In null-hypothesis significance testing, the -value[note 1] is the probability of obtaining test results at least as extreme as the result actually observed, under the assumption that the null hypothesis is correct.[2][3] A very small p-value means that such an extreme observed outcome would be very unlikely under the null hypothesis. Even though reporting p-values of statistical tests is common practice in academic publications of many quantitative fields, misinterpretation and misuse of p-values is widespread and has been a major topic in mathematics and metascience.[4][5] In 2016, the American Statistical Association (ASA) made a formal statement that "p-values do not measure the probability that the studied hypothesis is true, or the probability that the data were produced by random chance alone" and that "a p-value, or statistical significance, does not measure the size of an effect or the importance of a result" or "evidence regarding a model or hypothesis."[6] That said, a 2019 task force by ASA has issued a statement on statistical significance and replicability, concluding with: "p-values and significance tests, when properly applied and interpreted, increase the rigor of the conclusions drawn from data."[7]

Basic concepts edit

In statistics, every conjecture concerning the unknown probability distribution of a collection of random variables representing the observed data   in some study is called a statistical hypothesis. If we state one hypothesis only and the aim of the statistical test is to see whether this hypothesis is tenable, but not to investigate other specific hypotheses, then such a test is called a null hypothesis test.

As our statistical hypothesis will, by definition, state some property of the distribution, the null hypothesis is the default hypothesis under which that property does not exist. The null hypothesis is typically that some parameter (such as a correlation or a difference between means) in the populations of interest is zero. Our hypothesis might specify the probability distribution of   precisely, or it might only specify that it belongs to some class of distributions. Often, we reduce the data to a single numerical statistic, e.g.,  , whose marginal probability distribution is closely connected to a main question of interest in the study.

The p-value is used in the context of null hypothesis testing in order to quantify the statistical significance of a result, the result being the observed value of the chosen statistic  .[note 2] The lower the p-value is, the lower the probability of getting that result if the null hypothesis were true. A result is said to be statistically significant if it allows us to reject the null hypothesis. All other things being equal, smaller p-values are taken as stronger evidence against the null hypothesis.

Loosely speaking, rejection of the null hypothesis implies that there is sufficient evidence against it.

As a particular example, if a null hypothesis states that a certain summary statistic   follows the standard normal distribution N(0,1), then the rejection of this null hypothesis could mean that (i) the mean of   is not 0, or (ii) the variance of   is not 1, or (iii)   is not normally distributed. Different tests of the same null hypothesis would be more or less sensitive to different alternatives. However, even if we do manage to reject the null hypothesis for all 3 alternatives, and even if we know the distribution is normal and variance is 1, the null hypothesis test does not tell us which non-zero values of the mean are now most plausible. The more independent observations from the same probability distribution one has, the more accurate the test will be, and the higher the precision with which one will be able to determine the mean value and show that it is not equal to zero; but this will also increase the importance of evaluating the real-world or scientific relevance of this deviation.

Definition and interpretation edit

Definition edit

The p-value is the probability under the null hypothesis of obtaining a real-valued test statistic at least as extreme as the one obtained. Consider an observed test-statistic   from unknown distribution  . Then the p-value   is what the prior probability would be of observing a test-statistic value at least as "extreme" as   if null hypothesis   were true. That is:

  •   for a one-sided right-tail test-statistic distribution,
  •   for a one-sided left-tail test-statistic distribution,
  •   for a two-sided test-statistic distribution. If the distribution of   is symmetric about zero, then  

Interpretations edit

The error that a practising statistician would consider the more important to avoid (which is a subjective judgment) is called the error of the first kind. The first demand of the mathematical theory is to deduce such test criteria as would ensure that the probability of committing an error of the first kind would equal (or approximately equal, or not exceed) a preassigned number α, such as α = 0.05 or 0.01, etc. This number is called the level of significance.

— Jerzy Neyman, "The Emergence of Mathematical Statistics"[8]

In a significance test, the null hypothesis   is rejected if the p-value is less than or equal to a predefined threshold value  , which is referred to as the alpha level or significance level.   is not derived from the data, but rather is set by the researcher before examining the data.   is commonly set to 0.05, though lower alpha levels are sometimes used. In 2018, a group of statisticians led by Daniel Benjamin proposed the adoption of the 0.005 value as standard value for statistical significance worldwide.[9]

Different p-values based on independent sets of data can be combined, for instance using Fisher's combined probability test.

Distribution edit

The p-value is a function of the chosen test statistic   and is therefore a random variable. If the null hypothesis fixes the probability distribution of   precisely (e.g.   where   is the only parameter), and if that distribution is continuous, then when the null-hypothesis is true the p-value is uniformly distributed between 0 and 1. Regardless of the truth of the  , the p-value is not fixed; if the same test is repeated independently with fresh data one will typically obtain a different p-value in each iteration.

Usually only a single p-value relating to a hypothesis is observed, so the p-value is interpreted by a significance test and no effort is made to estimate the distribution it was drawn from. When a collection of p-values are available (e.g. when considering a group of studies on the same subject), the distribution of p-values is sometimes called a p-curve.[10] A p-curve can be used to assess the reliability of scientific literature, such as by detecting publication bias or p-hacking. [10][11]

Distribution for composite hypothesis edit

In parametric hypothesis testing problems, a simple or point hypothesis refers to a hypothesis where the parameter's value is assumed to be a single number. In contrast, in a composite hypothesis the parameter's value is given by a set of numbers. When the null-hypothesis is composite (or the distribution of the statistic is discrete), then when the null-hypothesis is true the probability of obtaining a p-value less than or equal to any number between 0 and 1 is still less than or equal to that number. In other words, it remains the case that very small values are relatively unlikely if the null-hypothesis is true, and that a significance test at level   is obtained by rejecting the null-hypothesis if the p-value is less than or equal to  .[12][13]

For example, when testing the null hypothesis that a distribution is normal with a mean less than or equal to zero against the alternative that the mean is greater than zero ( , variance known), the null hypothesis does not specify the exact probability distribution of the appropriate test statistic. In this example that would be the Z-statistic belonging to the one-sided one-sample Z-test. For each possible value of the theoretical mean, the Z-test statistic has a different probability distribution. In these circumstances the p-value is defined by taking the least favorable null-hypothesis case, which is typically on the border between null and alternative. This definition ensures the complementarity of p-values and alpha-levels:   means one only rejects the null hypothesis if the p-value is less than or equal to  , and the hypothesis test will indeed have a maximum type-1 error rate of  .

Usage edit

The p-value is widely used in statistical hypothesis testing, specifically in null hypothesis significance testing. In this method, before conducting the study, one first chooses a model (the null hypothesis) and the alpha level α (most commonly 0.05). After analyzing the data, if the p-value is less than α, that is taken to mean that the observed data is sufficiently inconsistent with the null hypothesis for the null hypothesis to be rejected. However, that does not prove that the null hypothesis is false. The p-value does not, in itself, establish probabilities of hypotheses. Rather, it is a tool for deciding whether to reject the null hypothesis.[14]

Misuse edit

According to the ASA, there is widespread agreement that p-values are often misused and misinterpreted.[3] One practice that has been particularly criticized is accepting the alternative hypothesis for any p-value nominally less than .05 without other supporting evidence. Although p-values are helpful in assessing how incompatible the data are with a specified statistical model, contextual factors must also be considered, such as "the design of a study, the quality of the measurements, the external evidence for the phenomenon under study, and the validity of assumptions that underlie the data analysis".[3] Another concern is that the p-value is often misunderstood as being the probability that the null hypothesis is true.[3][15]

Some statisticians have proposed abandoning p-values and focusing more on other inferential statistics,[3] such as confidence intervals,[16][17] likelihood ratios,[18][19] or Bayes factors,[20][21][22] but there is heated debate on the feasibility of these alternatives.[23][24] Others have suggested to remove fixed significance thresholds and to interpret p-values as continuous indices of the strength of evidence against the null hypothesis.[25][26] Yet others suggested to report alongside p-values the prior probability of a real effect that would be required to obtain a false positive risk (i.e. the probability that there is no real effect) below a pre-specified threshold (e.g. 5%).[27]

That said, in 2019 a task force by ASA had convened to consider the use of statistical methods in scientific studies, specifically hypothesis tests and p-values, and their connection to replicability.[7] It states that "Different measures of uncertainty can complement one another; no single measure serves all purposes.", citing p-value as one of these measures. They also stress that p-values can provide valuable information when considering the specific value as well as when compared to some threshold. In general, it stresses that "p-values and significance tests, when properly applied and interpreted, increase the rigor of the conclusions drawn from data."

Calculation edit

Usually,   is a test statistic. A test statistic is the output of a scalar function of all the observations. This statistic provides a single number, such as a t-statistic or an F-statistic. As such, the test statistic follows a distribution determined by the function used to define that test statistic and the distribution of the input observational data.

For the important case in which the data are hypothesized to be a random sample from a normal distribution, depending on the nature of the test statistic and the hypotheses of interest about its distribution, different null hypothesis tests have been developed. Some such tests are the z-test for hypotheses concerning the mean of a normal distribution with known variance, the t-test based on Student's t-distribution of a suitable statistic for hypotheses concerning the mean of a normal distribution when the variance is unknown, the F-test based on the F-distribution of yet another statistic for hypotheses concerning the variance. For data of other nature, for instance categorical (discrete) data, test statistics might be constructed whose null hypothesis distribution is based on normal approximations to appropriate statistics obtained by invoking the central limit theorem for large samples, as in the case of Pearson's chi-squared test.

Thus computing a p-value requires a null hypothesis, a test statistic (together with deciding whether the researcher is performing a one-tailed test or a two-tailed test), and data. Even though computing the test statistic on given data may be easy, computing the sampling distribution under the null hypothesis, and then computing its cumulative distribution function (CDF) is often a difficult problem. Today, this computation is done using statistical software, often via numeric methods (rather than exact formulae), but, in the early and mid 20th century, this was instead done via tables of values, and one interpolated or extrapolated p-values from these discrete values[citation needed]. Rather than using a table of p-values, Fisher instead inverted the CDF, publishing a list of values of the test statistic for given fixed p-values; this corresponds to computing the Quantile function (inverse CDF).

Example edit

Testing the fairness of a coin edit

As an example of a statistical test, an experiment is performed to determine whether a coin flip is fair (equal chance of landing heads or tails) or unfairly biased (one outcome being more likely than the other).

Suppose that the experimental results show the coin turning up heads 14 times out of 20 total flips. The full data   would be a sequence of twenty times the symbol "H" or "T." The statistic on which one might focus could be the total number   of heads. The null hypothesis is that the coin is fair, and coin tosses are independent of one another. If a right-tailed test is considered, which would be the case if one is actually interested in the possibility that the coin is biased towards falling heads, then the p-value of this result is the chance of a fair coin landing on heads at least 14 times out of 20 flips. That probability can be computed from binomial coefficients as

 

This probability is the p-value, considering only extreme results that favor heads. This is called a one-tailed test. However, one might be interested in deviations in either direction, favoring either heads or tails. The two-tailed p-value, which considers deviations favoring either heads or tails, may instead be calculated. As the binomial distribution is symmetrical for a fair coin, the two-sided p-value is simply twice the above calculated single-sided p-value: the two-sided p-value is 0.115.

In the above example:

  • Null hypothesis (H0): The coin is fair, with Pr(heads) = 0.5
  • Test statistic: Number of heads
  • Alpha level (designated threshold of significance): 0.05
  • Observation O: 14 heads out of 20 flips; and
  • Two-tailed p-value of observation O given H0 = 2 × min(Pr(no. of heads ≥ 14 heads), Pr(no. of heads ≤ 14 heads)) = 2 × min(0.058, 0.978) = 2*0.058 = 0.115.

The Pr (no. of heads ≤ 14 heads) = 1 - Pr(no. of heads ≥ 14 heads) + Pr (no. of head = 14) = 1 - 0.058 + 0.036 = 0.978; however, the symmetry of this binomial distribution makes it an unnecessary computation to find the smaller of the two probabilities. Here, the calculated p-value exceeds .05, meaning that the data falls within the range of what would happen 95% of the time, if the coin were fair. Hence, the null hypothesis is not rejected at the .05 level.

However, had one more head been obtained, the resulting p-value (two-tailed) would have been 0.0414 (4.14%), in which case the null hypothesis would be rejected at the 0.05 level.

Multistage experiment design edit

The difference between the two meanings of "extreme" appear when we consider a multistage experiment for testing the fairness of the coin. Suppose we design the experiment as follows:

  • Flip the coin twice. If both comes up heads or tails, end the experiment.
  • Else, flip the coin 4 more times.

This experiment has 7 types of outcomes: 2 heads, 2 tails, 5 heads 1 tail..., 1 head 5 tails. We now calculate the p-value of the "3 heads 3 tails" outcome .

If we use the test statistic  , then under the null hypothesis is exactly 1 for two-sided p-value, and exactly   for one-sided left-tail p-value, and same for one-sided right-tail p-value.

If we consider every outcome that has equal or lower probability than "3 heads 3 tails" as "at least as extreme", then the p-value is exactly  .

However, suppose we have planned to simply flip the coin 6 times no matter what happens, then the second definition of p-value would mean that the p-value of "3 heads 3 tails" is exactly 1.

Thus, the "at least as extreme" definition of p-value is deeply contextual, and depends on what the experimenter planned to do even in situations that did not occur.

History edit

 
John Arbuthnot
 
Pierre-Simon Laplace
 
Karl Pearson
 
Ronald Fisher

P-value computations date back to the 1700s, where they were computed for the human sex ratio at birth, and used to compute statistical significance compared to the null hypothesis of equal probability of male and female births.[28] John Arbuthnot studied this question in 1710,[29][30][31][32] and examined birth records in London for each of the 82 years from 1629 to 1710. In every year, the number of males born in London exceeded the number of females. Considering more male or more female births as equally likely, the probability of the observed outcome is 1/282, or about 1 in 4,836,000,000,000,000,000,000,000; in modern terms, the p-value. This is vanishingly small, leading Arbuthnot that this was not due to chance, but to divine providence: "From whence it follows, that it is Art, not Chance, that governs." In modern terms, he rejected the null hypothesis of equally likely male and female births at the p = 1/282 significance level. This and other work by Arbuthnot is credited as "… the first use of significance tests …"[33] the first example of reasoning about statistical significance,[34] and "… perhaps the first published report of a nonparametric test …",[30] specifically the sign test; see details at Sign test § History.

The same question was later addressed by Pierre-Simon Laplace, who instead used a parametric test, modeling the number of male births with a binomial distribution:[35]

In the 1770s Laplace considered the statistics of almost half a million births. The statistics showed an excess of boys compared to girls. He concluded by calculation of a p-value that the excess was a real, but unexplained, effect.

The p-value was first formally introduced by Karl Pearson, in his Pearson's chi-squared test,[36] using the chi-squared distribution and notated as capital P.[36] The p-values for the chi-squared distribution (for various values of χ2 and degrees of freedom), now notated as P, were calculated in (Elderton 1902), collected in (Pearson 1914, pp. xxxi–xxxiii, 26–28, Table XII).

The use of the p-value in statistics was popularized by Ronald Fisher,[37][full citation needed] and it plays a central role in his approach to the subject.[38] In his influential book Statistical Methods for Research Workers (1925), Fisher proposed the level p = 0.05, or a 1 in 20 chance of being exceeded by chance, as a limit for statistical significance, and applied this to a normal distribution (as a two-tailed test), thus yielding the rule of two standard deviations (on a normal distribution) for statistical significance (see 68–95–99.7 rule).[39][note 3][40]

He then computed a table of values, similar to Elderton but, importantly, reversed the roles of χ2 and p. That is, rather than computing p for different values of χ2 (and degrees of freedom n), he computed values of χ2 that yield specified p-values, specifically 0.99, 0.98, 0.95, 0,90, 0.80, 0.70, 0.50, 0.30, 0.20, 0.10, 0.05, 0.02, and 0.01.[41] That allowed computed values of χ2 to be compared against cutoffs and encouraged the use of p-values (especially 0.05, 0.02, and 0.01) as cutoffs, instead of computing and reporting p-values themselves. The same type of tables were then compiled in (Fisher & Yates 1938), which cemented the approach.[40]

As an illustration of the application of p-values to the design and interpretation of experiments, in his following book The Design of Experiments (1935), Fisher presented the lady tasting tea experiment,[42] which is the archetypal example of the p-value.

To evaluate a lady's claim that she (Muriel Bristol) could distinguish by taste how tea is prepared (first adding the milk to the cup, then the tea, or first tea, then milk), she was sequentially presented with 8 cups: 4 prepared one way, 4 prepared the other, and asked to determine the preparation of each cup (knowing that there were 4 of each). In that case, the null hypothesis was that she had no special ability, the test was Fisher's exact test, and the p-value was   so Fisher was willing to reject the null hypothesis (consider the outcome highly unlikely to be due to chance) if all were classified correctly. (In the actual experiment, Bristol correctly classified all 8 cups.)

Fisher reiterated the p = 0.05 threshold and explained its rationale, stating:[43]

It is usual and convenient for experimenters to take 5 per cent as a standard level of significance, in the sense that they are prepared to ignore all results which fail to reach this standard, and, by this means, to eliminate from further discussion the greater part of the fluctuations which chance causes have introduced into their experimental results.

He also applies this threshold to the design of experiments, noting that had only 6 cups been presented (3 of each), a perfect classification would have only yielded a p-value of   which would not have met this level of significance.[43] Fisher also underlined the interpretation of p, as the long-run proportion of values at least as extreme as the data, assuming the null hypothesis is true.

In later editions, Fisher explicitly contrasted the use of the p-value for statistical inference in science with the Neyman–Pearson method, which he terms "Acceptance Procedures".[44] Fisher emphasizes that while fixed levels such as 5%, 2%, and 1% are convenient, the exact p-value can be used, and the strength of evidence can and will be revised with further experimentation. In contrast, decision procedures require a clear-cut decision, yielding an irreversible action, and the procedure is based on costs of error, which, he argues, are inapplicable to scientific research.

Related indices edit

The E-value can refer to two concepts, both of which are related to the p-value and both of which play a role in multiple testing. First, it corresponds to a generic, more robust alternative to the p-value that can deal with optional continuation of experiments. Second, it is also used to abbreviate "expect value", which is the expected number of times that one expects to obtain a test statistic at least as extreme as the one that was actually observed if one assumes that the null hypothesis is true.[45] This expect-value is the product of the number of tests and the p-value.

The q-value is the analog of the p-value with respect to the positive false discovery rate.[46] It is used in multiple hypothesis testing to maintain statistical power while minimizing the false positive rate.[47]

The Probability of Direction (pd) is the Bayesian numerical equivalent of the p-value.[48] It corresponds to the proportion of the posterior distribution that is of the median's sign, typically varying between 50% and 100%, and representing the certainty with which an effect is positive or negative.

Second-generation p-values extend the concept of p-values by not considering extremely small, practically irrelevant effect sizes as significant.[49]

See also edit

Notes edit

  1. ^ Italicisation, capitalisation and hyphenation of the term vary. For example, AMA style uses "P value", APA style uses "p value", and the American Statistical Association uses "p-value". In all cases, the "p" stands for probability.[1]
  2. ^ The statistical significance of a result does not imply that the result also has real-world relevance. For instance, a medicine might have a statistically significant effect that is too small to be interesting.
  3. ^ To be more specific, the p = 0.05 corresponds to about 1.96 standard deviations for a normal distribution (two-tailed test), and 2 standard deviations corresponds to about a 1 in 22 chance of being exceeded by chance, or p ≈ 0.045; Fisher notes these approximations.

References edit

  1. ^ "ASA House Style" (PDF). Amstat News. American Statistical Association.
  2. ^ Aschwanden C (2015-11-24). "Not Even Scientists Can Easily Explain P-values". FiveThirtyEight. Archived from the original on 25 September 2019. Retrieved 11 October 2019.
  3. ^ a b c d e Wasserstein RL, Lazar NA (7 March 2016). "The ASA's Statement on p-Values: Context, Process, and Purpose". The American Statistician. 70 (2): 129–133. doi:10.1080/00031305.2016.1154108.
  4. ^ Hubbard R, Lindsay RM (2008). "Why P Values Are Not a Useful Measure of Evidence in Statistical Significance Testing". Theory & Psychology. 18 (1): 69–88. doi:10.1177/0959354307086923. S2CID 143487211.
  5. ^ Munafò MR, Nosek BA, Bishop DV, Button KS, Chambers CD, du Sert NP, et al. (January 2017). "A manifesto for reproducible science". Nature Human Behaviour. 1: 0021. doi:10.1038/s41562-016-0021. PMC 7610724. PMID 33954258. S2CID 6326747.
  6. ^ Wasserstein, Ronald L.; Lazar, Nicole A. (2016-04-02). "The ASA Statement on p -Values: Context, Process, and Purpose". The American Statistician. 70 (2): 129–133. doi:10.1080/00031305.2016.1154108. ISSN 0003-1305. S2CID 124084622.
  7. ^ a b Benjamini, Yoav; De Veaux, Richard D.; Efron, Bradley; Evans, Scott; Glickman, Mark; Graubard, Barry I.; He, Xuming; Meng, Xiao-Li; Reid, Nancy M.; Stigler, Stephen M.; Vardeman, Stephen B.; Wikle, Christopher K.; Wright, Tommy; Young, Linda J.; Kafadar, Karen (2021-10-02). "ASA President's Task Force Statement on Statistical Significance and Replicability". Chance. Informa UK Limited. 34 (4): 10–11. doi:10.1080/09332480.2021.2003631. ISSN 0933-2480.
  8. ^ Neyman, Jerzy (1976). "The Emergence of Mathematical Statistics: A Historical Sketch with Particular Reference to the United States". In Owen, D.B. (ed.). On the History of Statistics and Probability. Textbooks and Monographs. New York: Marcel Dekker Inc. p. 161.
  9. ^ Benjamin, Daniel J.; Berger, James O.; Johannesson, Magnus; Nosek, Brian A.; Wagenmakers, E.-J.; Berk, Richard; Bollen, Kenneth A.; Brembs, Björn; Brown, Lawrence; Camerer, Colin; Cesarini, David; Chambers, Christopher D.; Clyde, Merlise; Cook, Thomas D.; De Boeck, Paul; Dienes, Zoltan; Dreber, Anna; Easwaran, Kenny; Efferson, Charles; Fehr, Ernst; Fidler, Fiona; Field, Andy P.; Forster, Malcolm; George, Edward I.; Gonzalez, Richard; Goodman, Steven; Green, Edwin; Green, Donald P.; Greenwald, Anthony G.; Hadfield, Jarrod D.; Hedges, Larry V.; Held, Leonhard; Hua Ho, Teck; Hoijtink, Herbert; Hruschka, Daniel J.; Imai, Kosuke; Imbens, Guido; Ioannidis, John P. A.; Jeon, Minjeong; Jones, James Holland; Kirchler, Michael; Laibson, David; List, John; Little, Roderick; Lupia, Arthur; Machery, Edouard; Maxwell, Scott E.; McCarthy, Michael; Moore, Don A.; Morgan, Stephen L.; Munafó, Marcus; Nakagawa, Shinichi; Nyhan, Brendan; Parker, Timothy H.; Pericchi, Luis; Perugini, Marco; Rouder, Jeff; Rousseau, Judith; Savalei, Victoria; Schönbrodt, Felix D.; Sellke, Thomas; Sinclair, Betsy; Tingley, Dustin; Van Zandt, Trisha; Vazire, Simine; Watts, Duncan J.; Winship, Christopher; Wolpert, Robert L.; Xie, Yu; Young, Cristobal; Zinman, Jonathan; Johnson, Valen E. (1 September 2017). "Redefine statistical significance". Nature Human Behaviour. 2 (1): 6–10. doi:10.1038/s41562-017-0189-z. eISSN 2397-3374. hdl:10281/184094. PMID 30980045. S2CID 256726352.
  10. ^ a b Head ML, Holman L, Lanfear R, Kahn AT, Jennions MD (March 2015). "The extent and consequences of p-hacking in science". PLOS Biology. 13 (3): e1002106. doi:10.1371/journal.pbio.1002106. PMC 4359000. PMID 25768323.
  11. ^ Simonsohn U, Nelson LD, Simmons JP (November 2014). "p-Curve and Effect Size: Correcting for Publication Bias Using Only Significant Results". Perspectives on Psychological Science. 9 (6): 666–681. doi:10.1177/1745691614553988. PMID 26186117. S2CID 39975518.
  12. ^ Bhattacharya B, Habtzghi D (2002). "Median of the p value under the alternative hypothesis". The American Statistician. 56 (3): 202–6. doi:10.1198/000313002146. S2CID 33812107.
  13. ^ Hung HM, O'Neill RT, Bauer P, Köhne K (March 1997). "The behavior of the P-value when the alternative hypothesis is true". Biometrics (Submitted manuscript). 53 (1): 11–22. doi:10.2307/2533093. JSTOR 2533093. PMID 9147587.
  14. ^ Nuzzo R (February 2014). "Scientific method: statistical errors". Nature. 506 (7487): 150–152. Bibcode:2014Natur.506..150N. doi:10.1038/506150a. PMID 24522584.
  15. ^ Colquhoun D (November 2014). "An investigation of the false discovery rate and the misinterpretation of p-values". Royal Society Open Science. 1 (3): 140216. arXiv:1407.5296. Bibcode:2014RSOS....140216C. doi:10.1098/rsos.140216. PMC 4448847. PMID 26064558.
  16. ^ Lee DK (December 2016). "Alternatives to P value: confidence interval and effect size". Korean Journal of Anesthesiology. 69 (6): 555–562. doi:10.4097/kjae.2016.69.6.555. PMC 5133225. PMID 27924194.
  17. ^ Ranstam J (August 2012). "Why the P-value culture is bad and confidence intervals a better alternative". Osteoarthritis and Cartilage. 20 (8): 805–808. doi:10.1016/j.joca.2012.04.001. PMID 22503814.
  18. ^ Perneger TV (May 2001). "Sifting the evidence. Likelihood ratios are alternatives to P values". BMJ. 322 (7295): 1184–1185. doi:10.1136/bmj.322.7295.1184. PMC 1120301. PMID 11379590.
  19. ^ Royall R (2004). "The Likelihood Paradigm for Statistical Evidence". The Nature of Scientific Evidence. pp. 119–152. doi:10.7208/chicago/9780226789583.003.0005. ISBN 9780226789576.
  20. ^ Schimmack U (30 April 2015). "Replacing p-values with Bayes-Factors: A Miracle Cure for the Replicability Crisis in Psychological Science". Replicability-Index. Retrieved 7 March 2017.
  21. ^ Marden JI (December 2000). "Hypothesis Testing: From p Values to Bayes Factors". Journal of the American Statistical Association. 95 (452): 1316–1320. doi:10.2307/2669779. JSTOR 2669779.
  22. ^ Stern HS (16 February 2016). "A Test by Any Other Name: P Values, Bayes Factors, and Statistical Inference". Multivariate Behavioral Research. 51 (1): 23–29. doi:10.1080/00273171.2015.1099032. PMC 4809350. PMID 26881954.
  23. ^ Murtaugh PA (March 2014). "In defense of P values". Ecology. 95 (3): 611–617. Bibcode:2014Ecol...95..611M. doi:10.1890/13-0590.1. PMID 24804441.
  24. ^ Aschwanden C (7 March 2016). "Statisticians Found One Thing They Can Agree On: It's Time To Stop Misusing P-Values". FiveThirtyEight.
  25. ^ Amrhein V, Korner-Nievergelt F, Roth T (2017). "The earth is flat (p > 0.05): significance thresholds and the crisis of unreplicable research". PeerJ. 5: e3544. doi:10.7717/peerj.3544. PMC 5502092. PMID 28698825.
  26. ^ Amrhein V, Greenland S (January 2018). "Remove, rather than redefine, statistical significance". Nature Human Behaviour. 2 (1): 4. doi:10.1038/s41562-017-0224-0. PMID 30980046. S2CID 46814177.
  27. ^ Colquhoun D (December 2017). "The reproducibility of research and the misinterpretation of p-values". Royal Society Open Science. 4 (12): 171085. doi:10.1098/rsos.171085. PMC 5750014. PMID 29308247.
  28. ^ Brian E, Jaisson M (2007). "Physico-Theology and Mathematics (1710–1794)". The Descent of Human Sex Ratio at Birth. Springer Science & Business Media. pp. 1–25. ISBN 978-1-4020-6036-6.
  29. ^ Arbuthnot J (1710). "An argument for Divine Providence, taken from the constant regularity observed in the births of both sexes" (PDF). Philosophical Transactions of the Royal Society of London. 27 (325–336): 186–190. doi:10.1098/rstl.1710.0011. S2CID 186209819.
  30. ^ a b Conover WJ (1999). "Chapter 3.4: The Sign Test". Practical Nonparametric Statistics (Third ed.). Wiley. pp. 157–176. ISBN 978-0-471-16068-7.
  31. ^ Sprent P (1989). Applied Nonparametric Statistical Methods (Second ed.). Chapman & Hall. ISBN 978-0-412-44980-2.
  32. ^ Stigler SM (1986). The History of Statistics: The Measurement of Uncertainty Before 1900. Harvard University Press. pp. 225–226. ISBN 978-0-67440341-3.
  33. ^ Bellhouse P (2001). "John Arbuthnot". In Heyde CC, Seneta E (eds.). Statisticians of the Centuries. Springer. pp. 39–42. ISBN 978-0-387-95329-8.
  34. ^ Hald A (1998). "Chapter 4. Chance or Design: Tests of Significance". A History of Mathematical Statistics from 1750 to 1930. Wiley. p. 65.
  35. ^ Stigler SM (1986). The History of Statistics: The Measurement of Uncertainty Before 1900. Harvard University Press. p. 134. ISBN 978-0-67440341-3.
  36. ^ a b Pearson K (1900). "On the criterion that a given system of deviations from the probable in the case of a correlated system of variables is such that it can be reasonably supposed to have arisen from random sampling" (PDF). Philosophical Magazine. Series 5. 50 (302): 157–175. doi:10.1080/14786440009463897.
  37. ^ Inman 2004.
  38. ^ Hubbard R, Bayarri MJ (2003), "Confusion Over Measures of Evidence (p′s) Versus Errors (α′s) in Classical Statistical Testing", The American Statistician, 57 (3): 171–178 [p. 171], doi:10.1198/0003130031856, S2CID 55671953
  39. ^ Fisher 1925, p. 47, Chapter III. Distributions.
  40. ^ a b Dallal 2012, Note 31: Why P=0.05?.
  41. ^ Fisher 1925, pp. 78–79, 98, Chapter IV. Tests of Goodness of Fit, Independence and Homogeneity; with Table of χ2, Table III. Table of χ2.
  42. ^ Fisher 1971, II. The Principles of Experimentation, Illustrated by a Psycho-physical Experiment.
  43. ^ a b Fisher 1971, Section 7. The Test of Significance.
  44. ^ Fisher 1971, Section 12.1 Scientific Inference and Acceptance Procedures.
  45. ^ "Definition of E-value". National Institutes of Health.
  46. ^ Storey JD (2003). "The positive false discovery rate: a Bayesian interpretation and the q-value". The Annals of Statistics. 31 (6): 2013–2035. doi:10.1214/aos/1074290335.
  47. ^ Storey JD, Tibshirani R (August 2003). "Statistical significance for genomewide studies". Proceedings of the National Academy of Sciences of the United States of America. 100 (16): 9440–9445. Bibcode:2003PNAS..100.9440S. doi:10.1073/pnas.1530509100. PMC 170937. PMID 12883005.
  48. ^ Makowski D, Ben-Shachar MS, Chen SH, Lüdecke D (10 December 2019). "Indices of Effect Existence and Significance in the Bayesian Framework". Frontiers in Psychology. 10: 2767. doi:10.3389/fpsyg.2019.02767. PMC 6914840. PMID 31920819.
  49. ^ An Introduction to Second-Generation p-Values Jeffrey D. Blume,Robert A. Greevy,Valerie F. Welty,Jeffrey R. Smith &William D. Dupont https://www.tandfonline.com/doi/full/10.1080/00031305.2018.1537893

Further reading edit

External links edit