Sunday Bayes: A brief history of Bayesian stats

The following discussion is essentially nontechnical; the aim is only to convey a little introductory “feel” for our outlook, purpose, and terminology, and to alert newcomers to common pitfalls of understanding.

Sometimes, in our perplexity, it has seemed to us that there are two basically different kinds of mentality in statistics; those who see the point of Bayesian inference at once, and need no explanation; and those who never see it, however much explanation is given.

–Jaynes, 1986 (pdf link)

Sunday Bayes

The format of this series is short and simple: Every week I will give a quick summary of a paper while sharing a few excerpts that I like. If you’ve read our eight easy steps paper and you’d like to follow along on this extension, I think a pace of one paper per week is a perfect way to ease yourself into the Bayesian sphere.

Bayesian Methods: General Background

The necessity of reasoning as best we can in situations where our information is incomplete is faced by all of us, every waking hour of our lives. (p. 2)

In order to understand Bayesian methods, I think it is essential to have some basic knowledge of their history. This paper by Jaynes (pdf) is an excellent place to start.

[Herodotus] notes that a decision was wise, even though it led to disastrous consequences, if the evidence at hand indicated it as the best one to make; and that a decision was foolish, even though it led to the happiest possible consequences, if it was unreasonable to expect those consequences. (p. 2)

Jaynes traces the history of Bayesian reasoning all the way back to Herodotus in 500BC. Herodotus could hardly be called a Bayesian, but the above quote captures the essence of Bayesian decision theory: take the action that maximizes your expected gain. It may turn out to be the wrong choice in the end, but if your reasoning that leads to your choice is sound then you took the correct course.

After all, our goal is not omniscience, but only to reason as best we can with whatever incomplete information we have. To demand more than this is to demand the impossible; neither Bernoulli’s procedure nor any other that might be put in its place can get something for nothing. (p. 3)

Much of the foundation for Bayesian inference was actually laid down by James Bernoulli, in his work Ars Conjectandi (“the art of conjecture”) in 1713. Bernoulli was the first to really invent a rational way of specifying a state of incomplete information. He put forth the idea that one can enumerate all “equally possible” cases N, and then count the number of cases for which some event A can occur. Then the probability of A, call it p(A), is just M/N, or the number of cases on which A can occur (M) to the total number of cases (N).

Jaynes gives only a passing mention to Bayes, noting his work “had little if any direct influence on the later development of probability theory” (p. 5). Laplace, Jeffreys, Cox, and Shannon all get a thorough discussion, and there is a lot of interesting material in those sections.

Despite the name, Bayes’ theorem was really formulated by Laplace. By all accounts, we should all be Laplacians right now.

The basic theorem appears today as almost trivially simple; yet it is by far the most important principle underlying scientific inference. (p. 5)

Laplace used Bayes’ theorem to estimate the mass of Saturn, and, by the best estimates when Jaynes was writing, his estimate was correct within .63%. That is very impressive for work done in the 18th century!

This strange history is only one of the reasons why, today [speaking in 1984], we Bayesians need to take the greatest pains to explain our rationale, as I am trying to do here. It is not that it is technically complicated; it is the way we have all been thinking intuitively from childhood. It is just so different from what we were all taught in formal courses on “orthodox” probability theory, which paralyze the mind into an inability to see a distinction between probability and frequency. Students who come to us free of that impediment have no difficulty in understanding our rationale, and are incredulous to anyone that could fail to understand it. (p. 7)

The sections on Laplace, Jeffreys, Cox and Shannon are all very good, but I will skip most of them because I think the most interesting and illuminating section of this paper is “Communication Difficulties” beginning on page 10.

Our background remarks would be incomplete without taking note of a serious disease that has afflicted probability theory for 200 years. There is a long history of confusion and controversy, leading in some cases to a paralytic inability to communicate. (p.10)

Jaynes is concerned in this section with the communication difficulties that Bayesians and frequentists have historically encountered.

[Since the 1930s] there has been a puzzling communication block that has prevented orthodoxians [frequentists] from comprehending Bayesian methods, and Bayesians from comprehending orthodox criticisms of our methods. (p. 10)

On the topic of this disagreement, Jaynes gives a nice quote from L.J. Savage: “there has seldom been such complete disagreement and breakdown of communication since the tower of Babel.” I wrote about one kind of communication breakdown in last week’s Sunday Bayes entry.

So what is the disagreement that Jaynes believes underlies much of the conflict between Bayesians and frequentists?

For decades Bayesians have been accused of “supposing that an unknown parameter is a random variable”; and we have denied hundreds of times with increasing vehemence, that we are making any such assumption. (p. 11)

Jaynes believes the confusion can be made clear by rephrasing the criticism as George Barnard once did.

Barnard complained that Bayesian methods of parameter estimation, which present our conclusions in the form of a posterior distribution, are illogical; for “How could the distribution of a parameter possibly become known from data which were taken with only one value of the parameter actually present?” (p. 11)

Aha, this is a key reformulation! This really illuminates the confusions between frequentists and Bayesians. To show why I’ll give one long quote to finish this Sunday Bayes entry.

Orthodoxians trying to understand Bayesian methods have been caught in a semantic trap by their habitual use of the phrase “distribution of the parameter” when one should have said “distribution of the probability”. Bayesians had supposed this to be merely a figure of speech; i.e., that those who used it did so only out of force of habit, and really knew better. But now it seems that our critics  have been taking that phraseology quite literally all the time.

Therefore, let us belabor still another time what we had previously thought too obvious to mention. In Bayesian parameter estimation, both the prior and posterior distributions represent, not any measurable property of the parameter, but only our own state of knowledge about it. The width of the distribution is not intended to indicate the range of variability of the true values of the parameter, as Barnards terminology had led him to suppose. It indicates the range of values that are consistent with our prior information and data, and which honesty therefore compels us to admit as possible values. What is “distributed” is not the parameter, but the probability. [emphasis added]

Now it appears that, for all these years, those who have seemed immune to all Bayesian explanation have just misunderstood our purpose. All this time, we had thought it clear from our subject-matter context that we are trying to estimate the value that the parameter had at the time the data were taken. [emphasis original] Put more generally, we are trying to draw inferences about what actually did happen in the experiment; not about the things that might have happened but did not. (p. 11)

I think if you really read the section on communication difficulties closely, then you will see that a lot of the conflict between Bayesians and frequentists can be boiled down to deep semantic confusion. We are often just talking past one another, getting ever more frustrated that the other side doesn’t understand our very simple points. Once this is sorted out I think a lot of the problems frequentists see with Bayesian methods will go away.

Sunday Bayes: Optional stopping is no problem for Bayesians

Optional stopping does not affect the interpretation of posterior odds. Even with optional stopping, a researcher can interpret the posterior odds as updated beliefs about hypotheses in light of data.

–Rouder, 2014 (pdf link)

Sunday Bayes

The format of this series is short and simple: Every week I will give a quick summary of a paper while sharing a few excerpts that I like. If you’ve read our eight easy steps paper and you’d like to follow along on this extension, I think a pace of one paper per week is a perfect way to ease yourself into the Bayesian sphere.

Optional stopping: No problem for Bayesians

Bayesian analysts use probability to express a degree of belief. For a flipped coin, a probability of 3/4 means that the analyst believes it is three times more likely that the coin will land heads than tails. Such a conceptualization is very convenient in science, where researchers hold beliefs about the plausibility of theories, hypotheses, and models that may be updated as new data become available. (p. 302)

It is becoming increasingly common to evaluate statistical procedures by way of simulation. Instead of doing formal analyses, we can use flexible simulations to tune many different parameters and immediately see the effect it has on the behavior of a procedure.

Simulation results have a tangible, experimental feel; moreover, if something is true mathematically, we should be able to see it in simulation as well. (p. 303)

But this brings with it a danger that the simulations performed might be doing the wrong thing, and unless we have a good grasp of the theoretical background of what is being simulated we can easily be misled. In this paper, Rouder (pdf) shows that common intuitions we have for evaluating simulations of frequentist statistics often do not translate to simulations of Bayesian statistics.

The critical element addressed here is whether optional stopping is problematic for Bayesians. My  argument is that both sets of authors use the wrong criteria or lens to draw their conclusions. They evaluate and interpret Bayesian statistics as if they were frequentist statistics. The more germane question is whether Bayesian statistics are interpretable as Bayesian statistics even if data are collected under optional stopping. (p. 302)

When we evaluate a frequentist procedure via simulation, it is common to set a parameter to a certain value and evaluate the number of times certain outcomes occur. For example, we can set the difference between two group means to zero, simulate a bunch of p values, and see how many fall below .05. Then we can set the difference to some nonzero number, simulate a bunch of p values, and again see how many are below .05. The first gives you the type-1 error rate for the procedure, and the second gives you the statistical power. This is appropriate for frequentist procedures because the probabilities calculated are always conditional on one or the other hypothesis being true.

One might be tempted to evaluate Bayes factors in the same way; that is, set the difference between two groups to zero and see how many BFs are above some threshold, and then set the difference to something nonzero and see how many BFs are again above some threshold.

The critical error … is studying Bayesian updating conditional on some hypothetical truth rather than conditional on data. This error is easy to make because it is what we have been taught and grown familiar with in our frequentist training. (p. 308)

Evaluating simulations of Bayes factors in this way is incorrect. Bayes factors (and posterior odds) are conditional on only the data observed. In other words, the appropriate evaluation is: “Given that I have observed this data (i.e., BF = x), what is the probability the BF was generated by H1 vs H0?”

Rouder visualizes this as follows. Flip a coin to choose the true hypothesis, then simulate a Bayes factor, and repeat these two steps many many times. At the end of the simulation, whenever BF=x is observed, check and see how many of these came from one model vs the other. The simulation shows that in this scenario if we look at all the times BF=3 is observed, there will be 3 BFs from the true model to every 1 BF from the false model. Since the prior odds are 1 to 1, the posterior odds equals the Bayes factor.

Screenshot 2016-03-06 12.33.59

You can see in the figure above (taken from Rouder’s figure 2), the distribution of Bayes factors observed when the null is true (purple, projected downwards) vs when the alternative is true (pink, projected upwards). Remember, the true hypothesis was chosen by coin flip. You can clearly see that when a BF of 3 to 1 in favor of the null is observed, the purple column is three times bigger than the pink column (shown with the arrows).

Below (taken from Rouder’s figure 2) you see what happens when one employs optional stopping (e.g., flip a coin to pick underlying true model, then sample until BF favors one model to another by at least 10 or you reach a maximum n). The distribution of Bayes factors generated by each model becomes highly skewed, which is often taken as evidence that conclusions drawn from Bayes factors depend on the stopping rule. The incorrect interpretation would be: Given the null is true, the number of times I find BF=x in favor of the alternative (i.e., in favor of the wrong model) has gone up, therefore the BF is sensitive to optional stopping. This is incorrect because it conditions on one model being true and checks the number of times a BF is observed, rather than conditioning on the observed BF and checking how often it came from H0 vs. H1.

Look again at what matters: What is the ratio of observed BFs that come from H1 vs. H0 for a given BF? No matter what stopping rule is used, the answer is always the same: If the true hypothesis is chosen by a coin flip, and a BF of 10 in favor of the alternative is observed, there will be 10 times as many observed BFs in the alternative column (pink) than in the null column (purple).

Screenshot 2016-03-06 12.40.31

In Rouder’s simulations he always used prior odds of 1 to 1, because then the posterior odds equal the Bayes factor. If one were to change the prior odds then the Bayes factor would no longer equal the posterior odds, and the shape of the distribution would again change; but importantly, while the absolute number of Bayes factors that end up in each bin would change, but the ratios of each pink column to purple column would not. No matter what stopping rule you use, the conclusions we draw from Bayes factors and posterior odds are unaffected by the stopping rule.

Feel free to employ any stopping rule you wish.

This result was recently shown again by Deng, Lu, and Chen in a paper posted to arXiv (pdf link) using similar simulations, and they go further in that they prove the theorem.

A few choice quotes

Page 308:

Optional-stopping protocols may be hybrids where sampling occurs until the Bayes factor reaches a certain level or a certain number of samples is reached. Such an approach strikes me as justifiable and reasonable, perhaps with the caveat that such protocols be made explicit before data collection. The benefit of this approach is that more resources may be devoted to more ambiguous experiments than to clear ones.

Page 308:

The critical error … is studying Bayesian updating conditional on some hypothetical truth rather than conditional on data. This error is easy to make because it iswhat we have been taught and grown familiar with in our frequentist training. In my opinion, the key to understanding Bayesian analysis is to focus on the degree of belief for considered models, which need not and should not be calibrated relative to some hypothetical truth.

Page 306-307:

When we update relative beliefs about two models, we make an implicit assumption that they are worthy of our consideration. Under this assumption, the beliefs may be updated regardless of the stopping rule. In this case, the models are dramatically wrong, so much so that the posterior odds contain no useful information whatsoever. Perhaps the more important insight is not that optional stopping is undesirable, but that the meaningfulness of posterior odds is a function of the usefulness of the models being compared.

Sunday Bayes: Testing precise hypotheses

First and foremost, when testing precise hypotheses, formal use of P-values should be abandoned. Almost anything will give a better indication of the evidence provided by the data against Ho.

–Berger & Delampady, 1987 (pdf link)

Sunday Bayes series intro:

After the great response to the eight easy steps paper we posted, I started a recurring series, where each week I highlight one of the papers that we included in the appendix of the paper. The format is short and simple: I will give a quick summary of the paper while sharing a few excerpts that I like. If you’ve read our eight easy steps paper and you’d like to follow along on this extension, I think a pace of one paper per week is a perfect way to ease yourself into the Bayesian sphere. At the end of the post I will list a few suggestions for the next entry, so vote in the comments or on twitter (@alxetz) for which one you’d like next. This paper was voted to be the next in the series.

(I changed the series name to Sunday Bayes, since I’ll be posting these on every Sunday.)

Testing precise hypotheses

This would indicate that say, claiming that a P-value of .05 is significant evidence against a precise hypothesis is sheer folly; the actual Bayes factor may well be near 1, and the posterior probability of Ho near 1/2 (p. 326)

Berger and Delampady (pdf link) review the background and standard practice for testing point null hypotheses (i.e., “precise hypotheses”). The paper came out nearly 30 years ago, so some parts of the discussion may not be as relevant these days, but it’s still a good paper.

They start by reviewing the basic measures of evidence — p-values, Bayes factors, posterior probabilities — before turning to an example. Rereading it, I remember why we gave this paper one of the highest difficulty ratings in the eight steps paper. There is a lot of technical discussion in this paper, but luckily I think most of the technical bits can be skipped in lieu of reading their commentary.

One of the main points of this paper is to investigate precisely when it is appropriate to approximate a small interval null hypothesis by using a point null hypothesis. They conclude, that most of the time, the error of approximation for Bayes factors will be small (<10%),

these numbers suggest that the point null approximation to Ho will be reasonable so long as [the width of the null interval] is one-half a [standard error] in width or smaller. (p. 322)

A secondary point of this paper is to refute the claim that classical answers will typically agree with some “objective” Bayesian analyses. Their conclusion is that such a claim

is simply not the case in the testing of precise hypotheses. This is indicated in Table 1 where, for instance, P(Ho | x) [NB: the posterior probability of the null] is from 5 to 50 times larger than the P-value. (p. 318)

They also review some lower bounds on the amount of Bayesian evidence that corresponds to significant p-values. They sum up their results thusly,

The message is simple: common interpretation of P-values, in terms of evidence against precise [null] hypotheses, are faulty (p. 323)

and

the weighted likelihood of H1 is at most [2.5] times that of Ho. A likelihood ratio [NB: Bayes factor] of [2.5] is not particularly strong evidence, particularly when it is [an upper] bound. However, it is customary in practice to view [p] = .05 as strong evidence against Ho. A P-value of [p] = .01, often considered very strong evidence against Ho, corresponds to [BF] = .1227, indicating that H1 is at most 8 times as likely as Ho. The message is simple: common interpretation of P-values, in terms of evidence against precise [null] hypotheses, are faulty (p. 323)

A few choice quotes

Page 319:

[A common opinion is that if] θ0 [NB: a point null] is not in [a confidence interval] it can be rejected, and looking at the set will provide a good indication as to the actual magnitude of the difference between θ and θ0. This opinion is wrong, because it ignores the supposed special nature of θo. A point can be outside a 95% confidence set, yet not be so strongly contraindicated by the data. Only by calculating a Bayes factor … can one judge how well the data supports a distinguished point θ0.

Page 327:

Of course, every statistician must judge for himself or herself how often precise hypotheses actually occur in practice. At the very least, however, we would argue that all types of tests should be able to be properly analyzed by statistics

Page 327 (emphasis original, since that text is a subheading):

[It is commonly argued that] The P-Value Is Just a Data Summary, Which We Can Learn To Properly Calibrate … One can argue that, through experience, one can learn how to interpret P-values. … But if the interpretation depends on Ho, the sample size, the density and the stopping rule, all in crucial ways, it becomes ridiculous to argue that we can intuitively learn to properly calibrate P-values.

page 328:

we would urge reporting both the Bayes factor, B, against [H0] and a confidence or credible region, C. The Bayes factor communicates the evidence in the data against [H0], and C indicates the magnitude of the possible discrepancy.

Page 328:

Without explicit alternatives, however, no Bayes factor or posterior probability could be calculated. Thus, the argument goes, one has no recourse but to use the P-value. A number of Bayesian responses to this argument have been raised … here we concentrate on responding in terms of the discussion in this paper. If, indeed, it is the case that P-values for precise hypotheses essentially always drastically overstate the actual evidence against Ho when the alternatives are known, how can one argue that no problem exists when the alternatives are not known?


Vote for the next entry:

  1. Edwards, Lindman, and Savage (1963) — Bayesian Statistical Inference for Psychological Research (pdf)
  2. Rouder (2014) — Optional Stopping: No Problem for Bayesians (pdf)
  3. Gallistel (2009) — The Importance of Proving the Null (pdf)
  4. Lindley (2000) — The philosophy of statistics (pdf)

The next steps: Jerome Cornfield and sequential analysis

This is equivalent to saying that if the application of a principle to given evidence leads to an absurdity then the evidence must be discarded. It is reminiscent of the heavy smoker, who, worried by the literature relating smoking to lung cancer, decided to give up reading.

— Cornfield, 1966 (pdf link)

The next steps series intro:

After the great response to the eight easy steps paper we posted, I have decided to start a recurring series, where each week I highlight one of the papers that we included in the appendix of the paper. The format will be short and simple: I will give a quick summary of the paper while sharing a few excerpts that I like. If you’ve read our eight easy steps paper and you’d like to follow along on this extension, I think a pace of one paper per week is a perfect way to ease yourself into the Bayesian sphere. At the end of the post I will list a few suggestions for the next entry, so vote in the comments or on twitter (@alxetz) for which one you’d like next.


Sequential trials, sequential analysis and the likelihood principle

Theoretical focus, low difficulty

Cornfield (1966) begins by posing a question:

Do the conclusions to be drawn from any set of data depend only on the data or do they depend also on the stopping rule which led to the data? (p. 18)

The purpose of his paper is to discuss this question and explore the implications of answering “yes” versus “no.” This paper is a natural followup to entries one and three in the eight easy steps paper.

If you have read the eight easy steps paper (or at least the first and third steps), you’ll know that the answer to the above question for classical statistics is “yes”, while the answer for Bayesian statistics is “no.”

Cornfield introduces a concepts he calls the “α-postulate,” which states,

All hypotheses rejected at the same critical level [i.e., p<.05] have equal amounts of evidence against them. (p. 19)

Through a series of examples, Cornfield shows that the α-postulate appears to be false.

Cornfield then introduces a concept called the likelihood principle, which comes up in a few of the eight easy steps entries. The likelihood principle says that the likelihood function contains all of the information relevant to the evaluation of statistical evidence. Other facets of the data that do not factor into the likelihood function are irrelevant to the evaluation of the strength of the statistical evidence.

He goes on to show how subscription to the likelihood principle minimizes a linear combination of type-I (α) and type-II (β) error rates, as opposed to the Neyman-Pearson procedure that minimizes type-II error rates (i.e., maximizes power) for a fixed type-I error rate (usually 5%).

Thus, if instead of minimizing β for a given α, we minimize [their linear combination], we must come to the same conclusion for all sample points which have the same likelihood function, no matter what the design. (p. 21)


A few choice quotes

page 19 (emphasis added):

The following example will be recognized by statisticians with consulting experience as a simplified version of a very common situation. An experimenter, having made n observations in the expectation that they would permit the rejection of a particular hypothesis, at some predesignated significance level, say .05, finds that he has not quite attained this critical level. He still believes that the hypothesis is false and asks how many more observations would be required to have reasonable certainty of rejecting the hypothesis if the means observed after n observations are taken as the true values. He also makes it clear that had the original n observations permitted rejection he would simply have published his findings. Under these circumstances it is evident that there is no amount of additional observation, no matter how large, which would permit rejection at the .05 level. If the hypothesis being tested is true, there is a .05 chance of its having been rejected after the first round of observations. To this chance must be added the probability of rejecting after the second round, given failure to reject after the first, and this increases the total chance of erroneous rejection to above .05. In fact … no amount of additional evidence can be collected which would provide evidence against the hypothesis equivalent to rejection at the P =.05 level

page 19-20 (emphasis added):

I realize, of course, that practical people tend to become impatient with counter-examples of this type. Quite properly they regard principles as only approximate guides to practice, and not as prescriptions that must be literally followed even when they lead to absurdities. But if one is unwilling to be guided by the α-postulate in the examples given, why should he be any more willing to accept it when analyzing sequential trials? The biostatistician’s responsibility for providing biomedical scientists with a satisfactory explication of inference cannot, in my opinion, be satisfied by applying certain principles when he agrees with their consequences and by disregarding them when he doesn’t.

page 22 (emphasis added):

The stopping rule is this: continue observations until a normal mean differs from the hypothesized value by k standard errors, at which point stop. It is certain, using the rule, that one will eventually differ from the hypothesized value by at least k standard errors even when the hypothesis is true. … The Bayesian viewpoint of the example is as follows. If one is seriously concerned about the probability that a stopping rule will certainly result in the rejection of a true hypothesis, it must be because some possibility of the truth of the hypothesis is being entertained. In that case it is appropriate to assign a non-zero prior probability to the hypothesis. If this is done, differing from the hypothesized value by k standard errors will not result in the same posterior probability for the hypothesis for all values of n. In fact for fixed k the posterior probability of the hypothesis monotonically approaches unity as n increases, no matter how small the prior probability assigned, so long as it is non-zero, and how large the k, so long as it is finite. Differing by k standard errors does not therefore necessarily provide any evidence against the hypothesis and disregarding the stopping rule does not lead to an absurd conclusion. The Bayesian viewpoint thus indicates that the hypothesis is certain to be erroneously rejected-not because the stopping rule was disregarded-but because the hypothesis was assigned zero prior probability and that such assignment is inconsistent with concern over the possibility that the hypothesis will certainly be rejected when true.


Vote for the next entry:

  1. Edwards, Lindman, and Savage (1963) — Bayesian Statistical Inference for Psychological Research (pdf)
  2. Rouder (2014) — Optional Stopping: No Problem for Bayesians (pdf)
  3. Gallistel (2009) — The Importance of Proving the Null (pdf)
  4. Berger and Delampady (1987) — Testing Precise Hypotheses (pdf)

 

Understanding Bayes: How to become a Bayesian in eight easy steps

How to become a Bayesian in eight easy steps: An annotated reading list

(TLDR: We wrote an annotated reading list to get you started in learning Bayesian statistics. Published version. Researchgate. PsyArxiv.)

It can be hard to know where to start when you want to learn about Bayesian statistics. I am frequently asked to share my favorite introductory resources to Bayesian statistics, and my go-to answer has been to share a dropbox folder with a bunch of PDFs that aren’t really sorted or cohesive. In some sense I was acting as little more than a glorified Google Scholar search bar.

It seems like there is some tension out there with regard to Bayes, in that many people want to know more about it, but when they pick up, say, Andrew Gelman and colleagues’ Bayesian Data Analysis they get totally overwhelmed. And then they just think, “Screw this esoteric B.S.” and give up because it doesn’t seem like it is worth their time or effort.

I think this happens a lot. Introductory Bayesian texts usually assume a level of training in mathematical statistics that most researchers simply don’t have time (or otherwise don’t need) to learn. There are actually a lot of accessible Bayesian resources out there that don’t require much math stat background at all, but it just so happens that they are not consolidated anywhere so people don’t necessarily know about them.

Enter the eight step program

Beth Baribault, Peter Edelsbrunner (@peter1328), Fabian Dablander (@fdabl), Quentin Gronau, and I have just finished a new paper that tries to remedy this situation, titled, “How to become a Bayesian in eight easy steps: An annotated reading list.” We were invited to submit this paper for a special issue on Bayesian statistics for Psychonomic Bulletin and Review. Each paper in the special issue addresses a specific question we often hear about Bayesian statistics, and ours was the following:

I am a reviewer/editor handling a manuscript that uses Bayesian methods; which articles should I read to get a quick idea of what that means?

So the paper‘s goal is not so much to teach readers how to actually perform Bayesian data analysis — there are other papers in the special issue for that — but to facilitate readers in their quest to understand basic Bayesian concepts. We think it will serve as a nice introductory reading list for any interested researcher.

The format of the paper is straightforward. We highlight eight papers that had a big impact on our own understanding of Bayesian statistics, as well as short descriptions of an additional 28 resources in the Further reading appendix. The first four papers are focused on theoretical introductions, and the second four have a slightly more applied focus.

We also give every resource a ranking from 1–9 on two dimensions: Focus (theoretical vs. applied) and Difficulty (easy vs. hard). We tried to provide a wide range of resources, from easy applications (#14: Wagenmakers, Lee, and Morey’s “Bayesian benefits for the pragmatic researcher”) to challenging theoretical discussions (#12: Edwards, Lindman and Savage’s “Bayesian statistical inference for psychological research”) and others in between.

The figure below (Figure A1, available on the last page of the paper) summarizes our rankings:

Readinglist.png

The emboldened numbers (1–8) are the papers that we’ve commented on in detail, numbers in light text (9–30) are papers we briefly describe in the appendix, and the italicized numbers (31–36) are our recommended introductory books (also listed in the appendix).

This is how we chose to frame the paper,

Overall, the guide is designed such that a researcher might be able to read all eight of the highlighted articles and some supplemental readings within a few days. After readers acquaint themselves with these sources, they should be well-equipped both to interpret existing research and to evaluate new research that relies on Bayesian methods.

The list

Here’s the list of papers we chose to cover in detail:

  1.  Lindley (1993): The analysis of experimental data: The appreciation of tea and wine. PDF.
  2. Kruschke (2015, chapter 2): Introduction: Credibility, models, and parameters. Available on the DBDA website.
  3. Dienes (2011): Bayesian versus orthodox statistics: Which side are you on? PDF.
  4. Rouder, Speckman, Sun, Morey, & Iverson (2009): Bayesian t tests for accepting and rejecting the null hypothesis. PDF.
  5. Vandekerckhove, Matzke, & Wagenmakers (2014): Model comparison and the principle of parsimony. PDF.
  6. van de Schoot, Kaplan, Denissen, Asendorpf, Neyer, & Aken (2014): A gentle introduction to Bayesian analysis: Applications to developmental research. PDF.
  7. Lee and Vanpaemel (from the same special issue): Determining priors for cognitive models. PDF.
  8. Lee (2008): Three case studies in the Bayesian analysis of cognitive models. PDF.

You’ll have to check out the paper to see our commentary and to find out what other articles we included in the Further reading appendix. We provide urls (web archived when possible; archive.org/web/) to PDFs of the eight main papers (except #2, that’s on the DBDA website), and wherever possible for the rest of the resources (some did not have free copies online; see the References).

I thought this was a fun paper to write, and if you think you might want to learn some Bayesian basics I hope you will consider reading it.

Oh, and I should mention that we wrote the whole paper collaboratively on Overleaf.com. It is a great site that makes it easy to get started using LaTeX, and I highly recommend trying it out.

This is the fifth post in the Understanding Bayes series. Until next time,

bill-murray-youre-awesome