem}[1]{\textbf{#1}}
5: \newcommand{\ve}[1]{\mathbf{#1}}
6: \newcommand{\dd}[2]{\frac{\partial #1}{\partial #2}}
7: \newcommand{\tdd}[2]{\partial #1/\partial #2}
8: \newcommand{\avg}[1]{\left\langle #1 \right\rangle}
9: 
10: \sloppy
11: 
12: 
13: \begin{document}
14: 
15: \bibliographystyle{apsrev}
16: 
17: %\preprint{}
18:  
19: \title{Minimal model for aeolian sand dunes}
20: 
21: \author{Klaus Kroy}
22: \affiliation{Dept.\ of Physics and Astronomy, University of
23: Edinburgh, EH9 3JZ, United Kingdom}
24: \author{Gerd Sauermann, and Hans J. Herrmann}
25: \affiliation{ICA-1, Universit\"at Stuttgart, Pfaffenwaldring 27, 70569
26: Stuttgart, Germany} 
27: 
28: \date{\today} 
29:              
30:  
31: \begin{abstract}
32:  We present a minimal model for the formation and migration of aeolian
33:   sand dunes. It combines a perturbative description of the turbulent
34:   wind velocity field above the dune with a continuum saltation model
35:   that allows for saturation transients in the sand flux. The latter
36:   are shown to provide the characteristic length scale. The model can
37:   explain the origin of important features of dunes, such as the
38:   formation of a slip face, the broken scale invariance, and the
39:   existence of a minimum dune size. It also predicts the longitudinal
40:   shape and aspect ratio of dunes and heaps, their migration velocity
41:   and shape relaxation dynamics. Although the minimal model employs
42:   non--local expressions for the wind shear stress as well as for the
43:   sand flux, it is simple enough to serve as a very efficient tool for
44:   analytical and numerical investigations and to open up the way to
45:   simulations of large scale desert topographies.
46: \end{abstract}
47:  
48: \pacs{45.70.Mg, 45.70.Qj, 47.27.-i, 51.10.+y}
49: 
50: %\keywords{Suggested keywords}
51: 
52: \maketitle
53: 
54: 
55: \section{Introduction}
56:   
57: Sand dunes develop wherever sand is exposed to an agitating medium
58: such as air or water that lifts grains from the ground and entrains
59: them into a surface flow. The diverse conditions of wind and of sand
60: supply in different regions on Earth give rise to a large variety of
61: shapes of aeolian dunes \cite{Bagnold41,Pye90,Lancaster95}. Moreover,
62: dunes have been found on the sea--bottom and even on Mars
63: \cite{malin-etal:98,Thomas99}.  Despite the long history of the
64: subject, the underlying physical mechanisms of dune formation are
65: still not very well understood.  How are aerodynamics (hydrodynamics)
66: and the particular properties of granular matter acting together to
67: create dunes?  How is the shape of a dune maintained when it moves?
68: Since the macroscopic phenomena of interest are separated by many
69: orders of magnitude from the grain scale and involve various coupled
70: nonlinear processes such as turbulent air flow and grain hopping
71: (``saltation''), one is bound to devise some simplified models in
72: order to address such questions. We will argue that approximate
73: numerical models can only be successful if based on a sound
74: qualitative understanding of the problem. Therefore, our main aim is
75: to identify the key mechanisms underlying dune formation and migration
76: and incorporate them into a working minimal model of aeolian sand
77: dunes, and we will emphasize generic aspects over the more specific
78: details. For definiteness, the reader may find it helpful to think of
79: isolated transverse dunes or crescent--shaped barchan dunes as major
80: applications of the model. A schematic sketch of the height profile of
81: a barchan is shown in Fig.~\ref{fig:barchan}. The broad phenomenology
82: of aeolian and submarine land forms provides a large number of
83: different characteristic structures that can certainly not all be
84: described by the same simple model developed with the specific
85: examples of barchan or transverse dunes in mind. However, we expect
86: that our approach is amenable to future adaptations that make it
87: applicable to a broader class of sand topographies on the one hand,
88: and for quantitative investigations of more specific questions on the
89: other hand. Although the minimal model refers only to rather generic
90: properties of the wind velocity field and the laws of aeolian sand
91: transport, it can make interesting predictions about the surface
92: profile, the development and position of the slip face, dune migration
93: etc.\ that are insensitive to the simplifying assumptions. The main
94: features of the model were already briefly presented in a recent
95: Letter \cite{kroy-sauermann-herrmann:2002}.  The present contribution
96: gives a more comprehensive discussion of the model and tries to
97: communicate its precise definition as well as its major predictions to
98: an interdisciplinary readership. The model, as presented here, is
99: restricted to a two--dimensional ($2d$) slice of a dune parallel to
100: the unidirectional wind.  (A generalization to $3d$ problems is in
101: preparation.) A further restriction is the neglect of ripples and
102: direct slope effects onto the sand transport outside slip
103: faces. Although they have successfully been incorporated into
104: continuum sand transport models
105: \cite{hoyle-woods:97,hoyle-mehta:99,prigozhin:99} similar to our own
106: \cite{sauermann-kroy-herrmann:2001}, we chose to disregard them for
107: the present purpose and leave their integration to future work.
108: 
109: The paper is organized as follows. In the next introductory section we
110: summarize some background knowledge and basic definitions. We will
111: also introduce a naive ``zeroth order'' description of the wind shear
112: stress and the induced aeolian sediment transport. Its instructive
113: failure to produce dune--like steady--state solutions will be a guide
114: for identifying two relatively small effects (the upwind shift of the
115: maximum of the shear stress with respect to the topography and the
116: saturation transients in the sand flux) as key ingredients of a proper
117: description of structure formation by aeolian sand transport.  We will
118: moreover derive the scaling behavior of the migration velocity for
119: translation invariant heaps and dunes of different size but similar
120: shape based on very general grounds. Sections~\ref{sec:wind} and
121: \ref{sec:saltation} are devoted to the definition of the minimal
122: model, i.e., to the modeling of the air shear stress exerted onto a
123: heap of sand and the induced sand transport, respectively. The first
124: step builds on turbulent boundary layer calculations developed in a
125: series of publications mainly by Hunt and coworkers
126: \cite{jackson-hunt:75,%
127: sykes:80,zeman-jensen:88,hunt-leibovich-richards:88,Carruthers90,weng-etal:91},
128: the second one on a previous contribution
129: \cite{sauermann-kroy-herrmann:2001} by the present authors. Only the
130: most pertinent results of these earlier developments will be
131: summarized here.  In the remainder, we will derive some important
132: predictions of the model for the central slice of a barchan dune or
133: transverse dune. In particular, we will demonstrate that there is a
134: minimal dune size. Although we will thereby gain interesting results,
135: these are rather meant to be illustrative examples of possible
136: applications of the model. By no means do we attempt to provide a
137: complete analysis of its predictions, and it should become obvious
138: that much more remains still to be done.  Finally we will summarize
139: our main results and speculate about probable consequences of the
140: present $2d$ theory for $3d$ topographies.
141: 
142: \begin{figure}[t]
143:   \includegraphics[width=\columnwidth]{sketch.eps} 
144: \caption{Sketch of a barchan dune. Sand is eroded by the wind on the
145: upwind or ``stoss'' side and transported to the brink. Strong
146: deposition occurs due to flow separation behind the brink. On the
147: downwind or ``lee'' side, sand slides down at the angle of repose
148: (about $32^{\circ}-35^{\circ}$) over a concave slip face.}
149: \label{fig:barchan}
150: \end{figure}
151: 
152: 
153: \section{General}\label{sec:general}
154: \subsection{Aeolian sand transport}\label{sec:aeolian}
155: Before going into the description of the model, we want to recall some
156: general background and to introduce some quantities of major interest.
157: First of all, for convenience, we will usually refer to dunes without
158: slip face as heaps. Further, we will sometimes find it helpful to
159: focus on isolated heaps or dunes on bedrock, although most of our
160: discussion is not restricted to this situation.  
161: 
162: The key quantity for the description of the formation and migration of
163: sand dunes and heaps is the local horizontal surface velocity $v(x,t)$
164: of a sand height profile $h(x,t)$ at all positions $x$ and times
165: $t$. Via mass conservation it can be related to the erosion rate
166: $\nabla q(x,t)$ (negative erosion is deposition), where the sand flux
167: $q(x,t)$ is defined as the mass of sand transported per unit of time
168: across a hyper--plane transverse to the wind direction. More
169: precisely, since we want to specialize our discussion to a $2d$ slice
170: parallel to the unidirectional wind velocity, the hyperplane is a
171: vertical line and $q$ is a flux per unit width. Mass conservation then
172: takes the form of a continuity equation for the height profile
173: \begin{equation}\label{eq:mass}
174: \varrho_{s}\frac{d h(x,t)}{dt} = -\frac{\partial q(x,t)}{\partial x} 
175: \end{equation}
176: with $\varrho_{s}$ the density of the sand bed. 
177: 
178: With Eq.(\ref{eq:mass}) one can write the position dependent migration
179: velocity at a given time $t$ as
180: \begin{equation}\label{eq:v_def}
181: v(x) = \varrho_{s}^{-1} \; \frac{q'}{h'}
182: % = \varrho_{s}^{-1} \; \frac{dq}{dh} 
183: \;,
184: \end{equation}
185: where we have introduced the shorthand notation $f'(x)\equiv
186: df(x)/dx$.  At this stage we can already get some physical insight by
187: observing that this equation needs special attention at the top of a
188: heap or dune, where we expect the denominator to vanish. For $v$ to
189: remain finite at the crest as required in the steady state, there are
190: in general only two possibilities. Either the sand flux $q$ is fine
191: tuned so that the erosion $q'$ vanishes in exactly the same way as the
192: slope $h'$, or the profile $h(x)$ is not differentiable at the
193: crest. As the reader may already anticipate and will be verified
194: below, both cases have their physical realizations, the former
195: in heaps or small dunes with smooth crests and the latter in large
196: dunes with a slip face that terminates in a sharp brink.
197: 
198: The problem we face, if we want to calculate the dynamic evolution of
199: desert topographies, is the closure of Eq.(\ref{eq:mass}) or
200: (\ref{eq:v_def}) by expressing the flux $q(x,t)$ in terms of the
201: height profile $h(x,t)$ and the external wind and boundary
202: conditions. Since for the applications we have in mind, the migration
203: velocity is very small compared to the speed of elementary sand
204: transport processes (grain hopping etc.) and the wind speed, the
205: topography can be assumed to be stationary for considerations
206: concerning the wind and sand transport dynamics. This allows one to
207: subdivide the problem of calculating $q(x)$ into two independent
208: steps. First, one needs to know the stationary wind velocity above a
209: given topography. More precisely, what is required is the shear stress
210: $\tau$ exerted by the wind onto the ground. And secondly, one needs a
211: model that predicts the stationary sand flux $q(x)$ for a given
212: stationary $\tau(x)$, schematically
213: \begin{align}
214: h(x) &\to \tau(x) \label{eq:task1} \\
215: \tau(x) &\to q(x)\;. 
216: \label{eq:task2}
217: \end{align} 
218: Computing the derivative $q'$ and integrating the mass conservation
219: Eq.(\ref{eq:mass}) then closes the model and allows one to predict the
220: development of the surface profile in time.  Since aeolian dunes
221: typically have relatively gentle slopes outside their slip face, we
222: will at the present stage restrict the scope of the minimal model to
223: this case and disregard in Eq.(\ref{eq:task2}) the direct slope
224: effects $h'(x) \to q(x)$ onto the flux outside the slip face.
225: 
226: In special cases, the relations (\ref{eq:task1}), (\ref{eq:task2}) are
227: phenomenologically and theoretically well established.  For a flat
228: surface, $h(x)\equiv$ const., it is well known
229: \cite{landau-lifshitz:fm} that the mean turbulent wind velocity
230: increases logarithmically with height above the surface. It can be
231: characterized by a single characteristic velocity, the ``shear
232: velocity'' $u_*$ defined by $u_*^2\equiv\tau_0/\varrho_{a}$ with
233: $\tau_0$ the (suitably time averaged) shear stress and $\varrho_{a}$
234: the density of air. Since the shear stress of the air is transmitted
235: to the surface, the latter can mobilize grains on a surface covered
236: with sand, if it exceeds a threshold value $\tau_t$. As a result, the
237: wind entrains some grains into a surface layer flow.  The grains
238: advance mainly by an irregular hopping process (``saltation''),
239: thereby reducing the wind velocity in the surface layer. Via this
240: feed--back mechanism a unique relation between the shear stress $\tau$
241: and the sand flux $q$ is established in the equilibrium state.  If
242: $\tau$ is not too close to the threshold, this relation can
243: approximately be represented as \cite{Bagnold41}
244: \begin{equation}\label{eq:bagnold} 
245: q_s \propto \tau^{3/2} \;.     
246: \end{equation} 
247: Although a host of more accurate descriptions have been discussed in
248: the literature
249: \cite{Lettau78,Sorensen91,Pye90,ActaMechanica91,sauermann-kroy-herrmann:2001}
250: and one of them will be part of our definition of the minimal model
251: below, the simpler Eq.(\ref{eq:bagnold}) will be sufficient for our
252: qualitative discussion in the first part of the paper. The index $s$
253: in Eq.(\ref{eq:bagnold}) emphasizes that such local relations are
254: restricted to situations where the flux is saturated, that is, equal
255: to its equilibrium transport capacity. This is certainly not the case
256: near a boundary between uncovered and covered ground or on sloped
257: beds. Neglecting this restriction for the moment,
258: Eq.(\ref{eq:bagnold}) predicts that the shear stress perturbation
259: \begin{equation}\label{eq:tauhat} 
260:   \hat\tau(x) \equiv \tau(x)/\tau_0-1
261: \end{equation} 
262: above a modulated topography $h(x)$ is responsible for flux gradients
263: $dq_s/dx$ that cause erosion and deposition and thus --- according to
264: Eqs.(\ref{eq:mass}), (\ref{eq:v_def}) --- migration of the sand surface.
265: Explicitly closing the model by assuming that the shear stress is an
266: affine function of the modulation of the topography ($\hat \tau
267: \propto h$) leads to what we call the ``zeroth order'' model, which
268: will briefly be analyzed in the next paragraph.
269: 
270: \subsection{The ``zeroth order'' model}\label{sec:zero}
271: The zeroth order model is given by
272: \begin{align}
273: \hat\tau \{h(x)\} & \to \hat\tau(h)\propto h(x)/L \label{eq:tau0} \\ 
274: q\{\tau(x)\}  & \to q(\tau)= q_s(\tau) \label{eq:q0}
275: \end{align}
276: where we have used the curly brackets to indicate a general functional
277: dependence and introduced a characteristic length scale $L$ of the
278: topography to normalize the height profile. (The motivation for the
279: latter step will become clear in the next section.) The zeroth order
280: model assumes \emph{local} relations in
281: Eqs.(\ref{eq:task1}),(\ref{eq:task2}). It approximates the wind shear
282: stress perturbation by its ``affine'' contribution (proportional to
283: the profile $h$ that causes the perturbation) and replaces the true
284: sand flux $q$ by its saturated value $q_s$, thereby neglecting
285: saturation transients.  This model is so simple that its qualitative
286: predictions for an arbitrary smooth heap of sand can easily be
287: anticipated without doing any actual calculations. 
288: 
289: Combining Eq.(\ref{eq:v_def}) with
290: Eqs.(\ref{eq:bagnold})--(\ref{eq:q0}) one obtains a surface velocity
291: that increases with height ($dv/dh \geq 0$) due to the nonlinearity
292: of Eq.(\ref{eq:bagnold}). This implies that the upwind (or ``stoss'')
293: slope tends to decrease and the downwind (or ``lee'') slope tends to
294: increase. Since $dq/dx \propto dh/dx$ by the chain rule, there is no
295: erosion or deposition at the top of a smooth heap, which therefore
296: keeps its initial height. Obviously, integrating forward in time will
297: eventually increase the lee slope up to the angle of repose, where
298: surface avalanches have to be introduced and a slip face of constant
299: slope develops.  If the latter reaches the crest the above argument
300: for the persistence of the height can no longer be applied, because
301: the slope at the crest is then ill defined. Since there is so far
302: nothing to stop a further decrease of the windward slope, the model
303: dune will then start to decrease in height and finally flatten
304: out. The steady--state solution is a flat surface.
305: 
306: The simple argument shows that the zeroth order model --- although it
307: gives some clue as to the origin of the slip face --- is insufficient
308: for a proper qualitative understanding of dunes. However, some
309: important lessons can be learned from it that will be helpful in our
310: further investigation of the problem. First, even with a very
311: simplistic model any reasonably heap--like initial condition will
312: quickly develop into a dune--like shape with a slip face. Secondly,
313: although the latter may seem to converge to a steady--state solution
314: for intermediate times, it finally turns out to be unstable and
315: flattens out. The discussion of the migration velocity in
316: Section~\ref{sec:aeolian} suggests that small deviations from
317: Eqs.(\ref{eq:tau0}) and (\ref{eq:q0}) at the brink can make an
318: important difference. Obviously some caution is needed in judging the
319: success of numerical models of dune formation. Unless stability has
320: explicitly been demonstrated, they may be suspected to fail in a
321: similar way as the zeroth order model when integrated over
322: sufficiently long times (which has actually not been checked for some
323: models that can be found in the literature) or to be sensitive to
324: numerical errors at the brink. Detailed numerical modeling should
325: therefore be preceded by a sound qualitative understanding of the
326: mechanisms underlying dune formation.  We will argue in
327: Sections~\ref{sec:wind}, \ref{sec:saltation} that to this end a subtle
328: balance between two small deviations from
329: Eqs.(\ref{eq:tau0}),(\ref{eq:q0}) and especially non--local
330: contributions in Eq.~(\ref{eq:task2}) have to be taken into account.
331: 
332: \subsection{Migration velocity}\label{sec:v}
333: Before entering a detailed discussion of the minimal model, it is
334: worth pausing for some general thoughts as to what can be said about
335: the shear stress and the speed--up of the wind above an obstacle,
336: without actually doing the (somewhat involved) calculation. 
337: 
338: A basic property of strongly developed turbulence is its dilation
339: invariance or scale free structure. Whereas general Navier--Stokes
340: flow is invariant under a scale transformation that keeps the Reynolds
341: number constant, strongly turbulent flow (for ``infinite'' Reynolds
342: number) allows for infinitely many such similarity
343: transformations. Landau and Lifshitz \cite{landau-lifshitz:fm} took
344: advantage of this fact for deriving the logarithmic velocity profile
345: mentioned above by an elegant scaling argument. The logarithmic
346: velocity profile suggests that the speed--up of the wind and therefore
347: also the shear stress perturbation above a heap of given shape should
348: itself be logarithmically dependent on its size. But how do they
349: depend on the shape of the obstacle?  Since the flow itself does not
350: provide any characteristic length scale, the dimensionless quantity
351: $\hat \tau$ defined in Eq.(\ref{eq:tauhat}) can only depend on a
352: dimensionless characterization of the profile $h(x)$. In other words,
353: to lowest order in the perturbation, it must be a linear functional of
354: the derivative $h'$ and can be written as
355: \begin{equation}\label{eq:scaling}
356: \hat \tau(\xi) = 
357: 	\varepsilon \,{\cal T}\left\{f'(\xi)\right\}\;, \qquad
358: 	\varepsilon \equiv H/\! L \;,
359: \end{equation}
360: with a dimensionless profile function
361: \begin{equation}\label{eq:profile}
362: f(\xi)\equiv h(x)/H \qquad \xi \equiv x/\! L \;. 
363: \end{equation}
364: and a scale--free (and necessarily non--local) linear functional
365: ${\cal T}$.  This reasoning can be repeated for the dimensionless
366: velocity and pressure perturbations.  Intuitively, the scaling $\hat
367: \tau \propto \varepsilon$ for a flat smooth obstacle ($\varepsilon \ll
368: 1$) can be understood from Fig.~\ref{fig:scaling}. When the air flows
369: over the obstacle, the velocity close to the obstacle is deflected by
370: an angle $\varepsilon$ whereas it remains constant far above the
371: obstacle. For incompressible flow, continuity translates this into a
372: speed--up of order $\varepsilon$ and (via Bernoulli's law) into a
373: corresponding pressure drop near the top of the heap. This in turn
374: causes a shear stress perturbation $\hat \tau$ of the same order.
375: 
376: \begin{figure}[t]
377: \psfrag{H}{$H$}
378: \psfrag{L}{$L$}
379: \psfrag{e}{$\varepsilon$}
380: \psfrag{u}{$u$}
381: \psfrag{ee}{$u+\varepsilon u $}
382: \begin{center}
383: \includegraphics[width=\columnwidth]{airflow.eps}
384: \end{center}
385: \caption{Schematic sketch of the deflection of the wind velocity $u$
386: above a flat heap of aspect ratio $\varepsilon \equiv H\!/\!
387: L\ll1$. The characteristic length scale $L$ is in this context
388: conventionally often identified with  the half length at half
389: height of the heap. The vertical deflection causes a speed--up above
390: the top of the heap. This is accompanied by a pressure perturbation
391: that is negative above the top of the heap and positive at its tails.
392: Due to turbulence, the flow pattern is asymmetric even above a
393: symmetric heap.}
394: \label{fig:scaling}
395: \end{figure}
396: 
397: These general considerations already allow us to predict the scaling
398: of the migration velocity $v$ with dune size if we assume that dunes
399: of different size have roughly similar shapes $f(\xi)$ and aspect ratios
400: $\varepsilon$, which is indeed suggested by the scale invariance of
401: the turbulent wind field and by observations.  Inserting
402: Eq.(\ref{eq:scaling}) into Eq.(\ref{eq:v_def}), and again
403: approximating Eq.(\ref{eq:task2}) by a local sand transport law $q
404: \equiv q(\tau)$, we find
405: \begin{equation}\label{eq:v}
406:  v  \frac{dq}{d\hat\tau}
407: 	\frac{d\hat\tau/dx}{\varrho_{s}\varepsilon f'}
408: 	=\frac{dq}{d\hat\tau}\frac{{\cal
409: 	T'}\{f'\}}{\varrho_{s} L f'} \propto \frac1L \;.
410: \end{equation}
411: The final proportionality strictly holds only if the steady--state
412: shape $f(\xi)$ and aspect ratio $\varepsilon$ are scale
413: invariant. However, it can be expected to be robust and rather
414: insensitive against violations of exact scale invariance.  First, the
415: normalized steady--state shapes $f(\xi)$ are strongly constrained by
416: the requirement that they render $v(x)\equiv v$, independent of $x$,
417: along the heap. Therefore, they should to a first approximation be
418: independent of size, which is indeed borne out by the minimal model
419: (Fig.~\ref{fig:shapes}) and empirical observations
420: \cite{sauermann-etal:2000}.  Moreover, the dependence $v\{ f'(\xi) \}$
421: is rather indirect and can therefore be expected to be weak. Secondly,
422: Eq.(\ref{eq:bagnold}) suggests that for gently sloped obstacles
423: ($\varepsilon \ll 1$) the dependence of $dq/d\tau$ on the aspect ratio
424: $\varepsilon$ also is not very pronounced. And finally, --- due to the
425: above mentioned scale invariance of turbulence --- a scale invariant
426: aspect ratio can reasonably be expected for large dunes. In fact, we
427: will show below that the minimal model predicts that the aspect ratio
428: of small heaps is not constant but rather decreases proportional to
429: their height. But this also implies that the latter becomes too small
430: to have a very significant effect on the above argument.  Note,
431: however, that only for strictly scale invariant dunes, Eq.(\ref{eq:v})
432: becomes identical to the often quoted observation that dunes migrate
433: with a speed inversely proportional to their height
434: \cite{speed}. Since the deviations of large dunes from scale
435: invariance are not very pronounced, the difference between these
436: predictions is not very strong except for small dunes and heaps.
437: Presently available field data are maybe not accurate enough to
438: clearly distinguish between $v\propto 1/L$ and $v\propto 1/H$, though
439: some data support $v\propto1/L$, most notably the comprehensive study
440: of barchan dunes in southern Peru by Finkel \cite{Finkel59}. As we
441: will show below, our numerical results for the minimal model clearly
442: favor $v\propto1/L$.
443: 
444: We also note that together with Eq.(\ref{eq:bagnold}), Eq.(\ref{eq:v})
445: moreover predicts that the migration velocity grows non--linearly with
446: (as the third power of) the wind velocity. A more accurate relation
447: can be obtained from the minimal model as described below, but the
448: qualitative conclusion is the same. Dunes can migrate farther in a
449: short period of exceptionally strong wind than during much longer
450: periods of gentle winds. Finally, we should mention that some caution
451: is needed when identifying the characteristic length scale $L$ in
452: Eq.(\ref{eq:v}). In our discussion, we have so far assumed that
453: $f(\xi)$ is a smooth function, which is not the case for dunes with
454: slip face. Below we will argue that in this case $f$ should be
455: identified with the envelope of the dune and its separation bubble and
456: $L$ with the characteristic length scale of this envelope. For a
457: barchan dune the latter practically coincides with the total length of
458: the dune from its windward end to the tips of its horns (cf.\
459: Fig.~\ref{fig:barchan}).
460: 
461: In contrast to the overall migration velocity of a translation
462: invariant dune, the position dependent migration velocity $v(x)$ that
463: determines that shape is much harder to obtain since it requires a
464: precise knowledge of the non--local functional ${\cal T}$ in
465: Eq.(\ref{eq:scaling}). This will be provided in the following section.
466: 
467: \section{Wind shear stress}\label{sec:wind}
468: \subsection{Surface shear stress on a smooth heap}\label{sec:theory}
469: The discussion in the preceding paragraph showed that --- in contrast
470: to the assumption in Eq.(\ref{eq:tau0}) --- the dependence of the
471: shear stress on the height profile is non--local. Although it will
472: turn out that this shortcoming of Eq.(\ref{eq:tau0}) is not
473: responsible for the failure of the zeroth order model, it should by
474: now have become apparent that further progress can hardly be achieved
475: without a rather detailed understanding of the turbulent wind field
476: above heaps and dunes. For dunes with a slip face that typically has a
477: slope of about $32^{\circ}-35^{\circ}$ and terminates in a sharp
478: brink, the situation is similar to the textbook example of a backward
479: facing step, which has the reputation of a test case for numerical
480: turbulent models. Even if a commercial turbulent solver is used, the
481: accurate calculation of the shear stress e.g.\ on a barchan dune is a
482: non--trivial task and quite demanding in computer time and memory, and
483: the most interesting long--time dynamics of dunes is therefore
484: difficult to access. For this reason, we want to focus on flat smooth
485: heaps, first. In this case, one can apply an analytical perturbation
486: theory for turbulent boundary layer flow over smooth hills that has
487: been developed over the last decades \cite{jackson-hunt:75,sykes:80,%
488: zeman-jensen:88,hunt-leibovich-richards:88,Carruthers90,weng-etal:91}.
489: Though the calculation is essentially a formalization of the intuitive
490: description accompanying Fig.~\ref{fig:scaling}, it requires a highly
491: non--trivial boundary layer construction that we will not recapitulate
492: here. The interested reader is referred to the original literature. We
493: merely quote the final result for the $x-$component (along the main
494: wind direction) $\hat \tau_x$ of the surface shear stress perturbation
495: above a profile $h(x,y)$
496: \cite{hunt-leibovich-richards:88,weng-etal:91},
497: \begin{equation}
498: \label{eq:hunt} 
499: {\cal F}_{xy}\{\hat\tau_x\} = \frac{A k_x( k_x +
500: i B |k_x|)}{(k_x^2 + k_y^2)^{1/2}} \, {\cal F}_{xy} \{ h(x,y) \}
501: \;.  
502: \end{equation} 
503: We have abbreviated the Fourier transformation from the space
504: variables $x$, $y$ to the respective wave numbers $k_x$, $k_y$ by
505: ${\cal F}_{xy}$. For simplicity the logarithmic $k-$dependence of the
506: parameters $A$ and $B$ was neglected.  The latter are then given
507: by
508: \begin{equation}\label{eq:a_b}
509: \begin{split}
510:  A &= 
511: \frac{\ln \left(\Phi^2/\ln \Phi\right)^2}{2(\ln \phi)^3}
512:  [1+\ln \phi+2\ln(\pi/2)+4\gamma] \\
513:  B &= \pi/[1+\ln\phi+2\ln(\pi/2)+4\gamma_E] \\
514:   \phi & \equiv 2\kappa^2\Phi/\ln\phi 
515: \end{split} 
516: \end{equation}
517: and depend logarithmically on the ratio $\Phi\equiv L/z_0$, where $L$
518: is the characteristic length of $h(x,0)$ (for this purpose
519: conventionally often identified with half the length at half height or
520: about one fourth of the characteristic wavelength) and $z_0$ is a
521: measure of the surface roughness (typically an effective length
522: somewhat below the linear dimension of the latter). We also have
523: introduced the von K\'arm\'an constant $\kappa\approx 0.4$ and Euler's
524: constant $\gamma_E\approx 0.577$.  A practical approximation for
525: $\phi$ is obtained by iterating (twice) the implicit equation for
526: $\phi$ and closing it by dropping the remaining $\ln\phi$.  The
527: dependence of $A$ and $B$ on $L$ is depicted in
528: Fig.~\ref{fig:ab}. Obviously, as long as $L/z_0$ does not change by
529: orders of magnitude (e.g.\ due to vegetation), this extremely weak
530: scale dependence is negligible for our purposes, and $A$ and $B$ can
531: be regarded as constant theoretical or phenomenological parameters.
532: For definiteness we often work with the values $A=4$ and $B=0.25$
533: (approximately obtained for $L/z_0 = 10^5 \dots 10^6$). Although these
534: values may differ somewhat depending on the particular application in
535: mind, or on the presence or absence of ripples, and may
536: phenomenologically be somewhat different from the theoretical
537: prediction, this does not affect our general conclusions.
538: 
539: \begin{figure}[tb]
540:    \psfrag{L}{$10^{-5} L/z_0$}
541:    \psfrag{A}{$A$, $B$} 
542: \begin{center}  
543:    \includegraphics[width=\columnwidth]{a_b_log.eps} 
544: \caption{The
545:    theoretical prediction for the dependence of the parameters $A$
546:    (solid line) and $B$ (dashed line) of Eq.(\ref{eq:a_b}) on the
547:    ratio of the dune size $L$ to the roughness length $z_0$.}  
548: \label{fig:ab}
549: \end{center}
550: \end{figure} 
551: 
552: For the following discussion we want to specialize Eq.(\ref{eq:hunt})
553: to the central slice of a transverse or barchan dune along the wind
554: direction. To this end we evaluate Eq.(\ref{eq:hunt}) for the central
555: slice $h(x)$ of a heap that has a Gaussian shape with standard
556: deviation $\sigma$ in the transverse direction parameterized by $y$,
557: \begin{equation}\label{eq:gauss}
558: h(x,y)=h(x)e^{-y^2/2\sigma^2} \;.
559: \end{equation}
560: This approximation is technically useful, and although it may seem
561: relatively crude for a particular real dune, it typically does not
562: introduce any noteworthy derogation of the results compared to a more
563: accurate description. The Fourier coefficients of the shear stress
564: $\hat\tau(x)\equiv \hat\tau_x(x,0)$ on the central slice along the
565: wind direction become
566: %\begin{equation}
567: \begin{multline}
568: {\cal F}\{\hat\tau(x)\}= \frac{\sigma A}{\sqrt{2\pi}} k(k +
569: iB |k|) \times
570: \\
571:  e^{-\frac14k^2\sigma^2}
572: K_0\!\left(\frac{k^2\sigma^2}4\right){\cal F}\{h(x)\}
573: \end{multline}
574: %\end{equation}
575: Here, $K_0$ denotes a modified Bessel function and the Fourier
576: transforms are one--dimensional, so that we can drop the redundant
577: $x-$subscripts.  
578: 
579: For transverse dunes ($\sigma/L\to \infty$), we obtain the two
580: equivalent expressions
581: \begin{subequations}\label{eq:wind}
582: \begin{align}
583: {\cal F}\{\hat\tau^\infty (x)\} &= A(|k|+iB\, k){\cal F}\{h(x)\} \;. 
584: \label{eq:fourier_wind} \\
585:  \hat\tau^\infty(x) &= A \left[ h'(x)\otimes (\pi x)^{-1} + B h'(x)\right]
586:  \;. \label{eq:real_wind}
587: \end{align}
588: \end{subequations}
589: For the real space version we have abbreviated a convolution integral
590: according to
591: \begin{equation}
592: \label{eq:convolution}
593: f\otimes g \equiv \int_{-\infty}^{\infty}\!\!\!\!  d\xi \; f(\xi)\,g(x-\xi) \;.
594: \end{equation} 
595: Evaluation for arbitrary $\sigma$ gives two correction terms
596: \begin{equation}
597: \label{eq:correction}
598: \hat\tau^\sigma=\hat\tau^\infty - A (h\otimes \Delta_1 + B
599: h'\otimes \Delta_2) \;,
600: \end{equation}
601: with 
602: \begin{align}
603: \label{eq:deltas}
604: \Delta_1 &=
605: \frac{U\left(\frac32,1,\frac{x^2}{2\sigma^2}\right)}{4\sqrt \pi
606: \sigma^2}+\frac{\frac{\sqrt{\pi}}2
607: U\left(\frac12,0,\frac{x^2}{2\sigma^2}\right)-1}{\pi x^2} \\ 
608: \Delta_2 &=\frac1{\sigma \pi}\int_{0}^{\infty}
609: \!\!\!\!d\xi\,\cos\left(\frac{\xi x}\sigma\right) \left[1-\frac{\xi
610: e^{\xi^2/4}}{\sqrt{2\pi}} K_0\left(\frac{\xi^2}4\right)\right]
611: \end{align}
612: two even functions depicted in Fig.~\ref{fig:corrections} that are
613: flat for $\sigma/L\to\infty$ and become peaked for $\sigma\simeq
614: L$. (The confluent hypergeometric $U$ functions
615: \cite{abramowitz-stegun} have been introduced to rephrase the
616: sine--part of the Fourier integrals.)
617: 
618: 
619: \begin{figure}[t]
620: \psfrag{x}{$x/L$}
621: \psfrag{De}{$\Delta_1$} 
622: \psfrag{Ds}{$\Delta_2$} 
623: \begin{center}
624:  
625: \includegraphics[width=\columnwidth]{d2.eps} 
626: \includegraphics[width=\columnwidth]{d1.eps}
627: \caption{The peaked functions $\Delta_1$ and $\Delta_2$ of
628: Eq.(\ref{eq:deltas}) for $\sigma=1$, 2, 5. The area under the peaks
629: remains constant (0 and $1/2\pi$), while the peak heights decrease
630: proportional to $\sigma^{-2}$ and $\sigma^{-1}$, respectively.}
631: \label{fig:corrections} 
632: \end{center} 
633: \end{figure} 
634: 
635: Since the correction terms in Eq.(\ref{eq:correction}) are numerically
636: small, we may --- given a reasonable localized heap shape in the wind
637: direction --- approximately replace both functions $\Delta_1$ and
638: $\Delta_2$ by delta functions, thus arriving at
639: \begin{equation}\label{eq:approxcorrect}
640: \hat\tau_\sigma\approx \hat\tau_\infty - A [c_1(\sigma) h + B
641: c_2(\sigma) h'] \;.
642: \end{equation}
643: In this approximation they are seen to give merely a
644: $\sigma-$dependent renormalization of the asymmetry parameter $B\to
645: B(\sigma) \lesssim B(\infty)= B$ and to add a (trivial) term
646: $c_1(\sigma) h(x)$ within the brackets of
647: Eq.(\ref{eq:real_wind}). Numerically, one can estimate $L\,c_1(
648: L/\sqrt2)$ and $c_2( L/\sqrt2)$ to be about 0.2 (cf.\
649: Fig.~\ref{fig:hunt_jackson}). The exact $\sigma-$dependence of the
650: coefficients is determined by the shape and extension of the heap in
651: the $x-$direction, because the area under the peaks $\Delta_1$ and
652: $\Delta_2$ is constant and independent of $\sigma$, while the peak
653: height decreases proportional to $\sigma^{-2}$ and $\sigma^{-1}$,
654: respectively. Since both corrections vanish for $\sigma/L\to\infty$
655: and do not contribute any substantial new effects to
656: Eq.(\ref{eq:wind}), they may for simplicity be omitted altogether in
657: the following discussion that mainly aims at a qualitative
658: understanding.  Fig.~\ref{fig:hunt_jackson} moreover shows that they
659: can approximately be mimicked by a renormalized parameter $A$ in
660: Eq.(\ref{eq:wind}) for the central slice of a symmetric heap. This
661: leads to the important conclusion that the wind shear stress on the
662: central slice of a $3d$ symmetric heap and on a heap with a profile
663: that is constant in the transverse direction, are qualitatively the
664: same and quantitatively similar, which was not \emph{a priori}
665: obvious.  Together with the fact that on a gently sloped obstacle the
666: transverse components of the shear stress are small compared to its
667: longitudinal components, this suggests that the predictions of
668: Eq.(\ref{eq:wind}) apply in a first approximation to any slice of a
669: dune parallel to the wind direction. In this sense, the study of
670: Eq.(\ref{eq:wind}) is representative.  Summarizing the foregoing
671: discussion we can say that to gain a qualitative understanding of dune
672: formation by aeolian sand transport one may focus on
673: Eq.(\ref{eq:wind}) as a model for the wind shear stress. We therefore
674: analyze this equation in some detail in the next paragraph.
675: 
676: \begin{figure}[t]
677: \psfrag{x}{$x/L$}
678: \psfrag{ta}{$\hat \tau(x)$} 
679: \begin{center} 
680: \includegraphics[width=\columnwidth]{hunt_jackson.eps} 
681: \caption{Shear stress perturbation above the central slice of a $3d$
682: symmetric ($\sigma=L/\sqrt 2$) Gaussian heap. The plot compares two
683: approximations to Eqs.(\ref{eq:hunt}), (\ref{eq:correction}) (points):
684: (i) Eq.(\ref{eq:approxcorrect}) with $Lc_1\approx c_2\approx0.2$
685: (solid line), and (ii) Eq.(\ref{eq:wind}) with $A$ renormalized by a
686: factor 0.8 (dashed line). While (i) is practically indistinguishable
687: of Eq.(\ref{eq:hunt}) on the present level of accuracy, the simpler
688: approximation (ii) already captures most of the $3d$ effects.}
689: \label{fig:hunt_jackson} 
690: \end{center} 
691: \end{figure} 
692: 
693: \subsection{Properties and consequences of Eq.(\ref{eq:wind})}\label{sec:properties}
694: 
695: A scaling analysis of Eq.(\ref{eq:wind}) immediately reveals
696: that $\hat \tau$ is indeed of the general form anticipated on general
697: grounds in Eq.(\ref{eq:scaling}). The amplification of the shear
698: stress at the top of a smooth profile is thus determined by its aspect
699: ratio $\varepsilon=H\!/\!L$ and is essentially independent of the
700: absolute height $H$. It only has a very weak logarithmic dependence on
701: the absolute size of the dune through the prefactors $A$ and $B$ given
702: in Eq.(\ref{eq:a_b}). Moreover, for a symmetric profile
703: $f(-\xi)=f(\xi)$, $\hat \tau$ is the sum of a symmetric part and an
704: anti--symmetric part, i.e., the flow over the heap has a symmetry
705: breaking component that is a consequence of turbulence.  The origin of
706: the symmetric and antisymmetric parts of $\hat \tau$ can intuitively
707: be understood as follows. As we have pointed out in
708: Section~\ref{sec:v} (see Fig.~\ref{fig:scaling}), the streamlines have
709: to be compressed above the heap if the perturbation is not to be
710: transmitted to infinite height, and as a consequence, there is a
711: corresponding increase in the shear stress. For the laminar average
712: flow, this speed--up and the associated decrease in atmospheric
713: pressure above the heap are symmetric for a symmetric heap as is the
714: corresponding shear stress perturbation, which accounts for the
715: dominant symmetric part of $\hat \tau$. On the other hand, the inertia
716: of the turbulent velocity fluctuations around this laminar main flow
717: contributes an asymmetric resistance to deflections of the flow. It
718: counteracts the upturn of the streamlines on the windward side and the
719: downturn on the lee side. Formally, this effect enters the
720: perturbative calculation of $\hat \tau$ through the Reynolds stress.
721: 
722: Further insight can be gained from special analytical
723: solutions to Eq.(\ref{eq:wind}).  For the normalized heap
724: profiles
725: \begin{equation}
726:    \label{eq:hills}
727: \begin{split}
728:    f_L(\xi) & = \frac1{1+\xi^2}\\
729:    f_G(\xi) & = \exp\left(-\xi^2\right)\\ 
730:    f^n_C(\xi) & = S(\xi) \cos^n \xi  
731: \end{split}
732: \end{equation}
733: with
734: \begin{equation}
735:     S(\xi)\equiv  
736:    \begin{cases}
737:      1 & |\xi| \leq \pi/2 \\
738:      0 & |\xi| \geq \pi/2   \;, 
739:    \end{cases} 
740: \end{equation}
741:  we obtain
742: \begin{equation}
743: \label{eq:solutions}
744: \begin{split}
745: {\hat\tau_L}  
746: &= A (1 - 2B \xi -\xi^2)f_L^2(\xi) \\ 
747: {\hat\tau_G}
748: &=2A \left[\pi^{-1/2} -  \xi(B + \text{erfi}\, \xi) 
749: f_G(\xi)\right] \\
750: {\hat\tau_C}
751:  &=\frac{A}{\pi} \cos(2\xi) \left[\,\text{Si}(\pi+2\xi) +
752: \text{Si}(\pi-2\xi)\right] \\ & \quad
753: -\frac{A}{2\pi}
754: \sin(2\xi)\left[\,\text{Ci}(\pi+2\xi)+\text{Ci}(-\pi-2\xi) \right.  \\ 
755: & \quad \phantom{-A\sin(2x)[}\left.
756: -\text{Ci}(\pi-2x) - \text{Ci}(-\pi+2x) \right] \\
757: & \quad  -2B \, S(\xi)\,\cos \xi\,\sin \xi \qquad (n=2)
758: \end{split}
759: \end{equation}
760: with erfi the imaginary error function and Si and Ci the sine and
761: cosine integral functions, respectively \cite{abramowitz-stegun}. The
762: result given for $\hat \tau_C$ is for the special case $n=2$.  Both
763: the profiles of Eq.(\ref{eq:hills}) and the corresponding solutions of
764: Eq.(\ref{eq:wind}) given in Eq.(\ref{eq:solutions}) are shown in
765: Fig.~\ref{fig:hills}.
766: 
767: \begin{figure}[t]
768: \psfrag{x}{$x$}
769: \psfrag{h}{\mbox{}\;\; $f$} 
770: \psfrag{t}{$\!\!\!\!\tau/\tau_0$} 
771: \begin{center}
772:  
773: \includegraphics[width=\columnwidth]{hills2.eps} 
774: \caption{\emph{Lower curves:} The normalized profiles of
775: Eq.(\ref{eq:hills}), the Lorentzian $f_L$ (dashed), Gaussian $f_G$
776: (solid), and cosine $f^2_C$ (dotted). \emph{Upper curves:} The
777: corresponding surface shear stresses $\tau(x)/\tau_0$ from
778: Eq.(\ref{eq:solutions}) with $A \varepsilon = 0.8$, $B=0.25$.}
779: \label{fig:hills} 
780: \end{center} 
781: \end{figure} 
782: 
783: \begin{figure}[tb]
784:     %\psfrag{T}[l]{$\hat \tau_G$, $f_G$} 
785:     \psfrag{X}{$\xi$}
786:     \psfrag{f}{$f_G$} 
787:     \psfrag{t1}{$\hat \tau_{\rm sym}$}
788:     \psfrag{t2}{$\hat \tau_{\rm asy}$} 
789:     \psfrag{t3}{$\hat \tau_G$}
790: \begin{center}  
791:     \includegraphics[width=\columnwidth]{sym_asym.eps}
792:     \caption{The symmetric and asymmetric parts $\hat \tau_{\rm sym}$
793:     and $\hat \tau_{\rm asy}$ of the shear stress perturbation $\hat
794:     \tau_G$ of Eq.(\ref{eq:solutions}) for the Gaussian profile $f_G$
795:     of Eq.(\ref{eq:hills}) with $A\varepsilon=0.8$ and $B=0.25$. Note
796:     the small windward shift of the maximum of the shear stress with
797:     respect to the crest of the heap caused by the asymmetric
798:     contribution proportional to $B$.}
799: \label{fig:sym_asym}
800: \end{center}
801: \end{figure}
802: 
803: The plots of $\tau$ show that as a rule of thumb one can estimate the
804: relative magnitude of the shear stress perturbation at the top of the
805: heap by $A\varepsilon$. The plots share several crucial properties not
806: present in the affine approximation $\hat\tau\propto h$ of the zeroth
807: order model. At the tails of the profiles in Fig.~\ref{fig:hills}, the
808: shear stress decreases below its asymptotic value $\tau_0$ on the
809: plane.  This effect is particularly pronounced for the profile $f^n_C$
810: that has a discontinuity in its second derivative.  Further, as a
811: consequence of the second term in Eq.(\ref{eq:wind}), the surface
812: shear stress is not symmetric even for symmetric profiles like those
813: in Eq.(\ref{eq:hills}). Fig.~\ref{fig:sym_asym} displays the symmetric
814: and asymmetric parts $\hat \tau_{\rm sym}$ and $\hat \tau_{\rm asy}$
815: of the shear stress perturbation $\hat \tau_G$ for the Gaussian
816: profile $f_G$, separately. The asymmetric contribution to $\hat \tau$
817: is small compared to the symmetric one. For the profile $f_L$ the
818: corresponding shift $\delta x_\tau$ of the location of the maximum of
819: the shear stress with respect to that of the maximum of $f_L(x)$ can
820: be calculated analytically,
821: \begin{equation}\label{eq:shift}
822: \delta x_\tau/\! L=2\,(1 + B^2)^{1/2}\,\sin [\arctan (B)/3] -B
823: \;.
824: \end{equation}
825: It is indeed found to be very small, because $B$ is small and thus
826: $\delta x_\tau/\!  L\approx -B/3$ typically amounts to a length of
827: about a few percent of the total heap length. Nevertheless, it is a
828: crucial element in the modeling of aeolian sand transport, as will now
829: be demonstrated.
830: 
831: \begin{figure}[t]
832:   \psfrag{v}[lr]{$v(x)$} \psfrag{x}{$x/L$}
833: \begin{center}   
834:   \includegraphics[width=\columnwidth]{v_s2.eps}
835:   \includegraphics[width=\columnwidth]{v_s3.eps} 
836: \caption{The position
837:   dependent surface migration velocity $v(x)$ in arbitrary units
838:   according to Eqs.(\ref{eq:v_def}), (\ref{eq:wind}) with $A=4$,
839:   $B=0.25$. \emph{Upper part:} For the profile $f^2_C$ (gray) and
840:   varying aspect ratios $\varepsilon = 0.01$ (dashed), $0.1$ (solid),
841:   $0.19$ (dotted) and the local flux relation Eq.(\ref{eq:q0}).
842:   \emph{Lower part:} For the profile $f^{1.65}_C$ (gray) with
843:   $\varepsilon=0.1$ (solid), $0.05$ (dashed) and the non--local flux
844:   relation Eq.(\ref{eq:flux}). For simplicity, $q_s$ was represented
845:   by Eq.(\ref{eq:bagnold}) and $l_s\approx 0.1$ was taken constant.}
846: \label{fig:v}
847: \end{center}
848: \end{figure} \noindent
849: 
850: For a qualitative estimate of the effects of Eq.(\ref{eq:wind}) onto
851: the sand transport over a dune, it is useful to consider once again
852: the local zeroth order model for aeolian sand transport,
853: Eq.(\ref{eq:q0}), i.e.\ a completely saturated flux $q=q_s(\tau)$ with
854: $q_s$ given by Eq.(\ref{eq:bagnold}). (Below, we will show that this
855: is asymptotically valid on large dunes in strong winds.) The distinct
856: features of Eq.(\ref{eq:wind}) that are missing in the zeroth order
857: approximation for the shear stress, Eq.(\ref{eq:tau0}), are then
858: easily seen to have potentially profound effects on the shape
859: evolution.  First, due to the depression of the shear stress at the
860: tails of the profiles, deposition rather than erosion may occur at the
861: windward foot. Secondly, due to the asymmetric contribution in
862: $\tau(x)$ there can be a net deposition on a symmetric heap of sand.
863: In particular, the shift of the position of the maximum shear stress
864: with respect to the top of the heap allows deposition at the top of
865: the heap. For an initially flat heap of sand there is thus the
866: possibility of a steepening of the windward slope and mass
867: growth. This implies that a plane sand surface is unstable against
868: modulations. To illustrate this effect, we used
869: Eq.(\ref{eq:solutions}) to calculate the migration velocity $v(x)$ of
870: a cosine--shaped heap of sand $f^2_C(x)$ according to
871: Eqs.(\ref{eq:v_def}), (\ref{eq:bagnold}), (\ref{eq:q0}), and
872: (\ref{eq:wind}). The latter is shown in the upper part of
873: Fig.~\ref{fig:v}.  The decrease of $v(x)$ on the lee side reveals the
874: anticipated self--amplifying tendency of the unstable lee slope to
875: steepen, since $v'<0$ implies that $v$ increases with height. This
876: gives rise to the formation of the slip face.  More interestingly,
877: Eq.(\ref{eq:wind}) renders $v(x)$ approximately constant over almost
878: the whole windward side if $\varepsilon=H\!/\!L$ is close to a certain
879: value determined by the values of the coefficients $A$ and $B$ in
880: Eq.(\ref{eq:wind}).  Slightly better constancy can be achieved for
881: slightly lower $n$ (with slightly larger $\varepsilon$) but \emph{not}
882: for the profiles $f_G$ and $f_L$, for which $v(x)$ is always
883: non--uniform.  Let us finally consider the dashed and dotted lines in
884: the upper part of Fig.~\ref{fig:v}. They were obtained for a smaller
885: and a larger aspect ratio and represent a steepening ($v'<0$) and
886: flattening ($v'>0$) of the windward side, respectively. In other
887: words, profiles with larger/smaller windward slopes are driven towards
888: the solution with constant windward $v(x)$ and a stable optimum
889: windward slope different from zero.  Altogether, Fig.~\ref{fig:v} thus
890: suggests that the coupled Eqs.(\ref{eq:bagnold}) and (\ref{eq:wind})
891: drive a heap of sand towards a ``dune'' with a cosine--like windward
892: profile of a preferred aspect ratio, and a slip face on the lee side.
893: 
894: 
895: 
896: From the qualitative analysis presented so far, it is not yet obvious
897: that this process converges to a translation invariant steady--state
898: solution.  Several previous studies using similar descriptions either
899: did not scrutinize the long time behavior of their models
900: \cite{Wippermann86,van_dijk-arens-boxel:99}, or failed to obtain
901: stable dunes \cite{zeman-jensen:88,Stam97}.  To obtain a consistent
902: general model for dune formation under general influx and wind
903: conditions, the present wind model Eqs.(\ref{eq:hunt}),
904: (\ref{eq:wind}) has to be appropriately adapted to situations
905: with flow separation above slip faces. And, most importantly, the
906: saturated--flux approximation Eq.(\ref{eq:q0}) of the ``zeroth order''
907: model has to be abandoned. These steps will be discussed in the
908: following subsection and in Section \ref{sec:saltation}, where we will
909: also explain the lower part of Fig~\ref{fig:v}. This will complete the
910: definition of the minimal model.  Its numerical solutions will be
911: presented in Section \ref{sec:solution}.
912: 
913: 
914: \subsection{Flow separation}\label{sec:separation}
915: The wind model as discussed so far works fine for smooth heaps with
916: gentle slopes. However, as we have already mentioned, its application
917: to dune profiles with slip faces and sharp brink lines is not
918: straightforward. The perturbative turbulent boundary layer approach
919: leading to Eq.(\ref{eq:hunt}) does not account for flow separation, a
920: phenomenon that occurs at sharp edges and steep slopes to prevent an
921: extreme bending of the streamlines \cite{landau-lifshitz:fm}. (For
922: some of the technical terms involved in this section, the reader is
923: referred to Fig.~\ref{fig:separation}.)  Instead of bending the
924: streamlines around sharp edges, re--circulating eddies separate from
925: the (on average) laminar main flow, thereby creating an effective
926: envelope that diverts the main flow on a smooth detour around the
927: obstacle. See Figs.~\ref{fig:separation} and \ref{fig:fluent} for a
928: schematic sketch and a numerical calculation of a typical velocity
929: field, respectively. Fortunately, it turns out that dune formation and
930: migration do not in general depend very sensitively on the details of
931: this complicated process. Or in other words, there is a large number
932: of interesting problems of aeolian sand transport for which these
933: details are largely irrelevant, and for which their somewhat realistic
934: physical representation would create a huge overhead in complexity
935: (especially in $3d$) to an otherwise tractable problem. It was
936: therefore suggested earlier \cite{zeman-jensen:88} that for the
937: purpose of calculating the shear stress on the windward side of a
938: dune, one may to a good approximation represent flow separation on the
939: lee side by the following heuristic method. A wind model such as
940: Eq.(\ref{eq:wind}) restricted to smooth, gently sloped objects is
941: applied to the envelope
942: \begin{equation}\label{eq:envelope}
943: \tilde h(x)=\text{max} \{h(x),s(x)\}
944: \end{equation}
945: of a dune $h(x)$ and a phenomenologically defined separation bubble
946: $s(x)$. This disregards the fact that the separating streamline does
947: not represent a solid boundary of the same roughness as the original
948: object, but the corresponding errors are expected to be small.
949: Typically, one wants $s(x)$ to be a mathematically simple smooth
950: continuation of the dune profile. It is, however, crucial that the
951: latter respects some major phenomenological properties of flow
952: separation \cite{badbubble}.  Although this is by no means a rigorous
953: procedure, one can test its predictions for selected cases against
954: numerical solutions of various turbulence models. The hope is that via
955: this approach, one can eventually get a qualitative understanding of
956: the mechanisms and phenomena involved in dune formation and migration,
957: leaving certain quantitative aspects to a more elaborate (and much
958: more laborious) future analysis.
959: 
960: \begin{figure}[t]
961:   \psfrag{s}[b]{streamlines} 
962:   \psfrag{sb}[cb]{separation bubble}
963:   \psfrag{ss}[cb]{separating streamline}
964:   \psfrag{d}{dune}
965: \begin{center}
966:   \includegraphics[width=\columnwidth]{separation.eps}
967:   \caption{Sketch of the central slice of a barchan dune and the
968:   separation bubble. The shear stress on the windward side of the dune
969:   is calculated by applying Eq.(\ref{eq:wind}) to the
970:   phenomenologically defined envelope of the dune and the separation
971:   zone.}
972: \label{fig:separation}
973: \end{center}
974: \end{figure}
975: 
976: In the spirit of the minimal model we want to parameterize the
977: separating streamline $s(x)$ in the simplest form that obeys
978: physically motivated boundary conditions at its detachment and
979: reattachment points $x_d$ and $x_r$. At detachment, the slope of the
980: separating streamline must match the slope of the dune. Moreover, also
981: the curvature must be continuous there, since discontinuities in
982: curvature are detected by Eq.(\ref{eq:wind}) and cause kinks in $\tau$
983: and discontinuous steps in the erosion/deposition as is e.g.\ the case
984: for the profile $f^2_C$. For the reattachment point, there are no
985: comparable restrictions to the slope and curvature, since the
986: separation bubble is not sharply defined there, and the model aims at
987: a realistic description of the conditions in the wake region only
988: insofar as they affect the shear stress on the \emph{windward}
989: side. On the lee side, inside the separation bubble, the shear stress
990: can simply be set to zero \cite{rainout}, since it is typically below
991: the threshold for aeolian sand transport.  Therefore, the choice of
992: the reattachment matching condition is a matter of convenience rather
993: than physical significance in the present model. However, we want
994: $s(x)$ to reproduce some common phenomenological knowledge about flow
995: separation. First, from many numerical calculations it is known that,
996: at high Reynolds numbers, the turbulent boundary layer reattaches at a
997: distance of about $6H$ after a backward--facing step of height
998: $H$. Secondly, it has often been observed experimentally that in
999: strongly turbulent flows over hills and symmetric triangular
1000: obstacles, flow separation sets in if the backward slope exceeds an
1001: angle of about $14^\circ$. Although, in both cases the exact numerical
1002: values depend on various factors such as the surface roughness and the
1003: Reynolds number, they shall be treated as fixed phenomenological
1004: constants at the present stage. A model that fulfills all the above
1005: requirements is a third order polynomial with continuous slopes at the
1006: boundaries and a maximum negative slope of $\tan 14^\circ$. The
1007: boundary conditions
1008: \begin{equation}\label{eq:bc}
1009: \begin{split}
1010: s(x_d) & = h_d \equiv h(x_d) \qquad \; s(x_r) = 0 \\
1011: s'(x_d) & = h_d' \equiv h'(x_d) \qquad s'(x_r) = 0 \;. \\
1012: \bar s' & \equiv \text{max}\{-s'(x)\} = \tan 14^\circ = 0.25
1013: \end{split}
1014: \end{equation}
1015: constrain the third order bubble parameterization to be of the form
1016: \begin{equation}\label{eq:sepbubble}
1017:   s(z)= (2 h_d + h_d' L_b)z^3 -(3 h_d + 2 h_d' L_b) z^2 + h'_d L_bz +
1018:   h_d
1019: \end{equation}
1020: with $z\equiv (x-x_d)/L_b \in [0,1]$.  With the further abbreviation
1021: $\nu\equiv h_d'/\bar s'$, we can express the length $L_b\equiv
1022: x_r-x_d $ of the bubble as
1023: \begin{equation}\label{eq:sx}
1024: L_b =\frac{3h_d}{h_d'} \frac{1 - \nu - \sqrt{1 +  \nu}}{3- \nu} \approx
1025:   \frac{3h_d}{2\bar s'}\left(1+ \frac  \nu4 + \frac{ \nu^2}8\right) \;,
1026: \end{equation}
1027: where the final approximation for small $h_d'$ is sufficient for our
1028: purpose (and numerically better behaved as the exact expression). A
1029: subtlety of such a separation bubble parameterization is the fact that
1030: the slope at $x_d$ determines the length of the bubble, which in turn,
1031: via Eq.(\ref{eq:wind}) influences the curvature at $x_d$. In other
1032: words, the presence of the bubble introduces a non--local feed--back
1033: between the slope and the curvature at the brink, which we believe is
1034: physically reasonable. In Fig.~\ref{fig:bubbles} we give some examples
1035: of separation bubbles for different boundary slopes $h_d'$, while
1036: Fig.~\ref{fig:tau_bubble} illustrates the application of the above
1037: discussion for the calculation of the shear stress.  It shows an
1038: example for a dune profile $h(x)$ with slip face and the separation
1039: bubble $s(x)$, together with the shear stress $\tau(x)$ resulting from
1040: Eq.(\ref{eq:wind}) if $h$ is replaced by the envelope $\tilde h$.
1041: 
1042: \begin{figure}
1043:    % \psfrag{L in h0}[t]{$(x-x_d)/h_d$}
1044:    % \psfrag{s in h0}{$s(x)/h_d$}
1045:      \psfrag{X}[t]{$(x-x_d)/h_d$}
1046:      \psfrag{T}{$\frac{\displaystyle s(x)}{\displaystyle h_d}$}
1047:  \begin{center}
1048:     \includegraphics[width=\columnwidth]{bubbles.eps} 
1049:     \caption{Separation bubbles with a maximum negative slope of 0.25 
1050:     according to Eq.(\ref{eq:sx}) for varying initial slopes $-0.25 \leq 
1051:     h_d' \leq 0.25$. (The aspect ratio of the plot was stretched for 
1052:     presentation.)}
1053:       \label{fig:bubbles} 
1054: \end{center}
1055: \end{figure}
1056: 
1057: 
1058: \begin{figure}
1059:     \psfrag{x in L}[t]{$x / L$}
1060:     \psfrag{h in h0, tau in tau0}{$h / H$, $\tau / \tau_0$}
1061:     \psfrag{hxxxxxx}[lc]{\footnotesize $h$}
1062:     \psfrag{sxxxxxx}[lc]{\footnotesize $s$}
1063:     \psfrag{tauxxxx}[lc]{\footnotesize $\tau$}
1064:  \begin{center}   
1065:     \includegraphics[width=\columnwidth]{sepbub_tau.eps}
1066:     \caption{The windward profile $h(x)$ of a dune with slip face and
1067: the separation bubble $s(x)$ form together a smooth effective
1068: obstacle, defined by the envelope $\tilde h (x)$. To calculate the shear stress
1069: $\tau(x)$ on the windward side of the dune, $\tilde h$ is substituted
1070: for $h$ in Eq.(\ref{eq:wind}). In the region of re--circulation
1071: the surface shear stress $\tau$ is set to zero \cite{rainout}. 
1072: Without the separation bubble, $\tau(x)$ would develop a sharp
1073: singularity at the brink.}
1074: \label{fig:tau_bubble}
1075: \end{center}
1076: \end{figure}
1077: 
1078: We have performed several series of numerical fluid dynamics
1079: calculations in $2d$ and $3d$ with the commercial fluid dynamics
1080: solver Fluent~5 \cite{Fluent5} using the $k\epsilon$ and
1081: \emph{large--eddy} turbulent closure models to confirm the general
1082: picture outlined above and our particular implementation of the
1083: separation bubble. The differences between numerical and theoretical
1084: predictions for the shear stress on the windward side of various
1085: dune-- and heap--like objects in $2d$ and $3d$ were quantitatively
1086: small and not more significant than other neglected terms. Moreover, a
1087: comparison of predictions obtained from Eq.(\ref{eq:sepbubble}) with
1088: wind measurements on a barchan dune in Brazil \cite{sauermann:phd}
1089: showed good agreement. Therefore, we are confident that the proposed
1090: mathematical description of the wind shear stress captures the
1091: relevant aspects in the spirit of the minimal model. As an example for
1092: the numerical fluid dynamics calculation, we show in
1093: Fig.~\ref{fig:fluent} the flow velocity in the symmetry plane of a
1094: $3d$ barchan dune obtained with Fluent~5 \cite{Fluent5}. The wind is
1095: blowing from left to right. The boundaries were chosen to be periodic
1096: in the transverse direction. At the influx boundary, the velocity was
1097: fixed by imposing a logarithmic velocity profile. The wind profile at
1098: the outflux boundary is not known \emph{a priori}. Although, for high
1099: Reynolds numbers the latter is expected to affect the solution only
1100: close to the boundary, it is well known that different choices for the
1101: outflux boundary condition as well as different discretization schemes
1102: may lead to quantitatively different results \cite{wesseling:2000}.
1103: Here, we chose to set the derivative of the velocity normal to the
1104: outflux boundary to zero. The surface profile was represented as a
1105: solid boundary with constant roughness length. Finally, along the top
1106: boundary we imposed the velocity of the undisturbed logarithmic inflow
1107: profile. The whole calculation was performed on a grid that had an
1108: exponentially growing mesh size in the vertical direction. A
1109: considerable grid refinement was necessary in the wedge--like region
1110: of the separation bubble close to the brink.
1111:  
1112: \begin{figure}[tb]
1113:   \begin{center}
1114:    
1115:     \includegraphics[width=\columnwidth]{velocity_cut_black.eps}
1116:     \caption{Cut along the symmetry plane of a $3d$ barchan dune. The
1117:       velocity vectors calculated numerically with a commercial 
1118:       fluid dynamics solver \cite{Fluent5} clearly display the flow 
1119:       separation at the brink and a large eddy in the wake region.}
1120:       \label{fig:fluent}
1121: \end{center}
1122: \end{figure}
1123: 
1124: These remarks complete the first task of constructing a model for the
1125: calculation of the wind shear stress on a given dune profile as
1126: outlined in Eq.(\ref{eq:task1}). By deriving the linear
1127: Eq.(\ref{eq:wind}) for the shear stress and combining it with
1128: the heuristic separation bubble, we have obtained an approximate but
1129: numerically extremely efficient model for the wind shear stress on
1130: dunes. This is a crucial step in the construction of a minimal model
1131: of aeolian sand dunes, since the enormous complexity of the turbulent
1132: air flow over structured terrain otherwise severely restricts the
1133: possible applications of the model.
1134: 
1135: Going back to the upper part of Fig.~\ref{fig:v} with the above
1136: discussion in mind, we can re--interpret this figure in order to
1137: anticipate the behavior of the surface migration velocity $v(x)$ of a
1138: dune with slip face. If, for qualitative purposes, $f^2_C$ is
1139: interpreted as the envelope of a dune and its separation bubble, we
1140: can conclude that the slip face must be located near the sharp drop of
1141: $v(x)$ slightly upwind from the top of the envelope. This is indeed
1142: consistent with observations for large dunes. Together with the good
1143: representation of the windward profiles of large dunes
1144: \cite{sauermann-etal:2000} by $f^n_C$ ($n\approx 2$), it suggests that
1145: the given description becomes qualitatively correct in the limit of
1146: large dune sizes.  The next section is devoted to the discussion of
1147: important subtleties related to the fact that dunes are not typically
1148: in this limit.
1149: 
1150: 
1151: \section{Sand flux}\label{sec:saltation}
1152: As outlined in Eq.(\ref{eq:task2}), the second task in the
1153: specification of the minimal model is to find a prescription for
1154: calculating the sand flux $q(x)$ for a given topography $h(x)$ and
1155: shear stress $\tau(x)$. So far, we have been using the local
1156: saturated--flux approximation Eq.(\ref{eq:q0}) in our qualitative
1157: arguments. However, a closer look at the predictions obtained within
1158: this approximation reveals a number of inconsistencies.  First, as we
1159: have already noted in the discussion of Fig.~\ref{fig:v}, the use of
1160: Eq.(\ref{eq:q0}) together with the complete wind model of
1161: Section~\ref{sec:wind} leads to the odd prediction of deposition at
1162: the windward foot of an isolated heap or dune, where the shear stress
1163: decreases. This defect of Eq.(\ref{eq:q0}) has been noticed in the
1164: literature before (see e.g.\ Refs.\
1165: \cite{Wiggs96,sauermann-kroy-herrmann:2001}). Previous numerical
1166: studies tried to avoid this problem by focusing onto the short time
1167: behavior and by introducing \emph{ad hoc} heuristic methods such as a
1168: ``smoothing operator'' \cite{Wippermann86} or an ``adaptation length''
1169: \cite{van_dijk-arens-boxel:99}. The reason for the problem is that the
1170: saturated--flux approximation breaks down at a boundary
1171: ground/sand. As another shortcoming, we want to mention that the model
1172: as discussed so far predicts a universal scale invariant dune shape
1173: with a brink that is displaced slightly upwind from the maximum of the
1174: envelope, leading always to a positive slope at the brink.  A glance
1175: at a real dune field proves that the latter is not always the case and
1176: careful measurements \cite{sauermann-etal:2000} have revealed
1177: systematic deviations from scale invariance.  Though less obvious, it
1178: turns out that the reason for this discrepancy lies again in the
1179: saturated--flux approximation. Both mentioned problems are thus
1180: naturally resolved by introducing a slightly more general sand
1181: transport law that allows for saturation transients.
1182: 
1183: \subsection{Saturation transients}
1184: The saturated--flux approximation Eq.(\ref{eq:q0}) assumes that the
1185: flux is everywhere equal to the equilibrium transport capacity $q_s$
1186: of the wind. However, due to variable wind or sand conditions, the
1187: actual sand flux $q$ is in general different from $q_s$. These
1188: deviations are called saturation transients, because they quickly
1189: relax to zero under homogeneous conditions. We have recently
1190: demonstrated \cite{sauermann-kroy-herrmann:2001} that this relaxation
1191: occurs within a characteristic length scale, called the
1192: \emph{saturation length} $\ell_s$, which is related to (but distinct
1193: from) the mean saltation length of the grains. It was moreover shown
1194: how the introduction of saturation transients cures the problem of
1195: deposition at the windward foot of an isolated sand dune.  Here, we
1196: only summarize the most pertinent results of this earlier development
1197: in order to demonstrate how a size dependence of the dune shape
1198: naturally results as a consequence of saturation transients.
1199: 
1200: The sand transport model of Ref~\cite{sauermann-kroy-herrmann:2001} is
1201: based on a mean--field like description of saltation. It models a
1202: typical grain that is accelerated by friction with the air and slowed
1203: down by dissipative interactions with the bed. The average properties
1204: of the complicated splash process
1205: \cite{Anderson91,Nalpanis93,Rioual2000} are subsumed into two
1206: dimensionless parameters, an effective restitution coefficient
1207: $\alpha$ for collisions with the bed, and a kinetic coefficient
1208: $\gamma$ that characterizes the relaxation of the density of saltating
1209: grains to its saturated value. Together with an effective height for
1210: the wind--grain interaction that enters only logarithmically, these
1211: are the only phenomenological parameters of the model. They have been
1212: determined by a comparison with experiments and grain scale
1213: simulations.  Formally, the model consists of two coupled differential
1214: equations for mass and momentum conservation, and a modified turbulent
1215: closure relation that accounts for the feedback of the saltating
1216: grains on the wind velocity.
1217: 
1218: 
1219: For the present purpose, the model can be simplified by taking
1220: advantage of the fact that the prevailing conditions in applications
1221: to dunes are typically well described by the steady--state
1222: ($\partial/\partial t\simeq 0$) version. Further, the relaxation of
1223: the typical sand transport velocity can be assumed to be fast compared to
1224: the variations in the density of mobilized grains in the saltation
1225: layer. Approximating the latter by its saturated value for the
1226: calculation of the effective wind speed $u_{\rm eff}$ via the modified
1227: turbulent closure, one can decouple the mass and momentum conservation
1228: equations.  The whole model can then in a reasonable approximation be
1229: reduced to a single differential equation
1230: \begin{equation}
1231:   \label{eq:flux}
1232:   \ell_s\partial q/\partial x = q( 1-q/q_s)  
1233: \end{equation}
1234: for the sand flux $q(x)$.  The shear stress dependent parameters
1235: \begin{equation}\label{eq:ls}
1236:   \ell_s= l/(\tau/\tau_t-1)\;, \qquad q_s = \rho_s u_s 
1237: \end{equation}
1238: are immediately identified as the saturation length and the saturated
1239: flux, respectively.  The equation for $q_s$ generalizes
1240: Eq.(\ref{eq:bagnold}) to arbitrary wind speeds. In the following we
1241: specify the explicit expressions for both quantities as they result
1242: from the sand transport model of
1243: Ref.~\cite{sauermann-kroy-herrmann:2001}, but the structure of
1244: Eq.(\ref{eq:flux}) is thought to be more general and independent of
1245: the precise form of Eq.(\ref{eq:ls}). Again, $\tau(x)$ is the position
1246: dependent shear stress discussed in Section~\ref{sec:wind} and $\tau_t
1247: \approx 0.1\; \text{kg\,m}^{-1}\text{s}^{-1}$ is the estimated impact
1248: shear stress threshold that corresponds to a critical shear velocity
1249: $u_{*t}\approx 0.28 \;\text{m\,s}^{-1}$ \cite{Owen64}.  (For
1250: simplicity, we do not introduce the additional threshold for purely
1251: aerodynamic entrainment here, but allow instead for a small residual
1252: influx even if the latter is nominally zero.) To make the underlying
1253: structure of the model more palpable, we have expressed $\ell_s$ in
1254: terms of another characteristic length scale $l\equiv 2\alpha
1255: u_s^2/(g\gamma)$, which (up to a numerical factor) is the average
1256: saltation length of the grains. The latter --- but not $\ell_s$ ---
1257: must always be considerably smaller than the dune length for the model
1258: to be applicable.  Further, we have decomposed $q_s$ into the
1259: saturated density $\rho_s = 2\alpha(\tau-\tau_t)/g$ and the effective
1260: sand transport velocity at saturation $u_s = u_{\rm eff} - \delta u$
1261: with $u_{\rm eff}$ the effective wind velocity that accelerates the
1262: grains, given by
1263: \begin{equation}\label{eq:ueff}
1264:  u_{\rm eff}\kappa\sqrt{\varrho_{a}} = 2\sqrt{\tau_t+(\tau
1265: -\tau_t)/\zeta} + (\ln\zeta'-2)\sqrt{\tau_t}\;. 
1266: \end{equation}
1267: By $g$ we have denoted the gravitational acceleration and from
1268: Ref~\cite{sauermann-kroy-herrmann:2001} we adopt the (approximate)
1269: numerical values $\alpha=0.35$, $\gamma=0.2$, $\zeta= 8$, $\zeta'=
1270: 200$, and $\delta u= 1.8\;$m/s for the lag velocity of the grains. We
1271: note that these numerical values are not completely independent of
1272: each other and of the mentioned value for the impact threshold
1273: $\tau_t$, due to the calibration of the sand transport model with
1274: experimental data \cite{sauermann-kroy-herrmann:2001}.  For
1275: convenience we show a plot of the saturation length $\ell_s$ obtained
1276: with these values in Fig.~\ref{fig:ls}.  This completes the definition
1277: of the sand transport part of the minimal model on gently sloped
1278: ground.
1279: 
1280: \begin{figure}[tb]
1281:  \psfrag{T}{$\tau/\tau_t$}
1282:  \psfrag{L}{$\ell_s \;\text{[m]}$}
1283:   \begin{center}
1284:     \includegraphics[width=\columnwidth]{ls.eps} 
1285:     \caption{The
1286:     saturation length $\ell_s$ in meters as a function of the shear
1287:     stress exerted by the wind onto the sand bed. This function 
1288:     sets the natural length scale for dunes and heaps.}
1289: \label{fig:ls}
1290: \end{center}
1291: \end{figure}
1292: 
1293: 
1294: \subsection{Consequences}
1295: Before we complete the general model definition by a brief paragraph
1296: on slip faces, we want to point out some implications of the model as
1297: developed so far.  First, note that the full expression for $q_s$
1298: given in Eq.(\ref{eq:ls}), contains Eq.(\ref{eq:bagnold}) as a
1299: limiting case for strong winds but is better approximated by
1300: $q_s\propto \tau-\tau_t$ for moderate wind speeds. For weak flux
1301: gradients and strong winds, one may set the left hand side of
1302: Eq.(\ref{eq:flux}) to zero, leading to $q=q_s$. This is
1303: typically the case on most of the windward slope of a large dune,
1304: where the left hand side of Eq.(\ref{eq:flux}) can roughly be
1305: estimated by $\ell_s q_s/L \ll q_s$.  The local saturated--flux
1306: approximation with Eq.(\ref{eq:bagnold}) for $q_s$, which we have
1307: applied throughout our qualitative discussion so far, is thus
1308: asymptotically valid for large dunes and strong winds (except near the
1309: windward foot of an isolated dune). This is what one might have
1310: expected in the first place, and the reader may wonder at this point
1311: how the saturation transients and their characteristic length scale
1312: $\ell_s$ can have the claimed importance. How can $\ell_s$ affect the
1313: shape of a dune that is typically about two orders of magnitude
1314: larger? This apparent puzzle is now easily resolved by going back to
1315: Figs.~\ref{fig:hills}, \ref{fig:sym_asym} and Eq.(\ref{eq:shift}) and
1316: by observing that the symmetry breaking shift $\delta x_\tau$ of the
1317: location of the maximum of the shear stress with respect to that of
1318: the maximum of the height profile (or envelope) that is responsible
1319: for the finite windward slope and growth of dunes, is also of the
1320: order of some per cent of the total dune length. In summary, the
1321: longitudinal profile of dunes and heaps is determined by the
1322: competition of two quantitatively small but qualitatively crucial
1323: effects, one related to turbulent wind flow and the other to sediment
1324: transport. This may be the reason, why its explanation proved elusive
1325: for a long time.
1326: 
1327: To get a qualitative idea of the consequences of the introduction of
1328: the generalized non--local flux law in Eq.(\ref{eq:flux}) as
1329: replacement for Eq.(\ref{eq:q0}), we want again to go back to our
1330: discussion of the surface migration velocity of the cosine shaped heap
1331: $f^n_C$ in Fig.~\ref{fig:v}. Let us for the moment adopt a crude
1332: approximation and replace the expression for $q_s$ given in
1333: Eq.(\ref{eq:ls}) by its simpler limiting form $q_s \propto \tau^{3/2}$
1334: introduced in Eq.(\ref{eq:bagnold}). We also neglect the variation of
1335: $\ell_s$ on the dune and replace it by a (fine--tuned) constant
1336: $\ell_s\approx 0.1\, L$.  With an influx (about $0.7\, q_s$ for the
1337: solid and $0.8\, q_s$ for the dashed line in the lower part of
1338: Fig.~\ref{fig:v}) one can thus achieve a fairly constant surface
1339: velocity over the \emph{whole} length of a cosine shaped heap.  Again
1340: the constancy is slightly better for $n<2$ than for $n=2$. It is
1341: further improved by reducing the slope of the heap well below the
1342: optimum windward slope of the dune obtained for $\ell_s\to0$, as seen
1343: from a comparison of the solid line and the dashed line. We also note
1344: in passing that the influx needed to maintain the shape is increasing
1345: with decreasing slope. Together, the plots in Fig.~\ref{fig:v}
1346: confirms our claim that even an $\ell_s\ll L$ may visibly affect the
1347: overall shape of aeolian dunes.  Although, for the example shown in
1348: the lower part of Fig.~\ref{fig:v}, the saturation length is only
1349: about $1/30$ of the heap length, the slip face instability is
1350: evidently completely washed out. Altogether, this strongly suggests
1351: the existence of translation invariant cosine shaped heap solutions
1352: for the model.  The ultimate proof will be provided by the numerical
1353: results presented in Section~\ref{sec:solution}, where the full form
1354: of $q_s$ and $\ell_s$ according to Eq.(\ref{eq:ls}) will be used, but
1355: the present crude approximation already illustrates the main point,
1356: and also demonstrates that the behavior is a generic consequence of
1357: Eq.(\ref{eq:ls}) and insensitive to the detailed form of the
1358: parameters $\ell_s(x)$ and $q_s(x)$ that may phenomenologically be
1359: somewhat different from the model prediction without affecting our
1360: general conclusions.
1361: 
1362: An immediate consequence of the foregoing discussion is the existence
1363: of a minimum dune size. For small enough dunes, the slip face
1364: instability is washed out by the saturation transients. One may also
1365: arrive at this conclusion from an analysis of heaps. To this end we
1366: observe that the value of the saturation length $\ell_s$ is a property
1367: of the wind velocity and the saltation kinetics and depends on the
1368: topography only indirectly through the variable shear stress
1369: $\tau$. Moreover, it is apparent from Fig.~\ref{fig:ls} that this
1370: dependence becomes weak for strong winds. On the other hand, the
1371: symmetry breaking shift $\delta x_\tau$ is proportional to the
1372: absolute size of the heap (or envelope) and not directly dependent on
1373: the wind velocity. For the special profile $f_L$, this was verified
1374: analytically in Eq.(\ref{eq:shift}). As we have seen, a smooth heap
1375: can only be a translation invariant solution of the model if the lag
1376: (of order $\ell_s$) of $q(x)$ with respect to $q_s(x)$ and the shift
1377: $\delta x_\tau$ are fine--tuned to guarantee a vanishing erosion rate
1378: at the top of the heap. From this we expect heaps to obey
1379: \begin{equation}\label{eq:heap_aspect}
1380: \ell_s \simeq \delta x_\tau \propto L \approx \text{const.}
1381: \end{equation}
1382: to a first approximation. This condition can only be fulfilled if the
1383: aspect ratio $\varepsilon$ of heaps grows proportional to their size
1384: (i.e.\ roughly $\varepsilon\propto H$). Hence, in contrast to large
1385: dunes with slip face, for which we have argued that they are
1386: asymptotically scale invariant ($\varepsilon \sim$ const.), heaps must
1387: have a strongly size dependent aspect ratio. As a consequence,
1388: translation invariant heap solutions obviously cannot exist beyond a
1389: certain critical size. A slip face will develop when the shear stress
1390: on the lee side of the heap drops below the threshold value $\tau_t$,
1391: or at the latest, when the lee slope exceeds the critical slope for
1392: flow separation. This will be further analyzed in
1393: Section~\ref{sec:solution}.  Finally, we note that the steady--state
1394: flux of a heap can be estimated by the observation that the outflux is
1395: essentially determined by the strength of the reduction
1396: $\tau_0-\tau_{\rm min}$ of the shear stress at the lee end of the
1397: heap. According to Eq.(\ref{eq:scaling}), the latter is (for a given
1398: shape) proportional to the aspect ratio $\varepsilon$. For qualitative
1399: purposes, the outflux $q^{\rm out}$ may thus be estimated in the
1400: saturated flux approximation with
1401: Eqs.(\ref{eq:ls}),(\ref{eq:ueff}),(\ref{eq:scaling}) as
1402: \begin{equation}\label{eq:outflux}
1403: q_s^{\rm out} \propto \tau_{\rm min} -\tau_t \propto
1404: \varepsilon_c-\varepsilon \;,
1405: \end{equation}
1406: where we have assumed $\tau_{\rm min}/\tau_t \lesssim 2$ (fulfilled
1407: for moderate wind speeds and/or heaps near the critical heap size)
1408: to linearize the expression for $q_s(\tau)$ given in
1409: Eq.(\ref{eq:ls}). Here, $\varepsilon_c \propto \tau_0-\tau_t$ is the
1410: critical aspect ratio for which the shear stress on the lee drops
1411: below the threshold and the outflux vanishes. Note that the latter
1412: increases with increasing shear stress whereas the heap length
1413: decreases according to Eqs.(\ref{eq:ls}), (\ref{eq:heap_aspect}). The
1414: effects of the two trends onto the critical heap mass could therefore
1415: partially cancel unless the lee slope exceeds the critical slope for
1416: flow separation.
1417: 
1418: \subsection{Slip face}\label{sec:slipface}
1419: We have argued above that for large heaps ($L\gg\ell_s$), aeolian
1420: sediment transport tends to increase the downwind slope until it
1421: reaches the angle of repose of the grains. At this point, any further
1422: increase of the lee slope initiates avalanches that restore a slope
1423: slightly below the static angle of repose and eventually create a slip
1424: face of a roughly uniform slope of about $32^\circ - 35^\circ$. Since
1425: the physical modeling of this process itself is not a major objective
1426: of the present contribution, we can choose between different possible
1427: implementations for this phenomenon. In $2d$ it is possible to
1428: represent the slip face as boundary condition for the sand transport.
1429: It is uniquely determined by its fixed uniform slope and mass
1430: conservation. However, with regard to a future generalization of the
1431: present $2d$ model to $3d$ we chose a more realistic implementation
1432: based on a widely--used avalanche model \cite{Bouchaud94}. The
1433: formulation of this model bears some close similarities with the sand
1434: transport model presented in the preceding paragraph, and thus
1435: suggests itself as a natural extension of the latter to the slip
1436: face. This completes the definition of the minimal model that will be
1437: solved numerically in the next section.
1438: 
1439: \section{Solution of the model}\label{sec:solution}
1440: Apart from the model definition, the preceding sections have provided
1441: some qualitative insights into the main mechanisms responsible for
1442: dune formation and migration. Now we are prepared to study numerically
1443: the quantitative predictions of the model. Again, we emphasize that we
1444: only can explore some major features of the model in the present
1445: report, leaving many interesting questions and more systematic and
1446: quantitative parameter studies for future work.
1447: 
1448: For convenience, the solution procedure of the minimal model is
1449: summarized as a flow chart in Fig.~\ref{fig:flowchart}. One starts
1450: with an initial profile $h(x,t=0)$ (typically $f_G$ or $f_C^2$),
1451: checks whether a separation bubble has to be added for the calculation
1452: of the shear stress, then obtains the latter from
1453: Eq.(\ref{eq:fourier_wind}) and uses the result as input for the
1454: iterative solution of the sand transport equation
1455: Eq.(\ref{eq:flux}). This finally gives the erosion/deposition needed
1456: to update the surface profile. Technically, Eq.(\ref{eq:fourier_wind})
1457: is implemented as a Fast--Fourier--Transform algorithm, and for the
1458: integration of Eqs.(\ref{eq:flux}) and (\ref{eq:mass}) an upwind
1459: discretization scheme is used. Simulation times can be reduced by
1460: using an adaptive time step.
1461: 
1462: \begin{figure}[tb]
1463:   \psfrag{initial}[c][c]{\tiny $h(x,t=0)$}
1464:   \psfrag{bubble}[c][c]{\small Eq.(\ref{eq:sepbubble})}
1465:   \psfrag{tau}[c][c]{\small Eq.(\ref{eq:fourier_wind})}
1466:   \psfrag{flux}[c][c]{\small Eq.(\ref{eq:flux})}
1467:   \psfrag{mass}[c][c]{$\genfrac{}{}{0pt}{1}{\text{Eq.(\ref{eq:mass})}\; \&}
1468:     {\text{\tiny avalanches}}$}
1469:   \begin{center}   
1470:     \includegraphics[width=\columnwidth]{flowchart.eps}
1471:     \caption{Solution of the minimal model.}
1472:       \label{fig:flowchart} 
1473: \end{center}
1474: \end{figure}
1475: 
1476:  
1477: \subsection{Steady--state shapes}
1478: The scheme of Fig.~\ref{fig:flowchart} can be iterated for different
1479: influx boundary conditions. For all of the numerical calculations
1480: presented below, we chose periodic boundary conditions. They are the
1481: natural choice for studies of the steady--state shapes.  To
1482: investigate the mass balance under prescribed influx conditions, on
1483: the other hand, one has to apply open boundary conditions. 
1484: 
1485: 
1486: \begin{figure}[tb]
1487:   \psfrag{h in m}[][90]{$h(x)$ [m]} 
1488:   \psfrag{x in m}{$x$ [m]}
1489:   \begin{center} 
1490: \includegraphics[width=\columnwidth]{heaps.eps}
1491: \includegraphics[width=\columnwidth]{dunes.eps}
1492:   \caption{Steady--state heaps (upper plot) and dunes (lower
1493:   plot). The aspect ratio is stretched for better visualization.}
1494:   \label{fig:shapes}
1495: \end{center}
1496: \end{figure}
1497: 
1498: 
1499: Fig.~\ref{fig:shapes} shows steady--state solutions of the model for
1500: initial profiles $f_G$ of different mass. These solutions are obtained
1501: for fixed wind conditions with parameters $A=3.2$ and $B=0.25$
1502: appropriate for the central slice of a $3d$ (symmetric) heap or of a
1503: barchan dune. The shear velocity $u_*=0.4$ m/s lies well above the
1504: impact threshold. (The situation very close to or below the threshold
1505: would need special attention.) As anticipated above, large dunes
1506: become asymptotically scale invariant. The asymptotic master curve for
1507: the windward profile is practically indistinguishable from the profile
1508: $f_C^n$ ($n\lesssim 2$), and the slope at the brink is indeed
1509: positive. Its average windward slope is inversely proportional to the
1510: value of the parameter $A$ given in Eq.(\ref{eq:a_b}). Due to the
1511: additional terms in the expression Eq.(\ref{eq:correction}) for the
1512: shear stress on dunes with a finite width, somewhat steeper average
1513: windward slopes are predicted for barchan dunes than for transverse
1514: dunes under identical influx and wind conditions. However, a detailed
1515: quantitative comparison is probably beyond the scope of the present
1516: semi--quantitative implementation. More important are the remarkable
1517: qualitative predictions of the model. In particular, the fact that
1518: dunes with slip face are only stable above a certain (wind dependent)
1519: critical size, whereas smooth steady--state heaps only exist below a
1520: critical size, deserves attention. We also note that the steady state
1521: is not always unique. There is a hysteretic regime, where the initial
1522: conditions can select one of two possible steady--state shapes and
1523: accordingly the masses for the two sets of profiles in
1524: Fig.~\ref{fig:shapes} are not all distinct. The largest heaps in the
1525: upper plot were obtained from flat initial profiles $f_G$, whereas the
1526: smallest dunes with slip face in the lower plot were obtained from
1527: steeper initial profiles $f_G$ of the same mass. Especially, the dune
1528: with a negative slope at the brink could only be obtained from steep
1529: initial conditions. Since under natural wind and sand conditions, the
1530: initial conditions themselves will generally be heaps or dunes close
1531: to the steady state, one can say that the model predicts a critical
1532: heap size for slip face formation and a critical dune size for slip
1533: face destruction. In both cases the slip face is finite as a
1534: consequence of flow separation.  The latter also allows a dune to be
1535: somewhat higher than a heap of the same mass, since its effective
1536: volume as seen by the average air flow is increased by the separation
1537: zone. As anticipated, the aspect ratio of the dunes is asymptotically
1538: constant, whereas it is strongly size--dependent for heaps. This
1539: effect can be seen more quantitatively in the representation of
1540: Fig.~\ref{fig:H_V}, where the height $H$ of the steady--state heaps
1541: and dunes is plotted versus the product $HL$ of their height and
1542: length $L$. Clearly, heaps are better described by $H\propto H L$ as
1543: predicted in Eq.(\ref{eq:heap_aspect}), whereas large dunes approach
1544: the scaling limit $H\propto \sqrt{H L}$.
1545: 
1546: \begin{figure}[tb]
1547:   \psfrag{flat}{\small flat}
1548:   \psfrag{steep}{\small steep}
1549:   \psfrag{H}[rl]{$H$ [m]}
1550:   \psfrag{V}[t][b]{$H L$ [m$^2$]}
1551:   \psfrag{infit}{\small $\propto H L$}
1552:   \psfrag{asfit}{\small $\propto \sqrt{H L}$}
1553:   \begin{center}   
1554:     \includegraphics[width=\columnwidth]{H_HLrl.eps}
1555:    \caption{Steady--state heights $H$ versus the product of the height
1556:   and length of the heaps and dunes. In the hysteretic
1557:   regime, flat and steep initial conditions have to be distinguished.}
1558:     \label{fig:H_V}
1559: \end{center}
1560: \end{figure}
1561: 
1562: \begin{figure}[tb]
1563:   \begin{center} 
1564: 
1565:   \psfrag{ustar035xxxx}{\tiny $u_*=0.35$}
1566:   \psfrag{ustar040xxxx}{\tiny $u_*=0.40$} 
1567:   \psfrag{ustar045xxxx}{\tiny $u_*=0.45$} 
1568:   \psfrag{ustar050xxxx}{\tiny $u_*=0.50$}
1569:   \psfrag{propto1hxxxx}{\tiny $\propto 1/H$}
1570:   \psfrag{propto1Lxxxx}{\tiny $\propto 1/L$} 
1571:   \psfrag{height in m}[t][b]{$H$ [m]} 
1572:   \psfrag{length in m}[t][b]{$L$ [m]} 
1573:   \psfrag{vd in m per year}{$v$ [m/yr]} 
1574:   \psfrag{Lvd in m per year}{}   
1575:   \includegraphics[width=0.45\columnwidth]{2d_v_vs_h.eps}
1576:   \includegraphics[width=0.45\columnwidth]{2d_v_vs_L.eps}
1577:   \caption{Migration velocities predicted by the minimal model for
1578:   steady--state dunes of different size at various wind
1579:   velocities. The caption gives the shear velocity $u_*$ in m/s. The
1580:   numerical data are compared to the scaling laws $v\propto H^{-1}$
1581:   (left) and $v\propto L^{-1}$, where $L$ is the length of the
1582:   envelope of the dune and its separation bubble as described in the
1583:   main text. (Note that the migration of real dunes is
1584:   substantially slower due to the small fraction of wind days per
1585:   year.)}  \label{fig:v_H_L}
1586: \end{center}
1587: \end{figure}
1588: 
1589: \subsection{Migration velocity}
1590: For the overall migration velocity of steady--state dunes with a scale
1591: invariant profile, we derived on general grounds the simple scaling
1592: prediction $v\propto L^{-1}$ in Section~\ref{sec:v}. We have also
1593: given some arguments why this prediction should be rather robust
1594: against relaxing the condition of shape invariance, in contrast to the
1595: relation $v\propto H^{-1}$ that can only be inferred from it if the
1596: scaling assumption holds exactly. Here, we check these predictions
1597: for the steady--state solutions numerically. Fig.~\ref{fig:v_H_L}
1598: shows the numerically obtained migration velocity for dunes fitted to
1599: the scaling relations $v\propto H^{-1}$ (left) and $v\propto L^{-1}$
1600: (right).  As we have mentioned above, one has to take for $L$ the
1601: characteristic length of the envelope rather than that of the dune
1602: alone. For simplicity, we estimate $L$ by adding $6H$ to the
1603: horizontal length from the windward foot of the dune to its crest,
1604: thus neglecting the weak slope dependence of the size of the
1605: separation bubble. Obviously, the $L^{-1}-$scaling is superior for
1606: moderate winds and small dunes where $v\propto H^{-1}$ systematically
1607: fails to describe the data. This is also supported by field data
1608: \cite{Finkel59}. Both fits become identical in the scaling limit $L\gg
1609: \ell_s$. Due to the decrease of $\ell_s$ with the wind speed, the
1610: latter is reached for smaller dunes at stronger winds.
1611: 
1612: \subsection{Stability}
1613: We have already pointed out that the choice of different boundary
1614: conditions for the flux allow a separate discussion of shape and mass
1615: stability. This is of practical importance, since (in $2d$) all
1616: steady--state shapes are unstable with respect to mass changes. If the
1617: influx of a steady--state solution deviates slightly from its
1618: corresponding steady--state flux, this solution will start to either
1619: shrink until it has flattened out or grow without bound.  Though the
1620: latter effect could (but need not) be a peculiarity of the vanishing
1621: outflux for $2d$ dunes with slip face, at least the former generalizes
1622: to $3d$ heaps. Despite the fact that the steady--state shapes are
1623: (locally) stable attractors for the shape evolution under periodic
1624: flux conditions, mass stability can in general not be achieved under
1625: open boundary conditions. The situation is clarified in
1626: Fig.~\ref{fig:q_V}. It depicts the steady--state sand flux $q$ over
1627: the bedrock as a function of aspect ratio. The numerical results
1628: nicely confirm our theoretical expectation from
1629: Eq.(\ref{eq:outflux}). For all dunes with slip face the flux vanishes
1630: identically in $2d$, whereas in general it grows with decreasing size
1631: for smooth heaps.  For open boundary conditions, the line in
1632: Fig.~\ref{fig:q_V} can be interpreted as an unstable phase boundary
1633: (with hysteresis) between infinitely growing and shrinking
1634: solutions. For example, a heap with influx slightly below the
1635: steady--state, will shrink a bit. To remain close to the steady--state
1636: shape, it will therefore mainly reduce its height, whereas its length
1637: will stay almost constant. Due to the reduced aspect ratio
1638: $\varepsilon$, $\hat \tau$ decreases in magnitude and the shear stress
1639: depression at the lee boundary is less pronounced. As a consequence
1640: there is less deposition on the downwind slope and the outflux is
1641: higher, so that the heap shrinks even more etc. A completely analogous
1642: reasoning applies to the opposite case of higher influx. The
1643: corresponding shape attractors are the scale invariant asymptotic dune
1644: shape and the flat surface, respectively.
1645: 
1646: The above discussion explains, why isolated smooth aeolian sand heaps
1647: are rarely observed as distinct features of desert topographies. Under
1648: approximately stationary wind and influx conditions they only exist as
1649: transient states that either vanish or develop into dunes with slip
1650: face. Under variable wind and influx conditions, the situation is less
1651: clear and deserves a detailed study of its own. For example, the model
1652: predicts that during a period of strong wind all dunes are driven
1653: towards the asymptotic shape. After a subsequent period of weak winds,
1654: finite size effects become more pronounced, and small dunes may
1655: develop longitudinal profiles like those in the hysteretic regime or
1656: even loose their slip face. Again, the case $\tau_0 \approx \tau_t$ of
1657: a shear stress close to or below the threshold shear stress needs
1658: special attention.
1659: 
1660: The prevailing wind conditions as well as recent changes in the wind
1661: velocity are thus encoded in a complicated but comprehensible way in
1662: the shapes of the dunes in a dune field. This is a promising direction
1663: for further studies. One may hope that by systematic studies
1664: along these lines one will in the future be able to infer flow
1665: conditions in remote or uncomfortable places (e.g.\ on the sea bottom
1666: or on other planets) by analyzing dune
1667: shapes.
1668: 
1669: \begin{figure}[tb]
1670:   \psfrag{flat}{ flat} 
1671:   \psfrag{steep}{ steep} 
1672:   \psfrag{qin}[c]{$q^{\rm out}$ [kg\,m$^{-1}$s$^{-1}$]}  
1673:   \psfrag{eps}{$H/L$}
1674:   \psfrag{growing}{$\stackrel{\displaystyle
1675:   \text{growing}}{\text{solutions}}$}
1676:   \psfrag{shrinking}{$\stackrel{\displaystyle
1677:   \text{shrinking}}{\text{solutions}}$} 
1678: \begin{center}
1679:   \includegraphics[width=\columnwidth]{q_aspect.eps}
1680:   \caption{Steady--state outflux under periodic boundary
1681:   conditions. In the hysteretic regime, steep and flat initial
1682:   conditions have to be distinguished, as in Fig.~\ref{fig:H_V} The
1683:   figure may also be read as a phase diagram for the situation with
1684:   open boundary conditions. In this case the steady--state solutions
1685:   --- though attractors for the shape --- are unstable with respect to
1686:   mass fluctuations.}  \label{fig:q_V}
1687: \end{center}
1688: \end{figure}
1689: 
1690: 
1691: \subsection{Relaxation dynamics}\label{sec:relaxation}
1692: As a first step towards an understanding of the effects of variable
1693: wind speeds (for constant wind direction), this section is devoted to
1694: an exploratory study of the transient shape evolution. We restrict
1695: ourselves to periodic boundary conditions leaving the richer phase
1696: space of open boundary conditions for future
1697: studies. Fig.~\ref{fig:growth} shows two extreme scenarios. A flat
1698: initial condition with a mass greater than the critical mass for slip
1699: face formation (a), and a steep initial condition with a mass below
1700: the critical mass for slip face destruction (b). The steep initial
1701: condition gives rise to the temporary formation of a slip face that is
1702: finally washed out by the saturation transients, whereas the flat heap
1703: steepens until the shear stress on the lee falls to the threshold
1704: value.  This causes complete deposition on the lee side of the
1705: heap. Whether this happens before or with the onset of flow separation
1706: depends on wind and influx conditions. As it also depends on the
1707: precise numerical values of some of the phenomenological parameters of
1708: the model, a detailed parametric study is again beyond the scope of
1709: the present contribution.
1710: 
1711: \begin{figure}[t]
1712:   \begin{center}  
1713:     \includegraphics[width=\columnwidth]{comb_slipface.eps}
1714:     \caption{Growth histories for two Gaussian heaps of different mass
1715:       and initial aspect ratio for periodic flux boundary conditions.
1716:       Note that the distances to reach the steady--state shape are
1717:       different.  (The aspect ratio of both plots was rescaled by a
1718:       common factor for better visualization.)}
1719: \label{fig:growth} 
1720: \end{center}
1721: \end{figure}\noindent
1722: 
1723: Although the times to reach the steady state are apparently somewhat
1724: longer for the flat initial condition, it is evident from
1725: Fig.~\ref{fig:growth} that the relaxation dynamics is in general
1726: relatively fast even if the initial condition is far from the steady
1727: state shape. Large dunes under low influx conditions as they prevail
1728: e.g.\ in fields of isolated dunes should therefore be well described
1729: by an adiabatic approximation assuming that (except after drastic
1730: changes in the wind and sand conditions as they occur during sand
1731: storms) the dune is practically in a steady state.  Apart from the low
1732: influx, this also relies on the fact that virtually no sand is lost
1733: over the slip face. For a large isolated $3d$ barchan dune this
1734: implies that most of the sediment transported over the dune is
1735: actually trapped in a tread--milling flux, and only a small portion of
1736: the total flux is contributed by and contributes to the external flux.
1737: Hence, under steady wind conditions these dunes are in a quasi steady
1738: state and thus very close to their true steady--state
1739: shape. Investigation of the steady--state properties is therefore the
1740: starting point also for the study of their time evolution. Moreover,
1741: this suggests that a comparison of our steady--state shapes to shapes
1742: obtained in field measurements is justified. In fact, the calculated
1743: shapes agree nicely with recent measurements for barchan dunes
1744: \cite{sauermann-etal:2000}. The situation is less clear for small
1745: heaps, where mass losses can be of the order of the total flux and may
1746: thus lead to significant differences between the steady--state and the
1747: transient shapes under vanishing influx.
1748: 
1749: In the remainder of this section, we want to investigate more closely
1750: the mechanism that drives the shape relaxation. As we have pointed
1751: out, the positions of the maximum of the sand flux and of the maximum
1752: of the profile must coincide in the steady state to make the
1753: erosion/deposition vanish at the crest. We have shown that for small
1754: heaps, this can be achieved by a fine--tuning of $\delta x_\tau$ to
1755: about $\ell_s$. In contrast, for large dunes and strong winds, $\delta
1756: x_\tau \gg \ell_s$, and the steady--state condition can only be met
1757: with a singularity at the crest. This important difference is
1758: exemplified by Figs.~\ref{fig:deltaheap} and \ref{fig:deltadune}. Both
1759: figures show the evolution of the height and the displacements $\delta
1760: x_\tau$ and $\delta x_q$ of the locations of the maximum of the shear
1761: stress and of the maximum of the sand flux from the location of the
1762: top of the sand profile, respectively. The distance between both
1763: displacements is the lag of the flux with respect to the shear stress
1764: due to the saturation transients, and is therefore closely related to
1765: the saturation length $\ell_s$ for smooth surface profiles.  It
1766: guarantees the proper vanishing of the erosion rate $q'$ at the top of
1767: a steady--state heap where $\delta x_\tau$ is finite, but vanishes for
1768: large steady--state dunes, where the slip face ends in a sharp brink
1769: singularity at which the grains fall into the wake and are quickly
1770: deposited.
1771: 
1772: \begin{figure}[tb]
1773:   \begin{center}  
1774:   \psfrag{h in m}{$H$ [m]}
1775:   \psfrag{delta in m}{$\delta x_{\tau, q}$ [m]}
1776:   \psfrag{time}{time}
1777:   \psfrag{xmaxtau}{$\delta x_\tau$}
1778:   \psfrag{xmaxqxx}{$\delta x_q$}
1779:   \includegraphics[width=\columnwidth]{2d_displacement_vs_time_bump.eps} 
1780:  \caption{The figure shows the transient evolution of various
1781:   interesting length scales for a heap. \emph{Lower part:} Height of
1782:   the heap. \emph{Upper part:} Distance of the locations of the
1783:   maximum of the shear stress and of the maximum of the sand flux from
1784:   the position of the top of the heap. In the steady state, the
1785:   erosion/deposition vanishes at the crest.}  \label{fig:deltaheap}
1786: \end{center}
1787: \end{figure}
1788: 
1789: \begin{figure}[tb]
1790:   \begin{center}  
1791:   \psfrag{h in m}{$H$ [m]}
1792:   \psfrag{delta in m}{$\delta x_{\tau, q}$ [m]}
1793:   \psfrag{time}{time}
1794:   \psfrag{xmaxtau}{$\delta x_\tau$}
1795:   \psfrag{xmaxqxx}{$\delta x_q$}
1796: 
1797:   \includegraphics[width=\columnwidth]{2d_displacement_vs_time.eps}
1798:   \caption{The figure shows the transient evolution of various
1799:   interesting length scales for a dune that develops out of a smooth
1800:   heap as in Fig.~\ref{fig:growth}. \emph{Lower part:} Height of the
1801:   dune. \emph{Upper part:} Distance of the locations of the maximum of
1802:   the shear stress and of the maximum sand flux from the position of
1803:   the top of the crest. The lag between shear stress and sand flux
1804:   vanishes, when the slip face reaches the crest.}
1805:   \label{fig:deltadune}
1806: \end{center}
1807: \end{figure} \noindent
1808: 
1809: \section{Summary and outview}\label{sec:sum}
1810: In summary, we have shown that a simple minimal model for the
1811: wind--driven sediment transport over a sand dune is capable of
1812: explaining several important features of desert topographies. Among
1813: them are the migration velocities of heaps and dunes, their shape
1814: along the wind direction and the existence of a minimal dune size and
1815: a maximum heap size.
1816: 
1817: As we have emphasized throughout this contribution and demonstrated by
1818: the numerical solutions in the preceding section, the symmetry
1819: breaking part of the shear stress exerted by turbulent air flow on an
1820: obstacle, and local deviations of the sediment flux from its
1821: equilibrium transport capacity (``saturation transients''), are the
1822: essential ingredients in the modeling of aeolian sand dunes. It is
1823: exactly the balance of these two relatively small effects that is
1824: responsible for the relaxation of arbitrary initial conditions into a
1825: characteristic dune or heap shape. Their neglect was responsible for
1826: the failure of the naive zeroth order model discussed in
1827: Section~\ref{sec:zero}. In hindsight we can say that it is not so much
1828: the quantitative errors but the omission of this \emph{qualitatively}
1829: important mechanism, what makes the zeroth order model an insufficient
1830: description. In contrast, taking this balance properly into account,
1831: makes the minimal model structurally stable against the neglect of
1832: less significant quantitative details of the same order of magnitude.
1833: 
1834: \begin{figure}[tb]
1835:   \psfrag{L}{$L$ [m]}
1836:   \psfrag{ustar}{$u_*/u_{*t}$}
1837:   \begin{center}
1838:   \includegraphics[width=\columnwidth]{shape_diagram.eps}
1839:   \caption{Qualitative shape diagram that could be useful
1840:   in the analysis and comparison of field studies. The migration
1841:   velocity is constant along rising lines, whereas falling lines
1842:   indicate invariant dune shape.} \label{fig:shapediagram}
1843: \end{center}
1844: \end{figure}
1845: 
1846: This direction was recently pursued further in an effort to calculate
1847: analytically certain steady state shapes of dunes and heaps by
1848: ``linearizing'' the minimal model \cite{andreotti-claudin:2002}. One
1849: may as well wish to proceed also into the opposite direction. After
1850: the basic mechanism is understood, more elaborate dune models can be
1851: constructed by putting some of the neglected details back into the
1852: description. Detailed parametric studies of such a refined model for a
1853: certain dune type and comparison to field data would be very useful to
1854: test some of the less generic predictions of the underlying sand
1855: transport model \cite{sauermann-kroy-herrmann:2001}, such as the shear
1856: stress dependence of the saturation length $\ell_s$
1857: (Fig.~\ref{fig:ls}). This is important, since, as we have shown, the
1858: variable parameter $\ell_s$ sets the characteristic length scale with
1859: respect to which dunes and heaps can be said to be large or
1860: small. Phenomenological knowledge about $\ell_s$ is still very
1861: limited. More detailed studies could further be helpful to map out
1862: quantitative shape diagrams, of the type sketched qualitatively in
1863: Fig.~\ref{fig:shapediagram}. These diagrams could be useful not only
1864: for the validation of the model, but also for the comparison of field
1865: data from different places with different prevailing wind and sand
1866: conditions. The migration velocity is constant along the rising lines
1867: in Fig.~\ref{fig:shapediagram}, which were obtained from
1868: Eq.(\ref{eq:v}) using $q\approx q_s$ together with
1869: Eq.(\ref{eq:ls}). They allow for example a comparison of the migration
1870: velocities of dunes of different sizes that are exposed to (on
1871: average) identical winds. Or one may infer the average wind speed from
1872: measurements of sizes and migration velocities in a dune field. The
1873: falling lines in Fig.~\ref{fig:shapediagram} are lines of constant
1874: shape, assuming that the latter is determined by $\ell_s/L$. They may
1875: thus be used for correlating wind speeds with dune shapes. In general
1876: (in particular for the full $3d$ problem), such shape diagrams will be
1877: more complex since the influx is an additional important variable that
1878: we have neglected here, as it vanishes for $2d$ dunes in the steady
1879: state.
1880: 
1881: Moreover, as we pointed out, there are still many consequences of the
1882: present model that await a systematic investigation.  And a major
1883: future task is finally the generalization of the present discussion to
1884: the $3d$ case. A promising route could be the construction of an
1885: effectively sliced model that allows one to use the proposed model for
1886: the separation bubble and to keep the time--limiting calculation (the
1887: integration of the flux equation) effectively one--dimensional. The
1888: smaller transverse currents could be inferred from the sliced
1889: solution. A generalization of the flux equation to the $2d$ surface of
1890: a $3d$ dune is also feasible \cite{sauermann:phd}.  A more ambitious
1891: task will eventually be the simulation of dune fields.  Whereas the
1892: existence of a minimum dune size could be obtained by an analysis of
1893: the shape stability alone, the question of a possible existence of a
1894: characteristic or maximum dune size in a dune field, depends on the
1895: mass balance of a dune in the complicated environment provided by the
1896: other dunes, and is much more difficult to answer
1897: \cite{lima-etal:2002}.
1898: 
1899: \begin{acknowledgments}
1900: We gratefully acknowledge financial support by the Deutsche
1901: Forschungsgemeinschaft through contract No HE 2732/1-1, and helpful
1902: discussions with Ken Andersen and Philippe Claudin.
1903: \end{acknowledgments}
1904: 
1905: %\bibliography{journals,basics,dune,special,unpub}
1906: 
1907: \begin{thebibliography}{40}
1908: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1909: \expandafter\ifx\csname bibnamefont\endcsname\relax
1910:   \def\bibnamefont#1{#1}\fi
1911: \expandafter\ifx\csname bibfnamefont\endcsname\relax
1912:   \def\bibfnamefont#1{#1}\fi
1913: \expandafter\ifx\csname citenamefont\endcsname\relax
1914:   \def\citenamefont#1{#1}\fi
1915: \expandafter\ifx\csname url\endcsname\relax
1916:   \def\url#1{\texttt{#1}}\fi
1917: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1918: \providecommand{\bibinfo}[2]{#2}
1919: \providecommand{\eprint}[2][]{\url{#2}}
1920: 
1921: \bibitem[{\citenamefont{Bagnold}(1941)}]{Bagnold41}
1922: \bibinfo{author}{\bibfnamefont{R.~A.} \bibnamefont{Bagnold}},
1923:   \emph{\bibinfo{title}{The physics of blown sand and desert dunes.}}
1924:   (\bibinfo{publisher}{Methuen}, \bibinfo{address}{London},
1925:   \bibinfo{year}{1941}).
1926: 
1927: \bibitem[{\citenamefont{Pye and Tsoar}(1990)}]{Pye90}
1928: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Pye}} \bibnamefont{and}
1929:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Tsoar}},
1930:   \emph{\bibinfo{title}{Aeolian sand and sand dunes}}
1931:   (\bibinfo{publisher}{Unwin Hyman}, \bibinfo{address}{London},
1932:   \bibinfo{year}{1990}).
1933: 
1934: \bibitem[{\citenamefont{Lancaster}(1995)}]{Lancaster95}
1935: \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Lancaster}},
1936:   \emph{\bibinfo{title}{Geomorphology of desert dunes}}
1937:   (\bibinfo{publisher}{Routledge}, \bibinfo{address}{London},
1938:   \bibinfo{year}{1995}).
1939: 
1940: \bibitem[{\citenamefont{Malin et~al.}(1998)\citenamefont{Malin, Carr,
1941:   Danielson, Davies, Hartmann, Ingersoll, James, Masursky, McEwen, Soderblom
1942:   et~al.}}]{malin-etal:98}
1943: \bibinfo{author}{\bibfnamefont{M.~C.} \bibnamefont{Malin}},
1944:   \bibinfo{author}{\bibfnamefont{M.~H.} \bibnamefont{Carr}},
1945:   \bibinfo{author}{\bibfnamefont{G.~E.} \bibnamefont{Danielson}},
1946:   \bibinfo{author}{\bibfnamefont{M.~E.} \bibnamefont{Davies}},
1947:   \bibinfo{author}{\bibfnamefont{W.~K.} \bibnamefont{Hartmann}},
1948:   \bibinfo{author}{\bibfnamefont{A.~P.} \bibnamefont{Ingersoll}},
1949:   \bibinfo{author}{\bibfnamefont{P.~B.} \bibnamefont{James}},
1950:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Masursky}},
1951:   \bibinfo{author}{\bibfnamefont{A.~S.} \bibnamefont{McEwen}},
1952:   \bibinfo{author}{\bibfnamefont{L.~A.} \bibnamefont{Soderblom}},
1953:   \bibnamefont{et~al.}, \bibinfo{journal}{Science}
1954:   \textbf{\bibinfo{volume}{279}}, \bibinfo{pages}{1681} (\bibinfo{year}{1998}).
1955: 
1956: \bibitem[{\citenamefont{Thomas et~al.}(1999)\citenamefont{Thomas, Malin, Carr,
1957:   Danielson, Davies, Hartmann, Ingersoll, James, McEwen, Soderblom
1958:   et~al.}}]{Thomas99}
1959: \bibinfo{author}{\bibfnamefont{P.~C.} \bibnamefont{Thomas}},
1960:   \bibinfo{author}{\bibfnamefont{M.~C.} \bibnamefont{Malin}},
1961:   \bibinfo{author}{\bibfnamefont{M.~H.} \bibnamefont{Carr}},
1962:   \bibinfo{author}{\bibfnamefont{G.~E.} \bibnamefont{Danielson}},
1963:   \bibinfo{author}{\bibfnamefont{M.~E.} \bibnamefont{Davies}},
1964:   \bibinfo{author}{\bibfnamefont{W.~K.} \bibnamefont{Hartmann}},
1965:   \bibinfo{author}{\bibfnamefont{A.~P.} \bibnamefont{Ingersoll}},
1966:   \bibinfo{author}{\bibfnamefont{P.~B.} \bibnamefont{James}},
1967:   \bibinfo{author}{\bibfnamefont{A.~S.} \bibnamefont{McEwen}},
1968:   \bibinfo{author}{\bibfnamefont{L.~A.} \bibnamefont{Soderblom}},
1969:   \bibnamefont{et~al.}, \bibinfo{journal}{Nature}
1970:   \textbf{\bibinfo{volume}{397}}, \bibinfo{pages}{592} (\bibinfo{year}{1999}).
1971: 
1972: \bibitem[{\citenamefont{Kroy et~al.}(2002)\citenamefont{Kroy, Sauermann, and
1973:   Herrmann}}]{kroy-sauermann-herrmann:2002}
1974: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Kroy}},
1975:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Sauermann}},
1976:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.~J.}
1977:   \bibnamefont{Herrmann}}, \bibinfo{journal}{Phys.~Rev.~Lett.}
1978:   \textbf{\bibinfo{volume}{88}}, \bibinfo{pages}{054301}
1979:   (\bibinfo{year}{2002}).
1980: 
1981: \bibitem[{\citenamefont{Hoyle and Woods}(1997)}]{hoyle-woods:97}
1982: \bibinfo{author}{\bibfnamefont{R.~B.} \bibnamefont{Hoyle}} \bibnamefont{and}
1983:   \bibinfo{author}{\bibfnamefont{A.~W.} \bibnamefont{Woods}},
1984:   \bibinfo{journal}{Phys.~Rev.~E} \textbf{\bibinfo{volume}{56}},
1985:   \bibinfo{pages}{6861} (\bibinfo{year}{1997}).
1986: 
1987: \bibitem[{\citenamefont{Hoyle and Mehta}(1999)}]{hoyle-mehta:99}
1988: \bibinfo{author}{\bibfnamefont{R.~B.} \bibnamefont{Hoyle}} \bibnamefont{and}
1989:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Mehta}},
1990:   \bibinfo{journal}{Phys.~Rev.~Lett.} \textbf{\bibinfo{volume}{83}},
1991:   \bibinfo{pages}{5170} (\bibinfo{year}{1999}).
1992: 
1993: \bibitem[{\citenamefont{Prigozhin}(1999)}]{prigozhin:99}
1994: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Prigozhin}},
1995:   \bibinfo{journal}{Phys.~Rev.~E} \textbf{\bibinfo{volume}{60}},
1996:   \bibinfo{pages}{729} (\bibinfo{year}{1999}).
1997: 
1998: \bibitem[{\citenamefont{Sauermann et~al.}(2001)\citenamefont{Sauermann, Kroy,
1999:   and Herrmann}}]{sauermann-kroy-herrmann:2001}
2000: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Sauermann}},
2001:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Kroy}}, \bibnamefont{and}
2002:   \bibinfo{author}{\bibfnamefont{H.~J.} \bibnamefont{Herrmann}},
2003:   \bibinfo{journal}{Phys.~Rev.~E} \textbf{\bibinfo{volume}{64}},
2004:   \bibinfo{pages}{031305} (\bibinfo{year}{2001}).
2005: 
2006: \bibitem[{\citenamefont{Jackson and Hunt}(1975)}]{jackson-hunt:75}
2007: \bibinfo{author}{\bibfnamefont{P.~S.} \bibnamefont{Jackson}} \bibnamefont{and}
2008:   \bibinfo{author}{\bibfnamefont{J.~C.~R.} \bibnamefont{Hunt}},
2009:   \bibinfo{journal}{Q. J. R. Meteorol. Soc.} \textbf{\bibinfo{volume}{101}},
2010:   \bibinfo{pages}{929} (\bibinfo{year}{1975}).
2011: 
2012: \bibitem[{\citenamefont{Sykes}(1980)}]{sykes:80}
2013: \bibinfo{author}{\bibfnamefont{R.~I.} \bibnamefont{Sykes}},
2014:   \bibinfo{journal}{J.~Fluid Mech.} \textbf{\bibinfo{volume}{101}},
2015:   \bibinfo{pages}{647} (\bibinfo{year}{1980}).
2016: 
2017: \bibitem[{\citenamefont{Zeman and Jensen}(1988)}]{zeman-jensen:88}
2018: \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Zeman}} \bibnamefont{and}
2019:   \bibinfo{author}{\bibfnamefont{N.~O.} \bibnamefont{Jensen}},
2020:   \bibinfo{type}{Tech. Rep.} \bibinfo{number}{M-2738},
2021:   \bibinfo{institution}{Ris{\o} National Laboratory} (\bibinfo{year}{1988}).
2022: 
2023: \bibitem[{\citenamefont{Hunt et~al.}(1988)\citenamefont{Hunt, Leibovich, and
2024:   Richards}}]{hunt-leibovich-richards:88}
2025: \bibinfo{author}{\bibfnamefont{J.~C.~R.} \bibnamefont{Hunt}},
2026:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Leibovich}},
2027:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{K.~J.}
2028:   \bibnamefont{Richards}}, \bibinfo{journal}{Q. J. R. Meteorol. Soc.}
2029:   \textbf{\bibinfo{volume}{114}}, \bibinfo{pages}{1435} (\bibinfo{year}{1988}).
2030: 
2031: \bibitem[{\citenamefont{Carruthers and Hunt}(1990)}]{Carruthers90}
2032: \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Carruthers}}
2033:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~C.~R.}
2034:   \bibnamefont{Hunt}}, in \emph{\bibinfo{booktitle}{Atmospheric Processes over
2035:   Complex Terrain}}, edited by
2036:   \bibinfo{editor}{\bibfnamefont{W.}~\bibnamefont{Blumen}}
2037:   (\bibinfo{publisher}{Am. Meteorological Soc.}, \bibinfo{address}{Boston},
2038:   \bibinfo{year}{1990}), vol.~\bibinfo{volume}{23}.
2039: 
2040: \bibitem[{\citenamefont{Weng et~al.}(1991)\citenamefont{Weng, Hunt, Carruthers,
2041:   A., Wiggs, Livingstone, and Castro}}]{weng-etal:91}
2042: \bibinfo{author}{\bibfnamefont{W.~S.} \bibnamefont{Weng}},
2043:   \bibinfo{author}{\bibfnamefont{J.~C.~R.} \bibnamefont{Hunt}},
2044:   \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Carruthers}},
2045:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{A.}},
2046:   \bibinfo{author}{\bibfnamefont{G.~F.~S.} \bibnamefont{Wiggs}},
2047:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Livingstone}},
2048:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Castro}},
2049:   \bibinfo{journal}{Acta Mech. Suppl.} \textbf{\bibinfo{volume}{2}},
2050:   \bibinfo{pages}{1} (\bibinfo{year}{1991}).
2051: 
2052: \bibitem[{\citenamefont{Barndorff-Nielsen and
2053:   Willetts}(1991)}]{ActaMechanica91}
2054: \bibinfo{editor}{\bibfnamefont{O.~E.} \bibnamefont{Barndorff-Nielsen}}
2055:   \bibnamefont{and} \bibinfo{editor}{\bibfnamefont{B.~B.}
2056:   \bibnamefont{Willetts}}, eds., \emph{\bibinfo{title}{Aeolian Grain Transport
2057:   1 {\&} 2}}, Acta Mechanica Suppl. (\bibinfo{publisher}{Springer},
2058:   \bibinfo{address}{Wien, New York}, \bibinfo{year}{1991}).
2059: 
2060: \bibitem[{\citenamefont{Lettau and Lettau}(1978)}]{Lettau78}
2061: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Lettau}} \bibnamefont{and}
2062:   \bibinfo{author}{\bibfnamefont{H.~H.} \bibnamefont{Lettau}}, in
2063:   \emph{\bibinfo{booktitle}{Exploring the world's driest climate}}, edited by
2064:   \bibinfo{editor}{\bibfnamefont{H.~H.} \bibnamefont{Lettau}} \bibnamefont{and}
2065:   \bibinfo{editor}{\bibfnamefont{K.}~\bibnamefont{Lettau}}
2066:   (\bibinfo{publisher}{Madison}, \bibinfo{address}{Center for Climatic
2067:   Research, Univ. Wisconsin}, \bibinfo{year}{1978}).
2068: 
2069: \bibitem[{\citenamefont{S{\o}rensen}(1991)}]{Sorensen91}
2070: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{S{\o}rensen}},
2071:   \bibinfo{journal}{Acta Mechanica (Suppl.)} \textbf{\bibinfo{volume}{1}},
2072:   \bibinfo{pages}{67} (\bibinfo{year}{1991}).
2073: 
2074: \bibitem[{\citenamefont{Landau and Lifshitz}(1963)}]{landau-lifshitz:fm}
2075: \bibinfo{author}{\bibnamefont{Landau}} \bibnamefont{and}
2076:   \bibinfo{author}{\bibnamefont{Lifshitz}}, \emph{\bibinfo{title}{Fluid
2077:   Mechanics}}, vol.~\bibinfo{volume}{6} of \emph{\bibinfo{series}{Course of
2078:   Theoretical Physics}} (\bibinfo{publisher}{Pergamon Press},
2079:   \bibinfo{address}{London}, \bibinfo{year}{1963}).
2080: 
2081: \bibitem[{spe()}]{speed}
2082: \bibinfo{note}{This has to be distinguished from the statement that the
2083:   migration velocity of a dune with slip face is proportional to the flux over
2084:   the brink divided by the height, which is (trivially) true if losses over the
2085:   slip face are negligible, but useless unless the flux over the brink is
2086:   specified.}
2087: 
2088: \bibitem[{\citenamefont{Finkel}(1959)}]{Finkel59}
2089: \bibinfo{author}{\bibfnamefont{H.~J.} \bibnamefont{Finkel}},
2090:   \bibinfo{journal}{Journal of Geology} \textbf{\bibinfo{volume}{67}},
2091:   \bibinfo{pages}{614} (\bibinfo{year}{1959}).
2092: 
2093: \bibitem[{\citenamefont{Abramowitz and Stegun}(1965)}]{abramowitz-stegun}
2094: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Abramowitz}} \bibnamefont{and}
2095:   \bibinfo{author}{\bibfnamefont{I.~A.} \bibnamefont{Stegun}},
2096:   \emph{\bibinfo{title}{Handbook of Mathematical Functions}}
2097:   (\bibinfo{publisher}{Dover}, \bibinfo{address}{New York},
2098:   \bibinfo{year}{1965}).
2099: 
2100: \bibitem[{\citenamefont{Wippermann and Gross}(1986)}]{Wippermann86}
2101: \bibinfo{author}{\bibfnamefont{F.~K.} \bibnamefont{Wippermann}}
2102:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Gross}},
2103:   \bibinfo{journal}{Boundary Layer Meteorology} \textbf{\bibinfo{volume}{36}},
2104:   \bibinfo{pages}{319} (\bibinfo{year}{1986}).
2105: 
2106: \bibitem[{\citenamefont{van Dijk et~al.}(1999)\citenamefont{van Dijk, Arens,
2107:   and van Boxel}}]{van_dijk-arens-boxel:99}
2108: \bibinfo{author}{\bibfnamefont{P.~M.} \bibnamefont{van Dijk}},
2109:   \bibinfo{author}{\bibfnamefont{S.~M.} \bibnamefont{Arens}}, \bibnamefont{and}
2110:   \bibinfo{author}{\bibfnamefont{J.~H.} \bibnamefont{van Boxel}},
2111:   \bibinfo{journal}{Earth Surf. Process. Landforms}
2112:   \textbf{\bibinfo{volume}{24}}, \bibinfo{pages}{319} (\bibinfo{year}{1999}).
2113: 
2114: \bibitem[{\citenamefont{Stam}(1997)}]{Stam97}
2115: \bibinfo{author}{\bibfnamefont{J.~M.~T.} \bibnamefont{Stam}},
2116:   \bibinfo{journal}{Sedimentology} \textbf{\bibinfo{volume}{44}},
2117:   \bibinfo{pages}{127} (\bibinfo{year}{1997}).
2118: 
2119: \bibitem[{bad()}]{badbubble}
2120: \bibinfo{note}{An inapropriate boundary condition appears to be the main reason
2121:   for the failure of the original attempts to model flow separation in this way
2122:   \cite{zeman-jensen:88}.}
2123: 
2124: \bibitem[{rai()}]{rainout}
2125: \bibinfo{note}{For most applications it is a matter of convenience how exactly
2126:   the shear stress is set to zero inside the separation zone. For example, one
2127:   may choose a form that mimics the theoretically expected
2128:   \cite{sauermann-kroy-herrmann:2001} deposition of sand on the slip face
2129:   without further adaptation of the model. In $2d$ one may also treat the brink
2130:   as a boundary for the flux equation (see Section~\ref{sec:slipface}).}
2131: 
2132: \bibitem[{\citenamefont{{Fluent Inc.}}(1999)}]{Fluent5}
2133: \bibinfo{author}{\bibnamefont{{Fluent Inc.}}}, \emph{\bibinfo{title}{{FLUENT
2134:   5}}} (\bibinfo{year}{1999}), \bibinfo{note}{finite volume solver}.
2135: 
2136: \bibitem[{\citenamefont{Sauermann}(2001)}]{sauermann:phd}
2137: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Sauermann}}, Ph.D. thesis,
2138:   \bibinfo{school}{{U}niversit{\"a}t {S}tuttgart} (\bibinfo{year}{2001}).
2139: 
2140: \bibitem[{\citenamefont{Wesseling}(2000)}]{wesseling:2000}
2141: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Wesseling}},
2142:   \emph{\bibinfo{title}{Principles of Computational Fluid Dynamics}}
2143:   (\bibinfo{publisher}{Springer}, \bibinfo{address}{Berlin, Heidelberg, New
2144:   York}, \bibinfo{year}{2000}).
2145: 
2146: \bibitem[{\citenamefont{Sauermann et~al.}(2000)\citenamefont{Sauermann, Rognon,
2147:   Poliakov, and Herrmann}}]{sauermann-etal:2000}
2148: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Sauermann}},
2149:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Rognon}},
2150:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Poliakov}}, \bibnamefont{and}
2151:   \bibinfo{author}{\bibfnamefont{H.~J.} \bibnamefont{Herrmann}},
2152:   \bibinfo{journal}{Geomorphology} \textbf{\bibinfo{volume}{36}},
2153:   \bibinfo{pages}{47} (\bibinfo{year}{2000}).
2154: 
2155: \bibitem[{\citenamefont{Wiggs et~al.}(1996)\citenamefont{Wiggs, Livingstone,
2156:   and Warren}}]{Wiggs96}
2157: \bibinfo{author}{\bibfnamefont{G.~F.~S.} \bibnamefont{Wiggs}},
2158:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Livingstone}},
2159:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Warren}},
2160:   \bibinfo{journal}{Geomorphology} \textbf{\bibinfo{volume}{17}},
2161:   \bibinfo{pages}{29} (\bibinfo{year}{1996}).
2162: 
2163: \bibitem[{\citenamefont{Anderson}(1991)}]{Anderson91}
2164: \bibinfo{author}{\bibfnamefont{R.~S.} \bibnamefont{Anderson}},
2165:   \bibinfo{journal}{Acta Mechanica (Suppl.)} \textbf{\bibinfo{volume}{1}},
2166:   \bibinfo{pages}{21} (\bibinfo{year}{1991}).
2167: 
2168: \bibitem[{\citenamefont{Nalpanis et~al.}(1993)\citenamefont{Nalpanis, Hunt, and
2169:   Barrett}}]{Nalpanis93}
2170: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Nalpanis}},
2171:   \bibinfo{author}{\bibfnamefont{J.~C.~R.} \bibnamefont{Hunt}},
2172:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{C.~F.}
2173:   \bibnamefont{Barrett}}, \bibinfo{journal}{J.~Fluid Mech.}
2174:   \textbf{\bibinfo{volume}{251}}, \bibinfo{pages}{661} (\bibinfo{year}{1993}).
2175: 
2176: \bibitem[{\citenamefont{Rioual et~al.}(2000)\citenamefont{Rioual, Valance, and
2177:   Bideau}}]{Rioual2000}
2178: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Rioual}},
2179:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Valance}}, \bibnamefont{and}
2180:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Bideau}},
2181:   \bibinfo{journal}{Phys.~Rev.~E} \textbf{\bibinfo{volume}{62}},
2182:   \bibinfo{pages}{2450} (\bibinfo{year}{2000}).
2183: 
2184: \bibitem[{\citenamefont{Owen}(1964)}]{Owen64}
2185: \bibinfo{author}{\bibfnamefont{P.~R.} \bibnamefont{Owen}}, \bibinfo{journal}{J.
2186:   Fluid. Mech.} \textbf{\bibinfo{volume}{20}}, \bibinfo{pages}{225}
2187:   (\bibinfo{year}{1964}).
2188: 
2189: \bibitem[{\citenamefont{{J. P. Bouchaud and M. E. Cates and J. Ravi Prakash and
2190:   S. F. Edwards}}(1994)}]{Bouchaud94}
2191: \bibinfo{author}{\bibnamefont{{J. P. Bouchaud and M. E. Cates and J. Ravi
2192:   Prakash and S. F. Edwards}}}, \bibinfo{journal}{J. Phys. France I}
2193:   \textbf{\bibinfo{volume}{4}}, \bibinfo{pages}{1383} (\bibinfo{year}{1994}).
2194: 
2195: \bibitem[{\citenamefont{Andreotti and Claudin}()}]{andreotti-claudin:2002}
2196: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Andreotti}} \bibnamefont{and}
2197:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Claudin}},
2198:   \emph{\bibinfo{title}{Selection of dune shapes and velocities. part 2: A
2199:   two-dimensional modelling}}, \eprint{cond-mat/0201105}.
2200: 
2201: \bibitem[{\citenamefont{Lima et~al.}()\citenamefont{Lima, Sauermann, Herrmann,
2202:   and Kroy}}]{lima-etal:2002}
2203: \bibinfo{author}{\bibfnamefont{A.~R.} \bibnamefont{Lima}},
2204:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Sauermann}},
2205:   \bibinfo{author}{\bibfnamefont{H.~J.} \bibnamefont{Herrmann}},
2206:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Kroy}},
2207:   \emph{\bibinfo{title}{Modelling a dune field}}, \eprint{cond-mat/0202528}.
2208: 
2209: \end{thebibliography}
2210: 
2211: 
2212: \end{document}
2213: