You are on page 1of 16

Mon. Not. R. Astron. Soc. 000, 1–16 (0000) Printed 10 August 2012 (MN LATEX style file v2.

2)

Scale dependent halo bias in the excursion set approach

Marcello Musso1⋆ , Aseem Paranjape2† & Ravi K. Sheth2,3


1 CP3-IMP3, Université Catholique de Louvain, 2 Chemin du Cyclotron, 1348 Louvain-la-Neuve, Belgium
2 The Abdus Salam International Center for Theoretical Physics, Strada Costiera, 11, Trieste 34151, Italy
3 Center for Particle Cosmology, University of Pennsylvania, 209 S. 33rd St., Philadelphia, PA 19104, USA
arXiv:1205.3401v2 [astro-ph.CO] 9 Aug 2012

10 August 2012

ABSTRACT
If one accounts for correlations between scales, then nonlocal, k-dependent halo
bias is part and parcel of the excursion set approach, and hence of halo model
predictions for galaxy bias. We present an analysis that distinguishes between
a number of different effects, each one of which contributes to scale-dependent
bias in real space. We show how to isolate these effects and remove the scale de-
pendence, order by order, by cross-correlating the halo field with suitably trans-
formed versions of the mass field. These transformations may be thought of as
simple one-point, two-scale measurements that allow one to estimate quantities
which are usually constrained using n-point statistics. As part of our analysis,
we present a simple analytic approximation for the first crossing distribution
of walks with correlated steps which are constrained to pass through a specified
point, and demonstrate its accuracy. Although we concentrate on nonlinear, non-
local bias with respect to a Gaussian random field, we show how to generalize
our analysis to more general fields.
Key words: large-scale structure of Universe

1 INTRODUCTION position (Fry & Gaztañaga 1993; Manera & Gaztañaga


2012; Pollack, Smith & Porciani 2012; Chan & Scocci-
Galaxy clustering depends on galaxy type (Zehavi et al.
marro 2012). Even in this case, there are a number of
2011 and references therein). Therefore, not all galaxies
ways in which scale dependence can arise, even for the
are fair tracers of the dark matter distribution. Precise
simplest case of Gaussian initial conditions and standard
constraints on cosmological models require a good under-
gravity. Since the measured bias will generally be a com-
standing of this galaxy bias (Sefusatti et al. 2006; More et
bination of all these effects, we present some ideas on how
al. 2012). In the simplest models, galaxies are linearly bi-
to disentagle them from one another.
ased tracers (Kaiser 1984), but, even at the linear level,
this bias may depend on physical scale or wavenumber In general, of course, δh might depend on the value
k (e.g. Desjacques et al. 2010; Matsubara 2011). This of δ at different locations, on its derivatives (Desjacques
scale-dependence, which is clearly detected in simulations et al. 2010; Musso & Sheth 2012), on other higher or-
of hierarchical clustering models (Sheth & Tormen 1999; der statistics of the field (e.g. Sheth, Mo & Tormen 2001;
Smith et al. 2007; Manera et al. 2010), contains important Sheth, Chan & Scoccimarro 2012) at the same or at dif-
information about the statistics of the initial fluctuation ferent positions, etc.; the dependence might even not be
field, and the nature of gravity (Parfrey, Hui & Sheth deterministic (e.g., Sheth & Lemson 1999; Dekel & Lahav
2011; Lam & Li 2012). 1999). Our final goal is to present methods which are able
The most common galaxy bias model – the local bias to pinpoint this relation even when the bias is nonlinear,
model – assumes that the galaxy overdensity field δh (x) nonlocal and stochastic.
is a local, possibly nonlinear, monotonic, deterministic We study insights which arise from the simplest
transformation of the dark matter field δ(x) at the same treatment of halo bias: that based on the excursion set
approach (Press & Schechter 1974). This approach maps
the problem of counting the number of collapsed halos to
⋆ E-mail: marcello.musso@uclouvain.be that of the first crossing of a suitable threshold (the ‘bar-
† E-mail: aparanja@ictp.it rier’) by random walks in density generated by smoothing

c 0000 RAS

2 M. Musso et al.
the initial matter density field using a sequence of filters surement at finite scale which returns these limiting val-
of decreasing scales (Bond et al. 1991). In addition to ues, and quantifies the importance of computing averages
depending on the ‘barrier’ shape, the first crossing dis- over the correct ensemble.
tribution also depends on how far from the ‘origin’ the Section 3 extends these results to the case of cor-
walks happen to be for the largest smoothing scale S0 . related steps. We first derive the conditional distribution
Walks that do not start from the origin have modi- f (s|δ0 , S0 ) that a walk crosses the barrier for the first time
fied first crossing distributions (Lacey & Cole 1993). This at scale s having taken up the value δ0 at scale S0 , and
introduces a dependence of the abundance of halos 1 + δh demonstrate its accuracy by comparing with the results
on the initial matter density field δ smoothed on the much of a Monte Carlo treatment of the problem. We then turn
larger scale S0 , and hence leads to a prediction for halo to the problem of halo bias, and highlight some impor-
bias (Mo & White 1996). tant differences from the uncorrelated case: the question
The excursion set approach greatly simplifies when of the correct pdf is shown to be much less important,
the smoothing filter is sharp in Fourier space, because whereas the scale dependence of bias becomes more dra-
in this case the steps in each walk are uncorrelated with matic. We discuss some of the implications of our analysis
each other. Since most analyses to date have relied on this and conclude in section 4. Appendix A collects proofs of
choice, we use it to illustrate many of our key points. E.g., some results quoted in the text, while Appendix B con-
if the bias is deterministic and nonlinear in real-space, it nects the bias coefficients defined using cross-correlation
will be stochastic in k-space. And, estimates of cross- measurements to other definitions in the literature.
correlations between the halo and mass fields depend on Throughout we will present results for a constant
the assumed form of the probability distribution function barrier of height δc . Moving barriers pose no conceptual
of the mass: one must be careful to use the appropriate difficulty for the first crossing distributions we are inter-
probability density function (pdf). One of the key insights ested in. Also, while our analytical results are generally
of this paper is to show that suitably defined real-space valid for any smoothing filter and power spectrum, for
cross-correlation measurements allow one to extract the ease of implementation, the explicit comparisons with nu-
different bias coefficients, order by order. merical solutions will use the Gaussian filter and a power
Recently, however, there has been renewed interest law power spectrum. Again, we do not expect our final
in studying the effects of smoothing with more realistic conclusions to depend on this choice.
filters such as the TopHat in real space or the Gaussian.
The problem is complicated in this case by the presence
of nontrivial correlations between the steps of the ran- 2 THE EXCURSION SET APPROACH:
dom walks (Peacock & Heavens 1990; Bond et al. 1991), UNCORRELATED STEPS
and a number of different approximations for the effect on
The excursion set ansatz relates the number of halos in
the first crossing distribution have been introduced (Mag-
a mass range (m, m + dm) to the fraction f (s) of walks
giore & Riotto 2010; Paranjape, Lam & Sheth 2012). We
that first cross the barrier in the scale range (s, s + ds)
show that the most accurate of these, due to Musso &
through the relation
Sheth (2012), can be extended to provide a very accurate
model for walks which do not start from the origin. m dn(m)
dm = f (s) ds, (1)
We then show that correlations between steps generi- ρ̄ dm
cally introduce two additional sources of scale-dependent where s = s(m) ≡ δ 2 (m) is the variance of the mat-


bias into the predictions. One is relatively benign, and ter density field smoothed on a Lagrangian length scale
simply arises from the fact that the excursion set predic- corresponding to mass m and linearly extrapolated to
tion is for a real-space quantity, but the halo bias in N - present day, and ρ̄ is the background density.
body simulations is typically measured in Fourier space, In this approach, the influence of the underlying dark
through ratios of power spectra. That this matters reit- matter field on the abundance of halos of mass m (i.e. the
erates a point first made by Paranjape & Sheth (2012), bias) can be estimated from the fraction f (s|δ0 , S0 ) of
but it is easily accounted for by using a more appro- walks that first cross the barrier at s starting from some
priate normalization of the bias coefficients. The second prescribed height δ0 on some prescribed scale S0 , rather
is more pernicious and is a genuinely new source of k- than from the origin (Mo & White 1996; Sheth & Tor-
dependent bias (a point made in Musso & Sheth 2012, men 1999). The mean number overdensity of halos can
but not studied further). Although this complicates dis- be defined as
cussion of scale-dependent bias, our method of measur-
ing suitably defined real-space cross-correlations between f (s|δ0 , S0 )
h 1 + δh |δ0 , S0 i ≡ , (2)
the halo and mass fields can be used to extract the k- f (s)
dependence of halo bias order by order. which is explicitly a prediction in real space, and valid on
This paper is organised as follows. Section 2 briefly scale S0 in the Lagrangian initial conditions.
summarizes known excursion set results for uncorrelated Typically, the bias is characterised by expanding the
steps, defines the halo bias factors as a ratio of real space above expression in powers of δ0 . The coefficients of this
measurements, derives their large scale limiting values, expansion will in general depend on S0 (besides obvi-
uses these to motivate a real-space cross-correlation mea- ously depending on s). Moreover, the evaluation of f (s)

c 0000 RAS, MNRAS 000, 1–16



Scale dependent bias 3
and f (s|δ0 , S0 ) (and therefore of the bias coefficients) is where pG (δ0 ; S0 ) is a Gaussian distribution with zero
rather different depending on whether or not the steps mean and variance S0 . Although this result is formally
in the walk are correlated. In what follows, we elucidate correct, we argue in the next subsection that the appro-
the issue of the scale dependence of the bias coefficients in priate distribution over which to average should not be a
the simpler case of walks with uncorrelated steps. We also Gaussian (nor even a Gaussian chopped at δ0 > δc ). But
argue that the same coefficients can be obtained as the if we continue to ignore this detail, then we find
mean value of the product of h 1 + δh |δ0 , S0 i and poly- Z ∞
fu (s|δ0 , S0 )
nomials in δ0 , weighted by the probability distribution of dδ0 pG (δ0 ; S0 ) δ0 = bu1 S0
−∞ fu (s)
δ0 . This alternative definition as an expectation value will
be more suitable to be extended to the case of correlated (Desjacques et al. 2010), and more generally, the orthogo-
steps (section 3), and to make contact with the defini- nality of the Hermite polynomials implies (Appendix A1)
tion of bias in generic models other than the excursion that
set approach (Appendix B).
δcn
  
fu (s|δ0 , S0 ) δ0
n/2
H n √ = δcn bun . (7)
S0 fu (s) S0
2.1 Large scale Lagrangian bias factors
This exact result is remarkable because the left hand side
The conditional first crossing distribution of a constant involves quantities for an arbitrary S0 , whereas the right
barrier δc for walks with uncorrelated steps is (Bond et hand side, which is independent of S0 , is simply the S0 →
al. 1991; Lacey & Cole 1993) 0 limit of the appropriate bias coefficient. This is not at
all obvious if one had viewed the local bias expansion as
1 δc − δ0 2
fu (s|δ0 , S0 ) = √ 3/2
e−(δc −δ0 ) /2(s−S0 ) , (3) a formal Taylor series: one would naively have thought
2π (s − S0 )
that, at the very least, the cross correlation h (1 + δh )δ0 i
(the subscript in fu standing for “uncorrelated”), where should involve the bias coefficients of all (odd) orders (for
δc > δ0 and s > S0 . The corresponding unconditional a further discussion, see Frusciante & Sheth 2012).
2
distribution is sfu (s) = (2π)−1/2 νe−ν /2 , where ν 2 ≡ Strictly speaking, this is only a mathematical curi-
2
δc /s. In this case, setting S0 = 0 and expanding around ousity, since the conditional distribution fu (s|δ0 , S0 ) is
δ0 = 0 leads to (Mo & White 1996; Mo, Jing & White formally zero for δ0 > δc , but the identity above holds
1997) only when (incorrectly) averaging the expression in (3)
∞ over the full (Gaussian) distribution of δ0 . However, if we
fu (s|δ0 , S0 = 0) X δ0n u forget for the moment about how the bias factors in equa-
=1+ b (ν) , (4)
fu (s) n=1
n! n tion (5) were determined, then the analysis above shows
with the bias coefficients given by that the S0 → 0 limit of the bias coefficients can be recov-
ered by cross-correlating the halo overdensity field with a
δcn bun = ν n−1 Hn+1 (ν) , (5) suitably transformed version of the mass field (the trans-
x2 /2 m −x2 /2
formation uses Hermite polynomials). In particular, our
where Hm (x) = e (−d/dx) e are the “proba- cross-correlation method works for any smoothing scale
bilist’s” Hermite polynomials. For example, n = 1 re- S0 ; there is no requirement that this scale be large (al-
turns the familiar expression for the linear halo bias though, strictly speaking, one does require that S0 < s,
bu1 = (ν 2 − 1)/δc . Note that these coefficients are pure i.e., that the smoothing scale be larger than that used for
numbers, independent of wavenumber k, and (by defini- defining the halos in the first place).
tion) of S0 . It is these scale-independent numbers which There are two important lessons here. First, treat-
are most often used to derive cosmological constraints. ing the S0 → 0 limit of the bias coefficients as though
(Of course, for non-negligible S0 , the Taylor series ex- they are arbitrary is risky: one must be careful to ensure
pansion of equation (2) will yield bias coefficients that that the implied conditional distribution function is sen-
depend on S0 , but this dependence is almost never cal- sible (e.g. positive definite). Except for the coefficients
culated or used.) Since the S0 → 0 limit of equation (3) which come from the more physically motivated excur-
corresponds to setting δc → δc − δ0 in the unconditional sion set approach, this is rarely ever done. We return to
crossing distribution, these bn are simply related to the this point in the next subsection. The second lesson is
nth derivative of sfu (s) with respect to δc . This makes that cross-correlating with appropriate transformations
it easy to see why the Hermite polynomials feature so of the mass field may be an efficient way of isolating the
prominently in much of what follows. different large scale bias coefficients from one another.
One view of this second lesson is to contrast it with the
2.2 A weighted-average definition of bias usual probe of higher order bias factors: 2-point statistics
constrain b1 , 3-point statistics constraint both b1 and b2 ,
If we ignore the fact that the conditional distribution in and so on (Sefusatti & Scoccimarro 2005; Smith et al.
equation (3) should really have δ0 < δc , then it is easy to 2007; Pollack et al. 2012). Since the Hermite polynomials
check that here are polynomials, one may think of the transforma-
Z ∞
tion as picking out that combination of n-point functions
fu (s) = dδ0 fu (s|δ0 , S0 ) pG (δ0 ; S0 ). (6)
−∞
which isolates the dependence on bn . The analysis above

c 0000 RAS, MNRAS 000, 1–16



4 M. Musso et al.
suggests that, once the appropriate transformation has
been made, bn can be determined by a real-space cross- 2
ν -1
correlation measurement alone, and this cross-correlation q-avg
4
be made on any smoothing scale S0 ; there is no require- MC
ment that this scale be large.
sharp-k
∆σ/δc = 0.025
2.3 Scale dependence from appropriate
2 S0/δc2 = 0.25
averaging

δc b1
The previous subsection noted that a naive averaging of
h 1 + δh |δ0 i over a Gaussian distribution appeared to re-
turn the large-scale S0 → 0 bias factors. However, the
correct distribution over which to average is not a Gaus- 0
sian, but
1 h 2 2
i
qu (δ0 , S0 ; δc ) = √ e−δ0 /2S0 − e−(2δc −δ0 ) /2S0 ,
2πS0
(8)
-2
where δ0 < δc (Sheth & Lemson 1999). This is be- -1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6
cause qu (δ0 , S0 ; δc ) gives the probability that the walk lg(y = δc2/s)
had height δ0 at scale S0 , and remained below the bar-
rier δc on all scales S < S0 (Chandrasekhar 1943). It is 10
easy to check that 2 2
ν (ν -3)
Z δc q-avg
fu (s) = dδ0 fu (s|δ0 , S0 ) qu (δ0 , S0 ; δc ) , (9) MC
−∞

as it should.
For similar reasons, whenever one deals with the con- 5 sharp-k
ditional mean h 1 + δh |δ0 i, the appropriate way to com- ∆σ/δc = 0.025
pute cross correlations between the halo overdensity field S0/δc2 = 0.25
δc2 b2

and the mass is by averaging over q and not p, and this


generically makes the measured coefficients depend on
scale S0 as we discuss below. For n = 1, this yields 0
δc 2
 √ 
h δ0 h 1 + δh |δ0 i iq = H2 (ν) + (ν10 + 1)erfc ν10 / 2
S0
q 2
− 2ν10 2
/π e−ν10 /2
(10)
-5
2 2
where ν10 = ν (s/S0 − 1) (equation 17 in Sheth & Lem- -1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6
son 1999). Note that, in contrast to the previous calcula- lg(y = δc2/s)
tion, this quantity yields H2 (ν) only in the limit S0 → 0.
Similarly, averaging (1 + δh )Hn over qu yields a more Figure 1. Monte-Carlo measurement of the cross-correlation
complicated function of S0 . between δh and δ0 (top) and H2 (bottom), for walks with
We have verified these analytical arguments in a uncorrelated steps. See the main text for a description of the
comparison with numerical results. The symbols in Fig- measurement. Solid curves show the prediction associated with
ure 1 show a measurement of the cross-correlation be- averaging over a Gaussian distribution, as is commonly done,
tween δ0 and (1 + δh ) in a Monte Carlo simulation of and which the main text argued was inappropriate, and dashed
curves show the result of averaging using q of equation (8).
random walks with uncorrelated steps. We generate these
walks by accumulating independent Gaussian draws, each
with zero mean and variance (∆σ)2 . For each such walk, δc for these walks is at s > S0 by construction, this mea-
we note the scale s at which it first crossed a constant surement is a q-averaged one.
barrier δc , as well as its height δ0 at a chosen scale The two panels show the measurements for n = 1
S0 . The Figure shows results for ∆σ/δc = 0.025 and and 2, and the solid and dashed curves show the ana-
S0 /δc2 = 0.25. To measure the correlation in a given bin lytic result of averaging using p and q, respectively. The
in y = δc2 /s, we identify those walks that first cross δc solid curve remains the same for all S0 (equation 7) but
in this bin. If these are Ny in number, we compute the the dashed one does not (e.g. equation 10). Therefore,
PNy √
mean j=1 Hn (δ0j / S0 )/Ny where δ0j is the height at the difference between the solid and dashed predictions
n/2
S0 of the j th such walk. Dividing this mean by S0 gives depends on S0 ; we have checked that averaging over qu
the numerical estimate of bn . Since the first-crossing of always yields the correct, scale-dependent value.

c 0000 RAS, MNRAS 000, 1–16



Scale dependent bias 5
While this agreement demonstrates that we have a which is due to a correlation with x3 . Its components
are c̃11 = h x1 x3 i2 /
x23 , c̃22 = h x2 x3 i2 / x23 and



good understanding of just what it is that the excursion
2
set returns, and of the S0 scale-dependence of the bias c̃12 = h x1 x3 i h x3 x2 i / x3 .
coefficients returned by a cross-correlation measurement In the excursion set framework one is interested in
– it has also shown that averaging equation (3) over a p(δ, δ ′ ) and p(δ, δ ′ |δ0 ), where δ ′ is the “curvature” of the
Gaussian distribution (rather than qu ) will lead to incor- walk at scale s, δ ′ = dδ/ds. Since all three quantities δ,
rect estimates of the bias factors and their scale depen- δ ′ and δ0 are essentially linear combinations of the un-
dence. derlying Gaussian-distributed Fourier modes,
both these
The fact that q 6= p leads to measureable differences distributions are also
Gaussian. In this

2 case δ 2 = s,
′ 2

suggests that unless one has a good model of how both h δδ i = (1/2)(d/ds) δ = 1/2 and δ0 = S0 , and the
f and q depend on scale, one must use large survey vol- relevant quantities read
umes (to ensure one is safely in the S0 → 0 limit) if   2  
s 1/2 S× S0 s ǫ× /2
halo bias (e.g. equation 4) is to constrain parameters. C= , c̃ = 2
S0 s ǫ× /2 ǫ2× /4s

At the moment, this understanding exists only for the 1/2 δ ′2


(14)
special case of predictions based on walks with uncorre-
lated steps. Unfortunately, for the q-averaging which we and
have argued is the more appropriate, it is not straightfor- S×  ǫ× 
x̄ = δ0 1, , (15)
ward to separate out the scale independent terms Hn (ν) S0 2s
from those which depend on S0 (e.g., through ν10 ). If where
it were, we would be able to derive cosmological con-
straints from smaller volumes. As it stands, if walks with h δ ′ δ0 i
S× ≡ h δδ0 i and ǫ× ≡ 2s . (16)
uncorrelated steps were a realistic model, then for ha- h δδ0 i
los with ν ∼ 1.3 (mass m ∼ 1013 h−1 M⊙ or Lagrangian For a Gaussian filter, W (kR) = exp(−k2 R2 /2), one has
scales R ∼ 3h−1 Mpc), in order to achieve percent level 2
S× = σ0× and ǫ× = σ1× 2
σ02 /σ0×
2
σ12 , where
accuracy in predicting the scale-independent b1 (b2 ), one
would need to work at scales S0 /δc2 ≃ 0.155 (0.115) or dk k3 P (k) 2j
Z
2
σj× = k W (kR)W (kR0 ) , (17)
Lagrangian scales R0 ∼ 10h−1 Mpc (14h−1 Mpc). We now k 2π 2
turn to a study of these issues for the more realistic case Z
dk k3 P (k) 2j 2
of walks with correlated steps. σj2 = k W (kR) . (18)
k 2π 2
If, in addition, P (k) ∝ kn , then
3 THE EXCURSION SET APPROACH S×  −(n+3)/2
WITH CORRELATED STEPS = 2(n+3)/2 1 + (S0 /s)2/(n+3) , (19)
S0
 −1
We would like to extend the analysis of the previous sec- ǫ× = 2(S0 /s) 1 + (S0 /s)2/(n+3) . (20)
tion to include the effects of correlated steps. To do so,
we must first set up some notation. We will also use the same notation pG (z; σ 2 ) to denote a
one-dimensional Gaussian distribution when there is no
scope for confusion.
3.1 Notation
Let us recall some standard results regarding Gaussian
distributions, which we will use frequently. If the joint 3.2 The unconditional distribution
distribution p(x1 , x2 ) for two variables is the bivariate
Although our goal is to write down the analogue of equa-
Gaussian with zero mean, then
tion (3) for the first crossing distribution associated with
1 T −1
e− 2 x C x walks which are conditioned to pass through δ0 on scale
p(x1 , x2 ) = pG (x; C) ≡ p , (11) S0 , our first step is to write down the unconstrained dis-
(2π)2 Det[C]
tribution. As shown by Musso & Sheth (2012), for a con-
where C is the covariance matrix Cij = h xi xj i. stant barrier of height δc the latter is well-approximated
If the joint distribution p(x1 , x2 , x3 ) for three vari- by
ables is a trivariate Gaussian, then the conditional distri- Z ∞
bution p(x1 , x2 |x3 ) is also a bivariate Gaussian: f (s) = dδ ′ δ ′ p(δc , δ ′ ) , (21)
0
p(x1 , x2 |x3 ) = pG (x − x̄; C − c̃) , (12)
where p(δc , δ ) is the bivariate Gaussian pG (δc , δ ′ ; C) with

where the conditional mean x̄ is proportional to x3 , covariance matrix given in equation (14).
The integral in equation (21) can be performed an-
 
h x1 x3 i h x2 x3 i
x̄ = x3 2
, 2
. (13) alytically and leads to
h x3 i h x3 i
2 √  2 2
The “correction” to the covariance matrix c̃ accounts νe−ν /2 1 + erf Γν/ 2 e−Γ ν /2
 
for that part of the correlation between x1 and x2 sf (s) = √ + √ , (22)
2 2π 2 2πΓν

c 0000 RAS, MNRAS 000, 1–16



6 M. Musso et al.
with
2
γ2 h δδ ′ i S0 / δc2 = 0.075
Γ2 ≡ and γ2 ≡ . (23)
(1 − γ 2 ) h δ 2 ih δ ′2 i 0.6
(Equation 22 corrects a typo in equation 6 of the pub-
lished version of Musso & Sheth 2012.) For later conve-
nience, the same can also be written as
2 2 δ0 = 0

ln(10) yf(y)
e−ν (1+Γ )/2 0.4
sf (s) = (1 + A) , (24)
(2π)(2Γ)
where δ0 = √S0

  
1 Γν 2 2
A ≡ A(ν) = 1 + erf √ 2π Γν eΓ ν /2 . (25) 0.2
2 2
Musso & Sheth (2012) showed that this approximation
δ0 = − √S0
(as well as its generalisation to moving barriers) works
extremely well over a large range of scales for a range
of choices of power spectra and filters (including TopHat 0
filtered LCDM) when compared with Monte Carlo solu- -0.5 0 0.5 1
tions of the first crossing problem. (It breaks down in the lg(y = δc2/s)
limit in which the walks must have taken many steps to
cross the barrier.) Our final analytic results in this paper
will be valid for arbitrary power spectra and filters for a S0 / δc2 = 0.25
constant barrier. However, since explicit expressions for 1.2
various quantities greatly simplify for the choice of Gaus-
sian smoothing of power law power spectra, we will show 1 δ0 = √S0
comparisons with Monte Carlo solutions for the latter.
For Gaussian smoothing, γ = σ12 /σ0 σ2 .
0.8
ln(10) yf(y)

3.3 The conditional distribution


0.6 δ0 = 0
Musso & Sheth (2012) argued that the conditional dis-
tribution corresponding to (21) is simply 0.4
Z ∞
f (s|δ0 , S0 ) = dδ ′ δ ′ p(δc , δ ′ |δ0 ) , (26)
0 0.2

where p(δc , δ |δ0 ) is the probability that the walk had a δ0 = − √S0
height δc and curvature δ ′ at scale s, given that it passed 0
through δ0 at scale S0 < s. In principle, one is really in- -0.5 0 0.5 1
terested in imposing the stronger condition that the walk lg(y = δc2/s)
must have passed through (δ0 , S0 ) without having crossed
δc before S0 . We will return to this point later and argue
Figure 2. First crossing distribution of a barrier of height δc
that the effects of ignoring this stronger requirement are by the subset of walks which are conditioned to pass through
small. (δ0 , S0 ), for a few choices of δ0 (as labelled). Short dashed,
The conditional distribution p(δc , δ ′ |δ0 ) is the bivari- solid and long-dashed curves show the analytic prediction from
ate Gaussian equation (28) for Gaussian smoothing of a Gaussian field with
P (k) ∝ k −1.2 .
p(δc , δ ′ |δ0 ) = pG (∆ − x̄; C − c̃) , (27)
with ∆ ≡ (δc , δ ′ ), C and c̃ given by equation (14) and the
conditional mean x̄ given by equation (15). For generic
power spectra and filters, the integral in equation (26)
where
can be performed analytically, exactly as in the case of
equation (21), and expressed in terms of S× /S0 and ǫ× . S×


2
S0
The result is δc× ≡ δc − δ0 ; Q≡1− , (29)
S0 S0 s
2
/2sQ
δ̄ ′ e −δc×
f (s|δ0 , S0 ) = √
2πsQ
" √  ′2 2
#
1



S× S0

1 + erf δ̄ ′ / 2σ̄ e−δ̄ /2σ̄ δ̄ ′ ≡ δ ′ |δc , δ0 =


δc× + ǫ× δ0 − δ c ,
× +√ , (28) 2sQ S0 S0 s
2 2π(δ̄ ′ /σ̄) (30)

c 0000 RAS, MNRAS 000, 1–16



Scale dependent bias 7
fact that the values of δ0 being considered in Figure 2
are significantly smaller than δc , so that very few of the
10 S0 / δc2 = 0.25 walks would have reached the barrier prior to S0 and
then returned to pass through δ0 at S0 . One can then
δ0 = 1.8√S0 ask whether the expression in equation (28) would con-
tinue to be accurate even for δ0 . δc , since this is the
regime of interest for calculations of merger rates.
1
δ0 = 0 We test this in Figure 3, which compares equa-
ln(10) yf(y)

tion (28) with the Monte Carlo solution for the same
choice of conditioning scale S0 as in the lower panel of
Figure 2, but with a larger magnitude for δ0 which is now
|δ0 | = 0.9δc . We see that for δ0 = +0.9δc , the numerical
0.1
solution has a sharp peak which is very well described
by equation (28). The latter starts overestimating the
numerical answer around log10 ν 2 ≃ 0.4, which can be
δ0 = -1.8√S0 understood as follows.
Paranjape et al. (2012) demonstrated in their Fig-
0.01
-0.5 0 0.5 1 ure 7 that the numerical conditional distributions are, to
lg(y = δc2/s) a good approximation, related to the corresponding un-
conditional one by √ a simple scaling relation which sends
ν → ν10 = δc× / sQ in the unconditional distribution.
Figure 3. Same as lower panel of Figure 2, but for a larger This is also approximately true of the analytic expres-
value of |δ0 |, and note that now the y-axis is on a log scale. sion in equation (28). Since equation (24) is not a good
The analytic prediction equation (28) for δ0 = 0.9δc describes
approximation to the unconditional first crossing distri-
the sharp peak in the numerical solution remarkably accu-
rately. It begins overestimating the numerical solution around
bution at small values of ν (Musso & Sheth 2012), it
log10 ν 2 ≃ 0.4, at which point about 75% of the probability follows that the corresponding analytic conditional dis-
has been accounted for (see text for why this happens). tribution will not be a good approximation at small ν10 .
One can check that, for the choices of S0 and δ0 in Fig-
ure 3, ν10 actually passes through zero and becomes neg-
and ative around log 10 ν 2 ≃ 0.5. So the mismatch between
2
Γ2 S 0 S × (1 − ǫ× )2 the analytic prediction Rand the numerical solution is not
 
1
σ̄ 2 ≡ Var(δ ′ |δc , δ0 ) =1 − . − ln S0
4Γ2 s Qs S02 surprising. In practice, 0.4 ln 10 d ln y yf (y|S0 ) = 0.75, in-
(31) dicating that the prediction is inaccurate only for the 25%
Note that, in contrast to equation (3), this expression which cross at the largest values of s (smallest values of
for the conditional distribution remains positive definite y).
even when δ0 > δc , although it is understood that only
δ0 ≤ δc is sensible. For future reference, the sharp k-space
filter has S× /S0 = 1 and ǫ× = 0; its conditional crossing 3.5 Halo bias with correlated steps
distribution, equation (24), corresponds to inserting these Now that we have in hand a good approximation to the
values in equation (28) and replacing the term in square conditional first crossing distribution, we can turn to the
brackets with a factor of 2. associated description of halo bias.
The first issue that we would like to address is if Her-
mite polynomials of the smoothed matter density field
3.4 Comparison with Monte Carlo solution
are still special. Appendix B suggests that they are, as
Figure 2 compares the prediction in equation (28) with long as the underlying matter density field is Gaussian.
a Monte Carlo solution of the conditional first crossing More formally, we show √ there that the rôle of the Hermite
distribution. The comparison is for Gaussian filtered ran- polynomial Hn (δ0 / S0 ) in the average is that of remov-
dom walks using a power spectrum P (k) ∝ k−1.2 . The ing from it all the disconnected parts, so that only the
numerical treatment uses the algorithm of Bond et al. connected part of the expectation value of δ0n remains.
(1991) and was described in Paranjape et al. (2012). The Secondly, for reasons discussed in section 2.3, in
histograms are the same as in Figure 6 of Paranjape et principle we must specify the probability distribution
al. and show the distribution of first crossing scales for q(δ0 , S0 ; δc ) to be used in the average. In the present
a constant barrier, for walks that were required to pass case, q is not known analytically. However, in the spirit
through the indicated values of δ0 at scale S0 , for two of Musso & Sheth (2012), we can argue that the error
choices of S0 . We see that the analytic prediction works in ignoring the difference between p and q is of the same
very well in describing the numerical solution. order as that already included in f (s|δ0 , S0 ). Indeed, the
This good agreement is despite the fact that equa- fact that the conditional distributions shown in Figure 2
tion (26) formally ignores walks which might have crossed are such an accurate description of the numerical solution
the barrier prior to S0 . This can be understood by the means that, in this case, the approximation is consistent.

c 0000 RAS, MNRAS 000, 1–16



8 M. Musso et al.

1.5
brings these into the form (see Appendix A2)
δc pG(δ0;S0) (−S× /S0 )n ∞ ′ ′
Z 
∂ ǫ× ∂
n
δc q(δ0,S0;δc) 2 bn = dδ δ + p(δc , δ ′ ) ,
S0 = 30∆σ f (s) 0 ∂δc 2s ∂δ ′
(33)
Gaussian, n=-1.2 with f (s) given in equation (21). Appendix A3 shows that
∆σ/δc = 0.05
probability distribution

1 ∞ n Z ∞
δ0n

X S×
f (s|δ0 , S0 ) = − dδ ′ δ ′
n=0
n! S 0 0
 n
∂ ǫ× ∂
× + pG (∆; C − c̃) , (34)
∂δc 2s ∂δ ′
2
S0 = 100∆σ
0.5 holds exactly for the distribution (26), where ∆ = (δc , δ ′ )
and the matrices C and c̃ were defined in equation (14).
S0 = 300∆σ
2 Since the bivariate Gaussian pG (∆; C) is precisely the
distribution p(δc , δ ′ ) that appears in equation (33), we
clearly have

0 f (s|δ0 , c̃ = 0) X δ0n
-2 -1.5 -1 -0.5 0 0.5 1 =1+ bn . (35)
δ0 / δc f (s) n=1
n!

Setting c̃ = 0 corresponds to the following assignments


Figure 4. Distribution of the height δ0 on scale S0 of walks in equation (28):
which have not crossed δc prior to S0 , for a range of choices √
of S0 (histograms), measured in the same Monte-Carlo simu- Q→1 ; σ̄ s → (2Γ)−1 ;
lations which were used to make the Figures 2 and 3. Dashed δ̄ ′ /σ̄ → Γν [1 − (δ0 /δc )(S× /S0 )(1 − ǫ× )] . (36)
curves show that a Gaussian, truncated at δ0 = δc , provides
a good approximation. Note that S0 /δc2 = 300 × 0.052 = 3/4 As a result (see Appendix A4) the bias coefficients can
corresponds to smoothing scales which are of order that asso- be reduced to:
ciated with a typical halo: therefore, if one restricts attention  n

to smaller S0 , then ignoring the truncation of the Gaussian δcn bn = (αn + βn + γn ) , (37)
should be a good approximation. S0
where
αn = ν n Hn (ν) , n ≥ 1 , (38)
(
1 −A (1 − ǫ× ) ,n=1
This is a consequence of the fact that for correlated steps βn = (39)
1 + A (1 − ǫ× )n (Γν)n Hn−2 (Γν) ,n≥2
zig-zags are exponentially rare at small S0 ; in this limit,
p ≈ q. We can test this explicitly by looking directly at
(
0 , n=1
the distribution of q in our Monte-Carlos. Figure 4 shows γn = Pn−1 n (40)
k=1 k αk βn−k , n ≥ 2,
that, for S0 values which are smaller than those associ-
ated with typical halos, the difference between q and the where A was defined in equation (25).
Gaussian is almost negligible. There are some interesting parallels with the calcu-
Motivated by this simplification, let us define the lation for sharp-k walks, and some important differences.
real-space bias coefficients associated with the condi- There is obviously a close analogy between the S0 → 0
tional distribution f (s|δ0 , S0 ) using limit of sharp-k walks and the c̃ → 0 limit for correlated
steps, especially since the matrix c̃ is proportional to S0
1 D √ E
(c.f. equation 14, noting that the factor S× /S0 becomes
bn ≡ n/2
(1 + δh )Hn (δ0 / S0 )
S constant as S0 → 0). However, in the present case one is
Z 0∞ not throwing away all the dependence on S0 , since fac-

= dδ0 pG (δ0 ; S0 ) h 1 + δh |δ0 , S0 i Hn (δ0 / S0 ) , tors of ǫ× explicitly appear in the expression for the bn . In
−∞
(32) particular, these factors of ǫ× would not have appeared
if we had simply taken derivatives of the unconditional
distribution (equation 24) with respect to δc . This has an
with h 1 + δh |δ0 , S0 i given in equation (2). Below we will
important consequence: for sharp-k filtering, the quanti-
show comparisons between numerical measurements of
ties bn were independent of S0 , whereas here they depend
these quantities (q-averaged by construction) with ana-
explicitly on S0 . If we write equation (37) as
lytic results using the p-averaged expression in the second
n
line of (32). From the discussion above, we expect these X
to match well at least for the smallest S0 shown in Fig- bn ∼ (S× /S0 )n bnk ǫk× , (41)
k=0
ure 4.
For f (s|δ0 , S0 ) given by equation (26), some algebra then it is the quantities bnk (rather than bn ) which are

c 0000 RAS, MNRAS 000, 1–16



Scale dependent bias 9

8 related, one might expect equation (33) to be quite ac-


S0 = 30∆σ
2 curate. We test this explicitly in Figure 5 by compar-
S0 = 100∆σ
2 ing the results of evaluating the r.h.s. of equation (33)
S0 = 300∆σ
2 for n = 1 and n = 2 with corresponding measurements
6
(performed as described in section 2.3) using the same
Gaussian, n=-1.2 Monte Carlo simulations that were used in Figures 2
∆σ/δc = 0.05 and 3. By construction, the numerically estimated quan-
4 tity is q-averaged, whereas the analytic curves show the
Gaussian-averaged coefficients in equation (32). The an-
δc b1

alytic predictions closely track the measurements over a


2 range of s-values for several choices of S0 . There are small
systematic deviations which are likely due to a combina-
tion of the facts that q 6= p at large S0 and that the
analytic prediction fails to be a good approximation at
0
large s.
Since ignoring the difference between p and q is a
good approximation, one might wonder if the effect of ǫ×
-2 can also be ignored; naively one expects the q-averaging
-0.5 0 0.5
to be irrelevant at small S0 /s where ǫ× is also likely to
lg(y = δc2/s)
be small. Figure 6 shows the results for b1 and b2 for one
of the choices of S0 from Figure 5, comparing the same
60
2 measurements as in that figure with analytic expressions
S0 = 30∆σ
2 in which ǫ× is retained as per equations (39) and (40)
S0 = 100∆σ
2 (solid curves) or set to zero by hand (dashed curves). We
S0 = 300∆σ
Gaussian, n=-1.2 see that the terms involving ǫ× contribute significantly
40 ∆σ/δc = 0.05 and must be retained to get an accurate description of
the bias.
δc2 b2

3.6 Recovery of scale-independent bias factors


20 The bias coefficients in Figure 5 show a strong depen-
dence on the scale S0 . This is rather different from the
case of sharp-k filtering, for which the bn recovered from
p-averaging (equation 7) were independent of S0 . Indeed,
the scale-independence of the recovered bn was one of
0
our motivations for cross-correlating with the Hermite-
transformed field in the first place, so it is interesting to
ask if the dependence on S0 can be removed.
-0.5 0 0.5 This turns out to be possible because of the fol-
lg(y = δc2/s) lowing. First, the scale dependence of bn is almost en-
tirely due to the factors of S× and ǫ× (the other effect
Figure 5. Comparison of bias coefficients b1 and b2 of equa- comes from the small difference between p and q averag-
tion (37) (smooth curves) with corresponding measurements ing). And secondly, equation (37) shows that the scale-
(points with Poisson errors) in the same Monte Carlo sim- independent bnk are linearly related to each other in such
ulations used in Figures 2 and 3, for a range of S0 values. a way that measuring b1 , . . . , bn is sufficient to recover all
The measurements were performed as described in section 2.3.
the b1k , . . . bnk .
The analytic prediction clearly tracks both the s- and S0 -
dependence fairly accurately. There are small systematic dif- We demonstrate this explicitly for n = 1 and 2. For
ferences, especially at large s, which arise because q 6= p at n = 1, we can write
large S0 , and our analytic approximation to f (s|δ0 , S0 ) stops 1 S×

A

A

being a good approximation when s ≫ S0 . b1 = ν2 − + ǫ×
δc S 0 1+A 1+A

≡ (b10 + ǫ× b11 ) . (42)
S0
scale-independent. This will be important below when in-
terpreting our results in terms of Fourier-space bias. Note Since
that the bn0 are the peak-background split parameters δc b11 = ν 2 − δc b10 , (43)
f −1 (−∂/∂δc )n f which are of most interest in cosmologi-
cal applications. This is obvious upon setting ǫ× → 0 in we can estimate
equation (33). δc (S0 /S× )b1 − ǫ× ν 2
Since p ≈ q, in contrast to when steps are uncor- δc b10 = . (44)
1 − ǫ×

c 0000 RAS, MNRAS 000, 1–16



10 M. Musso et al.
discussion, we find
with εcross
δc2 b21 = ν 2 (δc b10 − 1) − δc2 b20
10 w/o εcross
peak-bg split δc2 b22 = δc2 b20 + ν 2 (ν 2 − 2δc b10 + 1) . (46)
8 Hence,
  2
1 S0
6 δc2 b20 = δ 2
c b2
Gaussian, n=-1.2 (1 − ǫ× )2 S×
∆σ/δc = 0.05
δc b1

 
S0 = 30(∆σ)2
S0
4 − ǫ× ν 2 2δc b1 − ǫ× (ν 2 − 1) − 2 . (47)

We have deliberately isolated the peak-background split
2
parameters bn0 above. From the structure of the coeffi-
cients in equation (37) it is clear that this reconstruction
0 can be extended to the higher order coefficients as well.
The dotted curves in Figure 6 show the analytic pre-
-2 dictions for b10 and b20 from equation (37), while the
-0.5 0 0.5 1 triangular symbols show the numerical estimates using
lg(y = δc2/s) equations (44), (47) and the corresponding measurements
of b1 and b2 . Clearly, the reconstruction works well. More-
over, since we are working at finite S0 , our procedure
110 with εcross has allowed a simple and direct estimate of the peak-
w/o εcross background split parameters bn0 from a measurement of
peak-bg split scale-dependent bias, without having to access very large
90 scales. E.g., the Figure shows results for S0 = 0.075 δc2 ,
which corresponds to the scale associated with a ν ≈ 3.7
70 halo and a Lagrangian length scale of R0 ∼ 17h−1 Mpc;
most other analyses of halo bias are restricted to length
δc2 b2

scales which are several times larger.


50 Gaussian, n=-1.2 Another way to see this is to notice that, in the ex-
∆σ/δc = 0.05 pressions above, bn → bn0 when ǫ× → 0. Since ǫ× → 0 on
30 S0 = 30(∆σ)2 large scales, the analysis above shows explicitly that our
method for reconstructing bn0 works even on the smaller
scales where ǫ× 6= 0. Indeed, although we have concen-
10
trated on isolating bn0 , the analysis above shows that we
can isolate the other bnk as well. For example, having
-10 measured b1 and b2 using our Hermite-weighting scheme,
-0.5 0 0.5 1
and having used equations (44) and (47) to estimate b10
lg(y = δc2/s)
and b20 , equation (46) furnishes estimates of b21 and b22 .
The expressions above show that our method will
Figure 6. Same as Figure 5 for one choice of S0 , but compar- break if ǫ× = 1, which happens when s → S0 . This is
ing the numerical answer for b1 and b2 with the analytic pre- not surprising since this is the limit in which the large
diction equation (37) when the dependence on ǫ× is retained scale environment is the same as that on which the halo
(solid red) or set to zero by hand (dashed blue). Clearly, re- was defined, so our expressions for the conditional dis-
taining the dependence on ǫ× is important, indicating that
tribution are becoming ill-defined. Since this regime is
our method is sensitive to the k-dependence of bias (see text).
substantially smaller than the one of most interest in
Black triangles show the result of implementing the recursive
procedure described in the text for reconstructing the usual (k- cosmology, we conclude that our method allows a sub-
independent) peak-background split parameters b10 and b20 stantial range of interesting scales to provide estimates
(dotted curves) from these measurements. Although defined of the bias factors bnk .
at finite S0 , the procedure works well in reproducing the S0 -
independent bn0 .
3.7 Real and Fourier-space bias
Similarly, The appearance of ǫ× in the real-space expressions for bn


2 generically indicates that the bias in Fourier space must
b20 + 2ǫ× b21 + ǫ2× b22 ,

b2 = (45) be k-dependent. This is most easily seen with b1 using a
S0 2 2
Gaussian filter W (kR) = e−k R /2 .
where the excursion set predictions for the coefficients b2j
Suppose that
can be read off from equation (37). For example, δc2 b21 =
ν 2 (A − Γ2 )/(1 + A). But, more relevant to the present δ0 (k) = δ(k)W (kR0 ) (48)

c 0000 RAS, MNRAS 000, 1–16



Scale dependent bias 11
and constrained to pass through a specified position (equa-
tion 28), and showed that it was accurate (Figure 2).
δh (k) = b1 (k)δ(k)W (kR) , (49)
Although this is interesting in its own right, we did not
so that in real space h δh |δ0 i = δ0 h δh δ0 i /S0 . Then equa- explore this further. Rather, we used it to provide a sim-
tion (42) implies that ple analytic expression for the large scale halo bias fac-
tors (equation 37), showing that, as a result of corre-
k2 s
b1 (k) = b10 + b11 . (50) lations between scales, real space measures of halo bias
σ12
are scale dependent (equation 41 and Figure 5), but this
This shows that the excursion set analysis makes a pre- scale dependence is best thought of as arising from k-
diction for how the Fourier -space coefficients b10 and b11 dependent bias in Fourier space (Section 3.7). Although
should depend on ν = δc /σ and Γ. we presented comparisons with numerical results for a
It is remarkable that peaks theory predicts this same specific choice of filter (Gaussian) and power spectrum
structure (constant plus k2 ) for the linear Fourier space (P (k) ∝ k−1.2 ), the results of Musso & Sheth (2012) lead
bias factor (Desjacques et al. 2010). Although the coef- us to expect that our analytical results will be equally
ficients b10 and b11 for peaks differ from that for the ex- accurate for other filters and power spectra, including
cursion set halos studied here, the relation (43) between TopHat filtered ΛCDM.
these coefficients is the same. We have checked explic- For correlations which arise because of a Gaussian
itly that peaks also satisfy the relationships between the smoothing filter, the linear bias factor b1 is a constant
second order bias coefficients as shown in equation (47) plus a term which is proportional to k2 . This is a con-
(although the actual values of b20 , b21 and b22 are differ- sequence of the fact that our analysis is based on the
ent), and so we expect this correspondence between the approximation of Musso & Sheth (2012), which asso-
k-dependence of peak and halo-bias will hold for all n. ciates halos with places where the height of the smoothed
Because this correspondence is seen in two very different field and its first derivative with respect to smoothing
analyses (excursion sets and peaks), there is likely to be scale satisfy certain constraints. If constraining the sec-
a deeper reason for its existence. ond derivative as well leads to an even more accurate
We explore this further in Appendix B where we dis- model of the first crossing distribution, then this would
cuss the relation between our analysis and the work of give rise to k4 -dependence in the bias. It is in this sense
Matsubara (2011) who has argued that k-dependent bias that k-dependent halo bias is part and parcel of the
factors are generically associated with nonlocal biasing excursion set approach. Such k-dependence will lead to
schemes. He provides a number of generic results for such stochasticity in real space measures of bias (Desjacques
nonlocal bias, noting that the Fourier-space structure at & Sheth 2010); we have not pursued this further.
order n which can be written in terms of what he calls We also provided an algorithm for estimating the
renormalized bias coefficients cn (k1 , . . . , kn ). For peaks scale-independent coefficients of the k-dependent bias
theory, factors from real space measurements (Section 3.6).
X 2 X 2 2
cn (k1 , . . . , kn ) = bn0 + bn1 ki + bn2 ki kj + . . . . Although the method uses cross-correlations between
i i<j the halo field and suitably transformed versions of the
(51) smoothed mass field at the same spatial position (equa-
tion 32), the bias factors it returns are independent of the
In this case, for Gaussian initial conditions, the Hermite-
scale on which this transformation is done (Figure 6). In
weighted averages (with a Gaussian filter as per equa-
particular, the coefficient of the k-independent part of the
tion B2) show a structure that is identical to our excur-
bias which our algorithm returns equals that associated
sion set predictions of (41). More generally, our Hermite-
with the peak-background split argument, even though
weighting scheme provides a practical way of measuring
our algorithm can be applied on scales for which the usual
integrals of Matsubara’s renormalized bias coefficients cn .
formulation of the peak-background split argument does
We therefore conclude that our real-space Hermite-
not apply.
weighted prescription for measuring halo bias can allow
For Gaussian fields, the transformation we advocate
us to separate the scale-dependent contribution to bias as
uses the Hermite polynomials. Therefore, our work has an
well as isolate the scale-independent (peak-background
interesting connection to Szalay (1988) who noted that,
split) part arising from each order n, which traditional
instead of defining bias coefficients by writing δh as a
Fourier-space measurements cannot do. The specific re-
Taylor series in δ0 as is usually done, one could have
sults of our excursion set analysis (e.g., the relations be-
chosen to expand the mass field in Hermite polynomials.
tween the bnk ) are then predictions that can be tested in
Our analysis shows that this is indeed a fruitful way to
more realistic settings such as N -body simulations. But
proceed, even when the bias factors are k-dependent.
this is beyond the scope of the present work.
There are two reasons why this is remarkable. First,
our analysis shows that, for the excursion set model, the
coefficients of the expansion in δ0 are the same as those
4 CONCLUSIONS AND DISCUSSION
for the expansion in Hermite polynomials (equations 32
We provided an analytic approximation for the first cross- and 35). There is no reason why this should be true in
ing distribution for walks with correlated steps which are general. And second, Szalay explicitly assumed that halo

c 0000 RAS, MNRAS 000, 1–16



12 M. Musso et al.
bias was ‘local’: δh was a function of δ0 only. For local relation in the δh -δ0 scatter plot. A more detailed com-
bias, the bias factors are k-independent; k-dependent bias parison with traditional techniques is complicated by the
factors are a signature that the bias is nonlocal (Matsub- fact that we have made predictions for Lagrangian bias
ara 2011, with the k-dependence of peak bias discussed whereas analyses such as Manera & Gaztañaga’s typi-
in Desjacques et al. 2010 being a specific example), so cally work in the final, Eulerian field. We leave such a
it is not a priori obvious that an expansion in Hermite comparison to future work.
polynomials would have been useful. In this context, it is worth noting that our algo-
In Appendix B we showed why, even for nonlocally rithm is more than just a simple way of estimating the
biased tracers of a Gaussian field, the Hermites are so nonlinear bias coefficients bn . For example, there has
special. For completeness, we also provided an analysis been recent interest in reducing the stochasticity between
of the general case, in which the underlying field is not the underlying mass field and that defined by the bi-
necessarily Gaussian (equation B18). This more general ased tracers (Hamaus et al. 2010; Cai et al. 2011). Some
analysis may prove useful should it turn out that the of this stochasticity is due to the nonlinear nature of
primordial fluctuation field was non-Gaussian, or if one the bias (Hamaus et al. 2011). Our demonstration that
wishes to describe halo bias with respect to the nonlinear the nonlinear bias factors measure the amplitude of the
Eulerian field rather than with respect to the initial one. cross-correlation function between the halo field and the
In the former case, primordial non-Gaussianity is ex- Hermite-transformed mass field will simplify such analy-
pected to be sufficiently weak that the Edgeworth ex- ses.
pansion can be used to provide insight into the expected
modifications to halo abundances. Since Hermites play an
important role in the Edgeworth expansion, it is likely
that our Hermite-based algorithm for halo bias will be ACKNOWLEDGEMENTS
useful for constraining fNL . We are grateful to E. Sefusatti and M. Simonovic for
Recent work has emphasized the advantages of us- discussions. MM is supported by ESA under the Bel-
ing cross- rather than auto-correlations to estimate halo gian Federal PRODEX program N◦ 4000103071. This
bias (Smith et al. 2007; Pollack et al. 2012). Since Hn work is supported in part by NSF 0908241 and NASA
is an n-th order polynomial in the mass field, one may NNX11A125G.
think of our algorithm as an extension of this program:
it uses two-scale halo-mass cross-correlations at the same
real-space position to extract information which is usu-
ally obtained from n-point statistics. However, in addi- REFERENCES
tion to being simpler, our algorithm is able to estimate
Bond J. R., Cole S., Efstathiou G., Kaiser N., 1991, ApJ, 379,
the bias coefficients on smaller scales than those on which
440
the more traditional analyses n-point (Fourier or real- Cai Y.-C., Bernstein G., Sheth R. K., 2011, MNRAS, 412, 995
space) analyses are performed. So we expect it to find Chan K. C., Scoccimarro R., 2012, arXiv:1204.5770
use in analyses of halo bias in simulations, and galaxy Chandrasekhar S., 1943, Rev. Mod. Phys., 15, 1
bias in real datasets. Dekel A., Lahav O., 1999, ApJ, 520, 24
For example, one can compare our prescription with Desjacques V., Crocce M., Scoccimarro R., Sheth R. K., 2010,
traditional methods of estimating bias in real space, e.g. PRD, 82, 103529
Manera & Gaztañaga (2012). Here, instead of comput- Desjacques V., Sheth R. K., 2010, PRD, 81, 023526
ing averages of the matter field centered at locations of Frusciante N., Sheth R. K., 2012, JCAP, submitted
(arXiv:1208.0229)
halos (as is natural in the excursion set approach), one
Fry J., Gaztañaga E., 1993, ApJ, 413, 447
explicitly defines a halo field δh (x) smoothed on a grid
Hamaus N., Seljak U., Desjacques V., Smith R. E., Baldauf
of cell-size R0 and uses the matter field δ0 (x) smoothed T., 2010, PRD, 82, 043515
on the same grid. One then fits a polynomial of the Hamaus N., Seljak U., Desjacques V., 2011, PRD, 84, 083509
type δh = b0 + b1 δ0 + b2 δ02 /2 to a scatter plot of δh Kaiser N., 1984, ApJ, 284, L9
vs. δ0 using a least squares prescription. This is concep- Lacey C., Cole S., 1993, MNRAS, 262, 627
tually the same as approximating the function h δh |δ0 i Lam T. Y., Li B., 2012, MNRAS, submitted, arXiv:1205.0059
(which is most easily seen by considering linear biasing Maggiore M., Riotto A., 2010, ApJ, 711, 907
of a Gaussian field, for which the statement is exact). Manera M., Scoccimarro R., Sheth R. K., 2010, MNRAS, 402,
This can be compared with the excursion set prediction 589
h 1 + δh |δ0 i = f (s|δ0 , S0 )/f (s), and we see that the coeffi- Manera M., Gaztañaga E., 2012, MNRAS (to appear),
arXiv:0912.0446
cients obtained from the fit will generically depend on S0 .
Matsubara T., 1995, ApJS, 101, 1
As Manera & Gaztañaga show, one needs to define a grid
Matsubara T., 2011, PRD, 83, 083518
on very large scales (R0 & 40h−1 Mpc) in order to recover Mo H. J, Jing Y. P., White S. D. M., 1997, MNRAS, 284, 189
scale independent bias coefficients. On the other hand, Mo H. J., White S. D. M., 1996, MNRAS, 282, 347
our prescription can in principle operate at much smaller More S., van den Bosch F., Cacciato M., More A., Mo H.,
scales (c.f. section 3.6) and remove this scale dependence Yang X., 2012, arXiv:1204.0786
by basically computing weighted integrals of the mean Musso M., Sheth R. K., 2012, MNRAS, 423, L102

c 0000 RAS, MNRAS 000, 1–16



Scale dependent bias 13
Paranjape A., Lam. T. Y., Sheth R. K., 2012, MNRAS, 420, then follows from equation (A1). Together with sfu (s) =
2
1429 (2π)−1/2 ν e−ν /2 , this gives the result.
Paranjape A., Sheth R. K., 2012, MNRAS, 419, 132
Parfrey K., Hui L., Sheth R. K., 2011, PRD, 83, 063511
Peacock J. A., Heavens A. F., 1990, MNRAS, 243, 133
Pollack J. E., Smith R. E., Porciani C., 2012, MNRAS, 420, A2 Form of bias coefficients in equation (33)
3469
Press W. H., Schechter P., 1974, ApJ, 187, 425 The weighted average of the distribution (26) is
Sefusatti E., Crocce M., Pueblas S., Scoccimarro R., 2006, D √ E
PRD, 74, 023522 f (s|δ0 , S0 )Hn (δ0 / S0 )
Sefusatti E., Scoccimarro R., 2005, PRD, 71, 063001 Z ∞ Z  
Sheth R. K., Lemson G., 1999, MNRAS, 304, 767 δ0
= dδ ′ δ ′ dδ0 pG (δ0 ; S0 )Hn √ p(δc , δ ′ |δ0 ) .
Sheth R. K., Mo H. J., Tormen G., 2001, MNRAS, 323, 1 0 S0
Sheth R. K., Tormen G., 1999, MNRAS, 308, 119 (A4)
Sheth R. K., Chan K.-C., Scoccimarro R., 2012, PRD, sub-

mitted (arXiv:1207.7117) The product pG (δ0 ; S0 )Hn (δ0 / S0 ) and the bivariate
Smith R. E., Scoccimarro R., Sheth R. K., 2007, PRD, 75, ′
Gaussian p(δc , δ |δ0 ) (equation 27) can be expressed in
063512 terms of their Fourier transforms: i.e., we use equa-
Szalay A. S., 1988, ApJ, 333, 21
tion (A1) and
Zehavi I., et al., 2011, ApJ, 736, 59
d2 k ikT (∆−x̄) − 21 kT (C−c̃)k
Z
p(δc , δ ′ |δ0 ) = e e , (A5)
(2π)2
APPENDIX A: DETAILS OF
with ∆ = (δc , δ ′ ) and x̄, C and c̃ given by
CALCULATIONS
equation (15) and (14), respectively. The integral
In this Appendix we sketch the proofs of various identities over δ0 then gives a one-dimensional Dirac delta
stated in the main text. δD (k0 − kS× /S0 − k′ ǫ× S× /2sS0 ) where k0 , k and k′ are
the Fourier variables corresponding to δ0 , δc and δ ′ , re-
spectively. Performing the k0 integral gives an expression
A1 Proof of equation (7) in which the contribution of the “correction” matrix c̃
To prove equation (7) for sharp-k walks, it is useful to exactly cancels. The result can be expressed as
consider the following Fourier transform relations involv- 1 D √ E
ing the Hermite polynomials, which follow from the def- n/2
f (s|δ 0 , S 0 )H n (δ0 / S 0 )
S0
inition of the Hn :  n Z ∞  n
S× ′ ′ ∂ ǫ× ∂
2
e−x /2
Z ∞
dk ikx = − dδ δ + p(δc , δ ′ ) , (A6)
√ Hn (x) = e (−ik)n e−k /2
2
S0 0 ∂δc 2s ∂δ ′
2π −∞ (2π)
Z ∞ and using h 1 + δh |δ0 , S0 i ≡ f (s|δ0 , S0 )/f (s) gives the re-
n −k2 /2 dx 2
(−ik) e = √ e−ikx e−x /2 Hn (x) . (A1) sult (33).
−∞ 2π
For the conditional first crossing distribution of equa-
tion (3), we use the relation
  A3 Taylor expansion of the conditional first
∂ crossing distribution in equation (34)
sfu (s|δ0 , S0 ) = s − pG (δ0 − δc ; s − S0 ) . (A2)
∂δc
√ √ Using equation (27) and the shorthand notation pG for
Using y0 ≡ δ0 / S0 and ν = δc / s one can write pG (∆; C − c̃) where ∆ = (δc , δ ′ ) and the matrices C and
D √ E c̃ were defined in equation (14), straightforward algebra
sfu (s|δ0 , S0 )Hn (δ0 / S0 ) shows that
2 √
−ν s/S0 )2
dy0 e−y0 /2 Hn (y0 ) − (y02(s/S
Z ∞

=− e 0 −1) p(δc , δ ′ |δ0 )
∂ν −∞ 2π
p
1 − S0 /s ∞ m  n
(−δ0 S× /S0 )m+n

 n/2 X ∂ ǫ× ∂
S0 ∂
Z ∞
dk 2 = pG
=− (−ik)n eikν−k /2 m,n=0
m!n! ∂δc 2s ∂δ ′
s ∂ν −∞ 2π
k
∞ k X ! n  k−n
δ0k

 n/2
S0 1 2
X S× k ∂ ǫ× ∂
= √ e−ν /2 Hn+1 (ν) , (A3) = − pG
s k! S0 n ∂δc 2s ∂δ ′
2π k=0 n=0
∞ k  k
δ0k

where the second equality follows from writing the X S× ∂ ǫ× ∂
= − + pG (∆; C − c̃) .
Fourier integrals corresponding topthe Hermite polyno- k=0
k! S0 ∂δc 2s ∂δ ′
mial and the Gaussian in (y0 − ν s/S0 ), doing the in- (A7)
tegral over y0 to give a Dirac delta and using this to
perform one Fourier-space integral. The third equality Using this in the definition (26) proves equation (34).

c 0000 RAS, MNRAS 000, 1–16



14 M. Musso et al.
A4 Explicit expressions for the bias coefficients ρh (x) ≡ 1 + δh (x). In this case,
in equation (37) √
1 D E
bn = n/2 (1 + δh )Hn (δ0 / S0 )
The explicit form of the conditional distribution (26) in S0
the limit c̃ → 0 allows for a more convenient calculation
d 3 k1 d 3 kn
Z
1
of the bias coefficients than computing the derivatives in = n 3
... P1 . . . Pn W1 . . . Wn
S0 (2π) (2π)3
equation (33). Using the relations (36) in equation (28)
brings the conditional first crossing distribution to the × cn (k1 , . . . , kn ) , (B2)
form
2
(1−δ̄0 S×/S0 )2/2 where RPi = P (ki ), Wi = W (ki R0 ) and S0 =
e−ν
sf (s|δ0 , c̃ = 0) = √ (2π)−3 d3 kP (k)W (kR0 )2 .
2Γ 2π
Z ∞ We demonstrate this in section B1 by working in
dy y 2
× √ e−(y−Γν+δ̄0 ν1 ) /2 , (A8) Fourier space and explicitly evaluating the integral in the
0 2π
second line of equation (B2). In section B2 we work in
where δ̄0 = δ0 /δc and ν1 = Γν(S× /S0 )(1 − ǫ× ). The Tay- real space, repeating the calculation in field theoretic lan-
lor expansion of this expression in powers of δ̄0 can now guage and showing that the bias coefficients can be inter-
be used to read off the bias coefficients bn using equa- preted as connected expectation values. This real-space
tion (35). The Gaussian multiplying the integral can be calculation also shows how one might generalise our re-
expanded using the definition of the Hermite polynomials sults to the case when the distribution of the matter field
Hn (ν). The following relations are useful in simplifying is not Gaussian.
the integral:
1 2 2

e− 2 Γ ν
Z
dz z pG (z − Γν; 1) = √ (1 + A) , B1 Fourier space calculation
0 2π
1 2 2 To prove equation (B2), note that in the definition (B1),

e− 2 Γ ν A
Z
dz pG (z − Γν; 1) = √ , the functional derivatives can be transferred to the prob-
0 2π Γν ability density functional (which we denote as P[δk ]),
n 1 2 2
e− 2 Γ ν


[z p G (z − Γν; 1)] = √ nHn−1 (Γν) ,

δ n ρh (k)

∂z n
z=0 2π
1 2 2 δδk1 . . . δδkn
∂n e− 2 Γ ν

δ n ρh (k)
Z
p G (z − Γν; 1) = √ Hn (Γν) , (A9)
∂z n 2π = D[δk ]P[δk ]
z=0 δδk1 . . . δδkn
δ n P[δk ]
Z
where A was defined in equation (25). Some manipulation
= (−1)n D[δk ] ρh (k) , (B3)
then leads to the result quoted in equation (37). δδk1 . . . δδkn
R
where D[δk ] denotes a functional integral. Also, statis-
tical homogeneity allows us to introduce 1 = ei(k+k1...n )·x
where k1...n = k1 + . . . + kn and hence write the second
APPENDIX B: RELATION BETWEEN
line of (B2) as
MATSUBARA’S RENORMALISED
COEFFICIENTS AND WEIGHTED Z Z
d3 k ik·x
Z
d 3 k1 d3 kn ik1...n ·x
AVERAGES OF THE MATTER DENSITY D[δk ] 3
e ρ h (k) 3
... e
(2π) (2π) (2π)3
Matsubara (2011) has argued that k-dependent bias W1 . . . Wn δ n P[δk ]
× n
(−1)n (2π)3n P1 . . . Pn . (B4)
factors are generically associated with nonlocal bias- S0 δδk1 . . . δδkn
ing schemes and has provided a number of generic re-
For  Gaussian initial conditions, P[δk ] ∝
sults for such nonlocal bias. In this appendix we show
exp −(1/2) d3 k δk δk∗ /((2π)3 P (k)) . The functional
R 
the connection between the “renormalised” coefficients
derivative of P[δk ] can then be understood as follows
cn (k1 , . . . , kn ) defined by him in terms of functional
(see also Matsubara 1995). Consider the action of a
derivatives of the Fourier-space halo field δh (k) with re-
single functional derivative δ/δδki . When this acts
spect to the matter field δk ,
on the distribution P[δk ], it brings down a factor
d3 k δ n δh (k) (−1)δk∗i (2π)−3 P (ki )−1 . On the other hand, when it
Z  
cn (k1 , . . . , kn ) = (2π)3n ,
(2π)3 δδk1 . . . δδkn acts on an existing factor of δk∗j , it gives a Dirac
(B1) delta δD (ki + kj ) (since δk∗j = δ−kj ). The result of n
and the real-space weighted averages of the matter den- derivatives on P[δk ] can be organised as an alternating
sity field which we discuss in the main text. In particular, sum over terms containing an increasing number of
for Gaussian initial conditions, we show that the Hermite- Dirac deltas or connections between pairs of vectors
weighted bias coefficients bn of equation (32) are just the ki , kj . The alternation arises because each connected
integrals of the cn , provided one formally uses the quan- pair carries a minus sign. For the n-th derivative, the
tity ρh (k) rather than δh (k) in defining the cn , where term containing p connected pairs (when multiplied by

c 0000 RAS, MNRAS 000, 1–16



Scale dependent bias 15
(−1)n (2π)3n P1 . . . Pn ) looks like vertex with k legs. The correlation function h ρh (x)δ0 (z) i
 can be computed using Wick’s theorem to isolate the two-
(−1)p δk1 . . . δkn−2p (2π)3p point correlation functions connecting δ0 (z) to any of the
δ(yj )’s in the sum, and get
× Pn−2p+2 δD (kn−2p+1 + kn−2p+2 ) . . .
∞ Z
1
 X
× Pn δD (kn−1 + kn ) + perms. , (B5) d3 y1 . . . d3 yk bk (x − y1 , . . . , x − yk )
(k − 1)!
k=1

where “perms.” indicates all permutations of the vectors


kj . Since we integrate over all the kj with a totally sym- × h δ(y1 ) . . . δ(yk−1 ) i h δ(yk )δ0 (z) i .
metric prefactor eik1...n ·x (W1 . . . Wn ), all these permuta-
(B10)
tions lead to identical contributions.
The product of (2π)3n P1 . . . Pn with the nth deriva- Since one also has
tive of P[δk ] therefore equals P[δk ] multiplied by
∞ Z
δρh (x) X 1
[n/2]
d3 y1 . . . d3 yk−1
!
X p n =
(−1) (2p − 1)!! (2π)3p δk1 . . . δkn−2p δδ(y) k=1
(k − 1)!
p=0
n − 2p
p−1
Y × bk (x − y1 , . . . , x − yk−1 , x − y)
× Pn−2j δD (kn−2j−1 + kn−2j ) , (B6)
j=0
× δ(y1 ) . . . δ(yk−1 ) , (B11)
where [n/2] is the floor of n/2 and the combinatorial fac-
tor counts the number of partitions of n distinct objects
into (n − 2p) singletons and p pairs, which is precisely the then one obtains
coefficient of xn−2p in the Hermite polynomial Hn (x). Z  
δρh (x)
On performing the integrals over ki in the term with h ρh (x)δ0 (z) i = d3 y h δ(y)δ0 (z) i . (B12)
p connected pairs, the factors of δki will contribute (n − δδ(y)
2p) powers of δ0 (x) and the Dirac deltas will contribute
Similarly, in order to compute any “connected” correla-
p powers of S0 . Further identifying the inverse Fourier
tion function h ρh (x)δ0 (z1 ) . . . δ0 (zn ) ic one should retain
transform of ρh (k) in the first line of (B4), we can write
only those terms where each of the n external field is
the expression in (B4) as
Z connected to any of the internal fields of Equation (B8).
Since the combinatorial factors generated by the action
D[δk ]P[δk ] ρh (x)
of Wick’s theorem are the same as those obtained from
[n/2]
!  n−2p differentiation, one gets
1 X n δ0
× n/2
(−1)p (2p − 1)!! √
S0 p=0
n − 2p S0 h ρh (x)δ0 (z1 ) . . . δ0 (zn ) ic
1 D √ E Z 
δ n ρh (x)

= n/2
(1 + δh )Hn (δ0 / S0 ) , (B7) = d 3 y1 . . . d 3 yn
S0 δδ(y1 ) . . . δ(yn )
n
which completes the proof. Y
× h δ(yj )δ0 (zj ) i . (B13)
j=1

B2 Real space calculation: bias as connected


Going Rto Fourier space one has h δ(y)δ0 (z) i =
expectation values
(2π)−3 R d3 k eik·(y−z) P (k)W (kR0 ) and δ/δδk =
In real space, the statement that ρh (k) can be expressed (2π)−3 d3 y eik·y (δ/δδ(y)), so that
in terms of the modes δk of the matter field translates to
the generic expansion h ρh (x)δ0 (z1 ) . . . δ0 (zn ) ic
∞ Z Z n h
1
Y i
= d 3 k1 . . . d 3 kn e−ikj ·zj P (kj )W (kj R0 )
X
ρh (x) = d3 y1 . . . d3 yk bk (x − y1 , . . . , x − yk )
k! j=1
k=0
δ n ρh (x)
 
× δ(y1 ) . . . δ(yk ) , (B8) × . (B14)
δδk1 . . . δkn
where the bk are the coefficients of the Taylor expansion
If we write
of ρh in powers of δ,
δ n ρh (x) ei(k1 +···+kn )·x
 
δ k ρh (x)

bk (x − y1 , . . . , x − yk ) ≡

, (B9) ≡ cn (k1 , . . . , kn ) ,
δδk1 . . . δkn (2π)3n

δδ(y1 ) . . . δδ(yk ) δ(yi )=0
(B15)
which are totally symmetric in their arguments. then it is not hard to see that the cn above agrees with
Each term of Equation (B8) can be considered as a Matsubara’s definition (equation B1, with δh → ρh ) upon

c 0000 RAS, MNRAS 000, 1–16



16 M. Musso et al.
requiring statistical homogeneity, and moreover,
h ρh (x)δ0 (z1 ) . . . δ0 (zn ) ic
n
d 3 k1 d3 kn Y h ikj ·(x−zj )
Z i
= 3
... 3
e P (kj )W (kj R0 )
(2π) (2π) j=1

× cn (k1 , . . . , kn ). (B16)

In other words, the second line of equation (B2) corre-


n n
sponds

2 to the quantity h ρh (x)δ0 (x) ic /S0 , where S0 =


δ0 (x) .
The connected n-point expectation value can be re-
cursively obtained using
h ρh δ0 i = h ρh δ0 ic + h ρh i h δ0 i
ρh δ02 = ρh δ02 c + 2 h ρh δ0 ic h δ0 i + h ρh i δ02



ρh δ03 = ρh δ03 c + 3 ρh δ02 c h δ0 i





+ 3 h ρh δ0 ic δ02 + h ρh i δ03 , (B17)




and in general
n−1
!
X n
h ρh δ0n ic h ρh δ0n h ρh δ0m ic δ0n−m ,


= i−
m=0
m
(B18)
to remove the disconnected contributions from the av-
erage. Since δ0 is Gaussian-distributed, one has h δ0r i =
r/2
(r − 1)!!S0 for r even and h δ0r i = 0 for r odd; writing
back h ρh δ0m ic in terms of h ρh δ0m i in the above expression
for m < n, one recovers
h ρh δ0n ic D √ E
n/2
= ρ h H n (δ 0 / S 0 ) . (B19)
S0
This therefore justifies the interpretation of the bias fac-
tors bn as the connected parts of the n-point expectation
values.
Moreover, it is clear that the scale dependence of
h ρh δ0n ic comes from the presence of the n mixed corre-
lation functions h δ(yj )δ0 (zj ) i in Equation (B13), intro-
ducing n occurrences of the filter W (kj R0 ) in Equation
(B14). Therefore one can expect the ratio h ρh δ0n ic /S× n

to be approximately scale invariant.


Similar considerations hold when the distribution
of the matter field δ is non-Gaussian. This would in-
clude both the presence of non-Gaussian initial condi-
tions and non-linear gravitational evolution. In this case,
each external field δ0 (zi ) is connected to ρh (x) by the full
non-Gaussian renormalized propagator, while the coeffi-
cients cn (k1 , . . . , kn ) should be defined in terms of what
in quantum field theory is usually called the 1-PI cor-
relation function (that is, the sum of all the diagrams
that cannot be split in two pieces by cutting one single
line) amputated of the external legs. The bias coefficients
in this case will not, in general, correspond to Hermite-
weighted averages, but must be recursively constructed
using equation (B18) (which still involves only 2-point
measurements).

c 0000 RAS, MNRAS 000, 1–16

You might also like