Rafael S. de Souza
Emille E. O. Ishida
Alberto Krone-Martins
Communicated by Notices Associate Editor Richard A. Levine
In this short review, we trace the evolution of inference in astronomy, highlighting key milestones rather than providing an exhaustive survey. We focus on the shift from classical optimization to Bayesian inference, the rise of gradient-based methods fueled by advances in deep learning, and the emergence of adaptive models that shape the very design of scientific datasets. Understanding this shift is essential for appreciating the current landscape of astronomical research and the future it is helping to build.
Inference—the process of drawing conclusions about a model or its parameters from data—is central to scientific inquiry into the Universe. Since ancient times, humans have built what we now call predictive models, aimed at estimating future outcomes under given conditions in the world, and explanatory models, to elucidate relationships among variables and the mechanisms underlying observed phenomena. In all cases, models depend on parameters to be inferred or distributions to be determined. Throughout history, methods of inference have been deeply linked to advances in astronomy, dating back to Babylonian times 18.
According to Πτολεμαῖος (Ptolemy), in his second-century AD treatise Μαϑηματιχὴ Σύνταξις (later known through Arabic translations as the Almagest), the Greek astronomer ῞Ιππαρχος (Hipparchus) improved the precision of his astronomical measurements—such as the timing of solstices, equinoxes, and lunar eclipses—by averaging the earliest and latest observations of the same event. This approach mitigates observational noise by assuming that timing errors are roughly symmetric, allowing deviations to cancel and yielding a more accurate estimate of the true event time.
In the sixteenth century, the Portuguese mathematician and cosmographer Pedro Nunes discovered the loxodrome, or rhumb line,Footnote1 on the surface of a sphere and, perhaps more importantly, invented the NoniusFootnote2 By combining readings from several closely spaced, independent scales, the Nonius enabled more precise measurements of quantities such as the altitude angles of celestial bodies.
A century later, a simplified version of Nunes’ method gave rise to the Vernier scales still used in modern calipers. With their fine divisions, these scales remain one of the key innovations behind the precision of modern absolute encoders, integral to computer numerically controlled machines and other robots essential to contemporary life. Nunes’ scales and methods inspired Tycho Brahe, as Tycho himself acknowledged in his book Astronomiæ instauratæ mechanica (Instruments for the restoration of astronomy). In the sixteenth century, Tycho Brahe appears to have been the first to deliberately make a large number of repeated observations of the same quantity—specifically, the positions of stars—which he later combined into a single inferred value. This process enabled Tycho to achieve unprecedented precision in determining the coordinates of individual stars and the motions of planets, laying the groundwork for Kepler’s laws of planetary motion—an achievement that definitively displaced Earth from the center of the Universe.
Copernicus’s 1543 heliocentric model, De revolutionibus orbium coelestium, had already challenged Ptolemy’s system, but it was Kepler’s laws, backed by Tycho’s precise data, that sealed the case. Not only did they rest on better observations, but they embodied what William of Ockham had axiomatized in his Summa Logicae: the principle of parsimony, or “Occam’s razor.” In practice, this meant that fewer adjustable parameters were needed to match observation, lending the heliocentric picture a decisive methodological advantage over its geocentric rival.
Kepler’s laws, inferred directly from Tycho’s precise data, served as predictive models. Their physical explanation later inspired Newton’s laws of motion and gravity, along with the development of infinitesimal calculus. In the same century, Galileo’s 1609 Dialogo sopra I due massimi sistemi del mondo (Dialogue on the Two Great World Systems) offered an early discussion of observational errors concerning the distance estimate to the 1572 supernova in the constellation of Cassiopeia. The formal understanding of how multiple observations improve inference would emerge only later.
In his Théorie analytique des probabilités (1812), Pierre-Simon Marquis de Laplace extended Abraham De Moivre’s formulation of the result later named the zentraler Grenzwertsatz (Central Limit Theorem) by George Pólya. De Moivre first introduced it in a 1733 article and later included it in the The Doctrine of Chances (1738). This cornerstone of probability theory was further refined by Augustin-Louis Cauchy, Siméon-Denis Poisson, and Friedrich Wilhelm Bessel.
In the same century, Carl Friedrich Gauss gained early recognition in astronomy by accurately predicting the orbit of the planetoid Ceres, briefly observed in early 1801 before disappearing. Ceres was located near Gauss’s predicted position later that year. He published his methods in Theoria motus corporum coelestium, introducing the method of least squares, which he claimed to have developed in 1795. This led to a controversy with Adrien-Marie Legendre, who had published a similar least squares method in his 1805 work Nouvelles méthodes pour la détermination des orbites des comètes. Regardless of this dispute, Laplace was the first to formalize the method as a probabilistic problem in his 1812 Théorie Analytique des Probabilités.
Curiously, inverse probability and Bayes’ theorem, developed by Thomas Bayes and Pierre-Simon Laplace from Laplace’s own astronomical work, had limited influence in astronomy until the late twentieth century. The broader adoption of Bayesian methods in the field owes much to Sir Harold Jeffreys, a geophysicist and Plumian Professor of Astronomy at Cambridge, who axiomatized Laplace’s theory in his 1939 book, Theory of Probability. By the 1990s, Bayesian methods saw broader use, especially in extragalactic astronomy and cosmology (e.g., 10). The rise of advanced Bayesian techniques and probabilistic programming is now enabling astronomers to address increasingly complex problems.
Linear Regression in Astronomy
Linear regression has long been central to astronomical data analysis 4, dating back to its early use by Isaac Newton around 1700 to study the timing of the equinoxes—points in Earth’s orbit when the Sun crosses the celestial equator. He achieved this by combining ancient data from Hipparchus with observations by his contemporary John Flamsteed.
An archetypal example includes early attempts to estimate the best-fit line in Hubble’s diagram, which relates galaxies’ recessional velocities (how fast they are moving away from us) to their distances. These analyses revealed that the Universe is expanding, with the slope of the line defining the Hubble constant , a measure of the current rate of cosmic expansion 69. Formally, , where
Among its many applications, linear regression also played a pivotal role in Hulse and Taylor’s Nobel Prize-winning research on binary pulsars 19. The Hulse–Taylor pulsar (PSR B1913+16) was the first binary pulsar discovered, a system of two neutron stars orbiting their common center of mass. Neutron stars are remnants of massive stars, typically about a dozen kilometers in diameter, whose cores collapsed to densities exceeding that of an atomic nucleus. In such a system, one of the stars appears as a highly regular radio pulsar: a rapidly spinning neutron star that emits beams of radio waves from its magnetic poles. These beams sweep across Earth at regular intervals, much like a lighthouse, allowing precise timing measurements. The companion star is often electromagnetically invisible and detectable only through its gravitational effects. Observations of orbital decay in the system closely matched Einstein’s prediction of energy loss via gravitational wave emission, providing indirect evidence for gravitational waves more than two decades before their direct detection.Footnote4
Beyond compact-object systems, linear regression also plays a crucial role in uncovering large-scale empirical relations in extragalactic astrophysics. Many such scaling laws, often expressed as power laws, can be linearized by taking logarithms. A classical example is the log-linear relationship between the mass of a galaxy’s central supermassive black hole (
where
These early applications exemplify a broader principle: many problems in astronomy reduce to estimating parameters that best explain observational data. Whether fitting a line to a Hubble diagram or deriving black hole scaling relations, the core task often involves comparing model predictions with observations. A common formulation of this task is the estimation of model parameters
Assuming independent, identically distributed (i.i.d.) Gaussian errors, one seeks to minimize the discrepancy between model and data:
where
Popular optimization techniques include the downhill simplex method, Levenberg–Marquardt, conjugate gradient, and gradient descent. The latter updates parameter estimates iteratively according to:
where
When latent variables or incomplete data are present, the expectation–maximization algorithm 2 is often preferred for maximum likelihood estimation. The likelihood function
which can also be approached using gradient-based updates:
Both EM and gradient ascent share a common iterative structure:
where
A classical example in astronomical image reconstruction is the Richardson–Lucy deconvolution 1115, which is used to recover images degraded by the point-spread function (PSF)Footnote5 under Poisson noise. The observed image
The corresponding likelihood function is:
Applying the expectation–maximization algorithm to this model leads to the classical Richardson–Lucy update rule:
Figure 1 shows the Richardson–Lucy deconvolution applied to a blurred, Poisson-noised image of the Whirlpool Galaxy (M51). The left panel shows the degraded image, while the right panel displays the deconvolved result, with finer structures—such as spiral arms—becoming more discernible. Note that although motivated by Bayesian reasoning, the Richardson–Lucy algorithm produces a maximum likelihood estimation rather than a full posterior distribution. As with other EM and gradient-based approaches, it iteratively refines parameter estimates by maximizing the (log-)likelihood or, equivalently, minimizing an objective function.
Starting with the next section, we follow a focused cosmological example as an illustrative case to trace how this conceptual framework has evolved in response to increasingly complex data and models.
Left: A blurred, Poisson-noised image of the Whirlpool Galaxy M51. Right: Denoised version obtained via the Richardson–Lucy deconvolution. The method enhances the contrast of fine details such as spiral arms.
Measuring the Universe Expansion Rate
As noted earlier, in the early decades of the twentieth century, Edwin Hubble provided the first observational evidence for cosmic expansion by showing that the recession velocities of extra-galactic nebulae are proportional to their distance. These nebulae were later recognized as galaxies beyond the Milky Way. Hubble inferred their velocities from their redshifts, the Doppler-induced shift of spectral lines toward longer wavelengths, and thus established that more distant objects move away faster 6.
This conclusion emerged from a linear regression analysis of measurements from 24 nebulae in our cosmic neighborhood (Figure 2). It had a profound impact on subsequent cosmological studies, sparking a century-long effort to collect increasingly precise data in the hope of getting more clues about the overall dynamics and composition of cosmic structures.
Nearly a century later, Type Ia supernovae provided the first compelling evidence for the Universe’s accelerated expansion 1316. This led to the discovery of dark energy—a mysterious form of energy that permeates space and drives cosmic acceleration. The 2011 Nobel Prize in Physics was awarded for this discovery.
Original plot of radial velocity-distance relation for extragalactic nebulae 6.
Type Ia supernovae mark the catastrophic endpoint in the life of a binary star system, with a white dwarf at the core of the process. In such systems, two stars orbit a shared center of mass, and over time, the white dwarf, a compact remnant left behind by a Sun-like star, can accrete material from its companion. White dwarfs are supported against gravitational collapse by electron degeneracy pressure, a quantum mechanical effect arising from the Pauli exclusion principle, which prevents electrons from occupying the same quantum state. As the white dwarf accumulates mass, it eventually approaches the Chandrasekhar limit (about 1.4 times the mass of our Sun), beyond which degeneracy pressure can no longer support it. This triggers a runaway thermonuclear reaction that tears the star apart in a brilliant supernova explosion.
These explosions are among the most energetic events in the Universe, releasing an enormous yet remarkably consistent amount of energy at peak brightness (or luminosity, the total energy emitted per unit time). This near-uniformity enables Type Ia supernovae to serve as standardizable candles—astronomical objects whose intrinsic brightness can be empirically calibrated. By comparing the observed brightness to this calibrated intrinsic brightness, astronomers can infer distances, making Type Ia supernovae invaluable objects for measuring cosmic distances on large scales.
This standardization is achieved through empirical corrections to the observed magnitude that account for variations in the light curve—the evolution of a supernova’s brightness over time—particularly its width (known as stretch) and its color (which reflects shifts in the observed spectrum toward shorter or longer wavelengths). These corrections substantially reduce the intrinsic scatter in the magnitude-distance relation and improve the precision of cosmological distance estimates 14.
The observed magnitude
where
The distance modulus
where
Here,
Assuming a flat Universe, the luminosity distance
Here, dark matter refers to a nonluminous form of matter that interacts gravitationally but neither emits nor absorbs light. It plays a crucial role in the formation of cosmic structures. The assumption of flatness corresponds to a Universe with zero spatial curvature, as predicted by inflationary cosmology and supported by measurements of the cosmic microwave background. In general, a positively curved (closed) Universe would cause light rays to converge, while a negatively curved (open) Universe would cause them to diverge, altering the relationship between
The likelihood for the observed magnitudes
where
Here,
Estimating cosmological parameters, such as
To harness the full potential of these ever-growing datasets, astronomers have increasingly adopted advanced statistical methods—ranging from sampling techniques such as Hamiltonian Monte Carlo and nested sampling to modern approaches in likelihood-free and amortized inference. The following sections examine how these tools are reshaping cosmological analysis, particularly when likelihoods are intractable, simulations costly, or adaptive observational strategies required.
Hamiltonian Monte Carlo for cosmology
Markov chain Monte Carlo (MCMC) has been a cornerstone of computational Bayesian inference for decades. Two widely used methods in astronomy are the Metropolis-Hastings 12 algorithm and Gibbs sampling 5. In the case of Metropolis-Hastings, each MCMC step proposes a new state
Hamiltonian Monte Carlo (HMC; 3) improves sampling efficiency over Metropolis-Hastings by reducing autocorrelation and avoiding the inefficiencies of random walk behavior. It is particularly effective for high-dimensional and strongly correlated posterior distributions, which are common in astrophysical applications.
The main idea behind HMC is to reinterpret the negative log-posterior as a potential energy function. The parameter space
where the potential energy is
and the kinetic energy is
with
This physical analogy enables the sampler to explore the posterior by simulating the motion of a particle through parameter space, following smooth, energy-conserving trajectories. In contrast to the random-walk proposals used in traditional Markov chain Monte Carlo, Hamiltonian Monte Carlo generates distant, low-autocorrelation proposals that more efficiently traverse the high-probability regions of complex, high-dimensional posteriors 1. Figure 3 illustrates this intuition in a cosmological context, showing a typical Hamiltonian Monte Carlo trajectory when estimating
Hamiltonian Monte Carlo trajectory.
Amortized Inference
Astronomical research often relies on computationally intensive simulation to model physical processes that lie beyond the reach of direct observation. These include hydrodynamical, radiative, and magnetohydrodynamical simulations that aim to reproduce the complex behavior of astrophysical systems. However, performing full Bayesian inference using Markov chain Monte Carlo methods is often computationally prohibitive, especially with datasets comprising millions or billions of observations.
To address these challenges, astronomers have increasingly adopted scalable inference techniques, such as Approximate Bayesian Computation (ABC). Initially developed in population genetics and rooted in Rubin’s 17 rejection sampling thought experiment, ABC has since been adapted for astronomical applications. Over the past decade, it has gained traction in contexts ranging from measuring the Hubble constant 8 to modeling galaxy populations. Recent efforts have increasingly focused on amortized inference, a class of simulation-based or likelihood-free methods that leverage learned surrogate models—typically neural networks—to approximate posterior distributions. Unlike approximate Bayesian computation, which performs inference separately for each dataset, amortized approaches learn a global mapping from observations to posteriors that can be reused across datasets.
Formally, let
In simulation-based inference, the true posterior
Figure 4 illustrates this approach. Starting from a prior
Schematic illustration of amortized inference.
Surrogate model
A common implementation for surrogate models employs normalizing flows as neural density estimators to enable efficient, tractable approximations of complex posterior distributions. These models transform a simple base distribution
Figure 5 presents a toy illustration demonstrating how complex distributions can be approximated by applying a sequence of transformations to a Gaussian distribution. One popular class of normalizing flows is the Masked Autoregressive Flow, which models the joint distribution of
where
Illustration of a normalizing flow transforming a simple base distribution into a complex target distribution through successive invertible mappings.
Active Inference in Cosmology
Traditional statistical analyses often operate under the assumption that, if more data is needed, it can reasonably be collected. In astronomy, however, this is rarely possible. Our vantage point in the Universe imposes inherent observational limits. As early as 1584, Giordano Bruno noted in De l’infinito, universo et mondi that the faintness or distance of celestial bodies could hinder our ability to observe them. This idea was later formalized by Eddington in his 1914 treatise Stellar Movements and the Structure of the Universe, where he highlighted that stars visible to the naked eye form a biased sample. This line of reasoning ultimately led to the concept of Malmquist bias: in flux-limited surveys, intrinsically brighter objects are overrepresented, skewing inferences about the true distribution of celestial sources.
Selection effects are particularly evident when comparing spectroscopic and photometric samples, the two main types of observations in cosmological studies. Spectroscopy involves dispersing an object’s light into its constituent wavelengths to produce a spectrum, which reveals physical properties such as chemical composition, temperature, and velocity. For supernovae, spectroscopy enables precise classification by detecting features such as the absence of hydrogen lines, which is characteristic of Type Ia events. In contrast, photometry measures the total brightness of an object through broad filters, capturing less detailed but more easily obtained information. Once a supernova is classified, its brightness is tracked photometrically over time to estimate its distance. However, spectroscopy is resource-intensive and only feasible for a small subset of objects, whereas photometric surveys can observe far more events at the expense of precision. This imbalance introduces selection effects that can bias cosmological analyses if not properly addressed.
These challenges have prompted a paradigm shift. Rather than passively analyzing fixed datasets, astronomers increasingly embrace active inference strategies 7, where observations are selected strategically to maximize scientific return under limited resources. In this framework, data are no longer static inputs but dynamic choices: their potential informativeness is explicitly considered when deciding which measurements to collect. For example, selecting supernovae for spectroscopic follow-up involves balancing observational cost against the expected scientific gain. The problem resembles a knapsack optimization, where each candidate observation has a different value and must be chosen within the constraints of finite telescope time.
In information-theoretic terms, active inference aims to minimize “surprise,” the extent to which new data update prior beliefs. Formally, surprise corresponds to the negative log-evidence, where the evidence is the probability of the observed data under the model, a quantity that is generally intractable. In practice, one minimizes a tractable upper bound: the variational free energy. Within this framework, the goal is to evaluate how much an unobserved measurement would reduce uncertainty about latent variables or parameters of interest—such as redshift, supernova type, or cosmological parameters. A common strategy is to select the next observation
This strategy can be formalized by minimizing the variational free energy:
where
This workflow illustrates an active learning loop in astronomy, where spectroscopic follow-up is prioritized for novel objects that maximize information gain.
As illustrated in Figure 6, this iterative process combines observational constraints with machine learning to prioritize spectroscopic follow-ups that most effectively reduce uncertainty and correct for selection biases. Over time, such strategies enhance model performance while ensuring that limited observational resources are used efficiently.
Remarks
From charting planetary motions by hand to modeling the expansion of the Universe with billions of data points, astronomy has evolved into a discipline where inference is not just a method, but its very foundation. This journey, from Tycho Brahe’s systematic use of empirical observations to today’s simulation-based and active inference techniques, reflects a deeper shift: from passively recording the heavens to actively interrogating them. This transformation offers immense opportunities, but also demands unprecedented integration of mathematics, statistics, computer science, and physics to advance our understanding of the cosmos.
A testament to this progress is the statistical inference of millions—and even billions—of parameters, once unimaginable. The Gaia satellite, which offers the most comprehensive catalog of stellar positions and motions to date, exemplifies this feat. Its dataset is used to infer tens of billions of astrometric parameters from trillions of time measurements, integrated through a carefully designed mathematical optimization framework. This approach extends the weighted least-squares method, employing a hybrid solution that integrates iterative optimization techniques—blending Jacobi’s and Gauss-Seidel block methods—with a modified conjugate gradient method. The result is a monumental dataset detailing the positions and motions of billions of stars, which now forms the bedrock of contemporary astronomical research.
The future of astronomical inference stands poised for profound transformation through the integration of advanced statistical models. These models, capable of discerning intricate patterns within vast and complex datasets, promise to unlock new pathways for knowledge discovery while significantly streamlining labor-intensive data processing tasks. As these tools continue to evolve, they will enhance the synergy between human expertise and computational power, allowing astronomers to focus on high-level guidance and interpretation while automation handles much of the underlying analysis. This shift will not only accelerate the pace of discovery, but will also deepen our understanding of the cosmos.
References
[1]Michael Betancourt, Simon Byrne, Sam Livingstone, and Mark Girolami, The geometric foundations of Hamiltonian Monte Carlo, Bernoulli 23 (2017), no. 4A, 2257–2298, DOI 10.3150/16-BEJ810. MR3648031,
AMSref \bib{Betancourt2017}{article}{ author={Betancourt, Michael}, author={Byrne, Simon}, author={Livingstone, Sam}, author={Girolami, Mark}, title={The geometric foundations of Hamiltonian Monte Carlo}, journal={Bernoulli}, volume={23}, date={2017}, number={4A}, pages={2257--2298}, issn={1350-7265}, review={\MR {3648031}}, doi={10.3150/16-BEJ810}, }A. P. Dempster, N. M. Laird, and D. B. Rubin, Maximum likelihood from incomplete data via the EM algorithm: With discussion, J. Roy. Statist. Soc. Ser. B 39 (1977), no. 1, 1–38. MR501537,
AMSref \bib{Dempster1977}{article}{ author={Dempster, A. P.}, author={Laird, N. M.}, author={Rubin, D. B.}, title={Maximum likelihood from incomplete data via the EM algorithm}, subtitle={With discussion}, journal={J. Roy. Statist. Soc. Ser. B}, volume={39}, date={1977}, number={1}, pages={1--38}, issn={0035-9246}, review={\MR {501537}}, }Simon Duane, A. D. Kennedy, Brian J. Pendleton, and Duncan Roweth, Hybrid Monte Carlo, Phys. Lett. B 195 (1987), no. 2, 216–222, DOI 10.1016/0370-2693(87)91197-x. MR3960671,
AMSref \bib{DUANE1987216}{article}{ author={Duane, Simon}, author={Kennedy, A. D.}, author={Pendleton, Brian J.}, author={Roweth, Duncan}, title={Hybrid Monte Carlo}, journal={Phys. Lett. B}, volume={195}, date={1987}, number={2}, pages={216--222}, issn={0370-2693}, review={\MR {3960671}}, doi={10.1016/0370-2693(87)91197-x}, }Eric D. Feigelson and G. Jogesh Babu, Modern statistical methods for astronomy: With R applications, Cambridge University Press, Cambridge, 2012, DOI 10.1017/CBO9781139015653. MR2963292,
AMSref \bib{Feigelson2012}{book}{ author={Feigelson, Eric D.}, author={Babu, G. Jogesh}, title={Modern statistical methods for astronomy}, subtitle={With R applications}, publisher={Cambridge University Press, Cambridge}, date={2012}, pages={xviii+476}, isbn={978-0-521-76727-9}, review={\MR {2963292}}, doi={10.1017/CBO9781139015653}, }S. Geman and D. Geman, Stochastic Relaxation, Gibbs Distributions, and the Bayesian Restoration of Images, IEEE Transactions on Pattern Analysis and Machine Intelligence PAMI-6 (1984), no. 6, 721–741, DOI 10.1109/TPAMI.1984.4767596.,
AMSref \bib{Geman1984}{article}{ author={Geman, S.}, author={Geman, D.}, journal={IEEE Transactions on Pattern Analysis and Machine Intelligence}, title={Stochastic Relaxation, Gibbs Distributions, and the Bayesian Restoration of Images}, date={1984}, volume={PAMI-6}, number={6}, pages={721--741}, doi={10.1109/TPAMI.1984.4767596}, }E. Hubble, A relation between distance and radial velocity among extra-galactic nebulae, Proceedings of the National Academy of Sciences 15 (1929), no. 3, 168–173.,
AMSref \bib{hub29}{article}{ author={Hubble, E.}, title={A relation between distance and radial velocity among extra-galactic nebulae}, journal={Proceedings of the National Academy of Sciences}, volume={15}, number={3}, pages={168--173}, date={1929}, publisher={National Acad Sciences}, }E. E. O. Ishida, Machine learning and the future of supernova cosmology, Nature Astronomy 3 (2019), 680–682, DOI 10.1038/s41550-019-0860-6.,
AMSref \bib{Ishida2019}{article}{ author={Ishida, E. E.~O.}, title={Machine learning and the future of supernova cosmology}, journal={Nature Astronomy}, date={2019}, volume={3}, pages={680--682}, doi={10.1038/s41550-019-0860-6}, }E. Ishida, S. Vitenti, M. Penna-Lima, J. Cisewski, R. de Souza, A. Trindade, E. Cameron, and V. Busti, cosmoabc: Likelihood-free inference via Population Monte Carlo Approximate Bayesian Computation, Astronomy and Computing 13 (2015), 1–11, DOI 10.1016/j.ascom.2015.09.001.,
AMSref \bib{ISHIDA20151}{article}{ author={Ishida, E.}, author={Vitenti, S.}, author={Penna-Lima, M.}, author={Cisewski, J.}, author={de Souza, R.}, author={Trindade, A.}, author={Cameron, E.}, author={Busti, V.}, title={cosmoabc: Likelihood-free inference via Population Monte Carlo Approximate Bayesian Computation}, journal={Astronomy and Computing}, volume={13}, pages={1--11}, date={2015}, issn={2213-1337}, doi={10.1016/j.ascom.2015.09.001}, }G. Lemaître, Un Univers homogène de masse constante et de rayon croissant rendant compte de la vitesse radiale des nébuleuses extra-galactiques, Annales de la Société Scientifique de Bruxelles 47 (1927), 49–59.,
AMSref \bib{1927ASSB4749L}{article}{ author={Lema{\^{\i }}tre, G.}, title={Un Univers homog{\`e}ne de masse constante et de rayon croissant rendant compte de la vitesse radiale des n{\'e}buleuses extra-galactiques}, journal={Annales de la Soci{\'e}t{\'e} Scientifique de Bruxelles}, date={1927}, volume={47}, pages={49--59}, }T. J. Loredo, The Promise of Bayesian Inference for Astrophysics (Eric D. Feigelson and G. Jogesh Babu, ed.), Springer-Verlag, New York, Statistical Challenges in Modern Astronomy, 1992, pp. 275–297.,
AMSref \bib{Loredo1992}{collection.article}{ author={Loredo, T. J.}, booktitle={Statistical Challenges in Modern Astronomy}, book={ editor={Eric D. Feigelson and G. Jogesh Babu}, publisher={Springer-Verlag}, address={New York}, }, title={The Promise of Bayesian Inference for Astrophysics}, pages={275--297}, date={1992}, }L. B. Lucy, An iterative technique for the rectification of observed distributions, Astronomical Journal 79 (1974), 745, DOI 10.1086/111605.,
AMSref \bib{Lucy1974}{article}{ author={Lucy, L.~B.}, title={An iterative technique for the rectification of observed distributions}, journal={Astronomical Journal}, date={1974}, volume={79}, pages={745}, doi={10.1086/111605}, }Nicholas Metropolis and S. Ulam, The Monte Carlo method, J. Amer. Statist. Assoc. 44 (1949), 335–341. MR31341,
AMSref \bib{metropolis1949monte}{article}{ author={Metropolis, Nicholas}, author={Ulam, S.}, title={The Monte Carlo method}, journal={J. Amer. Statist. Assoc.}, volume={44}, date={1949}, pages={335--341}, issn={0162-1459}, review={\MR {31341}}, } S. Perlmutter, G. Aldering, G. Goldhaber, R. A. Knop, and others, Measurements of
M. M. Phillips, The Absolute Magnitudes of Type IA Supernovae, The Astrophysical Journal Letters 413 (1993), L105, DOI 10.1086/186970.,
AMSref \bib{phillips1993}{article}{ author={Phillips, M.~M.}, title={The Absolute Magnitudes of Type IA Supernovae}, journal={The Astrophysical Journal Letters}, date={1993}, volume={413}, pages={L105}, doi={10.1086/186970}, }W. H. Richardson, Bayesian-Based Iterative Method of Image Restoration, Journal of the Optical Society of America (1917-1983) 62 (1972), no. 1, 55.,
AMSref \bib{Richardson1972}{article}{ author={Richardson, W. H.}, title={Bayesian-Based Iterative Method of Image Restoration}, journal={Journal of the Optical Society of America (1917-1983)}, date={1972}, volume={62}, number={1}, pages={55}, }A. G. Riess and others, Observational Evidence from Supernovae for an Accelerating Universe and a Cosmological Constant, The Astronomical Journal 116 (1998), no. 3, 1009–1038, DOI 10.1086/300499.,
AMSref \bib{riess1998}{article}{ author={Riess, A. G.}, author={{others}}, title={Observational Evidence from Supernovae for an Accelerating Universe and a Cosmological Constant}, journal={The Astronomical Journal}, date={1998}, volume={116}, number={3}, pages={1009--1038}, doi={10.1086/300499}, }Donald B. Rubin, Bayesianly justifiable and relevant frequency calculations for the applied statistician, Ann. Statist. 12 (1984), no. 4, 1151–1172, DOI 10.1214/aos/1176346785. MR760681,
AMSref \bib{Rubin1984}{article}{ author={Rubin, Donald B.}, title={Bayesianly justifiable and relevant frequency calculations for the applied statistician}, journal={Ann. Statist.}, volume={12}, date={1984}, number={4}, pages={1151--1172}, issn={0090-5364}, review={\MR {760681}}, doi={10.1214/aos/1176346785}, }Stephen M. Stigler, The history of statistics: The measurement of uncertainty before 1900, The Belknap Press of Harvard University Press, Cambridge, MA, 1990. Reprint of the 1986 original. MR1057346,
AMSref \bib{Stigler1990}{book}{ author={Stigler, Stephen M.}, title={The history of statistics}, subtitle={The measurement of uncertainty before 1900}, note={Reprint of the 1986 original}, publisher={The Belknap Press of Harvard University Press, Cambridge, MA}, date={1990}, pages={xviii+410}, isbn={0-674-40341-X}, review={\MR {1057346}}, }J. H. Taylor and J. M. Weisberg, A new test of general relativity - Gravitational radiation and the binary pulsar PSR 1913+16, Astrophysical Journal 253 (1982), 908–920, DOI 10.1086/159690.,
AMSref \bib{Taylor1982}{article}{ author={Taylor, J.~H.}, author={Weisberg, J.~M.}, title={A new test of general relativity - Gravitational radiation and the binary pulsar PSR 1913+16}, journal={Astrophysical Journal}, date={1982}, volume={253}, pages={908--920}, doi={10.1086/159690}, }J. M. Weisberg, D. J. Nice, and J. H. Taylor, Timing Measurements of the Relativistic Binary Pulsar PSR B1913+16, The Astrophysical Journal 722 (2010), no. 2, 1030–1034, DOI 10.1088/0004-637X/722/2/1030.,
AMSref \bib{Weisberg2010}{article}{ author={Weisberg, J.~M.}, author={Nice, D.~J.}, author={Taylor, J.~H.}, title={Timing Measurements of the Relativistic Binary Pulsar PSR B1913+16}, journal={The Astrophysical Journal}, date={2010}, volume={722}, number={2}, pages={1030--1034}, doi={10.1088/0004-637X/722/2/1030}, }Rafael S. de Souza is a senior lecturer in data science at the University of Hertfordshire. His email address is [email protected].
Emille E. O. Ishida is a CNRS research engineer at Université Clermont Auvergne. Her email address is [email protected].
Alberto Krone-Martins is a lecturer in the Department of Informatics at the University of California, Irvine. His email address is [email protected].
Article DOI: 10.1090/noti3267
Credits
Figure 1 is courtesy of NASA/ESA Hubble Space Telescope.
Figure 2 is public domain.
Figures 3–6 are courtesy of the authors.
Photo of Rafael S. de Souza is courtesy of Rafael S. de Souza.
Photo of Emille E. O. Ishida is courtesy of Natan Chikaoka.
Photo of Alberto Krone-Martins is courtesy of Alberto Krone-Martins.
.png)
