Top Banner
PROBABILITY THEORY THE LOGIC OF SCIENCE E. T. Jaynes edited by G. Larry Bretthorst
31

PROBABILITY THEORY THE LOGIC OF SCIENCE

May 12, 2023

Download

Documents

Khang Minh
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: PROBABILITY THEORY THE LOGIC OF SCIENCE

PROBABILITY THEORY

THE LOGIC OF SCIENCE

E . T . Jaynes

edited by G. Larry Bretthorst

Page 2: PROBABILITY THEORY THE LOGIC OF SCIENCE

published by the press syndicate of the university of cambridgeThe Pitt Building, Trumpington Street, Cambridge, United Kingdom

cambridge university pressThe Edinburgh Building, Cambridge CB2 2RU, UK

40 West 20th Street, New York, NY 10011-4211, USA477 Williamstown Road, Port Melbourne, VIC 3207, Australia

Ruiz de Alarcon 13, 28014 Madrid, SpainDock House, The Waterfront, Cape Town 8001, South Africa

http://www.cambridge.org

C© E. T. Jaynes 2003

This book is in copyright. Subject to statutory exceptionand to the provisions of relevant collective licensing agreements,

no reproduction of any part may take place withoutthe written permission of Cambridge University Press.

First published 2003

Printed in the United Kingdom at the University Press, Cambridge

TypefaceTimes 10/13 pt SystemLATEX 2ε [tb]

A catalogue record for this book is available from the British Library

Library of Congress Cataloguing in Publication data

Jaynes, E. T. (Edwin T.)Probability theory: the logic of science / by E.T. Jaynes; edited by G. Larry Bretthorst.

p. cm.Includes bibliographical references and index.

ISBN 0 521 59271 21. Probabilities. I. Bretthorst, G. Larry. II. Title.

QA273 .J36 2003519.2 – dc21 2002071486

ISBN 0 521 59271 2 hardback

Page 3: PROBABILITY THEORY THE LOGIC OF SCIENCE

Contents

Editor’s foreword pagexviiPreface xix

Part I Principles and elementary applications1 Plausible reasoning 3

1.1 Deductive and plausible reasoning 31.2 Analogies with physical theories 61.3 The thinking computer 71.4 Introducing the robot 81.5 Boolean algebra 91.6 Adequate setsof operations 121.7 The basic desiderata 171.8 Comments 19

1.8.1 Common language vs. formal logic 211.8.2 Nitpicking 23

2 The quantitative rules 242.1 The product rule 242.2 The sum rule 302.3 Qualitative properties 352.4 Numericalvalues 372.5 Notation and finite-sets policy 432.6 Comments 44

2.6.1 ‘Subjective’ vs. ‘objective’ 442.6.2 Godel’s theorem 452.6.3 Venn diagrams 472.6.4 The ‘Kolmogorov axioms’ 49

3 Elementary sampling theory 513.1 Sampling without replacement 523.2 Logic vs. propensity 603.3 Reasoning from less precise information 643.4 Expectations 663.5 Other forms and extensions 68

vii

Page 4: PROBABILITY THEORY THE LOGIC OF SCIENCE

viii Contents

3.6 Probability as a mathematical tool 683.7 The binomial distribution 693.8 Sampling with replacement 72

3.8.1 Digression: a sermon on reality vs. models 733.9 Correction for correlations 753.10 Simplification 813.11 Comments 82

3.11.1 A look ahead 844 Elementary hypothesis testing 86

4.1 Prior probabilities 874.2 Testing binary hypotheses with binary data 904.3 Nonextensibility beyond the binary case 974.4 Multiple hypothesis testing 98

4.4.1 Digression on another derivation 1014.5 Continuous probability distributionfunctions 1074.6 Testing an infinite number of hypotheses 109

4.6.1 Historical digression 1124.7 Simple and compound (or composite) hypotheses 1154.8 Comments 116

4.8.1 Etymology 1164.8.2 What have we accomplished? 117

5 Queer uses for probability theory 1195.1 Extrasensory perception 1195.2 Mrs Stewart’s telepathic powers 120

5.2.1 Digression on the normal approximation 1225.2.2 Back to Mrs Stewart 122

5.3 Converging and diverging views 1265.4 Visual perception – evolution into Bayesianity? 1325.5 The discovery of Neptune 133

5.5.1 Digression on alternative hypotheses 1355.5.2 Back to Newton 137

5.6 Horse racing and weather forecasting 1405.6.1 Discussion 142

5.7 Paradoxes of intuition 1435.8 Bayesian jurisprudence 1445.9 Comments 146

5.9.1 What is queer? 1486 Elementary parameter estimation 149

6.1 Inversion of the urn distributions 1496.2 BothN andR unknown 1506.3 Uniform prior 1526.4 Predictive distributions 154

Page 5: PROBABILITY THEORY THE LOGIC OF SCIENCE

Contents ix

6.5 Truncated uniform priors 1576.6 A concave prior 1586.7 The binomial monkey prior 1606.8 Metamorphosis into continuous parameter estimation 1636.9 Estimation with a binomial sampling distribution 163

6.9.1 Digression on optional stopping 1666.10 Compound estimation problems 1676.11 A simple Bayesian estimate: quantitative prior information 168

6.11.1 From posterior distribution function to estimate 1726.12 Effects of qualitative prior information 1776.13 Choice of a prior 1786.14 On with the calculation! 1796.15 The Jeffreys prior 1816.16 The point of it all 1836.17 Interval estimation 1866.18 Calculation of variance 1866.19 Generalization and asymptotic forms 1886.20 Rectangular sampling distribution 1906.21 Small samples 1926.22 Mathematical trickery 1936.23 Comments 195

7 The central, Gaussian or normal distribution 1987.1 The gravitating phenomenon 1997.2 The Herschel–Maxwell derivation 2007.3 The Gauss derivation 2027.4 Historical importance of Gauss’s result 2037.5 The Landon derivation 2057.6 Why the ubiquitous use of Gaussian distributions? 2077.7 Why the ubiquitous success? 2107.8 What estimator should we use? 2117.9 Error cancellation 2137.10 The near irrelevance of sampling frequency distributions 2157.11 The remarkable efficiency of information transfer 2167.12 Other sampling distributions 2187.13 Nuisance parameters as safety devices 2197.14 More general properties 2207.15 Convolution of Gaussians 2217.16 The central limit theorem 2227.17 Accuracy of computations 2247.18 Galton’s discovery 2277.19 Population dynamics and Darwinian evolution 2297.20 Evolution of humming-birds and flowers 231

Page 6: PROBABILITY THEORY THE LOGIC OF SCIENCE

x Contents

7.21 Application to economics 2337.22 The great inequality of Jupiter and Saturn 2347.23 Resolution of distributions into Gaussians 2357.24 Hermite polynomial solutions 2367.25 Fourier transform relations 2387.26 There is hope after all 2397.27 Comments 240

7.27.1 Terminology again 2408 Sufficiency, ancillarity, and all that 243

8.1 Sufficiency 2438.2 Fisher sufficiency 245

8.2.1 Examples 2468.2.2 The Blackwell–Rao theorem 247

8.3 Generalized sufficiency 2488.4 Sufficiency plus nuisance parameters 2498.5 The likelihood principle 2508.6 Ancillarity 2538.7 Generalized ancillary information 2548.8 Asymptotic likelihood: Fisher information 2568.9 Combining evidence from different sources 2578.10 Pooling the data 260

8.10.1 Fine-grained propositions 2618.11 Sam’s broken thermometer 2628.12 Comments 264

8.12.1 The fallacy of sample re-use 2648.12.2 A folk theorem 2668.12.3 Effect of prior information 2678.12.4 Clever tricks and gamesmanship 267

9 Repetitive experiments: probability and frequency 2709.1 Physical experiments 2719.2 The poorly informed robot 2749.3 Induction 2769.4 Are there general inductive rules? 2779.5 Multiplicity factors 2809.6 Partition function algorithms 281

9.6.1 Solution by inspection 2829.7 Entropy algorithms 2859.8 Another way of looking at it 2899.9 Entropy maximization 2909.10 Probability and frequency 2929.11 Significance tests 293

9.11.1 Implied alternatives 296

Page 7: PROBABILITY THEORY THE LOGIC OF SCIENCE

Contents xi

9.12 Comparison of psi and chi-squared 3009.13 The chi-squared test 3029.14 Generalization 3049.15 Halley’s mortality table 3059.16 Comments 310

9.16.1 The irrationalists 3109.16.2 Superstitions 312

10 Physics of ‘random experiments’ 31410.1 An interesting correlation 31410.2 Historical background 31510.3 How to cheat at coin and die tossing 317

10.3.1 Experimental evidence 32010.4 Bridge hands 32110.5 General random experiments 32410.6 Induction revisited 32610.7 But what about quantum theory? 32710.8 Mechanics under the clouds 32910.9 More on coins and symmetry 33110.10 Independence of tosses 33510.11 The arrogance of the uninformed 338

Part II Advanced applications11 Discrete prior probabilities: the entropy principle 343

11.1 A new kind of prior information 34311.2 Minimum

∑p2i 345

11.3 Entropy: Shannon’s theorem 34611.4 The Wallis derivation 35111.5 An example 35411.6 Generalization: a more rigorous proof 35511.7 Formal properties of maximum entropy

distributions 35811.8 Conceptual problems – frequency correspondence 36511.9 Comments 370

12 Ignorance priors and transformation groups 37212.1 What are we trying to do? 37212.2 Ignorance priors 37412.3 Continuous distributions 37412.4 Transformation groups 378

12.4.1 Location and scale parameters 37812.4.2 A Poisson rate 38212.4.3 Unknown probability for success 38212.4.4 Bertrand’s problem 386

12.5 Comments 394

Page 8: PROBABILITY THEORY THE LOGIC OF SCIENCE

xii Contents

13 Decision theory, historical background 39713.1 Inference vs. decision 39713.2 Daniel Bernoulli’s suggestion 39813.3 The rationale of insurance 40013.4 Entropy and utility 40213.5 The honest weatherman 40213.6 Reactions to Daniel Bernoulli and Laplace 40413.7 Wald’s decision theory 40613.8 Parameter estimation for minimum loss 41013.9 Reformulation of the problem 41213.10 Effect of varying loss functions 41513.11 General decision theory 41713.12 Comments 418

13.12.1 ‘Objectivity’ of decision theory 41813.12.2 Loss functions in human society 42113.12.3 A new look at the Jeffreys prior 42313.12.4 Decision theory is not fundamental 42313.12.5 Another dimension? 424

14 Simple applications of decision theory 42614.1 Definitions and preliminaries 42614.2 Sufficiency and information 42814.3 Loss functions and criteria of optimum

performance 43014.4 A discrete example 43214.5 How would our robot do it? 43714.6 Historical remarks 438

14.6.1 The classical matched filter 43914.7 The widget problem 440

14.7.1 Solution for Stage 2 44314.7.2 Solution for Stage 3 44514.7.3 Solution for Stage 4 449

14.8 Comments 45015 Paradoxes of probability theory 451

15.1 How do paradoxes survive and grow? 45115.2 Summing a series the easy way 45215.3 Nonconglomerability 45315.4 The tumbling tetrahedra 45615.5 Solution for a finite number of tosses 45915.6 Finite vs. countable additivity 46415.7 The Borel–Kolmogorov paradox 46715.8 The marginalization paradox 470

15.8.1 On to greater disasters 474

Page 9: PROBABILITY THEORY THE LOGIC OF SCIENCE

Contents xiii

15.9 Discussion 47815.9.1 The DSZ Example #5 48015.9.2 Summary 483

15.10 A useful result after all? 48415.11 How to mass-produce paradoxes 48515.12 Comments 486

16 Orthodox methods: historical background 49016.1 The early problems 49016.2 Sociology of orthodox statistics 49216.3 Ronald Fisher, Harold Jeffreys, and Jerzy Neyman 49316.4 Pre-data and post-data considerations 49916.5 The sampling distribution for an estimator 50016.6 Pro-causal and anti-causal bias 50316.7 What is real, the probability or the phenomenon? 50516.8 Comments 506

16.8.1 Communication difficulties 50717 Principles and pathology of orthodox statistics 509

17.1 Information loss 51017.2 Unbiased estimators 51117.3 Pathology of an unbiased estimate 51617.4 The fundamental inequality of the sampling variance 51817.5 Periodicity: the weather in Central Park 520

17.5.1 The folly of pre-filtering data 52117.6 A Bayesian analysis 52717.7 The folly of randomization 53117.8 Fisher: common sense at Rothamsted 532

17.8.1 The Bayesian safety device 53217.9 Missing data 53317.10 Trend and seasonality in time series 534

17.10.1 Orthodox methods 53517.10.2 The Bayesian method 53617.10.3 Comparison of Bayesian and orthodox

estimates 54017.10.4 An improved orthodox estimate 54117.10.5 The orthodox criterion of performance 544

17.11 The general case 54517.12 Comments 550

18 TheAp distribution and rule of succession 55318.1 Memory storage for old robots 55318.2 Relevance 55518.3 A surprising consequence 55718.4 Outer and inner robots 559

Page 10: PROBABILITY THEORY THE LOGIC OF SCIENCE

xiv Contents

18.5 An application 56118.6 Laplace’s rule of succession 56318.7 Jeffreys’ objection 56618.8 Bass or carp? 56718.9 So where does this leave the rule? 56818.10 Generalization 56818.11 Confirmation and weight of evidence 571

18.11.1 Is indifference based on knowledge or ignorance? 57318.12 Carnap’s inductive methods 57418.13 Probability and frequency in exchangeable sequences 57618.14 Prediction of frequencies 57618.15 One-dimensional neutron multiplication 579

18.15.1 The frequentist solution 57918.15.2 The Laplace solution 581

18.16 The de Finetti theorem 58618.17 Comments 588

19 Physical measurements 58919.1 Reduction of equations of condition 58919.2 Reformulation as a decision problem 592

19.2.1 Sermon on Gaussian error distributions 59219.3 The underdetermined case:K is singular 59419.4 The overdetermined case:K can be made nonsingular 59519.5 Numerical evaluation of the result 59619.6 Accuracy of the estimates 59719.7 Comments 599

19.7.1 A paradox 59920 Model comparison 601

20.1 Formulation of the problem 60220.2 The fair judge and the cruel realist 603

20.2.1 Parameters known in advance 60420.2.2 Parameters unknown 604

20.3 But where is the idea of simplicity? 60520.4 An example: linear response models 607

20.4.1 Digression: the old sermon still another time 60820.5 Comments 613

20.5.1 Final causes 61421 Outliers and robustness 615

21.1 The experimenter’s dilemma 61521.2 Robustness 61721.3 The two-model model 61921.4 Exchangeable selection 62021.5 The general Bayesian solution 622

Page 11: PROBABILITY THEORY THE LOGIC OF SCIENCE

Contents xv

21.6 Pure outliers 62421.7 One receding datum 625

22 Introduction to communication theory 62722.1 Origins of the theory 62722.2 The noiseless channel 62822.3 The information source 63422.4 Does the English language have statistical properties? 63622.5 Optimum encoding: letter frequencies known 63822.6 Better encoding from knowledge of digram frequencies 64122.7 Relation to a stochastic model 64422.8 The noisy channel 648

Appendix A Other approaches to probability theory 651A.1 The Kolmogorov system of probability 651A.2 The de Finetti system of probability 655A.3 Comparative probability 656A.4 Holdouts against universal comparability 658A.5 Speculations about lattice theories 659

Appendix B Mathematical formalities and style 661B.1 Notation and logical hierarchy 661B.2 Our ‘cautious approach’ policy 662B.3 Willy Feller on measure theory 663B.4 Kronecker vs. Weierstrasz 665B.5 What is a legitimate mathematical function? 666

B.5.1 Delta-functions 668B.5.2 Nondifferentiable functions 668B.5.3 Bogus nondifferentiable functions 669

B.6 Counting infinite sets? 671B.7 The Hausdorff sphere paradox and mathematical

diseases 672B.8 What am I supposed to publish? 674B.9 Mathematical courtesy 675

Appendix C Convolutions and cumulants 677C.1 Relation of cumulants and moments 679C.2 Examples 680

References 683Bibliography 705Author index 721Subject index 724

Page 12: PROBABILITY THEORY THE LOGIC OF SCIENCE

1

Plausible reasoning

The actual science of logic is conversant at present only with things eithercertain, impossible, or entirelydoubtful, none of which (fortunately) wehave to reason on. Therefore the true logic for this world is the calculusof Probabilities, which takes account of the magnitude of the probabilitywhich is, or ought to be, in a reasonable man’s mind.

James Clerk Maxwell (1850)

Suppose somedark night a policemanwalks downa street, apparently deserted. Suddenly hehears a burglar alarm, looks across the street, and sees a jewelry store with a brokenwindow.Then a gentlemanwearing amask comes crawling out through the brokenwindow, carryinga bag which turns out to be full of expensive jewelry. The policeman doesn’t hesitate at allin deciding that this gentleman is dishonest. But by what reasoning process does he arriveat this conclusion? Let us first take a leisurely look at the general nature of such problems.

1.1 Deductive and plausible reasoning

A moment’s thought makes it clear that our policeman’s conclusion was not a logicaldeduction from the evidence; for there may have been a perfectly innocent explanationfor everything. It might be, for example, that this gentleman was the owner of the jewelrystore and he was coming home from a masquerade party, and didn’t have the key with him.However, just as he walked by his store, a passing truck threw a stone through the window,and he was only protecting his own property.Now, while the policeman’s reasoning process was not logical deduction, we will grant

that it had a certain degree of validity. The evidence did notmake the gentleman’s dishonestycertain, but it did make it extremelyplausible. This is an example of a kind of reasoningin which we have all become more or less proficient, necessarily, long before studyingmathematical theories. We are hardly able to get through one waking hour without facingsome situation (e.g. will it rain or won’t it?) where we do not have enough information topermit deductive reasoning; but still we must decide immediately what to do.In spite of its familiarity, the formation of plausible conclusions is a very subtle process.

Although history records discussions of it extending over 24 centuries, probably nobody has

3

Page 13: PROBABILITY THEORY THE LOGIC OF SCIENCE

4 Part 1 Principles and elementary applications

ever produced an analysis of the process which anyone else finds completely satisfactory.In this work we will be able to report some useful and encouraging new progress, in whichconflicting intuitive judgments are replaced by definite theorems, andad hocproceduresare replaced by rules that are determined uniquely by some very elementary – and nearlyinescapable – criteria of rationality.All discussions of these questions start by giving examples of the contrast between

deductive reasoning and plausible reasoning. As is generally credited to theOrganonofAristotle (fourth centurybc)1 deductive reasoning (apodeixis) can be analyzed ultimatelyinto the repeated application of two strong syllogisms:

if A is true, thenB is true

A is true (1.1)

therefore,B is true,

and its inverse:if A is true, thenB is true

B is false (1.2)

therefore,A is false.

This is the kind of reasoning we would like to use all the time; but, as noted, in almost allthe situations confronting us we do not have the right kind of information to allow this kindof reasoning. We fall back on weaker syllogisms (epagoge):

if A is true, thenB is true

B is true (1.3)

therefore,A becomes more plausible.

The evidence does not prove thatA is true, but verification of one of its consequences doesgive us more confidence inA. For example, let

A ≡ it will start to rain by 10am at the latest;

B ≡ the sky will become cloudy before 10am.

Observing clouds at 9:45am does not give us a logical certainty that the rain will follow;nevertheless our common sense, obeying the weak syllogism, may induce us to change ourplans and behaveas ifwe believed that it will, if those clouds are sufficiently dark.

This example shows also that the major premise, ‘ifA thenB’ expressesB only as alogical consequence ofA; and not necessarily a causal physical consequence, which couldbe effective only at a later time. The rain at 10am is not the physical cause of the clouds at

1 Today, several different views are held about the exact nature of Aristotle’s contribution. Such issues are irrelevant to our presentpurpose, but the interested reader may find an extensive discussion of them in Lukasiewicz (1957).

Page 14: PROBABILITY THEORY THE LOGIC OF SCIENCE

1 Plausible reasoning 5

9:45am. Nevertheless, the proper logical connection is not in the uncertain causal direction(clouds=⇒ rain), but rather (rain=⇒ clouds), which is certain, although noncausal.

We emphasize at the outset that we are concerned here withlogicalconnections, becausesome discussions and applications of inference have fallen into serious error through failureto see the distinction between logical implication and physical causation. The distinctionis analyzed in some depth by Simon and Rescher (1966), who note that all attempts tointerpret implication as expressing physical causation founder on the lack of contrapositionexpressed by the second syllogism (1.2). That is, if we tried to interpret the major premiseas ‘A is the physical cause ofB’, then we would hardly be able to accept that ‘not-B isthe physical cause of not-A’. In Chapter 3 we shall see that attempts to interpret plausibleinferences in terms of physical causation fare no better.Another weak syllogism, still using the same major premise, is

If A is true, thenB is true

A is false (1.4)

therefore,B becomes less plausible.

In this case, the evidence does not prove thatB is false; but one of the possible reasons forits being true has been eliminated, and so we feel less confident aboutB. The reasoning of ascientist, by which he accepts or rejectshis theories, consists almost entirely of syllogismsof the second and third kind.Now, the reasoning of our policeman was not even of the above types. It is best described

by a still weaker syllogism:

If A is true, thenB becomes more plausible

B is true (1.5)

therefore,A becomes more plausible.

But in spite of the apparent weakness of this argument, when stated abstractly in terms ofAandB, we recognize that the policeman’s conclusion has a very strong convincing power.There is something which makes us believe that, in this particular case, his argument hadalmost the power of deductive reasoning.These examples show that the brain, in doing plausible reasoning, not only decides

whether something becomesmore plausible or less plausible, but that it evaluates thedegreeof plausibility in some way. The plausibility for rain by 10am depends very much on thedarkness of those clouds at 9:45. And the brain also makes use of old informationas wellas the specific new data of the problem; in deciding what to do we try to recall our pastexperience with clouds and rain, and what the weatherman predicted last night.To illustrate that the policeman was also making use of the past experience of policemen

in general, we have only to change that experience. Suppose that events like these happenedseveral times every night to every policeman – and that in every case the gentleman turned

Page 15: PROBABILITY THEORY THE LOGIC OF SCIENCE

6 Part 1 Principles and elementary applications

out to be completely innocent. Very soon, policemen would learn to ignore such trivialthings.Thus, in our reasoningwe depend verymuch onprior informationto help us in evaluating

the degree of plausibility in a new problem. This reasoning process goes on unconsciously,almost instantaneously, and we conceal how complicated it really is by calling itcommonsense.

The mathematician George P´olya (1945, 1954) wrote three books about plausible rea-soning, pointing out a wealth of interesting examples and showing that there are definiterules by which we do plausible reasoning (although in his work they remain in qualitativeform). The above weak syllogisms appear in his third volume. The reader is strongly urgedto consult P´olya’s exposition, which was the original source ofmany of the ideas underlyingthe present work. We show below how P´olya’s principles may be made quantitative, withresulting useful applications.Evidently, the deductive reasoning described above has the property that we can go

through long chains of reasoning of the type (1.1) and (1.2) and the conclusions have just asmuch certainty as the premises.With the other kinds of reasoning, (1.3)–(1.5), the reliabilityof the conclusion changes as we go through several stages. But in their quantitative formweshall find that in many cases our conclusions can still approach the certainty of deductivereasoning (as the example of the policeman leads us to expect). P´olya showed that evena pure mathematician actually uses these weaker forms of reasoning most of the time. Ofcourse, on publishing a new theorem, the mathematician will try very hard to invent anargument which uses only the first kind; but the reasoning process which led to the theoremin the first place almost always involves one of the weaker forms (based, for example, onfollowing up conjectures suggested by analogies). The same idea is expressed in a remarkof S. Banach (quoted by S. Ulam, 1957):

Good mathematicians see analogies between theorems; great mathematicians see analogies betweenanalogies.

As a first orientation, then, let us note some very suggestive analogies to another field –which is itself based, in the last analysis, on plausible reasoning.

1.2 Analogies with physical theories

In physics, we learn quickly that the world is too complicated for us to analyze it all at once.We can make progress only if we dissect it into little pieces and study them separately.Sometimes, we can invent a mathematical model which reproduces several features of oneof these pieces, and whenever this happens we feel that progress has been made. Thesemodels are calledphysical theories. As knowledge advances, we are able to invent betterandbettermodels,which reproducemoreandmore featuresof the realworld,moreandmoreaccurately. Nobody knows whether there is some natural end to this process, or whether itwill go on indefinitely.

Page 16: PROBABILITY THEORY THE LOGIC OF SCIENCE

1 Plausible reasoning 7

In trying to understand common sense, we shall take a similar course. We won’t tryto understand it all at once, but we shall feel that progress has been made if we areable to construct idealized mathematical models which reproduce a few of its features.We expect that any model we are now able to construct will be replaced by more com-plete ones in the future, and we do not know whether there is any natural end to thisprocess.The analogy with physical theories is deeper than a mere analogy of method. Often, the

things which are most familiar to us turn out to be the hardest to understand. Phenomenawhose very existence is unknown to the vast majority of the human race (such as the differ-ence in ultraviolet spectra of iron and nickel) can be explained in exhaustive mathematicaldetail – but all of modern science is practically helpless when faced with the complicationsof such a commonplace fact as growth of a blade of grass. Accordingly, we must not expecttoomuch of our models; wemust be prepared to find that some of themost familiar featuresof mental activity may be ones for which we have the greatest difficulty in constructing anyadequate model.There are many more analogies. In physics we are accustomed to finding that any ad-

vance in knowledge leads to consequences of great practical value, but of an unpredictablenature. R¨ontgen’s discovery of X-rays led to important new possibilities of medical diag-nosis; Maxwell’s discovery of one more term in the equation for curlH led to practicallyinstantaneous communication all over the earth.Our mathematical models for common sense also exhibit this feature of practical useful-

ness. Any successful model, even though it may reproduce only a few features of commonsense, will prove to be a powerful extension of common sense in some field of application.Within this field, it enables us to solve problems of inference which are so involved incomplicated detail that we would never attempt to solve them without its help.

1.3 The thinking computer

Models have practical uses of a quite different type. Many people are fond of saying, ‘Theywill never make a machine to replace the human mind – it does many things which nomachine could ever do.’ A beautiful answer to this was given by J. von Neumann in a talkon computers given in Princeton in 1948, which the writer was privileged to attend. In replyto the canonical question from the audience (‘But of course, a mere machine can’t reallythink, can it?’), he said:

You insist that there is something a machine cannot do. If you will tell me precisely what it is that amachine cannot do, then I can always make a machine which will do just that!

In principle, the only operations which a machine cannot perform for us are those whichwe cannot describe in detail, or which could not be completed in a finite number of steps.Of course, some will conjure up images of G¨odel incompleteness, undecidability, Turingmachines which never stop, etc. But to answer all such doubts we need only point to the

Page 17: PROBABILITY THEORY THE LOGIC OF SCIENCE

8 Part 1 Principles and elementary applications

existence of the human brain, whichdoesit. Just as von Neumann indicated, the onlyreal limitations on making ‘machines which think’ are our own limitations in not knowingexactly what ‘thinking’ consists of.But in our study of common sense we shall be led to some very explicit ideas about

the mechanism of thinking. Every time we can construct a mathematical model whichreproduces a part of common sense by prescribing a definite set of operations, this showsus how to ‘build a machine’, (i.e. write a computer program) which operates on incompleteinformation and, by applying quantitative versions of the above weak syllogisms, doesplausible reasoning instead of deductive reasoning.Indeed, the development of such computer software for certain specialized problems of

inference is oneof themostactive anduseful current trends in thisfield.One kind of problemthus dealt with might be: given a mass of data, comprising 10 000 separate observations,determine in the light of these data and whatever prior information is at hand, the relativeplausibilities of 100 different possible hypotheses about the causes at work.Our unaided common sense might be adequate for deciding between two hypotheses

whose consequences are very different; but, in dealing with 100 hypotheses which are notvery different, wewould be helplesswithout a computerandawell-developedmathematicaltheory that shows us how to program it. That is, what determines, in the policeman’ssyllogism (1.5), whether the plausibility forA increases by a large amount, raising it almostto certainty; or only a negligibly small amount, making the dataB almost irrelevant? Theobject of the present work is to develop themathematical theory which answers suchquestions, in the greatest depth and generality now possible.While we expect a mathematical theory to be useful in programming computers, the

idea of a thinking computer is also helpful psychologicallyin developing the mathematicaltheory. The question of the reasoning process used by actual human brains is charged withemotion and grotesque misunderstandings. It is hardly possible to say anything about thiswithout becoming involved in debates over issues that are not only undecidable in ourpresent state of knowledge, but are irrelevant to our purpose here.Obviously, the operation of real human brains is so complicated that we can make no

pretense of explaining its mysteries; and in any event we are not trying to explain, much lessreproduce, all the aberrations and inconsistencies of human brains. That is an interesting andimportant subject; but it is not the subject we are studying here. Our topic is thenormativeprinciples of logic, and not the principles of psychology or neurophysiology.

To emphasize this, instead of asking, ‘How can we build a mathematical model of humancommon sense?’, let us ask, ‘How could we build a machine which would carry out usefulplausible reasoning, following clearly defined principles expressing an idealized commonsense?’

1.4 Introducing the robot

In order to direct attention to constructive things and away from controversial irrelevancies,we shall invent an imaginary being. Its brain is to be designedby us, so that it reasons

Page 18: PROBABILITY THEORY THE LOGIC OF SCIENCE

1 Plausible reasoning 9

according to certain definite rules. These rules will be deduced from simple desideratawhich, it appears to us, would be desirable in human brains; i.e. we think that a rationalperson, on discovering that they were violating one of these desiderata, would wish to revisetheir thinking.In principle, we are free to adopt any rules we please; that is our way ofdefiningwhich

robot we shall study. Comparing its reasoning with yours, if you find no resemblance youare in turn free to reject our robot and design a different one more to your liking. But if youfind a very strong resemblance, and decide that you want and trust this robot to help you inyour own problems of inference, then that will be an accomplishment of the theory, not apremise.Our robot is going to reason about propositions. As already indicated above, we shall

denote various propositions by italicized capital letters,{A, B,C,etc.}, and for the timebeing we must require that any proposition used must have, to the robot, an unambiguousmeaning andmust be of the simple, definite logical type thatmust be either true or false. Thatis, until otherwise stated, we shall be concerned only with two-valued logic, or Aristotelianlogic. We do not require that the truth or falsity of such an ‘Aristotelian proposition’ beascertainable by any feasible investigation; indeed, our inability to do this is usually justthe reason why we need the robot’s help. For example, the writer personally considers bothof the following propositions to be true:

A ≡ Beethoven and Berlioz never met.

B ≡ Beethoven’s music has a better sustained quality than that of

Berlioz, although Berlioz at his best is the equal of anybody.

PropositionB is not a permissible one for our robot to think about at present, whereaspropositionA is, although it is unlikely that its truth or falsity could be definitely establishedtoday.2 After our theory is developed, it will be of interest to see whether the presentrestriction to Aristotelian propositions such asA can be relaxed, so that the robot might helpus also with more vague propositions such asB (see Chapter 18 on theAp-distribution).3

1.5 Boolean algebra

To state these ideas more formally, we introduce some notation of the usual symbolic logic,or Boolean algebra, so called because George Boole (1854) introduced anotationsimilarto the following. Of course, the principles of deductive logic itself were well understoodcenturies before Boole, and, as we shall see, all the results that follow from Boolean al-gebra were contained already as special cases in the rules of plausible inference given

2 Their meeting is a chronological possibility, since their lives overlapped by 24 years; my reason for doubting it is the failure ofBerlioz to mention any such meeting in his memoirs – on the other hand, neither does he come out and say definitely that theydid notmeet.

3 The question of how one is to make amachine in some sense ‘cognizant’ of the conceptual meaning that a proposition likeA hasto humans, might seem very difficult, and much of the subject of artificial intelligence is devoted to inventingad hocdevices todeal with this problem. However, we shall find in Chapter 4 that for us the problem is almost nonexistent; our rules for plausiblereasoning automatically provide the means to do the mathematical equivalent of this.

Page 19: PROBABILITY THEORY THE LOGIC OF SCIENCE

10 Part 1 Principles and elementary applications

by (1812). The symbol

AB, (1.6)

called thelogical productor theconjunction, denotes the proposition ‘bothA andB aretrue’. Obviously, the order in which we state them does not matter;AB andBA say thesame thing. The expression

A+ B, (1.7)

called thelogical sumor disjunction, stands for ‘at least one of the propositions,A, B istrue’ and has the same meaning asB + A. These symbols are only a shorthand way ofwriting propositions, and do not stand for numerical values.Given two propositionsA, B, it may happen that one is true if and only if the other is true;

we then say that they have the sametruth value. This may be only a simpletautology (i.e.A andB are verbal statements which obviously say the same thing), or it may be that onlyafter immense mathematical labor is it finally proved thatA is the necessary and sufficientcondition for B. From the standpoint of logic it does not matter; once it is established,by any means, thatA andB have the same truth value, then they are logically equivalentpropositions, in the sense that any evidence concerning the truth of one pertains equallywell to the truth of the other, and they have the same implications for any further reasoning.Evidently, then, it must be the most primitive axiom of plausible reasoning that two

propositions with the same truth value are equally plausible. This might appear almost tootrivial to mention, were it not for the fact that Boole himself (Boole, 1854, p. 286) fell intoerror on this point, bymistakenly identifying two propositionswhichwere in fact different –and then failing to see any contradictionin their different plausibilities. Three years later,Boole (1857) gave a revised theory which supersedes that in his earlier book; for furthercomments on this incident, see Keynes (1921, pp. 167–168); Jaynes (1976, pp. 240–242).In Boolean algebra, the equal sign is used to denote not equal numerical value, but equal

truth value:A = B, and the ‘equations’ of Boolean algebra thus consist of assertions thatthe proposition on the left-hand side has the same truth value as the one on the right-handside. The symbol ‘≡’ means, as usual, ‘equals by definition’.In denoting complicated propositions we use parentheses in the same way as in ordinary

algebra, i.e. to indicate the order in which propositions are to be combined(at timeswe shalluse them also merely for clarity of expression although they are not strictly necessary). Intheir absence we observe the rules of algebraic hierarchy, familiar to those who use handcalculators: thusAB+ C denotes (AB) + C; and notA(B + C).

Thedenialof a proposition is indicated by a bar:

A ≡ A is false. (1.8)

The relation betweenA, A is a reciprocal one:

A = A is false, (1.9)

Page 20: PROBABILITY THEORY THE LOGIC OF SCIENCE

1 Plausible reasoning 11

and it does not matter which proposition we denote by the barred and which by the unbarredletter. Note that some care is needed in the unambiguous use of the bar. For example,according to the above conventions,

AB = AB is false; (1.10)

A B = bothA andB are false. (1.11)

These are quite different propositions; in fact,AB is not the logical productA B, but thelogical sum:AB = A+ B.

With these understandings, Boolean algebra is characterized by some rather trivial andobvious basic identities, which express the properties of:

Idempotence:

{AA= A

A+ A = A

Commutativity:

{AB = BA

A+ B = B + A

Associativity:

{A(BC) = (AB)C = ABC

A+ (B + C) = (A+ B) + C = A+ B + C

Distributivity:

{A(B + C) = AB+ AC

A+ (BC) = (A+ B)(A+ C)

Duality:

{If C = AB, thenC = A+ B

If D = A+ B, thenD = A B

(1.12)

but by their application one can prove any number of further relations, some highly non-trivial. For example, we shall presently have use for the rather elementary theorem:

if B = AD thenA B = B andB A = A. (1.13)

Implication

The proposition

A ⇒ B (1.14)

to be read as ‘A impliesB’, does not assert that eitherA or B is true; it means only thatA Bis false, or, what is the same thing, (A+ B) is true. This can be written also as the logicalequationA = AB. That is, given (1.14), ifA is true thenB must be true; or, ifB is falsethenAmust be false. This is just what is stated in the strong syllogisms (1.1) and (1.2).

Page 21: PROBABILITY THEORY THE LOGIC OF SCIENCE

12 Part 1 Principles and elementary applications

On the other hand, ifA is false, (1.14) says nothing aboutB: and ifB is true, (1.14) saysnothing aboutA. But these are just the cases in which our weak syllogisms (1.3), (1.4) dosay something. In one respect, then, the term ‘weak syllogism’ is misleading. The theoryof plausible reasoning based on weak syllogisms is not a ‘weakened’ form of logic; it isanextensionof logic with new content not present at all in conventional deductive logic. Itwill become clear in the next chapter (see (2.69) and (2.70)) that our rules include deductivelogic as a special case.

A tricky point

Note carefully that in ordinary language one would take ‘A implies B’ to mean thatBis logically deducible fromA. But, in formal logic, ‘A implies B’ means only that thepropositionsA and AB have the same truth value. In general, whetherB is logicallydeducible fromA does not depend only on the propositionsA andB; it depends on thetotality of propositions (A, A′, A′′, . . .) that we accept as true and which are thereforeavailable to use in the deduction.Devinatz (1968, p. 3) and Hamilton (1988, p. 5) give thetruth table for the implication as a binary operation, illustrating thatA ⇒ B is false only ifA is true andB is false; in all other casesA ⇒ B is true!This may seem startling at first glance; however, notethat, indeed, ifA andB are both

true, thenA = AB and soA ⇒ B is true; in formal logic every true statement impliesevery other true statement. On the other hand, ifA is false, thenAQ is also false for allQ, thusA = AB andA = AB are both true, soA ⇒ B andA ⇒ B are both true; a falseproposition implies all propositions. If we tried to interpret this as logical deducibility(i.e. bothB andB are deducible fromA), it would follow that every false proposition islogically contradictory. Yet the proposition: ‘Beethoven outlived Berlioz’ is false but hardlylogically contradictory (for Beethoven did outlive many people who were the same age asBerlioz).Obviously, merely knowing that propositionsA andB are both true does not provide

enough information to decide whether either is logically deducible from the other, plussome unspecified ‘toolbox’ of other propositions. The question of logical deducibility ofone proposition from a set of others arises in a crucial way in theG¨odel theoremdiscussed atthe end of Chapter 2. This great difference in the meaning of the word ‘implies’ in ordinarylanguage and in formal logic is a tricky point that can lead to serious error if it is not properlyunderstood; it appears to us that ‘implication’ is an unfortunate choice of word, and thatthis is not sufficiently emphasized in conventional expositions of logic.

1.6 Adequate sets of operations

We note some features of deductive logic which will be needed in the design of our robot.We have defined four operations, or ‘connectives’, by which, starting from two propositionsA, B, other propositions may be defined: the logical product or conjunctionAB, the logical

Page 22: PROBABILITY THEORY THE LOGIC OF SCIENCE

1 Plausible reasoning 13

sum or disjunctionA+ B, the implicationA ⇒ B, and the negationA. By combiningthese operations repeatedly in every possible way, one can generate any number of newpropositions, such as

C ≡ (A+ B)(A+ A B) + A B(A+ B). (1.15)

Many questions then occur to us: How large is the class of new propositions thus generated?Is it infinite, or is thereafinite set that is closedunder theseoperations?Caneverypropositiondefined fromA, B be thus represented, or does this require further connectives beyond theabove four? Or are these four already overcomplete so that some might be dispensed with?What is the smallest set of operations that is adequate to generate all such ‘logic functions’of A and B? If instead of two starting propositionsA, B we have an arbitrary number{A1, . . . , An}, is this set of operationsstill adequate to generate all possible logic functionsof {A1, . . . , An}?

All these questions are answered easily, with results useful for logic, probability theory,and computer design. Broadly speaking, we are asking whether, starting from our presentvantage point, we can (1) increase the number of functions, (2) decrease the number ofoperations. The first query is simplified by noting that two propositions, although they mayappear entirely differentwhenwritten out in themanner (1.15), are not different propositionsfrom the standpoint of logic if they have the same truth value. For example, it is left forthe reader to verify thatC in (1.15) is logically the same statement as the implicationC = (B ⇒ A).

Since we are, at this stage, restricting our attention to Aristotelian propositions, any logicfunctionC = f (A, B) such as (1.15) has only two possible ‘values’, true and false; andlikewise the ‘independent variables’A andB can take ononly those two values.

At this point, a logician might object to our notation, saying that the symbolA hasbeen defined as standing for some fixed proposition, whose truth cannot change; so if wewish to consider logic functions, then instead of writingC = f (A, B) we should introducenew symbols and writez= f (x, y), wherex, y, z, are ‘statement variables’ for whichvarious specific statementsA, B,C may be substituted. But ifA stands for some fixed butunspecifiedproposition, then it canstill beeither trueor false.Weachieve thesameflexibilitymerely by the understanding that equations like (1.15) which define logic functions are tobe true for all ways of definingA, B ; i.e. instead of a statement variable we use a variablestatement.In relations of the formC = f (A, B), we are concerned with logic functions defined

on a discrete ‘space’ S consisting of only 22 = 4 points; namely those at whichA andB take on the ‘values’{TT,TF,FT,FF}, respectively; and, at each point, the functionf (A, B) can take on independently either of two values{T,F}. There are, therefore, exactly24 = 16 different logic functionsf (A, B), and nomore. An expressionB = f (A1, . . . , An)involving n propositions is a logic function on a space S ofM = 2n points; and there areexactly 2M such functions.

Page 23: PROBABILITY THEORY THE LOGIC OF SCIENCE

14 Part 1 Principles and elementary applications

In the casen = 1, there are four logic functions{ f1(A), . . . , f4(A)}, which we can defineby enumeration, listing all their possible values in a truth table:

A T F

f1(A) T Tf2(A) T Ff3(A) F Tf4(A) F F

But it is obvious by inspection that these are just

f1(A) = A+ A

f2(A) = A

f3(A) = A

f4(A) = A A,

(1.16)

soweprove by enumeration that the three operations: conjunction, disjunction, and negationare adequate to generate all logic functions of a single proposition.For the case of generaln, consider first the special functions, each of which is true at one

and only one point of S. Forn = 2 there are 2n = 4 such functions,

A, B TT TF FT FF

f1(A, B) T F F Ff2(A, B) F T F Ff3(A, B) F F T Ff4(A, B) F F F T

It is clear by inspection that these are just the four basic conjunctions,

f1(A, B) = A B

f2(A, B) = A B

f3(A, B) = A B

f4(A, B) = A B.

(1.17)

Consider now any logic function which is true on certain specified points of S; for example,f5(A, B) and f6(A, B), defined by

A, B TT TF FT FF

f5(A, B) F T F Tf6(A, B) T F T T

Page 24: PROBABILITY THEORY THE LOGIC OF SCIENCE

1 Plausible reasoning 15

We assert that each of these functions is the logical sum of the conjunctions (1.17) that aretrue on the same points (this is not trivial; the reader should verify it in detail). Thus,

f5(A, B) = f2(A, B) + f4(A, B)

= A B + A B

= (A+ A) B

= B,

(1.18)

and, likewise,

f6(A, B) = f1(A, B) + f3(A, B) + f4(A, B)

= AB+ A B+ A B

= B + A B

= A+ B.

(1.19)

That is, f6(A, B) is the implication f6(A, B) = (A ⇒ B), with the truth table discussedabove. Any logic functionf (A, B) that is true on at least one point of S can be constructed inthis way as a logical sum of the basic conjunctions (1.17). There are 24 − 1 = 15 such func-tions. For the remainingfunction, which is always false, it suffices to take the contradiction,f16(A, B) ≡ A A.This method (called ‘reduction todisjunctive normal form’ in logic textbooks) will work

for anyn. For example, in the casen = 5 there are 25 = 32 basic conjunctions,

{ABCDE, ABCDE, ABCDE, . . . , A BC D E}, (1.20)

and232 = 4294967296different logic functionsfi (A, B,C, D, E); ofwhich4294967295can be written as logical sums of the basic conjunctions, leaving only the contradiction

f4294967296(A, B,C, D, E) = A A. (1.21)

Thus one can verify by ‘construction in thought’ that the three operations

{conjunction, disjunction, negation}, i.e. {AND, OR, NOT}, (1.22)

suffice to generate all possible logic functions; or, more concisely, they form anadequate set.The duality property (1.12) shows that a smaller set will suffice; for disjunction ofA, B

is the same as denying that they are both false:

A+ B = (A B). (1.23)

Therefore, the two operations (AND, NOT) already constitute an adequate set for deductivelogic.4 This fact will be essential in determining when we have an adequate set of rules forplausible reasoning; see Chapter 2.

4 For you to ponder: Does it follow that these two commands are the only ones needed to write any computer program?

Page 25: PROBABILITY THEORY THE LOGIC OF SCIENCE

16 Part 1 Principles and elementary applications

It is clear that we cannot now strike out either of these operations, leaving only theother; i.e. the operation ‘AND’ cannot be reduced to negations; and negation cannot beaccomplished by any number of ‘AND’ operations. But this still leaves open the possibilitythat both conjunction and negation might be reducible to some third operation, not yetintroduced, so that a single logic operation would constitute an adequate set.It comes as a pleasant surprise to find that there is not only one but two such operations.

The operation ‘NAND’ is defined as the negation of ‘AND’:

A ↑ B ≡ AB = A+ B (1.24)

which we can read as ‘ANAND B’. But then we have at once

A = A ↑ A

AB = (A ↑ B) ↑ (A ↑ B)

A+ B = (A ↑ A) ↑ (B ↑ B).

(1.25)

Therefore, every logic function can be constructed with NAND alone. Likewise, theoperation NOR defined by

A ↓ B ≡ A+ B = A B (1.26)

is also powerful enough to generate all logic functions:

A = A ↓ A

A+ B = (A ↓ B) ↓ (A ↓ B)

AB = (A ↓ A) ↓ (B ↓ B).

(1.27)

One can take advantage of this in designing computer and logic circuits. A‘logic gate’ is acircuit having, besides a common ground, two input terminals and one output. The voltagerelative to ground at any of these terminals can take on only two values; say+3 volts, or‘up’, representing ‘true’; and 0 volts or ‘down’, representing ‘false’. A NAND gate is thusone whose output is up if and only if at least one of the inputs is down; or, what is the samething, down if and only if both inputs are up; while for a NOR gate the output is upif andonly if both inputs are down.One of the standard components of logic circuits is the ‘quad NAND gate’, an integrated

circuit containing four independent NAND gates on one semiconductor chip. Given a suffi-cient number of these and no other circuit components, it is possible to generate any requiredlogic function by interconnecting them in various ways.This short excursion into deductive logic is as far as we need go for our purposes. Further

developments are given inmany textbooks; for example, amodern treatment of Aristotelianlogic is given by Copi (1994). For non-Aristotelian forms with special emphasis on G¨odelincompleteness, computability, decidability, Turing machines, etc., see Hamilton (1988).We turn now to our extension of logic, which is to follow from the conditions discussed

next.We call them ‘desiderata’ rather than ‘axioms’ because they do not assert that anythingis ‘true’ but only state what appear to be desirable goals. Whether these goals are attainable

Page 26: PROBABILITY THEORY THE LOGIC OF SCIENCE

1 Plausible reasoning 17

without contradictions, and whether they determine any unique extension of logic, arematters of mathematical analysis, given in Chapter 2.

1.7 The basic desiderata

To each proposition about which it reasons, our robot must assign some degree of plausi-bility, based on the evidence we have given it; and whenever it receives new evidence itmust revise these assignments to take that new evidence into account. In order that theseplausibility assignments can be stored and modified in the circuits of its brain, they mustbe associated with some definite physical quantity, such as voltage or pulse duration or abinary coded number, etc. – however our engineers want to design the details. For presentpurposes, this means that there will have to be some kind of association between degreesof plausibility and real numbers:

(I) Degrees of plausibility are represented by real numbers. (1.28)

Desideratum (I) is practically forced on us by the requirement that the robot’s brain mustoperate by the carrying out of some definite physical process. However, it will appear(Appendix A) that it is also required theoretically; we do not see the possibility of anyconsistent theory without a property that is equivalent functionally to desideratum (I).Weadopt anatural but nonessential convention: that agreater plausibility shall correspond

to a greater number. It will also be convenient to assume a continuity property, which is hardto state precisely at this stage; to say it intuitively: an infinitesimally greater plausibilityought to correspond only to an infinitesimally greater number.The plausibility that the robot assigns to some propositionAwill, in general, depend on

whether we told it that some other propositionB is true. Following the notation of Keynes(1921) and Cox (1961), we indicate this by the symbol

A|B, (1.29)

which we may call ‘the conditional plausibility thatA is true, given thatB is true’ or just‘ A givenB’. It stands for some real number. Thus, for example,

A|BC (1.30)

(which we may read as ‘A givenBC’) represents the plausibility thatA is true, given thatbothB andC are true. Or,

A+ B|CD (1.31)

represents the plausibility that at least one of the propositionsA andB is true, given thatbothC andD are true; and so on. We have decided to represent a greater plausibility by agreater number, so

(A|B) > (C|B) (1.32)

Page 27: PROBABILITY THEORY THE LOGIC OF SCIENCE

18 Part 1 Principles and elementary applications

says that, givenB, A is more plausible thanC. In this notation, while the symbol forplausibility is just of the formA|B without parentheses, we often add parentheses forclarity of expression. Thus, (1.32) says the same thing as

A|B > C|B, (1.33)

but its meaning is clearer to the eye.In the interest of avoiding impossible problems, we are not going to ask our robot to

undergo the agony of reasoning from impossible or mutually contradictory premises; therecould be no ‘correct’ answer. Thus, we make no attempt to defineA|BCwhenB andC aremutually contradictory. Whenever such a symbol appears, it is understood thatB andC arecompatible propositions.Also, we do not want this robot to think in a way that is directly opposed to the way you

and I think. So we shall design it to reason in a way that is at leastqualitativelylike the wayhumans try to reason, as described by the above weak syllogisms and a number of othersimilar ones.Thus, if it hasold informationCwhichgetsupdated toC′ in suchaway that theplausibility

for A is increased:

(A|C′) > (A|C); (1.34)

but the plausibility forB givenA is not changed:

(B|AC′) = (B|AC). (1.35)

This can, of course, produce only an increase, never a decrease, in the plausibility that bothA andB are true:

(AB|C′) ≥ (AB|C); (1.36)

and it must produce a decrease in the plausibility thatA is false:

(A|C′) < (A|C). (1.37)

This qualitative requirement simply gives the ‘sense of direction’ in which the robot’sreasoning is to go; it says nothing abouthow muchthe plausibilities change, except thatour continuity assumption (which is also a condition for qualitative correspondence withcommon sense) now requires that ifA|C changes only infinitesimally, it can induce only aninfinitesimal change inAB|C andA|C. The specific ways in which we use these qualitativerequirements will be given in the next chapter, at the point where it is seen why we needthem. For the present we summarize them simply as:

(II) Qualitative correspondence with common sense. (1.38)

Finally, we want to give our robot another desirable property for which honest people strivewithout always attaining: that it always reasonsconsistently. By this we mean just the three

Page 28: PROBABILITY THEORY THE LOGIC OF SCIENCE

1 Plausible reasoning 19

common colloquial meanings of the word ‘consistent’:

(IIIa)If a conclusion can be reasoned out in more than one way, thenevery possible way must lead to the same result.

(1.39a)

(IIIb)

The robot always takes into account all of the evidence it hasrelevant to a question. It does not arbitrarily ignore some ofthe information, basing its conclusions only on what remains.In other words, the robot is completely nonideological.

(1.39b)

(IIIc)

The robot always represents equivalent states of knowledge byequivalent plausibility assignments. That is, if in two problemsthe robot’s state of knowledge is the same (except perhaps forthe labeling of the propositions), then it must assign the sameplausibilities in both.

(1.39c)

Desiderata (I), (II), and (IIIa) are the basic ‘structural’ requirements on the inner workingsof our robot’s brain, while (IIIb) and(IIIc) are ‘interface’ conditions which show how therobot’s behavior should relate to the outer world.At this point,most students are surprised to learn that our search for desiderata is at an end.

The above conditions, it turnsout, uniquely determine the rules by which our robot mustreason; i.e. there is only one set of mathematical operations for manipulating plausibilitieswhich has all these properties. These rules are deduced in Chapter 2.

(At the end of most chapters, we insert a sectionof informal Comments in whicharecollected various side remarks, backgroundmaterial, etc. The readermay skip themwithoutlosing the main thread of the argument.)

1.8 Comments

As politicians, advertisers, salesmen, and propagandists for various political, economic,moral, religious, psychic, environmental, dietary, and artistic doctrinaire positions knowonly too well, fallible human minds are easily tricked, by clever verbiage, into committingviolations of the above desiderata. We shall try to ensure that they do not succeed with ourrobot.We emphasize another contrast between the robot and a human brain. By Desideratum

I, the robot’s mental state about any proposition is to be represented by a real number.Now, it is clear that our attitude toward any given proposition may have more than one‘coordinate’. You and I form simultaneous judgments about a proposition not only as towhether it is plausible, but also whether it is desirable, whether it is important, whether itis useful, whether it is interesting, whether it is amusing, whether it is morally right, etc.If we assume that each of these judgments might be represented by a number, then a fullyadequate description of a human state of mind would be represented by a vector in a spaceof a rather large number of dimensions.

Page 29: PROBABILITY THEORY THE LOGIC OF SCIENCE

20 Part 1 Principles and elementary applications

Not all propositions require this. For example, the proposition ‘The refractive index ofwater is less than 1.3’ generates no emotions; consequently the state of mind which itproduces has very few coordinates. On the other hand, the proposition, ‘Your mother-in-law just wrecked your new car’ generates a state of mind with many coordinates. Quitegenerally, the situations of everyday life are those involving many coordinates. It is just forthis reason, we suggest, that the most familiar examples of mental activity are often themost difficult to reproduce by a model. Perhaps we have here the reason why science andmathematics are the most successful of human activities: they deal with propositions whichproduce the simplest of all mental states. Such states would be the ones least perturbed bya given amount of imperfection in the human mind.Of course, for many purposes we would not want our robot to adopt any of these more

‘human’ features arising from the other coordinates. It is just the fact that computers donotget confused by emotional factors, donotget bored with a lengthy problem, donotpursuehidden motives opposed to ours, that makes them safer agents than men for carrying outcertain tasks.These remarks are interjected to point out that there is a large unexplored area of possible

generalizations and extensions of the theory to be developed here; perhaps this may inspireothers to try their hand at developing ‘multidimensional theories’ of mental activity, whichwould more and more resemble the behavior of actual human brains – not all of whichis undesirable. Such a theory, if successful, might have an importance beyond our presentability to imagine.5

For the present, however, we shall have to be content with amuchmoremodest undertak-ing. Is it possible to develop a consistent ‘one-dimensional’ model of plausible reasoning?Evidently, our problemwill be simplest if wecanmanage to represent adegreeof plausibilityuniquely by a single real number, and ignore the other ‘coordinates’ just mentioned.We stress that we are in no way asserting that degrees of plausibility in actual human

minds have a unique numericalmeasure. Our job is not to postulate – or indeed to conjectureabout – any such thing; it is toinvestigatewhether it is possible, in our robot, to set up sucha correspondence without contradictions.But to some it may appear that we have already assumed more than is necessary, thereby

puttinggratuitous restrictionson thegeneralityof our theory.Whymustwe representdegreesof plausibility by real numbers? Would not a ‘comparative’ theory based on a system ofqualitative ordering relations suchas (A|C) > (B|C) suffice? This point is discussed furtherinAppendixA,wherewedescribe other approaches to probability theory and note that someattempts have been made to develop comparative theories which it was thought would belogically simpler, or more general. But this turned out not to be the case; so, although it isquite possible to develop the foundations in other ways than ours, the final results will notbe different.

5 Indeed, some psychologists think that as few as five dimensions might suffice to characterize a human personality; that is, thatwe all differ only in having different mixes of five basic personality traits which may be genetically determined. But it seems tous that this must be grossly oversimplified; identifiable chemical factors continuously varying in both space and time (such asthe distribution of glucose metabolism in the brain) affect mental activity but cannot be represented faithfully in a space of onlyfive dimensions. Yet it may be that five numbers can capture enough of the truth to be useful for many purposes.

Page 30: PROBABILITY THEORY THE LOGIC OF SCIENCE

1 Plausible reasoning 21

1.8.1 Common language vs. formal logic

We should note the distinction between the statements of formal logic and those of ordinarylanguage. It might be thought that the latter is only a less precise form of expression; but onexamination of details the relation appears different. It appears to us that ordinary language,carefully used, need not be less precise than formal logic; but ordinary language is morecomplicated in its rules and has consequently richer possibilities of expression than weallow ourselves in formal logic.In particular, common language, being in constant use for other purposes than logic, has

developed subtle nuances – means of implying something without actually stating it – thatare lost on formal logic. Mr A, to affirm his objectivity, says, ‘I believe what I see.’ Mr Bretorts: ‘He doesn’t see what he doesn’t believe.’ From the standpoint of formal logic, itappears that they have said the same thing; yet from the standpoint of common language,those statements had the intent and effect of conveying opposite meanings.Here is a less trivial example, taken from amathematics textbook. Let L be a straight line

in a plane, and S an infinite set of points in that plane, each of which is projected onto L.Now consider the following statements:

(I) The projection of the limit is the limit of the projections.(II) The limit of the projections is the projection of the limit.

These have the grammatical structures ‘A is B’ and ‘B is A’, and so they might appearlogically equivalent. Yet in that textbook, (I) was held to be true, and (II) not true in general,on the grounds that the limit of the projections may exist when the limit of the set does not.As we see from this, in common language – even in mathematics textbooks – we have

learned to readsubtle nuancesofmeaning into theexact phrasing, probablywithout realizingit until an example like this is pointed out. We interpret ‘A is B’ as asserting first of all,as a kind of major premise, thatA exists; and the rest of the statement is understood tobe conditional on that premise. Put differently, in common grammar the verb ‘is’ impliesa distinction between subject and object, which the symbol ‘=’ does not have in formallogic or in conventional mathematics. (However, in computer languages we encounter suchstatements as ‘J= J+ 1’, which everybody seems to understand, but in which the ‘=’ signhas now acquired that implied distinction after all.)Another amusing example is the old adage ‘knowledge is power’, which is a very cogent

truth, both in human relations and in thermodynamics. An ad writer for a chemical tradejournal6 fouled this up into ‘power is knowledge’, an absurd – indeed, obscene – falsity.

These examples remind us that the verb ‘is’ has, like any other verb, a subject and apredicate; but it is seldom noted that this verb has two entirely different meanings. A personwhose native language is English may require some effort to see the different meanings inthe statements: ‘The room is noisy’ and ‘There is noise in the room’. But in Turkish thesemeanings are rendered by different words, whichmakes the distinction so clear that a visitor

6 LC-CG Magazine, March 1988, p. 211.

Page 31: PROBABILITY THEORY THE LOGIC OF SCIENCE

22 Part 1 Principles and elementary applications

who uses the wrong word will not be understood. The latter statement is ontological, assert-ing the physical existence of something, while the former is epistemological, expressingonly the speaker’s personal perception.Common language – or, at least, the English language – has an almost universal tendency

to disguise epistemological statements by putting them into a grammatical form which sug-gests to the unwary an ontological statement. A major source of error in current probabilitytheory arises from an unthinking failure to perceive this. To interpret the first kind of state-ment in the ontological sense is to assert that one’s own private thoughts and sensations arerealities existing externally in Nature. We call this the ‘mind projection fallacy’, and notethe trouble it causesmany times in what follows. But this trouble is hardly confined to prob-ability theory; as soon as it is pointed out, it becomes evident that much of the discourse ofphilosophers and Gestalt psychologists, and the attempts of physicists to explain quantumtheory, are reduced to nonsense by the author falling repeatedly into the mind projectionfallacy.These examples illustrate the care thatis needed when we try to translate the complex

statements of common language into the simpler statements of formal logic. Of course,common language is often less precise than we should want in formal logic. But everybodyexpects this and is on the lookout for it, so it is less dangerous.It is too much to expect that our robot will grasp all the subtle nuances of common

language, which a human spends perhaps 20 years acquiring. In this respect, our robot willremain like a small child – it interprets all statements literally and blurts out the truthwithoutthought of whom this may offend.It is unclear to the writer how difficult – and even less clear how desirable – it would be

to design a newer model robot with the ability to recognize these finer shades of meaning.Of course, the question of principle is disposed of at once by the existence of the humanbrain, which does this. But, in practice, von Neumann’s principle applies; a robot designedby us cannot do it until someone develops a theory of ‘nuance recognition’, which reducesthe process to a definitely prescribed set of operations. This we gladly leave to others.In any event, our present model robot is quite literally real, because today it is almost

universally true that any nontrivial probability evaluation is performed by a computer. Theperson who programmed that computer was necessarily, whether or not they thought of itthat way, designing part of the brain of a robot according to some preconceived notion ofhow the robot should behave. But very few of the computer programs now in use satisfy allour desiderata; indeed, most are intuitivead hocprocedures that were not chosen with anywell-defined desiderata at all in mind.Any such adhockery is presumably usable within some special area of application –

that was the criterion for choosing it – but as the proofs of Chapter 2 will show, anyadhockery which conflicts with the rules of probability theory must generate demonstrableinconsistencies when we try to apply it beyond some restricted area. Our aim is to avoidthis by developing the general principles of inference once and for all, directly from therequirement of consistency, and in a form applicable to any problem of plausible inferencethat is formulated in a sufficiently unambiguous way.