0804.4172/ms.tex
1: \newcommand{\hmpc}{\mbox{ } h^{-1} \mbox{ Mpc}}
2: \newcommand{\hkpc}{\mbox{ } h^{-1} \mbox{ kpc}}
3: \newcommand{\hmsol}{\mbox{ } h^{-1}  M_{\odot}}
4: \newcommand{\lcdm}{$\Lambda$CDM }
5: 
6: %\documentclass[12pt, preprint]{aastex}
7: \documentclass[floatfix]{emulateapj}
8: \usepackage{epsfig,subfigure}
9: 
10: \bibpunct[; ]{(}{)}{;}{a}{}{,}
11: 
12: \begin{document}
13: \title{Detecting dark matter-dark energy coupling with the halo mass function}
14: 
15: \author{P.~M.~Sutter} \email{psutter2@uiuc.edu}
16: \affil{Department of Physics,
17: 	     University of Illinois at Urbana-Champaign,
18:             Urbana, IL 61801-3080}
19: 
20: \and
21: 
22: \author{P.~M.~Ricker} \email{pmricker@uiuc.edu}
23: \affil{Department of Astronomy,
24:        University of Illinois at Urbana-Champaign,
25:              Urbana, IL 61801\\
26: 		National Center for Supercomputing Applications,
27:       University of Illinois at Urbana-Champaign,
28:             Urbana, IL 61801}
29:             
30: \begin{abstract}
31: We use high-resolution simulations of large-scale structure 
32: formation to analyze the effects of interacting dark matter and 
33: dark energy on the evolution of the halo mass function.
34: Using a chi-square likelihood analysis, we find significant 
35: differences in the mass function 
36: between models of coupled dark matter-dark energy and standard 
37: concordance cosmology (\lcdm) out to redshift $z=1.5$. We also 
38: find a preliminary indication that the Dark Energy 
39: Survey should be able to distinguish these models 
40: from \lcdm
41: within its mass and 
42: redshift contraints. While we can distinguish the effects 
43: of these models from \lcdm cosmologies with different 
44: fundamental parameters, DES will require independent 
45: measurements of $\sigma_8$ to confirm these effects.
46: \end{abstract}
47: \keywords{cosmology:theory, dark matter, dark energy, structure formation, methods: N-body simulations}
48: \maketitle
49: 
50: \section{Introduction}
51: \label{sec:introduction}
52: Multiple independent lines of evidence, including 
53: observations of the large-scale matter distribution~\citep[eg.][]{Percival},
54: cosmic microwave background fluctuations~\citep[eg.][]{WMAP5}, 
55: and type Ia supernovae~\citep{Perlmutter,Riess},
56:  suggest that our universe 
57: is dominated by two components: dark matter, which is probably 
58: a form of nonbaryonic matter, and dark energy, which is a name 
59: for the presently unknown cause of the 
60: observed acceleration of expansion. However, we still lack 
61: an understanding of any possible interactions between these two 
62: principal constituents~\citep{Bean}. In a previous paper 
63: \citep[][hereafter SR08]{Sutter}, we examined 
64: the role that interacting dark matter and dark energy would play 
65: in the development of one-dimensional Zel'dovich pancakes, an 
66: important idealized case useful for understanding structure formation. 
67: Here we extend that 
68: preliminary work to a more realistic three-dimensional 
69: simulation of the growth of dark matter halos
70: in an attempt to find ways to distinguish these 
71: models from standard cosmology.
72:  
73: In this paper, we study the effects of a Yukawa interaction 
74: between a single family of nonrelativistic dark matter (DM) 
75: particles and a scalar field 
76: that is responsible for the dark energy (DE). We follow closely the formalism 
77: developed by~\cite*{Farrar}. Such an interaction is initially attractive 
78: because it is motivated by particle physics~\citep{Amendola2}  and 
79: might provide a way to explain the apparent emptiness of 
80: the voids, as demonstrated numerically by~\cite*{Nusser}.
81: 
82: There has been considerable interest recently in studying the 
83: effects of these interactions on structure, both using 
84: an analytic approach~\citep{Mainini} and using 
85: direct simulations~\citep[eg.][and others]{Maccio, Manera}. 
86: However, the current numerical 
87: studies suffer from poor resolution, and these results cannot 
88: reliably be compared to simulations of standard cosmological 
89: structure formation. In this paper, we use high 
90: spatial resolution and careful analysis 
91: to accurately capture many dark matter halos 
92: for use in comparison.
93:  
94: The following is a brief 
95: summary of the equations we solve and our numerical techniques. 
96: In Section~\ref{sec:massFunction} we discuss modifications to the halo 
97: mass function. We use these mass functions to distinguish 
98: interacting DM-DE from standard concordance cosmology 
99: using a $\chi^2$ likelihood test. 
100: Additionally, we discuss the feasibility of using the Dark Energy 
101: Survey~\citep{Annis} to detect this coupling within its mass 
102: and redshift contraints. Finally, we determine the 
103: extent to which two specific models of interacting DM-DE 
104: can be differentiated from each other.
105: 
106: \subsection{Analytical Methods}
107: Compared to simulations of the full non-linear theory, we found 
108: in SR08 that the perturbation theory presented by 
109: Farrar and Peebles is very accurate in determining the evolution of 
110: structure, and hence we will maintain the perturbative approach 
111: and assume fluctuations in the scalar field are small.
112: Under perturbation theory, the homogenous part of the 
113: dark energy scalar field, $\phi_b$, evolves as
114: \begin{equation}
115: \label{eq:phiEoM}
116: 	\ddot{\phi_b} + 3 \frac{\dot{a}}{a} \dot{\phi_b} 
117: 	  + \frac{d V}{d \phi_b} + 3 \frac{\Omega_{m0} H_0^2}{8 \pi G}
118: 	    \frac{1}{\phi_b} a^{-3}= 0,
119: \end{equation}
120: where $\Omega_{m0}$ is the dark matter 
121: particle fraction of the critical density  
122: and $H_0$ is the Hubble constant.  
123: A subscript of $0$ denotes the present-day value.
124: The dark matter particle equation of motion is
125: \begin{equation}
126: \label{eq:dmEoM}
127: \dot{\bf{v}}  + \left( 2 \frac{\dot{a}}{a} + \frac{\dot{\phi_b}}{\phi_b} \right)
128:  \bf{v} = - \left( 1 + \frac{1}{4 \pi G} \frac{1}{\phi_b^2} \right) 
129:    \nabla \Phi.
130: \end{equation}
131: Here $\Phi$ is the normal comoving gravitational potential, 
132: $a$ is the scale factor, $\bf{v}$ is 
133: the comoving particle peculiar velocity, 
134: and $\bf{x}$ is the comoving position. Throughout, dots 
135: refer to derivatives with respect to proper time $t$. 
136: Perturbations in the scalar field give rise to the fifth force 
137: on the right-hand side in the equation above.
138: 
139: The comoving potential satisfies the Poisson equation:
140: \begin{equation}
141: \label{eq:poisson}
142:  \nabla^2 \Phi = \frac{4 \pi G}{a^3} \left( \rho - \overline{\rho} \right),
143: \end{equation}
144: with $\rho$ as the comoving DM particle density. 
145: Here and throughout, an overline indicates a spatial average.
146: 
147: At the present epoch, the field 
148: behaves as a cosmological constant, so the potential term 
149: in Eq.(\ref{eq:phiEoM} dominates and has a value 
150: \begin{equation}
151: 	V(\phi_{b,0}) = \Omega_{\Lambda 0} \rho_{crit}. 
152: \label{eq:potentialToday}
153: \end{equation}
154: At early times, the coupling to matter dominates the 
155: scalar field equation of motion, and Eq.~(\ref{eq:phiEoM}) reduces to
156: \begin{equation}
157: \label{eq:initialPhiDot}
158: 	\frac{d \phi_b} {d t} = - \frac{H_0^2}{G} 
159: 	\frac{3 \Omega_{m0} }{8 \pi \phi_0} \frac{1}{a^3} t,
160: \end{equation}
161: which we use to set the initial condition for $\dot \phi_b$. 
162: 
163: The DM particle also has a field-dependent mass
164: \begin{equation}
165: \label{eq:mass}
166:    m_{DM} = {m_{DM,0}} \frac{\phi_b}{\phi_{b,0}}.
167: \end{equation}
168: The Friedmann equation, neglecting radiation, curvature, and 
169: baryonic terms, becomes 
170: \begin{equation}
171: \label{eq:friedmann}
172: 	\left( \frac{\dot{a}}{a} \right)^2 = 
173: 	H_0^2 \Omega_{m0} \frac{\phi_b}{\phi_{b,0}} a^{-3} + 
174: 	\frac{8 \pi G}{3} \left[ \frac{1}{2} \left( \frac{d \phi_b}{d t} 
175: 	\right)^2 + V(\phi_b) \right].
176: \end{equation}
177: 
178: To simulate a comparative $\Lambda$CDM cosmology, 
179: we fixed $\phi_b$ to the value in Eq.(\ref{eq:potentialToday}) 
180: and prevented any interactions between the field and particles.
181: 
182: \subsection{Numerical Methods}
183: We study a general power-law potential:
184: \begin{equation}
185: \label{eq:powerLaw}
186: 	V(\phi) = K / {\phi}^{\alpha},
187: \end{equation}
188: where we are free to choose the constants $K$ and $\alpha$. Based 
189: on the comments made by Farrar and Peebles and our analysis 
190: in SR08, we chose two 
191: combinations of parameters. 
192: These were selected for behaviors that were significantly different 
193: from standard cosmologies, but not drastic enough to rule them out 
194: with current observational contraints.
195: Table~\ref{tab:FandP} lists the parameter values, the guessed 
196: initial field value at our simulation initial redshift, 
197: and the field value today as calculated 
198: from Eq.~(\ref{eq:phiEoM}). 
199: 
200: \begin{table}
201:   \centering
202: 	\begin{tabular}{|c|c|c|c|c|} 
203: 		\hline
204: 	Label & $ \alpha $ & $K (G^{1+\alpha/2} / H_0^2)$ 
205: 	      & $\phi_{\mbox{init}} (G^{1/2})$ & $\phi_0 (G^{1/2})$\\
206: 	\hline
207: 	A & $-2$ & $0.03$   & $1.89$   & $1.72$  \\
208: 	B & $6$  & $2.0$    & $1.80$  & $1.68$  \\
209: 	\hline
210: 	\end{tabular}
211: 	\caption{Simulation potential parameter value choices.}
212: 	\label{tab:FandP}
213: \end{table}
214: 
215: For our simulations we chose FLASH v2.5, an adaptive-mesh refinement (AMR) 
216: code for astrophysics and cosmology~\citep{Fryxell}. FLASH solves the 
217: N-body potential 
218: problem with a particle-mesh multigrid FFT method~\citep{Ricker}.
219: FLASH uses cloud-in-cell mapping for interpolating between 
220: the mesh and particles~\citep{Hockney} and 
221: a second-order leapfrog integration scheme 
222: for variable-timestep particle advancement. 
223: We modified the standard FLASH code by adding 
224: the additional drag and force terms 
225: from Eq.~(\ref{eq:dmEoM}). 
226: At each timestep, the particle mass is updated according to 
227: Eq.~(\ref{eq:mass}). We calculate the scale factor and scalar field 
228: value in-code by numerically solving Eqs.~(\ref{eq:friedmann}) 
229: and (\ref{eq:phiEoM}), respectively.
230: For a more detailed explanation of solving the scalar field 
231: equation , see SR08.
232: 
233: For all calculations, 
234: we used concordance parameter values of
235:  $\Omega_{m0} = 0.26$, $\Omega_{\Lambda 0}=0.74$, and 
236:  $H_0 = 100 h = 71 \mbox{ km s}^{-1} \mbox{ Mpc}^{-1}$. 
237:  All runs took place in a three-dimensional box 
238: measuring $128 \hmpc$ per
239: side with $256^3$ particles. For each model, including 
240: a $\Lambda$CDM reference, we performed 10 simulations 
241: with $512$ zones per side and an additional 4 
242: simulations with $1024$ zones per side to study lower mass ranges. 
243: There was no refinement of grid spacing. 
244: All simulations used the same initial conditions: 
245: unperturbed particle positions were situated on a grid, 
246: and the initial velocities and positions were perturbed 
247: using Guassian fluctuations normalized to 
248: $\sigma_8 = 0.751$. We assumed P(k) from a \lcdm cosmology. 
249: We used the GRAFIC2 code~\citep{Bertschinger}
250: to generate these initial conditions. 
251: All computations started at a redshift of $z=56.8$.
252: 
253: We used a friends-of-friends (FOF) routine to find halos. This algorithm 
254: builds lists of all particles that are within a certain maximum 
255: distance of their neighbors. For all results, we chose a linking 
256: distance of $1/5$ of the unperturbed particle spacing, which is $500 \hkpc$. 
257: At every analysis redshift, we calculated the minimum resolvable halo 
258: particle count according to the prescription in~\citet{Lukic}:
259: \begin{equation}
260: n_{h,min} = \frac{\Delta (1.62 n_p/n_g)^3}{\Omega_{m0} (1+z)^3 } 
261: \left[ \Omega_{m0}(1+z)^3 + \Omega_{\Lambda 0} \right],
262: \label{eq:minMass}
263: \end{equation}
264: where $n_g$ and $n_p$ are the number of zones and particles 
265: per side, respectively. We chose an overdensity factor of $\Delta=200$.
266: To examine the halo mass function, we corrected the halo FOF 
267: particle counts
268: by the factor given in~\citet{Warren}:
269: \begin{equation}
270: {n_h}^{corr} = n_h (1-{n_h}^{-0.6}).
271: \label{eq:haloCorrection}
272: \end{equation}
273: 
274: \section{The Halo Mass Function}
275: \label{sec:massFunction}
276: Figure~\ref{fig:massFunction} shows the relative mass function from 
277: redshift $z=0$ to $z=1.5$ for the two models listed above. Both 
278: of these are compared to the $\Lambda$CDM simulation mass function. 
279: We will use the frequency density definition of the mass function,
280: \begin{equation}
281: F(M,z) \equiv \log{\frac{dn}{d \log{M}}},
282: \label{eq:massFunc}
283: \end{equation}
284: so that for model $i$ we may define the relative mass function as
285: \begin{equation}
286: RF(M,z)_i \equiv \left. \log{\frac{dn}{d \log{M}}} \right|_{i}
287: 		- \left. \log{\frac{dn}{d \log{M}}} \right|_{\Lambda CDM}.
288: \label{eq:relativeMassFunction}
289: \end{equation}
290: We analyzed relative mass functions to reduce any systematic 
291: errors in the simulations, including those due to small box 
292: effects, such as missing tidal forces. We binned our distributions into 10
293: fixed logarithmic intervals from $10^{11.5}$ 
294: to $10^{15}$ $\hmsol$.
295: We only display values in bins for which we have complete data 
296: (i.e. the bin does not contain the minimum resolvable mass).
297: The uncertainties shown are obtained by summing in 
298: quadrature the 
299: individual statistical counting errors in the interacting and 
300: \lcdm cases. We see that at the present epoch, both models 
301: produce a greater number of the most massive halos while 
302: underproducing low-mass objects. At higher redshifts, both 
303: models produce greater numbers of all objects.
304: Since the relative mass function does not remain constant 
305: with redshift, it is distinguishable from a concordance 
306: cosmology with different fundamental parameters. 
307: 
308: \begin{figure*}
309: 	\centering
310:   	\subfigure[Model A - $\Lambda$CDM]
311: 		{\epsfig{figure=fig1a.eps,width=\columnwidth}}
312:   	\subfigure[Model B - $\Lambda$CDM]
313: 		{\epsfig{figure=fig1b.eps,width=\columnwidth}}
314:         \caption{Relative mass functions for the power-law potentials.. 
315:                  The plots are labeled as in Table~\ref{tab:FandP}. 
316:                  Error bars are at one standard deviation and come 
317: 		 from statistical counting uncertainties. The solid 
318: 		 curve is the relative mass function 
319: 		 of two \lcdm cosmologies with $\sigma_8=0.775$ 
320: 		 and $\sigma_8=0.751$.}
321: \label{fig:massFunction}
322: \end{figure*}
323: 
324: To determine the significance of these mass function differences, 
325: we performed  
326: a $\chi^2$ likelihood test at each redshift.
327: For two indepedent frequency distributions $R$ and $S$,
328: \begin{equation}
329: \chi^2 = \sum_i \frac{(S_i-R_i)^2}{S_i+R_i},
330: \label{eq:chi-square}
331: \end{equation} 
332: where the sum takes place over all bins and the 
333: number of degrees of freedom is the total number 
334: of non-zero bins.
335: Figure~\ref{fig:pvalComp} shows the probability 
336: at each redshift that the 
337: frequency distributions from the interacting cases are consistent 
338: with the \lcdm case. We see that for almost all redshifts, the  
339: probability is exceedingly small, indicating that these 
340: coupled DM-DE models are well distinguished from \lcdm cosmology 
341: with this sample of objects.
342: As a function of redshift, the probability 
343: for both models generally decreases. 
344: For model $A$ at $z=0.5$, however, 
345: these distributions are not well separated, since this is 
346: the redshift at which the mass functions come closest.
347: Model $B$ remains indistinct from \lcdm until $z=0.5$, 
348: at which point the separation becomes progressively more evident.
349: The probabilities for both models drop below $10^{-10}$ at 
350: redshifts $z=0.75$ and $z=1.5$ and are not shown at higher redshifts.
351: 
352: To determine the feasibility of using a survey like DES to 
353: detect these models, we need to know the expected frequency 
354: distribution of observed objects in the survey, since the 
355: fundamental uncertainty arises from counting errors.
356: While the exact selection function for DES has not yet 
357: been determined, the survey is expected to capture $\sim10,000$ 
358: objects of mass $> 10^{13.5} \hmsol$ out to redshift $z\sim1.5$~\citep{DES}.
359: Our combined simulations produced roughly this many objects in this mass 
360: and redshift range.
361: Figure~\ref{fig:pvalComp} shows the $\chi^2$ 
362: probability when only considering mass bins above $10^{13.5} \hmsol$.
363: With this cut the probability suffers; however, 
364: we maintain the general trend of increased 
365: disparity with higher redshift. For both models, the 
366: greatest deviations occur at $z=0.75$ and $z=1.5$. The 
367: probabilities at these redshifts are below the 
368: common significance threshold and indicate that DES 
369: is capable of detecting these models. At lower redshifts, 
370: there is not a significant difference in the number of high-mass 
371: objects. Between these redshifts, 
372: the uncertainties in the largest mass bins are too large to make 
373: a confident distinction between the distributions.
374: 
375: \begin{figure}
376: 	\epsfig{figure=fig2.eps, width=\columnwidth}
377: 	\caption{Probability for the chi-square statistic 
378: 		as a function of redshift for model A (top) and 
379: 		model B (bottom) compared to $\Lambda$CDM. 
380: 		The solid lines are from 
381: 		including all resolvable masses, and the dashed lines 
382: 		are from only considering objects 
383: 		with $M_{FOF} > 10^{13.5} \hmsol$.}
384: \label{fig:pvalComp}
385: \end{figure}
386: 
387: It is possible to observe similar deviations in the 
388: mass function by changing the fundamental parameters 
389: of a \lcdm cosmology, such as $\Omega_{m0}$ 
390: and $\sigma_8$. We could not find any combination 
391: of fundamental parameters that reproduced these 
392: relative mass functions for all mass bins and redshifts. 
393: However, when examining masses within the DES limit, there 
394: are degeneracies. For example, Figure~\ref{fig:massFunction} 
395: shows the relative mass function, obtained using 
396: the~\citet{Warren} analytic mass function, of 
397: two \lcdm cosmologies with different values of $\sigma_8$. In this 
398: case, we compared a $\sigma_8$ of $0.751$, which we 
399: used in our simulations, to a cosmology with $\sigma_8=0.775$.
400: We chose this value to mimic the entire relative mass function at high 
401: redshift, but it does not capture the counts of lower-mass 
402: objects at low redshift. However, above $10^{13.5} \hmsol$, 
403: the uncertainties are large enough to permit this modified 
404: $\sigma_8$ to adequately fit the data. Thus, DES alone 
405: may not be able to distinguish between a universe with interacting DM-DE 
406: and a universe with higher $\sigma_8$.
407: 
408: While these two models are easily distinguishable from \lcdm cosmology, 
409: they are more difficult to differentiate from each other.
410: Figure~\ref{fig:pvalDiff} shows the $\chi^2$ probability when comparing 
411: these two models, both for all masses and DES-accessible masses.
412: When including all masses, the probabilities are significant
413: at redshifts $z=0.0$, $1.0$, and $1.5$. However, when examining 
414: only masses available to DES, the probabilities maintain a 
415: roughly constant performance and never reach a significant level.
416: 
417: \begin{figure}
418: 	\epsfig{figure=fig3.eps, width=\columnwidth}
419: 	\caption{Probability for the chi-square statistic 
420: 		as a function of redshift for model A compared to 
421: 		model B. 
422: 		The solid line is from 
423: 		including all resolvable masses, and the dashed line 
424: 		is from only considering objects 
425: 		with $M_{FOF} > 10^{13.5} \hmsol$.}
426: \label{fig:pvalDiff}
427: \end{figure}
428: 
429: \section{Conclusions}
430: \label{sec:Conclusion}
431: We have found that coupling dark matter to a dark energy scalar 
432: field produces significantly different mass functions at 
433: redshifts as high as $z=1.5$ relative to a \lcdm cosmology 
434: with the same set of fundamental parameters. 
435: This difference in the mass function follows from 
436: our analysis in SR08: an additional fifth force 
437: and a reduced particle Hubble drag lead to more 
438: structures than in \lcdm cosmologies at early times, and at late 
439: times will cause an overabundance of high-mass objects and 
440: a subsquent reduction in low-mass cluster counts.
441: By examining the mass function, we have developed a 
442: simple way of distinguishing these models. This 
443: analysis allows us to discover ways of 
444: further constraining different parameters of DM-DE coupling.
445: 
446: We have found that the statistical uncertainties in the mass 
447: function do not prevent the Dark Energy Survey from detecting 
448: this form of coupled dark matter and dark energy. 
449: Once the selection function for DES is known, 
450: a galaxy formation model can be applied 
451: and a more detailed study will need to take place. 
452: However, 
453: we have found that DES alone will have difficulty  
454: differentiating among different sets of 
455: parameters that control the coupling.
456: 
457: At high masses the statistical uncertainties 
458: may prevent DES from distinguishing between 
459: coupled DM-DE and \lcdm cosmologies with different values 
460: of $\sigma_8$. We can overcome this degeneracy in several ways. 
461: First, missions such as Planck~\citep{Planck} can 
462: independently constrain $\sigma_8$ and $\Omega_{m0}$. 
463: If DES prefers a higher value of $\sigma_8$ through 
464: the mass function, this may 
465: be explained by interacting dark matter and dark energy..
466: Secondly, more detailed measurements of the halo mass function 
467: from projects such as LSST~\citep{LSST} will tightly constrain the mass 
468: function at multiple redshifts. Also, DES itself may detect more 
469: clusters than our estimated 10,000. A universe with coupled DM-DE 
470: will then produce an apparently redshift-dependent $\sigma_8$.
471: 
472: It will also be necessary to compare these mass functions 
473: to those produced by modified General Relativity 
474: (such as those found in~\citet{Stabenau}), as 
475: both theories modify the Poisson equation, and hence 
476: can in principle have similar effects.
477: 
478: \section*{Acknowledgments}
479: The authors acknowledge support under a Presidential Early 
480: Career Award from the U.S. Department of Energy, 
481: Lawrence Livermore National Laboratory (contract B532720).
482: Additional support was provided by a DOE 
483: Computational Science Graduate Fellowship 
484: (DE-FG02-97ER25308) and the National Center for 
485: Supercomputing Applications.
486: The software used in this work was in part developed by the DOE-supported ASC 
487: / Alliance Center for Astrophysical Thermonuclear Flashes at the University of Chicago.
488: \bibliography{ms}		
489: \bibliographystyle{apj}	
490: \nocite{*}
491: 
492: \end{document}
493: 
494: