0707.2968/ms.tex
1: \documentclass[floatfix]{emulateapj}
2: %\documentclass[12pt,preprint]{aastex}
3: %\usepackage{graphicx,amsmath,epsfig,amssymb,natbib}
4: \usepackage{epsfig}
5: 
6: \citestyle{aa}
7: \begin{document}
8: 
9: \title{Structure and evolution of Zel'dovich pancakes as 
10: 				probes of dark energy models}
11: 
12: \author{P.~M.~Sutter} \email{psutter2@uiuc.edu}
13: \affil{Department of Physics,
14: 	     University of Illinois at Urbana-Champaign,
15:             Urbana, IL 61801-3080}
16: 
17: \and
18: 
19: \author{P.~M.~Ricker} \email{pmricker@uiuc.edu}
20: \affil{Department of Astronomy,
21:        University of Illinois at Urbana-Champaign,
22:              Urbana, IL 61801\\
23: 		National Center for Supercomputing Applications,
24:       University of Illinois at Urbana-Champaign,
25:             Urbana, IL 61801}
26: 
27: \begin{abstract}  
28: We examine how coupled dark matter and dark energy modify the 
29: development of Zel'dovich pancakes.
30: We study how the various effects of these theories, 
31: such as a fifth force in the dark sector and a modified 
32: particle Hubble drag, produce variations in the 
33: redshifts of caustic formation and the present-day 
34: density profiles of pancakes.
35: We compare our results in direct simulation to 
36: a perturbation theory approach for the dark energy 
37: scalar field.
38: We determine the range of initial 
39: scalar field amplitudes 
40: for which perturbation theory is accurate 
41: in describing the development of the pancakes.
42: Notably, we find that perturbative methods which neglect 
43: kinetic terms in the scalar field equation of motion 
44: are not valid for arbitrarily small perturbations. 
45: We also examine whether models that 
46: have been tuned to match the constraints of 
47: current observations can produce new observable effects
48: in the nonlinear structure of pancakes.
49: Our results suggest that a fully realistic three-dimensional 
50: simulation will produce significant new observable features, 
51: such as modifications to the mass function and halo 
52: radial density profile shapes, that can be used to 
53: distinguish these models from standard concordance cosmology and from each other.
54: \end{abstract}
55: 
56: \keywords{cosmology:theory, dark matter, dark energy, structure formation, methods: N-body simulations}
57: \maketitle
58: 
59: \section{Introduction}
60: \label{sec:introduction}
61: Dark energy is perhaps the most profound and essential 
62: mystery in modern cosmology. 
63: While the $\Lambda$CDM 
64: cosmological model has proven very successful in 
65: explaining and predicting many features of our universe, 
66: such as the 
67: fluctuations in the cosmic microwave background~\citep[eg.][]{deBernardis}, 
68: the large-scale matter distribution~\citep[eg.][]{Percival}, 
69: and 
70: distance measurements to type Ia supernovae~\citep{Perlmutter,Riess}, the physics 
71: of the dark sector (dark matter and dark energy), which comprises roughly ninety-six 
72: percent 
73: of the energy density of the universe, is 
74: largely unknown. 
75: 
76:  Currently, there are too few 
77: observational constraints to determine the precise 
78: nature of the dark energy 
79: and any possible interactions it might have with dark and baryonic 
80: matter~\citep{Bean}.
81: However, we can use 
82: simulations of nonlinear structure 
83: formation to explore the consequences of plausible
84: dark energy models, 
85:  including those that propose 
86: couplings between dark matter (DM) and dark energy (DE)~
87: \citep[see][and references therein]{Alcaniz}. 
88: Models that have particle physics motivations 
89: often predict such couplings~\citep{Amendola2}.
90: Such theories are 
91: intriguing because they might provide a resolution to current 
92: cosmological problems, 
93: such as the so-called coincidence problem~\citep{Zimdahl,Amendola3}, 
94: and the observed emptiness of the voids~\citep{Farrar}, 
95: the latter of which was confirmed numerically by~\cite*{Nusser}.
96: 
97: An important goal of coupled DM-DE simulations 
98: is to identify observational methods to test these theories. 
99: The observables studied to date include 
100: luminosity distances~\citep{Amendola}, 
101: the growth of matter perturbations
102: ~\citep{Olivares,Koivisto}, the abundance of clusters~\citep{Manera}, 
103: and the Sandage-Loeb test~\citep{Corasaniti}.
104: Not only can we use such simulations to search for additional 
105: observable features, 
106: but we can also use them to find ways to distinguish models and determine 
107: the validity of perturbation methods. 
108: 
109: Much of the work to date has followed the framework 
110: established by~\cite*{Farrar}.
111: This model uses a 
112: dynamical scalar field to provide the dark energy, and it allows that field 
113: to couple to dark matter via a Yukawa interaction.
114: Although there are some 
115: issues with models of this type~\citep{Doran}, 
116: they provide useful foils 
117: for studying the possible effects that other 
118: models, which are complicated but more robust, 
119: might predict.
120: These models would include the model discussed 
121: by~\cite*{Huey} and the two-family model considered in Farrar and Peebles.  
122: 
123: Some of the previous work has examined structure formation 
124: with modifications due to a fifth force, albeit at low 
125: spatial and force resolution~\citep{Nusser,Maccio}. 
126: However, we believe 
127: that an incremental approach that analyzes the various 
128: effects of DM-DE interactions on simpler initial conditions is crucial 
129: before tackling more realistic initial conditions. This approach 
130: allows the effects observed in more realistic simulations 
131: to be understood and generalized.
132: Therefore, in this paper, we will study the effects of 
133: coupling between dark matter and dark energy 
134: on the development of Zel'dovich pancakes. Zel'dovich pancakes are 
135: the well-known solutions to the problem of 
136: the gravitational collapse of
137: one-dimensional, sinusoidal, plane-wave, 
138: adiabatic density perturbations~\citep{Zel}. Since they
139: are well-studied in a variety of contexts 
140: not involving DM-DE interactions 
141: \citep[for example,][and others]{More,Gnedin,Yuan,Valinia,Anninos1,Anninos2}, 
142: we can more easily understand the effects of 
143: additional physics on structure formation, laying the 
144: groundwork for a more complete three-dimensional study.
145: 
146: In Section~\ref{sec:model} we discuss
147: the relevant equations, the effects we will study, 
148: and our numerical techniques.
149: Section~\ref{sec:effects} 
150: discusses the role that the exotic physics 
151: introduced by the Farrar and Peebles model, such as a DM particle fifth force and 
152: time-dependent DM particle mass, 
153: plays in structure formation.
154: We are careful to separate the various effects predicted by the 
155: model. Even if this particular example does not withstand 
156: closer scrutiny, unrelated theories may predict 
157: similar effects.
158: 
159: In addition, so far many researchers have exploited perturbation theory to 
160: treat fluctuations in the DE scalar field. However, we do not know 
161: \emph{a priori} the validity of this approach, especially as we begin to 
162: explore the nonlinear consequences of this theory. 
163: In Section~\ref{sec:validity}
164: we will compare the structure formation results 
165: from perturbation theory alone  
166: to the results from doing a complete nonlinear analysis. 
167: We will also discuss 
168: the regime where perturbation theory is most valid in 
169: accurately predicting structure.
170: 
171: Many of the theories of this type have adjustable parameters, and 
172: these parameters must be adjusted to match $\Lambda$CDM predictions, at 
173: the risk of being indistinguishable from it. There are also 
174: usually several unique combinations of parameters that provide similar, 
175: if not identical, results.
176: In Section~\ref{sec:distinguish} we attempt to find ways 
177: to distinguish models that remain 
178: indistinct in perturbation theory. 
179: Also, we will determine if nonlinear effects can distinguish 
180: models from standard cosmology even when effects based on 
181: perturbative methods cannot.
182: 
183:   
184: 
185: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
186: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
187: \section{The Model}
188: \label{sec:model}
189: \subsection{Coupled Dark Matter and Dark Energy}
190: Following Farrar and Peebles, we will consider 
191: a dark energy (DE) scalar field $\phi$ with action
192: \begin{equation}
193: \label{eq:DEaction}
194: 	S_{DE} = \int d^4 x \sqrt{-g} \left[ \frac{1}{2} 
195: 		\phi_{,\nu} \phi ^ {,\nu} - V(\phi) \right],
196: \end{equation}
197: and a single classical nonrelativistic dark matter (DM) 
198: particle family with action for the $i$th particle
199: \begin{equation}
200: \label{eq:DMaction}
201: 	{S_{DM}}_i = - \int y \left| 
202: 		\phi\left( x_i \right) - \phi_\ast \right| 
203: 		\sqrt{g_{\mu \nu} dx^{\mu}_{i} dx^{\nu}_{i}},
204: \end{equation}
205: where $y$ is the dimensionless Yukawa interaction 
206: strength and $\phi_\ast$ is a constant providing an 
207: intrinsic mass to the DM 
208: particle. We will set $\phi_\ast = 0$, so that 
209: the DM particle mass is due entirely to the field value.
210: In both of the above actions and throughout, we have set $\hbar = c = 1$, and 
211: we will ignore baryons. 
212: 
213: Assuming a spatially flat Friedmann universe, we can 
214: use equation~(\ref{eq:DMaction}) to
215: obtain a comoving DM particle 
216: equation of motion in one dimension:
217: \begin{equation}
218: \label{eq:dmEoM}
219: \dot v  + \left( 2 \frac{\dot{a}}{a} + \frac{\dot{\phi}}{\phi} \right)
220:  v = - \frac{\partial \Phi}{\partial x} - \frac{1}{a^2} \frac{1}{\phi} \frac{\partial \phi}{\partial x}  .
221: \end{equation}
222: Here, $\Phi$ is the normal comoving gravitational potential, 
223: $a$ is the scale factor, and $v$ is 
224: the comoving particle velocity, 
225: and $x$ is the comoving position. Throughout, dots 
226: refer to derivatives with respect to proper time $t$. 
227: The field action in equation (\ref{eq:DEaction}) gives the evolution of $\phi$:
228: \begin{equation}
229: \label{eq:phiEoM}
230: 	\ddot{\phi} - \frac{1}{a^2} \nabla^2 \phi + 3 \frac{\dot{a}}{a} \dot{\phi} 
231: 	  + \frac{d V}{d \phi} + \frac{\rho}{\phi}a^{-3}= 0,
232: \end{equation}
233: where $\rho$ is the DM particle comoving density. 
234: 
235: The comoving potential satisfies the Poisson equation:
236: \begin{equation}
237: \label{eq:poisson}
238:  \nabla^2 \Phi = \frac{4 \pi G}{a^3} \left( \rho - \overline{\rho} \right).
239: \end{equation}
240: Here and throughout, an overline indicates a spatial average.
241: 
242: The Friedmann equation, neglecting radiation, curvature, and 
243: baryonic terms, is now
244: \begin{equation}
245: \label{eq:friedmann}
246: 	\left( \frac{\dot{a}}{a} \right)^2 = 
247: 	H_0^2 \Omega_m \frac{\overline{\phi}}{\overline{\phi}_0} a^{-3} + 
248: 	\frac{8 \pi G}{3} \left[ \frac{1}{2} \left( \frac{d \overline{\phi}}{d t} 
249: 	\right)^2 + V(\overline{\phi}) \right].
250: \end{equation}
251: A subscript of $0$ denotes the present-day value.
252: The first term on the right-hand side reflects the contribution 
253: of the DM with its time-dependent mass. 
254: The terms in brackets are, respectively, the kinetic 
255: and potential energies of the DE scalar field.
256: 
257: We do not know the 
258: initial conditions of the field, but we do know that today the field 
259: behaves as a cosmological constant, so the potential term dominates and 
260: has a value 
261: \begin{equation}
262: 	V(\overline{\phi}_0) = \Omega_\Lambda \rho_{crit}. 
263: \label{eq:potentialToday}
264: \end{equation}
265: Also, at early 
266: enough times, Farrar and Peebles found that 
267: equation (\ref{eq:phiEoM}) reduces to
268: \begin{equation}
269: \label{eq:initialPhiDot}
270: 	\frac{d \phi} {d t} = - \frac{H_0^2}{G} 
271: 	\frac{3 \Omega_m }{8 \pi \phi_0} \frac{1}{a^3} t,
272: \end{equation}
273: which we use to set the initial condition for $\dot \phi$. 
274: 
275: We can identify four unique ways in which the 
276: extra interactions modify structure formation.
277: First, the DM particle mass directly depends on 
278: the field value, so we may write 
279: the ratio of the modified mass to its present-day value as
280: \begin{equation}
281: \label{eq:mass}
282:  \eta \equiv \frac{m_{DM}}{m_{DM,0}} = \frac{\phi}{\phi_0}
283: \end{equation}
284: Secondly, the interactions modify the Hubble drag found 
285: in the DM equation of motion, so that its 
286: ratio to the drag in standard cosmology is
287: \begin{equation}
288: \label{eq:drag}
289:  \gamma \equiv \frac{2 \dot{a}/a + \dot{\phi}/ \phi}{2\dot{a}/a}.
290: \end{equation}
291: Next, we notice a modified particle acceleration due 
292: to a fifth force in equation (\ref{eq:dmEoM}). 
293: We will define 
294: \begin{equation}
295: \label{eq:force}
296:  \beta \equiv \frac{\frac{\partial \Phi}{\partial x}  + 
297:  								\frac{1}{a^2} \frac{1}{\phi} \frac{\partial \phi}{\partial x} }
298:  									{\frac{\partial \Phi}{\partial x} }.
299: \end{equation}
300: Finally, the dynamic scalar field itself 
301: indirectly affects structure formation 
302: via a time-varying $\Omega_\Lambda$ in the Friedmann equation. We define
303: \begin{equation}
304: \label{eq:field}
305:  \delta \equiv \frac{
306:     \frac{1}{2} \left( \frac{d \overline{\phi}}{d t} 
307: 	\right)^2 + V(\overline{\phi})
308:  }{V \left( \overline{\phi}_0\right)}.
309: \end{equation}
310: 
311: If the fluctuations in the field are small enough, we may do 
312: perturbation theory. Farrar and Peebles found that in this regime, 
313: we may replace $\phi(x)$ with 
314: a single spatial average, $\phi_b$ and a 
315: sufficiently small perturbation field $\phi_1(x)$. 
316: We may then write equation~(\ref{eq:phiEoM}) as
317: \begin{equation}
318: \label{eq:phibEoM}
319: 	\ddot{\phi_b} + 3 \frac{\dot{a}}{a} \dot{\phi_b} 
320: 	  + \frac{d V}{d \phi_b} + \frac{3 \Omega_m H_0^2}
321: 		{8 \pi G \phi_{b,0}}a^{-3}= 0.
322: \end{equation}
323: Also, the fifth force ratio appears instead as
324: \begin{equation}
325: \label{eq:fifthPert}
326: 	\beta_{pert} \equiv 1 + \frac{1}{4 \pi G \phi_b^2}.
327: \end{equation}
328: 
329: We will contrast our results with standard concordance cosmology, which we 
330: achieve by setting $\phi$ to a single value satisfying 
331: equation (\ref{eq:potentialToday}) and by  
332: preventing $\phi$ from evolving dynamically.
333: 
334: We have the freedom to choose an appropriate potential $V(\phi)$. 
335: Although there are many potentials in the literature, such as the 
336: exponential~\citep{RatraAndPeebles}, power-law~\citep{PeeblesAndRatra}, 
337: and power-law and sine~\citep{Dodelson},
338: we will adopt the power-law potential found in Farrar and Peebles:
339: \begin{equation}
340: \label{eq:powerLaw}
341: 	V(\phi) = K / {\phi}^{\alpha},
342: \end{equation}
343: where we are also free to choose the constants $K$ and $\alpha$.
344: Potentials like this lead to reasonable behavior, i.e. 
345: potential-dominated solutions at the present epoch. 
346: Again, we chose this potential merely as an example.
347: 
348: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
349: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
350: \subsection{Numerical Techniques}
351: For our simulations we developed a one-dimensional 
352: $N$-body particle-mesh code. 
353: We used cloud-in-cell mapping for interpolating between 
354: the mesh and particles~\citep{Hockney}, and 
355: a second order leapfrog integration scheme for particle advancement. 
356: In one dimension and with finite differencing, 
357: we can solve the Poisson equation exactly using a 
358: Thomas algorithm~\citep{Thomas}, modified for 
359: periodic boundaries via the Sherman-Morrison formula~\citep{Sherman}.
360: We discuss the details of our scheme for solving the scalar field 
361: and scale factor in the appendix.
362: 
363: For all calculations, 
364: we used $\Omega_m = 0.26$, $\Omega_\Lambda=0.74$, 
365: and $H_0 = 100 \mbox{ }h = 71 \mbox{ km s}^{-1} 
366: \mbox{ Mpc}^{-1}$. All runs took place in a 
367: one-dimensional box of length $10 \mbox{ } h^{-1} \mbox{ Mpc}$ per
368: side, with $65,536$ particles to represent 
369: the dark matter and $8,192$ zones 
370: for the Poisson solver and the scalar field.
371: Since we used finite-differencing when forming the gradient 
372: to construct the fifth force, we required a 
373: large particle-zone ratio to dampen noise 
374: in the density field, which couples to the scalar field.
375: 
376: All simulations used the same initial conditions. 
377: We distributed particles evenly throughout the grid (described by 
378: position $q_i \equiv i \Delta x$) and perturbed them 
379: using the Zel'dovich approximation:
380: \begin{eqnarray}
381: 	x & = & q + \frac{2}{5} a A \sin{(k q)} \\
382: 	v & = & \frac{2}{5} \dot{a} A \sin{(k q)},
383: \end{eqnarray}
384: where $k=2\pi / \lambda$ is the comoving wavenumber 
385: of the perturbation.
386: The amplitude $A$ can be written in terms of the redshift 
387: $z_c$ of the formation 
388: of the first caustic: 
389: \begin{equation}
390: 	A = -5(1+z_c)/(2k).
391: \end{equation}
392: We chose the initial perturbation amplitude 
393: such that the first caustic would 
394: form at $z_c=5.0$ for an $\Omega_m = 1.0$ universe, and 
395: we chose the comoving wavelength 
396: of perturbations to be $\lambda = 10 \mbox{ }h^{-1} \mbox{ Mpc}$.
397: These choices are arbitrary, but commonly used in the literature. 
398: All computations start 
399: at a redshift of $z=50$ with DM particle masses 
400: determined by equation~(\ref{eq:mass}).
401: 
402: We do not know the initial conditions of $\phi$ in advance, 
403: so we must guess initial conditions 
404: and iterate until we meet the condition $\dot{a} = H_0$ 
405: at the present epoch.
406: Based on the comments made by Farrar and Peebles, we chose four 
407: combinations of the potential parameters $K$ and $\alpha$ 
408: that yield reasonable behavior. 
409: Table~\ref{tab:FandP} lists the parameters, the guessed 
410: initial field value at $z=50$, and the field value today as calculated 
411: from equation~(\ref{eq:phiEoM}). We performed these calculations in 
412: perturbation theory. 
413: We selected these parameters for behaviors that were 
414: consistent with current observations, but provided 
415: unique evolutions of the scalar field for 
416: comparison.
417:  	         
418: \begin{table}
419:   \centering
420: 	\begin{tabular}{|c|c|c|c|c|} 
421: 		\hline
422: 	Label & $ \alpha $ & $K (G^{1+\alpha/2} / H_0^2)$ 
423: 	      & $\phi_{\mbox{init}} (G^{1/2})$ & $\phi_0 (G^{1/2})$\\
424: 	\hline
425: 	A & $-2$ & $0.03$   & $1.89$   & $1.72$  \\
426: 	B & $-2$ & $0.0057$ & $4.02$   & $3.94$  \\
427: 	C & $6$  & $2.0$    & $1.80$  & $1.68$  \\
428: 	D & $6$  & $280.0$  & $3.89$   & $3.83$  \\
429: 	\hline
430: 	\end{tabular}
431: 	\caption{Simulation parameter choices.}
432: 	\label{tab:FandP}
433: \end{table}
434:  
435: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
436: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
437: \section{Analysis of Effects in Perturbation Theory}
438: \label{sec:effects}
439: We now illustrate the effects of these 
440: additional interactions on structure formation. 
441: To highlight each individual consequence of the DM-DE interaction, 
442: we separately analyze the resulting structure when including the 
443: effects of the variable mass, 
444: the modified Hubble drag, the fifth force, and the 
445: dynamical field.
446: Finally, we will examine the final structure when 
447: these effects are combined.
448: In all that follows, we highlight the behavior 
449: when using the potential 
450: parameters $K = 0.03$ and $\alpha = -2$ 
451: (parameter set $A$) as an example.
452: 
453: Figure~\ref{fig:compareEffects} shows the 
454: evolution of $\eta$, $\gamma$, $\beta_{pert}$, 
455: and $\delta$ as
456:  functions of scale factor for parameter set $A$. 
457: On average, the fifth force is 
458: about three percent of the gravitational force. 
459: We expect the fifth force to play a significant role 
460: throughout cosmic history. 
461: The Hubble drag is reduced by about one percent. The particle 
462: mass starts at ten percent greater than its present-day value, 
463: but by a scale factor of $0.5$ it is only 
464: two percent larger. We expect the modified mass to contribute to 
465: structure mostly in the early universe, 
466: but by the present day its effects will not be significant. 
467: In the early universe, the dynamic 
468: field gives a large $\Omega_\Lambda$, and hence a 
469: more rapid expansion, which should delay 
470: caustic formation relative to standard cosmology. 
471: Notice that our calculations do not give 
472: precisely the required values of $\Omega_\Lambda$, 
473: but are less than two percent off. 
474: We fixed our field at late times 
475: such that the energy density of the field was purely 
476: in the potential, when in reality 
477: the kinetic term, while very small, 
478: is not entirely negligible. Fortunately, this small 
479: error does not in any way significantly alter structure formation.
480:   
481: \begin{figure}
482: 	\plotone{f1.eps}
483: 	%\epsfig{figure=figCompareEffects.eps,width=\columnwidth}
484: 	\caption{Comparison of additional effects due to the 
485: 		DM-DE interactions. Shown are relative values defined 
486: 	        above as $\eta$ (solid line), $\gamma$ (line with hatches), 
487: 		$\beta_{pert}$ (line with stars), 
488: 	        and $\delta$ (line with boxes). 
489: 	        The symbols are only to aid in distinguishing the lines.}
490: \label{fig:compareEffects}
491: \end{figure}
492:  
493: We show the density profile at $z=0$ in 
494: Figure~\ref{fig:compareDensityProfile}, modified 
495: by the individual additional effects of the DM-DE interaction.
496: We do not show the result from an evolving mass and 
497: dynamical field, because these profiles are 
498: hard to distinguish from standard cosmology. 
499: %Figure~\ref{fig:compareCentralDensity} displays 
500: %the evolution of the density at the center of the pancake,
501: %showing the redshifts of caustic formation. 
502: When isolating the individual effects 
503: of the drag, fifth force, and dynamic field, 
504: we forced the scalar field to evolve as if the scale factor 
505: obeyed the modified 
506: Friedmann equation in equation~(\ref{eq:friedmann}); however, the actual scale 
507: factor evolved according to concordance cosmology in these comparisons. 
508: We did this because without the dynamic mass in 
509: the Friedmann equation, the 
510: scalar field would not reach its required value today.
511:  
512: \begin{figure}
513: 	\plotone{f2.eps}
514: 	%\epsfig{figure=figCompareDensityProfile.eps,width=\columnwidth}
515: 	\caption{Dark matter density profile at $z=0$. 
516: 		The solid line in all plots is the standard 
517: 		cosmology. The dotted line indicates the 
518: 		effects of the additional interactions: 
519: 		the top panel shows only the result with 
520: 		the drag term included, and the bottom panel
521: 		shows only the fifth force included.
522: 		%modified mass, the second panel shows only the drag term, 
523: 		%the third shows only the fifth force, 
524: 		%and the bottom shows only the dynamic field.
525: 		Here, $x$ is the distance from the midplane 
526: 		of each pancake.
527: 		The numerical shot noise in the density near $x=1$ 
528: 		is negligible and unimportant.}
529: \label{fig:compareDensityProfile}
530: \end{figure}
531:  
532: %\begin{figure}
533: %	\epsfig{figure=figCompareCentralDensity.eps,width=\columnwidth}
534: %	\caption{Dark matter density at the center of the pancake 
535: %		as a function of redshift. The panels are 
536: %		arranged the same as in 
537: %		Figure~\ref{fig:compareDensityProfile}.
538: %		Note that the peak heights are not realistic.}
539: %\label{fig:compareCentralDensity}
540: %\end{figure}
541: 
542: We can easily understand the modifications due to the 
543: fifth force: an additional attractive 
544: force causes the caustics to form earlier, 
545: and hence each peak in the density profile at $z=0$ 
546: is farther away from the center. The effect on each peak is similar:
547: a fifth force only in the early universe will 
548: affect mostly the first two caustics, 
549: but since the force ratio remains roughly constant 
550: even in late times, the latest peak also shifts.
551: The overall size of the pancake is then larger. 
552: 
553: Due to the modified drag term alone, the first two caustics form earlier, 
554: but the latest caustic forms at the same time as in 
555: standard cosmology.
556: This requires more explanation.
557: The Hubble drag term is most important in the early universe. 
558: Due to the reduced drag, particles 
559: move faster, and reach a greater turnaround radius. 
560: So the first two caustics form earlier for the 
561: modified cosmology. However, at later times, 
562: the drag is no longer significant, and 
563: the third caustic forms at nearly the same 
564: redshift as in standard cosmology.
565: Also, it appears that modifications to the drag 
566: are very important: even small deviations in the 
567: redshift of caustic formation lead to significant 
568: difference in the final peak locations.
569: 
570: The increased mass creates tension between 
571: two opposing influences on structure formation. 
572: Since the particles already in halos are more massive, 
573: they create a deeper potential 
574: well and encourage other particles to fall in faster, 
575: thereby causing caustics to form 
576: earlier. On the other hand, a larger $\Omega_m$ in the 
577: Friedmann equation accelerates 
578: the expansion in the early universe, which increases 
579: the Hubble drag and dampens 
580: structure formation. 
581: We found that the expansion effect dominates, but 
582: not significantly, and caustics form only slightly 
583: later than in standard cosmology.
584: %As Figure~\ref{fig:compareCentralDensity} 
585: %shows, the expansion effects dominates 
586: %the redshift of caustic formation. 
587: Also, since the DM particle is more massive in the past, 
588: the regions near early-forming caustics have a higher density, 
589: but by $z=0$ the mass is the same as in standard cosmology, so the overall 
590: amplitude of the density profile does not change. 
591: 
592: We must contrast this evolving-mass 
593: universe with a $\Lambda$CDM cosmology having simply a larger $\Omega_m$. 
594: In such a universe, the competing 
595: effects of increased mass roughly cancel each other out, 
596: and caustics form at 
597: similar redshifts to those in standard $\Lambda$CDM. 
598: In an evolving-mass universe, 
599: however, as the mass decreases, 
600: the halo potential wells 
601: get shallower, but the universe has already expanded 
602: more, so outlying 
603: particles take longer to reach the halo, 
604: and caustics correspondingly form later. 
605: %This also causes the first caustic, which 
606: %forms later than in standard cosmology, to reach a
607: %slightly greater final position. 
608: At lower redshifts, 
609: the particle masses are nearly their present-day 
610: values, so this effect is not as significant, and the final 
611: peak locations are slightly closer to the pancake midplane.
612: 
613: Finally, the dynamic field has the expected result: 
614: more rapid early-universe expansion 
615: pulls particles away from each other, and caustics take longer to form.
616: As expected, this effect is nearly negligible 
617: in the late universe, as the field 
618: approaches its present-day value. 
619: Unlike the differences between an evolving-mass 
620: universe and a constant-mass universe, 
621: the structure modifications from a dynamic 
622: field are, in this case, nearly indistinguishable from a universe with 
623: a roughly ten percent larger, but still constant, $\Omega_\Lambda$.
624: 
625: We can also examine the effects on particle 
626: velocities due to these additional interactions. 
627: Figure~\ref{fig:comparePhase} is a DM particle phase plot at $z=0$ with 
628: modifications due to the reduced Hubble drag and the fifth force, 
629: compared to standard cosmology. 
630: The dynamic field and variable mass terms have no effect, but
631: as expected, the reduced drag and fifth force increase particle velocities.
632:  
633: \begin{figure}
634: 	\plotone{f3.eps}
635: 	%\epsfig{figure=figComparePhase.eps,width=\columnwidth}
636: 	\caption{Dark matter particle phase plots at $z=0$.
637: 		The solid line in all plots is from concordance cosmology, 
638: 		and the dotted lines with open squares 
639: 		are the results from including the 
640: 		effects indicated in each plot. 
641: 		The open squares only represent select data points, to aid 
642: 		in distinguishing the lines.
643: 		We have omitted the plots 
644: 		with dynamic field and dynamic mass effects, since these 
645: 		have no discernible influence.}
646: \label{fig:comparePhase}
647: \end{figure}
648:  
649: When we combine results, in Figure~\ref{fig:compareStandardPertFull}, 
650: we see that the effects of the modified mass 
651: and the dynamic field tend to cancel out the fifth force. 
652: So, the final density profile most closely resembles the 
653: results from the drag term alone. 
654: The innermost caustic is the same distance from the midplane
655: of the pancake, 
656: but the outermost and middle peaks are roughly 
657: $15\%$ farther away.
658: The slope of the density profile 
659: remains undisturbed and the total number of peaks formed by the 
660: present epoch are the same.
661:  
662: \begin{figure}
663: 	\plotone{f4.eps}
664: 	%\epsfig{figure=figCompareStandardPertFull.eps,width=\columnwidth}
665: 	\caption{Dark matter density profile at $z=0$.
666: 		The solid line is from standard cosmology. The dashed line is 
667: 		the density from using perturbation theory, and the dotted line 
668: 		is from using full theory. 
669: 		This is including all effects.
670: 		Note that perturbation theory is nearly 
671: 		indistinct from the full theory.
672: 		The numerical shot noise in the density near $x=1$ 
673: 		is negligible and unimportant.}
674: \label{fig:compareStandardPertFull}
675: \end{figure} 
676:  
677: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
678: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
679: \section{The Validity of Perturbation Theory}
680: \label{sec:validity}
681: So far, we have only considered perturbation theory, 
682: relying on a single value to describe the DE scalar field at 
683: a specific time. When we relax this constraint,  
684: each particle feels its own force, has its own drag, and has 
685: an independent mass. When solving the Friedmann equation, 
686: we must take a spatial average of the field, rather 
687: than using a single background value. Also, the expression for 
688: the fifth force changes 
689: to be directly proportional to gradients in the scalar field.
690: 
691: There is some ambiguity when considering the initialization 
692: of the scalar field. While 
693: the average value must match the perturbation theory background 
694: value, we are left with 
695: few clues to the initial wavelength and amplitude of the 
696: perturbations. So, using adiabatic reasoning, 
697: we initialized $\phi$ to have 
698: the same spectrum as the density field, which in our case 
699: is a single perturbation mode with 
700: wavelength $\lambda = 10 h^{-1} \mbox{ Mpc}$ at $z=0$. 
701: However, we freely chose the amplitude, and 
702: we found that to make an attractive fifth force,
703: which is required for a Yukawa-type interaction, 
704: we needed $\phi$ to have a phase opposite to that of the density field.
705: 
706: We must take care in the initialization of both the background 
707: value and amplitude of perturbations of the field when 
708: considering the full theory. First, the addition of the 
709: gradient term in the equation of motion for the field will 
710: affect the field's background evolution, and may force us to modify 
711: the initial value, so that $\phi$ still satisfies the 
712: present-day condition 
713: in equation~(\ref{eq:potentialToday}).
714: Fortunately, the gradients are not large, and 
715: this term does not modify the evolution of $\phi$ greatly. 
716: Hence, we constructed the field 
717: so its average value matched the background perturbation value.
718: It appears that perturbation theory is perfectly appropriate for probing the linear 
719: results of this model, such as estimates of the 
720: Hubble time and the location of peaks in the CMB power spectrum, 
721: which Farrar and Peebles discuss. 
722: 
723: Secondly, by varying the initial amplitude, we can change the 
724: strength of the fifth force at high 
725: redshift.~\cite*{Kesden,Kesden2} have shown that a constraint 
726: on any dark matter fifth force to within only a few percent 
727: of the gravitational force is observationally possible. 
728: The fifth force under perturbation theory agrees with this 
729: constraint: $\beta_{pert} \approx 1.03$ throughout cosmic history. 
730: Figure~\ref{fig:compareFifth} shows the space-averaged 
731: and time-averaged value of 
732: $\beta_{pert}-1$ and $\beta-1$, which 
733: we denote as $\langle \overline{\beta} \rangle - 1$, versus redshift for varying 
734: initial amplitudes. This is the average fraction of 
735: gravitational acceleration experienced by the DM particles due to the 
736: fifth force up to the given scale factor.
737: The field fluctuates rapidly in the early universe, and by 
738: time-averaging we can better examine the general 
739: force trends. 
740: The fluctuations in the field decay rapidly and 
741: asymptotically approach a constant value by redshift 
742: $z=45$. 
743:  
744: \begin{figure}
745: 	\plotone{f5.eps}
746: 	%\epsfig{figure=figCompareFifth.eps,width=\columnwidth}
747: 	\caption{Average ratio of fifth force to gravitational 
748: 		force ($\langle \overline{\beta} \rangle-1$)
749: 		for perturbation theory (solid line), 
750: 		compared to the full result with 
751: 		varying initial amplitudes 
752: 		(dotted lines with various marks, labeled in figure).
753: 		The marks on the lines only represent select data points, to aid 
754: 		in distinguishing the lines. 
755: 		Shown are the average strengths up to the 
756: 		given scale factor. }
757: \label{fig:compareFifth}
758: \end{figure}
759:  
760: 
761: No matter the initial amplitude, the average 
762: perturbation theory fifth force ratio
763: agrees to within ten percent with the full result by the present epoch. 
764: However, the strength of the fifth force matters much 
765: more at high redshifts: different forces here can 
766: greatly affect structure formation. We found that an initial amplitude 
767: contrast of $\delta_\phi(z=50)=10^{-8}$ provides 
768: roughly the same order of magnitude additional 
769: acceleration as perturbation theory 
770: predicts at early times. 
771: 
772: For amplitudes smaller than $10^{-8}$, the results 
773: for the fifth force do not change. Since the scalar field 
774: couples directly to the dark matter density, the field 
775: will always contain perturbations. If we initialize 
776: the amplitudes below the threshold of $10^{-8}$, 
777: the density coupling will produce fluctuations in the 
778: scalar field spectrum, creating a fifth force.
779: However, larger amplitudes 
780: provide too much of an acceleration at early times. 
781: For example, a contrast of $\delta_\phi(z=50)=10^{-6}$ 
782: provides an average fifth force that is almost half as strong as 
783: gravity at early times.  
784: We found that if the initial contrast is larger than 
785: can be handled adequately by 
786: perturbation theory, then we get radically different structure formation, 
787: such as a very early first caustic and the development of four 
788: caustics by the present epoch. 
789: Note that these results are largely redshift-independent: 
790: the DM coupling will always end up dominating the spectrum 
791: of scalar field fluctuations after a short period of time. 
792: The large fifth force for amplitudes above $10^{-8}$ 
793: is due to the large artificial initial scalar field amplitude.
794:  
795: However, if the amplitude of 
796: perturbations is small in the early universe, 
797: then perturbation theory under-emphasizes the 
798: fifth force at high redshift.  
799: The fact that perturbation theory 
800: cannot describe large fluctuations makes sense.
801: However, we would expect that perturbation 
802: theory should accurately capture the behavior 
803: of all initial amplitudes below a certain value. 
804: 
805: %\begin{figure}
806: %	\epsfig{figure=figCompareDensityFifth.eps,width=\columnwidth}
807: %	\caption{The evolution of caustic formation for 
808: %		perturbation theory with only 
809: %		the fifth force (solid line) compared to 
810: %		the full theory with only fifth force 
811: %		(dashed lines) with varying initial amplitudes of the 
812: %		scalar field. Drag, mass, and dynamic field 
813: %		effects are ignored. 
814: %		The initial contrast of the DE scalar 
815: %		field perturbations is shown in each panel.}
816: %\label{fig:compareDensityFifth}
817: %\end{figure} 
818: 
819: With the above considerations, we examined the case of an initial contrast of $\delta_\phi(z=50)=10^{-8}$,
820: since this provides a fifth force large enough to affect structure, but 
821: which is not ruled out by observations. 
822: We compare perturbation theory 
823: to the results from a full dynamical field. 
824: We found that for this amplitude, perturbation theory 
825: is excellent in capturing the structure 
826: and evolution of halos. By construction, the average value 
827: of the dynamical field is the same as the background field 
828: in perturbation theory. Hence, the modified drag term and 
829: modified mass terms are almost identical.
830: Also, 
831: perturbation theory appears to be adequate for 
832: describing the evolution of the scalar field, 
833: so the effects of the dynamic field itself agree in the full theory.
834: Additionally, since fluctuations dampen with time, 
835: we only need to satisfy perturbation theory 
836: constraints in the early universe. We see in Figure~\ref{fig:compareStandardPertFull} 
837: that, for this particular amplitude of fluctuations, perturbation theory 
838: is excellent in describing the overall structure of the pancakes.
839: 
840: Overall, as Figure~\ref{fig:compareStandardPertFull} showed, 
841: perturbation theory seems to be adequate in describing
842: the larger features of the pancakes: 
843: the number of caustics and the location of all three caustics at $z=0$. 
844: With the notable exception of the fifth force, 
845: as we decrease initial scalar field amplitudes, 
846: perturbation theory becomes more accurate 
847: in providing an accurate solution. Even though the perturbation 
848: theory fifth force disagrees with the full theory by several 
849: percent, this is not enough to affect structure formation 
850: at our resolution.
851: 
852: %\begin{figure}
853: %	\epsfig{figure=figComparePertFull.eps,width=\columnwidth}
854: %	\caption{Dark matter density profile at $z=0$. 
855: %		The solid line in all plots is from 
856: %		perturbation theory. The dashed line is 
857: %		the density from using a full dynamical field. 
858: %		The topmost panel compares with only 
859: %		the mass term included, 
860: %		the second panel compares only the drag term, 
861: %		the third compares with only the fifth force, 
862: %		and the bottom compares only the dynamic field.}
863: %\label{fig:comparePertFull}
864: %\end{figure} 
865: 
866: It is interesting to note that we can produce significant 
867: changes in structure formation even with a negligible fifth force, 
868: such as might happen with a sufficiently screened potential. 
869: Figure~\ref{fig:compareStandardNoForce} shows the results 
870: from the full theory, but with no fifth force. In this case, 
871: we get stronger deviations in the innermost pancake substructure. 
872: Also, the second caustic forms slightly later than in 
873: standard cosmology, while the largest 
874: caustic is no different than when the fifth force is large.
875: 
876:  
877: \begin{figure}
878: 	\plotone{f6.eps}
879: 	%\epsfig{figure=figCompareStandardNoForce.eps,width=\columnwidth}
880: 	\caption{Dark matter density profile at $z=0$. 
881: 		The solid line is the standard $\Lambda$CDM  
882: 		cosmology, and the dashed line is the final result in full 
883: 		theory when the fifth force is negligible.
884: 		The numerical shot noise in the density near $x=1$ 
885: 		is negligible and unimportant.}
886: \label{fig:compareStandardNoForce}
887: \end{figure}
888:  
889: 
890: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
891: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
892: \section{Distinguishability of Models} 
893: \label{sec:distinguish} 
894: The models chosen in Table~\ref{tab:FandP} are 
895: constructed to be consistent with current observations. 
896: For example, all these 
897: models predict values for the CMB peak locations that are 
898: within observational constraints. 
899: 
900: Figure~\ref{fig:comparePhi} 
901: shows how each of our chosen potential parameters in Table~\ref{tab:FandP} 
902: affects the evolution of $\phi$. 
903: These results agree with the perturbation results from Farrar and Peebles.
904: There are in general two classes 
905: of viable $\phi$ evolution tracks.
906: We see from Table~\ref{tab:FandP} that models such as  
907: $A$ and $C$ have a smaller initial value, but 
908: drop by roughly ten percent by the current epoch.
909: On the other hand, models such as $B$ and $D$ have a high 
910: initial value and do not change much throughout cosmic history.
911: 
912:  
913: \begin{figure}
914: 	\plotone{f7.eps}
915: 	%\epsfig{figure=figComparePhi.eps,width=\columnwidth}
916: 	\caption{Evolution of $\phi_b$, relative to the 
917: 	  present-day value, for the various 
918: 		potential parameters labeled in 
919: 		Table~\ref{tab:FandP}. 
920: 		Note that the symbols are only to aid in distinguishing the lines.
921: 		These 
922: 		calculations are from perturbation theory.}
923: \label{fig:comparePhi}
924: \end{figure} 
925:  
926: 
927: These two classes of parameter sets have interesting 
928: consequences for the behaviors of 
929: the DM particle mass, the modified Hubble drag, 
930: the fifth force, and the dynamical field itself. 
931: Figure~\ref{fig:comparePhiEffects} shows how 
932: the mass ratio ($\eta$) , 
933: the modified Hubble drag ($\gamma$), 
934: the fifth-force acceleration ($\beta_{pert}$), and 
935: the field energy density ratio ($\delta$) 
936: vary across the models in perturbation theory. For models of the second type, 
937: in which $\phi$ does not change much, the mass at $z=50$ is 
938: much closer to its present-day value, and hence this effect
939: on structure formation will be smaller than in models 
940: where $\phi$ varies strongly. Also, if $\phi$ does not 
941: vary much, then the Hubble drag will not be modified greatly.
942: Together, this makes sense: a set of parameters 
943: that favors a slightly-varying scalar field 
944: will look more like standard $\Lambda$CDM than those which do not.
945: Finally, for larger values of $\phi$, the fifth 
946: force is much weaker, since the 
947: force is proportional to $\phi^{-2}$ in perturbation theory. 
948: 
949:  
950: \begin{figure}
951: 	\plotone{f8.eps}
952: 	%\epsfig{figure=figComparePhiEffects.eps,width=\columnwidth}
953: 	\caption{Evolution of 
954: 		mass ($\eta$), 
955: 		Hubble drag ($\gamma$), 
956: 		fifth-force acceleration ($\beta_{pert}$), 
957: 		and dynamical field ($\delta$) 
958: 		for the various potential parameters labeled in 
959: 		Table~\ref{tab:FandP}. 
960: 		The symbols are only to aid in distinguishing the lines.
961: 		These 
962: 		calculations are from perturbation theory.}
963: \label{fig:comparePhiEffects}
964: \end{figure} 
965:  
966: 
967: These behaviors play out accordingly in the final structure at $z=0$, as 
968: Figure~\ref{fig:comparePhiDensityProfileFull} shows. 
969: These calculations were done using the full theory. 
970: When performing the analysis in the full dynamical theory, 
971: we again have to 
972: be careful when setting the initial scalar field amplitude. 
973: For all cases, 
974: we set an initial contrast of 
975: $\delta_\phi(z=50)=10^{-8}$.
976: 
977: Models $B$ and $D$ do not differ 
978: much from standard cosmology or each other. Models $A$ and $C$, 
979: while both different from 
980: $\Lambda$CDM, remain mostly indistinguishable, 
981: except that parameter set $A$, whose 
982: changes in $\phi$ are most drastic, produces caustics 
983: that are slightly farther away from the pancake midplane. 
984: Even though $\phi$ 
985: is increasing at the present epoch in parameter set $C$, 
986: this change happens 
987: at late times, and so it does not have a significant 
988: effect on the resulting structure. 
989: 
990: We also performed this study in perturbation theory, and 
991: we found that accounting for the effects of the full 
992: dynamics do not increase the distinguishability of the models.
993: 
994: 
995: %\begin{figure}
996: %	\epsfig{figure=figComparePhiDensityProfilePert.eps,width=\columnwidth}
997: %	\caption{Dark matter density profile at $z=0$. 
998: %		The solid line in all plots is the standard 
999: %		cosmology. The dashed line indicates 
1000: %		the result in perturbation theory, 
1001: %		for the various potential parameters labeled in 
1002: %		Table~\ref{tab:FandP}.}
1003: %\label{fig:comparePhiDensityProfilePert}
1004: %\end{figure}
1005: 
1006:  
1007: \begin{figure}
1008: 	\plotone{f9.eps}
1009: 	%\epsfig{figure=figComparePhiDensityProfileFull.eps,width=\columnwidth}
1010: 	\caption{Dark matter density profile at $z=0$. 
1011: 		The solid line in all plots is the standard 
1012: 		cosmology. The dashed line indicates 
1013: 		the result in full theory, 
1014: 		for the various potential parameters labeled in 
1015: 		Table~\ref{tab:FandP}.}
1016: \label{fig:comparePhiDensityProfileFull}
1017: \end{figure}
1018:  
1019: 
1020: Surprisingly, we do notice some additional distinguishing features 
1021: when examining the phase plots, as in Figure~\ref{fig:comparePhaseModels}. Since models $B$ and $D$ 
1022: remain nearly identical to concordance cosmology, we do not display them. 
1023: While models $A$ and $C$ share many common features, 
1024: the drag term $\gamma$ 
1025: increases at late times in model $C$. 
1026: This eventually slows down particles at low redshifts and pushes 
1027: the peak velocities closer to concordance cosmology.
1028: Hence, the higher peak particle velocities 
1029: distinguish model $A$, whereas simply considering 
1030: pancake density profiles may not.
1031: 
1032:  
1033: \begin{figure}
1034: 	\plotone{f10.eps}
1035: 	%\epsfig{figure=figComparePhaseModels.eps,width=\columnwidth}
1036: 	\caption{Dark matter particle phase plots at $z=0$.
1037: 		The solid line in all plots is from concordance cosmology, 
1038: 		and the dotted lines are the results from various 
1039: 		model parameters, which are labeled in the plots.
1040: 		These were calculated in the full theory.
1041: 		The symbols are only to aid in distinguishing the lines.}
1042: \label{fig:comparePhaseModels}
1043: \end{figure}
1044:  
1045: 
1046: \section{Conclusions}
1047: \label{sec:Conclusion}
1048: These one-dimensional simulations have clearly demonstrated that 
1049: models of interacting dark matter and dark energy affect the 
1050: growth and structure of plane parallel perturbations. Various consequences of 
1051: these theories play important roles at different stages in halo 
1052: evolution. Larger fifth forces and modified Hubble drag terms 
1053: in the high-redshift universe greatly alter early structure formation, 
1054: while an evolving particle mass can affect even late-forming structures.   
1055: Ordinary quintessence, which is a dynamic field alone, 
1056: does not significantly alter the formation 
1057: of pancakes. Of all the effects, a modification to the Hubble 
1058: drag gives the largest deviations in the resulting structure.
1059: 
1060: We have found that models that are specifically tuned 
1061: to match current observational constraints, such as the 
1062: CMB peak locations and equation of state parameters, 
1063: produce significant deviations from standard 
1064: $\Lambda$CDM in the formation of pancakes.
1065: 
1066: We have also found that an approach  
1067: based on perturbation theory is adequate for understanding the 
1068: general evolution of the scalar field.
1069: However, perturbation theory does not seem to be appropriate 
1070: at high redshift. Even for arbitrarily small amplitudes, 
1071: the fifth force in the full theory is larger than that obtained 
1072: by perturbative techniques.
1073: This behavior deviates from the general 
1074: pattern of perturbation methods, in that 
1075: it does not accurately describe all characteristics
1076: for amplitudes smaller than a threshold value. 
1077: 
1078: This discrepancy arises from the fact that the 
1079: fifth force directly depends on the gradient of the 
1080: field. In the perturbative method described 
1081: by Farrar and Peebles, the kinetic terms of the perturbation 
1082: field equation of motion are dropped, producing a 
1083: Poisson equation for the perturbation field and leading to the 
1084: simplification of the fifth force expression found in equation~(\ref{eq:fifthPert}). 
1085: This simplification is valid at low redshifts, but at high 
1086: redshifts the kinetic terms are still important.
1087: Thus, any realistic spectrum of perturbations 
1088: for the scalar field which 
1089: ignores the kinetic terms may be invalid for the 
1090: very largest-scale perturbations.
1091: It also appears that the initial scalar field 
1092: amplitude does not serve as a suitable ``small'' 
1093: parameter for governing the appropriateness 
1094: of perturbation theory.
1095: 
1096: Because we can freely choose the initial amplitude, 
1097: full theory gives us more freedom to study the effects 
1098: of various strengths of a fifth force. 
1099: In the context of this model, 
1100: this can also work in reverse: 
1101: working within these frameworks,
1102: constraints on a fifth force in the dark sector could 
1103: lead to limitations on the amplitudes 
1104: of a DE scalar field in the early universe.
1105: 
1106: Anninos and Norman have extensively studied the effects on baryons in the formation of 
1107: pancakes~\citep{Anninos1,Anninos2}. 
1108: While baryons appear to modify only the last-forming caustic, DM-DE 
1109: interactions affect all caustics. 
1110: Hence, the modifications to the last caustic may be masked by the baryon dynamics. 
1111: Also, 
1112: baryons do not appear to significantly modify dark matter particle velocities, while 
1113: some of the models considered above produce significant variations. 
1114: 
1115: The structural differences described above are relatively small. 
1116: However, in three dimensions these effects should appear as 
1117: percent-level differences in the statistical properties of large samples of dark 
1118: matter halos. The various effects could produce different halo mass 
1119: functions, which could be compared to other 
1120: high-resolution mass functions, 
1121: such as the $\Lambda$CDM ones considered by~\citet{Warren}.  
1122: In fact,~\cite*{Mainini} have shown that 
1123: DM-DE interactions do modify analytic mass functions, 
1124: and a direct simulation could be compared against 
1125: these results. DM-DE 
1126: interactions may also affect halo substructure, 
1127: in which case they would impact 
1128: NFW profiles~\citep{NFW} and distributions of 
1129: concentrations in halo catalogs~\citep{Lukic, Reed}.
1130: 
1131: The present work focused on a single model and a 
1132: single potential, simply to analyze the 
1133: feasibility of a direct simulation approach.
1134: With a full three-dimensional simulation, we can 
1135: also examine the results of using other potentials, or 
1136: even more complicated models.
1137: With future simulations, we 
1138: can also contrast these structure formation results to the 
1139: \emph{N}-body results from other theories of DE, 
1140: such as modified General Relativity, which 
1141: have been examined in~\citet{Stabenau}.
1142: 
1143: For the models with the largest deviations from $\Lambda$CDM, we notice 
1144: that the sign of the difference of the caustic peak location changes.
1145: This implies that the shape and evolution of the mass function for these
1146: cosmologies will change relative to $\Lambda$CDM, 
1147: not simply the amplitude. This holds the most hope for 
1148: observations, particularly for surveys such as the DES~\citep{Annis}.
1149: Most notably, we found that we will still get
1150: significant modifications to pancake structure even if the fifth force is negligible. 
1151: Indeed, the modifications to the mass function may even 
1152: be more significant without a fifth force.
1153: Since many recent efforts have concentrated mostly on 
1154: constraining the fifth force~\citep{Kesden,Farrar2,Bertolami,Guo}, this result clearly shows 
1155: that we should not ignore the other 
1156: consequences, such as modifications to the 
1157: Hubble drag, of these models.
1158: Also, the phase plots revealed another observable effect: the large-scale matter velocity field, 
1159: which is also accessible to the DES survey. 
1160: The feasibility of constraining these models with this behavior could be tested with 
1161: velocity correlations in full three-dimensional simulations.
1162: 
1163: However, there are several computational and 
1164: theoretical challenges in developing 
1165: and analyzing full 
1166: three dimensional simulations. For example, there should be a stronger 
1167: theoretical understanding of the initial conditions of the field. 
1168: While implicitly solving the scalar field is 
1169: relatively straightforward in one dimension,
1170: requiring a simple Thomas algorithm, 
1171: a full three dimensional solver would be much more complex, 
1172: probably requiring a multigrid solver that could easily be 
1173: combined with existing parallel adaptive-mesh codes
1174: such as FLASH~\citep{Fryxell} or GADGET-2~\citep{Springel}. Since there 
1175: might be interesting changes in halo substructure, simulations 
1176: will require very high resolution. 
1177: Also, simulations will require many halos 
1178: to get significant statistical results.
1179: 
1180: Fortunately, none of these issues are intractable, and this 
1181: approach holds much hope for providing a strong, consistent 
1182: method of analyzing the nonlinear effects of these myriad 
1183: dark energy models.
1184: 
1185: \acknowledgements
1186: The authors would like to thank Luke Olson, Ben Wandelt, 
1187: and Greg Huey for enlightening and valuable discussions.
1188: 
1189: The authors acknowledge support under a Presidential Early 
1190: Career Award from the U.S. Department of Energy, 
1191: Lawrence Livermore National Laboratory (contract B532720).
1192: Additional support was provided by a DOE 
1193: Computational Science Graduate Fellowship 
1194: (DE-FG02-97ER25308) and the National Center for 
1195: Supercomputing Applications.
1196: 
1197: \appendix
1198: \section{Numerically Solving the Scalar Field Equation}
1199: \label{app:numerics}
1200: We have
1201: \begin{eqnarray}
1202: \label{eq:phi}
1203: \ddot{\phi} - \frac{c^2}{a^2} \nabla^2 \phi + 3 \frac{\dot{a}}{a} \dot{\phi} + \frac{d V}{d \phi} 
1204:  + \frac{\rho}{\phi}a^{-3}& = & 0.  
1205: \end{eqnarray}
1206: 
1207: We will let $V(\phi) = K \phi^{-\alpha}$, 
1208: and we have re-introduced $c$ for clarity. We may break up equation~(\ref{eq:phi}) 
1209: using standard operator splitting techniques:
1210: \begin{eqnarray}
1211: 	\label{eq:splitPhi1}
1212: 		{\phi}^n \rightarrow {\phi}^{(1)} & : & \frac{\partial^2 \phi}{\partial t^2} - \frac{c^2}{a^2} \nabla^2 \phi = 0 \\		
1213: 	\label{eq:splitPhi2}
1214: 		{\phi}^{(1)} \rightarrow {\phi}^{n+1} & : & \frac{\partial^2 \phi}{\partial t^2} + \frac{d V}{d \phi} 
1215: 								+ 3 \frac{\dot{a}}{a} \frac{\partial \phi}{\partial t} + \frac{\rho}{\phi}a^{-3} = 0.
1216: \end{eqnarray}
1217: 
1218: Here and below, superscripts are temporal indices and subscripts 
1219: will denote the index of the location on the mesh.  
1220: We solve Eq.(\ref{eq:splitPhi1}) by reducing it to a set 
1221: of two 1st-order equations: 
1222: \begin{eqnarray}
1223: 	\label{eq:splitWave1}
1224:   	\frac{\partial \phi^{(1)}}{\partial t} & = & \dot{\phi}^{n} \\
1225:   \label{eq:splitWave2}	
1226:   	\frac{\partial \dot {\phi}^{(1)}}{\partial t} & = & - \frac{c^2}{a^2} \nabla^2 \phi^{n}.
1227: \end{eqnarray}
1228:  
1229: Since the sound speed is so high ($v_s = c/a$), 
1230: we found that a direct integration scheme was highly unstable 
1231: except for unworkably small $\Delta t$. So, 
1232: we use a midpoint method for Eq.(\ref{eq:splitWave1}) 
1233: and a fully implicit method for equation~(\ref{eq:splitWave2}) :
1234: \begin{eqnarray}
1235: 	\label{eq:midpoint}
1236:   	\phi^{1}_{i} & = & \phi^{n}_i  + \frac{1}{2} \Delta t \left(  \dot{\phi}^{n}_i + \dot{\phi}^{(1)}_i \right)\\
1237:   \label{eq:implicit}	
1238:   	\dot{\phi}^{(1)}_i & = & \dot{\phi}^{n}_i
1239:   			 + \frac{c^2 \Delta t}{\Delta x^2 \left( a^{(1)}\right)^2} 
1240:   													\left( \phi^{(1)}_{i+1} - 2 \phi^{(1)}_{i} + \phi^{(1)}_{i-1} \right).
1241: \end{eqnarray}
1242: 
1243: We substitute our expression for $\phi^{(1)}_{i}$ into equation~(\ref{eq:implicit}), yielding the matrix equation $Ax=b$, where
1244: \begin{equation}
1245: x_i \equiv \dot{\phi}_i^{(1)},
1246: \end{equation}
1247: \begin{eqnarray}
1248: b_i \equiv \dot{\phi}^{n}_{i} & + & \sigma \phi^{n}_{i+1} + \gamma \sigma \dot{\phi}^{n}_{i+1} \nonumber \\
1249: 															& - & 2 \sigma \phi^{n}_{i} -2 \gamma \sigma \dot{\phi}^{n}_{i} \\
1250: 															& + & \sigma \phi^{n}_{i-1} + \gamma \sigma \dot{\phi}^{n}_{i-1} \nonumber,
1251: \end{eqnarray}
1252: and
1253: \begin{eqnarray}
1254: A = \left( \begin{array}{ccccc}
1255: 											1+2 \sigma \gamma & -\sigma \gamma & 0 & \cdots & -\sigma \gamma  \\
1256: 											-\sigma \gamma & 1+2 \sigma \gamma & -\sigma \gamma & \cdots & 0  \\
1257: 											0 & -\sigma \gamma & 1+2 \sigma \gamma & -\sigma \gamma & 0  \\
1258: 											\cdots & \cdots & \cdots & \cdots & \cdots  \\ 
1259: 											-\sigma \gamma & \cdots & 0 & -\sigma \gamma & 1+2 \sigma \gamma \end{array} 
1260: 														\right) \nonumber.
1261: \end{eqnarray}
1262: Note the periodic boundary conditions. We have defined
1263: \begin{eqnarray}
1264: \sigma & \equiv & \frac{c^2}{\left( a^{(1)}\right)^2} \frac{\Delta t}{\Delta x^2}, \\
1265: \gamma & \equiv & \frac{1}{2} \Delta t. \nonumber
1266: \end{eqnarray}
1267: We solved this matrix-vector equation exactly using a Thomas algorithm~\citep{Thomas}, modified for 
1268: periodic boundaries via the Sherman-Morrison formula~\citep{Sherman}.
1269: 
1270: We may now update our solution through equation~(\ref{eq:splitPhi2}) by expanding in a Taylor series:
1271: \begin{equation}
1272: \dot{\phi}^{n+1} = \dot{\phi}^{(1)} + \ddot{\phi}^{(1)} \Delta t 
1273: 									+ \frac{1}{2} \frac{d^3 \phi^{(1)}}{dt^3} \Delta t^2 + O(\Delta t^3),
1274: \end{equation}
1275: where
1276: \begin{eqnarray*}
1277: \ddot{\phi}^{(1)} & = & K \alpha \left( {\phi^{n}}\right)^{-\alpha-1} 
1278: 											- 3 \frac{\dot{a}^{n}}{a^{n}} \dot{\phi}^{(1)} 
1279: 											- \frac{\rho}{\phi^{(1)}}(a^{n})^{-3}
1280: \end{eqnarray*}
1281: and
1282: \begin{eqnarray*}
1283: \frac{d^3 \phi^{(1)}}{dt^3}& = & -K \alpha (\alpha+1) \left(\phi^{n} \right)^{-\alpha-2} \dot{\phi}^{(1)} 
1284: 											- 3 \frac{\ddot{a}^{n}}{a^{n}} \dot{\phi}^{(1)} \\
1285:  									&\mbox{ }&			+ 12 \left( \frac{\dot{a}^{n}}{a^{n}}  \right)^2 \dot{\phi}^{(1)} 
1286: 													- \frac{\dot \rho}{\phi^{(1)}}(a^{n})^{-3} \\ 
1287: 									&\mbox{ }&			- \frac{\rho}{\left( \phi^{(1)} \right)^2}(a^{n})^{-3} \dot{\phi}^{(1)}
1288: 												+3 \frac{\rho}{\phi^{(1)}}(a^{n})^{-4} (\dot{a}^{n}).
1289: \end{eqnarray*}
1290: Here, $\dot \rho = \left( \rho^{n} - \rho^{n-1}\right)/dt $.
1291: 
1292: At any step $n$, after computing $\phi^{n}$ across the grid, 
1293: we may update the scale factor using a Taylor series:
1294: \begin{equation}
1295: a^{n+1} = a^n + \dot{a}^n \Delta t + \frac{1}{2} \ddot{a}^n \Delta t^2 + O(\Delta t^3),
1296: \end{equation}
1297: 
1298: where $\dot{a}$ and $\ddot{a}$ are given by the Friedmann equation and Friedmann energy equation, respectively:
1299: \begin{eqnarray}
1300: \label{eq:a}
1301: \dot{a}^{n} & = & \left[ H_0^2 \Omega_m \frac{ \overline{\phi} }{ \overline{\phi}_0} a^{-1} 
1302: 									+ \frac{8 \pi G}{3} a^{2} \rho_\phi \right]^{1/2} \\
1303: \ddot{a}^{n} & = & -\frac{1}{2} H_0^2 \Omega_m \frac{\overline{\phi} }{\overline{\phi}_0} a^{-2} - \frac{4 \pi G}{3} (1+3w)\rho_\phi a.
1304: \end{eqnarray}
1305: 
1306: Ignoring fluctuations, the equation of state parameter, $w$, is 
1307: \begin{equation}
1308: w \equiv \frac{p_\phi}{\rho_\phi} = \frac{\frac{1}{2}  \overline{\dot \phi}^2 -  V(\overline{\phi})}
1309: 																		{\frac{1}{2} \overline{\dot \phi}^2 + V(\overline{\phi})}.
1310: \end{equation}
1311: 
1312: \bibliographystyle{apj}	
1313: \bibliography{ms}		
1314: \nocite{*}
1315: 
1316: \end{document}