1: \documentclass[usenatbib]{mn2e}
2: \usepackage{amsmath}
3: %\usepackage{apjfonts}
4: \usepackage{graphicx}
5: \usepackage{natbib}
6: \include{aas_journals}
7: \usepackage{epstopdf}
8: \usepackage{color}
9:
10:
11: \title[effects of correlation on halo formation]{
12: Effects of correlation between merging steps on the global halo
13: formation}
14:
15: \begin{document}
16: \author[Pan et al.]{
17: Jun Pan$^{1}$\thanks{jpan@pmo.ac.cn},
18: Yougang Wang$^{2}$,
19: Xuelei Chen$^{2}$,
20: and Lu\'{i}s Teodoro$^{3}$\\
21: $^1$ The Purple Mountain Observatory, 2 West Beijing Road,
22: Nanjing 210008, China\\
23: $^2$ National Astronomical Observatories, Chinese Academy of Sciences,
24: Beijing 100012, China\\
25: $^3$ Department of Physics and Astronomy, University of Glasgow,
26: Glasgow G12 8QQ, UK
27: }
28: \maketitle
29:
30: \begin{abstract}
31: The excursion set theory of halo formation is modified by
32: adopting the fractional Brownian motion,
33: to account for possible correlation between merging steps.
34: We worked out analytically the conditional
35: mass function, halo merging rate and formation time distribution
36: in the spherical collapse model. We also developed an approximation
37: for the ellipsoidal collapse model and applied it to the calculation of
38: the conditional mass function and the halo formation time distribution.
39: For models in which the steps are positively
40: correlated, the halo merger rate is enhanced when the accreted
41: mass is less than $\sim 25M^*$, while for the negatively correlated
42: case this rate is reduced. Compared with the
43: standard model in which the steps are uncorrelated,
44: the models with positively correlated steps produce more
45: aged population in small mass halos and more younger population
46: in large mass halos, while for the models with negatively correlated steps
47: the opposite is true. An examination of simulation results shows that
48: a weakly positive correlation between successive merging steps
49: appears to fit best. We have also found a systematic
50: effect in the measured mass function due to the finite volume of simulations.
51: In future work, this will be included in the halo model to
52: accurately predict the three
53: point correlation function estimated from simulations.
54:
55: \end{abstract}
56:
57: \begin{keywords}
58: cosmology: theory -- large scale structure of the Universe --
59: galaxies : halos -- methods : analytical
60: \end{keywords}
61:
62: %=========================================================================
63: \section{introduction}
64:
65: The excursion set theory provides a simple and intuitive model for
66: cosmic structure formation. In this theory, as one varies the smoothing
67: scale $R$, the linear density fluctuation $\delta_R$ obtained with
68: smoothing window $W(R)$ forms a one dimensional random walk.
69: It is deemed that the non-linear halo formation and evolution
70: history can be treated with the corresponding excursion set theory:
71: a halo is formed when a pre-set barrier $\delta_c(z)$ is exceeded by
72: a tracjectory of the random walk.
73:
74: It is convinient to take the variance of the random density field
75: $S(R)=\sigma^2(R)$ as the pseudo-time variable.
76: Properties of the random walk rely on the window function
77: used and the nature of
78: the primordial fluctuation. For Gaussian fluctuations
79: and using the sharp k-space filter, the random walk produced by
80: smoothing is a normal Brownian motion
81: (hereafter NBM), i.e. there is no correlation between steps.
82: The density of trajectories $Q(S,\delta)$ passing through $(S,\delta)$
83: then satisfies a diffusion equation
84: \begin{equation}
85: \frac{\partial Q}{\partial S}=\frac{1}{2} \frac{\partial^2 Q}{\partial \delta^2}
86: \end{equation}
87: From $Q(S,\delta)$ one could derive the halo mass function and merger
88: rates, etc. \citep{BondEtal1991,LaceyCole1993}.
89:
90:
91: However, although the N-body simulation results generally agree with
92: the predictions of the excursion set theory, there are significant
93: deviations in the details. The discrepancy is particularly severe in
94: the description of small mass halos. This discrepancy
95: can be partly overcomed by replacing the spherical collapse
96: model with ellisoidal collapse model, i.e. by
97: introducing a moving barrier instead of a fixed barrier imposed on
98: the random walk \citep[e.g.][]{ShethTormen2002}.
99: This practical approach provides a reasonably good fitting formula
100: for halo mass function \citep{WarrenEtal2006}, but its prediction
101: on the formation time distribution of low- and intemediate-mass
102: halo remains unsatisfactory \citep{GiocoliEtal2007}.
103: One suspects that if correlations between steps of the
104: random walk is introduced, the excursion set theory might be improved.
105: Indeed, if the smoothing window function is not a sharp k-space
106: filter but a Gaussian filter or a real space tophat filter, the excursion
107: steps would be correlated.
108:
109:
110: In a previous work by the leading author,
111: the fractional Brownian motion (FBM), the simplest random walk with
112: steps correlated in long range, was introduced to generalize
113: the excursion set theory. The correlation between steps of the
114: random walk was shown to be capable of modifying the final halo
115: mass function in a non-trivial way \citep{Pan2007}.
116: The model presented there was incomplete though, as the solution
117: for ellipsoidal collapses and the treatment of halo merging history were
118: not discussed. These are the topics of the present report.
119:
120: This paper is organized as follows. In section 2 we fit the FBM into
121: the excursion set theory to account for the possible correlated halo formation
122: process. Then in section 3 the diffusion equation of FBM is solved in
123: the spherical collapse model, and we calculate the conditional mass function,
124: halo merger rate and the halo formation time distribution with the
125: modified theory. Treatment to
126: the ellipsoidal collapse is given
127: in section 4, together with a comparison between theoretical predictions
128: and measurements with simulations. The final section contains
129: our conclusions and discussion. We adopt the flat $\Lambda$CDM model with
130: the following set of cosmological parameter values: $\Omega_m=0.3$,
131: $\Omega_\Lambda=0.7$, $h=0.7$ and $\sigma_8=0.9$.
132:
133: \section{Modeling correlated merging steps}
134: \subsection{The diffusion equation}
135:
136:
137: To work out the halo conditional mass function, the central
138: element is the diffusion equation which governs the behavior of
139: the random walk. Such diffusion equation in turn depends on
140: the understanding of the random walk with which the physical
141: problem is concerned with. The conditional mass function is
142: in fact a two-barriers crossing problem of a random walk, which means
143: the key object we shall check is the scaling relation between
144: $\delta(S_1)-\delta(S_0)$ and $S_1-S_0$.
145:
146:
147: The FBM is a random process $X(t)$ on
148: some probability space such that:
149: \begin{enumerate}
150: \item $X(t)$ is continuous and $X(0)\equiv 0$;
151: \item for any $t\ge 0$ and $\tau> 0$, the increment $X(t+\tau)-X(t)$ follows
152: a normal distribution with mean
153: zero and variance $\tau^{2\alpha}$, so that
154: \begin{equation}
155: P\left( X(t+\tau)-X(t)\le x \right)=\frac{\tau^{-\alpha}}{\sqrt{2\pi}}
156: \int_{-\infty}^{x} e^{-u^2/2\tau^{2\alpha} }du \ .
157: \end{equation}
158: \end{enumerate}
159: The parameter $\alpha$ is the Hurst exponent, if $\alpha=1/2$, it is reduced to
160: the normal Brownian motion \citep[c.f.][]{Feder1988}.
161: It is also easy to see that the following is satisfied by the FBM:
162: \begin{equation}
163: \begin{aligned}
164: \langle \left[ X(t+\tau)-X(t) \right]^2 \rangle &=\tau^{2\alpha}\\
165: \langle X(t) \left[X(t+\tau)-X(t)\right] \rangle &=\frac{ (
166: t+\tau)^{2\alpha}-t^{2\alpha}-\tau^{2\alpha}}{2} ,
167: \end{aligned}
168: \label{eq:cov}
169: \end{equation}
170: With this definition of FBM, the trajectory density
171: $Q_\alpha(X, t)$ at time $t$ in interval $(X, X+dX)$
172: follows the diffusion equation \citep[c.f.][]{Lutz2001},
173: \begin{equation}
174: \frac{\partial Q_\alpha}{\partial t}=
175: {\mathcal D}\frac{\partial^2Q_\alpha}{{\partial X}^2}\ ,\
176: {\mathcal D}=\frac{1}{2}\frac{d}{dt}\langle X(t)^2\rangle=\alpha t^{2\alpha-1}.
177: \label{eq:diff}
178: \end{equation}
179:
180: Why invoke FBM?
181: The trajectory of $\delta(S)$ is
182: characterized by properties of the increments
183: $\delta(S_1)-\delta(S_0)$ between any pairs of $(S_0, S_1>S_0)$.
184: If the smoothing window function corresponding to halo definition is not
185: a sharp k-space filter but e.g. a Gaussian or top-hat, it is
186: easy to check that $\langle [\delta(S_1)-\delta(S_0)]^2 \rangle$ is
187: not $ S_1-S_0$ but proportional to $(S_1-S_0)^\alpha$ with a constant
188: $\alpha$ in broad scale range.
189: The scaling relation is the flagging attribute of the FBM\footnote{There
190: are many anomalous random walks such as the fractal time
191: process \citep{Lutz2001} which have the same scaling feature and
192: are classified as sub-diffusion. The FBM is the simplest one of them.},
193: so it is reasonable to install the FBM into the excursion set theory
194: as the simplest approach to analytically inspect halo
195: models constructed with random walks with correlated steps.
196:
197: On the other hand, we know in practice that the boundary, and
198: subsequently the mass of a halo identified in a simulation is
199: rather arbitrary, e.g. the mass picked up is often defined
200: by a halo's virial radius while the mass outside is simply
201: not counted. The $S$ from the virial mass $M$ is very
202: likely not the real place where the particular random walk corresponding
203: to the halo hit the barrier.
204: In fact \citet{CuestaEtal2007} argued that if one replaces the
205: virial mass with the static mass, the mass function agrees with the
206: Press-Schechter formula remarkably well, rather than the Sheth-Tormen
207: one, and the ratio of the virial mass to
208: the static mass depends on redshift and halo mass.
209:
210: While the true scale can be smaller or larger than the $S$ inferred
211: from $R(M)$, one can always parametrize the true scale with $S(R)$ so that
212: $\langle \delta^2 \rangle = S^{2\alpha}$. Taking the approximation that
213: $\alpha$ is a constant not too far away from $1/2$, we can
214: comfortably assume the
215: applicability of the FBM.
216:
217: For an FBM with $\langle \delta^2 \rangle = S^{2\alpha}$, by Eq.~(\ref{eq:cov})
218: , the variance of the increments at two points
219: $(S_0, S_1>S_0>0)$ satisfies
220: \begin{equation}
221: \begin{aligned}
222: \langle \left[ \delta(S_1)-\delta(S_0)\right]^2\rangle = & (S_1-S_0)^{2\alpha} \ ,\\
223: & {\rm with} \ \alpha \in (0, 1)\ , \ S_1>S_0>0\ ,
224: \end{aligned}
225: \end{equation}
226: so that a new trajectory is formed by
227: $\widetilde{\delta}(\Delta S)=\delta(S_1)-\delta(S_0)$
228: along $\Delta S=S_1-S_0$, given any source
229: point $(S_0, \omega_0=\delta_0(S_0))$.
230: Apparently the new random walk also
231: complies with the definition of FBM, its diffusion
232: equation being
233: \begin{equation}
234: \frac{\partial Q_\alpha}{\partial \Delta S}=
235: {\mathcal D}_{\Delta S} \frac{\partial^2 Q_\alpha}{ {\partial \widetilde{\delta}}^2}\ , \quad
236: {\mathcal D}_{\Delta S}=\frac{1}{2}\frac{d\langle {\widetilde{\delta}}^2\rangle}{d\Delta S}
237: =\alpha \Delta S^{2\alpha-1}\ .
238: \label{eq:dfFBM}
239: \end{equation}
240: It is this diffusion equation that ought to be solved to figure out the
241: conditional mass function, which can be further transformed to the familiar
242: diffusion equation of normal Brownian motion
243: \begin{equation}
244: \frac{\partial Q_\alpha }{ \partial \widetilde{S} }= \frac{1}{2}
245: \frac{\partial^2 Q_\alpha }{{\partial \widetilde{\delta}}^2}
246: \label{eq:dfNBM}
247: \end{equation}
248: by the substitution $\widetilde{S}=\Delta S^{2\alpha}$.
249:
250: %-----------------------------------
251: \subsection{The correlation}
252: Imagine a halo is formed by the collapsing condition $\delta(S_1)=\omega_1$
253: at $S_1=\sigma^2(M_1)$ with $M_1$ being the halo mass. After some time the halo
254: merges into a bigger halo with mass $M_0>M_1$ at $S_0=\sigma^2(M_0)<S_1$
255: by another collapsing condition $\delta(S_0)=\omega_0 < w_1$. The following
256: question needs to be addressed:
257: how the formation event by $\omega_0$ at $S_0$ is correlated with
258: the past merging process of $\omega_1-\omega_0$ within $S_1-S_0$?
259:
260: If this process is approximated by the FBM, the correlation function
261: can be easily calculated with Eq.~(\ref{eq:cov}). We have
262: \begin{equation}
263: \begin{aligned}
264: \xi_{0,0\rightarrow 1}&=\langle (\omega_0-0)(\omega_1-\omega_0) \rangle\\ &
265: =\frac{S_1^{2\alpha}}{2}
266: \left[ 1-\left( \frac{S_0}{S_1} \right)^{2\alpha}-\left( 1-\frac{S_0}{S_1}\right)^{2\alpha}
267: \right] \ .
268: \end{aligned}
269: \end{equation}
270:
271: \begin{figure}
272: \resizebox{\hsize}{!}{\includegraphics{corr_steps.eps}}
273: \caption{Correlation functions between successive halo merging steps
274: (Eq.\ref{eq:corr}).}
275: \label{fig:corr}
276: \end{figure}
277:
278: It is more instructive to capture the correlation between two successive merging steps:
279: $\omega_2-\omega_1$ within $S_2-S_1$, and $\omega_1-\omega_0$
280: within $S_1-S_0$ ($S_2>S_1>S_0$, $\omega_2>\omega_1>\omega_0$).
281: The correlation function is similarly
282: \begin{equation}
283: \begin{aligned}
284: \xi_{0\rightarrow1, 1\rightarrow 2}& = \langle (\omega_2-\omega_1)(\omega_1-\omega_0)
285: \rangle = \frac{(S_2-S_0)^{2\alpha}}{2} \\
286: \times & \left[ 1-
287: \left( \frac{S_1-S_0}{S_2-S_0}\right)^{2\alpha} -
288: \left(1-\frac{S_1-S_0}{S_2-S_0}\right)^{2\alpha}\right]\ ,
289: \end{aligned}
290: \label{eq:corr}
291: \end{equation}
292: which is positive when $\alpha>1/2$ and negative if $\alpha<1/2$.
293: Apparently, $\alpha=1/2$ causes null correlation (Fig.~\ref{fig:corr}).
294: Thus a unified paradigm is
295: given by a single parameter controlled process:
296: an anti-persistent FBM predicts that a merging step is anti-correlated
297: with its immediate early occurrence of merging, and persistent FBM models
298: positive correlation.
299:
300:
301:
302: %=============================================================================
303: \section{Spherical collapse}
304:
305: %----------------------------------------------------
306: \subsection{Conditional mass function}
307: It is quite straightforward to solve the diffusion equation
308: in spherical collapse model, we can actually copy
309: the solution in the literatures
310: \citep[e.g.][]{BondEtal1991, LaceyCole1993}.
311: Spherical collapse is equivalent to employ the boundary condition that
312: there is a fixed absorbing barrier of height
313: $\widetilde{\delta}_c=\omega_1-\omega_0 > 0$ to the reformed random walk
314: described by Eq.~(\ref{eq:dfNBM}).
315: The number of trajectories of the reformed walk within
316: $(\widetilde{\delta}, \widetilde{\delta}+d\widetilde{\delta})$ at
317: $\widetilde{S}$ is
318: \begin{equation}
319: Q_\alpha d\widetilde{\delta}=
320: \frac{1}{\sqrt{2\pi \widetilde{S}}} \left[ e^{-\widetilde{\delta}^2/2\widetilde{S}}-
321: e^{-(\widetilde{\delta}-2\widetilde{\delta}_c)^2/2\widetilde{S}} \right]
322: d\widetilde{\delta}\ ,
323: \end{equation}
324: and the number of trajectories absorbed by barrier within
325: $(\widetilde{S}, \widetilde{S}+d\widetilde{S})$ is given by
326: \begin{equation}
327: \begin{aligned}
328: f(\widetilde{S}, \widetilde{\delta}_c)d\widetilde{S}= &- d\widetilde{S}
329: \frac{\partial}{\partial \widetilde{S} }
330: \int_{-\infty}^{\widetilde{\delta}_c} Q_\alpha d \widetilde{\delta} \\
331: = &\frac{\widetilde{\delta}_c}{\sqrt{2\pi}} \widetilde{S}^{-3/2} \exp\left(
332: \frac{-\widetilde{\delta}_c^2}{2\widetilde{S}} \right) d\widetilde{S}\ ,
333: \end{aligned}
334: \end{equation}
335: which directly yields the universal conditional halo mass function
336: \begin{equation}
337: \begin{aligned}
338: f(S_1 & -S_0 , \omega_1-\omega_0)dS_1=f(S_1, \omega_1| S_0, \omega_0)dS_1 \\
339: = &\frac{2\alpha}{\sqrt{2\pi}}
340: \frac{\omega_1-\omega_0}{(S_1-S_0)^{\alpha+1}}
341: \exp\left( -\frac{(\omega_1-\omega_0)^2}{2(S_1-S_0)^{2\alpha}} \right)
342: dS_1\ .
343: \end{aligned}
344: \label{eq:sccmf}
345: \end{equation}
346: If $\alpha=1/2$ we recover the Eq.~(2.15) in \citet{LaceyCole1993}.
347:
348: The conditional mass function is in fact
349: more fundamental than the mass function,
350: since the mass function can be recovered from the conditional mass function
351: by setting the limit $S_0\rightarrow 0, \omega_0 \rightarrow 0$ \citep{Pan2007},
352: \begin{equation}
353: f(S, \delta_c)dS=\frac{2\alpha}{\sqrt{2\pi}} \frac{\delta_c}{S^{\alpha+1}}
354: \exp\left( {-\frac{\delta_c^2}{2 S^{2\alpha}} } \right)dS\ .
355: \label{eq:scmf}
356: \end{equation}
357:
358: Thus, prompted by the relation between mass function and conditional
359: mass function, we recognize that there is a systematical effect due
360: to the finite volume of simulations. This can lead to an
361: underestimate of the halo mass function in the large mass regime
362: (more details are discussed in Appendix A).
363: Recall that the halo model tends to over-estimate the amplitude
364: of the three point correlation function of dark matter, which can
365: be corrected (at least partly) by
366: applying certain arbitrary high mass cut-off to the halo mass
367: function \citep{WangEtal2004, FosalbaPanSzapudi2005}.
368: To predict the three point correlation function of
369: a simulation, a better approach would be to use the halo mass function
370: of Eq.~(\ref{eq:simmf}) to include the finite volume effects.
371:
372: %--------------------------------------------------------------------------------------
373: \subsection{Merger rate}
374:
375: \begin{figure*}
376: \resizebox{\hsize}{!}{\includegraphics{mr.eps}\includegraphics{ar.eps}}
377: \caption{Halo merger rate (left panel) and accretion rate (right panel) computed
378: with Eq.~(\ref{eq:mr}). The accretion rate is simply
379: $\Delta M/M_1 \times d^2p/d\ln\Delta M/d\ln t$. Hereafter
380: $M^*=10^{13} M_{\sun}$.}
381: \label{fig:mr}
382: \end{figure*}
383:
384: Now we will
385: consider the merger rate function in line with \citet{LaceyCole1993}.
386: For a trajectory which has experienced a first up-crossing
387: over $\omega_1$ at $S_1$, the conditional probability of having
388: a first up-crossing over $\omega_2 (\omega_2 < \omega_1)$ at $S_2 (S_2 < S_1)$
389: in the interval $dS_2$ is given by the Bayes formula,
390: \begin{equation}
391: \begin{aligned}
392: f(S_2, \omega_2|& S_1,\omega_1)dS_2=\frac{f(S_1, \omega_1|S_2, \omega_2)dS_1 f(S_2, \omega_2)dS_2}
393: {f(S_1, \omega_1)dS_1} \\
394: = &\frac{2\alpha}{\sqrt{2\pi}}\frac{\omega_2(\omega_1-\omega_2)}{\omega_1}\left[
395: \frac{S_1}{S_2(S_1-S_2)} \right]^{\alpha+1} \\
396: \times & \exp \left[ - \frac{1}{2}\left(\frac{(\omega_1-\omega_2)^2}{(S_1-S_2)^{2\alpha}} +
397: \frac{\omega_2^2}{ S_2^{2\alpha}}
398: -\frac{\omega_1^2}{S_1^{2\alpha}}\right)\right] \ ,\\
399: (\omega_1>\omega_2\ , & \ S_1>S_2)\ .
400: \end{aligned}
401: \end{equation}
402: This function is interpreted as the probability that a halo of mass
403: $M_1$ at time $t$ will merge to build a halo of mass between $M_2$ and
404: $M_2+dM_2$ at time $t_2>t_1$. The mean transition rate is obtained from it
405: by setting $t_2\rightarrow t_1$ (equivalently $\omega_2\rightarrow \omega_1=\omega$),
406: \begin{equation}
407: \begin{aligned}
408: \frac{d^2 p(S_1\rightarrow S_2|\omega)}{dS_2 d\omega} & dS_2 d\omega=
409: \frac{2\alpha}{\sqrt{2\pi}}
410: \left[ \frac{S_1}{S_2(S_1-S_2)} \right]^{\alpha+1} \\
411: \times & \exp\left[ -\frac{\omega^2}{2}
412: \frac{S_1^{2\alpha} - S_2^{2\alpha}}{S_1^{2\alpha}S_2^{2\alpha}} \right]
413: dS_2 d\omega \ .
414: \end{aligned}
415: \end{equation}
416: Therefore the merger rate, i.e. the probability that a halo
417: of mass $M_1$ accretes a clump of mass $\Delta M=M_2-M_1$ within
418: time $d\ln t$ (corresponding to $d\omega$), is
419: \begin{equation}
420: \begin{aligned}
421: d^2p&( M_1\rightarrow M_2|t)/d\ln \Delta M d\ln t \\
422: &= 2 \sigma(M_2) \Delta M \left| \frac{d\sigma_2}{dM_2} \right|
423: \left| \frac{d\omega}{d\ln t} \right|
424: \frac{d^2p(S_1\rightarrow S_2|\omega)}{dS_2d\omega}\\
425: &=2\alpha\sqrt{\frac{2}{\pi}}\frac{\Delta M}{M_2}
426: \left| \frac{d\ln\delta_c}{d\ln t} \right|
427: \left| \frac{d\ln \sigma_2}{d\ln M_2} \right|
428: \frac{\delta_c(t)}{\sigma_2^{2\alpha}} \\
429: &\times \left( \frac{\sigma_1^2}{\sigma_1^2-\sigma_2^2} \right)^{\alpha+1}
430: \exp\left[ -\frac{\delta_c^2}{2}
431: \left( \frac{1}{\sigma_2^{2\alpha}}- \frac{1}{\sigma_1^{2\alpha}} \right)
432: \right] \ .
433: \end{aligned}
434: \label{eq:mr}
435: \end{equation}
436:
437: In Figure~\ref{fig:mr} the halo merger rates of different masses as predicted by
438: Eq.~(\ref{eq:mr}) are plotted for comparison. Three models are presented:
439: an anti-persistent FBM of $\alpha=0.4$, the standard normal Brownian motion
440: of $\alpha=0.5$ and a persistent FBM of $\alpha=0.6$.
441:
442: For halo progenitors of the typical mass scale $M_1=M^*$, the merger rate and
443: halo accretion rate increases with increasing $\alpha$ at small $\Delta M/M_1$,
444: but at very large mass ratios ($\Delta M/M_1>25$) the case is reversed.
445:
446: This is also true for more massive progenitor masses ($M_1=10M^*$), but here
447: the transition point in the mass ratio is smaller: $\Delta M/M_1 =3$.
448:
449: For the less massive progenitors, $M_1=0.1 M^*$, the transition point is very
450: high, that in the whole plotted range (up to $\Delta M/M_1=100$), the models
451: with greater Hurst exponent always have greater merger and accretion rate.
452:
453: %---------------------------------------------------------
454: \subsection{Halo formation time distribution}
455:
456: \begin{figure*}
457: \resizebox{\hsize}{!}{
458: \includegraphics{halo_cft.eps}
459: \includegraphics{halo_ft.eps}}
460: \caption{Halo formation time distributions for halos at $z=0$
461: with mass of $0.1$, 1 and $10M^{\star}$, which are predicted by the modified
462: excursion set theory of different $\alpha$ assuming spherical collapse. The
463: left panel is the cumulative distribution.}
464: \label{fig:ft}
465: \end{figure*}
466:
467: The formation time (redshift) of a halo is the time (redshift) when its major
468: progenitor contains half of the halo mass. According to
469: the counter argument of LC93, the cumulative distribution of halo formation
470: time, i. e. the probability that a halo of mass $M_0$ at redshift
471: $z_0$ is formed at redshift larger than $z_f$, can be acquired
472: by
473: \begin{equation}
474: P(>z_f)=\int_{S_0}^{S_h}\frac{M_0}{M(S_1)}
475: f\left(S_1,\delta_c(z_f)|S_0,\delta_c(z_0)\right)d S_1
476: \end{equation}
477: where $S_0=S(M_0)$ and $S_h=S(M_0/2)$. The halo formation time
478: distribution is simply given by the differentiation
479: $-dP(z_f)/dz_f$.
480:
481: Figure~\ref{fig:ft} shows the predictions of the halo formation
482: redshift distribution for difference masses.
483: In general smaller halos have more extended formation time distribution
484: than that of larger halos, %this is
485: in agreement with the results of
486: \citet{LinEtal2003}. For $\alpha>1/2$, the halo
487: formation redshift distribution is more concentrated than the
488: $\alpha=1/2$ case, and for $\alpha<1/2$ the halo formation
489: redshift distribution is more extended. This seems to be in
490: accordance with our finding that the merger rate is greater for
491: larger $\alpha$.
492:
493: As can be seen, the detailed
494: effects of incorporating correlation between merging steps are
495: intricate, varying with the halo mass. Impact on the halo formation time
496: distribution is small for those halos with masses around $M^*$,
497: and it becomes apparent only when halo
498: mass deviates significantly from $M^*$. The impact is
499: also different for halos with
500: small masses and large masses. Compared with the results given by
501: the standard excursion set theory ($\alpha=1/2$),
502: \begin{enumerate}
503: \item $M \ll M^*$:
504: negative correlation ($\alpha<1/2$) shifts the halo formation time distribution
505: curve to the side of smaller $z_f$, which means that the younger halos are more
506: abundant and the older ones are less abundant; whilst for
507: positive correlation ($\alpha>1/2$) the older ones are more abundant;
508: \item $M > M^*$: the impact of the correlation is opposite to the case of small
509: halo mass, negative correlation ($\alpha<1/2$) boosts more halos
510: to form at an earlier time while positive correlation ($\alpha>1/2$)
511: induces more halos to form at an later time.
512: \end{enumerate}
513:
514: %==================================================================
515: \section{Ellipsoidal collapse}
516:
517: \subsection{Moving barriers}
518: The spherical collapse model is perhaps accurate at high redshift e.g.
519: the re-ionization era. However, as shown by simulations, at low
520: redshift the collapse of a clump of mass is ellipsoidal.
521: For the excursion set theory, the general collapse condition for
522: halo formation is not a constant $\delta_c$ any more, but a
523: moving barrier ${\mathcal B}(S)$.
524:
525: Imposing a moving barrier on the random walk of
526: $(\widetilde{\delta},\Delta S)$ described by Eq.~(\ref{eq:dfFBM})
527: is equivalent to the case of placing a constant barrier boundary condition to
528: the diffusion equation with an extra drifting term \citep{Zentner2007}. In
529: our case of FBM, the Fokker-Planck equation turns out to be
530: \begin{equation}
531: \frac{\partial Q_\alpha}{\partial \Delta S}=
532: \alpha {\Delta S}^{2\alpha-1} \frac{\partial^2 Q_\alpha}{ \partial \widetilde{\delta}^2}
533: +\frac{\partial \Delta{\mathcal B}}{\partial \Delta S}
534: \frac{\partial Q_\alpha}{\partial \widetilde{\delta}}
535: \label{eq:dfECFBM}
536: \end{equation}
537: where $\Delta {\mathcal B}={\mathcal B}(S_1)-{\mathcal B}(S_0)$. Unfortunately
538: this equation can be solved analytically only for a few very
539: special cases (see Appendix B).
540:
541: %----------------------------------------------------------------------
542: \subsection{Conditional mass function: the Sheth-Tormen approximation}
543:
544: \begin{figure*}
545: \resizebox{\hsize}{!}{\includegraphics{ec_cmf.eps}}
546: \caption{Conditional mass functions of progenitor halos at redshift
547: $z_1=$0.5, 1, 2, 4 (identical for each row) for three different halo masses
548: $M_0$ (each column has the same mass). Histograms are the results
549: from the Millennium Simulation given by
550: \citet{ColeEtal2008}, while lines are predictions by Eq.~(\ref{eq:ecCMF})
551: of three different values of $\alpha$ as indicated.}
552: \label{fig:ecCMF}
553: \end{figure*}
554:
555: Although we do not have analytical solution of Eq.~(\ref{eq:dfECFBM}),
556: it could be solved with the numerical method of \citet{ZhangHui2006}. However,
557: this is inconvenient for general explorations. Here
558: we borrow the pragmatic approach of \citet{ShethTormen2002},
559: the conditional mass function is then approximated by
560: \begin{equation}
561: \begin{aligned}
562: f(S_1-S_0)dS_1& =dS_1 \frac{2\alpha}{\sqrt{2\pi}}
563: \frac{\left|T(S_1, z_1|S_0,z_0)\right|}{(S_1-S_0)^{\alpha+1}}\\
564: \times &
565: \exp\left\{ - \frac{\left[ {\mathcal B}(S_1, z_1)-{\mathcal B}(S_0,z_0) \right]^2}
566: {2 (S_1-S_0)^{2\alpha}} \right\}\ ,
567: \end{aligned}
568: \label{eq:ecCMF}
569: \end{equation}
570: in which
571: \begin{equation}
572: T(S_1|S_0)=\sum_{n=0}^5 \frac{(S_0-S_1)^n}{n!}
573: \frac{\partial^n \left[{\mathcal B}(S_1)-{\mathcal B}(S_0) \right]}{{\partial S_1}^n}
574: \label{eq:ecT}
575: \end{equation}
576: and
577: \begin{equation}
578: {\mathcal B}(S, z)=\sqrt{q} \delta_c(z) \left[ 1+ b \left( \frac{S}{q \delta_c^2(z)}
579: \right)^\gamma \right]
580: \end{equation}
581: with $q=0.707$, $b=0.485$ and $\gamma=0.615$
582: \citep{ShethMoTormen2001, ShethTormen2002}.
583:
584: Inspired by the results given in Appendix B,
585: it is probably more appropriate to use
586: \begin{equation}
587: T=\sum_n \frac{ (-\widetilde{S})^n}{n!}
588: \frac{\partial^n \Delta {\mathcal B}}{{\partial \widetilde{S}}^n}
589: \end{equation}
590: with $\widetilde{S}=(S_1-S_0)^{2\alpha}$ in Eq.~\ref{eq:ecCMF}.
591: Testing the two approximation methods numerically, we found that
592: the actual difference between the two is in fact very small.
593: So we opted to use the numerically simpler
594: Eq.~(\ref{eq:ecT}).
595:
596: In figure~\ref{fig:ecCMF} we compare our prediction on
597: the conditional mass function with that measured from the
598: Millennium Simulation \citep{ColeEtal2008} \footnote{The data is publicly
599: available at {\em http://star-www.dur.ac.uk/$\sim$cole/merger\_trees}} at
600: $z_0=0, z_1=0.5, 1, 2, 4$, for halos of masses $M_0$=0.1, 3.16, 100$M^*$.
601: It is indeed very interesting to notice that the performance of
602: the original Sheth-Tormen approach (Eq.~\ref{eq:ecCMF} of $\alpha=1/2$) is
603: fairly good though not as remarkable as the fitting function
604: of \citet{ColeEtal2008}. It is surprising that
605: \citet{ColeEtal2008} extrapolated the Sheth-Tormen mass function for
606: the conditional mass function in ellipsodal collapse model, in spite of
607: the one actually proposed in \citet{ShethTormen2002}.
608:
609: From Figure~\ref{fig:ecCMF} we can see that the effects of correlation
610: between merging steps gradually decrease with time, with halo formation
611: becoming sensitive to the correlation at high redshift.
612: Except in cases of high
613: redshifts and of extremely high halo mass, it appears that
614: positive correlation ($\alpha>1/2$) is preferred by the simulation
615: in $\Lambda$CDM universe. Nevertheless, one must be cautious to
616: this result, as the measurement presented is rather
617: crude and lacks error bars. Its accuracy is not sufficient to justify with
618: confidence whether the correlation between merging steps is positive,
619: negative or zero.
620:
621: %------------------------------------------------------------
622: \subsection{Halo formation time distribution}
623: \begin{figure*}
624: \resizebox{\hsize}{!}{
625: \includegraphics{echalo_cft.eps}
626: \includegraphics{echalo_ft.eps}}
627: \caption{Halo formation time distributions for halos of masses
628: 0.1, 1 and 10$M^{\star}$ at $z=0$, which are predicted by the modified
629: excursion set theory of different $\alpha$ in ellipsoidal collapse model.
630: The left panel is the cumulative distribution.}
631: \label{fig:ecft}
632: \end{figure*}
633:
634: The halo formation time distribution is recalculated with the modified
635: excursion set theory in ellipsoidal collapse model and the results
636: displayed in Figure~\ref{fig:ecft}. Although the halo formation time distributions
637: are modulated significantly by the non-constant barrier,
638: its dependence on $\alpha$ does not change, and demonstrates similar trends as
639: the spherical models predicts:
640: positive correlation results in more old small mass halos and more young
641: large mass halos, while the effects of the
642: negative correlation are to
643: the contrary.
644:
645: \citet{GiocoliEtal2007} calculated the cumulative distribution of halo formations
646: time and compared with GIF2 simulation. They found that the median formation
647: redshift of halos of mass less than $M^*$ extracted from the simulation
648: is larger than that predicted by the Sheth-Tormen formula, even after some
649: numerical improvement. Combined with our discovery of effects of correlation,
650: it seems that the discrepancy can be explained by the existence
651: of a positive correlation between adjacent merging steps.
652:
653:
654: %================================================================
655: \section{Conclusions and discussion}
656: In this work the FBM is formally incorporated in the excursion
657: set theory, to account for possible correlations between merging steps
658: of halo formation. Such modification
659: is minimal but provides a unified theoretical frame to investigate the
660: effects of correlation. More specifically, the correlation depends on the
661: Hurst exponent $\alpha$, for $\alpha<1/2$, the correlation
662: between adjacent merging steps is negative, for $\alpha>1/2$ the
663: correlation is positive, while $\alpha=1/2$ is reduced to the standard case
664: of null correlation.
665:
666: We calculated the conditional mass functions in models
667: of the spherical collapse and the ellipsoidal collapse. Different
668: collapsing conditions do not change the relative effects of
669: correlations. It is revealed that for a typical progenitor of $M^*$,
670: a positive correlation will boost
671: the merger rate when the mass to accrete is less than
672: $\sim 25M^*$, and negative correlation will significantly reduce
673: such a rate.
674: Accordingly we find that compared with the
675: standard excursion set theory without correlation,
676: positive correlation increases the aged population for small mass halos
677: but produces more young members for large mass halos; while the negative
678: correlation projects opposite effects to the halo formation. The same trend is
679: also seen for other progenitor masses, although the transition mass ratio
680: is greater for small progenitor masses and smaller for greater progenitor
681: masses.
682:
683:
684: The results of the modified excursion set theory in ellipsoidal
685: collapse model are checked with those measurements from simulations of
686: \citet{GiocoliEtal2007} and \citet{ColeEtal2008}. The comparison
687: indicates that there is a sign of weak positive
688: correlation, although not yet of significant confidence, as it is
689: limited by the accuracy and dynamic ranges of their analysis.
690: The conclusion appears to be in conflict with the claim of
691: \citet{Pan2007}, but note that in that work
692: the mass function was calculated with the spherical
693: collapse model, which is not valid at most redshifts of concern.
694: Besides, the main purpose of \citet{Pan2007} is to show that
695: the effect of correlation is not trivial.
696:
697: The accuracy of the Sheth-Tormen approximation to the ellipsoidal
698: collapse in the modified (FBM) excursion set theory has not been
699: checked against rigorous numerical solutions. However,
700: since the actual correlation is weak,
701: i.e. $\alpha$ only deviates from $1/2$ slightly, the
702: approximation suffices for general qualitative discussion.
703: Calibration and subsequent improvement of the approximation
704: with numerical computation is of course needed
705: for more precise analytical modeling and implementation
706: in semi-analytical models of galaxy formation.
707:
708: In this paper we explored the effects of correlations between
709: mergers on the global halo formation process. It is of
710: interests to actually check if such a correlation will
711: lead to the assembly bias discovered by \citet{GaoEtal2005}, and whether it
712: can explain the dependence of halo formation on the large scale environment.
713:
714: A by product of our work is the discovery of a systematic effect
715: in the mass function due to the finite volume of simulation. It explains
716: why applying a high mass cut-off to the mass function can improve
717: the halo model prediction on three point correlation function. A better
718: approach in the halo model calculation of three point function
719: is to shift the theoretical mass function
720: according to the formulae given in Appendix A, this may lead to better
721: agreement between simulation and analytical results.
722:
723:
724:
725: %================================================================
726: \section*{Acknowledgment}
727: The authors acknowledge stimulating discussions with
728: Pengjie Zhang, Jun Zhang, Weipeng Lin and Longlong Feng.
729: JP is supported by the China Ministry of Science \& Technology
730: through 973 grant of No. 2007CB815402 and the NSFC through
731: grants of Nos. 10643002, 10633040. YGW and XLC
732: are supported by NSFC via grants 1052314, 10533010, the CAS
733: under grant KJCX3-SYW-N2, and the Ministry of
734: Science \& Technology via 973 grant of 2007CB815401.
735: LT acknowledges the financial support of the Leverhulme Trust (UK).
736:
737:
738:
739:
740: %================================================================
741: %%\bibliography{cc}
742: %%\bibliographystyle{mn2e}
743: \input ms.bbl
744: %================================================================
745: %=======================================================================
746:
747: \appendix
748: \section{Finite volume effects in small-box simulations}
749:
750: N-body simulations are always performed in cubic boxes of finite size
751: $L_{sim}$. By definition, the density contrast of whole the simulation box is
752: zero, but the variance is not. Henceforth the measured mass function
753: of simulation is
754: conditional, determined by $f(S_1=S, \omega_1=\delta_c| S_0(L_{sim})\neq 0, \omega_0=0)$.
755: From Eq.~(\ref{eq:sccmf}) the explicit halo mass function of a simulation is then
756: \begin{equation}
757: \begin{aligned}
758: f(\sigma, L_{sim})& d\ln \sigma = \frac{4\alpha}{\sqrt{2\pi}}
759: \frac{\delta_c}{(\sigma^2-\sigma_0^2)^\alpha}
760: \frac{\sigma^2}{\sigma^2-\sigma_0^2} \\
761: & \times \exp\left[ -\frac{\delta_c^2}{2(\sigma^2-\sigma_0^2)^{2\alpha}}\right]
762: d\ln \sigma\ ,
763: \end{aligned}
764: \label{eq:simmf}
765: \end{equation}
766: where $$S=\sigma^2=(2\pi)^{-3}\int_{2\pi/L_{sim}}^\infty P(k)\widetilde{W}4\pi k^2dk,$$ and
767: $$S_0=\sigma_0^2(L_{sim})=(2\pi)^{-3}\int_{2\pi/L_{sim}}^\infty P(k)
768: \widetilde{W}(k, L_{sim})4\pi k^2dk.$$
769: The infrared cutoff in the integration for $\sigma^2$ accounts for the
770: deficiency of large scale power in finite box, which has already
771: been identified as a systematic effect \citep[e.g.][]{ReedEtal2007}.
772:
773: \begin{figure}
774: \resizebox{\hsize}{!}{\includegraphics{finiteV.eps}}
775: \caption{Deviation of mass functions in finite-sized simulations
776: (Eq.~\ref{eq:simmf}) to Eq.~(\ref{eq:scmf}). For the fiducial
777: model $\alpha=0.436$.
778: $L_{sim}$ is the size of the cubic box for simulation. $\sigma_0$ is
779: calculated with infrared cutoff in integration (see text), and a
780: spherical top-hat window function in real space of radius
781: $L_{sim}/(4\pi/3)^{1/3}$. Deviation is negligible in
782: regime $-\ln \sigma <0.2$.}
783: \label{fig:simmf}
784: \end{figure}
785:
786: The difference between Eq.~(\ref{eq:simmf}) and Eq.~(\ref{eq:scmf}) is shown in
787: Fig.~\ref{fig:simmf} for different box sizes. The fiducial model has
788: $\alpha=0.436$, by which the spherical model provides good approximation to
789: the mass function of simulations in high mass region \citep{Pan2007}. We
790: can see that the finite volume effect becomes apparent only when
791: $-\ln \sigma >0.2$, i.e. in high mass regime, and the resulting
792: fractional error
793: quickly drops down to $<10\%$ for simulations of $L_{sim}>100h^{-1}$Mpc.
794:
795: Nowadays it is a very common practice to use
796: simulations with boxes as small as $< 50h^{-1}$Mpc to
797: extend the dynamic range. The finite volume
798: effect demonstrated in Fig.~\ref{fig:simmf} has to be taken into account
799: for precision measurement. The necessary
800: correction is actually very simple: the true measured mass
801: function $f(S)dS$ can be recovered
802: by shifting the raw mass function $f(S|S_0)dS$ along $S$ by $S_0$. Note
803: that there is a cosmic variance for $S_0$, for each individual
804: realization of simulation one may need to calculate the individual
805: $S_0$ from the realization \citep{ReedEtal2007, LukicEtal2007}.
806:
807: %---------------------------------------------------------------
808: \section{{\em pseudo-linear} barrier:
809: ${\mathcal B}=\delta_c+\beta S^{2\alpha}$}
810: Starting with Eq.~(\ref{eq:dfECFBM}), let $S_0=0$ and $S_1=S>0$,
811: a variable substitution $\widetilde{S}=S^{2\alpha}$ yields
812: \begin{equation}
813: \frac{\partial Q_\alpha}{\partial \widetilde{S}}=
814: \frac{1}{2}\frac{\partial^2 Q_\alpha}{ \partial \widetilde{\delta}^2}
815: +\frac{1}{2\alpha S^{2\alpha-1}}\frac{\partial {\mathcal B}}{\partial S}
816: \frac{\partial Q_\alpha}{\partial \widetilde{\delta}}
817: \end{equation}
818: If the barrier is ${\mathcal B}=\delta_c+\beta S^{2\alpha}$, the above
819: nonlinear Fokker-Planck equation is further simplified to
820: \begin{equation}
821: \frac{\partial Q_\alpha}{\partial \widetilde{S}}=
822: \frac{1}{2}\frac{\partial^2 Q_\alpha}{ \partial \widetilde{\delta}^2}
823: +\beta \frac{\partial Q_\alpha}{\partial \widetilde{\delta}}\ ,
824: \end{equation}
825: which is the well known problem of a linear
826: barrier upon the normal Brownian motion \citep{Sheth1998, Zentner2007}.
827:
828: The probability of a trajectory that has the first barrier passage
829: within $(\widetilde{S}, \widetilde{S}+d\widetilde{S})$ in the solution
830: is given by
831: \begin{equation}
832: f(\widetilde{S})d\widetilde{S}=\frac{\delta_c}{\sqrt{2\pi} {\widetilde{S}}^{3/2}}
833: \exp\left[ - \frac{(\beta \widetilde{S}+\delta_c)^2}{2 \widetilde{S}} \right]
834: d\widetilde{S}\ ,
835: \end{equation}
836: from which the universal mass function is easily obtained
837: \begin{equation}
838: f(\sigma)d\ln \sigma=\frac{4\alpha}{\sqrt{2\pi}}\frac{\delta_c}{\sigma^{2\alpha}}
839: \exp\left[ - \frac{(\delta_c+\beta \sigma^{4\alpha})^2}{2\sigma^{4\alpha}}\right]
840: d\ln \sigma\ .
841: \label{eq:linBmf}
842: \end{equation}
843:
844: However, there is no analogous analytical expression for the
845: conditional mass function under this type of barrier, as the
846: barrier does not have the symmetry that a true linear barrier possesses:
847: for any pair of points at $S_0$ and $S_1>S_0$
848: \begin{equation}
849: {\mathcal B}_1-{\mathcal B}_0=\beta (S_1^{2\alpha}-S_0^{2\alpha})\neq
850: \beta(S_1-S_0)^{2\alpha}\ .
851: \end{equation}
852: This is why we call it {\em pseudo-linear}.
853:
854: \end{document}
855: