0% found this document useful (0 votes)
99 views27 pages

Bayesian Astrostatistics: A Backward Look To The Future: Loredo@astro - Cornell.edu

Uploaded by

Vane Piedrahita
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
99 views27 pages

Bayesian Astrostatistics: A Backward Look To The Future: Loredo@astro - Cornell.edu

Uploaded by

Vane Piedrahita
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 27

Bayesian astrostatistics: a backward look to the

future†

Thomas J. Loredo
arXiv:1208.3036v1 [astro-ph.IM] 15 Aug 2012

This perspective chapter briefly surveys: (1) past growth in the use of Bayesian methods
in astrophysics; (2) current misconceptions about both frequentist and Bayesian statistical
inference that hinder wider adoption of Bayesian methods by astronomers; and (3) multi-
level (hierarchical) Bayesian modeling as a major future direction for research in Bayesian
astrostatistics, exemplified in part by presentations at the first ISI invited session on astro-
statistics, commemorated in this volume. It closes with an intentionally provocative rec-
ommendation for astronomical survey data reporting, motivated by the multilevel Bayesian
perspective on modeling cosmic populations: that astronomers cease producing catalogs of
estimated fluxes and other source properties from surveys. Instead, summaries of likelihood
functions (or marginal likelihood functions) for source properties should be reported (not
posterior probability density functions), including nontrivial summaries (not simply upper
limits) for candidate objects that do not pass traditional detection thresholds.

This volume contains presentations from the first invited session on astrostatistics to
be held at an International Statistical Institute (ISI) World Statistics Congress. This
session was a major milestone for astrostatistics as an emerging cross-disciplinary
research area. It was the first such session organized by the ISI Astrostatistics Com-
mittee, whose formation in 2010 marked formal international recognition of the
importance and potential of astrostatistics by one of its information science parent
disciplines. It was also a significant milestone for Bayesian astrostatistics, as this
research area was chosen as a (non-exclusive) focus for the session.
As an early (and elder!) proponent of Bayesian methods in astronomy, I have
been asked to provide a “perspective piece” on the history and status of Bayesian
astrostatistics. I begin by briefly documenting the rapid rise in use of the Bayesian

Thomas J. Loredo
Center for Radiophysics & Space Research, Cornell University, Ithaca, NY 14853-6801, e-mail:
loredo@astro.cornell.edu
† This paper is a lightly revised version of an invited chapter for Astrostatistical Challenges
for the New Astronomy (Joseph M. Hilbe, ed., Springer, New York, 2012), the inaugural volume
for the Springer Series in Astrostatistics. The volume commemorates the first invited session on
astrostatistics held at an International Statistical Institute (ISI) World Statistics Congress, which
took place at the Congress held in Dublin, Ireland in August 2011.

1
2 Thomas J. Loredo

approach by astrostatistics researchers over the past two decades. Next, I describe
three misconceptions about both frequentist and Bayesian methods that hinder
wider adoption of the Bayesian approach across the broader community of as-
tronomer data analysts. Then I highlight the emerging role of multilevel (hierarchi-
cal) Bayesian modeling in astrostatistics as a major future direction for research in
Bayesian astrostatistics. I end with a provocative recommendation for survey data
reporting, motivated by the multilevel Bayesian perspective on modeling cosmic
populations.

1 Looking back

Bayesian ideas entered modern astronomical data analysis in the late 1970s, when
Steve Gull & Geoff Daniell (1978, 1979) framed astronomical image deconvolution
in Bayesian terms.1 Motivated by Harold Jeffreys’ Bayesian Theory of Probabil-
ity (Jeffreys 1961), and Edwin Jaynes’s introduction of Bayesian and information
theory methods into statistical mechanics and experimental physics (Jaynes 1959),
they addressed image estimation by writing down Bayes’s theorem for the poste-
rior probability for candidate images, adopting an entropy-based prior distribution
for images. They focused on finding a single “best” image estimate based on the
posterior: the maximum entropy (MaxEnt) image (maximizing the entropic prior
probability subject to a somewhat ad hoc likelihood-based constraint expressing
goodness of fit to the data). Such estimates could also be found using frequentist
penalized likelihood or regularization approaches. The Bayesian underpinnings of
MaxEnt image deconvolution thus seemed more of a curiosity than the mark of a
major methodological shift.
By about 1990, genuinely Bayesian data analysis—in the sense of reporting
Bayesian probabilities for statistical hypotheses, or samples from Bayesian poste-
rior distributions—began appearing in astronomy. The Cambridge MaxEnt group of
astronomers and physicists, led by Gull and John Skilling, began developing “quan-
tified MaxEnt” methods to quantify uncertainty in image deconvolution (and other
inverse problems), rather than merely reporting a single best-fit image. On the statis-
tics side, Brian Ripley’s group began using Gibbs sampling to sample from posterior
distributions for astronomical images based on Markov random field priors (Ripley
1992). My PhD thesis (defended in 1990) introduced parametric Bayesian model-
ing of Poisson counting and point processes (including processes with truncation or
thinning, and measurement error) to high-energy astrophysics (X-ray and gamma-
ray astronomy) and to particle astrophysics (neutrino astronomy). Bayesian methods
were just beginning to be used for parametric modeling of ground- and space-based
cosmic microwave background (CMB) data (e.g., Readhead & Lawrence 1992).

1 Notably, Peter Sturrock (1973) earlier introduced astronomers to the use of Bayesian probabilities

for “bookkeeping” of subjective beliefs about astrophysical hypotheses, but he did not discuss
statistical modeling of measurements per se.
Bayesian astrostatistics 3

It was in this context that the first session on Bayesian methods to be held at an as-
tronomy conference (to my knowledge) took place, at the first Statistical Challenges
in Modern Astronomy conference (SCMA I), hosted by statistician G. Jogesh Babu
and astronomer Eric Feigelson at Pennsylvania State University in August 1991.
Bayesian methods were not only new, but also controversial in astronomy at that
time. Of the 22 papers published in the SCMA I proceedings volume (Feigelson &
Babu 1992), only two were devoted to Bayesian methods (Ripley 1992 and Loredo
1992a; see also the unabridged version of the latter, Loredo 1992b).2 Both papers
had a strong pedagogical component (and a bit of polemic). Of the 131 SCMA I par-
ticipants (about 60% astronomers and 40% statisticians), only two were astronomers
whose research prominently featured Bayesian methods (Gull and me).
Twenty years later, the role of Bayesian methods in astrostatistics research is
dramatically different. The 2008 Joint Statistical Meetings included two sessions
on astrostatistics predominantly devoted to Bayesian research. At SCMA V, held in
June 2011, two sessions were devoted entirely to Bayesian methods in astronomy:
“Bayesian analysis across astronomy,” with eight papers and two commentaries,
and “Bayesian cosmology,” including three papers with individual commentaries.
Overall, 14 of 32 invited SCMA V presentations (not counting commentaries) fea-
tured Bayesian methods, and the focus was on calculations and results rather than
on pedagogy and polemic. About two months later, the ISI World Congress session
on astrostatistics commemorated in this volume was held; as already noted, its focus
was Bayesian astrostatistics.
On the face of it, these events seem to indicate that Bayesian methods are not
only no longer controversial, but are in fact now widely used, even favored for some
applications (most notably for parametric modeling in cosmology). But how repre-
sentative are the conference presentations of broader astrostatistical practice?
Fig. 1 shows my amateur attempt at bibliometric measurement of the growing
adoption of Bayesian methods in both astronomy and physics, based on queries of
publication data in the NASA Astrophysics Data System (ADS). Publication counts
indicate significant and rapidly growing use of Bayesian methods in both astron-
omy and physics.3 Cursory examination of the publications reveals that Bayesian
methods are being developed across a wide range of astronomical subdisciplines.
It is tempting to conclude from the conference and bibliometric indicators that
Bayesian methods are now well-established and well-understood across astronomy.
But the conference metrics reflect the role of Bayesian methods in the astrostatis-
tics research community, not in bread-and-butter astronomical data analysis. And as
impressive as the trends in the bibliometric metrics may be, the absolute numbers
remain small in comparison to all astronomy and physics publications, even limit-

2 A third paper (Nousek 1992) had some Bayesian content but focused on frequentist evaluation

criteria, even for the one Bayesian procedure considered; these three presentations, with discussion,
comprised the Bayesian session.
3 Roberto Trotta and Martin Hendry have shown similar plots in various venues, helpfully noting

that the recent rate of growth apparent in Fig. 1 is much greater than the rate of growth in the
number of all publications; i.e., not just the amount but also the prevalence of Bayesian work is
rapidly rising.
4 Thomas J. Loredo

300 1250

250 Pioneer Propagation Popularity


1000

200

AST+PHY Abstracts
AST Abstracts
750

150

500
100

250
50

0 0
1985 1990 1995 2000 2005 2010
Year

Fig. 1 Simple bibliometrics measuring the growing use of Bayesian methods in astronomy and
physics, based on queries of the NASA ADS database in October 2011. Thick (blue) curves
(against the left axis) are from queries of the astronomy database; thin (red) curves (against the
right axis) are from joint queries of the astronomy and physics databases. For each case the dashed
lower curve indicates the number of papers each year that include “Bayes” or “Bayesian” in the
title or abstract. The upper curve is based on the same query, but also counting papers that use
characteristically Bayesian terminology in the abstract (e.g., the phrase “posterior distribution” or
the acronym “MCMC”); it is meant to capture Bayesian usage in areas where the methods are
well-established, with the “Bayesian” appellation no longer deemed necessary or notable.

ing consideration to data-based studies. Although their impact is growing, Bayesian


methods are not yet in wide use by astronomers.
My interactions with colleagues indicate that significant misconceptions persist
about fundamental aspects of both frequentist and Bayesian statistical inference,
clouding understanding of how these rival approaches to data analysis differ and
relate to one another. I believe these misconceptions play no small role in hinder-
ing broader adoption of Bayesian methods in routine data analysis. In the following
section I highlight a few misconceptions I frequently encounter. I present them here
as a challenge to the Bayesian astrostatistics community; addressing them may ac-
celerate the penetration of sound Bayesian analysis into routine astronomical data
analysis.
Bayesian astrostatistics 5

2 Misconceptions

For brevity, I will focus on just three important misconceptions I repeatedly en-
counter about Bayesian and frequentist methods, posed as (incorrect!) “conceptual
equations.” They are:
• Variability = Uncertainty: This is a misconception about frequentist statistics
that leads analysts to think that good frequentist statistics is easier to do than it
really is.
• Bayesian computation = Hard: This is a misconception about Bayesian meth-
ods that leads analysts to think that implementing them is harder than it really
is, in particular, that it is harder than implementing a frequentist analysis with
comparable capability.
• Bayesian = Frequentist + Priors: This is a misconception about the role of
prior probabilities in Bayesian methods that distracts analysts from more essen-
tial distinguishing features of Bayesian inference.
I will elaborate on these faulty equations in turn.
Variability = Uncertainty: Frequentist statistics gets its name from its reliance on
the long-term frequency conception of probability: frequentist probabilities describe
the long-run variability of outcomes in repeated experimentation. Astronomers who
work with data learn early in their careers how to quantify the variability of data
analysis procedures in quite complicated settings using straightforward Monte Carlo
methods that simulate data. How to get from quantification of variability in the out-
come of a procedure applied to an ensemble of simulated data, to a meaningful
and useful statement about the uncertainty in the conclusions found by applying
the procedure to the one actually observed data set, is a subtle problem that has
occupied the minds of statisticians for over a century. My impression is that many
astronomers fail to recognize the distinction between variability and uncertainty, and
thus fail to appreciate the achievements of frequentist statistics and their relevance
to data analysis practice in astronomy. The result can be reliance on overly simplis-
tic “home-brew” analyses that at best may be suboptimal, but that sometimes can be
downright misleading. A further consequence is a failure to recognize fundamental
differences between frequentist and Bayesian approaches to quantification of un-
certainty (e.g., that Bayesian probabilities for hypotheses are not statements about
variability of results in repeated experiments).
To illustrate the issue, consider estimation of the parameters of a model being fit
to astronomical data, say, the parameters of a spectral model. It is often straightfor-
ward to find best-fit parameters with an optimization algorithm, e.g., minimizing a
χ 2 measure of misfit, or maximizing a likelihood function. A harder but arguably
more important task is quantifying uncertainty in the parameters. For abundant data,
an asymptotic Gaussian approximation may be valid, justifying use of the Hessian
matrix returned by many optimization codes to calculate an approximate covariance
matrix for defining confidence regions. But when uncertainties are significant and
models are nonlinear, we must use more complicated procedures to find accurate
confidence regions.
6 Thomas J. Loredo

Bootstrap resampling is a powerful framework statisticians use to develop meth-


ods to accomplish this. There is a vast literature on applying the bootstrap idea in
various settings; much of it is devoted to the nontrivial problem of devising algo-
rithms that enable useful and accurate uncertainty statements to be derived from sim-
ple bootstrap variability calculations. Unfortunately, this literature is little-known in
the astronomical community, and too often astronomers misuse bootstrap ideas. The
variability-equals-uncertainty misconception appears to be at the root of the prob-
lem.
As a cartoon example, suppose we have spectral data from a source that we
wish to fit with a simple thermal spectral model with two parameters, an amplitude,
A (e.g., proportional to the source area and inversely proportional to its distance
squared), and a temperature, T (determining the shape of the spectrum as a function
of energy or wavelength); we denote the parameters jointly by P = (A, T ). Fig. 2
depicts the two-dimensional (A, T ) parameter space, with the best-fit parameters,
P̂(Dobs ), indicated by the blue four-pointed star. We can use simulated data to find
the variability of the estimator (i.e., of the function P̂(D) defined by the optimizer)
were we to repeatedly observe the source. But how should we simulate data when
we do not know the true nature of the signal (and perhaps of noise and instrumental
distortions)? And how should the variability of simulation results be used to quantify
the uncertainty in inferences based on the one observed dataset?
The underlying idea of the bootstrap is to use the observed dataset to define the
ensemble of hypothetical data to use in variability calculations, and to find functions
of the data (statistics) whose variability can be simply used to quantify uncertainty
(e.g., via moments or a histogram). Normally using the observed data to define the
ensemble for simulations would be cheating and would invalidate one’s inferences;
all frequentist probabilities formally must be “pre-observation” calculations. A ma-
jor achievement of the bootstrap literature is showing how to use the observed data
in a way that gives approximately
√ valid results (hopefully with a rate of conver-
gence better than the O(1/ N) rate achieved by simple Gaussian approximations,
for sample size N).
One way to proceed is to use the full model we are fitting (both the signal model
and the instrument and noise model) to simulate data, with the parameters fixed
at P̂(Dobs ) as a surrogate for the true values. Statisticians call this the parametric
bootstrap; it was popularized to astronomers in a well-known paper by Lampton,
Margon & Bowyer (1976, hereafter LMB76; statisticians introduced the “paramet-
ric bootstrap” terminology later). Alternatively, if some probabilistic aspects of the
model are not trusted (e.g., the error distribution is considered unrealistic), an al-
ternative approach is the nonparametric bootstrap, which “recycles” the observed
data to generate simulated data (in some simple cases, this may be done by sam-
pling from the observed data with replacement to generate each simulated data set).
Whichever approach we adopt, we will generate a set of simulated data, {Di }, to
which we can apply our fitting procedure to generate a set of best-fit parameter
points {P̂(Di )} that together quantify the variability of our estimator. The black
dots in Fig. 2 show a scatterplot or “point cloud” of such parameter estimates.
Bayesian astrostatistics 7

A
P̂(Dobs )

T
Fig. 2 Illustration of the nontrivial relationship between variability of an estimator, and uncertainty
of an estimate as quantified by a frequentist confidence region. Shown is a two-dimensional pa-
rameter space with a best-fit estimate to the observed data (blue 4-pointed star), best-fit estimates
to boostrapped data (black dots) showing variability of the estimator, and a contour bounding a
parametric bootstrap confidence region quantifying uncertainty in the estimate.

What does the point cloud tell us about the uncertainty we should associate with
the estimate from the observed data (the blue star)? The practice I have seen in too
many papers is to interpret the points as samples from a probability distribution
in parameter space.4 The cloud itself might be shown, with a contour enclosing a
specified fraction of the points offered as a joint confidence region. One-dimensional
histograms may be used to find “1σ ” (i.e., 68.3%) confidence regions for particular
parameters of interest. Such procedures naively equate variability with uncertainty:
the uncertainty of the estimate is identified with the variability of the estimator.
Regions created this way are wrong, plainly and simply. They will not cover the true
parameters with the claimed probability, and they are skewed in the wrong direction.
This is apparent from the figure; the black points are skewed down and to the right
of the star indicating the parameter values used to produce the simulated data; the
parameters that produced the observed data are thus likely to be up and to the left of
the star.
In a correct parametric bootstrapping calculation (i.e., with a trusted model), one
can use simulated data to calibrate χ 2 or likelihood contours for defining confidence
regions. The procedure is not very complicated, and produces regions like the one
shown by the contour in Fig. 2, skewed in just the right way; LMB76 described the
construction. But frequently investigators are attempting a nonparametric bootstrap,
corresponding to a trusted signal model but an untrusted noise model (often this is
done without explicit justification, as if nonparametric bootstrapping is the only type

4 I am not providing references to publications exhibiting the problem for diplomatic reasons and

for a more pragmatic and frustrating reason: In the field where I have repeatedly encountered
the problem—analysis of infrared exoplanet transit data—authors routinely fail to describe their
analysis methods with sufficient detail to know what was done, let alone to enable readers to verify
or duplicate the analysis. While there are clear signs of statistical impropriety in many of the papers,
I only know the details from personal communications with exoplanet transit scientists.
8 Thomas J. Loredo

of bootstrapping). In this case devising a sound bootstrap confidence interval algo-


rithm is not so simple. Indeed, there is a large statistics literature on nonparametric
bootstrap confidence intervals, its size speaking to nontrivial challenges in making
the bootstrap idea work. In particular, no simple procedure is currently known for
finding accurate joint confidence regions for multiple parameters in nonlinear mod-
els using the nonparametric bootstrap; yet results purporting to come from such
a procedure appear in a number of astronomical publications. Strangely, the most
cited reference on bootstrap confidence regions in the astronomical literature is a
book on numerical methods, authored by astronomers, with an extremely brief and
misleadingly simplistic discussion of the nonparametric bootstrap (and no explicit
discussion of parametric bootstrapping). Sadly, this is not the only area where our
community appears content disregarding a large body of relevant statistics research
(frequentist or otherwise).
What explains such disregard of relevant and nontrivial expertise? I am sure there
are multiple factors, but I suspect an important one is the misconception that vari-
ability may be generically identified with uncertainty. If one believes this, then doing
(frequentist) statistics appears to be a simple matter of simulating data to quantify
the variability of procedures. For linear models with normally-distributed errors,
the identification is practically harmless; it is conceptually flawed but leads to cor-
rect results by accident. But more generally, determining how to use variability to
quantify uncertainty can be subtle and challenging. Statisticians have worked for a
century to establish how to map variability to uncertainty; when we seek frequentist
quantifications of uncertainty in nontrivial settings, we need to mine their expertise.
Why devote so much space to a misconception about frequentist statistics in a
commentary on Bayesian methods? The variability-equals-uncertainty misconcep-
tion leads data analysts to think that frequentist statistics is easier than it really is, in
fact, to think that they already know what they need to know about it. If astronomers
realize that sound statistical practice is nontrivial and requires study, they may be
more likely to study Bayesian methods, and more likely to come to understand the
differences between the frequentist and Bayesian approaches. Also, for the example
just described, the way some astronomers misuse the bootstrap is to try to use it
to define a probability distribution over parameter space. Frequentist statistics de-
nies such a concept is meaningful, but it is exactly what Bayesian methods aim to
provide, e.g., with point clouds produced via Markov chain Monte Carlo (MCMC)
posterior sampling algorithms. This brings us to the topic of Bayesian computation.
Bayesian computation = Hard: A too-common (and largely unjustified) com-
plaint about Bayesian methods is that their computational implementation is dif-
ficult, or more to the point, that Bayesian computation is harder than frequentist
computation. Analysts wanting a quick and useable result are thus dissuaded from
considering a Bayesian approach. It is certainly true that posterior sampling via
MCMC—generating pseudo-random parameter values distributed according to the
posterior—is harder to do well than is generating pseudo-random data sets from a
best-fit model. (In fact, I would argue that our community may not appreciate how
hard it can be to do MCMC well.) But this is an apples-to-oranges comparison be-
tween methods making very different types of approximations. Fair comparisons,
Bayesian astrostatistics 9

between Bayesian and frequentist methods with comparable capabilities and ap-
proximations, tell a different story. Advocates of Bayesian methods need to tell this
story, and give this complaint a proper and public burial.
Consider first basic “textbook” problems that are analytically accessible. Exam-
ples include estimating the mean of a normal distribution, estimating the coefficients
of a linear model (for data with additive Gaussian noise), similar estimation when
the noise variance is unknown (leading to Student’s t distribution), estimating the
intensity of a Poisson counting process, etc.. In such problems, the Bayesian cal-
culation is typically easier than its frequentist counterpart, sometimes significantly
so. Nowhere is this more dramatically demonstrated than in Jeffreys’s classic text,
Theory of Probability (1961). In chapter after chapter, Jeffreys solves well-known
statistics problems with arguments significantly more straightforward and mathe-
matics significantly more accessible than are used in the corresponding frequentist
treatments. The analytical tractability of such foundational problems is an aid to
developing sound statistical intuition, so this is not a trivial virtue of the Bayesian
approach. (Of course, ease of use is no virtue at all if you reject an approach—
Bayesian, frequentist, or otherwise—on philosophical grounds.)
Turn now to problems of more realistic complexity, where approximate numer-
ical methods are necessary. The main computational challenge for both frequen-
tist and Bayesian inference is multidimensional integration, over the sample space
for frequentist methods, and over parameter space for Bayesian methods. In high
dimensions, both approaches tend to rely on Monte Carlo methods for their com-
putational implementation. The most celebrated approach to Bayesian computation
is MCMC, which builds a multivariate pseudo-random number generator producing
dependent samples from a posterior distribution. Frequentist calculation instead uses
Monte Carlo methods to simulate data (i.e., draw samples from the sampling distri-
bution). Building an MCMC posterior sampler, and properly analyzing its output, is
certainly more challenging than simulating data from a model, largely because data
are usually independently distributed; nonparametric bootstrap resampling of data
may sometimes be simpler still. But the comparison is not fair. In typical frequentist
Monte Carlo calculations, one simulates data from the best-fit model (or a boot-
strap surrogate), not from the true model (or other plausible models). The resulting
frequentist quantities are approximate, not just because of Monte Carlo error, but
because the integrand (or family of integrands) that would appear in an exact fre-
quentist calculation is being approximated (asymptotically). In some cases, an exact
finite-sample frequentist formulation of the problem may not even be known. In con-
trast, MCMC posterior sampling makes no approximation of integrands; results are
approximate only due to Monte Carlo sampling error. No large-sample asymptotic
approximations need be invoked.
This point deserves amplification. Although the main computational challenge
for frequentist statistics is integration over sample space, there are additionally seri-
ous theoretical challenges for finite-sample inference in many realistically complex
settings. Such challenges do not typically arise for Bayesian inference. These the-
oretical challenges, and the analytical approximations that are adopted to address
them, are ignored in comparisons that pit simple Monte Carlo simulation of data, or
10 Thomas J. Loredo

bootstrapping, against MCMC or other nontrivial Bayesian computation techniques.


Arguably, accurate finite-sample parametric inference is often computationally sim-
pler for Bayesian methods, because an accurate frequentist approach is simply im-
possible, and an approximate calculation can quantify only the rate of convergence
of the approximation, not the actual accuracy of the specific calculation being per-
formed.
If one is content with asymptotic approximations, the fairer comparison is be-
tween asymptotic frequentist and asymptotic Bayesian methods. At the lowest or-
der, asymptotic Bayesian computation using the Laplace approximation is not sig-
nificantly harder than asymptotic frequentist calculation. It uses the same basic nu-
merical quantities—point estimates and Hessian matrices from an optimizer—but
in different ways. It provides users with nontrivial new capabilities, such as the
ability to marginalize over nuisance parameters, or to compare models using Bayes
factors that include an “Ockham’s razor” penalty not present in frequentist signif-
icance tests, and that enable straightforward comparison of rival models that need
not be nested (with one being a special case
√ of the other). And the results are some-
times accurate to one order higher (in 1/ N) than corresponding frequentist results
(making them potentially competitive with some bootstrap methods seeking similar
asymptotic approximation rates).
Some elaboration of these issues is in Loredo (1999), including references to
literature on Bayesian computation up to that time. A more recent discussion of
this misconception about Bayesian methods (and other misconceptions) is in an
insightful essay by the Bayesian economist Christopher Sims, winner of the 2011
Nobel Prize in economics (Sims 2010).
Bayesian = Frequentist + Priors: I recently helped organize and present two days
of tutorial lectures on Bayesian computation, as a prelude to the SCMA V con-
ference mentioned above. As I was photocopying lecture notes for the tutorials, a
colleague walked into the copy room and had a look at the table of contents for the
tutorials. “Why aren’t all of the talks about priors?” he asked. In response to my
puzzled look, he continued, “Isn’t that what Bayesian statistics is about, accounting
for prior probabilities?”
Bayesian statistics gets its name from Bayes’s theorem, establishing that the pos-
terior probability for a hypothesis is proportional to the product of its prior proba-
bility, and the probability for the data given the hypothesis (i.e., the sampling dis-
tribution for the data). The latter factor is the likelihood function when considered
as a function of the hypotheses being considered (with the data fixed to their ob-
served values). Frequentist methods that directly use the likelihood function, and
their least-squares cousins (e.g., χ 2 minimization), are intuitively appealing to as-
tronomers and widely used. On the face of it, Bayes’s theorem appears merely to
add modulation by a prior to likelihood methods. By name, Bayesian statistics is
evidently about using Bayes’s theorem, so it would seem it must be about how fre-
quentist results should be altered to account for prior probabilities.
It would be hard to overstate how wrong this conception of Bayesian statistics is.
The name is unfortunate; Bayesian statistics uses all of probability theory, not
just Bayes’s theorem, and not even primarily Bayes’s theorem. What most funda-
Bayesian astrostatistics 11

mentally distinguishes Bayesian calculations from frequentist calculations is not


modulation by priors, but the key role of probability distributions over parame-
ter (hypothesis) space in the former, and the complete absence of such distribu-
tions in the latter. Via Bayes’s theorem, a prior enables one to use the likelihood
function—describing a family of measures over the sample space—to build the
posterior distribution—a measure over the parameter (hypothesis) space (where by
“measure” I mean an additive function over sets in the specified space). This con-
struction is just one step in inference. Once it happens, the rest of probability theory
kicks in, enabling one to assess scientific arguments directly by calculating probabil-
ities quantifying the strengths of those arguments, rather than indirectly, by having
to devise a way that variability of a cleverly chosen statistic across hypothetical data
might quantify uncertainty across possible choices of parameters or models for the
actually observed data.
Perhaps the most important theorem for doing Bayesian calculations is the law of
total probability (LTP) that relates marginal probabilities to sums of joint and con-
ditional probabilities. To display its role, suppose we are analyzing some observed
data, Dobs , using a parametric model with parameters θ . Let M denote all the mod-
eling assumptions—definitions of the sample and parameter spaces, description of
the connection between the model and the data, and summaries of any relevant prior
information (parameter constraints or results of other measurements). Now consider
some of the common uses of LTP in Bayesian analysis:
• Calculating the probability in a credible region, R, for θ :
Z
p(θ ∈ R|Dobs ) = d θ p(θ ∈ R|θ ) p(θ |Dobs ) || M, (1)
R

where p(θ |Dobs , M) is the posterior probability density for θ . Here I have in-
troduced a convenient shorthand due to John Skilling: “||M” indicates that M is
conditioning information common to all displayed probabilities.
• Calculating a marginal posterior distribution when a vector parameter θ =
(ψ , η ) has both an interesting subset of parameters, ψ , and uninteresting “nui-
sance” parameters, η . The uncertainty in ψ (with the η uncertainty fully propa-
gated) is quantified by the marginal density,
Z
p(ψ |Dobs ) = d η p(ψ , η |Dobs ) || M. (2)

• Predicting future data, D′ , with the posterior predictive distribution,


Z
p(D′ |Dobs ) = d θ p(D′ |θ ) p(θ |Dobs ) || M, (3)

with the integration accounting for parameter uncertainty in the prediction.


• Comparing rival parametric models Mi (each with parameters θi ) via posterior
odds or Bayes factors, which requires computation of the marginal likelihood for
each model given by
12 Thomas J. Loredo
Z
p(Dobs |Mi ) = d θi p(θi |Mi ) p(Dobs |θi , Mi ) || M1 ∨ M2 . . . . (4)

In words, this says that the likelihood for a model is the average of the likelihood
function for that model’s parameters.
Arguably, if this approach to inference is to be named for a theorem, “total probabil-
ity inference” would be a more appropriate appellation than “Bayesian statistics.” It
is probably too late to change the name. But it is not too late to change the emphasis.
In axiomatic developments of Bayesian inference, priors play no fundamental
role; rather, they emerge as a required ingredient when one seeks a consistent or co-
herent calculus for the strengths of arguments that reason from data to hypotheses.
Sometimes priors are eminently useful, as when one wants to account for a posi-
tivity constraint on a physical parameter, or to combine information from different
experiments or observations. Other times they are frankly a nuisance, but alas still a
necessity.
A physical analogy I find helpful for elucidating the role of priors in Bayesian
inference appeals to the distinction between intensive and extensive quantities in
thermodynamics. Temperature is an intensive property; in a volume of space it is
meaningful to talk about the temperature T (x) at a point x, but not about the “total
temperature” of the volume; temperature does not add or integrate across space. In
contrast, heat is an extensive property, an additive property of volumes; in mathe-
matical parlance, it may be described by a measure (a mapping from regions, rather
than points, to a real number).
R Temperature and heat are related; the heat in a vol-
ume V is given by Q = V dx [ρ (x)c(x)]T (x), where ρ (x) is the density and c(x) is
the specific heat capacity. The product ρ c is extensive, and serves to convert the
intensive temperature to its extensive relative, heat. In Bayesian inference, the prior
plays an analogous role, not just “modulating” likelihood, but converting intensive
likelihood to extensive probability. In thermodynamics, a region with a high temper-
ature may have a small amount of heat if its volume is small, or if, despite having a
large volume, the value of ρ c is small. In Bayesian statistics, a region of parameter
space with high likelihood may have a small probability if its volume is small, or if
the prior assigns low probability to the region.
This accounting for volume in parameter space is a key feature of Bayesian
methods. What makes it possible is having a measure over parameter space. Pri-
ors are important, not so much as modulators of likelihoods, but as converters from
intensities (likelihoods) to measures (probabilities). With poetic license, one might
say that frequentist statistics focuses on the “hottest” (highest likelihood) hypothe-
ses, while Bayesian inference focuses on hypotheses with the most “heat” (proba-
bility).
Incommensurability: The growth in the use of Bayesian methods in recent decades
has sometimes been described as a “revolution,” presumably alluding to Thomas
Kuhn’s concept of scientific revolutions (Kuhn 1970). Although adoption of Bayesian
methods in many disciplines has been growing steadily and sometimes dramatically,
Bayesian methods have yet to completely or even substantially replace frequentist
methods in any broad discipline I am aware of (although this has happened in some
Bayesian astrostatistics 13

subdisciplines). I doubt the pace and extent of change qualifies for a Kuhnian revolu-
tion. Also, the Bayesian and frequentist approaches are not rival scientific theories,
but rather rival paradigms for a part of the scientific method itself (how to build and
assess arguments from data to scientific hypotheses). Nevertheless, the competition
between Bayesian and frequentist approaches to inference does bear one hallmark
of a Kuhnian revolution: incommensurability. I believe Bayesian-frequentist incom-
mensurability is not well-appreciated, and that it underlies multiple misconceptions
about the approaches.
Kuhn insisted that there could be no neutral or objective measure allowing com-
parison of competing paradigms in a budding scientific revolution. He based this
claim on several features he found common to scientific revolutions, including
the following: (1) Competing paradigms often adopt different meanings for the
same term or statement, making it very difficult to effectively communicate across
paradigms (a standard illustration is the term “mass,” which takes on different mean-
ings in Newtonian and relativistic physics). (2) Competing paradigms adopt differ-
ent standards of evaluation; each paradigm typically “works” when judged by its
own standards, but the standards themselves are of limited use in comparing across
paradigms.
These aspects of Kuhnian incommensurability are evident in the frequentist and
Bayesian approaches to statistical inference. (1) The term “probability” takes dif-
ferent meanings in frequentist and Bayesian approaches to uncertainty quantifica-
tion, inviting misunderstanding when comparing frequentist and Bayesian answers
to a particular inference problem. (2) Long-run performance is the gold standard
for frequentist statistics; frequentist methods aim for specified performance across
repeated experiments by construction, but make no probabilistic claims about the
result of application of a procedure to a particular observed dataset. Bayesian meth-
ods adopt more abstract standards, such as coherence or internal consistency, that
apply to inference for the case-at-hand, with no fundamental role for frequency-
based long-run performance. A frequentist method with good long-run performance
can violate Bayesian coherence or consistency requirements so strongly as to be ob-
viously unacceptable for inference in particular cases.5 On the other hand, Bayesian
algorithms do not have guaranteed frequentist performance; if it is of interest, it
must be separately evaluated, and priors may need adjustment to improve frequen-
tist performance.6
Kuhn considered rival paradigms to be so “incommensurable” that “proponents
of competing paradigms practice their trades in different worlds.” He argued that
5 Efron (2003) describes some such cases by saying the frequentist result can be accurate but not

correct. Put another way, the performance claim is valid, but the long-run performance can be
irrelevant to the case-at-hand, e.g., due to the existance of so-called recognizable subsets in the
sample space (see Loredo (1992) and Efron (2003) for elaboration of this notion). This is a further
example of how nontrivial the relationship between variability and uncertainty can be.
6 There are theorems linking single-case Bayesian probabilities and long-run performance in some

general settings, e.g., establishing that, for fixed-dimension parametric inference, Bayesian cred-
ible regions
√ with probability P have frequentist coverage close to P (the rate of convergence is
O(1/ N) for flat priors, and faster for so-called reference priors). But the theorems do not apply
in some interesting classes of problems, e.g., nonparametric problems.
14 Thomas J. Loredo

incommensurability is often so stark that for an individual scientist to adopt a new


paradigm requires a psychological shift that could be termed a “conversion experi-
ence.” Following Kuhn, philosophers of science have debated how extreme incom-
mensurability really is between rival paradigms, but the concept is widely consid-
ered important for understanding significant changes in science. In this context, it
is notable that statisticians, and scientists more generally, often adopt a particular
almost-religious terminology in frequentist vs. Bayesian discussions: rather than a
method being described as frequentist or Bayesian, the investigator is so described.
This seems to me to be an unfortunate tradition that should be abandoned. Never-
theless, it does highlight the fundamental incommensurability between these rival
paradigms for uncertainty quantification. Advocates of one approach or the other
(or of a nuanced combination) need to more explicitly note and discuss this incom-
mensurability, especially with nonexperts seeking to choose between approaches.
The fact that both paradigms remain in broad use suggests that ideas from both
approaches may be relevant to inference; perhaps they are each suited to address-
ing different types of scientific questions. For example, my discussion of miscon-
ceptions has been largely from the perspective of parametric modeling (parameter
estimation and model comparison). Nonparametric inference raises more subtle is-
sues regarding both computation and the role of long-term performance in Bayesian
inference; see Bayarri & Berger (2004) and Sims (2010) for insightful discussions
of some of these issues. Also, Bayesian model checking (assessing the adequacy
of a parametric model without an explicitly specified class of alternatives) typically
invokes criteria based on predictive frequencies (Bayarri & Berger 2004; Gelman
et al. 2004; Little 2006). A virtue of the Bayesian approach is that one may predict
or estimate frequencies when they are deemed relevant; explicitly distinguishing
probability (as degree of strength of an argument) from frequency (in finite or hy-
pothetical infinite samples) enables this. This suggests some kind of unification of
approaches may be easier to achieve from the Bayesian direction. This is a worth-
while task for research; see Loredo (2012a) for a brief overview of some recent work
on the Bayesian/frequentist interface.
No one would claim that the Bayesian approach is a data analysis panacea, pro-
viding the best way to address all data analysis questions. But among astronomers
outside of the community of astrostatistics researchers, Bayesian methods are sig-
nificantly underutilized. Clearing up misconceptions should go a long way toward
helping astronomers appreciate what both frequentist and Bayesian methods have
to offer for both routine and research-level data analysis tasks.

3 Looking forward

Having looked at the past growth of interest in Bayesian methods and present mis-
conceptions, I will now turn to the future. As inspiration, I cite Mike West’s com-
mentary on my SCMA I paper (West 1992). In his closing remarks he pointed to an
especially promising direction for future Bayesian work in astrostatistics:
Bayesian astrostatistics 15

On possible future directions, it is clear that Bayesian developments during recent years
have much to offer—I would identify prior modeling developments in hierarchical mod-
els as particularly noteworthy. Applications of such models have grown tremendously in
biomedical and social sciences, but this has yet to be paralleled in the physical sciences.
Investigations involving repeat experimentation on similar, related systems provide the
archetype logical structure for hierarchical modeling. . . There are clear opportunities for
exploitation of these (and other) developments by astronomical investigators. . . .

However clear the opportunities may have appeared to West, for over a decade
following SCMA I, few astronomers pursued hierarchical Bayesian modeling. A
particularly promising application area is modeling of populations of astronomical
sources, where hierarchical models can naturally account for measurement error,
selection effects, and “scatter” of properties across a population. I discussed this at
some length at SCMA IV in 2006 (Loredo 2007), but even as of that time there was
relatively little work in astronomy using hierarchical Bayesian methods, and for the
most part only the simplest such models were used.
The last few years mark a change point in this respect, and evidence of the
change is apparent in the contributions to the summer 2011 Bayesian sessions at
both SCMA V and the ISI World Congress. Several presentations in both forums
described recent and ongoing research developing sophisticated hierarchical mod-
els for complex astronomical data. Other papers raised issues that may be addressed
with hierarchical models. Together, these papers point to hierarchical Bayesian mod-
eling as an important emerging research direction for astrostatistics.
To illustrate the notion of a hierarchical model—also known as a multilevel model
(MLM)—we start with a simple parametric density estimation problem, and then
promote it to a MLM by adding measurement error.
Suppose we would like to estimate parameters θ defining a probability density
function f (x; θ ) for an observable x. A concrete example might be estimation of a
galaxy luminosity function, where x would be two-dimensional, x = (L, z) for lu-
minosity L and redshift z, and f (x; θ ) would be the normalized luminosity function
(i.e., a probability density rather than a galaxy number density). Consider first the
case where we have a set of precise measurements of the observables, {xi } (and no
selection effects). Panel (a) in Fig. 3 depicts this simple setting. The likelihood func-
tion for θ is L (θ ) ≡ p({xi }|θ , M) = ∏i f (xi ; θ ). Bayesian estimation of θ requires
a prior density, π (θ ), leading to a posterior density p(θ |{xi }, M) ∝ π (θ )L (θ ).
An alternative way to write Bayes’s theorem expresses the posterior in terms of
the joint distribution for parameters and data:

p(θ , {xi }|M)


p(θ |{xi }, M) = . (5)
p({xi }|M)

This “probability for everything” version of Bayes’s theorem changes the process of
modeling from separate specification of a prior and likelihood, to specification of the
joint distribution for everything; this proves helpful for building models with com-
plex dependencies. Panel (c) depicts the dependencies in the joint distribution with
a graph—a collection of nodes connected by edges—where each node represents
a probability distribution for the indicated variable, and the directed edges indicate
16 Thomas J. Loredo

Fig. 3 Illustration of multilevel model approach to handling measurement error. (a) and (b) (top
row): Measurements of a two-dimensional observable and its probability distribution (contours); in
(a) the measurements are precise (points); in (b) they are noisy (filled circles depict uncertainties).
(c) and (d): Graphical models corresponding to Bayesian estimation of the density in (a) and (b),
respectively.

dependences between variables. Shaded nodes indicate variables whose values are
known (here, the data); we may manipulate the joint to condition on these quanti-
ties. The graph structure visually displays how the joint distribution may be factored
as a sequence of independent and conditional distributions: the θ node represents
the prior, and the xi nodes represent f (xi ; θ ) = p(xi |θ , M) factors, dependent on θ
but independent of other xi values when θ is given (i.e., conditionally independent).
The joint distribution is thus p(θ , {xi }|M) = π (θ ) ∏i f (xi ; θ ). In a sense, the most
important edges in the graph are the missing edges; they indicate independence that
makes factors simpler than they might otherwise be.
Now suppose that, instead of precise xi measurements, for each observation we
get noisy data, Di , producing a measurement likelihood function ℓi (xi ) ≡ p(Di |xi , M)
describing the uncertainties in xi (we might summarize it with the mean and standard
deviation of a Gaussian). Panel (b) depicts the situation; instead of points in x space,
we now have likelihood functions (depicted as “1σ ” error circles). Panel (d) shows
a graph describing this measurement error problem, which adds a {Di } level to the
previous graph; we now have a multilevel model.7 The xi nodes are now unshaded;
they are no longer known, and have become latent parameters. From the graph we
can read off the form of the joint distribution:

7The convention is to reserve the term for models with three or more levels of nodes, i.e., two or
more levels of edges, or two or more levels of nodes for uncertain variables (i.e., unshaded nodes).
The model depicted in panel (d) would be called a two-level model.
Bayesian astrostatistics 17

p(θ , {xi }, {Di }|M) = π (θ ) ∏ f (xi ; θ )ℓi (xi ). (6)


i

From this joint distribution we can make inferences about any quantity of interest.
To estimate θ , we use the joint to calculate p(θ , {xi }|{Di }, M) (i.e., we condition
on the known data using Bayes’s theorem), and then we marginalize over all of the
latent xi variables. We can estimate all the xi values jointly by instead marginalizing
over θ . Note that this produces a joint marginal distribution for {xi } that is not a
product of independent factors; although the xi values are conditionally indepen-
dent given θ , they are marginally dependent. If we do not know θ , each xi tells us
something about all the others through what it tells us about θ . Statisticians use the
phrase “borrowing strength” to describe this effect, from John Tukey’s evocative de-
scription of “mustering and borrowing strength” from related data in multiple stages
of data analysis (see Loredo and Hendry 2010 for a tutorial discussion of this effect
and the related concept of shrinkage estimators). Note the prominent role of LTP in
inference with MLMs, where inference at one level requires marginalization over
unknowns at other levels.
The few Bayesian MLMs used by astronomers through the 1990s and early
2000s did not go much beyond this simplest hierarchical structure. For example,
unbeknownst to West, at the time of his writing my thesis work had already de-
veloped a MLM for analyzing the arrival times and energies of neutrinos detected
from SN 1987A; the multilevel structure was needed to handle measurement error
in the energies (an expanded version of this work appears in Loredo & Lamb 2002).
Panel (a) of Fig. 4 shows a graph describing the model. The rectangles are “plates”
indicating substructures that are repeated; the integer variable in the corner indi-
cates the number of repeats. There are two plates because neutrino detectors have
a limited (and energy-dependent) detection efficiency. The plate with a known re-
peat count, N, corresponds to the N detected neutrinos with times t and energies
ε ; the plate with an unknown repeat count, N, corresponds to undetected neutrinos,
which must be considered in order to constrain the total signal rate; D denotes the
nondetection data, i.e., reports of zero events in time intervals between detections.
Other problems tackled by astronomers with two-level MLMs include: model-
ing of number-size distributions (“logN–log S” or “number counts”) of gamma-ray
bursts and trans-Neptunian objects (e.g., Loredo & Wasserman 1998; Petit et al.
2008); performing linear regression with measurement error along both axes, e.g.,
for correlating quasar hardness and luminosity (Kelly 2007; see his contribution
in Feigelson & Babu 2012 for an introduction to MLMs for measurement error);
accounting for Eddington and Malmquist biases in cosmology (Loredo & Hendry
2010); statistical assessment of directional coincidences with gamma-ray bursts
(Luo, Loredo & Wasserman 1996; Graziani & Lamb 1996) and cross-matching cata-
logs produced by large star and galaxy surveys (Budavári & Szalay 2008; see Loredo
2012b for discussion of the underlying MLM); and handling multivariate measure-
ment error when estimating stellar velocity distributions from proper motion survey
data (Bovy, Hogg, & Roweis 2011).
Beginning around 2000, interdisciplinary teams of astronomers and information
scientists began developing significantly more sophisticated MLMs for astronom-
18 Thomas J. Loredo

Fig. 4 Graphs describing multilevel models used in astronomy, as described in the text.

ical data. The most advanced such work has come from a collaboration of as-
tronomers and statisticians associated with the Harvard-Smithsonian Center for As-
trophysics (CfA). Much of their work has been motivated by analysis of data from
the Chandra X-ray observatory satellite, whose science operations are managed by
CfA. van Dyk et al. (2001) developed a many-level MLM for fitting Chandra X-ray
spectral data; a host of latent parameters enable accurate accounting for uncertain
backgrounds and instrumental effects such as pulse pile-up. Esch et al. (2004) de-
veloped a Bayesian image reconstruction algorithm for Chandra imaging data that
uses a multiscale hierarchical prior to build spatially-adaptive smoothing into im-
age estimation and uncertainty quantification. van Dyk et al. (2009) showed how
to analyze stellar cluster color-magnitude diagrams (CMDs) using finite mixture
models (FMMs) to account for contamination of the data from stars not lying in
the targeted cluster. In FMMs, categorical class membership variables appear as la-
tent parameters; the mixture model effectively averages over many possible graphs
(corresponding to different partitions of the data into classes; such averaging over
partitions also appears in the Bayesian cross-matching framework of Luo, Loredo
& Wasserman 1996). FMMs have long been used to handle outliers and contam-
ination in Bayesian regression and density estimation. This work showed how to
implement it with computationally expensive models and informative class mem-
bership priors. In the area of time series, the astronomer-engineer collaboration of
Dobigeon, Tourneret, & Scargle (2007) developed a three-level MLM to tackle joint
Bayesian astrostatistics 19

segmentation of astronomical arrival time series (a multivariate extension of Scar-


gle’s well-known Bayesian Blocks algorithm).
Cosmology is a natural arena for multilevel modeling, because of the indirect
link between theory and observables. For example, in modeling both the cosmic
microwave background (CMB) and the large scale structure (LSS) of the galaxy
distribution, theory does not predict a specific temperature map or set of galaxy
locations (these depend on unknowable initial conditions), but instead predicts sta-
tistical quantities, such as angular or spatial power spectra. Modeling observables
given theoretical parameters typically requires introducing these quantities as latent
parameters. In Loredo (1995) I described a highly simplified hierarchical treatment
of CMB data, with noisy CMB temperature difference time series data at the lowest
level, l = 2 spherical harmonic coefficients in the middle, and a single physical pa-
rameter of interest, the cosmological quadrupole moment Q, at the top. While a use-
ful illustration of the MLM approach, I noted there were enormous computational
challenges facing a more realistic implementation. It took a decade for such an im-
plementation to be developed, in the pioneering work of Wandelt et al. (2004). And
only recently have explicit hierarchical models been implemented for LSS modeling
(e.g., Kitaura & Enßlin 2008).
This brings us to the present. Contributions in this volume and in the SCMA V
proceedings (Feigelson & Babu 2012) document burgeoning interest in Bayesian
MLMs among astrostatisticians. I discuss the role of MLMs in the SCMA V contri-
butions elsewhere (Loredo 2012c). Four of the contributions in the present volume
rely on MLMs. Two address complex problems and help mark an era of new com-
plexity and sophistication in astrophysical MLMs. Two highlight the potential of
basic MLMs for addressing common astronomical data analysis problems that defy
accurate analysis with conventional methods taught to astronomers. I will highlight
the MLM aspects of these contributions in turn.
Some of the most impressive new Bayesian multilevel modeling in astronomy
addresses the analysis of measurements of multicolor light curves (brightness vs.
time in various wavebands) from Type Ia supernovae (SNe Ia). In the late 1990s,
astronomers discovered that these enormous stellar thermonuclear exoplosions are
“standardizable candles;” the shapes of their light curves are strongly correlated
with their luminosities (the intrinsic amplitudes of the light curves). This enables
use of SNe Ia to measure cosmic distances (via a generalized inverse-square law)
and to trace the history of cosmic expansion. The 2011 Nobel Prize in physics went
to three astronomers who used this capability to show that the expansion rate of
the universe is growing with time (“cosmic acceleration”), indicating the presence
of “dark energy” that somehow prevents the deceleration one would expect from
gravitational attraction.
A high-water mark in astronomical Bayesian multilevel modeling was set by
Mandel et al. (2011), who address the problem of estimating supernova luminosi-
ties from light curve data. Their model has three levels, complex connections be-
tween latent variables (some of them random functions—light curves—rather than
scalars), and jointly describes three different types of data (light curves, spectra, and
host galaxy spectroscopic redshifts). Panel (b) of Fig. 4 shows the graph for their
20 Thomas J. Loredo

MLM (the reader will have to consult Mandel et al. (2011), or Mandel’s more tuto-
rial overview in Feigelson & Babu (2012), for a description of the variables and the
model). In this volume, March et al. tackle a subsequent SNe Ia problem: how to
use the output of light curve models to estimate cosmological parameters, including
the density of dark energy, and an “equation of state” parameter aiming to capture
how the dark energy density may be evolving. Their framework can fuse informa-
tion from SN Ia with information from other sources, such as the power spectrum
of CMB fluctuations, and characterization of the baryon acoustic oscillation (BAO)
seen in the large scale spatial distribution of galaxies. Panel (c) of Fig. 4 shows the
graph for their model, also impressive for its complexity (see their contribution in
this volume for an explanation of this graph). I display the graphs (without much
explanation) to show how much emerging MLM research in astronomy is leapfrog-
ging the simple models of the recent past, exemplified by Panel (a) in the figure.
Equally impressive is Wandelt’s contribution, describing a hierarchical Bayes ap-
proach (albeit without explicit MLM language) for reconstructing the galaxy den-
sity field from noisy photometric redshift data measuring the line-of-sight velocities
of galaxies (which includes a component from cosmic expansion and a “peculiar
velocity” component from gravitational attraction of nearby galaxies and dark mat-
ter). His team’s framework (described in detail by Jasche & Wandelt 2011) includes
nonparametric estimation of the density field at an upper level, which adaptively
influences estimation of galaxy distances and peculiar velocities at a lower level,
borrowing strength in the manner described above. The nonparametric upper level,
and the size of the calculation, make this work groundbreaking.
The contributions by Andreon and by Kunz et al. describe simpler MLMs, but
with potentially broad applicability. Andreon describes regression with measure-
ment error (fitting lines and curves using data with errors in both the abscissa and
ordinate) using basic MLMs, including testing model adequacy with a relatively
simple predictive distribution test resembling the tail-area p-value (“significance
level”) tests familiar to astronomers. Such problems are common in astronomy.
Kelly (2007; see also Kelly’s more introductory treatment in Feigelson & Babu
2012) provided a more formal account of the Bayesian treatment of such prob-
lems, drawing on the statistics literature on measurement error problems (e.g., Car-
roll et al. 2006). The complementary emphasis of Andreon’s account is on how
straightforward—indeed, almost automatic—implementation can be using modern
software packages such as BUGS and JAGS.8 The contribution by Kunz et al. further
develops the Bayesian estimation applied to multiple species (BEAMS) framework
first described by Kunz et al. (2007). BEAMS aims to improve parameter estima-
tion in nonlinear regression when the data may come from different types of sources
(“species” or classes), with different error or population distributions, but with un-
certainty in the type of each datum. The classification labels for the data become
discrete latent parameters in a MLM; marginalizing over them (and possibly esti-
mating parameters in the various error distributions) can greatly improve inferences.
They apply the approach to estimating cosmological parameters using SNe Ia data,

8 http://www.mrc-bsu.cam.ac.uk/bugs/
Bayesian astrostatistics 21

and show that accounting for uncertainty in supernova classification has the poten-
tial to significantly improve the precision and accuracy of estimates. In the context
of this volume, one cannot help but wonder what would come of integrating some-
thing like BEAMS into the MLM framework of March et al..
In Section 2 I noted how the “variability = uncertainty” misconception leads
many astronomers to ignore relevant frequentist statistics literature; it gives the an-
alyst the impression that the ability to quantify variability is all that is needed to
devise a sound frequentist procedure. There is a similar danger for Bayesian in-
ference. Once one has specified a model for the data (embodied in the likelihood
function), Bayesian inference appears to be automatic in principle; one just follows
the rules of probability theory to calculate probabilities for hypotheses of interest
(after assigning priors). But despite the apparent simplicity of the sum and product
rules, probability theory can exhibit a sophisticated subtlety, with apparently in-
nocuous assumptions and calculations sometimes producing surprising results. The
huge literature on Bayesian methods is more than mere crank turning; it amasses
significant experience with this sophisticated machinery that astronomers should
mine to guide development of new Bayesian methods, or to refine existing ones. For
this non-statistician reader, it is the simpler of the contributions in this volume—
those by Andreon and by Kunz et al.—that point to opportunities for “borrowing
and mustering of strength” from earlier work by statisticians.
Andreon’s “de-TeXing” approach to Bayesian modeling (i.e., transcribing model
equations to BUGS or JAGS code, and assigning simple default priors) is both ap-
pealing and relatively safe for simple models, such as standard regression (paramet-
ric curve fitting with errors only in the ordinate) or density estimation from precise
point measurements in low dimensions. But the implications of modeling assump-
tions become increasingly subtle as model complexity grows, particularly when one
starts increasing the number of levels or the dimensions of model components (or
both). This makes me wary of an automated style of Bayesian modeling. Let us
focus here on some MLM subtlety that can complicate matters.
Information gain from the data tends to weaken as one considers parameters at
increasingly high levels in a multilevel model (Goel & DeGroot 1981). On the one
hand, if one is interested in quantities at lower levels, this weakens dependence on
assumptions made at high levels. On the other hand, if one is interested in high-
level quantities, sensitivity to the prior becomes an issue. The weakened impact of
data on high levels has the effect that improper (i.e., non-normalized) priors that are
safe to use in simple models (because the likelihood makes the posterior proper)
can be dangerous in MLMs, producing improper posteriors; proper but “vague”
default priors may hide the problem without truly ameliorating it (Hadjicostas &
Berry 1999; Gelman 2006). Paradoxically, in some settings one may need to as-
sign very informative upper-level priors to allow lower level distributions to adapt
to the data (see Esch et al. 2004 for an astronomical example). Also, the impact
of the graph structure on a model’s predictive ability becomes less intuitively ac-
cessible as complexity grows, making predictive tests of MLMs important, but also
nontrivial; simple posterior predictive tests may be insensitive to significant discrep-
ancies (Sinharay & Stern 2003; Gelman et al. 2004; Bayarri & Castellanos 2007).
22 Thomas J. Loredo

An exemplary feature of the SNe Ia MLM work of Mandel et al. is the use of care-
ful predictive checks, implemented via a frequentist cross-validation procedure, to
quantitatively assess the adequacy of various aspects of the model (notably, Man-
del audited graduate-level statistics courses to learn the ins and outs of MLMs for
this work, comprising his PhD thesis). In the context of nonlinear regression with
measurement error—Andreon’s topic—Carroll et al. (2006) provides a useful en-
try point to both Bayesian and frequentist literature, incidentally also describing a
number of frequentist approaches to such problems that would be more fair com-
petitors to Bayesian MLMs than the “χ 2 ” approach that serves as Andreon’s straw
man competitor.
The problem addressed by the BEAMS framework—essentially the problem of
data contamination, or mixed data—is not unique to astronomy, and there is sig-
nificant statistics literature addressing similar problems with lessons to offer as-
tronomers. BEAMS is a form of Bayesian regression using FMM error distributions.
Statisticians first developed such models a few decades ago, to treat outliers (points
evidently not obeying the assumed error distribution) using simple two-component
mixtures (e.g., normal distributions with two different variances). More sophisti-
cated versions have since been developed for diverse applications. An astronomical
example building on some of this expertise is the finite mixture modeling of stel-
lar populations by van Dyk et al. (2009), mentioned above. The most immediate
lessons astronomers may draw from this literature are probably computational; for
example, algorithms using data augmentation (which involves a kind of guided, iter-
ative Monte Carlo sampling of the class labels and weights) may be more effective
for implementing BEAMS than the weight-stepping approach described by Kunz et
al. in this volume.

4 Provocation

I will close this commentary with a provocative recommendation I have offered at


meetings since 2005 but not yet in print, born of my experience using multilevel
models for astronomical populations. It is that astronomers cease producing cata-
logs of estimated fluxes and other source properties from surveys. This warrants
explanation and elaboration.
As noted above, a consequence of the hierarchical structure of MLMs is that
the values of latent parameters at low levels cannot be estimated independently of
each other. In a survey context, this means that the flux (and potentially other prop-
erties) of a source cannot be accurately or optimally estimated considering only
the data for that source. This may initially seem surprising, but at some level as-
tronomers already know this to be true. We know—from Eddington, Malmquist,
and Lutz & Kelker—that simple estimates of source properties will be misleading
if we do not take into account information besides the measurements and selection
effects; we also must specify the population distribution of the property. The stan-
dard Malmquist and Lutz-Kelker corrections adopt an a priori fixed (e.g., spatially
Bayesian astrostatistics 23

homogeneous) population distribution, and produce a corrected estimate indepen-


dently for each object. What the fully Bayesian MLM approach adds to the picture
is the ability to handle uncertainty in the population distribution. After all, a prime
reason for performing surveys is to learn about populations. When the population
distribution is not well-known a priori, each source property measurement bears
on estimation of the population distribution, and thus indirectly, each measurement
bears on the estimation of the properties of every other source, via a kind of adaptive
bias correction (Loredo & Hendry 2010).9 This is Tukey’s “mustering and borrow-
ing of strength” at work again.
To enable this mustering and borrowing, we have to stop thinking of a catalog
entry as providing all the information needed to infer a particular source’s proper-
ties (even in the absence of auxiliary information from outside a particular survey).
Such a complete summary of information is provided by the marginal posterior
distribution for that source, which depends on the data from all sources—and on
population-level modeling assumptions. However, in the MLM structure (e.g., panel
(d) of Fig. 3), the likelihood function for the properties of a particular source may
be independent of information about other sources. The simplest output of a survey
that would enable accurate and optimal subsequent analysis is thus a catalog of like-
lihood functions (or possibly marginal likelihood functions when there are uncertain
survey-specific backgrounds or other “nuisance” effects the surveyor must account
for).
For a well-measured source, the likelihood function may be well-approximated
by a Gaussian that can be easily summarized with a mean and standard deviation.
But these should not be presented as point estimates and uncertainties.10 For sources
near the “detection limit,” more complicated summaries may be justified. Counter-
part surveys should cease reporting upper limits when a known source is not se-
curely detected; instead they should report a more informative non-Gaussian likeli-
hood summary. Discovery surveys (aiming to detect new sources rather than coun-
terparts) could potentially devise likelihood summaries that communicate informa-
tion about sources with fluxes below a nominal detection limit, and about uncertain
source multiplicty in crowded fields. Recent work on maximum-likelihood fitting
of “pixel histograms” (also known as “probability of deflection” or P(D) distribu-
tions), which contain information about undetected sources, hints at the science such
summaries might enable in a MLM setting (e.g., Patanchon et al. 2009).
In this approach to survey reporting, the notion of a detection limit as a decision
boundary identifying sources disappears. In its place there will be decision bound-
aries, driven by both computational and scientific considerations, that determine

9 It is worth pointing out that this is not a uniquely Bayesian insight. Eddington, Malmquist, and

Lutz & Kelker used frequentist arguments to justify their corrections; Eddington even offered adap-
tive corrections. The large and influential statistics literature on shrinkage estimators leads to sim-
ilar conclusions; see Loredo (2007) for further discussion and references.
10 I am tempted to recommend that, even in this regime, the likelihood summary be chosen so as

to deter misuse as an estimate, say by tabulating the +1σ and −2σ points rather than means and
standard deviations. I am only partly facetious about this!
24 Thomas J. Loredo

what type of likelihood summary is associated with each possible candidate source
location.
Coming at this issue from another direction, Hogg & Lang (2011) have recently
made similar suggestions, including some specific ideas for how likelihoods may be
summarized. Multilevel models provide a principled framework, both for motivat-
ing such a thoroughgoing revision of current practice, and for guiding its detailed
development. Perhaps by the 2015 ISI World Congress in Brazil we will hear reports
of analyses of the first survey catalogs providing such more optimal, MLM-ready
summaries.
But even in the absence of so revolutionary a development, I think one can
place high odds in favor of a bet that Bayesian multilevel modeling will become in-
creasingly prevalent (and well-understood) in forthcoming astrostatistics research.
Whether Bayesian methods (multilevel and otherwise) will start flourishing out-
side the astrostatistics research community is another matter, dependent on how
effectively astrostatisticians can rise to the challenge of correcting misconceptions
about both frequentist and Bayesian statistics, such as those outlined above. The
abundance of young astronomers with enthusiasm for astrostatistics makes me op-
timistic.

Acknowledgements I gratefully acknowledge NSF and NASA for support of current research un-
derlying this commentary, via grants AST-0908439, NNX09AK60G and NNX09AD03G. I thank
Martin Weinberg for helpful discussions on information propagation within multilevel models.
Students of Ed Jaynes’s writings on probability theory in physics may recognize the last part of my
title, borrowed from a commentary by Jaynes on the history of Bayesian and maximum entropy
ideas in the physical sciences (Jaynes 1993). This bit of plagiarism is intended as a homage to
Jaynes’s influence on this area—and on my own research and thinking.

References

1. Bayarri, M.J., Berger, J.O.: The interplay of Bayesian and frequentist analysis.
Statist. Sci. 19(1), 58–80 (2004). DOI 10.1214/088342304000000116. URL
http://dx.doi.org/10.1214/088342304000000116
2. Bayarri, M.J., Castellanos, M.E.: Bayesian checking of the second levels of hierarchical mod-
els. Statist. Sci. 22(3), 322–343 (2007). DOI 10.1214/07-STS235.
3. Bovy, J., Hogg, D.W., Roweis, S.T.: Extreme deconvolution: Inferring complete distribution
functions from noisy, heterogeneous and incomplete observations. Ann. Appl. Stat. 5(2B),
1657–1677 (2011)
4. Budavári, T., Szalay, A.S.: Probabilistic Cross-Identification of Astronomical Sources. Astro-
physical Journal 679, 301–309 (2008). DOI 10.1086/587156
5. Carroll, R.J., Ruppert, D., Stefanski, L.A., Crainiceanu, C.M.: Measurement error in non-
linear models, Monographs on Statistics and Applied Probability, vol. 105, second edn.
Chapman & Hall/CRC, Boca Raton, FL (2006). DOI 10.1201/9781420010138. URL
http://dx.doi.org/10.1201/9781420010138. A modern perspective
6. Dobigeon, N., Tourneret, J.Y., Scargle, J.D.: Joint segmentation of multivariate astronomical
time series: Bayesian sampling with a hierarchical model. IEEE Trans. Signal Process. 55(2),
414–423 (2007). DOI 10.1109/TSP.2006.885768.
Bayesian astrostatistics 25

7. Efron, B.: Bayesians, frequentists, and physicists. In: L. Lyons (ed.) PHYSTAT2003: Statis-
tical Problems in Particle Physics, Astrophysics, and Cosmology, SLAC eConf C030908, pp.
17–24 (2003)
8. Esch, D.N., Connors, A., Karovska, M., van Dyk, D.A.: An Image Restoration Technique with
Error Estimates. Astrophysical Journal 610, 1213–1227 (2004). DOI 10.1086/421761
9. Feigelson, E.D., Babu, G.J. (eds.): Statistical Challenges in Modern Astronomy. Springer
(1992)
10. Feigelson, E.D., Babu, G.J. (eds.): Statistical Challenges in Modern Astronomy V. Springer
(2012)
11. Gelman, A.: Prior distributions for variance parameters in hierarchical models (comment on
article by Browne and Draper). Bayesian Anal. 1(3), 515–533 (electronic) (2006)
12. Gelman, A., Carlin, J.B., Stern, H.S., Rubin, D.B.: Bayesian data analysis, second edn. Texts
in Statistical Science Series. Chapman & Hall/CRC, Boca Raton, FL (2004)
13. Goel, P.K., DeGroot, M.H.: Information about hyperparameters in hierarchical models. J.
Amer. Statist. Assoc. 76(373), 140–147 (1981).
14. Graziani, C., Lamb, D.Q.: Likelihood methods and classical burster repetition. In: R. E. Roth-
schild & R. E. Lingenfelter (ed.) High Velocity Neutron Stars, American Institute of Physics
Conference Series, vol. 366, pp. 196–200 (1996). DOI 10.1063/1.50246
15. Gull, S.F., Daniell, G.J.: Image reconstruction from incomplete and noisy data. Nature 272,
686–690 (1978). DOI 10.1038/272686a0
16. Gull, S.F., Daniell, G.J.: The Maximum Entropy Method (invited Paper). In: C. van Schoon-
eveld (ed.) IAU Colloq. 49: Image Formation from Coherence Functions in Astronomy, As-
trophysics and Space Science Library, vol. 76, p. 219 (1979)
17. Hadjicostas, P., Berry, S.M.: Improper and proper posteriors with improper priors in a Poisson-
gamma hierarchical model. Test 8(1), 147–166 (1999). DOI 10.1007/BF02595867.
18. Hogg, D.W., Lang, D.: Telescopes don’t make catalogues! In: EAS Publications Series, EAS
Publications Series, vol. 45, pp. 351–358 (2011). DOI 10.1051/eas/1045059
19. Jasche, J., Wandelt, B.D.: Bayesian inference from photometric redshift surveys.
ArXiv/1106.2757 (2011)
20. Jaynes, E.T.: Probability theory in science and engineering. Colloquium lectures in pure
and applied science. Socony Mobil Oil Co. Field Research Laboratory (1959). URL
http://books.google.com/books?id=Ft4-AAAAIAAJ
21. Jaynes, E.T.: A Backward Look to the Future. In: W. T. Grandy, Jr. & P. W. Milonni (ed.)
Physics and Probability, pp. 261–276 (1993)
22. Jeffreys, H.: Theory of probability. Third edition. Clarendon Press, Oxford (1961)
23. Kelly, B.C.: Some Aspects of Measurement Error in Linear Regression of Astronomical Data.
Astrophysical Journal 665, 1489–1506 (2007). DOI 10.1086/519947
24. Kitaura, F.S., Enßlin, T.A.: Bayesian reconstruction of the cosmological large-scale structure:
methodology, inverse algorithms and numerical optimization. Mon. Not. Roy. Astron. Soc.
389, 497–544 (2008). DOI 10.1111/j.1365-2966.2008.13341.x
25. Kuhn, T.S.: The structure of scientific revolutions. Second edition. University of Chicago
Press, Chicago (1970)
26. Kunz, M., Bassett, B.A., Hlozek, R.A.: Bayesian estimation applied to multiple species. Phys-
ical Review D 75(10), 103508 (2007). DOI 10.1103/PhysRevD.75.103508
27. Lampton, M., Margon, B., Bowyer, S.: Parameter estimation in X-ray astronomy. Astrophys-
ical Journal 208, 177–190 (1976). DOI 10.1086/154592
28. Little, R.J.: Calibrated Bayes: a Bayes/frequentist roadmap. Amer.
Statist. 60(3), 213–223 (2006). DOI 10.1198/000313006X117837. URL
http://dx.doi.org/10.1198/000313006X117837
29. Loredo, T.J.: Promise of Bayesian inference for astrophysics. In: E. D. Feigelson & G. J. Babu
(ed.) Statistical Challenges in Modern Astronomy, pp. 275–306 (1992)
30. Loredo, T.J.: The promise of bayesian inference for astrophysics (unabridged). Tech. rep.,
Department of Astronomy, Cornell University (1992). CiteSeer DOI 10.1.1.56.1842
26 Thomas J. Loredo

31. Loredo, T.J.: The return of the prodigal: Bayesian inference For astrophysics. In: J. M.
Bernardo, J. O. Berger, A. P. Dawid and A. F. M. Smith (eds.) Bayesian Statistics 5 Prelimi-
nary Proceedings, volume distributed to participants of the 5th Valencia Meeting on Bayesian
Statistics (1995). CiteSeer DOI 10.1.1.55.3616
32. Loredo, T.J.: Computational Technology for Bayesian Inference. In: D. M. Mehringer,
R. L. Plante, & D. A. Roberts (ed.) Astronomical Data Analysis Software and Systems VIII,
Astronomical Society of the Pacific Conference Series, vol. 172, p. 297 (1999)
33. Loredo, T.J.: Analyzing Data from Astronomical Surveys: Issues and Directions. In:
G. J. Babu & E. D. Feigelson (ed.) Statistical Challenges in Modern Astronomy IV, Astro-
nomical Society of the Pacific Conference Series, vol. 371, p. 121 (2007)
34. Loredo, T.J.: Statistical foundations and statistical practice (contribution to a panel discussion
on the future of astrostatistics). In: E. D. Feigelson & G. J. Babu (ed.) Statistical Challenges
in Modern Astronomy, p. 7pp. Springer (2012a, in press)
35. Loredo, T.J.: Commentary on Bayesian coincidence assessment (cross-matching). In:
E. D. Feigelson & G. J. Babu (ed.) Statistical Challenges in Modern Astronomy, p. 6pp.
Springer (2012b, in press)
36. Loredo, T.J.: Commentary on Bayesian analysis across astronomy. In: E. D. Feigelson &
G. J. Babu (ed.) Statistical Challenges in Modern Astronomy, p. 12pp. Springer (2012c, in
press)
37. Loredo, T.J., Hendry, M.A.: Bayesian multilevel modelling of cosmological populations. In:
Hobson, M. P., Jaffe, A. H., Liddle, A. R., Mukeherjee, P., & Parkinson, D. (ed.) Bayesian
Methods in Cosmology, p. 245. Cambridge University Press (2010)
38. Loredo, T.J., Lamb, D.Q.: Bayesian analysis of neutrinos observed from supernova SN 1987A.
Physical Review D 65(6), 063002 (2002). DOI 10.1103/PhysRevD.65.063002
39. Loredo, T.J., Wasserman, I.M.: Inferring the Spatial and Energy Distribution of Gamma-Ray
Burst Sources. II. Isotropic Models. Astrophysical Journal 502, 75 (1998). DOI 10.1086/
305870
40. Luo, S., Loredo, T., Wasserman, I.: Likelihood analysis of GRB repetition. In: C. Kouve-
liotou, M. F. Briggs, & G. J. Fishman (ed.) American Institute of Physics Conference Se-
ries, American Institute of Physics Conference Series, vol. 384, pp. 477–481 (1996). DOI
10.1063/1.51706
41. Mandel, K.S., Narayan, G., Kirshner, R.P.: Type Ia Supernova Light Curve Inference: Hier-
archical Models in the Optical and Near-infrared. Astrophysical Journal 731, 120 (2011).
DOI 10.1088/0004-637X/731/2/120
42. Nousek, J.A.: Source existence and parameter fitting when few counts are available. In:
E. D. Feigelson & G. J. Babu (ed.) Statistical Challenges in Modern Astronomy, pp. 307–
327 (1992)
43. Patanchon, G., et al.: Submillimeter Number Counts from Statistical Analysis of BLAST
Maps. Astrophysical Journal 707, 1750–1765 (2009). DOI 10.1088/0004-637X/707/2/1750
44. Petit, J.M., Kavelaars, J.J., Gladman, B., Loredo, T.: Size Distribution of Multikilometer
Transneptunian Objects. In: Barucci, M. A., Boehnhardt, H., Cruikshank, D. P., Morbidelli,
A., & Dotson, R. (ed.) The Solar System Beyond Neptune, pp. 71–87. University of Arizona
Press (2008)
45. Readhead, A.C.S., Lawrence, C.R.: Observations of the isotropy of the cosmic microwave
background radiation. Annual Review of Astronomy & Astrophysics 30, 653–703 (1992).
DOI 10.1146/annurev.aa.30.090192.003253
46. Ripley, B.D.: Bayesian methods of deconvolution and shape classification. In: E. D. Feigelson
& G. J. Babu (ed.) Statistical Challenges in Modern Astronomy, pp. 329–346 (1992)
47. Sims, C.: Understanding non-bayesians. Unpublished chapter, De-
partment of Economics, Princeton University (2010). URL
http://www.princeton.edu/˜ sims/#UndstndngNnBsns
48. Sinharay, S., Stern, H.S.: Posterior predictive model checking in hierarchical models. J. Statist.
Plann. Inference 111(1-2), 209–221 (2003). DOI 10.1016/S0378-3758(02)00303-8.
49. Sturrock, P.A.: Evaluation of Astrophysical Hypotheses. Astrophysical Journal 182, 569–580
(1973). DOI 10.1086/152165
Bayesian astrostatistics 27

50. van Dyk, D.A., Connors, A., Kashyap, V.L., Siemiginowska, A.: Analysis of Energy Spectra
with Low Photon Counts via Bayesian Posterior Simulation. Astrophysical Journal 548, 224–
243 (2001). DOI 10.1086/318656
51. van Dyk, D.A., DeGennaro, S., Stein, N., Jefferys, W.H., von Hippel, T.: Statistical analysis of
stellar evolution. Ann. Appl. Stat. 3(1), 117–143 (2009). DOI 10.1214/08-AOAS219. URL
http://dx.doi.org/10.1214/08-AOAS219
52. Wandelt, B.D., Larson, D.L., Lakshminarayanan, A.: Global, exact cosmic microwave back-
ground data analysis using Gibbs sampling. Physical Review D 70(8), 083511 (2004). DOI
10.1103/PhysRevD.70.083511
53. West, M.: Commentary. In: E. D. Feigelson & G. J. Babu (ed.) Statistical Challenges in
Modern Astronomy, p. 328ff (1992)

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy