Issue |
A&A
Volume 514, May 2010
|
|
---|---|---|
Article Number | A46 | |
Number of page(s) | 14 | |
Section | Cosmology (including clusters of galaxies) | |
DOI | https://doi.org/10.1051/0004-6361/200912636 | |
Published online | 12 May 2010 |
Mass function and bias of dark matter halos for non-Gaussian initial conditions
P. Valageas
Institut de Physique Théorique, CEA Saclay, 91191 Gif-sur-Yvette, France
Received 4 June 2009 / Accepted 31 January 2010
Abstract
Aims. We revisit the derivation of the mass function and the bias of dark matter halos for non-Gaussian initial conditions.
Methods. We use a steepest-descent approach to point out that
exact results can be obtained for the high-mass tail of the halo mass
function and the two-point correlation of massive halos. Focusing on
primordial non-Gaussianity of the local type, we check that these
results agree with numerical simulations.
Results. The high-mass cutoff of the halo mass function takes
the same form as the one obtained from the Press-Schechter formalism,
but with a linear threshold
that depends on the definition of the halo (i.e.
for
a nonlinear density contrast of 200). We show that a simple
formula, which obeys this high-mass asymptotic and uses the fit
obtained for Gaussian initial conditions, matches numerical simulations
while keeping the mass function normalized to unity. Next, by deriving
the real-space halo two-point correlation in the spirit of Kaiser
(1984, ApJ, 284, L9) and taking a Fourier transform, we obtain good
agreement with simulations for the correction to the halo bias,
,
due to primordial non-Gaussianity. Therefore, neither the halo mass function nor the bias require an ad-hoc parameter q (such as
), provided one uses the correct linear threshold
and pays attention to halo displacements. The nonlinear real-space
expression can be useful for checking that the ``linearized'' bias is a
valid approximation. Moreover, it clearly shows how the baryon acoustic
oscillation at
Mpc is amplified by the bias of massive halos and modified by primordial non-Gaussianity. On smaller scales,
30 <x< 90 h-1 Mpc, the correction to the real-space bias roughly scales as
.
The low-k
behavior of the halo bias does not imply a divergent real-space
correlation, so that one does not need to introduce counterterms
that depend on the survey size.
Key words: gravitation - methods: analytical - large-scale structure of Universe
1 Introduction
Standard single-field slow-roll inflationary models predict a nearly
scale-invariant and Gaussian spectrum of primordial curvature
fluctuations (e.g., Bartolo et al. 2004). This agrees with current observations of the cosmic microwave background (CMB) anisotropies (Komatsu et al. 2009) and of large-scale structures (Slosar et al. 2008). Nevertheless, several inflationary models predict a
potentially observable level of non-Gaussianity (e.g., Bartolo et al. 2004,
for a review), so that constraining or detecting primordial
non-Gaussianity is an important task for current cosmological studies.
This would allow one to rule out some of the many inflationary models
that have already been proposed. In particular, in many
cases, the non-Gaussianity is of the local type, meaning that it only
depends on the local value of Bardeen's potential .
That is, the latter can be decomposed as
where






The effects of primordial non-Gaussianity on large-scale structures can be seen, for instance, through the mass function of virialized halos, especially in the high-mass tail as the steep falloff magnifies the sensitivity to initial conditions (Lucchin & Matarrese 1988; Colafrancesco et al. 1989; Grossi et al. 2007; Maggiore & Riotto 2009). This allows using the X-ray luminosity function of clusters to constrain the amount of non-Gaussianity (Amara & Refregier 2004).
A second probe of non-Gaussianity is provided by the clustering of
these halos, as measured through their many-body correlations.
In particular, the halo two-point correlation can be significantly
increased if the underlying primordial density field is non-Gaussian
(Grinstein & Wise 1986). More specifically, Dalal et al. (2008) have recently shown that primordial non-Gaussianity of the local type (1)
gives rises to a strongly scale-dependent bias on large scales, whereas
in the Gaussian case the bias is roughly constant in this range. Thus,
at linear order over
they obtain in Fourier space a correction of the form
where bM(k,0) is the Gaussian-case bias (defined as the ratio of the halo and matter power spectra, bM2(k)=PM(k)/P(k), for objects of mass M). Here




In this article, following a previous work devoted to the Gaussian case (Valageas 2009b),
we revisit the derivations of the halo mass function and of the bias
for primordial non-Gaussianity. Although we focus on the local
type (1), our approach also
applies to any non-Gaussian model where
Bardeen's potential can be written as the sum of linear and quadratic
terms over an auxiliary Gaussian field, that is, where
becomes a convolution kernel. (It also extends to cases that contain higher order terms and multiple Gaussian fields.)
After introducing our notations and the quantities needed for our calculations in Sect. 2, we consider the halo mass function in Sect. 3.
Here, our aim is to argue that the exponential cutoff of the high-mass
tail can be obtained exactly from a saddle-point approach. This is
equivalent to the saddle-point computation of Matarrese et al. (2000),
which is often used to model the non-Gaussian halo mass function.
However, with a different treatment, we simultaneously derive the
linear density profile of this saddle point, which allows us to check
that the latter is almost insensitive to
primordial non-Gaussianity, so that shell crossing is not
amplified and exact results can be obtained provided one uses the
correct linear density threshold, rather than the usual one. We also
propose a simple recipe to match the dependence on
of the high-mass tail while keeping the mass function normalized to unity. Then, in Sect. 4 we consider the two-point correlation of dark matter halos in real space, following the spirit of Kaiser (1984).
Next, taking a Fourier transform we obtain the halo bias in
Fourier space. Here, our aim is to show that one does not need to
introduce free parameters to match the results of numerical
simulations. Moreover, the nonlinear real-space expression is of
interest by itself and it also allows one to check whether the
``linearized'' bias is valid on the range of interest. Finally, we
conclude in Sect. 5.
2 Non-Gaussian initial conditions
We focus in this paper on non-Gaussianities of the local type, where Bardeen's potential
is of the form (1), with
a Gaussian random field. On scales smaller than the Hubble
radius,
equals minus the Newtonian gravitational potential and the Poisson equation gives in Fourier space (Slosar et al. 2008)
where


Note that we define





we can write the linear density field at redshift z as
with
This reads in real space as
where the kernel


as long as the system remains statistically homogeneous, as for the local model (1). The real-space and Fourier-space kernels are related by (note the different normalization from (4))
The relationships (6) and (8) describe any homogeneous model where the Bardeen potential can be expressed as the sum of linear and quadratic terms over some Gaussian field. Thus, our analytical results also apply to other ``

For
we recover Gaussian initial conditions,
,
with a linear density power spectrum
and a two-point linear density correlation
As usual, it is convenient to introduce the smoothed linear density contrast,


with a top-hat window that reads in Fourier space as
Then, in the linear regime, the cross-correlation of the smoothed linear density contrasts on scales q1 and q2 and positions


In particular,





In the following sections, where we define dark matter halos as
spherical overdensities, we shall need the average of the kernel
over spherical cells, weighted by the linear correlation (12). Thus, omitting the superscript
for simplicity, we define the quantity
where the spheres of volumes V, V1 and V2, and radii q, q1 and q2, are centered on the points






Thanks to statistical homogeneity and isotropy, the kernel


which does not depend on the position



and
Their Fourier transforms with respect to the separation

and
3 Mass function of dark matter halos
We now extend the analysis of Valageas (2009b) to obtain the mass function of dark matter halos for non-Gaussian initial conditions.
3.1 Rare-event saddle point for the density distribution
In a fashion similar to Valageas (2009b), we note that the exponential falloffs of the high-mass tail of the halo mass function n(M), and of the overdensity tail of the linear density contrast distribution
,
can be exactly obtained from the constrained maximum,
Here we introduced the probability distribution

In the Gaussian case,












![[*]](/icons/foot_motif.png)


and
where the function








In the Gaussian case, the constrained weight (23) simply reads as
,
and we recover the exponential tail of the usual Press-Schechter mass function (Press & Schechter 1974), except that the standard threshold
must be replaced by
,
with
for
.
In the non-Gaussian case, that is for
,
we can compute the weight (23) by using a Lagrange multiplier
.
Thus, we define the action
,
where
![$\delta_{{\rm L}q}[\chi]$](/articles/aa/full_html/2010/06/aa12636-09/img135.png)






![]() |
(28) |
whence, using Eq. (8),
Differentiating the action (27) with respect to

whence
Next, we solve the system (29)-(31) as a perturbative series over the non-Gaussianity kernel


At first order we obtain
where the quantity fq;qq is given by Eq. (18).
![]() |
Figure 1:
The radial profile (36, 37) of the linear density contrast
|
Open with DEXTER |
From Eqs. (32)-(35) we also obtain the radial linear density profile of the saddle-point, up to first order,
We can check that at q'=q it verifies the constraint (30), and at order zero we recover the Gaussian profile (Valageas 2009b). Equations (36, 37) give the integrated density profile, that is,


![]() |
(38) |
whence
where f0;qq(q') and fq;0q(q') are given by Eqs. (19, 20).
We show in Fig. 1 the integrated linear density profile (36, 37) obtained for the masses M=1011 and
,
for a
CDM cosmology.
The dependence on mass is due to the change of slope of the linear
power spectrum with scale. We plot our results for the Gaussian case (
,
solid lines), large positive
(
103 and
104, dashed lines) and large negative
(
103 and
104, dotted lines), for the local model (7). Thus, we can see that a positive
increases the relative density (i.e. with respect to the density at radius q) both at small and large radii. The very large values of
required to be able to distinguish the curves in the figure imply that for realistic cases (
)
the perturbation of the density profile is very small. Therefore, the values of the upper boundary
,
which marks the onset of shell-crossing, obtained in Valageas (2009b)
for the Gaussian case remain valid up to a very good accuracy. We can
note that to obtain a similar deviation from the Gaussian profile
we need a larger value of the parameter
on a smaller scale. This can be understood from the expression (7), which scales as
and grows as k-2
on very large scales. The same behavior (i.e. a higher
sensitivity to local-type non-Gaussianity on large scales) is obtained
for the bias of dark matter halos, see Eq. (2) above and Sect. 4 below.
Next, we define the constrained weight (23) as
at the relevant saddle-point

Therefore, the tails of the probability distribution (23) read as
where we introduced the skewness of the linear density contrast, at first order over

Indeed, from Eq. (8) we have at first order
![]() |
= | ![]() |
|
![]() |
(46) | ||
= | 6 fq;qq. | (47) |
Hereafter, since S3(0)=0, we simply note S3=S3(1).
3.2 Mass function
Following the Press-Schechter approach (Press & Schechter 1974), the mass function that is obtained in the Gaussian case from the probability distribution
reads as
with (using the subscript ``PS'' to distinguish the Press-Schechter prediction)
and
Here





but the power-law prefactor and the low-mass tail of (49) have no reason to be valid (and numerical simulations indeed show that they are not exact). Then, in order to match numerical simulations, one needs to use fitting formulae. One such fit to simulations, which obeys the exact tail (51), is (Valageas 2009b)
Both mass functions (49) and (52) satisfy the normalization
which ensures that all the mass is contained in such halos:
In the non-Gaussian case, that is

As explained above, this yields the exact high-mass tail (up to first order over

we may use for the non-Gaussian mass function
where we use the same scaling function f as for the Gaussian case (48). This recovers the high-mass tail (55) and satisfies the normalization (54). Note that Afshordi & Tolley (2008) have recently proposed a similar rescaling as (56), which they re-interprate as a modified effective variance



Using Eqs. (50), (56), we obtain
If we use the Press-Schechter mass function (49), Eqs. (57)-(59) give back the result obtained in Matarrese et al. (2000), except for the fact that we use

Note that Matarrese et al. (2000)
also use a saddle-point approach to derive the halo mass function for
non-Gaussian initial conditions, expanding at linear order over
(or S3), so that their computation is equivalent to the one described above. However, since they first compute the cumulant
generating function (
in their notations, or
in Valageas 2009b), the constraint (30) is expressed through a Dirac function, written as an exponential by introducing an auxiliary variable
,
so that they can first integrate over the Gaussian field
and next expand over
the expression obtained for
.
The somewhat simpler method described in this article has the advantage of simultaneously giving the density profile (36), (37) of the underlying saddle point. As explained above in Fig. 1,
this allows us to check that realistic amounts of primordial
non-Gaussianity
have a negligible effect on this profile, so that the onset of
shell-crossing appears for almost the same nonlinear density
threshold
.
This ensures that the rare-event and high-mass tails (44) and (55) are exact (at leading order), as long as halos are defined by
a nonlinear threshold
below the upper bound
.
The simplicity of the method presented in this article also allows a
straightforward application to two-point distributions, as shown
in Sect. 4
below, or to more complex primordial non-Gaussianities, which may
involve several fields or higher order polynomials as in Eq. (95) below.
Another approach presented in Lo Verde et al. (2008) is to use the Edgeworth expansion, which writes the probability distribution
as a series over the cumulants of the non-Gaussian variable
.
In practice, one truncates at the lowest order beyond the Gaussian, that is, at the third cumulant
described by S3. Thus, expanding the exponentials (44) and (55) one recovers the results of Lo Verde et al. (2008) at large
(i.e. low
).
However, in the rare-event limit, where computations rest on firm
grounds as explained above, the Edgeworth expansion does not fare very
well. In particular, although we only derived the
saddle-point
and the argument
of the exponential up to linear order over
,
see Eqs. (42), (43), it is best to keep the exponential as in Eqs. (55) or (57). Indeed, in the rare-event and small-
limits, the tail (55) can be a good approximation even when the corrective factor
is much larger than unity (i.e. we can have the hierarchy
).
Since the low-mass tail (49) does not match numerical simulations, it is sometimes proposed to keep the ratio (58)
given by the Press-Schechter mass function, and to multiply the fit
from simulations of the Gaussian mass function by this factor (Grossi
et al. 2007, 2009; Lo Verde et al. 2008). However, this procedure clearly violates the normalization condition (54). Therefore, we suggest to use (58) with the fitting formula obtained from Gaussian simulations, that is, to use Eq. (57), which automatically satisfies the normalization (54).
However, there is no reason to expect that the low-mass tail can be
exactly recovered by any such procedure, even though by construction it
gives the right behavior for the Gaussian case, as the low-mass slope
might also depend on
in some specific manner.
![]() |
Figure 2:
The ratio
|
Open with DEXTER |
We compare in Fig. 2 our results for the halo mass function with numerical simulations from Grossi et al. (2009). They use the LSS convention for
,
so that in terms of the CMB convention used in this article this corresponds to
.
We show the ratio of the non-Gaussian mass function to the Gaussian one, as given by the multiplicative factor (55), the ratio (58) computed with the Press-Schechter mass function (44) or with the fitting function (52). In agreement with Grossi et al. (2009), we find that the ratio (58) computed with the Press-Schechter mass function, which also corresponds to the result of Matarrese et al. (2000) (but with
instead of 1.686 as explained above), agrees reasonably well with simulations. Using the fitting function (52) or the simple multiplicative factor (55) yields close results in this regime and also agrees with simulations. However, using Eq. (58) with the fitting function (52)
appears to agree somewhat better with simulations, especially at low
masses. This could be expected from the fact that this procedure
ensures that the mass function is properly normalized
(in contrast, the simple multiplicative factor (55) is greater than unity and does not reproduce the crossing of both mass functions for
).
On the other hand, let us point out that, contrary to some previous
works, we do not need to introduce any ad-hoc parameter q (e.g., through a change of the form
as in Grossi et al. 2009) to obtain a good match with numerical simulations. This decrease in the linear threshold with respect to the standard value
(for
)
is actually obtained in our approach by using the exact linear threshold
,
as explained above. Therefore, the advantage of this procedure is
that we do not need to run new simulations for other cosmologies to
obtain a fit for such a q-factor, since the value
can always be computed from the spherical collapse dynamics.
![]() |
Figure 3:
Same as Fig. 2, but with
|
Open with DEXTER |
Next, we compare our results with numerical simulations from Dalal et al. (2008) in Fig. 3. We again show the ratio
as a function of mass, but for a greater primordial non-Gaussianity,
.
For
(upper panel), we note as in Dalal et al. (2008) that the prediction of (58) computed with the Press-Schechter mass function (44) (i.e. the result of Matarrese et al. 2000, but with
)
tends to overestimate the deviations from the Gaussian case (and the simple multiplicative factor (55) fares somewhat worse). However, using Eq. (58) with the correct
Gaussian mass function (52) decreases this ratio somewhat (as in Fig. 2) and provides good agreement with the simulations. For
(lower panel) the match is not as good. However, in that case the
agreement might improve at higher masses, where there are no data
points but where the predictions (55) or (58)
are asymptotically exact. Moreover, the discrepancy between theoretical
predictions and numerical simulations has the same order as the
deviation between the different theoretical curves. Since the latter
show the same exact high-mass behavior (at the leading order given
by the exponential cutoff (44)),
these deviations show the sensitivity of the mass functions to the
details of the theoretical prescriptions. These power-law prefactors
have not been rigorously derived (and are expected not to be exact for
all formulae used here). Therefore, the deviation between these
theoretical predictions estimates the theoretical uncertainty for the
ratio of the halo mass functions. Then, taking this theoretical
uncertainty into account, we can see that the agreement with the
numerical results is still reasonable. For practical purposes,
when one tries to derive constraints on cosmology from observations of
halo mass functions, it would be useful to consider several theoretical
prescriptions in addition to the best prediction (58)-(52), as in Figs. 2 and 3, so as to take the theoretical uncertainty into account in the analysis.
4 Bias of dark matter halos
4.1 Two-cell saddle point and real-space bias
We now consider the bias of dark matter halos, or more precisely their two-point correlation function. As in Valageas (2009b), following Kaiser (1984),
we identify rare massive halos with positive density fluctuations in
the linear density field. Thus, we first consider the bivariate
probability distribution,
,
of the linear density contrasts
and
within two spheres V1 and V2, of radii q1 and q2, and separated by the distance s. Note that we distinguish the Lagrangian distance s between the halos measured in the linear density field from their Eulerian distance x
measured in the nonlinear density field. Indeed, since halos have moved
through their mutual gravitational attraction, these two distances are
usually different. Proceeding as in Sect. 3.1 to obtain the rare-event tails,
we are led to introduce the action
which now involves the two Lagrange multipliers








where we note


and
where we note for instance f1;22=fa;bb(s) and f2;12=fb;ab(s), with Va=V1, Vb=V2, from Eqs. (19), (20). Next, the Gaussian weight

and at first order,
Therefore, in the rare-event limit the tail of the bivariate distribution (60) reads as
with
where



where the factor







where

which provides an explicit expression for s.
Finally, we define the real-space halo bias as the ratio of the halo and matter two-point correlations,
Since at a large distance the matter correlation is within the linear regime,

which fully determines the halo bias from Eq. (72).
For equal-mass halos of radius q, defined by the same threshold
,
the two Lagrange multipliers are equal,
,
and Eqs. (63, 64) and (66, 67) simplify as
and
where we note


and
Then, the difference

and
We can check that



As stressed in Politzer & Wise (1984), real-space formulae such as (72), which are obtained in the rare event (
)
and large separation (
)
limits, do not assume that the exponent
is small. In fact, as shown in Valageas (2009b),
at high redshift one can probe a regime where this exponent is large,
so that one needs to keep the nonlinear form (72), which yields a nonlinear bias. As seen from Eq. (81), this regime corresponds to very massive halos,
,
at fixed (low) ratio
.
Nevertheless, in the regime where
is small (i.e. in the large separation limit
), which covers the cases of interest encountered at low redshift, we can expand the exponential in Eq. (72). Then, at the lowest order over the terms that vanish in the large separation limit, we obtain
![]() |
(83) |
for equal-mass halos, whence
Hereafter we call the bias b2M(x) obtained from Eqs. (84) and (76) the ``linearized'' bias.
Following the approach of Kaiser (1984), Matarrese & Verde (2008) also computed the effect of local-type non-Gaussianity (1) on the halo two-point correlation. Drawing on earlier work by Matarrese et al. (1986), who computed the n-point
halo correlations by expanding the relevant path-integrals and next
resumming the series within a large-distance and rare-event
approximation, they obtained expressions of the form (72) without the prefactor
and (84) without the terms in the first line, which arise from this prefactor, and the last term f1;11 in the second bracket. This term arises from the factor
in the denominator of expression (80), associated with the effect of primordial non-Gaussianity on the one-point distribution (43). It can be seen as a renormalization of the Gaussian term
,
since it shows the same scale dependence through the function
.
The presence of such a term has already been noticed in Slosar et al. (2008) in Fourier space, and Desjacques et al. (2009)
points out that it needs to be included to obtain good agreement with
numerical simulations. This will give rise to the last term in
Eq. (85) and the second term in Eqs. (91) and (93) below. On the other hand, the terms in the first line of Eq. (84), associated with the prefactor
in Eq. (72), stem from
the mapping from Lagrangian to Eulerian space, and the first bracket in Eq. (84) expresses the (weak) effect of primordial non-Gaussianity on this mapping. Another difference between Eqs. (72) and (84)
and previous works is that (within some approximation) we pay attention
to the difference between Lagrangian and Eulerian distances s and x. This can play a non-negligible role as seen in Fig. 4 below.
![]() |
Figure 4:
The halo bias bM(x), as a function of |
Open with DEXTER |
![]() |
Figure 5:
The halo bias bM(x) as a function of distance x, at redshifts z=0 ( upper panel) and z=1 ( lower panel) for several masses. We show the cases
|
Open with DEXTER |
![]() |
Figure 6:
The real-space ratio
|
Open with DEXTER |
We show in Fig. 4 the halo bias bM(x) as a function of
at fixed distance
x=50 h-1 Mpc and redshift z=0, for
and
.
We display both the nonlinear result of Eq. (72) and the linear result of Eq. (84). In addition, for the Gaussian case we also display the bias obtained by setting s=x in Eq. (72). As in Valageas (2009b), for the Gaussian case (
)
we obtain good agreement with the fits to numerical simulations of
Sheth et al. (2001) and Pillepich (2010). Moreover, we can see
that it is important to take the displacement of the halos into account
through Eq. (73), as the approximation s=x significantly overestimates the bias. We checked that using the simpler Eq. (74) gives a close result to the one obtained with Eq. (73) and also agrees with the simulations. Thus, for practical purposes it is sufficient to use Eq. (74). We can see that for all cases shown in Fig. 4 the linear bias from Eq. (84) gives results that are very close to the fully nonlinear expression (72). This justifies the use of such linearized expressions in this regime. This will be especially useful in Sect. 4.2,
where we consider the bias of dark matter halos in Fourier space.
Indeed, it is easier to take the Fourier transform of Eq. (84), which allows us to recover the
results obtained in previous works. Thus, one interest of the real-space results (72)-(84) is to provide a check on whether linearized predictions
(i.e. where the halo correlation only involves the matter power spectrum at linear order)
are valid. Then, in agreement with those studies (which were mostly performed in Fourier space and led to Eq. (2), see Dalal et al. 2008; Slosar et al. 2008), and with Eq. (84), we can see that the deviation from the Gaussian bias,
,
grows linearly with bM(x,0) at large bias and has the same sign as
.
![]() |
Figure 7:
The halo (
|
Open with DEXTER |
Next, we display in Fig. 5 the dependence on the distance x of the bias obtained for several masses at redshifts z=0 and z=1. More precisely, we show the ratio
,
since the halo and matter correlations do not change sign at the same point. We plot the cases
,
as well as the Gaussian case
.
While the Gaussian bias is roughly constant on large scales, up to
Mpc (in agreement with previous studies, Mo & White 1996; Mo et al. 1997), the non-Gaussian bias shows a strong scale dependence, with a deviation from the Gaussian bias that roughly grows as x2 up to
Mpc. This agrees with the k-2 behavior observed in Fourier space, see Eq. (2) above (Dalal et al. 2008; Slosar et al. 2008) and Eq. (93) below.
To see the scaling of the real-space correction
to the Gaussian bias more clearly, we show in Fig. 6 the ratio
,
with
.
We display the results obtained from Eq. (72) for several masses at z=0 (upper panel) and z=1 (lower panel) for
.
We can see that over the range
30<x<90 h-1 Mpc all curves roughly collapse
onto one another. This means that the real-space correction roughly scales as
over this range, which roughly agrees with the Fourier-space scaling (2) (here we neglected any constant offset, such as the factor -1 in Eq. (2), see the discussion of Eq. (94) below). It appears that our predictions scale more closely as x2, as shown in Fig. 6, than as
,
which would be suggested by Eqs. (2), (93)
(at these scales the transfer function already deviates from
unity). This agrees with the behavior observed in Fourier space in
Fig. 10 below. The masses shown in Fig. 6 span the range
1.2<bM(x,0)<5.9 at
x=50 h-1 Mpc and z=0, and
2.4<bM(x,0)<15.4 at z=1, so that the linear scaling with bM(x,0) of the correction
appears to be a good approximation.
Below 30 h-1 Mpc higher masses show steeper scale dependence for
.
At very large distance,
x > 100 h-1 Mpc, the oscillations seen in Figs. 5 and 6
are caused by the baryon acoustic oscillation. Indeed, the baryon
oscillations seen in the halo and matter two-point correlations are not
exactly proportional, since the halo correlation is not exactly
proportional to
,
even in the Gaussian case and in the linear regime (for instance it involves the smoothing scale q, see Eq. (84)). This yields the non-monotonic behavior seen in Figs. 5 and 6 around
100 h-1 Mpc.
For the same reason, the halo and matter correlations do not exactly
vanish at the same distance, which gives rise to the divergent spike at
Mpc. These features simply mean that it is no longer useful to work with the bias bM
on these scales, which only makes sense if the halo and matter
correlations are roughly proportional. In this range, where the
correlations show some oscillations and change sign, it is no
longer a good approximation to write the halo correlation in terms of
the matter correlation multiplied by some slowly varying bias factor.
Then, one instead needs to directly study the halo and matter
correlations themselves.
Thus, we compare in Fig. 7
the halo and matter two-point correlations. We focus on large scales to
see how the baryon acoustic oscillation is modified when one uses
massive halos as a tracer of the initial matter power spectrum.
In agreement with previous works (Desjacques 2008), we can see that the oscillation is strongly amplified for massive halos that have a strong bias.
This amplification still holds for significant primordial non-Gaussianity (
), although it appears to be slightly lower for positive
.
Moreover, the peak of the oscillation shows no significant shift,
so that a measure of its position appears to be a robust ruler for
constraining cosmology, independently of the halo bias and of the
primordial non-Gaussianity. In contrast, the distance at which the
two-point correlation changes sign is not significantly modified as one
goes from the matter to the halo correlation in the Gaussian case, but
it is fairly sensitive to the primordial non-Gaussianity.
In particular, a positive
shifts this point to a greater distance. However, theoretical and observational error bars may be too large to use this
effect to constrain
in a competitive manner compared to other probes.
4.2 Fourier-space bias
Rather than the real-space two-point correlation, recent works have mostly studied the effect of primordial non-Gaussianity on the halo power spectrum, where at lowest order the Poisson Eq. (3) directly gives an estimate of the form (2) for the deviation from the bias obtained with Gaussian initial conditions (Dalal et al. 2008; Slosar et al. 2008).
It is not convenient to take the Fourier transform of the nonlinear correlation (72), but at moderate redshifts, the linearized form (84) provides a very good approximation. Then, if we also make the approximation
,
which is valid at the lowest order, the Fourier transform of Eq. (84) readily gives the halo power spectrum as
where the quantities



where we used


where we use the explicit expression (74) for s(x). This expresses that Lagrangian-space wavelengths


For large negative
the halo power spectrum (85)
can become negative, because we looked for an expression of the halo
correlation, or of the halo power spectrum, at linear order over
as in Eq. (85). However, the power spectrum PM(k) must be positive by definition. Then, in cases where expression (85) turns negative one should consider higher-order terms over
,
which would ensure that the power spectrum remains positive.
Nevertheless, since such high-order terms are beyond the scope of this
article, we consider below the following simple procedure that ensures
that PM(k), or the squared bias bM2(k), remain positive. Up to linear order over
,
the bias (86) reads as
where

where the two sides only differ by higher-order terms over



![]() |
Figure 8:
The correction to the halo power spectrum due to primordial
non-Gaussianity. We show the ratio
|
Open with DEXTER |
![]() |
Figure 9:
Same as Fig. 8, but for M=1.5 |
Open with DEXTER |
If we take the limit of very rare events, which is
in Eq. (85), we can only keep the last two terms (note that
,
see Eq. (92) below),
At low k, with


The first term,


which yields from Eq. (91)
Thus, we recover the k-2 dependence at low k brought by the local-type non-Gaussianity (1), through the

In spite of the k-2 dependence at low k obtained in Eq. (93) for the halo bias, the real-space halo two-point correlation is well-defined and finite, as seen in Sect. 4.1. Indeed, Eq. (93) only applies to a limited range, and one cannot
write the real-space two-point correlation as a Fourier transform of the form
,
which would diverge at low k.
Thus, the advantage of the real-space approaches, such as the one
described in this paper, is that we obtain well-defined results in
both real space and Fourier space, and we do not need to regularize
integrals by introducing a counterterm associated with a survey-size
window, as in Wands & Slosar (2009).
This is reassuring, since one does not expect the halo correlation on a
given scale to depend on the size of the survey. Mathematically, the
lack of divergence in our approach comes from the fact that it is the
halo power spectrum itself which contains a term of the form
,
see Eqs. (85) and (92),
and it is only by expanding the square-root as in (88),
,
that b can be written as in Eq. (93). As we shall see below, in Figs. 8 and 9, the expression (85) is sufficient to explain the behavior observed in numerical simulations, without introducing worrying divergences.
We compare in Figs. 8 and 9 our results for the Fourier-space bias bM2(k) with numerical simulations from Desjacques et al. (2009).
Since the dependence on mass is rather weak, we show our results in Fig. 8 for M=2
,
whereas the data points are for M>2
,
and we show our results in Fig. 9 for M=1.5
,
whereas the data points are for
1013<M<2
.
The predictions (85) and (87) are almost indistinguishable in this regime, and they agree reasonably well with the simulations, except at low k for
where they give a negative halo power spectrum. Equation (89), gives a much better fit to simulations at low k for negative
,
as could be expected from the
fact that it always gives a positive halo power spectrum. However, a
priori one should not give too much weight to this improved accuracy in
this regime. Indeed, as is clear from Eqs. (88), (89), the solid and dot-dashed curves in Figs. 8, 9 only differ by terms of order
and beyond. Since all our results have been derived at linear order over
,
one can expect that Eq. (89) does not include all terms of order
.
Then, although for practical purposes it is better to use Eq. (89) in this regime (i.e. negative
at low k), it is still
useful to also consider Eq. (85), as the deviation between both predictions should give an estimate of the theoretical uncertainty.
In contrast to some previous approaches (e.g., Grossi et al. 2009; Desjacques et al. 2009), the good agreement with numerical simulations shown in Figs. 8 and 9 is obtained from Eq. (85) without any fitting parameter (such as the rescaling parameter q in Grossi et al. 2009; or the mass function parameters in Desjacques et al. 2009). As for the halo mass function studied in Sect. 3.2, the role of this parameter is partly played by the use of the exact linear threshold
,
which is predicted by the spherical dynamics of the rare-event saddle points. This makes formulae such as Eq. (85) fully predictive for any values of cosmological parameters.
![]() |
Figure 10:
The Fourier-space ratio
|
Open with DEXTER |
As in Fig. 6, in order to see the scaling of the correction
to the Fourier-space Gaussian bias more clearly, we show the ratio
in Fig. 10, where we now define
The correction obtained from a simple peak-background split argument instead gives
![$\Delta b_M(k,f_{\rm NL}) = f_{\rm NL}[b_M(k,0)-1] (2\delta_{\rm L}/\alpha(k))$](/articles/aa/full_html/2010/06/aa12636-09/img344.png)













5 Conclusion
We have shown in this article how to extend to non-Gaussian initial
conditions the computation of the mass function and of the bias of dark
matter halos presented in Valageas (2009b)
for the Gaussian case. This relies on a saddle-point approach that
allows to derive the high-mass asymptotic tails of the quantities of
interest from the statistical weight of the initial conditions,
supplemented by additional nonlinear constraints. Then, focusing on the
case of ``
-type''
primordial non-Gaussianity, where the linear gravitational potential
can be written as the sum of linear and quadratic terms over an
auxiliary Gaussian field
,
we explained how to obtain the relevant saddle-points as a perturbative series over the nonlinear parameter
.
This method is very general, and it applies to any case of small
primordial non-Gaussianity, where Bardeen's potential
(or equivalently, below the Hubble radius, the gravitational
potential or the density field) can be written as a polynomial over a
Gaussian field
as
where the nonlinear parameters fi are small. Then, following the method presented in Sects. 3, 4, the one-cell and two-cell saddle points (associated with the one-point and two-point density distributions, whence the halo mass function and bias) can be computed as a perturbative series over the coefficients fi, which need not be of the same order. This includes the case of a cubic term




Focusing on the case of local-type primordial non-Gaussianity, we described how to obtain, up to linear order over
,
the one-cell saddle point associated with the probability distributions
and
of the linear and nonlinear density contrasts
within spherical cells. This gives the quasi-linear limit of these
distributions, as well
as the high-mass exponential falloff of the halo mass function. One
advantage of our method is that it allows us to explicitly check that
realistic amounts of primordial non-Gaussianity have no significant
effect on the density profile of this saddle point. This ensures that
shell crossing appears for (almost) the same nonlinear density
(see Valageas 2009b),
so that the high-mass tail of the halo mass function can be derived
provided halos are defined by a nonlinear density threshold that is
below this upper bound (which is indeed the case). Although this
procedure only gives the high-mass tail, we proposed a simple change of
variable, applied to the mass function fitted to Gaussian numerical
simulations, that obeys this high-mass asymptotic while keeping the
mass function normalized to unity. If one uses the Press-Schechter mass
function, this gives back the result of Matarrese et al. (2000),
but we argue that this procedure is somewhat more natural if one wishes
to recover a more accurate mass function in the Gaussian case. We also
checked that this agrees with results from non-Gaussian numerical
simulations.
Next, we applied this method to the two-point correlation of massive halos, following the approach of Kaiser (1984). As in Valageas (2009b)
we take the displacement of halo pairs under their mutual gravitational
attraction into account. This gives the real-space halo
correlation ,
whence the real-space bias bM(x). Since this approach does not assume that the halo correlation is weak,
the nonlinear formula it yields can be used to check whether the ``linearized'' form
(where one only keeps the linear term over the matter correlation
)
is a good approximation in the regime of interest. As expected, we find that the correction
to the Gaussian bias grows with bM(x,0) and with scale, roughly as
,
up to
Mpc.
Beyond this scale, the baryon acoustic oscillation and the fact that
the halo and matter correlations do not change sign at the same point
lead to strong oscillations and divergent spikes for bM(x). This means that, for
x > 100 h-1 Mpc, the bias is no
longer a useful quantity, and one should directly work with the halo
and matter correlations. In agreement with Desjacques (2008),
we find that the two-point correlation of massive halos, which have a
large bias, strongly amplifies the baryon acoustic oscillation. In
addition we also obtain the modifications associated with primordial
non-Gaussianity. The baryon oscillation remains strongly amplified,
with a small shift, but somewhat less so for positive
.
Finally, we used the ``linearized'' form of the halo two-point
correlation to derive the halo power spectrum and the halo bias in
Fourier space. We also give a simple recipe that ensures that the halo
power spectrum always remains positive. (This only differs from the
direct prediction by terms of order
and
higher.) We obtain good agreement with numerical simulations without
introducing any free parameter. Moreover, the two formulae described
above allow one to estimate the range over which linear approximations
over
are sufficient. Thus, we find that terms of order
start playing a role at low k (
k < 0.01 h Mpc-1) for large negative
(
),
where the direct formula would give a negative power spectrum.
We also pointed out that the k-2 behavior observed at low kfor the halo bias does not imply any divergence for the real-space two-point
correlation. Indeed, this behavior is only obtained within a certain limit,
and it is the halo power spectrum itself (i.e.
rather than
)
that shows this k-2 factor. We showed that this is
sufficient to explain the behavior observed in numerical simulations.
Moreover, it avoids the need to introduce counterterms, that depend on the
size of the survey, so as to obtain finite real-space correlations. This is an advantage of
real-space approaches, such as the one presented in this paper, which are better
suited to describing the nonlinear effects associated with the bias of massive
halos.
These results, which do not involve free parameters except for
the mass function (if one requires its full shape, where one needs
the fit to numerical simulations for Gaussian initial conditions)
should be useful for constraining primordial non-Gaussianities from
observations of large-scale structures. Thus, neither the high-mass
tail of the halo mass function nor the bias require
rescaling parameters (such as
), because such a
correction to the linear threshold
is achieved through the use of
the exact linear threshold
predicted by the spherical dynamics of rare-event saddle points. This
makes this approach more predictive than some of the previous works,
since one does not need to run new simulations to fit for such q-factors in order to investigate other cosmologies. In particular, as discussed above for Eq. (95),
the method presented in this article is quite general and can be
applied to a large class of models. Moreover, since it provides results
in both real space and Fourier space (i.e. the halo two-point
correlation and power spectrum), it gives a complete and consistent
description of halo clustering. As for previous approaches, the most
reliable use of these models to constrain cosmology is to take
advantage of the
specific shape of the dependence on mass (for the mass function) or
scale (for the bias) brought by primordial non-Gaussianity to
constrain
,
rather than the change in the amplitude at a given mass or scale.
References
- Afshordi, N., & Tolley, A. J. 2008, Phys. Rev. D, 78, 123507 [NASA ADS] [CrossRef] [Google Scholar]
- Amara, A., & Refregier, A. 2004, MNRAS, 351, 375 [NASA ADS] [CrossRef] [Google Scholar]
- Arkani-Hamed, N., Creminelli, P., Mukohyama, S., & Zaldarriaga, M. 2004, JCAP, 4, 1 [NASA ADS] [CrossRef] [Google Scholar]
- Bardeen, J., Bond, J. R., Kaiser, N., & Szalay, A. S. 1986, ApJ, 304, 15 [NASA ADS] [CrossRef] [Google Scholar]
- Barnaby, N., & Cline, J. M. 2006, Phys. Rev. D, 73, 106012 [NASA ADS] [CrossRef] [Google Scholar]
- Bartolo, N., Matarrese, S., & Riotto, A. 2002, Phys. Rev. D, 65, 103505 [NASA ADS] [CrossRef] [Google Scholar]
- Bartolo, N., Komatsu, E., Matarrese, S., & Riotto, A. 2004, Phys. Rep., 402, 103 [NASA ADS] [CrossRef] [Google Scholar]
- Colafrancesco, S., Lucchin, F., & Matarrese, S. 1989, ApJ, 345, 3 [NASA ADS] [CrossRef] [Google Scholar]
- Dalal, N., Dore, O., Huterer, D., & Shirokov, A. 2008, Phy. Rev. D, 77, 123514 [Google Scholar]
- Desjacques, V. 2008, Phys. Rev. D, 78, 103503 [NASA ADS] [CrossRef] [Google Scholar]
- Desjacques, V., Seljak, U., & Iliev, I. T. 2009, MNRAS, 396, 85 [Google Scholar]
- Falk, T., Rangarajan, R., & Srednicki, M. 1993, ApJ, 403, L1 [NASA ADS] [CrossRef] [Google Scholar]
- Grinstein, B., & Wise, M. B. 1986, ApJ, 310, 19 [NASA ADS] [CrossRef] [Google Scholar]
- Grossi, M., Dolag, K., Branchini, E., Matarrese, S., & Moscardini, L. 2007, MNRAS, 382, 1261 [NASA ADS] [CrossRef] [Google Scholar]
- Grossi, M., Verde, L., Carbone, C., et al. 2009, MNRAS, 398, 321 [NASA ADS] [CrossRef] [Google Scholar]
- Hamilton, A. J. S., Kumar, P., Lu, E., & Matthews, A. 1991, ApJ, 374, L1 [NASA ADS] [CrossRef] [Google Scholar]
- Kaiser, N. 1984, ApJ, 284, L9 [NASA ADS] [CrossRef] [Google Scholar]
- Komatsu, E., Dunkley, J., Nolta, M. R., et al. 2009, ApJS, 180, 330 [NASA ADS] [CrossRef] [Google Scholar]
- Lucchin, F., & Matarrese, S. 1988, ApJ, 330, L535 [NASA ADS] [CrossRef] [Google Scholar]
- Lo Verde, M., Miller, A., Shandera, S., & Verde, L. 2008, JCAP, 4, 14 [Google Scholar]
- Lyth, D. H., Ungarelli, C., & Wands, D. 2000, Phys. Rev. D, 67, 023503 [Google Scholar]
- Maggiore, M., & Riotto, A. 2009 [arXiv:0903.1251] [Google Scholar]
- Matarrese, S., & Verde, L. 2008, ApJ, 677, L77 [NASA ADS] [CrossRef] [Google Scholar]
- Matarrese, S., Lucchin, F., & Bonometto, S. A. 1986, ApJ, 310, L21 [NASA ADS] [CrossRef] [Google Scholar]
- Matarrese, S., Verde, L., & Jimenez, R. 2000, ApJ, 541, L10 [NASA ADS] [CrossRef] [Google Scholar]
- Mo, H. J., & White, S. D. M. 1996, MNRAS, 282, 347 [NASA ADS] [CrossRef] [Google Scholar]
- Mo, H. J., Jing, Y. P., & White, S. D. M. 1997, MNRAS, 284, 189 [NASA ADS] [Google Scholar]
- Peacock, J. A., & Dodds, S. J. 1996, MNRAS, 280, L19 [NASA ADS] [CrossRef] [Google Scholar]
- Pillepich, A., Porciani, C., & Hahn, O. 2010, MNRAS, 402, 191 [NASA ADS] [CrossRef] [Google Scholar]
- Politzer, H. D., & Wise, M. B. 1984, ApJ, 285, L1 [NASA ADS] [CrossRef] [Google Scholar]
- Press, W. H., & Schechter, P. 1974, ApJ, 187, 425 [NASA ADS] [CrossRef] [Google Scholar]
- Senatore, L., Tassev, S., & Zaldarriaga, M. 2009, JCAP, 9, 38 [NASA ADS] [CrossRef] [Google Scholar]
- Slosar, A., Hirata, C., Seljak, U., Ho, S., & Padmanabhan, N. 2008, JCAP, 8, 31 [NASA ADS] [Google Scholar]
- Valageas, P. 2002a, A&A, 382, 412 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Valageas, P. 2002b, A&A, 382, 431 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Valageas, P. 2009a, Phys. Rev. E, 80, 016305 [NASA ADS] [CrossRef] [Google Scholar]
- Valageas, P. 2009b, A&A, 508, 93 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Wands, D., & Slosar, A. 2009, Phys. Rev. D, 79, 123507 [NASA ADS] [CrossRef] [Google Scholar]
Footnotes
- ...2009a,b)
- As shown in Valageas (2009a),
for the closely related adhesion model, where the same procedure can be
applied, one can explicitly check that the asymptotic results obtained
by this approach agree with the complete distribution
that is exactly known for two cases (Brownian and white-noise linear velocity in 1D, corresponding to a power-law linear density power spectrum with n=-2 and n=0).
All Figures
![]() |
Figure 1:
The radial profile (36, 37) of the linear density contrast
|
Open with DEXTER | |
In the text |
![]() |
Figure 2:
The ratio
|
Open with DEXTER | |
In the text |
![]() |
Figure 3:
Same as Fig. 2, but with
|
Open with DEXTER | |
In the text |
![]() |
Figure 4:
The halo bias bM(x), as a function of |
Open with DEXTER | |
In the text |
![]() |
Figure 5:
The halo bias bM(x) as a function of distance x, at redshifts z=0 ( upper panel) and z=1 ( lower panel) for several masses. We show the cases
|
Open with DEXTER | |
In the text |
![]() |
Figure 6:
The real-space ratio
|
Open with DEXTER | |
In the text |
![]() |
Figure 7:
The halo (
|
Open with DEXTER | |
In the text |
![]() |
Figure 8:
The correction to the halo power spectrum due to primordial
non-Gaussianity. We show the ratio
|
Open with DEXTER | |
In the text |
![]() |
Figure 9:
Same as Fig. 8, but for M=1.5 |
Open with DEXTER | |
In the text |
![]() |
Figure 10:
The Fourier-space ratio
|
Open with DEXTER | |
In the text |
Copyright ESO 2010
Current usage metrics show cumulative count of Article Views (full-text article views including HTML views, PDF and ePub downloads, according to the available data) and Abstracts Views on Vision4Press platform.
Data correspond to usage on the plateform after 2015. The current usage metrics is available 48-96 hours after online publication and is updated daily on week days.
Initial download of the metrics may take a while.